This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Products in spinc-cobordism

Hassan Abdallah  and  Andrew Salch Department of Mathematics, Wayne State University, Detroit, MI, USA [email protected], [email protected]
Abstract.

We calculate the mod 22 spinc-cobordism ring up to uniform FF-isomorphism (i.e., inseparable isogeny). As a consequence we get the prime ideal spectrum of the mod 22 spinc-cobordism ring. We also calculate the mod 22 spinc-cobordism ring “on the nose” in degrees 33\leq 33. We construct an infinitely generated nonunital subring of the 22-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 2424-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 nn-dimensional manifold MM is a reduction of its structure group from O(n)O(n) to Spinc(n)Spin^{c}(n). We find the following perspective illuminating: a compact smooth manifold is

  • orientable if its first Stiefel–Whitney class w1w_{1} vanishes

  • and admits a spin structure if its first two Stiefel–Whitney classes, w1w_{1} and w2w_{2}, both vanish.

A spinc-structure is intermediate between an orientation and a spin structure. Specifically, a compact smooth manifold MM admits a spinc structure if its first Stiefel–Whitney class w1w_{1} vanishes, and its second Stiefel–Whitney class w2w_{2} is a reduction of an integral class. That is, w2H2(M;𝔽2)w_{2}\in H^{2}(M;\mathbb{F}_{2}) is in the image of the reduction-of-coefficients map H2(M;)H2(M;𝔽2)H^{2}(M;\mathbb{Z})\rightarrow H^{2}(M;\mathbb{F}_{2}). For these and many other relevant facts, consult Stong’s book [24].

The spinc-cobordism ring, written ΩSpinc\Omega^{Spin^{c}}_{*}, 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 ΩSpinc\Omega^{Spin^{c}}_{*} 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 22-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 ΩSpinc\Omega^{Spin^{c}}_{*}, which is the goal of this paper..

Since spinc-cobordism is an example of a “(B,f)(B,f)-cobordism theory” in the sense of Thom, the general results of [25] ensure that there exists a spectrum MSpincMSpin^{c} such that π(MSpinc)ΩSpinc\pi_{*}(MSpin^{c})\cong\Omega^{Spin^{c}}_{*}. The homotopy type of MSpincMSpin^{c} is understood as follows.

Away from 22:

The map π:BSpincBSO×K(,2)\pi:BSpin^{c}\longrightarrow BSO\times K(\mathbb{Z},2) is an odd-primary homotopy equivalence and induces an isomorphism ΩSpinc[12]ΩSO(K(,2))[12]\Omega^{Spin^{c}}_{*}[\frac{1}{2}]\cong\Omega^{SO}_{*}(K(\mathbb{Z},2))[\frac{1}{2}], and consequently MSpinc[12]MSO[12]PMSpin^{c}[\frac{1}{2}]\cong MSO[\frac{1}{2}]\wedge\mathbb{C}P^{\infty}. Since MSpincMSpin^{c} is complex-oriented, we have an isomorphism of graded rings π(MSpinc[12])π(MSO)[12][[x]]\pi_{*}(MSpin^{c}[\frac{1}{2}])\cong\pi_{*}(MSO)[\frac{1}{2}][[x]] with xπ2(MSpinc)x\in\pi_{2}(MSpin^{c}). The ring π(MSO)[12]π(MSp)[12]π(MSpin)[12]\pi_{*}(MSO)[\frac{1}{2}]\cong\pi_{*}(MSp)[\frac{1}{2}]\cong\pi_{*}(MSpin)[\frac{1}{2}] is understandable in various ways; see Wall [26] or Baker–Morava [6] for different approaches.

At 22:

In 1966, Anderson, Brown, and Peterson [2],[3] proved that MSpincMSpin^{c} splits 22-locally as a wedge of suspensions of the connective complex KK-theory spectrum kuku and the mod 22 Eilenberg-Mac Lane spectrum H𝔽2H\mathbb{F}_{2}:

(1) MSpin(2)c\displaystyle MSpin^{c}_{(2)} ZJΣ4|J|ku(2)\displaystyle\simeq Z\vee\coprod_{J}\Sigma^{4\left|J\right|}ku_{(2)}

where the coproduct (i.e., wedge sum) is taken over all partitions (i.e., unordered finite tuples of positive integers) JJ, and |J|\left|J\right| denotes the sum of the entries of JJ.

Not much is known about the summand ZZ in (1), other than that

  • •:

    it is a coproduct of suspensions of copies of H𝔽2H\mathbb{F}_{2},

  • •:

    and from a Poincaré series [2], it is known how to solve inductively for the number of copies of ΣnH𝔽2\Sigma^{n}H\mathbb{F}_{2} in ZZ, for each nn.

In that sense, ZZ is understood additively.

This purely additive understanding of ZZ, and consequently of 22-local ΩSpinc\Omega^{Spin^{c}}_{*}, is not entirely satisfying. To see the problem, consider the following table, which we reproduce from Bahri–Gilkey [5]:

nn dim𝔽2πnZ\dim_{\mathbb{F}_{2}}\pi_{n}Z nn dim𝔽2πnZ\dim_{\mathbb{F}_{2}}\pi_{n}Z nn dim𝔽2πnZ\dim_{\mathbb{F}_{2}}\pi_{n}Z nn dim𝔽2πnZ\dim_{\mathbb{F}_{2}}\pi_{n}Z
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
Table 1.

The 𝔽2\mathbb{F}_{2}-linear dimension of πnZ\pi_{n}Z, as recorded in table 1, is equivalently the number of copies of ΣnH𝔽2\Sigma^{n}H\mathbb{F}_{2} in 22-local MSpincMSpin^{c}, and equivalently the 𝔽2\mathbb{F}_{2}-rank of the 22-torsion subgroup of ΩnSpinc\Omega^{Spin^{c}}_{n}. Hence this table is telling us about the 22-torsion in the spinc-cobordism ring. One has the sense that some deep pattern is present in the distribution of the 22-torsion, but whatever it is, it cannot be seen clearly from these 𝔽2\mathbb{F}_{2}-ranks, nor from the Poincaré series used to inductively compute them.

However, since π(Z)\pi_{*}(Z) is precisely the 22-torsion in ΩSpinc\Omega^{Spin^{c}}_{*}, π(Z)\pi_{*}(Z) is not only a summand but also an ideal in ΩSpinc\Omega^{Spin^{c}}_{*}. One wants to understand π(Z)\pi_{*}(Z) multiplicatively, i.e., one wants to be able to describe the ring structure on ΩSpinc\Omega^{Spin^{c}}_{*}, including its 22-torsion elements. A reasonably clear description of ΩSpinc\Omega^{Spin^{c}}_{*} as a ring would yield a far more illuminating understanding of π(Z)\pi_{*}(Z) than the inductive formula for its 𝔽2\mathbb{F}_{2}-rank in each degree, which is presently all we have.

Fifty years after the additive structure of MSpincMSpin^{c} was described by Anderson–Brown–Peterson, the problem of calculating the ring structure of ΩSpinc\Omega^{Spin^{c}}_{*} remains open. The purpose of this paper is to make progress towards a solution to this problem, restricting to the 22-local case, which is the most difficult.

1.2. The mod 22 spinc-cobordism ring in low degrees

A traditional notation for the unoriented bordism ring ΩOMO\Omega^{O}_{*}\cong MO_{*} is 𝔑\mathfrak{N}_{*}. In the 1968 book [24, pg. 351], Stong asks:

Open question: Can one determine these images nicely as subrings of 𝔑\mathfrak{N}_{*}?

By “these images,” Stong refers to the images of the natural maps ΩSpin𝔑\Omega^{Spin}_{*}\rightarrow\mathfrak{N}_{*} and ΩSpinc𝔑\Omega^{Spin^{c}}_{*}\rightarrow\mathfrak{N}_{*}. Our approach to understanding the mod 22 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 𝔑\mathfrak{N}_{*} using symmetric polynomials [25] to develop a method for calculating the image of the map ΩSpinc𝔑\Omega^{Spin^{c}}_{*}\rightarrow\mathfrak{N}_{*} through degree dd, for any fixed choice of integer dd. Our method gives a presentation for ΩSpinc/(2,β)\Omega^{Spin^{c}}/(2,\beta) through degree dd, since the map ΩSpinc/(2,β)𝔑\Omega^{Spin^{c}}/(2,\beta)\longrightarrow\mathfrak{N}_{*} 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 22 spinc-cobordism ring ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} generated by all homogeneous elements of degree 33\leq 33 is isomorphic to

𝔽2[β,Z4,Z8,Z10,Z12,Z16,Z18,Z20,Z22,Z24,Z26,Z28,Z32,\displaystyle\mathbb{F}_{2}[\beta,Z_{4},Z_{8},Z_{10},Z_{12},Z_{16},Z_{18},Z_{20},Z_{22},Z_{24},Z_{26},Z_{28},Z_{32},
T24,T29,T31,T32,T33]/I\displaystyle T_{24},T_{29},T_{31},T_{32},T_{33}]/I

where II is the ideal generated by the relations:

  • βZi=0\beta Z_{i}=0 for each i2mod4i\equiv 2\mod 4,

  • and βTi=0\ \beta T_{i}=0 and Ti2=U2iT_{i}^{2}=U_{2i} for i{24,29,31,32,33}i\in\{24,29,31,32,33\}, where each UiU_{i} is a particular polynomial in the generators ZjZ_{j} with ji20j\leq i-20. The polynomial UiU_{i} is described explicitly preceding Theorem 3.3.

The degrees of the generators are as follows: β=[P1]\beta=[\mathbb{C}P^{1}] is in degree 22, while ZiZ_{i} and TiT_{i} are each in degree ii.

With Theorem A in hand, the patterns in table 1 become completely clear: in each degree in this range, one can see why the 𝔽2\mathbb{F}_{2}-linear dimension of the 22-torsion subgroup of ΩSpinc\Omega^{Spin^{c}}_{*} takes the particular value it takes, as follows. Since Anderson–Brown–Peterson proved that the 22-torsion coincides with the β\beta-torsion in ΩSpinc\Omega^{Spin^{c}}_{*}, in degrees n33n\leq 33 the 22-torsion in ΩnSpinc\Omega^{Spin^{c}}_{n} is simply the 𝔽2\mathbb{F}_{2}-linear combinations of the monomials in the generators Zi,TiZ_{i},T_{i} such that at least one of the factors is β\beta-torsion, i.e., at least one of the factors is either a generator ZiZ_{i} with i2mod4i\equiv 2\mod 4, or a generator TiT_{i}. Here is the same table as table 1, but augmented with an 𝔽2\mathbb{F}_{2}-linear basis in each degree, using the multiplicative structure from Theorem A. We start in degree 1010 since there is no nontrivial 22-torsion in ΩSpinc\Omega^{Spin^{c}}_{*} below degree 1010.

nn dim𝔽2πnZ\dim_{\mathbb{F}_{2}}\pi_{n}Z 𝔽2\mathbb{F}_{2}-linear basis for πnZ\pi_{n}Z
10 1 Z10Z_{10}
11,12,13 0
14 1 Z4Z10Z_{4}Z_{10}
15, 16, 17 0
18 3 Z42Z10,Z8Z10,Z18Z_{4}^{2}Z_{10},Z_{8}Z_{10},Z_{18}
19 0
20 1 Z102Z_{10}^{2}
21 0
22 5 Z43Z10,Z4Z8Z10,Z12Z10,Z4Z18,Z22Z_{4}^{3}Z_{10},Z_{4}Z_{8}Z_{10},Z_{12}Z_{10},Z_{4}Z_{18},Z_{22}
23 0
24 2 Z4Z102,T24Z_{4}Z_{10}^{2},T_{24}
25 0
26 9 Z44Z10,Z42Z8Z10,Z82Z10,Z4Z12Z10,Z16Z10,Z_{4}^{4}Z_{10},Z_{4}^{2}Z_{8}Z_{10},Z_{8}^{2}Z_{10},Z_{4}Z_{12}Z_{10},Z_{16}Z_{10},
Z42Z18,Z8Z18,Z4Z22,Z26Z_{4}^{2}Z_{18},Z_{8}Z_{18},Z_{4}Z_{22},Z_{26}
27 0
28 4 Z42Z102,Z8Z102,Z10Z18,Z4T24Z_{4}^{2}Z_{10}^{2},Z_{8}Z_{10}^{2},Z_{10}Z_{18},Z_{4}T_{24}
29 1 T29T_{29}
30 14 Z45Z10,Z43Z8Z10,Z4Z82Z10,Z42Z12Z10,Z8Z12Z10,Z_{4}^{5}Z_{10},Z_{4}^{3}Z_{8}Z_{10},Z_{4}Z_{8}^{2}Z_{10},Z_{4}^{2}Z_{12}Z_{10},Z_{8}Z_{12}Z_{10},
Z4Z16Z10,Z20Z10,Z43Z18,Z4Z8Z18,Z12Z18,Z_{4}Z_{16}Z_{10},Z_{20}Z_{10},Z_{4}^{3}Z_{18},Z_{4}Z_{8}Z_{18},Z_{12}Z_{18},
Z42Z22,Z8Z22,Z4Z26,Z103Z_{4}^{2}Z_{22},Z_{8}Z_{22},Z_{4}Z_{26},Z_{10}^{3}
31 1 T31T_{31}
32 8 Z43Z102,Z4Z8Z102,Z12Z102,Z4Z10Z18,Z10Z22,Z_{4}^{3}Z_{10}^{2},Z_{4}Z_{8}Z_{10}^{2},Z_{12}Z_{10}^{2},Z_{4}Z_{10}Z_{18},Z_{10}Z_{22},
Z42T24,Z8T24,T32Z_{4}^{2}T_{24},Z_{8}T_{24},T_{32}
33 2 Z4T29,T33Z_{4}T_{29},T_{33}
Table 2.

One can also read off the product structure on the 22-torsion in ΩSpinc\Omega^{Spin^{c}}_{*} in degrees 33\leq 33 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 33\leq 33, ΩSpinc\Omega^{Spin^{c}}_{*} has a subring generated by elements Z4,Z8,Z12,Z16,Z_{4},Z_{8},Z_{12},Z_{16},\dots and by elements Z2iZ_{2i} with ii odd and not one less than a power of 22, subject to the relations 2Z2i=0=βZ2i2Z_{2i}=0=\beta Z_{2i} for all odd ii. We are able to show that this pattern extends into all degrees, and goes some way to describing the ideal π(Z)\pi_{*}(Z) of 22-torsion elements of ΩSpinc\Omega^{Spin^{c}}_{*} in multiplicative terms:

Theorem B (Theorem 4.3).

Consider the spinc-cobordism ring as a graded algebra over the graded ring S:=(2)[β,Z2j:j+1 not a power of 2]/(βZ2j,2Z2j for odd j).S:=\mathbb{Z}_{(2)}[\beta,Z_{2j}:j+1\mbox{ not\ a\ power\ of\ }2]/(\beta Z_{2j},2Z_{2j}\mbox{\ for\ odd\ }j). Let JJ be the ideal of SS generated by all the elements Z2jZ_{2j} with jj odd. Then JJ embeds, as a non-unital graded SS-algebra, into the 22-torsion ideal π(Z)\pi_{*}(Z) of the spinc-cobordism ring.

Theorem B describes the multiplicative structure of some, but not all, of the 22-torsion in ΩSpinc\Omega^{Spin^{c}}_{*}. For example, in degrees 33\leq 33, it accounts for precisely those monomials in table 2 which are not divisible by the elements TiT_{i}. In particular, the lowest-degree 22-torsion element of ΩSpinc\Omega^{Spin^{c}}_{*} which is not described by Theorem 4.3 is T24Ω24SpincT_{24}\in\Omega^{Spin^{c}}_{24}.

1.3. Milnor’s 2424-dimensional spin manifold.

We also calculate the image of the map ΩSpin𝔑\Omega^{Spin}_{*}\rightarrow\mathfrak{N}_{*} through degree 31 in Proposition 3.2 via a similar method to the one used to calculate the image of the map ΩSpinc𝔑\Omega^{Spin^{c}}_{*}\rightarrow\mathfrak{N}_{*} 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 Σ\Sigma of dimension 2424 so that s6(p1,,p6)[Σ]1mod2s_{6}(p_{1},\dots,p_{6})[\Sigma]\equiv 1\mod 2?

Here s6s_{6} is a certain symmetric polynomial, and p1,,p6p_{1},\dots,p_{6} are Pontryagin classes. The reason for Milnor’s question is that, in [14], Milnor proves that, for a compact smooth manifold MM of dimension 23\leq 23, the following conditions are equivalent:

  1. (1)

    MM is unorientedly cobordant to a spin manifold.

  2. (2)

    The Stiefel–Whitney numbers of MM involving w1w_{1} and w2w_{2} are all zero.

  3. (3)

    MM is unorientedly cobordant to N×NN\times N, with NN an orientable compact manifold.

Milnor points out that, if there exists a compact spin manifold Σ\Sigma whose Pontryagin number s6(p1,,p6)[Σ]s_{6}(p_{1},\dots,p_{6})[\Sigma] is odd, then these conditions would fail to be equivalent in dimension 2424. Anderson–Brown–Peterson [2],[3] established that, as a consequence of their splitting of 22-local MSpinMSpin, there does indeed exist such a compact spin manifold Σ\Sigma. However, it seems that no explicit description of that 2424-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 Σ\Sigma, 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 Ω24Spin𝔑24\Omega^{Spin}_{24}\rightarrow\mathfrak{N}_{24} and the manifold representatives calculated in Proposition 3.5.

1.4. Determination of the mod 22 spinc-cobordism ring up to inseparable isogeny

Thom’s famous calculation [25] established that the unoriented bordism ring ΩO=𝔑\Omega^{O}_{*}=\mathfrak{N}_{*} is isomorphic to a polynomial algebra over 𝔽2\mathbb{F}_{2}. A theorem of Stong [23, Proposition 14] shows that the spinc cobordism ring, reduced modulo torsion and then reduced modulo 22, is also isomorphic to a polynomial 𝔽2\mathbb{F}_{2}-algebra.

By contrast, the spinc cobordism ring cannot itself be isomorphic to a polynomial algebra, since by [2], it has 22-torsion but is not an 𝔽2\mathbb{F}_{2}-algebra, hence it has nontrivial zero divisors. Similarly, since the mod 22 spinc-cobordism ring has nontrivial β\beta-torsion, it cannot be isomorphic to a polynomial 𝔽2\mathbb{F}_{2}-algebra.

It follows as a trivial consequence of Theorem A that the mod (2,β)(2,\beta) spinc-cobordism ring still cannot be a polynomial 𝔽2\mathbb{F}_{2}-algebra. One can, with a bit of calculation, deduce the same fact from the additive structure of 22-local MSpincMSpin^{c}, by verifying that the Poincaré series of the mod (2,β)(2,\beta) 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 ΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta) is instead uniformly FF-isomorphic to a polynomial algebra.

As far as we know, the terms “FF-isomorphism” (perhaps better known as “inseparable isogeny”) and “uniform FF-isomorphism” originated with Quillen [19]:

Definition 1.1.

Given a prime pp, a homomorphism of 𝔽p\mathbb{F}_{p}-algebras f:ABf:A\rightarrow B is said to be an FF-isomorphism if

  • for each akerfa\in\ker f, some power ana^{n} is zero, and

  • for each element bBb\in B, some power bpnb^{p^{n}} of bb is in the image of ff.

The FF-isomorphism ff is said to be uniform if nn can be chosen independently of aa and bb.

The notion of FF-isomorphism is applied only to algebras over a field of positive characteristic, so we had better reduce modulo 22 in order to apply this idea to the spinc-cobordism ring. We get a positive result:

Theorem C (Theorem 4.4).

The mod 22 spinc-cobordism ring ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} is uniformly FF-isomorphic to the graded 𝔽2\mathbb{F}_{2}-algebra

(2) 𝔽2[β,y4i,Z4j2:i1,j1,j not a power of 2]/(βZ4j2),\mathbb{F}_{2}\left[\beta,y_{4i},Z_{4j-2}:i\geq 1,\ j\geq 1,\ j\mbox{\ not\ a\ power\ of\ }2\right]/(\beta Z_{4j-2}),

with β\beta the Bott element in degree 22, with y4iy_{4i} in degree 4i4i, and with Z4j2Z_{4j-2} in degree 4j24j-2.

Corollary D.

The mod (2,β)(2,\beta) spinc-cobordism ring is uniformly FF-isomorphic to a graded polynomial 𝔽2\mathbb{F}_{2}-algebra on

  • a generator in degree 4i4i for all positive integers ii,

  • and a generator in degree 4j24j-2 for all positive integers jj such that jj is not a power of 22.

An FF-isomorphism induces a homeomorphism on prime ideal spectra, so Theorem C yields a description of all prime ideals in the mod 22 spinc cobordism ring. That is, we have

Corollary E (Corollary 4.5).

The topological space SpecΩSpinc/(2)\operatorname{{\rm Spec}}\Omega^{Spin^{c}}_{*}/(2) is homeomorphic to Spec\operatorname{{\rm Spec}} of the 𝔽2\mathbb{F}_{2}-algebra (2). The topological space SpecΩSpinc/(2,β)\operatorname{{\rm Spec}}\Omega^{Spin^{c}}_{*}/(2,\beta) is homeomorphic to Spec\operatorname{{\rm Spec}} of the 𝔽2\mathbb{F}_{2}-algebra described in Corollary D.

1.5. Conventions

  • Given a ring RR and symbols x1,,xnx_{1},\dots,x_{n}, we write R{x1,,xn}R\{x_{1},\dots,x_{n}\} for the free RR-module with basis x1,,xnx_{1},\dots,x_{n}.

  • We write β\beta for the Bott element in π2(ku)\pi_{2}(ku), and also for its corresponding element β=[P1]Ω2Spinc\beta=[\mathbb{C}P^{1}]\in\Omega^{Spin^{c}}_{2} under the Anderson–Brown–Peterson splitting of 22-local MSpincMSpin^{c}.

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 MSpincMSpin^{c} and MSpinMSpin

There is an exact sequence of Lie groups

1U(1)Spinc(n)SO(n)1\displaystyle 1\longrightarrow U(1)\longrightarrow Spin^{c}(n)\longrightarrow SO(n)\longrightarrow 1

that gives rise to the fiber sequence

(3) BU(1)BSpincBSO.\displaystyle BU(1)\longrightarrow BSpin^{c}\longrightarrow BSO.

Using this fibration, Harada and Kono [11] computed the mod 2 cohomology of the space BSpincBSpin^{c}:

Theorem 2.1.

[11]

(4) H(BSpinc;𝔽2)\displaystyle H^{*}(BSpin^{c};\mathbb{F}_{2}) 𝔽2(w2,w3,w4,w5,)/I\displaystyle\cong\mathbb{F}_{2}(w_{2},w_{3},w_{4},w_{5},...)/I

where II is the ideal w3,Sq2(w3),Sq2(Sq4(w3)),Sq8(Sq4(Sq2(w3)),\langle w_{3},\operatorname{{\rm Sq}}^{2}(w_{3}),\operatorname{{\rm Sq}}^{2}(Sq^{4}(w_{3})),\operatorname{{\rm Sq}}^{8}(Sq^{4}(Sq^{2}(w_{3})),...\rangle.

The triviality of the ideal II in the cohomology of BSpincBSpin^{c} is a consequence of the first d2d_{2} differential in the Serre spectral sequence associated to the fiber sequence (3). It is not practical to write down a presentation for the 𝔽2\mathbb{F}_{2}-algebra H(BSpinc;𝔽2)H^{*}(BSpin^{c};\mathbb{F}_{2}) which is more explicit than (4), since the difficulty of calculating iterated Steenrod squares applied to the Stiefel–Whitney class w3w_{3} grows rapidly as the number of Steenrod squares grows. For example, Sq8(Sq4(Sq2(w3))\operatorname{{\rm Sq}}^{8}(\operatorname{{\rm Sq}}^{4}(\operatorname{{\rm Sq}}^{2}(w_{3})) has 38 monomials when expressed as a polynomial in the Stiefel–Whitney classes.

There is also an exact sequence

1/2Spin(n)SO(n)1\displaystyle 1\longrightarrow\mathbb{Z}/2\mathbb{Z}\longrightarrow Spin(n)\longrightarrow SO(n)\longrightarrow 1

which gives rise to the fiber sequence

B/2BSpin(n)BSO(n).\displaystyle B\mathbb{Z}/2\mathbb{Z}\longrightarrow BSpin(n)\longrightarrow BSO(n).

Using this, Quillen calculated:

Theorem 2.2.

[18]

H(BSpin;𝔽2)𝔽2(w2,w3,w4,w5,)/J\displaystyle H^{*}(BSpin;\mathbb{F}_{2})\cong\mathbb{F}_{2}(w_{2},w_{3},w_{4},w_{5},...)/J

where JJ is the ideal w2,w3,Sq2(w3),Sq2(Sq4(w3)),Sq8(Sq4(Sq2(w3)),\langle w_{2},w_{3},\operatorname{{\rm Sq}}^{2}(w_{3}),\operatorname{{\rm Sq}}^{2}(Sq^{4}(w_{3})),\operatorname{{\rm Sq}}^{8}(Sq^{4}(Sq^{2}(w_{3})),...\rangle.

By the Thom isomorphism, we have that H(MSpinc;𝔽2)H(BSpinc;𝔽2){U}H^{*}(MSpin^{c};\mathbb{F}_{2})\cong H^{*}(BSpin^{c};\mathbb{F}_{2})\{U\} and H(MSpin;𝔽2)H(BSpin;𝔽2){U}H^{*}(MSpin;\mathbb{F}_{2})\cong H^{*}(BSpin;\mathbb{F}_{2})\{U\} as graded 𝔽2\mathbb{F}_{2}-vector spaces. Since
H(BSpinc;𝔽2)H^{*}(BSpin^{c};\mathbb{F}_{2}) and H(BSpin;𝔽2)H^{*}(BSpin;\mathbb{F}_{2}) are quotients of H(BO;𝔽2)𝔽2[w1,w2,w3,]H^{*}(BO;\mathbb{F}_{2})\cong\mathbb{F}_{2}[w_{1},w_{2},w_{3},\dots], the action of Steenrod squares on H(BSpinc;𝔽2)H^{*}(BSpin^{c};\mathbb{F}_{2}) and H(BSpin;𝔽2)H^{*}(BSpin;\mathbb{F}_{2}) is determined by the Wu formula Sqiwj=k=0i(j+ki1k)wikwj+k\operatorname{{\rm Sq}}^{i}w_{j}=\sum_{k=0}^{i}\binom{j+k-i-1}{k}w_{i-k}w_{j+k} and the Cartan formula. This, together with the formula SqnU=wnU\operatorname{{\rm Sq}}^{n}U=w^{n}U for the action of Steenrod squares on the Thom class UU, determines the action of the Steenrod squares on the cohomology H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) of the spinc-bordism spectrum MSpincMSpin^{c}.

2.2. Review of MOMO_{*} and symmetric polynomials in the Stiefel–Whitney classes.

The following definitions are classical (see e.g. chapter 1 of [13]):

Definition 2.3.

Let nn be a nonnegative integer.

  • Suppose λ=(a1,,an)\lambda=(a_{1},\dots,a_{n}) is an unordered nn-tuple of nonnegative integers. The monomial symmetric polynomial associated to λ\lambda is the symmetric polynomial

    mλ(X1,,Xn)[X1,,Xn]Σnm_{\lambda}(X_{1},\dots,X_{n})\in\mathbb{Z}[X_{1},\dots,X_{n}]^{\Sigma_{n}}

    which has the fewest nonzero monomial terms among all those which have Xa1Xa2XanX^{a_{1}}X^{a_{2}}\dots X^{a_{n}} as a monomial term.

  • Given a nonnegative integer mnm\leq n, the mmth elementary symmetric polynomial is the symmetric polynomial

    em(X1,,Xn)[X1,,Xn]Σne_{m}(X_{1},\dots,X_{n})\in\mathbb{Z}[X_{1},\dots,X_{n}]^{\Sigma_{n}}

    given by

    em(X1,,Xn)\displaystyle e_{m}(X_{1},\dots,X_{n}) =1d1<d2<<dmnXd1Xd2Xdm.\displaystyle=\sum_{1\leq d_{1}<d_{2}<\dots<d_{m}\leq n}X_{d_{1}}X_{d_{2}}\dots X_{d_{m}}.

The monomial symmetric polynomials form a \mathbb{Z}-linear basis for the ring of symmetric polynomials. The set of all finite products of elementary symmetric polynomials also famously (by Newton) forms a \mathbb{Z}-linear basis for the ring of symmetric polynomials. Consequently, for each λ\lambda, there exists a unique polynomial Pλ(X1,,Xn)P_{\lambda}(X_{1},\dots,X_{n}) such that

Pλ(e1(X1,,Xn),e2(X1,,Xn),,en(X1,,Xn))\displaystyle P_{\lambda}\left(e_{1}(X_{1},\dots,X_{n}),e_{2}(X_{1},\dots,X_{n}),\dots,e_{n}(X_{1},\dots,X_{n})\right) =mλ(X1,,Xn).\displaystyle=m_{\lambda}(X_{1},\dots,X_{n}).

For more details about the polynomials Pλ(X1,,Xn)P_{\lambda}(X_{1},\dots,X_{n}), see the material on the transition matrix M(m,e)M(m,e) and its inverse M(e,m)M(e,m) 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 Λ\Lambda be the set of unordered finite-length tuples of positive integers, each of which is not equal to 2a12^{a}-1 for any integer aa. Such integers are called “non-dyadic,” and such partitions are called “non-dyadic partitions.” For each λΛ\lambda\in\Lambda, write |λ|\left|\lambda\right| for the length of λ\lambda, and write λ\left|\left|\lambda\right|\right| for the sum of the elements of λ\lambda. Consider the polynomial

Pλ(w1,,w|λ|)\displaystyle P_{\lambda}(w_{1},\dots,w_{\left|\lambda\right|}) H(BO;𝔽2)\displaystyle\in H^{*}(BO;\mathbb{F}_{2})
𝔽2[w1,w2,]\displaystyle\cong\mathbb{F}_{2}[w_{1},w_{2},\dots]

in the Stiefel–Whitney classes w1,w2,w_{1},w_{2},\dots. Then (see page 96 of [24], or pages 301-302 of [26]) the set

{Pλ(w1,,w|λ|)U):λΛ}\{P_{\lambda}(w_{1},\dots,w_{\left|\lambda\right|})U):\lambda\in\Lambda\}

is a homogeneous AA-linear basis for the graded free AA-module H(MO;𝔽2)H^{*}(MO;\mathbb{F}_{2}), where UH0(MO;𝔽2)U\in H^{0}(MO;\mathbb{F}_{2}) denotes the Thom class, and AA is the mod 22 Steenrod algebra. Consequently, for each nonnegative integer nn, πn(MO)\pi_{n}(MO) is the 𝔽2\mathbb{F}_{2}-linear dual of the 𝔽2\mathbb{F}_{2}-vector space with basis the set

(5) {Pλ(w1,,w|λ|)U:λΛ,λ=n}.\left\{P_{\lambda}(w_{1},\dots,w_{\left|\lambda\right|})U:\lambda\in\Lambda,\left|\left|\lambda\right|\right|=n\right\}.

Given two tuples (a1,,am)(a_{1},\dots,a_{m}) and (b1,,bn)(b_{1},\dots,b_{n}), we have their concatenation

(a1,,am)(b1,,bn)\displaystyle(a_{1},\dots,a_{m})\coprod(b_{1},\dots,b_{n}) :=(a1,,am,b1,,bn).\displaystyle:=(a_{1},\dots,a_{m},b_{1},\dots,b_{n}).

The coproduct on H(MO;𝔽2)H^{*}(MO;\mathbb{F}_{2}) is then given by

Δ(Pλ(w1,,w|λ|)U)\displaystyle\Delta(P_{\lambda}(w_{1},\dots,w_{\left|\lambda\right|})U) =λ,λ′′Λ:λ=λλ′′Pλ(w1,,w|λ|)UPλ′′(w1,,w|λ′′|)U.\displaystyle=\sum_{\lambda^{\prime},\lambda^{\prime\prime}\in\Lambda:\ \lambda=\lambda^{\prime}\coprod\lambda^{\prime\prime}}P_{\lambda^{\prime}}(w_{1},\dots,w_{\left|\lambda^{\prime}\right|})U\otimes P_{\lambda^{\prime\prime}}(w_{1},\dots,w_{\left|\lambda^{\prime\prime}\right|})U.

Consequently, if we write {Yλ}\{Y_{\lambda}\} for the basis of π(MO)\pi_{*}(MO) dual to the basis (5), then YλYλ=YλλY_{\lambda}Y_{\lambda^{\prime}}=Y_{\lambda\coprod\lambda^{\prime}}, and 𝔑π(MO)𝔽2[Y(2),Y(4),Y(5),Y(6),Y(8),]\mathfrak{N}_{*}\cong\pi_{*}(MO)\cong\mathbb{F}_{2}[Y_{(2)},Y_{(4)},Y_{(5)},Y_{(6)},Y_{(8)},\dots], with Y(i)Y_{(i)} in degree ii. We will sometimes write YiY_{i} as an abbreviation for Thom’s generator Y(i)Y_{(i)} of 𝔑\mathfrak{N}_{*}.

2.3. Maps between bordism theories

The first stages of the Whitehead tower for the orthogonal group are:

BString=BO8BSpin=BO4BSO=BO2BO.\displaystyle BString=BO\langle 8\rangle\longrightarrow BSpin=BO\langle 4\rangle\longrightarrow BSO=BO\langle 2\rangle\longrightarrow BO.

While BSpincBSpin^{c} does not fit into this sequence via a connective cover, the map BSpinBSOBSpin\longrightarrow BSO factors through BSpincBSpin^{c}. There is a commutative diagram whose rows and columns are fiber sequences:

K(/2,0){K(\mathbb{Z}/2\mathbb{Z},0)}Spin(n){Spin(n)}SO(n){SO(n)}K(,1){K(\mathbb{Z},1)}Spinc(n){Spin^{c}(n)}SO(n){SO(n)}K(,1)U(1){K(\mathbb{Z},1)\cong U(1)}U(1){U(1)}{*}

On the level of spectra, we have maps

(6) MStringMSpinMSpincMSOMO.\displaystyle MString\longrightarrow MSpin\longrightarrow MSpin^{c}\longrightarrow MSO\longrightarrow MO.

The maps induced in homotopy give the maps of respective cobordism rings. We will specifically consider the images of MSpincMSpin^{c}_{*} and MSpinMSpin_{*} in MOMO_{*}.

By the Anderson–Brown–Peterson splitting of 22-local MSpincMSpin^{c}, the cohomology H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) splits as a direct sum of suspensions of H(ku;𝔽2)A//E(1)H^{*}(ku;\mathbb{F}_{2})\cong A//E(1) and of H(H𝔽2;𝔽2)AH^{*}(H\mathbb{F}_{2};\mathbb{F}_{2})\cong A. Here we are using the standard notation AA for the mod 22 Steenrod algebra, and A//E(1)A//E(1) for its quotient AE(1)𝔽2A\otimes_{E(1)}\mathbb{F}_{2}, where E(1)E(1) is the subalgebra of AA generated by Sq1\operatorname{{\rm Sq}}^{1} and by Q1=[Sq1,Sq2]Q_{1}=[\operatorname{{\rm Sq}}^{1},\operatorname{{\rm Sq}}^{2}]. Hence the s=0s=0-line in the 22-primary Adams spectral sequence for MSpincMSpin^{c},

(7) E2s,tExtAs,t(H(MSpinc;𝔽2),𝔽2)\displaystyle E_{2}^{s,t}\cong\operatorname{{\rm Ext}}_{A}^{s,t}\left(H^{*}(MSpin^{c};\mathbb{F}_{2}),\mathbb{F}_{2}\right) πts(MSpinc)2\displaystyle\Rightarrow\pi_{t-s}(MSpin^{c})^{\wedge}_{2}
dr:Ers,t\displaystyle d_{r}:E_{r}^{s,t} Ers+r,t+r1\displaystyle\rightarrow E_{r}^{s+r,t+r-1}

is a direct sum of suspensions of 𝔽2\mathbb{F}_{2}, with one summand Σt𝔽2\Sigma^{t}\mathbb{F}_{2} for each summand Σtku(2)\Sigma^{t}ku_{(2)} in MSpin(2)cMSpin^{c}_{(2)}, and also with one summand Σt𝔽2\Sigma^{t}\mathbb{F}_{2} for each summand ΣtH𝔽2\Sigma^{t}H\mathbb{F}_{2} in MSpin(2)cMSpin^{c}_{(2)}.

Anderson–Brown–Peterson prove in [3] that all differentials in the Adams spectral sequence (7) are zero. Consequently the s=0s=0-line homA(H(MSpinc;𝔽2),𝔽2)\hom_{A}(H^{*}(MSpin^{c};\mathbb{F}_{2}),\mathbb{F}_{2}) is the reduction of π(MSpinc)\pi_{*}(MSpin^{c}) modulo the ideal generated by 22 and by the Bott element βπ2(ku)\beta\in\pi_{2}(ku).

This means we can calculate π(MSpinc)/(2,β)\pi_{*}(MSpin^{c})/(2,\beta) simply by calculatinghomA(H(MSpinc;𝔽2),𝔽2)\hom_{A}(H^{*}(MSpin^{c};\mathbb{F}_{2}),\mathbb{F}_{2}), i.e., the AA_{*}-comodule primitives 𝔽2AH(MSpinc;𝔽2)\mathbb{F}_{2}\Box_{A_{*}}H_{*}(MSpin^{c};\mathbb{F}_{2}). The advantage of thinking in terms of comodule primitives is that the Adams spectral sequence respects ring structure: if we calculate the homology H(MSpinc;𝔽2)H_{*}(MSpin^{c};\mathbb{F}_{2}) as a ring, then by simply restricting to the comodule primitives in H(MSpinc;𝔽2)H_{*}(MSpin^{c};\mathbb{F}_{2}), we have calculated ΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta).

The same remarks apply mutatis mutandis for the spin bordism spectrum MSpinMSpin, the oriented bordism spectrum MSOMSO or for the unoriented bordism spectrum MOMO in place of MSpincMSpin^{c}. The Anderson–Brown–Peterson splitting for MSpinMSpin is as a wedge of suspensions of koko, ko2ko\langle 2\rangle, and H𝔽2H\mathbb{F}_{2}. The analogue of the Anderson–Brown–Peterson splitting for MOMO is Thom’s splitting of MOMO as a wedge of suspensions of H𝔽2H\mathbb{F}_{2}, while MSO(2)MSO_{(2)} splits as a wedge of suspensions of H(2)H\mathbb{Z}_{(2)} and H𝔽2H\mathbb{F}_{2}; as far as we know, the latter splitting was originally proven by Wall [26]. Since H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) is a quotient AA-module of H(MSO;𝔽2)H^{*}(MSO;\mathbb{F}_{2}), which is in turn a quotient AA-module of H(MO;𝔽2)H^{*}(MO;\mathbb{F}_{2}), dualizing yields that H(MSpinc;𝔽2)H_{*}(MSpin^{c};\mathbb{F}_{2}) is a subcomodule of H(MSO;𝔽2)H_{*}(MSO;\mathbb{F}_{2}), which is in turn a subcomodule of H(MO;𝔽2)H_{*}(MO;\mathbb{F}_{2}).

Hence our broad strategy for calculating ΩSpinc/(2,β)MSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta)\cong MSpin^{c}_{*}/(2,\beta) and ΩSpin/(2,η,α,β)MSpin/(2,η,α,β)\Omega^{Spin}_{*}/(2,\eta,\alpha,\beta)\cong MSpin_{*}/(2,\eta,\alpha,\beta), and the natural maps MSpin/(2,η,α,β)MSpinc/(2,β)MOMSpin_{*}/(2,\eta,\alpha,\beta)\rightarrow MSpin^{c}_{*}/(2,\beta)\rightarrow MO_{*}, is to calculate the AA_{*}-comodule primitives in H(MSpinc;𝔽2)H_{*}(MSpin^{c};\mathbb{F}_{2}) and in H(MSpin;𝔽2)H_{*}(MSpin;\mathbb{F}_{2}), regarding each as AA_{*}-subcomodule algebras of H(MO;𝔽2)𝔽2[w1,w2,w3,]{U}H_{*}(MO;\mathbb{F}_{2})\cong\mathbb{F}_{2}[w_{1},w_{2},w_{3},\dots]\{U\}. The resulting information will describe ΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta) as a subring of 𝔑\mathfrak{N}_{*}. 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:

MSpin/(2,η,α,B){{\text{MSpin}}_{*}/(2,\eta,\alpha,B)}H(MSpin;𝔽2){{H_{*}(\text{MSpin};\mathbb{F}_{2})}}MSpinc/(2,β){{\text{MSpin}^{c}_{*}/(2,\beta)}}H(MSpinc;𝔽2){{H_{*}(\text{MSpin}^{c};\mathbb{F}_{2})}}MSO/(2){{\text{MSO}_{*}/(2)}}H(MSO;𝔽2){{H_{*}(\text{MSO};\mathbb{F}_{2})}}MO{{\text{MO}_{*}}}H(MO;𝔽2){{H_{*}(\text{MO};\mathbb{F}_{2})}}

3. The SpincSpin^{c} bordism ring in low degrees

In the statement of Proposition 3.1, we use Thom’s presentation 𝔽2[Y2,Y4,Y5,]\mathbb{F}_{2}[Y_{2},Y_{4},Y_{5},...] for the unoriented cobordism ring 𝔑\mathfrak{N}_{*}.

Proposition 3.1.

The image of the map ΩSpinc𝔑\Omega^{Spin^{c}}_{*}\longrightarrow\mathfrak{N}_{*} agrees, in degrees 33\leq 33, with subring of 𝔑\mathfrak{N}_{*} generated by the elements

Y22,Y42,Y52,Y62,Y92,Y102,Y112,Y122,Y132,Y142,Y152,Y162,T24,T29,T31,T32, and T33,Y_{2}^{2},Y_{4}^{2},Y_{5}^{2},Y_{6}^{2},Y_{9}^{2},Y_{10}^{2},Y_{11}^{2},Y_{12}^{2},Y_{13}^{2},Y_{14}^{2},Y_{15}^{2},Y_{16}^{2},T_{24},T_{29},T_{31},T_{32},\mbox{\ and\ }T_{33},

where

(8) T24\displaystyle T_{24} =Y14Y52+Y13Y11+Y13Y9Y2+Y13Y6Y5+Y13Y5Y23+Y12Y52Y2+Y112Y2\displaystyle=Y_{14}Y_{5}^{2}+Y_{13}Y_{11}+Y_{13}Y_{9}Y_{2}+Y_{13}Y_{6}Y_{5}+Y_{13}Y_{5}Y_{2}^{3}+Y_{12}Y_{5}^{2}Y_{2}+Y_{11}^{2}Y_{2}
+Y11Y9Y4+Y11Y8Y5+Y11Y6Y5Y2+Y11Y5Y42+Y11Y5Y4Y22+Y10Y52Y4\displaystyle\ \ \ +Y_{11}Y_{9}Y_{4}+Y_{11}Y_{8}Y_{5}+Y_{11}Y_{6}Y_{5}Y_{2}+Y_{11}Y_{5}Y_{4}^{2}+Y_{11}Y_{5}Y_{4}Y_{2}^{2}+Y_{10}Y_{5}^{2}Y_{4}
+Y10Y52Y22+Y92Y4Y2+Y92Y23+Y9Y8Y5Y2+Y9Y6Y5Y4+Y9Y6Y5Y22\displaystyle\ \ \ +Y_{10}Y_{5}^{2}Y_{2}^{2}+Y_{9}^{2}Y_{4}Y_{2}+Y_{9}^{2}Y_{2}^{3}+Y_{9}Y_{8}Y_{5}Y_{2}+Y_{9}Y_{6}Y_{5}Y_{4}+Y_{9}Y_{6}Y_{5}Y_{2}^{2}
+Y9Y52+Y9Y5Y42Y2+Y62Y52Y2+Y54Y4+Y52Y43Y2+Y52Y42Y23,\displaystyle\ \ \ +Y_{9}Y_{5}^{2}+Y_{9}Y_{5}Y_{4}^{2}Y_{2}+Y_{6}^{2}Y_{5}^{2}Y_{2}+Y_{5}^{4}Y_{4}+Y_{5}^{2}Y_{4}^{3}Y_{2}+Y_{5}^{2}Y_{4}^{2}Y_{2}^{3},
T29\displaystyle T_{29} =Y19Y52+Y17Y52Y2+Y14Y53+Y13Y6Y52+Y13Y52Y23+Y11Y92+Y11Y8Y52\displaystyle=Y_{19}Y_{5}^{2}+Y_{17}Y_{5}^{2}Y_{2}+Y_{14}Y_{5}^{3}+Y_{13}Y_{6}Y_{5}^{2}+Y_{13}Y_{5}^{2}Y_{2}^{3}+Y_{11}Y_{9}^{2}+Y_{11}Y_{8}Y_{5}^{2}
+Y11Y52Y4Y22+Y10Y9Y52+Y10Y53Y22+Y93Y2+Y92Y6Y5+Y92Y5Y23\displaystyle\ \ \ +Y_{11}Y_{5}^{2}Y_{4}Y_{2}^{2}+Y_{10}Y_{9}Y_{5}^{2}+Y_{10}Y_{5}^{3}Y_{2}^{2}+Y_{9}^{3}Y_{2}+Y_{9}^{2}Y_{6}Y_{5}+Y_{9}^{2}Y_{5}Y_{2}^{3}
+Y9Y6Y52Y22+Y9Y54+Y62Y53Y2+Y55Y4+Y53Y42Y23,\displaystyle\ \ \ +Y_{9}Y_{6}Y_{5}^{2}Y_{2}^{2}+Y_{9}Y_{5}^{4}+Y_{6}^{2}Y_{5}^{3}Y_{2}+Y_{5}^{5}Y_{4}+Y_{5}^{3}Y_{4}^{2}Y_{2}^{3},
T31\displaystyle T_{31} =Y21Y52+Y19Y52Y2+Y17Y52Y4+Y17Y52Y22+Y16Y53+Y13Y92+Y13Y8Y52\displaystyle=Y_{21}Y_{5}^{2}+Y_{19}Y_{5}^{2}Y_{2}+Y_{17}Y_{5}^{2}Y_{4}+Y_{17}Y_{5}^{2}Y_{2}^{2}+Y_{16}Y_{5}^{3}+Y_{13}Y_{9}^{2}+Y_{13}Y_{8}Y_{5}^{2}
+Y13Y52Y4Y22+Y12Y9Y52+Y12Y53Y22+Y11Y10Y52+Y11Y92Y2+Y11Y6Y52Y22\displaystyle\ \ \ +Y_{13}Y_{5}^{2}Y_{4}Y_{2}^{2}+Y_{12}Y_{9}Y_{5}^{2}+Y_{12}Y_{5}^{3}Y_{2}^{2}+Y_{11}Y_{10}Y_{5}^{2}+Y_{11}Y_{9}^{2}Y_{2}+Y_{11}Y_{6}Y_{5}^{2}Y_{2}^{2}
+Y93Y4+Y93Y22+Y92Y8Y5+Y92Y5Y42+Y92Y5Y4Y22+Y92Y5Y24\displaystyle\ \ \ +Y_{9}^{3}Y_{4}+Y_{9}^{3}Y_{2}^{2}+Y_{9}^{2}Y_{8}Y_{5}+Y_{9}^{2}Y_{5}Y_{4}^{2}+Y_{9}^{2}Y_{5}Y_{4}Y_{2}^{2}+Y_{9}^{2}Y_{5}Y_{2}^{4}
+Y9Y8Y52Y22+Y9Y62Y52+Y9Y52Y42Y22+Y82Y53+Y62Y53Y4+Y53Y43Y22+Y53Y42Y24,\displaystyle\ \ \ +Y_{9}Y_{8}Y_{5}^{2}Y_{2}^{2}+Y_{9}Y_{6}^{2}Y_{5}^{2}+Y_{9}Y_{5}^{2}Y_{4}^{2}Y_{2}^{2}+Y_{8}^{2}Y_{5}^{3}+Y_{6}^{2}Y_{5}^{3}Y_{4}+Y_{5}^{3}Y_{4}^{3}Y_{2}^{2}+Y_{5}^{3}Y_{4}^{2}Y_{2}^{4},
T32\displaystyle T_{32} =Y22Y52+Y21Y11+Y21Y9Y2+Y21Y6Y5+Y21Y5Y23+Y20Y52Y2+Y19Y13\displaystyle=Y_{22}Y_{5}^{2}+Y_{21}Y_{11}+Y_{21}Y_{9}Y_{2}+Y_{21}Y_{6}Y_{5}+Y_{21}Y_{5}Y_{2}^{3}+Y_{20}Y_{5}^{2}Y_{2}+Y_{19}Y_{13}
+Y19Y9Y4+Y19Y8Y5+Y19Y6Y5Y2+Y19Y5Y42+Y19Y5Y4Y22+Y18Y52Y4\displaystyle\ \ \ +Y_{19}Y_{9}Y_{4}+Y_{19}Y_{8}Y_{5}+Y_{19}Y_{6}Y_{5}Y_{2}+Y_{19}Y_{5}Y_{4}^{2}+Y_{19}Y_{5}Y_{4}Y_{2}^{2}+Y_{18}Y_{5}^{2}Y_{4}
+Y18Y52Y22+Y17Y13Y2+Y17Y11Y4+Y17Y8Y5Y2+Y17Y6Y5Y4+Y17Y6Y5Y22\displaystyle\ \ \ +Y_{18}Y_{5}^{2}Y_{2}^{2}+Y_{17}Y_{13}Y_{2}+Y_{17}Y_{11}Y_{4}+Y_{17}Y_{8}Y_{5}Y_{2}+Y_{17}Y_{6}Y_{5}Y_{4}+Y_{17}Y_{6}Y_{5}Y_{2}^{2}
+Y17Y53+Y17Y5Y42Y2+Y16Y11Y5+Y16Y9Y5Y2+Y14Y13Y5+Y14Y11Y5Y2\displaystyle\ \ \ +Y_{17}Y_{5}^{3}+Y_{17}Y_{5}Y_{4}^{2}Y_{2}+Y_{16}Y_{11}Y_{5}+Y_{16}Y_{9}Y_{5}Y_{2}+Y_{14}Y_{13}Y_{5}+Y_{14}Y_{11}Y_{5}Y_{2}
+Y14Y92+Y14Y9Y5Y4+Y14Y9Y5Y22+Y132Y6+Y132Y23+Y13Y11Y6Y2\displaystyle\ \ \ +Y_{14}Y_{9}^{2}+Y_{14}Y_{9}Y_{5}Y_{4}+Y_{14}Y_{9}Y_{5}Y_{2}^{2}+Y_{13}^{2}Y_{6}+Y_{13}^{2}Y_{2}^{3}+Y_{13}Y_{11}Y_{6}Y_{2}
+Y13Y10Y9+Y13Y10Y5Y22+Y13Y9Y8Y2+Y13Y9Y6Y4+Y13Y62Y5Y2+Y13Y53Y4\displaystyle\ \ \ +Y_{13}Y_{10}Y_{9}+Y_{13}Y_{10}Y_{5}Y_{2}^{2}+Y_{13}Y_{9}Y_{8}Y_{2}+Y_{13}Y_{9}Y_{6}Y_{4}+Y_{13}Y_{6}^{2}Y_{5}Y_{2}+Y_{13}Y_{5}^{3}Y_{4}
+Y12Y11Y9+Y12Y11Y5Y22+Y12Y10Y52+Y12Y9Y6Y5+Y12Y54+Y112Y10\displaystyle\ \ \ +Y_{12}Y_{11}Y_{9}+Y_{12}Y_{11}Y_{5}Y_{2}^{2}+Y_{12}Y_{10}Y_{5}^{2}+Y_{12}Y_{9}Y_{6}Y_{5}+Y_{12}Y_{5}^{4}+Y_{11}^{2}Y_{10}
+Y112Y8Y2+Y112Y6Y22+Y112Y4Y23+Y11Y10Y6Y5+Y11Y9Y8Y4+Y11Y9Y62\displaystyle\ \ \ +Y_{11}^{2}Y_{8}Y_{2}+Y_{11}^{2}Y_{6}Y_{2}^{2}+Y_{11}^{2}Y_{4}Y_{2}^{3}+Y_{11}Y_{10}Y_{6}Y_{5}+Y_{11}Y_{9}Y_{8}Y_{4}+Y_{11}Y_{9}Y_{6}^{2}
+Y11Y62Y5Y4+Y11Y6Y53+Y102Y52Y2+Y10Y9Y8Y5+Y10Y9Y5Y42+Y10Y62Y52\displaystyle\ \ \ +Y_{11}Y_{6}^{2}Y_{5}Y_{4}+Y_{11}Y_{6}Y_{5}^{3}+Y_{10}^{2}Y_{5}^{2}Y_{2}+Y_{10}Y_{9}Y_{8}Y_{5}+Y_{10}Y_{9}Y_{5}Y_{4}^{2}+Y_{10}Y_{6}^{2}Y_{5}^{2}
+Y93Y5+Y92Y8Y6+Y92Y6Y42+Y92Y52Y4+Y9Y82Y5Y2+Y9Y8Y53+Y9Y63Y5\displaystyle\ \ \ +Y_{9}^{3}Y_{5}+Y_{9}^{2}Y_{8}Y_{6}+Y_{9}^{2}Y_{6}Y_{4}^{2}+Y_{9}^{2}Y_{5}^{2}Y_{4}+Y_{9}Y_{8}^{2}Y_{5}Y_{2}+Y_{9}Y_{8}Y_{5}^{3}+Y_{9}Y_{6}^{3}Y_{5}
+Y82Y52Y4Y2+Y82Y52Y23,and\displaystyle\ \ \ +Y_{8}^{2}Y_{5}^{2}Y_{4}Y_{2}+Y_{8}^{2}Y_{5}^{2}Y_{2}^{3},\ \ \ \mbox{and}
T33\displaystyle T_{33} =Y23Y52+Y21Y52Y2+Y19Y52Y4+Y18Y53+Y17Y6Y52+Y14Y9Y52+Y13Y10Y52\displaystyle=Y_{23}Y_{5}^{2}+Y_{21}Y_{5}^{2}Y_{2}+Y_{19}Y_{5}^{2}Y_{4}+Y_{18}Y_{5}^{3}+Y_{17}Y_{6}Y_{5}^{2}+Y_{14}Y_{9}Y_{5}^{2}+Y_{13}Y_{10}Y_{5}^{2}
+Y13Y52Y42Y2+Y12Y11Y52+Y11Y11Y11+Y11Y11Y9Y2+Y11Y11Y6Y5+Y11Y11Y5Y23\displaystyle\ \ \ +Y_{13}Y_{5}^{2}Y_{4}^{2}Y_{2}+Y_{12}Y_{11}Y_{5}^{2}+Y_{11}Y_{11}Y_{11}+Y_{11}Y_{11}Y_{9}Y_{2}+Y_{11}Y_{11}Y_{6}Y_{5}+Y_{11}Y_{11}Y_{5}Y_{2}^{3}
+Y11Y52Y42Y4+Y10Y53Y42+Y92Y53+Y9Y6Y52Y42+Y82Y53Y2+Y55Y42+Y53Y44Y2.\displaystyle\ \ \ +Y_{11}Y_{5}^{2}Y_{4}^{2}Y_{4}+Y_{10}Y_{5}^{3}Y_{4}^{2}+Y_{9}^{2}Y_{5}^{3}+Y_{9}Y_{6}Y_{5}^{2}Y_{4}^{2}+Y_{8}^{2}Y_{5}^{3}Y_{2}+Y_{5}^{5}Y_{4}^{2}+Y_{5}^{3}Y_{4}^{4}Y_{2}.
Proof.

This proposition is proven using computer calculation. We will describe our method for calculating im(ΩSpincΩO)\operatorname{{\rm im\>}}(\Omega^{Spin^{c}}_{*}\rightarrow\Omega^{O}_{*}) in degrees d\leq d, for any fixed choice of dd. 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 3333, 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 22 cohomology of BOBO detailed in section 2. Let λ\lambda be a non-dyadic partition of a nonnegative integer nn. We want to know whether its corresponding element YλMOnY_{\lambda}\in MO_{n} is in the image of the map MSpinncMOnMSpin^{c}_{n}\rightarrow MO_{n}. The element YλMOnY_{\lambda}\in MO_{n} has a Hurewicz image, i.e., the image of YλY_{\lambda} under the Hurewicz map πn(MO)Hn(MO;𝔽2)\pi_{n}(MO)\rightarrow H_{n}(MO;\mathbb{F}_{2}). In section 2.2, we described the dual element PλUHn(MO;𝔽2)P_{\lambda}U\in H^{n}(MO;\mathbb{F}_{2}) to the Hurewicz image of YλY_{\lambda}, using Thom’s basis for MOMO_{*}. The element PλUP_{\lambda}U can be written as UU 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 PλUP_{\lambda}U is calculated, we see that YλY_{\lambda} is in the image if and only if, when reduced modulo w1w_{1} and the relations in the H(BO;𝔽2)H^{*}(BO;\mathbb{F}_{2})-module H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}), PλUP_{\lambda}U is an AA-module primitive in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}).

Consequently our method for calculating im(ΩSpincΩO)\operatorname{{\rm im\>}}(\Omega^{Spin^{c}}_{*}\rightarrow\Omega^{O}_{*}) is merely a method for building up a basis for the 𝔽2\mathbb{F}_{2}-vector space of AA-module primitives through some fixed degree dd, 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 0, where there is nothing to say: the empty partition \varnothing yields the unique AA-module primitive U\varnothing\cdot U in H0(MSpinc;𝔽2)H^{0}(MSpin^{c};\mathbb{F}_{2}). For each integer n[1,d]n\in[1,d], the product SqnUHn(MSpinc;𝔽2)\operatorname{{\rm Sq}}^{n}\cdot\varnothing\cdot U\in H^{n}(MSpin^{c};\mathbb{F}_{2}) is simply wnUw_{n}U modulo the relations in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}), by the classical formula SqnU=wnU\operatorname{{\rm Sq}}^{n}U=w_{n}U for the action of Steenrod squares on the Thom class in H0(MO;𝔽2)H^{0}(MO;\mathbb{F}_{2}). Record the elements {Sq1U,Sq2U,,SqdU}\{\operatorname{{\rm Sq}}^{1}U,\operatorname{{\rm Sq}}^{2}U,\dots,\operatorname{{\rm Sq}}^{d}U\} in an unordered list DD. Here the symbol DD stands for “decomposable,” as we will use it to build up a list of AA-module decomposables in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) in degrees d\leq d.

We are not done with the initial step in the induction: for each nonzero element SqnU\operatorname{{\rm Sq}}^{n}U, we calculate {Sq1SqnU,Sq2SqnU,,SqdnSqnU}\{\operatorname{{\rm Sq}}^{1}\operatorname{{\rm Sq}}^{n}U,\operatorname{{\rm Sq}}^{2}\operatorname{{\rm Sq}}^{n}U,\dots,\operatorname{{\rm Sq}}^{d-n}\operatorname{{\rm Sq}}^{n}U\} using the Thom formula SqnU=wnU\operatorname{{\rm Sq}}^{n}U=w_{n}U and the Wu formula ([27], but see [15, pg. 94] for a textbook reference) for the action of Sqn\operatorname{{\rm Sq}}^{n} on Stiefel–Whitney classes, and we include the results in DD. Now DD contains an 𝔽2\mathbb{F}_{2}-linear basis for the AA-submodule of H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) generated by the Thom class UU, in all degrees d\leq d.

Now we are ready for the inductive step:

Inductive hypothesis at the nnth step:

We have produced a list GG of 𝔽2\mathbb{F}_{2}-linear combinations of non-dyadic partitions of degree <n<n, such that the 𝔽2\mathbb{F}_{2}-linear span of {PλU:λG}\{P_{\lambda}U:\lambda\in G\} is a basis for the set of AA-module primitives in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) in all degrees <n<n. We have also produced a list DD of 𝔽2\mathbb{F}_{2}-linear combinations of non-dyadic partitions of degree d\leq d, such that the 𝔽2\mathbb{F}_{2}-linear span of {PλU:λD}\{P_{\lambda}U:\lambda\in D\} is precisely the AA-submodule of H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) generated by GG in degrees d\leq d.

Calculation for the nnth inductive step:

Write DnD_{n} for the 𝔽2\mathbb{F}_{2}-linear span of the degree nn elements in DD. Calculate an 𝔽2\mathbb{F}_{2}-linear basis BB forHn(MSpinc;𝔽2)/DnH^{n}(MSpin^{c};\mathbb{F}_{2})/D_{n}. Let GG^{\prime} be GBG\cup B. Use the calculated transition matrix to convert the members of GG^{\prime} 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 GG^{\prime}, then all Steenrod squares on those, etc., in degrees d\leq d. Use the transition matrix to convert back to the partition basis, and DD^{\prime} for the resulting list of linear combinations of non-dyadic partitions. Now we are ready to iterate, with GG^{\prime} in place of GG, and with DD^{\prime} in place of DD.

Once we complete the n=dn=d step, we have an 𝔽2\mathbb{F}_{2}-linear basis for the image of the map ΩSpinc𝔑\Omega^{Spin^{c}}_{*}\rightarrow\mathfrak{N}_{*} in all degrees d\leq d, expressed in terms of Thom’s partition basis for 𝔑\mathfrak{N}_{*}. Consequently we have a description of ΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta), in all degrees d\leq d, as a subring of 𝔑𝔽2[Y2,Y4,Y5,Y6,Y8,Y9,]\mathfrak{N}_{*}\cong\mathbb{F}_{2}[Y_{2},Y_{4},Y_{5},Y_{6},Y_{8},Y_{9},\dots]. ∎

In principle there is no obstruction to using the same method to make calculations of products in ΩSpinc\Omega^{Spin^{c}}_{*} in degrees >33>33. We stopped at degree 3333 simply because, around the time we completed degree 3333, we could see enough of the ring structure of ΩSpinc\Omega^{Spin^{c}}_{*} to prove that the mod (2,β)(2,\beta) spinc-cobordism ring is not a polynomial algebra, and to suggest the right statements for Proposition 4.2 and Theorem C. The products in ΩSpinc\Omega^{Spin^{c}}_{*} required for the proof of Theorem 3.6 are known as soon as one computes the ring structure through degree 2424.

The same computational method described in the proof of Proposition 3.1, applied to MSpinMSpin rather than MSpincMSpin^{c}, yields:

Proposition 3.2.

The image of the map ΩSpin𝔑\Omega^{Spin}_{*}\longrightarrow\mathfrak{N}_{*} agrees, in degrees 31\leq 31, with subring of MOMO_{*} generated by the elements

Y24,Y52,Y44,Y92+Y52Y42,Y112+Y92Y22+Y52Y42Y22,Y132+Y112Y22+Y92Y42+Y82Y52,Y64\displaystyle\langle Y_{2}^{4},Y_{5}^{2},Y_{4}^{4},Y_{9}^{2}+Y_{5}^{2}Y_{4}^{2},Y_{11}^{2}+Y_{9}^{2}Y_{2}^{2}+Y_{5}^{2}Y_{4}^{2}Y_{2}^{2},Y_{13}^{2}+Y_{11}^{2}Y_{2}^{2}+Y_{9}^{2}Y_{4}^{2}+Y_{8}^{2}Y_{5}^{2},Y_{6}^{4}
T24+Y122+Y102Y22+Y82Y42+Y82Y24+Y62Y42Y22+Y54Y22+Y46,T29.\displaystyle T_{24}+Y_{12}^{2}+Y_{10}^{2}Y_{2}^{2}+Y_{8}^{2}Y_{4}^{2}+Y_{8}^{2}Y_{2}^{4}+Y_{6}^{2}Y_{4}^{2}Y_{2}^{2}+Y_{5}^{4}Y_{2}^{2}+Y_{4}^{6},T_{29}\rangle.

Each of the elements TiT_{i} defined in Proposition 3.2 is a linear combination of monomials in 𝔑\mathfrak{N}_{*}. Those monomials are generally not individually members of ΩSpinc\Omega^{Spin^{c}}_{*}: for example, Y14Y52𝔑24Y_{14}Y_{5}^{2}\in\mathfrak{N}_{24} does not lift to an element of Ω24Spinc\Omega^{Spin^{c}}_{24}, even though a linear combination of Y14Y52Y_{14}Y_{5}^{2} with other monomials in degree 2424 does lift to the element T24Ω24SpincT_{24}\in\Omega^{Spin^{c}}_{24}.

However, Yi2𝔑2iY_{i}^{2}\in\mathfrak{N}_{2i} lifts to the element Z2iΩ2iSpincZ_{2i}\in\Omega^{Spin^{c}}_{2i}, and consequently the squares of each of the monomials in each of the elements TiT_{i} lift to ΩSpinc\Omega^{Spin^{c}}_{*}. For i=24,29,31,32i=24,29,31,32, and 3333, let U2iU_{2i} denote the element of ΩSpinc\Omega^{Spin^{c}}_{*} obtained by taking the definition of TiT_{i} in Proposition 3.1 and replacing each instance of YnY_{n} with Z2nZ_{2n}. For example, (8) yields that

U48\displaystyle U_{48} =Z28Z102+Z26Z22+Z26Z18Z4+Z26Z12Z10+Z26Z10Z43+Z24Z102Z4+Z222Z4\displaystyle=Z_{28}Z_{10}^{2}+Z_{26}Z_{22}+Z_{26}Z_{18}Z_{4}+Z_{26}Z_{12}Z_{10}+Z_{26}Z_{10}Z_{4}^{3}+Z_{24}Z_{10}^{2}Z_{4}+Z_{22}^{2}Z_{4}
+Z22Z18Z8+Z22Z16Z10+Z22Z12Z10Z4+Z22Z10Z82+Z22Z10Z8Z42\displaystyle\ \ \ +Z_{22}Z_{18}Z_{8}+Z_{22}Z_{16}Z_{10}+Z_{22}Z_{12}Z_{10}Z_{4}+Z_{22}Z_{10}Z_{8}^{2}+Z_{22}Z_{10}Z_{8}Z_{4}^{2}
+Z20Z102Z8+Z20Z102Z42+Z182Z8Z4+Z182Z43+Z18Z16Z10Z4+Z18Z12Z10Z8\displaystyle\ \ \ +Z_{20}Z_{10}^{2}Z_{8}+Z_{20}Z_{10}^{2}Z_{4}^{2}+Z_{18}^{2}Z_{8}Z_{4}+Z_{18}^{2}Z_{4}^{3}+Z_{18}Z_{16}Z_{10}Z_{4}+Z_{18}Z_{12}Z_{10}Z_{8}
+Z18Z12Z10Z42+Z18Z102+Z18Z10Z82Z4+Z122Z102Z4+Z104Z8+Z102Z83Z4\displaystyle\ \ \ +Z_{18}Z_{12}Z_{10}Z_{4}^{2}+Z_{18}Z_{10}^{2}+Z_{18}Z_{10}Z_{8}^{2}Z_{4}+Z_{12}^{2}Z_{10}^{2}Z_{4}+Z_{10}^{4}Z_{8}+Z_{10}^{2}Z_{8}^{3}Z_{4}
+Z102Z82Z43.\displaystyle\ \ \ +Z_{10}^{2}Z_{8}^{2}Z_{4}^{3}.

Then, as a consequence of Proposition 3.1, we have:

Theorem 3.3.

The subring of ΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta) generated by all homogeneous elements of degree 33\leq 33 is isomorphic to:

𝔽2[Z4,Z8,Z10,Z12,Z16,Z18,Z20,Z22,Z24,Z26,Z28,Z32,T24,T29,T31,T32,T33]/I,\displaystyle\mathbb{F}_{2}[Z_{4},Z_{8},Z_{10},Z_{12},Z_{16},Z_{18},Z_{20},Z_{22},Z_{24},Z_{26},Z_{28},Z_{32},T_{24},T_{29},T_{31},T_{32},T_{33}]/I,

where II is the ideal generated by T242U48,T292U58,T312U62,T322U64,T_{24}^{2}-U_{48},\ T_{29}^{2}-U_{58},\ T_{31}^{2}-U_{62},\ T_{32}^{2}-U_{64}, and T332U66T_{33}^{2}-U_{66}.

The relations Ti2=U2iT_{i}^{2}=U_{2i}, with TiT_{i} indecomposable in ΩiSpinc\Omega^{Spin^{c}}_{i} and with U2iU_{2i} a polynomial in the indecomposable elements ZnZ_{n}, immediately implies that ΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta) is not a polynomial algebra.

Theorem 3.4.

The subring of the mod 22 spinc-cobordism ring ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} generated by all homogeneous elements of degree 33\leq 33 is isomorphic to:

𝔽2[β,Z4,Z8,Z10,Z12,Z16,Z18,Z20,Z22,Z24,Z26,Z28,Z32,T24,T29,T31,T32,T33]/I,\displaystyle\mathbb{F}_{2}[\beta,Z_{4},Z_{8},Z_{10},Z_{12},Z_{16},Z_{18},Z_{20},Z_{22},Z_{24},Z_{26},Z_{28},Z_{32},T_{24},T_{29},T_{31},T_{32},T_{33}]/I,

where II is the ideal generated by

  • βZi\beta Z_{i} for each i2mod4i\equiv 2\mod 4,

  • and βTi\beta T_{i} and Ti2U2iT_{i}^{2}-U_{2i} for i=24,29,31,32,33i=24,29,31,32,33.

Proof.

Let TT denote the ideal of ΩSpinc\Omega^{Spin^{c}}_{*} consisting of 22-torsion elements, and let T~\tilde{T} denote the kernel of the ring map

ΩSpinc𝔽2\displaystyle\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} (ΩSpinc/T)𝔽2.\displaystyle\rightarrow\left(\Omega^{Spin^{c}}_{*}/T\right)\otimes_{\mathbb{Z}}\mathbb{F}_{2}.

The ring (ΩSpinc/T)𝔽2\left(\Omega^{Spin^{c}}_{*}/T\right)\otimes_{\mathbb{Z}}\mathbb{F}_{2} was calculated by Stong [23, Proposition 11]: it is a polynomial 𝔽2[β]\mathbb{F}_{2}[\beta]-algebra on generators in degrees 4,8,12,16,4,8,12,16,\dots. Since T~\tilde{T} is an ideal in ΩSpinc\Omega^{Spin^{c}}_{*}, to calculate the product in the ring ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2}, it suffices to calculate

  • the products between generators of T~\tilde{T},

  • and the products between generators of T~\tilde{T} and lifts, to ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} of generators of (ΩSpinc/T)𝔽2\left(\Omega^{Spin^{c}}_{*}/T\right)\otimes_{\mathbb{Z}}\mathbb{F}_{2}.

Both of these types of products land in T~\tilde{T}. Since T~\tilde{T} maps injectively under the map ΩSpinc𝔑\Omega^{Spin^{c}}_{*}\rightarrow\mathfrak{N}_{*}, we can embed ΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta) into 𝔑\mathfrak{N}_{*} 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 ΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta) through degree 3333, a maximal set of linear combinations of products of generators which generate β\beta-torsion elements in ΩSpinc/(2)\Omega^{Spin^{c}}_{*}/(2), i.e., copies of H𝔽2H\mathbb{F}_{2} rather than ku(2)ku_{(2)} in the Anderson–Brown–Peterson splitting.

As a consequence of the Anderson–Brown–Peterson splitting, generators ofΩSpinc/(2,β)\Omega^{Spin^{c}}_{*}/(2,\beta) that are 2-torsion, and hence β\beta-torsion, in ΩSpinc\Omega^{Spin^{c}}_{*} are those whose corresponding AA-module primitive in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) is not Q0Q_{0} or Q1Q_{1} 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 DD be a list of all Stiefel–Whitney monomials in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) in degrees d\leq d which are (Q0,Q1)(Q_{0},Q_{1})-torsion, together with all words in the Steenrod squares applied to such (Q0,Q1)(Q_{0},Q_{1})-torsion Stiefel–Whitney monomials333In principle, this step in the calculation could go wrong, failing to identify all the (Q0,Q1)(Q_{0},Q_{1})-torsion in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}), as follows: suppose there is some 𝔽2\mathbb{F}_{2}-linear combination of Stiefel–Whitney monomials in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) which is (Q0,Q1)(Q_{0},Q_{1})-torsion, but none of its summands are (Q0,Q1)(Q_{0},Q_{1})-torsion. If this occurs, it would not be noticed by the method we describe, since our method only checks the (Q0,Q1)(Q_{0},Q_{1})-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 ΩSpinc\Omega^{Spin^{c}}_{*}, as follows. After running our method, we compare the rank of our calculated (Q0,Q1)(Q_{0},Q_{1})-torsion in each degree to the expected rank, using the known Poincaré series for the 22-torsion in π(MSpinc)\pi_{*}(MSpin^{c}). If our method has failed to notice a linear combination of Stiefel–Whitney monomials which was (Q0,Q1)(Q_{0},Q_{1})-torsion despite its summands not being (Q0,Q1)(Q_{0},Q_{1})-torsion, then the rank of the (Q0,Q1)(Q_{0},Q_{1})-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 3333 in MSpincMSpin^{c}_{*}. If the rank mismatch were to ever occur, the fix is conceptually trivial, but computationally very hard: instead of populating DD with all the homogeneous (Q0,Q1)(Q_{0},Q_{1})-torsion Stiefel–Whitney monomials at the start of the torsion calculation, we simply populate DD with the all the (Q0,Q1)(Q_{0},Q_{1})-torsion Stiefel–Whitney polynomials at the start of the calculation, then re-run the calculation. This of course cannot miss any (Q0,Q1)(Q_{0},Q_{1})-torsion in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2})! Its disadvantage is simply that it is extremely computationally expensive, since the total number of homogeneous Stiefel–Whitney polynomials (not just monomials) in degrees d\leq d in H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) grows extremely quickly as dd grows. We carry out the calculations in the way we describe—i.e., initially populating DD by only the (Q0,Q1)(Q_{0},Q_{1})-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 DD begin as the empty set. Consequently, as we proceed through the induction, DD is not only the set of AA-module decomposables, but also the set of Stiefel–Whitney monomials which generate copies of A//E(1)A//E(1).

Re-running our inductive calculation from Proposition 3.1, but with this initial list for DD, yields a set of AA-module generators for H(MSpinc;𝔽2)H^{*}(MSpin^{c};\mathbb{F}_{2}) modulo (Q0,Q1)(Q_{0},Q_{1})-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 𝔑\mathfrak{N}_{*}), gives us a set of generators for im(ΩSpinc𝔑)\operatorname{{\rm im\>}}\left(\Omega^{Spin^{c}}_{*}\rightarrow\mathfrak{N}_{*}\right) in degrees 0,1,,n0,1,\dots,n, and tells us, for each generator, whether it corresponds to a copy of ku(2)ku_{(2)} or of H𝔽2H\mathbb{F}_{2} under the Anderson–Brown–Peterson splitting.

In degrees 33\leq 33, we find that an element Yλ𝔑Y_{\lambda}\in\mathfrak{N}_{*} which is in the image of the map ΩSpinc𝔑\Omega^{Spin^{c}}_{*}\rightarrow\mathfrak{N}_{*} is 22-torsion as long as the partition λ\lambda includes an odd number. As described in section 1.1, the β\beta-torsion elements of ΩSpinc\Omega^{Spin^{c}}_{*} are exactly the 22-torsion elements. This yields the presentation for ΩSpinc\Omega^{Spin^{c}}_{*} in degrees 33\leq 33 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 ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} and of ΩSpin𝔽2\Omega^{Spin}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} in Proposition 3.1 and in Proposition 3.2, respectively. The following table was calculated through degree 66 by Thom [25]. We extend the calculation through degree 1717. The symbol DiD_{i} denotes the ii-dimensional Dold manifold, defined in [10].

Proposition 3.5 (Thom [25]).

Manifold representatives for elements YnY_{n} in Thom’s partition basis for 𝔑\mathfrak{N}_{*} are as follows:

Element Manifold Representative
Y2Y_{2} P2\mathbb{R}P^{2}
Y4Y_{4} P4P2×P2\mathbb{R}P^{4}\sqcup\mathbb{R}P^{2}\times\mathbb{R}P^{2}
Y5Y_{5} D5D_{5}
Y6Y_{6} P6\mathbb{R}P^{6}
Y8Y_{8} P8(P4)2P4×(P2)2(P2)4\mathbb{R}P^{8}\sqcup(\mathbb{R}P^{4})^{2}\sqcup\mathbb{R}P^{4}\times(\mathbb{R}P^{2})^{2}\sqcup(\mathbb{R}P^{2})^{4}
Y9Y_{9} D9D5×P4D5×(P2)2D_{9}\sqcup D_{5}\times\mathbb{R}P^{4}\sqcup D_{5}\times(\mathbb{R}P^{2})^{2}
Y10Y_{10} P10(P2)5\mathbb{R}P^{10}\sqcup(\mathbb{R}P^{2})^{5}
Y11Y_{11} 𝔻11𝔻9×P2\mathbb{D}_{11}\sqcup\mathbb{D}_{9}\times\mathbb{R}P^{2}
Y12Y_{12} P12(P6)2P8×(P2)2(D52×P2(P4)3\mathbb{R}P^{12}\sqcup(\mathbb{R}P^{6})^{2}\sqcup\mathbb{R}P^{8}\times(\mathbb{R}P^{2})^{2}\sqcup(D_{5}^{2}\times\mathbb{R}P^{2}\sqcup(\mathbb{R}P^{4})^{3}
Y13Y_{13} D13D11×P2D9×P4D8×P5D5×P4×(P2)2D_{13}\sqcup D_{11}\times\mathbb{R}P^{2}\sqcup D_{9}\times\mathbb{R}P^{4}\sqcup D_{8}\times\mathbb{R}P^{5}\sqcup D_{5}\times\mathbb{R}P^{4}\times(\mathbb{R}P^{2})^{2}
Y14Y_{14} P14\mathbb{R}P^{14}
Y16Y_{16} P16P12×(P2)2(P8)2P8×(P4)2\mathbb{R}P^{16}\sqcup\mathbb{R}P^{12}\times(\mathbb{R}P^{2})^{2}\sqcup(\mathbb{R}P^{8})^{2}\sqcup\mathbb{R}P^{8}\times(\mathbb{R}P^{4})^{2}
P8×(P2)4(P6)2×P4P6×(D5)2\sqcup\mathbb{R}P^{8}\times(\mathbb{R}P^{2})^{4}\sqcup(\mathbb{R}P^{6})^{2}\times\mathbb{R}P^{4}\sqcup\mathbb{R}P^{6}\times(D_{5})^{2}
(D5)2×(P2)3(P4)4(P4)2×(P2)4(P2)8\sqcup(D_{5})^{2}\times(\mathbb{R}P^{2})^{3}\sqcup(\mathbb{R}P^{4})^{4}\sqcup(\mathbb{R}P^{4})^{2}\times(\mathbb{R}P^{2})^{4}\sqcup(\mathbb{R}P^{2})^{8}
Y17Y_{17} D17D13×P4D13×(P2)2P12×D5D11×P6D_{17}\sqcup D_{13}\times\mathbb{R}P^{4}\sqcup D_{13}\times(\mathbb{R}P^{2})^{2}\sqcup\mathbb{R}P^{12}\times D_{5}\sqcup D_{11}\times\mathbb{R}P^{6}
D11×P4×P2D11×(P2)3D9×P8D9×P6×P2\sqcup D_{11}\times\mathbb{R}P^{4}\times\mathbb{R}P^{2}\sqcup D_{11}\times(\mathbb{R}P^{2})^{3}\sqcup D_{9}\times\mathbb{R}P^{8}\sqcup D_{9}\times\mathbb{R}P^{6}\times\mathbb{R}P^{2}
P8×D5×(P2)2(P6)2×D5(D5)3×P2\sqcup\mathbb{R}P^{8}\times D_{5}\times(\mathbb{R}P^{2})^{2}\sqcup(\mathbb{R}P^{6})^{2}\times D_{5}\sqcup(D_{5})^{3}\times\mathbb{R}P^{2}
D5×(P4)2×(P2)2D5×P4×(P2)4D5×(P2)6\sqcup D_{5}\times(\mathbb{R}P^{4})^{2}\times(\mathbb{R}P^{2})^{2}\sqcup D_{5}\times\mathbb{R}P^{4}\times(\mathbb{R}P^{2})^{4}\sqcup D_{5}\times(\mathbb{R}P^{2})^{6}
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 23\leq 23. The ambiguity in dimension 2424 stems from the existence of an orientable manifold whose only nonzero Stiefel–Whitney numbers are w4w6w72w_{4}w_{6}w_{7}^{2}, w64w_{6}^{4}, w46w_{4}^{6}, w43w62w_{4}^{3}w_{6}^{2}, and w42w82w_{4}^{2}w_{8}^{2}. Milnor then poses the problem of whether a spin manifold of dimension 2424 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 Im(ΩSpin𝔑)\text{Im}(\Omega^{Spin}_{*}\rightarrow\mathfrak{N}_{*}) which is not the square of an orientable manifold is 2424 [3]. In Proposition 3.2 we calculated ΩSpin\Omega^{Spin}_{*} in degrees through 3131 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 Ω24Spin\Omega^{Spin}_{24}. Solving for an element of Ω24Spin\Omega^{Spin}_{24} with Milnor’s prescribed Stiefel–Whitney numbers, we find an explicit manifold of the kind Milnor asked for:

Theorem 3.6.

The cobordism class T24+Y122+Y102Y22+Y82Y42+Y82Y24+Y62Y42Y22+Y54Y22+Y46Im(ΩSpin𝔑)T_{24}+Y_{12}^{2}+Y_{10}^{2}Y_{2}^{2}+Y_{8}^{2}Y_{4}^{2}+Y_{8}^{2}Y_{2}^{4}+Y_{6}^{2}Y_{4}^{2}Y_{2}^{2}+Y_{5}^{4}Y_{2}^{2}+Y_{4}^{6}\in\text{Im}(\Omega^{Spin}_{*}\rightarrow\mathfrak{N}_{*}) has nonzero Stiefel–Whitney numbers w4w6w72w_{4}w_{6}w_{7}^{2}, w64w_{6}^{4}, w46w_{4}^{6}, w43w62w_{4}^{3}w_{6}^{2}, and w42w82w_{4}^{2}w_{8}^{2}, and is represented by the manifold:

(P2)6×(P6)2(P4)6P2×(P4)3×(D5)2(P2)2×(P4)2×(P6)2\displaystyle(\mathbb{R}P^{2})^{6}\times(\mathbb{R}P^{6})^{2}\sqcup(\mathbb{R}P^{4})^{6}\sqcup\mathbb{R}P^{2}\times(\mathbb{R}P^{4})^{3}\times(D^{5})^{2}\sqcup(\mathbb{R}P^{2})^{2}\times(\mathbb{R}P^{4})^{2}\times(\mathbb{R}P^{6})^{2}
(P2)4×(P8)2(P2)3×P4×D5×(D9)(P4)2×(D5)2×(P6)\displaystyle\ \ \ \sqcup(\mathbb{R}P^{2})^{4}\times(\mathbb{R}P^{8})^{2}\sqcup(\mathbb{R}P^{2})^{3}\times\mathbb{R}P^{4}\times D^{5}\times(D^{9})\sqcup(\mathbb{R}P^{4})^{2}\times(D^{5})^{2}\times(\mathbb{R}P^{6})
(P2)2×D5×P6×(D9)(P6)4(D5)2×P6×(P8)\displaystyle\ \ \ \sqcup(\mathbb{R}P^{2})^{2}\times D^{5}\times\mathbb{R}P^{6}\times(D^{9})\sqcup(\mathbb{R}P^{6})^{4}\sqcup(D^{5})^{2}\times\mathbb{R}P^{6}\times(\mathbb{R}P^{8})
(P4)2×(P8)2(D5)3×(D9)P4×(D5)2×(P10)\displaystyle\ \ \ \sqcup(\mathbb{R}P^{4})^{2}\times(\mathbb{R}P^{8})^{2}\sqcup(D^{5})^{3}\times(D^{9})\sqcup\mathbb{R}P^{4}\times(D^{5})^{2}\times(\mathbb{R}P^{10})
(P2)2×(P10)2(P4)2×D5×(D11)(P2)2×D9×(D11)\displaystyle\ \ \ \sqcup(\mathbb{R}P^{2})^{2}\times(\mathbb{R}P^{10})^{2}\sqcup(\mathbb{R}P^{4})^{2}\times D^{5}\times(D^{11})\sqcup(\mathbb{R}P^{2})^{2}\times D^{9}\times(D^{11})
P2×(D5)2×(P12)P2×P4×D5×(D13)D5×P6×(D13)\displaystyle\ \ \ \sqcup\mathbb{R}P^{2}\times(D^{5})^{2}\times(\mathbb{R}P^{12})\sqcup\mathbb{R}P^{2}\times\mathbb{R}P^{4}\times D^{5}\times(D^{13})\sqcup D^{5}\times\mathbb{R}P^{6}\times(D^{13})
(D5)2×(P14)(P12)2(D11)×(D13)(P2)6×(P6)2(P4)6\displaystyle\ \ \ \sqcup(D^{5})^{2}\times(\mathbb{R}P^{14})\sqcup(\mathbb{R}P^{12})^{2}\sqcup(D^{11})\times(D^{13})\sqcup(\mathbb{R}P^{2})^{6}\times(\mathbb{R}P^{6})^{2}\sqcup(\mathbb{R}P^{4})^{6}
(P2)2×(P4)2×(P6)2(P2)4×(P8)2(P6)4\displaystyle\ \ \ \sqcup(\mathbb{R}P^{2})^{2}\times(\mathbb{R}P^{4})^{2}\times(\mathbb{R}P^{6})^{2}\sqcup(\mathbb{R}P^{2})^{4}\times(\mathbb{R}P^{8})^{2}\sqcup(\mathbb{R}P^{6})^{4}
(P4)2×(P8)2(P2)2×(P10)2(P12)2.\displaystyle\ \ \ \sqcup(\mathbb{R}P^{4})^{2}\times(\mathbb{R}P^{8})^{2}\sqcup(\mathbb{R}P^{2})^{2}\times(\mathbb{R}P^{10})^{2}\sqcup(\mathbb{R}P^{12})^{2}.

4. The spinc-cobordism ring in all degrees.

4.1. A nonunital subring of the 22-torsion in the spinc-cobordism ring.

In Theorem 3.4, we showed that, in degrees 33\leq 33, the cobordism classes Y(5,5),Y(9,9),Y(11,11),Y_{(5,5)},Y_{(9,9)},Y_{(11,11)}, and Y(13,13)𝔑Y_{(13,13)}\in\mathfrak{N}_{*} lift to indecomposable 22-torsion elements Z10,Z18,Z22,Z_{10},Z_{18},Z_{22}, and Z26Z_{26} in ΩSpinc\Omega^{Spin^{c}}_{*}. The elements Y(5,5),Y(9,9),Y_{(5,5)},Y_{(9,9)},\dots are precisely those of the form Yi2𝔑Y_{i}^{2}\in\mathfrak{N}_{*} with ii odd and non-dyadic. It is natural to ask whether this pattern extends above degree 3333 as well. In Proposition 4.2, we show that something of this kind is true, if we use the Dold manifold DiD_{i}, from [10], rather than the cobordism class YiY_{i} from Thom’s partition basis for 𝔑\mathfrak{N}_{*}. In low degrees, the algebraic relation between the Dold manifolds and the Thom generators for 𝔑\mathfrak{N}_{*} is as follows:

Proposition 4.1.

The squares of odd-dimensional Dold manifolds represent the following polynomials in Thom’s generators Y2,Y4,Y5,Y6,Y8,Y_{2},Y_{4},Y_{5},Y_{6},Y_{8},\dots for 𝔑\mathfrak{N}_{*}:

Element Manifold Representative
Y52Y_{5}^{2} D52D_{5}^{2}
Y92+Y52Y42Y_{9}^{2}+Y_{5}^{2}Y_{4}^{2} D92D_{9}^{2}
Y112+Y92Y22+Y52Y42Y22Y_{11}^{2}+Y_{9}^{2}Y_{2}^{2}+Y_{5}^{2}Y_{4}^{2}Y_{2}^{2} D112D_{11}^{2}
Y132+Y112Y22+Y92Y42+Y82Y52Y_{13}^{2}+Y_{11}^{2}Y_{2}^{2}+Y_{9}^{2}Y_{4}^{2}+Y_{8}^{2}Y_{5}^{2} D132D_{13}^{2}

Furthermore, each of these elements of ΩO\Omega^{O}_{*} is in the image of the map ΩSpinΩO\Omega^{Spin}_{*}\rightarrow\Omega^{O}_{*}.

Proof.

The manifold representatives are straightforwardly calculated from Proposition 3.5. By Proposition 3.2, these elements are in the image of the map ΩSpinΩO\Omega^{Spin}_{*}\rightarrow\Omega^{O}_{*}. ∎

Proposition 4.2.

For each odd integer ii such that i+1i+1 is not a power of 22, the Dold manifold DiD_{i} has the property that Di×DiD_{i}\times D_{i} lifts to an indecomposable (2,β)(2,\beta)-torsion element of Ω2iSpinc\Omega^{Spin^{c}}_{2i}. It furthermore lifts to an indecomposable 22-torsion element of Ω2iSpin\Omega^{Spin}_{2i}.

Proof.

Dold [10] proves that there exists a minimal set of generators for the cobordism ring 𝔑\mathfrak{N}_{*} whose odd-degree elements are Di=P(2r1,s2r)D_{i}=P(2^{r}-1,s2^{r}), where i+1=2r(2s+1)i+1=2^{r}(2s+1) and ii is odd. Milnor proved that the map ΩU𝔑\Omega^{U}_{*}\longrightarrow\mathfrak{N}_{*} maps onto all squares of elements in 𝔑\mathfrak{N}_{*}. This map factors through ΩSpinc\Omega^{Spin^{c}}_{*}, so Di×DiD_{i}\times D_{i} lifts to Ω2iSpinc\Omega^{Spin^{c}}_{2i} for all non-dyadic ii. The mod 22 cohomology of the Dold manifold P(m,n)P(m,n) is given as a graded ring by

H(P(m,n);𝔽2)\displaystyle H^{*}(P(m,n);\mathbb{F}_{2}) 𝔽2[c,d]/(cm+1,dn+1),\displaystyle\cong\mathbb{F}_{2}[c,d]/(c^{m+1},d^{n+1}),

with cH1(P(m,n);𝔽2)c\in H^{1}(P(m,n);\mathbb{F}_{2}) and dH2(P(m,n);𝔽2)d\in H^{2}(P(m,n);\mathbb{F}_{2}). The total Stiefel–Whitney class of P(m,n)P(m,n) is

w(P(m,n))\displaystyle w(P(m,n)) =(1+c)m(1+c+d)n+1.\displaystyle=(1+c)^{m}(1+c+d)^{n+1}.

Setting m=2r1m=2^{r}-1 and n=s2rn=s2^{r}, the total Stiefel–Whitney class of DiD_{i} as in the statement of the theorem is then given by:

w(Di)\displaystyle w(D_{i}) =(1+c)2r1(1+c+d)s2r,\displaystyle=(1+c)^{2^{r}-1}(1+c+d)^{s2^{r}},

and in particular, w1=0w_{1}=0. Hence DiD_{i} is orientable.

To show that Di×DiD_{i}\times D_{i} lifts to the spin cobordism ring, one can carry out an algebraic calculation to show that none of the nonzero Stiefel–Whitney numbers of Di×DiD_{i}\times D_{i} are divisible by w2w_{2}. 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 w1w_{1} vanishes on DiD_{i}, its square Di×DiD_{i}\times D_{i} must be in the image of the map Ω2iSpin𝔑2i\Omega^{Spin}_{2i}\rightarrow\mathfrak{N}_{2i}.

It follows easily from the structure of koko_{*}, and the Anderson–Brown–Peterson splitting of MSpin into a wedge of suspensions of koko, ko2ko\langle 2\rangle, and H𝔽2H\mathbb{F}_{2}, that all elements of ΩSpin\Omega^{Spin}_{*} in degrees 2mod4\equiv 2\mod 4 are 22-torsion. Hence, for odd ii, any spin-cobordism class that lifts Di×Di𝔑2iD_{i}\times D_{i}\in\mathfrak{N}_{2i} must be 22-torsion.

Let Di2~\widetilde{D_{i}^{2}} be a lift of Di×Di𝒩2iD_{i}\times D_{i}\in\mathcal{N}_{2i} to Ω2iSpin\Omega^{Spin}_{2i}. The image of Di2~\widetilde{D_{i}^{2}} under the map Ω2iSpinΩ2iSpinc\Omega^{Spin}_{2i}\rightarrow\Omega^{Spin^{c}}_{2i} is then a lift of Di2D_{i}^{2} to Ω2iSpinc\Omega^{Spin^{c}}_{2i}, and it is 22-torsion since it is the image of a 22-torsion element. It is furthermore β\beta-torsion in ΩSpinc\Omega^{Spin^{c}}_{*}, since by the Anderson–Brown–Peterson splitting of 22-local MSpincMSpin^{c}, every 22-torsion element of ΩSpinc\Omega^{Spin^{c}}_{*} is also β\beta-torsion.

To see that Di2Ω2iOD_{i}^{2}\in\Omega^{O}_{2i} lifts to an indecomposable element of ΩSpin\Omega^{Spin}_{*}, it is enough to observe that ΩSpin/(2,η,α,β)\Omega^{Spin}_{*}/(2,\eta,\alpha,\beta) embeds into 𝔑\mathfrak{N}_{*}, and since DiD_{i} has second Stiefel–Whitney class d0d\neq 0, DiD_{i} does not lift to ΩSpin\Omega^{Spin}_{*}. Hence the unique lift of Di2D_{i}^{2} to ΩSpin/(2,η,α,β)\Omega^{Spin}_{*}/(2,\eta,\alpha,\beta) is indecomposable, hence any lift of Di2D_{i}^{2} to ΩSpin\Omega^{Spin}_{*} is indecomposable. A completely analogous argument establishes that any lift of Di2D_{i}^{2} to ΩSpinc\Omega^{Spin^{c}}_{*} is also indecomposable. ∎

Since Dold [10] showed that Di𝔑iD_{i}\in\mathfrak{N}_{i} can be written as Thom’s generator YiY_{i} 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 33\leq 33, and indeed extend to all degrees. Namely, for odd non-dyadic ii, if we write Z2iZ_{2i} for a lift of Di×Di𝔑2iD_{i}\times D_{i}\in\mathfrak{N}_{2i} to an indecomposable 22-torsion element of Ω2iSpinc\Omega^{Spin^{c}}_{2i} (guaranteed to exist by Proposition 4.2), and for even ii we write Z2iZ_{2i} for a lift of Yi2𝔑2iY_{i}^{2}\in\mathfrak{N}_{2i} to an element of Ω2iSpinc\Omega^{Spin^{c}}_{2i}, then we have:

Theorem 4.3.

Consider the spinc-cobordism ring as a graded algebra over the graded ring S:=(2)[β,Z2j:j2,j non-dyadic]/(βZ2j, 2Z2j for odd j).S:=\mathbb{Z}_{(2)}[\beta,Z_{2j}:j\geq 2,\ \mbox{j\ non-dyadic}]/(\beta Z_{2j},\ 2Z_{2j}\mbox{\ for\ odd\ }j). Then the ideal (Z2j:jodd)\left(Z_{2j}:j\ \mbox{odd}\right) of SS embeds, as a non-unital graded SS-algebra, into the 22-torsion ideal π(Z)\pi_{*}(Z) of the spinc-cobordism ring.

4.2. Mod 22 spinc-cobordism, up to uniform FF-isomorphism.

In this section, we show that the torsion-free generators Z4,Z8,Z12,Z16,Z_{4},Z_{8},Z_{12},Z_{16},\dots, together with the large nonunital subring of the 22-torsion in ΩSpinc\Omega^{Spin^{c}}_{*}, constructed in Theorem 4.3, suffice to generate a subring of the mod 22 spinc-cobordism ring 𝔽2ΩSpinc\mathbb{F}_{2}\otimes_{\mathbb{Z}}\Omega^{Spin^{c}}_{*} which, although smaller than 𝔽2ΩSpinc\mathbb{F}_{2}\otimes_{\mathbb{Z}}\Omega^{Spin^{c}}_{*}, is nevertheless uniformly FF-isomorphic to all of 𝔽2ΩSpinc\mathbb{F}_{2}\otimes_{\mathbb{Z}}\Omega^{Spin^{c}}_{*}. See Definition 1.1 for the definition of a uniform FF-isomorphism.

Theorem 4.4.

The mod 22 spinc-cobordism ring is uniformly FF-isomorphic to the graded 𝔽2\mathbb{F}_{2}-algebra

(9) 𝔽2[β,y4i,Z4j2:i1,j1, 2j not a power of 2]/(βZ4j2),\mathbb{F}_{2}\left[\beta,y_{4i},Z_{4j-2}:i\geq 1,\ j\geq 1,\ 2j\mbox{\ not\ a\ power\ of\ }2\right]/(\beta Z_{4j-2}),

with β\beta the Bott element in degree 22, with y4iy_{4i} in degree 4i4i, and with Z4j2Z_{4j-2} in degree 4j24j-2.

Proof.

Write TT for the ideal of ΩSpinc\Omega^{Spin^{c}}_{*} consisting of the 22-torsion elements. Stong [23, Proposition 14] proved that (ΩSpinc/T)𝔽2\left(\Omega^{Spin^{c}}_{*}/T\right)\otimes_{\mathbb{Z}}\mathbb{F}_{2} is a polynomial 𝔽2\mathbb{F}_{2}-algebra on generators y2y_{2} and y4,y8,y12,y_{4},y_{8},y_{12},\dots. Since Ω2Spinc\Omega^{Spin^{c}}_{2}\cong\mathbb{Z} generated by β=[P1]\beta=[\mathbb{C}P^{1}], Stong’s generator y2y_{2} agrees modulo 22 with the Bott element β\beta. Hence (ΩSpinc/T)𝔽2\left(\Omega^{Spin^{c}}_{*}/T\right)\otimes_{\mathbb{Z}}\mathbb{F}_{2} is isomorphic as a graded 𝔽2\mathbb{F}_{2}-algebra to 𝔽2[β,y4i:i1]\mathbb{F}_{2}\left[\beta,y_{4i}:i\geq 1\right].

Now let BB denote the graded subring of ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} generated by β\beta, by y4iy_{4i} for all i>1i>1, and by the mod 22 reductions of the (2,β)(2,\beta)-torsion elements in ΩSpinc\Omega^{Spin^{c}}_{*} from Theorem 4.2 which lift Di×DiD_{i}\times D_{i} for all odd non-dyadic ii. Since BB contains all the squares of elements in ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2}, the graded 𝔽2\mathbb{F}_{2}-algebra map

B\displaystyle B ΩSpinc𝔽2\displaystyle\hookrightarrow\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2}

is a uniform FF-isomorphism.

Let T~\tilde{T} denote the kernel of the ring map ΩSpinc𝔽2(ΩSpinc/T)𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2}\rightarrow\left(\Omega^{Spin^{c}}_{*}/T\right)\otimes_{\mathbb{Z}}\mathbb{F}_{2}. Filter BB by powers of the ideal BT~B\cap\tilde{T}, i.e., equip BB with the (BT~)(B\cap\tilde{T})-adic filtration. By the Anderson–Brown–Peterson splitting and by Theorem 4.2, the associated graded ring E0BE_{0}B is isomorphic to 𝔽2[β]\mathbb{F}_{2}[\beta] tensored with the image of BB in 𝔑\mathfrak{N}_{*} and reduced modulo the relations βx=0\beta\cdot x=0 for all xBT~x\in B\cap\tilde{T}, i.e., E0BE_{0}B is isomorphic to (9).

We claim that BB itself is isomorphic to (9). The (BT~)(B\cap\tilde{T})-adic filtration on BB is additively split, so BE0BB\cong E_{0}B as graded 𝔽2\mathbb{F}_{2}-vector spaces. In principle, the ring structure on E0BE_{0}B could differ from the ring structure on BB if the multiplication on BB were to exhibit (BT~)(B\cap\tilde{T})-adic filtration jumps, i.e., when we multiply two elements x,yx,y of BB, with xx of BT~B\cap\tilde{T}-adic filtration ii and with yy of (BT~)(B\cap\tilde{T})-adic filtration jj, we could perhaps get an element of (BT~)(B\cap\tilde{T})-adic filtration >i+j>i+j.

However, even if filtration jumps occur, E0BE_{0}B is still isomorphic to the graded 𝔽2\mathbb{F}_{2}-algebra with presentation (9). This is by a freeness argument similar to the classical argument that, if the associated graded of a filtered commutative kk-algebra is a polynomial (i.e., free commutative) kk-algebra, then the original filtered commutative kk-algebra must also have been free commutative. The argument is as follows. Let 𝒞\mathcal{C} be the category of pairs (A,S)(A,S), where AA is a graded-commutative 𝔽2[β]\mathbb{F}_{2}[\beta]-algebra, and SS is a set of homogeneous elements of AA such that βx=0\beta\cdot x=0 for all xSx\in S. There is a forgetful functor from 𝒞\mathcal{C} to the category 𝒮ub\operatorname{{\mathcal{S}\mbox{\em{ub}}}} of pairs (S0,S1)(S_{0},S_{1}) in which S0,S1S_{0},S_{1} are sets and S1S0S_{1}\subseteq S_{0}. The forgetful functor 𝒞𝒮ub\mathcal{C}\rightarrow\operatorname{{\mathcal{S}\mbox{\em{ub}}}} sends (A,S)(A,S) to the underlying sets of AA and of SS. The graded 𝔽2[β]\mathbb{F}_{2}[\beta]-algebra (9) is the free object of 𝒞\mathcal{C} on the pair

({y4,y8,y12,y16,}{Z2,Z6,Z10,Z18,Z22,},{Z2,Z6,Z10,Z18,Z22,}).\left(\{y_{4},y_{8},y_{12},y_{16},\dots\}\cup\{Z_{2},Z_{6},Z_{10},Z_{18},Z_{22},\dots\},\ \{Z_{2},Z_{6},Z_{10},Z_{18},Z_{22},\dots\}\right).

By Proposition 4.2, the elements {Z2,Z6,Z10,Z18,Z22,}\{Z_{2},Z_{6},Z_{10},Z_{18},Z_{22},\dots\} are β\beta-torsion in BB, not merely in E0BE_{0}B. Hence there are no relations on E0BE_{0}B except those which make it an object of the category of 𝒞\mathcal{C}, and BB lives in 𝒞\mathcal{C} as well, i.e., E0BE_{0}B and BB are isomorphic in the category 𝒞\mathcal{C}. Hence E0BE_{0}B and BB are isomorphic as graded 𝔽2\mathbb{F}_{2}-algebras, and hence ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} is uniformly FF-isomorphic to the 𝔽2\mathbb{F}_{2}-algebra (9), as claimed. ∎

In section 3, we showed that ΩSpinc𝔽2\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2} is not isomorphic to a polynomial algebra. Nevertheless, since an FF-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 Spec(ΩSpinc/(2,β))\operatorname{{\rm Spec}}\left(\Omega^{Spin^{c}}_{*}/(2,\beta)\right) is homeomorphic to Spec\operatorname{{\rm Spec}} of a polynomial 𝔽2\mathbb{F}_{2}-algebra on countably infinity many generators.

Furthermore, the topological space Spec(ΩSpinc𝔽2)\operatorname{{\rm Spec}}\left(\Omega^{Spin^{c}}_{*}\otimes_{\mathbb{Z}}\mathbb{F}_{2}\right) is homeomorphic to Spec\operatorname{{\rm Spec}} of the ring (9).

The last two sentences in Stong’s 1968 book [24] before the appendices begin are:

One may relate the pair (Spin,Spinc)(Spin,Spin^{c}) through exact sequences in precisely the same way as (SU,U)(SU,U) are related (or as (SO,O)(SO,O) are related). Computationally this is not of much use since one has no way to nicely describe the torsion in ΩSpinc\Omega^{Spin^{c}}_{*}.

We regard Theorems 4.3 and 4.4 as progress toward nicely describing the torsion in ΩSpinc\Omega^{Spin^{c}}_{*} 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, Pinc{\rm Pin}^{c} bordism, and equivariant Spinc{\rm Spin}^{c} bordism for cyclic 22-groups. Pacific J. Math., 128(1):1–24, 1987.
  • [6] Andrew Baker and Jack Morava. MSp{MS}p localized away from 22 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 𝔑{\mathfrak{N}}. 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 mod{\rm mod} 22 cohomology rings of extra-special 22-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 ii-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.