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

Periodicity Uncovered: A Deep Dive into Bott’s Theorems in K-Theory and Fiber Bundles

Ivan Z. Feng ifeng@usc.edu  | dornsife.usc.edu/ivan
(May 10, 2023)
Abstract

This paper presents a comparison between two versions of Bott Periodicity Theorems: one in topological K-theory and the other in stable homotopy groups of classical groups. It begins with an introduction to K-theory, discussing vector bundles and their role in understanding the algebraic and topological aspects of these spaces. Then the two versions of Bott periodicity, as well as the topological notions necessary to understand them, are further explored. The aim is to illustrate the connections and distinctions between these two theorems, deepening our understanding of their underlying mathematical structures such as topological K-theory and fiber bundles.

1 Introduction to Essentials of K-Theory

In this section, we aim to introduce two versions of K-theory, K(X)K(X) and K~(X)\widetilde{K}(X) for a fixed base space XX, as well as their related motivations and properties. They are the foundation of further development of the Bott periodicity.

1.1 Vector Bundles

Let’s begin with the foundation of topological K-theory: vector bundles. The concept of vector bundles originated in the early 20th century, with significant contributions from mathematicians such as Hermann Weyl, Hassler Whitney, and Norman Steenrod. It has since become a fundamental object of study in algebraic and differential topology, as well as a central tool in areas such as gauge theory and index theory.

The main idea behind vector bundles is to associate a vector space to every point of a topological space XX in such a way that these vector spaces fit together in a locally trivial manner. This structure allows us to study local properties of the base space XX by analyzing the corresponding properties of the associated vector spaces. With this motivation in mind, we can now introduce the formal definition of vector bundles:

Definition 1.1 (Vector Bundle).

A vector bundle is a family of vector spaces p:EXp:E\rightarrow X that has a local strong basis at every point of XX. In other words, for each xXx\in X there is a neighborhood xUXx\in U\subseteq X, an n0{}n\in\mathbb{Z}_{\geq 0}\cup\{\infty\}, and an isomorphism111In this paper we adopt the symbol \approx to mean isomorphism as used in Hatcher’s [8] and [9], instead of the more ubiquitous symbol \cong as used in [1] and [4]. In general, the symbol \approx is usually used to denote a weaker notion of equivalence that may allow for more flexibility or ambiguity in the choice of isomorphisms, while \cong is a stronger notion of equivalence that implies a more canonical choice of isomorphism. of families of vector spaces

p1(U){p^{-1}(U)}U×n{{U\times\mathbb{R}^{n}}}U{U}\approx

Note that when n=n=\infty the space n\mathbb{R}^{n} has the colimit topology. The isomorphism in the above diagram is called a local trivialization. Usually, one simply says that a vector bundle is a family of vector spaces that is locally trivial. In particular, the trivial bundle (or product bundle) is given by E=B×nE=B\times\mathbb{R}^{n} with pp the projection onto the first factor. Let’s provide some more typical examples.

Example 1.2 (Line Bundle).

Let EE be the quotient space of I×I\times\mathbb{R} under the identifications (0,t)(1,t)(0,t)\sim(1,-t). The projection I×II\times\mathbb{R}\rightarrow I induces a map p:ES1p:E\rightarrow S^{1}, which is a 1-dimensional vector bundle. We call this bundle the line bundle. Here since EE is homeomorphic to a Möbius band with its boundary circle deleted, we also call this bundle the Möbius bundle.

Example 1.3 (Canonical Line Bundle).

Consider the real projective nn-space, denoted as Pn\mathbb{R}\mathrm{P}^{n}, which represents the set of lines in n+1\mathbb{R}^{n+1} that pass through the origin. Since each such line intersects the nn-dimensional unit sphere SnS^{n} in a pair of antipodal points, we can view Pn\mathbb{R}\mathrm{P}^{n} as the quotient space of SnS^{n}, where antipodal pairs of points are identified. Corresponding to this space, we have the canonical line bundle p:EPnp:E\rightarrow\mathbb{R}\mathrm{P}^{n}. The total space of this bundle, EE, is a subspace of Pn×n+1\mathbb{R}\mathrm{P}^{n}\times\mathbb{R}^{n+1}, consisting of pairs (,v)(\ell,v), where vv\in\ell. The projection map pp is defined as p(,v)=p(\ell,v)=\ell. Local trivializations can be defined by orthogonal projection.

Now, to further examine this structure, let’s proceed with the following definition.

Definition 1.4 (Vectn(B)\operatorname{Vect}^{n}(B) and Vectn(B)\operatorname{Vect}_{\mathbb{C}}^{n}(B)).

The set of isomorphism classes of nn-dimensional real vector bundles over BB is denoted by Vectn(B)\operatorname{Vect}^{n}(B)222As a convention, we usually omit \mathbb{R} as it is a common default assumption in math.. Similarly, the set of isomorphism classes of nn-dimensional complex vector bundles over BB is denoted by Vectn(B)\operatorname{Vect}_{\mathbb{C}}^{n}(B).

Now let’s use the collection of vector bundles over a base space to explore how a continuous map between two spaces can relate their corresponding vector bundles. This leads us to the concept of pullback bundles.

Definition 1.5 (Pullback Bundle).

Given a map f:ABf:A\rightarrow B and a vector bundle p:EBp:E\rightarrow B, there exists a unique (up to isomorphism) vector bundle p:EAp^{\prime}:E^{\prime}\rightarrow A with a map f:EEf^{\prime}:E^{\prime}\rightarrow E such that the fiber of EE^{\prime} over each point aAa\in A is isomorphic onto the fiber of EE over f(a)f(a). This vector bundle EE^{\prime} is called the pullback of EE by ff and is often denoted as f(E)f^{*}(E).

Notice that, in this definition, we assumed the existence and uniqueness of the vector bundles pp^{\prime} and EE^{\prime}. In fact, this can be easily proven as shown in Section 1.2 of [9]. The pullback bundle allows us to study how the structure of vector bundles is preserved or transformed under continuous maps between base spaces. By considering the uniqueness of the pullback bundle, we can establish a function f:Vect(B)Vect(A)f^{*}:\operatorname{Vect}(B)\rightarrow\operatorname{Vect}(A), taking the isomorphism class of EE to the isomorphism class of EE^{\prime}. This function gives us a way to relate the vector bundles of two spaces through a continuous map, and it further enriches our understanding of the interplay between topology and vector bundles.

1.2 Two Versions of K-theory: K~(X)\widetilde{K}(X) and K(X){K}(X)

Building upon vector bundles, K-theory, which was developed by Michael Atiyah and Friedrich Hirzebruch in the late 1950s, is a powerful tool in algebraic topology and has since found applications in various branches of mathematics.

The reduced K-theory group K~(X)\widetilde{K}(X) is introduced to better understand and organize the structure of vector bundles over a fixed base space XX. To define it, we first need to introduce a related concept and establish some inspiration.

Definition 1.6 (Stable Isomorphism).

Two vector bundles E1E_{1} and E2E_{2} over XX are called stably isomorphic, written E1sE2E_{1}\approx_{s}E_{2}, if E1εnE2εnE_{1}\oplus\varepsilon^{n}\approx E_{2}\oplus\varepsilon^{n} for some nn, where εn\varepsilon^{n} is the trivial nn-dimensional vector bundle over XX.

The motivation for this definition is to identify bundles that differ only by a trivial bundle, focusing on their essential structure. This idea of identification is widely used in math; for example, we identify spaces that are homotopy equivalent or homeomorphic, thereby focusing on their intrinsic topological properties rather than the specific details of their construction.

Now, let’s consider the set of stable isomorphism classes of vector bundles over XX. We can define an abelian group structure on this set, as shown in the following proposition:

Proposition 1.7.

If XX is compact Hausdorff, then the set of stable isomorphism classes of vector bundles over XX forms an abelian group with respect to the direct sum operation \oplus.

The motivation for this proposition is to organize vector bundles over a fixed base space XX into a group structure, which allows us to study their properties and interactions more systematically. Its proof relies on showing that the set of stable isomorphism classes of vector bundles satisfies the group axioms (closure, associativity, existence of identity, and existence of inverses) under the direct sum operation.

Once we have established this fact, we can now introduce the reduced K-theory group K~(X)\widetilde{K}(X).

Definition 1.8 (Reduced K-theory Group).

Given a compact Hausdorff space XX, the abelian group formed by the set of stable isomorphism classes of vector bundles over XX under the direct sum operation \oplus is called the reduced K-theory group, denoted by K~(X)\widetilde{K}(X).

Let us provide an example for a better understanding of these definitions above.

Example 1.9.

Let X=:S1X=:S^{1}, the circle. Consider the following two complex line bundles over XX: the trivial bundle E1=ε1E_{1}=\varepsilon^{1} and the Möbius bundle E2E_{2}. By Example 1.2, the Möbius bundle E2E_{2} is not isomorphic to the trivial bundle E1E_{1}. However, we can show that E2E2ε1E_{2}\oplus E_{2}\approx\varepsilon^{1}, which means, by definition, that E2E_{2} and the trivial bundle ε1\varepsilon^{1} are stably isomorphic and belong to the same stable isomorphism class in K~(X)\widetilde{K}(X).

By defining the reduced K-theory group, we create a structure that allows us to analyze vector bundles in a systematic way. This group helps us uncover relationships between vector bundles and other algebraic and topological concepts, paving the way for motivation of the following definition.

Definition 1.10 (K-theory Group K(X)K(X)).

The K-theory group K(X)K(X) consists of formal differences EEE-E^{\prime} of complex vector bundles EE and EE^{\prime} over XX, modulo an equivalence relation defined by "E1E1E2E2E_{1}-E_{1}^{\prime}\sim E_{2}-E_{2}^{\prime} if and only if E1E2sE2E1E_{1}\oplus E_{2}^{\prime}\approx_{s}E_{2}\oplus E_{1}^{\prime}". The group operation is given by the direct sum of vector bundles, i.e., (E1E1)(E2E2)=(E1E2)(E1E2)(E_{1}-E_{1}^{\prime})\oplus(E_{2}-E_{2}^{\prime})=(E_{1}\oplus E_{2})-(E_{1}^{\prime}\oplus E_{2}^{\prime}).

The motivation for this definition is to create a more refined group structure that captures differences between vector bundles.

Example 1.11 (K-theory Group of a Single Point).

Consider the base space XX to be a single point, denoted as ptpt. We will explore the K-theory group K(pt)K(pt) for this space.

Since XX consists of a single point, the only vector bundles over XX are trivial bundles. Let nn and mm be non-negative integers representing the dimensions of the trivial bundles εnpt\varepsilon^{n}\rightarrow pt and εmpt\varepsilon^{m}\rightarrow pt, respectively. Notice that the direct sum of these bundles is another trivial bundle of dimension n+mn+m, i.e., εnεmεn+m\varepsilon^{n}\oplus\varepsilon^{m}\approx\varepsilon^{n+m}.

By the definition above, the K-theory group K(pt)K(pt) consists of formal differences of trivial bundles, such as εnεm\varepsilon^{n}-\varepsilon^{m}. We can see that εnεmεkεl\varepsilon^{n}-\varepsilon^{m}\sim\varepsilon^{k}-\varepsilon^{l} if and only if εnεlsεkεm\varepsilon^{n}\oplus\varepsilon^{l}\approx_{s}\varepsilon^{k}\oplus\varepsilon^{m}. Since all the bundles are trivial, it follows that εnεlεn+l\varepsilon^{n}\oplus\varepsilon^{l}\approx\varepsilon^{n+l} and εkεmεk+m\varepsilon^{k}\oplus\varepsilon^{m}\approx\varepsilon^{k+m}. Thus, the equivalence relation simply reduces to n+l=k+mn+l=k+m. In addition, the group operation in K(pt)K(pt) is given by the direct sum of the differences of trivial bundles, i.e., (εnεm)(εkεl)=(εnεk)(εmεl)=εn+kεm+l(\varepsilon^{n}-\varepsilon^{m})\oplus(\varepsilon^{k}-\varepsilon^{l})=(\varepsilon^{n}\oplus\varepsilon^{k})-(\varepsilon^{m}\oplus\varepsilon^{l})=\varepsilon^{n+k}-\varepsilon^{m+l}.

From these observations, we can see that K(pt)K(pt) is isomorphic to the group of integers \mathbb{Z} under the operation of addition, with the isomorphism given by εnεmnm\varepsilon^{n}-\varepsilon^{m}\mapsto n-m.

Again, the motivation behind the reduced K-theory group K~(X)\widetilde{K}(X) and the K-theory group K(X)K(X) is to provide a systematic way of studying vector bundles and their interactions. By organizing vector bundles into these groups, we are enabled to uncover deep results and explore various applications in algebraic topology, geometry, and mathematical physics.

1.3 The Fundamental Product Theorem

In Example 1.11, we have gained an in-depth understanding of K-theory groups of a single point. In this subsection, we shall focus on the Fundamental Product Theorem, a key result that allows us to compute K(X)K(X) for nontrivial cases. Specifically, we aim to derive a formula for calculating K(X×S2)K\left(X\times S^{2}\right) in terms of K(X)K(X). This formula will be crucial for deducing Bott Periodicity in the subsequent section.

Recall the canonical line bundle HH over S2S^{2} is isomorphic to P1\mathbb{C}P^{1}. It is easy to show the canonical line bundle HP1H\rightarrow\mathbb{C}P^{1} satisfies the relation (HH)1HH(H\otimes H)\oplus 1\approx H\oplus H. (Details see Example 1.13 in [9]). In K(S2)K\left(S^{2}\right), this translates to the formula H2+1=2HH^{2}+1=2H, i.e., H2=2H1H^{2}=2H-1, or, (H1)2=0(H-1)^{2}=0. Consequently, we have a natural ring homomorphism [H]/(H1)2K(S2)\mathbb{Z}[H]/(H-1)^{2}\rightarrow K\left(S^{2}\right), with the domain being the quotient ring of the polynomial ring [H]\mathbb{Z}[H] by the ideal generated by (H1)2(H-1)^{2}. Note that an additive basis for [H]/(H1)2\mathbb{Z}[H]/(H-1)^{2} is {1,H}\{1,H\}.

Definition 1.12 (External Product).

The external product is a map that combines the KK-theory groups of two spaces into a single KK-theory group of their Cartesian product. It is defined as

K(X)K(Y)K(X×Y).K(X)\otimes K(Y)\rightarrow K(X\times Y).

With the external product in place, we can define a homomorphism μ\mu as the composition:

μ:K(X)[H]/(H1)2K(X)K(S2)K(X×S2),\mu:K(X)\otimes\mathbb{Z}[H]/(H-1)^{2}\rightarrow K(X)\otimes K\left(S^{2}\right)\rightarrow K\left(X\times S^{2}\right),

where the second map is the external product. This homomorphism indeed can be further shown to be an isomorphism:

Theorem 1.13 (Fundamental Product Theorem [9]).

The homomorphism μ:K(X)[H]/(H1)2K(X×S2)\mu:K(X)\otimes\mathbb{Z}[H]/(H-1)^{2}\rightarrow K\left(X\times S^{2}\right) is an isomorphism of rings for all compact Hausdorff spaces XX.

By taking XX as a point, we obtain the following corollary:

Corollary 1.14.

The map [H]/(H1)2K(S2)\mathbb{Z}[H]/(H-1)^{2}\rightarrow K\left(S^{2}\right) is an isomorphism of rings.

Thus, when we regard K~(S2)\widetilde{K}\left(S^{2}\right) as the kernel of K(S2)K(x0)K\left(S^{2}\right)\rightarrow K\left(x_{0}\right), it is generated as an abelian group by H1H-1. Since we have the relation (H1)2=0(H-1)^{2}=0, this means that the multiplication in K~(S2)\widetilde{K}\left(S^{2}\right) is completely trivial: the product of any two elements is zero.333Here the situation is similar to that in H(S2;)H\left(S^{2};\mathbb{Z}\right) and H~(S2;)\widetilde{H}\left(S^{2};\mathbb{Z}\right), with H1H-1 behaving like the generator of H2(S2;)H^{2}\left(S^{2};\mathbb{Z}\right).

2 Bott Periodicity for Topological K-theory

In this section, we delve into the intricate world of the first version of Bott Periodicity: Bott Periodicity for Topological K-theory, and explore its mathematical foundations, properties, theorems, and applications.

2.1 Exact Sequences and Splitness

One of the key properties of the reduced groups K~(X)\tilde{K}(X) is the existence of exact sequences. Exact sequences are central to understanding the structure of the groups and will be useful in proving the periodicity results in K\mathrm{K}-theory. We begin with the most general definition of exact sequences.

Definition 2.1 (Exact Sequence).

An exact sequence is a sequence of morphisms in an abelian category (whose objects are groups, rings, modules, and so on) where the image of one morphism is equal to the kernel of the next.

In particular, a short exact sequence is an exact sequence of the form

0AfBgC0,0\longrightarrow A\stackrel{{\scriptstyle f}}{{\longrightarrow}}B\stackrel{{\scriptstyle g}}{{\longrightarrow}}C\longrightarrow 0,

where f:ABf:A\rightarrow B and g:BCg:B\rightarrow C. Here our universe is the abelian category of abelian groups. By definition, the short sequence is exact if ff is injective, gg is surjective, and im(f)=ker(g)\operatorname{im}(f)=\operatorname{ker}(g). We also call this short exact sequence an extension of CC by AA.444See the following example for the explanation of another convention ”extension of AA by CC”. Moreover, if AZ(B)A\subset Z(B), namely AA is included in the center of the group BB, we call this sequence a central extension. Let’s use the following example about the second K-theory group K2(R)K_{2}(R) of rings555Notice in Section 1, the K-theory group we defined is for a compact Hausdorff (topological) space XX that doesn’t necessarily have a ring structure. in algebraic geometry to better understand the intuition behind central extensions.

Example 2.2 (Second K-theory Group of Rings).

We define K2(R):=ker(φ:St(R)GL(R))K_{2}(R):=\operatorname{ker}(\varphi:{St}(R)\rightarrow GL(R)) for a ring RR, called the second K-theory group. Here St(R)St(R), called the (stable) Steinberg group of RR , is defined as the direct limit of Stn(R):=xij(r),1<k,jn | 𝐑St_{n}\left(R\right):=\langle x_{ij}\left(r\right),1<k,j\leqslant n\textbf{ }|\textbf{ }\mathbf{R}\rangle, where rRr\in R, and the bold R refers to the Steinberg relations. By definition, we can show K2(R)=Z(St(R))K_{2}(R)=Z(St(R)), i.e., K2(R)K_{2}(R) is precisely the center of St(R)St(R). (See [5] for more details and examples of this group.)

First, in particular, we obtain K2(R)K_{2}(R) is always an abelian group. Also, by definition, 1K2(R)St(R)E(R)11\rightarrow K_{2}(R)\rightarrow St(R)\rightarrow E(R)\rightarrow 1 is a central extension of E(R)E(R) by K2(R)K_{2}(R); shortly, we say St(R)St(R) or St(R)E(R)St(R)\rightarrow E(R) (not the whole exact sequence) is a central extension. Here “central”, of course, means the subgroup K2(R)K_{2}(R) lies in the center of St(R)St(R). As for the phrase “extension of E(R)E(R) by K2(R)K_{2}(R)”, in fact, the idea of “central extension” is partially motivated and regularly used in considering homotopy classes in homotopical algebra, and it’s natural to say a central extension is of the “base” by the “fiber”. As a result, we sometimes extend this idea to all extensions in general; that is, in the short exact sequence 0AfBgC00\rightarrow A\stackrel{{\scriptstyle f}}{{\rightarrow}}B\stackrel{{\scriptstyle g}}{{\rightarrow}}C\rightarrow 0, we may call this sequence an extension of CC by AA as defined above. However, the same sequence can also be called an extension of AA by CC in the literature, for example, in [10]. This latter naming is motivated and based on different origins, such as Galois theory.666Intuitively, in this example of the second K-group, the map St(R)E(R)St(R)\to E(R) is surjective, namely St(R)St(R) is “bigger” than E(R)E(R), so we can call St(R)St(R) an extension of E(R)E(R). On the other hand, K2(R)St(R)K_{2}(R)\rightarrow St(R) is injective, so it also makes sense when we call St(R)St(R) an extension of K2(R)K_{2}(R).

Furthermore, we can give short exact sequences another structure:

Definition 2.3 (Splitness).

A short exact sequence

0AfBgC00\rightarrow A\stackrel{{\scriptstyle f}}{{\longrightarrow}}B\stackrel{{\scriptstyle g}}{{\longrightarrow}}C\rightarrow 0

is split if there exists a map h:CBh:C\rightarrow B with gh=1Cgh=1_{C}

We have the following direct consequence from abstract algebra:

Proposition 2.4.

If an exact sequence

0AfBgC00\rightarrow A\stackrel{{\scriptstyle f}}{{\longrightarrow}}B\stackrel{{\scriptstyle g}}{{\longrightarrow}}C\rightarrow 0

is split, then BACB\approx A\oplus C.

2.2 Long Exact Sequences in K-theory

Motivated by the example in the last subsection and the general notion of a sequence being exact, it’s natural to form the following definition of exact sequences of homomorphisms as a particular case:

Definition 2.5 (Exact Sequence of Homomorphisms).

A sequence of homomorphisms G1G2GnG_{1}\rightarrow G_{2}\rightarrow\cdots\rightarrow G_{n} is said to be exact if at each intermediate group GiG_{i}, the kernel of the outgoing map equals the image of the incoming map.

With this definition in mind, now it’s time to dive into the following main property:

Proposition 2.6 (Exact Sequence Theorem for Compact Hausdorff Spaces).

If XX is compact Hausdorff and AXA\subset X is a closed subspace, then the inclusion and quotient maps AiXqX/AA\stackrel{{\scriptstyle i}}{{\longrightarrow}}X\stackrel{{\scriptstyle q}}{{\longrightarrow}}X/A induce an exact sequence of homomorphisms K~(X/A)q\tilde{K}(X/A)\stackrel{{\scriptstyle q^{*}}}{{\longrightarrow}} K~(X)iK~(A)\tilde{K}(X)\stackrel{{\scriptstyle i^{*}}}{{\longrightarrow}}\widetilde{K}(A). That is, the inclusion map i:AXi:A\hookrightarrow X and quotient map q:XX/Aq:X\to X/A induce homomorphisms q:K~(X/A)K~(X)q^{*}:\tilde{K}(X/A)\to\tilde{K}(X) and i:K~(X)K~(A)i^{*}:\tilde{K}(X)\to\tilde{K}(A) such that Ker(i)=Im(q)\mathrm{Ker}(i^{*})=\mathrm{Im}(q^{*}).

The motivation behind this proposition is to relate the K\mathrm{K}-theory of XX, AA, and X/AX/A through exact sequences. This will help us understand how K\mathrm{K}-theory behaves under various operations on topological spaces. Below is a sketch of the proof; details can be found in [9].

Proof.

Naturally, to show two sets are equal, we need to prove the two directions: showing that ImqKeri\operatorname{Im}q^{*}\subset\operatorname{Ker}i^{*} and KeriImq\operatorname{Ker}i^{*}\subset\operatorname{Im}q^{*}.

For the first part (easier direction), we need to show that iq=0i^{*}q^{*}=0. This follows directly from the fact that qiqi is equal to the composition AA/AX/AA\rightarrow A/A\hookrightarrow X/A, and K~(A/A)=0\tilde{K}(A/A)=0.

The second part (harder direction) involves showing that if the restriction over AA of a vector bundle p:EXp:E\rightarrow X is stably trivial, then there exists a vector bundle over X/AX/A such that q(E/h)Eq(E/h)\approx E. This is done by constructing a quotient space E/hE/h of EE under certain identifications and proving that it is a vector bundle. The key observation is that EE is trivial over a neighborhood of AA, and we can use this fact to construct local trivializations over a neighborhood of the point A/AA/A. Finally, we easily verify that Eq(E/h)E\approx q^{*}(E/h). ∎

Building upon Proposition 2.6, we now aim to extend the exact sequence K~(X/A)K~(X)K~(A)\tilde{K}(X/A)\rightarrow\tilde{K}(X)\rightarrow\widetilde{K}(A) to the left, using a diagram involving cones and suspensions denoted by CC and SS. This extension will provide a deeper understanding of the interrelationships between the K~\tilde{K} groups of different spaces:

AXXCA(XCA)CX((XCA)CX)C(XCA)\begin{array}[]{ccc}A\hookrightarrow X\hookrightarrow X\cup CA\hookrightarrow(X\cup CA)\cup CX\hookrightarrow((X\cup CA)\cup CX)\cup C(X\cup CA)\\ \end{array}

In this sequence, each space is obtained from its predecessor by attaching a cone on the subspace two steps back in the sequence. This process illustrates how the spaces are related to each other, with the cone construction allowing us to capture local information about the spaces.

Moreover, for the last three spaces we have the following quotient maps:

{XCAX/A(XCA)CXSA((XCA)CX)C(XCA)SX\begin{cases}X\cup CA\rightarrow X/A\\ (X\cup CA)\cup CX\rightarrow SA\\ ((X\cup CA)\cup CX)\cup C(X\cup CA)\rightarrow SX\end{cases}

These quotient maps are obtained by collapsing the most recently attached cone to a point. Oftentimes, the quotient map collapsing a contractible subspace to a point is a homotopy equivalence, thus inducing an isomorphism on K~\tilde{K}. This observation leads us to the following lemma, which will be a key ingredient in obtaining a long exact sequence of K~\tilde{K} groups:

Lemma 2.7 (Cone Quotient Lemma).

Let AA be a contractible subspace of XX. If q:XX/Aq:X\rightarrow X/A is the quotient map obtained by collapsing AA to a point, then the induced map q:Vectn(X/A)Vectn(X)q^{*}:\operatorname{Vect}^{n}(X/A)\rightarrow\operatorname{Vect}^{n}(X) is a bijection for all n0n\geq 0.

The Cone Quotient Lemma illustrates the importance of understanding how contractible subspaces relate to the sequence above. By collapsing a contractible subspace to a point, we can obtain useful information about the topological properties of the spaces and their associated vector bundles.

Proof.

The idea of the proof is as follows. We first consider a vector bundle EXE\rightarrow X. It must be trivial over AA since AA is contractible. By using a trivialization hh, we can construct a vector bundle E/hX/AE/h\rightarrow X/A and demonstrate that the isomorphism class of E/hE/h does not depend on hh. Finally, we establish a well-defined map between the vector bundle spaces and show that it is an inverse to qq^{*}. ∎

In conclusion, let’s combine the results of Lemma 2.7 and Proposition 2.6. That gives us the main theorem in this subsection.

Theorem 2.8 (Long Exact Sequence in K-theory).

Let XX be a topological space, and let AXA\subset X be a closed subspace. Then there exists a long exact sequence of K-theory groups:

K~(SX)K~(SA)K~(X/A)K~(X)K~(A)\cdots\rightarrow\widetilde{K}(SX)\rightarrow\widetilde{K}(SA)\rightarrow\widetilde{K}(X/A)\rightarrow\widetilde{K}(X)\rightarrow\widetilde{K}(A)\rightarrow\cdots

This theorem is significant in the sense that it provides a way to compute the K-theory groups of more complicated spaces, such as spheres, which is essential for the development of the Bott Periodicity Theorem in the next subsection.

Example 2.9.

In particular, let X:=ABX:=A\vee B, be the wedge sum (or wedge product) of the spaces AA and BB. That is, XX is formed by identifying a basepoint from AA and a basepoint from BB. When we consider the quotient space X/AX/A, we are essentially collapsing the entire subspace AA to a single point. Since XX is formed by identifying a basepoint in AA with a basepoint in BB, collapsing AA to a single point also identifies the basepoint of BB with this collapsed point. The resulting space is (homeomorphic to) BB because it retains the same structure as BB, except that its basepoint has been identified with the collapsed point. Thus, we have X/A=BX/A=B.

In this case, the sequence in Theorem 2.8 breaks up into split short exact sequences. By Proposition 2.4, this implies that we have the following isomorphism

K~(AB)K~(A)K~(B)\widetilde{K}(A\vee B)\approx\widetilde{K}(A)\oplus\widetilde{K}(B)

obtained by restriction to AA and BB. This result demonstrates the close relationship between the K~\tilde{K} groups of AA, BB, and XX and highlights the additivity property of K-theory for wedge sums.

2.3 Deducing Periodicity from Exact Sequences

In this subsection, we aim to finally deduce periodicity properties of the K-theory using the Product Theorem as discussed Section 1.3. To achieve this, we will utilize the smash product and reduced external product, ultimately leading us to the Bott Periodicity Theorem.

Definition 2.10 (Smash Product).

Given topological spaces XX and YY with chosen basepoints x0Xx_{0}\in X and y0Yy_{0}\in Y, the smash product of XX and YY, denoted XYX\wedge Y, is the quotient space X×Y/XYX\times Y/X\vee Y, where XY:=X×{y0}{x0}×YX×YX\vee Y:=X\times\{y_{0}\}\cup\{x_{0}\}\times Y\subset X\times Y is the wedge sum for chosen basepoints x0Xx_{0}\in X and y0Yy_{0}\in Y as discussed in Example 2.9.

The smash product combines properties of both XX and YY, while simplifying their topology by identifying certain parts of the product. This construction will help us in formulating the reduced external product.

Definition 2.11 (Reduced Suspension).

Given a space ZZ with basepoint z0z_{0}, the reduced suspension of ZZ, denoted ΣZ\Sigma Z, is the quotient space of the suspension SZSZ obtained by collapsing the segment {z0}×I\{z_{0}\}\times I to a point, where I:=[0,1]I:=[0,1] is the unit interval.

The reduced suspension is also a functional construction that preserves many properties of the original space while simplifying the topology. For this concept, we have the following natural (linear) property:

Proposition 2.12 (Distributive Property of Reduced Suspension).

Given compact Hausdorff spaces XX and YY, the reduced suspension distributes over the wedge sum:

Σ(XY)=ΣXΣY.\Sigma(X\vee Y)=\Sigma X\vee\Sigma Y.
Proof.

This property can be intuitively understood by definition. The reduced suspension can be viewed as a quotient of the ordinary suspension, collapsing an interval to a point. When applied to the wedge sum, the intervals corresponding to the base points of XX and YY are identified, resulting in the wedge sum of the reduced suspensions. ∎

Now, let us consider the long exact sequence for the pair (X×Y,XY)(X\times Y,X\vee Y):

K~(S(X×Y))K~(S(XY))K~(XY)K~(X×Y)K~(XY)\begin{gathered}\widetilde{K}(S(X\times Y))\longrightarrow\widetilde{K}(S(X\vee Y))\longrightarrow\widetilde{K}(X\wedge Y)\longrightarrow\widetilde{K}(X\times Y)\longrightarrow\widetilde{K}(X\vee Y)\end{gathered} (1)

We can easily observe:

Proposition 2.13.

For compact Hausdorff spaces XX and YY as shown in the sequence above, we have the following isomorphisms:

  1. 1.

    (Second) K~(S(XY))K~(SX)K~(SY)\widetilde{K}(S(X\vee Y))\approx\widetilde{K}(SX)\oplus\widetilde{K}(SY).

  2. 2.

    (Last) K~(XY)K~(X)K~(Y)\widetilde{K}(X\vee Y)\approx\widetilde{K}(X)\oplus\widetilde{K}(Y).

Proof.

2. The last isomorphism is a direct consequence of the Example 2.9 in the previous subsection.

1. To prove the first isomorphism, we begin by the reduced suspension distributes over the wedge sum Σ(XY)=ΣXΣY\Sigma(X\vee Y)=\Sigma X\vee\Sigma Y by Proposition 2.12. Now, let us consider the K-theory of both sides of this equation; we have

K~(Σ(XY))=K~(ΣXΣY).\widetilde{K}(\Sigma(X\vee Y))=\widetilde{K}(\Sigma X\vee\Sigma Y).

Applying the K-theory direct sum property ("the last isomorphism" proven above) to the right-hand side, we obtain:

K~(ΣXΣY)=K~(ΣX)K~(ΣY).\widetilde{K}(\Sigma X\vee\Sigma Y)=\widetilde{K}(\Sigma X)\oplus\widetilde{K}(\Sigma Y).

Thus

K~(Σ(XY))=K~(ΣX)K~(ΣY).\widetilde{K}(\Sigma(X\vee Y))=\widetilde{K}(\Sigma X)\oplus\widetilde{K}(\Sigma Y).

By Lemma 2.7, we know that the quotient map q:ΣZSZq:\Sigma Z\rightarrow SZ induces an isomorphism K~(ΣZ)K~(SZ)\widetilde{K}(\Sigma Z)\approx\widetilde{K}(SZ) for Z=X,Y,XYZ=X,Y,X\vee Y. In other words, the reduced suspensions induces isomorphisms:

{K~(ΣX)K~(SX)K~(ΣY)K~(SY)K~(Σ(XY))K~(S(XY)).\begin{cases}\widetilde{K}(\Sigma X)\approx\widetilde{K}(SX)\\ \widetilde{K}(\Sigma Y)\approx\widetilde{K}(SY)\\ \widetilde{K}(\Sigma(X\vee Y))\approx\widetilde{K}(S(X\vee Y)).\end{cases}

Hence, we have:

K~(S(XY))K~(SX)K~(SY).\widetilde{K}(S(X\vee Y))\approx\widetilde{K}(SX)\oplus\widetilde{K}(SY).

This establishes the first isomorphism in this proposition. ∎

Building upon the previous long exact sequence (1) and Proposition 2.13, we can easily obtain the following facts:

Proposition 2.14.

For compact Hausdorff spaces XX and YY, the last map in the long exact sequence for the pair (X×Y,XY)(X\times Y,X\vee Y) is a split surjection with splitting

K~(X)K~(Y)K~(X×Y).\widetilde{K}(X)\oplus\widetilde{K}(Y)\rightarrow\widetilde{K}(X\times Y).
Proof.

By Proposition 2.13 (Example 2.9), we have established the isomorphism:

K~(XY)K~(X)K~(Y).\widetilde{K}(X\vee Y)\approx\widetilde{K}(X)\oplus\widetilde{K}(Y).

Since the last map in the long exact sequence (1) is a surjection, there exists a splitting map from K~(X)K~(Y)\widetilde{K}(X)\oplus\widetilde{K}(Y) to K~(X×Y)\widetilde{K}(X\times Y) given by (a,b)p1(a)+p2(b)(a,b)\mapsto p_{1}^{*}(a)+p_{2}^{*}(b), where p1p_{1} and p2p_{2} are the projections of X×YX\times Y onto XX and YY, respectively. ∎

Proposition 2.15.

For compact Hausdorff spaces XX and YY, the first map in the long exact sequence for the pair (X×Y,XY)(X\times Y,X\vee Y) splits. As a result, we obtain a splitting

K~(X×Y)K~(XY)K~(X)K~(Y).\tilde{K}(X\times Y)\approx\widetilde{K}(X\wedge Y)\oplus\tilde{K}(X)\oplus\tilde{K}(Y).
Proof.

Continuing from the last proof of Proposition 2.14, we know that the first map in the long exact sequence splits via (Sp1)+(Sp2)\left(Sp_{1}\right)^{*}+\left(Sp_{2}\right)^{*}. By Proposition 2.4, this splitting implies that K~(X×Y)\tilde{K}(X\times Y) is isomorphic to the direct sum K~(XY)K~(X)K~(Y)\widetilde{K}(X\wedge Y)\oplus\tilde{K}(X)\oplus\tilde{K}(Y). ∎

Proposition 2.16.

For aK~(X)a\in\widetilde{K}(X) and bK~(Y)b\in\widetilde{K}(Y), the external product aba*b lies in K~(X×Y)\tilde{K}(X\times Y) and induces a unique element in K~(XY)\widetilde{K}(X\wedge Y), defining a reduced external product K~(X)K~(Y)K~(XY)\tilde{K}(X)\otimes\tilde{K}(Y)\rightarrow\widetilde{K}(X\wedge Y).

Proof.

Recall that p1(a)p_{1}^{*}(a) restricts to zero in K(Y)K(Y) and p2(b)p_{2}(b) restricts to zero in K(X)K(X). As a result, p1(a)p2(b)p_{1}^{*}(a)p_{2}^{*}(b) restricts to zero in both K(X)K(X) and K(Y)K(Y), and therefore in K(XY)K(X\vee Y). Consequently, aba*b lies in K~(X×Y)\tilde{K}(X\times Y), and from the short exact sequence given by Proposition 2.15, aba*b pulls back to a unique element of K~(XY)\widetilde{K}(X\wedge Y). This defines the reduced external product. ∎

Now combining these propositions above, we obtain the following commutative diagram:

K(X)K(Y)(K~(X)K~(Y))K~(X)K~(Y)K(X×Y)K~(XY)K~(X)K~(Y)\begin{array}[]{ccc}K(X)\otimes K(Y)&\approx(\widetilde{K}(X)\otimes\widetilde{K}(Y))\oplus\widetilde{K}(X)\oplus\widetilde{K}(Y)\oplus\mathbb{Z}\\ \downarrow&\downarrow\\ K(X\times Y)&\approx\widetilde{K}(X\wedge Y)\oplus\widetilde{K}(X)\oplus\widetilde{K}(Y)\oplus\mathbb{Z}\end{array}

Since the reduced external product is essentially a restriction of the unreduced external product, it also preserves the ring structure and is therefore also a ring homomorphism. With this in mind, let’s use the notation aba*b for both reduced and unreduced external products.

From Proposition 2.16, we defined the reduced external product. Next, we will show that this reduced external product gives rise to the homomorphism β\beta as described in the following theorem.

Lemma 2.17.

The reduced external product gives rise to a homomorphism

β:K~(X)K~(S2X),β(a)=(H1)a,\beta:\tilde{K}(X)\rightarrow\widetilde{K}\left(S^{2}X\right),\quad\beta(a)=(H-1)*a,

where HH is the canonical line bundle over S2=1S^{2}=\mathbb{CP}^{1}.

Proof.

Consider the nn-fold iterated reduced suspension ΣnX\Sigma^{n}X of the space XX. This space is a quotient of the ordinary nn-fold suspension SnXS^{n}X, which is obtained by collapsing an nn-disk in SnXS^{n}X to a point. Consequently, the quotient map SnXSnXS^{n}X\rightarrow S^{n}\wedge X is defined. By Lemma 2.7, we know that this quotient map SnXSnXS^{n}X\rightarrow S^{n}\wedge X induces an isomorphism on K~\tilde{K}. In other words, we have an isomorphism between the reduced K-theories of SnXS^{n}X and SnXS^{n}\wedge X.

Therefore, we can now define the homomorphism β:K~(X)K~(S2X)\beta:\tilde{K}(X)\rightarrow\widetilde{K}\left(S^{2}X\right).777The motivation behind defining β\beta is to study the relationship between the K-theory of a space and its double suspension. Specifically, we define β(a):=(H1)a\beta(a):=(H-1)*a, where HH is the canonical line bundle over S2=1S^{2}=\mathbb{CP}^{1}, and the operation * denotes the reduced external product as defined in Proposition 2.16. ∎

So far, we have analyzed the reduced external product and its relation to the K-theory of these spaces, and shown that it leads to a homomorphism between the reduced K-theories of XX and the double suspension of XX. As we did for deducing the fundamental product theorem, motivated by the previous lemma, it is natural to ask: wouldn’t it be great if the homomorphism β:K~(X)K~(S2X)\beta:\tilde{K}(X)\rightarrow\widetilde{K}\left(S^{2}X\right) was further an isomorphism? If that is true, we will obtain a powerful tool to examine and compute reduced K groups since this isomorphism will exhibit a periodic property. This idea leads us to the discovery of Bott Periodicity Theorem, a central result in K-theory and one of our two main results in this paper:

Theorem 2.18 (Bott Periodicity Theorem for Topological K-theory).

K~(X)βK~(S2X)\tilde{K}(X)\stackrel{{\scriptstyle\beta}}{{\approx}}\tilde{K}\left(S^{2}X\right) by β(a):=(H1)a\beta(a):=(H-1)*a, \forall compact Hausdorff XX. That is, the homomorphism β:K~(X)K~(S2X)\beta:\tilde{K}(X)\rightarrow\tilde{K}\left(S^{2}X\right), defined by β(a)=(H1)a\beta(a)=(H-1)*a as in the Lemma above, is an isomorphism for all compact Hausdorff spaces XX.

Proof.

One way (a sinuous way) to show that β:K~(X)K~(S2X)\beta:\widetilde{K}(X){\rightarrow}\widetilde{K}\left(S^{2}X\right) is an isomorphism is to decompose β\beta into two maps β1β2\beta_{1}\beta_{2} and then show they are both isomorphisms. That is, we write

K~(X)β1K~(S2)K~(X)β2K~(S2X).\widetilde{K}(X)\stackrel{{\scriptstyle\beta_{1}}}{{\rightarrow}}\widetilde{K}\left(S^{2}\right)\otimes\widetilde{K}(X)\stackrel{{\scriptstyle\beta_{2}}}{{\rightarrow}}\widetilde{K}\left(S^{2}X\right).

The first map β1\beta_{1} sends aa to (H1)a(H-1)\otimes a, where HH is the canonical line bundle over S2=1S^{2}=\mathbb{CP}^{1}. This map is an isomorphism because K~(S2)\tilde{K}\left(S^{2}\right) is infinite cyclic generated by H1H-1.

The second map β2\beta_{2} is defined as the reduced external product. Since K~(S2)\widetilde{K}\left(S^{2}\right) is generated by H1H-1 as an abelian group and has the relation (H1)2=0(H-1)^{2}=0, we can rewrite the map β2\beta_{2} as a composition:

β2:K~(X)K~(S2)K(X)[H]/(H1)2K(X×S2)K~(S2X).\beta_{2}:\widetilde{K}(X)\otimes\widetilde{K}\left(S^{2}\right)\rightarrow K(X)\otimes\mathbb{Z}[H]/(H-1)^{2}\rightarrow K\left(X\times S^{2}\right)\rightarrow\widetilde{K}\left(S^{2}X\right).

The first map here is an isomorphism due to Corollary 1.14. The second map here is the homomorphism μ\mu, which is an isomorphism according to Theorem 1.13. The third map here is the natural map induced by the quotient map X×S2S2XX\times S^{2}\rightarrow S^{2}X that collapses an nn-disk in X×S2X\times S^{2} to a point. This map induces an isomorphism on K~\tilde{K}. Since all the maps in the composition are isomorphisms, the composition β2\beta_{2} is also an isomorphism. Thus, β2:K~(S2)K~(X)K~(S2X)\beta_{2}:\widetilde{K}\left(S^{2}\right)\otimes\widetilde{K}(X)\rightarrow\widetilde{K}\left(S^{2}X\right) is an isomorphism, as required.

Therefore, we have concluded that the homomorphism β:K~(X)K~(S2X)\beta:\tilde{K}(X)\rightarrow\tilde{K}\left(S^{2}X\right), defined by β(a)=(H1)a\beta(a)=(H-1)*a, is an isomorphism for all compact Hausdorff spaces XX. ∎

The Bott Periodicity Theorem is so named due to the periodic behavior it reveals in K-theory. Specifically, it shows that there is a repeating pattern in the K-theory groups as we move through suspensions, with a period of two. In other words, as we see above, the K-theory of a space XX and its double suspension S2XS^{2}X are isomorphic. This remarkable periodicity property plays a fundamental role in the study of K-theory and has far-reaching consequences in algebraic topology and other areas of mathematics.

This theorem is named after Raoul Bott (1923 – 2005), an American mathematician who first discovered this periodicity phenomenon in the late 1950s in [2] and [3]. His work was groundbreaking, as it provided a new and powerful tool for studying the homotopy groups of spheres and other topological spaces. Bott’s discovery of this periodicity also led to deeper connections between K-theory and other areas of mathematics, such as representation theory and algebraic geometry. His work laid the foundation for subsequent developments in the field, and the Bott Periodicity Theorem remains a cornerstone of K-theory to this day.

Refer to caption
Figure 1: Raoul Bott999Image source: Wikimedia Commons, licensed under GFDL 1.2.

As a consequence of the Bott Periodicity Theorem, we obtain the following corollary, which gives us a deeper understanding of the periodicity properties of K-theory and highlights the significance of the smash product and reduced external product in connecting the K-theory of different spaces.

Corollary 2.19.

K~(S2n+1)=0\widetilde{K}\left(S^{2n+1}\right)=0 and K~(S2n)\widetilde{K}\left(S^{2n}\right)\approx\mathbb{Z}, generated by the nn-fold reduced external product (H1)(H1)(H-1)*\cdots*(H-1).

Notice by this corollary of the Bott Periodicity Theorem, we have obtained a very generalized proposition: the K-theory of odd-dimensional spheres is trivial (i.e., consisting only of the zero element). As well, K~(S2n)\widetilde{K}\left(S^{2n}\right)\approx\mathbb{Z} demonstrates that the K-theory of the 2n2n-sphere exhibits a periodic structure.

Example 2.20.

Consider the case when n=1n=1. According to the corollary, we have: K~(S2)\widetilde{K}\left(S^{2}\right)\approx\mathbb{Z} (generated by the reduced external product (H1)(H-1)), and K~(S3)=0\widetilde{K}\left(S^{3}\right)=0 (which is trivial since 3 is odd). That is, we have easily computed the reduced K-groups of the 2-sphere, S2S^{2}, and the 3-sphere, S3S^{3} by this consequence of the Bott Periodicity Theorem.

3 Bott Periodicity and Fiber Bundles

In this section, we explore fiber bundles and their associated properties, ultimately discussing the second version Bott Periodicity: Bott Periodicity for stable homotopy groups of classical groups. This was originally discovered to prove periodicity of homotopy groups for the infinite unitary group.

3.1 Fiber Bundles and Basic Properties

At the beginning of the paper, we introduced vector bundles. In fact, a vector bundle is a specific type of a more generalized construction as shown in the following definition:

Definition 3.1 (Fiber Bundle).

A fiber bundle over BB is a space EE with a map π:EB\pi:E\rightarrow B (continuous), satisfying "local triviality": bB\forall b\in B, denoting Eb:=π1(b),E_{b}:=\pi^{-1}(b),\exists open UbU\ni b in BB and a map E|U:=π1(U)tEb\left.E\right|_{U}:=\pi^{-1}(U)\stackrel{{\scriptstyle t}}{{\longrightarrow}}E_{b} such that the map EU(π,t)=φU×EbE_{U}\stackrel{{\scriptstyle(\pi,t)=\varphi}}{{\longrightarrow}}U\times E_{b} is a homeomorphism. Note that any two fibers of a fiber bundle in the same connected component of BB must be homeomorphic.

In other words, a fiber bundle is a short exact sequence101010In this context, by Proposition 2.4, “exact” is used loosely as an intuitive way to capture the idea that locally, EE behaves like the product B×FB\times F of spaces FEpBF\rightarrow E\stackrel{{\scriptstyle p}}{{\longrightarrow}}B, in which all the subspaces p1(b)Ep^{-1}(b)\subset E, called fibers, are homeomorphic. The fiber bundle structure is characterized by the projection map p:EBp:E\rightarrow B. To denote the specific fiber, we also represent a fiber bundle using the notation FEBF\rightarrow E\rightarrow B, which forms a short exact sequence of spaces. The last space BB is called the base space of the bundle, while the middle space EE is called the total space.

Comparing their definitions, we can see a vector bundle is a fiber bundle where the fibers are vector spaces, and the projection map is compatible with the vector space structure. Fiber bundles can also be thought of as twisted products. Familiar examples of fiber bundles are the Möbius band (which is a twisted annulus with line segments as fibers) and the Klein bottle (which is a twisted torus with circles as fibers). Fiber bundles provide a rich structure to study topological spaces and their relationships.

Definition 3.2 (Homotopy Lifting Property).

A map p:EBp:E\rightarrow B is said to have the homotopy lifting property with respect to a space XX if, given a homotopy gt:XBg_{t}:X\rightarrow B and a map g~0:XE\tilde{g}_{0}:X\rightarrow E lifting g0g_{0}, so that pg~0=g0p\tilde{g}_{0}=g_{0}, there exists a homotopy g~t:XE\tilde{g}_{t}:X\rightarrow E lifting gtg_{t}.

The homotopy lifting property plays a critical role in studying the relationship between spaces involved in a fiber bundle, as it enables the construction of homotopies between them:

Theorem 3.3 (Homotopy Lifting Property Isomorphism Theorem).

Let p:EBp:E\rightarrow B be a map satisfying the homotopy lifting property with respect to disks DkD^{k} for all k0k\geq 0. Choose basepoints b0Bb_{0}\in B and x0F=p1(b0)x_{0}\in F=p^{-1}(b_{0}). Then, we have the isomorphism πn(E,F,x0)pπn(B,b0)\pi_{n}(E,F,x_{0})\stackrel{{\scriptstyle p_{*}}}{{\approx}}\pi_{n}(B,b_{0}) by the induced map p:πn(E,F,x0)πn(B,b0)p_{*}:\pi_{n}(E,F,x_{0})\rightarrow\pi_{n}(B,b_{0}) for all n1n\geq 1. Consequently, if BB is path-connected, there exists a long exact sequence of homotopy groups:

πn(F,x0)πn(E,x0)pπn(B,b0)πn1(F,x0)π0(E,x0)0\cdots\rightarrow\pi_{n}\left(F,x_{0}\right)\rightarrow\pi_{n}\left(E,x_{0}\right)\stackrel{{\scriptstyle p_{*}}}{{\rightarrow}}\pi_{n}\left(B,b_{0}\right)\rightarrow\pi_{n-1}\left(F,x_{0}\right)\rightarrow\cdots\rightarrow\pi_{0}\left(E,x_{0}\right)\rightarrow 0

This theorem establishes the connection between the homotopy groups of the spaces involved in a fiber bundle when the homotopy lifting property is satisfied. The proof of the theorem can be found in Chapter 4 of [8].

With the definition of the homotopy lifting property, we define the term fibration to be a map p:EBp:E\rightarrow B that satisfies this homotopy lifting property with respect to all spaces XX. In particular, if the map pp satisfies the homotopy lifting property for disks, then it’s usually called a Serre fibration.

3.2 Stiefel Manifolds in Real, Complex, and Quaternionic Cases

Definition 3.4 (Stiefel Manifold).

A Stiefel manifold, denoted by Vn(k)V_{n}(\mathbb{R}^{k}), or Vn,kV_{n,k}, is a space consisting of all orthonormal kk-frames in n\mathbb{R}^{n}; in other words, it is the set of all orthonormal kk-tuples of vectors in n\mathbb{R}^{n}. It is given the subspace topology as a subset of the product of kk copies of the (n1)(n-1)-sphere, Sn1S^{n-1}.

By this definition, Vn(k)V_{n}(\mathbb{R}^{k}) is a space of orthonormal kk-tuples of vectors, which can be thought of as a subspace of the product of kk copies of the (n1)(n-1)-sphere. Oftentimes, the Stiefel manifold is used to study various properties of orthonormal frames, and its homotopy and cohomology groups are of particular interest in algebraic topology. In addition, we have the following property (as discussed in Ganatra’s [6]):

Proposition 3.5 (Compactness of Vn(k)V_{n}(\mathbb{R}^{k})).

The Stiefel manifold Vn(k)V_{n}(\mathbb{R}^{k}) is a compact manifold.

Proof.

To start, observe O(n)O(n) acts on Vk(n)V_{k}\left(\mathbb{R}^{n}\right) by composition: transitive action. Also, the isotropy group of basepoints {e1,,ek}\left\{e_{1},...,e_{k}\right\} is Ik×O(nk)I_{k}\times O(n-k). Using this, we can easily show Vk(n)O(n)/(Ik×O(nk))V_{k}\left(\mathbb{R}^{n}\right)\approx O(n)/(I_{k}\times O{(n-k)}). ∎

As an example, let us consider the general case of fiber bundles for m<nkm<n\leq k:

Vnm(km)Vn(k)pVm(k)V_{n-m}\left(\mathbb{R}^{k-m}\right)\rightarrow V_{n}\left(\mathbb{R}^{k}\right)\stackrel{{\scriptstyle p}}{{\longrightarrow}}V_{m}\left(\mathbb{R}^{k}\right) (2)

Here, we have three Stiefel manifolds involved in the fiber bundle. The base space is Vm(k)V_{m}(\mathbb{R}^{k}), which consists of all orthonormal mm-frames in k\mathbb{R}^{k}. The total space is Vn(k)V_{n}(\mathbb{R}^{k}), which consists of all orthonormal nn-frames in k\mathbb{R}^{k}. The fiber over a point in the base space is Vnm(km)V_{n-m}(\mathbb{R}^{k-m}), which consists of all orthonormal (nm)(n-m)-frames in km\mathbb{R}^{k-m}. In this fiber bundle, the projection map p:Vn(k)Vm(k)p:V_{n}(\mathbb{R}^{k})\rightarrow V_{m}(\mathbb{R}^{k}) takes an orthonormal nn-frame in k\mathbb{R}^{k} and maps it onto an orthonormal mm-frame in k\mathbb{R}^{k}.

First, let’s restrict (2) to the case m=1m=1, so we have bundles

Vn1(k1)Vn(k)pV1(k).V_{n-1}\left(\mathbb{R}^{k-1}\right)\rightarrow V_{n}\left(\mathbb{R}^{k}\right)\stackrel{{\scriptstyle p}}{{\longrightarrow}}V_{1}\left(\mathbb{R}^{k}\right).

Notice that V1(k)V_{1}\left(\mathbb{R}^{k}\right) can be identified with the (k1)(k-1)-dimensional sphere Sk1S^{k-1}. This is because the 1-frames in k\mathbb{R}^{k} are essentially the unit vectors in k\mathbb{R}^{k}, and the unit vectors form the (k1)(k-1)-dimensional sphere Sk1S^{k-1}. Thus, we can rewrite the fiber bundle as

Vn1(k1)Vn(k)pSk1.V_{n-1}\left(\mathbb{R}^{k-1}\right)\rightarrow V_{n}\left(\mathbb{R}^{k}\right)\stackrel{{\scriptstyle p}}{{\longrightarrow}}S^{k-1}.

That allows us to deduce the connectivity of Vn(k)V_{n}\left(\mathbb{R}^{k}\right) by induction on nn.

Next, let’s (only) restrict (2) to the case k=nk=n, so we obtain fiber bundles

Vnm(nm)Vn(n)pVm(n).V_{n-m}\left(\mathbb{R}^{n-m}\right)\rightarrow V_{n}\left(\mathbb{R}^{n}\right)\stackrel{{\scriptstyle p}}{{\longrightarrow}}V_{m}\left(\mathbb{R}^{n}\right).

Recall that Vn(n)V_{n}\left(\mathbb{R}^{n}\right) represents the set of all orthonormal nn-frames in n\mathbb{R}^{n}, which can be identified with the orthogonal group O(n)O(n). Likewise, the fiber Vnm(nm)V_{n-m}\left(\mathbb{R}^{n-m}\right) can be identified with the orthogonal group O(nm)O(n-m) because it represents the set of orthonormal (nm)(n-m)-frames in nm\mathbb{R}^{n-m}. The projection map pp sends an nn-frame in O(n)O(n) to the mm-frame formed by its first mm vectors, and the fibers consist of the cosets αO(nm)\alpha O(n-m) for αO(n)\alpha\in O(n). Therefore, we can rewrite the fiber bundle as

O(nm)O(n)pVm(n).O(n-m)\rightarrow O(n)\stackrel{{\scriptstyle p}}{{\longrightarrow}}V_{m}\left(\mathbb{R}^{n}\right).

Now, let’s combine the preceding two cases together: taking m=1m=1 and k=nk=n to obtain bundles

O(n1)O(n)pSn1.O(n-1)\rightarrow O(n)\stackrel{{\scriptstyle p}}{{\longrightarrow}}S^{n-1}.

Before we further explore these particular bundles, let’s analogously define fiber bundles for the complex and quaternionic cases:

Definition 3.6 (Unitary and Symplectic Groups).

The unitary group U(n)U(n) is the group of n×nn\times n unitary matrices, i.e., complex matrices UU such that UU=IU^{*}U=I, where UU^{*} is the conjugate transpose of UU. The symplectic group Sp(n)Sp(n) consists of 2n×2n2n\times 2n matrices SS that preserve a skew-symmetric bilinear form, i.e., STJS=JS^{T}JS=J, where JJ is a skew-symmetric matrix.

To understand their fiber bundle structure, we first need to connect the Stiefel manifolds to the coset spaces of these groups. The complex and quaternionic versions of Stiefel manifolds can be identified with the coset spaces of the unitary and symplectic groups, respectively. For example, Vn(k)V_{n}(\mathbb{C}^{k}) is identifiable with the coset space U(k)/U(kn)U(k)/U(k-n), and Vn(k)V_{n}(\mathbb{H}^{k}) is identifiable with the coset space Sp(k)/Sp(kn)Sp(k)/Sp(k-n).

Now let’s consider their corresponding Stiefel manifolds in the complex and quaternionic cases. Employing the same reasoning as above, by restricting to the case m=1m=1 and k=nk=n, we obtain fiber bundles in the complex and quaternionic cases as follows:

U(n1)\displaystyle U(n-1) U(n)pS2n1\displaystyle\rightarrow U(n)\stackrel{{\scriptstyle p}}{{\longrightarrow}}S^{2n-1}
Sp(n1)\displaystyle Sp(n-1) Sp(n)pS4n1\displaystyle\rightarrow Sp(n)\stackrel{{\scriptstyle p}}{{\longrightarrow}}S^{4n-1}

These bundles are analogous to the real case, with the projection map pp in each case sending an element of U(n)U(n) or Sp(n)Sp(n) to the corresponding coset in the Stiefel manifold, which is homeomorphic to a sphere of dimension 2n12n-1 or 4n14n-1, respectively. Unfortunately, these bundles also show that computing homotopy groups of O(n),U(n)O(n),U(n), and Sp(n)Sp(n) should be at least as difficult as computing homotopy groups of spheres.

3.3 Deducing Periodicity from Fiber Bundles

In the preceding subsection, we have examined the three bundles:

O(n1)\displaystyle O(n-1) O(n)pSn1\displaystyle\rightarrow O(n)\stackrel{{\scriptstyle p}}{{\longrightarrow}}S^{n-1}
U(n1)\displaystyle U(n-1) U(n)pS2n1\displaystyle\rightarrow U(n)\stackrel{{\scriptstyle p}}{{\longrightarrow}}S^{2n-1}
Sp(n1)\displaystyle Sp(n-1) Sp(n)pS4n1\displaystyle\rightarrow Sp(n)\stackrel{{\scriptstyle p}}{{\longrightarrow}}S^{4n-1}

They actually lead to an interesting stability property. In the real case, the inclusion O(n1)O(n)O(n-1)\hookrightarrow O(n) induces an isomorphism on πi\pi_{i} for i<n2i<n-2 from the long exact sequence of the first bundle. As a consequence, the groups πiO(n)\pi_{i}O(n) are independent of nn if nn is sufficiently large. Similarly, the same is true for the groups πiU(n)\pi_{i}U(n) and πiSp(n)\pi_{i}Sp(n) through the other two bundles.

With these observations in mind, it is time to conclude one of the most remarkable results in algebraic topology, the Bott Periodicity Theorem for Stable Homotopy Groups of Classical Groups, which states that these stable groups repeat periodically, with a period of eight for OO and SpSp, and a period of two for UU. The theorem can be formally stated as follows:

Theorem 3.7 (Bott Periodicity for Stable Homotopy Groups of Classical Groups).

The stable homotopy groups of O(n)O(n), U(n)U(n), and Sp(n)Sp(n) exhibit periodic behavior with periods of eight for OO and SpSp, and a period of two for UU. The periodic values are given in the following table:

imod8i\bmod 8 0 11 22 33 44 55 66 77
πiO(n)\pi_{i}O(n) 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} 0 \mathbb{Z} 0 0 0 \mathbb{Z}
πiU(n)\pi_{i}U(n) 0 \mathbb{Z} 0 \mathbb{Z} 0 \mathbb{Z} 0 \mathbb{Z}
πiSp(n)\pi_{i}Sp(n) 0 0 0 \mathbb{Z} 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} 0 \mathbb{Z}

This theorem not only highlights a deep connection between the homotopy groups of the orthogonal, unitary, and symplectic groups, but, most importantly, reveals an unexpected periodic structure in these groups. As the Bott Periodicity Theorem for Topological K-theory, this version of Bott Periodicity Theorem also has far-reaching applications in algebraic topology, and it enables us to easily classify homotopy groups of O(n)O(n), U(n)U(n), and Sp(n)Sp(n).

Example 3.8.

Consider the stable homotopy groups of the unitary group U(n)U(n). According to the Bott Periodicity Theorem above, these groups have a period of two. This means that the homotopy groups πiU(n)\pi_{i}U(n) and πi+2U(n)\pi_{i+2}U(n) are isomorphic for all ii. Namely, πiU(n)πi+2U(n)\pi_{i}U(n)\approx\pi_{i+2}U(n) for all ii. This shows a periodic behavior for the homotopy groups πiU(n)\pi_{i}U(n) i\forall i.

4 Conclusion

In the previous two sections, we have delved into two vibrant versions of Bott Periodicity that exhibit periodic behavior: the Bott Periodicity Theorem for topological K-theory, and the Bott Periodicity Theorem for stable homotopy groups of classical groups (orthogonal, unitary, and symplectic groups). The first version establishes an isomorphism between the reduced K-theories of a compact Hausdorff space and its double suspension, emphasizing the algebraic and geometric properties of vector bundles over topological spaces; whereas the second version reveals periodic patterns in stable homotopy groups, providing valuable insights into the global structure of these groups.

Although these two versions of Bott Periodicity focus on different aspects of topology, they are deeply interconnected, and both demonstrate the fascinating periodic structure in topology. They showcase the rich interplay between algebra, geometry, and topology. The beauty and elegance of these periodic phenomena lie not only in their symmetric simplicity but also in the insights they provide into the boundless creativity that inspires the evolution of topology and mathematics in general, reaching far and wide across dimensions yet untold.

References