Products in spinc-cobordism
Abstract.
We calculate the mod spinc-cobordism ring up to uniform -isomorphism (i.e., inseparable isogeny). As a consequence we get the prime ideal spectrum of the mod spinc-cobordism ring. We also calculate the mod spinc-cobordism ring “on the nose” in degrees . We construct an infinitely generated nonunital subring of the -torsion in the spinc-cobordism ring. We use our calculations of product structure in the spin and spinc cobordism rings to give an explicit example, up to cobordism, of a compact -dimensional spin manifold which is not cobordant to a sum of squares, which was asked about in a 1965 question of Milnor.
1. Introduction and summary of results
1.1. Spinc cobordism
A spinc-structure on a compact smooth -dimensional manifold is a reduction of its structure group from to . We find the following perspective illuminating: a compact smooth manifold is
-
•
orientable if its first Stiefel–Whitney class vanishes
-
•
and admits a spin structure if its first two Stiefel–Whitney classes, and , both vanish.
A spinc-structure is intermediate between an orientation and a spin structure. Specifically, a compact smooth manifold admits a spinc structure if its first Stiefel–Whitney class vanishes, and its second Stiefel–Whitney class is a reduction of an integral class. That is, is in the image of the reduction-of-coefficients map . For these and many other relevant facts, consult Stong’s book [24].
The spinc-cobordism ring, written , is the ring of spinc-cobordism classes of compact smooth spinc-manifolds. The addition is given by disjoint union of manifolds, while the multiplication is Cartesian product. There are several reasons to care about spinc-cobordism: aside from its applications to mathematical physics, e.g. [7] and [28], spinc-cobordism is of particular interest because it is one of the complex-oriented cobordism theories, and consequently there exists a one-dimensional group law on which describes how the first Chern class in spinc-cobordism behaves on a tensor product of complex line bundles. See [1] or [12] for these classical ideas, whose consequences for complex cobordism (as in [20]) have been enormous, but whose consequences for spinc-cobordism have apparently never been fully explored111In future work, the authors hope to apply the results about the ring structure of the spinc-cobordism ring obtained in this paper to the problem of describing the formal group law on the spinc-cobordism ring in formal-group-law theoretic terms, similar to what Quillen did for complex and unoriented cobordism in [16],[17], what Baker–Morava did for -inverted symplectic bordism in [6], and what Buchstaber did for symplectic bordism [9]. It seems impossible to get much understanding of the formal group law of spinc-cobordism without first coming to some understanding of the structure of its coefficient ring , which is the goal of this paper..
Since spinc-cobordism is an example of a “-cobordism theory” in the sense of Thom, the general results of [25] ensure that there exists a spectrum such that . The homotopy type of is understood as follows.
- Away from :
- At :
-
In 1966, Anderson, Brown, and Peterson [2],[3] proved that splits -locally as a wedge of suspensions of the connective complex -theory spectrum and the mod Eilenberg-Mac Lane spectrum :
(1) where the coproduct (i.e., wedge sum) is taken over all partitions (i.e., unordered finite tuples of positive integers) , and denotes the sum of the entries of .
This purely additive understanding of , and consequently of -local , is not entirely satisfying. To see the problem, consider the following table, which we reproduce from Bahri–Gilkey [5]:
0 | 0 | 8 | 0 | 16 | 0 | 24 | 2 |
1 | 0 | 9 | 0 | 17 | 0 | 25 | 0 |
2 | 0 | 10 | 1 | 18 | 3 | 26 | 9 |
3 | 0 | 11 | 0 | 19 | 0 | 27 | 0 |
4 | 0 | 12 | 0 | 20 | 1 | 28 | 4 |
5 | 0 | 13 | 0 | 21 | 0 | 29 | 1 |
6 | 0 | 14 | 1 | 22 | 5 | 30 | 14 |
7 | 0 | 15 | 0 | 23 | 0 | 31 | 1 |
The -linear dimension of , as recorded in table 1, is equivalently the number of copies of in -local , and equivalently the -rank of the -torsion subgroup of . Hence this table is telling us about the -torsion in the spinc-cobordism ring. One has the sense that some deep pattern is present in the distribution of the -torsion, but whatever it is, it cannot be seen clearly from these -ranks, nor from the Poincaré series used to inductively compute them.
However, since is precisely the -torsion in , is not only a summand but also an ideal in . One wants to understand multiplicatively, i.e., one wants to be able to describe the ring structure on , including its -torsion elements. A reasonably clear description of as a ring would yield a far more illuminating understanding of than the inductive formula for its -rank in each degree, which is presently all we have.
Fifty years after the additive structure of was described by Anderson–Brown–Peterson, the problem of calculating the ring structure of remains open. The purpose of this paper is to make progress towards a solution to this problem, restricting to the -local case, which is the most difficult.
1.2. The mod spinc-cobordism ring in low degrees
A traditional notation for the unoriented bordism ring is . In the 1968 book [24, pg. 351], Stong asks:
Open question: Can one determine these images nicely as subrings of ?
By “these images,” Stong refers to the images of the natural maps and . Our approach to understanding the mod spinc-cobordism ring begins by answering Stong’s open question in a range of degrees. We use the Anderson–Brown–Peterson splitting [2], product structure in the Adams spectral sequences, and Thom’s determination of using symmetric polynomials [25] to develop a method for calculating the image of the map through degree , for any fixed choice of integer . Our method gives a presentation for through degree , since the map is injective. We carry out computer calculation using our method to obtain our first main theorem:
Theorem A (Theorem 3.4).
The subring of the mod spinc-cobordism ring generated by all homogeneous elements of degree is isomorphic to
where is the ideal generated by the relations:
-
•
for each ,
-
•
and and for , where each is a particular polynomial in the generators with . The polynomial is described explicitly preceding Theorem 3.3.
The degrees of the generators are as follows: is in degree , while and are each in degree .
With Theorem A in hand, the patterns in table 1 become completely clear: in each degree in this range, one can see why the -linear dimension of the -torsion subgroup of takes the particular value it takes, as follows. Since Anderson–Brown–Peterson proved that the -torsion coincides with the -torsion in , in degrees the -torsion in is simply the -linear combinations of the monomials in the generators such that at least one of the factors is -torsion, i.e., at least one of the factors is either a generator with , or a generator . Here is the same table as table 1, but augmented with an -linear basis in each degree, using the multiplicative structure from Theorem A. We start in degree since there is no nontrivial -torsion in below degree .
-linear basis for | ||
---|---|---|
10 | 1 | |
11,12,13 | 0 | |
14 | 1 | |
15, 16, 17 | 0 | |
18 | 3 | |
19 | 0 | |
20 | 1 | |
21 | 0 | |
22 | 5 | |
23 | 0 | |
24 | 2 | |
25 | 0 | |
26 | 9 | |
27 | 0 | |
28 | 4 | |
29 | 1 | |
30 | 14 | |
31 | 1 | |
32 | 8 | |
33 | 2 |
One can also read off the product structure on the -torsion in in degrees from this table, since it is given by multiplication of monomials along with the relations from Theorem A.
It is evident from Theorem A that, in degrees , has a subring generated by elements and by elements with odd and not one less than a power of , subject to the relations for all odd . We are able to show that this pattern extends into all degrees, and goes some way to describing the ideal of -torsion elements of in multiplicative terms:
Theorem B (Theorem 4.3).
Consider the spinc-cobordism ring as a graded algebra over the graded ring Let be the ideal of generated by all the elements with odd. Then embeds, as a non-unital graded -algebra, into the -torsion ideal of the spinc-cobordism ring.
Theorem B describes the multiplicative structure of some, but not all, of the -torsion in . For example, in degrees , it accounts for precisely those monomials in table 2 which are not divisible by the elements . In particular, the lowest-degree -torsion element of which is not described by Theorem 4.3 is .
1.3. Milnor’s -dimensional spin manifold.
We also calculate the image of the map through degree 31 in Proposition 3.2 via a similar method to the one used to calculate the image of the map in Theorem A. There is a noteworthy geometric consequence of Proposition 3.2. In the 1965 paper [14], Milnor asks this question:
Problem. Does there exist a spin manifold of dimension so that ?
Here is a certain symmetric polynomial, and are Pontryagin classes. The reason for Milnor’s question is that, in [14], Milnor proves that, for a compact smooth manifold of dimension , the following conditions are equivalent:
-
(1)
is unorientedly cobordant to a spin manifold.
-
(2)
The Stiefel–Whitney numbers of involving and are all zero.
-
(3)
is unorientedly cobordant to , with an orientable compact manifold.
Milnor points out that, if there exists a compact spin manifold whose Pontryagin number is odd, then these conditions would fail to be equivalent in dimension . Anderson–Brown–Peterson [2],[3] established that, as a consequence of their splitting of -local , there does indeed exist such a compact spin manifold . However, it seems that no explicit description of that -dimensional compact spin manifold has been given in the literature (or anywhere else, as far as we know).
In Theorem 3.6, we give an explicit formula for the unoriented bordism class of such a compact spin manifold , as a disjoint union of products of real projective spaces and squares of Dold manifolds. We refer the reader to the Theorem 3.6 for a statement of that formula, which is lengthy. The formula is obtained using our calculation of the image of the map and the manifold representatives calculated in Proposition 3.5.
1.4. Determination of the mod spinc-cobordism ring up to inseparable isogeny
Thom’s famous calculation [25] established that the unoriented bordism ring is isomorphic to a polynomial algebra over . A theorem of Stong [23, Proposition 14] shows that the spinc cobordism ring, reduced modulo torsion and then reduced modulo , is also isomorphic to a polynomial -algebra.
By contrast, the spinc cobordism ring cannot itself be isomorphic to a polynomial algebra, since by [2], it has -torsion but is not an -algebra, hence it has nontrivial zero divisors. Similarly, since the mod spinc-cobordism ring has nontrivial -torsion, it cannot be isomorphic to a polynomial -algebra.
It follows as a trivial consequence of Theorem A that the mod spinc-cobordism ring still cannot be a polynomial -algebra. One can, with a bit of calculation, deduce the same fact from the additive structure of -local , by verifying that the Poincaré series of the mod spinc-cobordism ring is not the Poincaré series of any polynomial algebra. This avoids the use of our multiplicative methods. The advantage of our multiplicative methods is that we are able to prove that is instead uniformly -isomorphic to a polynomial algebra.
As far as we know, the terms “-isomorphism” (perhaps better known as “inseparable isogeny”) and “uniform -isomorphism” originated with Quillen [19]:
Definition 1.1.
Given a prime , a homomorphism of -algebras is said to be an -isomorphism if
-
•
for each , some power is zero, and
-
•
for each element , some power of is in the image of .
The -isomorphism is said to be uniform if can be chosen independently of and .
The notion of -isomorphism is applied only to algebras over a field of positive characteristic, so we had better reduce modulo in order to apply this idea to the spinc-cobordism ring. We get a positive result:
Theorem C (Theorem 4.4).
The mod spinc-cobordism ring is uniformly -isomorphic to the graded -algebra
(2) |
with the Bott element in degree , with in degree , and with in degree .
Corollary D.
The mod spinc-cobordism ring is uniformly -isomorphic to a graded polynomial -algebra on
-
•
a generator in degree for all positive integers ,
-
•
and a generator in degree for all positive integers such that is not a power of .
An -isomorphism induces a homeomorphism on prime ideal spectra, so Theorem C yields a description of all prime ideals in the mod spinc cobordism ring. That is, we have
1.5. Conventions
-
•
Given a ring and symbols , we write for the free -module with basis .
-
•
We write for the Bott element in , and also for its corresponding element under the Anderson–Brown–Peterson splitting of -local .
1.6. Acknowledgements
The first author would like to thank Bob Bruner for many helpful conversations related to this work, and the Simons Foundation for providing the license for a copy of Magma [8] used in calculations. The first author was partially supported by the electronic Computational Homotopy Theory (eCHT) research community, funded by National Science Foundation Research Training Group in the Mathematical Sciences grant 2135884.
2. Preliminaries
In this section we present an extended review of some well-known facts about spin and spinc cobordism, including the relationships various cobordism spectra, their homotopy groups, homology and cohomology groups, including the Steenrod algebra action on cohomology and the Pontryagin product in homology. This background material is necessary in order to understand the proofs of the results in the rest of the paper. Readers confident in their knowledge of this background material can skip to section 3, where we begin proving new results.
2.1. Review of the cohomology of the spectra and
There is an exact sequence of Lie groups
that gives rise to the fiber sequence
(3) |
Using this fibration, Harada and Kono [11] computed the mod 2 cohomology of the space :
Theorem 2.1.
The triviality of the ideal in the cohomology of is a consequence of the first differential in the Serre spectral sequence associated to the fiber sequence (3). It is not practical to write down a presentation for the -algebra which is more explicit than (4), since the difficulty of calculating iterated Steenrod squares applied to the Stiefel–Whitney class grows rapidly as the number of Steenrod squares grows. For example, has 38 monomials when expressed as a polynomial in the Stiefel–Whitney classes.
There is also an exact sequence
which gives rise to the fiber sequence
Using this, Quillen calculated:
Theorem 2.2.
By the Thom isomorphism, we have that and as graded -vector spaces. Since
and are quotients of , the action of Steenrod squares on and is determined by the Wu formula and the Cartan formula. This, together with the formula for the action of Steenrod squares on the Thom class , determines the action of the Steenrod squares on the cohomology of the spinc-bordism spectrum .
2.2. Review of and symmetric polynomials in the Stiefel–Whitney classes.
The following definitions are classical (see e.g. chapter 1 of [13]):
Definition 2.3.
Let be a nonnegative integer.
-
•
Suppose is an unordered -tuple of nonnegative integers. The monomial symmetric polynomial associated to is the symmetric polynomial
which has the fewest nonzero monomial terms among all those which have as a monomial term.
-
•
Given a nonnegative integer , the th elementary symmetric polynomial is the symmetric polynomial
given by
The monomial symmetric polynomials form a -linear basis for the ring of symmetric polynomials. The set of all finite products of elementary symmetric polynomials also famously (by Newton) forms a -linear basis for the ring of symmetric polynomials. Consequently, for each , there exists a unique polynomial such that
For more details about the polynomials , see the material on the transition matrix and its inverse in section 1.6 of [13], particularly (6.7)(i).
See [24], particularly pages 71 and 96 and surrounding material, for a nice exposition of the following result, which dates back to Thom [25]: let be the set of unordered finite-length tuples of positive integers, each of which is not equal to for any integer . Such integers are called “non-dyadic,” and such partitions are called “non-dyadic partitions.” For each , write for the length of , and write for the sum of the elements of . Consider the polynomial
in the Stiefel–Whitney classes . Then (see page 96 of [24], or pages 301-302 of [26]) the set
is a homogeneous -linear basis for the graded free -module , where denotes the Thom class, and is the mod Steenrod algebra. Consequently, for each nonnegative integer , is the -linear dual of the -vector space with basis the set
(5) |
Given two tuples and , we have their concatenation
The coproduct on is then given by
Consequently, if we write for the basis of dual to the basis (5), then , and , with in degree . We will sometimes write as an abbreviation for Thom’s generator of .
2.3. Maps between bordism theories
The first stages of the Whitehead tower for the orthogonal group are:
While does not fit into this sequence via a connective cover, the map factors through . There is a commutative diagram whose rows and columns are fiber sequences:
On the level of spectra, we have maps
(6) |
The maps induced in homotopy give the maps of respective cobordism rings. We will specifically consider the images of and in .
By the Anderson–Brown–Peterson splitting of -local , the cohomology splits as a direct sum of suspensions of and of . Here we are using the standard notation for the mod Steenrod algebra, and for its quotient , where is the subalgebra of generated by and by . Hence the -line in the -primary Adams spectral sequence for ,
(7) | ||||
is a direct sum of suspensions of , with one summand for each summand in , and also with one summand for each summand in .
Anderson–Brown–Peterson prove in [3] that all differentials in the Adams spectral sequence (7) are zero. Consequently the -line is the reduction of modulo the ideal generated by and by the Bott element .
This means we can calculate simply by calculating, i.e., the -comodule primitives . The advantage of thinking in terms of comodule primitives is that the Adams spectral sequence respects ring structure: if we calculate the homology as a ring, then by simply restricting to the comodule primitives in , we have calculated .
The same remarks apply mutatis mutandis for the spin bordism spectrum , the oriented bordism spectrum or for the unoriented bordism spectrum in place of . The Anderson–Brown–Peterson splitting for is as a wedge of suspensions of , , and . The analogue of the Anderson–Brown–Peterson splitting for is Thom’s splitting of as a wedge of suspensions of , while splits as a wedge of suspensions of and ; as far as we know, the latter splitting was originally proven by Wall [26]. Since is a quotient -module of , which is in turn a quotient -module of , dualizing yields that is a subcomodule of , which is in turn a subcomodule of .
Hence our broad strategy for calculating and , and the natural maps , is to calculate the -comodule primitives in and in , regarding each as -subcomodule algebras of . The resulting information will describe as a subring of . Details of this strategy are given in the description of the computational method in the proof of Proposition 3.1.
The relationships between the spin, spinc, oriented, and unoriented cobordism rings and their homologies is summarized in the following diagram, in which hooked arrows represent one-to-one maps:
3. The bordism ring in low degrees
In the statement of Proposition 3.1, we use Thom’s presentation for the unoriented cobordism ring .
Proposition 3.1.
The image of the map agrees, in degrees , with subring of generated by the elements
where
(8) | ||||
Proof.
This proposition is proven using computer calculation. We will describe our method for calculating in degrees , for any fixed choice of . The first author wrote a Magma [8] program which implements this method, and we have made its source code available at https://github.com/hassan-abdallah/spinc_cobordism. Once the reader is convinced of the correctness of the method, the proof of this proposition consists of simply running the calculation through degree , either by using our software, or by writing their own software implementation of the method, if desired.
We freely use the relationship between the spinc-cobordism ring, the unoriented cobordism ring, and the mod cohomology of detailed in section 2. Let be a non-dyadic partition of a nonnegative integer . We want to know whether its corresponding element is in the image of the map . The element has a Hurewicz image, i.e., the image of under the Hurewicz map . In section 2.2, we described the dual element to the Hurewicz image of , using Thom’s basis for . The element can be written as times a polynomial in the Stiefel–Whitney classes by applying an appropriate transition matrix222We remark that the computation of this transition matrix is one of the most computationally expensive parts of this process, despite its being a simple combinatorial problem. It is in fact the inverse of a Kostka matrix [22].. Once is calculated, we see that is in the image if and only if, when reduced modulo and the relations in the -module , is an -module primitive in .
Consequently our method for calculating is merely a method for building up a basis for the -vector space of -module primitives through some fixed degree , in terms of non-dyadic partitions. We work one degree at a time, but via induction, assuming we have already completed the calculation at all lower degrees.
The induction begins at degree , where there is nothing to say: the empty partition yields the unique -module primitive in . For each integer , the product is simply modulo the relations in , by the classical formula for the action of Steenrod squares on the Thom class in . Record the elements in an unordered list . Here the symbol stands for “decomposable,” as we will use it to build up a list of -module decomposables in in degrees .
We are not done with the initial step in the induction: for each nonzero element , we calculate using the Thom formula and the Wu formula ([27], but see [15, pg. 94] for a textbook reference) for the action of on Stiefel–Whitney classes, and we include the results in . Now contains an -linear basis for the -submodule of generated by the Thom class , in all degrees .
Now we are ready for the inductive step:
- Inductive hypothesis at the th step:
-
We have produced a list of -linear combinations of non-dyadic partitions of degree , such that the -linear span of is a basis for the set of -module primitives in in all degrees . We have also produced a list of -linear combinations of non-dyadic partitions of degree , such that the -linear span of is precisely the -submodule of generated by in degrees .
- Calculation for the th inductive step:
-
Write for the -linear span of the degree elements in . Calculate an -linear basis for. Let be . Use the calculated transition matrix to convert the members of from the Thom/partition basis to the Stiefel–Whitney monomial basis, and then use the Thom formula and the Wu formula to calculate all Steenrod squares on the members of , then all Steenrod squares on those, etc., in degrees . Use the transition matrix to convert back to the partition basis, and for the resulting list of linear combinations of non-dyadic partitions. Now we are ready to iterate, with in place of , and with in place of .
Once we complete the step, we have an -linear basis for the image of the map in all degrees , expressed in terms of Thom’s partition basis for . Consequently we have a description of , in all degrees , as a subring of . ∎
In principle there is no obstruction to using the same method to make calculations of products in in degrees . We stopped at degree simply because, around the time we completed degree , we could see enough of the ring structure of to prove that the mod spinc-cobordism ring is not a polynomial algebra, and to suggest the right statements for Proposition 4.2 and Theorem C. The products in required for the proof of Theorem 3.6 are known as soon as one computes the ring structure through degree .
The same computational method described in the proof of Proposition 3.1, applied to rather than , yields:
Proposition 3.2.
The image of the map agrees, in degrees , with subring of generated by the elements
Each of the elements defined in Proposition 3.2 is a linear combination of monomials in . Those monomials are generally not individually members of : for example, does not lift to an element of , even though a linear combination of with other monomials in degree does lift to the element .
However, lifts to the element , and consequently the squares of each of the monomials in each of the elements lift to . For , and , let denote the element of obtained by taking the definition of in Proposition 3.1 and replacing each instance of with . For example, (8) yields that
Then, as a consequence of Proposition 3.1, we have:
Theorem 3.3.
The subring of generated by all homogeneous elements of degree is isomorphic to:
where is the ideal generated by and .
The relations , with indecomposable in and with a polynomial in the indecomposable elements , immediately implies that is not a polynomial algebra.
Theorem 3.4.
The subring of the mod spinc-cobordism ring generated by all homogeneous elements of degree is isomorphic to:
where is the ideal generated by
-
•
for each ,
-
•
and and for .
Proof.
Let denote the ideal of consisting of -torsion elements, and let denote the kernel of the ring map
The ring was calculated by Stong [23, Proposition 11]: it is a polynomial -algebra on generators in degrees . Since is an ideal in , to calculate the product in the ring , it suffices to calculate
-
•
the products between generators of ,
-
•
and the products between generators of and lifts, to of generators of .
Both of these types of products land in . Since maps injectively under the map , we can embed into and bring to bear our calculations of the image of this map, from Proposition 3.1. All we need to do is to determine, in our set of generators for through degree , a maximal set of linear combinations of products of generators which generate -torsion elements in , i.e., copies of rather than in the Anderson–Brown–Peterson splitting.
As a consequence of the Anderson–Brown–Peterson splitting, generators of that are 2-torsion, and hence -torsion, in are those whose corresponding -module primitive in is not or torsion. We identify such generators by re-running the entire process from the proof of Proposition 3.1, but with the following modification: at the start of the calculation, before the induction on degree, we begin by letting be a list of all Stiefel–Whitney monomials in in degrees which are -torsion, together with all words in the Steenrod squares applied to such -torsion Stiefel–Whitney monomials333In principle, this step in the calculation could go wrong, failing to identify all the -torsion in , as follows: suppose there is some -linear combination of Stiefel–Whitney monomials in which is -torsion, but none of its summands are -torsion. If this occurs, it would not be noticed by the method we describe, since our method only checks the -torsion status of Stiefel–Whitney monomials. We handle this by a very simple idea: we make the calculation as described, and after making the full calculation, we “check our answer” by comparing to the known additive structure of , as follows. After running our method, we compare the rank of our calculated -torsion in each degree to the expected rank, using the known Poincaré series for the -torsion in . If our method has failed to notice a linear combination of Stiefel–Whitney monomials which was -torsion despite its summands not being -torsion, then the rank of the -torsion from our calculation will be too small. We have never observed this mismatched rank to happen, i.e., it does not happen through degree in . If the rank mismatch were to ever occur, the fix is conceptually trivial, but computationally very hard: instead of populating with all the homogeneous -torsion Stiefel–Whitney monomials at the start of the torsion calculation, we simply populate with the all the -torsion Stiefel–Whitney polynomials at the start of the calculation, then re-run the calculation. This of course cannot miss any -torsion in ! Its disadvantage is simply that it is extremely computationally expensive, since the total number of homogeneous Stiefel–Whitney polynomials (not just monomials) in degrees in grows extremely quickly as grows. We carry out the calculations in the way we describe—i.e., initially populating by only the -torsion Stiefel–Whitney monomials, rather than polynomials—to dramatically speed up calculation, and because we are able to check that the resulting answer in the end agrees with the answer we would have gotten with the much slower calculation using all the homogeneous Stiefel–Whitney polynomials., instead of letting begin as the empty set. Consequently, as we proceed through the induction, is not only the set of -module decomposables, but also the set of Stiefel–Whitney monomials which generate copies of .
Re-running our inductive calculation from Proposition 3.1, but with this initial list for , yields a set of -module generators for modulo -torsion. Comparison of the lists produced by the first calculation and the second calculation, then using the translation matrix to translate back from the basis of Stiefel–Whitney monomials to the dual basis of partitions (i.e., Thom’s basis for ), gives us a set of generators for in degrees , and tells us, for each generator, whether it corresponds to a copy of or of under the Anderson–Brown–Peterson splitting.
In degrees , we find that an element which is in the image of the map is -torsion as long as the partition includes an odd number. As described in section 1.1, the -torsion elements of are exactly the -torsion elements. This yields the presentation for in degrees in the statement of the theorem. ∎
Our next result determines explicit manifolds that represent some of the bordism classes whose powers occurs as ring-theoretic generators of and of in Proposition 3.1 and in Proposition 3.2, respectively. The following table was calculated through degree by Thom [25]. We extend the calculation through degree . The symbol denotes the -dimensional Dold manifold, defined in [10].
Proposition 3.5 (Thom [25]).
Manifold representatives for elements in Thom’s partition basis for are as follows:
Element | Manifold Representative |
---|---|
Proof.
Routine calculation using Stiefel–Whitney numbers. ∎
In [14], Milnor investigates whether every spin manifold is unorientedly cobordant to the square of an orientable manifold. He shows it is true for spin manifolds of dimension . The ambiguity in dimension stems from the existence of an orientable manifold whose only nonzero Stiefel–Whitney numbers are , , , , and . Milnor then poses the problem of whether a spin manifold of dimension exists with these nonzero Stiefel–Whitney numbers. Anderson–Brown–Peterson stated two years later [3] that, as a corollary of their main theorem, the lowest dimension in which there exists an element of which is not the square of an orientable manifold is [3]. In Proposition 3.2 we calculated in degrees through in terms of Thom’s partition basis, and in proposition 3.5 we have manifold representatives for ring-theoretic generators which suffice to generate everything in . Solving for an element of with Milnor’s prescribed Stiefel–Whitney numbers, we find an explicit manifold of the kind Milnor asked for:
Theorem 3.6.
The cobordism class has nonzero Stiefel–Whitney numbers , , , , and , and is represented by the manifold:
4. The spinc-cobordism ring in all degrees.
4.1. A nonunital subring of the -torsion in the spinc-cobordism ring.
In Theorem 3.4, we showed that, in degrees , the cobordism classes and lift to indecomposable -torsion elements and in . The elements are precisely those of the form with odd and non-dyadic. It is natural to ask whether this pattern extends above degree as well. In Proposition 4.2, we show that something of this kind is true, if we use the Dold manifold , from [10], rather than the cobordism class from Thom’s partition basis for . In low degrees, the algebraic relation between the Dold manifolds and the Thom generators for is as follows:
Proposition 4.1.
The squares of odd-dimensional Dold manifolds represent the following polynomials in Thom’s generators for :
Element | Manifold Representative |
---|---|
Furthermore, each of these elements of is in the image of the map .
Proof.
The manifold representatives are straightforwardly calculated from Proposition 3.5. By Proposition 3.2, these elements are in the image of the map . ∎
Proposition 4.2.
For each odd integer such that is not a power of , the Dold manifold has the property that lifts to an indecomposable -torsion element of . It furthermore lifts to an indecomposable -torsion element of .
Proof.
Dold [10] proves that there exists a minimal set of generators for the cobordism ring whose odd-degree elements are , where and is odd. Milnor proved that the map maps onto all squares of elements in . This map factors through , so lifts to for all non-dyadic . The mod cohomology of the Dold manifold is given as a graded ring by
with and . The total Stiefel–Whitney class of is
Setting and , the total Stiefel–Whitney class of as in the statement of the theorem is then given by:
and in particular, . Hence is orientable.
To show that lifts to the spin cobordism ring, one can carry out an algebraic calculation to show that none of the nonzero Stiefel–Whitney numbers of are divisible by . This is not a difficult calculation, but it is simpler to invoke the main result of Anderson’s paper [4]: the square of any orientable compact manifold is unorientedly cobordant to a spin manifold. Since vanishes on , its square must be in the image of the map .
It follows easily from the structure of , and the Anderson–Brown–Peterson splitting of MSpin into a wedge of suspensions of , , and , that all elements of in degrees are -torsion. Hence, for odd , any spin-cobordism class that lifts must be -torsion.
Let be a lift of to . The image of under the map is then a lift of to , and it is -torsion since it is the image of a -torsion element. It is furthermore -torsion in , since by the Anderson–Brown–Peterson splitting of -local , every -torsion element of is also -torsion.
To see that lifts to an indecomposable element of , it is enough to observe that embeds into , and since has second Stiefel–Whitney class , does not lift to . Hence the unique lift of to is indecomposable, hence any lift of to is indecomposable. A completely analogous argument establishes that any lift of to is also indecomposable. ∎
Since Dold [10] showed that can be written as Thom’s generator plus decomposables in the same degree, Proposition 4.2 tells us that some of the patterns exhibited in table (2) are not limited to degrees , and indeed extend to all degrees. Namely, for odd non-dyadic , if we write for a lift of to an indecomposable -torsion element of (guaranteed to exist by Proposition 4.2), and for even we write for a lift of to an element of , then we have:
Theorem 4.3.
Consider the spinc-cobordism ring as a graded algebra over the graded ring Then the ideal of embeds, as a non-unital graded -algebra, into the -torsion ideal of the spinc-cobordism ring.
4.2. Mod spinc-cobordism, up to uniform -isomorphism.
In this section, we show that the torsion-free generators , together with the large nonunital subring of the -torsion in , constructed in Theorem 4.3, suffice to generate a subring of the mod spinc-cobordism ring which, although smaller than , is nevertheless uniformly -isomorphic to all of . See Definition 1.1 for the definition of a uniform -isomorphism.
Theorem 4.4.
The mod spinc-cobordism ring is uniformly -isomorphic to the graded -algebra
(9) |
with the Bott element in degree , with in degree , and with in degree .
Proof.
Write for the ideal of consisting of the -torsion elements. Stong [23, Proposition 14] proved that is a polynomial -algebra on generators and . Since generated by , Stong’s generator agrees modulo with the Bott element . Hence is isomorphic as a graded -algebra to .
Now let denote the graded subring of generated by , by for all , and by the mod reductions of the -torsion elements in from Theorem 4.2 which lift for all odd non-dyadic . Since contains all the squares of elements in , the graded -algebra map
is a uniform -isomorphism.
Let denote the kernel of the ring map . Filter by powers of the ideal , i.e., equip with the -adic filtration. By the Anderson–Brown–Peterson splitting and by Theorem 4.2, the associated graded ring is isomorphic to tensored with the image of in and reduced modulo the relations for all , i.e., is isomorphic to (9).
We claim that itself is isomorphic to (9). The -adic filtration on is additively split, so as graded -vector spaces. In principle, the ring structure on could differ from the ring structure on if the multiplication on were to exhibit -adic filtration jumps, i.e., when we multiply two elements of , with of -adic filtration and with of -adic filtration , we could perhaps get an element of -adic filtration .
However, even if filtration jumps occur, is still isomorphic to the graded -algebra with presentation (9). This is by a freeness argument similar to the classical argument that, if the associated graded of a filtered commutative -algebra is a polynomial (i.e., free commutative) -algebra, then the original filtered commutative -algebra must also have been free commutative. The argument is as follows. Let be the category of pairs , where is a graded-commutative -algebra, and is a set of homogeneous elements of such that for all . There is a forgetful functor from to the category of pairs in which are sets and . The forgetful functor sends to the underlying sets of and of . The graded -algebra (9) is the free object of on the pair
By Proposition 4.2, the elements are -torsion in , not merely in . Hence there are no relations on except those which make it an object of the category of , and lives in as well, i.e., and are isomorphic in the category . Hence and are isomorphic as graded -algebras, and hence is uniformly -isomorphic to the -algebra (9), as claimed. ∎
In section 3, we showed that is not isomorphic to a polynomial algebra. Nevertheless, since an -isomorphism induces a homeomorphism on the prime spectra (see [19, Proposition B.8] or [21, Lemma 29.46.9]), we have:
Corollary 4.5.
The topological space is homeomorphic to of a polynomial -algebra on countably infinity many generators.
Furthermore, the topological space is homeomorphic to of the ring (9).
The last two sentences in Stong’s 1968 book [24] before the appendices begin are:
One may relate the pair through exact sequences in precisely the same way as are related (or as are related). Computationally this is not of much use since one has no way to nicely describe the torsion in .
We regard Theorems 4.3 and 4.4 as progress toward nicely describing the torsion in by means of ring structure.
References
- [1] J. F. Adams. Stable homotopy and generalised homology. Chicago Lectures in Mathematics. University of Chicago Press, Chicago, IL, 1995. Reprint of the 1974 original.
- [2] D. W. Anderson, E. H. Brown, Jr., and F. P. Peterson. Spin cobordism. Bull. Amer. Math. Soc., 72:256–260, 1966.
- [3] D. W. Anderson, E. H. Brown, Jr., and F. P. Peterson. The structure of the Spin cobordism ring. Ann. of Math. (2), 86:271–298, 1967.
- [4] Peter G. Anderson. Cobordism classes of squares of orientable manifolds. Bull. Amer. Math. Soc., 70:818–819, 1964.
- [5] Anthony Bahri and Peter Gilkey. The eta invariant, bordism, and equivariant bordism for cyclic -groups. Pacific J. Math., 128(1):1–24, 1987.
- [6] Andrew Baker and Jack Morava. localized away from and odd formal group laws. arXiv preprint 1403.2596, 2014.
- [7] Ralph Blumenhagen, Niccolò Cribiori, Christian Kneißl, and Andriana Makridou. Dimensional reduction of cobordism and K-theory. Journal of High Energy Physics, 2023(3), 2023.
- [8] Wieb Bosma, John Cannon, and Catherine Playoust. The Magma algebra system. I. The user language. J. Symbolic Comput., 24(3-4):235–265, 1997. Computational algebra and number theory (London, 1993).
- [9] V. M. Buchstaber and S. P. Novikov. Formal groups, power systems and Adams operators. Mat. Sb. (N.S.), 84(126):81–118, 1971.
- [10] Albrecht Dold. Erzeugende der Thomschen Algebra . Math. Z., 65:25–35, 1956.
- [11] Masana Harada and Akira Kono. Cohomology mod 2 of the classifying space of Spinc(n). Publ. Res. Inst. Math. Sci., 22(3):543–549, sep 1986.
- [12] Mike Hopkins. Complex oriented cohomology theories and the language of stacks. unpublished course notes, available on the Web, 1999.
- [13] I. G. Macdonald. Symmetric functions and Hall polynomials. Oxford Classic Texts in the Physical Sciences. The Clarendon Press, Oxford University Press, New York, second edition, 2015. With contribution by A. V. Zelevinsky and a foreword by Richard Stanley, Reprint of the 2008 paperback edition.
- [14] J. Milnor. On the Stiefel-Whitney numbers of complex manifolds and of spin manifolds. Topology, 3:223–230, 1965.
- [15] John W. Milnor and James D. Stasheff. Characteristic classes, volume No. 76 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ; University of Tokyo Press, Tokyo, 1974.
- [16] Daniel Quillen. On the formal group laws of unoriented and complex cobordism theory. Bull. Amer. Math. Soc., 75:1293–1298, 1969.
- [17] Daniel Quillen. Elementary proofs of some results of cobordism theory using Steenrod operations. Advances in Math., 7:29–56, 1971.
- [18] Daniel Quillen. The cohomology rings of extra-special -groups and the spinor groups. Math. Ann., 194:197–212, 1971.
- [19] Daniel Quillen. The spectrum of an equivariant cohomology ring. I, II. Ann. of Math. (2), 94:549–572; ibid. (2) 94 (1971), 573–602, 1971.
- [20] Douglas C. Ravenel. Complex cobordism and stable homotopy groups of spheres, volume 121 of Pure and Applied Mathematics. Academic Press Inc., Orlando, FL, 1986.
- [21] The Stacks Project Authors. Stacks Project. https://stacks.math.columbia.edu, 2024.
- [22] Richard P. Stanley. Enumerative combinatorics. Volume 1, volume 49 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, second edition, 2012.
- [23] R. E. Stong. Relations among characteristic numbers. II. Topology, 5:133–148, 1966.
- [24] Robert E. Stong. Notes on cobordism theory. Princeton University Press, Princeton, NJ; University of Tokyo Press, Tokyo, 1968. Mathematical notes.
- [25] René Thom. Quelques propriétés globales des variétés différentiables. Comment. Math. Helv., 28:17–86, 1954.
- [26] C. T. C. Wall. Determination of the cobordism ring. Ann. of Math. (2), 72:292–311, 1960.
- [27] Wen-tsün Wu. Les -carrés dans une variété grassmannienne. C. R. Acad. Sci. Paris, 230:918–920, 1950.
- [28] Ümit Ertem. Weyl semimetals and spinc cobordism. arXiv preprint 2003.04082, 2020.