dimension theory (algebra)
   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 ...
, dimension theory is the study in terms of
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 ...
of the notion
dimension of an algebraic variety In mathematics and specifically in algebraic geometry, the dimension of an algebraic variety may be defined in various equivalent ways. Some of these definitions are of geometric nature, while some other are purely algebraic and rely on commut ...
(and by extension that of a
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 ...
). The need of a ''theory'' for such an apparently simple notion results from the existence of many definitions of the dimension that are equivalent only in the most regular cases (see
Dimension of an algebraic variety In mathematics and specifically in algebraic geometry, the dimension of an algebraic variety may be defined in various equivalent ways. Some of these definitions are of geometric nature, while some other are purely algebraic and rely on commut ...
). A large part of dimension theory consists in studying the conditions under which several dimensions are equal, and many important classes of
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 ...
s may be defined as the rings such that two dimensions are equal; for example, a
regular ring In commutative algebra, a regular local ring is a Noetherian local ring having the property that the minimal number of generators of its maximal ideal is equal to its Krull dimension. In symbols, let ''A'' be a Noetherian local ring with maximal ide ...
is a commutative ring such that the homological dimension is equal to the
Krull dimension In commutative algebra, the Krull dimension of a commutative ring ''R'', named after Wolfgang Krull, is the supremum of the lengths of all chains of prime ideals. The Krull dimension need not be finite even for a Noetherian ring. More generally th ...
. The theory is simpler for
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 ...
s that are finitely generated algebras over a field, which are also
quotient ring In ring theory, a branch of abstract algebra, a quotient ring, also known as factor ring, difference ring or residue class ring, is a construction quite similar to the quotient group in group theory and to the quotient space in linear algebra. ...
s of
polynomial ring In mathematics, especially in the field of algebra, a polynomial ring or polynomial algebra is a ring (which is also a commutative algebra) formed from the set of polynomials in one or more indeterminates (traditionally also called variables ...
s in a finite number of indeterminates over a field. In this case, which is the algebraic counterpart of the case of
affine algebraic set Affine may describe any of various topics concerned with connections or affinities. It may refer to: * Affine, a relative by marriage in law and anthropology * Affine cipher, a special case of the more general substitution cipher * Affine com ...
s, most of the definitions of the dimension are equivalent. For general commutative rings, the lack of geometric interpretation is an obstacle to the development of the theory; in particular, very little is known for non-noetherian rings. (Kaplansky's ''Commutative rings'' gives a good account of the non-noetherian case.) Throughout the article, \dim denotes
Krull dimension In commutative algebra, the Krull dimension of a commutative ring ''R'', named after Wolfgang Krull, is the supremum of the lengths of all chains of prime ideals. The Krull dimension need not be finite even for a Noetherian ring. More generally th ...
of a ring and \operatorname the
height Height is measure of vertical distance, either vertical extent (how "tall" something or someone is) or vertical position (how "high" a point is). For example, "The height of that building is 50 m" or "The height of an airplane in-flight is ab ...
of a prime ideal (i.e., the Krull dimension of the localization at that prime ideal.) Rings are assumed to be commutative except in the last section on dimensions of non-commutative rings.


Basic results

Let ''R'' be a
noetherian ring In mathematics, a Noetherian ring is a ring that satisfies the ascending chain condition on left and right ideals; if the chain condition is satisfied only for left ideals or for right ideals, then the ring is said left-Noetherian or right-Noethe ...
or
valuation ring In abstract algebra, a valuation ring is an integral domain ''D'' such that for every element ''x'' of its field of fractions ''F'', at least one of ''x'' or ''x''−1 belongs to ''D''. Given a field ''F'', if ''D'' is a subring of ''F'' suc ...
. Then \dim R = \dim R + 1. If ''R'' is noetherian, this follows from the fundamental theorem below (in particular,
Krull's principal ideal theorem In commutative algebra, Krull's principal ideal theorem, named after Wolfgang Krull (1899–1971), gives a bound on the height of a principal ideal in a commutative Noetherian ring. The theorem is sometimes referred to by its German name, ''Krull ...
), but it is also a consequence of a more precise result. For any prime ideal \mathfrak in ''R'', \operatorname(\mathfrak R = \operatorname(\mathfrak). \operatorname(\mathfrak) = \operatorname(\mathfrak) + 1 for any prime ideal \mathfrak \supsetneq \mathfrak R /math> in R /math> that contracts to \mathfrak. This can be shown within basic ring theory (cf. Kaplansky, commutative rings). In addition, in each fiber of \operatorname R \to \operatorname R, one cannot have a chain of primes ideals of length \ge 2. Since an
artinian ring In mathematics, specifically abstract algebra, an Artinian ring (sometimes Artin ring) is a ring that satisfies the descending chain condition on (one-sided) ideals; that is, there is no infinite descending sequence of ideals. Artinian rings are ...
(e.g., a field) has dimension zero, by induction one gets a formula: for an artinian ring ''R'', \dim R _1, \dots, x_n= n.


Local rings


Fundamental theorem

Let (R, \mathfrak) be a noetherian local ring and ''I'' a \mathfrak-
primary ideal In mathematics, specifically commutative algebra, a proper ideal ''Q'' of a commutative ring ''A'' is said to be primary if whenever ''xy'' is an element of ''Q'' then ''x'' or ''y'n'' is also an element of ''Q'', for some ''n'' > 0. Fo ...
(i.e., it sits between some power of \mathfrak and \mathfrak). Let F(t) be the Poincaré series of the associated graded ring \operatorname_I R = \oplus_0^\infty I^n / I^. That is, F(t) = \sum_0^\infty \ell(I^n / I^) t^n where \ell refers to the
length of a module In abstract algebra, the length of a module is a generalization of the dimension of a vector space which measures its size. page 153 In particular, as in the case of vector spaces, the only modules of finite length are finitely generated modules. It ...
(over an artinian ring (\operatorname_I R)_0 = R/I). If x_1, \dots, x_s generate ''I'', then their image in I/I^2 have degree 1 and generate \operatorname_I R as R/I-algebra. By the Hilbert–Serre theorem, ''F'' is a rational function with exactly one pole at t=1 of order d \le s. Since (1-t)^ = \sum_0^\infty \binom t^j, we find that the coefficient of t^n in F(t) = (1-t)^d F(t) (1 - t)^ is of the form \sum_0^N a_k \binom = (1 - t)^d F(t)\big, _ + O(n^). That is to say, \ell(I^n / I^) is a polynomial P in ''n'' of degree d - 1. ''P'' is called the
Hilbert polynomial In commutative algebra, the Hilbert function, the Hilbert polynomial, and the Hilbert series of a graded commutative algebra finitely generated over a field are three strongly related notions which measure the growth of the dimension of the homog ...
of \operatorname_I R. We set d(R) = d. We also set \delta(R) to be the minimum number of elements of ''R'' that can generate an \mathfrak-primary ideal of ''R''. Our ambition is to prove the fundamental theorem: \delta(R) = d(R) = \dim R. Since we can take ''s'' to be \delta(R), we already have \delta(R) \ge d(R) from the above. Next we prove d(R) \ge \dim R by induction on d(R). Let \mathfrak_0 \subsetneq \cdots \subsetneq \mathfrak_m be a chain of prime ideals in ''R''. Let D = R/\mathfrak_0 and ''x'' a nonzero nonunit element in ''D''. Since ''x'' is not a zero-divisor, we have the exact sequence 0 \to D \overset\to D \to D/xD \to 0. The degree bound of the Hilbert-Samuel polynomial now implies that d(D) > d(D/xD) \ge d(R/\mathfrak_1). (This essentially follows from the Artin-Rees lemma; see Hilbert-Samuel function for the statement and the proof.) In R/\mathfrak_1, the chain \mathfrak_i becomes a chain of length m - 1 and so, by inductive hypothesis and again by the degree estimate, m-1 \le \dim (R/\mathfrak_1) \le d(R/\mathfrak_1) \le d(D) - 1 \le d(R) - 1. The claim follows. It now remains to show \dim R \ge \delta(R). More precisely, we shall show: (Notice: (x_1, \dots, x_d) is then \mathfrak-primary.) The proof is omitted. It appears, for example, in Atiyah–MacDonald. But it can also be supplied privately; the idea is to use
prime avoidance In algebra, the prime avoidance lemma says that if an ideal ''I'' in a commutative ring ''R'' is contained in a union of finitely many prime ideals ''P'is, then it is contained in ''P'i'' for some ''i''. There are many variations of the ...
.


Consequences of the fundamental theorem

Let (R, \mathfrak) be a noetherian local ring and put k = R/\mathfrak. Then *\dim R \le \dim_k \mathfrak/\mathfrak^2, since a basis of \mathfrak/\mathfrak^2 lifts to a generating set of \mathfrak by Nakayama. If the equality holds, then ''R'' is called a
regular local ring In commutative algebra, a regular local ring is a Noetherian local ring having the property that the minimal number of generators of its maximal ideal is equal to its Krull dimension. In symbols, let ''A'' be a Noetherian local ring with maximal ide ...
. *\dim \widehat = \dim R, since \operatornameR = \operatorname\widehat. *(
Krull's principal ideal theorem In commutative algebra, Krull's principal ideal theorem, named after Wolfgang Krull (1899–1971), gives a bound on the height of a principal ideal in a commutative Noetherian ring. The theorem is sometimes referred to by its German name, ''Krull ...
) The height of the ideal generated by elements x_1, \dots, x_s in a noetherian ring is at most ''s''. Conversely, a prime ideal of height ''s'' is minimal over an ideal generated by ''s'' elements. (Proof: Let \mathfrak be a prime ideal minimal over such an ideal. Then s \ge \dim R_\mathfrak = \operatorname \mathfrak. The converse was shown in the course of the proof of the fundamental theorem.) Proof: Let x_1, \dots, x_n generate a \mathfrak_A-primary ideal and y_1, \dots, y_m be such that their images generate a \mathfrak_B/\mathfrak_A B-primary ideal. Then ^s \subset (y_1, \dots, y_m) + \mathfrak_A B for some ''s''. Raising both sides to higher powers, we see some power of \mathfrak_B is contained in (y_1, \dots, y_m, x_1, \dots, x_n); i.e., the latter ideal is \mathfrak_B-primary; thus, m + n \ge \dim B. The equality is a straightforward application of the going-down property.
Q.E.D. Q.E.D. or QED is an initialism of the Latin phrase , meaning "which was to be demonstrated". Literally it states "what was to be shown". Traditionally, the abbreviation is placed at the end of mathematical proofs and philosophical arguments in pri ...
Proof: If \mathfrak_0 \subsetneq \mathfrak_1 \subsetneq \cdots \subsetneq \mathfrak_n are a chain of prime ideals in ''R'', then \mathfrak_iR /math> are a chain of prime ideals in R /math> while \mathfrak_nR /math> is not a maximal ideal. Thus, \dim R + 1 \le \dim R /math>. For the reverse inequality, let \mathfrak be a maximal ideal of R /math> and \mathfrak = R \cap \mathfrak. Clearly, R \mathfrak = R_ \mathfrak. Since R / \mathfrak R_ R = (R_/\mathfrakR_) is then a localization of a principal ideal domain and has dimension at most one, we get 1 + \dim R \ge 1 + \dim R_\mathfrak \ge \dim R \mathfrak by the previous inequality. Since \mathfrak is arbitrary, it follows 1 + \dim R \ge \dim R /math>. Q.E.D.


Nagata's altitude formula

Proof: First suppose R' is a polynomial ring. By induction on the number of variables, it is enough to consider the case R' = R /math>. Since ''R'' is flat over ''R'', \dim R'_ = \dim R_ + \dim \kappa(\mathfrak) \otimes_R _. By Noether's normalization lemma, the second term on the right side is: \dim \kappa(\mathfrak) \otimes_R R' - \dim \kappa(\mathfrak) \otimes_R R'/\mathfrak' = 1 - \operatorname_ \kappa(\mathfrak') = \operatorname_R R' - \operatorname \kappa(\mathfrak'). Next, suppose R' is generated by a single element; thus, R' = R I. If ''I'' = 0, then we are already done. Suppose not. Then R' is algebraic over ''R'' and so \operatorname_R R' = 0. Since ''R'' is a subring of ''R'', I \cap R = 0 and so \operatorname I = \dim R I = \dim Q(R) I = 1 - \operatorname_ \kappa(I) = 1 since \kappa(I) = Q(R') is algebraic over Q(R). Let \mathfrak^ denote the pre-image in R /math> of \mathfrak'. Then, as \kappa(\mathfrak^) = \kappa(\mathfrak), by the polynomial case, \operatorname = \operatorname \le \operatorname - \operatorname = \dim R_ - \operatorname_ \kappa(\mathfrak'). Here, note that the inequality is the equality if ''R'' is catenary. Finally, working with a chain of prime ideals, it is straightforward to reduce the general case to the above case. Q.E.D.


Homological methods


Regular rings

Let ''R'' be a noetherian ring. The
projective dimension In mathematics, particularly in algebra, the class of projective modules enlarges the class of free modules (that is, modules with basis vectors) over a ring, by keeping some of the main properties of free modules. Various equivalent characterizat ...
of a finite ''R''-module ''M'' is the shortest length of any projective resolution of ''M'' (possibly infinite) and is denoted by \operatorname_R M. We set \operatorname R = \sup \; it is called the
global dimension In ring theory and homological algebra, the global dimension (or global homological dimension; sometimes just called homological dimension) of a ring ''A'' denoted gl dim ''A'', is a non-negative integer or infinity which is a homological invariant ...
of ''R''. Assume ''R'' is local with residue field ''k''. Proof: We claim: for any finite ''R''-module ''M'', \operatorname_R M \le n \Leftrightarrow \operatorname^R_(M, k) = 0. By dimension shifting (cf. the proof of Theorem of Serre below), it is enough to prove this for n = 0. But then, by the local criterion for flatness, \operatorname^R_1(M, k) = 0 \Rightarrow M\text \Rightarrow M\text \Rightarrow \operatorname_R(M) \le 0. Now, \operatorname R \le n \Rightarrow \operatorname_R k \le n \Rightarrow \operatorname^R_(-, k) = 0 \Rightarrow \operatorname_R - \le n \Rightarrow \operatorname R \le n, completing the proof. Q.E.D. Remark: The proof also shows that \operatorname_R K = \operatorname_R M - 1 if ''M'' is not free and K is the kernel of some surjection from a free module to ''M''. Proof: If \operatorname_R M = 0, then ''M'' is ''R''-free and thus M \otimes R_1 is R_1-free. Next suppose \operatorname_R M > 0. Then we have: \operatorname_R K = \operatorname_R M - 1 as in the remark above. Thus, by induction, it is enough to consider the case \operatorname_R M = 1. Then there is a projective resolution: 0 \to P_1 \to P_0 \to M \to 0, which gives: \operatorname^R_1(M, R_1) \to P_1 \otimes R_1 \to P_0 \otimes R_1 \to M \otimes R_1 \to 0. But \operatorname^R_1(M, R_1) = _f M = \ = 0. Hence, \operatorname_R (M \otimes R_1) is at most 1. Q.E.D. Proof: If ''R'' is regular, we can write k = R/(f_1, \dots, f_n), f_i a regular system of parameters. An exact sequence 0 \to M \overset\to M \to M_1 \to 0, some ''f'' in the maximal ideal, of finite modules, \operatorname_R M < \infty, gives us: 0 = \operatorname^R_(M, k) \to \operatorname^R_(M_1, k) \to \operatorname^R_i(M, k) \overset\to \operatorname^R_i(M, k), \quad i \ge \operatorname_R M. But ''f'' here is zero since it kills ''k''. Thus, \operatorname^R_(M_1, k) \simeq \operatorname^R_i(M, k) and consequently \operatorname_R M_1 = 1 + \operatorname_R M. Using this, we get: \operatorname_R k = 1 + \operatorname_R (R/(f_1, \dots, f_)) = \cdots = n. The proof of the converse is by induction on \dim R. We begin with the inductive step. Set R_1 = R/f_1 R, f_1 among a system of parameters. To show ''R'' is regular, it is enough to show R_1 is regular. But, since \dim R_1 < \dim R, by inductive hypothesis and the preceding lemma with M = \mathfrak, \operatorname R < \infty \Rightarrow \operatorname R_1 = \operatorname_ k \le \operatorname_ \mathfrak / f_1 \mathfrak < \infty \Rightarrow R_1 \text. The basic step remains. Suppose \dim R = 0. We claim \operatornameR = 0 if it is finite. (This would imply that ''R'' is a semisimple local ring; i.e., a field.) If that is not the case, then there is some finite module M with 0 < \operatorname_R M < \infty and thus in fact we can find ''M'' with \operatorname_R M = 1. By Nakayama's lemma, there is a surjection F \to M from a free module ''F'' to ''M'' whose kernel ''K'' is contained in \mathfrak F. Since \dim R = 0, the maximal ideal \mathfrak is an
associated prime In abstract algebra, an associated prime of a module ''M'' over a ring ''R'' is a type of prime ideal of ''R'' that arises as an annihilator of a (prime) submodule of ''M''. The set of associated primes is usually denoted by \operatorname_R(M), ...
of ''R''; i.e., \mathfrak = \operatorname(s) for some nonzero ''s'' in ''R''. Since K \subset \mathfrak F, s K = 0. Since ''K'' is not zero and is free, this implies s = 0, which is absurd. Q.E.D. Proof: Let ''R'' be a regular local ring. Then \operatornameR \simeq k _1, \dots, x_d/math>, which is an integrally closed domain. It is a standard algebra exercise to show this implies that ''R'' is an integrally closed domain. Now, we need to show every divisorial ideal is principal; i.e., the divisor class group of ''R'' vanishes. But, according to Bourbaki, ''Algèbre commutative,'' chapitre 7, §. 4. Corollary 2 to Proposition 16, a divisorial ideal is principal if it admits a finite free resolution, which is indeed the case by the theorem. Q.E.D.


Depth

Let ''R'' be a ring and ''M'' a module over it. A sequence of elements x_1, \dots, x_n in R is called an ''M''-
regular sequence In commutative algebra, a regular sequence is a sequence of elements of a commutative ring which are as independent as possible, in a precise sense. This is the algebraic analogue of the geometric notion of a complete intersection. Definitions Fo ...
if x_1 is not a zero-divisor on M and x_i is not a zero divisor on M/(x_1, \dots, x_)M for each i = 2, \dots, n. ''A priori'', it is not obvious whether any permutation of a regular sequence is still regular (see the section below for some positive answer.) Let ''R'' be a local Noetherian ring with maximal ideal \mathfrak and put k = R/\mathfrak. Then, by definition, the depth of a finite ''R''-module ''M'' is the supremum of the lengths of all ''M''-regular sequences in \mathfrak. For example, we have \operatorname M = 0 \Leftrightarrow \mathfrak consists of zerodivisors on ''M'' \Leftrightarrow \mathfrak is associated with ''M''. By induction, we find \operatorname M \le \dim R/ for any associated primes \mathfrak of ''M''. In particular, \operatorname M \le \dim M. If the equality holds for ''M'' = ''R'', ''R'' is called a
Cohen–Macaulay ring In mathematics, a Cohen–Macaulay ring is a commutative ring with some of the algebro-geometric properties of a smooth variety, such as local equidimensionality. Under mild assumptions, a local ring is Cohen–Macaulay exactly when it is a fin ...
. Example: A regular Noetherian local ring is Cohen–Macaulay (since a regular system of parameters is an ''R''-regular sequence.) In general, a Noetherian ring is called a Cohen–Macaulay ring if the localizations at all maximal ideals are Cohen–Macaulay. We note that a Cohen–Macaulay ring is universally catenary. This implies for example that a polynomial ring k _1, \dots, x_d/math> is universally catenary since it is regular and thus Cohen–Macaulay. Proof: We first prove by induction on ''n'' the following statement: for every ''R''-module ''M'' and every ''M''-regular sequence x_1, \dots, x_n in \mathfrak, The basic step ''n'' = 0 is trivial. Next, by inductive hypothesis, \operatorname_R^(N, M) \simeq \operatorname_R(N, M/(x_1, \dots, x_) M). But the latter is zero since the annihilator of ''N'' contains some power of x_n. Thus, from the exact sequence 0 \to M \overset \to M \to M_1 \to 0 and the fact that x_1 kills ''N'', using the inductive hypothesis again, we get \operatorname^n_R(N, M) \simeq \operatorname^_R(N, M/x_1 M) \simeq \operatorname_R(N, M/(x_1, \dots, x_n) M), proving (). Now, if n < \operatorname M, then we can find an ''M''-regular sequence of length more than ''n'' and so by () we see \operatorname_R^n(N, M) = 0. It remains to show \operatorname_R^n(N, M) \ne 0 if n = \operatorname M. By () we can assume ''n'' = 0. Then \mathfrak is associated with ''M''; thus is in the support of ''M''. On the other hand, \mathfrak \in \operatorname(N). It follows by linear algebra that there is a nonzero homomorphism from ''N'' to ''M'' modulo \mathfrak; hence, one from ''N'' to ''M'' by Nakayama's lemma. Q.E.D. The Auslander–Buchsbaum formula relates depth and projective dimension. Proof: We argue by induction on \operatorname_R M, the basic case (i.e., ''M'' free) being trivial. By Nakayama's lemma, we have the exact sequence 0 \to K \overset\to F \to M \to 0 where ''F'' is free and the image of ''f'' is contained in \mathfrak F. Since \operatorname_R K = \operatorname_R M - 1, what we need to show is \operatorname K = \operatorname M + 1. Since ''f'' kills ''k'', the exact sequence yields: for any ''i'', \operatorname_R^i(k, F) \to \operatorname_R^i(k, M) \to \operatorname_R^(k, K) \to 0. Note the left-most term is zero if i < \operatorname R. If i < \operatorname K - 1, then since \operatorname K \le \operatorname R by inductive hypothesis, we see \operatorname_R^i(k, M) = 0. If i = \operatorname K - 1, then \operatorname_R^(k, K) \ne 0 and it must be \operatorname_R^i(k, M) \ne 0. Q.E.D. As a matter of notation, for any ''R''-module ''M'', we let \Gamma_(M) = \ = \. One sees without difficulty that \Gamma_ is a left-exact functor and then let H^j_ = R^j \Gamma_ be its ''j''-th
right derived functor In mathematics, certain functors may be ''derived'' to obtain other functors closely related to the original ones. This operation, while fairly abstract, unifies a number of constructions throughout mathematics. Motivation It was noted in va ...
, called the
local cohomology In algebraic geometry, local cohomology is an algebraic analogue of relative cohomology. Alexander Grothendieck introduced it in seminars in Harvard in 1961 written up by , and in 1961-2 at IHES written up as SGA2 - , republished as . Given a fu ...
of ''R''. Since \Gamma_(M) = \varinjlim \operatorname_R(R/\mathfrak^j, M), via abstract nonsense, H^i_(M) = \varinjlim \operatorname^i_R (R/^j, M). This observation proves the first part of the theorem below. Proof: 1. is already noted (except to show the nonvanishing at the degree equal to the depth of ''M''; use induction to see this) and 3. is a general fact by abstract nonsense. 2. is a consequence of an explicit computation of a local cohomology by means of Koszul complexes (see below). \square


Koszul complex

Let ''R'' be a ring and ''x'' an element in it. We form the
chain complex In mathematics, a chain complex is an algebraic structure that consists of a sequence of abelian groups (or modules) and a sequence of homomorphisms between consecutive groups such that the image of each homomorphism is included in the kernel of t ...
''K''(''x'') given by K(x)_i = R for ''i'' = 0, 1 and K(x)_i = 0 for any other ''i'' with the differential d: K_1(R) \to K_0(R), \, r \mapsto xr. For any ''R''-module ''M'', we then get the complex K(x, M) = K(x) \otimes_R M with the differential d \otimes 1 and let \operatorname_*(x, M) = \operatorname_*(K(x, M)) be its homology. Note: \operatorname_0(x, M) = M/xM , \operatorname_1(x, M) = _x M = \. More generally, given a finite sequence x_1, \dots, x_n of elements in a ring ''R'', we form the
tensor product of complexes In mathematics, the tensor product of modules is a construction that allows arguments about bilinear maps (e.g. multiplication) to be carried out in terms of linear maps. The module construction is analogous to the construction of the tensor produ ...
: K(x_1, \dots, x_n) = K(x_1) \otimes \dots \otimes K(x_n) and let \operatorname_*(x_1, \dots, x_n, M) = \operatorname_*(K(x_1, \dots, x_n, M)) its homology. As before, \operatorname_0(\underline, M) = M/(x_1, \dots, x_n)M, \operatorname_n(\underline, M) = \operatorname_M((x_1, \dots, x_n)). We now have the homological characterization of a regular sequence. A Koszul complex is a powerful computational tool. For instance, it follows from the theorem and the corollary \operatorname^i_(M) \simeq \varinjlim \operatorname^i(K(x_1^j, \dots, x_n^j; M)) (Here, one uses the self-duality of a Koszul complex; see Proposition 17.15. of Eisenbud, ''Commutative Algebra with a View Toward Algebraic Geometry''.) Another instance would be Remark: The theorem can be used to give a second quick proof of Serre's theorem, that ''R'' is regular if and only if it has finite global dimension. Indeed, by the above theorem, \operatorname^R_s(k, k) \ne 0 and thus \operatorname R \ge s. On the other hand, as \operatorname R = \operatorname_R k, the Auslander–Buchsbaum formula gives \operatorname R = \dim R. Hence, \dim R \le s \le \operatorname R = \dim R. We next use a Koszul homology to define and study
complete intersection ring In commutative algebra, a complete intersection ring is a commutative ring similar to the coordinate rings of varieties that are complete intersections. Informally, they can be thought of roughly as the local rings that can be defined using the "min ...
s. Let ''R'' be a Noetherian local ring. By definition, the
first deviation In commutative algebra, the deviations of a local ring ''R'' are certain invariants ε''i''(''R'') that measure how far the ring is from being regular. Definition The deviations ε''n'' of a local ring ''R'' with residue field ''k'' are non-neg ...
of ''R'' is the vector space dimension \epsilon_1(R) = \dim_k \operatorname_1(\underline) where \underline = (x_1, \dots, x_d) is a system of parameters. By definition, ''R'' is a complete intersection ring if \dim R + \epsilon_1(R) is the dimension of the tangent space. (See Hartshorne for a geometric meaning.)


Injective dimension and Tor dimensions

Let ''R'' be a ring. The
injective dimension In mathematics, especially in the area of abstract algebra known as module theory, an injective module is a module ''Q'' that shares certain desirable properties with the Z-module Q of all rational numbers. Specifically, if ''Q'' is a submodule of ...
of an ''R''-module ''M'' denoted by \operatorname_R M is defined just like a projective dimension: it is the minimal length of an injective resolution of ''M''. Let \operatorname_R be the category of ''R''-modules. Proof: Suppose \operatorname R \le n. Let ''M'' be an ''R''-module and consider a resolution 0 \to M \to I_0 \overset\to I_1 \to \dots \to I_ \overset\to N \to 0 where I_i are injective modules. For any ideal ''I'', \operatorname^1_R(R/I, N) \simeq \operatorname^2_R(R/I, \operatorname(\phi_)) \simeq \dots \simeq \operatorname_R^(R/I, M), which is zero since \operatorname_R^(R/I, -) is computed via a projective resolution of R/I. Thus, by Baer's criterion, ''N'' is injective. We conclude that \sup \ \le n. Essentially by reversing the arrows, one can also prove the implication in the other way. Q.E.D. The theorem suggests that we consider a sort of a dual of a global dimension: \operatorname = \inf \ . It was originally called the weak global dimension of ''R'' but today it is more commonly called the Tor dimension of ''R''. Remark: for any ring ''R'', \operatorname R \le \operatorname R.


Multiplicity theory


Dimensions of non-commutative rings

Let ''A'' be a graded algebra over a field ''k''. If ''V'' is a finite-dimensional generating subspace of ''A'', then we let f(n) = \dim_k V^n and then put \operatorname(A) = \limsup_ . It is called the Gelfand–Kirillov dimension of ''A''. It is easy to show \operatorname(A) is independent of a choice of ''V''. Example: If ''A'' is finite-dimensional, then gk(''A'') = 0. If ''A'' is an affine ring, then gk(''A'') = Krull dimension of ''A''.


See also

* Bass number *
Perfect complex In algebra, a perfect complex of modules over a commutative ring ''A'' is an object in the derived category of ''A''-modules that is quasi-isomorphic to a bounded complex of finite projective ''A''-modules. A perfect module is a module that is per ...
*
amplitude The amplitude of a periodic variable is a measure of its change in a single period (such as time or spatial period). The amplitude of a non-periodic signal is its magnitude compared with a reference value. There are various definitions of am ...


Notes


References

* * Part II of . * Chapter 10 of . * Kaplansky, Irving, ''Commutative rings'', Allyn and Bacon, 1970. * H. Matsumura ''Commutative ring theory.'' Translated from the Japanese by M. Reid. Second edition. Cambridge Studies in Advanced Mathematics, 8. * * {{cite book , last=Weibel , first=Charles A. , author-link=Charles Weibel , title=An Introduction to Homological Algebra , year=1995 , publisher=Cambridge University Press Dimension Commutative algebra