<<

NEW ZEALAND JOURNAL OF MATHEMATICS Volume 26 (1997), 263-273

FLAT DIMENSION, CONSTRUCTIVITY, AND THE HILBERT SYZYGY THEOREM

F r e d R i c h m a n

(Received November 1995)

Abstract. A constructive theory of flat dimension is developed to prove the abstract Hilbert syzygy theorem in terms of flat rather than projective di­ mension. This allows a smooth constructive treatment that does not appeal to countable choice, while still generalizing the concrete theorem about the global projective dimension of polynomial rings over a .

1. Constructive Mathematics

This paper is written in the spirit of constructive mathematics in the sense of Errett Bishop [3]. For a general introduction to constructive algebra see [11]. An algebraist with no background in constructive mathematics should have no trouble understanding the arguments, but might not see what obstacles are being overcome or sidestepped. From a procedural point of view, the difference is that we do not use the law of excluded middle. From an intuitive point of view, the difference is that our theorems are true under a computational interpretation. For example, a set isdiscrete if, for any two elementsx and y, eitherx = y or x ± y. Assuming the law of excluded middle, all sets are discrete; but in the absence of that law, there is no reason to believe that all sets are discrete. Moreover, the computational interpretation of a discrete set is one for which you have an algorithm that decides whether or notx = y. So, for example, the set of binary sequences is not discrete because (1) we cannot prove it is without using some form of the law of excluded middle, or (2) there is no algorithm that will decide, given an algorithm for a binary sequence, whether or not the sequence consists entirely of zeros. This is not to say that it is a theorem in constructive mathematics that the set of binary sequences is not discrete - on the contrary, any theorem in constructive mathematics is also a theorem in classical mathematics. As we can reduce questions about the equality of binary sequences to questions about the equality of real numbers, the real numbers constitute another example of a set that is not discrete. Arguments that require the axiom of choice are not constructively valid; in fact, the axiom of choice implies the law of excluded middle [6]. The failure of the axiom of choice comes up in the present context in that free modules need not be projective.

1991 AMS Mathematics Subject Classification: 13D02, 03F65, 16E10. 264 FRED RICHMAN

2. Projective Dimension

LetR be a . Although our interest is in the case whenR is commutative, none of our results will require this. However, we will not investigate the relationship between the categories of left i?-modules and of right i2-modules. The default is that “” means left module, but this will rarely play any role. If A is an ^-module, with generators (a*)tej, then we can map the free i?-module R ^ \ which consists of finite formal sumsr\i\ H------b rnin, onto A by mapping the elementr\i\ + ------1-rnin to r \ a + • • • + rna,in. The kernelK of this map is the relation module, or module of syzygies. In the nicest situation,K = 0, and A is free on the given generators. Almost as good is whenK is a summand, in which caseA is projective - ifI is not a finite set, then this holds only classically, because the projectivity of free modules requires some sort of choice axiom. With the same proviso, the moduleK is unique up to a free summand, independent of the choice of generating set. Indeed, given another generating set indexed by J', then Schanuel’s trick, again assuming the free modules are projective, shows that

K ® R ^ £* K ' e R W .

If A is projective, we say that theprojective dimension of A is zero, and write pd A = 0. If K is projective, we say that pdA < 1. If pd K < 1, we say that pdA < 2, etc. Theglobal projective dimension of R is the supremum of the projective dimensions of its modules. M. Auslander [1] showed that it is enough to consider cyclic modules, so the idea of computing is not so outlandish as it might appear. Let k be a discrete field. The (inhomogeneous)Hilbert syzygy theorem says that the R = k[x i , . . . , x n] has global projective dimension n. The original (homogeneous) Hilbert syzygy theorem,[10, Theorem VII.6.4], says that graded modules overR have free resolutions of length n. We are interested here in a constructive treatment of the usual (inhomogeneous)abstract Hilbert syzygy theorem, [10, Theorem VII.4.2], that if A; is a of global dimension r, then k[x] has global dimension r + 1. Or at least that is our starting point; we will see that a proper constructive treatment requires a modification of this, while still subsuming the concrete theorem. We have to be careful when defining global dimension for a constructive develop­ ment. Consider the casen = 0 of the concrete theorem. It is too much to ask that every module, or even every finitely generated module, over the rationalQ numbers be projective. an If is a binary sequence with at most one 1, let 5 be the subspace of V = Q ® Q generated by the countable {(l,n) set : an = 1}. Then M = V /S is a finitely generated discrete Q-module, which locally looks two-dimensional if an = 0 for all n up to some large value. But M if were projective, thenS would be a summand of V, hence finite-dimensional, and this would imply that either an = 1 for some n, or an = 0 for all n. So we have to limit the class of modules that we consider for global dimension. One possibility is the class ofdiscrete cyclic modules. This works for fields, but we can not compute the dimension of an arbitrary discrete over the integers - we have to know whether or not it is torsion. This latter observation leads us to consider the class offinitely presented modules - we can compute the FLAT DIMENSION, CONSTRUCTIVITY, AND THE HILBERT SYZYGY THEOREM 265 dimension of any finitely presented abelian group. This takes care of the case n = 0 because finitely presented vector spaces are finite-dimensional, hence free. In order to be able to compute the projective dimensions of modules in a certain class, it would obviously be desirable, and probably essential, for the class to be closed under taking modules of relations. For this to hold for finitely presented modules, we need the ring to coherent be , that is, finitely generated ideals are finitely presented. While this follows classically from the ascending chain condition, it is an additional hypothesis from a constructive point of view. For R coherent and Noetherian, and global dimension defined in terms of the finitely presented modules only, we can prove the abstract inhomogeneous theorem. A central role is played by the countably generated, coherent modules. The reason for this is that R[x] is such an i?-module, and that global dimension computed for such modules is no greater than for finitely presented modules. The factR that is Noetherian plays very little part in the argument: it is used to show that H[x] is coherent([12], [14], [11], [15]), and that coherentR[x ]-modules are coherent /^-modules. But restricting the class of modules changes the classical meaning of global di­ mension. In addition, by requiringR to be Noetherian, we have placed a classically significant restriction on the ring. Less significant, but still annoying, is that we need to invoke the countable axiom of choice - a principle used by all constructive schools - because we need countable rank free modules to be projective. A clue to the proper treatment is that a finitely presented module is projective if and only if it is flat ([4, Exercise 2.23.a] and[11]). In [2] this is used to decide whether a finitely presented module is projective (for commutative rings). It follows thatR if is coherent, then the projective dimension of a finitely presented module is the same as its flat dimension. Thus what we are really computing here is flat dimension. Moreover, for the concrete inhomogeneous Hilbert syzygy theorem, we might as well be dealing with flat dimension.

3. Flat Dimension

What is the definition of flat dimension? If we define it in terms of free resolutions that end up in flat modules, we must worry about dependence on the . The usual proof of independence requires projectivity of the free modules, which requires a choice axiom. In fact this definition turns out to be equivalent to the one we adopt (see Theorem 3, keeping in mind that free modules are flat). We will consider other alternative definitions at the end of the paper. Here is a rephrasing of the usual definition of flat. We say thatA is flat if, whenever we are given a map / from a finite-rank free moduleF into A , and an element9 in the kernel of / , then / factors through a finite-rank , F —> F ' —> A, such that 9 goes to zero inF'. • Vector spaces over discrete fields are flat [11, Exercise III.5.5]. Note, however, that vector spaces over the real numbers need not be flat. In fact,R- all modules are flat if and only ifR is von Neumann regular[11, Exercise III.5.6], and if the real numbers were von Neumann regular, they would be discrete: axa = a implies ax( 1 — ax) = 0, and eitherax ^ 0, whence a / 0, or 1 — ax ^ 0, whenceax = 0, and a = axa = 0. (If r + s = 1 in the real numbers, then either r ^ 0 s or ^ 0. This can be shown by approximating 266 FRED RICHMAN

r and s by rational numbers. It is essentially the statement that the real numbers form a .) • Free modules are flat[11, III.5.2]. This follows rather easily from the defi­ nition of equality in the construction of a free module on an arbitrary index set as a set of formal sums[11, p. 54]. The reason this is not completely trivial is that the index set need not be discrete, and so we must be a little careful in defining equality of formal sums. Notice that every module is the homomorphic image of a free module - there are enough free modules - but the same cannot be said for projective modules. • Summands of flat modules are flat. This follows immediately from the defi­ nition. • Projective modules are flat - they are summands of free modules. • Finitely presented flat modules are projective[11].

We say that A has flat dimension at most n, and write id A < n , if given any complex

Fn'—> F n -i —>•••—> .Fo —> .A (1) where the F* are finite-rank free, and an element9 in Fn that goes to zero in Fn_i, then there exists a commutative diagram Fn —* Fn- 1 —> .. . —► Fo —> A I I I II (2) FL - K - 1 —► • • ■ Fq A where the F/ are finite-rank free,9 and e Fn goes to zero in F^. Some observations:

• By iteration, we can arrange for any finitely enumerable subsetFn ofthat goes to zero in Fn_i to go to zero inF'n . (A set { x i ,x 2, .. . ,x m} is finitely enumerable; it may fail to be finite because it may not be discrete.) • Hence fdA < n if and only if whenever n+1 F —> Fn —► • • • —> Fo —> A is a complex, with theFi finite-rank free, then we can construct a commutative diagram Fn+i —► Fn —¥ Fn-1 —> • • • —> Fo —» A ill i II 0 - K - K-1 - ••• - n - A with the F[ finite-rank free. It is not hard to see that we could replace “finite-rank free” by “finitely generated projective,” which has the virtue of being definable purely in terms of the of modules. Note also that, for n = €, this characterization gives an alternative definition of “flat” which amounts to the characterization in [5, 1.6, Theorem l(iii)] that every map from a finitely presented module factors through a finite-rank free module.

• If fd A < n, then fd.4 < n + 1. This follows easily from the preceding remark. So if we defineidA = {n£N :idA

any event, we will not need to interpret the symbol “fd ^4” standing alone; it will always occur in the expression “fdA < n.”

• We can simplify the notation by lettingF stand for a complex, indexed by the nonnegative integers. map A from a complex F to a module A is a map F0 A so that Fi —* Fo —> A is a complex. This amounts to a map of complexes whereA is considered as a complex withAi — 0 for i 0.

• We say that degF < n if Fm = 0 for m > n. Then fdA < n if and only if every map of a finite-rank free complex intoA factors through one of degree at most n.

• We can think of our complexes as indexed by the integers, Fn with = 0 for n < 0. • The internal maps of a complex will be denoted genericallyd. by

• If F is a complex, the translates F - and F + of F are defined by (F _ )n = Fn- 1 and (F + ) n = Fn+ 1 for n > 0. Note that F _ + = F, and (F-) 0 = 0.

• Given a map / : F —► F f between two complexes, themapping cylinder M ( f ) of / is the complexF © F'+ where the mapd : Fi ® F 'i+i F i-i © F- takes (x ,y) to (~dx,dy + fx). It is readily checked thatM ( /) is a complex, and that the natural map from F'+ to M { f ) is a map of complexes, soF'+ is a subcomplex of M ( f ) in a natural way. In our applications, Fq will always be zero, so nothing gets lost in passing to F'+ .

What does fdA > n mean? For discrete rings it undoubtedly should mean that there exists a complex (1), and an element6 € Fn that goes to zero in Fn_i, so that whenever we have a commutative diagram (2), then6 goes to a nonzero element in F„. For arbitrary rings it is not clear what the good meaning of fdA > n is. We will formulate all our theorems in terms ofA fd < n. So what is the flat Hilbert syzygy theorem? We want to show thatA fd < n for all i?-modules A , if and only if fd A < n -f 1 for all R[x]-modules A. The main tool is a lemma relating the flat dimensions of modules in a short . In proving this we will use the following construction of a map into a mapping cylinder.

Lemma 1. Let F , F ' and F " be complexes. Suppose we are given maps f : F —* F'_ and g : M ( f ) - F " . Let g' be the restriction of g to F ', viewed as a map from F'_ —> F'L • Then there is a map F —► M (g ') that takes x € Fi to ( -1 y(-fx,gx)eFU ®F''.

Proof. We have to show that the square

Fi+1 $ Fi i i F( © F ''+1 F '_ x © F" commutes, where the maps down takex € Fi+\ to (—1 )l(fx, —gx) and x € Fi to (—1 )l(—fx,gx), and the bottom map takes(x ,y) to (—dx ,d y + gx). 268 FRED RICHMAN

Let x e Fi+ 1. Going over and down takesx to dx to (—1 )l(—fdx,gdx). Going down and over takesx to (—1 Y (fx , —gx) to (—1 )l(—dfx, — dgx + gfx). But fd x = d fx because / is a map of complexes, andgdx = —dgx + gfx because g : M ( f ) —> F" is a map of complexes, and the boundary ofx in M (/) is {—dx, fx ), so dgx = g(-dx, fx) = —gdx + g fx . □

Here is the main lemma.

Lemma 2. Let 0—►A— ►O&ean exact sequence of R-modules. If fd A < n and fdB < n + 1, then fdC < n + 1. IffdB

Proof. Throughout the proof, use of the bold letterF will indicate a complex of finite-rank free modules. We will considerA to be a submodule ofB. To show that fdC < n + 1, suppose we have a mapF —* C. Lift the map Fo —» C to a map (3 : Fq —> B. The composition Fi —► Fo —» B maps F\ into A , giving a map (3d :F + ^ A. As id. A < n, there is a complexF with degF' < n, and maps F + —> F' and a : F ' —>• A, so that the diagram

(3d F + A 1 II (3) a F' A commutes. The mapping cylinderM ( f ) of the map / : F —» F'_ corresponding to F + -> F' maps into B via the map Fo ©Fq —> B that is the sum of the given maps {3 : Fo —+ B and a : Fq —> A. We need to show thatM ( f ) —> B is a complex at F o ® F q. But (x, y) e F i ® F[ goes to ( -d x , dy + fx ) e F 0 © F q which goes to —(3dx + a fx G B which is zero as (3d = a f from diagram (3). As fd B < n + 1 there is a commutative diagram

M ( f ) B 91 II (4) F ” B with degF " < n + 1 . Let g' : F'_ —> F'L be the map induced by the restriction of g to F'. Lemma 1 gives us a mapF —> M (g ') that is equal to g on Fo. We can map M (g ') to C by the composite Fq' —►B —> C. This makes the diagram

F -► C i II M (g ') ^ C commute because diagram (4) commutes. The bottom row is a complex because the last three terms are

Fq © F[' —> Fo' —>• C and (x ,y) € Fq © F" goes to - d y -I- gx e Fq . The term- d y goes to 0 already in B, while x is taken by a into A, so gx goes to 0 in C - see diagram (4). As degM (g ') < n + 1, this proves the first part of the lemma. FLAT DIMENSION, CONSTRUCTIVITY, AND THE HILBERT SYZYGY THEOREM 269

To show that fd A < n, suppose we are given a mapa : F —> A. As fd B < n we can construct a commutative diagram

F A A 1 i F' A B with degF ' < n. Let / : F_ —> F'_ be the map induced by the mapF —>■ F'. Let 7r denote the given map fromB onto C. The mapping cylinderM (/) maps into C via the compositeFq A B A C. We need to show thatM ( f ) ^ C is a complex at Fq. The element(x, y) G Fq ® F{ goes to dy + f x G Fq, and Ady = 0, because F ' B is a complex, while f A x = ax G A. As fd C < n + 1 there is a commutative diagram

M ( f ) ^ C 9 i II F " C

The last two squares inthis diagram are 7rA F 0 © F i' - K C 1 1 II d T?n F i" 0 c from which we see that x if € Fo, then g fx = dgx. Letg' : F'_ —> F'L be the map induced byg restricted toF'. Lemma1 gives us a map F _ —> M (g '), and hence a mapF —> M (g ')+ , that takes x E Fq = (F-) i to (fx , —gx). Lift the map Fq —► C to a map fi : Fq —> B. The difference^ = fig — X of the compositesF q Fq' A B and Fq A B goes into A because it goes to 0 in C by the last square in the diagram above. The composite77: F " —> Fq' A B also goes into A, becauseF " —> C is a complex, so we get a commutative diagram of complexes

F A A i II

A M (g ') +

The only problems are observing that the bottom row is a complexF qat ® F " , and that the right-hand square commutes. The last two squares of this diagram look like:

Fi —► Fo A A i 1 II F i © F'2 Fq © F " A

In the bottom row, (x,y) G F{ © F2 goes to (—dx ,d y + gx) which goes to Hgdx — Xdx + fiddy -f [idgx = 0 as g commutes with d and Xd = 0. To see that the last square commutes, letx be in Fo. Down and right takesx to (fx , —gx) to — £/x + rjgx = —figfx + A f x + \idgx = A f x = ax. □ 270 FRED RICHMAN

We leave to the reader to prove that A if

Theorem 3.Let 0 —► iiT —► Fn_ i —► • • • —* Fo —* A —* 0 be an exact sequence with each Fi a . Then fd A < n if and only if K is flat.

Proof. If K is flat, then repeated application of Lemma 2, starting by proving that the image of Fn_i in n_2 F has flat dimension at most one, shows that fdA < n. Conversely, if fd^4 < n, then repeated application of Lemma 2, starting by proving that the kernel of the map Fo A —> has flat dimension at most n — 1, shows that K is flat. □

Corollary 4. If A is a finitely presented module over a coherent ring, then i & A < n if and only if the projective dimension of A is at most n.

Proof. As finitely presented modules are flat if and only if they are projective, this is an immediate consequence of Theorem 3. □

4. The Flat Hilbert Syzygy Theorem

If R is a ring, then R[x] denotes the polynomial ring in one variablex with coefficients inR, wherex commutes with each elementR. of If M is an i?-module, then M[x\ is the set of polynomials with coefficients M. in Clearly M[x] is an i?-module, and an #[x]-module, in a natural way. In [11, Theorem III.5.3] it is shown that a (left) moduleM is flat if and only if wheneverA is a submodule of a right moduleB , then the natural map from A M to B ® M is one-to-one - a standard classical result. As a consequence, if R —► S is a map of rings, and M is a flat /2-module, thenS r M is a flat S- module. In particular, using this observation with the natural ring mapsR —► R[x] and R[x] —► R, we see thatM is a flat i?-module if and only if M[x] is a flat i?[x]-module.

Theorem 5. If A is an R-module, then fd^A < n if and only if fd^^jAfx] < n.

Proof. We observed the case n = 0 above. n For > 0, take a free resolution 0 —> K —> F A —►OofA, giving a free resolution 0 —►K[x] —> F[x] —► A[x) —> 0 of A[x\. By induction, fdr K < n — 1 if and only if fd#[x] K[x] < n — 1, so the theorem follows from Theorem 3. □

The following result uses the key construction in the proof (by contradiction) of [13, Theorem 9.33]. By fd^A <00 we mean that fd^A

Lemma 6. Let A be an R[x\-module such that xA = 0 and fd^-A < 00. If X]A < n + 1 , then fd^-A < n. FLAT DIMENSION, CONSTRUCTIVITY, AND THE HILBERT SYZYGY THEOREM 271

Proof. Let 0—> K —> F —> A —> 0 be a free i?[x]-resolution A. of Then we have an exact sequence ' x F K F A n 0 —* —— —► —— —► —— —► A —► 0. x K x K x F If n = 0, then K is flat over .R[x], so K /x K = R <8>fl[a;] K is flat over R. As A is isomorphic to x F /x K , if fd^-A < m, then fd^A < max(0, m — 2), so A is a flat .R-module. If n > 0, then fdr [x]K < n — 1 and, as F is also a free ^-module, r fdK < oo. By induction, fd^if < n — 1, so fd^-A < n . □

The next theorem is [13, Lemma 9.29].

Theorem 7.If A is an R[x]-module, then the sequence

0 —► A[x] —> A[x] —* A —> 0 is exact, where the second map takes ax 1 to x la, and the first map takes ax 1 to xax I —__ ax /Tf 1

Proof. The first map takesY^i=o aixl to

xao + (xai — ao)x + (xa2 - a\)x2 ------akXk+1.

The kernel of the first map is zero, and the composition is zero. Supposex lai — 0. For i > 0 define

—b{ = (Xi+i + xai+2 + x 2 ai+3 4 - . . . .

Then xbi — 6i_i = ai for i > 1, and xbo = ao- □

Corollary 8 (Flat Hilbert Syzygy). Let R be a ring and n a nonnegative integer. Then the following conditions are equivalent.

1. Every R-module has flat dimension at most n.

2. Every R[x]-module has flat dimension at most n + 1.

Proof. Suppose (1) holds andA is an i?[x]-module. Then fd^A < n, so fdfi[xjyl[x] < n. Hence fd^fx]i4 < n -I-1 by Theorem 7 and Lemma 2. Conversely, suppose (2) holds and A is an .R-module. ThenA becomes an i?[x]-module if we setxA = 0, and fdft[x]A < n 4 1. Hence fd^A < n by Lemma6. □

5. Flat Dimension via Tor

Another definition of flat dimension is that Torn(A ,B ) = 0 for all modules A and B, where the Torn is the one constructed without projective resolutions by MacLane [10, V.7]. The elements of MacLane’s Torn are pairs of complexes

Fn * -^n—1 * * ‘ * * Fq * A

Fn —» Fn- 1 —> • • • —> Fo —> B where theFi are finite-rank free, andF* — HornR(Fj, R). Thus the basic structures involved in this definition of Torn are the same as in our definition of flat dimension. MacLane’s treatment appears to go through without a constructive hitch. 272 FRED RICHMAN

Alternatively we could define Torn by taking a free resolutionB, oftensoring it with A , and showing that the homology of the resulting complex is a functor (and so does not depend on the resolution B). of The usual constructions use arbitrary projective resolution. The existence of projective resolutions derives from the existence of free resolutions, which is constructive, and the projectivity of free modules, which is not. We can circumvent this by takingcanonical free resolutions: for any i?-module B there is a canonical free module mapping onto it, namely the free moduleFi (B) on the underlying set ofB. It follows from the universal property of free modules thatF i(B ) is a functor (see [11] for constructive details). Set Fo(B) = B and F -i( B ) = 0, and inductively defineFi(B) to be F\ of the kernel of the map fromF i-i(B ) to F i-2(B) for i > 0. The result is a functorial free resolution ofB. Given a map B —* B ' we get a commutative diagram of exact sequences

. .. F3(B) -> F2{B) Fi(B) —► B —► 0 I I 111 ... -H. F3{B') -> F2{B') -> Fi(B') -> 5' -> 0 hence a commutative diagram

••• -+ A ® F 3(B) A ® F 2(B) -¥ A®Fi(B) -*■ A®B I 1 I i . .. A ® F 3(B') -+ A ® F 2{B’) -> A ® F i{ B ') -> A ® B ' from which it is easy to see that Torn(.A,B) is a functor. The question of dependence on resolution does not come up because we take a canonical one. Thus we can construct Torn(A, B), but its key properties are more difficult to verify. Clearly Torn(A,B) = 0 if A is flat, but it’s not so clear that Torn(^4,B) = 0 if B is flat. This can be established by showing that a short exact sequence—* K 0 —* F —> F /K —> 0 remains exact after tensoring provided thatF/K is flat, and that K is flat if F and F/K are (the latter is part of Lemma 2). We also have to show that, conversely, if Tori (A,B) = 0 for everyA , then B is flat. This may be deduced from the long exact sequence

• Tori(vl2, B) —> Tori (A3, B) —» A\ ® B —* A 2 ® B —► A 3 ® B —* 0 which is easy to construct because the free resolutionB consistsof of flat modules, so tensoring with an exact sequence of modules gives an exact sequence of complexes from which to derive the long exact sequence of homology. However to prove Lemma 2 in this development we need the long exact sequence

• • • —> Tori (A, B 2) —► Tori (A B 3) —► A ® B\ —► A ® B 2 —> A ® B 3 —► 0 which is also what we would use to prove that if Tori(A, B) = 0 for everyB, then A is flat. Here the connecting homomorphism Tori(vl,B 3) —> A ® B\ is defined in the classical case by building a resolution ofB 2 from the resolutions ofB\ and B 3, an option not available to us. Perhaps we can prove that if Tori(A, B) = 0 for everyB, then A is flat, without using this latter sequence, in which case we could define the flat dimension A of in terms of the vanishing of Torn(A,B) for all B. But we would want both exact sequences eventually. FLAT DIMENSION, CONSTRUCTIVITY, AND THE HILBERT SYZYGY THEOREM 273

References

1. M. Auslander, On the dimension of modules and algebras (III), Nagoya Math. J. 9 (1955), 67-77. 2. G. Baumslag, F. Cannonito and C. Miller, Computable algebra and group embeddings, J. Algebra 69 (1981), 186-212. 3. E. Bishop, Foundations of Constructive Analysis, McGraw-Hill, 1967. 4. N. Bourbaki, Algebre Commutative: I. Modules Plats, Hermann, Paris, 1961. 5. N. Bourbaki, Algebre: 10. Algebre homologique, Masson, Paris, 1980. 6. N. Goodman and J. Myhill, Choice implies excluded middle, Z. Math. Logic Grundlagen Math. 23 (1978), 461. 7. W . Grobner, Uber die Syzgyien-Theorie der Polynomideale, Monatsh. Math. 53 (1949), 1-16. 8. W. Grobner, Modeme Algebraische Geometrie, Springer-Verlag, 1949. 9. D. Hilbert, Uber die Theorie der algebraischen Formen, Math. Ann. 36 (1890), 473-534. 10. S. MacLane, Homology, Springer-Verlag, 1963. 11. R. Mines, F. Richman and W. Ruitenburg,A Course in Constructive Algebra, Springer-Verlag, 1988. 12. F. Richman, Constructive aspects of Noetherian rings, Proc. Amer. Math. Soc. 44 (1974), 436-441. 13. J.J. Rotman, An Introduction to , Academic Press, 1979. 14. A. Seidenberg,What is Noetherian!, Rend. Sem. Mat. e Fis. Milano44 (1974), 55-61. 15. J-P. Soublin, Anneaux et modules coherents, J. Algebra 15 (1970), 455-472.

Fred Richman Department of Mathematics Florida Atlantic University Boca Raton, FL 33431 Florida U.S.A [email protected]