Intersection forms of spin 4-manifolds and the pin(2)-equivariant Mahowald invariant

By Michael J. Hopkins, Jianfeng Lin, XiaoLin Danny Shi, and Zhouli Xu

Abstract

In studying the “11/8-Conjecture” on the Geography Problem in 4-dimensional topology, Furuta proposed a question on the existence of -equivariant stable maps between certain representation spheres. A precise answer of Furuta’s problem was later conjectured by Jones. In this paper, we completely resolve Jones conjecture by analyzing the -equivariant Mahowald invariants. As a geometric application of our result, we prove a “10/8+4”-Theorem.

We prove our theorem by analyzing maps between certain finite spectra arising from and various Thom spectra associated with it. To analyze these maps, we use the technique of cell diagrams, known results on the stable homotopy groups of spheres, and the -based Atiyah–Hirzebruch spectral sequence.

1. Introduction

1.1. The classification problem of simply connected 4-manifolds

A fundamental question in four-dimensional topology is the following:

Question 1.1.

How to classify all closed simply connected topological 4-manifolds?

To start our discussion, let be a simply connected topological 4-manifold. There are two important invariants of :

(1)

The intersection form : this is a symmetric unimodular bilinear form over given by the cup-product

(2)

The Kirby–Siebenmann invariant (defined in Reference KS77): this is an element in .

Question 1.1 was resolved by the following famous work of Freedman:

Theorem 1.2 (Freedman Reference Fre82).
(1)

Two closed simply connected topological 4-manifolds are homeomorphic if and only if their intersection forms are isomorphic and their Kirby–Siebenmann invariants are the same.

(2)

When the form is not even, any combination of the symmetric unimodular bilinear form and Kirby–Siebenmann invariant can be realized by a closed simply connected topological 4-manifold.

(3)

When the form is even, the combination can be realized if and only if the Kirby–Siebenmann invariant is equal to the signature of the form divided by 8 modulo 2. (Note that the signature of an even form must be divisible by . See Reference DK90, Section 1.1.3 for example.)

Therefore, given two manifolds, one can deduce whether they are homeomorphic or not by computing their intersection forms and Kirby–Siebenmann invariants. Moreover, Theorem 1.2 implies that any symmetric unimodular bilinear form can be realized by exactly two non-homeomorphic closed simply connected topological 4-manifolds if it is non-even, and by exactly one manifold if it is even.

We will now move on to the smooth category.

Question 1.3.

How to classify all closed simply connected smooth 4-manifolds?

By the works of Cairns, Whitehead, Munkres, Hirsch, and Kirby–Siebenmann Reference Cai35Reference Whi40Reference Mun60Reference Mun64bReference Mun64aReference Hir63Reference KS77, the Kirby–Siebenmann invariant of any smooth manifold, and in particular, a smooth 4-manifold, is zero. This fact, combined with Theorem 1.2, shows that two closed simply connected smooth 4-manifolds are homeomorphic if and only if they have isomorphic intersection forms. Therefore, Question 1.3 naturally breaks down into Questions 1.4 and 1.5:

Question 1.4.

Given a symmetric unimodular bilinear form , can it be realized as the intersection form of a closed simply connected smooth 4-manifold?

Question 1.5.

Suppose that the answer to Question 1.4 is yes; then how many non-diffeomorphic 4-manifolds can realize the given form?

In other words, Question 1.4 is asking which closed simply connected topological 4-manifolds admit a smooth structure. Question 1.5 is asking that if they do, how many different smooth structures do they admit. Topologists often refer Question 1.4 as the “Geography Problem” and Question 1.5 as the “Botany Problem”.

The main motivation of our work comes from the Geography Problem. In the past thirty years, starting with Donaldson’s groundbreaking work in Reference Don83, significant progress towards the resolution of the Geography Problem has been made.

Let’s divide symmetric unimodular bilinear forms over into two categories: the definite ones and the indefinite ones. For definite forms, a complete algebraic classification is still unknown. Nevertheless, Donaldson proved the following seminal theorem.

Theorem 1.6 (Donaldson’s diagonalizability theorem Reference Don83).

A definite symmetric unimodular bilinear form can be realized as the intersection form of a closed simply connected smooth 4-manifold if and only if can be represented by the matrix or .

This gives a complete answer to Question 1.4 in the case when is definite.

For indefinite forms, a powerful algebraic theorem of Hasse and Minkowski (see Reference Ser77) states that if is not even, it must be isomorphic to a diagonal form with entries , and if is even, it must be isomorphic to

for some and (for negative , denotes the direct sum of copies of ).

When the bilinear form is not even, by the theorem of Hasse and Minkowski, can always be realized by a connected sum of copies of and .

When the bilinear form is even, by Wu’s formula Reference Wu50, the closed simply connected 4-manifold realizing must be spin. Furthermore, by Rokhlin’s theorem Reference Roh52, the integer in Equation 1.1 must be even. By reversing the orientation of , we may assume that .

To this end, the following celebrated conjecture of Matsumoto Reference Mat82 serves as the last missing piece to this puzzle:

Conjecture 1.7 (The -conjecture, version 1).

The form

can be realized as the intersection form of a closed smooth spin -manifold if and only if .

Remark 1.8.

Note that Conjecture 1.7 is for general closed smooth spin -manifolds, which are not necessarily simply connected.

The “if” part of Conjecture 1.7 is straightforward: if , then the form can be realized by

Recall that the intersection forms of and are

respectively.

The “only if” part of Conjecture 1.7 can be reformulated as follows:

Conjecture 1.9 (The -conjecture, version 2).

Any closed smooth spin 4-manifold must satisfy the inequality

where and are the second Betti number and the signature of , respectively.

Definition 1.10.

An even symmetric unimodular bilinear form is spin realizable if it can be realized as the intersection form of a closed smooth spin 4-manifold.

By studying anti-self-dual Yang–Mills equations, Donaldson proved Conjecture 1.7 in the case when , under the additional assumption that has no -torsions Reference Don86Reference Don87. The condition on was later removed by Kronheimer Reference Kro94, who made use of the -symmetries in Seiberg–Witten theory. Later, Furuta combined Kronheimer’s approach with a technique called the “finite dimensional approximation” and proved the following significant result:

Theorem 1.11 (Furuta’s -theorem Reference Fur01).

For , the bilinear form

is spin realizable only if .

As we will explain in Section 1.2, Furuta proved Theorem 1.11 by studying a problem in equivariant stable homotopy theory (Question 1.17), which concerns the existence of certain stable -equivariant maps between representation spheres. The main purpose of this paper is to provide a complete answer to this -equivariant problem. A consequence of our main theorem (Theorem 1.21) is the following:

Theorem 1.12.

For , the bilinear form

is spin realizable only if

Corollary 1.13.

Any closed simply connected smooth spin -manifold that is not homeomorphic to , , or must satisfy the inequality

Proof.

Recall that the rank of is 8, and that the signatures of and are and , respectively. Therefore, Equation 1.2 is equivalent to the inequality

By Theorem 1.12, this is true when . By Theorem 1.6 and Theorem 1.11, the only exceptional cases are the following:

These cases correspond to , and by Theorem 1.2.

As we will see in Section 1.3, Corollary 1.13 is the “limit” of current methods towards resolving the -Conjecture using Bauer–Furuta invariants (see Remark 1.23).

1.2. Finite dimensional approximation in Seiberg–Witten theory

In this subsection, we will give a brief summary of Furuta’s proof of Theorem 1.11.

Let be a smooth spin 4-manifold. By doing surgery along essential loops in (which does not change its intersection form), we may assume that . The Seiberg–Witten equations (a set of first order nonlinear elliptic differential equations), together with the Coulomb gauge fixing condition, can be combined to produce a nonlinear continuous map

between two Hilbert spaces and . (See Reference BF04, Section 3, where the map is denoted by .) Instead of describing the map explicitly, we list three of its key properties:

(I)

can be decomposed into the sum , where is a Fredholm operator and is a nonlinear map that send any closed bounded subset of to a compact subset of .

(II)

There exist constants such that

where denotes the closed ball in with center and given radius.

(III)

The Lie group

acts on both and . Under these actions, the map is a -equivariant map.

By choosing a finite dimensional subspace of that is transverse to the image of and invariant under the -action, one can define the “approximated Seiberg–Witten map”

Here, and is the orthogonal projection. For , consider the set . By property (II) above and elliptic bootstrapping arguments, one can show that whenever is large enough, the following condition holds

Now, consider the representation spheres

and

Then by Equation 1.4, the map induces a -equivariant map

Applying , the map represents an element in , the -graded equivariant stable homotopy group of spheres. It was proved by Bauer and Furuta Reference BF04 that this element is independent with respect to the choices of auxiliary data (e.g., the Riemann metric and the spaces ) and is an invariant of the smooth structure on . This invariant is called the Bauer–Furuta invariant and is denoted by .

Theorem 1.14 is due to Furuta Reference Fur01. We include a sketch of proof for completeness.

Theorem 1.14 (Furuta Reference Fur01).
(1)

Suppose . Then

Here, is the four-dimensional representation of , with acting on it via left multiplication, and is a -dimensional representation such that the unit component acts as identity and the other component acts as negative identity.

(2)

The element fits into the commutative diagram

where and are stable homotopy classes that represents the inclusions and of fixed points.

Sketch of proof.
(1)

The -grading of is . This is the index of the operator and can be computed by the Atiyah–Singer index theorem.

(2)

By the specific definitions of , and are direct sums of and . Therefore, the elements and are both represented by some unstable maps from the space to the space for some . By the equivariant Hopf theorem Reference tD79, Section 8.4, the equivariant, unstable homotopy classes of such maps are determined by the mapping degrees of the induced maps on the -fixed points for the isotropy groups and (the isotropy group is not relevant because its Weyl group has dimension ). Therefore, to prove that , it suffices to show that they are equal when passing to the -geometric fixed points.

To prove this is indeed true, we note that when we restrict to the -fixed points, the nonlinear term vanishes and the map is the standard inclusion Therefore, the element , when passing to the -geometric fixed points, is represented by the standard inclusion from the space to the space . Since the same conclusion holds for , we see that and are equal to each other when passing to the -geometric fixed points.

Definition 1.15.

For , a Furuta–Mahowald class of level- is a stable map

that fits into the diagram

Using equivariant -theory, Furuta proved Theorem 1.16, from which Theorem 1.11 directly follows.

Theorem 1.16 (Furuta Reference Fur01).

A level- Furuta–Mahowald class exists only if .

1.3. Main theorem

At this point, it is natural to ask Question 1.17:

Question 1.17.

What is the necessary and sufficient condition for the existence of a level- Furuta–Mahowald class?

Remark 1.18.

We would now like to discuss the choice of the universe (i.e. the -representations that one stabilizes with respect to when passing from the space level to the spectrum level). In Furuta’s original proof of Theorem 1.16 Reference Fur01, he used the universe consisting of only the representations and , because this universe is the most relevant to the geometric problem. Modified proofs by Manolescu Reference Man14 and Bryan Reference Bry98, using divisibilities of the -theoretic Euler classes, show that the statement of Theorem 1.16 holds for any universe.

For Question 1.17, the answer could potentially depend on the choice of the universe. By works of Schmidt Reference Sch03, Theorem 2.6, Theorem 3.2 and Minami Reference Min, any Furuta–Maholwald class can be desuspended to the same diagram on the space level as long as . By the discussions in the previous paragraph, the bound in Theorem 1.16 holds for any universe. Therefore, a level- Furuta–Mahowald class in one universe can be desuspended to a space-level map , and then be further suspended to a level- Furuta–Mahowald class in any other universe. It follows that the answer to Question 1.17 does not depend on the choice of the universe.

Without loss of generality, we always work with the complete universe.

One might hope that the answer to Question 1.17 is because this would directly imply the -conjecture (Conjecture 1.7). Unfortunately, John Jones showed that this is false by exhibiting a counter-example for . See Reference FKMM07 for a more conceptual explanation of why such counter-examples must exist.

Subsequently, Jones proposed Conjecture 1.19:

Conjecture 1.19 (Jones Reference FKMM07).

For , a level- Furuta–Mahowald class exists if and only if

For the necessary condition, various progress has been made by Stolz Reference Sto89, Schmidt Reference Sch03, and Minami Reference Min. Before our paper, the best result is given by Furuta–Kametani:

Theorem 1.20 (Furuta–Kametani Reference FK).

For , a level- Furuta–Mahowald class exists only if

Much less is known about the sufficient condition for the existence of Furuta–Mahowald classes. So far, the best result is in Schmidt’s thesis Reference Sch03, in which Schmidt used -equivariant stable homotopy theory to attack Conjecture 1.19 for . In particular, Schmidt showed the existence of a level- Furuta–Mahowald class. This is also the first attempt to study this problem by using -equivariant stable homotopy theory.

In this paper, we completely resolve Question 1.17. Theorem 1.21 is the main result of our paper:

Theorem 1.21 (The limit is ).

For , a level- Furuta–Mahowald class exists if and only if

Remark 1.22.

The “only if” part of Theorem 1.21 directly implies Theorem 1.12 and Corollary 1.13.

Remark 1.23.

The “if” part of Theorem 1.21 implies that without further input from geometry or analysis, the best result one can achieve in proving Conjecture 1.9, using the existence of Furuta–Mahowald classes, is . Note that by Remark 1.18 this “limit” does not depend on the choice of the universe. In order to break this “limit” and to further attack the -conjecture, more delicate properties of the Seiberg–Witten map have to be studied. In particular, the Seiberg–Witten map should not be merely treated as a continuous map.

Remark 1.24.

Our answer differs from Conjecture 1.19 when . Note that in Reference Sch03, Schmidt proved that Conjecture 1.19 is true for . We came to a different conclusion for because there is a minor error in Schmidt’s computation (see Remark 10.2 for more details).

1.4. The -equivariant Mahowald invariant

Let be a finite group or a compact Lie group and let denote its real representation ring. One can consider , the -graded stable homotopy groups of spheres. Unlike the classical nonequivariant case, there are many non-nilpotent elements in . Here are some examples:

(1)

For each prime , the multiplication-by- map

between spheres with trivial -actions is non-nilpotent.

(2)

The geometric fix point functor induces a homomorphism

from the Burnside ring of to . Since preserves smash products, any preimage of the nonequivariant multiplication-by- map is also a non-nilpotent element in .

(3)

Let be a real irreducible representation of . The Euler class is the stable class in that represents the inclusion

of the fix points. Since all the powers of induce nonzero maps in equivariant homology, is non-nilpotent in .

Definition 1.25.

Suppose that and are elements in with non-nilpotent. The -equivariant Mahowald invariant of with respect to is the following set of elements in :

In other words, an element belongs to if the left diagram exists and the right diagram does not exist for any class .

Remark 1.26.

It is clear from Definition 1.25 that the -degree of each of the elements in is .

Historically, the -equivariant Mahowald invariant has been studied in many cases:

(1) Let be the cyclic group of order 2. The real representation ring of is

generated by the trivial representation 1 and the sign representation . The classical Borsuk–Ulam theorem in the unstable category is equivalent to the following statement when phrased in terms of the -equivariant Mahowald invariant:

Theorem 1.27 (Borsuk–Ulam).

For all , the -degree of is zero.

(2) Let . Consider the homomorphism

that is induced by the geometric fix point functor. For any non-equivariant class , consider all of its preimages under the map and their corresponding -equivariant Mahowald invariants with respect to the Euler class .

Among all the elements in , pick the element that has the highest degree in its -component. Then, apply the forgetful functor to the nonequivariant world. Bruner and Greenlees Reference BG95 proved that this construction produces the classical Mahowald invariant of , which has been studied extensively by Mahowald, Ravenel, and Behrens Reference MR93Reference Beh06Reference Beh07.

In particular, when and is a power of 2, Bredon Reference Bre67Reference Bre68 made conjectures about the degrees of the elements in for . His conjecture was proved by Landweber Reference Lan69, who used equivariant K-theory. Later, Bruner and Greenlees Reference BG95 translated Mahowald and Ravenel’s work Reference MR93 and obtained an independent proof of Bredon’s conjecture.

Theorem 1.28 (Landweber Reference Lan69, Mahowald–Ravenel Reference MR93).

For , the set contains the first nonzero element of Adams filtration . Moreover, the following 4-periodic result holds:

We would like to mention that Bredon–Löffler Reference Bre68Reference Bre67 and Mahowald–Ravenel Reference MR93 have independently made Conjecture 1.29:

Conjecture 1.29 (Bredon–Löffler, Mahowald–Ravenel).

For any non-equivariant class that is of positive degree, we have the inequality

Jones Reference Jon85 proved that for all non-equivariant classes of positive degrees. The -equivariant formulation of the classical Mahowald invariant gives a simpler proof of Jones’s result (see Reference BG95Reference Bru98, for example).

(3) Let , the cyclic group of order 4. The real representation ring of is

generated by the trivial representation 1, the sign representation , and the two-dimensional representation that corresponds to rotation by 90 degrees. The -equivariant Mahowald invariant of powers of with respect to has been studied by Crabb Reference Cra89, Schmidt Reference Sch03, and Stolz Reference Sto89.

Theorem 1.30 (Crabb Reference Cra89, Schmidt Reference Sch03, Stolz Reference Sto89).

For , the following 8-periodic result holds:

Since is a subgroup of , Theorem 1.30 was used by Minami Reference Min and Schmidt Reference Sch03 to deduce the existence of Furuta–Mahowald classes. Crabb Reference Cra89 also studied the -equivariant Mahowald invariant of powers of with respect to .

For our case, we are interested in the group and its irreducible representations and (defined in Theorem 1.14). By definition, it is clear that a level- Furuta–Mahowald class exists if and only if the -degree of

is greater than or equal to .

To prove our main theorem (Theorem 1.21), we translate it into a problem of analyzing the -equivariant Mahowald invariants of powers of with respect to . After this translation, our main theorem is equivalent to Theorem 1.31:

Theorem 1.31.

For , the following 16-periodic result holds:

Note that when ,

If the answer had been instead, then Theorem 1.31 would be an 8-periodic result and Jones conjecture (Conjecture 1.19) would be true. This deviation from Jones conjecture is explained in details in Step 6 of our proof (See Sections 2 and 10).

1.5. Summary of techniques

To resolve Question 1.17, which is a problem in -equivariant stable homotopy theory, we first translate it into a problem in non-equivariant stable homotopy theory. More specifically, we consider the sequence of maps

which are maps between certain Thom spectra over that are induced by inclusions of (virtual) subbundles. Given this sequence of maps, our -equivariant problem is equivalent to asking what is the maximal skeleton of each that maps trivially to . We call the “vanishing” line that connects these skeletons the Mahowald line. Intuitively, by drawing the cell diagrams for each , we can visualize the Mahowald line in Figure 1. See Section 2.1 for more details.

One can also form a Mahowald line for the computation of the classical Mahowald invariants for powers of 2. The analogous diagram to Figure 1 in the classical case has the cell diagram for in each column. Maps between the columns are the multiplication by 2 maps. The classical Mahowald line in this case is established by Mahowald–Ravenel by proving a lower bound and an upper bound for the line, and observing that they coincide. Our proof in the -equivariant case is in the same spirit as Mahowald–Ravenel. However, as we point out below, it is significantly more complicated and delicate than the classical arguments:

(1)

Classically, the lower bound is proved by using a theorem of Toda Reference Tod63, which states that 16 times the identity maps on certain 8-cell subquotients of are zero. This implies that the Mahowald line rises by at least 8 dimensions every time we move by four columns. In our situation, the analogue of Toda’s result does not hold. Therefore, our situation requires a more delicate inductive argument that gives us control over several cells above the Mahowald line (this control is not needed in the classical case).

(2)

Classically, the upper bound is proved via detection by the real connective -theory . In our case, this techniques does not work at , , which is the crux of the geometric application of our main theorem (Theorem 1.12 and Corollary 1.13). To handle this case, we need a careful study of both the -based and the sphere-based Atiyah–Hirzebruch spectral sequence of .

(3)

Classically, the lower bound and the upper bound are proven independently, and they happen to coincide. In our case, the proofs for the lower bound and the upper bound are not independent. More precisely, we first establish a rough lower bound in Step 1 (Section 2.3) and a rough upper bound in Step 2 (Section 2.4). These rough bounds do not coincide, but they do give us some information on the cells that are located in between them (Step 3, Section 2.5). Using this information, we refine the lower bound and the upper bound step-by-step, while updating information about the undetermined cells until the two bounds finally match each other (Steps 4–7, Sections 2.62.9).

1.6. Summary of contents

We now turn to give a summary of the paper. In Section 2, we provide an outline-of-proof for our main theorem (Theorem 1.21). We first reduce the -equivariant statement regarding the existence of a level- Furuta–Mahowald class into a non-equivariant statement (Proposition 2.2). The non-equivariant statement is determined by the location of the Mahowald line. Theorem 2.5 proves the exact location of the Mahowald line, from which our main theorem directly follows. Our proof of Theorem 2.5 consists of seven steps, described in Sections 2.32.9. The readers should regard Section 2 as a roadmap to the rest of the paper, as it contains all the main statements needed to prove Theorem 2.5.

In Section 3, we define maps between certain subquotients of that will be useful in the later sections. In Section 4, we prove certain attaching maps in . Sections 510 prove all the statements that are listed in Sections 2.32.9.

This paper has two appendices. Appendix A proves the combinatorial statements that are needed for the arguments in Sections 9 and 10. Appendix B recalls the definition of cell diagrams, a tool that we use for illustration purposes throughout the paper.

2. Outline of proof for main theorem

In this section, we give an outline of our proof for Theorem 1.21.

2.1. Equivariant to nonequivariant reduction

Consider the classifying space . There is a universal bundle

We let be the line bundle associated to the representation and set

Alternatively, there is a -action on the space , given by:

The quotient space of with respect to this -action is the classifying space . Given this, can also be defined as the line bundle that is associated to the principal bundle

Note that there is a fiber bundle

The cellular structure on (one cell in dimension for each ) and (one cell in dimensions , , ) induces a cellular structure on , and hence on . Given this cellular structure, we use to denote the subquotient of that contains all cells of dimensions between and . There are certain attaching maps between the cells in (see Figure 1). We prove these attaching maps in Section 4.

For , the inclusion of subbundles induces a map

Let

be the stabilization of the base-point preserving map that sends all of to the point in the space that is not the base-point. For , define the map to be the composition

We will also define the map to be the restriction of to the subcomplex :

Remark 2.1.

In general, there is no canonical choice of a cell decomposition of . For different choices, the skeleton (and hence the map ) will be different. However, we are only interested in whether is null homotopic or not. By cellular approximation, this does not depend on the choice of a cell decomposition. Hence we can use the specific cell decomposition described above.

Proposition 2.2.

A level- Furuta–Mahowald class exists if and only if the map

is zero.

Motivated by Proposition 2.2, we make Definition 2.3:

Definition 2.3.

The function is defined by setting to be the largest integer such that the map

is null-homotopic.

Definition 2.4.

The function can be visualized by drawing a line over the -cell in the cell-diagram of . When we connect these lines for all , the resulting “staircase” pattern is called the Mahowald line.

In light of Proposition 2.2, our goal is to find the exact location of the Mahowald line.

Theorem 2.5.

The function takes values as follows:

and for all ,

Theorem 2.5 directly implies Theorem 1.21. Our proof of Theorem 2.5 consists of seven steps, each giving a new bound on (see Figure 2):

(1)

Step 1 proves a lower bound for .

(2)

Step 2 proves an upper bound for . This upper bound agrees with the lower bound in Step 1 except at , .

(3)

Steps 3–5 prove that for all .

(4)

Step 6 proves that when is odd.

(5)

Step 7 proves that when is even.

Proof of Proposition 2.2.

Consider the diagram

In the diagram above, and . The left column is the cofiber sequence

where is the unit sphere of the representation . By our discussion in Section 1.2, a level- Furuta–Mahowald class exists if and only if there exists a map that makes diagram Equation 2.3 commute.

Since the first column is a cofiber sequence, exists if and only if the composition is null-homotopic. The Spanier–Whitehead dual of map is the map

Map is null-homotopic if and only if the map

is null-homotopic.

Map can be written as the composition

Note that is -free for all and acts trivially on . Therefore, is null-homotopic if and only if the nonequivariant map

is null-homotopic (see Theorem II.4.5 in Reference LMSM86). Here,

is the orbit. The maps and are induced by and , respectively.

Note that the restriction of the fiber bundle Equation 2.2 to gives the bundle

Therefore, the inclusion

is the inclusion of the -skeleton. This implies that

Under this identification, maps and are equal to and respectively. The map is exactly the composition map , which is null-homotopic if and only if a level- Furuta–Mahowald class exists.

2.2. The Mahowald line at odd primes and over

For each prime , we can localize the map at to obtain a map

Similar to Definition 2.3, we define the function as follows: is the largest integer such that the map

is null-homotopic. It is clear from this definition that for all ,

The line determined by the function is called the -local Mahowald line.

We show that, at any odd prime , the -local Mahowald line is above the 2-local Mahowald line (see Figures 1 and 3). This will reduce our problem to a 2-primary problem. After this subsection, we will focus on the case when we localize at the prime for the rest of the paper.

Recall the fiber bundle

As discussed in Section 2.1, the cell structures for and induce a cell structure for .

The standard cell structure for has one cell in dimensions , , and . The 2-cell is attached to the 1-cell by , which is invertible when localized at . Therefore,

This implies that when we localize at , there is a cellular structure for with only one cell in dimension 0, and no cells in other dimensions. Since the cell structure for has one cell in dimension for all , the induced cell structure for from the fiber bundle above also has one cell in dimension for all .

The bundle is orientable because its first Stiefel–Whitney class is 0. There is a Thom-isomorphism

This Thom-isomorphism implies that

It follows that there is a cell structure for with one cell in dimension for all . Note that by the cellular approximation theorem, Proposition 2.2 and Definition 2.3 do not depend on the cellular structure of . Therefore, we can use this specific cell structure to deduce a lower bound for the -local Mahowald line (see Figure 3). This lower bound is above the 2-local Mahowald line (shown in gray). Rationally, the lower bound for the rational Mahowald line is the same as the one for the -local Mahowald line.

2.3. Step 1: Lower bound

From now on, we localize at the prime . In the discussions below, the arrow denotes a map that induces an injection on -homology, and the arrow denotes a map that induces a surjection on -homology (see Definition 4.1).

Theorem 2.6.

For every , there exist maps

with the following properties (see Figure 4):

(i)

The diagram

commutes.

(ii)

The map induces an isomorphism on . In other words, is an -subcomplex of via the map (see Definition 4.1).

(iii)

The following diagram is commutative:

(iv)

Let be the restriction of to the bottom cell of . Then for , the map satisfies the inductive relation

where in and is some element in . We will show in Lemma 4.9 that and we set .

We prove Theorem 2.6 by using cell diagram chasing arguments.

Remark 2.7.

Property (i) immediately implies that the map

is null homotopic, and therefore it is the main property that we desire for . Properties (ii) and (iii) are added so that we can construct inductively from . Property (iv) is an additional requirement on that will be useful in Step 3.

Corollary 2.8.

For any and , we have the inequality

where

This line is shown in blue in Figure 4.

Proof.

When , the claim directly follows from diagram Equation 2.4. When , the claim follows from the case when and the following commutative diagram:

2.4. Step 2: Upper bound detected by

Using -equivariant theory, we prove Proposition 2.9:

Proposition 2.9.

For any , the composition

is nonzero.

Proposition 2.9 has Corollary 2.10:

Corollary 2.10.

The map is nontrivial.

Proof.

For the sake of contradiction, suppose that the map is trivial. Then the map

will factor through the quotient map via some map . Since no element in is detected by , the composition

is trivial. This is a contradiction to Proposition 2.9.

Corollary 2.11.

The equality

holds for all and . Here, is defined as in Corollary 2.8.

Proof.

Corollary 2.10 implies that

This directly implies that

for all . The claim follows by combining this inequality with the inequality in Corollary 2.8.

2.5. Step 3: Identifying the map on the first lock as

After establishing the lower bound for , the -cell and the -cell in will play significant roles for the rest of our argument. We call them the “first lock” and the “second lock”, respectively (see Figure 4).

In this step, we will focus on the first lock. Combining Theorem 2.6(iv) with an inductive Toda bracket computation, we prove Proposition 2.12, which will be essential in the proof of Proposition 2.16 and Proposition 2.20.

Proposition 2.12.

For all , we have the relations

Corollary 2.13 is a consequence of Proposition 2.12 and Theorem 2.6(i):

Corollary 2.13.

For all , the diagram

commutes.

Corollary 2.13 identifies the map on the first lock as .

2.6. Step 4: A technical lemma for the upper bound

To prove an upper bound for , we make use of the spectrum , which is defined as the fiber of the map

Here, is the 1-connected cover of . Proposition 2.14 is proved by analyzing the interactions between and the spectrum .

Proposition 2.14.

For any , the map

induced by the quotient map is injective.

Terminology 2.15.

Let be a CW spectrum that has at most one cell in each dimension. Recall that the cohomological -based Atiyah–Hirzebruch spectral sequence for has the following form:

Here, is the indexing set containing the dimensions of the cells of , is the cellular filtration of . The degrees for the -differentials are as follows:

Similarly, the homological -based Atiyah–Hirzebruch spectral sequence for has the following form:

Here, is the indexing set containing the dimensions of the cells of , is the cellular filtration of . The degrees for the -differentials are as follows:

Proposition 2.14 can be interpreted as follows: in the -based cohomological Atiyah–Hirzebruch spectral sequence of , any nonzero class of the form

survives. Using this, we can further show that in the -based Atiyah–Hirzebruch spectral sequence of , a nonzero class

with can only be killed by a differential of the form

where . Note that for , so this implies that a cell of dimension cannot support a differential with target .

2.7. Step 5: The second lock is not passed

Proposition 2.16.

There exists a map

with the following properties (see Figure 5):

(i)

The map

factors through the quotient map

via :

(ii)

The map factors through a quotient map

via a map

(iii)

The restriction of to its bottom cell is the map

(iv)

The map has order 2 in . In other words, the following composition is zero:

Properties (i) and (iii) in Proposition 2.16 are direct consequences of diagram Equation 2.6. Properties (ii) and (iv) are established by a local cell diagram chasing argument.

Lemma 2.17.

In the -based Atiyah–Hirzebruch spectral sequence of , there is a differential

where is a nonzero element in .

To prove Lemma 2.17, we first construct a map

that is of degree one on both the top and the bottom cell. Then, we prove a differential in by computing certain -invariants using the Chern character. Pulling back this differential to proves the desired differential.

Theorem 2.18.

The composition map

is not zero.

Proof.

For the sake of contradiction, suppose that is zero. Consider the composition

By Proposition 2.16(i), the map is the composition in the top row of the following diagram:

Since the sequence

is a cofiber sequence and ( has no negative homotopy groups), the map is zero.

Let be the pullback of under the composition

Let be the pullback of under the inclusion

Then the following three facts hold:

(i)

.

(ii)

pulls back to under the map

(iii)

.

Fact (i) is true by Proposition 2.16(iv). Fact (ii) is true because the map is zero. To see that fact (iii) is true, note that by Proposition 2.16(iii), can be represented as the map

Since is detected by , the composition

is nonzero. Proposition 2.14 then implies that .

Consider the following commutative diagram, where the rows are induced from cofiber sequences:

By fact (ii), for some . By the definition of and fact (iii), .

By Lemma 2.17, , where is the pullback of a nonzero element under the map

Since pulls pack to , . This implies that

(here denotes the 2-adic valuation). Therefore,

This is a contradiction because by Proposition 2.14.

Corollary 2.19.

We have the inequality

for all .

2.8. Step 6: The first lock is passed when is odd

In this step, we will show that when is odd, . To prove this, we first construct a spectrum for any . This spectrum is defined as the homotopy fiber of a certain map

The spectrum has bottom cell in dimension and top cell in dimension .

Proposition 2.20.

There exists a map

such that the following diagram commutes:

Proposition 2.21.

When is odd, the composition

is zero.

Proposition 2.21 is proven by considering , the -layer of the Adams tower for . Using the connectivity of the 0-connected cover of , we prove that there exists a differential of the form

in the -based Atiyah–Hirzebruch spectral sequence of . Moreover, is in the image of . By computing the -invariant of the element using Chern character, we show that .

It follows from Proposition 2.21 that the map

is also zero by the commutativity of diagram Equation 2.10.

Corollary 2.22.

When is odd, we have the inequality

2.9. Step 7: The first lock is not passed when is even

Proposition 2.23.

When is even, the class

is a permanent cycle in the -based Atiyah–Hirzebruch spectral sequence of .

The proof of Proposition 2.23 is sketched as follows: first, by restricting the map in Proposition 2.20 to the -skeleton, we obtain a map

where is constructed in Section 2.8. Then, we establish a permanent cycle

in the -based Atiyah–Hirzebruch spectral sequence for when is even via Chern character computations. This permanent cycle is then used to prove the desired permanent cycle.

Theorem 2.24.

When is even, the composition map

is not null.

Proof.

By Corollary 2.13, one can rewrite Equation 2.11 as the composition

For the sake of contradiction, suppose that Equation 2.12 is null-homotopic. By Proposition 2.14, there must exist a differential of the form

for some .

Recall that in Lemma 2.17, we established the differential

for some nonzero element . This, combined with differential Equation 2.13, shows that there exists a differential

Furthermore, the elements and , when considered as elements in , are equal. Since

and , must be the generator of .

Consider the exact sequence

that is induced from the cofiber sequence

Differential Equation 2.14 implies that the map

is zero. Therefore, the map

is injective. However, our induction hypothesis states that the composition map

is zero. The injection above will imply that the composition map

is also zero. This contradicts Theorem 2.18.

Corollary 2.25.

When is even, we have the equality

In light of Proposition 2.2, our main theorem (Theorem 1.21) follows directly from the various bounds that we have established for the Mahowald line (see Figure 2).

3. Preliminaries

In this section, we set up some preliminaries that will be useful in the later sections. In Section 3.1, we define maps between certain subquotients of . In Section 3.2, we discuss the transfer map.

3.1. Maps between subquotients

Definition 3.1.

Let , , and be integers with . The function is inductively defined as follows (see Figure 6):

when .

We also set .

Intuitively, the integer can be described as follows: start with the -cell in and walk to the right (towards ), moving down one cell every time we encounter an empty cell. The cell we reach at is .

Definition 3.2.

For and , define

where the function is defined as in Corollary 2.8. In other words, the -cell of is the first cell that is above the lower bound line proved in Section 2.3 (the blue line in Figure 6).

Proposition 3.3.

Let , , , be integers such that the following conditions hold:

(a)

and , where and ;

(b)

;

(c)

;

(d)

.

Then there exists a map

Furthermore, the maps are compatible with each other in the sense that the following three properties hold:

(1)

(Compatibility with respect to quotient). The following diagram commutes for all :

(2)

(Compatibility with respect to inclusion). If is another tuple satisfying the conditions above with and , then the following diagram commutes:

(3)

(Compatibility with respect to composition). If and are two tuples satisfying the conditions of the proposition, then

To avoid clustering the notations in the later sections, we will simply use the special arrow

to denote the map when the context is clear.

Proof.

We will construct the maps in four steps, increasing the level of generality at each step.

Step 1.

, . By our definition of and the cellular approximation theorem, there is always a map

Furthermore, this map makes the bottom square of the diagram

commute. Since both columns are cofiber sequences, there is an induced map

between the cofibers making the whole diagram commute. The top square of the commutative diagram above implies that property (1) holds for .

Step 2.

, . Note that by the definition of ,

We define the map to be the map

The map fits into the following commutative diagram:

Step 3.

. We define the map to be the composition

We now prove that property (2) holds when . The case when is directly implied by diagram Equation 3.2.

Suppose that . Consider the two compositions

and

in diagram Equation 3.1. We want to show that these two compositions are equal. After post-composing with the inclusion map

the maps and are homotopic to each other (this is because we have already verified Property (2) when ).

Consider the cofiber sequence

Since the difference is null after post-composing with the map

it factors through the fiber via a certain map

If the left vertical arrow in diagram Equation 3.1 is the identity map, then diagram Equation 3.1 commutes by definition. Otherwise, it is straightforward to check that the dimension of the top cell of is less than the dimension of the bottom cell in . Therefore, the map is zero by the cellular approximation theorem. This implies and that property (2) holds when .

Step 4.

General , , , . Choose a sequence , , …, such that

(1)

.

(2)

for all .

We define the map to be the composition

Note that by our discussion in Step 3, this composition does not depend on the choice of the sequence . Property (3) holds immediately by definition. Properties (1) and (2) hold by our discussions in Steps 1 and 3, respectively.

3.2. Transfer maps

Proposition 3.4.

There is a cofiber sequence

Proof.

The map can be rewritten as the map

which is induced by the map . The cofiber sequence

produces the cofiber sequence

Note that

This establishes the cofiber sequence Equation 3.3.

Let denote the rank-3 bundle over that is associated to the adjoint representation of on its Lie algebra .

Given a Lie group with a closed subgroup , there is a fiber bundle

Let (resp. ) be the vector bundle over (resp. ) associated to the adjoint representation on the Lie algebra. There is a well-known transfer map

that has been studied by Becker–Gottlieb Reference BG75, Becker–Schultz Reference BS78, and Bauer Reference Bau04. Now, set

(Recall that , as defined in Section 2.1, is the line bundle that is associated to the principal bundle .) We obtain a transfer map

Proposition 3.5.

The transfer map

induces an isomorphism on for all .

Proof.

Consider the pull back of under the inclusion map . We obtain the following commutative diagram:

Note that of all the spectra in the diagram above are .

Since map is induced by the inclusion of fiber of the bundle

and the Serre spectral sequence for this bundle collapses, map induces an isomorphism on . Moreover, map is the Pontryagin–Thom collapsing map, and it induces an isomorphism on . It follows from this that must induce an isomorphism on .

To prove that induces an isomorphism on for any , note that both and are modules over . Moreover, the induced map on -homology preserves this module structure. Therefore, the statement is reduced to proving an isomorphism for the case , which we have just proved.

We equip with the cell structure that has one cell in dimension for each .

Lemma 3.6.

is homotopy equivalent to .

Proof.

Let denote . We have the following equivalences:

Also,

where is the tautological bundle over . These equivalences imply that

Note the following general fact: given a vector bundle over , the attaching map in is given by . This fact can be proven by analyzing , which corresponds to the generator of .

We will now compute . By restricting the representations of to the subgroup , we deduce that under the map , the bundle pulls back to and the bundle pulls back to ( is the tautological bundle over ). Therefore,

and

It follows that . This completes the proof.

4. Attaching maps in

4.1. -subquotients

We recall Definition 4.1 and Lemma 4.2 from Reference WX17:

Definition 4.1.

Let , , and be CW spectra, and be maps

We say that is an -subcomplex of if the map induces an injection on mod 2 homology. An -subcomplex is denoted by a hooked arrow as above. Similarly, we say that is an -quotient complex of if the map induces a surjection on mod 2 homology. An -quotient complex is denoted by a double-headed arrow as above. When the maps involved are clear in the context, we may ignore the maps and and just say that is an -subcomplex of , and is an -quotient complex of .

Furthermore, is an -subquotient of if is either an -subcomplex of an -quotient complex of or an -quotient complex of an -subcomplex of .

Note that from Definition 4.1, -subcomplexes and -quotient complexes are not necessarily subcomplexes and quotient complexes on the point-set level. Our definitions should be thought of as in the homological or homotopical sense. A motivating example to illustrate this is the following: the top cell of the spectrum splits off, so there is a map from to that induces an injection on mod 2 homology. Therefore is an -subcomplex of in our sense. However, on the point-set level, the image of the attaching map is not a point and so is not a subcomplex of in the classical sense.

It follows directly from Definition 4.1 that if is an -subcomplex of , then the cofiber of is an -quotient complex of . We will often denote this quotient complex as . Dually, if is an -quotient complex of , then the fiber of is an -subcomplex of .

Lemma 4.2 is useful in constructing -subquotients.

Lemma 4.2.

Suppose that is an -subcomplex of . Let be the cofiber of and let be an -subcomplex of . Define to be the homotopy pullback of along . Then is an -subcomplex of . Moreover, is an -subcomplex of with quotient .

Dually, suppose is an -quotient complex of . Let be the fiber of . let be an -quotient complex of . Define to be the homotopy pushout of along . We have that is an -quotient complex of . Moreover, is an -quotient complex of with fiber .

Lemma 4.2 follows from the short exact sequences of homology induced by the following commutative diagrams of cofiber sequences and diagram chasing.

Definition 4.3.

For any element in the stable homotopy groups of spheres, we say that there is an -attaching map from dimension to dimension in a CW spectrum if is an -subquotient of . Here, is the degree of and is the cofiber of .

Lemma 4.4.

Suppose that is a CW spectrum, with only one cell in dimension . Then the following claims hold:

(1)

There is a 2-attaching map from dimension to dimension in if and only if the map

is nonzero.

(2)

There is an -attaching map from dimension to dimension in if and only if the map

is nonzero.

Proof.

This follows from naturality and the fact that in and in .

4.2. The and -attaching maps in

Recall that

Proposition 4.5.

The mod 2 homology of is as follows:

For ,

For ,

For ,

For ,

Proof.

When , , which is a bundle over with fiber . The corresponding Serre spectral sequence collapses at the -page, from which we obtain a computation for .

The homologies for all the other ’s follow from the homology of and the Thom isomorphism.

Recall from Proposition 3.4 that there is a cofiber sequence

for every .

Lemma 4.6.

The induced homomorphisms and on mod 2 homologies can be described as follows:

(1)

The map

is an isomorphism if and only if

and ;

and ;

and ;

and .

In other words, is an isomorphism when both the domain and the codomain are nonzero.

(2)

The map

is an isomorphism if and only if

and ;

and ;

and ;

and .

Intuitively, part of Lemma 4.6 is saying that for the cells in , the ones in dimensions come from , and the ones in dimensions go to .

Proof.

The proofs for both parts and follow from the associated long exact sequences on mod 2 homology groups from the cofiber sequence Equation 4.1.

Proposition 4.7.

In the mod 2 homology of ,

(1)

is nonzero if and only if

and ;

and ;

and ;

and .

(2)

is nonzero if and only if

and ;

and .

Proof.

Recall that is a bundle over with fiber . The existence of the ’s and the ’s in follows from the collapse of the Serre spectral sequence. More precisely,

where and . If we denote to be the total Steenrod squaring operation, then

To deduce the ’s and ’s in when , note that by the Thom isomorphism,

Here, is the Thom class associated with the virtual bundle . For any ,

where denotes the total Stiefel–Whitney class. Since

and , we have that

Substituting this into equation Equation 4.2 and letting take values from elements in produce all the ’s and ’s in .

Corollary 4.8.

There are 2 and -attaching maps in if and only if they are marked in Figure 7.

Proof.

The 2 and -attaching maps follow from Lemma 4.4 and Proposition 4.7.

Lemma 4.9.

Suppose that and satisfy one of following conditions:

and ;

and ;

and ;

and .

Then the map

is .

Proof.

By Lemma 4.6, the cofiber of the map is

Since there is a nonzero in its cohomology, this cofiber is indeed .

4.3. -Attaching maps in

Proposition 4.10.

There is an -attaching map in from dimension to dimension if and only if it is one of the following four cases (see Figure 7):

and ;

and ;

and ;

and .

Proof.

For dimension reasons, there are eight cases of possible -attaching maps in total. We need to show that of these eight cases, four cases have -attaching maps and four cases don’t. Recall that , generated by .

Case 1.

and . Consider the map

By Corollary 4.8, the cells in dimension are not attached to the lower skeletons of and . Therefore, they are -subcomplexes. Taking cofibers, we have the following commutative diagram:

Since is a 2 cell complex, it must be the cofiber of a class in the stable homotopy groups of spheres.

It is clear that we must have . If it is not, then would split as , and we would have a map

whose restriction to the bottom cell is by Lemma 4.9. This is not possible.

Case 2.

and . Consider the map

From the 2 and -attaching maps in Corollary 4.8, this map is the Spanier–Whitehead dual (up to suspension) of the map

in the case when and . Therefore, we must have the -attaching map.

Case 3.

and . The proof is similar to the case when and . Consider the map

By Corollary 4.8, the cells in dimension are not attached to the lower skeletons of and . Therefore, they are -subcomplexes. Taking the cofibers, we have the following commutative diagram:

Since is a 2 cell complex, it must be the cofiber of a class in the stable homotopy groups of spheres.

It is clear that we must have . If it is not, then would split as , and we would have a map

By Lemma 4.9, post-composing this map with the quotient map would give , which is not possible.

Case 4.

and . Consider the map

From the 2 and -attaching maps in Corollary 4.8, this is the Spanier–Whitehead dual (up to suspension) of the map

in the case when and . Therefore, we must have the -attaching map. Alternatively, one may also prove this -attaching map by considering the map

Now, we will show that in the other four cases, there do not exist -attaching maps.

Case 1.

and . Consider the map

By Corollary 4.8, the cells in dimension are not attached to the lower skeletons of and . Therefore, they are -subcomplexes. Taking the cofibers, we have the following commutative diagram:

Since is a 2 cell complex, it must be the cofiber of a class in the stable homotopy groups of spheres.

It is clear that we must have . Otherwise, we would have and there would be a map

Post-composing this map with the quotient map gives us the identity map. This is not possible.

Case 2.

and . Consider the map

From the 2 and -attaching maps in Corollary 4.8, this is the Spanier–Whitehead dual (up to suspension) of the map

in the case and . Therefore, there cannot be an -attaching map.

Case 3.

and . Consider the map

By Corollary 4.8, the cells in dimension are not attached to the lower skeletons of and . Therefore, they are -subcomplexes. Taking cofibers, we have the following commutative diagram:

Since is a 2 cell complex, it must be the cofiber of a class in the stable homotopy groups of spheres.

It is clear that we must have . Otherwise, if , we would have a map

whose restriction on the bottom cell is the identity. This is not possible.

Case 4.

and . Consider the map

From the 2 and -attaching maps in Corollary 4.8, this is the Spanier–Whitehead dual (up to suspension) of the map

in the case when and . Therefore, there cannot be an -attaching map.

4.4. Periodicity in

Proposition 4.11.

For any , there is an equivalence

Proof.

Given any two -representations and , there is a cofiber sequence

Let and . The cofiber sequence

produces the cofiber sequence

This cofiber sequence can be rewritten as

Here, and denote the bundles over that are associated to the representations and , respectively. From this, we deduce that

Let be the -skeleton of . We have the equality

To finish the proof, it suffices to show that the bundle is stably trivial. Note that since , this bundle is spin and can be classified by a stable map

Moreover, since , can be further be lifted to . It follows that because is -connected.

4.5. Some -subquotients of

In this subsection, we define and discuss some -subquotients of .

We start with the 3 cell complex and the 4 cell complex .

Lemma 4.12.

The 3 cell complex splits:

Proof.

By Corollary 4.8 and Proposition 4.10, there are no and -attaching maps in . The claim then follows from the fact that and are generated by and respectively.

Lemma 4.13.

The -cell complex splits:

Proof.

Consider the -skeleton of , which is the 3 cell complex . By Corollary 4.8 and Proposition 4.10, there are no and -attaching maps in . Since and are generated by and respectively, we have the following equivalence:

This gives as an -subcomplex of , and, therefore, as an -subcomplex of .

Now consider the attaching map

whose cofiber is . By Corollary 4.8, the cell in dimension is not attached to the cell in dimension by . It is also not attached to the cell in dimension by . Therefore, it is null homotopic and we have the following homotopy equivalence:

This gives as an -subcomplex of .

By Lemma 4.2, we can pullback along the quotient map

and obtain a 2 cell complex as an -subcomplex of .

We claim that this 2 cell complex must be . In fact, consider the map

induced by the map . Since there is a nontrivial on , we must have a nontrivial on and the 2 cell complex. This produces the -attaching map. Therefore, is an -subcomplex of .

In summary, we have shown that both and are -subcomplexes of . Their wedge gives an isomorphism on mod 2 homology and is therefore a homotopy equivalence. This completes the proof of the lemma.

Proposition 4.14.

There exists a 4 cell complex that is an -subcomplex of . It has cells in dimensions , , and .

Proof.

First, by Corollary 4.8, the cells in dimensions and are not attached by in . Therefore, there is an equivalence

In particular, we have as an -quotient complex of and , and as an -subcomplex of and .

Define to be the fiber of the following composition:

Then is a 3 cell complex with cells in dimensions and . This 3 cell complex is an -subcomplex of and . It is clear that we have the following commutative diagram in the homotopy category:

Therefore, we can identify the 4 cell complex

Now, we claim that the top cell of splits off. In fact, consider the attaching map

whose cofiber is . We will show that this attaching map is null-homotopic. Consider the -page of the Atiyah–Hirzebruch spectral sequence of the 3 cell complex that converges to its -homotopy groups:

The right hand side is generated by

By Corollary 4.8 and Proposition 4.10, there are no and -attaching maps in . This proves our claim.

Therefore, we have a splitting

In particular, this splitting exhibits as an -subcomplex of

Lastly, we pullback along the quotient map

By Lemma 4.2, is an -subcomplex of with cells in dimensions and . This concludes the proof of the proposition.

Definition 4.15.

Define to be the 4 cell complex in Proposition 4.14. Define to be the -skeleton of . Define

and to be its -skeleton.

It is clear from Proposition 4.10 that

Proposition 4.16.

There is a 2 cell complex with cells in dimensions and , such that it is an -quotient complex of .

Proof.

It suffices to show that has an -subcomplex with cells in dimensions and .

Firstly, by Corollary 4.8, we know that is an -subcomplex of . Secondly, by Corollary 4.8 and the fact that and , we know that is an -subcomplex of . Therefore, by Lemma 4.2, we have the following diagram and in particular we may define .

We then complete the proof by defining to be the cofiber of the map

5. Step 1: Proof of Theorem 2.6

In this section, we present the proof of Theorem 2.6, which states that: For every , there exist maps

with the following properties:

(i)

The diagram

commutes.

(ii)

The map induces an isomorphism on . In other words, is an -subcomplex of via the map .

(iii)

The following diagram is commutative:

(iv)

Let be the restriction of to the bottom cell of . Then for , the map satisfies the inductive relation

where in and is some element in . Note that by Lemma 4.9 and we set .

5.1. An outline of the proof

In this subsection, we list the main steps of our proof of Theorem 2.6 (see Figure 8). The intuition is explained later in Remark 5.6.

We need to show the existence of 4 families of maps

for all that satisfy two commutative diagrams, namely the ones in (i) and (iii) of Theorem 2.6, a property for , namely (ii) of Theorem 2.6 and a property for , namely (iv) of Theorem 2.6.

The strategy of our proof can be summarized as the following. We first prove the existence of the maps for all , and then construct the maps for all . We check that satisfies property (ii) in Theorem 2.6. This is Step 1.1 and Step 1.2 of our proof.

In the rest of the proof, we show inductively the existence of the maps and , and that the two diagrams in (i) and (iii) of Theorem 2.6 commute.

We first define to be the zero map and show the existence of . We check that the two diagrams in (i) and (ii) of Theorem 2.6 commute. This is Step 1.3 that gives the starting case .

Next, we assume the maps and exist and the two diagrams in (i) and (ii) of Theorem 2.6 commute for the 4 maps . We define the map and show the existence of , using information in the induction. Note that there are choices for . This is Step 1.4.

Then, we check that the two diagrams in (i) and (ii) of Theorem 2.6 commute for the 4 maps , for all choices of . This is Step 1.5.

Finally, in Step 1.6, we prove that there exists one choice of , such that it satisfies an inductive relation between the restriction of to the bottom cell of their domains. For this choice of , this establishes property (iv) and finishes the proof.

More precisely, the details of Steps 1.11.6 are stated as the following.

Step 1.1.

We establish the existence of the maps for all (see Figure 9).

Proposition 5.1.

For every , there exists a map that fits into the following commutative diagram

The proof of Proposition 5.1 is an extensive and careful study of the cell structures of the columns between and and in dimensions between and . It involves the computation of stable stems in the range . We define as the composition

Step 1.2.

Using Proposition 5.1 and the homotopy extension property, which is stated as Lemma 5.12 in Section 5.3, we show the existence of two maps and in Proposition 5.2.

Proposition 5.2.

For every , there exist maps that fit into the following commutative diagram:

Moreover, the map induces an isomorphism on . In other words, is an -subcomplex of .

We define the map as the following composite

Note here we use the octahedron axiom to identify with . It then follows from Proposition 5.2 that the map induces an isomorphism on , which establishes property (ii) in Theorem 2.6.

Step 1.3.

We define

to be the zero map. Note that the 3 cells of are in dimensions , , , so this is the only choice. Since , the following diagram (iii) in Theorem 2.6 for commutes regardless of the construction of .

For the existence of the map , it suffices to show the following composite is zero.

This is true because this map factors through by cellular approximation. This gives the following commutative diagram (i) in Theorem 2.6 for .

This gives the starting case of our inductive argument.

Step 1.4.

For , we assume the maps and exist, the two diagrams in (i) and (iii) of Theorem 2.6 commute for the 4 maps , and satisfies property (iv) in Theorem 2.6.

We define the map to be the composite

Using the commutative diagram Equation 2.5 in (iii) of Theorem 2.6 for the case , we have Proposition 5.3:

Proposition 5.3.

The following composite is zero.

Note that the first map is the inclusion of the bottom cell of , and that the map is established in Step 1.2 before the induction.

As a result, there exist maps

that fit into the following commutative diagram:

Note that there are many choices of that makes the diagram Equation 5.6 commute.

Step 1.5.

In this step, we prove Proposition 5.4.

Proposition 5.4.

For any choice of in Step 1.4, the two diagrams Equation 2.4 and Equation 2.5 in (i) and (iii) of Theorem 2.6 commute for the 4 maps .

The proof is a straightforward cell diagram chasing argument.

Step 1.6.

In this step, we prove Proposition 5.5.

Proposition 5.5.

Let be the restriction of to the bottom cell of . Then there exists one choice of in Step 1.4 such that the following property is satisfied:

where and . Note that by Lemma 4.9 and we set .

This proves that this choice of satisfies the relation in (iv) of Theorem 2.6 and therefore completes the induction.

Remark 5.6.

The critical part of Theorem 2.6 is the existence of the map . We want to prove it inductively. Namely, we assume that exists and want to show that exists. This induction would follow easily if the following map were zero:

However, this is not true. Intuitively, the -cell in maps nontrivially to the -cell in by . More precisely, one can show that the map Equation 5.8 factors through as an -quotient, and the latter map in the following composite

is detected by in the Atiyah–Hirzebruch spectral sequence of . Therefore, we have to show the composite

is zero. It turns out that we can show the composite of the latter two maps in Equation 5.9 is zero. This follows from a technical condition that can be chosen to satisfy:

factors through .

Here note that is an -subcomplex of . In fact, this is due to the composite

corresponding to an element in the group .

Now to complete the induction, we need to show that can be chosen to satisfy:

factors through .

Firstly, in , the -cell is only attached to the cells in dimensions and , all of which map trivially to . As a result, we can choose such that the restriction factors through .

Secondly, by some local arguments that involve attaching maps in for , …, , we can show that can be chosen such that factors through .

This allows us to complete the induction and to prove Theorem 2.6. See Figure 10 for an illustration of the discussion above.

We’d like to comment that our actual argument is a little different from our discussion above. We actually analyze instead of . This is used to deduce the inductive relation Equation 5.7, based on which we identify the first lock.

In the remaining subsections of this section, we will prove Propositions 5.1-5.4 one by one.

5.2. Proof of Proposition 5.1

The proof of Proposition 5.1 consists of many steps. The goal is to construct a map

such that it is compatible with the map

Since the top cell of is in dimension , we have the maps

So roughly speaking, we want to show that the bottom 3 cells of map trivially to , and the image of does not involve the cells in . Our strategy is to carefully study the cell structures of the intermediate columns of finite complexes, and to get rid of certain cells gradually.

Step 1.1.1.

In this step, we focus on column . We use the -attaching maps in column between cells in dimensions and , and , to get rid of the cell in dimension of , and to lower the upper bound of the image to dimension in column . More precisely, we prove Lemma 5.7.

Lemma 5.7.

There exists the following commutative diagram (see Figure 11):

Proof.

Firstly, we have the following commutative diagram:

By Lemma 4.9, we have that the map in middle of the top row of diagram Equation 5.11 is . By Corollary 4.8, we have an -attaching map in between the cells in dimensions and . This corresponds to an Atiyah–Hirzebruch differential

Therefore, the composition of the maps in the top row of diagram Equation 5.11 is zero. In particular, pre-composing with the map

is also zero. By the cofiber sequence of the right most column, we know that the map from to maps through .

Secondly, we have the following commutative diagram:

By Lemma 4.9, we have that the map in middle of the bottom row of diagram Equation 5.12 is . By Corollary 4.8, we have an -attaching map in between the cells in dimensions and . This corresponds to an Atiyah–Hirzebruch differential

Therefore, the composition of the maps in the bottom row of diagram Equation 5.12 is zero. In particular, post-composing with the map

is also zero. By the cofiber sequence of the left most column, we know that the map from to factors through .

This gives the required map

Remark 5.8.

We will use arguments similar to the ones in the proof of Lemma 5.7 many times in the rest of this paper. Instead of presenting all details in terms of commutative diagrams, we will simply refer them as “similar arguments as in the proof of Lemma Equation 5.10” or “cell diagram chasing arguments” due to certain attaching maps.

Step 1.1.2.

In this step, we focus on column . We show that in , the cells in dimensions and map through in column . More precisely, we have Lemma 5.9.

Lemma 5.9.

There exists the following commutative diagram:

Proof.

By Proposition 4.10, there is no -attaching map in . This shows that

We may therefore consider the cells in dimensions and separately.

For , it maps naturally through the -skeleton in column . By Proposition 4.10, there is an -attaching map in between the cells in dimensions and . A similar argument as in the proof of Lemma Equation 5.10 shows that maps through in column .

For , firstly note that by Corollary 4.8, there is an -attaching map in between the cells in dimensions and . A similar argument as in the proof of Lemma Equation 5.10 shows that maps through the -skeleton in column . Then it maps naturally through the -skeleton in column . To see that it actually maps through , we only need to show the following composite is zero.

This is in fact true, since .

Combining both parts, this gives the required map

We enlarge Diagram Equation 5.13 to Diagram Equation 5.14. We will establish the maps and in Steps 1.1.3, 1.1.4 and 1.1.5:

Step 1.1.3.

In this step, we establish the map , making the triangle under in Diagram Equation 5.14 commute.

By Lemma 4.9, we have that the map

is mapping into the bottom cell of . Since

the composition of maps in the bottom row of Diagram Equation 5.14 is zero. In particular, post-composing with the map

is also zero. By the cofiber sequence of the left most column, we know that the map from to factors through , which gives the desired map , making the triangle under commute.

Note that we haven’t shown the triangle above commutes. We will show it later in Step 1.1.5.

Step 1.1.4.

In this step, we establish the map , making the parallelogram below in Diagram Equation 5.14 commute.

By the cofiber sequence in the left most column, it suffices to show the following composite is zero.

Since both the triangle under and the upper rectangle in Diagram Equation 5.14 commute, it is equivalent to show that the following composite is zero.

This is in fact true, since the composition of the latter two maps is already zero.

Lemma 5.10.

The following composite in Diagram Equation 5.14 is zero.

Proof.

We first show that the left map factors through the bottom cell of the codomain. In fact, the composite

corresponds to an element in . Since , the group . Therefore, it must factor through the bottom cell . We have the following commutative diagram.

By Lemma 4.9, the map in the bottom row of Diagram Equation 5.15 is . Since

this completes the proof.

Step 1.1.5.

In this step, we establish the map , making all parts of Diagram Equation 5.14 commute.

It suffices to show Lemma 5.11.

Lemma 5.11.

The following composite is zero.

In fact, by Lemma 5.11 and Step 4, the following composite is zero.

Then by the cofiber sequence in the right most column of Diagram Equation 5.14, the map must map through , establishing the desired map .

To see that all parts of Diagram Equation 5.14 commute, first note that by Lemma 5.11 and Lemma 5.10, both the triangles above the map and under the map commute. Next, by the construction of the map , the triangles above it commute. Finally, by Step 1.1.3 and the cofiber sequence of the left most column in Diagram Equation 5.14, the triangle under the map commutes. Therefore, all parts of Diagram Equation 5.14 commute.

Now, let’s prove Lemma 5.11.

Proof of Lemma 5.11.

The composite in the statement splits into the following two composites.

For the first composite Equation 5.16, let’s study the second map . By Proposition 4.10 and Corollary 4.8, is a 3 cell complex, with cells in dimensions , and with a 2 and -attaching map. Since , there is a nonzero differential

in the Atiyah–Hirzebruch spectral sequence of . It follows that the second map must map through its -skeleton: . Since , the map must further map through and the composite Equation 5.16 can be decomposed as

Therefore, due to the relation

the first composite Equation 5.16 is zero.

For the second composite Equation 5.17, the second map must map through the -skeleton of , which is . Then it follows from the relation

that the second composite Equation 5.17 is zero. This completes the proof.

Now we claim that the map is our desired map in Proposition 5.1. In fact, part of Diagram Equation 5.14 gives us the following commutative diagram (see Figure 12).

Putting Diagrams Equation 5.10 and Equation 5.18 together, we have the following commutative diagram.

Forgetting some terms in this diagram, we obtain Diagram Equation 5.3 in Proposition 5.1.

5.3. Proof of Proposition 5.2

Lemma 5.12 is essentially the homotopy extension property.

Lemma 5.12.

Suppose that we have the following commutative diagram in the stable homotopy category

where and are the cofibers of the maps and respectively. Then it can be extended into the following commutative diagram:

Proof.

We can first extend the commutative diagram Equation 5.19 to the following commutative diagram:

Note that the map is not unique in general. We choose one and stick with our choice. Since the composite

is the zero map, there exists a map

making the diagram commute.

Now consider the map

The map is not zero in general. If it were zero, we then have the commutative diagram as requested.

The fix is to modify the map . Note that the composite

is the zero map. Therefore, by the cofiber sequence

there exists a map

such that . We define the map

Then the following diagram commutes as requested.

In fact, we have that

From the commutative diagram Equation 5.3 in Proposition 5.1 and the definitions of and , we have the following commutative diagram

By Lemma 5.12, we can extend it to the following commutative diagram

Removing the terms , we have the commutative diagram Equation 5.4 in Proposition 5.2. It is clear that the map induces an isomorphism on . In other words, is an -subcomplex of . This completes the proof of Proposition 5.2 (see Figure 13).

5.4. Proof of Proposition 5.3

In this subsection, we prove Proposition 5.3 that for , the following composite is zero.

We start with the commutative diagram Equation 2.5 for the case in (iii) of Theorem 2.6. We enlarge the commutative diagram Equation 2.5 for the case in the following way

We next state a lemma about the map , whose proof we postpone until the end of this subsection. This Lemma 5.13 will also be used in Section 5.6.

Lemma 5.13.

There exists a map

that fits into the following commutative diagram

Putting these two diagrams Equation 5.20 and Equation 5.21 together, we obtain the following commutative diagram (see Figure 14)

It is clear that Proposition 5.3 follows from Lemma 5.14, Lemma 5.15 and the above commutative diagram.

Lemma 5.14.

The following composite

factors through , giving the map in the diagram Equation 5.22.

Lemma 5.15.

The following composite is zero.

We first prove Lemma 5.14 and Lemma 5.15, and then prove Lemma 5.13.

Proof of Lemma 5.14.

By Lemma 4.12, the 3 cell complex splits as

To show that the map

maps through , we need to check the following composite is zero.

This composite corresponds to an element in the group

The last equation follows from the fact that . This completes the proof.

Proof of Lemma 5.15.

By Lemma 4.12, the 3 cell complex splits as

Therefore, the composite

corresponds to an element in the group

The last equation follows from the facts that

This completes the proof.

Now we present the proof of Lemma 5.13.

Proof of Lemma 5.13.

From the cofiber sequence

we need to show that the composite

is zero. By Proposition 4.10, is a 2 cell complex with an -attaching map:

Our strategy to show the composite Equation 5.23 being zero is to first deal with the bottom cell and then the top cell.

By the cellular approximation theorem, the restriction of the composite Equation 5.23 to the bottom cell of maps through the bottom cell of , by either or . The possibility of is ruled out by a cell diagram chasing argument due to the -attaching map between the cells in dimensions and in .

Therefore, the composite Equation 5.23 factors through the top cell of . We can further require it factor through the top 2 cells of , namely

By the cellular approximation theorem, it maps through the -skeleton of . Note that there is no cell in dimension in , so it maps through the 4 cell complex . We have the following commutative diagram.

To prove this lemma, it suffices to show the following composite is zero.

Firstly, post-composing with the quotient map

must be zero. This is due to the fact that it maps through the mod 2 Moore spectrum. Therefore, the composite Equation 5.24 must map through the -skeleton of , namely the 3 cell complex :

Now let’s consider the Atiyah–Hirzebruch filtration of this map Equation 5.25. It cannot be detected in filtration , since there is a nontrivial differential in the Atiyah–Hirzebruch spectral sequence of :

which is due to the -attaching map by Proposition 4.10. If it is detected in filtration , then it must be zero since . Therefore, if it is nonzero, then it must be detected by . In this case, post-composing with the inclusion to is zero, due to the -attaching map between the cells in dimensions and , and therefore the Atiyah–Hirzebruch differential

In sum, regardless of the actual Atiyah–Hirzebruch filtration of the map Equation 5.25, the following composite is always zero.

This completes the proof of the lemma.

5.5. Proof of Proposition 5.4

We check that the two diagrams Equation 2.4 and Equation 2.5 in (i) and (iii) of Theorem 2.6 commute for the 4 maps .

For the diagram Equation 2.4 in (i) of Theorem 2.6 for the case , we put together the following commutative diagrams

diagram Equation 5.6 in Step 1.4,

diagram Equation 2.4 in (i) of Theorem 2.6 for the case ,

the upper right corner of diagram Equation 5.4 in Proposition 5.2.

The commutativity of the upper left corner of this diagram is due to the compatibility of each columns.

For the diagram Equation 2.5 in (iii) of Theorem 2.6 for the case , we put together the following commutative diagrams

diagram Equation 5.6 in Step 1.4,

the lower half of diagram Equation 5.4 in Proposition 5.2.

By the definitions of in Step 1.2 and in Step 1.4, the composites in the left and right columns give us and respectively.

Therefore, we have the diagram Equation 2.5 in (iii) of Theorem 2.6 for the case . This completes the proof.

5.6. Proof of Proposition 5.5

In this subsection, we prove Proposition 5.5: There exists one choice of in Step 1.4 such that

where is the restriction of to the bottom cell of , and (see Figure 15). Note that by Lemma 4.9, and we set .

Consider the following composite

By Lemma 4.12, the 3 cell complex splits:

Therefore, the composite Equation 5.27 can be written as the sum of the following two composites Equation 5.28 and Equation 5.29.

For the composite Equation 5.28, first note that the map equals zero when restricted to bottom cell of . In fact, it corresponds to an element in

which follows from the fact that .

Next note that the composite

restricts to on the bottom cell of . Therefore, we have the following commutative diagram:

where with an element in that is annihilated by multiplication by 2, namely 0 or .

For the composite Equation 5.29, by the diagram Equation 2.5 for the case , we can rewrite it as

Using the splitting

we can rewrite the composite Equation 5.31 as the sum of the following two composites Equation 5.32 and Equation 5.33.

The composite Equation 5.32 is zero. In fact, since and

the composition of the first two maps in Equation 5.32 is already zero. Therefore, the composite Equation 5.31 can be identified as Equation 5.33.

For the composite Equation 5.33, we have Lemma 5.16.

Lemma 5.16.

The following composite is zero:

Proof.

Consider the following diagram.

Pre-composing the composite Equation 5.34 with the inclusion of the bottom cell of gives us the zero map. This is due to the fact that .

The map from to can be written as a Toda bracket of the form

where generated by , and generated by . For a precise argument of this fact, we refer to Lemma 5.3 of Reference WX18.

The indeterminacy of this Toda bracket is

since . We claim that this Toda bracket contains zero; therefore it is zero as a set. This completes the proof of the lemma.

In fact, the only potential nonzero element that this Toda bracket contains is

The corresponding Massey product

in filtration 9 of the Adams -page, which is higher than all nonzero elements in the Adams -page. Therefore, this potential nonzero element is also zero.

By Lemma 5.16, the composite Equation 5.33 maps through the bottom cell of , and we have the following commutative diagram:

Since , the following composite is zero.

Therefore, the composite Equation 5.33 further factors through the top cell of . We denote by the corresponding element in .

Removing some of the terms in Equation 5.35, we obtain the following diagram:

Adding the diagrams Equation 5.36 and Equation 5.30 together, we have the following commutative diagram

which can be enlarged into the following commutative diagram:

Using the homotopy extension property that we proved, namely Lemma 5.12, we have the following commutative diagram.

Note that the map induces an isomorphism on and therefore is an H-subcomplex. In sum, we have constructed a choice of the map that satisfies the condition Equation 5.7 in Proposition 5.5. This completes the proof of Proposition 5.5.

6. Step 2: Upper bound detected by

In this section, we prove Proposition 2.9:

Proposition 6.1 (Proposition 2.9).

For any , the composition

is nonzero.

Recall that is the homotopy orbit of the free -action on

Therefore, we have the following isomorphism:

6.1. Some results about the -equivariant -theory

In this subsection, we list some results about the group for various . These results are established in Reference Sch03, Section 5 (see also Reference Lin15).

(I)

There is a commutative and associative multiplication map (given by tensor product of virtual bundles)

(II)

There is a ring isomorphism

The generators are defined as follows:

(a)

.

(b)

, where is a -dimensional real representation. The representation space of is , with the unit component of acting via left multiplication and acting as reflection along the diagonal.

(c)

.

(III)

There are elements (called Euler classes)

They satisfy the following property: for any and , the map

equals the map on that is induced by the inclusion

(IV)

There are elements (called Bott classes)

such that the following maps are isomorphism for all and :

(V)

The relation

holds.

(VI)

The following relations hold:

(VII)

There is an isomorphism

generated by the elements and , .

6.2. Proof of Proposition 2.9

Let

be the base-point preserving map that sends the entire to the point in that is not the base-point. Consider the composition

where is induced by the inclusion

Lemma 6.2.

The map

sends to a nonzero element.

Proof.

Consider the map

that is induced by . By (III), . By (IV) and (VII), we have an isomorphism

generated by the elements and , . Here, is the unique element in such that . By (VI) and (V), we have

To finish the proof, it suffices to show that

We will prove this by contradiction. Suppose Equation 6.1 is not true. Consider the cofiber sequence

that is obtained from by taking . This cofiber sequence induces the sequence

which is exact in the middle. Since

there exists an element such that

By (IV) and (VII), can be written as

for some polynomial . By (VI) and (V), equation Equation 6.2 can be rewritten as

This implies that

By comparing the coefficients of and , we see that this is impossible.

By definition, under the isomorphism

the element corresponds to the element . Therefore, we have the following commutative diagram:

In the commutative diagram above, the left vertical map sends to . Therefore, Lemma 6.2 implies that the map

is nontrivial. This finishes the proof of Proposition 2.9.

Recall that the restriction of the map to the bottom cell of its domain is denoted

(see Theorem 2.6). Corollary 6.3 will be used in the next section:

Corollary 6.3.

For , the map is detected by .

Proof.

For the sake of contradiction, suppose that is not detected by . Then the composition

is zero. Since the map factors through , the composition

is zero. Moreover, since , the composition

is also zero.

By Proposition 2.9, the map is detected by . This maps factors through the map , which, as we have just shown, is not detected by . This is a contradiction.

7. Step 3: Identifying the map on the first lock as

In this section, we prove Proposition 2.12: For all , we have the relations

Combining Corollary 6.3 and part (iv) of Theorem 2.6, we have shown that the family

satisfies the following two properties:

(1)

For , can be detected by ;

(2)

For , we have that

for some in and . Here .

Since , generated by the Hurewicz image of the element in of the sphere spectrum, we make Definition 7.1 due to property of the family above.

Definition 7.1.

Define

and for ,

It is clear that the Hurewicz image of in is zero for all .

Then Proposition 2.12 follows from Lemma 7.2 for the elements in .

Lemma 7.2.

For all , the following relations hold:

Proof of Proposition 2.12.

By Definition 7.1 and Lemma 7.2, we have

Now we prove Lemma 7.2.

Proof of Lemma 7.2.

We first show that the elements are for all .

Suppose that for some , we have . Then we would have

where . Since no elements in and can be detected by the ring spectrum , mapping the relation Equation 7.2 to gives us in . This contradicts property that is detected by . Therefore, we must have

for all .

Substituting , the relation Equation 7.1 becomes

Here we set to unify the notation.

We have the Massey product

on the Adams -page with zero indeterminacy for all . Then by Moss’s theorem Reference Mos70, Theorem 1.2, we have the Toda bracket for all :

Therefore, we have

for all .

Using this relation, we complete the proof of Lemma 7.2 by induction on , which states that for all

The cases are trivial, since both and are zero.

For , suppose the lemma holds for and .

Multiplying , We have

Note that both and are divisible by . Since

we have

The indeterminacy

is zero, since and that

by induction. Therefore, we have that

for all . This completes the induction and therefore the proof of the lemma.

8. Step 4: A technical lemma for the upper bound

In this section, we prove Proposition 8.1, which is Proposition 2.14 in Section 2.

Proposition 8.1.

For any , the map

induced by the quotient map is injective.

The proof makes essential use of two spectra, and , which we review now.

8.1. The spectra and

By Atiyah–Bott–Shapiro Reference ABS64, Section 11, any spin bundle is -orientable. In other words, the spectrum is a module over the ring spectrum . By Ando–Hopkins–Rezk Reference AHR, the orientation map can be refined to an -map. Let be the cofiber of the rationalization map

Both and are modules over . Therefore, for any spin bundle of dimension over a space , we have the Thom isomorphism

which is induced by cup product with the Thom class.

Moreover, if is a subspace of , and is the restriction of to , we also have the relative Thom isomorphism

Lemma 8.2.

Let be a virtual bundle over a space , and let be a spin bundle of dimension over . Suppose that is a subspace of . Let and be the restrictions of and to . We have the Thom isomorphism

The isomorphism above is natural in the sense that if is a subspace of , and , are the restrictions of and to , then the following diagram commutes:

Proof.

The desired isomorphism follows from the isomorphism Equation 8.2 by setting

and letting be the pull-back of to (here, and denote the disc bundle and the sphere bundle of , respectively). Diagram Equation 8.4 follows from standard arguments on the point-set level.

Next, we introduce a slight variant of the spectrum : we define as the fiber of the map

Note that while . The map gives a map that induces isomorphism on for any . This proves the following simple lemma:

Lemma 8.3.

Let be a finite CW-spectrum with no cell of dimension . Then .

These two spectra and are related via Lemma 8.4:

Lemma 8.4.

Let be the 0-connected cover of . There is a map

that induces an injection on for any positive integer .

Proof.

Consider following commutative diagram

In the commutative diagram above, the columns form cofiber sequences. By the -Lemma Reference May01, Lemma 2.6, we can extend this diagram to the following diagram

where all the rows and columns are cofiber sequences.

Now, consider the commutative diagram

Since is -connected and for , the composition equals zero. Therefore, the composition factors through the fiber of , and there exists a map

making the diagram above commute. The composition

is our desired map.

To prove that induces an injection on , first note that induces an injection on because . Furthermore, since for all , the map induces an isomorphism on (just like the map ). Therefore, induces an isomorphism on . It follows that induces an injection on .

8.2. Proof of Proposition 2.14

Note that is the Thom spectrum

Set

Since is spin, is spin. By Lemma 8.2, we obtain Thom isomorphisms that fit into the following commutative diagram:

Set where is the bundle associated to the adjoint representation of . Recall that there is a transfer map

that induces isomorphism on (see Proposition 3.5) for any integer . Truncating this map, we obtain a commutative diagram:

For algebraic reasons, the -based Atiyah–Hirzebruch spectral sequence of collapses. Therefore, the map induces injection on . By diagram Equation 8.6, the pinch map also induces injection on . By Equation 8.5, the pinch map induces an injection

Now we relate and : the map

in Lemma 8.4 provides us with the following diagram:

Since both and are injective, the map is injective as well.

Finally, since both and have no and cells, and are identical for them. It follows that the map

is injective. By Lemma 8.3, the map

is also injective, as desired.

9. Step 5: Upper bound

9.1. Proving differentials using the Chern character

In this subsection, we introduce a useful technique for proving differentials in the -based Atiyah–Hirzebruch spectral sequence.

Definition 9.1.

A finite -spectrum is called -injective if the map

given by is injective. Here, denotes the complexification of .

Theorem 9.2.

Let be a finite CW-spectrum that satisfies the following properties:

(1)

has a single top cell in dimension ;

(2)

has no cells in dimension ;

(3)

The -skeleton of is -injective;

(4)

The -skeleton of is homotopy equivalent to .

Furthermore, suppose there is an element that satisfies the equality

Then in the -based Atiyah–Hirzebruch spectral sequence of , the following results hold:

(I)

If , then the class is a permanent cycle. Here, when is even and when is odd.

(II)

If , then there is a nontrivial differential

for some .

To prove Theorem 9.2, we first introduce some lemmas.

Lemma 9.3.

Let be the pull-back of under the inclusion map . Then .

Proof.

Recall that we have the equality

for all . Since ,

By our assumption, is -injective (property (3)). Therefore , as desired.

Lemma 9.4.

In the -based Atiyah–Hirzebruch spectral sequence for , the element is a permanent cycle.

Proof.

The cofiber sequences

and

induce the following commutative diagram:

Consider the element . By Lemma 9.3, . This implies that there exists an element such that

Furthermore, because of the commutative diagram

Since the map

is ,

for some .

We claim that . To see this, consider the composition

Since (property (4)), this map is induced by and sends to . Therefore, under the composition , is sent to

Therefore .

Consider the following commutative diagram:

The bottom horizontal arrow is an equivalence because of Lemma 8.3. By the previous discussion, . Therefore, the left vertical arrow sends the element to as well. This is equivalent to saying that element is permanent cycle in the -based Atiyah–Hirzebruch spectral sequence for .

Lemma 9.5.

is -injective.

Proof.

Let be an element in with . Since is -injective, the pulls-back of under the inclusion must be zero. Therefore, is the pull-back of some element

under the pinch map . Since

must be 0. It follows that and is -injective, as desired.

Proposition 9.6.

The element is a permanent cycle in the -based Atiyah–Hirzebruch spectral sequence of if and only if .

Proof.

If , then we can find an element such that

Given this element , we have the equality

where is induced from the pinch map . Using Lemma 9.5, we can prove that is a permanent cycle by the exact same argument as the proof of Lemma 9.4.

Now, suppose that . Consider the commutative diagram

To prove that is not a permanent cycle, it suffices to show that the element is not sent to under the right vertical map.

For the sake of contradiction, suppose that is sent to . Consider the following diagram:

Since , there exists an element such that by the exactness of the left column.

Let . Since the diagram is commutative,

It follows that .

Consider the element . We have

Since is -injective, the element equals for some

By comparing the Chern character, we obtain . This is impossible because .

Proof of Theorem 9.2.

The claim follows directly from Lemma 9.4 and Proposition 9.6.

9.2. Proof of Proposition 2.16

For , we define to be the composite

Then diagram Equation 2.8 follows directly from diagram Equation 2.4.

By Lemma 4.13, we have a splitting

Under this splitting, we can write

where and are the following two composites Equation 9.3 and Equation 9.4.

We will show the following claims on and . These claims directly imply Properties (ii) through (iv).

Claim 1.

.

Claim 2.

is of order 2 in . In other words, the following composite is zero.

Claim 3.

The restriction of to the bottom cell is

in .

It is clear that by Corollary 2.13 in Step 2 in Section 2.4 that Claim 3 is true. In the rest of this subsection, we first prove Claim 1, and then prove Claim 2.

For Claim 1, note that equals the composite

By exactly the same cell diagram chasing argument as the one in Step 1.1.2, we see that the restriction of the composite

to the bottom cell is zero. Therefore, we can rewrite as the composite

for some map . By cellular approximation theorem, the map maps through :

Moreover, due to the -attaching map in between the cells in dimensions and , the composite

must be zero. Therefore, the map maps through , and we can rewrite as the composite

for some map . By Theorem 2.6, there is an H-subcomplex

By Lemma 4.12, the 3 cell complex splits:

Since , we have

and the map maps through the H-subcomplex . In other words, we can rewrite the composite

as the composite

for some map in . Therefore we can rewrite as the composite

As in the proof of Proposition 5.3, the composite

is zero. Therefore, we have . This completes the proof of Claim 1.

For Claim 2, note that the composite maps through . Due to the -attaching map in between the cells in dimensions and , the composite

is zero. Therefore, we can rewrite as the composite

where is a map that induces a trivial homomorphism on . By the cellular approximation theorem and the -attaching map in between cells of dimensions and , the map maps through :

Therefore, we can rewrite as the composite

By Lemma 4.12, the 3 cell complex splits:

So we can write as the sum of the following two composites Equation 9.6 and Equation 9.7:

For the map in the composite Equation 9.6, it corresponds to an element in the group

which is generated by on the bottom cell of . Since is not detected by the spectrum , post-composing Equation 9.6 with the map is zero.

For the composite Equation 9.7, note that by Part (iii) of Theorem 2.6, the composite is

Therefore, the composite Equation 9.7 can be rewritten as

Using again the splitting

the composite Equation 9.8 can be written as the sum of the following two composites Equation 9.9 and Equation 9.10:

The composite Equation 9.9 is zero, since corresponds to an element in

The composite Equation 9.10 is zero, since corresponds to an element in

In fact, , which is generated by and . Both generators are annihilated by .

Therefore, the composite Equation 9.7, which equals the composite Equation 9.8, is zero.

In sum, we have that in . This finishes the proof of Claim 2.

9.3. Proof of Lemma 2.17

Recall that there is a map

that induces an isomorphism on for any (see formula Equation 3.3). Truncating this map, we obtain a map

where

Here, denotes the canonical bundle on .

The Thom isomorphism gives an identification

where and is the Thom class for homology.

In order to apply Theorem 9.2 to , we require Lemma 9.7:

Lemma 9.7.

For any odd integer and any , the spectrum is -injective. (See Definition 9.1.)

Proof.

We show that for the spectrum , where is odd and , the map

is injective. Since the Chern character map is injective for this spectrum, this would prove the lemma by Definition 9.1.

The complexification of real vector bundles corresponds to the following map on the spectra level

For degree reasons, the -based Atiyah–Hirzebruch spectral sequence for collapses at the -page. In particular, the group is a direct sum of copies of ’s.

Since is odd, the bottom two cells of are . More generally, we can decompose by its subquotients (with certain attaching maps among them) of the form for , and with one possible copy of when is odd. In this case, we have that is divisible by 4. Since

the -based Atiyah–Hirzebruch spectral sequence for collapses at the -page. This means that we only need to check that the following maps are injective

where and is divisible by 4.

Due to the compatibility of real and complex Bott periodicity, the map

maps to in . So in particular, it induces an isomorphism on for all . It is also well known that the generator of maps to in . So it induces an injective homomorphism on for all . This proves that the map Equation 9.12 is injective.

For the map Equation 9.11, since the Spanier–Whitehead dual of is , we may rewrite it as

which is an inclusion of a splitting summand.

Combining the injectivity of the maps Equation 9.11 and Equation 9.12 completes the proof of the lemma.

Lemma 9.8.

There exists an element such that

for some with .

Proof.

There is a Thom isomorphism

where and is the -theoretic Thom class for the virtual bundle . We have the relations

and

Now, suppose

where for all . Our goal is to determine the coefficients so that condition Equation 9.13 holds.

Applying to both sides of the equation and using the formulas above, we get

Now, make the substitution . Then and the above equation becomes

Condition Equation 9.13 requires

for some with . By comparing the constant terms in Equation 9.14, we deduce that and

Let the power series expansion of be . By comparing the coefficients of in the equation above, we obtain the relations

By Lemma A.2, we see that . By Lemma A.3, we see that for all . Therefore, belongs to .

Now, set . Then one has

By Lemma 9.7, we can apply Theorem 9.2 to and conclude the existence of the differential

in the -based Atiyah–Hirzebruch spectral sequence of , with in . By naturality of Atiyah–Hirzebruch spectral sequence, we can pullback this differential to using the map . This finishes the proof of Lemma 2.17.

10. Steps 6 and 7: First lock and second lock

In this section, we will prove the claims in Section 2.8 and Section 2.9.

10.1. Construction of

In this subsection, we will construct a spectrum for every . This spectrum will be crucial for proving Proposition 2.23 and Proposition 2.21. By Proposition 3.4, there is a cofiber sequence

By restricting to the subquotient , we obtain a cofiber sequence

Consider the quotient map

By Proposition 4.16, there is a 2 cell complex with cells in dimensions and such that it is an -quotient complex of . There is a commutative diagram

where the left vertical map is the composition

By the -Lemma Reference May01, Lemma 2.6, we can extend this commutative diagram to the following commutative diagram, where the rows and columns are cofiber sequences:

The complex is defined to be the cofiber of the map

By Lemma 4.6(2), the map induces an isomorphism on for all .

Lemma 10.1.

The complex satisfies the following properties:

(1)

;

(2)

Proof.

Property (1) is straightforward from the definition of . To prove property (2), note that by truncating the transfer map (see Equation 3.4)

we obtain an -sub map

Desuspending by and applying Proposition 4.11, we obtain the map

By truncating the map , we obtain an -quotient map

The composite

induces an isomorphism on -homology. Therefore, it is a homotopy equivalence. The claims now follow from Lemma 3.6.

Remark 10.2.

In the proof of Reference Sch03, Theorem 4.9, Schmidt made a minor error when computing . This error led to Schmidt’s proof of Jones Conjecture for .

Note that Lemma 10.1 is a crucial step in our proof of showing that the Jones conjecture is not true when . If Schmidt’s cohomotopy group computation were true, our statement of Lemma 10.1(2) would be different: . This would also lead to an affirmative answer for Jones conjecture for by using our subsequent arguments.

Lemma 10.3.

For any , the -skeleton of is -injective.

Proof.

Note that for some . Therefore, the claim follows from Lemma 9.7.

10.2. Proof of Proposition 2.20

Consider the map

in Proposition 2.16. By properties (ii) and (iii) in Proposition 2.16, there is a factorization of the map as follows:

Here, the vertical map is the restriction of the quotient map

to the -skeleton.

When restricted to the -skeleton, diagram Equation 2.8 becomes the diagram

This diagram, combined with the factorization above, produces the following diagram:

Given this commutative diagram, Proposition 2.20 follows from Lemma 10.4.

Lemma 10.4.

The following diagram commutes:

Proof.

Let denote the composition

and let denote the composition

We want to show that the map becomes 0 after post-composing with the map .

It is straightforward to see that when restricted to the subcomplex , . This is because both and become the quotient map

This implies that the map factors through the fiber of the inclusion map , which is :

Given any map , the composition map

is 0 because and . Since , the composition

is zero. This implies that , as desired.

10.3. Bundles with simple Chern character

In this subsection, we will construct virtual bundles over with simple Chern characters. This will allow us to use Theorem 9.2 to establish differentials in the Atiyah–Hirzebruch spectral sequence.

Recall from Section 9 the spectrum , which is defined as the Thom spectrum

By definition, is the fiber of a certain -quotient map

We denote the generator of by .

Lemma 10.5.

There exists an element such that

with and .

Proof.

There is a Thom isomorphism

where and is the -theoretic Thom class for the virtual bundle . We have the relations

and

Suppose

After taking Chern characters on both sides, we get

Just like before, we make the substitution . With this substitution, equation Equation 10.1 is equivalent to the following equation:

This equation is equivalent to the equation

By comparing coefficients on both sides of equation Equation 10.3, we obtain the relations

for all . By Lemma A.8, for all . Therefore, the coefficients and we have found a that satisfies equation Equation 10.1.

To show that the rest of the coefficients in satisfy the conditions of the lemma, note that by the definition of the coefficients ,

Subtracting equation Equation 10.2 from this equation and using the relation , we obtain the following equation:

Substituting back as , the above equation becomes

After rearranging, we get

Expanding the left hand side and comparing the coefficients of and on both sides of the equation, we obtain the relations

By Lemma A.6,

By Lemmas A.5, A.6, and A.7, when is odd, all three terms in the formula for are 2-local integers, so . When is even, the lemmas show that the first term is a 2-local integer while the other two terms are 2-local half-integers (they have 2-adic valuations ), and so again. This concludes the proof of the lemma.

Proposition 10.6.

There exists an element such that

(1)

When is even,

(2)

When is odd,

with .

Proof.

When is even, let be the pullback of under the map and let ( is the restriction map). By Lemma 10.5,

Recall from Lemma 10.1 that for even . Let

be the generators for the first and the second summand, respectively. Since the composition map

is multiplication by 1 when is even and multiplication by 2 when is odd, we have

and

Now, set

where

is the quotient map. Note that this construction is valid because both and belong to by Lemma 10.5. It follows that satisfies Equation 10.5.

When is odd, let be the pullback of under the map and let . By Lemma 10.5,

Recall from Lemma 10.1 that for odd. There is an element

such that

for some with (this is because the -invariant of has 2-adic evaluation ).

Now, set

where

is the quotient map. Then

By Lemma 10.5, has 2-adic valuation 0. Therefore, satisfies Equation 10.5, as desired.

10.4. First lock for odd

In this subsection, we will prove Proposition 2.21, which states that when is odd, the composition

is zero.

Proof of Proposition 2.21.

Let be the boundary map induced from the cofiber sequence

In other words, fits into the sequence

We will show that the following diagram is commutative:

Our proposition will follow from the commutativity of this diagram. This is because taking in the cofiber sequence

produces the sequence

In this sequence, the element

first maps to

and then maps to

by the commutativity of Equation 10.6. Since the sequence is exact at , we deduce that .

It remains for us to prove that diagram Equation 10.6 is commutative. Since has no 0-cells, the Adams filtration for the map is at least 1. This implies that the Adams filtration of the map is at least . Therefore, the map can be lifted through a map , where () is the th stage of the Adams tower of .

The cells of are in dimensions , , …, , and . Since for all , the -skeleton of maps trivially to under the composition map

Therefore, there exists a map such that the following diagram is commutative:

Let be the composition

To finish the proof of our proposition, it suffices to show that .

Since the Adams filtration of is at least , can be , , , or . We will compute the -invariant of and show that . This will finish the proof because the 2-adic valuations for the -invariants of the four possibilities above are

Consider the diagram

By the definition of the -invariant, there exists an element such that

This implies that when we pullback along the map , the Chern character is equal to

In Proposition 10.6, we constructed an element with Chern character

Subtracting equation Equation 10.8 from equation Equation 10.7, we get

In particular, this shows that when we restrict to the -skeleton ,

By Lemma 10.3,

Therefore,

for some . Here, is the quotient map . The Chern character of is

where . From the relation

we deduce that . Since and , . This concludes the proof of the proposition.

10.5. First lock for even

Proof of Proposition 2.23.

In Proposition 10.6, we showed that there exists an element such that

By Lemma 10.3, we can apply Theorem 9.2 to . Theorem 9.2 shows that the element

is a permanent cycle in the -based Atiyah–Hirzebruch spectral sequence of .

The map constructed in Section 10.1 induces a map of spectral sequences from the -based Atiyah–Hirzebruch spectral sequence of to that of . Therefore, the element

is also a permanent cycle in the -based Atiyah–Hirzebruch spectral sequence of and . This finishes the proof of the proposition.

Appendix A. Coefficients of

Let be the coefficient of in the power series expansion of

In this section, we prove several facts about the 2-adic valuations of that we are going to use in the rest of the paper.

Notation A.1.

For any , let be the 2-adic valuation of . For example, , , and .

In the power series expansion of

the coefficient for is

where the sum ranges through all tuples such that

(1)

for all ;

(2)

;

(3)

.

In all the cases that we are interested in, will always be at most , so the tuple will always be finite. Each tuple corresponds to the monomial

The number is the coefficient of this monomial, which is

Here,

In particular, this number is an integer.

Lemma A.2.

for all .

Proof.

For any tuple with and , the valuation

except when . Since

the valuation is equal to .

Lemma A.3.

The inequality holds for all and .

Proof.

For any positive integer , we have the inequality

Equality is achieved only when . This implies that

From this, we deduce that for all .

For , given any tuple with and , the valuation

except when . Since

the 2-adic valuation of the denominator is still at least . Therefore, .

Lemma A.4.

for all .

Proof.

The coefficient of the monomial in is

The valuation of this number is exactly . We will prove that the coefficients of all the other monomials in of degree have 2-adic valuations strictly larger than .

Consider the monomial

where only finitely many of the ’s are nonzero and . To prove our claim above, it suffices to show that the fraction

is an even 2-local integer.

This fraction is equal to

The condition essentially guarantees that the product of the first two terms is an even integer when differs from . There are two exception cases. They are and .

For the first exception case, the product is

The product of the first two terms is odd, but the last term is , which is even.

For the second exception case, the product is

The product of the first two terms is odd, but the last term is

which is even again. Therefore, , as desired.

Lemma A.5.

for all .

Proof.

The proof is very similar to the proof of Lemma A.4. Given a monomial in of degree , the smallest 2-adic valuation of its coefficient is achieved when . This coefficient is

Its 2-adic valuation is .

To prove that the 2-adic valuations of all the other coefficients are strictly bigger than this number, we make a similar computation to the proof of Lemma A.4 and reduce the problem into showing that the ratio

is even when and . The product of the first two terms is an even number.

Lemma A.6.

for all .

Proof.

The proof for this is again similar to the proof of Lemma A.4 and Lemma A.5. We claim that the smallest 2-adic valuation is achieved only when , , and for all . The corresponding coefficient is

The 2-adic valuation for this number is . To prove that all the other coefficients have bigger valuations, we need to show that the ratio

is even for all the other tuples such that . The product of the first two terms will always be an even number except when . For this exceptional case, the ratio is

The product of the first two terms is odd but the last term is

which is even.

Lemma A.7.

We have

Proof.

To prove the lemma, it suffices to consider all the coefficients in and whose valuation is at most . For , they are the following:

All the other coefficients have 2-adic valuations at least . For , only the term

will matter. All the other coefficients have 2-adic valuations at least .

We have

When is even, is odd, and the 2-adic valuation of the last expression is exactly . When is odd, is even, and the 2-adic valuation of the last expression is at least . This proves the lemma.

Lemma A.8.

For a fixed , the inequality holds for all .

Proof.

We claim that the 2-adic valuations of all the coefficients for satisfy . We will divide the proof into four cases:

Case 1.

There exist , such that , in the tuple . Consider the ratio

Since and ,

and the last expression is even. Therefore, the 2-adic valuation of the coefficient is at least .

Case 2.

There exists only one such that , and that is at least 2. Consider the ratio

Since and ,

and the last expression is even.

Case 3.

There exists only one such that , and that is 1. Consider the ratio

where we have used the facts that and . Let , and . Then and

The term

in the last expression is equal to . This number is an integer for all positive integers where and .

Case 4.

There exists no such that . Consider the ratio

Since exactly one of and is even and , the number

is always an integer.

Appendix B. Cell diagrams and the Atiyah–Hirzebruch spectral sequence

The theory of cell diagrams is a very powerful tool when thinking of finite CW spectra. See Reference BJM84Reference WX17Reference Xu16 for example. We use them as illustration purpose in our paper. In this section, we recall the definition of cell diagrams from Reference BJM84 and talk about its connection to the Atiyah–Hirzebruch spectral sequence.

Definition B.1.

Let be a finite CW spectrum. A cell diagram for consists of nodes and edges. The nodes are in 1-1 correspondence with a chosen basis of the mod 2 homology of , and may be labeled with symbols to indicate the dimension. When two nodes are joined by an edge, then it is possible to form an -subquotient

which is the cofiber of with certain suspension. Here , the attaching map, is an element in the stable homotopy groups of spheres. For simplicity, we do not draw an edge if the corresponding is null.

Suppose we have two nodes labeled and with , and there is no edge joining them. Then there are two possibilities.

The first one is that there is an integer , and a sequence of nodes labeled , with , and edges joining the nodes to the nodes . In this case we do not assert that there is an -subquotient of the form above; this does not imply that there is no such -subquotient.

The second one is that there is no such sequence as in the first case. In this case, there exists an -subquotient which a wedge of spheres .

Remark B.2.

In Reference BJM84’s original definition, they use subquotients instead of -subquotients.

Example B.3 shows the indeterminacy of cell diagrams associated to a given CW spectrum.

Example B.3.

Let be the composite of the following two maps:

where the second map is the inclusion of the bottom cell. Consider : the cofiber of , which is a 3 cell complex with the following cell diagram:

It is clear that the top cell of splits off, since can be divided by . So we do not have to draw any attaching map from the cell in dimension 3 to the one in dimension 0. Note that the cofiber of is in fact an -subcomplex of .

We give two more interesting examples.

Example B.4.

Consider the suspension spectrum of . It is a 3 cell complex with cells in dimensions 2, 4 and 6. It was shown by Adams Reference Ada58 that the secondary cohomology operation , which is associated to the relation

is nonzero on this spectrum. In other words, there exists an attaching map between the cells in dimensions 2 and 6, which is detected by in the 3-stem of the Adams page. Note that detects two homotopy classes: . Their difference is , which is divisible by . Therefore, we have its cell diagram as the following:

We can also consider the Spanier–Whitehead dual of the suspension spectrum of . It is a 3 cell complex with cells in dimensions -2, -4 and -6, with the following cell diagram

In a way, the attaching maps drawn in the cell diagram of a CW spectrum correspond to certain differentials in its Atiyah–Hirzebruch spectral sequence. We illustrate this idea through Example B.4. For notations regarding the Atiyah–Hirzebruch spectral sequence, we refer to Terminology 2.15 and Sections 3 and 6 of Reference WX17.

Example B.5.

For the suspension spectrum of , the attaching map corresponds to the -differential

and its multiples

for any element in the stable stems, in the Atiyah–Hirzebruch spectral sequence of . The -attaching map then corresponds to the -differential

and its multiples. Note that , which is already killed by a -differential. Therefore is a permanent cycle.

For its Spanier–Whitehead dual, the attaching map corresponds to the -differential

and its multiples. For the -attaching map, it does not correspond to a -differential

since already supports a nonzero -differential so it is not present at the -page anymore. However, this -differential still “exists”, in the sense that some of its multiples still exist. More precisely, suppose that is an element in the stable stems such that . Then survives to the -page and we have a -differential

which might or might not be zero, depending on whether is zero. For example, we have a nonzero -differential

Acknowledgments

The authors would like to thank Mark Behrens, Rob Bruner, Simon Donaldson, Houhong Fan, Dan Isaksen, Achim Krause, Peter Kronheimer, Ciprian Manolescu, Haynes Miller, Tom Mrowka, Doug Ravenel, and Guozhen Wang for helpful conversions. The authors would also like to thank Zilin Jiang and Yufei Zhao for writing a program in the early stages of the project to check their combinatorial results in Appendix A.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. 1.1. The classification problem of simply connected 4-manifolds
    2. Question 1.1.
    3. Theorem 1.2 (Freedman Fre82).
    4. Question 1.3.
    5. Question 1.4.
    6. Question 1.5.
    7. Theorem 1.6 (Donaldson’s diagonalizability theorem Don83).
    8. Conjecture 1.7 (The -conjecture, version 1).
    9. Conjecture 1.9 (The -conjecture, version 2).
    10. Definition 1.10.
    11. Theorem 1.11 (Furuta’s -theorem Fur01).
    12. Theorem 1.12.
    13. Corollary 1.13.
    14. 1.2. Finite dimensional approximation in Seiberg–Witten theory
    15. Theorem 1.14 (Furuta Fur01).
    16. Definition 1.15.
    17. Theorem 1.16 (Furuta Fur01).
    18. 1.3. Main theorem
    19. Question 1.17.
    20. Conjecture 1.19 (Jones FKMM07).
    21. Theorem 1.20 (Furuta–Kametani FK).
    22. Theorem 1.21 (The limit is ).
    23. 1.4. The -equivariant Mahowald invariant
    24. Definition 1.25.
    25. Theorem 1.27 (Borsuk–Ulam).
    26. Theorem 1.28 (Landweber Lan69, Mahowald–Ravenel MR93).
    27. Conjecture 1.29 (Bredon–Löffler, Mahowald–Ravenel).
    28. Theorem 1.30 (Crabb Cra89, Schmidt Sch03, Stolz Sto89).
    29. Theorem 1.31.
    30. 1.5. Summary of techniques
    31. 1.6. Summary of contents
  3. 2. Outline of proof for main theorem
    1. 2.1. Equivariant to nonequivariant reduction
    2. Proposition 2.2.
    3. Definition 2.3.
    4. Definition 2.4.
    5. Theorem 2.5.
    6. 2.2. The Mahowald line at odd primes and over
    7. 2.3. Step 1: Lower bound
    8. Theorem 2.6.
    9. Corollary 2.8.
    10. 2.4. Step 2: Upper bound detected by
    11. Proposition 2.9.
    12. Corollary 2.10.
    13. Corollary 2.11.
    14. 2.5. Step 3: Identifying the map on the first lock as
    15. Proposition 2.12.
    16. Corollary 2.13.
    17. 2.6. Step 4: A technical lemma for the upper bound
    18. Proposition 2.14.
    19. Terminology 2.15.
    20. 2.7. Step 5: The second lock is not passed
    21. Proposition 2.16.
    22. Lemma 2.17.
    23. Theorem 2.18.
    24. Corollary 2.19.
    25. 2.8. Step 6: The first lock is passed when is odd
    26. Proposition 2.20.
    27. Proposition 2.21.
    28. Corollary 2.22.
    29. 2.9. Step 7: The first lock is not passed when is even
    30. Proposition 2.23.
    31. Theorem 2.24.
    32. Corollary 2.25.
  4. 3. Preliminaries
    1. 3.1. Maps between subquotients
    2. Definition 3.1.
    3. Definition 3.2.
    4. Proposition 3.3.
    5. 3.2. Transfer maps
    6. Proposition 3.4.
    7. Proposition 3.5.
    8. Lemma 3.6.
  5. 4. Attaching maps in
    1. 4.1. -subquotients
    2. Definition 4.1.
    3. Lemma 4.2.
    4. Definition 4.3.
    5. Lemma 4.4.
    6. 4.2. The and -attaching maps in
    7. Proposition 4.5.
    8. Lemma 4.6.
    9. Proposition 4.7.
    10. Corollary 4.8.
    11. Lemma 4.9.
    12. 4.3. -Attaching maps in
    13. Proposition 4.10.
    14. 4.4. Periodicity in
    15. Proposition 4.11.
    16. 4.5. Some -subquotients of
    17. Lemma 4.12.
    18. Lemma 4.13.
    19. Proposition 4.14.
    20. Definition 4.15.
    21. Proposition 4.16.
  6. 5. Step 1: Proof of Theorem 2.6
    1. 5.1. An outline of the proof
    2. Proposition 5.1.
    3. Proposition 5.2.
    4. Proposition 5.3.
    5. Proposition 5.4.
    6. Proposition 5.5.
    7. 5.2. Proof of Proposition 5.1
    8. Lemma 5.7.
    9. Lemma 5.9.
    10. Lemma 5.10.
    11. Lemma 5.11.
    12. 5.3. Proof of Proposition 5.2
    13. Lemma 5.12.
    14. 5.4. Proof of Proposition 5.3
    15. Lemma 5.13.
    16. Lemma 5.14.
    17. Lemma 5.15.
    18. 5.5. Proof of Proposition 5.4
    19. 5.6. Proof of Proposition 5.5
    20. Lemma 5.16.
  7. 6. Step 2: Upper bound detected by
    1. Proposition 6.1 (Proposition 2.9).
    2. 6.1. Some results about the -equivariant -theory
    3. 6.2. Proof of Proposition 2.9
    4. Lemma 6.2.
    5. Corollary 6.3.
  8. 7. Step 3: Identifying the map on the first lock as
    1. Definition 7.1.
    2. Lemma 7.2.
  9. 8. Step 4: A technical lemma for the upper bound
    1. Proposition 8.1.
    2. 8.1. The spectra and
    3. Lemma 8.2.
    4. Lemma 8.3.
    5. Lemma 8.4.
    6. 8.2. Proof of Proposition 2.14
  10. 9. Step 5: Upper bound
    1. 9.1. Proving differentials using the Chern character
    2. Definition 9.1.
    3. Theorem 9.2.
    4. Lemma 9.3.
    5. Lemma 9.4.
    6. Lemma 9.5.
    7. Proposition 9.6.
    8. 9.2. Proof of Proposition 2.16
    9. 9.3. Proof of Lemma 2.17
    10. Lemma 9.7.
    11. Lemma 9.8.
  11. 10. Steps 6 and 7: First lock and second lock
    1. 10.1. Construction of
    2. Lemma 10.1.
    3. Lemma 10.3.
    4. 10.2. Proof of Proposition 2.20
    5. Lemma 10.4.
    6. 10.3. Bundles with simple Chern character
    7. Lemma 10.5.
    8. Proposition 10.6.
    9. 10.4. First lock for odd
    10. 10.5. First lock for even
  12. Appendix A. Coefficients of
    1. Lemma A.2.
    2. Lemma A.3.
    3. Lemma A.4.
    4. Lemma A.5.
    5. Lemma A.6.
    6. Lemma A.7.
    7. Lemma A.8.
  13. Appendix B. Cell diagrams and the Atiyah–Hirzebruch spectral sequence
    1. Definition B.1.
    2. Example B.3.
    3. Example B.4.
    4. Example B.5.
  14. Acknowledgments

Figures

Figure 1.

The Mahowald line. In each column, intuitively, the black dots represent cells of each , the black straight lines represent 2-attaching maps, the black curved lines represent -attaching maps, and the cyan curved lines represent -attaching maps. For the precise definitions of these attaching maps, see Section 4.

Graphic without alt text
Figure 2.

Various bounds for the Mahowald line

Graphic without alt text
Figure 3.

The lower bound of the -local Mahowald line at (black) is above the 2-local Mahowald line (gray)

Graphic without alt text
Figure 4.

Constructing and proving a lower bound for the Mahowald line

Graphic without alt text
Figure 5.

Proposition 2.16

Graphic without alt text
Figure 6.

Maps between subquotients

Graphic without alt text
Figure 7.

Some attaching maps in

Graphic without alt text
Figure 8.

Step 1 main picture

Graphic without alt text
Figure 9.

Step 1.1 picture

Graphic without alt text
Figure 10.

Intuition for Step 1

Graphic without alt text
Figure 11.

Step 1.1.1 picture

Graphic without alt text
Figure 12.

Step 1.1.5 picture

Graphic without alt text
Figure 13.

Step 1.2 picture

Graphic without alt text
Figure 14.

Step 1.4 picture

Graphic without alt text
Figure 15.

Step 1.6 picture

Graphic without alt text

Mathematical Fragments

Question 1.1.

How to classify all closed simply connected topological 4-manifolds?

Theorem 1.2 (Freedman Reference Fre82).
(1)

Two closed simply connected topological 4-manifolds are homeomorphic if and only if their intersection forms are isomorphic and their Kirby–Siebenmann invariants are the same.

(2)

When the form is not even, any combination of the symmetric unimodular bilinear form and Kirby–Siebenmann invariant can be realized by a closed simply connected topological 4-manifold.

(3)

When the form is even, the combination can be realized if and only if the Kirby–Siebenmann invariant is equal to the signature of the form divided by 8 modulo 2. (Note that the signature of an even form must be divisible by . See Reference DK90, Section 1.1.3 for example.)

Question 1.3.

How to classify all closed simply connected smooth 4-manifolds?

Question 1.4.

Given a symmetric unimodular bilinear form , can it be realized as the intersection form of a closed simply connected smooth 4-manifold?

Question 1.5.

Suppose that the answer to Question 1.4 is yes; then how many non-diffeomorphic 4-manifolds can realize the given form?

Theorem 1.6 (Donaldson’s diagonalizability theorem Reference Don83).

A definite symmetric unimodular bilinear form can be realized as the intersection form of a closed simply connected smooth 4-manifold if and only if can be represented by the matrix or .

Equation (1.1)
Conjecture 1.7 (The -conjecture, version 1).

The form

can be realized as the intersection form of a closed smooth spin -manifold if and only if .

Conjecture 1.9 (The -conjecture, version 2).

Any closed smooth spin 4-manifold must satisfy the inequality

where and are the second Betti number and the signature of , respectively.

Theorem 1.11 (Furuta’s -theorem Reference Fur01).

For , the bilinear form

is spin realizable only if .

Theorem 1.12.

For , the bilinear form

is spin realizable only if

Corollary 1.13.

Any closed simply connected smooth spin -manifold that is not homeomorphic to , , or must satisfy the inequality

Equation (1.4)
Theorem 1.14 (Furuta Reference Fur01).
(1)

Suppose . Then

Here, is the four-dimensional representation of , with acting on it via left multiplication, and is a -dimensional representation such that the unit component acts as identity and the other component acts as negative identity.

(2)

The element fits into the commutative diagram

where and are stable homotopy classes that represents the inclusions and of fixed points.

Theorem 1.16 (Furuta Reference Fur01).

A level- Furuta–Mahowald class exists only if .

Question 1.17.

What is the necessary and sufficient condition for the existence of a level- Furuta–Mahowald class?

Remark 1.18.

We would now like to discuss the choice of the universe (i.e. the -representations that one stabilizes with respect to when passing from the space level to the spectrum level). In Furuta’s original proof of Theorem 1.16 Reference Fur01, he used the universe consisting of only the representations and , because this universe is the most relevant to the geometric problem. Modified proofs by Manolescu Reference Man14 and Bryan Reference Bry98, using divisibilities of the -theoretic Euler classes, show that the statement of Theorem 1.16 holds for any universe.

For Question 1.17, the answer could potentially depend on the choice of the universe. By works of Schmidt Reference Sch03, Theorem 2.6, Theorem 3.2 and Minami Reference Min, any Furuta–Maholwald class can be desuspended to the same diagram on the space level as long as . By the discussions in the previous paragraph, the bound in Theorem 1.16 holds for any universe. Therefore, a level- Furuta–Mahowald class in one universe can be desuspended to a space-level map , and then be further suspended to a level- Furuta–Mahowald class in any other universe. It follows that the answer to Question 1.17 does not depend on the choice of the universe.

Without loss of generality, we always work with the complete universe.

Conjecture 1.19 (Jones Reference FKMM07).

For , a level- Furuta–Mahowald class exists if and only if

Theorem 1.21 (The limit is ).

For , a level- Furuta–Mahowald class exists if and only if

Remark 1.23.

The “if” part of Theorem 1.21 implies that without further input from geometry or analysis, the best result one can achieve in proving Conjecture 1.9, using the existence of Furuta–Mahowald classes, is . Note that by Remark 1.18 this “limit” does not depend on the choice of the universe. In order to break this “limit” and to further attack the -conjecture, more delicate properties of the Seiberg–Witten map have to be studied. In particular, the Seiberg–Witten map should not be merely treated as a continuous map.

Definition 1.25.

Suppose that and are elements in with non-nilpotent. The -equivariant Mahowald invariant of with respect to is the following set of elements in :

In other words, an element belongs to if the left diagram exists and the right diagram does not exist for any class .

Conjecture 1.29 (Bredon–Löffler, Mahowald–Ravenel).

For any non-equivariant class that is of positive degree, we have the inequality

Theorem 1.30 (Crabb Reference Cra89, Schmidt Reference Sch03, Stolz Reference Sto89).

For , the following 8-periodic result holds:

Theorem 1.31.

For , the following 16-periodic result holds:

Equation (2.2)
Proposition 2.2.

A level- Furuta–Mahowald class exists if and only if the map

is zero.

Definition 2.3.

The function is defined by setting to be the largest integer such that the map

is null-homotopic.

Theorem 2.5.

The function takes values as follows:

and for all ,

Equation (2.3)
Theorem 2.6.

For every , there exist maps

with the following properties (see Figure 4):

(i)

The diagram

commutes.

(ii)

The map induces an isomorphism on . In other words, is an -subcomplex of via the map (see Definition 4.1).

(iii)

The following diagram is commutative:

(iv)

Let be the restriction of to the bottom cell of . Then for , the map satisfies the inductive relation

where in and is some element in . We will show in Lemma 4.9 that and we set .

Corollary 2.8.

For any and , we have the inequality

where

This line is shown in blue in Figure 4.

Proposition 2.9.

For any , the composition

is nonzero.

Corollary 2.10.

The map is nontrivial.

Proposition 2.12.

For all , we have the relations

Corollary 2.13.

For all , the diagram

commutes.

Proposition 2.14.

For any , the map

induced by the quotient map is injective.

Terminology 2.15.

Let be a CW spectrum that has at most one cell in each dimension. Recall that the cohomological -based Atiyah–Hirzebruch spectral sequence for has the following form:

Here, is the indexing set containing the dimensions of the cells of , is the cellular filtration of . The degrees for the -differentials are as follows:

Similarly, the homological -based Atiyah–Hirzebruch spectral sequence for has the following form:

Here, is the indexing set containing the dimensions of the cells of , is the cellular filtration of . The degrees for the -differentials are as follows:

Proposition 2.16.

There exists a map

with the following properties (see Figure 5):

(i)

The map

factors through the quotient map

via :

(ii)

The map factors through a quotient map

via a map

(iii)

The restriction of to its bottom cell is the map

(iv)

The map has order 2 in . In other words, the following composition is zero:

Lemma 2.17.

In the -based Atiyah–Hirzebruch spectral sequence of , there is a differential

where is a nonzero element in .

Theorem 2.18.

The composition map

is not zero.

Proposition 2.20.

There exists a map

such that the following diagram commutes:

Proposition 2.21.

When is odd, the composition

is zero.

Proposition 2.23.

When is even, the class

is a permanent cycle in the -based Atiyah–Hirzebruch spectral sequence of .

Theorem 2.24.

When is even, the composition map

is not null.

Equation (2.12)
Equation (2.13)
Equation (2.14)
Proposition 3.3.

Let , , , be integers such that the following conditions hold:

(a)

and , where and ;

(b)

;

(c)

;

(d)

.

Then there exists a map

Furthermore, the maps are compatible with each other in the sense that the following three properties hold:

(1)

(Compatibility with respect to quotient). The following diagram commutes for all :

(2)

(Compatibility with respect to inclusion). If is another tuple satisfying the conditions above with and , then the following diagram commutes:

(3)

(Compatibility with respect to composition). If and are two tuples satisfying the conditions of the proposition, then

Step 1.

, . By our definition of and the cellular approximation theorem, there is always a map

Furthermore, this map makes the bottom square of the diagram

commute. Since both columns are cofiber sequences, there is an induced map

between the cofibers making the whole diagram commute. The top square of the commutative diagram above implies that property (1) holds for .

Step 2.

, . Note that by the definition of ,

We define the map to be the map

The map fits into the following commutative diagram:

Step 3.

. We define the map to be the composition

We now prove that property (2) holds when . The case when is directly implied by diagram Equation 3.2.

Suppose that . Consider the two compositions

and

in diagram Equation 3.1. We want to show that these two compositions are equal. After post-composing with the inclusion map

the maps and are homotopic to each other (this is because we have already verified Property (2) when ).

Consider the cofiber sequence

Since the difference is null after post-composing with the map

it factors through the fiber via a certain map

If the left vertical arrow in diagram Equation 3.1 is the identity map, then diagram Equation 3.1 commutes by definition. Otherwise, it is straightforward to check that the dimension of the top cell of is less than the dimension of the bottom cell in . Therefore, the map is zero by the cellular approximation theorem. This implies and that property (2) holds when .

Proposition 3.4.

There is a cofiber sequence

Proposition 3.5.

The transfer map

induces an isomorphism on for all .

Lemma 3.6.

is homotopy equivalent to .

Definition 4.1.

Let , , and be CW spectra, and be maps

We say that is an -subcomplex of if the map induces an injection on mod 2 homology. An -subcomplex is denoted by a hooked arrow as above. Similarly, we say that is an -quotient complex of if the map induces a surjection on mod 2 homology. An -quotient complex is denoted by a double-headed arrow as above. When the maps involved are clear in the context, we may ignore the maps and and just say that is an -subcomplex of , and is an -quotient complex of .

Furthermore, is an -subquotient of if is either an -subcomplex of an -quotient complex of or an -quotient complex of an -subcomplex of .

Lemma 4.2.

Suppose that is an -subcomplex of . Let be the cofiber of and let be an -subcomplex of . Define to be the homotopy pullback of along . Then is an -subcomplex of . Moreover, is an -subcomplex of with quotient .

Dually, suppose is an -quotient complex of . Let be the fiber of . let be an -quotient complex of . Define to be the homotopy pushout of along . We have that is an -quotient complex of . Moreover, is an -quotient complex of with fiber .

Lemma 4.4.

Suppose that is a CW spectrum, with only one cell in dimension . Then the following claims hold:

(1)

There is a 2-attaching map from dimension to dimension in if and only if the map

is nonzero.

(2)

There is an -attaching map from dimension to dimension in if and only if the map

is nonzero.

Equation (4.1)
Lemma 4.6.

The induced homomorphisms and on mod 2 homologies can be described as follows:

(1)

The map

is an isomorphism if and only if

and ;

and ;

and ;

and .

In other words, is an isomorphism when both the domain and the codomain are nonzero.

(2)

The map

is an isomorphism if and only if

and ;

and ;

and ;

and .

Proposition 4.7.

In the mod 2 homology of ,

(1)

is nonzero if and only if

and ;

and ;

and ;

and .

(2)

is nonzero if and only if

and ;

and .

Equation (4.2)
Corollary 4.8.

There are 2 and -attaching maps in if and only if they are marked in Figure 7.

Lemma 4.9.

Suppose that and satisfy one of following conditions:

and ;

and ;

and ;

and .

Then the map

is .

Proposition 4.10.

There is an -attaching map in from dimension to dimension if and only if it is one of the following four cases (see Figure 7):

and ;

and ;

and ;

and .

Proposition 4.11.

For any , there is an equivalence

Lemma 4.12.

The 3 cell complex splits:

Lemma 4.13.

The -cell complex splits:

Proposition 4.14.

There exists a 4 cell complex that is an -subcomplex of . It has cells in dimensions , , and .

Proposition 4.16.

There is a 2 cell complex with cells in dimensions and , such that it is an -quotient complex of .

Step 1.1.

We establish the existence of the maps for all (see Figure 9).

Proposition 5.1.

For every , there exists a map that fits into the following commutative diagram

The proof of Proposition 5.1 is an extensive and careful study of the cell structures of the columns between and and in dimensions between and . It involves the computation of stable stems in the range . We define as the composition

Step 1.2.

Using Proposition 5.1 and the homotopy extension property, which is stated as Lemma 5.12 in Section 5.3, we show the existence of two maps and in Proposition 5.2.

Proposition 5.2.

For every , there exist maps that fit into the following commutative diagram:

Moreover, the map induces an isomorphism on . In other words, is an -subcomplex of .

We define the map as the following composite

Note here we use the octahedron axiom to identify with . It then follows from Proposition 5.2 that the map induces an isomorphism on , which establishes property (ii) in Theorem 2.6.

Step 1.3.

We define

to be the zero map. Note that the 3 cells of are in dimensions , , , so this is the only choice. Since , the following diagram (iii) in Theorem 2.6 for commutes regardless of the construction of .

For the existence of the map , it suffices to show the following composite is zero.

This is true because this map factors through by cellular approximation. This gives the following commutative diagram (i) in Theorem 2.6 for .

This gives the starting case of our inductive argument.

Step 1.4.

For , we assume the maps and exist, the two diagrams in (i) and (iii) of Theorem 2.6 commute for the 4 maps , and satisfies property (iv) in Theorem 2.6.

We define the map to be the composite

Using the commutative diagram Equation 2.5 in (iii) of Theorem 2.6 for the case , we have Proposition 5.3:

Proposition 5.3.

The following composite is zero.

Note that the first map is the inclusion of the bottom cell of , and that the map is established in Step 1.2 before the induction.

As a result, there exist maps

that fit into the following commutative diagram:

Note that there are many choices of that makes the diagram 5.6 commute.

Step 1.5.

In this step, we prove Proposition 5.4.

Proposition 5.4.

For any choice of in Step 1.4, the two diagrams Equation 2.4 and Equation 2.5 in (i) and (iii) of Theorem 2.6 commute for the 4 maps .

The proof is a straightforward cell diagram chasing argument.

Step 1.6.

In this step, we prove Proposition 5.5.

Proposition 5.5.

Let be the restriction of to the bottom cell of . Then there exists one choice of in Step 1.4 such that the following property is satisfied:

where and . Note that by Lemma 4.9 and we set .

This proves that this choice of satisfies the relation in (iv) of Theorem 2.6 and therefore completes the induction.

Remark 5.6.

The critical part of Theorem 2.6 is the existence of the map . We want to prove it inductively. Namely, we assume that exists and want to show that exists. This induction would follow easily if the following map were zero:

However, this is not true. Intuitively, the -cell in maps nontrivially to the -cell in by . More precisely, one can show that the map 5.8 factors through as an -quotient, and the latter map in the following composite

is detected by in the Atiyah–Hirzebruch spectral sequence of . Therefore, we have to show the composite

is zero. It turns out that we can show the composite of the latter two maps in 5.9 is zero. This follows from a technical condition that can be chosen to satisfy:

factors through .

Here note that is an -subcomplex of . In fact, this is due to the composite

corresponding to an element in the group .

Now to complete the induction, we need to show that can be chosen to satisfy:

factors through .

Firstly, in , the -cell is only attached to the cells in dimensions and , all of which map trivially to . As a result, we can choose such that the restriction factors through .

Secondly, by some local arguments that involve attaching maps in for , …, , we can show that can be chosen such that factors through .

This allows us to complete the induction and to prove Theorem 2.6. See Figure 10 for an illustration of the discussion above.

We’d like to comment that our actual argument is a little different from our discussion above. We actually analyze instead of . This is used to deduce the inductive relation Equation 5.7, based on which we identify the first lock.

Step 1.1.1.

In this step, we focus on column . We use the -attaching maps in column between cells in dimensions and , and , to get rid of the cell in dimension of , and to lower the upper bound of the image to dimension in column . More precisely, we prove Lemma 5.7.

Lemma 5.7.

There exists the following commutative diagram (see Figure 11):

Proof.

Firstly, we have the following commutative diagram:

By Lemma 4.9, we have that the map in middle of the top row of diagram 5.11 is . By Corollary 4.8, we have an -attaching map in between the cells in dimensions and . This corresponds to an Atiyah–Hirzebruch differential

Therefore, the composition of the maps in the top row of diagram 5.11 is zero. In particular, pre-composing with the map

is also zero. By the cofiber sequence of the right most column, we know that the map from to maps through .

Secondly, we have the following commutative diagram:

By Lemma 4.9, we have that the map in middle of the bottom row of diagram 5.12 is . By Corollary 4.8, we have an -attaching map in between the cells in dimensions and . This corresponds to an Atiyah–Hirzebruch differential

Therefore, the composition of the maps in the bottom row of diagram 5.12 is zero. In particular, post-composing with the map

is also zero. By the cofiber sequence of the left most column, we know that the map from to factors through .

This gives the required map

Remark 5.8.

We will use arguments similar to the ones in the proof of Lemma 5.7 many times in the rest of this paper. Instead of presenting all details in terms of commutative diagrams, we will simply refer them as “similar arguments as in the proof of Lemma 5.10” or “cell diagram chasing arguments” due to certain attaching maps.

Step 1.1.2.

In this step, we focus on column . We show that in , the cells in dimensions and map through in column . More precisely, we have Lemma 5.9.

Lemma 5.9.

There exists the following commutative diagram:

Proof.

By Proposition 4.10, there is no -attaching map in . This shows that

We may therefore consider the cells in dimensions and separately.

For , it maps naturally through the -skeleton in column . By Proposition 4.10, there is an -attaching map in between the cells in dimensions and . A similar argument as in the proof of Lemma Equation 5.10 shows that maps through in column .

For , firstly note that by Corollary 4.8, there is an -attaching map in between the cells in dimensions and . A similar argument as in the proof of Lemma Equation 5.10 shows that maps through the -skeleton in column . Then it maps naturally through the -skeleton in column . To see that it actually maps through , we only need to show the following composite is zero.

This is in fact true, since .

Combining both parts, this gives the required map

We enlarge Diagram 5.13 to Diagram 5.14. We will establish the maps and in Steps 1.1.3, 1.1.4 and 1.1.5:

Step 1.1.3.

In this step, we establish the map , making the triangle under in Diagram Equation 5.14 commute.

By Lemma 4.9, we have that the map

is mapping into the bottom cell of . Since

the composition of maps in the bottom row of Diagram Equation 5.14 is zero. In particular, post-composing with the map

is also zero. By the cofiber sequence of the left most column, we know that the map from to factors through , which gives the desired map , making the triangle under commute.

Note that we haven’t shown the triangle above commutes. We will show it later in Step 1.1.5.

Step 1.1.4.

In this step, we establish the map , making the parallelogram below in Diagram Equation 5.14 commute.

By the cofiber sequence in the left most column, it suffices to show the following composite is zero.

Since both the triangle under and the upper rectangle in Diagram Equation 5.14 commute, it is equivalent to show that the following composite is zero.

This is in fact true, since the composition of the latter two maps is already zero.

Lemma 5.10.

The following composite in Diagram Equation 5.14 is zero.

Equation (5.15)
Step 1.1.5.

In this step, we establish the map , making all parts of Diagram Equation 5.14 commute.

It suffices to show Lemma 5.11.

Lemma 5.11.

The following composite is zero.

In fact, by Lemma 5.11 and Step 4, the following composite is zero.

Then by the cofiber sequence in the right most column of Diagram Equation 5.14, the map must map through , establishing the desired map .

To see that all parts of Diagram Equation 5.14 commute, first note that by Lemma 5.11 and Lemma 5.10, both the triangles above the map and under the map commute. Next, by the construction of the map , the triangles above it commute. Finally, by Step 1.1.3 and the cofiber sequence of the left most column in Diagram Equation 5.14, the triangle under the map commutes. Therefore, all parts of Diagram Equation 5.14 commute.

Now, let’s prove Lemma 5.11.

Proof of Lemma 5.11.

The composite in the statement splits into the following two composites.

For the first composite 5.16, let’s study the second map . By Proposition 4.10 and Corollary 4.8, is a 3 cell complex, with cells in dimensions , and with a 2 and -attaching map. Since , there is a nonzero differential

in the Atiyah–Hirzebruch spectral sequence of . It follows that the second map must map through its -skeleton: . Since , the map must further map through and the composite 5.16 can be decomposed as

Therefore, due to the relation

the first composite 5.16 is zero.

For the second composite 5.17, the second map must map through the -skeleton of , which is . Then it follows from the relation

that the second composite 5.17 is zero. This completes the proof.

Now we claim that the map is our desired map in Proposition 5.1. In fact, part of Diagram Equation 5.14 gives us the following commutative diagram (see Figure 12).

Putting Diagrams Equation 5.10 and 5.18 together, we have the following commutative diagram.

Forgetting some terms in this diagram, we obtain Diagram Equation 5.3 in Proposition 5.1.

Lemma 5.12.

Suppose that we have the following commutative diagram in the stable homotopy category

where and are the cofibers of the maps and respectively. Then it can be extended into the following commutative diagram:

Equation (5.20)
Lemma 5.13.

There exists a map

that fits into the following commutative diagram

Equation (5.22)
Lemma 5.14.

The following composite

factors through , giving the map in the diagram Equation 5.22.

Lemma 5.15.

The following composite is zero.

Equation (5.23)
Equation (5.24)
Equation (5.25)
Equation (5.27)
Equation (5.28)
Equation (5.29)
Equation (5.30)
Equation (5.31)
Equation (5.32)
Equation (5.33)
Lemma 5.16.

The following composite is zero:

Equation (5.35)
Equation (5.36)
Lemma 6.2.

The map

sends to a nonzero element.

Equation (6.1)
Equation (6.2)
Corollary 6.3.

For , the map is detected by .

Equation (7.1)
Definition 7.1.

Define

and for ,

Lemma 7.2.

For all , the following relations hold:

Equation (7.2)
Proposition 8.1.

For any , the map

induced by the quotient map is injective.

Equation (8.2)
Lemma 8.2.

Let be a virtual bundle over a space , and let be a spin bundle of dimension over . Suppose that is a subspace of . Let and be the restrictions of and to . We have the Thom isomorphism

The isomorphism above is natural in the sense that if is a subspace of , and , are the restrictions of and to , then the following diagram commutes:

Lemma 8.3.

Let be a finite CW-spectrum with no cell of dimension . Then .

Lemma 8.4.

Let be the 0-connected cover of . There is a map

that induces an injection on for any positive integer .

Equation (8.5)
Equation (8.6)
Definition 9.1.

A finite -spectrum is called -injective if the map

given by is injective. Here, denotes the complexification of .

Theorem 9.2.

Let be a finite CW-spectrum that satisfies the following properties:

(1)

has a single top cell in dimension ;

(2)

has no cells in dimension ;

(3)

The -skeleton of is -injective;

(4)

The -skeleton of is homotopy equivalent to .

Furthermore, suppose there is an element that satisfies the equality

Then in the -based Atiyah–Hirzebruch spectral sequence of , the following results hold:

(I)

If , then the class is a permanent cycle. Here, when is even and when is odd.

(II)

If , then there is a nontrivial differential

for some .

Lemma 9.3.

Let be the pull-back of under the inclusion map . Then .

Lemma 9.4.

In the -based Atiyah–Hirzebruch spectral sequence for , the element is a permanent cycle.

Lemma 9.5.

is -injective.

Proposition 9.6.

The element is a permanent cycle in the -based Atiyah–Hirzebruch spectral sequence of if and only if .

Equation (9.3)
Equation (9.4)
Claim 1.

.

Claim 2.

is of order 2 in . In other words, the following composite is zero.

Claim 3.

The restriction of to the bottom cell is

in .

Equation (9.6)
Equation (9.7)
Equation (9.8)
Equation (9.9)
Equation (9.10)
Lemma 9.7.

For any odd integer and any , the spectrum is -injective. (See Definition 9.1.)

Equation (9.11)
Equation (9.12)
Lemma 9.8.

There exists an element such that

for some with .

Equation (9.14)
Lemma 10.1.

The complex satisfies the following properties:

(1)

;

(2)

Remark 10.2.

In the proof of Reference Sch03, Theorem 4.9, Schmidt made a minor error when computing . This error led to Schmidt’s proof of Jones Conjecture for .

Note that Lemma 10.1 is a crucial step in our proof of showing that the Jones conjecture is not true when . If Schmidt’s cohomotopy group computation were true, our statement of Lemma 10.1(2) would be different: . This would also lead to an affirmative answer for Jones conjecture for by using our subsequent arguments.

Lemma 10.3.

For any , the -skeleton of is -injective.

Lemma 10.4.

The following diagram commutes:

Lemma 10.5.

There exists an element such that

with and .

Equation (10.2)
Equation (10.3)
Proposition 10.6.

There exists an element such that

(1)

When is even,

(2)

When is odd,

with .

Equation (10.6)
Equation (10.7)
Equation (10.8)
Lemma A.2.

for all .

Lemma A.3.

The inequality holds for all and .

Lemma A.4.

for all .

Lemma A.5.

for all .

Lemma A.6.

for all .

Lemma A.7.

We have

Lemma A.8.

For a fixed , the inequality holds for all .

Example B.3.

Let be the composite of the following two maps:

where the second map is the inclusion of the bottom cell. Consider : the cofiber of , which is a 3 cell complex with the following cell diagram:

It is clear that the top cell of splits off, since can be divided by . So we do not have to draw any attaching map from the cell in dimension 3 to the one in dimension 0. Note that the cofiber of is in fact an -subcomplex of .

Example B.4.

Consider the suspension spectrum of . It is a 3 cell complex with cells in dimensions 2, 4 and 6. It was shown by Adams Reference Ada58 that the secondary cohomology operation , which is associated to the relation

is nonzero on this spectrum. In other words, there exists an attaching map between the cells in dimensions 2 and 6, which is detected by in the 3-stem of the Adams page. Note that detects two homotopy classes: . Their difference is , which is divisible by . Therefore, we have its cell diagram as the following:

We can also consider the Spanier–Whitehead dual of the suspension spectrum of . It is a 3 cell complex with cells in dimensions -2, -4 and -6, with the following cell diagram

References

Reference [ABS64]
M. F. Atiyah, R. Bott, and A. Shapiro, Clifford modules, Topology 3 (1964), no. suppl, suppl. 1, 3–38, DOI 10.1016/0040-9383(64)90003-5. MR167985,
Show rawAMSref \bib{AtiyahBottShapiro}{article}{ label={ABS64}, author={Atiyah, M. F.}, author={Bott, R.}, author={Shapiro, A.}, title={Clifford modules}, journal={Topology}, volume={3}, date={1964}, number={suppl, suppl. 1}, pages={3--38}, issn={0040-9383}, review={\MR {167985}}, doi={10.1016/0040-9383(64)90003-5}, }
Reference [Ada58]
J. F. Adams, On the nonexistence of elements of Hopf invariant one, Bull. Amer. Math. Soc. 64 (1958), 279–282, DOI 10.1090/S0002-9904-1958-10225-6. MR97059,
Show rawAMSref \bib{Adams}{article}{ label={Ada58}, author={Adams, J. F.}, title={On the nonexistence of elements of Hopf invariant one}, journal={Bull. Amer. Math. Soc.}, volume={64}, date={1958}, pages={279--282}, issn={0002-9904}, review={\MR {97059}}, doi={10.1090/S0002-9904-1958-10225-6}, }
Reference [AHR]
Matthew Ando, Michael J. Hopkins, and Charles Rezk, Multiplicative orientations of -theory and the spectrum of topological modular forms, https://faculty.math.illinois.edu/~mando/papers/koandtmf.pdf, 2010.
Reference [Bau04]
Tilman Bauer, -Compact groups as framed manifolds, Topology 43 (2004), no. 3, 569–597, DOI 10.1016/j.top.2003.09.002. MR2041631,
Show rawAMSref \bib{BauerFramedManifolds}{article}{ label={Bau04}, author={Bauer, Tilman}, title={$p$-Compact groups as framed manifolds}, journal={Topology}, volume={43}, date={2004}, number={3}, pages={569--597}, issn={0040-9383}, review={\MR {2041631}}, doi={10.1016/j.top.2003.09.002}, }
Reference [Beh06]
Mark Behrens, Root invariants in the Adams spectral sequence, Trans. Amer. Math. Soc. 358 (2006), no. 10, 4279–4341, DOI 10.1090/S0002-9947-05-03773-6. MR2231379,
Show rawAMSref \bib{Behrens2}{article}{ label={Beh06}, author={Behrens, Mark}, title={Root invariants in the Adams spectral sequence}, journal={Trans. Amer. Math. Soc.}, volume={358}, date={2006}, number={10}, pages={4279--4341}, issn={0002-9947}, review={\MR {2231379}}, doi={10.1090/S0002-9947-05-03773-6}, }
Reference [Beh07]
Mark Behrens, Some root invariants at the prime 2, Proceedings of the Nishida Fest (Kinosaki 2003), Geom. Topol. Monogr., vol. 10, Geom. Topol. Publ., Coventry, 2007, pp. 1–40, DOI 10.2140/gtm.2007.10.1. MR2402775,
Show rawAMSref \bib{Behrens1}{article}{ label={Beh07}, author={Behrens, Mark}, title={Some root invariants at the prime 2}, conference={ title={Proceedings of the Nishida Fest (Kinosaki 2003)}, }, book={ series={Geom. Topol. Monogr.}, volume={10}, publisher={Geom. Topol. Publ., Coventry}, }, date={2007}, pages={1--40}, review={\MR {2402775}}, doi={10.2140/gtm.2007.10.1}, }
Reference [BF04]
Stefan Bauer and Mikio Furuta, A stable cohomotopy refinement of Seiberg-Witten invariants. I, Invent. Math. 155 (2004), no. 1, 1–19, DOI 10.1007/s00222-003-0288-5. MR2025298,
Show rawAMSref \bib{Bauer-Furuta2004}{article}{ label={BF04}, author={Bauer, Stefan}, author={Furuta, Mikio}, title={A stable cohomotopy refinement of Seiberg-Witten invariants. I}, journal={Invent. Math.}, volume={155}, date={2004}, number={1}, pages={1--19}, issn={0020-9910}, review={\MR {2025298}}, doi={10.1007/s00222-003-0288-5}, }
Reference [BG75]
J. C. Becker and D. H. Gottlieb, The transfer map and fiber bundles, Topology 14 (1975), 1–12, DOI 10.1016/0040-9383(75)90029-4. MR377873,
Show rawAMSref \bib{BeckerGottlieb}{article}{ label={BG75}, author={Becker, J. C.}, author={Gottlieb, D. H.}, title={The transfer map and fiber bundles}, journal={Topology}, volume={14}, date={1975}, pages={1--12}, issn={0040-9383}, review={\MR {377873}}, doi={10.1016/0040-9383(75)90029-4}, }
Reference [BG95]
Robert Bruner and John Greenlees, The Bredon-Löffler conjecture, Experiment. Math. 4 (1995), no. 4, 289–297. MR1387694,
Show rawAMSref \bib{BrunerGreenlees}{article}{ label={BG95}, author={Bruner, Robert}, author={Greenlees, John}, title={The Bredon-L\"{o}ffler conjecture}, journal={Experiment. Math.}, volume={4}, date={1995}, number={4}, pages={289--297}, issn={1058-6458}, review={\MR {1387694}}, }
Reference [BJM84]
M. G. Barratt, J. D. S. Jones, and M. E. Mahowald, Relations amongst Toda brackets and the Kervaire invariant in dimension , J. London Math. Soc. (2) 30 (1984), no. 3, 533–550, DOI 10.1112/jlms/s2-30.3.533. MR810962,
Show rawAMSref \bib{BJM}{article}{ label={BJM84}, author={Barratt, M. G.}, author={Jones, J. D. S.}, author={Mahowald, M. E.}, title={Relations amongst Toda brackets and the Kervaire invariant in dimension $62$}, journal={J. London Math. Soc. (2)}, volume={30}, date={1984}, number={3}, pages={533--550}, issn={0024-6107}, review={\MR {810962}}, doi={10.1112/jlms/s2-30.3.533}, }
Reference [Bre67]
Glen E. Bredon, Equivariant stable stems, Bull. Amer. Math. Soc. 73 (1967), 269–273, DOI 10.1090/S0002-9904-1967-11713-0. MR206947,
Show rawAMSref \bib{BredonStableStems}{article}{ label={Bre67}, author={Bredon, Glen E.}, title={Equivariant stable stems}, journal={Bull. Amer. Math. Soc.}, volume={73}, date={1967}, pages={269--273}, issn={0002-9904}, review={\MR {206947}}, doi={10.1090/S0002-9904-1967-11713-0}, }
Reference [Bre68]
Glen E. Bredon, Equivariant homotopy, Proc. Conf. on Transformation Groups (New Orleans, La., 1967), Springer, New York, 1968, pp. 281–292. MR0250303,
Show rawAMSref \bib{BredonEquivariant}{article}{ label={Bre68}, author={Bredon, Glen E.}, title={Equivariant homotopy}, conference={ title={Proc. Conf. on Transformation Groups}, address={New Orleans, La.}, date={1967}, }, book={ publisher={Springer, New York}, }, date={1968}, pages={281--292}, review={\MR {0250303}}, }
Reference [Bru98]
Robert R. Bruner, Some remarks on the root invariant, Stable and unstable homotopy (Toronto, ON, 1996), Fields Inst. Commun., vol. 19, Amer. Math. Soc., Providence, RI, 1998, pp. 31–37. MR1622335,
Show rawAMSref \bib{Bruner}{article}{ label={Bru98}, author={Bruner, Robert R.}, title={Some remarks on the root invariant}, conference={ title={Stable and unstable homotopy}, address={Toronto, ON}, date={1996}, }, book={ series={Fields Inst. Commun.}, volume={19}, publisher={Amer. Math. Soc., Providence, RI}, }, date={1998}, pages={31--37}, review={\MR {1622335}}, }
Reference [Bry98]
Jim Bryan, Seiberg-Witten theory and actions on spin -manifolds, Math. Res. Lett. 5 (1998), no. 1-2, 165–183, DOI 10.4310/MRL.1998.v5.n2.a3. MR1617929,
Show rawAMSref \bib{Bryan}{article}{ label={Bry98}, author={Bryan, Jim}, title={Seiberg-Witten theory and $\mathbf {Z}/2^p$ actions on spin $4$-manifolds}, journal={Math. Res. Lett.}, volume={5}, date={1998}, number={1-2}, pages={165--183}, issn={1073-2780}, review={\MR {1617929}}, doi={10.4310/MRL.1998.v5.n2.a3}, }
Reference [BS78]
J. C. Becker and R. E. Schultz, Fixed-point indices and left invariant framings, Geometric applications of homotopy theory (Proc. Conf., Evanston, Ill., 1977), I, Lecture Notes in Math., vol. 657, Springer, Berlin, 1978, pp. 1–31. MR513539,
Show rawAMSref \bib{BeckerSchultz}{article}{ label={BS78}, author={Becker, J. C.}, author={Schultz, R. E.}, title={Fixed-point indices and left invariant framings}, conference={ title={Geometric applications of homotopy theory (Proc. Conf., Evanston, Ill., 1977), I}, }, book={ series={Lecture Notes in Math.}, volume={657}, publisher={Springer, Berlin}, }, date={1978}, pages={1--31}, review={\MR {513539}}, }
Reference [Cai35]
S. S. Cairns, Triangulation of the manifold of class one, Bull. Amer. Math. Soc. 41 (1935), no. 8, 549–552, DOI 10.1090/S0002-9904-1935-06140-3. MR1563139,
Show rawAMSref \bib{Cairns}{article}{ label={Cai35}, author={Cairns, S. S.}, title={Triangulation of the manifold of class one}, journal={Bull. Amer. Math. Soc.}, volume={41}, date={1935}, number={8}, pages={549--552}, issn={0002-9904}, review={\MR {1563139}}, doi={10.1090/S0002-9904-1935-06140-3}, }
Reference [Cra89]
M. C. Crabb, Periodicity in -equivariant stable homotopy theory, Algebraic topology (Evanston, IL, 1988), Contemp. Math., vol. 96, Amer. Math. Soc., Providence, RI, 1989, pp. 109–124, DOI 10.1090/conm/096/1022677. MR1022677,
Show rawAMSref \bib{Crabb}{article}{ label={Cra89}, author={Crabb, M. C.}, title={Periodicity in $\mathbf {Z}/4$-equivariant stable homotopy theory}, conference={ title={Algebraic topology}, address={Evanston, IL}, date={1988}, }, book={ series={Contemp. Math.}, volume={96}, publisher={Amer. Math. Soc., Providence, RI}, }, date={1989}, pages={109--124}, review={\MR {1022677}}, doi={10.1090/conm/096/1022677}, }
Reference [DK90]
S. K. Donaldson and P. B. Kronheimer, The geometry of four-manifolds, Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 1990. Oxford Science Publications. MR1079726,
Show rawAMSref \bib{Donaldson-Kronheimer1990}{book}{ label={DK90}, author={Donaldson, S. K.}, author={Kronheimer, P. B.}, title={The geometry of four-manifolds}, series={Oxford Mathematical Monographs}, note={Oxford Science Publications}, publisher={The Clarendon Press, Oxford University Press, New York}, date={1990}, pages={x+440}, isbn={0-19-853553-8}, review={\MR {1079726}}, }
Reference [Don83]
S. K. Donaldson, An application of gauge theory to four-dimensional topology, J. Differential Geom. 18 (1983), no. 2, 279–315. MR710056,
Show rawAMSref \bib{Donaldson1983}{article}{ label={Don83}, author={Donaldson, S. K.}, title={An application of gauge theory to four-dimensional topology}, journal={J. Differential Geom.}, volume={18}, date={1983}, number={2}, pages={279--315}, issn={0022-040X}, review={\MR {710056}}, }
Reference [Don86]
S. K. Donaldson, Connections, cohomology and the intersection forms of -manifolds, J. Differential Geom. 24 (1986), no. 3, 275–341. MR868974,
Show rawAMSref \bib{Donnaldso1986}{article}{ label={Don86}, author={Donaldson, S. K.}, title={Connections, cohomology and the intersection forms of $4$-manifolds}, journal={J. Differential Geom.}, volume={24}, date={1986}, number={3}, pages={275--341}, issn={0022-040X}, review={\MR {868974}}, }
Reference [Don87]
S. K. Donaldson, The orientation of Yang-Mills moduli spaces and -manifold topology, J. Differential Geom. 26 (1987), no. 3, 397–428. MR910015,
Show rawAMSref \bib{Donaldson1987}{article}{ label={Don87}, author={Donaldson, S. K.}, title={The orientation of Yang-Mills moduli spaces and $4$-manifold topology}, journal={J. Differential Geom.}, volume={26}, date={1987}, number={3}, pages={397--428}, issn={0022-040X}, review={\MR {910015}}, }
Reference [FK]
Mikio Furuta and Yukio Kametani, Equivariant maps between sphere bundles over tori and KO-degree, Preprint, arXiv:math/0502511.
Reference [FKMM07]
Mikio Furuta, Yukio Kametani, Hirofumi Matsue, and Norihiko Minami, Homotopy theoretical considerations of the Bauer-Furuta stable homotopy Seiberg-Witten invariants, Proceedings of the Nishida Fest (Kinosaki 2003), Geom. Topol. Monogr., vol. 10, Geom. Topol. Publ., Coventry, 2007, pp. 155–166, DOI 10.2140/gtm.2007.10.155. MR2402782,
Show rawAMSref \bib{Furuta-Kametani-Matsue-Minami2007}{article}{ label={FKMM07}, author={Furuta, Mikio}, author={Kametani, Yukio}, author={Matsue, Hirofumi}, author={Minami, Norihiko}, title={Homotopy theoretical considerations of the Bauer-Furuta stable homotopy Seiberg-Witten invariants}, conference={ title={Proceedings of the Nishida Fest (Kinosaki 2003)}, }, book={ series={Geom. Topol. Monogr.}, volume={10}, publisher={Geom. Topol. Publ., Coventry}, }, date={2007}, pages={155--166}, review={\MR {2402782}}, doi={10.2140/gtm.2007.10.155}, }
Reference [Fre82]
Michael Hartley Freedman, The topology of four-dimensional manifolds, J. Differential Geometry 17 (1982), no. 3, 357–453. MR679066,
Show rawAMSref \bib{Freedman1982}{article}{ label={Fre82}, author={Freedman, Michael Hartley}, title={The topology of four-dimensional manifolds}, journal={J. Differential Geometry}, volume={17}, date={1982}, number={3}, pages={357--453}, issn={0022-040X}, review={\MR {679066}}, }
Reference [Fur01]
M. Furuta, Monopole equation and the -conjecture, Math. Res. Lett. 8 (2001), no. 3, 279–291, DOI 10.4310/MRL.2001.v8.n3.a5. MR1839478,
Show rawAMSref \bib{Furuta2001}{article}{ label={Fur01}, author={Furuta, M.}, title={Monopole equation and the $\frac {11}8$-conjecture}, journal={Math. Res. Lett.}, volume={8}, date={2001}, number={3}, pages={279--291}, issn={1073-2780}, review={\MR {1839478}}, doi={10.4310/MRL.2001.v8.n3.a5}, }
Reference [Hir63]
Morris W. Hirsch, Obstruction theories for smoothing manifolds and maps, Bull. Amer. Math. Soc. 69 (1963), 352–356, DOI 10.1090/S0002-9904-1963-10917-9. MR149493,
Show rawAMSref \bib{Hirsch1963}{article}{ label={Hir63}, author={Hirsch, Morris W.}, title={Obstruction theories for smoothing manifolds and maps}, journal={Bull. Amer. Math. Soc.}, volume={69}, date={1963}, pages={352--356}, issn={0002-9904}, review={\MR {149493}}, doi={10.1090/S0002-9904-1963-10917-9}, }
Reference [Jon85]
John D. S. Jones, Root invariants, cup--products and the Kahn-Priddy theorem, Bull. London Math. Soc. 17 (1985), no. 5, 479–483, DOI 10.1112/blms/17.5.479. MR806016,
Show rawAMSref \bib{Jones}{article}{ label={Jon85}, author={Jones, John D. S.}, title={Root invariants, cup-$r$-products and the Kahn-Priddy theorem}, journal={Bull. London Math. Soc.}, volume={17}, date={1985}, number={5}, pages={479--483}, issn={0024-6093}, review={\MR {806016}}, doi={10.1112/blms/17.5.479}, }
Reference [Kro94]
P. B. Kronheimer, Lecture at Cambridge in December 1994, 1994.
Reference [KS77]
Robion C. Kirby and Laurence C. Siebenmann, Foundational essays on topological manifolds, smoothings, and triangulations, Annals of Mathematics Studies, No. 88, Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1977. With notes by John Milnor and Michael Atiyah. MR0645390,
Show rawAMSref \bib{Kirby-Siebenmann1977}{book}{ label={KS77}, author={Kirby, Robion C.}, author={Siebenmann, Laurence C.}, title={Foundational essays on topological manifolds, smoothings, and triangulations}, series={Annals of Mathematics Studies, No. 88}, note={With notes by John Milnor and Michael Atiyah}, publisher={Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo}, date={1977}, pages={vii+355}, review={\MR {0645390}}, }
Reference [Lan69]
Peter S. Landweber, On equivariant maps between spheres with involutions, Ann. of Math. (2) 89 (1969), 125–137, DOI 10.2307/1970812. MR238313,
Show rawAMSref \bib{Landweber}{article}{ label={Lan69}, author={Landweber, Peter S.}, title={On equivariant maps between spheres with involutions}, journal={Ann. of Math. (2)}, volume={89}, date={1969}, pages={125--137}, issn={0003-486X}, review={\MR {238313}}, doi={10.2307/1970812}, }
Reference [Lin15]
Jianfeng Lin, -equivariant KO-theory and intersection forms of spin 4-manifolds, Algebr. Geom. Topol. 15 (2015), no. 2, 863–902, DOI 10.2140/agt.2015.15.863. MR3342679,
Show rawAMSref \bib{LinKO}{article}{ label={Lin15}, author={Lin, Jianfeng}, title={$\operatorname {Pin}(2)$-equivariant KO-theory and intersection forms of spin 4-manifolds}, journal={Algebr. Geom. Topol.}, volume={15}, date={2015}, number={2}, pages={863--902}, issn={1472-2747}, review={\MR {3342679}}, doi={10.2140/agt.2015.15.863}, }
Reference [LMSM86]
L. G. Lewis Jr., J. P. May, M. Steinberger, and J. E. McClure, Equivariant stable homotopy theory, Lecture Notes in Mathematics, vol. 1213, Springer-Verlag, Berlin, 1986. With contributions by J. E. McClure, DOI 10.1007/BFb0075778. MR866482,
Show rawAMSref \bib{LewisMaySteinberger}{book}{ label={LMSM86}, author={Lewis, L. G., Jr.}, author={May, J. P.}, author={Steinberger, M.}, author={McClure, J. E.}, title={Equivariant stable homotopy theory}, series={Lecture Notes in Mathematics}, volume={1213}, note={With contributions by J. E. McClure}, publisher={Springer-Verlag, Berlin}, date={1986}, pages={x+538}, isbn={3-540-16820-6}, review={\MR {866482}}, doi={10.1007/BFb0075778}, }
Reference [Man14]
Ciprian Manolescu, On the intersection forms of spin four-manifolds with boundary, Math. Ann. 359 (2014), no. 3-4, 695–728, DOI 10.1007/s00208-014-1010-1. MR3231012,
Show rawAMSref \bib{Manolescu}{article}{ label={Man14}, author={Manolescu, Ciprian}, title={On the intersection forms of spin four-manifolds with boundary}, journal={Math. Ann.}, volume={359}, date={2014}, number={3-4}, pages={695--728}, issn={0025-5831}, review={\MR {3231012}}, doi={10.1007/s00208-014-1010-1}, }
Reference [Mat82]
Yukio Matsumoto, On the bounding genus of homology -spheres, J. Fac. Sci. Univ. Tokyo Sect. IA Math. 29 (1982), no. 2, 287–318. MR672065,
Show rawAMSref \bib{Matsumoto1982}{article}{ label={Mat82}, author={Matsumoto, Yukio}, title={On the bounding genus of homology $3$-spheres}, journal={J. Fac. Sci. Univ. Tokyo Sect. IA Math.}, volume={29}, date={1982}, number={2}, pages={287--318}, issn={0040-8980}, review={\MR {672065}}, }
Reference [May01]
J. P. May, The additivity of traces in triangulated categories, Adv. Math. 163 (2001), no. 1, 34–73, DOI 10.1006/aima.2001.1995. MR1867203,
Show rawAMSref \bib{MayTraces}{article}{ label={May01}, author={May, J. P.}, title={The additivity of traces in triangulated categories}, journal={Adv. Math.}, volume={163}, date={2001}, number={1}, pages={34--73}, issn={0001-8708}, review={\MR {1867203}}, doi={10.1006/aima.2001.1995}, }
Reference [Min]
Norihiko Minami, G-join theorem: an unbased G-Freudenthal theorem, homotopy representations, and a Borsuk-Ulam theorem, Preprint.
Reference [Mos70]
R. Michael F. Moss, Secondary compositions and the Adams spectral sequence, Math. Z. 115 (1970), 283–310, DOI 10.1007/BF01129978. MR266216,
Show rawAMSref \bib{Moss}{article}{ label={Mos70}, author={Moss, R. Michael F.}, title={Secondary compositions and the Adams spectral sequence}, journal={Math. Z.}, volume={115}, date={1970}, pages={283--310}, issn={0025-5874}, review={\MR {266216}}, doi={10.1007/BF01129978}, }
Reference [MR93]
Mark E. Mahowald and Douglas C. Ravenel, The root invariant in homotopy theory, Topology 32 (1993), no. 4, 865–898, DOI 10.1016/0040-9383(93)90055-Z. MR1241877,
Show rawAMSref \bib{MahowaldRavenel}{article}{ label={MR93}, author={Mahowald, Mark E.}, author={Ravenel, Douglas C.}, title={The root invariant in homotopy theory}, journal={Topology}, volume={32}, date={1993}, number={4}, pages={865--898}, issn={0040-9383}, review={\MR {1241877}}, doi={10.1016/0040-9383(93)90055-Z}, }
Reference [Mun60]
James Munkres, Obstructions to the smoothing of piecewise-differentiable homeomorphisms, Ann. of Math. (2) 72 (1960), 521–554, DOI 10.2307/1970228. MR121804,
Show rawAMSref \bib{Munkres1960}{article}{ label={Mun60}, author={Munkres, James}, title={Obstructions to the smoothing of piecewise-differentiable homeomorphisms}, journal={Ann. of Math. (2)}, volume={72}, date={1960}, pages={521--554}, issn={0003-486X}, review={\MR {121804}}, doi={10.2307/1970228}, }
Reference [Mun64a]
James Munkres, Obstructions to extending diffeomorphisms, Proc. Amer. Math. Soc. 15 (1964), 297–299, DOI 10.2307/2034057. MR161344,
Show rawAMSref \bib{Munkres1964b}{article}{ label={Mun64a}, author={Munkres, James}, title={Obstructions to extending diffeomorphisms}, journal={Proc. Amer. Math. Soc.}, volume={15}, date={1964}, pages={297--299}, issn={0002-9939}, review={\MR {161344}}, doi={10.2307/2034057}, }
Reference [Mun64b]
James Munkres, Obstructions to imposing differentiable structures, Illinois J. Math. 8 (1964), 361–376. MR180979,
Show rawAMSref \bib{Munkres1964a}{article}{ label={Mun64b}, author={Munkres, James}, title={Obstructions to imposing differentiable structures}, journal={Illinois J. Math.}, volume={8}, date={1964}, pages={361--376}, issn={0019-2082}, review={\MR {180979}}, }
Reference [Roh52]
V. A. Rohlin, New results in the theory of four-dimensional manifolds (Russian), Doklady Akad. Nauk SSSR (N.S.) 84 (1952), 221–224. MR0052101,
Show rawAMSref \bib{Rohlin1952}{article}{ label={Roh52}, author={Rohlin, V. A.}, title={New results in the theory of four-dimensional manifolds}, language={Russian}, journal={Doklady Akad. Nauk SSSR (N.S.)}, volume={84}, date={1952}, pages={221--224}, review={\MR {0052101}}, }
Reference [Sch03]
Birgit Schmidt, Spin -manifolds and -equivariant homotopy theory, Ph.D. Thesis, University of Bielefeld, 2003.
Reference [Ser77]
Jean-Pierre Serre, Cours d’arithmétique (French), Le Mathématicien, No. 2, Presses Universitaires de France, Paris, 1977. Deuxième édition revue et corrigée. MR0498338,
Show rawAMSref \bib{Serre1977}{book}{ label={Ser77}, author={Serre, Jean-Pierre}, title={Cours d'arithm\'{e}tique}, language={French}, series={Le Math\'{e}maticien, No. 2}, note={Deuxi\`eme \'{e}dition revue et corrig\'{e}e}, publisher={Presses Universitaires de France, Paris}, date={1977}, pages={188}, review={\MR {0498338}}, }
Reference [Sto89]
Stephan Stolz, The level of real projective spaces, Comment. Math. Helv. 64 (1989), no. 4, 661–674, DOI 10.1007/BF02564700. MR1023002,
Show rawAMSref \bib{Stolz1989}{article}{ label={Sto89}, author={Stolz, Stephan}, title={The level of real projective spaces}, journal={Comment. Math. Helv.}, volume={64}, date={1989}, number={4}, pages={661--674}, issn={0010-2571}, review={\MR {1023002}}, doi={10.1007/BF02564700}, }
Reference [tD79]
Tammo tom Dieck, Transformation groups and representation theory, Lecture Notes in Mathematics, vol. 766, Springer, Berlin, 1979. MR551743,
Show rawAMSref \bib{tomDieck}{book}{ label={tD79}, author={tom Dieck, Tammo}, title={Transformation groups and representation theory}, series={Lecture Notes in Mathematics}, volume={766}, publisher={Springer, Berlin}, date={1979}, pages={viii+309}, isbn={3-540-09720-1}, review={\MR {551743}}, }
Reference [Tod63]
Hirosi Toda, Order of the identity class of a suspension space, Ann. of Math. (2) 78 (1963), 300–325, DOI 10.2307/1970345. MR156347,
Show rawAMSref \bib{Toda}{article}{ label={Tod63}, author={Toda, Hirosi}, title={Order of the identity class of a suspension space}, journal={Ann. of Math. (2)}, volume={78}, date={1963}, pages={300--325}, issn={0003-486X}, review={\MR {156347}}, doi={10.2307/1970345}, }
Reference [Whi40]
J. H. C. Whitehead, On -complexes, Ann. of Math. (2) 41 (1940), 809–824, DOI 10.2307/1968861. MR2545,
Show rawAMSref \bib{Whitehead}{article}{ label={Whi40}, author={Whitehead, J. H. C.}, title={On $C^1$-complexes}, journal={Ann. of Math. (2)}, volume={41}, date={1940}, pages={809--824}, issn={0003-486X}, review={\MR {2545}}, doi={10.2307/1968861}, }
Reference [Wu50]
Wen-tsün Wu, Classes caractéristiques et -carrés d’une variété (French), C. R. Acad. Sci. Paris 230 (1950), 508–511. MR35992,
Show rawAMSref \bib{Wu1950}{article}{ label={Wu50}, author={Wu, Wen-ts\"{u}n}, title={Classes caract\'{e}ristiques et $i$-carr\'{e}s d'une vari\'{e}t\'{e}}, language={French}, journal={C. R. Acad. Sci. Paris}, volume={230}, date={1950}, pages={508--511}, issn={0001-4036}, review={\MR {35992}}, }
Reference [WX17]
Guozhen Wang and Zhouli Xu, The triviality of the 61-stem in the stable homotopy groups of spheres, Ann. of Math. (2) 186 (2017), no. 2, 501–580, DOI 10.4007/annals.2017.186.2.3. MR3702672,
Show rawAMSref \bib{WangXu}{article}{ label={WX17}, author={Wang, Guozhen}, author={Xu, Zhouli}, title={The triviality of the 61-stem in the stable homotopy groups of spheres}, journal={Ann. of Math. (2)}, volume={186}, date={2017}, number={2}, pages={501--580}, issn={0003-486X}, review={\MR {3702672}}, doi={10.4007/annals.2017.186.2.3}, }
Reference [WX18]
Guozhen Wang and Zhouli Xu, Some extensions in the Adams spectral sequence and the 51-stem, Algebr. Geom. Topol. 18 (2018), no. 7, 3887–3906, DOI 10.2140/agt.2018.18.3887. MR3892234,
Show rawAMSref \bib{WangXu51Stem}{article}{ label={WX18}, author={Wang, Guozhen}, author={Xu, Zhouli}, title={Some extensions in the Adams spectral sequence and the 51-stem}, journal={Algebr. Geom. Topol.}, volume={18}, date={2018}, number={7}, pages={3887--3906}, issn={1472-2747}, review={\MR {3892234}}, doi={10.2140/agt.2018.18.3887}, }
Reference [Xu16]
Zhouli Xu, The strong Kervaire invariant problem in dimension 62, Geom. Topol. 20 (2016), no. 3, 1611–1624, DOI 10.2140/gt.2016.20.1611. MR3523064,
Show rawAMSref \bib{Xu}{article}{ label={Xu16}, author={Xu, Zhouli}, title={The strong Kervaire invariant problem in dimension 62}, journal={Geom. Topol.}, volume={20}, date={2016}, number={3}, pages={1611--1624}, issn={1465-3060}, review={\MR {3523064}}, doi={10.2140/gt.2016.20.1611}, }

Article Information

MSC 2020
Primary: 55P91 (Equivariant homotopy theory in algebraic topology), 57K41 (Invariants of 4-manifolds (e.g., Donaldson and Seiberg-Witten invariants)), 57K40 (General topology of 4-manifolds)
Author Information
Michael J. Hopkins
Department of Mathematics, Harvard University, 1 Oxford St., Cambridge, Massachusetts 02318
ORCID
MathSciNet
Jianfeng Lin
Yau Mathematical Sciences Center, Jing Zhai, Tsinghua University, Hai Dian District, Beijing 100084, People’s Republic of China
MathSciNet
XiaoLin Danny Shi
Department of Mathematics, University of Chicago, 5734 S University Ave, Chicago, Illinois 60637
MathSciNet
Zhouli Xu
Department of Mathematics, UC San Diego, 9500 Gilman Dr., La Jolla, California 92093
ORCID
MathSciNet
Additional Notes

The first author was supported by NSF grant DMS-1810917.

The second author was supported by NSF grant DMS-1707857; and the fourth author was supported by NSF grant DMS-1810638.

Journal Information
Communications of the American Mathematical Society, Volume 2, Issue 2, ISSN 2692-3688, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2022 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/cams/4
  • MathSciNet Review: 4385297
  • Show rawAMSref \bib{4385297}{article}{ author={Hopkins, Michael}, author={Lin, Jianfeng}, author={Shi, XiaoLin Danny}, author={Xu, Zhouli}, title={Intersection forms of spin 4-manifolds and the pin(2)-equivariant Mahowald invariant}, journal={Comm. Amer. Math. Soc.}, volume={2}, number={2}, date={2022}, pages={22-132}, issn={2692-3688}, review={4385297}, doi={10.1090/cams/4}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.