Theorems in Algebra

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 234

Theorems in Algebra

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information. PDF generated at: Sun, 16 Oct 2011 18:55:48 UTC

Contents
Articles
Abel's binomial theorem AbelRuffini theorem Abhyankar's conjecture Acyclic model Ado's theorem AlperinBrauerGorenstein theorem AmitsurLevitzki theorem Artin approximation theorem ArtinWedderburn theorem ArtinZorn theorem BaerSuzuki theorem BeauvilleLaszlo theorem Binomial inverse theorem Binomial theorem Birch's theorem Birkhoff's representation theorem Boolean prime ideal theorem BorelWeil theorem BorelWeilBott theorem Brauer's theorem on induced characters Brauer's three main theorems BrauerCartanHua theorem BrauerNesbitt theorem BrauerSiegel theorem BrauerSuzuki theorem BrauerSuzukiWall theorem Burnside theorem Cartan's theorem CartanDieudonn theorem Cauchy's theorem (group theory) Cayley's theorem CayleyHamilton theorem ChevalleyShephardTodd theorem ChevalleyWarning theorem 1 1 4 5 7 8 9 10 11 12 12 13 15 16 24 25 29 32 34 36 37 39 40 40 41 42 42 43 44 44 46 48 57 59

Classification of finite simple groups Cohn's irreducibility criterion Cramer's rule Crystallographic restriction theorem Descartes' rule of signs Dirichlet's unit theorem Engel theorem Factor theorem FeitThompson theorem Fitting's theorem Focal subgroup theorem Frobenius determinant theorem Frobenius theorem (real division algebras) Fundamental lemma (Langlands program) Fundamental theorem of algebra Fundamental theorem of cyclic groups Fundamental theorem of Galois theory Fundamental theorem of linear algebra Fundamental theorem on homomorphisms GilmanGriess theorem Going up and going down Goldie's theorem GolodShafarevich theorem GorensteinHarada theorem Gromov's theorem on groups of polynomial growth Grushko theorem Haboush's theorem Hahn embedding theorem Hajs's theorem Harish-Chandra isomorphism Hasse norm theorem HasseArf theorem Hilbert's basis theorem Hilbert's irreducibility theorem Hilbert's Nullstellensatz Hilbert's syzygy theorem Hilbert's Theorem 90 HopkinsLevitzki theorem

61 68 69 76 81 83 86 87 88 91 91 95 96 98 99 107 109 112 114 115 115 117 118 119 120 121 124 126 127 128 130 130 132 134 135 137 137 138

Hurwitz's theorem (normed division algebras) Isomorphism extension theorem Isomorphism theorem Jacobson density theorem Jordan's theorem (symmetric group) JordanSchur theorem Krull's principal ideal theorem KrullSchmidt theorem Knneth theorem Kurosh subgroup theorem Lagrange's theorem (group theory) LaskerNoether theorem Latimer-MacDuffee theorem Lattice theorem Levitzky's theorem Lie's third theorem LieKolchin theorem Maschke's theorem Milnor conjecture MordellWeil theorem Multinomial theorem NielsenSchreier theorem PerronFrobenius theorem PoincarBirkhoffWitt theorem Polynomial remainder theorem Primitive element theorem QuillenSuslin theorem Rational root theorem Regev's theorem Schreier refinement theorem SchurZassenhaus theorem SerreSwan theorem SkolemNoether theorem Specht's theorem Stone's representation theorem for Boolean algebras Structure theorem for finitely generated modules over a principal ideal domain Subgroup test Subring test

139 140 140 145 147 148 149 149 151 154 155 157 160 160 161 162 162 164 165 166 167 170 172 184 187 188 190 191 193 194 194 195 196 197 199 200 204 205

Sylow theorems Sylvester's determinant theorem Sylvester's law of inertia Takagi existence theorem Three subgroups lemma Trichotomy theorem Walter theorem Wedderburn's little theorem Weil conjecture on Tamagawa numbers Witt's theorem Z* theorem Zassenhaus lemma ZJ theorem

205 211 211 213 214 215 216 217 218 219 220 221 222

References
Article Sources and Contributors Image Sources, Licenses and Contributors 223 228

Article Licenses
License 229

Abel's binomial theorem

Abel's binomial theorem


Abel's binomial theorem, named after Niels Henrik Abel, states the following:

Example
m=2

References
Weisstein, Eric W., "Abel's binomial theorem [1]" from MathWorld.

References
[1] http:/ / mathworld. wolfram. com/ AbelsBinomialTheorem. html

AbelRuffini theorem
In algebra, the AbelRuffini theorem (also known as Abel's impossibility theorem) states that there is no general algebraic solutionthat is, solution in radicals to polynomial equations of degree five or higher.[1]

Interpretation
The content of this theorem is frequently misunderstood. It does not assert that higher-degree polynomial equations are unsolvable. In fact, the opposite is true: every non-constant polynomial equation in one unknown, with real or complex coefficients, has at least one complex number as solution; this is the fundamental theorem of algebra. Although the solutions cannot always be expressed exactly with radicals, they can be computed to any desired degree of accuracy using numerical methods such as the NewtonRaphson method or Laguerre method, and in this way they are no different from solutions to polynomial equations of the second, third, or fourth degrees. The theorem only concerns the form that such a solution must take. The theorem says that not all solutions of higher-degree equations can be obtained by starting with the equation's coefficients and rational constants, and repeatedly forming sums, differences, products, quotients, and radicals (n-th roots, for some integer n) of previously obtained numbers. This clearly excludes the possibility of having any formula that expresses the solutions of an arbitrary equation of degree 5 or higher in terms of its coefficients, using only those operations, or even of having different formulas for different roots or for different classes of polynomials, in such a way as to cover all cases. (In principle one could imagine formulas using irrational numbers as constants, but even if a finite number of those were admitted at the start, not all roots of higher-degree equations could be obtained.) However some polynomial equations, of arbitrarily high degree, are solvable with such operations. Indeed if the roots happen to be rational

AbelRuffini theorem numbers, they can trivially be expressed as constants. The simplest nontrivial example is the equation solutions are

2 , whose

Here the expression different possible values of

, which appears to involve the use of the exponential function, in fact just gives the (the n-th roots of unity), so it involves only extraction of radicals.

Lower-degree polynomials
The solutions of any second-degree polynomial equation can be expressed in terms of addition, subtraction, multiplication, division, and square roots, using the familiar quadratic formula: The roots of the following equation are shown below:

Analogous formulas for third- and fourth-degree equations, using cube roots and fourth roots, had been known since the 16th century.

Quintics and higher


The AbelRuffini theorem says that there are some fifth-degree equations whose solution cannot be so expressed. The equation is an example. (See Bring radical.) Some other fifth degree equations can be solved by radicals, for example , which factorizes to . The precise criterion that distinguishes between those equations that can be solved by radicals and those that cannot was given by variste Galois and is now part of Galois theory: a polynomial equation can be solved by radicals if and only if its Galois group (over the rational numbers, or more generally over the base field of admitted constants) is a solvable group. Today, in the modern algebraic context, we say that second, third and fourth degree polynomial equations can always be solved by radicals because the symmetric groups S2, S3 and S4 are solvable groups, whereas Sn is not solvable for n 5. This is so because for a polynomial of degree n with indeterminate coefficients (i.e., given by symbolic parameters), the Galois group is the full symmetric group Sn (this is what is called the "general equation of the n-th degree"). This remains true if the coefficients are concrete but algebraically independent values over the base field.

Proof
The following proof is based on Galois theory. Historically, Ruffini and Abel's proofs precede Galois theory. One of the fundamental theorems of Galois theory states that an equation is solvable in radicals if and only if it has a solvable Galois group, so the proof of the AbelRuffini theorem comes down to computing the Galois group of the general polynomial of the fifth degree. Let be a real number transcendental over the field of rational numbers , and so on to which is transcendental over and let : , and let be a real number . These numbers are

transcendental over

called independent transcendental elements over Q. Let Multiplying out yields the elementary symmetric functions of the

AbelRuffini theorem and so on up to

The coefficient of indeterminates over on that leaves

in

is thus

. Because our independent transcendentals in the symmetric group on 5 letters

act as

, every permutation

induces an automorphism

fixed and permutes the elements

. Since an arbitrary rearrangement of the roots of the

product form still produces the same polynomial, e.g.: is still the same polynomial as

the automorphisms

also leave

fixed, so they are elements of the Galois group

. Now, since . , and so

it must be that must be isomorphic to

, as there could possibly be automorphisms there that are not in

However, since the splitting field of a quintic polynomial has at most 5! elements, polynomial of degree is isomorphic to . And what of ? The only composition series of is not an abelian group, and so

. Generalizing this argument shows that the Galois group of every general is (where is the alternating group on (isomorphic to itself)

five letters, also known as the icosahedral group). However, the quotient group

is not solvable, so it must be that the general polynomial of the fifth degree has no

solution in radicals. Since the first nontrivial normal subgroup of the symmetric group on n letters is always the alternating group on n letters, and since the alternating groups on n letters for are always simple and non-abelian, and hence not solvable, it also says that the general polynomials of all degrees higher than the fifth also have no solution in radicals. Note that the above construction of the Galois group for a fifth degree polynomial only applies to the general polynomial, specific polynomials of the fifth degree may have different Galois groups with quite different properties, e.g. has a splitting field generated by a primitive 5th root of unity, and hence its Galois group is abelian and the equation itself solvable by radicals. However, since the result is on the general polynomial, it does say that a general "quintic formula" for the roots of a quintic using only a finite combination of the arithmetic operations and radicals in terms of the coefficients is impossible. Q.E.D.

History
Around 1770, Joseph Louis Lagrange began the groundwork that unified the many different tricks that had been used up to that point to solve equations, relating them to the theory of groups of permutations, in the form of Lagrange resolvents. This innovative work by Lagrange was a precursor to Galois theory, and its failure to develop solutions for equations of fifth and higher degrees hinted that such solutions might be impossible, but it did not provide conclusive proof. The theorem, however, was first nearly proved by Paolo Ruffini in 1799, but his proof was mostly ignored. He had several times tried to send it to different mathematicians to get it acknowledged, amongst them, French mathematician Augustin-Louis Cauchy, but it was never acknowledged, possibly because the proof was spanning 500 pages. The proof also, as was discovered later, contained an error. Ruffini assumed that a solution would necessarily be a function of the radicals (in modern terms, he failed to prove that the splitting field is one of the fields in the tower of radicals which corresponds to a solution expressed in radicals). While Cauchy felt that the assumption was minor, most historians believe that the proof was not complete until Abel proved this assumption. The theorem is thus generally credited to Niels Henrik Abel, who published a proof that required just six pages in 1824.[2] Insights into these issues were also gained using Galois theory pioneered by variste Galois. In 1885, John Stuart Glashan, George Paxton Young, and Carl Runge provided a proof using this theory.

AbelRuffini theorem

Notes
[1] Jacobson (2009), p. 211. [2] du Sautoy, Marcus. "January: Impossibilities". Symmetry: A Journey into the Patterns of Nature. ISBN978-0060789411.

References
Edgar Dehn. Algebraic Equations: An Introduction to the Theories of Lagrange and Galois. Columbia University Press, 1930. ISBN 0-486-43900-3. Jacobson, Nathan (2009), Basic algebra, 1 (2nd ed.), Dover, ISBN978-0-486-47189-1 John B. Fraleigh. A First Course in Abstract Algebra. Fifth Edition. Addison-Wesley, 1994. ISBN 0-201-59291-6. Ian Stewart. Galois Theory. Chapman and Hall, 1973. ISBN 0-412-10800-3. Abel's Impossibility Theorem at Everything2 (http://www.everything2.net/title/Abel%27s+Impossibility+ Theorem)

External links
MMOIRE SUR LES QUATIONS'ALGBRIQUES, OU L'ON DMONTRE. L'IMPOSSIBILIT DE LA RSOLUTION DE L'QUATION GNRALE. DU CINQUIME DEGR (http://www.abelprisen.no/ verker/oeuvres_1881_del1/oeuvres_completes_de_abel_nouv_ed_1_kap03_opt.pdf)PDF - the first proof on 1824 in French Dmonstration de l'impossibilit de la rsolution algbrique des quations gnrales qui passent le quatrime degr (http://www.abelprisen.no/verker/oeuvres_1839/oeuvres_completes_de_abel_1_kap02_opt.pdf)PDF the second proof on 1826 in French

Abhyankar's conjecture
In abstract algebra, Abhyankar's conjecture is a 1957 conjecture of Shreeram Abhyankar, on the Galois groups of function fields of characteristic p.[1] This problem was solved in 1994 by work of Michel Raynaud and David Harbater.[2] [3] The problem involves a finite group G, a prime number p, and a nonsingular integral algebraic curve C defined over an algebraically closed field K of characteristic p. The question addresses the existence of Galois extensions L of K(C), with G as Galois group, and with restricted ramification. From a geometric point of view L corresponds to another curve C, and a morphism : C C. Ramification geometrically, and by analogy with the case of Riemann surfaces, consists of a finite set S of points x on C, such that restricted to the complement of S in C is an tale morphism. In Abhyankar's conjecture, S is fixed, and the question is what G can be. This is therefore a special type of inverse Galois problem. The subgroup p(G) is defined to be the subgroup generated by all the Sylow subgroups of G for the prime number p. This is a normal subgroup, and the parameter n is defined as the minimum number of generators of G/p(G). Then for the case of C the projective line over K, the conjecture states that G can be realised as a Galois group of L, unramified outside S containing s + 1 points, if and only if n s. This was proved by Raynaud.

Abhyankar's conjecture For the general case, proved by Harbater, let g be the genus of C. Then G can be realised if and only if n s + 2 g.

References
[1] Abhyankar, Shreeram (1957), "Coverings of Algebraic Curves", American Journal of Mathematics 79 (4): 825856, doi:10.2307/2372438. [2] Raynaud, Michel (1994), "Revtements de la droite affine en caractristique p > 0", Inventiones Mathematicae 116 (1): 425462, doi:10.1007/BF01231568. [3] Harbater, David (1994), "Abhyankar's conjecture on Galois groups over curves", Inventiones Mathematicae 117 (1): 125, doi:10.1007/BF01232232.

External links
Weisstein, Eric W., " Abhyankar's conjecture (http://mathworld.wolfram.com/AbhyankarsConjecture.html)" from MathWorld. A layman's perspective of Abhyankar's conjecture (http://www.math.purdue.edu/about/purview/spring95/ conjecture.html) from Purdue University

Acyclic model
In algebraic topology, a discipline within mathematics, the acyclic models theorem can be used to show that two homology theories are isomorphic. The theorem was developed by topologists Samuel Eilenberg and Saunders MacLane. They discovered that, when topologists were writing proofs to establish equivalence of various homology theories, there were numerous similarities in the processes. Eilenberg and MacLane then discovered the theorem to generalize this process. It can be used to prove the EilenbergZilber theorem.

Statement of the theorem


Let be an arbitrary category and for be the category of chain complexes of -modules. Let be covariant functors such that: . for such that has a basis in -acyclic at these models, which means that

There are is - and

, so

is a free functor. for all and all

. Then the following assertions hold: Every natural transformation If for all models In particular the chain map is induced by a natural chain map . are natural transformations, are natural chain maps as before and , then there is a natural chain homotopy between is unique up to natural chain homotopy.[1] and .

Acyclic model

Generalizations
Projective and acyclic complexes
What is above is one of the earliest versions of the theorem. Another version is the one that says that if complex of projectives in an abelian category and extends to a chain map , unique up to homotopy. as the abelian category. Free being is a is an acyclic complex in that category, then any map

This specializes almost to the above theorem if one uses the functor category

functors are projective objects in that category. The morphisms in the functor category are natural transformations, so the constructed chain maps and homotopies are all natural. The difference is that in the above version, acyclic is a stronger assumption than being acyclic only at certain objects. On the other hand, the above version almost implies this version by letting the free functor is basically just free (and hence projective) module. is acyclic. one) means nothing else than that the complex

a category with only one object. Then

being acyclic at the models (there is only

Acyclic classes
Then there is the grand theorem that unifies them all. Let be an abelian category (for example or ). A class of chain complexes over will be called an acyclic class provided: The 0 complex is in . The complex belongs to if and only if the suspension of does. If the complexes and are homotopic and , then . Every complex in is acyclic. If is a double complex, all of whose rows are in , then the total complex of

belongs to

There are three natural examples of acyclic classes, although doubtless others exist. The first is that of homotopy contractible complexes. The second is that of acyclic complexes. In functor categories (e.g. the category of all functors from topological spaces to abelian groups), there is a class of complexes that are contractible on each object, but where the contractions might not be given by natural transformations. Another example is again in functor categories but this time the complexes are acyclic only at certain objects. Let denote the class of chain maps between complexes whose mapping cone belongs to . Although does not necessarily have a calculus of either right or left fractions, it has weaker properties of having homotopy classes of both left and right fractions that permit forming the class gotten by inverting the arrows in . Let be an augmented endofunctor on , meaning there is given a natural transformation is -presentable if for each , the chain complex (the identity functor on ). We say that the chain complex

belongs to

. The boundary operator is given by .

We say that the chain complex functor belongs to . Theorem. Let Suppose that be an acyclic class and is -presentable

is

-acyclic if the augmented chain complex

the corresponding class of arrows in the category of chain complexes. and is -acyclic. Then any natural transformation is to a natural transformation of chain functors

extends, in the category and this is unique in -presentable, that is -acyclic, and that

up to chain homotopies. If we suppose, in addition, that is an isomorphism, then is homotopy equivalence.

Acyclic model

Example
Here is an example of this last theorem in action. Let category of abelian group valued functors on simplicial chain complex functor. Let . Here, each is the . Let be the category of triangulable spaces and be the singular chain complex functor and be the functor that assigns to each space -simplex and this functor assigns to . Then let be defined by . It can be shown that both be the be the the space

the sum of as many copies of . There is an obvious and are both

-simplex as there are maps -acyclic (the proof that

augmentation -presentable and

and this induces one on

is not entirely straigtforward and uses a detour through simplicial is the class of homology and so we conclude that singular and simplicial

subdivision, which can also be handled using the above theorem). The class equivalences. It is rather obvious that

homology are isomorphic on . There are many other examples in both algebra and topology, some of which are described in M. Barr, Acyclic Models. AMS, 2002.

References
[1] Dold, Albrecht (1980), Lectures on Algebraic Topology, A Series of Comprehensive Studies in Mathematics, 200 (2nd ed.), Berlin, New York: Springer-Verlag, ISBN3-540-10369-4

Ado's theorem
In abstract algebra, Ado's theorem states that every finite-dimensional Lie algebra L over a field K of characteristic zero can be viewed as a Lie algebra of square matrices under the commutator bracket. More precisely, the theorem states that L has a linear representation over K, on a finite-dimensional vector space V, that is a faithful representation, making L isomorphic to a subalgebra of the endomorphisms of V. While for the Lie algebras associated to classical groups there is nothing new in this, the general case is a deeper result. Applied to the real Lie algebra of a Lie group G, it does not imply that G has a faithful linear representation (which is not true in general), but rather that G always has a linear representation that is a local isomorphism with a linear group. It was proved in 1935 by Igor Dmitrievich Ado of Kazan State University, a student of Nikolai Chebotaryov. The restriction on the characteristic was removed later, by Iwasawa and Harish-Chandra.

References
I. D. Ado, Note on the representation of finite continuous groups by means of linear substitutions, Izv. Fiz.-Mat. Obsch. (Kazan'), 7 (1935) pp.143 (Russian language) Ado, I. D. (1947), "The representation of Lie algebras by matrices" [1] (in Russian), Akademiya Nauk SSSR i Moskovskoe Matematicheskoe Obshchestvo. Uspekhi Matematicheskikh Nauk 2 (6): 159173, ISSN0042-1316, MR0027753 translation in Ado, I. D. (1949), "The representation of Lie algebras by matrices", American Mathematical Society Translations 1949 (2): 21, ISSN0065-9290, MR0030946 Iwasawa, Kenkichi (1948), "On the representation of Lie algebras", Jap. J. Math. 19: 405426, MR0032613 Harish-Chandra (1949), "Faithful representations of Lie algebras", Annals of Mathematics. Second Series 50: 6876, ISSN0003-486X, JSTOR1969352, MR0028829 Nathan Jacobson, Lie Algebras, pp.202203

Ado's theorem

References
[1] http:/ / mi. mathnet. ru/ eng/ umn/ v2/ i6/ p159

AlperinBrauerGorenstein theorem
In mathematics, the AlperinBrauerGorenstein theorem characterizes the finite simple groups with quasidihedral or wreathed[1] Sylow 2-subgroups. These are isomorphic either to three-dimensional projective special linear groups or projective special unitary groups over a finite fields of odd order, depending on a certain congruence, or to the Mathieu group . Alperin, Brauer & Gorenstein (1970) proved this in the course of 261 pages. The subdivision by 2-fusion is sketched there, given as an exercise in Gorenstein (1968, Ch. 7), and presented in some detail in Kwon et al. (1980).

Notes
[1] A 2-group is wreathed if it is a nonabelian semidirect product of a maximal subgroup that is a direct product of two cyclic groups of the same order, that is, if it is the wreath product of a cyclic 2-group with the symmetric group on 2 points.

References
Alperin, J. L.; Brauer, R.; Gorenstein, D. (1970), "Finite groups with quasi-dihedral and wreathed Sylow 2-subgroups.", Transactions of the American Mathematical Society (American Mathematical Society) 151 (1): 1261, doi:10.2307/1995627, ISSN0002-9947, JSTOR1995627, MR0284499 Gorenstein, D. (1968), Finite groups, Harper & Row Publishers, MR0231903 Kwon, T.; Lee, K.; Cho, I.; Park, S. (1980), "On finite groups with quasidihedral Sylow 2-groups" (http://kms. or.kr/home/journal/include/downloadPdfJournal. asp?articleuid={71EE4232-6997-4030-8CA7-85CDBCB5A2CC}), Journal of the Korean Mathematical Society 17 (1): 9197, ISSN0304-9914, MR593804

AmitsurLevitzki theorem

AmitsurLevitzki theorem
In algebra, the AmitsurLevitzki theorem states that the algebra of n by n matrices satisfies a certain identity of degree 2n. It was proved by Amitsur and Levitsky(1950). In particular matrix rings are PI rings such that the smallest identity they satisfy has degree exactly 2n.

Statement
If A1,...,A2n are n by n matrices then

where the sum is over all (2n)! elements of the symmetric group S2n. (This polynomial is called the standard polynomial of degree 2n.)

Proofs
Amitsur and Levitsky(1950) gave the first proof. Kostant (1958) deduced the AmitsurLevitzki theorem from the KoszulSamelson theorem about primitive cohomology of Lie algebras. Swan (1963) and Swan (1969) gave a simple combinatorial proof as follows. By linearity it is enough to prove the theorem when each matrix has only one nonzero entry, which is 1. In this case each matrix can be encoded as a directed edge of a graph with n vertices. So all matrices together give a graph on n vertices with 2n directed edges. The identity holds provided that for any two vertices A and B of the graph, the number of odd Eulerian paths from A to B is the same as the number of even ones. (Here a path is called odd or even depending on whether its edges taken in order give an odd or even permutation of the 2n edges.) Swan showed that this was the case provided the number of edges in the graph is at least 2n, thus proving the AmitsurLevitzki theorem. Razmyslov (1974) gave a proof related to the CayleyHamilton theorem. Rosset (1976) gave a short proof using the exterior algebra of a vector space of dimension 2n.

References
Amitsur, A. S.; Levitzki, Jakob (1950), "Minimal identities for algebras" [1], Proceedings of the American Mathematical Society 1: 449463, ISSN0002-9939, JSTOR2032312, MR0036751 Amitsur, A. S.; Levitzki, Jakob (1951), "Remarks on Minimal identities for algebras" [2], Proceedings of the American Mathematical Society 2: 320327, ISSN0002-9939, JSTOR2032509, MR? Formanek, E. (2001), "AmitsurLevitzki theorem" [3], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 Kostant, Bertram (1958), "A theorem of Frobenius, a theorem of AmitsurLevitski and cohomology theory", J. Math. Mech. 7: 237264, doi:10.1512/iumj.1958.7.07019, MR0092755 Razmyslov, Ju. P. (1974), "Identities with trace in full matrix algebras over a field of characteristic zero", Mathematics of the USSR-Izvestiya 8 (4): 727, doi:10.1070/IM1974v008n04ABEH002126, ISSN0373-2436, MR0506414 Rosset, Shmuel (1976), "A new proof of the AmitsurLevitski identity", Israel Journal of Mathematics 23 (2): 187188, doi:10.1007/BF02756797, ISSN0021-2172, MR0401804 Swan, Richard G. (1963), "An application of graph theory to algebra" [4], Proceedings of the American Mathematical Society 14: 367373, ISSN0002-9939, JSTOR2033801, MR0149468

AmitsurLevitzki theorem Swan, Richard G. (1969), "Correction to "An application of graph theory to algebra"" [5], Proceedings of the American Mathematical Society 21: 379380, ISSN0002-9939, JSTOR?, MR0255439

10

References
[1] [2] [3] [4] [5] http:/ / www. ams. org/ journals/ proc/ 1950-001-04/ S0002-9939-1950-0036751-9/ S0002-9939-1950-0036751-9. pdf http:/ / www. ams. org/ journals/ proc/ 1951-002-02/ S0002-9939-1951-0040285-6/ S0002-9939-1951-0040285-6. pdf http:/ / eom. springer. de/ a/ a110570. htm http:/ / www. ams. org/ journals/ proc/ 1963-014-03/ S0002-9939-1963-0149468-6/ S0002-9939-1963-0149468-6. pdf http:/ / www. ams. org/ journals/ proc/ 1969-021-02/ S0002-9939-1969-0255439-7/ S0002-9939-1969-0255439-7. pdf

Artin approximation theorem


In mathematics, the Artin approximation theorem is a fundamental result of Michael Artin in deformation theory which implies that formal power series with coefficients in a field k are well-approximated by the algebraic functions on k.

Statement of the theorem


Let x = x1, , xn denote a collection of n indeterminates, k[[x]] the ring of formal power series with indeterminates x over a field k, and y = y1, , ym a different set of indeterminates. Let f(x, y) = 0 be a system of polynomial equations in k[x, y], and c a positive integer. Then given a formal power series solution (x) k[[x]] there is an algebraic solution y(x) consisting of algebraic functions such that (x) y(x) mod (x)c.

Discussion
Given any desired positive integer c, this theorem shows that one can find an algebraic solution approximating a formal power series solution up to the degree specified by c. This leads to theorems that deduce the existence of certain formal moduli spaces of deformations as schemes.

References
Artin, Michael. Algebraic Spaces. Yale University Press, 1971.

ArtinWedderburn theorem

11

ArtinWedderburn theorem
In abstract algebra, the ArtinWedderburn theorem is a classification theorem for semisimple rings. The theorem states that an Artinian semisimple ring R is isomorphic to a product of finitely many ni-by-ni matrix rings over division rings Di, for some integers ni, both of which are uniquely determined up to permutation of the index i. In particular, any simple left or right Artinian ring is isomorphic to an n-by-n matrix ring over a division ring D, where both n and D are uniquely determined. As a direct corollary, the ArtinWedderburn theorem implies that every simple ring that is finite-dimensional over a division ring (a simple algebra) is a matrix ring. This is Joseph Wedderburn's original result. Emil Artin later generalized it to the case of Artinian rings. Note that if R is a finite-dimensional simple algebra over a division ring E, D need not be contained in E. For example, matrix rings over the complex numbers are finite-dimensional simple algebras over the real numbers. The ArtinWedderburn theorem reduces classifying simple rings over a division ring to classifying division rings that contain a given division ring. This in turn can be simplified: The center of D must be a field K. Therefore R is a K-algebra, and itself has K as its center. A finite-dimensional simple algebra R is thus a central simple algebra over K. Thus the ArtinWedderburn theorem reduces the problem of classifying finite-dimensional central simple algebras to the problem of classifying division rings with given center.

Examples
Let R be the field of real numbers, C be the field of complex numbers, and H the quaternions. Every finite-dimensional simple algebra over R must be a matrix ring over R, C, or H. Every central simple algebra over R must be a matrix ring over R or H. These results follow from the Frobenius theorem. Every finite-dimensional simple algebra over C must be a matrix ring over C and hence every central simple algebra over C must be a matrix ring over C. Every finite-dimensional central simple algebra over a finite field must be a matrix ring over that field. Every commutative semisimple ring must be a finite direct product of fields.[1]

References
[1] This is clear since matrix rings larger than 11 are never commutative.

P. M. Cohn (2003) Basic Algebra: Groups, Rings, and Fields, pages 1379. J.H.M. Wedderburn (1908). "On Hypercomplex Numbers". Proceedings of the London Mathematical Society 6: 77118. doi:10.1112/plms/s2-6.1.77.

ArtinZorn theorem

12

ArtinZorn theorem
In mathematics, the ArtinZorn theorem, named after Emil Artin and Max Zorn, states that any finite alternative division ring is necessarily a finite field. It was first published by Zorn, but in his publication Zorn credited it to Artin.[1] [2] The ArtinZorn theorem is a generalization of the Wedderburn theorem, which states that finite associative division rings are fields. As a geometric consequence, every finite Moufang plane is the classical projective plane over a finite field.[3] [4]

References
[1] Zorn, M. (1930), "Theorie der alternativen Ringe", Abh. Math. Sem. Hamburg 8: 123147. [2] Lneburg, Heinz (2001), "On the early history of Galois fields", in Jungnickel, Dieter; Niederreiter, Harald, Finite fields and applications: proceedings of the Fifth International Conference on Finite Fields and Applications Fq5, held at the University of Augsburg, Germany, August 26, 1999, Springer-Verlag, pp.341355, ISBN9783540411093, MR1849100. [3] Shult, Ernest (2011), Points and Lines: Characterizing the Classical Geometries, Universitext, Springer-Verlag, p.123, ISBN9783642156267. [4] McCrimmon, Kevin (2004), A taste of Jordan algebras, Universitext, Springer-Verlag, p.34, ISBN9780387954479.

BaerSuzuki theorem
In mathematical finite group theory, the BaerSuzuki theorem, proved by Baer (1957) and Suzuki (1965), states that if any two elements of a conjugacy class C of a finite group generate a nilpotent subgroup, then all elements of the conjugacy class C are contained in a nilpotent subgroup. Alperin & Lyons (1971) gave a short elementary proof.

References
Alperin, J. L.; Lyons, Richard (1971), "On conjugacy classes of p-elements", Journal of Algebra 19: 536537, ISSN0021-8693, MR0289622 Baer, Reinhold (1957), "Engelsche elemente Noetherscher Gruppen", Mathematische Annalen 133: 256270, doi:10.1007/BF02547953, ISSN0025-5831, MR0086815 Gorenstein, D. (1980), Finite groups [1] (2nd ed.), New York: Chelsea Publishing Co., ISBN978-0-8284-0301-6, MR569209 Suzuki, Michio (1965), "Finite groups in which the centralizer of any element of order 2 is 2-closed", Annals of Mathematics. Second Series 82: 191212, ISSN0003-486X, JSTOR1970569, MR0183773

References
[1] http:/ / www. ams. org/ bookstore-getitem/ item=CHEL-301-H

BeauvilleLaszlo theorem

13

BeauvilleLaszlo theorem
In mathematics, the BeauvilleLaszlo theorem is a result in commutative algebra and algebraic geometry that allows one to "glue" two sheaves over an infinitesimal neighborhood of a point on an algebraic curve. It was proved by Arnaud Beauville and Yves Laszlo(1995).

The theorem
Although it has implications in algebraic geometry, the theorem is a local result and is stated in its most primitive form for commutative rings. If A is a ring and f is a nonzero element of A, then we can form two derived rings: the localization at f, Af, and the completion at Af, ; both are A-algebras. In the following we assume that f is a non-zero divisor. Geometrically, A is viewed as a scheme X = Spec A and f as a divisor (f) on Spec A; then Af is its complement Df = Spec Af, the principal open set determined by f, while is an "infinitesimal neighborhood" D = Spec of (f). The intersection of Df and Spec is a "punctured infinitesimal neighborhood" D0 about (f), equal to Spec A Af = Spec f. Suppose now that we have an A-module M; geometrically, M is a sheaf on Spec A, and we can restrict it to both the principal open set Df and the infinitesimal neighborhood Spec , yielding an Af-module F and an -module G. Algebraically,

(Despite the notational temptation to write

, meaning the completion of the A-module M at the ideal Af,

unless A is noetherian and M is finitely-generated, the two are not in fact equal. This phenomenon is the main reason that the theorem bears the names of Beauville and Laszlo; in the noetherian, finitely-generated case, it is, as noted by the authors, a special case of Grothendieck's faithfully flat descent.) F and G can both be further restricted to the punctured neighborhood D0, and since both restrictions are ultimately derived from M, they are isomorphic: we have an isomorphism

Now consider the converse situation: we have a ring A and an element f, and two modules: an Af-module F and an -module G, together with an isomorphism as above. Geometrically, we are given a scheme X and both an open set Df and a "small" neighborhood D of its closed complement (f); on Df and D we are given two sheaves which agree on the intersection D0 = Df D. If D were an open set in the Zariski topology we could glue the sheaves; the content of the BeauvilleLaszlo theorem is that, under one technical assumption on f, the same is true for the infinitesimal neighborhood D as well. Theorem: Given A, f, F, G, and as above, if G has no f-torsion, then there exist an A-module M and isomorphisms

consistent with the isomorphism : is equal to the composition

The technical condition that G has no f-torsion is referred to by the authors as "f-regularity". In fact, one can state a stronger version of this theorem. Let M(A) be the category of A-modules (whose morphisms are A-module homomorphisms) and let Mf(A) be the full subcategory of f-regular modules. In this notation, we obtain a commutative diagram of categories (note Mf(Af) = M(Af)):

BeauvilleLaszlo theorem in which the arrows are the base-change maps; for example, the top horizontal arrow acts on objects by M M A . Theorem: The above diagram is a cartesian diagram of categories.

14

Global version
In geometric language, the BeauvilleLaszlo theorem allows one to glue sheaves on an affine scheme over an infinitesimal neighborhood of a point. Since sheaves have a "local character" and since any scheme is locally affine, the theorem admits a global statement of the same nature. The version of this statement that the authors found noteworthy concerns vector bundles: Theorem: Let X be an algebraic curve over a field k, x a k-rational smooth point on X with infinitesimal neighborhood D = Spec k[[t]], R a k-algebra, and r a positive integer. Then the category Vectr(XR) of rank-r vector bundles on the curve XR = X Spec k Spec R fits into a cartesian diagram:

This entails a corollary stated in the paper: Corollary: With the same setup, denote by Triv(XR) the set of triples (E, , ), where E is a vector bundle on XR, is a trivialization of E over (X \ x)R (i.e., an isomorphism with the trivial bundle O(X - x)R), and a trivialization over DR. Then the maps in the above diagram furnish a bijection between Triv(XR) and GLr(R((t))) (where R((t)) is the formal Laurent series ring). The corollary follows from the theorem in that the triple is associated with the unique matrix which, viewed as a "transition function" over D0R between the trivial bundles over (X \ x)R and over DR, allows gluing them to form E, with the natural trivializations of the glued bundle then being identified with and . The importance of this corollary is that it shows that the affine Grassmannian may be formed either from the data of bundles over an infinitesimal disk, or bundles on an entire algebraic curve.

References
Beauville, Arnaud; Laszlo, Yves (1995), "Un lemme de descente" [1], Comptes Rendus de l'Acadmie des Sciences Srie I. Mathmatique 320 (3): 335340, ISSN0764-4442, retrieved 2008-04-08

References
[1] http:/ / math1. unice. fr/ ~beauvill/ pubs/ descente. pdf

Binomial inverse theorem

15

Binomial inverse theorem


In mathematics, the Binomial Inverse Theorem is useful for expressing matrix inverses in different ways. If A, U, B, V are matrices of sizes pp, pq, qq, qp, respectively, then

provided A and B + BVA1UB are nonsingular. Note that if B is invertible, the two B terms flanking the quantity inverse in the right-hand side can be replaced with (B1)1, which results in

This is the matrix inversion lemma, which can also be derived using matrix blockwise inversion.

Verification
First notice that

Now multiply the matrix we wish to invert by its alleged inverse

which verifies that it is the inverse. So we get thatif A1 and theorem above.
[1]

exist, then

exists and is given by the

Special cases
If p = q and U = V = Ip is the identity matrix, then Remembering the identity

we can also express the previous equation in the simpler form as

If B = Iq is the identity matrix and q = 1, then U is a column vector, written u, and V is a row vector, written vT. Then the theorem implies

This is useful if one has a matrix quickly. If we set A = Ip and B = Iq, we get In particular, if q = 1, then

with a known inverse A1 and one needs to invert matrices of the form A+uvT

Binomial inverse theorem

16

References
[1] Gilbert Strang (2003). Introduction to Linear Algebra (3rd edition ed.). Wellesley-Cambridge Press: Wellesley, MA. ISBN0-9614088-98.

Binomial theorem
In elementary algebra, the binomial theorem describes the algebraic expansion of powers of a binomial. According to the theorem, it is possible to expand the power (x+y)n into a sum involving terms of the form axbyc, where the exponents b and c are nonnegative integers with b + c = n, and the coefficient a of each term is a specific positive integer depending on n and b. When an exponent is zero, the corresponding power is usually omitted from the term. For example,
The binomial coefficients appear as the entries of Pascal's triangle.

The coefficient a in the term of xbyc is known as the binomial coefficient

or

(the two have the same value).

These coefficients for varying n and b can be arranged to form Pascal's triangle. These numbers also arise in combinatorics, where gives the number of different combinations of b elements that can be chosen from an n-element set.

History
This formula and the triangular arrangement of the binomial coefficients are often attributed to Blaise Pascal, who described them in the 17th century, but they were known to many mathematicians who preceded him. The 4th century B.C. Greek mathematician Euclid mentioned the special case of the binomial theorem for exponent2[1] [2] as did the 3rd century B.C. Indian mathematician Pingala to higher orders. A more general binomial theorem and the so-called "Pascal's triangle" were known in the 10th-century A.D. to Indian mathematician Halayudha and Persian mathematician Al-Karaji,[3] , in the 11th century to Persian poet and mathematician Omar Khayyam[4] , and in the 13th century to Chinese mathematician Yang Hui, who all derived similar results.[5] Al-Karaji also provided a mathematical proof of both the binomial theorem and Pascal's triangle, using mathematical induction.[3]

Statement of the theorem


According to the theorem, it is possible to expand any power of x+y into a sum of the form

where each

is a specific positive integer known as binomial coefficient. This formula is also referred to as the

Binomial Formula or the Binomial Identity. Using summation notation, it can be written as

The final expression follows from the previous one by the symmetry of x and y in the first expression, and by comparison it follows that the sequence of binomial coefficients in the formula is symmetrical.

Binomial theorem A variant of the binomial formula is obtained by substituting 1 for x and x for y, so that it involves only a single variable. In this form, the formula reads

17

or equivalently

Examples
The most basic example of the binomial theorem is the formula for the square of x+y:

Pascal's triangle

The binomial coefficients 1, 2, 1 appearing in this expansion correspond to the third row of Pascal's triangle. The coefficients of higher powers of x+y correspond to later rows of the triangle:

Notice that 1. the powers of x go down until it reaches 0 (none of x),starting value is n (the n in 2. the powers of y go up from 0 (none of y)until it reaches n (also the n in 3. the nth row of the Pascal's Triangle will be the coefficients of the expanded binomial.(Note that the top is row 0.) The binomial theorem can be applied to the powers of any binomial. For example,

For a binomial involving subtraction, the theorem can be applied as long as the opposite of the second term is used. This has the effect of changing the sign of every other term in the expansion:

Binomial theorem

18

Geometrical explanation
For positive values of a and b, the binomial theorem with n=2 is the geometrically evident fact that a square of side a + b can be cut into a square of side a, a square of side b, and two rectangles with sides a and b. With n=3, the theorem states that a cube of side a + b can be cut into a cube of side a, a cube of side b, three aab rectangular boxes, and three abb rectangular boxes.

In calculus, this picture also gives a geometric proof of the derivative volume of an n-dimensional hypercube, the area of the n faces, each of dimension

[6]

if one sets ) is

and

interpreting b as an infinitesimal change in a, then this picture shows the infinitesimal change in the where the coefficient of the linear term (in

Substituting this into the definition of the derivative via a difference quotient and taking limits means that the higher order terms and higher become negligible, and yields the formula interpreted as "the infinitesimal change in volume of an n-cube as side length varies is the area of n of its -dimensional faces". If one integrates this picture, which corresponds to applying the fundamental theorem of calculus, one obtains Cavalieri's quadrature formula, the integral see proof of Cavalieri's quadrature formula for details.[6]

Binomial theorem

19

The binomial coefficients


The coefficients that appear in the binomial expansion are called binomial coefficients. These are usually written , and pronounced n choose k.

Formulas
The coefficient of xnkyk is given by the formula , which is defined in terms of the factorial function n!. Equivalently, this formula can be written

with k factors in both the numerator and denominator of the fraction. Note that, although this formula involves a fraction, the binomial coefficient is actually an integer.

Combinatorial interpretation
The binomial coefficient can be interpreted as the number of ways to choose k elements from an n-element set. This is related to binomials for the following reason: if we write (x+y)n as a product then, according to the distributive law, there will be one term in the expansion for each choice of either x or y from each of the binomials of the product. For example, there will only be one term xn, corresponding to choosing 'x from each binomial. However, there will be several terms of the form xn2y2, one for each way of choosing exactly two binomials to contribute a y. Therefore, after combining like terms, the coefficient of xn2y2 will be equal to the number of ways to choose exactly 2 elements from an n-element set.

Proofs
Combinatorial proof
Example The coefficient of xy2 in

equals

because there are three x,y strings of length 3 with exactly two y's, namely,

corresponding to the three 2-element subsets of {1,2,3}, namely,

where each subset specifies the positions of the y in a corresponding string.

Binomial theorem General case Expanding (x+y)n yields the sum of the 2n products of the form e1e2...en where each ei is x ory. Rearranging factors shows that each product equals xnkyk for some k between 0 andn. For a given k, the following are proved equal in succession: the number of copies of xnkyk in the expansion the number of n-character x,y strings having y in exactly k positions the number of k-element subsets of {1,2,...,n} (this is either by definition, or by a short combinatorial argument if one is defining as

20

). This proves the binomial theorem.

Inductive proof
Induction yields another proof of the binomial theorem(1). When n=0, both sides equal 1, since x0=1 for all x and . Now suppose that (1) holds for a given n; we will prove it for n+1. For j,k0, let [(x,y)]jk denote the coefficient of xjyk in the polynomial (x,y). By the inductive hypothesis, (x+y)n is a polynomial in x and y such that [(x+y)n]jk is if j+k=n, and 0 otherwise. The identity

shows that (x+y)n+1 also is a polynomial in x and y, and

If j+k=n+1, then (j1)+k=n and j+(k1)=n, so the right hand side is

by Pascal's identity. On the other hand, if j+kn+1, then (j1)+kn and j+(k1)n, so we get 0+0=0. Thus

which is the inductive hypothesis with n+1 substituted for n and so completes the inductive step.

Generalisations
Newton's generalised binomial theorem
Around 1665, Isaac Newton generalised the formula to allow real exponents other than nonnegative integers, and in fact it can be generalised further, to complex exponents. In this generalisation, the finite sum is replaced by an infinite series. In order to do this one needs to give meaning to binomial coefficients with an arbitrary upper index, which cannot be done using the above formula with factorials; however factoring out (nk)! from numerator and denominator in that formula, and replacing n by r which now stands for an arbitrary number, one can define

where |x|>|y|,
[7]

is the Pochhammer symbol here standing for a falling factorial. Then, if x and y are real numbers with and r is any complex number, one has

Binomial theorem

21

When r is a nonnegative integer, the binomial coefficients for k>r are zero, so (2) specializes to (1), and there are at most r+1 nonzero terms. For other values of r, the series (2) has infinitely many nonzero terms, at least if x and y are nonzero. This is important when one is working with infinite series and would like to represent them in terms of generalised hypergeometric functions. Taking r=s leads to a useful but non-obvious formula:

Further specializing to s=1 yields the geometric series formula. Generalisations Formula (2) can be generalised to the case where x and y are complex numbers. For this version, one should assume |x|>|y|[7] and define the powers of x+y and x using a holomorphic branch of log defined on an open disk of radius |x| centered at x. Formula (2) is valid also for elements x and y of a Banach algebra as long as xy=yx, xis invertible, and||y/x||<1.

The multinomial theorem


The binomial theorem can be generalised to include powers of sums with more than two terms. The general version is

where the summation is taken over all sequences of nonnegative integer indices k1 through km such that the sum of all ki isn. (For each term in the expansion, the exponents must add up ton). The coefficients are known as multinomial coefficients, and can be computed by the formula

Combinatorially, the multinomial coefficient n-element set into disjoint subsets of sizes k1,...,kn.

counts the number of different ways to partition an

The multi-binomial theorem


It is often useful, when working in more dimension, to deal with products of binomial expressions. By the binomial theorem this is equal to This may be written more concisely, by multi-index notation, as

Binomial theorem

22

Applications
Multiple angle identities
For the complex numbers the binomial theorem can be combined with De Moivre's formula to yield multiple-angle formulas for the sine and cosine. According to De Moivre's formula,

Using the binomial theorem, the expression on the right can be expanded, and then the real and imaginary parts can be taken to yield formulas for cos(nx) and sin(nx). For example, since

De Moivre's formula tells us that

which are the usual double-angle identities. Similarly, since

De Moivre's formula yields

In general,

and

Series for e
The number e is often defined by the formula

Applying the binomial theorem to this expression yields the usual infinite series for e. In particular:

The kth term of this sum is

As n, the rational expression on the right approaches one, and therefore

This indicates that e can be written as a series:

Indeed, since each term of the binomial expansion is an increasing function of n, it follows from the monotone convergence theorem for series that the sum of this infinite series is equal toe.

Binomial theorem

23

The binomial theorem in abstract algebra


Formula (1) is valid more generally for any elements x and y of a semiring satisfying xy=yx. The theorem is true even more generally: alternativity suffices in place of associativity. The binomial theorem can be stated by saying that the polynomial sequence {1,x,x2,x3,...} is of binomial type.

Notes
[1] Binomial Theorem (http:/ / mathworld. wolfram. com/ BinomialTheorem. html) [2] The Story of the Binomial Theorem, by J. L. Coolidge (http:/ / www. jstor. org/ pss/ 2305028), The American Mathematical Monthly 56:3 (1949), pp. 147157 [3] O'Connor, John J.; Robertson, Edmund F., "Abu Bekr ibn Muhammad ibn al-Husayn Al-Karaji" (http:/ / www-history. mcs. st-andrews. ac. uk/ Biographies/ Al-Karaji. html), MacTutor History of Mathematics archive, University of St Andrews, . [4] Sandler, Stanley (2011). An Introduction to Applied Statistical Thermodynamics. Hoboken NJ: John Wiley & Sons, Inc.. ISBN978-0-470-91347-5. [5] Landau, James A (1999-05-08). "Historia Matematica Mailing List Archive: Re: [HM] Pascal's Triangle" (http:/ / archives. math. utk. edu/ hypermail/ historia/ may99/ 0073. html) (mailing list email). Archives of Historia Matematica. . Retrieved 2007-04-13. [6] (Barth 2004) [7] This is to guarantee convergence. Depending on r, the series may also converge sometimes when |x|=|y|.

References
Bag, Amulya Kumar (1966). "Binomial theorem in ancient India". Indian J. History Sci 1 (1): 6874. Barth, Nils R. (November 2004). "Computing Cavalieri's Quadrature Formula by a Symmetry of the n-Cube". The American Mathematical Monthly (Mathematical Association of America) 111 (9): 811813. doi:10.2307/4145193. ISSN0002-9890. JSTOR4145193, author's copy (http://nbarth.net/math/papers/ barth-01-cavalieri.pdf), further remarks and resources (http://nbarth.net/math/papers/) Graham, Ronald; Donald Knuth, Oren Patashnik (1994). "(5) Binomial Coefficients". Concrete Mathematics (2nd ed.). Addison Wesley. pp.153256. ISBN0-201-55802-5. OCLC17649857. Solomentsev, E.D. (2001), "Newton binomial" (http://eom.springer.de/n/n066500.htm), in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104

External links
Binomial Theorem (http://demonstrations.wolfram.com/BinomialTheorem/) by Stephen Wolfram, and "Binomial Theorem (Step-by-Step)" (http://demonstrations.wolfram.com/BinomialTheoremStepByStep/) by Bruce Colletti and Jeff Bryant, Wolfram Demonstrations Project, 2007. Binomial Theorem Introduction (http://www.liftminds.com/lesson/23/Bionomial_theorem_introduction) This article incorporates material from inductive proof of binomial theorem on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.

Birch's theorem

24

Birch's theorem
In mathematics, Birch's theorem,[1] named for Bryan John Birch, is a statement about the representability of zero by odd degree forms.

Statement of Birch's theorem


Let K be an algebraic number field, k, l and n be natural numbers, r1,...,rk be odd natural numbers, and f1,...,fk be homogeneous polynomials with coefficients in K of degrees r1,...,rk respectively in n variables, then there exists a number (r1,...,rk,l,K) such that implies that there exists an l-dimensional vector subspace V of Kn such that

Remarks
The proof of the theorem is by induction over the maximal degree of the forms f1,...,fk. Essential to the proof is a special case, which can be proved by an application of the Hardy-Littlewood circle method, of the theorem which states that if n is sufficiently large and r is odd, then the equation

has a solution in integers x1,...,xn, not all of which are 0. The restriction to odd r is necessary, since even-degree forms, such as positive definite quadratic forms, may take the value 0 only at the origin.

References
[1] B. J. Birch, Homogeneous forms of odd degree in a large number of variables, Mathematika, 4, pages 102-105 (1957)

Birkhoff's representation theorem

25

Birkhoff's representation theorem


This is about lattice theory. For other similarly named results, see Birkhoff's theorem. In mathematics, Birkhoff's representation theorem for distributive lattices states that the elements of any finite distributive lattice can be represented as finite sets, in such a way that the lattice operations correspond to unions and intersections of sets. The theorem can be interpreted as providing a one-to-one correspondence between distributive lattices and partial orders, between quasi-ordinal knowledge spaces and preorders, or between finite topological spaces and preorders. It is named after Garrett Birkhoff, who published a proof of it in 1937.[1] The name Birkhoff's representation theorem has also been applied to two other results of Birkhoff, one from 1935 on the representation of Boolean algebras as families of sets closed under union, intersection, and complement (so-called fields of sets, closely related to the rings of sets used by Birkhoff to represent distributive lattices), and Birkhoff's HSP theorem representing algebras as products of irreducible algebras. Birkhoff's representation theorem has also been called the fundamental theorem for finite distributive lattices.[2]

Understanding the theorem


Many lattices can be defined in such a way that the elements of the lattice are represented by sets, the join operation of the lattice is represented by set union, and the meet operation of the lattice is represented by set intersection. For instance, the Boolean lattice defined from the family of all subsets of a finite set has this property. More generally any finite topological space has a lattice of sets as its family of open sets. Because set unions and intersections obey the distributive law, any lattice defined in this way is a distributive lattice. Birkhoff's theorem states that in fact all finite distributive lattices can be obtained this way, and later generalizations of Birkhoff's theorem state the same thing for infinite lattices.

Examples
Consider the divisors of some composite number, such as (in the figure) 120, partially ordered by divisibility. Any two divisors of 120, such as 12 and 20, have a unique greatest common factor 1220=4, the largest number that divides both of them, and a unique least common multiple 1220=60; both of these The distributive lattice of divisors of 120, and its representation as sets of prime powers. numbers are also divisors of 120. These two operations and satisfy the distributive law, in either of two equivalent forms: (xy)z=(xz)(yz) and (xy)z=(xz)(yz), for all x, y, and z. Therefore, the divisors form a finite distributive lattice. One may associate each divisor with the set of prime powers that divide it: thus, 12 is associated with the set {2,3,4}, while 20 is associated with the set {2,4,5}. Then 1220=4 is associated with the set {2,3,4}{2,4,5}={2,4}, while 1220=60 is associated with the set {2,3,4}{2,4,5}={2,3,4,5}, so the join and meet operations of the lattice correspond to union and intersection of sets. The prime powers 2, 3, 4, 5, and 8 appearing as elements in these sets may themselves be partially ordered by divisibility; in this smaller partial order, 2 4 8 and there are no order relations between other pairs. The 16 sets that are associated with divisors of 120 are the lower sets of this smaller partial order, subsets of elements such that if

Birkhoff's representation theorem x y and y belongs to the subset, then x must also belong to the subset. From any lower set L, one can recover the associated divisor by computing the least common multiple of the prime powers in L. Thus, the partial order on the five prime powers 2, 3, 4, 5, and 8 carries enough information to recover the entire original 16-element divisibility lattice. Birkhoff's theorem states that this relation between the operations and of the lattice of divisors and the operations and of the associated sets of prime powers is not coincidental, and not dependent on the specific properties of prime numbers and divisibility: the elements of any finite distributive lattice may be associated with lower sets of a partial order in the same way. As another example, the application of Birkhoff's theorem to the family of subsets of an n-element set, partially ordered by inclusion, produces the free distributive lattice with n generators. The number of elements in this lattice is given by the Dedekind numbers.

26

The partial order of join-irreducibles


In a lattice, an element x is join-irreducible if x is not the join of a finite set of other elements. Equivalently, x is join-irreducible if it is neither the bottom element of the lattice (the join of zero elements) nor the join of any two smaller elements. For instance, in the lattice of divisors of 120, there is no pair of elements whose join is 4, so 4 is join-irreducible. An element x is join-prime if, whenever xyz, either xy or xz. In the same lattice, 4 is join-prime: whenever lcm(y,z) is divisible by 4, at least one of y and z must itself be divisible by 4. In any lattice, a join-prime element must be join-irreducible. Equivalently, an element that is not join-irreducible is not join-prime. For, if an element x is not join-irreducible, there exist smaller y and z such that x=yz. But then xyz, and x is not less than or equal to either y or z, showing that it is not join-prime. There exist lattices in which the join-prime elements form a proper subset of the join-irreducible elements, but in a distributive lattice the two types of elements coincide. For, suppose that x is join-irreducible, and that xyz. This inequality is equivalent to the statement that x=x(yz), and by the distributive law x=(xy)(xz). But since x is join-irreducible, at least one of the two terms in this join must be x itself, showing that either x=xy (equivalently xy) or x=xz (equivalently xz). The lattice ordering on the subset of join-irreducible elements forms a partial order; Birkhoff's theorem states that the lattice itself can be recovered from the lower sets of this partial order.

Birkhoff's theorem
In any partial order, the lower sets form a lattice in which the lattice's partial ordering is given by set inclusion, the join operation corresponds to set union, and the meet operation corresponds to set intersection, because unions and intersections preserve the property of being a lower set. Because set unions and intersections obey the distributive law, this is a distributive lattice. Birkhoff's theorem states that any finite distributive lattice can be constructed in this way. Theorem. Any finite distributive lattice L is isomorphic to the lattice of lower sets of the partial order of the join-irreducible elements of L. That is, there is a one-to-one order-preserving correspondence between elements of L and lower sets of the partial order. The lower set corresponding to an element x of L is simply the set of join-irreducible elements of L that are less than or equal to x, and the element of L corresponding to a lower set S of join-irreducible elements is the join of S. If one starts with a lower set S, lets x be the join of S, and constructs lower set T of the join-irreducible elements less than or equal to x, then S=T. For, every element of S clearly belongs to T, and any join-irreducible element less than or equal to x must (by join-primality) be less than or equal to one of the members of S, and therefore must (by the assumption that S is a lower set) belong to S itself. Conversely, if one starts with an element x of L, lets S be the

Birkhoff's representation theorem join-irreducible elements less than or equal to x, and constructs y as the join of S, then x=y. For, as a join of elements less than or equal to x, y can be no greater than x itself, but if x is join-irreducible then x belongs to S while if x is the join of two or more join-irreducible items then they must again belong to S, so yx. Therefore, the correspondence is one-to-one and the theorem is proved.

27

Rings of sets and preorders


Birkhoff (1937) defined a ring of sets to be a family of sets that is closed under the operations of set unions and set intersections; later, motivated by applications in mathematical psychology, Doignon & Falmagne (1999) called the same structure a quasi-ordinal knowledge space. If the sets in a ring of sets are ordered by inclusion, they form a distributive lattice. The elements of the sets may be given a preorder in which xy whenever some set in the ring contains x but not y. The ring of sets itself is then the family of lower sets of this preorder, and any preorder gives rise to a ring of sets in this way.

Functoriality
Birkhoff's theorem, as stated above, is a correspondence between individual partial orders and distributive lattices. However, it can also be extended to a correspondence between order-preserving functions of partial orders and bounded homomorphisms of the corresponding distributive lattices. The direction of these maps is reversed in this correspondence. Let 2 denote the partial order on the two-element set {0, 1}, with the order relation 0 < 1, and (following Stanley) let J(P) denote the distributive lattice of lower sets of a finite partial order P. Then the elements of J(P) correspond one-for-one to the order-preserving functions from P to 2.[2] For, if is such a function, 1(0) forms a lower set, and conversely if L is a lower set one may define an order-preserving function L that maps L to 0 and that maps the remaining elements of P to 1. If g is any order-preserving function from Q to P, one may define a function g* from J(P) to J(Q) that uses the composition of functions to map any element L of J(P) to Lg. This composite function maps Q to 2 and therefore corresponds to an element g*(L)=(Lg)1(0) of J(Q). Further, for any x and y in J(P), g*(xy)=g*(x)g*(y) (an element of Q is mapped by g to the lower set xy if and only if belongs both to the set of elements mapped to x and the set of elements mapped to y) and symmetrically g*(xy)=g*(x)g*(y). Additionally, the bottom element of J(P) (the function that maps all elements of P to 0) is mapped by g* to the bottom element of J(Q), and the top element of J(P) is mapped by g* to the top element of J(Q). That is, g* is a homomorphism of bounded lattices. However, the elements of P themselves correspond one-for-one with bounded lattice homomorphisms from J(P) to 2. For, if x is any element of P, one may define a bounded lattice homomorphism jx that maps all lower sets containing x to 1 and all other lower sets to 0. And, for any lattice homomorphism from J(P) to 2, the elements of J(P) that are mapped to 1 must have a unique minimal element x (the meet of all elements mapped to 1), which must be join-irreducible (it cannot be the join of any set of elements mapped to 0), so every lattice homomorphism has the form jx for some x. Again, from any bounded lattice homomorphism h from J(P) to J(Q) one may use composition of functions to define an order-preserving map h* from Q to P. It may be verified that g**=g for any order-preserving map g from Q to P and that and h**=h for any bounded lattice homomorphism h from J(P) to J(Q). In category theoretic terminology, J is a contravariant hom-functor J=Hom(,2) that defines a duality of categories between, on the one hand, the category of finite partial orders and order-preserving maps, and on the other hand the category of finite distributive lattices and bounded lattice homomorphisms.

Birkhoff's representation theorem

28

Generalizations
In an infinite distributive lattice, it may not be the case that the lower sets of the join-irreducible elements are in one-to-one correspondence with lattice elements. Indeed, there may be no join-irreducibles at all. This happens, for instance, in the lattice of all integers, ordered with the reverse of the usual divisibility ordering (so xy when y divides x): any number x can be expressed as the join of numbers xp and xq where p and q are distinct prime numbers that do not divide x. However, elements in infinite distributive lattices may still be represented as sets via Stone's representation theorem for distributive lattices, a form of Stone duality in which each lattice element corresponds to a compact open set in a certain topological space. This generalized representation theorem can be expressed as a category-theoretic duality between distributive lattices and coherent spaces (sometimes called spectral spaces), topological spaces in which the compact open sets are closed under intersection and form a base for the topology.[3] Hilary Priestley showed that Stone's representation theorem could be interpreted as an extension of the idea of representing lattice elements by lower sets of a partial order, using Nachbin's idea of ordered topological spaces. Stone spaces with an additional partial order linked with the topology via Priestley separation axiom can also be used to represent bounded distributive lattices. Such spaces are known as Priestley spaces. Further, certain bitopological spaces, namely pairwise Stone spaces, generalize Stone's original approach by utilizing two topologies on a set to represent an abstract distributve lattice. Thus, Birkhoff's representation theorem extends to the case of infinite (bounded) distributive lattices in at least three different ways, summed up in duality theory for distributive lattices. Birkhoff's representation theorem may also be generalized to finite structures other than distributive lattices. In a distributive lattice, the self-dual median operation[4]

gives rise to a median algebra, and the covering relation of the lattice forms a median graph. Finite median algebras and median graphs have a dual structure as the set of solutions of a 2-satisfiability instance; Barthlemy & Constantin (1993) formulate this structure equivalently as the family of initial stable sets in a mixed graph.[5] For a distributive lattice, the corresponding mixed graph has no undirected edges, and the initial stable sets are just the lower sets of the transitive closure of the graph. Equivalently, for a distributive lattice, the implication graph of the 2-satisfiability instance can be partitioned into two connected components, one on the positive variables of the instance and the other on the negative variables; the transitive closure of the positive component is the underlying partial order of the distributive lattice. Another result analogous to Birkhoff's representation theorem, but applying to a broader class of lattices, is the theorem of Edelman (1980) that any finite join-distributive lattice may be represented as an antimatroid, a family of sets closed under unions but in which closure under intersections has been replaced by the property that each nonempty set has a removable element.

Notes
[1] [2] [3] [4] [5] Birkhoff (1937). (Stanley 1997). Johnstone (1982). Birkhoff & Kiss (1947). A minor difference between the 2-SAT and initial stable set formulations is that the latter presupposes the choice of a fixed base point from the median graph that corresponds to the empty initial stable set.

References
Barthlemy, J.-P.; Constantin, J. (1993), "Median graphs, parallelism and posets", Discrete Mathematics 111 (13): 4963, doi:10.1016/0012-365X(93)90140-O. Birkhoff, Garrett (1937), "Rings of sets", Duke Mathematical Journal 3 (3): 443454, doi:10.1215/S0012-7094-37-00334-X.

Birkhoff's representation theorem Birkhoff, Garrett; Kiss, S. A. (1947), "A ternary operation in distributive lattices" (http://projecteuclid.org/ euclid.bams/1183510977), Bulletin of the American Mathematical Society 52 (1): 749752, MR0021540. Doignon, J.-P.; Falmagne, J.-Cl. (1999), Knowledge Spaces, Springer-Verlag, ISBN3-540-64501-2. Edelman, Paul H. (1980), "Meet-distributive lattices and the anti-exchange closure", Algebra Universalis 10 (1): 290299, doi:10.1007/BF02482912. Johnstone, Peter (1982), "II.3 Coherent locales", Stone Spaces, Cambridge University Press, pp.6269, ISBN9780521337793. Priestley, H. A. (1970), "Representation of distributive lattices by means of ordered Stone spaces", Bulletin of the London Mathematical Society 2 (2): 186190, doi:10.1112/blms/2.2.186. Priestley, H. A. (1972), "Ordered topological spaces and the representation of distributive lattices", Proceedings of the London Mathematical Society 24 (3): 507530, doi:10.1112/plms/s3-24.3.507. Stanley, R. P. (1997), Enumerative Combinatorics, Volume I, Cambridge Studies in Advanced Mathematics 49, Cambridge University Press, pp.104112.

29

Boolean prime ideal theorem


In mathematics, a prime ideal theorem guarantees the existence of certain types of subsets in a given abstract algebra. A common example is the Boolean prime ideal theorem, which states that ideals in a Boolean algebra can be extended to prime ideals. A variation of this statement for filters on sets is known as the ultrafilter lemma. Other theorems are obtained by considering different mathematical structures with appropriate notions of ideals, for example, rings and prime ideals (of ring theory), or distributive lattices and maximal ideals (of order theory). This article focuses on prime ideal theorems from order theory. Although the various prime ideal theorems may appear simple and intuitive, they cannot be derived in general from the axioms of ZermeloFraenkel set theory (ZF). Instead, some of the statements turn out to be equivalent to the axiom of choice (AC), while othersthe Boolean prime ideal theorem, for instancerepresent a property that is strictly weaker than AC. It is due to this intermediate status between ZF and ZF+AC (ZFC) that the Boolean prime ideal theorem is often taken as an axiom of set theory. The abbreviations BPI or PIT (for Boolean algebras) are sometimes used to refer to this additional axiom.

Prime ideal theorems


Recall that an order ideal is a (non-empty) directed lower set. If the considered poset has binary suprema (a.k.a. joins), as do the posets within this article, then this is equivalently characterized as a lower set I which is closed for binary suprema (i.e. x, y in I imply x y in I). An ideal I is prime if, whenever an infimum x y is in I, one also has x in I or y in I. Ideals are proper if they are not equal to the whole poset. Historically, the first statement relating to later prime ideal theorems was in fact referring to filterssubsets that are ideals with respect to the dual order. The ultrafilter lemma states that every filter on a set is contained within some maximal (proper) filteran ultrafilter. Recall that filters on sets are proper filters of the Boolean algebra of its powerset. In this special case, maximal filters (i.e. filters that are not strict subsets of any proper filter) and prime filters (i.e. filters that with each union of subsets X and Y contain also X or Y) coincide. The dual of this statement thus assures that every ideal of a powerset is contained in a prime ideal. The above statement led to various generalized prime ideal theorems, each of which exists in a weak and in a strong form. Weak prime ideal theorems state that every non-trivial algebra of a certain class has at least one prime ideal. In contrast, strong prime ideal theorems require that every ideal that is disjoint from a given filter can be extended to a prime ideal which is still disjoint from that filter. In the case of algebras that are not posets, one uses different substructures instead of filters. Many forms of these theorems are actually known to be equivalent, so that the

Boolean prime ideal theorem assertion that "PIT" holds is usually taken as the assertion that the corresponding statement for Boolean algebras (BPI) is valid. Another variation of similar theorems is obtained by replacing each occurrence of prime ideal by maximal ideal. The corresponding maximal ideal theorems (MIT) are oftenthough not alwaysstronger than their PIT equivalents.

30

Boolean prime ideal theorem


The Boolean prime ideal theorem is the strong prime ideal theorem for Boolean algebras. Thus the formal statement is: Let B be a Boolean algebra, let I be an ideal and let F be a filter of B, such that I and F are disjoint. Then I is contained in some prime ideal of B that is disjoint from F. The weak prime ideal theorem for Boolean algebras simply states: Every Boolean algebra contains a prime ideal. We refer to these statements as the weak and strong BPI. The two are equivalent, as the strong BPI clearly implies the weak BPI, and the reverse implication can be achieved by using the weak BPI to find prime ideals in the appropriate quotient algebra. The BPI can be expressed in various different ways. For this purpose, recall the following theorem: For any ideal I of a Boolean algebra B, the following are equivalent: I is a prime ideal. I is a maximal proper ideal, i.e. for any proper ideal J, if I is contained in J then I = J. For every element a of B, I contains exactly one of {a, a}. This theorem is a well-known fact for Boolean algebras. Its dual establishes the equivalence of prime filters and ultrafilters. Note that the last property is in fact self-dualonly the prior assumption that I is an ideal gives the full characterization. It is worth mentioning that all of the implications within this theorem can be proven in classical Zermelo-Fraenkel set theory. Thus the following (strong) maximal ideal theorem (MIT) for Boolean algebras is equivalent to BPI: Let B be a Boolean algebra, let I be an ideal and let F be a filter of B, such that I and F are disjoint. Then I is contained in some maximal ideal of B that is disjoint from F. Note that one requires "global" maximality, not just maximality with respect to being disjoint from F. Yet, this variation yields another equivalent characterization of BPI: Let B be a Boolean algebra, let I be an ideal and let F be a filter of B, such that I and F are disjoint. Then I is contained in some ideal of B that is maximal among all ideals disjoint from F. The fact that this statement is equivalent to BPI is easily established by noting the following theorem: For any distributive lattice L, if an ideal I is maximal among all ideals of L that are disjoint to a given filter F, then I is a prime ideal. The proof for this statement (which can again be carried out in ZF set theory) is included in the article on ideals. Since any Boolean algebra is a distributive lattice, this shows the desired implication. All of the above statements are now easily seen to be equivalent. Going even further, one can exploit the fact the dual orders of Boolean algebras are exactly the Boolean algebras themselves. Hence, when taking the equivalent duals of all former statements, one ends up with a number of theorems that equally apply to Boolean algebras, but where every occurrence of ideal is replaced by filter. It is worth noting that for the special case where the Boolean algebra under consideration is a powerset with the subset ordering, the "maximal filter theorem" is called the ultrafilter lemma. Summing up, for Boolean algebras, the weak and strong MIT, the weak and strong PIT, and these statements with filters in place of ideals are all equivalent. It is known that all of these statements are consequences of the axiom of choice (the easy proof makes use of Zorn's lemma), but cannot be proven in classical Zermelo-Fraenkel set theory.

Boolean prime ideal theorem Yet, the BPI is strictly weaker than the axiom of choice, though the proof of this statement, due to J. D. Halpern and Azriel Levy is rather non-trivial.

31

Further prime ideal theorems


The prototypical properties that were discussed for Boolean algebras in the above section can easily be modified to include more general lattices, such as distributive lattices or Heyting algebras. However, in these cases maximal ideals are different from prime ideals, and the relation between PITs and MITs is not obvious. Indeed, it turns out that the MITs for distributive lattices and even for Heyting algebras are equivalent to the axiom of choice. On the other hand, it is known that the strong PIT for distributive lattices is equivalent to BPI (i.e. to the MIT and PIT for Boolean algebras). Hence this statement is strictly weaker than the axiom of choice. Furthermore, observe that Heyting algebras are not self dual, and thus using filters in place of ideals yields different theorems in this setting. Maybe surprisingly, the MIT for the duals of Heyting algebras is not stronger than BPI, which is in sharp contrast to the abovementioned MIT for Heyting algebras. Finally, prime ideal theorems do also exist for other (not order-theoretical) abstract algebras. For example, the MIT for rings implies the axiom of choice. This situation requires to replace the order-theoretic term "filter" by other conceptsfor rings a "multiplicatively closed subset" is appropriate.

The ultrafilter lemma


A filter on a set X is a collection of nonempty subsets of X that is closed under finite intersection and under superset. An ultrafilter is a maximal filter. The ultrafilter lemma states that every filter on a set X is a subset of some ultrafilter on X (a maximal filter of nonempty subsets of X).[1] This lemma is most often used in the study of topology. The existence of non-principal ultrafilters is due to Tarski in 1930. The ultrafilter lemma is equivalent to the Boolean prime ideal theorem, with the equivalence provable in ZF set theory without the axiom of choice. The idea behind the proof is that the subsets of any set form a Boolean algebra partially ordered by inclusion, and any Boolean algebra is representable as an algebra of sets by Stone's representation theorem.

Applications
Intuitively, the Boolean prime ideal theorem states that there are "enough" prime ideals in a Boolean algebra in the sense that we can extend every ideal to a maximal one. This is of practical importance for proving Stone's representation theorem for Boolean algebras, a special case of Stone duality, in which one equips the set of all prime ideals with a certain topology and can indeed regain the original Boolean algebra (up to isomorphism) from this data. Furthermore, it turns out that in applications one can freely choose either to work with prime ideals or with prime filters, because every ideal uniquely determines a filter: the set of all Boolean complements of its elements. Both approaches are found in the literature. Many other theorems of general topology that are often said to rely on the axiom of choice are in fact equivalent to BPI. For example, the theorem that a product of compact Hausdorff spaces is compact is equivalent to it. If we leave out "Hausdorff" we get a theorem equivalent to the full axiom of choice. A not too well known application of the Boolean prime ideal theorem is the existence of a non-measurable set[2] (the example usually given is the Vitali set, which requires the Axiom of Choice). From this and the fact that the BPI is strictly weaker than the Axiom of Choice, it follows that the existence of non-measurable sets is strictly weaker than the axiom of choice.

Boolean prime ideal theorem

32

Notes
[1] Halpern, James D. (1966), "Bases in Vector Spaces and the Axiom of Choice", Proceedings of the American Mathematical Society (American Mathematical Society) 17 (3): 670673, JSTOR2035388. [2] Sierpiski, Wacaw (1938), "Fonctions additives non compltement additives et fonctions non mesurables", Fundamenta Mathematicae 30: 9699

References
Davey, B. A.; Priestley, H. A. (2002), Introduction to Lattices and Order (2nd ed.), Cambridge University Press, ISBN9780521784511. An easy to read introduction, showing the equivalence of PIT for Boolean algebras and distributive lattices. Johnstone, Peter (1982), Stone Spaces, Cambridge studies in advanced mathematics, 3, Cambridge University Press, ISBN9780521337793. The theory in this book often requires choice principles. The notes on various chapters discuss the general relation of the theorems to PIT and MIT for various structures (though mostly lattices) and give pointers to further literature. Banaschewski, B. (1983), "The power of the ultrafilter theorem", Journal of the London Mathematical Society (2nd series) 27 (2): 193202, doi:10.1112/jlms/s2-27.2.193. Discusses the status of the ultrafilter lemma. Ern, M. (2000), "Prime ideal theory for general algebras", Applied Categorical Structures 8: 115144. Gives many equivalent statements for the BPI, including prime ideal theorems for other algebraic structures. PITs are considered as special instances of separation lemmas.

BorelWeil theorem
In mathematics, in the field of representation theory, the BorelWeil theorem, named after Armand Borel and Andr Weil, provides a concrete model for irreducible representations of compact Lie groups and complex semisimple Lie groups. These representations are realized in the spaces of global sections of holomorphic line bundles on the flag manifold of the group. Its generalization to higher cohomology spaces is called the BorelWeilBott theorem.

Statement of the theorem


The theorem can be stated either for a complex semisimple Lie group G or for its compact form K. Let G be a connected complex semisimple Lie group, B its Borel subgroup, and X=G/B the flag variety. In this picture, X is a complex manifold and a nonsingular algebraic G-variety. The flag variety can also be described as a compact homogeneous space K/T, where T=KB is a (compact) Cartan subgroup of K. An integral weight determines a G-equivariant holomorphic line bundle L on X and the group G acts on its space of global sections, The BorelWeil theorem states that if is a dominant integral weight then this representation is an irreducible highest weight representation of G with highest weight . Its restriction to K is an irreducible unitary representation of K with highest weight , and each irreducible unitary representations of K is obtained in this way for a unique value of .

BorelWeil theorem

33

Concrete description
The weight gives rise to a character (one-dimensional representation) of the Borel subgroup B, which is denoted . Holomorphic sections of the holomorphic line bundle L over G/B may be described more concretely as holomorphic maps

for all gG and bB. The action of G on these sections is given by

for g,hG.

Example
Let G be the complex special linear group SL(2,C), with a Borel subgroup consisting of upper triangular matrices with determinant one. Integral weights for G may be identified with integers, with dominant weights corresponding to nonnegative integers, and the corresponding characters n of B have the form

The flag variety G/B may be identified with the complex projective line P1 with homogeneous coordinates X, Y and the space of the global sections of the line bundle Ln is identified with the space of homogeneous polynomials of degree n on C2. For n0, this space has dimension n+1 and forms an irreducible representation under the standard action of G on the polynomial algebra C[X,Y]. Weight vectors are given by monomials of weights 2in, and the highest weight vector Xn has weight n.

History
The theorem dates back to the early 1950s and can be found in Serre (1995) and Tits (1955).

References
Serre, Jean-Pierre (1995), "Reprsentations linaires et espaces homognes khlriens des groupes de Lie compacts (d'aprs Armand Borel et Andr Weil)", Sminaire Bourbaki (Paris: Soc. Math. France) 2 (100): 447454. In French; translated title: Linear representations and Khler homogeneous spaces of compact Lie groups (after Armand Borel and Andr Weil. Tits, Jacques (1955), Sur certaines classes d'espaces homognes de groupes de Lie, Acad. Roy. Belg. Cl. Sci. Mm. Coll., 29 In French. Sepanski, Mark R. (2007), Compact Lie groups., Graduate Texts in Mathematics, 235, New York: Springer. Knapp, Anthony W. (2001), Representation theory of semisimple groups: An overview based on examples, Princeton Landmarks in Mathematics, Princeton, NJ: Princeton University Press. Reprint of the 1986 original.

BorelWeilBott theorem

34

BorelWeilBott theorem
In mathematics, the BorelWeil-Bott theorem is a basic result in the representation theory of Lie groups, showing how a family of representations can be obtained from holomorphic sections of certain complex vector bundles, and, more generally, from higher sheaf cohomology groups associated to such bundles. It is built on the earlier BorelWeil theorem of Armand Borel and Andr Weil, dealing just with the section case, the extension being provided by Raoul Bott. One can equivalently, through Serre's GAGA, view this as a result in complex algebraic geometry in the Zariski topology.

Formulation
Let G be a semisimple Lie group or algebraic group over , and fix a maximal torus T along with a Borel subgroup B which contains T. Let be an integral weight of T; defines in a natural way a one-dimensional representation C of B, by pulling back the representation on T = B/U, where U is the unipotent radical of B. Since we can think of the projection map G G/B as a principal B-bundle, for each C we get an associated fiber bundle L- on G/B (note the sign), which is obviously a line bundle. Identifying L with its sheaf of holomorphic sections, we consider the sheaf cohomology groups . Since G acts on the total space of the bundle by bundle automorphisms, this action naturally gives a G-module structure on these groups; and the BorelWeilBott theorem gives an explicit description of these groups as G-modules. We first need to describe the Weyl group action centered at W, we set weight is said to be dominant if , where . For any integral weight and in the Weyl group denotes the half-sum of positive roots of G. It is straightforward to for all simple roots such that . Let denote the length function on W. such that ; or (2) There is a unique is dominant, such

check that this defines a group action, although this action is not linear, unlike the usual Weyl group action. Also, a Given an integral weight that , one of two cases occur: (1) There is no

equivalently, there exists a nonidentity for all i; and in the second case, we have for all

is dominant. The theorem states that in the first case, we have

, while .

is the dual of the irreducible highest-weight representation of G with highest weight It is worth noting that case (1) above occurs if and only if identity element . for some positive root

. Also, we obtain to be the

the classical BorelWeil theorem as a special case of this theorem by taking

to be dominant and

Example
For example, consider G = SL2(C), for which G/B is the Riemann sphere, an integral weight is specified simply by an integer n, and = 1. The line bundle Ln is O(n), whose sections are the homogeneous polynomials of degree n (i.e. the binary forms). As a representation of G, the sections can be written as Symn(C2)*, and is canonically isomorphic to Symn(C2). This gives us at a stroke the representation theory of : (O(1)) is the standard representation, and (O(n)) is its n-th symmetric power. We even have a unified description of the action of the Lie algebra, derived from its realization as vector fields on the Riemann sphere: if H, X, Y are the standard generators of , then we can write

BorelWeilBott theorem

35

Positive characteristic
One also has a weaker form of this theorem in positive characteristic. Namely, let G be a semisimple algebraic group over an algebraically closed field of characteristic . Then it remains true that for all i if is a weight such that is non-dominant for all . However, the other statements of the theorem for all , do not remain valid in this setting. More explicitly, let be a dominant integral weight; then it is still true that as a G-submodule. If but it is no longer true that this G-module is simple in general, although it does contain the unique highest weight module of highest weight is an arbitrary integral weight, it is in fact a large unsolved in general. Unlike over , problem in representation theory to describe the cohomology modules it need not be the case for a fixed

that these modules are all zero except in a single degree i.

References
Fulton, William; Harris, Joe (1991), Representation theory. A first course, Graduate Texts in Mathematics, Readings in Mathematics, 129, New York: Springer-Verlag, ISBN978-0-387-97495-8, MR1153249, ISBN 978-0-387-97527-6. Baston, Robert J.; Eastwood, Michael G. (1989), The Penrose Transform: its Interaction with Representation Theory, Oxford University Press. Hazewinkel, Michiel, ed. (2001), "BottBorelWeil theorem" [1], Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 A Proof of the BorelWeilBott Theorem [2], by Jacob Lurie. Retrieved on Dec. 14, 2007. This article incorporates material from BorelBottWeil theorem on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.

References
[1] http:/ / eom. springer. de/ b/ b120400. htm [2] http:/ / www-math. mit. edu/ ~lurie/ papers/ bwb. pdf

Brauer's theorem on induced characters

36

Brauer's theorem on induced characters


Brauer's theorem on induced characters, often known as Brauer's induction theorem, and named after Richard Brauer, is a basic result in the branch of mathematics known as character theory, which is, in turn, part of the representation theory of a finite group. Let G be a finite group and let Char(G) denote the subring of the ring of complex-valued class functions of G consisting of integer combinations of irreducible characters. Char(G) is known as the character ring of G, and its elements are known as virtual characters (alternatively, as generalized characters, or sometimes difference characters). It is a ring by virtue of the fact that the product of characters of G is again a character of G. Its multiplication is given by the elementwise product of class functions. Brauer's induction theorem shows that the character ring can be generated (as an abelian group) by certain characters which are fairly easily understood. More precisely, the theorem states that every virtual character of G is expressible as an integer combination of induced characters of the form , , where H ranges over subgroups of G and ranges over linear characters (having degree 1) of H. In fact, Brauer showed that the subgroups H could be chosen from a very restricted collection, now called Brauer elementary subgroups. These are direct products of cyclic groups and groups whose order is a power of a prime. Using Frobenius reciprocity, Brauer's induction theorem leads easily to his fundamental characterization of characters, which asserts that a complex-valued class function of G is a virtual character if and only if its restriction to each Brauer elementary subgroup of G is a virtual character. This result, together with the fact that a virtual character is an irreducible character if and only if (1) > 0 and (where is the usual inner product on the ring of complex-valued class functions) gives a means of constructing irreducible characters without explicitly constructing the associated representations. An initial motivation for Brauer's induction theorem was application to Artin L-functions. It shows that those are built up from Dirichlet L-functions, or more general Hecke L-functions. Highly significant for that application is whether each character of G is a non-negative integer combination of characters induced from linear characters of subgroups. In general, this is not the case. In fact, by a theorem of Taketa, if all characters of G are so expressible, then G must be a solvable group (although solvability alone does not guarantee such expressions- for example, the solvable group SL(2,3) has an irreducible complex character of degree 2 which is not expressible as a non-negative integer combination of characters induced from linear characters of subgroups). An ingredient of the proof of Brauer's induction theorem is that when G is a finite nilpotent group, every complex irreducible character of G is induced from a linear character of some subgroup. A precursor to Brauer's induction theorem was Artin's induction theorem, which states that |G| times the trivial character of G is an integer combination of characters which are each induced from trivial characters of cyclic subgroups of G. Brauer's theorem removes the factor |G|, but at the expense of expanding the collection of subgroups used. Some years after the proof of Brauer's theorem appeared, J.A. Green showed (in 1955) that no such induction theorem (with integer combinations of characters induced from linear characters) could be proved with a collection of subgroups smaller than the Brauer elementary subgroups. The proof of Brauer's induction theorem exploits the ring structure of Char(G) (most proofs also make use of a slightly larger ring, Char*(G), which consists of -combinations of irreducible characters, where is a primitive complex |G|-th root of unity). The set of integer combinations of characters induced from linear characters of Brauer elementary subgroups is an ideal I(G) of Char(G), so the proof reduces to showing that the trivial character is in I(G). Several proofs of the theorem, beginning with a proof due to Brauer and John Tate, show that the trivial character is in the analogously defined ideal I*(G) of Char*(G) by concentrating attention on one prime p at a time, and constructing integer-valued elements of I*(G) which differ (elementwise) from the trivial character by (integer multiples of) a sufficiently high power of p. Once this is achieved for every prime divisor of |G|, some manipulations with congruences and algebraic integers, again exploiting the fact that I*(G) is an ideal of Ch*(G), place the trivial

Brauer's theorem on induced characters character in I(G). An auxiliary result here is that a

37 -valued class function lies in the ideal I*(G) if its values are all

divisible (in ) by |G|. Brauer's induction theorem was proved in 1946, and there are now many alternative proofs. In 1986, Victor Snaith gave a proof by a radically different approach, topological in nature (an application of the Lefschetz fixed-point theorem). There has been related recent work on the question of finding natural and explicit forms of Brauer's theorem, notably by Robert Boltje.

References
Isaacs, I.M. (1994). Character Theory of Finite Groups. Dover. ISBN0-486-68014-2. Corrected reprint of the 1976 original, published by Academic Press.

Brauer's three main theorems


Brauer's main theorems are three theorems in representation theory of finite groups linking the blocks of a finite group (in characteristic p) with those of its p-local subgroups, that is to say, the normalizers of its non-trivial p-subgroups. The second and third main theorems allow refinements of orthogonality relations for ordinary characters which may be applied in finite group theory. These do not presently admit a proof purely in terms of ordinary characters. All three main theorems are stated in terms of the Brauer correspondence.

Brauer correspondence
There are many ways to extend the definition which follows, but this is close to the early treatments by Brauer. Let G be a finite group, p be a prime, F be a field of characteristic p. Let H be a subgroup of G which contains

for some p-subgroup Q of G, and is contained in the normalizer . The Brauer homomorphism (with respect to H) is a linear map from the center of the group algebra of G over F to the corresponding algebra for H. Specifically, it is the restriction to of the (linear) projection from to whose kernel is spanned by the elements of G outside . The image of this map is contained in , and it transpires that the map is also a ring homomorphism. Since it is a ring homomorphism, for any block B of FG, the Brauer homomorphism sends the identity element of B either to 0 or to an idempotent element. In the latter case, the idempotent may be decomposed as a sum of (mutually orthogonal) primitive idempotents of Z(FH). Each of these primitive idempotents is the multiplicative identity of some block of FH. The block b of FH is said to be a Brauer correspondent of B if its identity element occurs in this decomposition of the image of the identity of B under the Brauer homomorphism.

Brauer's three main theorems

38

Brauer's first main theorem


Brauer's first main theorem (Brauer1944, 1956, 1970) states that if blocks of the normalizer is a finite group a is a -subgroup of and , each , then there is a bijection between the collections of (characteristic p) blocks of with defect group

with defect group D. This bijection arises because when

block of G with defect group D has a unique Brauer correspondent block of H, which also has defect group D.

Brauer's second main theorem


Brauer's second main theorem (Brauer1944, 1959) gives, for an element t whose order is a power of a prime p, a criterion for a (characteristic p) block of to correspond to a given block of , via generalized decomposition numbers. These are the coefficients which occur when the restrictions of ordinary characters of (from the given block) to elements of the form tu, where u ranges over elements of order prime to p in written as linear combinations of the irreducible Brauer characters of only necessary to use Brauer characters from blocks of of G. , are . The content of the theorem is that it is

which are Brauer correspondents of the chosen block

Brauer's third main theorem


Brauer's third main theorem (Brauer 1964, theorem3) states that when Q is a p-subgroup of the finite group G, and H is a subgroup of G, containing , and contained in , then the principal block of H is the only Brauer correspondent of the principal block of G (where the blocks referred to are calculated in characteristic p).

References
Brauer, R. (1944), "On the arithmetic in a group ring", Proceedings of the National Academy of Sciences of the United States of America 30: 109114, ISSN0027-8424, JSTOR87919, MR0010547 Brauer, R. (1946), "On blocks of characters of groups of finite order I", Proceedings of the National Academy of Sciences of the United States of America 32: 182186, ISSN0027-8424, JSTOR87578, MR0016418 Brauer, R. (1946), "On blocks of characters of groups of finite order. II", Proceedings of the National Academy of Sciences of the United States of America 32: 215219, ISSN0027-8424, JSTOR87838, MR0017280 Brauer, R. (1956), "Zur Darstellungstheorie der Gruppen endlicher Ordnung", Mathematische Zeitschrift 63: 406444, doi:10.1007/BF01187950, ISSN0025-5874, MR0075953 Brauer, R. (1959), "Zur Darstellungstheorie der Gruppen endlicher Ordnung. II", Mathematische Zeitschrift 72: 2546, doi:10.1007/BF01162934, ISSN0025-5874, MR0108542 Brauer, R. (1964), "Some applications of the theory of blocks of characters of finite groups. I", Journal of Algebra 1: 152167, doi:10.1016/0021-8693(64)90031-6, ISSN0021-8693, MR0168662 Brauer, R. (1970), "On the first main theorem on blocks of characters of finite groups." [1], Illinois Journal of Mathematics 14: 183187, ISSN0019-2082, MR0267010 Dade, Everett C. (1971), "Character theory pertaining to finite simple groups", in Powell, M. B.; Higman, Graham, Finite simple groups. Proceedings of an Instructional Conference organized by the London Mathematical Society (a NATO Advanced Study Institute), Oxford, September 1969., Boston, MA: Academic Press, pp.249327, ISBN978-0-12-563850-0, MR0360785 gives a detailed proof of the Brauer's main theorems. Ellers, H. (2001), "Brauer's first main theorem" [2], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 Ellers, H. (2001), "Brauer height-zero conjecture" [3], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 Ellers, H. (2001), "Brauer's second main theorem" [4], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104

Brauer's three main theorems Ellers, H. (2001), "Brauer's third main theorem" [5], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 Walter Feit, The representation theory of finite groups. North-Holland Mathematical Library, 25. North-Holland Publishing Co., Amsterdam-New York, 1982. xiv+502 pp.ISBN 0-444-86155-6

39

References
[1] [2] [3] [4] [5] http:/ / projecteuclid. org/ euclid. ijm/ 1256053174 http:/ / eom. springer. de/ b/ b120440. htm http:/ / eom. springer. de/ b/ b120450. htm http:/ / eom. springer. de/ b/ b120460. htm http:/ / eom. springer. de/ b/ b120470. htm

BrauerCartanHua theorem
The BrauerCartanHua theorem (named after Richard Brauer, lie Cartan, and Hua Luogeng) is a theorem in abstract algebra pertaining to division rings, which says that given two division rings K D such that xKx1 is contained in K for every x not equal to 0 in D, then either K is contained in Z, the center of D, or K=D. In other words, if the unit group of K is a normal subgroup of the unit group of D, then either K = D or K is central, (Lam 2001, p.211).

References
Herstein, I. N. (1975). Topics in algebra. New York: Wiley. p.368. ISBN0-471-01090-1. Lam, Tsit-Yuen (2001). A First Course in Noncommutative Rings (2nd ed.). Berlin, New York: Springer-Verlag. ISBN978-0-387-95325-0. MR1838439.

BrauerNesbitt theorem

40

BrauerNesbitt theorem
In mathematics, the Brauer-Nesbitt theorem can refer to several different theorems proved by Richard Brauer and Cecil J. Nesbitt in the representation theory of finite groups. In modular representation theory, the Brauer-Nesbitt theorem on blocks of defect zero states that a character whose order is divisible by the highest power of a prime p dividing the order of a finite group remains irreducible when reduced mod p and vanishes on all elements whose order is divisible by p. Moreover it belongs to a block of defect zero. A block of defect zero contains only one ordinary character and only one modular character. Another version states that if k is a field of characteristic zero, A is a k-algebra, V, W are semisimple A-modules which are finite dimensional over k, and TrV = TrW as elements of Homk(A,k), then V and W are isomorphic as A-modules.

References
Curtis, Reiner, Representation theory of finite groups and associative algebras, Wiley 1962. Brauer, R.; Nesbitt, C. On the modular characters of groups. Ann. of Math. (2) 42, (1941). 556-590.

BrauerSiegel theorem
In mathematics, the BrauerSiegel theorem, named after Richard Brauer and Carl Ludwig Siegel, is an asymptotic result on the behaviour of algebraic number fields, obtained by Richard Brauer and Carl Ludwig Siegel. It attempts to generalise the results known on the class numbers of imaginary quadratic fields, to a more general sequence of number fields

In all cases other than the rational field Q and imaginary quadratic fields, the regulator Ri of Ki must be taken into account, because Ki then has units of infinite order by Dirichlet's unit theorem. The quantitative hypothesis of the standard BrauerSiegel theorem is that if Di is the discriminant of Ki, then

Assuming that, and the algebraic hypothesis that Ki is a Galois extension of Q, the conclusion is that

where hi is the class number of Ki. This result is ineffective, as indeed was the result on quadratic fields on which it built. Effective results in the same direction were initiated in work of Harold Stark from the early 1970s.

References
Richard Brauer, On the Zeta-Function of Algebraic Number Fields, American Journal of Mathematics 69 (1947), 243250.

BrauerSuzuki theorem

41

BrauerSuzuki theorem
In mathematics, the BrauerSuzuki theorem, proved by Brauer & Suzuki (1959), Suzuki (1962), Brauer (1964), states that if a finite group has a generalized quaternion Sylow 2-subgroup and no non-trivial normal subgroups of odd order, then the group has a centre of order 2. In particular, such a group cannot be simple. A generalization of the BrauerSuzuki theorem is given by Glauberman's Z* theorem.

References
Brauer, R. (1964), "Some applications of the theory of blocks of characters of finite groups. II", Journal of Algebra 1: 307334, doi:10.1016/0021-8693(64)90011-0, ISSN0021-8693, MR0174636 Brauer, R.; Suzuki, Michio (1959), "On finite groups of even order whose 2-Sylow group is a quaternion group", Proceedings of the National Academy of Sciences of the United States of America 45: 17571759, ISSN0027-8424, JSTOR90063, MR0109846 Dade, Everett C. (1971), "Character theory pertaining to finite simple groups", in Powell, M. B.; Higman, Graham, Finite simple groups. Proceedings of an Instructional Conference organized by the London Mathematical Society (a NATO Advanced Study Institute), Oxford, September 1969., Boston, MA: Academic Press, pp.249327, ISBN978-0-12-563850-0, MR0360785 gives a detailed proof of the BrauerSuzuki theorem. Suzuki, Michio (1962), "Applications of group characters" [1], in Hall, M., 1960 Institute on finite groups: held at California Institute of Technology, Proc. Sympos. Pure Math., VI, American Mathematical Society, pp.101105, ISBN978-0821814062

References
[1] http:/ / books. google. com/ books?id=Nb8rT4rm0EUC& pg=PA101

BrauerSuzukiWall theorem

42

BrauerSuzukiWall theorem
In mathematics, the BrauerSuzukiWall theorem, proved by Brauer, Suzuki & Wall (1958), characterizes the one-dimensional unimodular projective groups over finite fields.

References
Brauer, R.; Suzuki, Michio; Wall, G. E. (1958), "A characterization of the one-dimensional unimodular projective groups over finite fields" [1], Illinois Journal of Mathematics 2: 718745, ISSN0019-2082, MR0104734

References
[1] http:/ / projecteuclid. org/ euclid. ijm/ 1255448336

Burnside theorem
In mathematics, Burnside's theorem in group theory states that if G is a finite group of order where p and q are prime numbers, and a and b are non-negative integers, then G is solvable. Hence each non-Abelian finite simple group has order divisible by three distinct primes. Furthermore, as a consequence of the Feit-Thompson theorem, one of those can be chosen to be 2.

History
The theorem was proved by William Burnside in the early years of the 20th century. Burnside's theorem has long been one of the best-known applications of representation theory to the theory of finite groups, though a proof avoiding the use of group characters was published by D. Goldschmidt around 1970.

Outline of Burnside's proof


1. By induction, it suffices to prove that a finite simple group G whose order has the form for primes p and q is cyclic. Suppose then that the order of G has this form, but G is not cyclic. Suppose for definiteness that b >0. 2. Using the modified class equation, G has a non-identity conjugacy class of size prime to q. Hence G either has a non-trivial center, or has a conjugacy class of size for some positive integer r. The first possibility is excluded since G is assumed simple, but not cyclic. Hence there is a non-central element x of G such that the conjugacy class of x has size . 3. Application of column orthogonality relations and other properties of group characters and algebraic integers lead to the existence of a non-trivial irreducible character of G such that . 4. The simplicity of G then implies that any non-trivial complex irreducible representation is faithful, and it follows that x is in the center of G, a contradiction.

Burnside theorem

43

References
1. James, Gordon; and Liebeck, Martin (2001). Representations and Characters of Groups (2nd ed.). Cambridge University Press. ISBN 0-521-00392-X. See chapter 31. 2. Fraleigh, John B. (2002) A First Course in Abstract Algebra (7th ed.). Addison Wesley. ISBN 0-201-33596-4.

Cartan's theorem
In mathematics, three results in Lie group theory are called Cartan's theorem, named after lie Cartan: 1. The theorem that for a Lie group G, any closed subgroup is a Lie subgroup.[1] 2. A theorem on highest weight vectors in the representation theory of a semisimple Lie group. 3. The equivalence between the category of connected real Lie groups and finite dimensional real Lie algebras is called usually (in the literature of the second half of 20th century) Cartan's or Cartan-Lie theorem as it is proved by lie Cartan whereas S. Lie has proved earlier just the infinitesimal version (local solvability of Maurer-Cartan equations (see Maurer-Cartan form) or the equivalence between the finite dimensional Lie algebras and the category of local Lie groups). Lie listed his results as 3 direct and 3 converse theorems, the infinitesimal variant of Cartan's theorem was essentially his 3rd converse theorem, hence Serre has called it in an influential book, the "third Lie theorem", the name which is historically somewhat misleading, but more often used in the recent decade in the connection to many generalizations. See also Cartan's theorems A and B, results of Henri Cartan, and Cartan's lemma for various other results attributed to lie and Henri Cartan.

Notes
[1] See 26 of Cartan's article La thorie des groups finis et continus et l'Analysis Situs.

References
Cartan, lie (1930), "La thorie des groupes finis et continus et l'Analysis Situs", Mmorial Sc. Math. XLII: 161 Helgason, Sigurdur (2001), Differential geometry, Lie groups, and symmetric spaces, Graduate Studies in Mathematics, 34, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-2848-9, MR1834454

CartanDieudonn theorem

44

CartanDieudonn theorem
In mathematics, the CartanDieudonn theorem, named after lie Cartan and Jean Dieudonn, is a theorem on the structure of the automorphism group of symmetric bilinear spaces.

Statement of the theorem


Let (V,b) be an n-dimensional, non-degenerate symmetric bilinear space over a field with characteristic not equal to 2. Then, every element of the orthogonal group O(V,b) is a composition of at most n reflections.

References
Sylvestre Gallot, Dominique Hulin, Jacques LaFontaine, Riemannian Geometry, Springer, 2004. ISBN 3540204938. Jean H Gallier, Geometric Methods and Applications, Springer, 2000. ISBN 0387950443.

Cauchy's theorem (group theory)


Cauchy's theorem is a theorem in the mathematics of group theory, named after Augustin Louis Cauchy. It states that if G is a finite group and p is a prime number dividing the order of G (the number of elements in G), then G contains an element of order p. That is, there is x in G so that p is the lowest non-zero number with xp = e, where e is the identity element. The theorem is related to Lagrange's theorem, which states that the order of any subgroup of a finite group G divides the order of G. Cauchy's theorem implies that for any prime divisor p of the order of G, there is a subgroup of G whose order is pthe cyclic group generated by the element in Cauchy's theorem. Cauchy's theorem is generalised by Sylow's first theorem, which implies that if pn is any prime power dividing the order of G, then G has a subgroup of order pn.

Statement and proof


Many texts appear to prove the theorem with the use of strong induction and the class equation, though considerably less machinery is required to prove the theorem in the abelian case. One can also invoke group actions for the proof. Theorem: Let G be a finite group and p be a prime. If p divides the order of G, then G has an element of order p. Proof 1: We induct on n = |G| and consider the two cases where G is abelian or G is nonabelian. Suppose G is abelian. If G is simple, then it must be cyclic of prime order and trivially contains an element of order p. Otherwise, there exists a nontrivial, proper normal subgroup . If p divides |H|, then H contains an element of order p by the inductive hypothesis, and thus G does as well. Otherwise, p must divide the index [G:H] by Lagrange's theorem, and we see the quotient group G/H contains an element of order p by the inductive hypothesis; that is, there exists an x in G such that (Hx)p = Hxp = H. Then there exists an element h1 in H such that h1xp = 1, the identity element of G. It is easily checked that for every element a in H there exists b in H such that bp = a, so there exists h2 in H so that h2 p = h1. Thus h2x has order p, and the proof is finished for the abelian case. Suppose that G is nonabelian, so that its center Z is a proper subgroup. If p divides the order of the centralizer CG(a) for some noncentral element a (i.e. a is not in Z), then CG(a) is a proper subgroup and hence contains an element of order p by the inductive hypothesis. Otherwise, we must have p dividing the index [G:CG(a)], again by Lagrange's Theorem, for all noncentral a. Using the class equation, we have p dividing the left side of the equation (|G|) and also dividing all of the summands on the right, except for possibly |Z|. However, simple arithmetic shows p must also

Cauchy's theorem (group theory) divide the order of Z, and thus the center contains an element of order p by the inductive hypothesis as it is a proper subgroup and hence of order strictly less than that of G. This completes the proof. Proof 2: This time we define the set of p-tuples whose elements are in the group G by . Note that we can choose only (p-1) of the identity. Thus Define the action group of order p. Then The stabilizer is . We have from the Orbit-Stabilizer Theorem that Take Hence we know that p divides |X| implies that there is at least one other Then we have Since xj is in G this completes the proof. and the distinct orbits. Then . with the property that its orbit has order 1. by the definition of X. for each . . independently, since we are constrained by the product equal to the , from which we deduce that p also divides by is the orbit of some element , from which we . can deduce the order, , where is the cyclic

45

Uses
A practically immediate consequence of Cauchy's Theorem is a useful characterization of finite p-groups, where p is a prime. In particular, a finite group G is a p-group (i.e. all of its elements have order pk for some natural number k) if and only if G has order pn for some natural number n. It is also typical to use Cauchy's Theorem to prove the first of Sylow's Theorems, though this is not required.

References
James McKay. Another proof of Cauchy's group theorem, American Math. Monthly, 66 (1959), p. 119.

External links
Cauchy's theorem [1] on PlanetMath Proof of Cauchy's theorem [2] on PlanetMath

References
[1] http:/ / planetmath. org/ ?op=getobj& amp;from=objects& amp;id=1569 [2] http:/ / planetmath. org/ ?op=getobj& amp;from=objects& amp;id=2186

Cayley's theorem

46

Cayley's theorem
In group theory, Cayley's theorem, named in honor of Arthur Cayley, states that every group G is isomorphic to a subgroup of the symmetric group acting on G.[1] This can be understood as an example of the group action of G on the elements of G.[2] A permutation of a set G is any bijective function taking G onto G; and the set of all such functions forms a group under function composition, called the symmetric group on G, and written as Sym(G).[3] Cayley's theorem puts all groups on the same footing, by considering any group (including infinite groups such as (R,+)) as a permutation group of some underlying set. Thus, theorems which are true for permutation groups are true for groups in general.

History
Although Burnside[4] attributes the theorem to Jordan,[5] Eric Nummela[6] nonetheless argues that the standard name"Cayley's Theorem"is in fact appropriate. Cayley, in his original 1854 paper,[7] showed that the correspondence in the theorem is one-to-one, but he failed to explicitly show it was a homomorphism (and thus an isomorphism). However, Nummela notes that Cayley made this result known to the mathematical community at the time, thus predating Jordan by 16 years or so.

Proof of the theorem


Where g is any element of G, consider the function fg : G G, defined by fg(x) = g*x. By the existence of inverses, this function has a two-sided inverse, . So multiplication by g acts as a bijective function. Thus, fg is a permutation of G, and so is a member of Sym(G). The set is a subgroup of Sym(G) which is isomorphic to G. The fastest way to establish this

is to consider the function T : G Sym(G) with T(g) = fg for every g in G. T is a group homomorphism because (using "" for composition in Sym(G)): for all x in G, and hence:

The homomorphism T is also injective since T(g) = idG (the identity element of Sym(G)) implies that g*x = x for all x in G, and taking x to be the identity element e of G yields g = g*e = e. Alternatively, T is also injective since, if g*x=g'*x implies g=g' (by post-multiplying with the inverse of x, which exists because G is a group). Thus G is isomorphic to the image of T, which is the subgroup K. T is sometimes called the regular representation of G.

Cayley's theorem

47

Alternative setting of proof


An alternative setting uses the language of group actions. We consider the group to have permutation representation, say Firstly, suppose with . . Then the group action is is injective, that is, if the kernel of by classification of G-orbits (also is trivial. Suppose ker Then, ker as a G-set, which can be shown

known as the orbit-stabilizer theorem). Now, the representation is faithful if , and thus ker by the equivalence of the permutation representation and the group action. But since is trivial. Then im isomorphism theorem.

and thus the result follows by use of the first

Remarks on the regular group representation


The identity group element corresponds to the identity permutation. All other group elements correspond to a permutation that does not leave any element unchanged. Since this also applies for powers of a group element, lower than the order of that element, each element corresponds to a permutation which consists of cycles which are of the same length: this length is the order of that element. The elements in each cycle form a left coset of the subgroup generated by the element.

Examples of the regular group representation


Z2 = {0,1} with addition modulo 2; group element 0 corresponds to the identity permutation e, group element 1 to permutation (12). Z3 = {0,1,2} with addition modulo 3; group element 0 corresponds to the identity permutation e, group element 1 to permutation (123), and group element 2 to permutation (132). E.g. 1 + 1 = 2 corresponds to (123)(123)=(132). Z4 = {0,1,2,3} with addition modulo 4; the elements correspond to e, (1234), (13)(24), (1432). The elements of Klein four-group {e, a, b, c} correspond to e, (12)(34), (13)(24), and (14)(23). S3 (dihedral group of order 6) is the group of all permutations of 3 objects, but also a permutation group of the 6 group elements:
* e a b c d f e a b c d f e a b c d f a e d f b c b f e d c a c d f e a b d c a b f e f b c a e d permutation e (12)(35)(46) (13)(26)(45) (14)(25)(36) (156)(243) (165)(234)

Cayley's theorem

48

Notes
[1] [2] [3] [4] [5] [6] Jacobson (2009), p. 38. Jacobson (2009), p. 72, ex. 1. Jacobson (2009), p. 31. Burnside, William (1911), Theory of Groups of Finite Order (2 ed.), Cambridge, ISBN0486495752 Jordan, Camille (1870), Traite des substitutions et des equations algebriques, Paris: Gauther-Villars Nummela, Eric (1980), "Cayley's Theorem for Topological Groups", American Mathematical Monthly (Mathematical Association of America) 87 (3): 202203, doi:10.2307/2321608, JSTOR2321608 [7] Cayley, Arthur (1854), "On the theory of groups as depending on the symbolic equation n=1", Phil. Mag. 7 (4): 4047

References
Jacobson, Nathan (2009), Basic algebra (2nd ed.), Dover, ISBN978-0-486-47189-1.

CayleyHamilton theorem
In linear algebra, the CayleyHamilton theorem (named after the mathematicians Arthur Cayley and William Hamilton) states that every square matrix over a commutative ring (including the real or complex field) satisfies its own characteristic equation. More precisely: If A is a given nn matrix and In is the nn identity matrix, then the characteristic polynomial of A is defined as where "det" is the determinant operation. Since the entries of the matrix are (linear or constant) polynomials in, the determinant is also a polynomial in . The CayleyHamilton theorem states that "substituting" the matrix A for in this polynomial results in the zero matrix:

The powers of that have become powers of A by the substitution should be computed by repeated matrix multiplication, and the constant term should be multiplied by the identity matrix (the zeroth power of A) so that it can be added to the other terms. The theorem allows An to be expressed as a linear combination of the lower matrix powers of A. The CayleyHamilton theorem is equivalent to the statement that the minimal polynomial of a square matrix divides its characteristic polynomial.

Example
As a concrete example, let . Its characteristic polynomial is given by

The CayleyHamilton theorem claims that, if we define

then

CayleyHamilton theorem

49

which one can verify easily.

Illustration for specific dimensions and practical applications


For a 11 matrix A=(a), the characteristic polynomial is given by p()=a, and so p(A)=(a)a(1)=(0) is obvious. For a 22 matrix,

the characteristic polynomial is given by p()=2(a+d)+(adbc), so the CayleyHamilton theorem states that

which is indeed always the case, evident by working out the entries of A2. For a general nn invertible matrix A, i.e., one with nonzero determinant, A1 can thus be written as an (n1)-th order polynomial expression in A: As indicated, the CayleyHamilton theorem amounts to the identity

with cn1=tr(A), etc., where tr(A) is the trace of the matrix A. This can then be written as and, by multiplying both sides by , one is led to the compact expression for the inverse

For larger matrices, the expressions for the coefficients ck of the characteristic polynomial in terms of the matrix components become increasingly complicated; but they can also be expressed in terms of traces of powers of the matrix A, using Newton's identities, thus resulting in more compact expressions (but which involve divisions by certain integers). For instance, the coefficient c1=a+d of above is just the trace of A, trA, while the constant coefficient c0=adbc can be written as ((trA)2tr(A2)). (Of course, it is also the determinant of A in this case.) In fact, this expression, ((trA)2tr(A2)), always gives the coefficient cn2 of n2 in the characteristic polynomial of any nn matrix; so, for a 33 matrix A, the statement of the CayleyHamilton theorem can also be written as

where the right-hand side designates a 33 matrix with all entries reduced to zero. Similarly, one can write for a 44 matrix A:

and so on for larger matrices, with the increasingly complex expressions for the coefficients deducible from Newton's identities. An alternate, practical method for obtaining these coefficients ck for a general nn matrix, yielding the above ones virtually by inspection, relies on . Hence,

CayleyHamilton theorem where the exponential only needs be expanded to order n, since p() is of order n. The CayleyHamilton theorem always provides a relationship between the powers of A (though not always the simplest one), which allows one to simplify expressions involving such powers, and evaluate them without having to compute the power An or any higher powers of A. For instance the concrete 22 Example above can be written as Then, for example, to calculate A4, observe

50

Proving the theorem in general


As the examples above show, obtaining the statement of the CayleyHamilton theorem for an nn matrix requires two steps: first the coefficients ci of the characteristic polynomial are determined by development as a polynomial in t of the determinant

and then these coeffcients are used in a linear combination of powers of A that is equated to the nn null matrix:

The left hand side can be worked out to an nn matrix whose entries are (enormous) polynomial expressions in the set of entries of A, so the CayleyHamilton theorem states that each of these expressions are equivalent to 0. For any fixed value of n these identities can be obtained by tedious but completely straightforward algebraic manipulations. None of these computations can show however why the CayleyHamilton theorem should be valid for matrices of all possible sizes n, so a uniform proof for all n is needed.

Preliminaries
If a vector v of size n happens to be an eigenvector of A with eigenvalue , in other words if , then

which is the null vector since

(the eigenvalues of A are precisely the roots of p(t)). This holds for all

possible eigenvalues , so the two matrices equated by the theorem certainly give the same (null) result when applied to any eigenvector. Now if A admits a basis of eigenvectors, in other words if A is diagonalizable, then the CayleyHamilton theorem must hold for A, since two matrices that give the same values when applied to each element of a basis must be equal. Not all matrices are diagonalizable, but for matrices with complex coefficients many of them are: the set of diagonalizable complex square matrices of a given size is dense in the set of all such square matrices (for a matrix to be diagonalizable it suffices for instance that its characteristic polynomial not have multiple roots). Now if any of the expressions that the theorem equates to 0 would not reduce to a null expression, in other words if it would be a nonzero polynomial in the coefficients of the matrix, then the set of complex matrices for which this expression happens to give 0 would not be dense in the set of all matrices, which would contradict the fact that the theorem holds for all diagonalizable matrices. Thus one can see that the

CayleyHamilton theorem CayleyHamilton theorem must be true. While this provides a valid proof, the argument is not very satisfactory, since the identities represented by the theorem do not in any way depend on the nature of the matrix (diagonalizable or not), nor on the kind of entries allowed (for matrices with real entries the diagonizable ones do not form a dense set, and it seems strange one would have to consider complex matrices to see that the CayleyHamilton theorem holds for them). We shall therefore now consider only arguments that prove the theorem directly for any matrix using algebraic manipulations only; these also have the benefit of working for matrices with entries in any commutative ring. There is a great variety of such proofs of the CayleyHamilton theorem, of which several will be given here. They vary in the amount of abstract algebraic notions required to understand the proof. The simplest proofs use just those notions needed to formulate the theorem (matrices, polynomials with numeric entries, determinants), but involve technical computations that render somewhat mysterious the fact that they lead precisely to the correct conclusion. It is possible to avoid such details, but at the price of involving more subtle algebraic notions: polynomials with coefficients in a non-commutative ring, or matrices with unusual kinds of entries. Adjugate matrices All proofs below use the notion of the adjugate matrix of an nn matrix M. This is a matrix whose

51

coefficients are given by polynomial expressions in the coefficients of M (in fact by certain (n1)(n1) determinants), in such a way that one has the following fundamental relations These relations are a direct consequence of the basic properties of determinants: evaluation of the (i,j) entry of the matrix product on the left gives the expansion by column j of the determinant of the matrix obtained from M by replacing column i by a copy of column j, which is if and zero otherwise; the matrix product on the right is similar, but for expansions by rows. Being a consequence of just algebraic expression manipulation, these relations are valid for matrices with entries in any commutative ring (commutativity must be assumed for determinants to be defined in the first place). This is important to note here, because these relations will be applied for matrices with non-numeric entries such as polynomials.

A direct algebraic proof


This proof uses just the kind of objects needed to formulate the CayleyHamilton theorem: matrices with polynomials as entries. The matrix whose determinant is the characteristic polynomial of A is such a matrix, and since polynomials form a commutative ring, it has an adjugate Then according to the right hand fundamental relation of the adjugate one has

Since B is also a matrix with polynomials in t as entries, one can for each i collect the coefficients of to form a matrix Bi of numbers, such that one has

in each entry

(the way the entries of B are defined makes clear that no powers higher than

occur). While this looks like a

polynomial with matrices as coefficients, we shall not consider such a notion; it is just a way to write a matrix with polynomial entries as linear combination of constant matrices, and the coefficient has been written to the left of the matrix to stress this point of view. Now one can expand the matrix product in our equation by bilinearity

CayleyHamilton theorem

52

Writing

, one obtains an equality of two matrices with is the same on both

polynomial entries, written as linear combinations of constant matrices with powers of t as coefficients. Such an equality can hold only if in any matrix position the entry that is multiplied by a given power sides; it follows that the constant matrices with coefficient equations for i from n down to 0 one finds We multiply the equation of the coefficients of ti from the left by Ai, and sum up; the left-hand sides form a telescoping sum and cancel completely, which results in the equation in both expressions must be equal. Writing these

This completes the proof.

A proof using polynomials with matrix coefficients


This proof is similar to the first one, but tries to give meaning to the notion of polynomial with matrix coefficients that was suggested by the expressions occurring in that proof. This requires considerable care, since it is somewhat unusual to consider polynomials with coefficients in a non-commutative ring, and not all reasoning that is valid for commutative polynomials can be appied in this setting. Notably, while arithmetic of polynomials over a commutative ring models the arithmetic of polynomial functions, this is not the case over a non-commutative ring (in fact there is no obvious notion of polynomial function in this case that is closed under multiplication). So when considering polynomials in t with matrix coefficients, the variable t must not be thought of as an "unknown", but as a formal symbol that is to manipulated according to given rules; in particular one cannot just set t to a specific value. Let M = Mn(R) be the ring of n n matrices with entries in some ring R (such as the real or complex numbers) that has A as an element. Matrices with as coefficients polynomials in t, such as or its adjugate B in the first proof, are elements of Mn(R[t]). By collecting like powers of t, such matrices can be written as "polynomials" in t with constant matrices as coefficients; write M[t] for the set of such polynomials. Since this set is in bijection with Mn(R[t]), one defines arithmetic operations on it correspondingly, in particular multiplication is given by

respecting the order of the coefficient matrices from the two operands; obviously this gives a non-commutative multiplication. Thus the identity

from the first proof can be viewed as one involving a multiplication of elements in M[t]. At this point, it is tempting to set t equal to the matrix A, which makes the first factor on the left equal to the null matrix, and the right hand side equal to p(A); however, this is not an allowed operation when coefficients do not

CayleyHamilton theorem

53

commute. It is possible to define a "right-evaluation map" evA : M[t] M, which replaces each ti by the matrix power Ai of A, where one stipulates that the power is always to be multiplied on the right to the corresponding coefficient. However this map is not a ring homomorphism: the right-evaluation of a product differs in general from the product of the right-evaluations. This is so because multiplication of polynomials with matrix coefficients does not model multiplication of expressions containing unknowns: a product is defined assuming that t co N, but this may fail if t is replaced by the matrix A. One can work around this difficulty in the particular situation at hand, since the above right-evaluation map does become a ring homomorphism if the matrix A is in the center of the ring of coefficients, so that it commutes with all the coefficients of the polynomials (the argument proving this is straightforward, exactly because commuting t with coefficients is now justified after evaluation). Now A is not always in the center of M, but we may replace M with a smaller ring provided it contains all the coefficients of the polynomials in question: , A, and the coefficients of the polynomial B. The obvious choice for such a subring is the centralizer Z of A, the subring of all matrices that commute with A; by definition A is in the center of Z. This centralizer obviously contains , and A, but one has to show that it contains the matrices out the adjugate B as a polynomial: . To do this one combines the two fundamental relations for adjugates, writing

Equating the coefficients shows that for each i, we have A Bi = Bi A as desired. Having found the proper setting in which evA is indeed a homomorphism of rings, one can complete the proof as suggested above:

This completes the proof.

A synthesis of the first two proofs


In the first proof, one was able to determine the coefficients Bi of B based on the right hand fundamental relation for the adjugate only. In fact the first n equations derived can be interpreted as determining the quotient B of the Euclidean division of the polynomial on the left by the monic polynomial , while the final equation expresses the fact that the remainder is zero. This division is performed in the ring of polynomials with matrix coefficients. Indeed, even over a non-commutative ring, Euclidean division by a monic polynomial P is defined, and always produces a unique quotient and remainder with the same degree condition as in the commutative case, provided it is specified at which side one wishes P to be a factor (here that is to the left). To see that quotient and remainder are unique (which is the important part of the statement here), it suffices to write as and observe that since P is monic, cannot have a degree less than that of P, unless But the dividend and divisor . used here both lie in the subring (R[A])[t], where R[A] is the subring

of the matrix ring M generated by A: the R-linear span of all powers of A. Therefore the Euclidean division can in fact be performed within that commutative polynomial ring, and of course it then gives the same quotient B and remainder 0 as in the larger ring; in particular this shows that B in fact lies in . But in this commutative setting it is valid to set t to A in the equation , in other words apply the evaluation map

CayleyHamilton theorem

54

which is a ring homomorphism, giving

just like in the second proof, as desired. In addition to proving the theorem, the above argument tells us that the coefficients of B are polynomials in A, while from the second proof we only knew that they lie in the centralizer Z of A; in general Z is a larger subring than R[A], and not necessarily commutative. In particular the constant term lies in R[A]. Since A is an arbitrary square matrix, this proves that that depend on identity can always be expressed as a polynomial in , ..., , (with coefficients ), something that is not obvious from the definition of the adjugate matrix. In fact the equations as polynomials in A, which leads to the

found in the first proof allow successively expressing

valid for all nn matrices, where

is the characteristic polynomial of A. Note to the right hand

that this identity implies the statement of the CayleyHamilton theorem: one may move side, multiply the resulting equation (on the left or on the right) by , and use the fact that

A proof using matrices of endomorphisms


As was mentioned above, the matrix in statement of the theorem is obtained by first evaluating the before is determinant and then substituting the matrix A for t; doing that subtitution into the matrix evaluating the determinant is not meaningful. Nevertheless, it is possible to give an interpretation where

obtained directly as the value of a certain deteminant, but this requires a more complicated setting, one of matrices over a ring in which one can interpret both the entries of A, and all of A itself. One could take for this the ring M of n n matrices over R, where the entry is realised as , and A as itself. But considering matrices with matrices as entries might cause confusion with block matrices, which is not intended, as that gives the wrong notion of determinant. It is clearer to distinguish A from the endomorphism of an n-dimensional vector space V (or free R-module if R is not a field) defined by it in a basis e1, ..., en, and to take matrices over the ring End(V) of all such endomorphisms. Then is a possible matrix entry, while A designates the element of whose interpreted as element of defined on entries of the matrix Then a determinant map entry is endomorphism of scalar multiplication by ; similarly In will be . However, since End(V) is not a commutative ring, no deteminant is ; this can only be done for matrices over a commutative subring of End(V). Now the all lie in the subring R[] generated by the identity and , which is commutative. is defined, and evaluates to the value p() of the

characteristic polynomial of A at (this holds independently of the relation between A and ); the CayleyHamilton In this form, the following proof can be obtained from that of (Atiyah & MacDonald1969, Prop. 2.4) (which in fact theorem states that p() is the null endomorphism. is the more general statement related to the Nakayama lemma; one takes for the ideal in that proposition the whole ring R). The fact that A is the matrix of in the basis e1, ..., enmeans that

One can interpret these as n components of one equation in Vn, whose members can be written using the matrix-vector product that is defined as usual, but with individual entries and being "multiplied" by forming ; this gives:

CayleyHamilton theorem where is the element whose component i is ei (in other words it is the basis e1, ..., en of V written as a

55

column of vectors). Writing this equation as

one recognizes the transpose of the matrix adjugate matrix of

considered above, and its determinant (as element of , one left-multiplies by the , giving

) is also p(). To derive from this equation that , which is defined in the matrix ring

the associativity of matrix-matrix and matrix-vector multiplication used in the first step is a purely formal property of those operations, independent of the nature of the entries. Now component i of this equation says that ; thus p() vanishes on all ei, and since these elements generate V it follows that , completing the proof. One additional fact that follows from this proof is that the matrix A whose characteristic polynomial is taken need not be identical to the value substituted into that polynomial; it suffices that be an endomorphism of V satisfying the initial equations (ei)=j Aj,iej for some sequence of elements e1,...,en that generate V (which space might have smaller dimension than n, or in case the ring R is not a field it might not be a free module at all).

A bogus "proof": p(A) = det(AInA) = det(AA) = 0


One elementary but incorrect argument for the theorem is to "simply" take the definition

and substitute

for

, obtaining

There are many ways to see why this argument is wrong. First, in CayleyHamilton theorem, p(A) is an nn matrix. However, the right hand side of the above equation is the value of a determinant, which is a scalar. So they cannot be equated unless n=1 (i.e. A is just a scalar). Second, in the expression , the variable actually occurs at the diagonal entries of the matrix previous example again: . To illustrate, consider the characteristic polynomial in the

If one substitutes the entire matrix

for

in those positions, one obtains

in which the "matrix" expression is simply not a valid one. Note, however, that if scalar multiples of identity matrices instead of scalars are subtracted in the above, i.e. if the substitution is performed as

then the determinant is indeed zero, but the expanded matrix in question does not evaluate to its determinant (a scalar) be compared to (a matrix). So the argument that

; nor can

CayleyHamilton theorem still does not apply. Actually, if such an argument holds, it should also hold when other multilinear forms instead of determinant is used. For instance, if we consider the permanent function and define , then by the same argument, we should be able to "prove" that q(A)=0. But this statement is demonstrably wrong. In the 2-dimensional case, for instance, the permanent of a matrix is given by

56

So, for the matrix

in the previous example,

Yet one can verify that One of the proofs for CayleyHamilton theorem above bears some similarity to the argument that . By introducing a matrix with non-numeric coefficients, one can actually let lives inside a matrix entry, but then is not equal to , and the conclusion is reached differently.

Abstraction and generalizations


The above proofs show that the CayleyHamilton theorem holds for matrices with entries in any commutative ring R, and that p() = 0 will hold whenever is an endomorphism of an R module generated by elements e1,...,en that satisfies for j = 1,...,n. This more general version of the theorem is the source of the celebrated Nakayama lemma in commutative algebra and algebraic geometry.

References
Atiyah, M. F.; MacDonald, I. G. (1969), Introduction to Commutative Algebra, Westview Press, ISBN0-201-40751-5

External links
A proof from PlanetMath. [1] The Cayley-Hamilton Theorem [2] at MathPages T. Kaczorek (2001), "CayleyHamilton theorem" [3], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104

References
[1] http:/ / planetmath. org/ ?op=getobj& from=objects& id=7308 [2] http:/ / www. mathpages. com/ home/ kmath640/ kmath640. htm [3] http:/ / eom. springer. de/ C/ c022410. htm

ChevalleyShephardTodd theorem

57

ChevalleyShephardTodd theorem
In mathematics, the ChevalleyShephardTodd theorem in invariant theory of finite groups states that the ring of invariants of a finite group acting on a complex vector space is a polynomial ring if and only if the group is generated by pseudoreflections. In the case of subgroups of the complex general linear group the theorem was first proved by G. C. Shephard and J. A. Todd(1954) who gave a case-by-case proof. Claude Chevalley(1955) soon afterwards gave a uniform proof. It has been extended to finite linear groups over an arbitrary field in the non-modular case by Jean-Pierre Serre.

Statement of the theorem


Let V be a finite-dimensional vector space over a field K and let G be a finite subgroup of the general linear group GL(V). An element s of GL(V) is called a pseudoreflection if it fixes a codimension one subspace of V and is not the identity transformation I, or equivalently, if the kernel Ker (s I) has codimension one in V. Assume that the order of G is relatively prime to the characteristic of K (the so-called non-modular case). Then the following three properties are equivalent: The group G is generated by pseudoreflections. The algebra of invariants K[V]G is a (free) polynomial algebra. The algebra K[V] is a free module over K[V]G. In the case when the field K is the field C of complex numbers, the first condition is usually stated as "G is a complex reflection group". Shephard and Todd derived a full classification of such groups.

Examples
Let V be one-dimensional. Then any finite group faithfully acting on V is a subgroup of the multiplicative group of the field K, and hence a cyclic group. It follows that G consists of roots of unity of order dividing n, where n is its order, so G is generated by pseudoreflections. In this case, K[V] = K[x] is the polynomial ring in one variable and the algebra of invariants of G is the subalgebra generated by xn, hence it is a polynomial algebra. Let V = Kn be the standard n-dimensional vector space and G be the symmetric group Sn acting by permutations of the elements of the standard basis. The symmetric group is generated by transpositions (ij), which act by reflections on V. On the other hand, by the main theorem of symmetric functions, the algebra of invariants is the polynomial algebra generated by the elementary symmetric functions e1, en. Let V = K2 and G be the cyclic group of order 2 acting by I. In this case, G is not generated by pseudoreflections, since the nonidentity element s of G acts without fixed points, so that dim Ker (s I) = 0. On the other hand, the algebra of invariants is the subalgebra of K[V] = K[x, y] generated by the homogeneous elements x2, xy, and y2 of degree 2. This subalgebra is not a polynomial algebra because of the relation x2y2 = (xy)2.

ChevalleyShephardTodd theorem

58

Generalizations
Broer (2007) gave an extension of the ChevalleyShephardTodd theorem to positive characteristic. There has been much work on the question of when a reductive algebraic group acting on a vector space has a polynomial ring of invariants. In the case when the algebraic group is simple and the representation is irreducible all cases when the invariant ring is polynomial have been classified by Schwarz (1978) In general, the ring of invariants of a finite group acting linearly on a complex vector space is Cohen-Macaulay, so it is a finite rank free module over a polynomial subring.

References
Broer, Abraham (2007), On Chevalley-Shephard-Todd's theorem in positive characteristic, [], arXiv:0709.0715 Chevalley, Claude (1955), "Invariants of finite groups generated by reflections", Amer. J. Of Math. 77 (4): 778782, doi:10.2307/2372597, JSTOR2372597 Neusel, Mara D.; Smith, Larry (2002), Invariant Theory of Finite Groups, American Mathematical Society, ISBN0-8218-2916-5 Shephard, G. C.; Todd, J. A. (1954), "Finite unitary reflection groups", Canadian J. Math. 6: 274304, doi:10.4153/CJM-1954-028-3 Schwarz, G. (1978), "Representations of simple Lie groups with regular rings of invariants", Invent. Math. 49 (2): 167191, doi:10.1007/BF01403085 Smith, Larry (1997), "Polynomial invariants of finite groups. A survey of recent developments" [1], Bull. Amer. Math. Soc. 34 (3): 211250, doi:10.1090/S0273-0979-97-00724-6, MR1433171 Springer, T. A. (1977), Invariant Theory, Springer, ISBN0-387-08242-5

References
[1] http:/ / www. ams. org/ bull/ 1997-34-03/ S0273-0979-97-00724-6/

ChevalleyWarning theorem

59

ChevalleyWarning theorem
In algebra, the ChevalleyWarning theorem implies that certain polynomial equations in sufficiently many variables over a finite field have solutions. It was proved by Ewald Warning(1936) and a slightly weaker form of the theorem, known as Chevalley's theorem, was proved by Chevalley(1936). Chevalley's theorem implied Artin's and Dickson's conjecture that finite fields are quasi-algebraically closed fields (Artin 1982, page x).

Statement of the theorems


Consider a system of polynomial equations

where the

are polynomials with coefficients in a finite field

and such that the number of variables satisfies

where

is the total degree of

. The ChevalleyWarning theorem states that the number of common solutions of . Chevalley's theorem states that if the system has the , i.e. if the polynomials have no constant terms, then the system also has a is at least 2. has total degree

is divisible by the characteristic trivial solution

non-trivial solution . Chevalley's theorem is an immediate consequence of the ChevalleyWarning theorem since Both theorems are best possible in the sense that, given any , the list

polynomial given by the norm of x1a1 + ... + xnan where the elements a form a basis of the finite field of order pn.

and only the trivial solution. Alternatively, using just one polynomial, we can take P1 to be the degree n

Proof of Warning's theorem


If i<p1 then

so the sum over Fn of any polynomial in x1,...,xn of degree less than n(p1) also vanishes. The total number of common solutions mod p of P1 = ... = Pr = 0 is

because each term is 1 for a solution and 0 otherwise. If the sum of the degrees of the polynomials Pi is less than n then this vanishes by the remark above.

ChevalleyWarning theorem

60

Artin's conjecture
It is a consequence of Chevalley's theorem that finite fields are quasi-algebraically closed. This had been conjectured by Emil Artin in 1935. The motivation behind Artin's conjecture was his observation that quasi-algebraically closed fields have trivial Brauer group, together with the fact that finite fields have trivial Brauer group by Wedderburn's theorem.

The AxKatz theorem


The AxKatz theorem, named after James Ax and Nicholas Katz, determines more accurately a power cardinality of dividing the number of solutions; here, if is the largest of the , then the exponent of the can

be taken as the ceiling function of

The AxKatz result has an interpretation in tale cohomology as a divisibility result for the (reciprocals of) the zeroes and poles of the local zeta-function. Namely, the same power of divides each of these algebraic integers.

References
Artin, Emil (1982), Lang, Serge.; Tate, John, eds., Collected papers, Berlin, New York: Springer-Verlag, ISBN978-0-387-90686-7, MR671416 Ax, James (1964), "Zeros of polynomials over finite fields", American Journal of Mathematics 86: 255261, doi:10.2307/2373163, MR0160775 Chevalley, Claude (1936), "Dmonstration d'une hypothse de M. Artin" (in French), Abhandlungen aus dem Mathematischen Seminar der Universitt Hamburg 11: 7375, doi:10.1007/BF02940714, JFM61.1043.01, Zbl0011.14504 Katz, Nicholas M. (1971), "On a theorem of Ax", Amer. J. Math. 93 (2): 485499, doi:10.2307/2373389 Warning, Ewald (1936), "Bemerkung zur vorstehenden Arbeit von Herrn Chevalley" (in German), Abhandlungen aus dem Mathematischen Seminar der Universitt Hamburg 11: 7683, doi:10.1007/BF02940715, JFM61.1043.02, Zbl0011.14601

Classification of finite simple groups

61

Classification of finite simple groups


In mathematics, the classification of the finite simple groups is a theorem stating that every finite simple group belongs to one of four categories described below. These groups can be seen as the basic building blocks of all finite groups, in much the same way as the prime numbers are the basic building blocks of the natural numbers. The JordanHlder theorem is a more precise way of stating this fact about finite groups. The proof of the theorem consists of tens of thousands of pages in several hundred journal articles written by about 100 authors, published mostly between 1955 and 2004. Gorenstein, Lyons, and Solomon are gradually publishing a simplified and revised version of the proof.

Statement of the classification theorem


Theorem. Every finite simple group is isomorphic to one of the following groups: A cyclic group with prime order; An alternating group of degree at least 5; A simple group of Lie type, including both the classical Lie groups, namely the groups of projective special linear, unitary, symplectic, or orthogonal transformations over a finite field; the exceptional and twisted groups of Lie type (including the Tits group which is not strictly a group of Lie type). The 26 sporadic simple groups. The classification theorem has applications in many branches of mathematics, as questions about the structure of finite groups (and their action on other mathematical objects) can sometimes be reduced to questions about finite simple groups. Thanks to the classification theorem, such questions can sometimes be answered by checking each family of simple groups and each sporadic group. Daniel Gorenstein announced in 1983 that the finite simple groups had all been classified, but this was premature as he had been misinformed about the proof of the classification of quasithin groups. The completed proof of the classification was announced by Aschbacher (2004) after Aschbacher and Smith published a 1221 page proof for the missing quasithin case.

Overview of the proof of the classification theorem


Gorenstein(1982, 1983) wrote two volumes outlining the low rank and odd characteristic part of the proof, and Michael Aschbacher, Richard Lyons, and Stephen D. Smith et al.(2011) wrote a 3rd volume covering the remaining characteristic 2 case. The proof can be broken up into several major pieces as follows:

Groups of small 2-rank


The simple groups of low 2-rank are mostly groups of Lie type of small rank over fields of odd characteristic, together with five alternating and seven characteristic 2 type and nine sporadic groups. The simple groups of small 2-rank include: Groups of 2-rank 0, in other words groups of odd order, which are all solvable by the Feit-Thompson theorem. Groups of 2-rank 1. The Sylow 2-subgroups are either cyclic, which is easy to handle using the transfer map, or generalized quaternion, which are handled with the Brauer-Suzuki theorem: in particular there are no simple groups of 2-rank 1. Groups of 2-rank 2. Alperin showed that the Sylow subgoup must be dihedral, quasidihedral, wreathed, or a Sylow 2-subgroup of U3(4). The first case was done by the GorensteinWalter theorem which showed that the

Classification of finite simple groups only simple groups are isomorphic to L2(q) for q odd or A7, the second and third cases were done by the AlperinBrauerGorenstein theorem which implies that the only simple groups are isomorphic to L3(q) or U3(q) for q odd or M11, and the last case was done by Lyons who showed that U3(4) is the only simple possibility. Groups of sectional 2-rank at most 4, classified by the GorensteinHarada theorem. The classification of groups of small 2-rank, especially ranks at most 2, makes heavy use of ordinary and modular character theory, which is almost never directly used elsewhere in the classification. All groups not of small 2 rank can be split into two major classes: groups of component type and groups of characteristic 2 type. This is because if a group has sectional 2-rank at least 5 then MacWilliams showed that its Sylow 2-subgroups are connected, and the balance theorem implies that any simple group with connected Sylow 2-subgroups is either of component type or characteristic 2 type. (For groups of low 2-rank the proof of this breaks down, because theorems such as the signalizer functor theorem only work for groups with elementary abelian subgroups of rank at least 3.)

62

Groups of component type


A group is said to be of component type if for some centralizer C of an involution, C/O(C) has a component (where O(C) is the core of C, the maximal normal subgroup of odd order). These are more or less the groups of Lie type of odd characteristic of large rank, and alternating groups, together with some sporadic groups. A major step in this case is to eliminate the obstruction of the core of an involution. This is accomplished by the B-theorem, which states that every component of C/O(C) is the image of a component of C. The idea is that these groups have a centralizer of an involution with a component that is a smaller quasisimple group, which can be assumed to be already known by induction. So to classify these groups one takes every central extension of every known finite simple group, and finds all simple groups with a centralizer of involution with this as a component. This gives a rather large number of different cases to check: there are not only 26 sporadic groups and 16 families of groups of Lie type and the alternating groups, but also many of the groups of small rank or over small fields behave differently from the general case and have to be treated separately, and the groups of Lie type of even and odd characteristic are also quite different.

Groups of characteristic 2 type


A group is of characteristic 2 type if the generalized Fitting subgroup F*(Y) of every 2-local subgroup Y is a 2-group. As the name suggests these are roughly the groups of Lie type over fields of characteristic 2, plus a handful of others that are alternating or sporadic or of odd characteristic. Their classification is divided into the small and large rank cases, where the rank is the largest rank of an odd abelian subgroup normalizing a nontrivial 2-subgroup, which is often (but not always) the same as the rank of a Cartan subalgebra when the group is a group of Lie type in characteristic 2. The rank 1 groups are the thin groups, classified by Aschbacher, and the rank 2 ones are the notorious quasithin groups, classified by Aschbacher and Smith. These correspond roughly to groups of Lie type of ranks 1 or 2 over fields of characteristic 2. Groups of rank at least 3 are further subdivided into 3 classes by the trichotomy theorem, proved by Aschbacher for rank 3 and by Gorenstein and Lyons for rank at least 4. The three classes are groups of GF(2) type (classified mainly by Timmesfeld), groups of "standard type" for some odd prime (classified by the Gilman-Griess theorem and work by several others), and groups of uniqueness type, where a result of Aschbacher implies that there are no simple groups. The general higher rank case consists mostly of the groups of Lie type over fields of characteristic 2 of rank at least 3 or 4.

Classification of finite simple groups

63

Existence and uniqueness of the simple groups


The main part of the classification produces a characterization of each simple group. It is then necessary to check that there exists a simple group for each characterization and that it is unique. This gives a large number of separate problems; for example, the original proofs of existence and uniqueness of the monster totaled about 200 pages, and the identification of the Ree groups by Thompson and Bombieri was one of the hardest parts of the classification. Many of the existence proofs and some of the uniqueness proofs for the sporadic proofs originally used computer calculations, some of which have since been replaced by shorter hand proofs.

History of the proof


Gorenstein's program
In 1972 Gorenstein (1979, Appendix) announced a program for completing the classification of finite simple groups, consisting of the following 16 steps: 1. Groups of low 2-rank. This was essentially done by Gorenstein and Harada, who classified the groups with sectional 2-rank at most 4. Most of the cases of 2-rank at most 2 had been done by the time Gorenstein announced his program. 2. The semisimplicity of 2-layers. The problem is to prove that the 2-layer of the centralizer of an involution in a simple group is semisimple. 3. Standard form in odd characteristic. If a group has an involution with a 2-component that is a group of Lie type of odd characteristic, the goal is to show that it has a centralizer of involution in "standard form" meaning that a centralizer of involution has a component that is of Lie type in odd characteristic and also has a centralizer of 2-rank 1. 4. Classification of groups of odd type. The problem is to show that if a group has a centralizer of involution in "standard form" then it is a group of Lie type of odd characteristic. This was solved by Aschbacher's classical involution theorem. 5. Quasi-standard form 6. Central involutions 7. Classification of alternating groups. More precisely, show that if a simple group has 8. Some sporadic groups 9. Thin groups. The simple thin finite groups, those with 2-local p-rank at most 1 for odd primes p, were classified by Aschbacher in 1978 10. Groups with a strongly p-embedded subgroup for p odd 11. The signalizer functor method for odd primes. The main problem is to prove a signalizer functor theorem for nonsolvable signalizer functors. This was solved by McBride in 1982. 12. Groups of characteristic p type. This is the problem of groups with a strongly p-embedded 2-local subgroup with p odd, which was handled by Aschbacher. 13. Quasithin groups. A quasithin group is one whose 2-local subgroups have p-rank at most 2 for all odd primes p, and the problem is to classify the simple ones of characteristic 2 type. This was completed by Aschbacher and Smith in 2004. 14. Groups of low 2-local 3-rank. This was essentially solved by Aschbacher's trichotomy theorem for groups with e(G)=3. The main change is that 2-local 3-rank is replaced by 2-local p-rank for odd primes. 15. Centralizers of 3-elements in standard form. This was essentially done by the Trichotomy theorem. 16. Classification of simple groups of characteristic 2 type. This was handled by the Gilman-Griess theorem, with 3-elements replaced by p-elements for odd primes.

Classification of finite simple groups

64

Timeline of the proof


Many of the items in the list below are taken from Solomon (2001). The date given is usually the publication date of the complete proof of a result, which is sometimes several years later than the proof or first announcement of the result, so some of the items appear in the "wrong" order.
Publication date 1832 1854 1861 1870 Galois introduces normal subgroups and finds the simple groups An (n5) and PSL2(Fp) (p5) Cayley defines abstract groups Mathieu finds the first two Mathieu groups M11, M12, the first sporadic simple groups. Jordan lists some simple groups: the alternating and projective special linear ones, and emphasizes the importance of the simple groups. Sylow proves the Sylow theorems Mathieu finds three more Mathieu groups M22, M23, M24. Otto Hlder proves that the order of any nonabelian finite simple group must be a product of at least 4 primes, and asks for a classification of finite simple groups. Cole classifies simple groups of order up to 660 Frobenius and Burnside begin the study of character theory of finite groups. Burnside classifies the simple groups such that the centralizer of every involution is a non-trivial elementary abelian 2-group. Frobenius proves that a Frobenius group has a Frobenius kernel, so in particular is not simple. Dickson defines classical groups over arbitrary finite fields, and exceptional groups of type G2 over fields of odd characteristic. Dickson introduces the exceptional finite simple groups of type E6. Burnside uses character theory to prove Burnside's theorem that the order of any non-abelian finite simple group must be divisible by at least 3 distinct primes. Dickson introduces simple groups of type G2 over fields of even characteristic Burnside conjectures that every non-abelian finite simple group has even order Hall proves the existence of Hall subgroups of solvable groups Hall begins his study of p-groups Brauer begins the study of modular characters. Zassenhaus classifies finite sharply 3-transitive permutation groups Fitting introduces the Fitting subgroup and proves Fitting's theorem that for solvable groups the Fitting subgroup contains its centralizer. Brauer describes the modular characters of a group divisible by a prime to the first power. Brauer classifies simple groups with GL2(Fq) as the centralizer of an involution. The BrauerFowler theorem implies that the number of finite simple groups with given centralizer of involution is finite, suggesting an attack on the classification using centralizers of involutions. Chevalley introduces the Chevalley groups, in particular introducing exceptional simple groups of types F4, E7, and E8. HallHigman theorem Suzuki shows that all finite simple CA groups of odd order are cyclic. The BrauerSuzukiWall theorem characterizes the projective special linear groups of rank 1, and classifies the simple CA groups. Steinberg introduces the Steinberg groups, giving some new finite simple groups, or types 3D4 and 2E6 (the latter were independently found at about the same time by Jacques Tits).

1872 1873 1892

1893 1896 1899 1901 1901 1901 1904

1905 1911 1928 1933 1935 1936 1938

1942 1954 1955

1955 1956 1957 1958 1959

Classification of finite simple groups

65

1959

The BrauerSuzuki theorem about groups with generalized quaternion Sylow 2-subgroups shows in particular that none of them are simple. Thompson proves that a group with a fixed-point-free automorphism of prime order is nilpotent. Feit, Hall, and Thompson show that all finite simple CN groups of odd order are cyclic. Suzuki introduces the Suzuki groups, with types 2B2. Ree introduces the Ree groups, with types 2F4 and 2G2. Feit and Thompson prove the odd order theorem. Tits introduces BN pairs for groups of Lie type and finds the Tits group The GorensteinWalter theorem classifies groups with a dihedral Sylow 2-subgroup. Glauberman proves the Z* theorem Janko introduces the Janko group J1, the first new sporadic group for about a century. Glauberman proves the ZJ theorem Higman and Sims introduce the HigmanSims group Conway introduces the Conway groups Walter's theorem classifies groups with abelian Sylow 2-subgroups Introduction of the Suzuki sporadic group, the Janko group J2, the Janko group J3, the McLaughlin group, and the Held group. Gorenstein introduces signalizer functors based on Thompson's ideas. Bender introduced the generalized Fitting subgroup The AlperinBrauerGorenstein theorem classifies groups with quasi-dihedral or wreathed Sylow 2-subgroups, completing the classification of the simple groups of 2-rank at most 2 Fischer introduces the three Fischer groups Thompson classifies quadratic pairs Bender classifies group with a strongly embedded subgroup Gorenstein proposes a 16-step program for classifying finite simple groups; the final classification follows his outline quite closely. Lyons introduces the Lyons group Rudvalis introduces the Rudvalis group Fisher discovers the baby monster group (unpublished), which Fischer and Griess use to discover the monster group, which in turn leads Thompson to the Thompson sporadic group and Norton to the HaradaNorton group (also found in a different way by Harada). Thompson classifies N-groups, groups all of whose local subgroups are solvable. The GorensteinHarada theorem classifies the simple groups of sectional 2-rank at most 4, dividing the remaining finite simple groups into those of component type and those of characteristic 2 type. Tits shows that groups with BN pairs of rank at least 3 are groups of Lie type Aschbacher classifies the groups with a proper 2-generated core Gorenstein and Walter prove the L-balance theorem Glauberman proves the solvable signalizer functor theorem Aschbacher proves the component theorem, showing roughly that groups of odd type satisfying some conditions have a component in standard form. The groups with a component of standard form were classified in a large collection of papers by many authors. O'Nan introduces the O'Nan group Janko introduces the Janko group J4, the last sporadic group to be discovered

1960 1960 1960 1961 1963 1964 1965 1966 1966 1968 1968 1968 1969 1969 1969 1970 1970

1971 1971 1971 1972 1972 1973 1973

1974 1974

1974 1974 1975 1976 1976

1976 1976

Classification of finite simple groups

66

1977

Aschbacher characterizes the groups of Lie type of odd characteristic in his classical involution theorem. After this theorem, which in some sense deals with "most" of the simple groups, it was generally felt that the end of the classification was in sight. Timmesfeld proves the O2 extraspecial theorem, breaking the classification of groups of GF(2)-type into several smaller problems. Aschbacher classifies the thin finite groups, which are mostly rank 1 groups of Lie type over fields of even characteristic. Bombieri uses elimination theory to complete Thompson's work on the characterization of Ree groups, one of the hardest steps of the classification. McBride proves the signalizer functor theorem for all finite groups. Griess constructs the monster group by hand The GilmanGriess theorem classifies groups groups of characteristic 2 type and rank at least 4 with standard components, one of the three cases of the trichotomy theorem. Aschbacher proves that no finite group satisfies the hypothesis of the uniqueness case, one of the three cases given by the trichotomy theorem for groups of characteristic 2 type. Gorenstein and Lyons prove the trichotomy theorem for groups of characteristic 2 type and rank at least 4, while Aschbacher does the case of rank 3. This divides these groups into 3 subcases: the uniqueness case, groups of GF(2) type, and groups with a standard component. Gorenstein announces the proof of the classification is complete, somewhat prematurely as the proof of the quasithin case was incomplete. Gorenstein, Lyons, and Solomon begin publication of the revised classification Aschbacher and Smith publish their work on quasithin groups (which are mostly groups of Lie type of rank at most 2 over fields of even characteristic), filling the last (known) gap in the classification.

1978 1978 1981

1982 1982 1983

1983

1983

1983

1994 2004

Second-generation classification
The proof of the theorem, as it stood around 1985 or so, can be called first generation. Because of the extreme length of the first generation proof, much effort has been devoted to finding a simpler proof, called a second-generation classification proof. This effort, called "revisionism", was originally led by Daniel Gorenstein. As of 2005, six volumes of the second generation proof have been published (Gorenstein, Lyons & Solomon1994, 1996, 1998, 1999, 2002, 2005), with most of the balance of the proof in manuscript. It is estimated that the new proof will eventually fill approximately 5,000 pages. (This length stems in part from second generation proof being written in a more relaxed style.) Aschbacher and Smith wrote their two volumes devoted to the quasithin case in such a way that those volumes can be part of the second generation proof. Gorenstein and his collaborators have given several reasons why a simpler proof is possible. The most important is that the correct, final statement of the theorem is now known. Simpler techniques can be applied that are known to be adequate for the types of groups we know to be finite simple. In contrast, those who worked on the first generation proof did not know how many sporadic groups there were, and in fact some of the sporadic groups (e.g., the Janko groups) were discovered while proving other cases of the classification theorem. As a result, many of the pieces of the theorem were proved using techniques that were overly general. Because the conclusion was unknown, the first generation proof consists of many stand-alone theorems, dealing with important special cases. Much of the work of proving these theorems was devoted to the analysis of numerous special cases. Given a larger, orchestrated proof, dealing with many of these special cases can be postponed until the most powerful assumptions can be applied. The price paid under this revised strategy is that these first generation theorems no longer have comparatively short proofs, but instead rely on the complete classification. Many first generation theorems overlap, and so divide the possible cases in inefficient ways. As a result, families and subfamiles of finite simple groups were identified multiple times. The revised proof eliminates these

Classification of finite simple groups redundancies by relying on a different subdivision of cases. Finite group theorists have more experience at this sort of exercise, and have new techniques at their disposal. Aschbacher (2004) has called the work on the classification problem by Ulrich Meierfrankenfeld, Bernd Stellmacher, Gernot Stroth, and a few others, a third generation program. One goal of this is to treat all groups in characteristic 2 uniformly using the amalgam method.

67

References
Aschbacher, Michael (2004). "The Status of the Classification of the Finite Simple Groups" (http://www.ams. org/notices/200407/fea-aschbacher.pdf). Notices of the American Mathematical Society. Aschbacher, Michael; Lyons, Richard; Smith, Stephen D.; Solomon, Ronald (2011), The Classification of Finite Simple Groups: Groups of Characteristic 2 Type (http://www.ams.org/bookstore?fn=20&ikey=SURV-172), Mathematical Surveys and Monographs, 172, ISBN978-0-8218-5336-8 Conway, John Horton; Curtis, Robert Turner; Norton, Simon Phillips; Parker, Richard A; Wilson, Robert Arnott (1985), Atlas of Finite Groups: Maximal Subgroups and Ordinary Characters for Simple Groups, Oxford University Press, ISBN0198531990 Gorenstein, D. (1979), "The classification of finite simple groups. I. Simple groups and local analysis", American Mathematical Society. Bulletin. New Series 1 (1): 43199, doi:10.1090/S0273-0979-1979-14551-8, ISSN0002-9904, MR513750 Gorenstein, D. (1982), Finite simple groups, University Series in Mathematics, New York: Plenum Publishing Corp., ISBN978-0-306-40779-6, MR698782 Gorenstein, D. (1983), The classification of finite simple groups. Vol. 1. Groups of noncharacteristic 2 type, The University Series in Mathematics, Plenum Press, ISBN978-0-306-41305-6, MR746470 Daniel Gorenstein (1985), "The Enormous Theorem", Scientific American, vol. 253, no. 6, pp.104115. Gorenstein, D. (1986), "Classifying the finite simple groups", American Mathematical Society. Bulletin. New Series 14 (1): 198, doi:10.1090/S0273-0979-1986-15392-9, ISSN0002-9904, MR818060 Gorenstein, D.; Lyons, Richard; Solomon, Ronald (1994), The classification of the finite simple groups (http:// www.ams.org/online_bks/surv401), Mathematical Surveys and Monographs, 40, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-0334-9, MR1303592 Gorenstein, D.; Lyons, Richard; Solomon, Ronald (1996), The classification of the finite simple groups. Number 2. Part I. Chapter G (http://www.ams.org/online_bks/surv402), Mathematical Surveys and Monographs, 40, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-0390-5, MR1358135 Gorenstein, D.; Lyons, Richard; Solomon, Ronald (1998), The classification of the finite simple groups. Number 3. Part I. Chapter A, Mathematical Surveys and Monographs, 40, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-0391-2, MR1490581 Gorenstein, D.; Lyons, Richard; Solomon, Ronald (1999), The classification of the finite simple groups. Number 4. Part II. Chapters 14, Mathematical Surveys and Monographs, 40, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-1379-9, MR1675976 Gorenstein, D.; Lyons, Richard; Solomon, Ronald (2002), The classification of the finite simple groups. Number 5. Part III. Chapters 16, Mathematical Surveys and Monographs, 40, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-2776-5, MR1923000 Gorenstein, D.; Lyons, Richard; Solomon, Ronald (2005), The classification of the finite simple groups. Number 6. Part IV, Mathematical Surveys and Monographs, 40, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-2777-2, MR2104668 Mark Ronan, Symmetry and the Monster, ISBN 978-0-19-280723-6, Oxford University Press, 2006. (Concise introduction for lay reader)

Classification of finite simple groups Marcus du Sautoy, Finding Moonshine, Fourth Estate, 2008, ISBN 978-0-00-721461-7 (another introduction for the lay reader) Ron Solomon (1995) " On Finite Simple Groups and their Classification, (http://www.ams.org/notices/ 199502/solomon.pdf)" Notices of the American Mathematical Society. (Not too technical and good on history) Solomon, Ronald (2001), "A brief history of the classification of the finite simple groups" (http://www.ams. org/bull/2001-38-03/S0273-0979-01-00909-0/S0273-0979-01-00909-0.pdf), American Mathematical Society. Bulletin. New Series 38 (3): 315352, doi:10.1090/S0273-0979-01-00909-0, ISSN0002-9904, MR1824893 article won Levi L. Conant prize (http://www.ams.org/notices/200604/comm-conant.pdf) for exposition Thompson, John G. (1984), "Finite nonsolvable groups", in Gruenberg, K. W.; Roseblade, J. E., Group theory. Essays for Philip Hall, Boston, MA: Academic Press, pp.112, ISBN978-0-12-304880-6, MR780566 Wilson, Robert A. (2009), The finite simple groups, Graduate Texts in Mathematics 251, 251, Berlin, New York: Springer-Verlag, doi:10.1007/978-1-84800-988-2, ISBN978-1-84800-987-5, Zbl05622792

68

External links
ATLAS of Finite Group Representations. (http://brauer.maths.qmul.ac.uk/Atlas/v3/) Searchable database of representations and other data for many finite simple groups. Elwes, Richard, " An enormous theorem: the classification of finite simple groups, (http://plus.maths.org/ issue41/features/elwes/index.html)" Plus Magazine, Issue 41, December 2006. For laypeople. Madore, David (2003) Orders of nonabelian simple groups. (http://www.eleves.ens.fr:8080/home/madore/ math/simplegroups.html) Includes a list of all nonabelian simple groups up to order 1010.

Cohn's irreducibility criterion


Arthur Cohn's irreducibility criterion is a test to determine whether a polynomial is irreducible in The criterion is often stated as follows: If a prime number is expressed in base 10 as (where .

) then the polynomial is irreducible in .

The theorem can be generalized to other bases as follows: Assume that is a natural number and is a

polynomial such that . If is a prime number then is irreducible in . The base-10 version of the theorem attributed to Cohn by Plya and Szeg in one of their books[1] while the generalization to any base, 2 or greater, is due to Brillhart, Filaseta, and Odlyzko[2] . In 2002, Ram Murty gave a simplified proof as well as some history of the theorem in a paper that is available online.[3] . The converse of this criterion is that, if p is an irreducible polynomial with integer coefficients that have greatest common divisor 1, then there exists a base such that the coefficients of p form the representation of a prime number in that base; this is the Bunyakovsky conjecture and its truth or falsity remains an open question.

Cohn's irreducibility criterion

69

Historical notes
Polya and Szeg gave their own generalization but it has many side conditions (on the locations of the roots, for instance) so it lacks the elegance of Brillhart's, Filaseta's, and Odlyzko's generalization. It is clear from context that the "A. Cohn" mentioned by Polya and Szeg is Arthur Cohn, a student of Issai Schur who was awarded his PhD in Berlin in 1921.[4]

References
[1] George Plya; Gbor Szeg (1925). Aufgaben und Lehrstze aus der Analysis, Bd 2. Springer, Berlin. OCLC73165700. English translation in: George Plya; Gabor Szeg (2004). Problems and theorems in analysis, volume 2. 2. Springer. p.137. ISBN3-540-63686-2. [2] Brillhart, John; Michael Filaseta, Andrew Odlyzko (1981). "On an irreducibility theorem of A. Cohn". Canadian Journal of Mathematics 33 (5): 10551059. doi:10.4153/CJM-1981-080-0. [3] Murty, Ram (2002). "Prime Numbers and Irreducible Polynomials" (http:/ / www. mast. queensu. ca/ ~murty/ polya4. dvi). American Mathematical Monthly (The American Mathematical Monthly, Vol. 109, No. 5) 109 (5): 452458. doi:10.2307/2695645. JSTOR2695645. . (dvi file) [4] Arthur Cohn's entry at the Mathematics Genealogy Project (http:/ / genealogy. math. ndsu. nodak. edu/ html/ id. phtml?id=17963)

External links
A. Cohn's irreducibility criterion (http://planetmath.org/?op=getobj&amp;from=objects&amp;id=6194) on PlanetMath

Cramer's rule
In linear algebra, Cramer's rule is a theorem, which gives an expression for the solution of a system of linear equations with as many equations as unknowns, valid in those cases where there is a unique solution. The solution is expressed in terms of the determinants of the (square) coefficient matrix and of matrices obtained from it by replacing one column by the vector of right hand sides of the equations. It is named after Gabriel Cramer (17041752), who published the rule in his 1750 Introduction l'analyse des lignes courbes algbriques (Introduction to the analysis of algebraic curves), although Colin Maclaurin also published the method in his 1748 Treatise of Algebra (and probably knew of the method as early as 1729).[1] [2]

General case
Consider a system of linear equations represented in matrix multiplication form as follows:

where the square matrix

is invertible and the vector

is the column vector of the variables.

Then the theorem states that:

where

is the matrix formed by replacing the ith column of

by the column vector

The rule holds for systems of equations with coefficients and unknowns in any field, not just in the real numbers. It has recently been shown that Cramer's rule can be implemented in O(n3) time,[3] which is comparable to more common methods of solving systems of linear equations, such as Gaussian elimination.

Cramer's rule

70

Proof
The proof for Cramer's rule is very simple, in fact it uses just two properties of determinants. The first property is that adding any multiple of one column to another does not change the value of the determinant, while the second property is that multiplying every element of one column by a factor will multiply the value of the determinant by the same factor. Given n linear equations with n variables .

Cramer's rule gives, for the value of

, the expression:

which can be verified using the aforementioned properties of determinants. In fact, substituting for from the equations of the system, this quotient is equal to

By subtracting from the first column the second multiplied by until the last column multiplied by , it is found to be equal to

, the third column multiplied by

, and so on

and according to the remaining property of determinants the common factor of numerator can be factored out of the determinant. Therefore this is equal to

in the first column of the

Cramer's rule

71

In the same way, if the columns of b's is used to replace the k-th column of the matrix of the system of equations the result will be equal to . As a result we get that

These expressions for can be put into matrix notation as follows. First do a Laplace expansion (aka cofactor expansion) on the determinants which are in the numerators using the columns which contain , , , . Thus Cramer's rule becomes;

this is equivalent to the matrix equation

Where

are the cofactors of the coefficient matrix [A]

is the determinant of the matrix formed by deleting row r and column c from is called the adjugate matrix of , written as adj[A].

Cramer's rule

72

Finding inverse matrix


Let A be an nn matrix. Then

where Adj(A) denotes the adjugate matrix of A, det(A) is the determinant, and I is the identity matrix. If det(A) is invertible in R, then the inverse matrix of A is

If R is a field (such as the field of real numbers), then this gives a formula for the inverse of A, provided det(A)0. In fact, this formula will work whenever R is a commutative ring, provided that det(A) is a unit. If det(A) is not a unit, then A is not invertible.

Applications
Explicit formulas for small systems
Consider the linear system be found with Cramer's rule as and which in matrix format is Then, x and y can

The rules for 33 are similar. Given

which in matrix format is

the values of x, y and z can be found as follows:

Cramer's rule

73

Differential geometry
Cramer's rule is also extremely useful for solving problems in differential geometry. Consider the two equations and . When u and v are independent variables, we can define and Finding an equation for is a trivial application of Cramer's rule.

First, calculate the first derivatives of F, G, x, and y:

Substituting dx, dy into dF and dG, we have:

Since u, v are both independent, the coefficients of du, dv must be zero. So we can write out equations for the coefficients:

Now, by Cramer's rule, we see that:

This is now a formula in terms of two Jacobians:

Similar formulae can be derived for

Cramer's rule

74

Algebra
Cramer's rule can be used to prove the CayleyHamilton theorem of linear algebra, as well as Nakayama's lemma, which is fundamental in commutative ring theory.

Integer programming
Cramer's rule can be used to prove that an integer programming problem whose constraint matrix is totally unimodular and whose right-hand side is integer, has integer basic solutions. This makes the integer program substantially easier to solve.

Ordinary differential equations


Cramer's rule is used to derive the general solution to an inhomogeneous linear differential equation by the method of variation of parameters.

Geometric interpretation
Cramer's rule has a geometric interpretation that can be considered also a proof or simply giving insight about its geometric nature. These geometric arguments work in general and not only in the case of two equations with two unknowns presented here. Given the system of equations

Geometric interpretation of Cramer's rule. The areas of the second and third shaded parallelograms are the same and the second is times the first. From this equality Cramer's rule follows.

it can be considered as an equation between vectors

Cramer's rule

75

The area of the parallelogram determined by equations:

and

is given by the determinant of the system of

In general, when there are more variables and equations, the determinant of vectors of length volume of the parallelepiped determined by those vectors in the -th dimensional Euclidean space. Therefore the area of the parallelogram determined by and has to be

will give the

times the area of the

first one since one of the sides has been multiplied by this factor. Now, this last parallelogram, by Cavalieri's principle, has the same area as the parallelogram determined by Equating the areas of this last and the second parallelogram gives the equation and .

from which Cramer's rule follows.

Incompatible and indeterminate cases


A system of equations is said to be incompatible when there are no solutions and it is called indeterminate when there is more than one solution. For linear equations, an indeterminate system will have infinitely many solutions (if it is over an infinite field), since the solutions can be expressed in terms of one or more parameters that can take arbitrary values. Cramer's rule applies to the case where the coefficient determinant is nonzero. In the contrary case the system is either incompatible or indeterminate, based on the values of the determinants only for 2x2 systems. For 3x3 or higher systems, the only thing one can say when the coefficient determinant equals zero is: if any of the "numerator" determinants are nonzero, then the system must be incompatible. However, the converse is false: having all determinants zero does not imply that the system is indeterminate. A simple example where all determinants vanish but the system is still incompatible is the 3x3 system x+y+z=1, x+y+z=2, x+y+z=3.

Notes
[1] Carl B. Boyer, A History of Mathematics, 2nd edition (Wiley, 1968), p. 431. [2] Victor Katz, A History of Mathematics, Brief edition (Pearson Education, 2004), pp. 378379. [3] Ken Habgood, Itamar Arel, A condensation-based application of Cramers rule for solving large-scale linear systems, Journal of Discrete Algorithms, ISSN 1570-8667, 10.1016/j.jda.2011.06.007.

External links
Proof of Cramer's Rule (http://planetmath.org/encyclopedia/ProofOfCramersRule.html) WebApp descriptively solving systems of linear equations with Cramer's Rule (http://sole.ooz.ie/) Cramer's Rule Solver (http://www.cramersrule.info/index.html) Online Calculator of System of linear equations (http://www.stud.feec.vutbr.cz/~xvapen02/vypocty/linrov. php?language=english)

Crystallographic restriction theorem

76

Crystallographic restriction theorem


The crystallographic restriction theorem in its basic form was based on the observation that the rotational symmetries of a crystal are usually limited to 2-fold, 3-fold, 4-fold, and 6-fold. However, quasicrystals can occur with other symmetries, such as 5-fold; these were not discovered until 1984[1] Prior to the discovery of quasicrystals, crystals were modeled as discrete lattices, generated by a list of independent finite translations (Coxeter 1989). Because discreteness requires that the spacings between lattice points have a lower bound, the group of rotational symmetries of the lattice at any point must be a finite group. The strength of the theorem is that not all finite groups are compatible with a discrete lattice; in any dimension, we will have only a finite number of compatible groups.

Dimensions 2 and 3
The special cases of 2D (wallpaper groups) and 3D (space groups) are most heavily used in applications, and we can treat them together.

Lattice proof
A rotation symmetry in dimension 2 or 3 must move a lattice point to a succession of other lattice points in the same plane, generating a regular polygon of coplanar lattice points. We now confine our attention to the plane in which the symmetry acts (Scherrer 1946). (We might call this a proof in the style of Busby Berkeley, with lattice vectors rather than pretty ladies dancing and swirling in geometric patterns.) Now consider an 8-fold rotation, and the displacement vectors between adjacent points of the polygon. If a displacement exists between any two lattice points, then that same displacement is repeated everywhere in the lattice. So collect all the edge displacements to begin at a single lattice point. The edge vectors become radial vectors, and their 8-fold symmetry implies a regular octagon of lattice points around the collection point. But this is impossible, because the new octagon is about 80% smaller than the original. The significance of the shrinking is that it is unlimited. The same construction can be repeated with the new octagon, and again and again until the distance between lattice points is as small as we like; thus no discrete lattice can have 8-fold symmetry. The same argument applies to any k-fold rotation, for k greater than 6.

Lattices restrict polygons Compatible: 6-fold (3-fold), 4-fold (2-fold) Incompatible: 8-fold, 5-fold

A shrinking argument also eliminates 5-fold symmetry. Consider a regular pentagon of lattice points. If it exists, then we can take every other edge displacement and (head-to-tail) assemble a 5-point star, with the last edge returning to the starting point. The vertices of such a star are again vertices of a regular pentagon with 5-fold symmetry, but about 60% smaller than the original. Thus the theorem is proved.

Crystallographic restriction theorem The existence of quasicrystals and Penrose tilings shows that the assumption of a linear translation is necessary. Penrose tilings may have 5-fold rotational symmetry and a discrete lattice, and any local neighborhood of the tiling is repeated infinitely many times, but there is no linear translation for the tiling as a whole. And without the discrete lattice assumption, the above construction not only fails to reach a contradiction, but produces a (non-discrete) counterexample. Thus 5-fold rotational symmetry cannot be eliminated by an argument missing either of those assumptions. A Penrose tiling of the whole (infinite) plane can only have exact 5-fold rotational symmetry (of the whole tiling) about a single point, however, whereas the 4-fold and 6-fold lattices have infinitely many centres of rotational symmetry.

77

Trigonometry proof
Consider two lattice points A and B separated by a translation vector r. Consider an angle such that a rotation of angle about any lattice point is a symmetry of the lattice. Rotating about point B by maps point A to a new point A'. Similarly, rotating about point A by , in the opposite direction, maps B to a point B'. Since both rotations mentioned are symmetry operations, A' and B' must both be lattice points. Due to periodicity of the crystal, the new vector r' which connects them must be equal to an integral multiple of r: r = mr' The four translation vectors (three of which are given by r, and one which connects A' and B' given by r') form a parallelogram. Therefore, the length of r' is also given by: r = -2rcos() + r' Combining the two equations gives: cos() = (1 - m)/2 = M/2 Where M = 1-m is also an integer. Since we always have: | cos() | 1 it follows that: |M|2 It can be shown that the only values of in the 0 to 180 range that satisfy the three equations above are 0, 60, 90, 120, and 180. In terms of radians, the only allowed rotations consistent with lattice periodicity are given by 2/n, where n = 1, 2, 3, 4, 6. This corresponds to 1-, 2-, 3-, 4-, and 6-fold symmetry, respectively, and therefore excludes the possibility of 5-fold or greater than 6-fold symmetry.

Short trigonometry proof


Consider a line of atoms A-O-B, separated by distance t. Rotate the entire row by +2/n and -2/n, with point O kept fixed. Due to the assumed periodicity of the lattice, the two lattice points C and D will be also in a directly below the initial row; moreover C and D will be separated by r = mt, with m an integer. But by the geometry, the separation between these points is 2tcos(2/n). Equating the two relations gives 2cos(2/n) = m. This is satisfied by only n = 1, 2, 3, 4, 6. (See also the sketch in section Preuve mathmatique simple of the French Wikipedia article (Thorme de restriction cristallographique)

Crystallographic restriction theorem

78

Matrix proof
For an alternative proof, consider matrix properties. The sum of the diagonal elements of a matrix is called the trace of the matrix. In 2D and 3D every rotation is a planar rotation, and the trace is a function of the angle alone. For a 2D rotation, the trace is 2 cos ; for a 3D rotation, 1 + 2 cos . Examples Consider a 60 (6-fold) rotation matrix with respect to an orthonormal basis in 2D.

The trace is precisely 1, an integer. Consider a 45 (8-fold) rotation matrix.

The trace is 2/2, not an integer. Using a lattice basis, neither orthogonality nor unit length is guaranteed, only independence. However, the trace is the same with respect to any basis. (Similarity transforms preserve trace.) In a lattice basis, because the rotation must map lattice points to lattice points, each matrix entry and hence the trace must be an integer. Thus, for example, wallpaper and crystals cannot have 8-fold rotational symmetry. The only possibilities are multiples of 60, 90, 120, and 180, corresponding to 6-, 4-, 3-, and 2-fold rotations. Example Consider a 60 (360/6) rotation matrix with respect to the oblique lattice basis for a tiling by equilateral triangles.

The trace is still 1. The determinant (always +1 for a rotation) is also preserved. The general crystallographic restriction on rotations does not guarantee that a rotation will be compatible with a specific lattice. For example, a 60 rotation will not work with a square lattice; nor will a 90 rotation work with a rectangular lattice.

Higher dimensions
When the dimension of the lattice rises to four or more, rotations need no longer be planar; the 2D proof is inadequate. However, restrictions still apply, though more symmetries are permissible. For example, the hypercubic lattice has an eightfold rotational symmetry, corresponding to an eightfold rotational symmetry of the hypercube. This is of interest, not just for mathematics, but for the physics of quasicrystals under the cut-and-project theory. In this view, a 3D quasicrystal with 8-fold rotation symmetry might be described as the projection of a slab cut from a 4D lattice. The following 4D rotation matrix is the aforementioned eightfold symmetry of the hypercube (and the cross-polytope):

Transforming this matrix to the new coordinates given by

Crystallographic restriction theorem

79

will produce:

This third matrix then corresponds to a rotation both by 45 (in the first two dimensions) and by 135 (in the last two). Projecting a slab of hypercubes along the first two dimensions of the new coordinates produces an AmmannBeenker tiling (another such tiling is produced by projecting along the last two dimensions), which therefore also has 8-fold rotational symmetry on average. The A4 lattice and F4 lattice have order 10 and order 12 rotational symmetries, respectively. To state the restriction for all dimensions, it is convenient to shift attention away from rotations alone and concentrate on the integer matrices (Bamberg, Cairns & Kilminster 2003). We say that a matrix A has order k when its k-th power (but no lower), Ak, equals the identity. Thus a 6-fold rotation matrix in the equilateral triangle basis is an integer matrix with order 6. Let OrdN denote the set of integers that can be the order of an NN integer matrix. For example, Ord2 = {1, 2, 3, 4, 6}. We wish to state an explicit formula for OrdN. Define a function based on Euler's totient function ; it will map positive integers to non-negative integers. For an odd prime, p, and a positive integer, k, set (pk) equal to the totient function value, (pk), which in this case is pkpk1. Do the same for (2k) when k > 1. Set (2) and (1) to 0. Using the fundamental theorem of arithmetic, we can write any other positive integer uniquely as a product of prime powers, m = pk ; set (m) = (pk ). This differs from the totient itself, because it is a sum instead of a product. The crystallographic restriction in general form states that OrdN consists of those positive integers m such that (m) N.

Smallest dimension for a given order


m

A080737

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 4 10 4 12 6 6 8 16 6 18 6 8 10 22 6 20 12 18 8 28 6 30

(m) 0 0 2 2 4 2 6 4 6

Note that these additional symmetries do not allow a planar slice to have, say, 8-fold rotation symmetry. In the plane, the 2D restrictions still apply. Thus the cuts used to model quasicrystals necessarily have thickness. Integer matrices are not limited to rotations; for example, a reflection is also a symmetry of order 2. But by insisting on determinant +1, we can restrict the matrices to proper rotations.

Formulation in terms of isometries


The crystallographic restriction theorem can be formulated in terms of isometries of Euclidean space. A set of isometries can form a group. By a discrete isometry group we will mean an isometry group that maps every point to a discrete subset of RN, i.e. a set of isolated points. With this terminology, the crystallographic restriction theorem in two and three dimensions can be formulated as follows. For every discrete isometry group in two- and three-dimensional space which includes translations spanning the whole space, all isometries of finite order are of order 1, 2, 3, 4 or 6. Note that isometries of order n include, but are not restricted to, n-fold rotations. The theorem also excludes S8, S12, D4d, and D6d (see point groups in three dimensions), even though they have 4- and 6-fold rotational symmetry only.

Crystallographic restriction theorem Note also that rotational symmetry of any order about an axis is compatible with translational symmetry along that axis. The result in the table above implies that for every discrete isometry group in four- and five-dimensional space which includes translations spanning the whole space, all isometries of finite order are of order 1, 2, 3, 4, 5, 6, 8, 10, or 12. All isometries of finite order in six- and seven-dimensional space are of order 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 12, 14, 15, 18, 20, 24 or 30 .

80

Notes
[1] Shechtman et al (1984)

References
Bamberg, John; Cairns, Grant; Kilminster, Devin (March 2003), "The crystallographic restriction, permutations, and Goldbach's conjecture" (http://www.latrobe.edu.au/mathstats/staff/cairns/papers/42.pdf), American Mathematical Monthly 110 (3): 202209, doi:10.2307/3647934, JSTOR3647934 Elliott, Stephen (1998), The Physics and Chemistry of Solids, Wiley, ISBN0-471-98194-X Coxeter, H. S. M. (1989), Introduction to Geometry (2nd ed.), Wiley, ISBN978-0-471-50458-0 Scherrer, W. (1946), "Die Einlagerung eines regulren Vielecks in ein Gitter" (http://www-gdz.sub. uni-goettingen.de/cgi-bin/digbib.cgi?PPN378850199_0001), Elemente der Mathematik 1 (6): 9798 Shechtman, D.; Blech, I.; Gratias, D.; Cahn, JW (1984), "Metallic phase with long-range orientational order and no translational symmetry", Physical Review Letters 53 (20): 19511953, Bibcode1984PhRvL..53.1951S, doi:10.1103/PhysRevLett.53.1951

External links
The crystallographic restriction (http://www-history.mcs.st-and.ac.uk/~john/geometry/Lectures/A2.html)

Descartes' rule of signs

81

Descartes' rule of signs


In mathematics, Descartes' rule of signs, first described by Ren Descartes in his work La Gomtrie, is a technique for determining the number of positive or negative real roots of a polynomial. The rule gives us an upper bound number of positive or negative roots of a polynomial. It is not a deterministic rule, i.e. it does not tell the exact number of positive or negative roots.

Descartes' rule of signs


Positive roots
The rule states that if the terms of a single-variable polynomial with real coefficients are ordered by descending variable exponent, then the number of positive roots of the polynomial is either equal to the number of sign differences between consecutive nonzero coefficients, or is less than it by a multiple of 2. Multiple roots of the same value are counted separately.

Negative roots
As a corollary of the rule, the number of negative roots is the number of negative integers after negating the coefficients of odd-power terms (otherwise seen as substituting the negation of the variable for the variable itself), or fewer than it by a multiple of 2.

Example
For example, the polynomial

has one sign change between the second and third terms (++, +, ). Therefore it has exactly one positive root. Note that the leading sign needs to be considered although it doesn't affect the answer in this case. In fact, this polynomial factors as

so the roots are 1 (twice) and 1. Now consider the polynomial

This polynomial has two sign changes (+, ++, +), meaning the original polynomial has two or zero negative roots and this second polynomial has two or zero positive roots. The factorization of the second polynomial is

So here, the roots are 1 (twice) and 1, the negation of the roots of the original polynomial. Since any nth degree polynomial equation has exactly n roots, the minimum number of complex roots is equal to

Where p denotes the maximum number of positive roots, and q denotes the maximum number of negative roots (both of which can be found out using descarte's rule of sign), and n denotes the degree of the equation.

Descartes' rule of signs

82

Special case
The exclusion of multiples of 2 is because the polynomial may have complex roots which always come in pairs. Thus if the polynomial is known to have all real roots, this rule allows one to find the exact number of positive and negative roots. Since it is easy to determine the multiplicity of zero as a root, the sign of all roots can be determined in this case.

Generalizations
If the real polynomial P has k real positive roots counted with multiplicity, then for every a > 0 there are at least k changes of sign in the sequence of coefficients of the Taylor series of the function eaxP(x).[1] In the 1970s Askold Georgevich Khovanski developed the theory of fewnomials that generalises Descartes' rule.[2] The rule of signs can be thought of as stating that the number of real roots of a polynomial is dependent on the polynomial's complexity, and that this complexity is proportional to the number of monomials it has, not its degree. Khovanski showed that this holds true not just for polynomials but for algebraic combinations of many transcendental functions, the so-called Pfaffian functions.

Notes
[1] Vladimir P. Kostov, A mapping defined by the Schur-Szeg composition, Comptes Rendus Acad. Bulg. Sci. tome 63, No. 7, 2010, 943 - 952. [2] A. G. Khovanskii, Fewnomials, Princeton University Press (1991) ISBN 0821845470.

External links
Descartes Rule of Signs (http://www.cut-the-knot.org/fta/ROS2.shtml) Proof of the Rule This article incorporates material from Descartes' rule of signs on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.

Dirichlet's unit theorem

83

Dirichlet's unit theorem


In mathematics, Dirichlet's unit theorem is a basic result in algebraic number theory due to Gustav Lejeune Dirichlet.[1] It determines the rank of the group of units in the ring OK of algebraic integers of a number field K. The regulator is a positive real number that determines how "dense" the units are.

Dirichlet's unit theorem


The statement is that the group of units is finitely generated and has rank (maximal number of multiplicatively independent elements) equal to r = r1 + r2 1 where r1 is the number of real embeddings and r2 the number of conjugate pairs of complex embeddings of K. This characterisation of r1 and r2 is based on the idea that there will be as many ways to embed K in the complex number field as the degree n = [K : Q]; these will either be into the real numbers, or pairs of embeddings related by complex conjugation, so that n = r1 + 2r2. Other ways of determining r1 and r2 are use the primitive element theorem to write K = Q(), and then r1 is the number of conjugates of that are real, 2r2 the number that are complex; write the tensor product of fields K QR as a product of fields, there being r1 copies of R and r2 copies of C. As an example, if K is a quadratic field, the rank is 1 if it is a real quadratic field, and 0 if an imaginary quadratic field. The theory for real quadratic fields is essentially the theory of Pell's equation. The rank is > 0 for all number fields besides Q and imaginary quadratic fields, which have rank 0. The 'size' of the units is measured in general by a determinant called the regulator. In principle a basis for the units can be effectively computed; in practice the calculations are quite involved when n is large. The torsion in the group of units is the set of all roots of unity of K, which form a finite cyclic group. For a number field with at least one real embedding the torsion must therefore be only {1,1}. There are number fields, for example most imaginary quadratic fields, having no real embeddings which also have {1,1} for the torsion of its unit group. Totally real fields are special with respect to units. If L/K is a finite extension of number fields with degree greater than 1 and the units groups for the integers of L and K have the same rank then K is totally real and L is a totally complex quadratic extension. The converse holds too. (An example is K equal to the rationals and L equal to an imaginary quadratic field; both have unit rank 0.) There is a generalisation of the unit theorem by Helmut Hasse (and later Claude Chevalley) to describe the structure of the group of S-units, determining the rank of the unit group in localizations of rings of integers. Also, the Galois module structure of has been determined.[2]

The regulator
Suppose that u1,...,ur are a set of generators for the unit group modulo roots of unity. If u is an algebraic number, write u1, ..., ur+1 for the different embeddings into R or C, and write Ni for the degree of the corresponding embedding over R (so it is 1 for real embeddings and 2 for complex ones). Then the r by r+1 matrix whose entries are has the property that the sum of any row is zero (because all units have norm 1, and the log of the norm is the sum of the entries of a row). This implies that the absolute value R of the determinant of the submatrix formed by deleting one column is independent of the column. The number R is called the regulator of the algebraic

Dirichlet's unit theorem number field (it does not depend on the choice of generators ui). It measures the "density" of the units: if the regulator is small, this means that there are "lots" of units. The regulator has the following geometric interpretation. The map taking a unit u to the vector with entries Nilog|ui| has image in the r-dimensional subspace of Rr+1 consisting of all vectors whose entries have sum 0, and by Dirichlet's unit theorem the image is a lattice in this subspace. The volume of a fundamental domain of this lattice is R(r+1). The regulator of an algebraic number field of degree greater than 2 is usually quite cumbersome to calculate, though there are now computer algebra packages that can do it in many cases. It is usually much easier to calculate the product hR of the class number h and the regulator using the class number formula, and the main difficulty in calculating the class number of an algebraic number field is usually the calculation of the regulator.

84

Examples
The regulator of an imaginary quadratic field, or of the rational integers, is 1 (as the determinant of a 00 matrix is 1). The regulator of the real quadratic field Q(5) is log((5+1)/2). This can be seen as follows. A fundamental unit is (5 + 1)/2, and its images under the two embeddings into R are (5+1)/2 and (5+1)/2. So the r by r+1 matrix is

A fundamental domain in logarithmic space of the group of units of the cyclic cubic field K obtained by adjoining to Q a root of f(x)=x3+x22x1. If denotes a root of f(x), then a set of fundamental units is {1,2} where 1=2+1 and 2=22. The area of the fundamental domain is approximately 0.910114, so the regulator of K is approximately 0.525455.

The regulator of the cyclic cubic field Q(), where is a root of x3+x22x1, is approximately 0.5255. A basis of the group of units modulo roots of unity is {1,2} where 1=2+1 and 2=22.[3]

Dirichlet's unit theorem

85

Higher regulators
A 'higher' regulator refers to a construction for an algebraic K-group with index n > 1 that plays the same role as the classical regulator does for the group of units, which is a group K1. A theory of such regulators has been in development, with work of Armand Borel and others. Such higher regulators play a role, for example, in the Beilinson conjectures, and are expected to occur in evaluations of certain L-functions at integer values of the argument.

Stark regulator
The formulation of Stark's conjectures led Harold Stark to define what is now called the Stark regulator, similar to the classical regulator as a determinant of logarithms of units, attached to any Artin representation.[4] [5]

Notes
[1] [2] [3] [4] [5] Elstrodt 2007, 8.D Neukirch, Schmidt & Wingberg 2000, proposition VIII.8.6.11. Cohen 1993, Table B.4 PDF (http:/ / www. math. tifr. res. in/ ~dprasad/ artin. pdf) PDF (http:/ / www. math. harvard. edu/ ~dasgupta/ papers/ Dasguptaseniorthesis. pdf)

References
Cohen, Henri (1993). A Course in Computational Algebraic Number Theory. Graduate Texts in Mathematics. 138. Berlin, New York: Springer-Verlag. ISBN978-3-540-55640-4. MR1228206 Elstrodt, Jrgen (2007). "The Life and Work of Gustav Lejeune Dirichlet (18051859)" (http://www.uni-math. gwdg.de/tschinkel/gauss-dirichlet/elstrodt-new.pdf) (PDF). Clay Mathematics Proceedings. Retrieved 2010-06-13. Serge Lang, Algebraic number theory, ISBN 0-387-94225-4 Neukirch, Jrgen (1999), Algebraic Number Theory, Grundlehren der mathematischen Wissenschaften, 322, Berlin: Springer-Verlag, ISBN978-3-540-65399-8, MR1697859 Neukirch, Jrgen; Schmidt, Alexander; Wingberg, Kay (2000), Cohomology of Number Fields, Grundlehren der Mathematischen Wissenschaften, 323, Berlin: Springer-Verlag, ISBN978-3-540-66671-4, MR1737196

Engel theorem

86

Engel theorem
In representation theory, Engel's theorem is one of the basic theorems in the theory of Lie algebras; it asserts that for a Lie algebra two concepts of nilpotency are identical. A useful form of the theorem says that if a Lie algebra L of matrices consists of nilpotent matrices, then they can all be simultaneously brought to a strictly upper triangular form. The theorem is named after the mathematician Friedrich Engel, who sketched a proof of it in a letter to Killing dated 20 July 1890 (Hawkins 2000, p. 176). Engel's student Umlauf gave a complete proof in his 1891 dissertation, reprinted as (Umlauf 2010). A linear operator T on a vector space V is defined to be nilpotent if there is a positive integer k such that Tk = 0. For example, any operator given by a matrix whose entries are zero on and below its diagonal is nilpotent.

An element x of a Lie algebra L is ad-nilpotent if and only if the linear operator on L defined by

is nilpotent. Note that in the Lie algebra L(V) of linear operators on V, the identity operator IV is ad-nilpotent (because ) but is not a nilpotent operator. A Lie algebra L is nilpotent if and only if the lower central series defined recursively by

eventually reaches {0}. Theorem. A finite-dimensional Lie algebra L is nilpotent if and only if every element of L is ad-nilpotent. Note that no assumption on the underlying base field is required. The key lemma in the proof of Engel's theorem is the following fact about Lie algebras of linear operators on finite dimensional vector spaces which is useful in its own right: Let L be a Lie subalgebra of L(V). Then L consists of nilpotent operators if and only if there is a sequence

of subspaces of V such that

and

Thus Lie algebras of nilpotent operators are simultaneously strictly upper-triangulizable.

References
Erdmann, Karin & Wildon, Mark. Introduction to Lie Algebras, 1st edition, Springer, 2006. ISBN 1-84628-040-0 Hawkins, Thomas (2000), Emergence of the theory of Lie groups [1], Sources and Studies in the History of Mathematics and Physical Sciences, Berlin, New York: Springer-Verlag, ISBN978-0-387-98963-1, MR1771134 G. Hochschild, The Structure of Lie Groups, Holden Day, 1965. J. Humphreys, Introduction to Lie Algebras and Representation Theory, Springer, 1972. Umlauf, Karl Arthur (2010) [1891] (in German), ber Die Zusammensetzung Der Endlichen Continuierlichen Transformationsgruppen, Insbesondre Der Gruppen Vom Range Null [2], Inaugural-Dissertation, Leipzig, Nabu Press, ISBN978-1-141-58889-3

Engel theorem

87

References
[1] http:/ / books. google. com/ books?isbn=978-0-387-98963-1 [2] http:/ / books. google. com/ books?isbn=978-1141588893

Factor theorem
In algebra, the factor theorem is a theorem linking factors and zeros of a polynomial. It is a special case of the polynomial remainder theorem. The factor theorem states that a polynomial has a factor if and only if .

Factorization of polynomials
Two problems where the factor theorem is commonly applied are those of factoring a polynomial and finding the roots of a polynomial equation; it is a direct consequence of the theorem that these problems are essentially equivalent. The factor theorem is also used to remove known zeros from a polynomial while leaving all unknown zeros intact, thus producing a lower degree polynomial whose zeros may be easier to find. Abstractly, the method is as follows: 1. "Guess" a zero of the polynomial . (In general, this can be very hard, but math textbook problems that

involve solving a polynomial equation are often designed so that some roots are easy to discover.) 2. Use the factor theorem to conclude that is a factor of . 3. Compute the polynomial , for example using polynomial long division. 4. Conclude that any root of is a root of . Since the polynomial degree of less than that of , it is "simpler" to find the remaining zeros by studying .

is one

Example
You wish to find the factors at

To do this you would use trial and error to find the first x value that causes the expression to equal zero. To find out if is a factor, substitute into the polynomial above:

As this is equal to 18 and not 0 this means (substituting This is equal to into the polynomial):

is not a factor of

. So, we next try

. Therefore

, which is to say

, is a factor, and by

is a root of to get a quadratic,

The next two roots can be found by algebraically dividing which can be solved directly, by the factor theorem or by the quadratic equation.

and therefore

and

are the factors of

Factor theorem

88

Formal version
Let be a polynomial with complex coefficients, and can be written in the form for which are precisely the roots of be in an integral domain (e.g. where ). Then is also a polynomial. is if and only if

determined uniquely. This indicates that those

. Repeated roots can be found by

application of the theorem to the quotient

, which may be found by polynomial long division.

FeitThompson theorem
In mathematics, the FeitThompson theorem, or odd order theorem, states that every finite group of odd order is solvable. It was proved by Walter Feit and John Griggs Thompson(1962, 1963)

History
The contrast that these results shew between groups of odd and even order suggests inevitably that simple groups of odd order do not exist. William Burnside(1911, p. 503 note M)

William Burnside(1911, p. 503 note M) conjectured that every nonabelian finite simple group has even order. Richard Brauer(1957) suggested using the centralizers of involutions of simple groups as the basis for the classification of finite simple groups, as the Brauer-Fowler theorem shows that there are only a finite number of finite simple groups with given centralizer of an involution. A group of odd order has no involutions, so to carry out Brauer's program it is first necessary to show that non-cyclic finite simple groups never have odd order. This is equivalent to showing that odd order groups are solvable, which is what Feit and Thompson proved. The attack on Burnside's conjecture was started by Michio Suzuki(1957), who studied CA groups; these are groups such that the Centralizer of every non-trivial element is Abelian. In a pioneering paper he showed that all CA groups of odd order are solvable. (He later classified all the simple CA groups, and more generally all simple groups such that the centralizer of any involution has a normal 2-Sylow subgroup, finding an overlooked family of simple groups of Lie type in the process, that are now called Suzuki groups.) Feit, Hall, and Thompson(1960) extended Suzuki's work to the family of CN groups; these are groups such that the Centralizer of every non-trivial element is Nilpotent. They showed that every CN group of odd order is solvable. Their proof is similar to Suzuki's proof. It was about 17 pages long, which at the time was thought to be very long for a proof in group theory. The FeitThompson theorem can be thought of as the next step in this process: they show that there is no non-cyclic simple group of odd order such that every proper subgroup is solvable. This proves that every finite group of odd order is solvable, as a minimal counterexample must be a simple group such that every proper subgroup is solvable. Although the proof follows the same general outline as the CA theorem and the CN theorem, the details are vastly more complicated. The final paper is 255 pages long.

Significance of the proof


The FeitThompson theorem showed that the classification of finite simple groups using centralizers of involutions might be possible, as every nonabelian simple group has an involution. Many of the techniques they introduced in their proof, especially the idea of local analysis, were developed further into tools used in the classification. Perhaps the most revolutionary aspect of the proof was its length: before the Feit-Thompson paper, few arguments in group theory were more than a few pages long and most could be read in a day. Once group theorists realized that such long arguments could work, a series of papers that were several hundred pages long started to appear. Some of these

FeitThompson theorem dwarfed even the FeitThompson paper; Aschbacher and Smith's paper on quasithin groups was 1221 pages long.

89

Revision of the proof


Many mathematicians have simplified parts of the original FeitThompson proof. However all of these improvements are in some sense local; the global structure of the argument is still the same, but some of the details of the arguments have been simplified. The simplified proof has been published in two books: (Bender & Glauberman 1995), which covers everything except the character theory, and (Peterfalvi 2000, part I) which covers the character theory. This revised proof is still very hard, and is longer than the original proof, but is written in a more leisurely style.

An outline of the proof


Instead of describing the FeitThompson theorem directly, it is easier to describe Suzuki's CA theorem and then comment on some of the extensions needed for the CN-theorem and the odd order theorem. The proof can be broken up into three steps. We let G be a non-abelian (minimal) simple group of odd order satisfying the CA condition. For a more detailed exposition of the odd order paper see Thompson (1963) or (Gorenstein 1980) or Glauberman (1999). Step 1. Local analysis of the structure of the group G. This is easy in the CA case because the relation "a commutes with b" is an equivalence relation on the non-identity elements. So the elements break up into equivalence classes, such that each equivalence class is the set of non-identity elements of a maximal abelian subgroup. The normalizers of these maximal abelian subgroups turn out to be exactly the maximal proper subgroups of G. These normalizers are Frobenius groups whose character theory is reasonably transparent, and well-suited to manipulations involving character induction. Also, the set of prime divisors of |G| is partitioned according to the primes which divide the orders of the distinct conjugacy classes of maximal abelian subgroups of |G|. This pattern of partitioning the prime divisors of |G| according to conjugacy classes of nilpotent Hall subgroups (a Hall subgroup is one whose order and index are relatively prime) whose normalizers give all the maximal subgroups of G (up to conjugacy) is repeated in both the proof of the Feit-Hall-Thompson CN-theorem and in the proof of the Feit-Thompson odd-order theorem. The proof of the CN-case is already considerably more difficult than the CA-case, while this part of the proof of the odd-order theorem takes over 100 journal pages. (Bender later simplified this part of the proof using Bender's method.) Whereas in the CN-case, the resulting maximal subgroups are still Frobenius groups, the maximal subgroups which occur in the proof of the odd-order theorem need no longer have this structure, and the analysis of their structure and interplay produces 5 very complicated possible configurations. Peterfalvi (2000) used the Dade isometry to simplify the character theory. Step 2. Character theory of G. If X is an irreducible character of the normalizer H of the maximal abelian subgroup A of the CA group G, not containing A in its kernel, we can induce X to a character Y of G, which is not necessarily irreducible. Because of the known structure of G, it is easy to find the character values of Y on all but the identity element of G. This implies that if X1 and X2 are two such irreducible characters of H and Y1 and Y2 are the corresponding induced characters, then Y1 Y2 is completely determined, and calculating its norm shows that it is the difference of two irreducible characters of G (these are sometimes known as exceptional characters of G with respect to H). A counting argument shows that each non-trivial irreducible character of G arises exactly once as an exceptional character associated to the normalizer of some maximal abelian subgroup of G. A similar argument (but replacing abelian Hall subgroups by nilpotent Hall subgroups) works in the proof of the CN-theorem. However, in the proof of the odd-order theorem, the arguments for constructing characters of G from characters of subgroups are far more delicate, and involve more subtle maps between character rings than character induction, since the maximal subgroups have a more complicated structure and are embedded in a less transparent way. Step 3. By step 2, we have a complete and precise description of the character table of the CA group G. From this, and using the fact that G has odd order, sufficient information is available to obtain estimates for |G| and arrive at a

FeitThompson theorem contradiction to the assumption that G is simple. This part of the argument works similarly in the CN-group case. In the proof of the FeitThompson theorem, however, this step is (as usual) vastly more complicated. The character theory only eliminates four of the possible five configurations left after step 1. To eliminate the final case, Thompson used some fearsomely complicated manipulations with generators and relations (which were later simplified by Peterfalvi (1984), whose argument is reproduced in (Bender & Glauberman 1994). The Feit-Thompson conjecture would simplify this step if it were proven.

90

References
Bender, Helmut; Glauberman, George (1994), Local analysis for the odd order theorem, London Mathematical Society Lecture Note Series, 188, Cambridge University Press, ISBN978-0-521-45716-3, MR1311244 Brauer, R. (1957), "On the structure of groups of finite order" [1], Proceedings of the International Congress of Mathematicians, Amsterdam, 1954, Vol. 1, Erven P. Noordhoff N.V., Groningen, pp.209217, MR0095203 Burnside, William (1911), Theory of groups of finite order [2], New York: Dover Publications, ISBN978-0-486-49575-0, MR0069818 Feit, Walter; Thompson, John G.; Hall, Marshall, Jr. (1960), "Finite groups in which the centralizer of any non-identity element is nilpotent", Math. Z. 74: 117, doi:10.1007/BF01180468, MR0114856 Feit, Walter; Thompson, John G. (1962), "A solvability criterion for finite groups and some consequences", Proc. Nat. Acad. Sci. 48 (6): 968970, doi:10.1073/pnas.48.6.968, JSTOR71265, MR0143802 Feit, Walter; Thompson, John G. (1963), "Solvability of groups of odd order" [3], Pacific Journal of Mathematics 13: 7751029, ISSN0030-8730, MR0166261 Glauberman, George (1999), "A new look at the Feit-Thompson odd order theorem" [4], Matemtica Contempornea 16: 7392, ISSN0103-9059, MR1756828 Gorenstein, D. (1980), Finite groups [1] (2nd ed.), New York: Chelsea Publishing Co., ISBN978-0-8284-0301-6, MR569209 Peterfalvi, Thomas (1984), "Simplification du chapitre VI de l'article de Feit et Thompson sur les groupes d'ordre impair", Comptes Rendus des Sances de l'Acadmie des Sciences. Srie I. Mathmatique 299 (12): 531534, ISSN0249-6291, MR770439 Peterfalvi, Thomas (2000), Character theory for the odd order theorem [5], London Mathematical Society Lecture Note Series, 272, Cambridge University Press, ISBN978-0-521-64660-4, MR1747393 Suzuki, Michio (1957), "The nonexistence of a certain type of simple groups of odd order", Proceedings of the American Mathematical Society (Proceedings of the American Mathematical Society, Vol. 8, No. 4) 8 (4): 686695, doi:10.2307/2033280, JSTOR2033280, MR0086818 Thompson, John G. (1963), "Two results about finite groups" [6], Proc. Internat. Congr. Mathematicians (Stockholm, 1962), Djursholm: Inst. Mittag-Leffler, pp.296300, MR0175972

References
[1] [2] [3] [4] [5] [6] http:/ / mathunion. org/ ICM/ ICM1954. 1/ http:/ / books. google. com/ books?isbn=1440035458 http:/ / projecteuclid. org/ Dienst/ UI/ 1. 0/ Journal?authority=euclid. pjm& issue=1103053941 http:/ / www. mat. unb. br/ ~matcont/ 16_5. ps http:/ / books. google. com/ books?isbn=052164660X http:/ / mathunion. org/ ICM/ ICM1962. 1/

Fitting's theorem

91

Fitting's theorem
Fitting's theorem is a mathematical theorem proved by Hans Fitting. It can be stated as follows: If M and N are nilpotent normal subgroups of a group G, then their product MN is also a nilpotent normal subgroup of G; if, moreover, M is nilpotent of class m and N is nilpotent of class n, then MN is nilpotent of class at most m + n. By induction it follows also that the subgroup generated by a finite collection of nilpotent normal subgroups is nilpotent. This can be used to show that the Fitting subgroup of certain types of groups (including all finite groups) is nilpotent. However, a subgroup generated by an infinite collection of nilpotent normal subgroups need not be nilpotent.

Order-theoretic statement
In terms of order theory, (part of) Fitting's theorem can be stated as: The set of nilpotent normal subgroups form a lattice of subgroups. Thus the nilpotent normal subgroups of a finite group also form a bounded lattice, and have a top element, the Fitting subgroup. However, nilpotent normal subgroups do not in general form a complete lattice, as a subgroup generated by an infinite collection of nilpotent normal subgroups need not be nilpotent, though it will be normal. The join of all nilpotent normal subgroups is still defined as the Fitting subgroup, but it need not be nilpotent.

Focal subgroup theorem


In abstract algebra, the focal subgroup theorem describes the fusion of elements in a Sylow subgroup of a finite group. The focal subgroup theorem was introduced in (Higman 1958) and is the "first major application of the transfer" according to (Gorenstein, Lyons & Solomon 1996, p.90). The focal subgroup theorem relates the ideas of transfer and fusion such as described in (Grn 1935). Various applications of these ideas include local criteria for p-nilpotence and various non-simplicity criterion focussing on showing that a finite group has a normal subgroup of index p.

Background
The focal subgroup theorem relates several lines of investigation in finite group theory: normal subgroups of index a power of p, the transfer homomorphism, and fusion of elements.

Subgroups
The following three normal subgroups of index a power of p are naturally defined, and arise as the smallest normal subgroups such that the quotient is (a certain kind of) p-group. Formally, they are kernels of the reflection onto the reflective subcategory of p-groups (respectively, elementary abelian p-groups, abelian p-groups). Ep(G) is the intersection of all index p normal subgroups; G/Ep(G) is an elementary abelian group, and is the largest elementary abelian p-group onto which G surjects. Ap(G) (notation from (Isaacs 2008, 5D, p. 164)) is the intersection of all normal subgroups K such that G/K is an abelian p-group (i.e., K is an index normal subgroup that contains the derived group ): G/Ap(G) is the largest abelian p-group (not necessarily elementary) onto which G surjects.

Focal subgroup theorem Op(G) is the intersection of all normal subgroups K of G such that G/K is a (possibly non-abelian) p-group (i.e., K is an index normal subgroup): G/Op(G) is the largest p-group (not necessarily abelian) onto which G surjects. Op(G) is also known as the p-residual subgroup. Firstly, as these are weaker conditions on the These are further related as: groups K, one obtains the containments

92

Ap(G) = Op(G)[G,G]. Op(G) has the following alternative characterization as the subgroup generated by all Sylow q-subgroups of G as qp ranges over the prime divisors of the order of G distinct from p. Op(G) is used to define the lower p-series of G, similarly to the upper p-series described in p-core.

Transfer homomorphism
The transfer homomorphism is a homomorphism that can be defined from any group G to the abelian group H/[H,H] defined by a subgroup H G of finite index, that is [G:H] < . The transfer map from a finite group G into its Sylow p-subgroup has a kernel that is easy to describe: The kernel of the transfer homomorphism from a finite group G into its Sylow p-subgroup P has Ap(G) as its kernel, (Isaacs 2008, Theorem 5.20, p. 165). In other words, the "obvious" homomorphism onto an abelian p-group is in fact the most general such homomorphism.

Fusion
The fusion pattern of a subgroup H in G is the equivalence relation on the elements of H where two elements h, k of H are fused if they are G-conjugate, that is, if there is some g in G such that h = kg. The normal structure of G has an effect on the fusion pattern of its Sylow p-subgroups, and conversely the fusion pattern of its Sylow p-subgroups has an effect on the normal structure of G, (Gorenstein, Lyons & Solomon 1996, p.89).

Focal subgroup
If one defines, as in (Gorenstein 1980, p.246), the focal subgroup of P in G as the intersection P[G,G] of the Sylow p-subgroup P of the finite group G with the derived subgroup [G,G] of G, then the focal subgroup is clearly important as it is a Sylow p-subgroup of the derived subgroup. However, more importantly, one gets the following result: There exists a normal subgroup K of G with G/K an abelian p-group isomorphic to P/P[G,G] (here K denotes Ap(G)), and if K is a normal subgroup of G with G/K an abelian p-group, then P[G,G] K, and G/K is a homomorphic image of P/P[G,G], (Gorenstein 1980, Theorem 7.3.1, p. 90). One can define, as in (Isaacs 2008, p.165) the focal subgroup of H with respect to G as: FocG(H) = x1 y | x,y in H and x is G-conjugate to y . This focal subgroup measures the extent to which elements of H fuse in G, while the previous definition measured certain abelian p-group homomorphic images of the group G. The content of the focal subgroup theorem is that these two definitions of focal subgroup are compatible.

Focal subgroup theorem

93

Statement of the theorem


The focal subgroup of a finite group X with Sylow p-subgroup P is given by: P[G,G] = PAp(G) = Pker(v) = FocG(P) = x1 y | x,y in P and x is G-conjugate to y where v is the transfer homomorphism from G to P/[P,P], (Isaacs 2008, Theorem 5.21, p. 165).

History and generalizations


This connection between transfer and fusion is credited to (Higman 1958),[1] where, in different language, the focal subgroup theorem was proved along with various generalizations. The requirement that G/K be abelian was dropped, so that Higman also studied Op(G) and the nilpotent residual (G), as so called hyperfocal subgroups. Higman also did not restrict to a single prime p, but rather allowed -groups for sets of primes and used Philip Hall's theorem of Hall subgroups in order to prove similar results about the transfer into Hall -subgroups; taking = {p} a Hall -subgroup is a Sylow p-subgroup, and the results of Higman are as presented above. Interest in the hyperfocal subgroups was renewed by work of (Puig 2000) in understanding the modular representation theory of certain well behaved blocks. The hyperfocal subgroup of P in G can defined as P(G) that is, as a Sylow p-subgroup of the nilpotent residual of G. If P is a Sylow p-subgroup of the finite group G, then one gets the standard focal subgroup theorem: P(G) = POp(G) = x1 y : x,y in P and y = xg for some g in G of order coprime to p and the local characterization: POp(G) = x1 y : x,y in Q P and y = xg for some g in NG(Q) of order coprime to p . This compares to the local characterization of the focal subgroup as: PAp(G) = x1 y : x,y in Q P and y = xg for some g in NG(Q) . Puig is interested in the generalization of this situation to fusion systems, a categorical model of the fusion pattern of a Sylow p-subgroup with respect to a finite group that also models the fusion pattern of a defect group of a p-block in modular representation theory. In fact fusion systems have found a number of surprising applications and inspirations in the area of algebraic topology known as equivariant homotopy theory. Some of the major algebraic theorems in this area only have topological proofs at the moment.

Other characterizations
Various mathematicians have presented methods to calculate the focal subgroup from smaller groups. For instance, the influential work (Alperin 1967) develops the idea of a local control of fusion, and as an example application shows that: PAp(G) is generated by the commutator subgroups [Q, NG(Q)] where Q varies over a family C of subgroups ofP The choice of the family C can be made in many ways (C is what is called a "weak conjugation family" in (Alperin 1967)), and several examples are given: one can take C to be all non-identity subgroups of P, or the smaller choice of just the intersections Q=PPg for g in G in which NP(Q) and NPg(Q) are both Sylow p-subgroups of NG(Q). The latter choice is made in (Gorenstein 1980, Theorem 7.4.1, p. 251). The work of (Grn 1935) studied aspects of the transfer and fusion as well, resulting in Grn's first theorem: PAp(G) is generated by P[N,N] and P[Q,Q] where N=NG(P) and Q ranges over the set of Sylow p-subgroups Q = Pg of G (Gorenstein 1980, Theorem 7.4.2, p. 252).

Focal subgroup theorem

94

Applications
The textbook presentations in (Rose 1978, pp.254264), (Isaacs 2008, Chapter 5), (Hall 1959, Chapter 14), (Suzuki 1986, 5.2, pp. 138165), all contain various applications of the focal subgroup theorem relating fusion, transfer, and a certain kind of splitting called p-nilpotence. During the course of the AlperinBrauerGorenstein theorem classifying finite simple groups with quasi-dihedral Sylow 2-subgroups, it becomes necessary to distinguish four types of groups with quasi-dihedral Sylow 2-subgroups: the 2-nilpotent groups, the Q-type groups whose focal subgroup is a generalized quaternion group of index 2, the D-type groups whose focal subgroup a dihedral group of index 2, and the QD-type groups whose focal subgroup is the entire quasi-dihedral group. In terms of fusion, the 2-nilpotent groups have 2 classes of involutions, and 2 classes of cyclic subgroups of order 4; the Q-type have 2 classes of involutions and one class of cyclic subgroup of order 4; the QD-type have one class each of involutions and cyclic subgroups of order 4. In other words, finite groups with quasi-dihedral Sylow 2-subgroups can be classified according to their focal subgroup, or equivalently, according to their fusion patterns. The explicit lists of groups with each fusion pattern are contained in (Alperin, Brauer & Gorenstein 1970).

Notes
[1] The focal subgroup theorem and/or the focal subgroup is due to (Higman 1958) according to (Gorenstein, Lyons & Solomon 1996, p.90), (Rose 1978, p.255), (Suzuki 1986, p.141); however, the focal subgroup theorem as stated there and here is quite a bit older and already appears in textbook form in (Hall 1959, p.215). There and in (Puig 2000) the ideas are credited to (Grn 1935); compare to (Grn 1935, Satz 5) in the special case of p-normal groups, and the general result in Satz 9 which is in some sense a refinement of the focal subgroup theorem.

References
Alperin, J. L. (1967), "Sylow intersections and fusion", Journal of Algebra 6: 222241, doi:10.1016/0021-8693(67)90005-1, ISSN0021-8693, MR0215913 Alperin, J. L.; Brauer, R.; Gorenstein, D. (1970), "Finite groups with quasi-dihedral and wreathed Sylow 2-subgroups.", Transactions of the American Mathematical Society (American Mathematical Society) 151: 1261, doi:10.2307/1995627, ISSN0002-9947, MR0284499 Gorenstein, D. (1980), Finite Groups, New York: Chelsea, ISBN978-0-8284-0301-6, MR81b:20002 Gorenstein, D.; Lyons, Richard; Solomon, Ronald (1996), The classification of the finite simple groups. Number 2. Part I. Chapter G (https://www.ams.org/online_bks/surv402), Mathematical Surveys and Monographs, 40, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-0390-5, MR1358135 Grn, Otto (1935), "Beitrge zur Gruppentheorie. I." (http://resolver.sub.uni-goettingen.de/ purl?GDZPPN002173409) (in German), Journal fr Reine und Angewandte Mathematik 174: 114, ISSN0075-4102, Zbl0012.34102 Hall, Marshall, Jr. (1959), The theory of groups, New York: Macmillan, MR0103215 Higman, Donald G. (1953), "Focal series in finite groups" (http://www.cms.math.ca/cjm/v5/p477), Canadian Journal of Mathematics 5: 477497, ISSN0008-414X, MR0058597 Puig, Lluis (2000), "The hyperfocal subalgebra of a block", Inventiones Mathematicae 141 (2): 365397, doi:10.1007/s002220000072, ISSN0020-9910, MR1775217 Rose, John S. (1994) [1978], A Course in Group Theory, New York: Dover Publications, ISBN978-0-486-68194-8, MR0498810 Suzuki, Michio (1986), Group theory. II, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], 248, Berlin, New York: Springer-Verlag, ISBN978-0-387-10916-9, MR815926

Frobenius determinant theorem

95

Frobenius determinant theorem


In mathematics, the Frobenius determinant theorem is a discovery made in 1896 by the mathematician Richard Dedekind, who wrote a letter to F. G. Frobenius about it (reproduced in (Dedekind 1968), with an English translation in (Curtis 2003, p.51)). If one takes the multiplication table of a group G and replaces each entry g with the variable xg, and subsequently takes the determinant, then the determinant factors as a product of n irreducible polynomials, where n is the number of conjugacy classes. Moreover, each polynomial is raised to a power equal to its degree. Frobenius proved this surprising fact, and this theorem became known as the Frobenius determinant theorem.

Formal statement
Let a finite group the matrix have elements . Then , and let be associated with each element of . Define with entries

where r is the number of conjugacy classes ofG.

References
Curtis, Charles W. (2003), Pioneers of Representation Theory: Frobenius, Burnside, Schur, and Brauer [1], History of Mathematics, Providence, R.I.: American Mathematical Society, doi:10.1090/S0273-0979-00-00867-3, ISBN978-0-8218-2677-5, MR1715145 Review [2] Dedekind, Richard (1968) [1931], Fricke, Robert; Noether, Emmy; Ore, ystein, eds., Gesammelte mathematische Werke. Bnde I--III, New York: Chelsea Publishing Co., JFM56.0024.05, MR0237282 Etingof, Pavel. Lectures on Representation Theory [3]. Frobenius, Ferdinand Georg (1968), Serre, J.-P., ed., Gesammelte Abhandlungen. Bnde I, II, III, Berlin, New York: Springer-Verlag, ISBN978-3-540-04120-7, MR0235974

References
[1] http:/ / books. google. com/ books?isbn=0821826778 [2] http:/ / www. ams. org/ journals/ bull/ 2000-37-03/ S0273-0979-00-00867-3/ [3] http:/ / www-math. mit. edu/ ~etingof/ cltrunc. pdf

Frobenius theorem (real division algebras)

96

Frobenius theorem (real division algebras)


In mathematics, more specifically in abstract algebra, the Frobenius theorem, proved by Ferdinand Georg Frobenius in 1877, characterizes the finite-dimensional associative division algebras over the real numbers. According to the theorem, every such algebra is isomorphic to one of the following: R (the real numbers) C (the complex numbers) H (the quaternions). These algebras have dimensions 1, 2, and 4, respectively. Of these three algebras, the real and complex numbers are commutative, but the quaternions are not. This theorem is closely related to Hurwitz's theorem, which states that the only normed division algebras over the real numbers are R, C, H, and the (non-associative) algebra O of octonions.

Proof
The main ingredients for the following proof are the CayleyHamilton theorem and the fundamental theorem of algebra. We can consider D as a finite-dimensional R-vector space. Any element d of D defines an endomorphism of D by left-multiplication and we will identify d with that endomorphism. Therefore we can speak about the trace ofd, the characteristic and minimal polynomials. Also, we identify the real multiples of 1 with R. When we write for an element a of D, we tacitly assume that a is contained inR. The key to the argument is the following Claim: The set V of all elements a of D such that is a vector subspace of D of codimension1. be the characteristic

To see that, we pick an a D. Let m be the dimension of D as an R-vector space. Let polynomial of a. By the fundamental theorem of algebra, we can write for some real ti and (non-real) complex numbers zj. We have and because D is a division algebra, it follows that either , for some i or that

. The polynomials

are irreducible over R. By the CayleyHamilton theorem, p(a)=0

for somej. The first case implies that aR. In the second case, it follows that

is

the minimal polynomial of a. Because p(x) has the same complex roots as the minimal polynomial and because it is real it follows that and m=2k. The coefficient of in is the trace of a (up to sign). Therefore we read from that equation: the

trace of a is zero if and only if , that is . Therefore V is the subset of all a with tra=0. In particular, it is a vector subspace (!). Moreover, V has codimension 1 since it is the kernel of a (nonzero) linear form. Also note that D is the direct sum of R and V (as vector spaces). Therefore, V generates D as an algebra. Define now for , it follows that is real and since Because of the identity if . Thus B is a positive definite

symmetric bilinear form, in other words, an inner product on V. Let W be a subspace of V which generated D as an algebra and which is minimal with respect to that property. Let be an orthonormal basis of W. These elements satisfy the following relations:

If n=0, then D is isomorphic to R.

Frobenius theorem (real division algebras) If n=1, then D is generated by 1 and e1 subject to the relation and (this only works if n>2). Therefore assumed to be a division algebra). But if u=1, then contradicts the minimality ofW. Remark: The fact that D is generated by . Hence it is isomorphic to C. subject to the relations . It is easy to see that implies that u=1 (because D is still and so generatesD. This . These are precisely the relations for H.

97

If n=2, it has been shown above that D is generated by If n>2, the D cannot be a division algebra. Assume that n > 2. Put

subject to above relation can be interpreted as the statement

that D is the Clifford algebra of Rn. The last step shows that the only real Clifford algebras which are division algebras are Cl0, Cl1 and Cl2. Remark: As a consequence, the only commutative division algebras are R and C. Also note that H is not a C-algebra. If it were, then the center of H has to contain C, but the center of H is R. Therefore, the only division algebra over C is C itself.

Pontryagin variant
If D is a connected, locally compact division ring, then either D = R, or D = C, or D = H.

References
Ray E. Artz (2009) Scalar Algebras and Quaternions [1], Theorem 7.1 "Frobenius Classification", page 26. Ferdinand Georg Frobenius (1878) "ber lineare Substitutionen und bilineare Formen [2]", Journal fr die reine und angewandte Mathematik 84:163 (Crelle's Journal). Reprinted in Gesammelte Abhandlungen Band I, pp.343405. Yuri Bahturin (1993) Basic Structures of Modern Algebra, Kluwer Acad. Pub. pp.302 ISBN 0-7923-2459-5 . Leonard Dickson (1914) Linear Algebras, Cambridge University Press. See 11 "Algebra of real quaternions; its unique place among algebras", pages 10 to 12. R.S. Palais (1968) "The Classification of Real Division Algebras" American Mathematical Monthly 75:3668. Lev Semenovich Pontryagin, Topological Groups, page 159, 1966.

References
[1] http:/ / www. math. cmu. edu/ ~wn0g/ noll/ qu1. pdf [2] http:/ / commons. wikimedia. org/ wiki/ File:%C3%9Cber_lineare_Substitutionen_und_bilineare_Formen. djvu

Fundamental lemma (Langlands program)

98

Fundamental lemma (Langlands program)


In the theory of automorphic forms, an area of mathematics, the term fundamental lemma refers to any of a group of results conjectured by Robert Langlands in the course of developing the Langlands program. A considerable body of subsequent work was conditional on their validity. The fundamental lemma was proved by Grard Laumon and Ng Bo Chu in the case of unitary groups and then by Ng for general reductive groups, building on a series of important reductions made by Jean-Loup Waldspurger to the case of Lie algebras. Time magazine placed Ng's proof on the list of the "Top 10 scientific discoveries of 2009".[1] [2] [3] In 2010 Ng was awarded the Fields medal for this proof.

Motivation and history


Robert Langlands outlined a strategy for proving local and global Langlands conjectures using the ArthurSelberg trace formula, but in order for this approach to work, the geometric sides of the trace formula for different groups must be related in a particular way. This relationship takes the form of identities between orbital integrals on reductive groups G and H over a nonarchimedean local field F, where the group H, called an endoscopic group of G, is constructed from G and some additional data. The first case considered was G = SL2 (Labesse and Langlands, 1979). Langlands and Diana Shelstad then developed the general framework for the theory of endoscopic transfer and formulated specific conjectures. However, during the next two decades only partial progress was made towards proving the fundamental lemma.[4] [5] Thus, it has been called a "bottleneck limiting progress on a host of arithmetic questions".[6] Langlands himself, writing on the origins of endoscopy, commented:

... it is not the fundamental lemma as such that is critical for the analytic theory of automorphic forms and for the arithmetic of Shimura varieties; it is the stabilized (or stable) trace formula, the reduction of the trace formula itself to the stable trace formula for a group and its endoscopic groups, and the stabilization of the GrothendieckLefschetz formula. None of these are possible without the fundamental lemma [7] and its absence rendered progress almost impossible for more than twenty years.

Approaches
A paper of George Lusztig and David Kazhdan pointed out that orbital integrals could be interpreted as counting points on certain algebraic varieties over finite fields.[3] Further, the integrals in question can be computed in a way that depends only on the residue field of F; and the issue can be reduced to the Lie algebra version of the orbital integrals. Then the problem was restated in terms of the Springer fiber of algebraic groups.[8] The circle of ideas was connected to a purity conjecture; Laumon gave a conditional proof based on such a conjecture, for unitary groups. Laumon and Ng then introduced the use of the Hitchin fibration, which is an abstract geometric analogue of the Hitchin system of mathematical physics, as a further technical tool, leading ultimately to a proof of the "function field" case; according to known previous reduction steps, this implies the general form of the conjecture.

Fundamental lemma (Langlands program)

99

Notes
[1] Not even wrong (http:/ / www. math. columbia. edu/ ~woit/ wordpress/ ?p=18) [2] Top 10 Scientific Discoveries of 2009 (http:/ / www. time. com/ time/ specials/ packages/ article/ 0,28804,1945379_1944416_1944435,00. html), Time [3] The fundamental lemma (http:/ / www. netera. ca/ seminars/ math/ fl-white. pdf), Bill Casselman [4] Kottwitz and Rogawski for U3, Wadspurger for SLn, Hales and Weissauer for Sp4. [5] http:/ / www. newton. ac. uk/ programmes/ ALT/ seminars/ 051316301. pdf [6] http:/ / www. institut. math. jussieu. fr/ projets/ fa/ bpFiles/ Introduction. pdf. p. 1. [7] http:/ / publications. ias. edu/ rpl/ series. php?series=55 [8] http:/ / www. claymath. org/ research_award/ Laumon-Ngo/ laumon. pdf, at p. 12.

References
Ng, Bo Chu (2008). "Le lemme fondamental pour les algebres de Lie". arXiv:0801.0446.. Laumon, G.; Ng, B. C. (2004). "Le lemme fondamental pour les groupes unitaires". arXiv:math/0404454.. Dat, Jean-Franois (Novembre 2004), Lemme fondamental et endoscopie, une approche gomtrique, d'aprs Grard Laumon et Ng Bao Chu (http://www.math.univ-paris13.fr/~dat/publis/lf.pdf), Sminaire Bourbaki, no 940. Labesse, J.-P.; Langlands, R. P. (1979), "L-indistinguishability for SL(2)", Can. Jour. Math. 31.

External links
Gerard Laumon lecture on the fundamental lemma for unitary groups (http://www.fields.utoronto.ca/audio/ 02-03/shimura/laumon/)

Fundamental theorem of algebra


The fundamental theorem of algebra states that every non-constant single-variable polynomial with complex coefficients has at least one complex root. Equivalently, the field of complex numbers is algebraically closed. Sometimes, this theorem is stated as: every non-zero single-variable polynomial with complex coefficients has exactly as many complex roots as its degree, if each root is counted up to its multiplicity. Although this at first appears to be a stronger statement, it is a direct consequence of the other form of the theorem, through the use of successive polynomial division by linear factors. In spite of its name, there is no purely algebraic proof of the theorem, since any proof must use the completeness of the reals (or some other equivalent formulation of completeness), which is not an algebraic concept. Additionally, it is not fundamental for modern algebra; its name was given at a time in which algebra was mainly about solving polynomial equations with real or complex coefficients.

History
Peter Rothe (Petrus Roth), in his book Arithmetica Philosophica (published in 1608), wrote that a polynomial equation of degree n (with real coefficients) may have n solutions. Albert Girard, in his book L'invention nouvelle en l'Algbre (published in 1629), asserted that a polynomial equation of degree n has n solutions, but he did not state that they had to be real numbers. Furthermore, he added that his assertion holds unless the equation is incomplete, by which he meant that no coefficient is equal to0. However, when he explains in detail what he means, it is clear that he actually believes that his assertion is always true; for instance, he shows that the equation x4=4x3, although incomplete, has four solutions (counting multiplicities): 1 (twice), 1+i2, and 1i2.

Fundamental theorem of algebra As will be mentioned again below, it follows from the fundamental theorem of algebra that every non-constant polynomial with real coefficients can be written as a product of polynomials with real coefficients whose degree is either 1 or 2. However, in 1702 Leibniz said that no polynomial of the type x4+a4 (with a real and distinct from0) can be written in such a way. Later, Nikolaus Bernoulli made the same assertion concerning the polynomial x44x3+2x2+4x+4, but he got a letter from Euler in 1742[1] in which he was told that his polynomial happened to be equal to

100

where is the square root of 4+27. Also, Euler mentioned that

A first attempt at proving the theorem was made by d'Alembert in 1746, but his proof was incomplete. Among other problems, it assumed implicitly a theorem (now known as Puiseux's theorem) which would not be proved until more than a century later, and furthermore the proof assumed the fundamental theorem of algebra. Other attempts were made by Euler (1749), de Foncenex (1759), Lagrange (1772), and Laplace (1795). These last four attempts assumed implicitly Girard's assertion; to be more precise, the existence of solutions was assumed and all that remained to be proved was that their form was a+bi for some real numbers a and b. In modern terms, Euler, de Foncenex, Lagrange, and Laplace were assuming the existence of a splitting field of the polynomial p(z). At the end of the 18th century, two new proofs were published which did not assume the existence of roots. One of them, due to James Wood and mainly algebraic, was published in 1798 and it was totally ignored. Wood's proof had an algebraic gap.[2] The other one was published by Gauss in 1799 and it was mainly geometric, but it had a topological gap, filled by Alexander Ostrowski in 1920, as discussed in Smale 1981 [3] (Smale writes, "...I wish to point out what an immense gap Gauss' proof contained. It is a subtle point even today that a real algebraic plane curve cannot enter a disk without leaving. In fact even though Gauss redid this proof 50 years later, the gap remained. It was not until 1920 that Gauss' proof was completed. In the reference Gauss, A. Ostrowski has a paper which does this and gives an excellent discussion of the problem as well..."). A rigorous proof was published by Argand in 1806; it was here that, for the first time, the fundamental theorem of algebra was stated for polynomials with complex coefficients, rather than just real coefficients. Gauss produced two other proofs in 1816 and another version of his original proof in 1849. The first textbook containing a proof of the theorem was Cauchy's Cours d'analyse de l'cole Royale Polytechnique (1821). It contained Argand's proof, although Argand is not credited for it. None of the proofs mentioned so far is constructive. It was Weierstrass who raised for the first time, in the middle of the 19th century, the problem of finding a constructive proof of the fundamental theorem of algebra. He presented his solution, that amounts in modern terms to a combination of the DurandKerner method with the homotopy continuation principle, in 1891. Another proof of this kind was obtained by Hellmuth Kneser in 1940 and simplified by his son Martin Kneser in 1981. Without using countable choice, it is not possible to constructively prove the fundamental theorem of algebra for complex numbers based on the Dedekind real numbers (which are not constructively equivalent to the Cauchy real numbers without countable choice[4] ). However, Fred Richman proved a reformulated version of the theorem that does work.[5]

Fundamental theorem of algebra

101

Proofs
All proofs below involve some analysis, at the very least the concept of continuity of real or complex functions. Some also use differentiable or even analytic functions. This fact has led some to remark that the Fundamental Theorem of Algebra is neither fundamental, nor a theorem of algebra. Some proofs of the theorem only prove that any non-constant polynomial with real coefficients has some complex root. This is enough to establish the theorem in the general case because, given a non-constant polynomial p(z) with complex coefficients, the polynomial

has only real coefficients and, if z is a zero of q(z), then either z or its conjugate is a root of p(z). A large number of non-algebraic proofs of the theorem use the fact (sometimes called growth lemma) that an n-th degree polynomial function p(z) whose dominant coefficient is 1 behaves like zn when |z| is large enough. A more precise statement is: there is some positive real number R such that:

when |z|>R.

Complex-analytic proofs
Find a closed disk D of radius r centered at the origin such that |p(z)|>|p(0)| whenever |z|r. The minimum of |p(z)| on D, which must exist since D is compact, is therefore achieved at some point z0 in the interior of D, but not at any point of its boundary. The minimum modulus principle implies then that p(z0)=0. In other words, z0 is a zero of p(z). A variation of this proof that does not require the use of the minimum modulus principle (most proofs of which in turn require the use of Cauchy's integral theorem or some of its consequences) is based on the observation that for the special case of a polynomial function, the minimum modulus principle can be proved directly using elementary arguments. More precisely, if we assume by contradiction that , then, expanding in powers of Here, the we can write 's are simply the coefficients of the polynomial , and we let be the index of the

first coefficient following the constant term that is non-zero. But now we see that for has behavior asymptotically similar to the simpler polynomial easy to check) the function . Therefore if we define small positive number (so that the bound is bounded by some positive constant and let

sufficiently close to this , in the sense that (as is in some neighborhood of , then for any sufficiently

mentioned above holds), using the triangle inequality we see that

When r is sufficiently close to 0 this upper bound for |p(z)| is strictly smaller than |a|, in contradiction to the definition of z0. (Geometrically, we have found an explicit direction 0 such that if one approaches z0 from that direction one can obtain values p(z) smaller in absolute value than |p(z0)|.) Another analytic proof can be obtained along this line of thought observing that, since |p(z)|>|p(0)| outside D, the minimum of |p(z)| on the whole complex plane is achieved at z0. If |p(z0)|>0, then 1/p is a bounded holomorphic function in the entire complex plane since, for each complex number z, |1/p(z)||1/p(z0)|. Applying Liouville's

Fundamental theorem of algebra theorem, which states that a bounded entire function must be constant, this would imply that 1/p is constant and therefore that p is constant. This gives a contradiction, and hence p(z0)=0. Yet another analytic proof uses the argument principle. Let R be a positive real number large enough so that every root of p(z) has absolute value smaller than R; such a number must exist because every non-constant polynomial function of degree n has at most n zeros. For each r>R, consider the number

102

where c(r) is the circle centered at 0 with radius r oriented counterclockwise; then the argument principle says that this number is the number N of zeros of p(z) in the open ball centered at 0 with radius r, which, since r>R, is the total number of zeros of p(z). On the other hand, the integral of n/z along c(r) divided by 2i is equal to n. But the difference between the two numbers is

The numerator of the rational expression being integrated has degree at most n1 and the degree of the denominator is n+1. Therefore, the number above tends to 0 as r tends to +. But the number is also equal to Nn and so N=n. Still another complex-analytic proof can be given by combining linear algebra with the Cauchy theorem. To establish that every complex polynomial of degree n>0 has a zero, it suffices to show that every complex square matrix of size n>0 has a (complex) eigenvalue.[6] The proof of the latter statement is by contradiction. Let A be a complex square matrix of size n>0 and let In be the unit matrix of the same size. Assume A has no eigenvalues. Consider the resolvent function

which is a meromorphic function on the complex plane with values in the vector space of matrices. The eigenvalues of A are precisely the poles of R(z). Since, by assumption, A has no eigenvalues, the function R(z) is an entire function and Cauchy's theorem implies that

On the other hand, R(z) expanded as a geometric series gives:

This formula is valid outside the closed disc of radius ||A|| (the operator norm of A). Let r>||A||. Then

(in which only the summand k=0 has a nonzero integral). This is a contradiction, and so A has an eigenvalue.

Topological proofs
Let z0C be such that the minimum of |p(z)| on the whole complex plane is achieved at z0; it was seen at the proof which uses Liouville's theorem that such a number must exist. We can write p(z) as a polynomial in zz0: there is some natural number k and there are some complex numbers ck, ck+1, ..., cn such that ck0 and that It follows that if a is a kth root of p(z0)/ck and if t is positive and sufficiently small, then |p(z0+ta)|<|p(z0)|, which is impossible, since |p(z0)| is the minimum of |p| on D. For another topological proof by contradiction, suppose that p(z) has no zeros. Choose a large positive number R such that, for |z|=R, the leading term zn of p(z) dominates all other terms combined; in other words, such that

Fundamental theorem of algebra |z|n>|an1zn1++a0|. As z traverses the circle given by the equation |z|=R once counter-clockwise, p(z), like zn, winds n times counter-clockwise around 0. At the other extreme, with |z|=0, the curve p(z) is simply the single (nonzero) point p(0), whose winding number is clearly 0. If the loop followed by z is continuously deformed between these extremes, the path of p(z) also deforms continuously. We can explicitly write such a deformation as where t is greater than or equal to 0 and less than or equal to 1. If one views the variable t as time, then at time zero the curve is p(z) and at time one the curve is p(0). Clearly at every point t, p(z) cannot be zero by the original assumption, therefore during the deformation, the curve never crosses zero. Therefore the winding number of the curve around zero should never change. However, given that the winding number started as n and ended as 0, this is absurd. Therefore, p(z) has at least one zero.

103

Algebraic proofs
These proofs use two facts about real numbers that require only a small amount of analysis (more precisely, the intermediate value theorem): every polynomial with odd degree and real coefficients has some real root; every non-negative real number has a square root. The second fact, together with the quadratic formula, implies the theorem for real quadratic polynomials. In other words, algebraic proofs of the fundamental theorem actually show that if R is any real-closed field, then its extension is algebraically closed. As mentioned above, it suffices to check the statement every non-constant polynomial p(z) with real coefficients has a complex root. This statement can be proved by induction on the greatest non-negative integer k such that 2k divides the degree n of p(z). Let a be the coefficient of zn in p(z) and let F be a splitting field of p(z) over C; in other words, the field F contains C and there are elements z1, z2, ..., zn in F such that If k=0, then n is odd, and therefore p(z) has a real root. Now, suppose that n=2km (with m odd and k>0) and that the theorem is already proved when the degree of the polynomial has the form 2k1m with m odd. For a real number t, define:

Then the coefficients of qt(z) are symmetric polynomials in the zi's with real coefficients. Therefore, they can be expressed as polynomials with real coefficients in the elementary symmetric polynomials, that is, in a1, a2, ..., (1)nan. So qt(z) has in fact real coefficients. Furthermore, the degree of qt(z) is n(n1)/2=2k1m(n1), and m(n1) is an odd number. So, using the induction hypothesis, qt has at least one complex root; in other words, zi+zj+tzizj is complex for two distinct elements i and j from {1,...,n}. Since there are more real numbers than pairs (i,j), one can find distinct real numbers t and s such that zi+zj+tzizj and zi+zj+szizj are complex (for the same i and j). So, both zi+zj and zizj are complex numbers. It is easy to check that every complex number has a complex square root, thus every complex polynomial of degree 2 has a complex root by the quadratic formula. It follows that zi and zj are complex numbers, since they are roots of the quadratic polynomial z2(zi+zj)z+zizj. J. Shipman showed in 2007 that the assumption that odd degree polynomials have roots is stronger than necessary; any field in which polynomials of prime degree have roots is algebraically closed (so "odd" can be replaced by "odd prime" and furthermore this holds for fields of all characteristics). For axiomatization of algebraically closed fields, this is the best possible, as there are counterexamples if a single prime is excluded. However, these counterexamples rely on 1 having a square root. If we take a field where 1 has no square root, and every polynomial of degree nI has a root, where I is any fixed infinite set of odd numbers, then every polynomial f(x) of odd degree has a root (since (x2 + 1)kf(x) has a root, where k is chosen so that deg(f) + 2k I).

Fundamental theorem of algebra Another algebraic proof of the fundamental theorem can be given using Galois theory. It suffices to show that C has no proper finite field extension.[7] Let K/C be a finite extension. Since the normal closure of K over R still has a finite degree over C (or R), we may assume without loss of generality that K is a normal extension of R (hence it is a Galois extension, as every algebraic extension of a field of characteristic 0 is separable). Let G be the Galois group of this extension, and let H be a Sylow 2-group of G, so that the order of H is a power of 2, and the index of H in G is odd. By the fundamental theorem of Galois theory, there exists a subextension L of K/R such that Gal(K/L)=H. As [L:R]=[G:H] is odd, and there are no nonlinear irreducible real polynomials of odd degree, we must have L= R, thus [K:R] and [K:C] are powers of 2. Assuming for contradiction [K:C]>1, the 2-group Gal(K/C) contains a subgroup of index 2, thus there exists a subextension M of C of degree2. However, C has no extension of degree2, because every quadratic complex polynomial has a complex root, as mentioned above.

104

Corollaries
Since the fundamental theorem of algebra can be seen as the statement that the field of complex numbers is algebraically closed, it follows that any theorem concerning algebraically closed fields applies to the field of complex numbers. Here are a few more consequences of the theorem, which are either about the field of real numbers or about the relationship between the field of real numbers and the field of complex numbers: The field of complex numbers is the algebraic closure of the field of real numbers. Every polynomial in one variable x with real coefficients is the product of a constant, polynomials of the form x+a with a real, and polynomials of the form x2+ax+b with a and b real and a24b<0 (which is the same thing as saying that the polynomial x2+ax+b has no real roots). Every rational function in one variable x, with real coefficients, can be written as the sum of a polynomial function with rational functions of the form a/(xb)n (where n is a natural number, and a and b are real numbers), and rational functions of the form (ax+b)/(x2+cx+d)n (where n is a natural number, and a, b, c, and d are real numbers such that c24d<0). A corollary of this is that every rational function in one variable and real coefficients has an elementary primitive. Every algebraic extension of the real field is isomorphic either to the real field or to the complex field.

Bounds on the zeroes of a polynomial


While the fundamental theorem of algebra states a general existence result, it is of some interest, both from the theoretical and from the practical point of view, to have information on the location of the zeroes of a given polynomial. The simpler result in this direction is a bound on the modulus: all zeroes of a monic polynomial satisfy an inequality where

Notice that, as stated, this is not yet an existence result but rather an example of what is called an a priori bound: it says that if there are solutions then they lie inside the closed disk of center the origin and radius . However, once coupled with the fundamental theorem of algebra it says that the disk contains in fact at least one solution. More generally, a bound can be given directly in terms of any p-norm of the n-vector of coefficients , that is , where is precisely the q-norm of the 2-vector , q being the conjugate exponent of p, 1/p + 1/q = 1, for any . Thus, the modulus of any solution is also bounded by

for

, and in particular

Fundamental theorem of algebra

105

(where we define to mean 1, which is reasonable since 1 is indeed the n-th coefficient of our polynomial). The case of a generic polynomial of degree n, , is of course reduced to the case of a monic, dividing all coefficients by . Also, in case that 0 is not a root, i.e. roots follow immediately as bounds from above on , that is, the roots of the distance from the roots to any point zeroes of the polynomial , whose coefficients are the Taylor expansion of We report the here the proof of the above bounds, which is short and elementary. Let ; in order to prove the inequality equation as if , this is , thus , and using the Hlder's inequality we find . In the case , bounds from below on the . Finally, as at be a root of the polynomial . Writing the . Now, , taking into

can be estimated from below and above, seeing

we can assume, of course,

account the summation formula for a geometric progression, we have

thus

and simplifying,

. Therefore

holds, for all

Notes
[1] [2] [3] [4] See section Le rle d'Euler in C. Gilain's article Sur l'histoire du thorme fondamental de l'algbre: thorie des quations et calcul intgral. Concerning Wood's proof, see the article A forgotten paper on the fundamental theorem of algebra, by Frank Smithies. http:/ / projecteuclid. org/ DPubS?service=UI& version=1. 0& verb=Display& handle=euclid. bams/ 1183547848 For the minimum necessary to prove their equivalence, see Bridges, Schuster, and Richman; 1998; A weak countable choice principle; available from (http:/ / www. math. fau. edu/ richman/ HTML/ DOCS. HTM). [5] See Fred Richman; 1998; The fundamental theorem of algebra: a constructive development without choice; available from (http:/ / www. math. fau. edu/ richman/ HTML/ DOCS. HTM). [6] A proof of the fact that this suffices can be seen here. [7] A proof of the fact that this suffices can be seen here.

References
Historic sources
Cauchy, Augustin Louis (1821), Cours d'Analyse de l'cole Royale Polytechnique, 1re partie: Analyse Algbrique (http://gallica.bnf.fr/ark:/12148/bpt6k29058v), Paris: ditions Jacques Gabay (published 1992), ISBN2-87647-053-5 (tr. Course on Analysis of the Royal Polytechnic Academy, part 1: Algebraic Analysis) Euler, Leonhard (1751), "Recherches sur les racines imaginaires des quations" (http://bibliothek.bbaw.de/ bbaw/bibliothek-digital/digitalequellen/schriften/anzeige/index_html?band=02-hist/1749&seite:int=228), Histoire de l'Acadmie Royale des Sciences et des Belles-Lettres de Berlin (Berlin) 5: 222288. English translation: Euler, Leonhard (1751), "Investigations on the Imaginary Roots of Equations" (http://www. mathsym.org/euler/e170.pdf) (PDF), Histoire de l'Acadmie Royale des Sciences et des Belles-Lettres de Berlin (Berlin) 5: 222288 Gauss, Carl Friedrich (1799), Demonstratio nova theorematis omnem functionem algebraicam rationalem integram unius variabilis in factores reales primi vel secundi gradus resolvi posse, Helmstedt: C.G.Fleckeisen (tr. New proof of the theorem that every integral rational algebraic function of one variable can be resolved into real factors of the first or second degree). C. F. Gauss, Another new proof of the theorem that every integral rational algebraic function of one variable can be resolved into real factors of the first or second degree (http://www.paultaylor.eu/misc/gauss-web.php), 1815

Fundamental theorem of algebra Kneser, Hellmuth (1940), "Der Fundamentalsatz der Algebra und der Intuitionismus" (http://www-gdz.sub. uni-goettingen.de/cgi-bin/digbib.cgi?PPN266833020_0046), Mathematische Zeitschrift 46: 287302, doi:10.1007/BF01181442, ISSN0025-5874 (The Fundamental Theorem of Algebra and Intuitionism). Kneser, Martin (1981), "Ergnzung zu einer Arbeit von Hellmuth Kneser ber den Fundamentalsatz der Algebra" (http://www-gdz.sub.uni-goettingen.de/cgi-bin/digbib.cgi?PPN266833020_0177), Mathematische Zeitschrift 177 (2): 285287, doi:10.1007/BF01214206, ISSN0025-5874 (tr. An extension of a work of Hellmuth Kneser on the Fundamental Theorem of Algebra). Ostrowski, Alexander (1920), "ber den ersten und vierten Gauschen Beweis des Fundamental-Satzes der Algebra" (http://gdz.sub.uni-goettingen.de/dms/load/img/?PPN=PPN236019856&DMDID=dmdlog53), Carl Friedrich Gauss Werke Band X Abt. 2 (tr. On the first and fourth Gaussian proofs of the Fundamental Theorem of Algebra). Weierstra, Karl (1891). "Neuer Beweis des Satzes, dass jede ganze rationale Function einer Vernderlichen dargestellt werden kann als ein Product aus linearen Functionen derselben Vernderlichen". Sitzungsberichte der kniglich preussischen Akademie der Wissenschaften zu Berlin. pp.10851101. (tr. New proof of the theorem that every integral rational function of one variable can be represented as a product of linear functions of the same variable).

106

Recent literature
Fine, Benjamin; Rosenber, Gerhard (1997), The Fundamental Theorem of Algebra, Undergraduate Texts in Mathematics, Berlin: Springer-Verlag, ISBN978-0-387-94657-3 Gersten, S.M.; Stallings, John R. (1988), "On Gauss's First Proof of the Fundamental Theorem of Algebra", Proceedings of the AMS 103 (1): 331332, doi:10.2307/2047574, ISSN0002-9939, JSTOR2047574 Gilain, Christian (1991), "Sur l'histoire du thorme fondamental de l'algbre: thorie des quations et calcul intgral", Archive for History of Exact Sciences 42 (2): 91136, doi:10.1007/BF00496870, ISSN0003-9519 (tr. On the history of the fundamental theorem of algebra: theory of equations and integral calculus.) Netto, Eugen; Le Vavasseur, Raymond (1916), "Les fonctions rationnelles 8088: Le thorme fondamental", in Meyer, Franois; Molk, Jules, Encyclopdie des Sciences Mathmatiques Pures et Appliques, tomeI, vol.2, ditions Jacques Gabay, 1992, ISBN2-87647-101-9 (tr. The rational functions 8088: the fundamental theorem). Remmert, Reinhold (1991), "The Fundamental Theorem of Algebra", in Ebbinghaus, Heinz-Dieter; Hermes, Hans; Hirzebruch, Friedrich, Numbers, Graduate Texts in Mathematics 123, Berlin: Springer-Verlag, ISBN978-0-387-97497-2 Shipman, Joseph (2007), "Improving the Fundamental Theorem of Algebra", Mathematical Intelligencer 29 (4): 914, doi:10.1007/BF02986170, ISSN0343-6993 Smale, Steve (1981), "The Fundamental Theorem of Algebra and Complexity Theory", Bulletin (new series) of the American Mathematical Society 4 (1) (http://projecteuclid.org/DPubS?service=UI&version=1.0& verb=Display&handle=euclid.bams/1183547848) Smith, David Eugene (1959), A Source Book in Mathematics, Dover, ISBN0-486-64690-4 Smithies, Frank (2000), "A forgotten paper on the fundamental theorem of algebra", Notes & Records of the Royal Society 54 (3): 333341, doi:10.1098/rsnr.2000.0116, ISSN0035-9149 van der Waerden, Bartel Leendert (2003), Algebra, I (7th ed.), Springer-Verlag, ISBN0-387-40624-7

Fundamental theorem of algebra

107

External links
Fundamental Theorem of Algebra (http://www.cut-the-knot.org/do_you_know/fundamental2.shtml) a collection of proofs D. J. Velleman: The Fundamental Theorem of Algebra: A Visual Approach, PDF (unpublished paper) (http:// www.cs.amherst.edu/~djv/), visualisation of d'Alembert's, Gauss's and the winding number proofs Fundamental Theorem of Algebra Module by John H. Mathews (http://math.fullerton.edu/mathews/c2003/ FunTheoremAlgebraMod.html) Bibliography for the Fundamental Theorem of Algebra (http://math.fullerton.edu/mathews/c2003/ FunTheoremAlgebraBib/Links/FunTheoremAlgebraBib_lnk_2.html) From the Fundamental Theorem of Algebra to Astrophysics: A "Harmonious" Path (http://www.ams.org/ notices/200806/tx080600666p.pdf)

Fundamental theorem of cyclic groups


In abstract algebra, the fundamental theorem of cyclic groups states that every subgroup of a cyclic group is cyclic. Moreover, the order of any subgroup of a cyclic group of order is a divisor of , and for each positive divisor of the group has exactly one subgroup of order .

Proof
Let be a cyclic group for some is cyclic. If . is cyclic every element in . . It follows immediately from the closure property that . is of the form , where is an integer. Let be the then If then since and with identity and order , and let be a subgroup of . We will now show that

least positive integer such that We will now show that To show that we let By the division algorithm, with . Since

we have that

for some positive integer

, and so , which yields .

Now since and But is the least positive integer such that and which means that and so . Thus Since . and is a divisor of Let . of order . We have already shown that it follows that and so is cyclic. , , it follows from closure that .

We will now show that the order of any subgroup of

be any subgroup of

Fundamental theorem of cyclic groups , where is the least positive integer such that , with Since , we must have : . It follows that for some integer . Thus . be any positive divisor of . We will show that We will now prove the last part of the theorem. Let , since is the smallest positive integer such that . . We know that and therefore we can write

108

is the one and only subgroup of

of order

. Note that

has order . Let where be any subgroup of , is a divisor of . So and Consequently and so and thus the theorem is proved. , . with order . We know that

Proof by homomorphism with integers


Let be a cyclic group, and let . Since of . If . Since and therefore , then is cyclic generated by be a subgroup of , to is surjective. Let . Since is a subgroup of , is . Define a morphism . by is a subgroup onto , is

is surjective, the restriction of is isomorphic to a quotient of , hence

defines a surjective homomorphism from is isomorphic to

for some integer . Therefore

, which is cyclic. Otherwise,

isomorphic to a quotient of

, and they are commonly known to be cyclic.

Fundamental theorem of cyclic groups

109

Converse
The following statements are equivalent. A group G of order is cyclic. For every divisor of a group G has exactly one subgroup of order For every divisor of a group G has at most one subgroup of order . .

Generalization
Suppose that R is a right Ore domain in which every left ideal is principal, and let M be a left R-module which is generated by n elements. Then each submodule of M can also be generated by n elements (and possibly fewer). This result implies the fundamental theorem of cyclic groups by observing that the ring of integers satisfies these conditions, and a cyclic group is precisely a left submodules are its subgroups.) -module which is generated by one element. (Its

Fundamental theorem of Galois theory


In mathematics, the fundamental theorem of Galois theory is a result that describes the structure of certain types of field extensions. In its most basic form, the theorem asserts that given a field extension E /F which is finite and Galois, there is a one-to-one correspondence between its intermediate fields and subgroups of its Galois group. (Intermediate fields are fields K satisfying F K E; they are also called subextensions of E /F.)

Proof
The proof of the fundamental theorem is not trivial. The crux in the usual treatment is a rather delicate result of Emil Artin which allows one to control the dimension of the intermediate field fixed by a given group of automorphisms. The automorphisms of a Galois extension K/F are linearly independent as functions over the field K. The proof of this fact follows from a more general notion, namely, the linear independence of characters. There is also a fairly simple proof using the primitive element theorem. This proof seems to be ignored by most modern treatments, possibly because it requires a separate (but easier) proof in the case of finite fields.[1] In terms of its abstract structure, there is a Galois connection; most of its properties are fairly formal, but the actual isomorphism of the posets requires some work.

Explicit description of the correspondence


For finite extensions, the correspondence can be described explicitly as follows. For any subgroup H of Gal(E /F ), the corresponding field, usually denoted EH, is the set of those elements of E which are fixed by every automorphism in H. For any intermediate field K of E /F, the corresponding subgroup is just Aut(E /K ), that is, the set of those automorphisms in Gal(E /F ) which fix every element of K. For example, the topmost field E corresponds to the trivial subgroup of Gal(E /F ), and the base field F corresponds to the whole group Gal(E /F ).

Fundamental theorem of Galois theory

110

Properties of the correspondence


The correspondence has the following useful properties. It is inclusion-reversing. The inclusion of subgroups H1 H2 holds if and only if the inclusion of fields EH1 EH2 holds. Degrees of extensions are related to orders of groups, in a manner consistent with the inclusion-reversing property. Specifically, if H is a subgroup of Gal(E /F ), then |H | = [E:EH] and [Gal(E /F ):H ] = [EH:F ]. The field EH is a normal extension of F (or, equivalently, Galois extension, since any subextension of a separable extension is separable) if and only if H is a normal subgroup of Gal(E /F ). In this case, the restriction of the elements of Gal(E /F ) to EH induces an isomorphism between Gal(EH/F ) and the quotient group Gal(E /F )/H.

Example
Consider the field K = Q(2, 3) = Q(2)(3). Since K is first determined by adjoining 2, then 3, each element of K can be written as:

Lattice of subgroups and subfields

where a, b, c, d are rational numbers. Its Galois group G = Gal (K/Q) can be determined by examining the automorphisms of K which fix a. Each such automorphism must send 2 to either 2 or 2, and must send 3 to either 3 or 3 since the permutations in a Galois group can only permute the roots of an irreducible polynomial. Suppose that f exchanges 2 and 2, so

and g exchanges 3 and 3, so

These are clearly automorphisms of K. There is also the identity automorphism e which does not change anything, and the composition of f and g which changes the signs on both radicals:

Therefore

and G is isomorphic to the Klein four-group. It has five subgroups, each of which correspond via the theorem to a subfield of K. The trivial subgroup (containing only the identity element) corresponds to all of K. The entire group G corresponds to the base field Q. The two-element subgroup {1, f } corresponds to the subfield Q(3), since f fixes 3. The two-element subgroup {1, g} corresponds to the subfield Q(2), again since g fixes 2. The two-element subgroup {1, fg} corresponds to the subfield Q(6), since fg fixes 6.

Fundamental theorem of Galois theory

111

Example
The following is the simplest case where the Galois group is not abelian. Consider the splitting field K of the polynomial x32 over Q; that is, K = Q (, ), where is a cube root of 2, and is a cube root of 1 (but not 1 itself). For example, if we imagine K to be inside the field of complex numbers, we may take to be the real cube root of 2, and to be

Lattice of subgroups and subfields

It can be shown that the Galois group G = Gal (K/Q) has six elements, and is isomorphic to the group of permutations of three objects. It is generated by (for example) two automorphisms, say f and g, which are determined by their effect on and ,

and then

The subgroups of G and corresponding subfields are as follows: As usual, the entire group G corresponds to the base field Q, and the trivial group {1} corresponds to the whole field K. There is a unique subgroup of order 3, namely {1, f, f 2}. The corresponding subfield is Q(), which has degree two over Q (the minimal polynomial of is x2 + x + 1), corresponding to the fact that the subgroup has index two in G. Also, this subgroup is normal, corresponding to the fact that the subfield is normal over Q. There are three subgroups of order 2, namely {1, g}, {1, gf } and {1, gf 2}, corresponding respectively to the three subfields Q(), Q(), Q(2). These subfields have degree three over Q, again corresponding to the subgroups having index 3 in G. Note that the subgroups are not normal in G, and this corresponds to the fact that the subfields are not Galois over Q. For example, Q() contains only a single root of the polynomial x32, so it cannot be normal over Q.

Applications
The theorem converts the difficult-sounding problem of classifying the intermediate fields of E /F into the more tractable problem of listing the subgroups of a certain finite group. For example, to prove that the general quintic equation is not solvable by radicals (see AbelRuffini theorem), one first restates the problem in terms of radical extensions (extensions of the form F() where is an n-th root of some element of F), and then uses the fundamental theorem to convert this statement into a problem about groups. That can then be attacked directly. Theories such as Kummer theory and class field theory are predicated on the fundamental theorem.

Fundamental theorem of Galois theory

112

Infinite case
There is also a version of the fundamental theorem that applies to infinite algebraic extensions, which are normal and separable. It involves defining a certain topological structure, the Krull topology, on the Galois group; only subgroups that are also closed sets are relevant in the correspondence.

References
[1] See Marcus, Daniel (1977). Number Fields. Appendix 2. New York: Springer-Verlag. ISBN0387902791.

Fundamental theorem of linear algebra


In mathematics, the fundamental theorem of linear algebra makes several statements regarding vector spaces. These may be stated concretely in terms of the rank r of an mn matrix A and its singular value decomposition:

First, each matrix fundamental subspaces are:


name of subspace column space, range or image nullspace or kernel row space or coimage left nullspace or cokernel

has

rows and

columns) induces four fundamental subspaces. These

definition or or or or

containing space

dimension (rank) (nullity) The first The last The first The last

basis columns of columns of rows of rows of

Secondly: 1. In 2. In , , , that is, the nullspace is the orthogonal complement of the row space , that is, the left nullspace is the orthogonal complement of the column space.

Fundamental theorem of linear algebra

113

The four subspaces associated to a matrix A.

The dimensions of the subspaces are related by the ranknullity theorem, and follow from the above theorem. Further, all these spaces are intrinsically defined they do not require a choice of basis in which case one rewrites this in terms of abstract vector spaces, operators, and the dual spaces as and : the kernel and image of are the cokernel and coimage of .

References
Strang, Gilbert. Linear Algebra and Its Applications. 3rd ed. Orlando: Saunders, 1988. Strang, Gilbert (1993), "The fundamental theorem of linear algebra" [1], American Mathematical Monthly 100 (9): 848855, doi:10.2307/2324660, JSTOR2324660

External links
Gilbert Strang, MIT Linear Algebra Lecture on the Four Fundamental Subspaces [2] at Google Video, from MIT OpenCourseWare

References
[1] http:/ / www. eng. iastate. edu/ ~julied/ classes/ CE570/ Notes/ strangpaper. pdf [2] http:/ / ocw. mit. edu/ OcwWeb/ Mathematics/ 18-06Spring-2005/ VideoLectures/ detail/ lecture10. htm

Fundamental theorem on homomorphisms

114

Fundamental theorem on homomorphisms


In abstract algebra, the fundamental theorem on homomorphisms, also known as the fundamental homomorphism theorem, relates the structure of two objects between which a homomorphism is given, and of the kernel and image of the homomorphism. The homomorphism theorem is used to prove the isomorphism theorems.

Group theoretic version


Given two groups G and H and a group homomorphism f : GH, let K be a normal subgroup in G and the natural surjective homomorphism GG/K. If K ker(f) then there exists a unique homomorphism h:G/KH such that f = h . The situation is described by the following commutative diagram:

By setting K = ker(f) we immediately get the first isomorphism theorem.

Other versions
Similar theorems are valid for monoids, vector spaces, modules, and rings.

External links
A proof at planetmath [1]

References
[1] http:/ / planetmath. org/ encyclopedia/ FundamentalHomomorphismTheorem. html

GilmanGriess theorem

115

GilmanGriess theorem
In finite group theory, a mathematical discipline, the GilmanGriess theorem, proved by (Gilman & Griess 1983), classifies the finite simple groups of characteristic 2 type with e(G)4 that have a "standard component", which covers one of the three cases of the trichotomy theorem.

References
Gilman, Robert H.; Griess, Robert L. (1983), "Finite groups with standard components of Lie type over fields of characteristic two", Journal of Algebra 80 (2): 383516, doi:10.1016/0021-8693(83)90007-8, ISSN0021-8693, MR691810

Going up and going down


In commutative algebra, a branch of mathematics, going up and going down are terms which refer to certain properties of chains of prime ideals in integral extensions. The phrase going up refers to the case when a chain can be extended by "upward inclusion", while going down refers to the case when a chain can be extended by "downward inclusion". The major results are the CohenSeidenberg theorems, which were proved by Irving S. Cohen and Abraham Seidenberg. These are colloquially known as the going-up and going-down theorems.

Going up and going down


Let AB be an extension of commutative rings. The going-up and going-down theorems give sufficient conditions for a chain of prime ideals in B, each member of which lies over members of a longer chain of prime ideals in A, can be extended to the length of the chain of prime ideals in A.

Lying over and incomparability


First, we fix some terminology. If and are prime ideals of A and B, respectively, such that

then we say that

lies under

and that

lies over

. In general, a ring extension AB of commutative rings is

said to satisfy the lying over property if every prime ideal P of A lies under some prime ideal Q of B. The extension AB is said to satisfy the incomparability property if whenever Q and Q' are distinct primes of B lying over prime P in A, then QQ' and Q' Q.

Going up and going down

116

Going-up
The ring extension AB is said to satisfy the going-up property if whenever

is a chain of prime ideals of A and

(m < n) is a chain of prime ideals of B such that for each 1 i m,

lies over

, then the chain

can be extended to a chain

such that for each 1 i n,

lies over

In (Kaplansky 1970) it is shown that if an extension AB satisfies the going-up property, then it also satsifies the lying-over property.

Going down
The ring extension AB is said to satisfy the going-down property if whenever

is a chain of prime ideals of A and

(m < n) is a chain of prime ideals of B such that for each 1 i m,

lies over

, then the chain

can be extended to a chain

such that for each 1 i n,

lies over

There is a generalization of the ring extension case with ring morphisms. Let f : A B be a (unital) ring homomorphism so that B is a ring extension of f(A). Then f is said to satisfy the going-up property if the going-up property holds for f(A) in B. Similarly, if f(A) is a ring extension, then f is said to satisfy the going-down property if the going-down property holds for f(A) in B. In the case of ordinary ring extensions such as AB, the inclusion map is the pertinent map.

Going-up and going-down theorems


The usual statements of going-up and going-down theorems refer to a ring extension AB: 1. (Going up) If B is an integral extension of A, then the extension satisfies the going-up property (and hence the lying over property), and the incomparability property. 2. (Going down) If B is an integral extension of A, and B is a domain, and A is integrally closed in its field of fractions, then the extension (in addition to going-up, lying-over and incomparability) satisfies the going-down property. There is another sufficient condition for the going-down property: If AB is a flat extension of commutative rings, then the going-down property holds[1] . Proof:[2] Let p1p2 be prime ideals of A and let q2 be a prime ideal of B such that q2 A = p2. We wish to prove that there is a prime ideal q1 of B contained in q2 such that q1 A = p1. Since AB is a flat extension of rings, it follows

Going up and going down that Ap2Bq2 is a flat extension of rings. In fact, Ap2Bq2 is a faithfully flat extension of rings since the inclusion map Ap2 Bq2 is a local homomorphism. Therefore, the induced map on spectra Spec(Bq2) Spec(Ap2) is surjective and there exists a prime ideal of Bq2 that contracts to the prime ideal p1Ap2 of Ap2. The contraction of this prime ideal of Bq2 to B is a prime ideal q1 of B contained in q2 that contracts to p1. The proof is complete. Q.E.D.

117

References
[1] This follows from a much more general lemma in Bruns-Herzog, Lemma A.9 on page 415. [2] Matsumura, page 33, (5.D), Theorem 4

Atiyah, M. F., and I. G. MacDonald, Introduction to Commutative Algebra, Perseus Books, 1969, ISBN 0-201-00361-9 MR242802 Winfried Bruns; Jrgen Herzog, CohenMacaulay rings. Cambridge Studies in Advanced Mathematics, 39. Cambridge University Press, Cambridge, 1993. xii+403 pp. ISBN 0-521-41068-1 Kaplansky, Irving, Commutative rings, Allyn and Bacon, 1970. Matsumura, Hideyuki (1970). Commutative algebra. W. A. Benjamin. ISBN978-0805370256. Sharp, R. Y. (2000). "13 Integral dependence on subrings (13.38 The going-up theorem, pp. 258259; 13.41 The going down theorem, pp. 261262)". Steps in commutative algebra. London Mathematical Society Student Texts. 51 (Second ed.). Cambridge: Cambridge University Press. pp.xii+355. ISBN0-521-64623-5. MR1817605.

Goldie's theorem
In mathematics, Goldie's theorem is a basic structural result in ring theory, proved by Alfred Goldie during the 1950s. What is now termed a right Goldie ring is a ring R that has finite uniform dimension (="finite rank") as a right module over itself, and satisfies the ascending chain condition on right annihilators of subsets of R. Goldie's theorem states that the semiprime right Goldie rings are precisely those that have a semisimple Artinian right classical ring of quotients. The structure of this ring of quotients is then completely determined by the ArtinWedderburn theorem. In particular, Goldie's theorem applies to semiprime right Noetherian rings, since by definition right Noetherian rings have the ascending chain condition on all right ideals. This is sufficient to guarantee that a ring is right Goldie.

References
Coutinho, S.C. & J.C. McConnell (2003) "The quest for quotient rings (of non-commutative Noetherian rings)", American Mathematical Monthly 110: 298313. Goldie, A.W. (1958). "The structure of prime rings under ascending chain conditions". Proc. London Math. Soc. 8 (4): 589608. doi:10.1112/plms/s3-8.4.589. Goldie, A.W. (1960). "Semi-prime rings with maximal conditions". Proc. London Math. Soc. 10: 201220. doi:10.1112/plms/s3-10.1.201. Herstein, I.N. (1969). Topics in ring theory. Chicago lectures in mathematics. Chicago, Ill.: Chicago Univ. Pr.. pp.6186. ISBN0-226-32802-3.

Goldie's theorem

118

External links
PlanetMath page on Goldie's theorem [1] PlanetMath page on Goldie ring [2]

References
[1] http:/ / planetmath. org/ encyclopedia/ GoldiesTheorem. html [2] http:/ / planetmath. org/ encyclopedia/ GoldieRing. html

GolodShafarevich theorem
In mathematics, the GolodShafarevich theorem was proved in 1964 by two Russian mathematicians, Evgeny Golod and Igor Shafarevich. It is a result in non-commutative homological algebra which has consequences in various branches of algebra.

The inequality
Let A = K<x1, ..., xn> be the free algebra over a field K in n = d+1 non-commuting variables xi. Let J be the 2-sided ideal of A generated by homogeneous elements fj of A of degree dj with 2 d1 d2 ... where dj tends to infinity. Let ri be the number of dj equal to i. Let B=A/J, a graded algebra. Let bj = dim Bj. The fundamental inequality of Golod and Shafarevich states that

As a consequence: B is infinite-dimensional if ri d2/4 for all i if B is finite-dimensional, then ri > d2/4 for some i.

Applications
This result has important applications in combinatorial group theory: If G is a nontrivial finite p-group, then r > d2/4 where d = dimH1(G,Z/pZ) and r = dimH2(G,Z/pZ) (the mod p cohomology groups of G). In particular if G is a finite p-group with minimal number of generators d and has r relators in a given presentation, then r > d2/4. For each prime p, there is an infinite group G generated by three elements in which each element has order a power of p. The group G provides a counterexample to the generalised Burnside conjecture: it is a finitely generated infinite torsion group, although there is no uniform bound on the order of its elements. In class field theory, the class field tower of a number field K is created by iterating the Hilbert class field construction. Another consequence of the construction is that such towers may be infinite (in other words, do not always terminate in a field equal to its Hilbert class field).

GolodShafarevich theorem

119

References
1. Golod, E.S; Shafarevich, I.R. (1964), "On the class field tower", Izv. Akad. Nauk SSSSR 28: 261272 (in Russian) MR0161852 2. Golod, E.S (1964), "On nil-algebras and finitely approximable p-groups.", Izv. Akad. Nauk SSSSR 28: 273276 (in Russian) MR0161878 3. Herstein, I.N. (1968), "Noncommutative rings," Carus Mathematical Monographs, MAA. ISBN 0-88385-039-7. See Chapter 8. 4. Johnson, D.L. (1980). "Topics in the Theory of Group Presentations" (1st ed.). Cambridge University Press. ISBN 0-521-23108-6. See chapter VI. 5. Roquette, P. (1967), On class field towers,pages 231249 in Algebraic number theory, Proceedings of the instructional conference held at the University of Sussex, Brighton, September 117 , 1965. Edited by J. W. S. Cassels and A. Frhlich. Reprint of the 1967 original. Academic Press, London, 1986. xviii+366 pp. ISBN 0-12-163251-2 6. Serre, J.-P. (2002), "Galois Cohomology," Springer-Verlag. ISBN 3-540-42192-0. See Appendix 2. (Translation of Cohomologie Galoisienne, Lecture Notes in Mathematics 5, 1973.)

GorensteinHarada theorem
In mathematical finite group theory, the GorensteinHarada theorem, proved by Gorenstein and Harada(1973, 1974) in a 464 page paper, classifies the simple finite groups of sectional 2-rank at most 4. It is part of the classification of finite simple groups. Finite simple groups of section 2 rank at least 5 have Sylow 2-subgroups with a self-centralizing normal subgroup of rank at least 3, which implies that they have to be of either component type or of characteristic 2 type. Therefore the GorensteinHarada theorem splits the problem of classifying finite simple groups into these two subcases.

References
Gorenstein, D.; Harada, Koichiro (1973), "Finite groups of sectional 2-rank at most 4", in Gagen, Terrence; Hale, Mark P. Jr.; Shult, Ernest E., Finite groups '72. Proceedings of the Gainesville Conference on Finite Groups, March 23-24, 1972, North-Holland Math. Studies, 7, Amsterdam: North-Holland, pp.5767, ISBN978-0-444-10451-9, MR0352243 Gorenstein, D.; Harada, Koichiro (1974), Finite groups whose 2-subgroups are generated by at most 4 elements [1] , Memoirs of the American Mathematical Society, 147, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-1847-3, MR0367048

References
[1] http:/ / books. google. com/ books?id=CzUZAQAAIAAJ

Gromov's theorem on groups of polynomial growth

120

Gromov's theorem on groups of polynomial growth


In geometric group theory, Gromov's theorem on groups of polynomial growth, named for Mikhail Gromov, characterizes finitely generated groups of polynomial growth, as those groups which have nilpotent subgroups of finite index. The growth rate of a group is a well-defined notion from asymptotic analysis. To say that a finitely generated group has polynomial growth means the number of elements of length (relative to a symmetric generating set) at most n is bounded above by a polynomial function p(n). The order of growth is then the least degree of any such polynomial function p. A nilpotent group G is a group with a lower central series terminating in the identity subgroup. Gromov's theorem states that a finitely generated group has polynomial growth if and only if it has a nilpotent subgroup that is of finite index. There is a vast literature on growth rates, leading up to Gromov's theorem. An earlier result of Joseph A. Wolf showed that if G is a finitely generated nilpotent group, then the group has polynomial growth. Yves Guivarc'h and independently Hyman Bass (with different proofs) computed the exact order of polynomial growth. Let G be a finitely generated nilpotent group with lower central series

In particular, the quotient group Gk/Gk+1 is a finitely generated abelian group. The BassGuivarch formula states that the order of polynomial growth of G is

where: rank denotes the rank of an abelian group, i.e. the largest number of independent and torsion-free elements of the abelian group. In particular, Gromov's theorem and the BassGuivarch formula imply that the order of polynomial growth of a finitely generated group is always either an integer or infinity (excluding for example, fractional powers). In order to prove this theorem Gromov introduced a convergence for metric spaces. This convergence, now called the GromovHausdorff convergence, is currently widely used in geometry. A relatively simple proof of the theorem was found by Bruce Kleiner. Later, Terence Tao and Yehuda Shalom modified Kleiner's proof to make an essentially elementary proof as well as a version of the theorem with explicit bounds.[1] [2]

References
[1] http:/ / terrytao. wordpress. com/ 2010/ 02/ 18/ a-proof-of-gromovs-theorem/ [2] Yehuda Shalom; Terence Tao (2009). "A finitary version of Gromov's polynomial growth theorem". arXiv:0910.4148[math.GR].

H. Bass, The degree of polynomial growth of finitely generated nilpotent groups, Proceedings London Mathematical Society, vol 25(4), 1972 M. Gromov, Groups of Polynomial growth and Expanding Maps, Publications mathematiques I.H..S., 53, 1981 (http://www.numdam.org/numdam-bin/feuilleter?id=PMIHES_1981__53_) Y. Guivarc'h, Groupes de Lie croissance polynomiale, C. R. Acad. Sci. Paris Sr. AB 272 (1971). (http:// www.numdam.org/item?id=BSMF_1973__101__333_0) Kleiner, Bruce (2007). "A new proof of Gromov's theorem on groups of polynomial growth". arXiv:0710.4593.

Gromov's theorem on groups of polynomial growth J. A. Wolf, Growth of finitely generated solvable groups and curvature of Riemannian manifolds, Journal of Differential Geometry, vol 2, 1968

121

Grushko theorem
In the mathematical subject of group theory, the Grushko theorem or the Grushko-Neumann theorem is a theorem stating that the rank (that is, the smallest cardinality of a generating set) of a free product of two groups is equal to the sum of the ranks of the two free factors. The theorem was first obtained in a 1940 article of Grushko[1] and then, independently, in a 1943 article of Neumann.[2]

Statement of the theorem


Let A and B be finitely generated groups and let AB be the free product of A and B. Then rank(AB) = rank(A) + rank(B). It is obvious that rank(AB) rank(A) + rank(B) since if X is a finite generating set of A and Y is a finite generating set of B then XY is a generating set for AB and that |XY||X| + |Y|. The opposite inequality, rank(AB) rank(A) + rank(B), requires proof. There is a more precise version of Grushko's theorem in terms of Nielsen equivalence. It states that if M = (g1, g2, ..., gn) is an n-tuple of elements of G = AB such that M generates G, <g1, g2, ..., gn> = G, then M is Nielsen equivalent in G to an n-tuple of the form M' = (a1, ..., ak, b1, ..., bnk) where {a1, ..., ak}A is a generating set for A and where {b1, ..., bnk}B is a generating set for B. In particular, rank(A) k, rank(B) nk and rank(A) + rank(B) k+(nk) = n. If one takes M to be the minimal generating tuple for G, that is, with n = rank(G), this implies that rank(A) + rank(B) rank(G). Since the opposite inequality, rank(G) rank(A) + rank(B), is obvious, it follows that rank(G)=rank(A) + rank(B), as required.

History and generalizations


After the original proofs of Grushko (1940) and Neumann(1943), there were many subsequent alternative proofs, simplifications and generalizations of Grushko's theorem. A close version of Grushko's original proof is given in the 1955 book of Kurosh.[3] Like the original proofs, Lyndon's proof (1965)[4] relied on length-functions considerations but with substantial simplifications. A 1965 paper of Stallings [5] gave a greatly simplified topological proof of Grushko's theorem. A 1970 paper of Zieschang[6] gave a Nielsen equivalence version of Grushko's theorem (stated above) and provided some generalizations of Grushko's theorem for amalgamated free products. Scott (1974) gave another topological proof of Grushko's theorem, inspired by the methods of 3-manifold topology[7] Imrich (1984)[8] gave a version of Grushko's theorem for free products with infinitely many factors. Modern techniques of Bass-Serre theory, particularly the machinery of foldings for group actions on trees and for graphs of groups provide a relatively straightforward proof of Grushko's theorem (see, for example [9] [10] ). Grushko's theorem is, in a sense, a starting point in Dunwoody's theory of accessibility for finitely generated and finitely presented groups. Since the ranks of the free factors are smaller than the rank of a free product, Grushko's theorem implies that the process of iterated splitting of a finitely generated group G as a free product must terminate in a finite number of steps (more precisely, in at most rank(G) steps). There is a natural similar question for iterating splittings of finitely generated groups over finite subgroups. Dunwoody proved that such a process must always terminate if a group G is finitely presented[11] but may go on forever if G is finitely generated but not finitely presented.[12]

Grushko theorem An algebraic proof of a substantial generalization of Grushko's theorem using the machinery of groupoids was given by Higgins (1966).[13] Higgins' theorem starts with groups G and B with free decompositions G = i Gi, B = i Bi and f : G B a morphism such that f(Gi) = Bi for all i. Let H be a subgroup of G such that f(H) = B. Then H has a decomposition H = i Hi such that f(Hi) = Bi for all i. Full details of the proof and applications may also be found in .[14]

122

Grushko decomposition theorem


A useful consequence of the original Grushko theorem is the so-called Grushko decomposition theorem. It asserts that any nontrivial finitely generated group G can be decomposed as a free product G = A1A2...ArFs, where s 0, r 0, where each of the groups Ai is nontrivial, freely indecomposable (that is, it cannot be decomposed as a free product) and not infinite cyclic, and where Fs is a free group of rank s; moreover, for a given G, the groups A1, ..., Ar are unique up to a permutation of their conjugacy classes in G (and, in particular, the sequence of isomorphism types of these groups is unique up to a permutation) and the numbers s and r are unique as well. More precisely, if G = B1...BkFt is another such decomposition then k = r, s = t, and there exists a permutation Sr such that for each i=1,...,r the subgroups Ai and B(i) are conjugate in G. The existence of the above decomposition, called the Grushko decomposition of G, is an immediate corollary of the original Grushko theorem, while the uniqueness statement requires additional arguments (see, for example[15] ). Algorithmically computing the Grushko decomposition for specific classes of groups is a difficult problem which primarily requires being able to determine if a given group is freely decomposable. Positive results are available for some classes of groups such as torsion-free word-hyperbolic groups, certain classes of relatively hyperbolic groups,[16] fundamental groups of finite graphs of finitely generated free groups[17] and others. Grushko decomposition theorem is a group-theoretic analog of the Kneser prime decomposition theorem for 3-manifolds which says that a closed 3-manifold can be uniquely decomposed as a connected sum of irreducible 3-manifolds.[18]

Sketch of the proof using Bass-Serre theory


The following is a sketch of the proof of Grushko's theorem based on the use of foldings techniques for groups acting on trees (see [9] [10] for complete proofs using this argument). Let S={g1,....,gn} be a finite generating set for G=AB of size |S|=n=rank(G). Realize G as the fundamental group of a graph of groups Y which is a single non-loop edge with vertex groups A and B and with the trivial edge group. Let be the Bass-Serre covering tree for Y. Let F=F(x1,....,xn) be the free group with free basis x1,....,xn and let 0:F G be the homomorphism such that 0(xi)=gi for i=1,...,n. Realize F as the fundamental group of a graph Z0 which is the wedge of n circles that correspond to the elements x1,....,xn. We also think of Z0 as a graph of groups with the underlying graph Z0 and the trivial vertex and edge groups. Then the universal cover of Z0 and the Bass-Serre covering tree for Z0 coincide. Consider a 0-equivariant map so that it sends vertices to vertices and edges to edge-paths. This map is non-injective and, since both the source and the target of the map are trees, this map "folds" some edge-pairs in the source. The graph of groups Z0 serves as an initial approximation for Y. We now start performing a sequence of "folding moves" on Z0 (and on its Bass-Serre covering tree) to construct a sequence of graphs of groups Z0, Z1, Z2, ...., that form better and better approximations for Y. Each of the graphs of groups Zj has trivial edge groups and comes with the following additional structure: for each nontrivial vertex group of it there assigned a finite generating set of that vertex group. The complexity c(Zj) of Zj is the sum of the sizes of the generating sets of its vertex groups and the rank of the free group 1(Zj). For the initial approximation graph we have c(Z0)=n.

Grushko theorem The folding moves that take Zj to Zj+1 can be of one of two types: folds that identify two edges of the underlying graph with a common initial vertex but distinct end-vertices into a single edge; when such a fold is performed, the generating sets of the vertex groups and the terminal edges are "joined" together into a generating set of the new vertex group; the rank of the fundamental group of the underlying graph does not change under such a move. folds that identify two edges, that already had common initial vertices and common terminal vertices, into a single edge; such a move decreases the rank of the fundamental group of the underlying graph by 1 and an element that corresponded to the loop in the graph that is being collapsed is "added" to the generating set of one of the vertex groups. One sees that the folding moves do not increase complexity but they do decrease the number of edges in Zj. Therefore the folding process must terminate in a finite number of steps with a graph of groups Zk that cannot be folded any more. It follows from the basic Bass-Serre theory considerations that Zk must in fact be equal to the edge of groups Y and that Zk comes equipped with finite generating sets for the vertex groups A and B. The sum of the sizes of these generating sets is the complexity of Zk which is therefore less than or equal to c(Z0)=n. This implies that the sum of the ranks of the vertex groups A and B is at most n, that is rank(A)+rank(B)rank(G), as required.

123

Notes
[1] [2] [3] [4] [5] [6] I. A. Grushko, On the bases of a free product of groups, Matematicheskii Sbornik, vol 8 (1940), pp. 169182. B. H. Neumann. On the number of generators of a free product. Journal of the London Mathematical Society, vol 18, (1943), pp. 1220. A. G. Kurosh, The theory of groups. Vol. I. Translated and edited by K. A. Hirsch. Chelsea Publishing Co., New York, N.Y., 1955 , Roger C. Lyndon, Grushko's theorem. Proceedings of the American Mathematical Society, vol. 16 (1965), pp. 822826. John R. Stallings. A topological proof of Grushko's theorem on free products. Mathematische Zeitschrift, vol. 90 (1965), pp. 18. Heiner Zieschang. ber die Nielsensche Krzungsmethode in freien Produkten mit Amalgam. Inventiones Mathematicae, vol. 10 (1970), pp. 437 [7] Scott, Peter. An introduction to 3-manifolds. Department of Mathematics, University of Maryland, Lecture Note, No. 11. Department of Mathematics, University of Maryland, College Park, Maryland, 1974 [8] Wilfried Imrich Grushko's theorem. Archiv der Mathematik (Basel), vol. 43 (1984), no. 5, pp. 385-387 [9] John R. Stallings. Foldings of G-trees. Arboreal group theory (Berkeley, California, 1988), pp. 355368, Mathematical Sciences Research Institute Publications, 19. Springer, New York, 1991; ISBN 0-387-97518-7 [10] Ilya Kapovich, Richard Weidmann, and Alexei Miasnikov. Foldings, graphs of groups and the membership problem. International Journal of Algebra and Computation, vol. 15 (2005), no. 1, pp. 95128 [11] Martin J. Dunwoody. The accessibility of finitely presented groups. Inventiones Mathematicae, vol. 81 (1985), no. 3, pp. 449457 [12] Martin J. Dunwoody. An inaccessible group. Geometric group theory, Vol. 1 (Sussex, 1991), pp. 7578, London Mathematical Society Lecture Notes Series, 181, Cambridge University Press, Cambridge, 1993. ISBN 0-521-43529-3 [13] P. J. Higgins. Grushko's theorem. Journal of Algebra, vol. 4 (1966), pp. 365372 [14] Higgins, Philip J., Notes on categories and groupoids. Van Nostrand Rienhold Mathematical Studies, No. 32. Van Nostrand Reinhold Co., London-New York-Melbourne, 1971. Reprinted as Theory and Applications of Categories Reprint No 7, 2005. (http:/ / www. tac. mta. ca/ tac/ reprints/ articles/ 7/ tr7abs. html) [15] John Stallings. Coherence of 3-manifold fundamental groups. (http:/ / www. numdam. org/ numdam-bin/ fitem?id=SB_1975-1976__18__167_0) Sminaire Bourbaki, 18 (1975-1976), Expos No. 481. [16] Franois Dahmani and Daniel Groves. Detecting free splittings in relatively hyperbolic groups. (http:/ / www. ams. org/ tran/ 0000-000-00/ S0002-9947-08-04486-3/ ) Transactions of the American Mathematical Society. Posted online July 21, 2008. [17] Guo-An Diao and Mark Feighn. The Grushko decomposition of a finite graph of finite rank free groups: an algorithm. (http:/ / msp. warwick. ac. uk/ gt/ 2005/ 09/ p041. xhtml) Geometry and Topology. vol. 9 (2005), pp. 18351880 [18] H. Kneser, Geschlossene Flchen in dreidimensionalen Mannigfaltigkeiten. Jahresber. Deutsch. Math. Verein., vol. 38 (1929), pp. 248260

Haboush's theorem

124

Haboush's theorem
In mathematics Haboush's theorem, often still referred to as the Mumford conjecture, states that for any semisimple algebraic group G over a field K, and for any linear representation of G on a K-vector space V, given v0 in V that is fixed by the action of G, there is a G-invariant polynomial F on V such that F(v) 0. The polynomial can be taken to be homogeneous, in other words an element of a symmetric power of the dual of V, and if the characteristic is p>0 the degree of the polynomial can be taken to be a power of p. When K has characteristic 0 this was well known; in fact Weyl's theorem on the complete reducibility of the representations of G implies that F can even be taken to be linear. Mumford's conjecture about the extension to prime characteristic p was proved by W. J. Haboush (1975), about a decade after the problem had been posed by David Mumford, in the introduction to the first edition of his book Geometric Invariant Theory.

Applications
Haboush's theorem can be used to generalize results of geometric invariant theory from characteristic 0, where they were already known, to characteristic p>0. In particular Nagata's earlier results together with Haboush's theorem show that if a reductive group (over an algebraically closed field) acts on a finitely generated algebra then the fixed subalgebra is also finitely generated. Haboush's theorem implies that if G is a reductive algebraic group acting regularly on an affine algebraic variety, then disjoint closed invariant sets X and Y can be separated by an invariant function f (this means that f is 0 on X and 1 on Y). C.S. Seshadri (1977) extended Haboush's theorem to reductive groups over schemes. It follows from the work of Nagata (1963), Haboush, and Popov that the following conditions are equivalent for an affine algebraic group G over a field K: G is reductive (its unipotent radical is trivial). For any non-zero invariant vector in a rational representation on G, there is an invariant homogeneous polynomial that does not vanish on it. For any finitely generated K algebra acted on rationally by G, the algebra of fixed elements is finitely generated.

Proof
The theorem is proved in several steps as follows: We can assume that the group is defined over an algebraically closed field K of characteristic p>0. Finite groups are easy to deal with as one can just take a product over all elements, so one can reduce to the case of connected reductive groups (as the connected component has finite index). By taking a central extension which is harmless one can also assume the group G is simply connected. Let A(G) be the coordinate ring of G. This is a representation of G with G acting by left translations. Pick an element v of the dual of V that has value 1 on the invariant vector v. The map V to A(G) by sending wV to the element aA(G) with a(g) = v(g(w)). This sends v to 1A(G), so we can assume that VA(G) and v=1. The structure of the representation A(G) is given as follows. Pick a maximal torus T of G, and let it act on A(G) by right translations (so that it commutes with the action of G). Then A(G) splits as a sum over characters of T of the subrepresentations A(G) of elements transforming according to . So we can assume that V is contained in the T-invariant subspace A(G) of A(G). The representation A(G) is an increasing union of subrepresentations of the form E+nEn, where is the Weyl vector for a choice of simple roots of T, n is a positive integer, and E is the space of sections of the line

Haboush's theorem bundle over G/B corresponding to a character of T, where B is a Borel subgroup containing T. If n is sufficiently large then En has dimension (n+1)N where N is the number of positive roots. This is because in characteristic 0 the corresponding module has this dimension by the Weyl character formula, and for n large enough that the line bundle over G/B is very ample, En has the same dimension as in characteristic 0. If q=pr for a positive integer r, and n=q1, then En contains the Steinberg representation of G(Fq) of dimension qN. (Here Fq K is the finite field of order q.) The Steinberg representation is an irreducible representation of G(Fq) and therefore of G(K), and for r large enough it has the same dimension as En, so there are infinitely many values of n such that En is irreducible. If En is irreducible it is isomorphic to its dual, so EnEn is isomorphic to End(En). Therefore the T-invariant subspace A(G) of A(G) is an increasing union of subrepresentations of the form End(E) for representations E (of the form E(q1))). However for representations of the form End(E) an invariant polynomial that separates 0 and 1 is given by the determinant. This completes the sketch of the proof of Haboush's theorem.

125

References
Demazure, Michel (1976), "Dmonstration de la conjecture de Mumford (d'aprs W. Haboush)", Sminaire Bourbaki (1974/1975: Exposs Nos. 453--470), Lecture Notes in Math., 514, Berlin: Springer, pp.138144, doi:10.1007/BFb0080063, ISBN978-3-540-07686-5, MR0444786 Haboush, W. J. (1975), "Reductive groups are geometrically reductive", Ann. Of Math. (The Annals of Mathematics, Vol. 102, No. 1) 102 (1): 6783, doi:10.2307/1970974, JSTOR1970974 Mumford, D.; Fogarty, J.; Kirwan, F. Geometric invariant theory. Third edition. Ergebnisse der Mathematik und ihrer Grenzgebiete (2) (Results in Mathematics and Related Areas (2)), 34. Springer-Verlag, Berlin, 1994. xiv+292 pp. MR1304906 ISBN 3-540-56963-4 Nagata, Masayoshi (1963), "Invariants of a group in an affine ring" [1], Journal of Mathematics of Kyoto University 3: 369377, ISSN0023-608X, MR0179268 M. Nagata, T. Miyata, "Note on semi-reductive groups" J. Math. Kyoto Univ. , 3 (1964) pp.379382 Popov, V.L. (2001), "Mumford hypothesis" [2], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 C.S. Seshadri, "Geometric reductivity over arbitrary base" Adv. Math. , 26 (1977) pp.225274

References
[1] http:/ / projecteuclid. org/ euclid. kjm/ 1250524787 [2] http:/ / eom. springer. de/ M/ m065570. htm

Hahn embedding theorem

126

Hahn embedding theorem


In mathematics, especially in the area of abstract algebra dealing with ordered structures on abelian groups, the Hahn embedding theorem gives a simple description of all linearly ordered abelian groups. The theorem states: Any linearly ordered abelian group

can be embedded as an ordered subgroup of the additive

group endowed with a lexicographical order, where is the additive group of real numbers (with its standard order), and is the set of Archimedean equivalence classes of . Let denote the identity element of ; denote this element by . For any nonzero and

, exactly one of the elements . (Heuristically: neither

or nor

is is

greater than

. Two nonzero elements

are Archimedean equivalent if there

exist natural numbers

such that

"infinitesimal" with respect to the other). The group

is Archimedean if all nonzero elements are

Archimedean-equivalent. In this case, is a singleton, so is just the group of real numbers. Then Hahn's Embedding Theorem reduces to Hlder's theorem (which states that a linearly ordered abelian group is Archimedean if and only if it is a subgroup of the ordered additive group of the real numbers). (Gravett 1956) gives a clear statement and proof of the theorem. The papers of (Clifford 1954) and (Hausner & Wendel 1952) together provide another proof. See also (Fuchs & Salce 2001, p.62).

References
Fuchs, Lszl; Salce, Luigi (2001), Modules over non-Noetherian domains, Mathematical Surveys and Monographs, 84, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-1963-0, MR1794715 Hahn, H. (1907), "ber die nichtarchimedischen Grensysteme." (in German), Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften, Wien, Mathematisch - Naturwissenschaftliche Klasse (Wien. Ber.) 116: 601655 Gravett, K. A. H. (1956), "Ordered Abelian Groups", Quarterly Journal of Mathematics of Oxford Series 2 7: 5763, doi:10.1093/qmath/7.1.57 Clifford, A.H. (1954), "Note on Hahn's Theorem on Ordered Abelian Groups", Proceedings of the American Mathematical Society 5 (6): 860863 Hausner, M.; Wendel, J.G. (1952), "Ordered vector spaces", Proceedings of the American Mathematical Society 3: 977982, doi:10.1090/S0002-9939-1952-0052045-1

Hajs's theorem

127

Hajs's theorem
In group theory, Hajs's theorem states that if a finite abelian group is expressed as the Cartesian product of simplexes, that is, sets of the form {e,a,a2,...,as-1} where e is the identity element, then at least one of the factors is a subgroup. The theorem was proved by the Hungarian mathematician Gyrgy Hajs in 1941 using group rings. Rdei later proved the statement when the factors are only required to contain the identity element and be of prime cardinality. An equivalent statement on homogeneous linear forms was originally conjectured by Hermann Minkowski. A consequence is Minkowski's conjecture on lattice tilings, which says that in any lattice tiling of space by cubes, there are two cubes that meet face to face. Keller's conjecture is the same conjecture for non-lattice tilings, which turns out to be false in high dimensions.

References
G. Hajs: ber einfache und mehrfache Bedeckung des 'n'-dimensionalen Raumes mit einem Wrfelgitter, Math. Z., 47(1941), 427467. H. Minkowski: Diophantische Approximationen, Leipzig, 1907.
In this lattice tiling of the plane by congruent squares, the green and violet squares meet edge-to-edge as do the blue and orange squares.

L. Rdei, Die neue Theorie der endlichen abelschen Gruppen und Verallgemeinerung des Hauptsatzes von Hajs, Acta Math. Acad. Sci. Hung., 16 (1965), 329373.

Stein, Sherman K. (1974), "Algebraic tiling" [1], The American Mathematical Monthly 81: 445462, ISSN0002-9890, MR0340063 Stein, Sherman K.; Szab, Sndor (1994), Algebra and tiling [2], Carus Mathematical Monographs, 25, Mathematical Association of America, ISBN978-0-88385-028-2, MR1311249

References
[1] http:/ / www. jstor. org/ stable/ 2318582 [2] http:/ / books. google. com/ books?id=QOa-mnX5Y4QC

Harish-Chandra isomorphism

128

Harish-Chandra isomorphism
In mathematics, the Harish-Chandra isomorphism, introduced by Harish-Chandra(1951), is an isomorphism of commutative rings constructed in the theory of Lie algebras. The isomorphism maps the center Z(U(g)) of the universal enveloping algebra U(g) of a reductive Lie algebra g to the elements S(h)W of the symmetric algebra S(h) of a Cartan subalgebra h that are invariant under the Weyl group W.

Fundamental invariants
Let n be the rank of g, which is the dimension of the Cartan subalgebra h. H. S. M. Coxeter observed that S(h)W is a polynomial algebra in n variables (see ChevalleyShephardTodd theorem for a more general statement). Therefore, the center of the universal enveloping algebra of a reductive Lie algebra is a polynomial algebra. The degrees of the generators are the degrees of the fundamental invariants given in the following table.
Lie algebra Coxeter number h Dual Coxeter number Degrees of fundamental invariants R An Bn Cn Dn E6 E7 E8 F4 G2 0 n+1 2n 2n 2n2 12 18 30 12 6 0 n+1 2n1 n+1 2n2 12 18 30 9 4 1 2, 3, 4, ..., n+1 2, 4, 6, ..., 2n 2, 4, 6, ..., 2n n; 2, 4, 6, ..., 2n2 2, 5, 6, 8, 9, 12 2, 6, 8, 10, 12, 14, 18 2, 8, 12, 14, 18, 20, 24, 30 2, 6, 8, 12 2, 6

For example, the center of the universal enveloping algebra of G2 is a polynomial algebra on generators of degrees 2 and 6.

Examples
If g is the Lie algebra sl2(R), then the center of the universal enveloping algebra is generated by the Casimir invariant of degree 2, and the ring of invariants of the Weyl group is also generated by an element of degree2.

Introduction and setting


Let be a semisimple Lie algebra, , resp. . its Cartan subalgebra and have been fixed. Let , resp. be two elements of the weight space and be highest weight modules with highest assume that a set of positive roots weight

Harish-Chandra isomorphism

129

Central characters
The -modules and and . are homomorphims to scalars called central characters. are representations of the universal enveloping algebra , and its center acts on the modules by scalar multiplication (this follows from the fact that the modules are generated by a highest weight vector). So, for and similarly for The functions

Statement of Harish-Chandra theorem


For any , the characters if and only if and are on the same orbit of the Weyl group of ). under the affine action (corresponding to the choice of the positive roots

Another closely related formulation is that the Harish-Chandra homomorphism from the centrum of the universal enveloping algebra to (invariant polynomials over the Cartan subalgebra fixed by the affine action of the Weyl group) is an isomorphism.

Applications
The theorem may be used to obtain a simple algebraic proof of Weyl's character formula for finite dimensional representations. Further, it is a necessary condition for the existence of a nonzero homomorphism of some highest weight moules (a homomorphism of such modules preserves central character). A simple consequence is that for Verma modules or generalized Verma modules with highest weight , there exist only finitely many weights such that a nonzero homomorphism exists.

References
Harish-Chandra (1951), "On some applications of the universal enveloping algebra of a semisimple Lie algebra", Transactions of the American Mathematical Society 70: 2896, ISSN0002-9947, JSTOR1990524, MR0044515 Humphreys, James E. (2000), Introduction to Lie algebras and representation theory, Birkhuser, p.126, ISBN978-0387900537 Humphreys, James E. (2008), Representations of semisimple Lie algebras in the BGG category O, AMS, p.26, ISBN978-0821846780 Knapp, Anthony W.; Vogan, David A. (1995), Cohomological induction and unitary representations, Princeton Mathematical Series, 45, Princeton University Press, ISBN978-0-691-03756-1, MR1330919 Knapp, Anthony, Lie groups beyond an introduction, Second edition, pages 300303.

Hasse norm theorem

130

Hasse norm theorem


In number theory, the Hasse norm theorem states that if L/K is a cyclic extension of number fields, then if a nonzero element of K is a local norm everywhere, then it is a global norm. Here to be a global norm means to be an element k of K such that there is an element l of L with ; in other words k is a relative norm of some element of the extension field L. To be a local norm means that for some prime p of K and some prime P of L lying over K, then k is a norm from LP; here the "prime" p can be an archimedean valuation, and the theorem is a statement about completions in all valuations, archimedean and non-archimedean. The theorem is no longer true in general if the extension is abelian but not cyclic. A counter-example is given by the field where every rational square is a local norm everywhere but is not a global norm. This is an example of a theorem stating a local-global principle, and is due to Helmut Hasse.

References
H. Hasse, "A history of class field theory", in J.W.S. Cassels and A. Frohlich (edd), Algebraic number theory, Academic Press, 1973. Chap.XI. G. Janusz, Algebraic number fields, Academic Press, 1973. Theorem V.4.5, p.156

HasseArf theorem
In mathematics, specifically in local class field theory, the HasseArf theorem is a result concerning jumps of a filtration of the Galois group of a finite Galois extension. A special case of it was originally proved by Helmut Hasse,[1] [2] and the general result was proved by Cahit Arf.[3]

Statement
Higher ramification groups
The theorem deals with the upper numbered higher ramification groups of a finite abelian extension L/K. So assume L/K is a finite Galois extension, and that vK is a discrete normalised valuation of K, whose residue field has characteristic p>0, and which admits a unique extension to L, say w. Denote by vL the associated normalised valuation ew of L and let be the valuation ring of L under vL. Let L/K have Galois group G and define the s-th ramification group of L/K for any real s1 by

So, for example, G1 is the Galois group G. To pass to the upper numbering one has to define the function L/K which in turn is the inverse of the function L/K defined by

The upper numbering of the ramification groups is then defined by Gt(L/K)=Gs(L/K) where s=L/K(t). These higher ramification groups Gt(L/K) are defined for any real t1, but since vL is a discrete valuation, the groups will change in discrete jumps and not continuously. Thus we say that t is a jump of the filtration {Gt(L/K):t1} if Gt(L/K)Gu(L/K) for any u>t. The HasseArf theorem tells us the arithmetic nature of these jumps

HasseArf theorem

131

Statement of the theorem


With the above set up, the theorem states that the jumps of the filtration {Gt(L/K):t1} are all rational integers.

Example
Suppose G is cyclic of order , residue characteristic and be the subgroup of such that of order . The the theorem says that there exist positive integers

...
[4]

Notes
[1] H. Hasse, Fhrer, Diskriminante und Verzweigunsgskrper relativ Abelscher Zahlkrper, J. Reine Angew. Math. 162 (1930), pp.169184. [2] H. Hasse, Normenresttheorie galoisscher Zahlkrper mit Anwendungen auf Fhrer und Diskriminante abelscher Zahlkrper, J. Fac. Sci. Tokyo 2 (1934), pp.477498. [3] C. Arf, Untersuchungen ber reinverzweigte Erweiterungen diskret bewerteter perfekter Krper, J. Reine Angew. Math. 181 (1940), pp.144. [4] Serre, 4.3

References
Jrgen Neukirch, Algebraic Number Theory, Springer (1999). (1980), Local Fields, Berlin, New York: Springer-Verlag, ISBN9780387904245, MR0554237

Hilbert's basis theorem

132

Hilbert's basis theorem


In mathematics, Hilbert's basis theorem, states that every ideal in the ring of multivariate polynomials over a Noetherian ring is finitely generated. This can be translated into algebraic geometry as follows: every algebraic set over a field can be described as the set of common roots of finitely many polynomial equations. Hilbert(1890) proved the theorem (for the special case of polynomial rings over a field) in the course of his proof of finite generation of rings of invariants. Hilbert produced an innovative proof by contradiction using mathematical induction; his method does not give an algorithm to produce the finitely many basis polynomials for a given ideal: it only shows that they must exist. One can determine basis polynomials using the method of Grbner bases.

Proof
The following more general statement will be proved. Theorem. If is a left- (respectively right-) Noetherian ring, then the polynomial ring is also a left-

(respectively right-) Noetherian ring. It suffices to consider just the "Left" case. Proof (Theorem) Suppose per contra that were a non-finitely generated left-ideal. Then it would be that by recursion of polynomials could be found so that, letting is a non-decreasing where the are the leading of minimal degree. It is clear that sequence of naturals. Now consider the left-ideal coefficients of the comprise an . Since some term is equal to that of contradicting minimality. is left-Noetherian, we have that Now consider moreover
(Thm)

(using the countable axiom of choice) that a sequence

over

must be finitely generated; and since the will suffice. So for example, whose leading

-basis, it follows that for a finite amount of them, say

so

of degree

A constructive proof (not invoking the axiom of choice) also exists. Proof (Theorem): Let left-ideal over be a left-ideal. Let Let is As before, the say be the set of leading coefficients of members of Let are left-ideals over This is obviously a say whose degree

and so is finitely generated by the leading coefficients of finitely many members of be the set of leading coefficients of members of

and so are finitely generated by the leading coefficients of finitely with degrees Now let be the left-ideal generated

many members of

by We have Suppose per contra this were not so. Then let by Case 1:

and claim also be of minimal degree, and denote its leading coefficient so is a left-linear combination which has the same leading term as

Regardless of this condition, we have Consider of degree

of the coefficients of the moreover so

contradicting minimality.

Hilbert's basis theorem

133 Then Considering which is finitely generated.


(Thm)

Case 2: coefficients of the Case 1. Thus our claim holds, and

so is a left-linear combination

of the leading

we yield a similar contradiction as in

Note that the only reason we had to split into two cases was to ensure that the powers of were non-negative in the constructions.

multiplying the factors,

Applications
Let (i.e. a be a Ntherian commutative ring. Hilbert's basis theorem has some immediate corollaries. First, by will also be Ntherian. Second, since any affine variety over a collection of polynomials) may be written as the locus of an ideal and further as the locus of its generators, it follows that every affine variety is the were a (i.e. mod-ing out by is an ideal and thus is finitely generated. So generated by finitely many relations locus-set of induction we see that

locus of finitely many polynomials i.e. the interesection of finitely many hypersurfaces. Finally, if finitely-generated relations), where would be a -algebra, then we know that a set of polynomials. We can assume that free -algebra (on . generators)

Mizar System
The Mizar project has completely formalized and automatically checked a proof of Hilbert's basis theorem in the HILBASIS file [1].

References
Cox, Little, and O'Shea, Ideals, Varieties, and Algorithms, Springer-Verlag, 1997. Hilbert, David (1890), "Ueber die Theorie der algebraischen Formen", Mathematische Annalen 36 (4): 473534, doi:10.1007/BF01208503, ISSN0025-5831

References
[1] http:/ / www. mizar. org/ JFM/ Vol12/ hilbasis. html

Hilbert's irreducibility theorem

134

Hilbert's irreducibility theorem


In number theory, Hilbert's irreducibility theorem, conceived by David Hilbert, states that every finite number of irreducible polynomials in a finite number of variables and having rational number coefficients admit a common specialization of a proper subset of the variables to rational numbers such that all the polynomials remain irreducible. This theorem is a prominent theorem in number theory.

Formulation of the theorem


Hilbert's irreducibility theorem. Let

be irreducible polynomials in the ring

Then there exists an r-tuple of rational numbers (a1,...,ar) such that are irreducible in the ring

Remarks. It follows from the theorem that there are infinitely many r-tuples. In fact the set of all irreducible specialization, called Hilbert set, is large in many senses. For example, this set is Zariski dense in There are always (infinitely many) integer specializations, i.e., the assertion of the theorem holds even if we demand (a1,...,ar) to be integers. There are many Hilbertian fields, i.e., fields satisfying Hilbert's irreducibility theorem. For example, global fields are Hilbertian. The irreducible specialization property stated in the theorem is the most general. There are many reductions, e.g., it suffices to take in the definition. A recent result of Bary-Soroker shows that for a field K to be Hilbertian it suffices to consider the case of irreducible in the ring K [X,Y], where K
alg alg

and

absolutely irreducible, that is,

is the algebraic closure of K.

Applications
Hilbert's irreducibility theorem has numerous applications in number theory and algebra. For example: The inverse Galois problem, Hilbert's original motivation. The theorem almost immediately implies that if a finite group G can be realized as the Galois group of a Galois extension N of

then it can be specialized to a Galois extension N0 of the rational numbers with G as its Galois group. (To see this, choose a monic irreducible polynomial f(X1,,Xn,Y) whose root generates N over E. If f(a1,,an,Y) is irreducible for some ai, then a root of it will generate the asserted N0.) Construction of elliptic curves with large rank. Hilbert's irreducibility theorem is used as a step in the Andrew Wiles proof of Fermat's last theorem. If a polynomial polynomial in is a perfect square for all large integer values of x, then g(x) is the square of a . This follows from Hilbert's irreducibility theorem with . (More elementary proofs exist.) The same result is true when "square" is replaced by "cube", "fourth power", etc. and

Hilbert's irreducibility theorem

135

Generalizations
It has been reformulated and generalized extensively, by using the language of algebraic geometry. See thin set (Serre).

References
J. P. Serre, Lectures on The Mordell-Weil Theorem, Vieweg, 1989. M. D. Fried and M. Jarden, Field Arithmetic, Springer-Verlag, Berlin, 2005. H. Vlklein, Groups as Galois Groups, Cambridge University Press, 1996. G. Malle and B. H. Matzat, Inverse Galois Theory, Springer, 1999.

Hilbert's Nullstellensatz
Hilbert's Nullstellensatz (German: "theorem of zeros," or more literally, "zero-locus-theorem") is a theorem which makes precise a fundamental relationship between the geometric and algebraic sides of algebraic geometry, an important branch of mathematics. It relates algebraic sets to ideals in polynomial rings over algebraically closed fields. The theorem was first proved by David Hilbert, after whom it is named.

Formulation
Let k be a field (such as the rational numbers) and K be an algebraically closed field extension (such as the complex numbers), consider the polynomial ring k[X1,X2,..., Xn] and let I be an ideal in this ring. The affine variety V(I) defined by this ideal consists of all n-tuples x = (x1,...,xn) in Kn such that f(x) = 0 for all f in I. Hilbert's Nullstellensatz states that if p is some polynomial in k[X1,X2,..., Xn] which vanishes on the variety V(I), i.e. p(x) = 0 for all x in V(I), then there exists a natural number r such that pr is in I. An immediate corollary is the "weak Nullstellensatz": The ideal I in k[X1,X2,..., Xn] contains 1 if and only if the polynomials in I do not have any common zeros in Kn. When k=K the "weak Nullstellensatz" may also be stated as follows: if I is a proper ideal in K[X1,X2,..., Xn], then V(I) cannot be empty, i.e. there exists a common zero for all the polynomials in the ideal. This is the reason for the name of the theorem, which can be proved easily from the 'weak' form using the Rabinowitsch trick. The assumption that K be algebraically closed is essential here; the elements of the proper ideal (X2 + 1) in R[X] do not have a common zero. With the notation common in algebraic geometry, the Nullstellensatz can also be formulated as

for every ideal J. Here,

denotes the radical of J and I(U) is the ideal of all polynomials which vanish on the set

U. In this way, we obtain an order-reversing bijective correspondence between the affine varieties in Kn and the radical ideals of K[X1,X2,..., Xn]. In fact, more generally, one has a Galois connection between subsets of the space and subsets of the algebra, where "Zariski closure" and "radical of the ideal generated" are the closure operators.

Hilbert's Nullstellensatz

136

Generalization
This generalization is due to Bourbaki, and is the most general form of the Nullstellensatz. Let be a Jacobson ring. If is a finitely generated R-algebra, then is a maximal ideal of R, and is a Jacobson ring. Further, if . is a maximal ideal, then is a finite extension field of

Another generalization states that a faithfully flat morphism quasi-compact has a quasi-section, i.e. there exists an X-morphism .

locally of finite type with X

affine and faithfully flat and quasi-finite over X together with

Applications
Commuting matrices
The fact that commuting matrices have a common eigenvector and hence by induction stabilize a common flag and are simultaneously triangularizable can be interpreted as a result of the weak Nullstellensatz, as follows: commuting matrices form a commutative algebra over the matrices satisfy various polynomials such as their minimal polynomials, which form a proper ideal (because they are not all zero, in which case the result is trivial); one might call this the characteristic ideal, by analogy with the characteristic polynomial. One then defines an eigenvector for a commutative algebra as a vector v such that for all for a linear functional one has

This simply linearizes the definition of an eigenvalue, and is the correct definition for a common eigenvector, as if v is a common eigenvector, meaning then the functional is defined as

(treating scalars as multiples of the identity matrix conversely an eigenvector for such a functional Then the existence of an eigenvalue corresponds to the point in affine k-space with coordinates

, which has eigenvalue 1 for all vectors), and is a common eigenvector. Geometrically, the eigenvalue with respect to the basis given by being

is equivalent to the ideal generated by (the relations satisfied by)

non-empty, which exactly generalizes the usual proof of existence of an eigenvalue existing for a single matrix over an algebraically closed field by showing that the characteristic polynomial has a zero.

References
Shigeru Mukai; William Oxbury (translator) (2003). An Introduction to Invariants and Moduli. Cambridge studies in advanced mathematics. 81. p.82. ISBN0-521-80906-1. David Eisenbud, Commutative Algebra With a View Toward Algebraic Geometry, New York : Springer-Verlag, 1999.

Hilbert's syzygy theorem

137

Hilbert's syzygy theorem


In mathematics, Hilbert's syzygy theorem is a result of commutative algebra, first proved by David Hilbert (1890) in connection with the syzygy (relation) problem of invariant theory. Roughly speaking, starting with relations between polynomial invariants, then relations between the relations, and so on, it explains how far one has to go to reach a clarified situation. It is now considered to be an early result of homological algebra, and through the depth concept, to be a measure of the non-singularity of affine space.

Formal statement
A contemporary formal statement is the following. Let k be a field and M a finitely generated module over the polynomial ring

Hilbert's syzygy theorem then states that there exists a free resolution of M of length at most n.

References
David Eisenbud, Commutative algebra. With a view toward algebraic geometry. Graduate Texts in Mathematics, 150. Springer-Verlag, New York, 1995. xvi+785 pp. ISBN 0-387-94268-8; ISBN 0-387-94269-6 MR1322960

Hilbert's Theorem 90
In abstract algebra, Hilbert's Theorem 90 (or Satz 90) refers to an important result on cyclic extensions of fields (or to one of its generalizations) that leads to Kummer theory. In its most basic form, it tells us that if L/K is a cyclic extension of fields with Galois group G =Gal(L/K) generated by an element s and if a is an element of L of relative norm 1, then there exists b in L such that a = s(b)/b. The theorem takes its name from the fact that it is the 90th theorem in David Hilbert's famous Zahlbericht of 1897, although it is originally due to Kummer. Often a more general theorem due to Emmy Noether is given the name, stating that if L/K is a finite Galois extension of fields with Galois group G =Gal(L/K), then the first cohomology group is trivial: H1(G, L) = {1}

Examples
Let L/K be the quadratic extension conjugation: An element in L has norm
2 2

. The Galois group is cyclic of order 2, its generator s is acting via

. An element of norm one corresponds to a rational

solution of the equation a +b =1 or in other words, a point with rational coordinates on the unit circle. Hilbert's Theorem 90 then states that every element y of norm one can be parametrized (with integral c,d) as

which may be viewed as a rational parametrization of the rational points on the unit circle. Rational points on the unit circle correspond to Pythagorean triples, i.e. triples of

Hilbert's Theorem 90 integers satisfying .

138

Cohomology
The theorem can be stated in terms of group cohomology: if L is the multiplicative group of any (not necessarily finite) Galois extension L of a field K with corresponding Galois group G, then H1(G, L) = {1}. A further generalization using non-abelian group cohomology states that if H is either the general or special linear group over L, then H1(G,H) = {1}. This is a generalization since L = GL1(L). Another generalization is K-theory plays a role in Voevodsky's proof of the Milnor conjecture. for X a scheme, and another one to Milnor

References
Chapter II of J.S. Milne, Class Field Theory, available at his website [1]. Neukirch, Jrgen; Schmidt, Alexander; Wingberg, Kay (2000), Cohomology of Number Fields, Grundlehren der Mathematischen Wissenschaften, 323, Berlin: Springer-Verlag, ISBN978-3-540-66671-4, MR1737196

References
[1] http:/ / www. jmilne. org/ math

HopkinsLevitzki theorem
In the branch of abstract algebra called ring theory, the Akizuki-HopkinsLevitzki theorem connects the descending chain condition and ascending chain condition in modules over semiprimary rings. A ring R is called semiprimary if R/J(R) is semisimple and J(R) is a nilpotent ideal, where J(R) denotes the Jacobson radical. The theorem states that if R is a semiprimary ring and M is an R module, the three module conditions Noetherian, Artinian and "has a composition series" are equivalent. Without the semiprimary condition, the only true implication is that if M has a composition series, then M is both Noetherian and Artinian. The theorem takes its current form from a paper by Charles Hopkins and a paper by Jacob Levitzki, both in 1939. For this reason it is often cited as the HopkinsLevitzki theorem. However Yasuo Akizuki is sometimes included since he proved the result for commutative rings a few years earlier (Lam 2001). Since it is known that right Artinian rings are semiprimary, a direct corollary of the theorem is: a right Artinian ring is also right Noetherian. The analogous statement for left Artinian rings holds as well. This is not true in general for Artinian modules, because there are examples of Artinian modules which are not Noetherian. Another direct corollary is that if R is right Artinian, then R is left Artinian if and only if it is left Noetherian.

HopkinsLevitzki theorem

139

References
Charles Hopkins (1939) Rings with minimal condition for left ideals, Ann. of Math. (2) 40, pages 712730. T. Y. Lam (2001) A first course in noncommutative rings, Springer-Verlag. page 55 ISBN 0-387-95183-0 Jakob Levitzki (1939) On rings which satisfy the minimum condition for the right-hand ideals, Compositio Math. 7, pages 214222.

Hurwitz's theorem (normed division algebras)


In algebra, Hurwitz's theorem (also called the 1,2,4 8 Theorem), named after Adolf Hurwitz, who proved it in 1898, states: Every normed division algebra with an identity is isomorphic to one of the following four algebras: R, C, H and O, that is the real numbers, the complex numbers, the quaternions and the octonions.[1] [2] The classification of real division algebras began with Georg Frobenius,[3] continued with Hurwitz[4] and was set in general form by Max Zorn.[5] A brief historical summary may be found in Badger.[6] A full proof can be found in Kantor and Solodovnikov,[7] and in Shapiro.[8] As a basic idea, if an algebra A is proportional to 1 then it is isomorphic to the real numbers. Otherwise we extend the subalgebra isomorphic to 1 using the CayleyDickson construction and introducing a vector e which is orthogonal to 1. This subalgebra is isomorphic to the complex numbers. If this is not all of A then we once again use the CayleyDickson construction and another vector orthogonal to the complex numbers and get a subalgebra isomorphic to the quaternions. If this is not all of A then we double up once again and get a subalgebra isomorphic to the Cayley numbers (or Octonions). We now have a theorem which says that every subalgebra of A that contains 1 and is not A is associative. The Cayley numbers are not associative and therefore must be A. Hurwitz's theorem can be used to prove that the product of the sum of n squares by the sum of n squares is the sum of n squares in a bilinear way only when n is equal to 1, 2, 4 and 8.[9]

In-line references
[1] JA Nieto and LN Alejo-Armenta (2000). "Hurwitz theorem and parallelizable spheres from tensor analysis". Arxiv preprint hep-th/0005184. arXiv:hep-th/0005184. [2] Kevin McCrimmon (2004). "Hurwitz's theorem 2.6.2" (http:/ / books. google. com/ books?id=6YG4ycpKMYkC& pg=PA166). A taste of Jordan algebras. Springer. p.166. ISBN0387954473. . "Only recently was it established that the only finite-dimensional real nonassociative division algebras have dimensions 1,2,4,8; the algebras were not classified, and the proof was topological rather than algebraic." [3] Georg Frobenius (1878). "ber lineare Substitutionen und bilineare Formen". J. Reine Angew. Math. 84: 163. [4] Hurwitz, A. (1898). "Ueber die Composition der quadratischen Formen von beliebig vielen Variabeln (On the composition of quadratic forms of arbitrary many variables)" (in German). Nachr. Ges. Wiss. Gttingen: 309316. JFM29.0177.01. [5] Max Zorn (1930). "Theorie der alternativen Ringe". Abh. Math. Sem. Univ. Hamburg 8: 123147. [6] Matthew Badger. "Division algebras over the real numbers" (http:/ / www. math. washington. edu/ ~mbadger/ divalg3. pdf). . [7] IL Kantor and AS Solodovnikov (1989). "Normed algebras with an identity. Hurwitz's theorem." (http:/ / books. google. com/ books?as_q=& num=10& btnG=Google+ Search& as_epq=Normed+ algebras+ with+ an+ identity. + Hurwitz's+ theorem& as_oq=& as_eq=& as_brr=0& as_pt=ALLTYPES& lr=& as_vt=& as_auth=& as_pub=& as_sub=& as_drrb_is=q& as_minm_is=0& as_miny_is=& as_maxm_is=0& as_maxy_is=& as_isbn=& as_issn=). Hypercomplex numbers. An elementary introduction to algebras (2nd ed.). Springer-Verlag. p.121. ISBN0387969802. . [8] Daniel B. Shapiro (2000). "Appendix to Chapter 1. Composition algebras" (http:/ / books. google. com/ books?id=qrFhUda9JbkC& pg=PA21). Compositions of quadratic forms. Walter de Gruyter. pp.21 ff. ISBN311012629X. . [9] Joe Roberts (1992). "Square identities" (http:/ / books. google. com/ books?id=DvX90EKMxGwC& pg=PA93). Lure of the integers. Cambridge University Press. ISBN088385502X. .

Hurwitz's theorem (normed division algebras)

140

Background references
John H. Conway, Derek A. Smith On Quaternions and Octonions. A.K. Peters, 2003. John Baez, The Octonions (http://math.ucr.edu/home/baez/octonions/), AMS 2001.

Isomorphism extension theorem


In field theory, a branch of mathematics, the isomorphism extension theorem is an important theorem regarding the extension of a field isomorphism to a larger field.

Isomorphism extension theorem


The theorem states that given any field onto a field then , an algebraic extension field mapping of and an isomorphism mapping of can be extended to an isomorphism onto an algebraic extension

(a subfield of the algebraic closure of ). The proof of the isomorphism extension theorem depends on Zorn's lemma.

References
D.J. Lewis, Introduction to algebra, Harper & Row, 1965, Chap.IV.12, p.193.

Isomorphism theorem
In mathematics, specifically abstract algebra, the isomorphism theorems are three theorems that describe the relationship between quotients, homomorphisms, and subobjects. Versions of the theorems exist for groups, rings, vector spaces, modules, Lie algebras, and various other algebraic structures. In universal algebra, the isomorphism theorems can be generalized to the context of algebras and congruences.

History
The isomorphism theorems were formulated in some generality for homomorphisms of modules by Emmy Noether in her paper Abstrakter Aufbau der Idealtheorie in algebraischen Zahl- und Funktionenkrpern which was published in 1927 in Mathematische Annalen. Less general versions of these theorems can be found in work of Richard Dedekind and previous papers by Noether. Three years later, B.L. van der Waerden published his influential Algebra, the first abstract algebra textbook that took the now-traditional groups-rings-fields approach to the subject. Van der Waerden credited lectures by Noether on group theory and Emil Artin on algebra, as well as a seminar conducted by Artin, Wilhelm Blaschke, Otto Schreier, and van der Waerden himself on ideals as the main references. The three isomorphism theorems, called homomorphism theorem, and two laws of isomorphism when applied to groups, appear explicitly.

Isomorphism theorem

141

Groups
We first state the three isomorphism theorems in the context of groups. Note that some sources switch the numbering of the second and third theorems.[1] Sometimes, the lattice theorem is referred to as the fourth isomorphism theorem or the correspondence theorem.

Statement of the theorems


First isomorphism theorem Let G and H be groups, and let :GH be a homomorphism. Then: 1. The kernel of is a normal subgroup of G, 2. The image of is a subgroup of H, and 3. The image of is isomorphic to the quotient group G/ker(). In particular, if is surjective then H is isomorphic to G/ker(). Second isomorphism theorem Let G be a group. Let S be a subgroup of G, and let N be a normal subgroup of G. Then: 1. The product SN is a subgroup of G, 2. The intersection SN is a normal subgroup of S, and 3. The quotient groups (SN)/N and S/(SN) are isomorphic. Technically, it is not necessary for N to be a normal subgroup, as long as S is a subgroup of the normalizer of N. In this case, the intersection SN is not a normal subgroup of G, but it is still a normal subgroup of S. Third isomorphism theorem Let G be a group. Let N and K be normal subgroups of G, with KNG. Then 1. The quotient N/K is a normal subgroup of the quotient G/K, and 2. The quotient group (G/K)/(N/K) is isomorphic to G/N.

Discussion First isomorphism theorem

The first isomorphism theorem follows from the category theoretical fact that the category of groups is (normal epi, mono)-factorizable; in other words, the normal epimorphisms and the monomorphisms form a factorization system for the category. This is captured in the commutative diagram in the margin, which shows the objects and morphisms whose existence can be deduced from the morphism f: GH. The diagram shows that every morphism in the category of groups has a kernel in the category theoretical sense; the arbitrary morphism f factors into , where is a monomorphism and is an epimorphism (in a conormal category, all epimorphisms are normal). This is

Isomorphism theorem represented in the diagram by an object and a monomorphism

142

(kernels are always monomorphisms), which

the short exact sequence running from the lower left to the upper right of the diagram. The use of the exact sequence convention saves us from having to draw the zero morphisms from to H and . If the sequence is right split (i. e., there is a morphism that maps the semidirect product of the normal subgroup some such that and the subgroup to a -preimage of itself), then G is . If it is left split (i. e., there exists is a direct

), then it must also be right split, and

product decomposition of G. In general, the existence of a right split does not imply the existence of a left split; but in an abelian category (such as the abelian groups), left splits and right splits are equivalent by the splitting lemma, and a right split is sufficient to produce a direct sum decomposition . In an abelian category, all monomorphisms are also normal, and the diagram may be extended by a second short exact sequence . In the second isomorphism theorem, the product SN is the join of S and N in the lattice of subgroups of G, while the intersection SN is the meet. The third isomorphism theorem is generalized by the nine lemma to abelian categories and more general maps between objects. It is sometimes informally called the "freshman theorem", because "even a freshman could figure it out: just cancel out the Ks!"

Rings
The statements of the theorems for rings are similar, with the notion of a normal subgroup replaced by the notion of an ideal.

First isomorphism theorem


Let R and S be rings, and let :RS be a ring homomorphism. Then: 1. The kernel of is an ideal of R, 2. The image of is a subring of S, and 3. The image of is isomorphic to the quotient ring R/ker(). In particular, if is surjective then S is isomorphic to R/ker().

Second isomorphism theorem


Let R be a ring. Let S be a subring of R, and let I be an ideal of R. Then: 1. The sum S+I={s+i|sS,iI} is a subring of R, 2. The intersection SI is an ideal of S, and 3. The quotient rings (S+I)/I and S/(SI) are isomorphic.

Third isomorphism theorem


Let R be a ring. Let A and B be ideals of R, with BAR. Then 1. The set A/B is an ideal of the quotient R/B, and 2. The quotient ring (R/B)/(A/B) is isomorphic to R/A.

Isomorphism theorem

143

Modules
The statements of the isomorphism theorems for modules are particularly simple, since it is possible to form a quotient module from any submodule. The isomorphism theorems for vector spaces and abelian groups are special cases of these. For vector spaces, all of these theorems follow from the rank-nullity theorem. For all of the following theorems, the word module will mean R-module, where R is some fixed ring.

First isomorphism theorem


Let M and N be modules, and let :MN be a homomorphism. Then: 1. The kernel of is a submodule of M, 2. The image of is a submodule of N, and 3. The image of is isomorphic to the quotient module M/ker(). In particular, if is surjective then N is isomorphic to M/ker().

Second isomorphism theorem


Let M be a module, and let S and T be submodules of M. Then: 1. The sum S+T={s+t|sS,tT} is a submodule of M, 2. The intersection ST is a submodule of S, and 3. The quotient modules (S+T)/T and S/(ST) are isomorphic.

Third isomorphism theorem


Let M be a module. Let S and T be submodules of M, with TSM. Then 1. The quotient S/T is a submodule of the quotient M/T, and 2. The quotient (M/T)/(S/T) is isomorphic to M/S.

General
To generalise this to universal algebra, normal subgroups need to be replaced by congruences. Briefly, if is an algebra, a congruence on is an equivalence relation on which is a subalgebra when considered as a subset of equivalence classes well-defined since (the latter with the coordinate-wise operation structure). One can make the set of into an algebra of the same type by defining the operations via representatives; this will be is a subalgebra of .

First Isomorphism Theorem


If and are algebras, and if and only if , which is a subalgebra of . is a homomorphism is a congruence on , then the equivalence relation , and the algebra on defined by image of is isomorphic to the

Second Isomorphism Theorem


Given an algebra , a subalgebra of , and a congruence . Then . on , we let , and we let , be the subset of be the intersection of is a congruence on determined by all congruence classes that contain an element of (considered as a subset of , and the algebra ) with is isomorphic to the algebra

is a subalgebra of

Isomorphism theorem

144

Third Isomorphism Theorem


Let (where be an algebra, and let on represents the and be two congruence relations on defined by ), and if and only if is isomorphic to , with and contained in . . Then determines a congruence are equivalent modulo

-equivalence class of

Notes
[1] Jacobson (2009), p. 101, use "first" for the isomorphism of the modules (S+T)/T and S/(ST), and "second" for (M/T)/(S/T) and M/S.

References
Emmy Noether, Abstrakter Aufbau der Idealtheorie in algebraischen Zahl- und Funktionenkrpern, Mathematische Annalen 96 (1927) p.26-61 Colin McLarty, 'Emmy Noethers Set Theoretic Topology: From Dedekind to the rise of functors' in The Architecture of Modern Mathematics: Essays in history and philosophy (edited by Jeremy Gray and Jos Ferreirs), Oxford University Press (2006) p.21135. Jacobson, Nathan (2009), Basic algebra, 2 (2nd ed.), Dover, ISBN978-0-486-47187-7

External links
First isomorphism theorem (http://planetmath.org/?op=getobj&amp;from=objects&amp;id=1114) on PlanetMath. Proof of first isomorphism theorem (http://planetmath.org/?op=getobj&amp;from=objects& amp;id=2922) on PlanetMath Second isomorphism theorem (http://planetmath.org/?op=getobj&amp;from=objects&amp;id=1334) on PlanetMath. Proof of second isomorphism theorem (http://planetmath.org/?op=getobj&amp;from=objects& amp;id=3153) on PlanetMath Third isomorphism theorem (http://planetmath.org/?op=getobj&amp;from=objects&amp;id=1126) on PlanetMath. Proof of third isomorphism theorem (http://planetmath.org/?op=getobj&amp;from=objects& amp;id=7496) on PlanetMath

Jacobson density theorem

145

Jacobson density theorem


In mathematics, more specifically non-commutative ring theory, modern algebra, and module theory, the Jacobson density theorem is a theorem concerning simple modules over a ring R.[1] The theorem can be applied to show that any primitive ring can be viewed as a "dense" subring of the ring of linear transformations of a vector space[2] [3] . This theorem first appeared in the literature in 1945, in the famous paper "Structure Theory of Simple Rings Without Finiteness Assumptions" by Nathan Jacobson.[4] This can be viewed as a kind of generalization of the Artin-Wedderburn theorem's conclusion about the structure of simple Artinian rings.

Motivation and formal statement


Let R be a ring and let U be a simple right R-module. If u is a non-zero element of U, uR = U (where uR is the cyclic submodule of U generated by u). Therefore, if u and v are non-zero elements of U, there is an element of R that induces an endomorphism of U transforming u to v. The natural question now is whether this can be generalized to arbitrary (finite) tuples of elements. More precisely, find necessary and sufficient conditions on the tuple (x1, ..., xn) and (y1, ..., yn) separately, so that there is an element of R with the property that xir = yi for all i. If D is the set of all R-module endomorphisms of U, then Schur's lemma asserts that D is a division ring, and the Jacobson density theorem answers the question on tuples in the affirmative, provided that the x's are linearly independent over D. With the above in mind, theorem may be stated this way: The Jacobson Density Theorem Let U be a simple right R-module and write D = EndR(U). Let A be any D-linear transformation on U and let X be a finite D-linearly independent subset of U. Then there exists an element r of R such that A(x) = xr for all x in X.[5]

Proof
In the Jacobson density theorem, the right R-module U is simultaneously viewed as a left D-module where D=EndR(U) module in the natural way: the action gu is defined to be g(u). It can be verified that this is indeed a left module structure on U.[6] As noted before, Schur's lemma proves D is a division ring if U is simple, and so U is a vector space over D. The proof also relies on the following theorem proven in (Isaacs 1993) p. 185: Theorem Let U be a simple right R-module and let D = EndR(U) - the set of all R module endomorphisms of U. Let X be a finite subset of U and write I = annR(X) - the annihilator of X in R. Let u be in U with uI = 0. Then u is in XD; the D-span of X. Proof (of the Jacobson density theorem) We proceed by mathematical induction on the number n of elements in X. If n=0 so that X is empty, then the theorem is vacuously true and the base case for induction is verified. Now we assume that X is non-empty with cardinality n. Let x be an element of X and write Y = X \ {x}. If A is any D-linear transformation on U, the induction hypothesis guarantees that there exists an s in R such that A(y) = ys for all y in Y. Write I = annR(Y). It is easily seen that xI is a submodule of U. If it were the case that xI = 0, then the previous theorem would indicate that x would be in the D-span of Y. This would contradict the linear independence of X, so it must be that xI 0. So, by simplicity of U, the submodule xI = U. Since A(x) - xs is in U=xI, there exists i in I such that xi = A(x) - xs.

Jacobson density theorem After defining r = s + i, we compute that yr = y(s + i) = ys + yi = ys = A(y) for all y in Y.[7] Also, xr = x(s + i) = xs + A(x) - xs = A(x). Therefore, A(z) = zr for all z in X, as desired. This completes the inductive step of the proof. It follows now from mathematical induction that the theorem is true for finite sets X of any size.

146

Topological characterization
A ring R is said to act densely on a simple right R-module U if it satisfies the conclusion of the Jacobson density theorem.[8] There is a topological reason for describing R as "dense". Firstly, R can be identified with a subring of End(DU) by identifying each element of R with the D linear transformation it induces by right multiplication. If U is given the discrete topology, and if UU is given the product topology, and End(DU) is viewed as a subspace of UU and is given the subspace topology, then R acts densely on U if and only if R is dense set in End(DU) with this topology[9] .

Consequences
The Jacobson density theorem has various important consequences in the structure theory of rings.[10] Notably, the ArtinWedderburn theorem's conclusion about the structure of simple right Artinian rings is recovered. The Jacobson density theorem also characterizes right or left primitive rings as dense subrings of the ring of D-linear transformations on some D- vector space U, where D is a division ring.[11]

Relations to other results


This result is related to the Von Neumann bicommutant theorem, which states that, for a *-algebra A of operators on a Hilbert space H, the double commutant A can be approximated by A on any given finite set of vectors. See also the Kaplansky density theorem in the von Neumann algebra setting.

Notes
[1] Isaacs, p. 184 [2] Such rings of linear transformations are also known as full linear rings. [3] Isaacs, Corollary 13.16, p. 187 [4] Jacobson, Nathan "Structure Theory of Simple Rings Without Finiteness Assumptions" (http:/ / www. jstor. org/ pss/ 1990204) [5] Isaacs, Theorem 13.14, p. 185 [6] Incidentally it is also a D-R bimodule structure. [7] Of course, yi=0 by definition of I. [8] Herstein, Definition, p. 40 [9] It turns out this topology is the same as the compact-open topology in this case. Herstein, p. 41 uses this description. [10] Herstein, p. 41 [11] Isaacs, Corollary 13.16, p. 187

References
I.N. Herstein (1968). Noncommutative rings (1st edition ed.). The Mathematical Association of America. ISBN0-88385-015-X. I. Martin Isaacs (1993). Algebra, a graduate course (1st edition ed.). Brooks/Cole Publishing Company. ISBN0-534-19002-2. Jacobson, N. (1945), "Structure theory of simple rings without finiteness assumptions", Trans. Amer. Math. Soc. 57: 228245, ISSN0002-9947, MR0011680

Jacobson density theorem

147

External links
PlanetMath page (http://planetmath.org/encyclopedia/JacobsonDensityTheorem.html)

Jordan's theorem (symmetric group)


Jordan's theorem is a statement in finite group theory. It states that if a primitive permutation group GSn contains a p-cycle for a prime p<n-2, then G is either the whole symmetric group Sn or the alternating group An. The statement can be generalized for p being a prime power.

References
Griess, Robert L. (1998), Twelve sporadic groups, Springer, p.5, ISBN978-3540627784 Isaacs, I. Martin (2008), Finite group theory, AMS, p.245, ISBN978-0821843444 Neuman, Peter M. (1975), "Primitive permutation groups containing a cycle of prime power length" [1], Bulletin London Mathematical Society 7 (3): 298299

External links
Jordan's Symmetric Group Theorem [2], mathworld

References
[1] http:/ / blms. oxfordjournals. org/ content/ 7/ 3/ 298. extract [2] http:/ / mathworld. wolfram. com/ JordansSymmetricGroupTheorem. html

JordanSchur theorem

148

JordanSchur theorem
In mathematics, the JordanSchur theorem also known as Jordan's theorem on finite linear groups is a theorem in its original form due to Camille Jordan. In that form, it states that there is a function (n) such that given a finite group G that is a subgroup of the group of n-by-n complex matrices, then there is a subgroup H of G such that H is abelian, H is normal with respect to G and H has index at most(n). Schur proved a more general result that applies when G is assumed not to be finite but just periodic. Schur showed that (n) may be taken to be ((8n)1/2+1)2n2((8n)1/21)2n2.[1] A tighter bound (for n3) is due to Speiser who showed that as long as G is finite, one can take (n)=n!12n((n+1)+1) where (n) is the prime-counting function.[1] [2] This was subsequently improved by Blitchfeldt who replaced the "12" with a "6". Unpublished work on the finite case was also done by Boris Weisfeiler.[3] Subsequently, Michael Collins using the classification of finite simple groups showed that in the finite case, one can take f(n))=(n+1)! when n is at least 71, and gave near complete descriptions of the behavior for smaller n.

References
[1] Curtis, Charles; Reiner, Irving (1962). Representation Theory of Finite Groups and Associated Algebras. John Wiley & Sons. pp.258262. [2] Speiser, Andreas (1945). Die Theorie der Gruppen von endlicher Ordnung, mit Andwendungen auf algebraische Zahlen und Gleichungen sowie auf die Krystallographie, von Andreas Speiser. New York: Dover Publications. pp.216220. [3] Collins, Michael J. (2007). "On Jordans theorem for complex linear groups". Journal of Group Theory 10 (4): 411423. doi:10.1515/JGT.2007.032.

Ben Green lecture notes Analytic Topics in Group Theory, chapter 2 (http://www.dpmms.cam.ac.uk/~bjg23/ ATG.html)http://www.dpmms.cam.ac.uk/~bjg23/ATG/Chapter2.pdf

Krull's principal ideal theorem

149

Krull's principal ideal theorem


In commutative algebra, Krull's principal ideal theorem, named after Wolfgang Krull (18991971), gives a bound on the height of a principal ideal in a Noetherian ring. The theorem is sometimes referred to by its German name, Krulls Hauptidealsatz. Formally, if R is a Noetherian ring and I is a principal, proper ideal of R, then I has height at most one. This theorem can be generalized to ideals that are not principal, and the result is often called Krull's height theorem. This says that if R is a Noetherian ring and I is a proper ideal generated by n elements of R, then I has height at most n.

References
Matsumura, Hideyuki (1970), Commutative algebra, New York: Benjamin, see in particular section (12.I), p. 77

KrullSchmidt theorem
In mathematics, the KrullSchmidt theorem states that a group subjected to certain finiteness conditions on chains of subgroups, can be uniquely written as a finite direct product of indecomposable subgroups.

Definitions
We say that a group G satisfies the ascending chain condition (ACC) on subgroups if every sequence of subgroups of G:

is eventually constant, i.e., there exists N such that GN = GN+1 = GN+2=.... We say that G satisfies the ACC on normal subgroups if every such sequence of normal subgroups of G eventually becomes constant. Likewise, one can define the descending chain condition on (normal) subgroups, by looking at all decreasing sequences of (normal) subgroups:

Clearly, all finite groups satisfy both ACC and DCC on subgroups. The infinite cyclic group
2 3

satisfies ACC but

not DCC, since (2)>(2) >(2) >... is an infinite decreasing sequence of subgroups. On the other hand, the -torsion part of (the quasicyclic p-group) satisfies DCC but not ACC. We say a group G is indecomposable if it cannot be written as a direct product of non-trivial subgroups G = HK.

KrullSchmidt theorem
The theorem says: If is a group that satisfies ACC and DCC on normal subgroups, then there is a unique way of writing of finitely many indecomposable subgroups of as a direct product . Here, uniqueness

means direct decompositions into indecomposable subgroups have the exchange property. That is: suppose is another expression of as a product of indecomposable subgroups. Then and there is a reindexing of the and are isomorphic for each ; for each . 's satisfying

KrullSchmidt theorem

150

KrullSchmidt theorem for modules


If is a module that satisfies the ACC and DCC on submodules (that is, it is both Noetherian and Artinian or is a direct sum of indecomposable modules. Up to a permutation, the equivalently of finite length), then

indecomposable components in such a direct sum are uniquely determined up to isomorphism.[1] In general, the theorem fails, if one only assumes that the module is Noetherian.

History
The present-day KrullSchmidt theorem was first proved by Joseph Wedderburn (Ann. of Math (1909)), for finite groups, though he mentions some credit is due to an earlier study of G.A. Miller where direct products of abelian groups were considered. Wedderburn's theorem is stated as an exchange property between direct decompositions of maximum length. However, Wedderburn's proof makes no use of automorphisms. The thesis of Robert Remak (1911) derived the same uniqueness result as Wedderburn but also proved (in modern terminology) that the group of central automorphisms acts transitively on the set of direct decompositions of maximum length of a finite group. From that stronger theorem Remak also proved various corollaries including that groups with a trivial center and perfect groups have a unique Remak decomposition. Otto Schmidt (Sur les produits directs, S. M. F. Bull. 41 (1913), 161164), simplified the main theorems of Remak to the 3 page predecessor to today's textbook proofs. His method improves Remak's use of idempotents to create the appropriate central automorphisms. Both Remak and Schmidt published subsequent proofs and corollaries to their theorems. Wolfgang Krull (ber verallgemeinerte endliche Abelsche Gruppen, M. Z. 23 (1925) 161196), returned to G.A. Miller's original problem of direct products of abelian groups by extending to abelian operator groups with ascending and descending chain conditions. This is most often stated in the language of modules. His proof observes that the idempotents used in the proofs of Remak and Schmidt can be restricted to module homomorphisms; the remaining details of the proof are largely unchanged. O. Ore unified the proofs from various categories include finite groups, abelian operator groups, rings and algebras by proving the exchange theorem of Wedderburn holds for modular lattices with descending and ascending chain conditions. This proof makes no use of idempotents and does not reprove the transitivity of Remak's theorems. Kurosh's The Theory of Groups and Zassenhaus' The Theory of Groups include the proofs of Schmidt and Ore under the name of RemakSchmidt but acknowledge Wedderburn and Ore. Later texts use the title KrullSchmidt (Hungerford's Algebra) and KrullSchmidtAzumaya (CurtisReiner). The name KrullSchmidt is now popularly substituted for any theorem concerning uniqueness of direct products of maximum size. Some authors choose to call direct decompositions of maximum-size Remak decompositions to honor his contributions.

KrullSchmidt theorem

151

Notes
[1] Jacobson (2009), p. 115.

References
Jacobson, Nathan (2009), Basic algebra, 2 (2nd ed.), Dover, ISBN978-0-486-47187-7

Further reading
Hungerford, Thomas W. Algebra, Graduate Texts in Mathematics Volume 73. ISBN 0-387-90518-9 A. Facchini: Module theory. Endomorphism rings and direct sum decompositions in some classes of modules. Progress in Mathematics, 167. Birkhuser Verlag, Basel, 1998. ISBN 3-7643-5908-0 A. Facchini, D. Herbera, L.S. Levy, P. Vmos: KrullSchmidt fails for Artinian modules. Proc. Amer. Math. Soc. 123 (1995), no. 12, 35873592. C.M. Ringel: KrullRemakSchmidt fails for Artinian modules over local rings. Algebr. Represent. Theory 4 (2001), no. 1, 7786.

External links
Page at PlanetMath (http://planetmath.org/encyclopedia/KrullRemakSchmidtTheorem.html)

Knneth theorem
In mathematics, especially in homological algebra and algebraic topology, a Knneth theorem is a statement relating the homology of two objects to the homology of their product. The classical statement of the Knneth theorem relates the singular homology of two topological spaces X and Y and their product space X Y. In the simplest possible case the relationship is that of a tensor product, but for applications it is very often necessary to apply certain tools of homological algebra to express the answer. A Knneth theorem or Knneth formula is true in many different homology and cohomology theories, and the name has become generic. These many results are named for the German mathematician Hermann Knneth.

Singular homology with coefficients in a field


Let X and Y be two topological spaces. In general one uses singular homology; but if X and Y happen to be CW complexes, then this can be replaced by cellular homology, because that is isomorphic to singular homology. The simplest case is when the coefficient ring for homology is a field F. In this situation, the Knneth theorem (for singular homology) states that for any integer k,

Furthermore, the isomorphism is a natural isomorphism. The map from the sum to the homology group of the product is called the cross product. More precisely, there is a cross product operation by which an i-cycle on X and a j-cycle on Y can be combined to create an (i+j)-cycle on X Y; so that there is an explicit linear mapping defined from the direct sum to Hk(X Y). A consequence of this result is that the Betti numbers, the dimensions of the homology with Q coefficients, of X Y can be determined from those of X and Y. If pZ(t) is the generating function of the sequence of Betti numbers bk(Z) of a space Z, then

Knneth theorem Here when there are finitely many Betti numbers of X and Y, each of which is a natural number rather than , this reads as an identity on Poincar polynomials. In the general case these are formal power series with possibly infinite coefficients, and have to be interpreted accordingly. Furthermore, the above statement holds not only for the Betti numbers but also for the generating functions of the dimensions of the homology over any field. (If the integer homology is not torsion-free then these numbers may differ from the standard Betti numbers.)

152

Singular homology with coefficients in a PID


The above formula is simple because vector spaces over a field have very restricted behavior. As the coefficient ring becomes more general, the relationship becomes more complicated. The next simplest case is the case when the coefficient ring is a principal ideal domain. This case is particularly important because the integers are a PID. In this case the equation above is no longer always true. A correction factor appears to account for the possibility of torsion phenomena. This correction factor is expressed in terms of the Tor functor, the first derived functor of the tensor product. When R is a PID, then the correct statement of the Knneth theorem is that for any topological spaces X and Y there are natural short exact sequences

Furthermore these sequences split, but not canonically.

Example
The short exact sequences just described can easily be used to compute the homology groups with integer coefficients of the product RP2 x RP2 of two real projective planes. These spaces are CW complexes. Denoting the homology group by hi for brevity's sake, one knows from a simple calculation with cellular homology that and hi is zero for all other values of i. The only non-zero Tor group (torsion product) which can be formed from these values of hi is Therefore the Knneth short exact sequence reduces in every degree to an isomorphism, because there is a zero group in each case on either the left or the right side in the sequence. The result is

and all the other homology groups are zero.

Knneth theorem

153

The Knneth spectral sequence


For a general commutative ring R, the homology of X and Y is related to the homology of their product by a Knneth spectral sequence

In the cases described above, this spectral sequence collapses to give an isomorphism or a short exact sequence.

Relation with homological algebra, and idea of proof


The chain complex of the space X Y is related to the chain complexes of X and Y by a natural quasi-isomorphism

For singular chains this is the theorem of Eilenberg and Zilber. For cellular chains on CW complexes, it is a straightforward isomorphism. Then the homology of the tensor product on the right is given by the spectral Knneth formula of homological algebra.[1] The freeness of the chain modules means that in this geometric case it is not necessary to use any hyperhomology or total derived tensor product. There are analogues of the above statements for singular cohomology and sheaf cohomology. For sheaf cohomology on an algebraic variety, Grothendieck found six spectral sequences relating the possible hyperhomology groups of two chain complexes of sheaves and the hyperhomology groups of their tensor product.[2]

Knneth theorems in generalized homology and cohomology theories


There are many generalized or extraordinary homology and cohomology theories for topological spaces. K-theory and cobordism are the best-known. Their striking common feature (not their definition) is that they do not arise from ordinary chain complexes. Thus Knneth theorems can not be obtained by the above methods of homological algebra. Nevertheless Knneth theorems in just the same form have been proved in very many cases by various other methods. The first were Atiyah's Knneth theorem for complex K-theory and Conner and Floyd's result in cobordism.[3] [4] A general method of proof emerged, based upon a homotopical theory of modules over highly structured ring spectra.[5] The homotopy category of such modules closely resembles the derived categories of homological algebra.

References
[1] See final chapter of Mac Lane, Saunders (1963), Homology, Berlin: Springer, ISBN038703823X [2] Grothendieck, Alexander; Dieudonn, Jean (1963), "lments de gomtrie algbrique (rdigs avec la collaboration de Jean Dieudonn) : III. tude cohomologique des faisceaux cohrents, Seconde partie" (http:/ / www. numdam. org:80/ numdam-bin/ feuilleter?id=PMIHES_1963__17_), Publications Mathmatiques de l'IHS 17: 591, (EGA III2, Thorme 6.7.3.). [3] Atiyah, Michael F. (1967), K-theory, New York: W. A. Benjamin. [4] Conner, P. E.; Floyd, E. E. (1964), Differentiable periodic maps, Berlin: Springer. [5] Elmendorf, A. D.; Kriz, I.; Mandell, M. A. & May, J. P. (1997), Rings, modules and algebras in stable homotopy theory, Providence, RI: American Mathematical Society, ISBN0821806386.

External links
Hazewinkel, Michiel, ed. (2001), "Knneth formula" (http://eom.springer.de/k/k056010.htm), Encyclopaedia of Mathematics, Springer, ISBN978-1556080104

Kurosh subgroup theorem

154

Kurosh subgroup theorem


In the mathematical field of group theory, the Kurosh subgroup theorem describes the algebraic structure of subgroups of free products of groups. The theorem was obtained by Alexander Kurosh, a Russian mathematician, in 1934.[1] Informally, the theorem says that every subgroup of a free product is itself a free product of a free group and of groups conjugate to the subgroups of the factors of the original free product.

History and generalizations


After the original 1934 proof of Kurosh, there were many subsequent proofs of the Kurosh subgroup theorem, including proofs of Kuhn (1952),[2] Mac Lane (1958)[3] and others. The theorem was also generalized for describing subgroups of amalgamated free products and HNN extensions.[4] [5] Other generalizations include considering subgroups of free pro-finite products[6] and a version of the Kurosh subgroup theorem for topological groups.[7] In modern terms, the Kurosh subgroup theorem is a straightforward corollary of the basic structural results of Bass-Serre theory about groups acting on trees.[8]

Statement of the theorem


Let G = AB be the free product of groups A and B and let H G be a subgroup of G. Then there exist a family (Ai)i of subgroups Ai A, a family (Bj)j J of subgroups Bj B, families gi, i I and fj, j J of elements of G, and a I subset X G such that

This means that X freely generates a subgroup of G isomorphic to the free group F(X) with free basis X and that, moreover, giAigi1, fjBjfj1 and X generate H in G as a free product of the above form. There is a generalization of this to the case of free products with arbitrarily many factors.[9] Its formulation is: If H is a subgroup of iIGi = G, then where X G and J is some index set and gj G and each Hj is a subgroup of some Gi.

Proof using Bass-Serre theory


The Kurosh subgroup theorem easily follows from the basic structural results in BassSerre theory, as explained, for example in the book of Cohen (1987)[8] : Let G = AB and consider G as the fundamental group of a graph of groups Y consisting of a single non-loop edge with the vertex groups A and B and with the trivial edge group. Let X be the Bass-Serre universal covering tree for the graph of groups Y. Since H G also acts on X, consider the quotient graph of groups Z for the action of H on X. The vertex groups of Z are subgroups of G-stabilizers of vertices of X, that is, they are conjugate in G to subgroups of A and B. The edge groups of Z are trivial since the G-stabilizers of edges of X were trivial. By the fundamental theorem of BassSerre theory, H is canonically isomorphic to the fundamental group of the graph of groups Z. Since the edge groups of Z are trivial, it follows that H is equal to the free product of the vertex groups of Z and the free group F(X) which is the fundamental group (in the standard topological sense) of the underlying graph Z of Z. This implies the conclusion of the Kurosh subgroup theorem.

Kurosh subgroup theorem

155

Notes
[1] [2] [3] [4] [5] [6] [7] [8] [9] A. G. Kurosh, Die Untergruppen der freien Produkte von beliebigen Gruppen. Mathematische Annalen, vol. 109 (1934), pp. 647-660. H. W. Kuhn. Subgroup theorems for groups presented by generators and relations. Annals of Mathematics (2), vol. 56, (1952), pp. 22-46 S. Mac Lane. A proof of the subgroup theorem for free products. Mathematika, vol. 5 (1958), pp. 13-19 A. Karrass, and D. Solitar. The subgroups of a free product of two groups with an amalgamated subgroup. Transactions of the American Mathematical Society, vol. 150 (1970), pp. 227-255. A. Karrass, and D. Solitar. Subgroups of HNN groups and groups with one defining relation. Canadian Journal of Mathematics, vol. 23 (1971), pp. 627-643. Zalesskii, Pavel Aleksandrovich (1990). "[Open subgroups of free profinite products over a profinite space of indices]" (in Russian). Doklady Akademii Nauk SSSR 34 (1): 1720. P. Nickolas. A Kurosh subgroup theorem for topological groups. Proceedings of the London Mathematical Society (3), vol. 42 (1981), no. 3, pp. 461-477 Daniel Cohen. Combinatorial group theory: a topological approach. London Mathematical Society Student Texts, 14. Cambridge University Press, Cambridge, 1989. ISBN 0-521-34133-7; 0-521-34936-2 William S. Massey, Algebraic topology: an introduction. (http:/ / books. google. com/ books?id=IX0dhDDHezgC& pg=PA218& dq="Kurosh+ subgroup+ theorem"& as_brr=3& ei=dQ10S8zsKKasNaSNgJsE& cd=1#v=onepage& q="Kurosh subgroup theorem"& f=false) Graduate Texts in Mathematics, Springer-Verlag, New York, 1977, ISBN 0387902716; pp. 218225

Lagrange's theorem (group theory)


Lagrange's theorem, in the mathematics of group theory, states that for any finite group G, the order (number of elements) of every subgroup H of G divides the order of G. The theorem is named after Joseph Lagrange.

Proof of Lagrange's Theorem


This can be shown using the concept of left cosets of H in G. The left cosets are the equivalence classes of a certain equivalence relation on G and therefore form a partition of G. Specifically, x and y in G are related if and only if there exists h in H such that x = yh. If we can show that all cosets of H have the same number of elements, then each coset of H has precisely |H| elements. We are then done since the order of H times the number of cosets is equal to the number of elements in G, thereby proving that the order of H divides the order of G. Now, if aH and bH are two left cosets of H, we can define a map f : aH bH by setting f(x) = ba-1x. This map is bijective because its inverse is given by This proof also shows that the quotient of the orders |G| / |H| is equal to the index [G : H] (the number of left cosets of H in G). If we write this statement as

then, seen as a statement about cardinal numbers, it is equivalent to the Axiom of choice.

Using the theorem


A consequence of the theorem is that the order of any element a of a finite group (i.e. the smallest positive integer number k with ak = e, where e is the identity element of the group) divides the order of that group, since the order of a is equal to the order of the cyclic subgroup generated by a. If the group has n elements, it follows

This can be used to prove Fermat's little theorem and its generalization, Euler's theorem. These special cases were known long before the general theorem was proved. The theorem also shows that any group of prime order is cyclic and simple. This in turn can be used to prove Wilson's theorem, that if p is prime then p is a factor of (p-1)!+1.

Lagrange's theorem (group theory)

156

Existence of subgroups of given order


Lagrange's theorem raises the converse question as to whether every divisor of the order of a group is the order of some subgroup. This does not hold in general: given a finite group G and a divisor d of |G|, there does not necessarily exist a subgroup of G with order d. The smallest example is the alternating group G = A4 which has 12 elements but no subgroup of order 6. A CLT group is a finite group with the property that for every divisor of the order of the group, there is a subgroup of that order. It is known that a CLT group must be solvable and that every supersolvable group is a CLT group: however there exists solvable groups which are not CLT and CLT groups which are not supersolvable. There are partial converses to Lagrange's theorem. For general groups, Cauchy's theorem guarantees the existence of an element, and hence of a cyclic subgroup, of order any prime dividing the group order; Sylow's theorem extends this to the existence of a subgroup of order equal to the maximal power of any prime dividing the group order. For solvable groups, Hall's theorems assert the existence of a subgroup of order equal to any unitary divisor of the group order (that is, a divisor coprime to its cofactor).

History
Lagrange did not prove Lagrange's theorem in its general form. He stated, in his article Rflexions sur la rsolution algbrique des quations,[1] that if a polynomial in n variables has its variables permuted in all n ! ways, the number of different polynomials that are obtained is always a factor of n !. (For example if the variables x, y, and z are permuted in all 6 possible ways in the polynomial x + y - z then we get a total of 3 different polynomials: x + y z, x + z - y, and y + z x. Note that 3 is a factor of 6.) The number of such polynomials is the index in the symmetric group Sn of the subgroup H of permutations which preserve the polynomial. (For the example of x + y z, the subgroup H in S3 contains the identity and the transposition (xy).) So the size of H divides n !. With the later development of abstract groups, this result of Lagrange on polynomials was recognized to extend to the general theorem about finite groups which now bears his name. Lagrange did not prove his theorem; all he did, essentially, was to discuss some special cases. The first complete proof of the theorem was provided by Abbati and published in 1803.[2]

Notes
[1] Lagrange, J. L. (1771) "Rflexions sur la rsolution algbrique des quations" [Reflections on the algebraic solution of equations] (part II), Nouveaux Mmoires de lAcadmie Royale des Sciences et Belles-Lettres de Berlin, pages 138-254; see especially pages 202-203. Available on-line (in French, among Lagrange's collected works) at: http:/ / math-doc. ujf-grenoble. fr/ cgi-bin/ oeitem?id=OE_LAGRANGE__3_205_0 [Click on "Section seconde. De la rsolution des quations du quatrime degr 254-304"]. [2] P. Abbati (1803) "Lettera di Pietro Abbati Modenese al socio Paolo Ruffini da questo presentata il di 16. Dcembre 1802" [Letter from Pietro Abbati of Modena to the member Paolo Ruffini, who submitted it on the 16. December 1802], Memorie di Matematica e di Fisica della Societ Italiana delle Scienze, vol. 10 (part 2), pages 385-409. See also: Richard L. Roth (April 2001) "A history of Lagrange's theorem on groups," Mathematics Magazine, vol. 74, no. 2, pages 99-108.

References
Bray, Henry G. (1968), "A note on CLT groups", Pacific J. Math. 27 (2): 229231 Gallian, Joseph (2006), Contemporary Abstract Algebra (6th ed.), Boston: Houghton Mifflin, ISBN978-0-618-51471-7 Dummit, David S.; Foote, Richard M. (2004), Abstract algebra (3rd ed.), New York: John Wiley & Sons, ISBN978-0-471-43334-7, MR2286236 Roth, Richard R. (2001), "A History of Lagrange's Theorem on Groups", Mathematics Magazine 74 (2): 99108, doi:10.2307/2690624, JSTOR2690624

LaskerNoether theorem

157

LaskerNoether theorem
In mathematics, the LaskerNoether theorem states that every Noetherian ring is a Lasker ring, which means that every ideal can be written as an intersection of finitely many primary ideals (which are related to, but not quite the same as, powers of prime ideals). The theorem was first proven by Emanuel Lasker(1905) for the special case of polynomial rings and convergent power series rings, and was proven in its full generality by Emmy Noether(1921). The LaskerNoether theorem is an extension of the fundamental theorem of arithmetic, and more generally the fundamental theorem of finitely generated abelian groups to all Noetherian rings. It has a straightforward extension to modules stating that every submodule of a finitely generated module over a Noetherian ring is a finite intersection of primary submodules. This contains the case for rings as a special case, considering the ring as a module over itself, so that ideals are submodules. This also generalizes the primary decomposition form of the structure theorem for finitely generated modules over a principal ideal domain, and for the special case of polynomial rings over a field, it generalizes the decomposition of an algebraic set into a finite union of (irreducible) varieties. The first algorithm for computing primary decompositions for polynomial rings was published by Noether's student Grete Hermann(1926).

Definitions
Write R for a commutative ring, and M and N for modules over it. A zero divisor of a module M is an element x of R such that xm=0 for some non-zero m in M. An element x of R is called nilpotent in M if xnM=0 for some positive integer n. A module is called coprimary if every zero divisor of M is nilpotent in M. For example, groups of prime power order and free abelian groups are coprimary modules over the ring of integers. A submodule M of a module N is called a primary submodule if N/M is coprimary. An ideal I is called primary if it is a primary submodule of R. This is equivalent to saying that if ab is in I then either a is in I or bn is in I for some n, and to the condition that every zero-divisor of the ring R/I is nilpotent. A submodule M of a module N is called irreducible if it is not an intersection of two strictly larger submodules. An associated prime of a module M is a prime ideal that is the annihilator of some element of M.

Statement
The LaskerNoether theorem for modules states every submodule of a finitely generated module over a Noetherian ring is a finite intersection of primary submodules. For the special case of ideals it states that every ideal of a Noetherian ring is a finite intersection of primary ideals. An equivalent statement is: every finitely generated module over a Noetherian ring is contained in a finite product of coprimary modules. The LaskerNoether theorem follows immediately from the following three facts: Any submodule of a finitely generated module over a Noetherian ring is an intersection of a finite number of irreducible submodules. If M is an irreducible submodule of a finitely generated module N over a Noetherian ring then N/M has only one associated prime ideal. A finitely generated module over a Noetherian ring is coprimary if and only if it has at most one associated prime.

LaskerNoether theorem

158

Irreducible decomposition in rings


The study of the decomposition of ideals in rings began as a remedy for the lack of unique factorization in number fields like , in which . If a number does not factor uniquely into primes, then the ideal generated by the number may still factor into the intersection of powers of prime ideals. Failing that, an ideal may at least factor into the intersection of primary ideals. Let R be a Noetherian ring, and I an ideal in R. Then I has an irredundant primary decomposition into primary ideals.

Irredundancy means: Removing any of the changes the intersection, i.e.,

for all i, where the hat denotes omission. The associated prime ideals are distinct.

More over, this decomposition is unique in the following sense: the set of associated prime ideals is unique, and the primary ideal above every minimal prime in this set is also unique. However, primary ideals which are associated with non-minimal prime ideals are in general not unique. In the case of the ring of integers generated by , is , the LaskerNoether theorem is equivalent to the fundamental theorem of , then the primary decomposition of the ideal arithmetic. If an integer n has prime factorization

Minimal decompositions and uniqueness


In this section, all modules will be finitely generated over a Noetherian ring R. A primary decomposition of a submodule M of a module N is called minimal if it has the smallest possible number of primary modules. For minimal decompositions, the primes of the primary modules are uniquely determined: they are the associated primes of N/M. Moreover the primary submodules associated to the minimal or isolated associated primes (those not containing any other associated primes) are also unique. However the primary submodules associated to the non-minimal associated primes (called embedded primes for geometric reasons) need not be unique. Example: Let N=R=k[x,y] for some field k, and let M be the ideal (xy,y2). Then M has two different minimal primary decompositions M = (y) (x, y2) = (y) (x+y,y2). The minimal prime is (y) and the embedded prime is (x,y).

LaskerNoether theorem

159

When the conclusion does not hold


The decomposition does not hold in general for non-commutative Noetherian rings. Noether gave an example of a non-commutative Noetherian ring with a right ideal that is not an intersection of primary ideals.

Additive theory of ideals


This result is the first in an area now known as the additive theory of ideals, which studies the ways of representing an ideal as the intersection of a special class of ideals. The decision on the "special class", e.g., primary ideals, is a problem in itself. In the case of non-commutative rings, the class of tertiary ideals is a useful substitute for the class of primary ideals.

References
Danilov, V.I. (2001), "Lasker ring" [1], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 Eisenbud, David (1995), Commutative algebra, Graduate Texts in Mathematics, 150, Berlin, New York: Springer-Verlag, ISBN978-0-387-94268-1; 978-0-387-94269-8, MR1322960, esp. section 3.3. Hermann, Grete (1926), "Die Frage der endlich vielen Schritte in der Theorie der Polynomideale", Mathematische Annalen 95: 736788, doi:10.1007/BF01206635. English translation in Communications in Computer Algebra 32/3 (1998): 830. Lasker, E. (1905), "Zur Theorie der Moduln und Ideale", Math. Ann. 60: 19116, doi:10.1007/BF01447495 Markov, V.T. (2001), "Primary decomposition" [2], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 Noether, Emmy (1921), "Idealtheorie in Ringbereichen" [3], Mathematische Annalen 83 (1): 24, doi:10.1007/BF01464225 Curtis, Charles (1952), "On Additive Ideal Theory in General Rings", American Journal of Mathematics (The Johns Hopkins University Press) 74 (3): 687700, doi:10.2307/2372273, JSTOR2372273 Krull, Wolfgang (1928), [year=1928 "Zur Theorie der zweiseitigen Ideale in nichtkommutativen Bereichen"], Mathematische Zeitschrift 28 (1): 481503, doi:10.1007/BF01181179, year=1928

References
[1] http:/ / eom. springer. de/ L/ l057600. htm [2] http:/ / eom. springer. de/ P/ p074450. htm [3] http:/ / www. springerlink. com/ content/ m3457w8h62475473/ fulltext. pdf

Latimer-MacDuffee theorem

160

Latimer-MacDuffee theorem
The Latimer-MacDuffee theorem is a theorem in abstract algebra, a branch of mathematics. Let be a monic, irreducible polynomial of degree -similarity classes of . The Latimer-MacDuffee theorem gives a one-to-one matrices with characteristic polynomial and the ideal correspondence between classes in the order where ideals are considered equivalent if they are equal up to an overall (nonzero) rational scalar multiple. (Note that this order need not be the full ring of integers, so nonzero ideals need not be invertible.) Since an order in a number field has only finitely many ideal classes (even if it is not the maximal order, and we mean here ideals classes for all nonzero ideals, not just the invertible ones), it follows that there are only finitely many conjugacy classes of matrices over the integers with characteristic polynomial .

References
A Correspondence Between Classes of Ideals and Classes of Matrices, by Claiborne G. Latimer; C. C. MacDuffee The Annals of Mathematics. 1933.

Lattice theorem
In mathematics, the lattice theorem, sometimes referred to as the fourth isomorphism theorem or the correspondence theorem, states that if is a normal subgroup of a group , then there exists a bijection from the set of all subgroups containing of with of such that contains , onto the set of all subgroups of the quotient group . The structure of the subgroups of is exactly the same as the structure of the subgroups of and the lattice of subgroups

collapsed to the identity element. is

This establishes a monotone Galois connection between the lattice of subgroups of , where the associated closure operator on subgroups of Specifically, If G is a group, N is a normal subgroup of G, is the set of all subgroups A of G such that is the set of all subgroups of G/N, then there is a bijective map for all One further has that if A and B are in if if and only if then ; , and A' = A/N and B' = B/N, then such that , and

, where B:A is the index of A in B (the number of cosets bA of A in B);

where is the subgroup of generated by and is a normal subgroup of if and only if is a normal subgroup of

This list is far from exhaustive. In fact, most properties of subgroups are preserved in their images under the bijection onto subgroups of a quotient group.

Lattice theorem

161

References
W.R. Scott: Group Theory, Prentice Hall, 1964.

Levitzky's theorem
In mathematics, more specifically ring theory and the theory of nil ideals, Levitzky's theorem, named after Jacob Levitzki, states that in a right Noetherian ring, every nil one-sided ideal is necessarily nilpotent.[1] [2] Levitzky's theorem is one of the many results suggesting the veracity of the Kthe conjecture, and indeed provided a solution to one of Kthe's questions as described in (Levitzki 1945). The result was originally submitted in 1939 as (Levitzki 1950), and a particularly simple proof was given in (Utumi 1963).

Notes
[1] Herstein, Theorem 1.4.5, p. 37 [2] Isaacs, Theorem 14.38, p. 210

References
I. Martin Isaacs (1993), Algebra, a graduate course (1st ed.), Brooks/Cole Publishing Company, ISBN0-534-19002-2 I.N. Herstein (1968), Noncommutative rings (1st ed.), The Mathematical Association of America, ISBN0-88385-015-X J. Levitzki (1950). "On multiplicative systems" (http://www.numdam.org/item?id=CM_1951__8__76_0). Compositio Math. 8: 7680. MR0033799. Levitzki, Jakob (1945), "Solution of a problem of G. Koethe", American Journal of Mathematics (The Johns Hopkins University Press) 67 (3): 437442, doi:10.2307/2371958, ISSN0002-9327, JSTOR2371958, MR0012269 Utumi, Yuzo (1963), "Mathematical Notes: A Theorem of Levitzki", The American Mathematical Monthly (Mathematical Association of America) 70 (3): 286, doi:10.2307/2313127, ISSN0002-9890, JSTOR2313127, MR1532056

Lie's third theorem

162

Lie's third theorem


In mathematics, Lie's third theorem often means the result that states that any finite-dimensional Lie algebra g, over the real numbers, is the Lie algebra associated to some Lie group G. The relationship to the history has though become confused. There were (naturally) two other preceding theorems, of Sophus Lie. Those relate to the infinitesimal transformations of a transformation group acting on a smooth manifold. But, in fact, that language is anachronistic. The manifold concept was not clearly defined at the time, the end of the nineteenth century, when Lie was founding the theory. The conventional third theorem on the list was a result stating the Jacobi identity for the infinitesimal transformations, of a local Lie group. This result has a converse, stating that in the presence of a Lie algebra of vector fields, integration gives a local Lie group action. The result initially stated is an intrinsic and global converse to the original theorem, therefore.

External links
Encyclopaedia of Mathematics (EoM) article at Springer.de [1]

References
[1] http:/ / eom. springer. de/ l/ l058760. htm

LieKolchin theorem
In mathematics, the LieKolchin theorem is a theorem in the representation theory of linear algebraic groups; Lie's theorem is the analog for linear Lie algebras. It states that if G is a connected and solvable linear algebraic group defined over an algebraically closed field and

a representation on a nonzero finite-dimensional vector space V, then there is a one-dimensional linear subspace L of V such that

That is, (G) has an invariant line L, on which G therefore acts through a one-dimensional representation. This is equivalent to the statement that V contains a nonzero vector v that is a common (simultaneous) eigenvector for all . Because every (nonzero finite-dimensional) representation of G has a one-dimensional invariant subspace according to the LieKolchin theorem, every irreducible finite-dimensional representation of a connected and solvable linear algebraic group G has dimension one, which is another way to state the LieKolchin theorem. Lie's theorem states that any nonzero representation of a solvable Lie algebra on a finite dimensional vector space over an algebraically closed field of characteristic 0 has a one-dimensional invariant subspace. The result for Lie algebras was proved by Sophus Lie(1876) and for algebraic groups was proved by Ellis Kolchin(1948, p.19). The Borel fixed point theorem generalizes the LieKolchin theorem.

LieKolchin theorem

163

Triangularization
Sometimes the theorem is also referred to as the LieKolchin triangularization theorem because by induction it implies that with respect to a suitable basis of V the image has a triangular shape; in other words, the image group is conjugate in GL(n,K) (where n = dim V) to a subgroup of the group T of upper triangular matrices, the standard Borel subgroup of GL(n,K): the image is simultaneously triangularizable. The theorem applies in particular to a Borel subgroup of a semisimple linear algebraic group G.

Lie's theorem
Lie's theorem states that if V is a finite dimensional vector space over an algebraically closed field of characteristic 0, then for any solvable Lie algebra of endomorphisms of V there is a vector that is an eigenvector for every element of the Lie algebra. Applying this result repeatedly shows that there is a basis for V such that all elements of the Lie algebra are represented by upper triangular matrices. This is a generalization of the result of Frobenius that commuting matrices are simultaneously upper triangularizable, as commuting matrices form an abelian Lie algebra, which is a fortiori solvable. A consequence of Lie's theorem is that any finite dimensional solvable Lie algebra over a field of characteristic 0 has a nilpotent derived algebra.

Counter-examples
If the field K is not algebraically closed, the theorem can fail. The standard unit circle, viewed as the set of complex numbers of absolute value one is a one-dimensional commutative (and therefore solvable) linear algebraic group over the real numbers which has a two-dimensional representation into the special orthogonal group SO(2) without an invariant (real) line. Here the image of is the orthogonal matrix

For algebraically closed fields of characteristic p>0 Lie's theorem holds provided the dimension of the representation is less than p, but can fail for representations of dimension p. An example is given by the 3-dimensional nilpotent Lie algebra spanned by 1, x, and d/dx acting on the p-dimensional vector space k[x]/(xp), which has no eigenvectors. Taking the semidirect product of this 3-dimensional Lie algebra by the p-dimensional representation (considered as an abelian Lie algebra) gives a solvable Lie algebra whose derived algebra is not nilpotent.

References
Gorbatsevich, V.V. (2001), "LieKolchin theorem" [1], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 Kolchin, E. R. (1948), "Algebraic matric groups and the Picard-Vessiot theory of homogeneous linear ordinary differential equations", Annals of Mathematics. Second Series 49: 142, ISSN0003-486X, JSTOR1969111, MR0024884 Lie, Sophus (1876), "Theorie der Transformationsgruppen. Abhandlung II" [2], Archiv for Mathematik Og Naturvidenskab 1: 152193 William C. Waterhouse, Introduction to Affine Group Schemes, Graduate Texts in Mathematics vol. 66, Springer Verlag New York, 1979 (chapter 10, in particular section 10.2).

LieKolchin theorem

164

References
[1] http:/ / eom. springer. de/ l/ l058710. htm [2] http:/ / www. archive. org/ details/ archivformathem02sarsgoog

Maschke's theorem
In mathematics, Maschke's theorem,[1] [2] named after Heinrich Maschke,[3] is a theorem in group representation theory that concerns the decomposition of representations of a finite group into irreducible pieces. If (V,) is a finite-dimensional representation of a finite group G over a field of characteristic zero, and U is an invariant subspace of V, then the theorem claims that U admits an invariant direct complement W; in other words, the representation (V,) is completely reducible. More generally, the theorem holds for fields of positive characteristic p, such as the finite fields, if the prime p doesn't divide the order ofG.

Reformulation and the meaning


One of the approaches to representations of finite groups is through module theory. Representations of a group G are replaced by modules over its group algebraKG. Irreducible representations correspond to simple modules. Maschke's theorem addresses the question: is a general (finite-dimensional) representation built from irreducible subrepresentations using the direct sum operation? In the module-theoretic language, is an arbitrary module semisimple? In this context, the theorem can be reformulated as follows: Let G be a finite group and K a field whose characteristic does not divide the order of G. Then KG, the group algebra of G, is a semisimple algebra.[4] [5] The importance of this result stems from the well developed theory of semisimple rings, in particular, the ArtinWedderburn theorem (sometimes referred to as Wedderburn's Structure Theorem). When K is the field of complex numbers, this shows that the algebra KG is a product of several copies of complex matrix algebras, one for each irreducible representation.[6] If the field K has characteristic zero, but is not algebraically closed, for example, K is a field of real or rational numbers, then a somewhat more complicated statement holds: the group algebra KG is a product of matrix algebras over division rings over K. The summands correspond to irreducible representations of G over K.[7] Returning to representation theory, Maschke's theorem and its module-theoretic version allow one to make general conclusions about representations of a finite group G without actually computing them. They reduce the task of classifying all representations to a more manageable task of classifying irreducible representations, since when the theorem applies, any representation is a direct sum of irreducible pieces (constituents). Moreover, it follows from the JordanHlder theorem that, while the decomposition into a direct sum of irreducible subrepresentations may not be unique, the irreducible pieces have well-defined multiplicities. In particular, a representation of a finite group over a field of characteristic zero is determined up to isomorphism by its character.

Notes
[1] Maschke, Heinrich (1898-07-22). "Ueber den arithmetischen Charakter der Coefficienten der Substitutionen endlicher linearer Substitutionsgruppen [On the arithmetical character of the coefficients of the substitutions of finite linear substitution groups]" (http:/ / resolver. sub. uni-goettingen. de/ purl?GDZPPN002256975) (in German). Math. Ann. 50 (4): 492498. JFM29.0114.03. MR1511011. . [2] Maschke, Heinrich (1899-07-27). "Beweis des Satzes, dass diejenigen endlichen linearen Substitutionsgruppen, in welchen einige durchgehends verschwindende Coefficienten auftreten, intransitiv sind [Proof of the theorem that those finite linear substitution groups, in which some everywhere vanishing coefficients appear, are intransitive]" (http:/ / resolver. sub. uni-goettingen. de/ purl?GDZPPN002257599) (in German). Math. Ann. 52 (23): 363368. JFM30.0131.01. MR1511061. . [3] O'Connor, John J.; Robertson, Edmund F., "Heinrich Maschke" (http:/ / www-history. mcs. st-andrews. ac. uk/ Biographies/ Maschke. html), MacTutor History of Mathematics archive, University of St Andrews, . [4] It follows that every module over KG is a semisimple module.

Maschke's theorem
[5] The converse statement also holds: if the characteristic of the field divides the order of the group (the modular case), then the group algebra is not semisimple. [6] The number of the summands can be computed, and turns out to be equal to the number of the conjugacy classes of the group. [7] One must be careful, since a representation may decompose differently over different fields: a representation may be irreducible over the real numbers but not over the complex numbers.

165

References
Lang, Serge (2002-01-08). Algebra. Graduate Texts in Mathematics, 211 (Revised 3rd ed.). New York: Springer-Verlag. ISBN978-0-387-95385-4. MR1878556. Zbl0984.00001. Serre, Jean-Pierre (1977-09-01). Linear Representations of Finite Groups. Graduate Texts in Mathematics, 42. New YorkHeidelberg: Springer-Verlag. ISBN978-0-387-90190-9. MR0450380. Zbl0355.20006.

Milnor conjecture
In mathematics, the Milnor conjecture was a proposal by John Milnor(1970) of a description of the Milnor K-theory (mod2) of a general field F with characteristic different from 2, by means of the Galois (or equivalently tale) cohomology of F with coefficients in Z/2Z. It was proved by Vladimir Voevodsky(1996, 2003a, 2003b).

Statement of the theorem


Let F be a field of characteristic different from2. Then there is an isomorphism

for all n0.

About the proof


The proof of this theorem by Vladimir Voevodsky uses several ideas developed by Voevodsky, Andrei Suslin, Fabien Morel, Eric Friedlander, and others, including the newly-minted theory of motivic cohomology (a kind of substitute for singular cohomology for algebraic varieties) and the motivic Steenrod algebra.

Generalizations
The analogue of this result for primes other than 2 was known as the BlochKato conjecture. Work of Voevodsky, Markus Rost, and Charles Weibel yielded a complete proof of this conjecture in 2009; the result is now called the norm residue isomorphism theorem.

References
Mazza, Carlo; Voevodsky, Vladimir; Weibel, Charles (2006), Lecture notes on motivic cohomology [1], Clay Mathematics Monographs, 2, Providence, R.I.: American Mathematical Society, ISBN978-0-8218-3847-1; 978-0-8218-3847-1, MR2242284 Milnor, John Willard (1970), "Algebraic K-theory and quadratic forms", Inventiones Mathematicae 9: 318344, doi:10.1007/BF01425486, ISSN0020-9910, MR0260844 Voevodsky, V. (1996), The Milnor Conjecture [2], Preprint Voevodsky, Vladimir (2003a), "Reduced power operations in motivic cohomology" [3], Institut des Hautes tudes Scientifiques. Publications Mathmatiques (98): 157, doi:10.1007/s10240-003-0009-z, ISSN0073-8301, MR2031198

Milnor conjecture Voevodsky, Vladimir (2003b), "Motivic cohomology with Z/2-coefficients" [4], Institut des Hautes tudes Scientifiques. Publications Mathmatiques (98): 59104, doi:10.1007/s10240-003-0010-6, ISSN0073-8301, MR2031199

166

References
[1] [2] [3] [4] http:/ / math. rutgers. edu/ ~weibel/ motiviclectures. html http:/ / www. math. uiuc. edu/ K-theory/ 0170 http:/ / www. numdam. org/ item?id=PMIHES_2003__98__1_0 http:/ / www. numdam. org/ item?id=PMIHES_2003__98__59_0

MordellWeil theorem
In mathematics, the MordellWeil theorem states that for an abelian variety A over a number field K, the group A(K) of K-rational points of A is a finitely-generated abelian group, called the Mordell-Weil group. The case with A an elliptic curve E and K the rational number field Q is Mordell's theorem, answering a question apparently posed by Poincar around 1908; it was proved by Louis Mordell in 1922. The tangent-chord process (one form of addition theorem on a cubic curve) had been known as far back as the seventeenth century. The process of infinite descent of Fermat was well known, but Mordell succeeded in establishing the finiteness of the quotient group E(Q)/2E(Q) which forms a major step in the proof. Certainly the finiteness of this group is a necessary condition for E(Q) to be finitely-generated; and it shows that the rank is finite. This turns out to be the essential difficulty. It can be proved by direct analysis of the doubling of a point on E. Some years later Andr Weil took up the subject, producing the generalisation to Jacobians of higher genus curves over arbitrary number fields in his doctoral dissertation published in 1928. More abstract methods were required, to carry out a proof with the same basic structure. The second half of the proof needs some type of height function, in terms of which to bound the 'size' of points of A(K). Some measure of the co-ordinates will do; heights are logarithmic, so that (roughly speaking) it is a question of how many digits are required to write down a set of homogeneous coordinates. For an abelian variety, there is no a priori preferred representation, though, as a projective variety. Both halves of the proof have been improved significantly, by subsequent technical advances: in Galois cohomology as applied to descent, and in the study of the best height functions (which are quadratic forms). The theorem left unanswered a number of questions: Calculation of the rank (still a demanding computational problem, and not always effective, as far as it is currently known). Meaning of the rank: see Birch and Swinnerton-Dyer conjecture. For a curve C in its Jacobian variety as A, can the intersection of C with A(K) be infinite? (Not unless C = A, according to Mordell's conjecture, proved by Faltings.) In the same context, can C contain infinitely many torsion points of A? (No, according to the Manin-Mumford conjecture proved by Raynaud, other than in the elliptic curve case.)

MordellWeil theorem

167

References
A. Weil, L'arithmtique sur les courbes algbriques, Acta Math 52, (1929) p.281-315, reprinted in vol 1 of his collected papers ISBN 0387903305 L.J. Mordell, On the rational solutions of the indeterminate equations of the third and fourth degrees, Proc Cam. Phil. Soc. 21, (1922) p.179. J. H. Silverman, The arithmetic of elliptic curves, ISBN 0387962034 second edition

Multinomial theorem
In mathematics, the multinomial theorem says how to write a power of a sum in terms of powers of the terms in that sum. It is the generalization of the binomial theorem to polynomials.

Theorem
For any positive integer m and any nonnegative integer n, the multinomial formula tells us how a polynomial expands when raised to an arbitrary power:

The sum is taken over all sequences of nonnegative integer indices k1 through km such the sum of all ki is n. That is, for each term in the expansion, the exponents must add up to n. Also, as with the binomial theorem, quantities of the form x0 that appear are taken to equal 1 (even when x equals zero). Alternatively, this can be written concisely using multiindices as

where =(1,2,,m) and x=x11x22xmm.

Number of multinomial coefficients


The number of terms in multinomial sum, #n,m, is equal to the number of monomials of degree n on the variables x1,,xm:

The count can be easily performed using the Stars and bars (combinatorics) method.

Multinomial coefficients
The numbers

(which can also be written as:)

are the multinomial coefficients. Just like "n choose k" are the coefficients when you raise a binomial to the nth power (e.g. the coefficients are 1,3,3,1 for (a+b)3, where n=3), the multinomial coefficients appear when one raises a multinomial to the nth power (e.g. (a+b+c)3)

Multinomial theorem Sum of all multinomial coefficients The substitution of xi=1 for all i into:

168

gives immediately that

Central multinomial coefficients All of the multinomial coefficients for which the following holds true:

are central multinomial coefficients: the greatest ones and all of equal size. A special case for m=2 is central binomial coefficient.

Example multinomial coefficients


We could have calculated each coefficient by first expanding (a+b+c)2=a2+b2+c2+2ab+2bc+2ac, then self-multiplying it again to get (a+b+c)3 (and then if we were raising it to higher powers, we'd multiply it by itself even some more). However this process is slow, and can be avoided by using the multinomial theorem. The multinomial theorem "solves" this process by giving us the closed form for any coefficient we might want. It is possible to "read off" the multinomial coefficients from the terms by using the multinomial coefficient formula. For example: has the coefficient

has the coefficient

We could have also had a 'd' variable, or even more variableshence the multinomial theorem. The binomial theorem and binomial coefficients are special cases, for m=2, of the multinomial theorem and multinomial coefficients, respectively.

Interpretations
Ways to put objects into boxes
The multinomial coefficients have a direct combinatorial interpretation, as the number of ways of depositing n distinct objects into m distinct bins, with k1 objects in the first bin, k2 objects in the second bin, and so on.[1]

Number of ways to select according to a distribution


In statistical mechanics and combinatorics if one has a number distribution of labels then the multinomial coefficients naturally arise from the binomial coefficients. Given a number distribution {ni} on a set of N total items, ni represents the number of items to be given the label i. (In statistical mechanics i is the label of the energy state.) The number of arrangements is found by

Multinomial theorem

169

Choosing n1 of the total N to be labeled 1. This can be done

ways. ways. ways.

From the remaining Nn1 items choose n2 to label 2. This can be done From the remaining Nn1n2 items choose n3 to label 3. Again, this can be done Multiplying the number of choices at each step results in:

Upon cancellation, we arrive at the formula given in the introduction.

Number of unique permutations of words


In addition, the multinomial coefficient is also the number of distinct ways to permute a multiset of n elements, and ki are the multiplicities of each of the distinct elements. For example, the number of distinct permutations of the letters of the word MISSISSIPPI, which has 1 M, 4 Is, 4 Ss, and 2 Ps is

(This is just like saying that there are 11! ways to permute the lettersthe common interpretation of factorial as the number of unique permutations. However, we created duplicate permutations, due to the fact that some letters are the same, and must divide to correct our answer.)

Generalized Pascal's triangle


One can use the multinomial theorem to generalize Pascal's triangle or Pascal's pyramid to Pascal's simplex. This provides a quick way to generate a lookup table for multinomial coefficients. The case of n=3 can be easily drawn by hand. The case of n=4 can be drawn with effort as a series of growing pyramids.

Proof
This proof of the multinomial theorem uses the binomial theorem and induction on m. First, for m=1, both sides equal x1n since there is only one term k1=n in the sum. For the induction step, suppose the multinomial theorem holds for m. Then

by the induction hypothesis. Applying the binomial theorem to the last factor,

which completes the induction. The last step follows because

Multinomial theorem as can easily be seen by writing the three coefficients using factorials as follows:

170

References
[1] National Institute of Standards and Technology (May, 11, 2010). "NIST Digital Library of Mathematical Functions" (http:/ / dlmf. nist. gov/ ). Section 26.4 (http:/ / dlmf. nist. gov/ 26. 4). . Retrieved August 30, 2010.

NielsenSchreier theorem
In group theory, a branch of mathematics, the NielsenSchreier theorem is the statement that every subgroup of a free group is itself free.[1] [2] [3] It is named after Jakob Nielsen and Otto Schreier.

Statement of the theorem


A free group may be defined from a group presentation consisting of a set of generators and the empty set of relations (equations that the generators satisfy). That is, it is the unique group in which every element is a product of some sequence of generators and their inverses, and in which there are no equations between group elements that do not follow in a trivial way from the equations gg1 describing the relation between a generator and its inverse. The elements of a free group may be described as all of the possible reduced words formed by sequences of generators and their inverse that have no adjacent pair of a generator and the inverse of the same generator. The NielsenSchreier theorem states that if G is a subgroup of a free group, then G is itself isomorphic to a free group. That is, there exists a subset S of elements of G such that every element in S is a product of members of S and their inverses, and such that S satisfies no nontrivial relations.

Example
Let G be the free group with two generators, a and b, and let E be the subgroup consisting of all reduced words that are products of evenly many generators or their inverses. Then E is itself generated by the six elements p = aa, q = ab, r = ab1, s = ba, t = ba1, and u = bb. A factorization of any reduced word in E into these generators and their inverses may be constructed simply by taking consecutive pairs of symbols in the reduced word. However, this is not a free presentation of E because it satisfies the relations p = qr1 = rq1 and s = tu1 = ut1. Instead, E is generated as a free group by the three elements p = aa, q = ab, and s = ba. Any factorization of a word into a product of generators from the six-element generating set {p, q, r, s, t, u} can be transformed into a product of generators from this smaller set by replacing r with ps1, replacing t with sp1, and replacing u with sp1q. There are no additional relations satisfied by these three generators, so E is the free group generated by p, q, and s.[4] The NielsenScheier theorem states that this example is not a coincidence: like E, every subgroup of a free group can be generated as a free group, possibly with a larger set of generators.

Proof
It is possible to prove the NielsenScheier theorem using topology.[1] A free group G on a set of generators is the fundamental group of a bouquet of circles, a topological graph with a single vertex and with an edge for each generator.[5] Any subgroup H of the fundamental group is itself a fundamental group of a covering space of the bouquet, a (possibly infinite) topological Schreier coset graph that has one vertex for each coset of the subgroup.[6] And in any topological graph, it is possible to shrink the edges of a spanning tree of the graph, producing a bouquet of circles that has the same fundamental group H. Since H is the fundamental group of a bouquet of circles, it is

NielsenSchreier theorem itself free.[5] According to Schreier's subgroup lemma, a set of generators for a free presentation of H may be constructed from cycles in the covering graph formed by concatenating a spanning tree path from a base point (the coset of the identity) to one of the cosets, a single non-tree edge, and an inverse spanning tree path from the other endpoint of the edge back to the base point.[7]

171

Axiomatic foundations
Although several different proofs of the NielsenSchreier theorem are known, they all depend on the axiom of choice. In the proof based on fundamental groups of bouquets, for instance, the axiom of choice appears in the guise of the statement that every connected graph has a spanning tree. The use of this axiom is necessary, as there exist models of ZermeloFraenkel set theory in which the axiom of choice and the NielsenSchreier theorem are both false. The NielsenSchreier theorem in turn implies a weaker version of the axiom of choice, for finite sets.[8]

History
The NielsenSchreier theorem is a non-abelian analogue of an older result of Richard Dedekind, that every subgroup of a free abelian group is free abelian.[3] Jakob Nielsen(1921) originally proved a restricted form of the theorem, stating that any finitely-generated subgroup of a free group is free. His proof involves performing a sequence of Nielsen transformations on the subgroup's generating set that reduce their length (as reduced words in the free group from which they are drawn).[1] [9] Otto Schreier proved the NielsenSchreier theorem in its full generality in his 1926 habilitation thesis, Die Untergruppen der freien Gruppe, also published in 1927 in Abh. math. Sem. Hamburg. Univ.[10] [11] The topological proof based on fundamental groups of bouquets of circles is due to Reinhold Baer and Friedrich Levi(1936). Another topological proof, based on the BassSerre theory of group actions on trees, was published by Jean-Pierre Serre(1970).[12]

Notes
[1] [2] [3] [4] [5] [6] [7] [8] [9] Stillwell (1993), Section 2.2.4, The NielsenSchreier Theorem, pp. 103104. Magnus Solitar Karass (http:/ / v3. espacenet. com/ textdoc?DB=EPODOC& IDX=MagnusKarass), Corollary 2.9, p. 95. Johnson (1980), Section 2, The NielsenSchreier Theorem, pp. 923. Johnson (1997), ex. 15, p. 12. Stillwell (1993), Section 2.1.8, Freeness of the Generators, p. 97. Stillwell (1993), Section 2.2.2, The Subgroup Property, pp. 100101. Stillwell (1993), Section 2.2.6, Schreier Transversals, pp. 105106. Luchli (1962); Howard (1985). Magnus Solitar Karass (http:/ / v3. espacenet. com/ textdoc?DB=EPODOC& IDX=MagnusKarass), Section 3.2, A Reduction Process, pp. 121140. [10] O'Connor, John J.; Robertson, Edmund F., "NielsenSchreier theorem" (http:/ / www-history. mcs. st-andrews. ac. uk/ Biographies/ Schreier. html), MacTutor History of Mathematics archive, University of St Andrews, . [11] Hansen, Vagn Lundsgaard (1986), Jakob Nielsen, Collected Mathematical Papers: 1913-1932, Birkhuser, p.117, ISBN9780817631406. [12] Rotman (1995), The NielsenSchreier Theorem, pp. 383387.

NielsenSchreier theorem

172

References
Baer, Reinhold; Levi, Friedrich (1936), "Freie Produkte und ihre Untergruppen", Compositio Math. 3: 391398. Howard, Paul E. (1985), "Subgroups of a free group and the axiom of choice", The Journal of Symbolic Logic 50 (2): 458467, doi:10.2307/2274234, MR793126. Johnson, D. L. (1980), Topics in the Theory of Group Presentations, London Mathematical Society lecture note series, 42, Cambridge University Press, ISBN9780521231084. Johnson, D. L. (1997), Presentations of Groups, London Mathematical Society student texts, 15 (2nd ed.), Cambridge University Press, ISBN9780521585422. Luchli, Hans (1962), "Auswahlaxiom in der Algebra", Commentarii Mathematici Helvetici 37: 118, MR0143705. Magnus, Wilhelm; Karrass, Abraham; Solitar, Donald (1976), Combinatorial Group Theory (2nd revised ed.), Dover Publications. Nielsen, Jakob (1921), "Om regning med ikke-kommutative faktorer og dens anvendelse i gruppeteorien" (in Danish), Math. Tidsskrift B, 1921: 7894, JFM48.0123.03. Rotman, Joseph J. (1995), An Introduction to the Theory of Groups, Graduate Texts in Mathematics, 148 (4th ed.), Springer-Verlag, ISBN9780387942858. Serre, J.-P. (1970), Groupes Discretes, Extrait de I'Annuaire du College de France, Paris. Stillwell, John (1993), Classical Topology and Combinatorial Group Theory, Graduate Texts in Mathematics, 72 (2nd ed.), Springer-Verlag.

PerronFrobenius theorem
In linear algebra, the PerronFrobenius theorem, proved by Oskar Perron(1907) and Georg Frobenius(1912), asserts that a real square matrix with positive entries has a unique largest real eigenvalue and that the corresponding eigenvector has strictly positive components, and also asserts a similar statement for certain classes of nonnegative matrices. This theorem has important applications to probability theory (ergodicity of Markov chains); to the theory of dynamical systems (subshifts of finite type); to economics (Leontief's input-output model[1] 8.3.6 p. 681 [2]); to demography (Leslie population age distribution model[1] 8.3.7 p. 683 [2]); to mathematical background of the internet search engines[3] 15.2 p. 167 [4] and even to ranking of football teams[5] p. 80 [6]. "In addition to saying something useful, the PerronFrobenius theory is elegant. It is a testament to the fact that beautiful mathematics eventually tends to be useful, and useful mathematics eventually tends to be beautiful.[1] "

Statement of the PerronFrobenius theorem


A matrix in which all entries are positive real numbers is here called positive and a matrix whose entries are non-negative real numbers is here called non-negative. The eigenvalues of a real square matrix A are complex numbers and collectively they make up the spectrum of the matrix. The exponential growth rate of the matrix powers Ak as k is controlled by the eigenvalue of A with the largest absolute value. The PerronFrobenius theorem describes the properties of the leading eigenvalue and of the corresponding eigenvectors when A is a non-negative real square matrix. Early results were due to Oskar Perron(1907) and concerned positive matrices. Later, Georg Frobenius(1912) found their extension to certain classes of non-negative matrices.

PerronFrobenius theorem

173

Positive matrices
Let A=(aij) be an nn positive matrix: aij>0 for 1 i, j n. Then the following statements hold. 1. There is a positive real number r, called the Perron root or the PerronFrobenius eigenvalue, such that r is an eigenvalue of A and any other eigenvalue (possibly, complex) is strictly smaller than r in absolute value, || < r. Thus, the spectral radius (A) is equal to r. 2. The PerronFrobenius eigenvalue is simple: r is a simple root of the characteristic polynomial of A. Consequently, the eigenspace associated to r is one-dimensional. (The same is true for the left eigenspace, i.e., the eigenspace for AT.) 3. There exists an eigenvector v = (v1,,vn) of A with eigenvalue r such that all components of v are positive: A v = r v, vi > 0 for 1 i n. (Respectively, there exists a positive left eigenvector w : wT A = r wT, wi > 0.) 4. There are no other positive (moreover non-negative) eigenvectors except positive multiples of v (respectively, left eigenvectors except w), i.e. all other eigenvectors must have at least one negative or non-real component. 5. , where the left and right eigenvectors for A are normalized so that wTv = 1. Moreover, the matrix v wT is the projection onto the eigenspace corresponding tor. This projection is called the Perron projection. 6. CollatzWielandt formula: for all non-negative non-zero vectors x, let f(x) be the minimum value of [Ax]i / xi taken over all those i such that xi 0. Then f is a real valued function whose maximum is the PerronFrobenius eigenvalue. 7. A "Min-max" CollatzWielandt formula takes a form similar to the one above: for all strictly positive vectors x, let g(x) be the maximum value of [Ax]i / xi taken over i. Then g is a real valued function whose minimum is the PerronFrobenius eigenvalue. 8. The PerronFrobenius eigenvalue satisfies the inequalities

These claims can be found in Meyer[1] chapter 8 668-669.

[2]

claims 8.2.11-15 page 667 and exercises 8.2.5,7,9 pages

The left and right eigenvectors v and w are usually normalized so that the sum of their components is equal to 1; in this case, they are sometimes called stochastic eigenvectors.

Non-negative matrices
An extension of the theorem to matrices with non-negative entries is also available. In order to highlight the similarities and differences between the two cases the following points are to be noted: every non-negative matrix can be obviously obtained as a limit of positive matrices, thus one obtains the existence of an eigenvector with non-negative components; obviously the corresponding eigenvalue will be non-negative and greater or equal in absolute value than all other eigenvalues (see Meyer[1] chapter 8.3 [2] page 670 or Gantmacher[7] chapter XIII.3 theorem 3 page 66 [8] for details). However, the simple examples

show that for non-negative matrices there may exist eigenvalues of the same absolute value as the maximal one ((1) and (1) eigenvalues of the first matrix); moreover the maximal eigenvalue may not be a simple root of the characteristic polynomial, can be zero and the corresponding eigenvector (1,0) is not strictly positive (second example). So it may seem that most properties are broken for non-negative matrices, however Frobenius found the right way. The key feature of theory in the non-negative case is to find some special subclass of non-negative matrices irreducible matrices for which a non-trivial generalization is possible. Namely, although eigenvalues attaining

PerronFrobenius theorem maximal absolute value may not be unique, the structure of maximal eigenvalues is under control: they have the form ei2l/hr, where h is some integer number period of matrix, r is a real strictly positive eigenvalue, l=0,1,...,h1. The eigenvector corresponding to r has strictly positive components (in contrast with the general case of non-negative matrices, where components are only non-negative). Also all such eigenvalues are simple roots of the characteristic polynomial. Further properties are described below. Classification of matrices Let A be a square matrix (not necessarily positive or even real). The matrix A is irreducible if any of the following equivalent properties holds. Definition 1 : A does not have non-trivial invariant coordinate subspaces. Here a non-trivial coordinate subspace means a linear subspace spanned by any proper subset of basis vectors. More explicitly, for any linear subspace spanned by basis vectors ei1 , ..., eik, n> k>0 its image under the action of A is not contained in the same subspace. Definition 2: A cannot be conjugated into block upper triangular form by a permutation matrix P:

174

where E and G are non-trivial (i.e. of size greater than zero) square matrices. If A is non-negative other definitions exist: Definition 3: For every pair of indices i and j, there exists a natural number m such that (Am)ij is not equal to 0. Definition 4: One can associate with a matrix A a certain directed graph GA. It has exactly n vertices, where n is size of A, and there is an edge from vertex i to vertex j precisely when Aij > 0. Then the matrix A is irreducible if and only if its associated graph GA is strongly connected. A matrix is reducible if it is not irreducible. Let A be non-negative. Fix an index i and define the period of index i to be the greatest common divisor of all natural numbers m such that (Am)ii > 0. When A is irreducible, the period of every index is the same and is called the period of A. When A is also irreducible, the period can be defined as the greatest common divisor of the lengths of the closed directed paths in GA (see Kitchens[9] page 16). The period is also called the index of imprimitivity (Meyer[1] page 674) or the order of cyclicity. If the period is 1, A is aperiodic. A matrix A is primitive if it is non-negative and its mth power is positive for some natural number m (i.e. the same m works for all pairs of indices). It can be proved that primitive matrices are the same as irreducible aperiodic non-negative matrices. A positive square matrix is primitive and a primitive matrix is irreducible. All statements of the PerronFrobenius theorem for positive matrices remain true for primitive matrices. However, a general non-negative irreducible matrix A may possess several eigenvalues whose absolute value is equal to the spectral radius of A, so the statements need to be correspondingly modified. Actually the number of such eigenvalues is exactly equal to the period. Results for non-negative matrices were first obtained by Frobenius in 1912.

PerronFrobenius theorem PerronFrobenius theorem for irreducible matrices Let A be an irreducible non-negative nn matrix with period h and spectral radius (A)=r. Then the following statements hold. 1. The number r is a positive real number and it is an eigenvalue of the matrix A, called the PerronFrobenius eigenvalue. 2. The PerronFrobenius eigenvalue r is simple. Both right and left eigenspaces associated with r are one-dimensional. 3. A has a left eigenvector v with eigenvalue r whose components are all positive. 4. Likewise, A has a right eigenvector w with eigenvalue r whose components are all positive. 5. The only eigenvectors whose components are all positive are those associated with the eigenvalue r. 6. Matrix A has exactly h (where h is the period) complex eigenvalues with absolute value r. Each of them is a simple root of the characteristic polynomial and is the product of r with an hth root of unity. 7. Let = 2/h. Then the matrix A is similar to eiA, consequently the spectrum of A is invariant under multiplication by ei (corresponding to the rotation of the complex plane by the angle ). 8. If h > 1 then there exists a permutation matrix P such that

175

where the blocks along the main diagonal are zero square matrices. 9. CollatzWielandt formula: for all non-negative non-zero vectors x let f(x) be the minimum value of [Ax]i / xi taken over all those i such that xi 0. Then f is a real valued function whose maximum is the PerronFrobenius eigenvalue. 10. The PerronFrobenius eigenvalue satisfies the inequalities

The matrix

shows that the blocks on the diagonal may be of different sizes, the matrices Aj need not

be square, and h need not dividen.

Further properties
Let A be an irreducible non-negative matrix, then: 1. (1+A)n1 is a positive matrix. (Meyer[1] claim 8.3.5 p. 672 [2]). 2. Wielandt's theorem. If |B|<A, then (B)(A). If equality holds (i.e. if =(A)ei is eigenvalue for B), then B = ei D AD1 for some diagonal unitary matrix D (i.e. diagonal elements of D equals to eil, non-diagonal are zero). (Meyer[1] claim 8.3.11 p. 675 [2]). 3. If some power Aq is reducible, then it is completely reducible, i.e. for some permutation matrix P, it is true that:

, where Ai are irreducible matrices having the same maximal

eigenvalue. The number of these matrices d is the greatest common divisor of q and h, where h is period of A. (Gantmacher[7] section XIII.5 theorem 9).

PerronFrobenius theorem 4. If c(x)=xn+ck1 xn-k1 +ck2 xn-k2 + ... + cks xn-ks is the characteristic polynomial of A in which the only nonzero coefficients are listed, then the period of A equals to the greatest common divisor for k1, k2, ... , ks.(Meyer[1] page 679 [2]). 5. Cesro averages: where the left and right eigenvectors for A are normalized

176

so that wtv = 1. Moreover the matrix v wt is the spectral projection corresponding to r - Perron projection. (Meyer[1] example 8.3.2 p. 677 [2]). 6. Let r be the Perron-Frobenius eigenvalue, then the adjoint matrix for (r-A) is positive (Gantmacher[7] section XIII.2.2 page 62 [10]). 7. If A has at least one non-zero diagonal element, then A is primitive. (Meyer[1] example 8.3.3 p. 678 [2]). The following facts worth to be mentioned. If 0 A < B, the rA rB, moreover, if A is irreducible, then the inequality is strict: rA < rB. One of the definitions of primitive matrix requires A to be non-negative and there exists m, such that Am is positive. One may one wonder how big m can be, depending on the size of A. The following answers this question. Assume A is non-negative primitive matrix of size n, then An2-2n+2 is positive. Moreover there exists a matrix M given below, such that Mk remains not positive (just non-negative) for all k< n2-2n+2, in particular (Mn2-2n+1)11=0.

(Meyer[1] chapter 8 [2] example 8.3.4 page 679 and exercise 8.3.9 p.685).

Applications
Numerous books have been written on the subject of non-negative matrices, and PerronFrobenius theory is invariably a central feature. So there is a vast application area and the examples given below barely begin to scratch its surface.

Non-negative matrices
The PerronFrobenius theorem does not apply directly to non-negative matrices. Nevertheless any reducible square matrix A may be written in upper-triangular block form (known as the normal form of a reducible matrix)[11] PAP1 = where P is a permutation matrix and each Bi is a square matrix that is either irreducible or zero. Now if A is non-negative then so are all the Bi and the spectrum of A is just the union of their spectra. Therefore many of the spectral properties of A may be deduced by applying the theorem to the irreducible Bi. For example the Perron root is the maximum of the (Bi). Whilst there will still be eigenvectors with non-negative components it is quite possible that none of these will be positive.

PerronFrobenius theorem

177

Stochastic matrices
A row (column) stochastic matrix is a square matrix each of whose rows (columns) consists of non-negative real numbers whose sum is unity. The theorem cannot be applied directly to such matrices because they need not be irreducible. If A is row-stochastic then the column vector with each entry 1 is an eigenvector corresponding to the eigenvalue 1, which is also (A) by the remark above. It may not be the only eigenvalue on the unit circle: and the associated eigenspace can be multi-dimensional. If A is row-stochastic and irreducible then the Perron projection is also row-stochastic and all its rows are equal.

Algebraic graph theory


The theorem has particular use in algebraic graph theory. The "underlying graph" of a nonnegative n-square matrix is the graph with vertices numbered 1, ..., n and arc ij if and only if Aij 0. If the underlying graph of such a matrix is strongly connected, then the matrix is irreducible, and thus the theorem applies. In particular, the adjacency matrix of a strongly connected graph is irreducible.[12] [13]

Finite Markov chains


The theorem has a natural interpretation in the theory of finite Markov chains (where it is the matrix-theoretic equivalent of the convergence of an irreducible finite Markov chain to its stationary distribution, formulated in terms of the transition matrix of the chain; see, for example, the article on the subshift of finite type).

Compact operators
More generally, it can be extended to the case of non-negative compact operators, which, in many ways, resemble finite-dimensional matrices. These are commonly studied in physics, under the name of transfer operators, or sometimes RuellePerronFrobenius operators (after David Ruelle). In this case, the leading eigenvalue corresponds to the thermodynamic equilibrium of a dynamical system, and the lesser eigenvalues to the decay modes of a system that is not in equilibrium. Thus, the theory offers a way of discovering the arrow of time in what would otherwise appear to be reversible, deterministic dynamical processes, when examined from the point of view of point-set topology.[14]

Proof methods
A common thread in many proofs is the Brouwer fixed point theorem. Another popular method is that of Wielandt (1950). He used the CollatzWielandt formula described above to extend and clarify Frobenius's work (see Gantmacher[7] section XIII.2.2 page 54 [15] ). The proof based on the spectral theory can be found in[16] from which part of the arguments are borrowed.

Perron root is strictly maximal eigenvalue for positive (and primitive) matrices
If A is a positive (or more generally primitive) matrix, then there exists real positive eigenvalue r (Perron-Frobenius eigenvalue or Perron root), which is strictly greater in absolute value than all other eigenvalues, hence r is the spectral radius of A. Pay attention that claim is wrong for general non-negative irreducible matrices, which have h eigenvalues with the same absolute eigenvalue as r, where h is the period of A. Proof. Consider first the case of positive matrices. Let A be a positive matrix, assume that its spectral radius (A) = 1 (otherwise consider A/(A)). Hence, as it is known, there exists an eigenvalue on the unit circle, and all the other eigenvalues are less or equal 1 in absolute value. Assume that that 1. Then there exists a positive integer m such that Am is a positive matrix and the real part of m is negative. Let be half the smallest diagonal entry of Am and set

PerronFrobenius theorem T = Am1 which is yet another positive matrix. Moreover if Ax = x then Amx = mx thus m is an eigenvalue of T. Because of the choice of m this point lies outside the unit disk consequently (T) > 1. On the other hand all the entries in T are positive and less than or equal to those in Am so by Gelfand's formula (T) (Am) (A)m = 1. This contradiction means that =1 and there can be no other eigenvalues on the unit circle. Proof for positive matrices is finished . Absolutely the same arguments can be applied to the case of primitive matrices, after one just need to mention the following simple lemma, which clarifies the properties of primitive matrices Lemma: Consider a non-negative A. Assume there exists m, such that Am is positive, then Am+1, Am+2, Am+3,... are all positive. Indeed, Am+1= A Am, so it can have zero element only if some row of A is entirely zero, but in this case the same row of Am will be zero. Lemma is proved. After lemma is established one applies the same arguments as above for primitive matrices to prove the main claim.

178

Power method and the positive eigenpair


for a positive (or more generally irreducible non-negative) matrix A the dominant eigenvector is real and strictly positive (for non-negative A respectively non-negative) It can be argued by the power method, which states that for sufficiently generic (in the sense discussed below) matrix A the sequence of vectors bk+1=Abk / | Abk | converges to the eigenvector with the maximum eigenvalue. (Initial vector b0 can be chosen arbitrary except some measure zero set). Starting with a non-negative vector b0 one gets the sequence of non-negative vectors bk. Hence the limiting vector is also non-negative. By power method this limiting vector is the desired the dominant eigenvector for A, so assertion is proved. Corresponding eigenvalue is clearly non-negative. To accomplish the proof two arguments should be added. First, the power method converges for matrices which does not have several eigenvalues of the same absolute value as the maximal one. The previous section argument guarantees this. Second, is to ensure strict positivity of all of the components of the eigenvector for the case of irreducible matrices. This follows from the following simple fact, which is of independent interest: Lemma: consider a positive (or more generally irreducible non-negative) matrix A. Assume v is any non-negative eigenvector for A, then it is necessarily strictly positive and corresponding eigenvalue is also strictly positive. Proof. Recall that one of the definitions of irreducibility for non-negative matrices is that for all indexes i,j there exists m, such that (Am)ij is strictly positive. Consider a non-negative eigenvector v, assume that at least one of its components say j-th is strictly positive. Then one can deduce that the corresponding eigenvalue is strictly positive, indeed, consider n such that (An)ii >0, hence: rnvi = Anvi >= (An)iivi >0. Hence r is strictly positive. In the same manner one can deduce strict positivity of the eigenvector. To prove it, consider m, such that (Am)ij >0, hence: rmvj = (Amv)j >= (Am)ijvi >0, hence vj is strictly positive, i.e. eigenvector is strictly positive. Proof is finished.

PerronFrobenius theorem

179

Multiplicity one
The proof that the Perron-Frobenius eigenvalue is simple root of the characteristic polynomial is also elementary. The arguments here are close to that ones in Meyer[1] chapter 8 [2] page 665, where details can be found. Eigenspace associated to Perron-Frobenius eigenvalue r is one-dimensional Consider strictly positive eigenvector v corresponding to r. Assume there exists another eigenvector w with the same eigenvalue. (Vector w can be chosen to be real, because A and r are both real, so the null space of A-r has a basis consisting of real vectors). Assume at least one of the components of w is positive (otherwise multiply w by -1). Consider maximal possible such that u=v- w is non-negative. Then clearly one of the components of u is zero, otherwise is not maximum. Obviously vector u is an eigenvector. It is non-negative, hence by the lemma described in the previous section non-negativity implies strict positivity for any eigenvector. On the other hand it was stated just above that at least one component of u is zero. The contradiction implies, that w does not exist. The claim is proved. There is no Jordan cells corresponding to the Perron-Frobenius eigenvalue r and all other eigenvalues which has the same absolute value The idea is the following: if there is a Jordan cell, then the infinity norm ||(A/r)k|| tends to infinity for k , but it will actually contradict the existence of the positive eigenvector. The details are the following. Assume r=1, otherwise consider A/r. Let v be Perron-Frobenius strictly positive eigenvector, so Av=v, then: So ||Ak|| is bounded for all k. Actually it gives another proof that there is no eigenvalues which has greater absolute value then Perron-Frobenius one. And it gives more. It contradicts the existence of the Jordan cell for any eigenvalue which has absolute value equal to 1 (in particular for the Perron-Frobenius one), because existence of the Jordan cell implies that ||Ak|| is unbounded. For example for two by two matrix:

hence ||Jk|| = |k+| (for ||=1), so it tends to infinity when k does so. Since ||Jk|| = ||C-1 AkC ||, then || Ak || >= ||Jk||/ ( ||C1|| || C ||), so it also tends to infinity. Obtained contradiction implies that there is no Jordan cells for the corresponding eigenvalues. Proof is finished. Combining the two claims above together one gets: the Perron-Frobenius eigenvalue r is simple root of the characteristic polynomial. Actually in the case of non primitive matrices, there exists other eigenvalues which has the same absolute value as r. Same claim is true for them, but it requires more work.

No other non-negative eigenvectors


Consider positive (or more generally irreducible non-negative matrix) A. The Perron-Frobenius eigenvector is the only (up to multiplication by constant) non-negative eigenvector for A. So other eigenvectors should contain negative, or complex components. The idea of proof is the following eigenvectors for different eigenvalues should be orthogonal in some sense, however two positive eigenvectors cannot be orthogonal, so they correspond the same eigenvalue, but eigenspace for the Perron-Frobenius is one dimensional. The formal proof goes as follows. Assume there exists an eigenpair (, y) for A, such that vector y is positive. Consider (r, x), where x - is the right Perron-Frobenius eigenvector for A (i.e. eigenvector for At). Then: r xt y = (xt A) y= xt (A y)= xt y, also xt y >0, so one has: r=. But eigenspace for the Perron-Frobenius eigenvalue r is one dimensional, as it was established before. Hence non-negative eigenvector y is multiple of the Perron-Frobenius one. Assertion is proved.

PerronFrobenius theorem One can consult Meyer[1] chapter 8 [2] claim 8.2.10 page 666 for details.

180

CollatzWielandt formula
Consider positive (or more generally irreducible non-negative matrix) A. For all non-negative non-zero vectors x let f(x) be the minimum value of [Ax]i / xi taken over all those i such that xi 0. Then f is a real valued function whose maximum is the PerronFrobenius eigenvalue r. First observe that r is attained for x taken to be the Perron-Frobenius eigenvector v. So one needs to prove that values f on the other vectors are less or equal. Consider some vector x. Let =f(x), so 0<= x <=Ax. Consider w to be the right eigenvector for A. Hence wt x <= wt (Ax) = (wt A)x = r wt x . Hence <=r. So proof is finished. One can consult Meyer[1] chapter 8 [2] page 666 for details.

Perron projection as a limit: Ak/rk


Consider positive (or more generally primitive) matrix A, and r its Perron-Frobenius eigenvalue, then: 1. There exists a limit Ak/rk for k , denote it by P. 2. P is a projection operator: P2=P, which commutes with A: AP=PA. 3. The image of P is one-dimensional and spanned by the Perron-Frobenius eigenvector v (respectively for Pt by the Perron-Frobenius eigenvector w for At). 4. P= v wt, where v,w are normalized such that w^t v = 1. 5. Hence P is a positive operator. Hence P is a spectral projection for the Perron-Frobenius eigenvalue r, and it is called Perron projection. Pay attention: the assertion above is not true for general non-negative irreducible matrices. Actually the claims above (except claim 5) are valid for any matrix M such that there exists an eigenvalue r, which is strictly greater than the other eigenvalues in absolute value and it is simple root of the characteristic polynomial. (These requirements hold for primitive matrices as it was shown above). The proof of this more general fact is sketched below. Proof. One can demonstrate the existence of the limit as follows. Assume M is diagonalizable, so it is conjugate to diagonal matrix with eigenvalues r1, ... , rn on the diagonal (denote r1=r). The matrix Mk/rk will be conjugate (1, (r2/r)k, ... , (rn/r)k), which tends to (1,0,0,...,0), for k . So the limit exists. The same method works for general M (without assumption M is diagonalizable). The projection and commutativity properties are elementary corollaries of the definition: M Mk/rk= Mk/rk M ; P2 = lim M2k/r2k=P. The third fact is also elementary: M (P u)= M lim Mk/rk u = lim r Mk+1/rk+1 u, so taking the limit one gets M (P u)=r (P u), so image of P lies in the r-eigenspace for M, which is one-dimensional by the assumptions. Denote by v r-eigenvector for M (by w for Mt). Columns of P are multiples of v, because image of P is spanned by it. Respectively rows - of w. So P takes a form (a v wt), for some a. Hence its trace equals to (a wt v). On the other hand trace of projector equals to the dimension of its image. It was proved before that it is not more than one-dimensional. From the definition one sees that P acts identically on the r-eigenvector for M. So it is one-dimensional. So one concludes that choosing (wt v)=1, implies P=v wt. The proof is finished.

PerronFrobenius theorem

181

Inequalities for PerronFrobenius eigenvalue


For any non-nonegative matrix A its PerronFrobenius eigenvalue r satisfies the inequality:

Actually it is not specific to non-negative matrices: for any matrix A and any its eigenvalue it is true that . This is immediate corollary of the Gershgorin circle theorem. However there is another proof which is more direct. Any matrix induced norm satisfies the inequality ||A|| || for any eigenvalue , because ||A|| ||Ax||/||x|| = ||x||/||x|| = ||. The infinity norm of a matrix is the maximum of row sums: Hence the desired inequality is exactly ||A|| || applied to non-negative matrix A. Another inequality is:

This fact is specific for non-negative matrices, for general matrices there is nothing similar. To prove it one first suppose that A is positive (non just non-negative). Then there exists a positive eigenvector w such that Aw = rw and the smallest component of w (say wi) is 1. Then r = (Aw)i the sum of the numbers in row i of A. Thus the minimum row sum gives a lower bound for r and this observation can be extended to all non-negative matrices by continuity. Another way to argue it is via the Collatz-Wielandt formula. One takes the vector x=(1,1,...,1) and immediately obtains the inequality.

Further proofs
Perron projection The proof now proceeds using spectral decomposition. The trick here is to split the Perron root away from the other eigenvalues. The spectral projection associated with the Perron root is called the Perron projection and it enjoys the following property: The Perron projection of an irreducible non-negative square matrix is a positive matrix. Perron's findings and also (1)(5) of the theorem are corollaries of this result. The key point is that a positive projection always has rank one. This means that if A is an irreducible non-negative square matrix then the algebraic and geometric multiplicities of its Perron root are both one. Also if P is its Perron projection then AP = PA = (A)P so every column of P is a positive right eigenvector of A and every row is a positive left eigenvector. Moreover if Ax = x then PAx = Px = (A)Px which means Px = 0 if (A). Thus the only positive eigenvectors are those associated with (A). And there is yet more to come. If A is a primitive matrix with (A) = 1 then it can be decomposed as P (1P)A so that An = P + (1P)An. As n increases the second of these terms decays to zero leaving P as the limit of An as n. The power method is a convenient way to compute the Perron projection of a primitive matrix. If v and w are the positive row and column vectors that it generates then the Perron projection is just wv/vw. It should be noted that the spectral projections aren't neatly blocked as in the Jordan form. Here they are overlaid on top of one another and each generally has complex entries extending to all four corners of the square matrix. Nevertheless they retain their mutual orthogonality which is what facilitates the decomposition.

PerronFrobenius theorem Peripheral projection The analysis when A is irreducible and non-negative is broadly similar. The Perron projection is still positive but there may now be other eigenvalues of modulus (A) that negate use of the power method and prevent the powers of (1P)A decaying as in the primitive case whenever (A) = 1. So enter the peripheral projection which is the spectral projection of A corresponding to all the eigenvalues that have modulus (A) .... The peripheral projection of an irreducible non-negative square matrix is a non-negative matrix with a positive diagonal. Cyclicity Suppose in addition that (A) = 1 and A has h eigenvalues on the unit circle. If P is the peripheral projection then the matrix R = AP = PA is non-negative and irreducible, Rh = P, and the cyclic group P, R, R2, ...., Rh1 represents the harmonics of A. The spectral projection of A at the eigenvalue on the unit circle is given by the formula . All of these projections (including the Perron projection) have the same positive diagonal, moreover choosing any one of them and then taking the modulus of every entry invariably yields the Perron projection. Some donkey work is still needed in order to establish the cyclic properties (6)(8) but it's essentially just a matter of turning the handle. The spectral decomposition of A is given by A=R(1P)A so the difference between An and Rn is AnRn =(1P)An representing the transients of An which eventually decay to zero. P may be computed as the limit of Anh as n.

182

Caveats
The matrices L = ,P= ,T= ,M= provide simple examples of what

can go wrong if the necessary conditions are not met. It is easily seen that the Perron and peripheral projections of L are both equal to P, thus when the original matrix is reducible the projections may lose non-negativity and there is no chance of expressing them as limits of its powers. The matrix T is an example of a primitive matrix with zero diagonal. If the diagonal of an irreducible non-negative square matrix is non-zero then the matrix must be primitive but this example demonstrates that the converse is false. M is an example of a matrix with several missing spectral teeth. If = ei/3 then 6 = 1 and the eigenvalues of M are {1,2,3,4} so and 5 are both absent.

Terminology
A problem that causes confusion is a lack of standardisation in the definitions. For example, some authors use the terms strictly positive and positive to mean > 0 and 0 respectively. In this article positive means > 0 and non-negative means 0. Another vexed area concerns decomposability and reducibility: irreducible is an overloaded term. For avoidance of doubt a non-zero non-negative square matrix A such that 1+A is primitive is sometimes said to be connected. Then irreducible non-negative square matrices and connected matrices are synonymous.[17] The nonnegative eigenvector is often normalized so that the sum of its components is equal to unity; in this case, the eigenvector is a the vector of a probability distribution and is sometimes called a stochastic eigenvector. PerronFrobenius eigenvalue and dominant eigenvalue are alternative names for the Perron root. Spectral projections are also known as spectral projectors and spectral idempotents. The period is sometimes referred to as the index of imprimitivity or the order of cyclicity.

PerronFrobenius theorem

183

Notes
[1] Meyer, Carl (2000), Matrix analysis and applied linear algebra (http:/ / www. matrixanalysis. com/ Chapter8. pdf), SIAM, ISBN0-89871-454-0, [2] http:/ / www. matrixanalysis. com/ Chapter8. pdf [3] Langville, Amy; Meyer, Carl (2006), Google page rank and beyond (http:/ / pagerankandbeyond. com), Princeton University Press, ISBN0-691-12202-4, [4] http:/ / books. google. com/ books?id=hxvB14-I0twC& lpg=PP1& dq=isbn%3A0691122024& pg=PA167#v=onepage& q& f=false [5] Keener, James (1993), "The PerronFrobenius theorem and the ranking of football teams" (http:/ / links. jstor. org/ sici?sici=0036-1445(199303)35:1<80:TPTATR>2. 0. CO;2-O), SIAM Review (SIAM) 35 (1): 8093, [6] http:/ / links. jstor. org/ sici?sici=0036-1445%28199303%2935%3A1%3C80%3ATPTATR%3E2. 0. CO%3B2-O [7] Gantmacher, Felix (2000) [1959], The Theory of Matrices, Volume 2 (http:/ / books. google. com/ books?id=cyX32q8ZP5cC& lpg=PA178& vq=preceding section& pg=PA53#v=onepage& q& f=true), AMS Chelsea Publishing, ISBN0-8218-2664-6, (1959 edition had different title: "Applications of the theory of matrices". Also the numeration of chapters is different in the two editions.) [8] http:/ / books. google. com/ books?id=cyX32q8ZP5cC& lpg=PA178& vq=preceding%20section& pg=PA66#v=onepage& q& f=false [9] Kitchens, Bruce (1998), Symbolic dynamics: one-sided, two-sided and countable state markov shifts. (http:/ / books. google. ru/ books?id=mCcdC_5crpoC& lpg=PA195& ots=RbFr1TkSiY& dq=kitchens perron frobenius& pg=PA16#v=onepage& q& f=false), Springer, [10] http:/ / books. google. com/ books?id=cyX32q8ZP5cC& lpg=PA178& vq=preceding%20section& pg=PA62#v=onepage& q& f=true [11] See 2.43 (page 51) in Varga reference below [12] Brualdi, Richard A.; Ryser, Herbert John (1992). Combinatorial Matrix Theory. Cambridge: Cambridge UP. ISBN0521322650. [13] Brualdi, Richard A.; Cvetkovic, Dragos (2009). A Combinatorial Approach to Matrix Theory and Its Applications. Boca Raton, FL: CRC Press. ISBN9781420082234. [14] Mackey, Michael C. (1992). Time's Arrow: The origins of thermodynamic behaviour. New York: Springer-Verlag. ISBN0387977023. [15] http:/ / books. google. ru/ books?id=cyX32q8ZP5cC& lpg=PR5& dq=Applications%20of%20the%20theory%20of%20matrices& pg=PA54#v=onepage& q& f=false [16] Smith, Roger (2006), "A Spectral Theoretic Proof of PerronFrobenius" (ftp:/ / emis. maths. adelaide. edu. au/ pub/ EMIS/ journals/ MPRIA/ 2002/ pa102i1/ pdf/ 102a102. pdf), Mathematical Proceedings of the Royal Irish Academy (The Royal Irish Academy) 102 (1): 2935, doi:10.3318/PRIA.2002.102.1.29, [17] For surveys of results on irreducibility, see Olga Taussky-Todd and Richard A. Brualdi.

References
Original papers
Perron, Oskar (1907), "Zur Theorie der Matrices", Mathematische Annalen 64 (2): 248263, doi:10.1007/BF01449896 Frobenius, Georg (1912), "Ueber Matrizen aus nicht negativen Elementen", Sitzungsber. Knigl. Preuss. Akad. Wiss.: 456477 Frobenius, Georg (1908), "ber Matrizen aus positiven Elementen, 1", Sitzungsber. Knigl. Preuss. Akad. Wiss.: 471476 Frobenius, Georg (1909), "ber Matrizen aus positiven Elementen, 2", Sitzungsber. Knigl. Preuss. Akad. Wiss.: 514518 Romanovsky, V. (1933), "Sur les zros des matrices stocastiques" (http://www.numdam.org/ item?id=BSMF_1933__61__213_0), Bulletin de la Socit Mathmatique de France 61: 213219 Collatz, Lothar (1942), "Einschlieungssatz fr die charakteristischen Zahlen von Matrize", Mathematische Zeitschrift 48 (1): 221226, doi:10.1007/BF01180013 Wielandt, Helmut (1950), "Unzerlegbare, nicht negative Matrizen", Mathematische Zeitschrift 52 (1): 642648, doi:10.1007/BF02230720

PerronFrobenius theorem

184

Further reading
Abraham Berman, Robert J. Plemmons, Nonnegative Matrices in the Mathematical Sciences, 1994, SIAM. ISBN 0-89871-321-8. Chris Godsil and Gordon Royle, Algebraic Graph Theory, Springer, 2001. A. Graham, Nonnegative Matrices and Applicable Topics in Linear Algebra, John Wiley&Sons, New York, 1987. R. A. Horn and C.R. Johnson, Matrix Analysis, Cambridge University Press, 1990 S. P. Meyn and R. L. Tweedie, Markov Chains and Stochastic Stability (https://netfiles.uiuc.edu/meyn/www/ spm_files/book.html) London: Springer-Verlag, 1993. ISBN 0-387-19832-6 (2nd edition, Cambridge University Press, 2009) Henryk Minc, Nonnegative matrices, John Wiley&Sons, New York, 1988, ISBN 0-471-83966-3 Seneta, E. Non-negative matrices and Markov chains. 2nd rev. ed., 1981, XVI, 288 p., Softcover Springer Series in Statistics. (Originally published by Allen & Unwin Ltd., London, 1973) ISBN 978-0-387-29765-1 Suprunenko, D.A. (2001), "PerronFrobenius theorem" (http://eom.springer.de/P/p072350.htm), in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 (The claim that Aj has order n/h at the end of the statement of the theorem is incorrect.) Richard S. Varga, Matrix Iterative Analysis, 2nd ed., Springer-Verlag, 2002

PoincarBirkhoffWitt theorem
In the theory of Lie algebras, the PoincarBirkhoffWitt theorem (Poincar (1900), G. D. Birkhoff (1937), Witt (1937); frequently contracted to PBW theorem) is a result giving an explicit description of the universal enveloping algebra of a Lie algebra. The term 'PBW type theorem' or even 'PBW theorem' may also refer to various analogues of the original theorem, comparing a filtered algebra to its associated graded algebra, in particular, in the area of quantum groups.

Statement of the theorem


Recall that any vector space V over a field has a basis; this is a set S such that any element of V is a unique (finite) linear combination of elements of S. In the formulation of PoincarBirkhoffWitt theorem we consider bases of which the elements are totally ordered by some relation which we denote . If L is a Lie algebra over a field K, there is a canonical K-linear map h from L into the universal enveloping algebra U(L). Theorem. Let L be a Lie algebra over K and X a totally ordered basis of L. A canonical monomial over X is a finite sequence (x1, x2 ..., xn) of elements of X which is non-decreasing in the order , that is, x1 x2 ... xn. Extend h to all canonical monomials as follows: If (x1, x2, ..., xn) is a canonical monomial, let Then h is injective on the set of canonical monomials and its range is a basis of the K-vector space U(L). Stated somewhat differently, consider Y = h(X). Y is totally ordered by the induced ordering from X. The set of monomials

where y1 <y2 < ... < yn are elements of Y, and the exponents are non-negative, together with the multiplicative unit 1, form a basis for U(L). Note that the unit element 1 corresponds to the empty canonical monomial. The multiplicative structure of U(L) is determined by the structure constants in the basis X, that is, the coefficients cu,v,x such that

PoincarBirkhoffWitt theorem

185

This relation allows one to reduce any product of y's to a linear combination of canonical monomials: The structure constants determine yiyj yjyi, i.e. what to do in order to change the order of two elements of Y in a product. This fact, modulo an inductive argument on the degree of (non-canonical) monomials, shows one can always achieve products where the factors are ordered in a non-decreasing fashion. The PoincarBirkhoffWitt theorem can be interpreted as saying that the end result of this reduction is unique and does not depend on the order in which one swaps adjacent elements. Corollary. If L is a Lie algebra over a field, the canonical map L U(L) is injective. In particular, any Lie algebra over a field is isomorphic to a Lie subalgebra of an associative algebra.

More general contexts


Already at its earliest stages, it was known that K could be replaced by any commutative ring, provided that L is a free K-module, i.e., has a basis as above. To extend to the case when L is no longer a free K-module, one needs to make a reformulation that does not use bases. This involves replacing the space of monomials in some basis with the Symmetric algebra, S(L), on L. In the case that K contains the field of rational numbers, one can consider the natural map from S(L) to U(L), sending a monomial . for , to the element . Then, one has the theorem

that this map is an isomorphism of K-modules. Still more generally and naturally, one can consider U(L) as a filtered algebra, equipped with the filtration given by specifying that lies in filtered degree . The map of K-modules canonically extends to a map (actually, morphism of algebras, where is the tensor algebra on L (for example, by the universal with the filtration putting L in degree one , and hence descends to a canonical property of tensor algebras), and this is a filtered map equipping , which kills the elements for

is graded). Then, passing to the associated graded, one gets a canonical morphism . Then, the (graded) PBW theorem can be reformulated as the statement that, under

certain hypotheses, thisK andmorphism isexample, the last section of Cohn's 1961 paper), but is true in many cases. This is not true for all final L (see, for an isomorphism. These include the aforementioned ones, where either L is a free K-module, or K contains the field of rational numbers. More generally, the PBW theorem as formulated above extends to cases such as where (1) L is a flat K-module, (2) L is torsion-free as an abelian group, (3) L is a direct sum of cyclic modules (or all its localizations at prime ideals of K have this property), or (4) K is a Dedekind domain. See, for example, the 1969 paper by Higgins for these statements. Finally, it is worth noting that, in some of these cases, one also obtains the stronger statement that the canonical morphism lifts to a K-module isomorphism , without taking associated graded. This is true in the first cases mentioned, where L is a free K-module, or K contains the field of rational numbers, using the construction outlined here (in fact, the result is a coalgebra isomorphism, and not merely a K-module isomorphism, equipping both S(L) and U(L) with their natural coalgebra structures such that for ). This stronger statement, however, might not extend to all of the cases in the previous paragraph.

PoincarBirkhoffWitt theorem

186

History of the theorem


Ton-That and Tran have investigated the history of the theorem. They have found out that the majority of the sources before Bourbaki's 1960 book call it Birkhoff-Witt theorem. Following this old tradition, Fofanova in her encyclopaedic entry says that H. Poincar obtained the first variant of the theorem. She further says that the theorem was subsequently completely demonstrated by E. Witt and G.D. Birkhoff. It appears that pre-Bourbaki sources were not familiar with Poincar's paper. Birkhoff and Witt do not mention Poincar's work in their 1937 papers. Cartan and Eilenberg in their 1956 book call the theorem Poincar-Witt Theorem and attribute the complete proof to Witt. Bourbaki were the first to use all three names in their 1960 book. Knapp presents a clear illustration of the shifting tradition. In his 1986 book he calls it Birkhoff-Witt Theorem while in his later 1996 book he switches to Poincar-Birkhoff-Witt Theorem. It is not clear whether Poincar's result was complete. Ton-That and Tran conclude that Poincar had discovered and completely demonstrated this theorem at least thirty-seven years before Witt and Birkhoff. On the other hand, they point out that Poincar makes several statements without bothering to prove them. Their own proofs of all the steps are rather long according to their admission.

References
G.D. Birkhoff, Representability of Lie algebras and Lie groups by matrices Ann. of Math. (2) , 38 : 2 (1937) pp. 526532 T.S. Fofanova (2001), "BirkhoffWitt theorem" [1], in Hazewinkel, Michiel, Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 P.M. Cohn, A remark on the Birkho-Witt theorem, J. London Math. Soc. 38, 197-203, (1963) P.J. Higgins, Baer Invariants and the Birkhoff-Witt theorem, J. of Alg. 11, 469-482, (1969) G. Hochschild, The Theory of Lie Groups, Holden-Day, 1965. H. Poincar, Sur les groupes continus Trans. Cambr. Philos. Soc. , 18 (1900) pp. 220225 E. Witt, Treue Darstellung Liescher Ringe [2] J. Reine Angew. Math. , 177 (1937) pp. 152160 T. Ton-That, T.-D. Tran, Poincar's proof of the so-called Birkhoff-Witt theorem [3] Rev. Histoire Math., 5 (1999), pp. 249-284. N. Bourbaki, lments de mathmatique. XXVI. Groupes et algbres de Lie. Chapitre 1: Algbres de Lie. Hermann, Paris, 1960. A. W. Knapp, Representation theory of semisimple groups. An overview based on examples. Princeton University Press, 1986. A. W. Knapp, Lie groups beyond an introduction. Birkhuser Boston, 1996.

References
[1] http:/ / eom. springer. de/ B/ b016540. htm [2] http:/ / gdz. sub. uni-goettingen. de/ dms/ load/ img/ ?PPN=PPN243919689_0177& DMDID=dmdlog17 [3] http:/ / smf. emath. fr/ Publications/ RevueHistoireMath/ 5/ pdf/ smf_rhm_5_249-284. pdf

Polynomial remainder theorem

187

Polynomial remainder theorem


In algebra, the polynomial remainder theorem or little Bzout's theorem[1] is an application of polynomial long division. It states that the remainder of a polynomial divided by a linear divisor is equal to

Example
Let the remainder . Polynomial division of . Therefore, . by gives the quotient and

Proof
The polynomial remainder theorem follows from the definition of polynomial long division; denoting the divisor, quotient and remainder by, respectively, , , and , polynomial long division gives a solution of the equation where the degree of If we take Setting we obtain: is less than that of . as 0, i.e. :

as the divisor, giving the degree of

Applications
The polynomial remainder theorem may be used to evaluate by calculating the remainder, . Although polynomial long division is more difficult than evaluating the function itself, synthetic division is computationally easier. Thus, the function may be more "cheaply" evaluated using synthetic division and the polynomial remainder theorem. The factor theorem is another application of the remainder theorem: if the remainder is zero, then the linear divisor is a factor. Repeated application of the factor theorem may be used to factorize the polynomial.

References
[1] Piotr Rudnicki (2004). "Little Bzout Theorem (Factor Theorem)" (http:/ / mizar. org/ fm/ 2004-12/ pdf12-1/ uproots. pdf). Formalized Mathematics 12 (1): 4958. .

Primitive element theorem

188

Primitive element theorem


In mathematics, more specifically in the area of modern algebra known as field theory, the primitive element theorem or Artin's theorem on primitive elements is a result characterizing the finite degree field extensions that possess a primitive element. More specifically, the primitive element theorem characterizes those finite degree extensions such that there exists with .

Terminology
Let when In this situation, the extension in the form where for all i, and is fixed. That is, if is separable of degree n, there exists such that the set is referred to as a simple extension. Then every element x of E can be written be an arbitrary field extension. An element is said to be a primitive element for

is a basis for E as a vector space over F. For instance, the extensions and x, respectively ( and are simple extensions with primitive elements ).

denotes the field of rational functions in the indeterminate x over

Existence statement
The interpretation of the theorem changed with the formulation of the theory of Emil Artin, around 1930. From the time of Galois, the role of primitive elements had been to represent a splitting field as generated by a single element. This (arbitrary) choice of such an element was bypassed in Artin's treatment.[1] At the same time, considerations of construction of such an element receded: the theorem becomes an existence theorem. The following theorem of Artin then takes the place of the classical primitive element theorem. Theorem Let be a finite degree field extension. Then for some element if and only if there

exist only finitely many intermediate fields K with . A corollary to the theorem is then the primitive element theorem in the more traditional sense (where separability was usually tacitly assumed): Corollary Let be a finite degree separable extension. Then for some .

The corollary applies to algebraic number fields, i.e. finite extensions of the rational numbers Q, since Q has characteristic 0 and therefore every extension over Q is separable.

Primitive element theorem

189

Counterexamples
For non-separable extensions, necessarily in characteristic p with p a prime number, then at least when the degree [L:K] is p, then L/K has a primitive element, because there are no intermediate subfields. When [L:K] = p2 then there may not be a primitive element (and therefore there are infinitely many intermediate fields). This happens, for example if K is Fp(T,U), the field of rational functions in two indeterminates T and U over the finite field with p elements, and L is obtained from K by adjoining a p-th root of T, and of U. In fact one can see that for any in L, the element p lies in K, but a primitive element must have degree p2 over K.

Constructive results
Generally, the set of all primitive elements for a finite separable extension L/K is the complement of a finite collection of proper K-subspaces ofL, namely the intermediate fields. This statement says nothing for the case of finite fields, for which there is a computational theory dedicated to finding a generator of the multiplicative group of the field (a cyclic group), which is a fortiori a primitive element. Where K is infinite, a pigeonhole principle proof technique considers the linear subspace generated by two elements and proves that there are only finitely many linear combinations

with c in K in it, that fail to generate the subfield containing both elements. This is almost immediate as a way of showing how Artin's result implies the classical result, and a bound for the number of exceptional c in terms of the number of intermediate fields results (this number being something that can be bounded itself by Galois theory and a priori). Therefore in this case trial-and-error is a possible practical method to find primitive elements. See the Example.

Example
It is not, for example, immediately obvious that if one adjoins to the field Q of rational numbers roots of both polynomials

and

say and respectively, to get a field K=Q(,) of degree 4 over Q, that the extension is simple and there exists a primitive element in K so that K=Q(). One can in fact check that with the powers i for 0i3 can be written out as linear combinations of1,, and with integer coefficients. Taking these as a system of linear equations, or by factoring, one can solve for and overQ() (one gets, for instance, = ), which implies that this choice of is indeed a primitive element in this example. A simpler argument, assuming the knowledge of all the subfields as given by Galois theory, is to note the independence of 1,, and over the rationals; this shows that the subfield generated by cannot be that generated or , nor in fact that generated by , exhausting all the subfields of degree 2. Therefore it must be the whole field.

Primitive element theorem

190

References
The primitive element theorem at mathreference.com [2] The primitive element theorem at planetmath.org [3] The primitive element theorem on the site of Cornell's university (pdf file) [4]

Notes
[1] [2] [3] [4] Israel Kleiner, A History of Abstract Algebra (2007), p. 64. http:/ / www. mathreference. com/ fld-sep,pet. html http:/ / planetmath. org/ encyclopedia/ ProofOfPrimitiveElementTheorem2. html http:/ / www. math. cornell. edu/ ~kbrown/ 4340/ primitive. pdf

QuillenSuslin theorem
The QuillenSuslin theorem, also known as Serre's problem or Serre's conjecture, is a theorem in commutative algebra about the relationship between free modules and projective modules over polynomial rings. It states that every finitely generated projective module over a polynomial ring is free. Geometrically, finitely generated projective modules correspond to vector bundles over affine space, and free modules to trivial vector bundles. Affine space is topologically contractible, so it admits no non-trivial topological vector bundles. A simple argument using the exponential exact sequence and the d-bar Poincar lemma shows that it also admits no non-trivial holomorphic vector bundles. Jean-Pierre Serre, in his 1955 paper "Faisceaux algbriques cohrents", remarked that the equivalent question was not known for algebraic vector bundles: "It is not known if there exist projective A-modules of finite type which are not free."[1] Here A is a polynomial ring over a field, that is, A = k[x1, ..., xn]. To Serre's dismay, this problem quickly became known as Serre's conjecture. (Serre wrote, "I objected as often as I could [to the name]."[2] ) The statement is not immediately obvious from the topological and holomorphic cases, because these cases only guarantee that there is a continuous or holomorphic trivialization, not an algebraic trivialization. Instead, the problem turns out to be extremely difficult. Serre made some progress towards a solution in 1957 when he proved that every finitely generated projective module over a polynomial ring over a field was stably free, meaning that after forming its direct sum with a finitely generated free module, it became free. The problem remained open until 1976, when Daniel Quillen and Andrei Suslin independently proved that the answer was affirmative. Quillen was awarded the Fields Medal in 1978 in part for his proof of the Serre conjecture. Leonid Vaserten later gave a simpler and much shorter proof of the theorem which can be found in Serge Lang's Algebra.

Notes
[1] "On ignore s'il existe des A-modules projectifs de type fini qui ne soient pas libres." Serre, FAC, p. 243. [2] Lam, p. 1

References
Serre, Jean-Pierre (March 1955), "Faisceaux algbriques cohrents", Annals of Mathematics. Second Series. 61 (2): 197278, doi:10.2307/1969915, JSTOR1969915 Serre, Jean-Pierre (1958), "Modules projectifs et espaces fibrs fibre vectorielle" (in French), Sminaire P. Dubreil, M.-L. Dubreil-Jacotin et C. Pisot, 1957/58, Fasc. 2, Expos 23 Quillen, Daniel (1976), "Projective modules over polynomial rings", Inventiones Mathematicae 36: 167171, doi:10.1007/BF01390008

QuillenSuslin theorem Suslin, Andrei A. (1976), " " (in Russian), Doklady Akademii Nauk SSSR 229 (5): 10631066. Translated in Soviet Mathematics 17 (4): 11601164, 1976. Lang, Serge (2002), Algebra, Graduate Texts in Mathematics, 211 (Revised third ed.), New York: Springer-Verlag, ISBN978-0-387-95385-4, MR1878556 An account of this topic is provided by: Lam, T. Y. (2006), Serre's problem on projective modules, Berlin; New York: Springer Science+Business Media, pp.300pp., ISBN978-3540233176

191

Rational root theorem


In algebra, the rational root theorem (or rational root test) states a constraint on rational solutions (or roots) of the polynomial equation

with integer coefficients. If a0 and an are nonzero, then each rational solution x, when written as a fraction x=p/q in lowest terms (i.e., the greatest common divisor of p and q is 1), satisfies p is an integer factor of the constant term a0, and q is an integer factor of the leading coefficient an. Thus, a list of possible rational roots of the equation can be derived using the formula .

The rational root theorem is a special case (for a single linear factor) of Gauss's lemma on the factorization of polynomials. The integral root theorem is a special case of the rational root theorem if the leading coefficientan=1.

Proofs
An elementary proof
Let P(x) = anxn + an-1xn-1 + ... + a1x + a0 for some a0, ..., an Z, and suppose P(p/q) = 0 for some coprime p, q Z:

If we shift the constant term to the right hand side and multiply by qn, we get We see that p times the integer quantity in parentheses equals -a0qn, so p divides a0qn. But p is coprime to q and therefore to qn, so by (the generalized form of) Euclid's lemma it must divide the remaining factor a0 of the product. If we instead shift the leading term to the right hand side and multiply by qn, we get And for similar reasons, we can conclude that q divides an.[1]

Rational root theorem

192

Proof using Gauss's lemma


Should there be a nontrivial factor dividing all the coefficients of the polynomial, then one can divide by the greatest common divisor of the coefficients so as to obtain a primitive polynomial in the sense of Gauss's lemma; this does not alter the set of rational roots and only strengthens the divisibility conditions. That lemma says that if the polynomial factors in [X], then it also factors in [X] as a product of primitive polynomials. Now any rational root p/q corresponds to a factor of degree 1 in [X] of the polynomial, and its primitive representative is then qx p, assuming that p and q are coprime. But any multiple in [X] of qx p has leading term divisible by q and constant term divisible by p, which proves the statement. This argument shows that more generally, any irreducible factor of P can be supposed to have integer coefficients, and leading and constant coefficients dividing the corresponding coefficients ofP.

Example
For example, every rational solution of the equation

must be among the numbers symbolically indicated by which gives the list of possible answers:

These root candidates can be tested using the Horner scheme (for instance). In this particular case there is exactly one rational root. If a root candidate does not satisfy the equation, it can be used to shorten the list of remaining candidates. For example, x=1 does not satisfy the equation as the left hand side equals1. This means that substituting x=1+t yields a polynomial int with constant term1, while the coefficient of t3 remains the same as the coefficient of x3. Applying the rational root theorem thus yields the following possible roots fort:

Therefore,

Root candidates that do not occur on both lists are ruled out. The list of rational root candidates has thus shrunk to just x = 2 and x =2/3. If a root r1 is found, the Horner scheme will also yield a polynomial of degree n1 whose roots, together with r1, are exactly the roots of the original polynomial. It may also be the case that none of the candidates is a solution; in this case the equation has no rational solution. If the equation lacks a constant term a0, then 0 is one of the rational roots of the equation.

Rational root theorem

193

Notes
[1] D. Arnold, G. Arnold (1993). Four unit mathematics. Edward Arnold. pp.120121. ISBN0340543353.

References
Charles D. Miller, Margaret L. Lial, David I. Schneider: Fundamentals of College Algebra. Scott & Foresman/Little & Brown Higher Education, 3rd edition 1990, ISBN 0-673-38638-4, pp. 216-221 Phillip S. Jones, Jack D. Bedient: The historical roots of elementary mathematics. Dover Courier Publications 1998, ISBN 0486255638, pp. 116-117 ( online copy (http://books.google.com/books?id=7xArILpcndYC& pg=PA116) at Google Books) Ron Larson: Calculus: An Applied Approach. Cengage Learning 2007, ISBN 9780618958252, pp. 23-24 ( online copy (http://books.google.com/books?id=bDG7V0OV34C&pg=PA23) at Google Books)

External links
Weisstein, Eric W., " Rational Zero Theorem (http://mathworld.wolfram.com/RationalZeroTheorem.html)" from MathWorld. RationalRootTheorem (http://planetmath.org/encyclopedia/RationalRootTheorem.html) at PlanetMath Another proof that nth roots of integers are irrational, except for perfect nth powers (http://www.cut-the-knot. org/Generalization/RationalRootTheorem.shtml) by Scott E. Brodie The Rational Roots Test (http://www.purplemath.com/modules/rtnlroot.htm) at purplemath.com

Regev's theorem
In algebra, Regev's theorem, proved by Amitai Regev(1971, 1972), states that the tensor product of two PI algebras is a PI algebra.

References
Regev, Amitai (1971), "Existence of polynomial identities in AFB" [1], Bulletin of the American Mathematical Society 77 (6): 10671069, doi:10.1090/S0002-9904-1971-12869-0, ISSN0002-9904, MR0284468 Regev, Amitai (1972), "Existence of identities in AB", Israel Journal of Mathematics 11: 131152, doi:10.1007/BF02762615, ISSN0021-2172, MR0314893

References
[1] http:/ / projecteuclid. org/ euclid. bams/ 1183533194

Schreier refinement theorem

194

Schreier refinement theorem


In mathematics, the Schreier refinement theorem of group theory states that any two normal series of subgroups of a given group have equivalent refinements. The theorem is named after the Austrian mathematician Otto Schreier who proved it in 1928. It provides an elegant proof of the JordanHlder theorem.

References
Rotman, Joseph (1994). An introduction to the theory of groups. New York: Springer-Verlag. ISBN0-387-94285-8.

SchurZassenhaus theorem
The SchurZassenhaus theorem is a theorem in group theory which states that if normal subgroup whose order is coprime to the order of the quotient group of . It is clear that if we do not impose the coprime condition, the theorem is not true: consider for example the cyclic group and its normal subgroup . Then if were a semidirect product of and then would have to contain two elements of order 2, but it only contains one. The SchurZassenhaus theorem at least partially answers the question: "In a composition series, how can we classify groups with a certain set of composition factors?" The other part, which is where the composition factors do not have coprime orders, is tackled in extension theory. and . has a complement in is a finite group, and is a , then is a semidirect product

An alternative statement of the theorem is that any normal Hall subgroup of a finite group

References
Rotman, Joseph J. (1995). An Introduction to the Theory of Groups. New York: SpringerVerlag. ISBN978-0-387-94285-8. David S. Dummit & Richard M. Foote (2003). Abstract Algebra. Wiley. ISBN978-0-471-43334-7. J. S. Milne (2009). Group Theory [1]. Lecture notes.

References
[1] http:/ / www. jmilne. org/ math/ CourseNotes/ gt. html

SerreSwan theorem

195

SerreSwan theorem
In the mathematical fields of topology and K-theory, the SerreSwan theorem, also called Swan's theorem, relates the geometric notion of vector bundles to the algebraic concept of projective modules and gives rise to a common intuition throughout mathematics: "projective modules over commutative rings are like vector bundles on compact spaces". The two precise formulations of the theorems differ somewhat. The original theorem, as stated by Jean-Pierre Serre in 1955, is more algebraic in nature, and concerns vector bundles on an algebraic variety over an algebraically closed field (of any characteristic). The complementary variant stated by Richard Swan in 1962 is more analytic, and concerns (real, complex, or quaternionic) vector bundles on a smooth manifold or Hausdorff space.

Differential geometry
Suppose M is a compact smooth manifold, and a V is a smooth vector bundle over M. The space of smooth sections of V is then a module over C(M) (the commutative algebra of smooth real-valued functions on M). Swan's theorem states that this module is finitely generated and projective over C(M). In other words, every vector bundle is a direct summand of some trivial bundle: M Cn for some n. The theorem can be proved by constructing a bundle epimorphism from a trivial bundle M Cn onto V. This can be done by, for instance, exhibiting sections s1...sn with the property that for each point p, {si(p)} span the fiber over p. The converse is also true: every finitely generated projective module over C(M) arises in this way from some smooth vector bundle on M. Such a module can be viewed as a smooth function f on M with values in the n n idempotent matrices for some n. The fiber of the corresponding vector bundle over x is then the range of f(x). Therefore, the category of smooth vector bundles on M is equivalent to the category of finitely generated projective modules over C(M). Details may be found in (Nestruev 2003). This equivalence is extended to the case of a noncompact manifold M (Giachetta et al. 2005).

Topology
Suppose X is a compact Hausdorff space, and C(X) is the ring of continuous real-valued functions on X. Analogous to the result above, the category of real vector bundles on X is equivalent to the category of finitely generated projective modules over C(X). The same result holds if one replaces "real-valued" by "complex-valued" and "real vector bundle" by "complex vector bundle", but it does not hold if one replace the field by a totally disconnected field like the rational numbers. In detail, let Vec(X) be the category of complex vector bundles over X, and let ProjMod(C(X)) be the category of finitely generated projective modules over the C*-algebra C(X). There is a functor : Vec(X)ProjMod(C(X)) which sends each complex vector bundle E over X to the C(X)-module (X,E) of sections. Swan's theorem asserts that the functor is an equivalence of categories.

SerreSwan theorem

196

Algebraic geometry
The analogous result in algebraic geometry, due to Serre (1955, 50) applies to vector bundles in the category of affine varieties. Let X be an affine variety with structure sheaf OX, and F a coherent sheaf of OX-modules on X. Then F is the sheaf of germs of a finite-dimensional vector bundle if and only if the space of sections of F, (F,X), is a projective module over the commutative ring A=(OX,X).

References
Karoubi, Max (1978), K-theory: An introduction, Grundlehren der mathematischen Wissenschaften, Springer-Verlag, ISBN978-0387080901 Manoharan, P. (1995), "Generalized Swan's Theorem and its Application", Proceedings of the American Mathematical Society 123 (10): 32193223, doi:10.2307/2160685, JSTOR2160685. Serre, Jean-Pierre (1955), "Faisceaux Algebriques Coherents", Annals of Mathematics 61 (2): 197278, doi:10.2307/1969915, JSTOR1969915. Swan, Richard G. (1962), "Vector Bundles and Projective Modules", Transactions of the American Mathematical Society 105 (2): 264277, doi:10.2307/1993627, JSTOR1993627. Nestruev, Jet (2003), Smooth manifolds and observables, Graduate texts in mathematics, 220, Springer-Verlag, ISBN0-387-95543-7 Giachetta, G.; Mangiarotti, L.; Sardanashvily, G. (2005), Geometric and Algebraic Topological Methods in Quantum Mechanics, World Scientific, ISBN9812561293. This article incorporates material from Serre-Swan theorem on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.

SkolemNoether theorem
In mathematics, the SkolemNoether theorem, named after Thoralf Skolem and Emmy Noether, is an important result in ring theory which characterizes the automorphisms of simple rings. The theorem was first published by Skolem in 1927 in his paper Zur Theorie der assoziativen Zahlensysteme (German: On the theory of associative number systems) and later rediscovered by Noether.

SkolemNoether theorem
In a general formulation, let A and B be simple rings, and let K = Z(B) be the centre of B. Notice that K is a field since given x nonzero in K, the simplicity of B implies that the nonzero two-sided ideal Bx is the whole of B, and hence that x is a unit. Suppose further that the dimension of B over K is finite, i.e. that B is a central simple algebra. Then given K-algebra homomorphisms f, g : A B there exists a unit b in B such that for all a in A g(a) = b f(a) b1. In particular, every automorphism of a Brauer algebra is inner.

SkolemNoether theorem

197

References
Thoralf Skolem, Zur Theorie der assoziativen Zahlensysteme, 1927 A proof [1]

References
[1] http:/ / www. math. virginia. edu/ ~ww9c/ divalgebras. pdf

Specht's theorem
In mathematics, Specht's theorem gives a necessary and sufficient condition for two matrices to be unitarily equivalent. It is named after Wilhelm Specht, who proved the theorem in 1940.[1] Two matrices A and B are said to be unitarily equivalent if there exists a unitary matrix U such that B = U *AU.[2] Two matrices which are unitarily equivalent are also similar. Two similar matrices represent the same linear map, but with respect to a different basis; unitary equivalence corresponds to a change from an orthonormal basis to another orthonormal basis. If A and B are unitarily equivalent, then tr AA* = tr BB*, where tr denotes the trace (in other words, the Frobenius norm is a unitary invariant). This follows from the cyclic invariance of the trace: if B = U *AU, then tr BB* = tr U *AUU *A*U = tr AUU *A*UU * = tr AA*, where the second equality is cyclic invariance.[3] Thus, tr AA* = tr BB* is a necessary condition for unitary equivalence, but it is not sufficient. Specht's theorem gives infinitely many necessary conditions which together are also sufficient. The formulation of the theorem uses the following definition. A word in two variables, say x and y, is an expression of the form

where m1, n1, m2, n2, , mp are non-negative integers. The degree of this word is Specht's theorem: Two matrices A and B are unitarily equivalent if and only if tr W(A, A*) = tr W(B, B*) for all words W.[4] The theorem gives an infinite number of trace identities, but it can be reduced to a finite subset. Let n denote the size of the matrices A and B. For the case n = 2, the following three conditions are sufficient:[5]

For n = 3, the following seven conditions are sufficient:


[6]

For general n, it suffices to show that tr W(A, A*) = tr W(B, B*) for all words of degree at most [7] It has been conjectured that this can be reduced to an expression linear in n.[8]

Specht's theorem

198

Notes
[1] [2] [3] [4] [5] [6] [7] [8] Specht (1940) Horn & Johnson (1985), Definition 2.2.1 Horn & Johnson (1985), Theorem 2.2.2 Horn & Johnson (1985), Theorem 2.2.6 Horn & Johnson (1985), Theorem 2.2.8 Sibirski (1976), p. 260, quoted by okovi & Johnson (2007) Pappacena (1997), Theorem 4.3 Freedman, Gupta & Guralnick (1997), p. 160

References
okovi, Dragomir .; Johnson, Charles R. (2007), "Unitarily achievable zero patterns and traces of words in A and A*", Linear Algebra and its Applications 421 (1): 6368, doi:10.1016/j.laa.2006.03.002, ISSN0024-3795. Freedman, Allen R.; Gupta, Ram Niwas; Guralnick, Robert M. (1997), "Shirshov's theorem and representations of semigroups" (http://pjm.math.berkeley.edu/pjm/1997/181-3/p07.xhtml), Pacific Journal of Mathematics 181 (3): 159176, doi:10.2140/pjm.1997.181.159, ISSN0030-8730. Horn, Roger A.; Johnson, Charles R. (1985), Matrix Analysis, Cambridge University Press, ISBN978-0-521-38632-6. Pappacena, Christopher J. (1997), "An upper bound for the length of a finite-dimensional algebra", Journal of Algebra 197 (2): 535545, doi:10.1006/jabr.1997.7140, ISSN0021-8693. Sibirski, K. S. (1976) (in Russian), Algebraic Invariants of Differential Equations and Matrices, Izdat. "tiinca", Kishinev. Specht, Wilhelm (1940), "Zur Theorie der Matrizen. II" (http://gdz.sub.uni-goettingen.de/dms/load/toc/ ?PPN=PPN37721857X_0050&DMDID=dmdlog6), Jahresbericht der Deutschen Mathematiker-Vereinigung 50: 1923, ISSN0012-0456.

Stone's representation theorem for Boolean algebras

199

Stone's representation theorem for Boolean algebras


In mathematics, Stone's representation theorem for Boolean algebras states that every Boolean algebra is isomorphic to a field of sets. The theorem is fundamental to the deeper understanding of Boolean algebra that emerged in the first half of the 20th century. The theorem was first proved by Stone (1936), and thus named in his honor. Stone was led to it by his study of the spectral theory of operators on a Hilbert space.

Stone spaces
Each Boolean algebra B has an associated topological space, denoted here S(B), called its Stone space. The points in S(B) are the ultrafilters on B, or equivalently the homomorphisms from B to the two-element Boolean algebra. The topology on S(B) is generated by a basis consisting of all sets of the form

where b is an element of B. For any Boolean algebra B, S(B) is a compact totally disconnected Hausdorff space; such spaces are called Stone spaces (also profinite spaces). Conversely, given any topological space X, the collection of subsets of X that are clopen (both closed and open) is a Boolean algebra.

Representation theorem
A simple version of Stone's representation theorem states that any Boolean algebra B is isomorphic to the algebra of clopen subsets of its Stone space S(B). The isomorphism sends an element bB to the set of all ultrafilters that contain b. This is a clopen set because of the choice of topology on S(B) and because B is a Boolean algebra. The full statement of the theorem uses the language of category theory; it states that there is a duality between the category of Boolean algebras and the category of Stone spaces. This duality means that in addition to the isomorphisms between Boolean algebras and their Stone spaces, each homomorphism from a Boolean algebra A to a Boolean algebra B corresponds in a natural way to a continuous function from S(B) to S(A). In other words, there is a contravariant functor that gives an equivalence between the categories. This was an early example of a nontrivial duality of categories. The theorem is a special case of Stone duality, a more general framework for dualities between topological spaces and partially ordered sets. The proof requires either the axiom of choice or a weakened form of it. Specifically, the theorem is equivalent to the Boolean prime ideal theorem, a weakened choice principle which states that every Boolean algebra has a prime ideal.

References
Paul Halmos, and Givant, Steven (1998) Logic as Algebra. Dolciani Mathematical Expositions No. 21. The Mathematical Association of America. Johnstone, Peter T. (1982) Stone Spaces. Cambridge University Press. ISBN 0-521-23893-5. Marshall H. Stone (1936) "The Theory of Representations of Boolean Algebras, [1]" Transactions of the American Mathematical Society 40: 37-111. A monograph available free online: Burris, Stanley N., and H.P. Sankappanavar, H. P.(1981) A Course in Universal Algebra. [2] Springer-Verlag. ISBN 3-540-90578-2.

Stone's representation theorem for Boolean algebras

200

References
[1] http:/ / links. jstor. org/ sici?sici=0002-9947%28193607%2940%3A1%3C37%3ATTORFB%3E2. 0. CO%3B2-8 [2] http:/ / www. thoralf. uwaterloo. ca/ htdocs/ ualg. html

Structure theorem for finitely generated modules over a principal ideal domain
In mathematics, in the field of abstract algebra, the structure theorem for finitely generated modules over a principal ideal domain is a generalization of the fundamental theorem of finitely generated abelian groups and roughly states that finitely generated modules can be uniquely decomposed in much the same way that integers have a prime factorization. The result provides a simple framework to understand various canonical form results for square matrices over fields.

Statement
When a vector space over a field F has a finite generating set, then one may extract from it a basis consisting of a finite number n of vectors, and the space is therefore isomorphic to Fn. The corresponding statement with the F generalized to a principal ideal domain R is no longer true, as a finitely generated module over R need not have any basis. However such a module is still isomorphic to a quotient of some module Rn with n finite (to see this it suffices to construct the morphism that sends the elements of the canonical basis Rn to the generators of the module, and take the quotient by its kernel. By changing the choice of generating set, one can in fact describe the module as the quotient of some Rn by a particularly simple submodule, and this is the structure theorem. The structure theorem for finitely generated modules over a principal ideal domain has two statements, which are equivalent by the Chinese remainder theorem:

Invariant factor decomposition


Every finitely generated module M over a principal ideal domain R is isomorphic to a unique one of the form

where The elements

and

(up to unit) are a complete set of invariants for finitely generated R-modules, and are called are unique; the elements are unique up to multiplication by a unit, but the order is unique.

invariant factors. The ideals Some prefer to separate out the free part and write M as:

where

and

. The free part is where (in the first formulation)

; these occur at the end, as

everything divides zero.

Structure theorem for finitely generated modules over a principal ideal domain

201

Primary decomposition
Every finitely generated module M over a principal ideal domain R is isomorphic to a unique one of the form

where

and the

are primary ideals. The ideals

are unique (up to order); the elements .

are

unique up to multiplication by a unit, and are called the elementary divisors. Note that in a PID, primary ideals are powers of primes, so The summands are indecomposable, so the primary decomposition is a decomposition into indecomposable ) and write M as:

modules, and thus every finitely generated module over a PID is completely decomposable. Some prefer to separate out the free part (where

where

and the

are primary ideals.

Proofs
One proof proceeds as follows: Every finitely generated module over a PID is also finitely presented because a PID is Noetherian, an even stronger condition than coherence. Take a presentation, which is a map (relations to generators), and put it in Smith normal form. This yields the invariant factor decomposition, and the diagonal entries of Smith normal form are the invariant factors. Another outline of a proof: Denote by tM the torsion submodule of M. Then M/tM is a finitely generated torsion free module, and such a module over a commutative PID is a free module of finite rank. As a result, where for a positive integer n. For a prime p in R we can then speak of for each prime p. This is a submodule of tM, and it turns out that each Np is a direct sum of cyclic modules, and that tM is a direct sum of Np for a finite number of distinct primes p. Putting the previous two steps together, M is decomposed into cyclic modules of the indicated types.

Corollaries
This includes the classification of finite-dimensional vector spaces as a special case, where have no non-trivial ideals, every finitely generated vector space is free. Taking Taking yields the fundamental theorem of finitely generated abelian groups. classifies linear operators on a finite-dimensional vector space an operator on a vector space . Since fields

is the same as an algebra representation of the polynomial algebra in one variable where the last invariant factor is the minimal polynomial, and the product of invariant factors is the characteristic polynomial. Combined with a standard matrix form for , this yields various canonical forms: invariant factors + companion matrix yields Frobenius normal form (aka, rational canonical form) primary decomposition + companion matrix yields primary rational canonical form primary decomposition + Jordan blocks yields Jordan canonical form (this latter only holds over an algebraically closed field)

Structure theorem for finitely generated modules over a principal ideal domain

202

Uniqueness
While the invariants (rank, invariant factors, and elementary divisors) are unique, the isomorphism between M and its canonical form is not unique, and does not even preserve the direct sum decomposition. This follows because there are non-trivial automorphisms of these modules which do not preserve the summands. However, one has a canonical torsion submodule T, and similar canonical submodules corresponding to each (distinct) invariant factor, which yield a canonical sequence:

Compare composition series in JordanHlder theorem. For instance, if change of basis matrix , and is one basis, then is another basis, and the summand,

does not preserve the summand

. However, it does preserve the

as this is the torsion submodule (equivalently here, the 2-torsion elements).

Generalizations
Groups
The JordanHlder theorem is a more general result for finite groups (or modules over an arbitrary ring). In this generality, one obtains a composition series, rather than a direct sum. The KrullSchmidt theorem and related results give conditions under which a module has something like a primary decomposition, a decomposition as a direct sum of indecomposable modules in which the summands are unique up to order.

Primary decomposition
The primary decomposition generalizes to finitely generated modules over commutative Noetherian rings, and this result is called the LaskerNoether theorem.

Indecomposable modules
By contrast, unique decomposition into indecomposable submodules does not generalize as far, and the failure is measured by the ideal class group, which vanishes for PIDs. For rings that are not principal ideal domains, unique decomposition need not even hold for modules over a ring generated by two elements. For the ring R=Z[5], both the module R and its submodule M generated by 2 and 1+5 are indecomposable. While R is not isomorphic to M, RR is isomorphic to MM; thus the images of the M summands give indecomposable submodules L1,L2<RR which give a different decomposition of RR. The failure of uniquely factorizing RR into a direct sum of indecomposable modules is directly related (via the ideal class group) to the failure of the unique factorization of elements of R into irreducible elements ofR.

Structure theorem for finitely generated modules over a principal ideal domain

203

Non-finitely generated modules


Similarly for modules that are not finitely generated, one cannot expect such a nice decomposition: even the number of factors may vary. There are Z-submodules A of Q4 which are simultaneously direct sums of two indecomposable modules and direct sums of three indecomposable modules, showing the analogue of the primary decomposition cannot hold for infinitely generated modules, even over the integers, Z. Another issue that arises with non-finitely generated modules is that there are torsion-free modules which are not free. For instance, consider the ring Z of integers. A classical example of a torsion-free module which is not free is the BaerSpecker group, the group of all sequences of integers under termwise addition. In general, the question of which infinitely generated torsion-free abelian groups are free depends on which large cardinals exist. A consequence is that any structure theorem for infinitely generated modules depends on a choice of set theory axioms and may be invalid under a different choice.

References
Atiyah, Michael Francis; Macdonald, I.G. (1969), Introduction to Commutative Algebra, Westview Press, ISBN978-0-201-40751-8 Dummit, David S.; Foote, Richard M. (2004), Abstract algebra (3rd ed.), New York: Wiley, ISBN978-0-471-43334-7, MR2286236 Hungerford, Thomas W. (1980), Algebra, New York: Springer, pp.218226, Section IV.6: Modules over a Principal Ideal Domain, ISBN978-0-387-90518-1 Jacobson, Nathan (1985), Basic algebra. I (2 ed.), New York: W. H. Freeman and Company, pp.xviii+499, ISBN0-7167-1480-9, MR780184 Lam, T. Y. (1999), Lectures on modules and rings, Graduate Texts in Mathematics No. 189, Springer-Verlag, ISBN978-0-387-98428-5

Subgroup test

204

Subgroup test
In Abstract Algebra, the one-step subgroup test is a theorem that states that for any group, a nonempty subset of that group is itself a group if the inverse of any element in the subset multiplied with any other element in the subset is also in the subset. The two-step subgroup test is a similar theorem which requires the subset to be closed under the operation and taking of inverses.

One-step subgroup test


Let G be a group and let H be a nonempty subset of G. If for all a and b in H, ab-1 is in H, then H is a subgroup of G.

Proof
Let G be a group, let H be a nonempty subset of G and assume that for all a and b in H, ab-1 is in H. To prove that H is a subgroup of G we must show that H is associative, has an identity, has an inverse for every element and is closed under the operation. So, Since the operation of H is the same as the operation of G, the operation is associative since G is a group. Since H is not empty there exists an element x in H. Then the identity is in H since we can write it as e = x x-1 which is in H by the initial assumption. Let x be an element of H. Since the identity e is in H it follows that ex-1 = x-1 in H, so the inverse of an element in H is in H. Finally, let x and y be elements in H, then since y is in H it follows that y-1 is in H. Hence x(y-1)-1 = xy is in H and so H is closed under the operation. Thus H is a subgroup of G.

Two-step subgroup test


A corollary of this theorem is the two-step subgroup test which states that a nonempty subset of a group is itself a group if the subset is closed under the operation as well as under the taking of inverses.

Subring test

205

Subring test
In abstract algebra, the subring test is a theorem that states that for any ring, a nonempty subset of that ring is a subring if it is closed under multiplication and subtraction. Note that here that the terms ring and subring are used without requiring a multiplicative identity element. More formally, let for all be a ring, and let then be a nonempty subset of is a subring of . . If for all one has and

one has

If rings are required to have unity, then it must also be assumed that the multiplicative identity is in the subset.

Proof
Since Hence, is nonempty and closed under subtraction, by the subgroup test it follows that is closed under addition, addition is associative, are the same as those of is a group under addition. has has an additive identity, and every element in

an additive inverse. Since the operations of it immediately follows that addition is commutative,

multiplication is associative, multiplication is left distributive over addition, and multiplication is right distributive over addition. Thus, is a subring of .

Sylow theorems
In mathematics, specifically in the field of finite group theory, the Sylow theorems are a collection of theorems named after the Norwegian mathematician Ludwig Sylow (1872) that give detailed information about the number of subgroups of fixed order that a given finite group contains. The Sylow theorems form a fundamental part of finite group theory and have very important applications in the classification of finite simple groups. For a prime number p, a Sylow p-subgroup (sometimes p-Sylow subgroup) of a group G is a maximal p-subgroup of G, i.e., a subgroup of G which is a p-group (so that the order of any group element is a power of p), and which is not a proper subgroup of any other p-subgroup of G. The set of all Sylow p-subgroups for a given prime p is sometimes written Sylp(G). The Sylow theorems assert a partial converse to Lagrange's theorem that for any finite group G the order (number of elements) of every subgroup of G divides the order of G. For any prime factor p of the order of a finite group G, there exists a Sylow p-subgroup of G. The order of a Sylow p-subgroup of a finite group G is pn, where n is the multiplicity of p in the order of G, and any subgroup of order pn is a Sylow p-subgroup of G. The Sylow p-subgroups of a group (for fixed prime p) are conjugate to each other. The number of Sylow p-subgroups of a group for fixed prime p is congruent to 1 mod p.

Sylow theorems

206

Sylow theorems
Collections of subgroups which are each maximal in one sense or another are common in group theory. The surprising result here is that in the case of Sylp(G), all members are actually isomorphic to each other and have the largest possible order: if |G| = pnm with where p does not divide m, then any Sylow p-subgroup P has order |P| = pn. That is, P is a p-group and gcd(|G:P|, p) = 1. These properties can be exploited to further analyze the structure of G. The following theorems were first proposed and proven by Ludwig Sylow in 1872, and published in Mathematische Annalen. Theorem 1: For any prime factor p with multiplicity n of the order of a finite group G, there exists a Sylow p-subgroup of G, of order pn. The following weaker version of theorem 1 was first proved by Cauchy. Corollary: Given a finite group G and a prime number p dividing the order of G, then there exists an element of order p in G . Theorem 2: Given a finite group G and a prime number p, all Sylow p-subgroups of G are conjugate to each other, i.e. if H and K are Sylow p-subgroups of G, then there exists an element g in G with g1Hg = K. Theorem 3: Let p be a prime factor with multiplicity n of the order of a finite group G, so that the order of G can be written as pn m, where n > 0 and p does not divide m. Let np be the number of Sylow p-subgroups of G. Then the following hold: np divides m, which is the index of the Sylow p-subgroup in G. np 1 mod p. np = |G : NG(P)|, where P is any Sylow p-subgroup of G and NG denotes the normalizer.

Consequences
The Sylow theorems imply that for a prime number p every Sylow p-subgroup is of the same order, pn. Conversely, if a subgroup has order pn, then it is a Sylow p-subgroup, and so is isomorphic to every other Sylow p-subgroup. Due to the maximality condition, if H is any p-subgroup of G, then H is a subgroup of a p-subgroup of order pn. A very important consequence of Theorem 2 is that the condition np = 1 is equivalent to saying that the Sylow p-subgroup of G is a normal subgroup (there are groups which have normal subgroups but no normal Sylow subgroups, such as S4).

Sylow theorems for infinite groups


There is an analogue of the Sylow theorems for infinite groups. We define a Sylow p-subgroup in an infinite group to be a p-subgroup (that is, every element in it has p-power order) which is maximal for inclusion among all p-subgroups in the group. Such subgroups exist by Zorn's lemma. Theorem: If K is a Sylow p-subgroup of G, and np = |Cl(K)| is finite, then every Sylow p-subgroup is conjugate to K, and np 1 mod p, where Cl(K) denotes the conjugacy class of K.

Sylow theorems

207

Examples
A simple illustration of Sylow subgroups and the Sylow theorems are the dihedral group of the n-gon, For n odd, is the higher power of 2 dividing the order, and thus subgroups of order 2 are Sylow subgroups. These are the groups generated by a reflection, of which there are n, and they are all conjugate under rotations; geometrically the axes of symmetry pass through a vertex and a side. By contrast, if n is even, then 4 divides the order of the group, and these are no longer Sylow subgroups, and in fact they fall into two conjugacy classes, geometrically according to whether they pass through two vertices or two faces. These are related by an outer automorphism, which can be represented by rotation through half the minimal rotation in the dihedral group.
In all reflections are conjugate, as reflections correspond to Sylow 2-subgroups.

Example applications

In

reflections no longer correspond to

Sylow 2-subgroups, and fall into two conjugacy classes.

Cyclic group orders


Some numbers n are such that every group of order n is cyclic. One can show that n = 15 is such a number using the Sylow theorems: Let G be a group of order 15 = 3 5 and n3 be the number of Sylow 3-subgroups. Then and . The only value satisfying these constraints is 1; therefore, there is only one subgroup of order 3, and it must be normal (since it has no distinct conjugates). Similarly, n5 must divide 3, and n5 must equal 1 (mod 5); thus it must also have a single normal subgroup of order 5. Since 3 and 5 are coprime, the intersection of these two subgroups is trivial, and so G must be the internal direct product of groups of order 3 and 5, that is the cyclic group of order 15. Thus, there is only one group of order 15 (up to isomorphism).

Sylow theorems

208

Small groups are not simple


A more complex example involves the order of the smallest simple group which is not cyclic. Burnside's paqb theorem states that if the order of a group is the product of two prime powers, then it is solvable, and so the group is not simple, or is of prime order and is cyclic. This rules out every group up to order 30 (= 2 3 5). If G is simple, and |G| = 30, then n3 must divide 10 ( = 2 5), and n3 must equal 1 (mod 3). Therefore n3 = 10, since neither 4 nor 7 divides 10, and if n3 = 1 then, as above, G would have a normal subgroup of order 3, and could not be simple. G then has 10 distinct cyclic subgroups of order 3, each of which has 2 elements of order 3 (plus the identity). This means G has at least 20 distinct elements of order 3. As well, n5 = 6, since n5 must divide 6 ( = 2 3), and n5 must equal 1 (mod 5). So G also has 24 distinct elements of order 5. But the order of G is only 30, so a simple group of order 30 cannot exist. Next, suppose |G| = 42 = 2 3 7. Here n7 must divide 6 ( = 2 3) and n7 must equal 1 (mod 7), so n7 = 1. So, as before, G can not be simple. On the other hand for |G| = 60 = 22 3 5, then n3 = 10 and n5 = 6 is perfectly possible. And in fact, the smallest simple non-cyclic group is A5, the alternating group over 5 elements. It has order 60, and has 24 cyclic permutations of order 5, and 20 of order 3.

Fusion results
Frattini's argument shows that a Sylow subgroup of a normal subgroup provides a factorization of a finite group. A slight generalization known as Burnside's fusion theorem states that if G is a finite group with Sylow p-subgroup P and two subsets A and B normalized by P, then A and B are G-conjugate if and only if they are NG(P)-conjugate. The proof is a simple application of Sylow's theorem: If B=Ag, then the normalizer of B contains not only P but also Pg (since Pg is contained in the normalizer of Ag). By Sylow's theorem P and Pg are conjugate not only in G, but in the normalizer of B. Hence gh1 normalizes P for some h that normalizes B, and then Agh1 = Bh1 = B, so that A and B are NG(P)-conjugate. Burnside's fusion theorem can be used to give a more power factorization called a semidirect product: if G is a finite group whose Sylow p-subgroup P is contained in the center of its normalizer, then G has a normal subgroup K of order coprime to P, G = PK and PK = 1, that is, G is p-nilpotent. Less trivial applications of the Sylow theorems include the focal subgroup theorem, which studies the control a Sylow p-subgroup of the derived subgroup has on the structure of the entire group. This control is exploited at several stages of the classification of finite simple groups, and for instance defines the case divisions used in the AlperinBrauerGorenstein theorem classifying finite simple groups whose Sylow 2-subgroup is a quasi-dihedral group. These rely on J. L. Alperin's strengthening of the conjugacy portion of Sylow's theorem to control what sorts of elements are used in the conjugation.

Proof of the Sylow theorems


The Sylow theorems have been proved in a number of ways, and the history of the proofs themselves are the subject of many papers including (Waterhouse 1980), (Scharlau 1988), (Casadio & Zappa 1990), (Gow 1994), and to some extent (Meo 2004). One proof of the Sylow theorems exploit the notion of group action in various creative ways. The group G acts on itself or on the set of its p-subgroups in various ways, and each such action can be exploited to prove one of the Sylow theorems. The following proofs are based on combinatorial arguments of (Wielandt 1959). In the following, we use a | b as notation for "a divides b" and a b for the negation of this statement. Theorem 1: A finite group G whose order |G| is divisible by a prime power pk has a subgroup of order pk. Proof: Let |G| = pkm = pk+ru such that p does not divide u, and let denote the set of subsets of G of size pk. G acts on by left multiplication. The orbits G = {g | g G} of the are the equivalence classes under the action

Sylow theorems of G. For any consider its stabilizer subgroup G. For any fixed element the function [g g] maps G to injectively: for any two g,h G we have that g = h implies g = h, because G means that one may cancel on the right. Therefore pk = || |G|. On the other hand

209

and no power of p remains in any of the factors inside the product on the right. Hence p(||) = p(m) = r. Let R be a complete representation of all the equivalence classes under the action of G. Then,

Thus, there exists an element R such that s := p(|G|) p(||) = r. Hence |G| = psv where p does not divide v. By the stabilizer-orbit-theorem we have |G| = |G| / |G| = pk+r-su / v. Therefore pk | |G|, so pk |G| and G is the desired subgroup. Lemma: Let G be a finite p-group, let G act on a finite set , and let 0 denote the set of points of that are fixed under the action of G. Then || |0| mod p. Proof: Write as a disjoint sum of its orbits under G. Any element x not fixed by G will lie in an orbit of order |G|/|Gx| (where Gx denotes the stabilizer), which is a multiple of p by assumption. The result follows immediately. Theorem 2: If H is a p-subgroup of G and P is a Sylow p-subgroup of G, then there exists an element g in G such that g1Hg P. In particular, all Sylow p-subgroups of G are conjugate to each other (and therefore isomorphic), i.e. if H and K are Sylow p-subgroups of G, then there exists an element g in G with g1Hg = K. Proof: Let be the set of left cosets of P in G and let H act on by left multiplication. Applying the Lemma to H on , we see that |0| || = [G : P] mod p. Now p [G : P] by definition so p |0|, hence in particular |0| 0 so there exists some gP 0. It follows that for some g G and h H we have hgP = gP so g1hgP P and therefore g1Hg P. Now if H is a Sylow p-subgroup, |H| = |P| = |gPg1| so that H = gPg1 for some g G. Theorem 3: Let q denote the order of any Sylow p-subgroup of a finite group G. Then np | |G|/q and np 1 mod p. Proof: By Theorem 2, np = [G : NG(P)], where P is any such subgroup, and NG(P) denotes the normalizer of P in G, so this number is a divisor of |G|/q. Let be the set of all Sylow p-subgroups of G, and let P act on by conjugation. Let Q 0 and observe that then Q = xQx1 for all x P so that P NG(Q). By Theorem 2, P and Q are conjugate in NG(Q) in particular, and Q is normal in NG(Q), so then P = Q. It follows that 0 = {P} so that, by the Lemma, || |0| = 1 mod p.

Algorithms
The problem of finding a Sylow subgroup of a given group is an important problem in computational group theory. One proof of the existence of Sylow p-subgroups is constructive: if H is a p-subgroup of G and the index [G:H] is divisible by p, then the normalizer N = NG(H) of H in G is also such that [N:H] is divisible by p. In other words, a polycyclic generating system of a Sylow p-subgroup can be found by starting from any p-subgroup H (including the identity) and taking elements of p-power order contained in the normalizer of H but not in H itself. The algorithmic version of this (and many improvements) is described in textbook form in (Butler 1991, Chapter 16), including the algorithm described in (Cannon 1971). These versions are still used in the GAP computer algebra system. In permutation groups, it has been proven in (Kantor1985a, 1985b, 1990; Kantor & Taylor 1988) that a Sylow p-subgroup and its normalizer can be found in polynomial time of the input (the degree of the group times the

Sylow theorems number of generators). These algorithms are described in textbook form in (Seress 2003), and are now becoming practical as the constructive recognition of finite simple groups becomes a reality. In particular, versions of this algorithm are used in the Magma computer algebra system.

210

Notes References
Sylow, L. (1872), "Thormes sur les groupes de substitutions" (http://resolver.sub.uni-goettingen.de/ purl?GDZPPN002242052) (in French), Math. Ann. 5 (4): 584594, doi:10.1007/BF01442913, JFM04.0056.02

Proofs
Casadio, Giuseppina; Zappa, Guido (1990), "History of the Sylow theorem and its proofs" (in Italian), Boll. Storia Sci. Mat. 10 (1): 2975, ISSN0392-4432, MR1096350, Zbl0721.01008 Gow, Rod (1994), "Sylow's proof of Sylow's theorem", Irish Math. Soc. Bull. (33): 5563, ISSN0791-5578, MR1313412, Zbl0829.01011 Kammller, Florian; Paulson, Lawrence C. (1999), "A formal proof of Sylow's theorem. An experiment in abstract algebra with Isabelle HOL" (http://www.cl.cam.ac.uk/users/lcp/papers/Kammueller/sylow.pdf), J. Automat. Reason. 23 (3): 235264, doi:10.1023/A:1006269330992, ISSN0168-7433, MR1721912, Zbl0943.68149 Meo, M. (2004), "The mathematical life of Cauchy's group theorem", Historia Math. 31 (2): 196221, doi:10.1016/S0315-0860(03)00003-X, ISSN0315-0860, MR2055642, Zbl1065.01009 Scharlau, Winfried (1988), "Die Entdeckung der Sylow-Stze" (in German), Historia Math. 15 (1): 4052, doi:10.1016/0315-0860(88)90048-1, ISSN0315-0860, MR931678, Zbl0637.01006 Waterhouse, William C. (1980), "The early proofs of Sylow's theorem", Arch. Hist. Exact Sci. 21 (3): 279290, doi:10.1007/BF00327877, ISSN0003-9519, MR575718, Zbl0436.01006 Wielandt, Helmut (1959), "Ein Beweis fr die Existenz der Sylowgruppen" (in German), Arch. Math. 10 (1): 401402, doi:10.1007/BF01240818, ISSN0003-9268, MR0147529, Zbl0092.02403

Algorithms
Butler, G. (1991), Fundamental Algorithms for Permutation Groups, Lecture Notes in Computer Science, 559, Berlin, New York: Springer-Verlag, doi:10.1007/3-540-54955-2, ISBN978-3-540-54955-0, MR1225579, Zbl0785.20001 Cannon, John J. (1971), "Computing local structure of large finite groups", Computers in Algebra and Number Theory (Proc. SIAM-AMS Sympos. Appl. Math., New York, 1970), SIAM-AMS Proc., 4, Providence, RI: AMS, pp.161176, ISSN0160-7634, MR0367027, Zbl0253.20027 Kantor, William M. (1985a), "Polynomial-time algorithms for finding elements of prime order and Sylow subgroups", J. Algorithms 6 (4): 478514, doi:10.1016/0196-6774(85)90029-X, ISSN0196-6774, MR813589, Zbl0604.20001 Kantor, William M. (1985b), "Sylow's theorem in polynomial time", J. Comput. System Sci. 30 (3): 359394, doi:10.1016/0022-0000(85)90052-2, ISSN1090-2724, MR805654, Zbl0573.20022 Kantor, William M.; Taylor, Donald E. (1988), "Polynomial-time versions of Sylow's theorem", J. Algorithms 9 (1): 117, doi:10.1016/0196-6774(88)90002-8, ISSN0196-6774, MR925595, Zbl0642.20019 Kantor, William M. (1990), "Finding Sylow normalizers in polynomial time", J. Algorithms 11 (4): 523563, doi:10.1016/0196-6774(90)90009-4, ISSN0196-6774, MR1079450, Zbl0731.20005 Seress, kos (2003), Permutation Group Algorithms, Cambridge Tracts in Mathematics, 152, Cambridge University Press, ISBN978-0-521-66103-4, MR1970241, Zbl1028.20002

Sylvester's determinant theorem

211

Sylvester's determinant theorem


In matrix theory, Sylvester's determinant theorem is a theorem useful for evaluating certain types of determinants. It is named after James Joseph Sylvester. The theorem states that if A, B are matrices of size pn and np respectively, then where Ia is the identity matrix of order a.[1] It is closely related to the Matrix determinant lemma and its generalization. This theorem is useful in developing a Bayes estimator for multivariate Gaussian distributions. Sylvester (1857) stated this theorem without proof.

External links
[Related post [2]] on the blog of Terence Tao.

References
[1] David A. Harville. Matrix Algebra From a Statistician's Perspective. Springer, 2008, Pages 416 [2] https:/ / terrytao. wordpress. com/ 2010/ 12/ 17/ the-mesoscopic-structure-of-gue-eigenvalues/

Sylvester's law of inertia


Sylvester's law of inertia is a theorem in matrix algebra about certain properties of the coefficient matrix of a real quadratic form that remain invariant under a change of coordinates. Namely, if A is the symmetric matrix that defines the quadratic form, and S is any invertible matrix such that D=SAST is diagonal, then the number of negative elements in the diagonal of D is always the same, for all such S; and the same goes for the number of positive elements. This property is named after J. J. Sylvester who published its proof in 1852.[1] [2]

Statement of the theorem


Let A be a symmetric square matrix of order n with real entries. Any non-singular matrix S of the same size is said to transform A into another symmetric matrix B = SAST, also of order n, where ST is the transpose of S. If A is the coefficient matrix of some quadratic form of Rn, then B is the matrix for the same form after the change of coordinates defined by S. A symmetric matrix A can always be transformed in this way into a diagonal matrix D which has only entries 0, +1 and 1 along the diagonal. Sylvester's law of inertia states that the number of diagonal entries of each kind is an invariant of A, i.e. it does not depend on the matrix S used. The number of +1s, denoted n+, is called the positive index of inertia of A, and the number of 1s, denoted n, is called the negative index of inertia. The number of 0s, denoted n0, is the dimension of the kernel of A, and also the corank of A. These numbers satisfy an obvious relation

The difference sign(A) = n n+) is usually called the signature of A. (However, some authors use that term for the whole triple (n0, n+, n) consisting of the corank and the positive and negative indices of inertia of A.)

Sylvester's law of inertia If the matrix A has the property that every principal upper left kk minor k is non-zero then the negative index of inertia is equal to the number of sign changes in the sequence

212

Statement in terms of eigenvalues


The positive and negative indices of a symmetric matrix A are also the number of positive and negative eigenvalues of A. Any symmetric real matrix A has an eigendecomposition of the form QEQT where E is a diagonal matrix containing the eigenvalues of A, and Q is an orthonormal square matrix containing the eigenvectors. The matrix E can be written E=WDWT where D is diagonal with entries0,+1, or1, and W is diagonal with Wii=|Eii|. The matrix S=QW transforms D toA.

Law of inertia for quadratic forms


In the context of quadratic forms, a real quadratic form Q in n variables (or on an n-dimensional real vector space) can by a suitable change of basis be brought to the diagonal form

with each ai{0,1,1}. Sylvester's law of inertia states that the number of coefficients of a given sign is an invariant of Q, i.e. does not depend on a particular choice of diagonalizing basis. Expressed geometrically, the law of inertia says that all maximal subspaces on which the restriction of the quadratic form is positive definite (respectively, negative definite) have the same dimension. These dimensions are the positive and negative indices of inertia.

References
[1] Sylvester, J J (1852). "A demonstration of the theorem that every homogeneous quadratic polynomial is reducible by real orthogonal substitutions to the form of a sum of positive and negative squares" (http:/ / www. maths. ed. ac. uk/ ~aar/ sylv/ inertia. pdf). Philosophical Magazine (Ser. 4) 4 (23): 138142. doi:10.1080/14786445208647087. . Retrieved 2008-06-27. [2] Norman, C.W. (1986). Undergraduate algebra. Oxford University Press. pp.360361. ISBN0-19-853248-2.

External links
Sylvester's law (http://planetmath.org/encyclopedia/SylvestersLaw.html) on PlanetMath.

Takagi existence theorem

213

Takagi existence theorem


In class field theory, the Takagi existence theorem states that for any number field K there is a one-to-one inclusion reversing correspondence between the finite abelian extensions of K (in a fixed algebraic closure of K) and the generalized ideal class groups defined via a modulus of K. It is called an existence theorem because a main burden of the proof is to show the existence of enough abelian extensions of K.

Formulation
Here a modulus (or ray divisor) is a formal finite product of the valuations (also called primes or places) of K with positive integer exponents. The archimedean valuations that might appear in a modulus include only those whose completions are the real numbers (not the complex numbers); they may be identified with orderings on K and occur only to exponent one. The modulus m is a product of a non-archimedean (finite) part mf and an archimedean (infinite) part m. The non-archimedean part mf is a nonzero ideal in the ring of integers OK of K and the archimedean part m is simply a set of real embeddings of K. Associated to such a modulus m are two groups of fractional ideals. The larger one, Im, is the group of all fractional ideals relatively prime to m (which means these fractional ideals do not involve any prime ideal appearing in mf). The smaller one, Pm, is the group of principal fractional ideals (u/v) where u and v are nonzero elements of OK which are prime to mf, u v mod mf, and u/v > 0 in each of the orderings of m. (It is important here that in Pm, all we require is that some generator of the ideal has the indicated form. If one does, others might not. For instance, taking K to be the rational numbers, the ideal (3) lies in P4 because (3) = (3) and 3 fits the necessary conditions. But (3) is not in P4 since here it is required that the positive generator of the ideal is 1 mod 4, which is not so.) For any group H lying between Im and Pm, the quotient Im/H is called a generalized ideal class group. It is these generalized ideal class groups which correspond to abelian extensions of K by the existence theorem, and in fact are the Galois groups of these extensions. That generalized ideal class groups are finite is proved along the same lines of the proof that the usual ideal class group is finite, well in advance of knowing these are Galois groups of finite abelian extensions of the number field.

A well-defined correspondence
Strictly speaking, the correspondence between finite abelian extensions of K and generalized ideal class groups is not quite one-to-one. Generalized ideal class groups defined relative to different moduli can give rise to the same abelian extension of K, and this is codified a priori in a somewhat complicated equivalence relation on generalized ideal class groups. In concrete terms, for abelian extensions L of the rational numbers, this corresponds to the fact that an abelian extension of the rationals lying in one cyclotomic field also lies in infinitely many other cyclotomic fields, and for each such cyclotomic overfield one obtains by Galois theory a subgroup of the Galois group corresponding to the same field L. In the idelic formulation of class field theory, one obtains a precise one-to-one correspondence between abelian extensions and appropriate groups of ideles, where equivalent generalized ideal class groups in the ideal-theoretic language correspond to the same group of ideles.

Takagi existence theorem

214

Earlier work
A special case of the existence theorem is when m = 1 and H = P1. In this case the generalized ideal class group is the ideal class group of K, and the existence theorem says there exists a unique abelian extension L/K with Galois group isomorphic to the ideal class group of K such that L is unramified at all places of K. This extension is called the Hilbert class field. It was conjectured by David Hilbert to exist, and existence in this special case was proved by Furtwngler in 1907, before Takagi's general existence theorem. A further and special property of the Hilbert class field, not true of other abelian extensions of a number field, is that all ideals in a number field become principal in the Hilbert class field. It required Artin and Furtwngler to prove that principalization occurs.

History
The existence theorem is due to Takagi, who proved it in Japan during the isolated years of World War I. He presented it at the International Congress of Mathematicians in 1920, leading to the development of the classical theory of class field theory during the 1920s. At Hilbert's request, the paper was published in Mathematische Annalen in 1925.

References
Helmut Hasse, History of Class Field Theory, pp. 266279 in Algebraic Number Theory, eds. J. W. S. Cassels and A. Frhlich, Academic Press 1967. (See also the rich bibliography attached to Hasse's article.)

Three subgroups lemma


In mathematics, more specifically group theory, the three subgroups lemma is a result concerning commutators. It is a consequence of the HallWitt identity.

Notation
In that which follows, the following notation will be employed: If H and K are subgroups of a group G, the commutator of H and K will be denoted by [H,K]; if L is a third subgroup, the convention that [H,K,L] = [[H,K],L] will be followed. If x and y are elements of a group G, the conjugate of x by y will be denoted by . If H is a subgroup of a group G, then the centralizer of H in G will be denoted by CG(H).

Statement
Let X, Y and Z be subgroups of a group G, and assume and Then More generally, if .[1] , then if and , then .[2]

Proof and the HallWitt identity


HallWitt identity If , then

Three subgroups lemma Proof of the Three subgroups lemma Let , , and . Then and so . Therefore, , we conclude that , and by the HallWitt identity above, for all and and hence

215

it follows that

. Since these elements generate .

Notes
[1] Isaacs, Lemma 8.27, p. 111 [2] Isaacs, Corollary 8.28, p. 111

References
I. Martin Isaacs (1993). Algebra, a graduate course (1st edition ed.). Brooks/Cole Publishing Company. ISBN0-534-19002-2.

Trichotomy theorem
In mathematical finite group theory, the trichotomy theorem divides the simple groups of characteristic2 type and rank at least3 into three classes. It was proved by Aschbacher(1981, 1983) for rank3 and by Gorenstein & Lyons (1983) for rank at least4. The three classes are groups of GF(2) type (classified by Timmesfeld and others), groups of "standard type" for some odd prime (classified by the GilmanGriess theorem and work by several others), and groups of uniqueness type, where Aschbacher proved that there are no simple groups.

References
Aschbacher, Michael (1981), "Finite groups of rank 3. I", Inventiones Mathematicae 63 (3): 357402, doi:10.1007/BF01389061, ISSN0020-9910, MR620676 Aschbacher, Michael (1983), "Finite groups of rank 3. II", Inventiones Mathematicae 71 (1): 51163, doi:10.1007/BF01393339, ISSN0020-9910, MR688262 Gorenstein, D.; Lyons, Richard (1983), "The local structure of finite groups of characteristic 2 type" [1], Memoirs of the American Mathematical Society 42 (276): vii+731, ISBN978-0-8218-2276-0, ISSN0065-9266, MR690900

References
[1] http:/ / books. google. com/ books?isbn=978-0821822760

Walter theorem

216

Walter theorem
In mathematics, the Walter theorem, proved by Walter(1967, 1969), describes the finite groups whose Sylow 2-subgroup is abelian. Bender (1970) used Bender's method to give a simpler proof.

Statement
Walter's theorem states that if G is a finite group whose 2-sylow subgroups are abelian, then G/O(G) has a normal subgroup of odd index that is a product of groups each of which is a 2-group of one of the simple groups PSL2(q) for q = 2n or q = 3 or 5 mod 8, or the Janko group J1, or Ree groups 2G2(32n+1). The original statement of Walter's theorem did not quite identify the Ree groups, but only stated that the corresponding groups have a similar subgroup structure as Ree groups. Thompson(1967, 1972, 1977) and Bombieri, Odlyzko & Hunt (1980) later showed that they are all Ree groups, and Enguehard (1986) gave a unified exposition of this result.

References
Bender, Helmut (1970), "On groups with abelian Sylow 2-subgroups", Mathematische Zeitschrift 117: 164176, doi:10.1007/BF01109839, ISSN0025-5874, MR0288180 Bombieri, Enrico; Odlyzko, Andrew; Hunt, D. (1980), "Thompson's problem (2=3)", Inventiones Mathematicae 58 (1): 77100, doi:10.1007/BF01402275, ISSN0020-9910, MR570875 Enguehard, Michel (1986), "Caractrisation des groupes de Ree", Astrisque (142): 49139, ISSN0303-1179, MR873958 Thompson, John G. (1967), "Toward a characterization of E2*(q)", Journal of Algebra 7: 406414, doi:10.1016/0021-8693(67)90080-4, ISSN0021-8693, MR0223448 Thompson, John G. (1972), "Toward a characterization of E2*(q). II", Journal of Algebra 20: 610621, doi:10.1016/0021-8693(72)90074-9, ISSN0021-8693, MR0313377 Thompson, John G. (1977), "Toward a characterization ofE2*(q). III", Journal of Algebra 49 (1): 162166, doi:10.1016/0021-8693(77)90276-9, ISSN0021-8693, MR0453858 Walter, John H. (1967), "Finite groups with abelian Sylow 2-subgroups of order 8", Inventiones Mathematicae 2: 332376, doi:10.1007/BF01428899, ISSN0020-9910, MR0218445 Walter, John H. (1969), "The characterization of finite groups with abelian Sylow 2-subgroups.", Annals of Mathematics. Second Series 89: 405514, ISSN0003-486X, JSTOR1970648, MR0249504

Wedderburn's little theorem

217

Wedderburn's little theorem


In mathematics, Wedderburn's little theorem states that every finite domain is a field. In other words, for finite rings, there is no distinction between domains, skew-fields and fields. The ArtinZorn theorem generalizes the theorem to alternative rings.

History
The original proof was given by Joseph Wedderburn in 1905, who went on to prove it two other ways. Another proof was given by Leonard Eugene Dickson shortly after Wedderburn's original proof, and Dickson acknowledged Wedderburn's priority. However, as noted in (Parshall 1983), Wedderburn's first proof was incorrect it had a gap and his subsequent proofs came after he had read Dickson's correct proof. On this basis, Parshall argues that Dickson should be credited with the first correct proof. A simplified version of the proof was later given by Ernst Witt. Witt's proof is sketched below. Alternatively, the theorem is a consequence of the SkolemNoether theorem.

Sketch of proof
Let be a finite domain. For each nonzero , the map

is injective; thus, surjective. Hence, has a left inverse. By the same argument, skew-field. Since the center of is a field, is a vector space over objective is then to show the center, the centralizer . If is the order of , then A has order of x has order where d divides n. Viewing

has a right inverse. A is thus a with finite dimension n. Our . For each , and that is not in as groups

under multiplication, we can write the class equation

where the sum is taken over all representatives and that divides

that is not in

and d are the numbers discussed above. . After cancellation, we see .

both admit factorization in terms of cyclotomic polynomials and , so it must divide

. So we reach contradiction unless

References
Parshall, K. H. (1983), In pursuit of the finite division algebra theorem and beyond: Joseph H M Wedderburn, Leonard Dickson, and Oswald Veblen, Archives of International History of Science, 33, pp.27499

External links
Proof of Wedderburn's Theorem at Planet Math [1]

References
[1] http:/ / planetmath. org/ ?op=getobj& from=objects& id=3627

Weil conjecture on Tamagawa numbers

218

Weil conjecture on Tamagawa numbers


In mathematics, the Weil conjecture on Tamagawa numbers is a result about algebraic groups formulated by Andr Weil in the late 1950s and proved in 1989. It states that the Tamagawa number (G) is 1 for any simply connected semisimple algebraic group G defined over a number field K. Here simply connected is in the algebraic group theory sense of not having a proper algebraic covering, which is not always the topologists' meaning.

History
Weil checked this in enough classical group cases to propose the conjecture. In particular for spin groups it implies the known SmithMinkowskiSiegel mass formula. Robert Langlands (1966) introduced harmonic analysis methods to show it for Chevalley groups. J. G. M. Mars gave further results during the 1960s. K. F. Lai (1980) extended the class of known cases to quasisplit reductive groups. Kottwitz (1988) proved it for all groups satisfying the Hasse principle, which at the time was known for all groups without E8 factors. V. I. Chernousov (1989) removed this restriction, by proving the Hasse principle for the resistant E8 case (see strong approximation in algebraic groups), thus completing the proof of Weil's conjecture.

References
Hazewinkel, Michiel, ed. (2001), "Tamagawa number" [1], Encyclopaedia of Mathematics, Springer, ISBN978-1556080104 Chernousov, V. I. (1989), "The Hasse principle for groups of type E8", Soviet Math. Dokl. 39: 592596, MR1014762 Kottwitz, Robert E. (1988), "Tamagawa numbers", Ann. Of Math. (2) (Annals of Mathematics) 127 (3): 629646, doi:10.2307/2007007, JSTOR2007007, MR0942522. Lai, K. F. (1980), "Tamagawa number of reductive algebraic groups" [2], Compositio Mathematica 41 (2): 153188, MR581580 Langlands, R. P. (1966), "The volume of the fundamental domain for some arithmetical subgroups of Chevalley groups", Algebraic Groups and Discontinuous Subgroups, Proc. Sympos. Pure Math., Providence, R.I.: Amer. Math. Soc., pp.143148, MR0213362 Voskresenskii, V. E. (1991), Algebraic Groups and their Birational Invariants, AMS translation

References
[1] http:/ / eom. springer. de/ T/ t092060. htm [2] http:/ / www. numdam. org/ item?id=CM_1980__41_2_153_0

Witt's theorem

219

Witt's theorem
"Witt's theorem" or "the Witt theorem" may also refer to the BourbakiWitt fixed point theorem of order theory. Witt theorem, named after Ernst Witt, is a basic result in the algebraic theory of quadratic forms: any isometry between two subspaces of a nonsingular quadratic space over a field k may be extended to an isometry of the whole space. An analogous statement holds also for skew-symmetric, Hermitian and skew-Hermitian bilinear forms over arbitrary fields. The theorem applies to classification of quadratic forms over k and in particular allows one to define the Witt group W(k) which controls the "stable" theory of quadratic forms over the field k.

Statement of the theorem


Let (V, b) be a finite-dimensional vector space over an arbitrary field k together with a nondegenerate symmetric or skew-symmetric bilinear form. If f: UU' is an isometry between two subspaces of V then f extends to an isometry of V. Witt's theorem implies that the dimension of a maximal isotropic subspace of V is an invariant, called the index or Witt index of b, and moreover, that the isometry group of (V, b) acts transitively on the set of maximal isotropic subspaces. This fact plays an important role in the structure theory and representation theory of the isometry group and in the theory of reductive dual pairs.

Witt's cancellation theorem


Let (V, q), (V1, q1), (V2, q2) be three quadratic spaces over a field k. Assume that Then the quadratic spaces (V1, q1) and (V2, q2) are isometric: In other words, the direct summand (V, q) appearing in both sides of an isomorphism between quadratic spaces may be "cancelled".

Witt's decomposition theorem


Let (V, q) be a quadratic space over a field k. Then it admits a Witt decomposition:

where V0=ker q is the radical of q, (Va, qa) is an anisotropic quadratic space and (Vh, qh) is a hyperbolic quadratic space. Moreover, the anisotropic summand and the hyperbolic summand in a Witt decomposition of (V, q) are determined uniquely up to isomorphism.

References
O. Timothy O'Meara, Introduction to Quadratic Forms, Springer-Verlag, 1973

Z* theorem

220

Z* theorem
In mathematics, George Glauberman's Z* theorem states that if G is a finite group and T is a Sylow 2-subgroup of G containing an involution not conjugate in G to any other element of T, then the involution lies in Z*(G). The subgroup Z*(G) is the inverse image in G of the center of G/O(G), where O(G) is the maximal normal subgroup of G of odd order. This generalizes the BrauerSuzuki theorem (and the proof uses the Brauer-Suzuki theorem to deal with some small cases). The original paper (Glauberman 1966) gave several criteria for an element to lie outside Z*(G). Its theorem 4 states: For an element t in T, it is necessary and sufficient for t to lie outside Z*(G) that there is some g in G and abelian subgroup U of T satisfying the following properties: 1. 2. 3. 4. g normalizes both U and the centralizer CT(U), that is g is contained in N = NG(U)NG(CT(U)) t is contained in U and tg gt U is generated by the N-conjugates of t the exponent of U is equal to the order of t

Moreover g may be chosen to have prime power order if t is in the center of T, and g may be chosen in T otherwise. A simple corollary is that an element t in T is not in Z*(G) if and only if there is some s t such that s and t commute and s and t are G conjugate. A generalization to odd primes was recorded in (Guralnick & Robinson 1993): if t is an element of prime order p and the commutator [t,g] has order coprime to p for all g, then t is central modulo the p-core. This was also generalized to odd primes and to compact Lie groups in (Mislin & Thvenaz 1991), which also contains several useful results in the finite case.

References
Dade, Everett C. (1971), "Character theory pertaining to finite simple groups", in Powell, M. B.; Higman, Graham, Finite simple groups. Proceedings of an Instructional Conference organized by the London Mathematical Society (a NATO Advanced Study Institute), Oxford, September 1969., Boston, MA: Academic Press, pp.249327, ISBN978-0-12-563850-0, MR0360785 gives a detailed proof of the BrauerSuzuki theorem. Glauberman, George (1966), "Central elements in core-free groups", Journal of Algebra 4: 403420, doi:10.1016/0021-8693(66)90030-5, ISSN0021-8693, MR0202822, Zbl0145.02802 Guralnick, Robert M.; Robinson, Geoffrey R. (1993), "On extensions of the Baer-Suzuki theorem", Israel Journal of Mathematics 82 (1): 281297, doi:10.1007/BF02808114, ISSN0021-2172, MR1239051, Zbl0794.20029 Mislin, Guido; Thvenaz, Jacques (1991), "The Z*-theorem for compact Lie groups", Mathematische Annalen 291 (1): 103111, doi:10.1007/BF01445193, ISSN0025-5831, MR1125010

Zassenhaus lemma

221

Zassenhaus lemma
In mathematics, the butterfly lemma or Zassenhaus lemma, named after Hans Julius Zassenhaus, is a technical result on the lattice of subgroups of a group or the lattice of submodules of a module, or more generally for any modular lattice.[1] Lemma: Suppose operators and Suppose and is a group with are subgroups.

Hasse diagram of the Zassenhaus "butterfly" lemma - smaller subgroups are towards the top of the diagram

and are stable subgroups. Then, is isomorphic to Zassenhaus proved this lemma specifically to give the smoothest proof of the Schreier refinement theorem. The 'butterfly' becomes apparent when trying to draw the Hasse diagram of the various groups involved.

Notes
[1] See Pierce, p. 27, exercise 1.

References
Pierce, R. S., Associative algebras, Springer, pp.27, ISBN0387906932. Goodearl, K. R.; Warfield, Robert B. (1989), An introduction to noncommutative noetherian rings, Cambridge University Press, pp.51, 62, ISBN9780521369251. Lang, Serge, Algebra, Graduate Texts in Mathematics (Revised 3rd ed.), Springer-Verlag, pp.2021, ISBN9780387953854. Carl Clifton Faith, Nguyen Viet Dung, Barbara Osofsky. Rings, Modules and Representations. p.6. AMS Bookstore, 2009. ISBN 0821843702

Zassenhaus lemma

222

External links
Zassenhaus Lemma and proof at http://www.artofproblemsolving.com/Wiki/index.php/ Zassenhaus%27s_Lemma

ZJ theorem
In mathematics, George Glauberman's ZJ theorem states that if a finite group G is p-constrained and p-stable and has a normal p-subgroup for some odd prime p, then Op(G)Z(J(S)) is a normal subgroup of G, for any Sylow p-subgroup S.

Notation and definitions


J(S) is the Thompson subgroup of a p-group S: the subgroup generated by the abelian subgroups of maximal order. Z(H) means the center of a group H. Op is the maximal normal subgroup of G of order coprime to p, the p-core Op is the maximal normal p-subgroup of G, the p-core. Op,p(G) is the maximal normal p-nilpotent subgroup of G, the p,p-core, part of the upper p-series. For an odd prime p, a group G with Op(G) 1 is said to be p-stable if whenever P is a p-subgroup of G such that POp(G) is normal in G, and [P,x,x] = 1, then the image of x in NG(P)/CG(P) is contained in a normal p-subgroup of NG(P)/CG(P). For an odd prime p, a group G with Op(G) 1 is said to be p-constrained if the centralizer CG(P) is contained in Op,p(G) whenever P is a Sylow p-subgroup of Op,p(G).

References
Glauberman, George (1968), "A characteristic subgroup of a p-stable group" [1], Canadian Journal of Mathematics 20: 11011135, ISSN0008-414X, MR0230807 Gorenstein, D. (1980), Finite Groups, New York: Chelsea, ISBN978-0-8284-0301-6, MR81b:20002 Thompson, John G. (1969), "A replacement theorem for p-groups and a conjecture", Journal of Algebra 13: 149151, doi:10.1016/0021-8693(69)90068-4, ISSN0021-8693, MR0245683

References
[1] http:/ / www. cms. math. ca/ cjm/ v20/ p1101

Article Sources and Contributors

223

Article Sources and Contributors


Abel's binomial theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612000 Contributors: Brilliant trees, Cronholm144, Geometry guy, Giftlite, Kiensvay, Ktr101, Michael Hardy, PrimeHunter, Xanthoxyl, 5 anonymous edits AbelRuffini theorem Source: http://en.wikipedia.org/w/index.php?oldid=455715455 Contributors: Albmont, Andrewpmk, AxelBoldt, BeteNoir, BigDom, CRGreathouse, Cf. Hay, Charles Matthews, Cheesefondue, Colonies Chris, Crasshopper, Dmharvey, Doctormatt, Dominus, Duoduoduo, Edudobay, Ettrig, EverGreg, Fibonacci, Gene Nygaard, Geschichte, Giftlite, Helder.wiki, IRP, Icairns, Inquam, JackSchmidt, JamesBWatson, Kutu su, LamilLerran, Lunae, Marc Venot, Marc van Leeuwen, Mets501, Mh, Michael Hardy, Mike4ty4, Nbarth, Nsh, Oleg Alexandrov, Paddles, Psychonaut, Quantpole, RDBury, Rlinfinity, RodC, Salix alba, Sandrobt, Schneelocke, Simetrical, Stormwyrm, Sue Gardner, Swordsmankirby, Tbjablin, Thegeneralguy, Tide rolls, Ttwo, VKokielov, XJamRastafire, Zchenyu, 51 anonymous edits Abhyankar's conjecture Source: http://en.wikipedia.org/w/index.php?oldid=455857476 Contributors: Bender235, CRGreathouse, Charles Matthews, David Haslam, Giftlite, Mackdiesel5, MathMartin, Plclark, R.e.b., Rschwieb, 2 anonymous edits Acyclic model Source: http://en.wikipedia.org/w/index.php?oldid=455866493 Contributors: EmanWilm, Giftlite, GregorB, Inarchus, Jitse Niesen, Michael Hardy, Ringspectrum, Sodin, Techman224, Vaughan Pratt, 7 anonymous edits Ado's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612088 Contributors: CRGreathouse, Charles Matthews, Geometry guy, Giftlite, Jitse Niesen, Marc van Leeuwen, R.e.b., TakuyaMurata, Tarret, Tesseran, , 1 anonymous edits AlperinBrauerGorenstein theorem Source: http://en.wikipedia.org/w/index.php?oldid=455629144 Contributors: Algebraist, Davcrav, Giftlite, Headbomb, JackSchmidt, KathrynLybarger, Messagetolove, Nbarth, Sodin, Vanish2 AmitsurLevitzki theorem Source: http://en.wikipedia.org/w/index.php?oldid=455760783 Contributors: Darij, Giftlite, Headbomb, Michael Hardy, R.e.b., Rschwieb Artin approximation theorem Source: http://en.wikipedia.org/w/index.php?oldid=455767217 Contributors: Beetstra, BeteNoir, Charles Matthews, Gauge, Giftlite, Jowa fan, Oleg Alexandrov, Sodin, 9 anonymous edits ArtinWedderburn theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612164 Contributors: BeteNoir, Charles Matthews, Darij, Geometry guy, Giftlite, JamieVicary, John Baez, Justpasha, Psychonaut, Qutezuce, Rgdboer, Rschwieb, The Anome, Vivacissamamente, Waltpohl, 20 anonymous edits ArtinZorn theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609464 Contributors: David Eppstein, Giftlite, Headbomb, Michael Hardy, Sodin, Tobias Bergemann, Waltpohl, 1 anonymous edits BaerSuzuki theorem Source: http://en.wikipedia.org/w/index.php?oldid=455763227 Contributors: Gregbard, Headbomb, Krasnoludek, R.e.b., Rschwieb BeauvilleLaszlo theorem Source: http://en.wikipedia.org/w/index.php?oldid=455859427 Contributors: Giftlite, Jowa fan, Michael Hardy, R.e.b., Ryan Reich, Sodin, 1 anonymous edits Binomial inverse theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612331 Contributors: Ballista121, CBM, Cmansley, Entropeneur, Geometry guy, Giftlite, Headbomb, JRSpriggs, Jmath666, Jonas August, MaxSem, Michael Hardy, Neparis, O18, Robinh, TedPavlic, Thecheesykid, 3 anonymous edits Binomial theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612424 Contributors: -Ozone-, .:Ajvol:., 0, Acdx, Aciel, Adashiel, Aj.sujit, AlanUS, Alansohn, Andres, Anonymous Dissident, Antandrus, Arthur Rubin, AxelBoldt, Babababoshka, Badgernet, BeteNoir, Bharatveer, BiT, Binary TSO, Blaxthos, Bombshell, Borisblue, Bsodmike, CRGreathouse, Carbuncle, Cdang, Charibdis, Charles Matthews, Chemninja, Chuchogl, Conversion script, Cpoynton, Crabula, Cronholm144, DEMcAdams, Danno uk, Dirkbb, Discospinster, Dmcq, Duncharris, Ebony Jackson, EcneicsFlogCitanaf, Elockid, Eric119, FactSpewer, FlamingSilmaril, Geometry guy, Gesslein, Giants27, Giftlite, Gilliam, Goldencako, Gombang, Gregbard, Gsmgm, Gurch, Haham hanuka, Hawk8103, ILovePlankton, Iridescent, J.delanoy, JForget, Jagged 85, Jay Gatsby, Jim.belk, Joe056, JorgeGG, Jusdafax, Justin W Smith, KIMWOONGJI, Kbdank71, Keilana, Kenyon, Kiensvay, Kingpin13, Kloddant, Knight1993, La goutte de pluie, Leyo, Linas, Loadmaster, LouScheffer, Lupin, MC10, MER-C, MSGJ, Marc van Leeuwen, MarkSweep, MarnetteD, Matevzk, Meaghan, Meldor, Mets501, Michael Hardy, Michael Slone, Mikez, Minthellen, Molotron, Myrizio, NOrbeck, Najoj, Nbarth, Nirvana89, Nk, Nomet, Nonagonal Spider, Oceanwitch, Oleg Alexandrov, Opabinia regalis, PMDrive1061, PTP2009, PericlesofAthens, Pevernagie, Pion, Pleasantville, Poetaris, Poor Yorick, Quietbritishjim, R.e.b., RMFan1, RexNL, Rich Farmbrough, RickDC, Robinh, Salvatore Ingala, Saric, Sbealing, Shahab, SimonP, StandardizerII, Stebulus, Sawomir Biay, THEN WHO WAS PHONE?, TNeloms, TakuyaMurata, Tbsmith, Tcnuk, The Thing That Should Not Be, The absolute real deal, TheSuave, Thetruthseer, Tobias Bergemann, Ulfalizer, Urdutext, Verdy p, WardenWalk, Warut, Wiki alf, Wings Upon My Feet, Wroscel, Zchenyu, Zvika, 325 anonymous edits Birch's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455796486 Contributors: Akrabbim, Charles Matthews, Mon4, Sodin, 1 anonymous edits Birkhoff's representation theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612525 Contributors: Chris the speller, David Eppstein, Gabelaia, Geometry guy, Giftlite, Headbomb, Igorpak, Mhym Boolean prime ideal theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612600 Contributors: Aleph4, AugPi, AxelBoldt, CBM, CRGreathouse, Charles Matthews, Chinju, David Eppstein, Dysprosia, Eric119, Geometry guy, Giftlite, Headbomb, Hennobrandsma, Hugo Herbelin, Jon Awbrey, Kope, MarkSweep, Markus Krtzsch, Mhss, Michael Hardy, Ott2, RobHar, RoodyAlien, TexD, Tkuvho, Tobias Bergemann, Trioculite, Trovatore, Vivacissamamente, 15 anonymous edits BorelWeil theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612643 Contributors: Algebraist, Arcfrk, David Eppstein, Geometry guy, Giftlite, Michael Hardy, Nicolaisvendsen, Psychonaut, SemperBlotto, Tabletop, 1 anonymous edits BorelWeilBott theorem Source: http://en.wikipedia.org/w/index.php?oldid=455612716 Contributors: Algebraist, BeteNoir, Bunny Angel13, Charles Matthews, Fropuff, Gauge, Gene Nygaard, Geometry guy, Giftlite, M m hawk, Masnevets, Michael Hardy, Myasuda, Nbarth, Psychonaut, RobHar, Silly rabbit, Waltpohl, 4 anonymous edits Brauer's theorem on induced characters Source: http://en.wikipedia.org/w/index.php?oldid=455614459 Contributors: BeteNoir, Charles Matthews, Geffrey, Geometry guy, Giftlite, JackSchmidt, Masnevets, Messagetolove, 5 anonymous edits Brauer's three main theorems Source: http://en.wikipedia.org/w/index.php?oldid=455614539 Contributors: Arcfrk, Charles Matthews, Davcrav, DavidCBryant, Geometry guy, Headbomb, Messagetolove, Oleg Alexandrov, R.e.b., Thecheesykid BrauerCartanHua theorem Source: http://en.wikipedia.org/w/index.php?oldid=455614613 Contributors: 1ForTheMoney, Geometry guy, Headbomb, JackSchmidt, Omnipaedista, PMDrive1061, Pillsberry, Pmanderson, Sawomir Biay, THEN WHO WAS PHONE? BrauerNesbitt theorem Source: http://en.wikipedia.org/w/index.php?oldid=455614718 Contributors: Charles Matthews, Geometry guy, Michael Slone, R.e.b., 2 anonymous edits BrauerSiegel theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609585 Contributors: 3mta3, Charles Matthews, Giftlite, JackSchmidt, Michael Hardy, Sodin, Vanish2, Vesath BrauerSuzuki theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609550 Contributors: Bird of paradox, Charles Matthews, Davcrav, Headbomb, JackSchmidt, Kmhkmh, R.e.b., Sodin, 1 anonymous edits BrauerSuzukiWall theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609607 Contributors: Gregbard, Headbomb, Michael Hardy, R.e.b., Slawekb, Sodin Burnside theorem Source: http://en.wikipedia.org/w/index.php?oldid=455614918 Contributors: Ashsong, BeteNoir, Charles Matthews, Gebstadter, Geometry guy, Giftlite, Lucifer Anh, Messagetolove, Oleg Alexandrov, SchfiftyThree, Tobias Bergemann, 12 anonymous edits Cartan's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455619115 Contributors: Asmeurer, BeteNoir, Charles Matthews, Fropuff, Geometry guy, Giftlite, Headbomb, JCSantos, R'n'B, Silly rabbit, Sawomir Biay, Topology Expert, Unyoyega, Zoran.skoda CartanDieudonn theorem Source: http://en.wikipedia.org/w/index.php?oldid=455615149 Contributors: Charles Matthews, Crazy runner, Geometry guy, Keyi, MathMartin, Michael Hardy, Nbarth, Oleg Alexandrov, Silly rabbit, 2 anonymous edits

Article Sources and Contributors


Cauchy's theorem (group theory) Source: http://en.wikipedia.org/w/index.php?oldid=455615270 Contributors: AdamSmithee, AlexDenney, Algebraist, Andrei G Kustov, Blotwell, Egpetersen, Eric119, Fell Collar, Geometry guy, Giftlite, JackSchmidt, Kilva, LOL, Michael Hardy, Nm420, PV=nRT, R'n'B, Rghthndsd, Scottie 000, TakuyaMurata, Tango, Vanish2, 11 anonymous edits Cayley's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455615349 Contributors: Ambarish, Atyatya, AxelBoldt, BeteNoir, Calle, Charles Matthews, Chas zzz brown, ChickenMerengo, David Pierce, Dbenbenn, Dominus, Dr Zimbu, Dysprosia, Geometry guy, Giftlite, Hairy Dude, Headbomb, Helder.wiki, Inquam, JackSchmidt, Jane Bennet, Joen235, Joshuagmath, Looxix, Maxal, Mhss, Michael Hardy, Myasuda, Patrick, Scubbo, Simetrical, Thehotelambush, Tosha, Wapcaplet, Zvika, 38 anonymous edits CayleyHamilton theorem Source: http://en.wikipedia.org/w/index.php?oldid=455615451 Contributors: Albmont, Alexander Chervov, Algebraist, Aliotra, Attilios, AxelBoldt, BeteNoir, CRGreathouse, Carbuncle, Charles Matthews, Chetvorno, Choco.litt, Closedmouth, Cuzkatzimhut, David Shay, David Sneek, Dominus, Dysprosia, Eecs student, Equendil, Gene Nygaard, Geometry guy, Giftlite, Headbomb, Ivan tambuk, Jcobb, JdH, Jpkotta, Jujutacular, Lambiam, Lechatjaune, LilHelpa, Ling.Nut, Lionelbrits, Looxix, Malatinszky, Marc van Leeuwen, MathKnight, MehdiPedia, Merrybrit, Michael Hardy, Nschoem, NyAp, Obradovic Goran, Oleg Alexandrov, PAR, PhilDWraight, Pmdboi, Psychonaut, Quasicharacter, Rik Bos, Ryan Reich, SDaniel, Shreevatsa, Sawomir Biay, Tarquin, The suffocated, Turgidson, Txomin, Werdenwissen, Wshun, 82 anonymous edits ChevalleyShephardTodd theorem Source: http://en.wikipedia.org/w/index.php?oldid=455616003 Contributors: Arcfrk, Closedmouth, Geometry guy, Giftlite, Headbomb, Michael Hardy, R.e.b., Rjwilmsi, 4 anonymous edits ChevalleyWarning theorem Source: http://en.wikipedia.org/w/index.php?oldid=455767620 Contributors: BeteNoir, Charles Matthews, David Brink, David Eppstein, Dtrebbien, Giftlite, Headbomb, JackSchmidt, Jowa fan, LilHelpa, Michael Hardy, R.e.b., Sodin, 5 anonymous edits Classification of finite simple groups Source: http://en.wikipedia.org/w/index.php?oldid=455616173 Contributors: 63.162.153.xxx, 66.38.184.xxx, A3 nm, Aboriginal Noise, Akriasas, Andi5, Arneth, Astronautics, AxelBoldt, BananaFiend, Bruguiea, C S, Calabraxthis, Cambyses, Charles Matthews, Chas zzz brown, Conversion script, Dominus, Dratman, Drschawrz, Dysprosia, Eugene van der Pijll, Gareth Jones, Gene Ward Smith, Geometry guy, Giftlite, Gilliam, Gro-Tsen, Huppybanny, JackSchmidt, Jim.belk, Jitse Niesen, John of Reading, JoshuaZ, Jowa fan, Joy, Kidburla, Kilva, Lightmouse, Looxix, Loren Rosen, Maproom, Matt Crypto, Melchoir, Messagetolove, Michael Hardy, Michael Larsen, Mike40033, Nbarth, PierreAbbat, Pjacobi, Poulpy, R.e.b., Rjwilmsi, Salix alba, Schneelocke, Slightsmile, Tarquin, Timwi, Tobias Bergemann, Topbanana, Twri, VeryVerily, VladimirReshetnikov, Zundark, 59 anonymous edits Cohn's irreducibility criterion Source: http://en.wikipedia.org/w/index.php?oldid=455860476 Contributors: Cornflake pirate, David Eppstein, Dtrebbien, EdJohnston, Gandalf61, Giftlite, JaGa, John Vandenberg, Ntsimp, Oleg Alexandrov, Omid Hatami, RHaworth, Sodin, Ultimus, 7 anonymous edits Cramer's rule Source: http://en.wikipedia.org/w/index.php?oldid=455616575 Contributors: A19grey, AlanUS, Alansohn, Alphax, Anonymous Dissident, Arthur Rubin, Asmeurer, Barak Sh, Ben Spinozoan, Bender2k14, Benzi455, BeteNoir, Bh3u4m, Carmichael95, Ciphers, Cloudguitar, Cryptic, D0762, Dingenis, Djwinters, Dmaher, ERcheck, EconoPhysicist, Elphion, Epbr123, Flavio Guitian, FlowRate, Foxjwill, Franklin.vp, Gauge, Geometry guy, GermanX, Giftlite, Grafen, Hillbrand, Ioscius, J1e9n8s5, JHMM13, JakeVortex, Javanbakht, Jaysweet, Jcobb, Jks, Jmath666, Joriki, KathrynLybarger, Khabgood, Koavf, LOL, Lear's Fool, LilHelpa, Lleeoo, Lradrama, Malo, Marc van Leeuwen, Maxno, Mebden, Michael Hardy, Michael Slone, Mike Rosoft, Mormegil, Nemolus77, Obradovic Goran, Oleg Alexandrov, Oli Filth, Ooz dot ie, Patrick, Pharaoh of the Wizards, Protonk, Rludlow, Rogper, Saravask, Serebr, Silly rabbit, Snailwalker, Spoon!, Steve.jaramillov, Sawomir Biay, Tarquin, TedPavlic, TheMaestro, Thumperward, TotientDragooned, Tristanreid, User A1, Vivacissamamente, Vsb, Waggers, Wdspann, Wolfrock, Xaos, Yecril, Yoshiki Sunada, Yossiea, Zipcodeman, 163 anonymous edits Crystallographic restriction theorem Source: http://en.wikipedia.org/w/index.php?oldid=455616670 Contributors: Armando-Martin, Ben Standeven, BeteNoir, Bgm2011, Charles Matthews, Commander Keane, Dyaa, Euyyn, Geometry guy, Giftlite, Greg Kuperberg, Headbomb, Helder.wiki, Howard Landman, Joseph Myers, Jowa fan, KSmrq, Keenan Pepper, Michael Hardy, Mvpranav, NRLer, Oleg Alexandrov, Paolo.dL, Patrick, Rjwilmsi, Tesscass, Unco, Zvika, 16 anonymous edits Descartes' rule of signs Source: http://en.wikipedia.org/w/index.php?oldid=455616871 Contributors: 478jjjz, A. Pichler, Adam Field, Alansohn, Alethiareg, Bender235, Bento00, Charles Matthews, Chenxlee, Chuunen Baka, DavidMcKenzie, Dzordzm, Estudiarme, FHGJ, GNB, Gandalf61, Gazpacho, Geometry guy, Giftlite, Haihe, Jeekc, Jeepday, JimVC3, Lechatjaune, Lou Crazy, Magic Window, Mets501, Mglg, Michael Hardy, Mild Bill Hiccup, Nishantsah, PV=nRT, Palladinus, Pdebart, Reywas92, Salix alba, Silly rabbit, Sinblox, Spacepotato, Tide rolls, Tosha, Vroo, 51 anonymous edits Dirichlet's unit theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609753 Contributors: 4meter4, Bender235, BeteNoir, CRGreathouse, Charles Matthews, CharlesGillingham, Chenxlee, Crisfilax, Fropuff, Gene Ward Smith, Giftlite, Headbomb, Michael Hardy, R.e.b., Revolver, Rich Farmbrough, Ringspectrum, RobHar, Sodin, Vanish2, Vargenau, 11 anonymous edits Engel theorem Source: http://en.wikipedia.org/w/index.php?oldid=455801855 Contributors: Arcfrk, BeteNoir, CSTAR, Charles Matthews, Darij, Geometry guy, Giftlite, Headbomb, JackSchmidt, Mets501, Michael Hardy, Nbarth, R.e.b., Safemariner, Wjcook, 9 anonymous edits Factor theorem Source: http://en.wikipedia.org/w/index.php?oldid=455694654 Contributors: Arthena, Balrog30, Barak, BeteNoir, Bsadowski1, Celestianpower, Charles Matthews, Chuunen Baka, David Radcliffe, Discospinster, Geometry guy, Giftlite, Iain.dalton, Jacj, Kiensvay, Kurosuke88, LOL, La Pianista, La goutte de pluie, Magister Mathematicae, Mairi, MarSch, Mpatel, Qwfp, RMFan1, S243a, Static shock1994, Tooto, 59 anonymous edits FeitThompson theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609773 Contributors: AndrewWTaylor, BeteNoir, Bird of paradox, Charles Matthews, Gauge, Giftlite, JackSchmidt, Jim.belk, Kilva, Linas, Malatinszky, Messagetolove, Michael Hardy, Nicholas Jackson, PerryTachett, Psychonaut, R.e.b., Rjwilmsi, Sodin, Sullivan.t.j, Tyomitch, Woohookitty, Zundark, 9 anonymous edits Fitting's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455624594 Contributors: CBM, Charles Matthews, Geometry guy, MarSch, Michael Hardy, Nbarth, Paul Klenk, R.e.b., Zundark Focal subgroup theorem Source: http://en.wikipedia.org/w/index.php?oldid=455619355 Contributors: Giftlite, Headbomb, JHunterJ, JackSchmidt, Michael Hardy, Nbarth, Rschwieb Frobenius determinant theorem Source: http://en.wikipedia.org/w/index.php?oldid=455619264 Contributors: Bearcat, David Eppstein, Dd314, Ebe123, Geometry guy, Giftlite, Headbomb, LilHelpa, Michael Hardy, Pichpich, R.e.b., Sadads Frobenius theorem (real division algebras) Source: http://en.wikipedia.org/w/index.php?oldid=455618771 Contributors: BartekChom, Charles Matthews, Geometry guy, Giftlite, Incnis Mrsi, JAMnx, Jim.belk, JukoFF, KSmrq, MarkSweep, MarkusQ, MathMartin, Michael Hardy, Mild Bill Hiccup, OneWeirdDude, Rgdboer, RobHar, WestwoodMatt, 17 anonymous edits Fundamental lemma (Langlands program) Source: http://en.wikipedia.org/w/index.php?oldid=455796915 Contributors: Arcfrk, Charles Matthews, David Eppstein, Dominus, Giftlite, Headbomb, LJosil, Michael Hardy, Phanjuy, RobHar, Rumping, Slawekb, Sodin, Stevey7788, Sawomir Biay, TakuyaMurata, Tobias Bergemann, 11 anonymous edits Fundamental theorem of algebra Source: http://en.wikipedia.org/w/index.php?oldid=455618556 Contributors: .:Ajvol:., 64.12.102.xxx, Abovechief, Adam majewski, Ahoerstemeier, Alansohn, Alexb@cut-the-knot.com, Algebraist, Alink, Andy Fugard, Archelon, Arthena, Arthur Rubin, Arved, Aude, AugPi, AxelBoldt, BeteNoir, Bidabadi, BigJohnHenry, Blindsuperhero, Bob.v.R, Can't sleep, clown will eat me, Charles Matthews, Charleswallingford, Conversion script, Cybercobra, Daran, Darij, Deineka, Dmn, DonSiano, Doradus, Drilnoth, Dysprosia, EmilJ, Evil saltine, Fredrik, Furrykef, Gaius Cornelius, Gene Ward Smith, Geometry guy, Giftlite, Graham87, Greg Kuperberg, Hede2000, Helix84, Henning Makholm, Hesam7, Holger Blasum, Huddlebum, Icairns, JCSantos, Jacobolus, JdH, Jimbreed, Jimbryho, Jimp, Jna runn, Kartik J, Lakinekaki, Lambiam, Li-sung, LkNsngth, Lunchscale, Lupin, LutzL, MathMartin, MathsIsFun, Mav, Meni Rosenfeld, Michael Hardy, Michael Larsen, Michael Slone, Mike Segal, Monamip, Mpatel, Nic bor, Nsh, Nuno Tavares, Obradovic Goran, Ortonmc, Oxy86, PMajer, Paul D. Anderson, Paul Taylor, Philologer, Primalbeing, Pt, Qmwne235, Randomblue, Rgdboer, Rholton, Rich Farmbrough, Rjwilmsi, Robinh, Romanm, Shishir0610, Skomorokh, Smcinerney, Smimram, Snoyes, SoroSuub1, Syp, Tobias Bergemann, Toby Bartels, Toh, Trovatore, Tulcod, Unyoyega, Vladkornea, WikiUserPedia, Wmahan, Woohookitty, Wshun, XJamRastafire, Xantharius, Zfr, Zundark, Zvika, 109 anonymous edits Fundamental theorem of cyclic groups Source: http://en.wikipedia.org/w/index.php?oldid=455618448 Contributors: Arthur Rubin, Cybercobra, Geometry guy, Giftlite, Icairns, Joeldl, Lhf, Michael Hardy, Selfworm, Sigmundur, Tobias Bergemann, Zvika, 20 anonymous edits Fundamental theorem of Galois theory Source: http://en.wikipedia.org/w/index.php?oldid=455618329 Contributors: Bender235, Bender2k14, BeteNoir, Charles Matthews, Cwkmail, Cybercobra, Dfeuer, Dmharvey, Dysprosia, Frau Holle, Geometry guy, Giftlite, Hesam7, HorsePunchKid, Icairns, Jibbb, Jim.belk, MarkC77, MathMartin, Sandrobt, 15 anonymous edits Fundamental theorem of linear algebra Source: http://en.wikipedia.org/w/index.php?oldid=455618237 Contributors: BeteNoir, Charles Matthews, Cronholm144, Cybercobra, Flavio Guitian, Geometry guy, Giftlite, Harryboyles, Icairns, Keenan Pepper, Lowellian, Nbarth, Oleg Alexandrov, Qwfp, R'n'B, Silly rabbit, 17 anonymous edits Fundamental theorem on homomorphisms Source: http://en.wikipedia.org/w/index.php?oldid=455618150 Contributors: Andre Engels, Arthena, AugPi, AxelBoldt, BeteNoir, Charles Matthews, Chas zzz brown, Conversion script, Cybercobra, DefLog, Erud, Gelingvistoj, Geometry guy, Giftlite, Goochelaar, Graham87, Grubber, Icairns, Jay Gatsby, Linas, Magidin,

224

Article Sources and Contributors


MathMartin, Waltpohl, Weialawaga, 9 anonymous edits GilmanGriess theorem Source: http://en.wikipedia.org/w/index.php?oldid=455766095 Contributors: Giftlite, Headbomb, Michael Hardy, R.e.b., Rjwilmsi, Sodin, 1 anonymous edits Going up and going down Source: http://en.wikipedia.org/w/index.php?oldid=455619435 Contributors: Charles Matthews, David Shay, Discospinster, Giftlite, Jakob.scholbach, Jowa fan, Kiefer.Wolfowitz, Michael Slone, Paul August, Revolver, Rschwieb, 11 anonymous edits Goldie's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455618039 Contributors: Charles Matthews, Geometry guy, Giftlite, JackSchmidt, Jowa fan, Rgdboer, Rschwieb, Tobias Bergemann, Vanish2 GolodShafarevich theorem Source: http://en.wikipedia.org/w/index.php?oldid=455617950 Contributors: Academic Challenger, Bender235, Charles Matthews, Ebony Jackson, Geometry guy, Giftlite, JackSchmidt, Jackbarron, Mathsci, Michael Hardy, Michael Slone, RobHar, Turgidson, Yonatbe5, 1 anonymous edits GorensteinHarada theorem Source: http://en.wikipedia.org/w/index.php?oldid=455617855 Contributors: Geometry guy, R.e.b. Gromov's theorem on groups of polynomial growth Source: http://en.wikipedia.org/w/index.php?oldid=455619473 Contributors: BeteNoir, CSTAR, Charles Matthews, Efjb2, Geometry guy, Giftlite, Headbomb, Icairns, Jevansen, JoshuaZ, LarRan, Michael Hardy, Mosher, OdedSchramm, Teorth, The Anome, Tkuvho, Tosha, Zundark, 4 anonymous edits Grushko theorem Source: http://en.wikipedia.org/w/index.php?oldid=455766207 Contributors: Colonies Chris, Daniel5Ko, David Eppstein, Giftlite, HUnTeR4subs, JackSchmidt, Katzmik, Michael Hardy, Nsk92, RonnieBrown, Sodin, 2 anonymous edits Haboush's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455765461 Contributors: BeteNoir, Charles Matthews, David Eppstein, DavidCBryant, Giftlite, Jeff3000, R'n'B, R.e.b., Ringspectrum, Rjwilmsi, RobHar, Sodin, 4 anonymous edits Hahn embedding theorem Source: http://en.wikipedia.org/w/index.php?oldid=455619581 Contributors: BeteNoir, Charles Matthews, Diskz, Gene Ward Smith, Geometry guy, Giftlite, Iohannes Animosus, JackSchmidt, Marcus Pivato, 1 anonymous edits Hajs's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455765482 Contributors: Cronholm144, Giftlite, Kope, PigFlu Oink, R.e.b., Sodin, Tholly, Zachanter Harish-Chandra isomorphism Source: http://en.wikipedia.org/w/index.php?oldid=455629394 Contributors: Arcfrk, Charles Matthews, Franp9am, Giftlite, Headbomb, R.e.b., Sodin, Ulner, 6 anonymous edits Hasse norm theorem Source: http://en.wikipedia.org/w/index.php?oldid=455629296 Contributors: Arcfrk, BeteNoir, Charles Matthews, Dugwiki, Gene Ward Smith, Sodin, Vanish2, 1 anonymous edits HasseArf theorem Source: http://en.wikipedia.org/w/index.php?oldid=455765535 Contributors: Chenxlee, Giftlite, RobHar, Sodin, TakuyaMurata, Woohookitty, 2 anonymous edits Hilbert's basis theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609894 Contributors: Ad4m, Arcfrk, AxelBoldt, BeteNoir, Brockert, Bryan Derksen, Charles Matthews, Conversion script, Drusus 0, Gaius Cornelius, Giftlite, Guardian of Light, Hillman, ICPalm, Jxr, LilHelpa, MathMartin, Michael Slone, Oleg Alexandrov, R.e.b., Randomblue, Sodin, Tobias Bergemann, Vivacissamamente, Waltpohl, 17 anonymous edits Hilbert's irreducibility theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609923 Contributors: Aleph4, BUF4Life, Barylior, BeteNoir, Brookie, Charles Matthews, Giftlite, Mathtyke, Michael Hardy, Pearle, PoolGuy, Ringspectrum, Sodin, 3 anonymous edits Hilbert's Nullstellensatz Source: http://en.wikipedia.org/w/index.php?oldid=455690802 Contributors: Arcfrk, Atgnclk, AxelBoldt, BeteNoir, Bomazi, Charles Matthews, Chas zzz brown, Crisfilax, D.Lazard, D6, EmilJ, Fluxions, Fropuff, Giftlite, Jmath666, Michael Hardy, Mlm42, Nbarth, Ntsimp, Oleg Alexandrov, R.e.b., Ringspectrum, Rschwieb, Sannse, TomyDuby, Trevorgoodchild, Vanish2, Waltpohl, 21 anonymous edits Hilbert's syzygy theorem Source: http://en.wikipedia.org/w/index.php?oldid=455609953 Contributors: Arcfrk, BeteNoir, Charles Matthews, Giftlite, Grendelkhan, Myrizio, Nick Number, Sodin, Ylloh, Zaphod Beeblebrox, 12 anonymous edits Hilbert's Theorem 90 Source: http://en.wikipedia.org/w/index.php?oldid=455610006 Contributors: Bender235, BeteNoir, CRGreathouse, Charles Matthews, EmilJ, Four Dog Night, Gaius Cornelius, Gene Ward Smith, Giftlite, MathMartin, Myasuda, Pmanderson, R.e.b., Rich Farmbrough, Ringspectrum, RobHar, Set theorist, Sodin, Zundark, 13 anonymous edits HopkinsLevitzki theorem Source: http://en.wikipedia.org/w/index.php?oldid=455618651 Contributors: Anne Bauval, Bearcat, Giftlite, Michael Hardy, R.e.b., Rschwieb Hurwitz's theorem (normed division algebras) Source: http://en.wikipedia.org/w/index.php?oldid=455687589 Contributors: 3mta3, Algebraist, Anne Bauval, Arrataz, Brews ohare, Geometry guy, Giftlite, Headbomb, Ilmari Karonen, JCSantos, Mattbuck, RobHar, 15 anonymous edits Isomorphism extension theorem Source: http://en.wikipedia.org/w/index.php?oldid=455687893 Contributors: AtticusRyan, Cronholm144, Geometry guy, Giftlite, ImPerfectHacker, Michael Hardy, Vanish2, Zundark, 3 anonymous edits Isomorphism theorem Source: http://en.wikipedia.org/w/index.php?oldid=455688042 Contributors: AdamSmithee, Algebraist, Amdk8800, Avani, AxelBoldt, BeteNoir, Brighterorange, Bruguiea, Bryan Derksen, Csigabi, Cwkmail, Dysprosia, EmilyPeters, Flamingspinach, Frodo, Frozsyn, Geometry guy, Giftlite, Grubber, Helder.wiki, Isnow, JackSchmidt, Jcobb, Jeepday, JensMueller, Jim.belk, Karl-Henner, Konradek, Krasnoludek, Lethe, Magidin, MathMartin, Michael Hardy, Michael K. Edwards, Miyagawa, Mrajpkc, Nbarth, Rjgodoy, SJP, Sabbut, SetaLyas, Shiyang, Silly rabbit, Tarquin, Toby, Waltpohl, Weialawaga, Zundark, 41 anonymous edits Jacobson density theorem Source: http://en.wikipedia.org/w/index.php?oldid=455697398 Contributors: Bender235, Charles Matthews, David Eppstein, Giftlite, JamieVicary, Mct mht, Mike Peel, Philosopher, Point-set topologist, R.e.b., Rjwilmsi, Rschwieb, Tobias Bergemann, Vanish2, Waltpohl Jordan's theorem (symmetric group) Source: http://en.wikipedia.org/w/index.php?oldid=455688643 Contributors: Franp9am, Geometry guy JordanSchur theorem Source: http://en.wikipedia.org/w/index.php?oldid=455688735 Contributors: David Eppstein, Geometry guy, Giftlite, JoshuaZ, Matt me, Michael Hardy, 4 anonymous edits Krull's principal ideal theorem Source: http://en.wikipedia.org/w/index.php?oldid=455690389 Contributors: Anne Bauval, BeteNoir, Charles Matthews, Geometry guy, Giftlite, Jakob.scholbach, Kummini, Michael Hardy, Oleg Alexandrov, Silverfish, Vanish2, Vivacissamamente, 3 anonymous edits KrullSchmidt theorem Source: http://en.wikipedia.org/w/index.php?oldid=455695593 Contributors: Aholtman, Algebraist, Arcfrk, Arthur Rubin, Bogdangiusca, Charles Matthews, Dreadstar, Foobarnix, Fropuff, Gauge, Giftlite, Helder.wiki, JackSchmidt, Kevin Lamoreau, Masnevets, Matthew Fennell, Michael Hardy, Oleg Alexandrov, Omnipaedista, Rschwieb, The Anome, Tobias Bergemann, Zundark, 19 anonymous edits Knneth theorem Source: http://en.wikipedia.org/w/index.php?oldid=455744254 Contributors: Ambrose H. Field, Auntof6, Bender235, Charles Matthews, Cheesus, DVD R W, Dbenbenn, Giftlite, Gofors, KSmrq, LokiClock, Marhahs, Momotaro, Old Man Grumpus, Ozob, R.e.b., Rschwieb, Ryan Reich, Silly rabbit, 16 anonymous edits Kurosh subgroup theorem Source: http://en.wikipedia.org/w/index.php?oldid=455630389 Contributors: Dtrebbien, Giftlite, HUnTeR4subs, JackSchmidt, LilHelpa, Michael Hardy, Nsk92, Sodin, 3 anonymous edits Lagrange's theorem (group theory) Source: http://en.wikipedia.org/w/index.php?oldid=455610063 Contributors: Aboalbiss, AdamSmithee, Avik21, AxelBoldt, BeteNoir, Bryan Derksen, Calle, Charles Matthews, Chas zzz brown, Chromosome, Conversion script, Courcelles, Creidieki, Cwkmail, Dcoetzee, Dysprosia, Eric119, Giftlite, Goochelaar, Graham87, GregorB, Grubber, Hyju, Ixfd64, JCSantos, JackSchmidt, Joth, Kilva, Leycec, Lhf, Lowellian, Lupin, Mathsci, Miaow Miaow, Michael Hardy, Obradovic Goran, Plasticup, Quotient group, Reaper Eternal, Rghthndsd, Salvatore Ingala, Sodin, Superninja, Tarquin, Timwi, Tsemii, Xantharius, Youandme, Yuval Madar, 40 anonymous edits LaskerNoether theorem Source: http://en.wikipedia.org/w/index.php?oldid=455696265 Contributors: Arthur Rubin, Bender2k14, BeteNoir, Charles Matthews, David Eppstein, Expz, Gene Nygaard, Giftlite, Gwaihir, Ioannes Pragensis, Jakob.scholbach, KWRegan, Kummini, MathMartin, Michael Hardy, Mild Bill Hiccup, Nbarth, Oleg Alexandrov, Paisa, Psychonaut, R.e.b., Rich Farmbrough, Rschwieb, Silverfish, TakuyaMurata, WATARU, Waltpohl, XPEHOPE3, 14 anonymous edits

225

Article Sources and Contributors


Latimer-MacDuffee theorem Source: http://en.wikipedia.org/w/index.php?oldid=455884452 Contributors: Charles Matthews, Konstable, Oleg Alexandrov, Sodin, 2 anonymous edits Lattice theorem Source: http://en.wikipedia.org/w/index.php?oldid=455717612 Contributors: Algebraist, BeteNoir, Bruguiea, Cacadril, Caesura, Cwkmail, David Eppstein, DemonThing, E946, Error792, Frdrick Lacasse, Giftlite, GregorB, Hans Adler, JAK, MathMartin, MathMast, Nbarth, Patrick, RDBury, Silly rabbit, Thehotelambush, Tobias Bergemann, Vanish2, 4 anonymous edits Levitzky's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455630202 Contributors: Calaka, Giftlite, Headbomb, JackSchmidt, Kope, Michael Hardy, Point-set topologist, Rich Farmbrough, Rjwilmsi, Sodin, Zundark Lie's third theorem Source: http://en.wikipedia.org/w/index.php?oldid=455690122 Contributors: Charles Matthews, FrozenPurpleCube, Geometry guy, Giftlite, Jason Quinn LieKolchin theorem Source: http://en.wikipedia.org/w/index.php?oldid=455717004 Contributors: BeteNoir, Ceyockey, Charles Matthews, CrackerJack7891, David Eppstein, FactSpewer, Gauge, Gene Nygaard, Giftlite, Headbomb, Hillman, JackSchmidt, Linas, Michael Hardy, Natalya, Nbarth, Nowhither, Psychonaut, R.e.b., RDBury, West Brom 4ever, Zundark, 4 anonymous edits Maschke's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455618805 Contributors: Arcfrk, Aviados, BeteNoir, Calaka, CapitalR, Charles Matthews, Cwkmail, Dcoetzee, Fropuff, Giftlite, Gwaihir, Headbomb, Hesam7, Hillman, Jay Gatsby, Jtwdog, Michael Hardy, Natalya, PappyK, RobHar, Rschwieb, Solar-Poseidon, Tobias Bergemann, TommasoT, Undercrowdtroll, Xiaodai, Zundark, 18 anonymous edits Milnor conjecture Source: http://en.wikipedia.org/w/index.php?oldid=455716070 Contributors: C S, Charles Matthews, Exceptg, Gauge, Headbomb, Michael Hardy, Phantomsteve, R.e.b., RDBury, Semorrison, Singularity, 5 anonymous edits MordellWeil theorem Source: http://en.wikipedia.org/w/index.php?oldid=455630342 Contributors: BeteNoir, Charles Matthews, Giftlite, Hesam7, Jcobb, Patrick, Psychonaut, R.e.b., Reedy, Silenteuphony, Sodin, Thecheesykid, Vanish2, Zoicon5, Zundark, 3 anonymous edits Multinomial theorem Source: http://en.wikipedia.org/w/index.php?oldid=455690672 Contributors: AdjustShift, Altenmann, ArglebargleIV, Charles Matthews, Chenxlee, Chessimprov, Dysprosia, Endlessoblivion, Forge021, Geometry guy, Giftlite, Icairns, Jengelh, Kakila, LOL, Labus, Linas, McKay, MikeRumex, NeoUrfahraner, Pberndt, Quantling, Rar, Rich Farmbrough, RokerHRO, SPUI, Spoon!, Stephenb, WiiStation360, Wile E. Heresiarch, Yaleeconometrics, Zero0000, 70 , , anonymous edits NielsenSchreier theorem Source: http://en.wikipedia.org/w/index.php?oldid=455630131 Contributors: David Eppstein, Giftlite, HUnTeR4subs, Headbomb, JackSchmidt, Sodin, Zundark, 1 anonymous edits PerronFrobenius theorem Source: http://en.wikipedia.org/w/index.php?oldid=455759198 Contributors: Alexander Chervov, Arcfrk, B.wilson, BenFrantzDale, Bender235, BeteNoir, Bill luv2fly, Charles Matthews, Comfortably Paranoid, Cvdwoest, David Eppstein, Dcclark, Dima373, Doctorilluminatus, Flyingspuds, G.perarnau, Gdm, Giftlite, Justin Mauger, Kiefer.Wolfowitz, Kirbin, Linas, MRFS, Michael Hardy, Nbarth, Pavel Stanley, Psychonaut, R.e.b., Rschwieb, Shining Celebi, Sodin, Stigin, Tcnuk, TedPavlic, Urhixidur, Vinsz, 59 anonymous edits PoincarBirkhoffWitt theorem Source: http://en.wikipedia.org/w/index.php?oldid=455690945 Contributors: Arcfrk, AxelBoldt, BeteNoir, Burivykh, CSTAR, DR2006kl, Dan Gardner, Darij, Geometry guy, Giftlite, Hans Lundmark, Henning Makholm, Michael Hardy, Myasuda, Oleg Alexandrov, Psychonaut, R.e.b., Vanished user, 13 anonymous edits Polynomial remainder theorem Source: http://en.wikipedia.org/w/index.php?oldid=455691069 Contributors: Aakaalaar93, BeteNoir, Charles Matthews, Corwin., Fangz, Geometry guy, Gesslein, Giftlite, Jusdafax, KSmrq, Kenny TM~, Lambiam, MSGJ, Maxal, Michael Hardy, Mifter, Nonagonal Spider, Oli Filth, Pizza1512, Rommels, Samw, Silverfish, XMxWx, YUL89YYZ, 40 anonymous edits Primitive element theorem Source: http://en.wikipedia.org/w/index.php?oldid=455691200 Contributors: Algebraist, Austinmohr, Billymac00, Bo Jacoby, Charles Matthews, Dfeldmann, Gene Ward Smith, Geometry guy, Giftlite, IhorLviv, MathMartin, Michael Hardy, Oyd11, Point-set topologist, Sandrobt, Simetrical, Vanish2, William Avery, Zundark, 9 anonymous edits QuillenSuslin theorem Source: http://en.wikipedia.org/w/index.php?oldid=455696014 Contributors: BeteNoir, Charles Matthews, Dan Gardner, Dtrebbien, Giftlite, Itai, Jcobb, Joeldl, Michael Hardy, Myasuda, Pmanderson, Psychonaut, RobHar, Rschwieb, Ryan Reich, Silverfish, Singularity, 18 anonymous edits Rational root theorem Source: http://en.wikipedia.org/w/index.php?oldid=455691538 Contributors: 0, 478jjjz, 64.24.17.xxx, A. Pichler, ABF, Aisaac, Aleph4, Alexb@cut-the-knot.com, Aliotra, Anonymous Dissident, Apeiron, Asmeurer, AxelBoldt, BeteNoir, Blitz9, Charles Matthews, Count Iblis, Culix, Discospinster, Gak, Geometry guy, Giftlite, Henrygb, Jujutacular, Kmhkmh, Lambiam, Lindmere, Marc Venot, Marc van Leeuwen, Mets501, Michael Hardy, N5iln, Nousernamesleft, Postglock, Reyk, RoseParks, Salgueiro, Sam Hocevar, Sikory, Taw, The Cunctator, Tobias Hoevekamp, Zooplankton1972, Zundark, 42 anonymous edits Regev's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455610675 Contributors: Giftlite, Gregbard, Headbomb, Michael Hardy, R.e.b., Sodin Schreier refinement theorem Source: http://en.wikipedia.org/w/index.php?oldid=455691924 Contributors: BeteNoir, CBM, Charles Matthews, David Eppstein, Gauge, Geometry guy, ImPerfectHacker, Jaakko Seppl, Michael Slone, Spartanfox86, Tobias Bergemann, 1 anonymous edits SchurZassenhaus theorem Source: http://en.wikipedia.org/w/index.php?oldid=455697217 Contributors: Algebraist, Joe Decker, Kidburla, Marvoir, Mm06ahlf, Rich Farmbrough, Rschwieb, Wafulz, 2 anonymous edits SerreSwan theorem Source: http://en.wikipedia.org/w/index.php?oldid=455697339 Contributors: AxelBoldt, Bender235, Charles Matthews, Gene Nygaard, Giftlite, Hillman, JackSchmidt, JoergenB, John Baez, Matterink, Mct mht, Phys, Point-set topologist, Rausch, Rschwieb, Silly rabbit, Tesseran, Tosha, Waltpohl, 6 anonymous edits SkolemNoether theorem Source: http://en.wikipedia.org/w/index.php?oldid=455695953 Contributors: BeteNoir, Charles Matthews, Gaius Cornelius, Giftlite, JackSchmidt, MathMartin, Psychonaut, Rschwieb, The Rambling Man, Thehotelambush, 10 anonymous edits Specht's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455865798 Contributors: Bender235, Charles Matthews, Giftlite, Jitse Niesen, Michael Hardy, Sodin, 1 anonymous edits Stone's representation theorem for Boolean algebras Source: http://en.wikipedia.org/w/index.php?oldid=455744437 Contributors: Aleph0, Beroal, BeteNoir, Blotwell, CBM, Chalst, Chinju, David Eppstein, Falcor84, Fropuff, Giftlite, JanCK, Kuratowski's Ghost, Linas, Markus Krtzsch, Mhss, Michael Hardy, Naddy, Pjacobi, Porton, R'n'B, R.e.b., Rschwieb, Sharpcomputing, Smack, StevenJohnston, Tkuvho, Trovatore, Tsirel, Vivacissamamente, Zundark, 15 anonymous edits Structure theorem for finitely generated modules over a principal ideal domain Source: http://en.wikipedia.org/w/index.php?oldid=455693186 Contributors: Alecobbe, Algebraist, Altrevolte, Anne Bauval, ArnoldReinhold, Arthur Rubin, Auntof6, Bomazi, EmilJ, Expz, Foobarnix, Frap, Geometry guy, Giftlite, Grafen, Henning Makholm, JackSchmidt, Kundor, Marc van Leeuwen, Michael Hardy, Mulanhua, Nbarth, Oleg Alexandrov, Ozob, ReyBrujo, Rich Farmbrough, Rschwieb, Silly rabbit, Tobias Bergemann, Vincent Semeria, Zelmerszoetrop, Zundark, 17 anonymous edits Subgroup test Source: http://en.wikipedia.org/w/index.php?oldid=455693079 Contributors: Crazyjimbo, Geometry guy, JumpDiscont, Pt, Selfworm, Zvika, 4 anonymous edits Subring test Source: http://en.wikipedia.org/w/index.php?oldid=455693014 Contributors: Charles Matthews, Geometry guy, Joeldl, Oleg Alexandrov, Selfworm, Zvika, 1 anonymous edits Sylow theorems Source: http://en.wikipedia.org/w/index.php?oldid=455692905 Contributors: 01001, Aholtman, Alecobbe, Amitushtush, Ams80, Ank0ku, AxelBoldt, BenF, BeteNoir, CZeke, Charles Matthews, Chas zzz brown, Chochopk, Conversion script, Crisfilax, Cwkmail, David Eppstein, Derek Ross, Dominus, Druiffic, EmilJ, Eramesan, Functor salad, GTBacchus, Gauge, Geometry guy, Giftlite, Goochelaar, Graham87, Grubber, Haham hanuka, Hank hu, Headbomb, Hesam7, JackSchmidt, Japanese Searobin, Joelsims80, Jonathanzung, Kilva, Lzur, MathMartin, Mav, Michael Hardy, Nbarth, Ossido, PappyK, PierreAbbat, Pladdin, Pmanderson, Point-set topologist, Pyrop, R.e.b., Reedy, Schutz, Siroxo, Sl, Spoon!, Stove Wolf, Superninja, TakuyaMurata, Tarquin, Tobias Bergemann, Twilsonb, WLior, Welsh, Zundark, Zvika, 76 anonymous edits Sylvester's determinant theorem Source: http://en.wikipedia.org/w/index.php?oldid=455692773 Contributors: Bosmon, Exol, Gentsquash, Geometry guy, Giftlite, Guillaume.bouchard, Jitse Niesen, Michael Hardy, Michael Slone, Rich Farmbrough, Robinh, 9 anonymous edits Sylvester's law of inertia Source: http://en.wikipedia.org/w/index.php?oldid=455692673 Contributors: A. Pichler, Akriasas, Anne Bauval, Arcfrk, Bh3u4m, Can't sleep, clown will eat me, Charles Matthews, Choster, DanielJanzon, Geometry guy, Giftlite, Gryllida, Jorge Stolfi, JuJube, Michael Hardy, Nneonneo, Plclark, Randomblue, Ranicki, Seanwal111111, Shambolic Entity, Simplifix, The wub, Vanish2, 23 anonymous edits

226

Article Sources and Contributors


Takagi existence theorem Source: http://en.wikipedia.org/w/index.php?oldid=455768890 Contributors: Ambrose H. Field, BeteNoir, Biblbroks, Charles Matthews, Cobaltcigs, Dugwiki, Edward, Gene Ward Smith, Giftlite, R'n'B, R.e.b., Ringspectrum, Sodin, WhatamIdoing, 9 anonymous edits Three subgroups lemma Source: http://en.wikipedia.org/w/index.php?oldid=455692354 Contributors: Algebraist, Geometry guy, Giftlite, Michael Hardy, Omnipaedista, Point-set topologist Trichotomy theorem Source: http://en.wikipedia.org/w/index.php?oldid=455693272 Contributors: Charles Matthews, Geometry guy, Gregbard, Headbomb, Michael Hardy, R.e.b., Rjwilmsi Walter theorem Source: http://en.wikipedia.org/w/index.php?oldid=455765520 Contributors: Captain-tucker, Gregbard, Headbomb, R.e.b., Rjwilmsi, Rschwieb Wedderburn's little theorem Source: http://en.wikipedia.org/w/index.php?oldid=455693510 Contributors: CBM, Charles Matthews, Geometry guy, Giftlite, Howard McCay, Kammerer55, MarkSweep, Matikkapoika, Nbarth, Pcarpent1, Qwfp, Rgdboer, Ringspectrum, SantoBugito, TakuyaMurata, Thehotelambush, Tobias Bergemann, 5 anonymous edits Weil conjecture on Tamagawa numbers Source: http://en.wikipedia.org/w/index.php?oldid=455693645 Contributors: Bender235, BeteNoir, Charles Matthews, Geometry guy, Giftlite, LokiClock, R.e.b., Rjwilmsi, TakuyaMurata, 1 anonymous edits Witt's theorem Source: http://en.wikipedia.org/w/index.php?oldid=455693983 Contributors: Arcfrk, Blotwell, Charles Matthews, Geometry guy, Jdthomas, Michael Hardy, Nbarth, Oleg Alexandrov, S11-73-3-33, Tyrrell McAllister, Ulner, 4 anonymous edits Z* theorem Source: http://en.wikipedia.org/w/index.php?oldid=455765703 Contributors: Giftlite, Headbomb, JackSchmidt, Jitse Niesen, Messagetolove, R.e.b., Rschwieb, Woohookitty Zassenhaus lemma Source: http://en.wikipedia.org/w/index.php?oldid=455718572 Contributors: Ahills60, Charles Matthews, DRLB, David Eppstein, FF2010, Giftlite, Helder.wiki, Jason Recliner, Esq., Jeepday, Julien Tuerlinckx, Mat cross, MathMartin, Michael Hardy, Nbarth, RDBury, Schneelocke, Silly rabbit, Silverfish, Waltpohl, 5 anonymous edits ZJ theorem Source: http://en.wikipedia.org/w/index.php?oldid=455765215 Contributors: Giftlite, JackSchmidt, Jitse Niesen, R.e.b., Rschwieb, Turgidson

227

Image Sources, Licenses and Contributors

228

Image Sources, Licenses and Contributors


Image:Pascal's triangle 5.svg Source: http://en.wikipedia.org/w/index.php?title=File:Pascal's_triangle_5.svg License: GNU Free Documentation License Contributors: User:Conrad.Irwin originally User:Drini Image:Pascal triangle small.png Source: http://en.wikipedia.org/w/index.php?title=File:Pascal_triangle_small.png License: GNU Free Documentation License Contributors: user:gunther Image:BinomialTheorem.png Source: http://en.wikipedia.org/w/index.php?title=File:BinomialTheorem.png License: Public Domain Contributors: Danilo Guanabara Fernandes Image:Birkhoff120.svg Source: http://en.wikipedia.org/w/index.php?title=File:Birkhoff120.svg License: Public Domain Contributors: David Eppstein File:Cramer.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Cramer.jpg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Author:Franklin Vera Pacheco Image:Crystallographic restriction polygons.png Source: http://en.wikipedia.org/w/index.php?title=File:Crystallographic_restriction_polygons.png License: Creative Commons Attribution-Sharealike 2.5 Contributors: File:OEISicon light.svg Source: http://en.wikipedia.org/w/index.php?title=File:OEISicon_light.svg License: Public Domain Contributors: Lipedia Image:Discriminant49CubicFieldFundamentalDomainOfUnits.png Source: http://en.wikipedia.org/w/index.php?title=File:Discriminant49CubicFieldFundamentalDomainOfUnits.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: RobHar File:Lattice diagram of Q adjoin the positive square roots of 2 and 3, its subfields, and Galois groups.svg Source: http://en.wikipedia.org/w/index.php?title=File:Lattice_diagram_of_Q_adjoin_the_positive_square_roots_of_2_and_3,_its_subfields,_and_Galois_groups.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Self File:Lattice_diagram_of_Q_adjoin_a_cube_root_of_2_and_a_primitive_cube_root_of_1,_its_subfields,_and_Galois_groups.svg Source: http://en.wikipedia.org/w/index.php?title=File:Lattice_diagram_of_Q_adjoin_a_cube_root_of_2_and_a_primitive_cube_root_of_1,_its_subfields,_and_Galois_groups.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Self Image:The four subspaces.svg Source: http://en.wikipedia.org/w/index.php?title=File:The_four_subspaces.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Cronholm144 image:FundHomDiag.png Source: http://en.wikipedia.org/w/index.php?title=File:FundHomDiag.png License: GNU Free Documentation License Contributors: File:Shifted square tiling.svg Source: http://en.wikipedia.org/w/index.php?title=File:Shifted_square_tiling.svg License: Creative Commons Zero Contributors: User:David Eppstein Image:First-isomorphism-theorem.svg Source: http://en.wikipedia.org/w/index.php?title=File:First-isomorphism-theorem.svg License: Public Domain Contributors: Michael K. Edwards File:Labeled Triangle Reflections.svg Source: http://en.wikipedia.org/w/index.php?title=File:Labeled_Triangle_Reflections.svg License: Public Domain Contributors: Jim.belk File:Hexagon Reflections.png Source: http://en.wikipedia.org/w/index.php?title=File:Hexagon_Reflections.png License: Public Domain Contributors: Image:Butterfly lemma.svg Source: http://en.wikipedia.org/w/index.php?title=File:Butterfly_lemma.svg License: Creative Commons Attribution 3.0 Contributors: Claudio Rocchini

License

229

License
Creative Commons Attribution-Share Alike 3.0 Unported //creativecommons.org/licenses/by-sa/3.0/

You might also like