HOME

TheInfoList



OR:

In
mathematics Mathematics is an area of knowledge that includes the topics of numbers, formulas and related structures, shapes and the spaces in which they are contained, and quantities and their changes. These topics are represented in modern mathematics ...
, more specifically
abstract algebra In mathematics, more specifically algebra, abstract algebra or modern algebra is the study of algebraic structures. Algebraic structures include groups, rings, fields, modules, vector spaces, lattices, and algebras over a field. The ter ...
and
commutative algebra Commutative algebra, first known as ideal theory, is the branch of algebra that studies commutative rings, their ideals, and modules over such rings. Both algebraic geometry and algebraic number theory build on commutative algebra. Prom ...
, Nakayama's lemma — also known as the Krull–Azumaya theorem — governs the interaction between the Jacobson radical of a ring (typically a
commutative ring In mathematics, a commutative ring is a ring in which the multiplication operation is commutative. The study of commutative rings is called commutative algebra. Complementarily, noncommutative algebra is the study of ring properties that are not ...
) and its finitely generated
modules Broadly speaking, modularity is the degree to which a system's components may be separated and recombined, often with the benefit of flexibility and variety in use. The concept of modularity is used primarily to reduce complexity by breaking a s ...
. Informally, the
lemma Lemma may refer to: Language and linguistics * Lemma (morphology), the canonical, dictionary or citation form of a word * Lemma (psycholinguistics), a mental abstraction of a word about to be uttered Science and mathematics * Lemma (botany), ...
immediately gives a precise sense in which finitely generated modules over a commutative ring behave like
vector space In mathematics and physics, a vector space (also called a linear space) is a set whose elements, often called '' vectors'', may be added together and multiplied ("scaled") by numbers called ''scalars''. Scalars are often real numbers, but can ...
s over a field. It is an important tool in
algebraic geometry Algebraic geometry is a branch of mathematics, classically studying zeros of multivariate polynomials. Modern algebraic geometry is based on the use of abstract algebraic techniques, mainly from commutative algebra, for solving geometrical ...
, because it allows local data on algebraic varieties, in the form of modules over
local ring In abstract algebra, more specifically ring theory, local rings are certain rings that are comparatively simple, and serve to describe what is called "local behaviour", in the sense of functions defined on varieties or manifolds, or of algebrai ...
s, to be studied pointwise as vector spaces over the residue field of the ring. The lemma is named after the Japanese mathematician Tadashi Nakayama and introduced in its present form in , although it was first discovered in the special case of ideals in a commutative ring by Wolfgang Krull and then in general by
Goro Azumaya was a Japanese mathematician who introduced the notion of Azumaya algebra in 1951. His advisor was Shokichi Iyanaga. At the time of his death he was an emeritus professor at Indiana University Indiana University (IU) is a system of public ...
( 1951). In the commutative case, the lemma is a simple consequence of a generalized form of the Cayley–Hamilton theorem, an observation made by Michael Atiyah (
1969 This year is notable for Apollo 11's first landing on the moon. Events January * January 4 – The Government of Spain hands over Ifni to Morocco. * January 5 **Ariana Afghan Airlines Flight 701 crashes into a house on its approach to ...
). The special case of the
noncommutative In mathematics, a binary operation is commutative if changing the order of the operands does not change the result. It is a fundamental property of many binary operations, and many mathematical proofs depend on it. Most familiar as the name o ...
version of the lemma for right ideals appears in Nathan Jacobson ( 1945), and so the noncommutative Nakayama lemma is sometimes known as the Jacobson–Azumaya theorem. The latter has various applications in the theory of Jacobson radicals.


Statement

Let R be a
commutative ring In mathematics, a commutative ring is a ring in which the multiplication operation is commutative. The study of commutative rings is called commutative algebra. Complementarily, noncommutative algebra is the study of ring properties that are not ...
with identity 1. The following is Nakayama's lemma, as stated in : Statement 1: Let I be an
ideal Ideal may refer to: Philosophy * Ideal (ethics), values that one actively pursues as goals * Platonic ideal, a philosophical idea of trueness of form, associated with Plato Mathematics * Ideal (ring theory), special subsets of a ring considered ...
in R, and M a finitely-generated module over R. If IM=M, then there exists an r \in R with r \equiv 1\; (\operatorname I), such that rM = 0. This is proven
below Below may refer to: *Earth * Ground (disambiguation) *Soil *Floor * Bottom (disambiguation) *Less than *Temperatures below freezing *Hell or underworld People with the surname *Ernst von Below (1863–1955), German World War I general *Fred Below ...
. The following corollary is also known as Nakayama's lemma, and it is in this form that it most often appears. Statement 2: If ''M'' is a finitely-generated module over ''R'', J(R) is the Jacobson radical of R, and J(R)M=M, then M = 0. :''Proof'': r - 1 (with r as above) is in the Jacobson radical so r is invertible. More generally, one has that J(R)M is a superfluous submodule of M when M is finitely-generated. Statement 3: If M is a finitely-generated module over ''R'', ''N'' is a submodule of M, and ''M'' = ''N'' + J(''R'')''M'', then ''M'' = ''N''. :''Proof'': Apply Statement 2 to ''M''/''N''. The following result manifests Nakayama's lemma in terms of generators. Statement 4: If ''M'' is a finitely-generated module over ''R'' and the images of elements ''m''1,...,''m''''n'' of ''M'' in ''M'' / ''J''(''R'')''M'' generate ''M'' / ''J''(''R'')''M'' as an ''R''-module, then ''m''1,...,''m''''n'' also generate ''M'' as an ''R''-module. :''Proof'': Apply Statement 3 to ''N'' = Σ''i''''Rm''''i''. If one assumes instead that ''R'' is
complete Complete may refer to: Logic * Completeness (logic) * Completeness of a theory, the property of a theory that every formula in the theory's language or its negation is provable Mathematics * The completeness of the real numbers, which implies t ...
and ''M'' is separated with respect to the ''I''-adic topology for an ideal ''I'' in ''R'', this last statement holds with ''I'' in place of ''J''(''R'') and without assuming in advance that ''M'' is finitely generated. Here separatedness means that the ''I''-adic topology satisfies the ''T''1 separation axiom, and is equivalent to \textstyle


Consequences


Local rings

In the special case of a finitely generated module M over a
local ring In abstract algebra, more specifically ring theory, local rings are certain rings that are comparatively simple, and serve to describe what is called "local behaviour", in the sense of functions defined on varieties or manifolds, or of algebrai ...
R with maximal ideal \mathfrak, the quotient M/\mathfrakM is a vector space over the field R/\mathfrak. Statement 4 then implies that a basis of M/\mathfrakM lifts to a minimal set of generators of M. Conversely, every minimal set of generators of M is obtained in this way, and any two such sets of generators are related by an invertible matrix with entries in the ring.


Geometric interpretation

In this form, Nakayama's lemma takes on concrete geometrical significance. Local rings arise in geometry as the germs of functions at a point. Finitely generated modules over local rings arise quite often as germs of
sections Section, Sectioning or Sectioned may refer to: Arts, entertainment and media * Section (music), a complete, but not independent, musical idea * Section (typography), a subdivision, especially of a chapter, in books and documents ** Section sig ...
of
vector bundle In mathematics, a vector bundle is a topological construction that makes precise the idea of a family of vector spaces parameterized by another space X (for example X could be a topological space, a manifold, or an algebraic variety): to every p ...
s. Working at the level of germs rather than points, the notion of finite-dimensional vector bundle gives way to that of a
coherent sheaf In mathematics, especially in algebraic geometry and the theory of complex manifolds, coherent sheaves are a class of sheaves closely linked to the geometric properties of the underlying space. The definition of coherent sheaves is made with ref ...
. Informally, Nakayama's lemma says that one can still regard a coherent sheaf as coming from a vector bundle in some sense. More precisely, let \mathcal be a coherent sheaf of \mathcal_X-modules over an arbitrary
scheme A scheme is a systematic plan for the implementation of a certain idea. Scheme or schemer may refer to: Arts and entertainment * ''The Scheme'' (TV series), a BBC Scotland documentary series * The Scheme (band), an English pop band * ''The Schem ...
X. The stalk of \mathcal at a point p\in X, denoted by ''\mathcal_p'', is a module over the local ring (\mathcal_,) and the fiber of \mathcal at p is the vector space \mathcal(p) = \mathcal_p/\mathfrak_p\mathcal_p. Nakayama's lemma implies that a basis of the fiber \mathcal(p) lifts to a minimal set of generators of ''\mathcal_p''. That is: * Any basis of the fiber of a coherent sheaf ''\mathcal'' at a point comes from a minimal basis of local sections. Reformulating this geometrically, if \mathcal is a locally free \mathcal_X-module representing a vector bundle E \to X, and if we take a basis of the vector bundle at a point in the scheme X, this basis can be lifted to a basis of sections of the vector bundle in some neighborhood of the point. We can organize this data diagrammatically
\begin E, _p & \to & E, _U & \to & E \\ \downarrow & & \downarrow & & \downarrow \\ p & \to & U & \to & X \end
where E, _p is an n-dimensional vector space, to say a basis in E, _p (which is a basis of sections of the bundle E_p \to p) can be lifted to a basis of sections E, _U \to U for some neighborhood U of p.


Going up and going down

The going up theorem is essentially a corollary of Nakayama's lemma. It asserts: * Let R \hookrightarrow S be an
integral extension In commutative algebra, an element ''b'' of a commutative ring ''B'' is said to be integral over ''A'', a subring of ''B'', if there are ''n'' ≥ 1 and ''a'j'' in ''A'' such that :b^n + a_ b^ + \cdots + a_1 b + a_0 = 0. That is to say, ''b'' i ...
of commutative rings, and \mathfrak a
prime ideal In algebra, a prime ideal is a subset of a ring that shares many important properties of a prime number in the ring of integers. The prime ideals for the integers are the sets that contain all the multiples of a given prime number, together wi ...
of R. Then there is a prime ideal \mathfrak in S such that \mathfrak\cap R = \mathfrak. Moreover, \mathfrak can be chosen to contain any prime \mathfrak_1 of S such that \mathfrak_1\cap R \subset \mathfrak.


Module epimorphisms

Nakayama's lemma makes precise one sense in which finitely generated modules over a commutative ring are like vector spaces over a field. The following consequence of Nakayama's lemma gives another way in which this is true: *If M is a finitely generated R-module and f:M\to M is a surjective endomorphism, then f is an isomorphism. Over a local ring, one can say more about module epimorphisms: *Suppose that R is a local ring with maximal ideal \mathfrak, and M,N are finitely generated R-modules. If \phi:M\to N is an ''R''-linear map such that the quotient \phi_\mathfrak:M/\mathfrakM \to N/\mathfrakN is surjective, then \phi is surjective.


Homological versions

Nakayama's lemma also has several versions in
homological algebra Homological algebra is the branch of mathematics that studies homology in a general algebraic setting. It is a relatively young discipline, whose origins can be traced to investigations in combinatorial topology (a precursor to algebraic topolo ...
. The above statement about epimorphisms can be used to show: * Let M be a finitely generated module over a local ring. Then M is projective if and only if it is free. This can be used to compute the
Grothendieck group In mathematics, the Grothendieck group, or group of differences, of a commutative monoid is a certain abelian group. This abelian group is constructed from in the most universal way, in the sense that any abelian group containing a homomorphic ...
of any local ring R as K(R) = \mathbb. A geometrical and global counterpart to this is the Serre–Swan theorem, relating projective modules and coherent sheaves. More generally, one has * Let R be a local ring and M a finitely generated module over ''R''. Then the projective dimension of M over R is equal to the length of every minimal free resolution of M. Moreover, the projective dimension is equal to the global dimension of M, which is by definition the smallest integer i \geq 0 such that ::\operatorname_^R(k,M) = 0. :Here k is the residue field of ''R'' and \text is the
tor functor In mathematics, the Tor functors are the derived functors of the tensor product of modules over a ring. Along with the Ext functor, Tor is one of the central concepts of homological algebra, in which ideas from algebraic topology are used to co ...
.


Inverse function theorem

Nakayama's lemma is used to prove a version of the inverse function theorem in algebraic geometry: * Let f: X \to Y be a projective morphism between quasi-projective varieties. Then f is an isomorphism if and only if it is a bijection and the differential df_p is injective for all p \in X.


Proof

A standard proof of the Nakayama lemma uses the following technique due to . * Let ''M'' be an ''R''-module generated by ''n'' elements, and φ: ''M'' → ''M'' an ''R''-linear map. If there is an ideal ''I'' of ''R'' such that φ(''M'') ⊂ ''IM'', then there is a
monic polynomial In algebra, a monic polynomial is a single-variable polynomial (that is, a univariate polynomial) in which the leading coefficient (the nonzero coefficient of highest degree) is equal to 1. Therefore, a monic polynomial has the form: :x^n+c_x^+\ ...
::p(x) = x^n + p_1x^+\cdots + p_n :with ''p''''k'' ∈ ''I''''k'', such that ::p(\varphi)=0 :as an endomorphism of ''M''. This assertion is precisely a generalized version of the Cayley–Hamilton theorem, and the proof proceeds along the same lines. On the generators ''x''''i'' of ''M'', one has a relation of the form :\varphi(x_i) = \sum_^n a_x_j where ''a''''ij'' ∈ ''I''. Thus :\sum_^n\left(\varphi\delta_ - a_\right)x_j = 0. The required result follows by multiplying by the adjugate of the matrix (φδ''ij'' − ''a''''ij'') and invoking
Cramer's rule In linear algebra, Cramer's rule is an explicit formula for the solution of a system of linear equations with as many equations as unknowns, valid whenever the system has a unique solution. It expresses the solution in terms of the determinants o ...
. One finds then det(φδ''ij'' − ''a''''ij'') = 0, so the required polynomial is :p(t) = \det(t\delta_-a_). To prove Nakayama's lemma from the Cayley–Hamilton theorem, assume that ''IM'' = ''M'' and take φ to be the identity on ''M''. Then define a polynomial ''p''(''x'') as above. Then :r=p(1) = 1+p_1+p_2+\cdots+p_n has the required property.


Noncommutative case

A version of the lemma holds for right modules over non-commutative unital rings ''R''. The resulting theorem is sometimes known as the Jacobson–Azumaya theorem. Let J(''R'') be the Jacobson radical of ''R''. If ''U'' is a right module over a ring, ''R'', and ''I'' is a right ideal in ''R'', then define ''U''·''I'' to be the set of all (finite) sums of elements of the form ''u''·''i'', where · is simply the action of ''R'' on ''U''. Necessarily, ''U''·''I'' is a submodule of ''U''. If ''V'' is a
maximal submodule In mathematics, more specifically in ring theory, a maximal ideal is an ideal that is maximal (with respect to set inclusion) amongst all ''proper'' ideals. In other words, ''I'' is a maximal ideal of a ring ''R'' if there are no other ideals co ...
of ''U'', then ''U''/''V'' is
simple Simple or SIMPLE may refer to: *Simplicity, the state or quality of being simple Arts and entertainment * ''Simple'' (album), by Andy Yorke, 2008, and its title track * "Simple" (Florida Georgia Line song), 2018 * "Simple", a song by Johnn ...
. So ''U''·J(''R'') is necessarily a subset of ''V'', by the definition of J(''R'') and the fact that ''U''/''V'' is simple. Thus, if ''U'' contains at least one (proper) maximal submodule, ''U''·J(''R'') is a proper submodule of ''U''. However, this need not hold for arbitrary modules ''U'' over ''R'', for ''U'' need not contain any maximal submodules. Naturally, if ''U'' is a Noetherian module, this holds. If ''R'' is Noetherian, and ''U'' is finitely generated, then ''U'' is a Noetherian module over ''R'', and the conclusion is satisfied. Somewhat remarkable is that the weaker assumption, namely that ''U'' is finitely generated as an ''R''-module (and no finiteness assumption on ''R''), is sufficient to guarantee the conclusion. This is essentially the statement of Nakayama's lemma. Precisely, one has: :Nakayama's lemma: Let ''U'' be a finitely generated right module over a (unital) ring ''R''. If ''U'' is a non-zero module, then ''U''·J(''R'') is a proper submodule of ''U''.


Proof

Let X be a finite subset of U, minimal with respect to the property that it generates U. Since U is non-zero, this set X is nonempty. Denote every element of X by x_i for i\in \. Since X generates U,\sum_^n x_i R = U. Suppose U\cdot \operatorname J(R) = U, to obtain a contradiction. Then every element u \in Ucan be expressed as a finite combination u=\sum\limits_^u_j_ for some m\in\mathbb,\, u_s\in U,\, j_s \in \operatorname J(R), \,s=1,\dots,m. Each u_s can be further decomposed as u_s = \sum\limits_^ x_i r_ for some r_\in R. Therefore, we have u=\sum_^\left( \sum_^x_i r_ \right)j_s = \sum\limits_^x_i \left(\sum_^r_j_s\right). Since \operatorname J(R) is a (two-sided) ideal in R, we have \sum_^r_j_s \in \operatorname J(R) for every i\in\, and thus this becomes :u= \sum_^n x_i k_i for some k_i\in \operatorname J(R), i=1,\dots,n. Putting u=\sum_^x_i and applying distributivity, we obtain :\sum_^n x_i (1 - k_i) = 0. Choose some j\in\. If the right ideal (1-k_j) R were proper, then it would be contained in a maximal right ideal M\neq R and both 1-k_j and k_j would belong to M, leading to a contradiction (note that \operatorname J(R)\subseteq M by the definition of the Jacobson radical). Thus (1-k_j)R=R and 1-k_j has a right inverse (1-k_j)^ in R. We have :\sum_^n x_i (1 - k_i) (1 - k_j)^ = 0. Therefore, :\sum_ x_i (1 - k_i) (1 - k_j)^ = -x_j. Thus x_j is a linear combination of the elements from X\setminus\. This contradicts the minimality of X and establishes the result.;


Graded version

There is also a graded version of Nakayama's lemma. Let ''R'' be a ring that is graded by the ordered semigroup of non-negative integers, and let R_+ denote the ideal generated by positively graded elements. Then if ''M'' is a graded module over ''R'' for which M_i = 0 for ''i'' sufficiently negative (in particular, if ''M'' is finitely generated and ''R'' does not contain elements of negative degree) such that R_+M = M, then M = 0. Of particular importance is the case that ''R'' is a polynomial ring with the standard grading, and ''M'' is a finitely generated module. The proof is much easier than in the ungraded case: taking ''i'' to be the least integer such that M_i \ne 0, we see that M_i does not appear in R_+M, so either M \ne R_+M, or such an ''i'' does not exist, i.e., M = 0.


See also

* Module theory * Serre–Swan theorem


Notes


References

*. *. * * *. * *. *. *. *{{Citation , last1=Nakayama , first1=Tadasi , title=A remark on finitely generated modules , mr=0043770 , year=1951 , journal=Nagoya Mathematical Journal , issn=0027-7630 , volume=3 , pages=139–140 , doi=10.1017/s0027763000012265, doi-access=free .


Links


How to understand Nakayama's Lemma and its Corollaries
Theorems in ring theory Algebraic geometry Commutative algebra Lemmas in algebra