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

El Teorema de Gelfand Naimark desde una perspectiva Categórica
The Gelfand–Naimark Theorem from a Categorical Perspective

Sebastian Alvarez Avendaño AI/ML Associate, JPMorgan Chase & Co [email protected] 0000-0002-8022-4965 Breitner Ocampo Universidad de Antioquia
Medellin breitner.ocampo@@udea.edu.co
0000-0003-3507-4598
Pedro Rizzo Universidad de Antioquia
Medellin [email protected]
0000-0001-8793-0822
Abstract

Este artículo presenta como resultado principal la equivalencia entre, las categorías de espacios topológicos Hausdorff-Compactos y la categoría de las CC^{*}-álgebras conmutativas con unidad, producto de la “traducción” en este lenguaje del teorema de Gelfand–Naimark presentado en 1943 ([13]). Haremos un recorrido sobre las principales ideas del análisis y el álgebra, conjugadas con éxito, en el estudio de la teoría de Álgebras de Banach. Así mismo estableceremos, a forma de conclusión, diversas aplicaciones que resultan naturalmente posibles a la luz de la “analogía y generalización” que nos permiten la teoría de categorías.

Palabras claves: CC^{*}-algebras, Categorías, Espacios Topológicos, Teorema de Gelfand-Naimark, Teoría de Representaciones.

Abstract

The goal of this paper is to prove the categorical equivalence between the category of Hausdorff-Compact topological spaces and the category of Unital Commutative CC^{*}-algebras. This equivalence can be interpreted as a way of rewriting the well known Gelfand-Naimark Theorem [13] in a categorical language. We will present the basic concepts in the theory of Banach Algebras as a successful link between Analysis and Algebra. Likewise, we will show some applications due to this new perspective, highlighting the categorical connection through proofs of typical problems that don’t have an easy solution in CC^{*}-algebra.

Keywords: Category Theory, CC^{*}-algebras, Gelfand-Naimark Theorem, Topological Spaces, Representation Theory.

1 Introducción

In 1943, Israel Gelfand and Mark Naimark introduced the theory of CC^{*}-algebras in response to a problem proposed by von Neumann in the emerging field of quantum mechanics [13]. Their seminal article not only presented what is now known as the Gelfand-Naimark Theorem but also laid the foundation for the rapid development of Banach algebra theory and its applications in both mathematics and physics. For a comprehensive exploration of the far-reaching impact of [13] across various fields—including quantum field theory, statistical mechanics, knot theory, dynamical systems, group representation theory, and KK-theory—we recommend [9].

The Gelfand-Naimark Theorem can be broadly interpreted as the study of Hausdorff-compact topological spaces through the rich algebraic structure of their continuous function spaces. Conversely, it asserts that commutative CC^{\ast}-algebras with unity are determined by specific Hausdorff-compact topological spaces. Such correspondences between mathematical structures, often viewed as instances of duality, arise naturally within the framework of category theory. More precisely, if we denote by 𝒞\mathcal{HC} the category of Hausdorff-compact topological spaces and by 𝒞𝒜𝒰\mathcal{CAU} the category of commutative CC^{\ast}-algebras with unity, the categorical interpretation of the Gelfand-Naimark Theorem establishes an equivalence between these two categories.

One of the main objectives of this article is to define the context and demonstrate the equivalence mentioned above. To achieve this, we present all the necessary introductory content on commutative CC^{\ast}-algebras with unity in an accessible and engaging manner. Additionally, we introduce the fundamental concepts of basic category theory and provide concrete applications that illustrate the implications of these general correspondences.

A secondary objective of this article is to highlight the importance, simplicity, and relevance of results like that of Gelfand and Naimark by structuring the content in a way that allows readers to draw connections through analogies and generalizations provided by category theory. In line with this goal, we offer insights to help readers understand the fundamentals of more advanced topics, such as non-commutative geometry, which historically emerged from the correspondence between commutative CC^{\ast}-algebras with unity and Hausdorff-compact topological spaces.

To facilitate the reading of this article, we recommend having some familiarity with basic concepts from commutative algebra, functional analysis, topology, and complex analysis. Initially, we present the fundamental definitions to familiarize the reader with the terminology used throughout the text. In Sections 3 and 4, we explore the theory of CC^{*}-algebras, generalizing the concept of the spectrum and introducing the Gelfand Transform, which serves as the main tool of this work. In Section 5, we state the Gelfand-Naimark Theorem using the language of Category Theory, following a brief overview of its definitions and key concepts. Finally, in Section 6, we demonstrate applications of the main result, drawing inspiration from another duality theorem, the Hilbert Nullstellensatz. Additionally, we provide key characterizations of elements in a commutative CC^{*}-algebra with unity.

2 CC^{*}-algebras

In this section, we present the basic definitions necessary for the development of this work, including the formal definition of a CC^{*}-algebra and the notion of the spectrum of an element. We also provide some classical examples that will be useful in later sections. It is worth noting that many of the preliminary results apply to both Banach algebras and CC^{*}-algebras. However, since our focus is on describing CC^{*}-algebras, we will limit our discussion to this context. Throughout this article, all algebraic structures and analytic concepts will be considered over the field \mathbb{C} of complex numbers.

Recall that an algebra 𝒜\mathcal{A} is a \mathbb{C}-vector space equipped with a product “\cdot” that is compatible with the vector space structure. If this product satisfies ab=baa\cdot b=b\cdot a, the algebra is said to be commutative. If the algebra 𝒜\mathcal{A} contains an element ee such that ea=ae=ae\cdot a=a\cdot e=a for all a𝒜a\in\mathcal{A}, we say that the algebra has a unit. For algebras without a unit, it is possible to construct a unitized algebra 𝒜~\widetilde{\mathcal{A}} in which a “copy” of 𝒜\mathcal{A} exists. This construction is known as the unitalization process, which we will not discuss here, but we recommend [4] and [10] for further details.

We say that the algebra 𝒜\mathcal{A} is normed if there exists a norm :𝒜+{0}\|\cdot\|:\mathcal{A}\to\mathbb{R}^{+}\cup\{0\} that additionally satisfies abab\|a\cdot b\|\leq\|a\|\,\|b\| for all a,b𝒜a,b\in\mathcal{A}. As a normed vector space, 𝒜\mathcal{A} becomes a Hausdorff topological space, allowing us to discuss open sets, closed sets, sequences, convergence, and the uniqueness of limits. If the topological space (𝒜,)(\mathcal{A},\|\cdot\|) is complete (meaning that every Cauchy sequence converges), we say that 𝒜\mathcal{A} is a Banach algebra.

The simplest example of a set that satisfies these conditions is the set of complex numbers. The object of our interest, CC^{*}-algebras, can be understood as a natural generalization of the complex numbers. To clarify this idea, we will now define algebras with involution.

Definition 2.1 (\ast-algebra).

Let 𝒜\mathcal{A} be an algebra and :𝒜𝒜*:\mathcal{A}\rightarrow\mathcal{A} be a function. If for every a,b𝒜a,b\in\mathcal{A} and α,β\alpha,\beta\in\mathbb{C}, \ast satisfies:

  1. 1.

    (a)=a(a^{*})^{*}=a,

  2. 2.

    (αa+βb)=α¯a+β¯b(\alpha a+\beta b)^{*}=\overline{\alpha}a^{*}+\overline{\beta}b^{*},

  3. 3.

    (ab)=ba(ab)^{*}=b^{*}a^{*},

then * is called an involution, and the pair (𝒜,)(\mathcal{A},*) is called a \ast-algebra or algebra with involution.

Next, we will provide the definition of the object of study of this article, the CC^{*}-algebras. These were systematically introduced in 1943 in the article “On the embedding of normed rings into the ring of operators in Hilbert space” by Israel Gelfand and Mark Naimark [13], and they continue to be relevant today due to their connections with Harmonic Analysis, Operator Theory, Algebraic Topology, Quantum Physics, among others. To delve deeper into these relationships, we recommend, for example, [11], [14], [16], [20], [21], and the references cited therein.

Definition 2.2 (C*-algebra).

A normed algebra (𝒜,)(\mathcal{A},\|\cdot\|) is called a CC^{*}-algebra if:

  1. 1.

    (𝒜,)(\mathcal{A},\|\cdot\|) is a Banach algebra.

  2. 2.

    𝒜\mathcal{A} is an algebra with involution *.

  3. 3.

    (CC^{\ast}-Condition) For every a𝒜a\in\mathcal{A}, we have aa=a2\|a^{*}a\|=\|a\|^{2}.

In every CC^{*}-algebra, the involution is a continuous isometry. Indeed, since a2=aaaa\|a^{*}\|^{2}=\|aa^{*}\|\leq\|a\|\|a^{*}\|, it follows that aa\|a^{*}\|\leq\|a\|. Similarly, we have aa\|a\|\leq\|a^{*}\|, thus showing that a=a\|a\|=\|a^{*}\|.

From this point on, unless otherwise stated explicitly, the letter 𝒜\mathcal{A} will always denote a commutative CC^{*}-algebra with unity. Note that e=ee^{*}=e and e=1\|e\|=1, since the identity element is unique and e2=ee=e\|e\|^{2}=\|e^{*}e\|=\|e\|.

Example 2.3.

Complex numbers form a CC^{*}-algebra with the involution given by conjugation. Indeed, we can observe that in this case, for λ=a+bi\lambda=a+bi\in\mathbb{C}, we have λλ¯=(a+bi)(abi)=a2+b2=|λ|2\lambda\bar{\lambda}=(a+bi)(a-bi)=a^{2}+b^{2}=|\lambda|^{2}. Therefore, complex numbers satisfy the CC^{\ast}-condition of Definition 2.2.

Example 2.4.

Let XX be a Hausdorff-Compact space. The \mathbb{C}-vector space

C(X)={f:X|fis continuous}C(X)=\left\{f:X\rightarrow\mathbb{C}|\,f\,\text{is continuous}\right\}

equipped with the norm f=supxX|f(x)|\displaystyle\|f\|_{\infty}=\sup_{x\in X}|f(x)| and the involution f(x)=f(x)¯f^{*}(x)=\overline{f(x)}, is a non-trivial example of a commutative CC^{*}-algebra with unity, where the unity element is the constant function f(x)1f(x)\equiv 1.

The following two examples of CC^{\ast}-algebras are presented for the sake of completeness and due to their importance in this field. However, we emphasize that they will not be of interest to us, as the first example generally lacks a unit, and the second is not commutative.

Example 2.5.

Let XX be a Hausdorff locally compact space. The set C0(X)C_{0}(X) is defined as the set of functions f:Xf:X\to\mathbb{C} such that:

  • ff is a continuous function,

  • For every ϵ>0\epsilon>0, there exists a compact set KXK\subset X such that |f(x)|<ϵ|f(x)|<\epsilon for all xXKx\in X\setminus K.

With the norm f=supxX|f(x)|\displaystyle\|f\|_{\infty}=\sup_{x\in X}|f(x)| and involution f(x)=f(x)¯f^{*}(x)=\overline{f(x)}, the normed space C0(X)C_{0}(X) is a CC^{\ast}-algebra.

Example 2.6.

Let \mathcal{H} be a Hilbert space, with ,\langle\cdot,\cdot\rangle as its inner product. Consider 𝔅()\mathfrak{B}(\mathcal{H}) as the \mathbb{C}-vector space of continuous linear operators T:T:\mathcal{H}\to\mathcal{H}. The product of two elements T,S𝔅()T,S\in\mathfrak{B}(\mathcal{H}) is the composition TST\circ S. The norm of an operator TT is defined by

T𝔅()=sup{T(ζ):ζ,ζ1}.\|T\|_{\mathfrak{B}(\mathcal{H})}=\sup\left\{\|T(\zeta)\|_{\mathcal{H}}:\zeta\in\mathcal{H},\|\zeta\|_{\mathcal{H}}\leq 1\right\}.

The involution for the element TT in the set 𝔅()\mathfrak{B}(\mathcal{H}) is the unique operator T:T^{:}\mathcal{H}\to\mathcal{H} such that

T(ζ),η=ζ,T(η).\langle T(\zeta),\eta\rangle=\langle\zeta,T^{*}(\eta)\rangle.

It is well known that the space 𝔅()\mathfrak{B}(\mathcal{H}) is complete ([15, Theorem 2.10-2]). To conclude that 𝔅()\mathfrak{B}(\mathcal{H}) is a CC^{\ast}-algebra, we only need to prove the CC^{\ast}-condition from Definition 2.2. For this, we will calculate the norm of TT(ζ)TT^{*}(\zeta) as follows:

TT(ζ)2\displaystyle\|T^{*}T(\zeta)\|^{2} =\displaystyle= TT(ζ),TT(ζ)\displaystyle\langle T^{*}T(\zeta),T^{*}T(\zeta)\rangle
=\displaystyle= T(ζ),T(TT(ζ))\displaystyle\langle T(\zeta),T(T^{*}T(\zeta))\rangle
\displaystyle\leq T(ζ)T(TT(ζ))\displaystyle\|T(\zeta)\|\|T(T^{*}T(\zeta))\|
\displaystyle\leq TζTTT(ζ)\displaystyle\|T\|\|\zeta\|\|T\|\|T^{*}T(\zeta)\|

where

TT(ζ)ζT2\frac{\|T^{*}T(\zeta)\|}{\|\zeta\|}\leq\|T\|^{2}

and thus, we conclude that

TT=supζTT(ζ)ζT2.\|T^{*}T\|=\sup_{\zeta\in\mathcal{H}}\frac{\|T^{*}T(\zeta)\|}{\|\zeta\|}\leq\|T\|^{2}.

The other inequality is proved in a similar way. In particular, Mn()M_{n}(\mathbb{C}) is a CC^{\ast}-algebra (non-commutative if n2n\geq 2) if we consider the Euclidean norm 2\|\cdot\|_{2} on n\mathbb{C}^{n} and each matrix as an operator on 𝔅(n)\mathfrak{B}(\mathbb{C}^{n}). In this case, the involution is given by (ai,j)=(aj,i¯)(a_{i,j})^{*}=(\overline{a_{j,i}}) and the norm is defined by A=supx=1Ax2.\|A\|=\sup\limits_{\|x\|=1}\|Ax\|_{2}.

An element a𝒜a\in\mathcal{A}, non-zero, is called invertible if there exists b𝒜b\in\mathcal{A} such that ab=ba=eab=ba=e. Invertible elements play an important role in the development of the theory of CC^{*}-algebras, as will be shown with the sequence of results below.

Proposition 2.7.

(Geometric Series) Let a𝒜a\in\mathcal{A} such that a<1\|a\|<1. Then eae-a is invertible, and the series n=0an\sum_{n=0}^{\infty}a^{n}, where a0=ea^{0}=e, converges absolutely to (ea)1(e-a)^{-1}.

Proof.

For this proof, we will use a well-known fact in Functional Analysis ([15, §2.3\S 2.3 Problems, 7-9, p. 71]): “A normed vector space is a Banach space if and only if every absolutely convergent series is convergent.”

Now, since a<1\|a\|<1, then limNaNlimNaN=0\lim\limits_{N\to\infty}\|a^{N}\|\leq\lim\limits_{N\to\infty}\|a\|^{N}=0. Thus, we have, first, that limNaN=0\lim\limits_{N\to\infty}a^{N}=0, and second, that the numerical series n=0an\sum_{n=0}^{\infty}\|a\|^{n} converges to (1a)1(1-\|a\|)^{-1}. By applying the aforementioned result from Functional Analysis to the CC^{*}-algebra 𝒜\mathcal{A}, we conclude that the series b=n=0anb=\sum_{n=0}^{\infty}a^{n} converges absolutely. From

(ea)b=(ea)(limNn=0Nan)=limNeaN+1=e,(e-a)b=(e-a)\left(\lim_{N\to\infty}\sum_{n=0}^{N}a^{n}\right)=\lim_{N\to\infty}e-a^{N+1}=e,

and similarly proving the equality b(ea)=eb(e-a)=e, we conclude that eae-a is invertible. By the uniqueness of inverses, we deduce (ea)1=n=0an(e-a)^{-1}=\sum_{n=0}^{\infty}a^{n}, which is what we wanted to prove. ∎

We denote the set of invertible elements in 𝒜\mathcal{A} by G(𝒜)G(\mathcal{A}). An important consequence of the following result is that G(𝒜)G(\mathcal{A}) is a (multiplicative subgroup and) open subset of the CC^{*}-algebra 𝒜\mathcal{A}.

Corollary 2.8.

Let a𝒜a\in\mathcal{A} be an invertible element, b𝒜b\in\mathcal{A}, and λ\lambda\in\mathbb{C}.

  1. 1.

    If ab<1a1\|a-b\|<\frac{1}{\|a^{-1}\|}, then bb is invertible and the series

    n=0(a1(ab))na1\sum_{n=0}^{\infty}(a^{-1}(a-b))^{n}a^{-1} (1)

    converges absolutely to b1b^{-1}. In particular, G(𝒜)G(\mathcal{A}) is open.

  2. 2.

    If a<|λ|\|a\|<|\lambda|, then (λea)(\lambda e-a) is invertible and its inverse is given by

    1λn=0anλn\dfrac{1}{\lambda}\sum_{n=0}^{\infty}\dfrac{a^{n}}{\lambda^{n}} (2)
Proof.

For the first statement, note that the hypothesis implies ea1b<1\|e-a^{-1}b\|<1, and by Proposition 2.7, it follows that c=a1b=(e(ea1b))c=a^{-1}b=(e-(e-a^{-1}b)) is invertible. Thus, b=acb=ac is the product of two invertible elements, and therefore, b1b^{-1} exists. We have b1a=(a1b)1=n=0(ea1b)n=n=0(a1(ab))nb^{-1}a=\left(a^{-1}b\right)^{-1}=\sum_{n=0}^{\infty}\left(e-a^{-1}b\right)^{n}=\sum_{n=0}^{\infty}\left(a^{-1}(a-b)\right)^{n}. Hence, b1=n=0(a1(ab))na1b^{-1}=\sum_{n=0}^{\infty}(a^{-1}(a-b))^{n}a^{-1}. Therefore, for every aG(𝒜)a\in G(\mathcal{A}), we have B(a;1a1)G(𝒜)B(a;\frac{1}{\|a^{-1}\|})\subset G(\mathcal{A}), which proves that G(𝒜)G(\mathcal{A}) is an open set.

For the second statement, observe that the hypothesis is equivalent to aλ<1\|\frac{a}{\lambda}\|<1, and by a similar reasoning as before, applying Proposition 2.7, the conclusion follows.

Another important consequence of the invertibility of (ea)(e-a) is the following:

Corollary 2.9.

The group homomorphism

inv:G(𝒜)\displaystyle\mbox{inv}:G(\mathcal{A}) \displaystyle\to G(𝒜)\displaystyle G(\mathcal{A})
a\displaystyle a \displaystyle\mapsto inv(a)=a1\displaystyle\mbox{inv}(a)=a^{-1}

is a continuous function.

Proof.

Let ϵ>0\epsilon>0 be given. We define ϵ1:=ϵa1\epsilon_{1}:=\epsilon\|a^{-1}\| as an auxiliary quantity, and we choose δ=ϵ1(1+ϵ1)|a1|\delta=\dfrac{\epsilon_{1}}{(1+\epsilon_{1})|a^{-1}|}. Since ϵ11+ϵ1<1\dfrac{\epsilon_{1}}{1+\epsilon_{1}}<1, it follows that δ<1|a1|\delta<\dfrac{1}{|a^{-1}|}. Now, if bb satisfies |ab|<δ|a-b|<\delta, then |a1(ba)|<1|a^{-1}(b-a)|<1. We can calculate the difference |b1a1||b^{-1}-a^{-1}| using the first statement of Corollary 2.8, which gives us:

b1a1\displaystyle\|b^{-1}-a^{-1}\| =\displaystyle= n=0(a1(ab))na1a1\displaystyle\|\sum_{n=0}^{\infty}(a^{-1}(a-b))^{n}a^{-1}-a^{-1}\|
=\displaystyle= n=1(a1(ab))na1\displaystyle\|\sum_{n=1}^{\infty}(a^{-1}(a-b))^{n}a^{-1}\|
\displaystyle\leq a1n=1(a1(ab))n\displaystyle\|a^{-1}\|\sum_{n=1}^{\infty}\|(a^{-1}(a-b))^{n}\|
=\displaystyle= a1(a1(ab))1(a1(ab))\displaystyle\|a^{-1}\|\dfrac{\|(a^{-1}(a-b))\|}{1-\|(a^{-1}(a-b))\|}
\displaystyle\leq a1ϵ1=ϵ.\displaystyle\|a^{-1}\|\epsilon_{1}=\epsilon.

Thus, the inversion homomorphism is a continuous function. ∎

Definition 2.10.

For each element a𝒜a\in\mathcal{A}, we define the Resolvent of aa as the set

ρ(a)={λ:λea is invertible in 𝒜}.\rho(a)=\{\lambda\in\mathbb{C}:\lambda e-a\text{ is invertible in $\mathcal{A}$}\}. (3)

We define the Spectrum of aa as the complement in \mathbb{C} of the set ρ(a)\rho(a):

σ(a)=ρ(a).\sigma(a)=\mathbb{C}\setminus\rho(a). (4)
Remark 1.

The set of eigenvalues of a matrix MMn()M\in M_{n}(\mathbb{C}) is defined as

{λ:(MλIn)v=0 for some nonzero vector vn}.\{\lambda\in\mathbb{C}:(M-\lambda I_{n})v=0\text{ for some nonzero vector }v\in\mathbb{C}^{n}\}.

Thus, the spectrum of a matrix MM, as defined in Definition 2.10, coincides with the set of its eigenvalues. This is why the definition of spectrum can be seen as a natural generalization of the concept of eigenvalues for matrices.

Example 2.11.

Let XX be a Hausdorff-compact space, and let C(X)C(X) be the set of continuous functions with values in the complex numbers. The resolvent set of fC(X)f\in C(X) can be obtained as follows: Given fC(X)f\in C(X) and λ\lambda\in\mathbb{C}, it is easy to see that fλf-\lambda is invertible if and only if f(x)λ0f(x)-\lambda\neq 0 for every xXx\in X. This is equivalent to f(x)λf(x)\neq\lambda for each xXx\in X, or equivalently, λIm(f)\lambda\notin\text{Im}(f). Since the spectrum is the complement in \mathbb{C} of the resolvent set, we conclude that σ(f)=Im(f)\sigma(f)=\text{Im}(f).

The spectrum of an element plays a fundamental role in the theory of CC^{\ast}-algebras. One of its most important properties is the relationship between the spectrum of an element and a continuous function defined on that set. This relationship is known as the Image of Spectrum Theorem [12, Corollary 2.37, p. 41]. This connection is based on the following proposition.

Proposition 2.12.

For every a𝒜a\in\mathcal{A} and every polynomial with complex coefficients p(z)[z]p(z)\in\mathbb{C}[z], we have the equality of sets σ(p(a))=p(σ(a))\sigma(p(a))=p(\sigma(a)).

Proof.

If pp is a constant polynomial, the statement is trivial. Therefore, we can assume that p(z)[z]p(z)\in\mathbb{C}[z]\setminus\mathbb{C}. By the Fundamental Theorem of Algebra ([7, Theorem 3.5]), the polynomial q(z)=p(z)λq(z)=p(z)-\lambda can be factorized into linear polynomials: p(z)λ=ci=1n(zβi)p(z)-\lambda=c\prod_{i=1}^{n}(z-\beta_{i}). Now, let’s suppose that λ\lambda is an element of p(σ(a))p(\sigma(a)). Then, there must exist βσ(a)\beta\in\sigma(a) such that λ=p(β)\lambda=p(\beta), meaning that the polynomial q(z)=ci=1n(zβi)q(z)=c\prod_{i=1}^{n}(z-\beta_{i}) vanishes at β\beta. This implies that one of the factors vanishes at β\beta, thus β=βi\beta=\beta_{i} for some ii. The element aβe=aβiea-\beta e=a-\beta_{i}e is non-invertible, and therefore p(a)λe=ci=1n(aβie)p(a)-\lambda e=c\prod_{i=1}^{n}(a-\beta_{i}e) is non-invertible, which means λσ(p(a))\lambda\in\sigma(p(a)). On the other hand, if we take λσ(p(a))\lambda\in\sigma(p(a)), then the element p(a)λe=ci=1n(aβie)p(a)-\lambda e=c\prod_{i=1}^{n}(a-\beta_{i}e) is non-invertible. This implies that at least one of the factors aβiea-\beta_{i}e is non-invertible for some ii. Without loss of generality, we can assume that aβ1ea-\beta_{1}e is non-invertible, meaning that β1σ(a)\beta_{1}\in\sigma(a). As a consequence, the polynomial zβ1z-\beta_{1} vanishes at the spectrum of aa. Thus, we have 0=q(β1)=p(β1)λ0=q(\beta_{1})=p(\beta_{1})-\lambda, or equivalently, λ=p(β1)\lambda=p(\beta_{1}), which means λp(σ(a))\lambda\in p(\sigma(a)). ∎

Theorem 2.13.

The spectrum σ(a)\sigma(a) of an element a𝒜a\in\mathcal{A} is a compact set in \mathbb{C}.

Proof.

Indeed, as a consequence of Corollary 2.8, the spectrum of an element is bounded. Now, by the continuity of the function Fa(λ):=λea𝒜F_{a}(\lambda):=\lambda e-a\in\mathcal{A} defined for every λ\lambda\in\mathbb{C}, we have that Fa1(G(𝒜))=ρ(a)F_{a}^{-1}(G(\mathcal{A}))=\rho(a) is open in \mathbb{C} since G(𝒜)G(\mathcal{A}) is open, according to Corollary 2.8(1). Therefore, its complement σ(a)\sigma(a) is closed. In conclusion, σ(a)\sigma(a) is both closed and bounded in \mathbb{C}, making it a compact set. ∎

Definition 2.14.

The Resolvent Function of an element a𝒜a\in\mathcal{A} is the function

Ra:ρ(a)\displaystyle R_{a}:\rho(a) \displaystyle\to G(𝒜)\displaystyle G(\mathcal{A})
λ\displaystyle\lambda \displaystyle\mapsto (λea)1.\displaystyle(\lambda e-a)^{-1}.
Lemma 2.15.

For every a𝒜a\in\mathcal{A}, σ(a)\sigma(a) is a non-empty set.

Proof.

Consider the case a0a\neq 0, as if a=0a=0 then σ(a)=0\sigma(a)={0}. Suppose, by contradiction, that σ(a)\sigma(a) is empty, which is equivalent to ρ(a)=\rho(a)=\mathbb{C}. In this case, the function RaR_{a} is defined for all \mathbb{C} and, moreover, Ra=invFaR_{a}=\mathrm{inv}\circ F_{a} is continuous, where Fa(λ)=λeaF_{a}(\lambda)=\lambda e-a. Now, rewriting the difference Ra(μ)Ra(λ)R_{a}(\mu)-R_{a}(\lambda), we have:

Ra(μ)Ra(λ)\displaystyle R_{a}(\mu)-R_{a}(\lambda) =\displaystyle= (μea)1(λea)1\displaystyle(\mu e-a)^{-1}-(\lambda e-a)^{-1}
=\displaystyle= (μea)1((λea)\displaystyle(\mu e-a)^{-1}((\lambda e-a)
(μea))(λea)1\displaystyle-(\mu e-a))(\lambda e-a)^{-1}
=\displaystyle= (μea)1(λμ)(λea)1\displaystyle(\mu e-a)^{-1}(\lambda-\mu)(\lambda e-a)^{-1}
=\displaystyle= Ra(μ)(λμ)Ra(λ)\displaystyle R_{a}(\mu)(\lambda-\mu)R_{a}(\lambda)
=\displaystyle= (λμ)Ra(μ)Ra(λ).\displaystyle(\lambda-\mu)R_{a}(\mu)R_{a}(\lambda).

Therefore, for λμ\lambda\neq\mu, we have:

Ra(μ)Ra(λ)μλ=Ra(μ)Ra(λ).\dfrac{R_{a}(\mu)-R_{a}(\lambda)}{\mu-\lambda}=-R_{a}(\mu)R_{a}(\lambda). (5)

Claim 1: The function Ra:𝒜R_{a}:\mathbb{C}\to\mathcal{A} is an analytic function.

Following [6, §10.3], we need to prove that the function f=ϕRaf=\phi\circ R_{a} is a complex-analytic function, for each ϕ𝒜\phi\in\mathcal{A}^{\vee}. Here, 𝒜\mathcal{A}^{\vee} is the dual space (topological), i.e., the space of all continuous linear functionals ϕ:𝒜\phi:\mathcal{A}\to\mathbb{C}. From equation (5), we have:

f(μ)f(λ)μλ\displaystyle\dfrac{f(\mu)-f(\lambda)}{\mu-\lambda} =\displaystyle= ϕ(Ra(μ))ϕ(Ra(λ))μλ\displaystyle\dfrac{\phi(R_{a}(\mu))-\phi(R_{a}(\lambda))}{\mu-\lambda}
=\displaystyle= ϕ(Ra(μ))Ra(λ)μλ)\displaystyle\phi\left(\dfrac{R_{a}(\mu))-R_{a}(\lambda)}{\mu-\lambda}\right)
=\displaystyle= ϕ(Ra(μ)Ra(λ)).\displaystyle\phi(-R_{a}(\mu)R_{a}(\lambda)).

Taking limits in this expression as μλ\mu\to\lambda, we conclude, by the continuity of ϕ\phi, that f(λ)f^{\prime}(\lambda) exists, i.e., ff is analytic over \mathbb{C} and, therefore, the function RaR_{a} is analytic.

Claim 2: The function Ra:𝒜R_{a}:\mathbb{C}\to\mathcal{A} is a constant function.

First, note that if a<|λ|\|a\|<|\lambda|, then we have the inequality:

Ra(λ)=(λea)1|1λ|n=0aλ=1|λ|a.\|R_{a}(\lambda)\|=\|(\lambda e-a)^{-1}\|\leq|\dfrac{1}{\lambda}|\sum_{n=0}^{\infty}\|\dfrac{a}{\lambda}\|=\dfrac{1}{|\lambda|-\|a\|}.

This implies that limλ|Ra(λ)|=0\lim_{\lambda\to\infty}|R_{a}(\lambda)|=0, and in particular, RaR_{a} is bounded. The Uniform Boundedness Principle ([6, §6.7]) guarantees that ϕRa\phi\circ R_{a} is a bounded (entire analytic) function for any ϕ𝒜\phi\in\mathcal{A}^{\vee}. Consequently, by Claim 1 and the Liouville’s Theorem for functions of one complex variable ([7, Theorem 3.5]), ϕRa\phi\circ R_{a} is a constant function for each ϕ𝒜\phi\in\mathcal{A}^{\vee}. Finally, the Hahn-Banach Theorem ([6, §6.3]) guarantees that RaR_{a} is a constant function.

Now, from Claim 2 and limλ|Ra(λ)|=0\lim\limits_{\lambda\to\infty}|R_{a}(\lambda)|=0, it follows that Ra(λ)=0R_{a}(\lambda)=0 for each λ\lambda\in\mathbb{C}. In particular, Ra(0)=a1=0R_{a}(0)=a^{-1}=0, which contradicts our assumption and proves the lemma. ∎

An interesting consequence of Lemma 2.15 is the important Gelfand-Mazur Theorem. Before some definitions will be necessary, which we introduce below.

Definition 2.16 (\ast-homomorphism).

A function ϕ:𝒜\phi:\mathcal{A}\to\mathcal{B} between two commutative CC^{*}-algebras is a \ast-homomorphism if for every a,b𝒜a,b\in\mathcal{A} and λ\lambda\in\mathbb{C}, the following conditions hold:

  • ϕ(a+λb)=ϕ(a)+λϕ(b)\phi(a+\lambda b)=\phi(a)+\lambda\phi(b),

  • ϕ(ab)=ϕ(a)ϕ(b)\phi(ab)=\phi(a)\phi(b),

  • ϕ(a)=ϕ(a)\phi(a^{*})=\phi(a)^{*}.

  • The function ϕ\phi is a \ast-isomorphism if it is bijective. The function ϕ\phi is also an isometric \ast-isomorphism if ϕ(a)=a\|\phi(a)\|=\|a\|.

In particular, each \ast-homomorphism ϕ:𝒜\phi:\mathcal{A}\to\mathbb{C} is an element in the topological dual of 𝒜\mathcal{A}.

Theorem 2.17 (Gelfand-Mazur [2], Thm 14.7).

If 𝒜\mathcal{A} is a unital CC^{\ast}-algebra in which every nonzero element is invertible, then 𝒜\mathcal{A} is isometrically \ast-isomorphic to \mathbb{C}.

Proof.

Let aa be a non-zero element of 𝒜\mathcal{A}. By Lemma 2.15, there exists λa\lambda_{a} such that λaea\lambda_{a}e-a is non-invertible. Consequently, since the only non-invertible element is zero, we have 0=λaea0=\lambda_{a}e-a. Clearly, λa\lambda_{a} is unique, because if there existed λ1σ(a)\lambda_{1}\in\sigma(a) with this property, then λaea=0=λ1ea\lambda_{a}e-a=0=\lambda_{1}e-a, and therefore λa=λ1\lambda_{a}=\lambda_{1}. This allows us to define the function ρ:𝒜\rho:\mathcal{A}\to\mathbb{C}, where ρ(a)=λa\rho(a)=\lambda_{a}. Since a=λaea=\lambda_{a}e, we conclude that ρ\rho is an isometric \ast-isomorphism between 𝒜\mathcal{A} and \mathbb{C}, as desired. ∎

Definition 2.18.

Let a𝒜a\in\mathcal{A}, the Spectral Radius of aa is defined as

r(a):=supλσ(a)|λ|.r(a):=\sup_{\lambda\in\sigma(a)}|\lambda|.
Example 2.19.

Let XX be a Hausdorff-Compact space and fC(X)f\in C(X). In this case,

r(f)=sup{|λ|:λσ(f)}=sup{|λ|:λIm(f)}=f.r(f)=\sup\{|\lambda|:\lambda\in\sigma(f)\}=\sup\{|\lambda|:\lambda\in\text{Im}(f)\}=\|f\|_{\infty}.
Theorem 2.20.

For each a𝒜a\in\mathcal{A}, it holds that

r(a)=limnan1/n=lim infnan1/n.r(a)=\lim_{n\to\infty}\|a^{n}\|^{1/n}=\liminf_{n\in\mathbb{N}}\|a^{n}\|^{1/n}.

In particular, if a=aa=a^{*}, then r(a)=|a|r(a)=|a|.

Proof.

Let λσ(a)\lambda\in\sigma(a), by Proposition 2.12, we have λnσ(an)\lambda^{n}\in\sigma(a^{n}) for every nn\in\mathbb{N}. From Corollary 2.8, we have |λn||an||\lambda^{n}|\leq|a^{n}|, or equivalently,

|λ|an1/n.|\lambda|\leq\|a^{n}\|^{1/n}.

Taking the supremum over λσ(a)\lambda\in\sigma(a) and the lim inf\liminf over nn\in\mathbb{N} yields

r(a)lim infnan1/n.r(a)\leq\liminf_{n\in\mathbb{N}}\|a^{n}\|^{1/n}. (6)

For the other inequality, consider the function

h(λ):=λn=0anλn.h(\lambda):=\lambda\sum_{n=0}^{\infty}\dfrac{a^{n}}{\lambda^{n}}.

If |λ||\lambda| is such that |λ|>lim supnan1/n|\lambda|>\limsup\limits_{n\in\mathbb{N}}{\|a^{n}\|^{1/n}}, then by applying the root test, the series n=0an|λ|n\sum_{n=0}^{\infty}\dfrac{\|a^{n}\|}{|\lambda|^{n}} converges in \mathbb{R}. As a consequence of Corollary 2.9, the series λn=0anλn\lambda\sum_{n=0}^{\infty}\dfrac{a^{n}}{\lambda^{n}} converges absolutely in 𝒜\mathcal{A}. From the definition of h(λ)h(\lambda), we can conclude that h(λ)(λea)=(λea)h(λ)=eh(\lambda)(\lambda e-a)=(\lambda e-a)h(\lambda)=e, which means that if |λ|>lim supnan1/n|\lambda|>\limsup\limits_{n\in\mathbb{N}}{\|a^{n}\|^{1/n}}, then λea\lambda e-a is invertible. Taking the supremum over λ\lambda, we have r(a)lim supnan1/nr(a)\geq\limsup\limits_{n\in\mathbb{N}}{\|a^{n}\|^{1/n}}. In conclusion, r(a)lim supnan1/nlim infnan1/nr(a)r(a)\geq\limsup\limits_{n\in\mathbb{N}}{\|a^{n}\|^{1/n}}\geq\liminf\limits_{n\in\mathbb{N}}\|a^{n}\|^{1/n}\geq r(a), which proves the desired equality.

For the particular case, if a=aa=a^{*}, then a2=aa=a2\|a\|^{2}=\|a^{*}a\|=\|a^{2}\| and by induction a2n=a2n\|a\|^{2^{n}}=\|a^{2^{n}}\|. From the latter relation, we have r(a)=limna2n1/2n=limna2n/2n=ar(a)=\lim_{n\in\mathbb{N}}\|a^{2^{n}}\|^{1/2^{n}}=\lim_{n\in\mathbb{N}}\|a\|^{2^{n}/2^{n}}=\|a\|. ∎

Corollary 2.21.

The norm in a CC^{\ast}-algebra is unique.

Proof.

Suppose 𝒜\mathcal{A} has two norms 1\|\cdot\|_{1} and 2\|\cdot\|_{2} for which a12=aa1\|a\|_{1}^{2}=\|a^{*}a\|_{1} and a22=aa2\|a\|_{2}^{2}=\|a^{*}a\|_{2}. By Theorem 2.20, we have a12=aa1=r(aa)=aa2=a22\|a\|_{1}^{2}=\|a^{*}a\|_{1}=r(a^{*}a)=\|a^{*}a\|_{2}=\|a\|_{2}^{2}, and therefore a1=a2\|a\|_{1}=\|a\|_{2}. ∎

Remark 2.

As a consequence of Theorem 2.20 and Corollary 2.8, we have r(a)ar(a)\leq\|a\|. However, there are CC^{*}-algebras where the strict inequality holds. For example, consider the matrix a=(0100)a=\begin{pmatrix}0&1\\ 0&0\end{pmatrix}, which clearly has a norm greater than zero. Its spectrum consists only of the element zero, so r(a)=0r(a)=0, and thus r(a)<ar(a)<\|a\|. It is not a coincidence that this example is in a non-commutative CC^{*}-algebra. In fact, as a consequence of the categorical version of the Gelfand-Naimark Theorem (see Theorem 5.4), we will conclude that we always have equality for commutative CC^{*}-algebras with unity (see Theorem 6.3).

On the other hand, since λea\lambda e-a is invertible for every |λ|>a|\lambda|>\|a\|, the reasoning in the second part of the proof of the previous theorem guarantees that if λ>r(a)\lambda>r(a), then λea\lambda e-a is invertible.

3 The spectrum of a CC^{*}-algebra

In this section, we begin by gathering the necessary results to prove the main theorem of this work: the categorical version of the Gelfand–Naimark Theorem.

We start by extending the concept of the spectrum from an element to the spectrum of a CC^{\ast}-algebra. This set turns out to be a Hausdorff-compact topological space on which we can define continuous functions. Through the construction of continuous functions on the spectrum of an algebra 𝒜\mathcal{A} (via the Gelfand Transform, Theorem 4.3), we establish the connection between CC^{*}-algebras and topological spaces.

Definition 3.1 (Spectrum of a CC^{\ast}-algebra).

The spectrum of the CC^{\ast}-algebra 𝒜\mathcal{A} is defined as the set

𝒜^={ϕ:𝒜|ϕ it is a nonzero homomorphism}\widehat{\mathcal{A}}=\{\phi:\mathcal{A}\to\mathbb{C}|\,\phi\text{ it is a nonzero $\ast-$homomorphism}\}

We recall that all the CC^{\ast}-algebras of interest are unital, and thus, if ϕ𝒜^\phi\in\widehat{\mathcal{A}}, then ϕ(e)=1\phi(e)=1.

An important concept in the theory of CC^{\ast}-algebras, as will become evident from the following proposition and the further development of this article, is the notion of an ideal. We define an ideal II as a self-adjoint subspace of 𝒜\mathcal{A} such that abIab\in I for all a𝒜a\in\mathcal{A} and bIb\in I. An ideal II is called proper if it is a proper subset of 𝒜\mathcal{A}. Clearly, II is proper if and only if 1I1\notin I, which is also equivalent to II being disjoint from G(𝒜)G(\mathcal{A}).

Associated with each \ast-homomorphism ϕ:𝒜\phi:\mathcal{A}\to\mathbb{C}, the set ker(ϕ)\ker(\phi), defined as the elements a𝒜a\in\mathcal{A} such that ϕ(a)=0\phi(a)=0 and known as the kernel, is an important example of such sets.

Proposition 3.2.

If ϕ:𝒜\phi:\mathcal{A}\to\mathbb{C} is a \ast-homomorphism, then ϕ(a)a\|\phi(a)\|\leq\|a\| for all a𝒜a\in\mathcal{A}. In particular, ϕ\phi is continuous.

Proof.

Since ϕ\phi is a non-zero \ast-homomorphism, it follows that ker(ϕ)\ker(\phi) is a proper ideal of 𝒜\mathcal{A}. Now, for a𝒜a\in\mathcal{A}, we have ϕ(aϕ(a)e)=0\phi(a-\phi(a)e)=0, and therefore aϕ(a)eker(ϕ)a-\phi(a)e\in\ker(\phi). This implies that aϕ(a)ea-\phi(a)e is non-invertible. Thus, ϕ(a)σ(a)\phi(a)\in\sigma(a), and by Corollary 2.8, we conclude that ϕ(a)a\|\phi(a)\|\leq\|a\|. Furthermore, since ϕ(e)=1\phi(e)=1, we have ϕ=1\|\phi\|=1. ∎

To prove the next result, we need to recall two important theorems. The first one, from Functional Analysis, is the Banach-Alaoglu Theorem [3, Thm. 3.6, p. 66], which states that the closed unit ball of 𝒜\mathcal{A}^{\vee} is a compact set in the weak topology. Recall that the weak topology is the topology generated on 𝒜\mathcal{A}^{\vee} by all evaluation functionals. In this topology, the basic open sets Vϕ0V_{\phi_{0}} at an element ϕ0\phi_{0} are characterized as Vϕ0={ϕ𝒜:|(ϕ0ϕ)(ai)|<ϵ}V_{\phi_{0}}=\{\phi\in\mathcal{A}^{\vee}:|(\phi_{0}-\phi)(a_{i})|<\epsilon\} for a finite number of points ai𝒜a_{i}\in\mathcal{A} [3, Section 3.4, p. 64].

The second result, from Topology, states that every closed subspace of a compact space is compact [18, Ch. 3, Theorem 26.2, p. 165].

Theorem 3.3.

The spectrum 𝒜^\widehat{\mathcal{A}} is a compact space in the weak topology.

Proof.

Let S=𝒜^{0}𝒜S=\widehat{\mathcal{A}}\cup\{0\}\subseteq\mathcal{A}^{\vee}. By Proposition 3.2, SS is a subset of the unit ball.

Consider a sequence ϕn\phi_{n} of functionals in SS such that ϕnϕ\phi_{n}\to\phi. We have ϕ(a)ϕ(b)=limnϕn(a)limnϕn(b)=limnϕn(ab)=ϕ(ab)\phi(a)\phi(b)=\lim\limits_{n\to\infty}\phi_{n}(a)\lim\limits_{n\to\infty}\phi_{n}(b)=\lim\limits_{n\to\infty}\phi_{n}(ab)=\phi(ab), which shows that ϕ\phi is multiplicative. Similarly, we obtain ϕ(a)=ϕ(a)¯\phi(a^{*})=\overline{\phi(a)}, and thus ϕ\phi is an element of 𝒜^\widehat{\mathcal{A}}. Consequently, SS is a closed subset of the unit ball in the weak topology, and therefore, SS is compact. Since there are no sequences of norm-one \ast-homomorphisms converging to zero, zero is an isolated point of SS. Hence, 𝒜^\widehat{\mathcal{A}} is compact. ∎

Within the set of all ideals of 𝒜\mathcal{A}, the maximal ideals (in terms of set inclusion) play an important role in the theory of CC^{\ast}-algebras. For example, a fundamental fact is that every proper ideal J0J_{0} is contained in a maximal ideal JJ, as an application of Zorn’s Lemma. The maximality of JJ implies that it must be closed, since its closure is also an ideal.

Let 𝒜\mathcal{A} be an algebra and I𝒜I\subset\mathcal{A} a closed ideal. The set of equivalence classes a+I𝒜/Ia+I\in\mathcal{A}/I, defined by the relation a+I=b+Ia+I=b+I if and only if baIb-a\in I, admits a natural algebraic structure, and with the operation (a+I):=a+I(a+I)^{*}:=a^{*}+I, 𝒜/I\mathcal{A}/I becomes an algebra with involution. The function :𝒜/I+{0}\|\cdot\|:\mathcal{A}/I\to\mathbb{R}^{+}\cup\{0\} defined by a+I:=infa+b,bI\|a+I\|:=\inf{\|a+b\|,b\in I} defines a norm, and the fact that II is closed in 𝒜\mathcal{A} ensures that the normed algebra 𝒜/I\mathcal{A}/I is complete. This norm satisfies a+I=a+I\|a^{*}+I\|=\|a+I\|, and consequently, (a+I)(a+I)a+Ia+Ia+I2\|(a^{*}+I)(a+I)\|\leq\|a^{*}+I\|\|a+I\|\leq\|a+I\|^{2}. To prove that 𝒜/I\mathcal{A}/I is a CC^{*}-algebra, it suffices to show that a+I2(a+I)(a+I)\|a+I\|^{2}\leq\|(a^{*}+I)(a+I)\|. For this, consider bIb\in I and the norm a+b2\|a+b\|^{2}. Since 𝒜\mathcal{A} is a CC^{\ast}-algebra, we have a+b2=(a+b)(a+b)=aa+ab+ba+bb\|a+b\|^{2}=\|(a+b)(a^{*}+b^{*})\|=\|aa^{*}+ab^{*}+ba^{*}+bb^{*}\|. Here, c=ab+ba+bbIc=ab^{*}+ba^{*}+bb^{*}\in I for any aAa\in A, and therefore, infa+b2,bIinfaa+c,c=ab+ba+bb,bI\inf{\|a+b\|^{2},b\in I}\leq\inf{\|aa^{*}+c\|,c=ab^{*}+ba^{*}+bb^{*},b\in I}, which implies a+I2aa+I\|a+I\|^{2}\leq\|aa^{*}+I\|. This establishes the equality and shows that 𝒜/I\mathcal{A}/I is a CC^{*}-algebra.

In this case, the natural projection homomorphism π:𝒜𝒜/I\pi:\mathcal{A}\rightarrow\mathcal{A}/I, defined by aa+Ia\mapsto a+I, is a surjective \ast-homomorphism, it satisfies

Proposition 3.4.

Let ϕ:𝒜\phi:\mathcal{A}\rightarrow\mathcal{B} be a \ast-homomorphism and let II be a closed ideal of 𝒜\mathcal{A} such that Iker(ϕ)I\subseteq\ker(\phi). Then, there exists a unique \ast-homomorphism ψ:𝒜/I\psi:\mathcal{A}/I\rightarrow\mathcal{B} such that ϕ=ψπ\phi=\psi\circ\pi, where π\pi is the projection \ast-homomorphism. In particular, if ϕ\phi is surjective, then we have a \ast-isomorphism 𝒜/ker(ϕ)\mathcal{A}/\ker(\phi)\simeq\mathcal{B}.

Proof.

Indeed, consider ψ:𝒜/I\psi:\mathcal{A}/I\rightarrow\mathcal{B} defined by ψ(a+I)=ϕ(a)\psi(a+I)=\phi(a). This is well-defined: if a+I=b+Ia+I=b+I, then abIa-b\in I, and since Iker(ϕ)I\subseteq\ker(\phi), we have ϕ(ab)=0\phi(a-b)=0, i.e., ψ(a+I)=ϕ(a)=ϕ(b)=ψ(b+I)\psi(a+I)=\phi(a)=\phi(b)=\psi(b+I). Clearly, by its definition, ψ\psi is a \ast-homomorphism and ψπ=ϕ\psi\circ\pi=\phi.

Now, for the particular case, we need to prove that in the case of ψ:𝒜/ker(ϕ)\psi:\mathcal{A}/\ker(\phi)\rightarrow\mathcal{B}, it is injective. Indeed, if ψ(a+ker(ϕ))=0\psi(a+\ker(\phi))=0, it follows from the definition of ψ\psi that ϕ(a)=0\phi(a)=0, i.e., aker(ϕ)a\in\ker(\phi), which means a+ker(ϕ)=0+ker(ϕ)a+\ker(\phi)=0+\ker(\phi). This completes the proof. ∎

The following theorem demonstrates the important relationship between the spectrum of a CC^{\ast}-algebra and the spectrum of an element in that CC^{\ast}-algebra.

Theorem 3.5.

For every a𝒜a\in\mathcal{A}, it holds that

σ(a)={ϕ(a)|ϕ𝒜^}\sigma(a)=\{\phi(a)|\,\phi\in\widehat{\mathcal{A}}\}
Proof.

Let λσ(a)\lambda\in\sigma(a) be given. We define J0J_{0} as

J0=(λea)𝒜:={(λea)b|b𝒜}.J_{0}=(\lambda e-a)\mathcal{A}:=\{(\lambda e-a)b|\,b\in\mathcal{A}\}.

J0J_{0} is an ideal of 𝒜\mathcal{A}. Since λea\lambda e-a is not invertible, J0J_{0} is a proper ideal of 𝒜\mathcal{A}, and thus there exists a closed maximal ideal JJ containing J0J_{0}. Let =𝒜/J\mathcal{B}=\mathcal{A}/J be the quotient algebra. As we have seen before, \mathcal{B} is a CC^{\ast}-algebra.

Let [b][b] be a nonzero class in \mathcal{B}. The subset 𝒥:={j+ab|,jJ,,a𝒜}\mathcal{J}:=\{j+ab|,j\in J,,a\in\mathcal{A}\} of 𝒜\mathcal{A} is an ideal containing JJ. The maximality of JJ implies that 𝒥=𝒜\mathcal{J}=\mathcal{A}. Since 𝒜\mathcal{A} has a unit ee, there exist jj and aa such that e=j+abe=j+ab. Thus, the element eab=je-ab=j belongs to JJ, which in the quotient 𝒜/J\mathcal{A}/J means that [e]=[ab]=[a][b][e]=[ab]=[a][b], and therefore the class [b][b] is invertible. Applying the Gelfand-Mazur Theorem (Theorem 2.17), the CC^{\ast}-algebra 𝒜/J\mathcal{A}/J can be identified with the complex numbers \mathbb{C}.

Thus, the quotient map

π:𝒜𝒜/J=\pi:\mathcal{A}\to\mathcal{A}/J=\mathbb{C}

is a \ast-homomorphism (and hence an element of 𝒜^\widehat{\mathcal{A}}) with kernel JJ. Moreover, since λeaJ=ker(π)\lambda e-a\in J=\ker(\pi), it follows that π(λea)=0\pi(\lambda e-a)=0, or in other words, λ=π(a)\lambda=\pi(a). This shows that σ(a){ϕ(a)|,ϕ𝒜^}\sigma(a)\subseteq\{\phi(a)|,\phi\in\widehat{\mathcal{A}}\}.

We have previously shown that (ϕ(a)1a)ker(ϕ)(\phi(a)1-a)\in\ker(\phi). Since the kernel of a non-zero \ast-homomorphism is a proper ideal, it follows that ϕ(a)1a\phi(a)1-a is non-invertible, i.e., {ϕ(a)|,ϕ𝒜^}σ(a)\{\phi(a)|,\phi\in\widehat{\mathcal{A}}\}\subseteq\sigma(a). This establishes the equality σ(a)={ϕ(a)|,ϕ𝒜^}\sigma(a)=\{\phi(a)|,\phi\in\widehat{\mathcal{A}}\}, completing the proof. ∎

4 The Gelfand Transform

In this section, we introduce the central tool for proving the Gelfand-Naimark Theorem, known as the Gelfand Transform. The Gelfand Transform is an operator that assigns to each element of a CC^{\ast}-algebra a continuous function defined on a suitable compact space. Its importance lies in the fact that this operator acts isometrically and isomorphically on each CC^{\ast}-algebra, leading to a functorial assignment, as we will see later.

One of the examples of CC^{\ast}-algebras that we will consider in this section is the algebra of continuous functions, denoted by C(X)C(X), defined on a Hausdorff-compact space XX (see Example 2.4). In particular, according to Theorem 3.3, the set C(𝒜^)C(\widehat{\mathcal{A}}) is well-defined.

Definition 4.1.

The Gelfand Transform of 𝒜\mathcal{A} is the function κ:𝒜C(𝒜^)\kappa:\mathcal{A}\to C(\widehat{\mathcal{A}}) defined as follows:

κ:𝒜\displaystyle\kappa:\mathcal{A} C(𝒜^)\displaystyle\to C(\widehat{\mathcal{A}})
a\displaystyle a a^:𝒜^ϕa^(ϕ)=ϕ(a).\displaystyle\mapsto\!\begin{aligned} \widehat{a}:\widehat{\mathcal{A}}&\to\mathbb{C}\\ \phi&\mapsto\widehat{a}(\phi)=\phi(a).\end{aligned}

κ\kappa is an -homomorphism

Lemma 4.2.

For every a𝒜a\in\mathcal{A}, we have

κ(a)=a^=r(a)a.\|\kappa(a)\|_{\infty}=\|\hat{a}\|_{\infty}=r(a)\leq\|a\|.
Proof.

For any a𝒜a\in\mathcal{A}, we have:

a^\displaystyle\|\hat{a}\|_{\infty} =\displaystyle= sup{|a^(ϕ)|:ϕ𝒜^}\displaystyle\sup\{|\hat{a}(\phi)|:\phi\in\widehat{\mathcal{A}}\}
=\displaystyle= sup{|ϕ(a)|:ϕ𝒜^}\displaystyle\sup\{|\phi(a)|:\phi\in\widehat{\mathcal{A}}\}
=\displaystyle= sup{|λ|:λσ(a)}\displaystyle\sup\{|\lambda|:\lambda\in\sigma(a)\}
=\displaystyle= r(a),\displaystyle r(a),

where the third equality is guaranteed by Theorem 3.5. ∎

Theorem 4.3 (Gelfand).

Let 𝒜\mathcal{A} be a commutative CC^{\ast}-algebra with unity. The Gelfand Transform κ:𝒜C(𝒜^)\kappa:\mathcal{A}\to C(\widehat{\mathcal{A}}) is an isometric *-isomorphism.

Proof.

For each a𝒜a\in\mathcal{A}, due to the fact that aaa^{*}a is self-adjoint, we have that the spectral radius coincides with the norm of the element (Theorem 2.20), and therefore

a2\displaystyle\|a\|^{2} =\displaystyle= aa=r(aa)\displaystyle\|a^{*}a\|=r(a^{*}a) (7)
=\displaystyle= κ(aa)=κ(a)¯κ(a)\displaystyle\|\kappa(a^{*}a)\|_{\infty}=\|\overline{\kappa(a)}\kappa(a)\|_{\infty} (8)
=\displaystyle= κ(a)2\displaystyle\|\kappa(a)\|_{\infty}^{2} (9)

showing that κ\kappa is an isometry.

On the other hand, let ϕ,ψ𝒜^\phi,\psi\in\widehat{\mathcal{A}} such that ϕψ\phi\neq\psi. This implies that ϕ(a)ψ(a)\phi(a)\neq\psi(a) for some a𝒜a\in\mathcal{A}. Thus, κ(a)(ϕ)=ϕ(a)ψ(a)=κ(a)(ψ)\kappa(a)(\phi)=\phi(a)\neq\psi(a)=\kappa(a)(\psi), i.e., κ(𝒜)\kappa(\mathcal{A}) is a self-adjoint subalgebra of C(𝒜^)C(\widehat{\mathcal{A}}) with unity and separates points. Applying the Stone-Weierstrass Theorem [8, Thm. 8.1, p. 145], we conclude that κ\kappa is surjective. ∎

The main result of this section will pave the way to establish the equivalence between categories that leads to the Gelfand-Naimark Theorem. Specifically, we will present the construction of a CC^{*}-algebra from a locally compact topological space.

Theorem 4.4.

Let XX be a compact Hausdorff space. Then

α:X\displaystyle\alpha:X C(X)^\displaystyle\to\widehat{C(X)}
x\displaystyle x α(x)=ex:C(X)fex(f)=f(x)\displaystyle\mapsto\!\begin{aligned} \alpha(x)=e_{x}:C(X)&\to\mathbb{C}\\ f&\mapsto e_{x}(f)=f(x)\end{aligned}

is a surjective homeomorphism between topological spaces.

Proof.

To prove that α\alpha is a homeomorphism, we need to show that it is well-defined, continuous, injective, surjective, and has a continuous inverse.

First, let’s establish that α\alpha is well-defined. The function α(x)=ex\alpha(x)=e_{x} defined on C(X)C(X) is linear and multiplicative, as it corresponds to the evaluation of functions at the fixed point xx. If we take the constant function f(x)=1f(x)=1 for all xXx\in X, then α(x)(f)=ex(f)=f(x)=10\alpha(x)(f)=e_{x}(f)=f(x)=1\neq 0. Therefore, exe_{x} is non-zero. Moreover, we have ex(f)=f(x)¯=ex(f)e_{x}(f^{*})=\overline{f(x)}=e_{x}(f)^{*}, which shows that α(x)C(X)^\alpha(x)\in\widehat{C(X)}.

Next, we will show that α\alpha is continuous. Let {xi}i=1nX\{x_{i}\}_{i=1}^{n}\subseteq X be a convergent sequence to x0Xx_{0}\in X. We need to demonstrate that exie_{x_{i}} converges to ex0e_{x_{0}} as functionals defined on C(X)C(X). For any given function fC(X)f\in C(X), we have exi(f)=f(xi)e_{x_{i}}(f)=f(x_{i}) and ex0(f)=f(x0)e_{x_{0}}(f)=f(x_{0}). Since ff is continuous, f(xi)f(x_{i}) tends to f(x0)f(x_{0}), and therefore exi(f)e_{x_{i}}(f) tends to ex0(f)e_{x_{0}}(f).

To establish injectivity, consider x,yXx,y\in X with xyx\neq y. By the Urysohn Lemma ([18, Theorem 33.1]), there exists a continuous function ff such that f(x)=0f(x)=0 and f(y)=1f(y)=1. This implies that ex(f)ey(f)e_{x}(f)\neq e_{y}(f), and therefore α\alpha is injective.

To prove surjectivity, let us assume, by contradiction, that the function is not surjective. Therefore, there exists ϕC(X)^\phi\in\widehat{C(X)} for which, given any arbitrary xXx\in X, there exists fxC(X)f_{x}\in C(X) such that

|α(x)(fx)ϕ(fx)|=|fx(x)ϕ(fx)|>0.|\alpha(x)(f_{x})-\phi(f_{x})|=|f_{x}(x)-\phi(f_{x})|>0. (10)

Now, consider the continuous function

gx:X\displaystyle g_{x}:X \displaystyle\to \displaystyle\mathbb{C}
y\displaystyle y \displaystyle\mapsto gx(y)=fx(y)ϕ(fx).\displaystyle g_{x}(y)=f_{x}(y)-\phi(f_{x}).

Based on the inequality (10), we guarantee the existence of a neighborhood VxXV_{x}\subseteq X of xx such that |gx(y)|>0|g_{x}(y)|>0, for all yVxy\in V_{x}.

Clearly, X=xXVxX=\cup_{x\in X}V_{x}, and the compactness of XX guarantees the existence of finitely many x1,x2,,xnXx_{1},x_{2},\cdots,x_{n}\in X such that

X=i=1nVxi.X=\cup_{i=1}^{n}V_{x_{i}}.

Let g=i=1ngxigxi¯C(X)g=\sum_{i=1}^{n}g_{x_{i}}\overline{g_{x_{i}}}\in C(X), where the image of gg is real. For each yXy\in X, there exists xix_{i} such that yVxiy\in V_{x_{i}}, and therefore gxigxi¯(y)>0g_{x_{i}}\overline{g_{x_{i}}}(y)>0. This implies that g(y)>0g(y)>0 for all yXy\in X.

The functional ϕ\phi evaluated at gg can be calculated as follows:

ϕ(g)\displaystyle\phi(g) =\displaystyle= ϕ(i=1n(fxiϕ(fxi))(fxiϕ(fxi))¯)\displaystyle\phi\left(\sum_{i=1}^{n}(f_{x_{i}}-\phi(f_{x_{i}}))\overline{(f_{x_{i}}-\phi(f_{x_{i}}))}\right)
=\displaystyle= i=1n(ϕ(fxi)ϕ(fxi))ϕ(fxiϕ(fxi)¯)\displaystyle\sum_{i=1}^{n}(\phi(f_{x_{i}})-\phi(f_{x_{i}}))\phi(\overline{f_{x_{i}}-\phi(f_{x_{i}})})
=\displaystyle= 0.\displaystyle 0.

By Theorem 3.5, we have 0σ(g)=Im(g)0\in\sigma(g)=\text{Im}(g), which means gg vanishes, which is absurd. Therefore, we conclude that α\alpha is surjective.

Finally, by [18, Theorem 26.6], we conclude that α\alpha is a homeomorphism since it is bijective and continuous, as desired.

5 Gelfand-Naimark Theorem and Category Theory

A theory in mathematics is, in informal and simplistic terms, a collection of sets endowed with certain structure and functions that preserve that structure. For example, the theory of topological spaces consists of topological spaces and continuous functions on those spaces. In this sense, Category Theory can be interpreted as a “meta-theory” that provides the language and tools to deal with various theories in mathematics and establish connections between them through so-called “functors”. Thus, it is valid to interpret Category Theory as providing a panoramic view of mathematics, where certain theorems, when translated into the language of this theory, establish connections between different theories.

The main objective of this section is to present an example of such connections through the categorical version of the Gelfand-Naimark Theorem. As we will see, this theorem allows us to establish a connection between the theories of Hausdorff-compact spaces and the theory of commutative unital CC^{\ast}-algebras. In this context, the theorem can be interpreted, in a rough sense, as stating that all “geometric” information about a space is contained in its algebra of functions, and conversely, every CC^{\ast}-algebra is characterized by a certain space of continuous functions on a Hausdorff-compact space.

These types of global results, which link seemingly disparate contexts, bring notable benefits to the development of the involved theories. This will be demonstrated in the applications presented in the next section, where, for example, we will characterize the spectrum of self-adjoint, unitary, normal, and positive elements. Although this is not an elementary task in an arbitrary CC^{\ast}-algebra, it becomes much simpler in light of the Gelfand-Naimark theorem in its categorical version. Moreover, such global approaches have facilitated the construction of new, fruitful theories that would have otherwise been difficult to develop, such as noncommutative geometry, a topic that we will not delve into in this article but encourage curious readers to explore in references [5] and [22].

Before presenting the theorem, we will give a brief introduction to the language and tools of category theory. For this introductory part, we suggest references [1] and [17].

Definition 5.1 (Category).

A category 𝒞\mathcal{C} is formed by the following conditions:

  • A class of objects, denoted by Ob(𝒞)\text{Ob}(\mathcal{C}).

  • A set of morphisms, consisting of the sets Mor𝒞(A,B)\text{Mor}_{\mathcal{C}}(A,B) for each A,BOb(𝒞)A,B\in\text{Ob}(\mathcal{C}), which satisfy the following: On one hand, for each AOb(𝒞)A\in\text{Ob}(\mathcal{C}), there exists 1AMor𝒞(A,A)1_{A}\in\text{Mor}_{\mathcal{C}}(A,A). On the other hand, there is a function :Mor𝒞(A,B)×Mor𝒞(B,C)Mor𝒞(A,C)\circ:\text{Mor}_{\mathcal{C}}(A,B)\times\text{Mor}_{\mathcal{C}}(B,C)\to\text{Mor}_{\mathcal{C}}(A,C), (f,g)gf(f,g)\mapsto g\circ f, for any A,B,COb(𝒞)A,B,C\in\text{Ob}(\mathcal{C}), which satisfies the relation g(fh)=(gf)hg\circ(f\circ h)=(g\circ f)\circ h and f1A=f=1Bff\circ 1_{A}=f=1_{B}\circ f for each fMor𝒞(A,B)f\in\text{Mor}_{\mathcal{C}}(A,B).

Some of the main examples of categories are:

  • The category of sets, denoted by Set: whose objects are sets and morphisms are functions between these sets.

  • The category of topological spaces, denoted by Top: whose objects are topological spaces and morphisms are continuous functions between these spaces.

  • The previous example is the standard example of a category, i.e., collections of sets with some “structure” and morphisms being functions that “preserve this structure”. In this sense, we have a wealth of examples that satisfy these characteristics: groups and group homomorphisms, rings and ring homomorphisms, vector spaces over a certain field and linear transformations between them, smooth manifolds and smooth functions, and so on. These examples belong to the categories that we will define in this section.

  • Within the morphisms of a category 𝒞\mathcal{C}, there are some that deserve special attention, namely isomorphisms. An isomorphism f:ABf:A\to B is a morphism for which there exists another morphism g:BAg:B\to A such that fg=1Bf\circ g=1_{B} and gf=1Ag\circ f=1_{A}. In this case, the objects AA and BB of 𝒞\mathcal{C} are called isomorphic. Note that the relation “being isomorphic to” defines an equivalence relation on the objects of 𝒞\mathcal{C}.

  • For any category 𝒞\mathcal{C}, we define the opposite category, denoted by 𝒞op\mathcal{C}^{\text{op}}, as the category with the same objects as 𝒞\mathcal{C} and morphisms from 𝒞\mathcal{C} with the “inverted arrows,” i.e., Mor𝒞op(A,B):=Mor𝒞(B,A)\text{Mor}_{\mathcal{C}^{\text{op}}}(A,B):=\text{Mor}_{\mathcal{C}}(B,A) for each A,BOb(𝒞)A,B\in\text{Ob}(\mathcal{C}). Note that in this case, for any pair of morphisms ff and gg in 𝒞op\mathcal{C}^{\text{op}}, the composition “𝒞op\circ_{\mathcal{C}^{\text{op}}}” is naturally defined by g𝒞opf:=f𝒞gg\circ_{\mathcal{C}^{\text{op}}}f:=f\circ_{\mathcal{C}}g.

The categories that will be related through the categorical version of the Gelfand-Naimark Theorem, which we will study in this section, are as follows:

  • The category of Hausdorff-Compact spaces, denoted by 𝒞\mathcal{HC}: The objects are Hausdorff-Compact topological spaces. The morphisms are continuous functions between these spaces.

  • The category of commutative unital CC^{*}-algebras, denoted by 𝒞𝒜𝒰\mathcal{CAU}: The objects are commutative unital CC^{*}-algebras. The morphisms are \ast-homomorphisms ϕ:𝒜\phi:\mathcal{A}\rightarrow\mathcal{B} such that ϕ(e𝒜)=e\phi(e_{\mathcal{A}})=e_{\mathcal{B}}.

Remark 3.

In the categories of our interest, 𝒞\mathcal{HC} and 𝒞𝒜𝒰\mathcal{CAU}, the isomorphisms are, respectively, homeomorphisms and isometric \ast-isomorphisms that preserve the unit.

Now, it is natural to consider “morphisms” that preserve the category structure. This allows us to define the concept of a functor.

Definition 5.2.

Let 𝒞\mathcal{C} and 𝒟\mathcal{D} be two categories. A covariant functor (resp. contravariant functor) F:𝒞𝒟F:\mathcal{C}\rightarrow\mathcal{D} is determined by the following attributes:

  • F(A)Ob(𝒟)F(A)\in\text{Ob}(\mathcal{D}) for every AOb(𝒞)A\in\text{Ob}(\mathcal{C}).

  • If f:ABf:A\rightarrow B is a morphism in 𝒞\mathcal{C}, then F(f):F(A)F(B)F(f):F(A)\rightarrow F(B) (resp. F(f):F(B)F(A)F(f):F(B)\rightarrow F(A)) is a morphism in 𝒟\mathcal{D}.

  • F(1A)=1F(A)F(1_{A})=1_{F(A)} for every AOb(𝒞)A\in\text{Ob}(\mathcal{C}).

  • F(fg)=F(f)F(g)F(f\circ g)=F(f)\circ F(g) (resp. F(fg)=F(g)F(f)F(f\circ g)=F(g)\circ F(f)) for every pair of morphisms ff and gg in 𝒞\mathcal{C}.

Naturally, we define the identity functor Id𝒞:𝒞𝒞Id_{\mathcal{C}}:\mathcal{C}\rightarrow\mathcal{C} and the composition of functors GF:𝒞G\circ F:\mathcal{C}\rightarrow\mathcal{E} for functors F:𝒞𝒟F:\mathcal{C}\rightarrow\mathcal{D} and G:𝒟G:\mathcal{D}\rightarrow\mathcal{E}. We say that a functor F:𝒞𝒟F:\mathcal{C}\rightarrow\mathcal{D} is an isomorphism if there exists a functor G:𝒟𝒞G:\mathcal{D}\rightarrow\mathcal{C} such that FG=Id𝒟F\circ G=Id_{\mathcal{D}} and GF=Id𝒞G\circ F=Id_{\mathcal{C}}. In this case, the categories 𝒞\mathcal{C} and 𝒟\mathcal{D} are called isomorphic.

However, in some cases, the condition that two categories are isomorphic can be difficult to achieve. Therefore, it is necessary to introduce weaker conditions to “compare” two categories. To do so, we first introduce a natural way to “compare” functors:

Definition 5.3.

Let F,G:𝒞𝒟F,G:\mathcal{C}\rightarrow\mathcal{D} be two covariant functors between the categories 𝒞\mathcal{C} and 𝒟\mathcal{D}. A natural transformation μ:FG\mu:F\rightarrow G between the functors FF and GG is a collection of morphisms μ(A):F(A)G(A)\mu(A):F(A)\rightarrow G(A), for each AOb(𝒞)A\in\text{Ob}(\mathcal{C}), such that for every morphism ψ:AB\psi:A\rightarrow B in 𝒞\mathcal{C}, the following diagram

F(A)\textstyle{F(A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(A)\scriptstyle{\mu(A)}F(ψ)\scriptstyle{F(\psi)}G(A)\textstyle{G(A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G(ψ)\scriptstyle{G(\psi)}F(B)\textstyle{F(B)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(B)\scriptstyle{\mu(B)}G(B)\textstyle{G(B)}

commutes.

Analogous definitions hold for contravariant functors FF and GG and their possible combinations. We say that FF and GG are isomorphic if there exists a natural transformation μ:FG\mu:F\rightarrow G that is an isomorphism, i.e., μ(A):F(A)G(A)\mu(A):F(A)\rightarrow G(A) is an isomorphism for each AOb(𝒞)A\in\text{Ob}(\mathcal{C}). In this case, we use the notation FGF\simeq G.

Finally, we will define what we mean by equivalent categories. This notion of equivalence is inspired by the homotopy equivalence of two topological spaces. Indeed, we say that two categories 𝒞\mathcal{C} and 𝒟\mathcal{D} are equivalent if there exists a functor F:𝒞𝒟F:\mathcal{C}\rightarrow\mathcal{D} that has a pseudo-inverse, i.e., there exists a functor G:𝒟𝒞G:\mathcal{D}\rightarrow\mathcal{C} such that FG1𝒟F\circ G\simeq 1_{\mathcal{D}} and GF1𝒞G\circ F\simeq 1_{\mathcal{C}}.

As is easy to see, any pair of isomorphic categories are equivalent. However, in general, equivalent categories do not have to be isomorphic. The richness of this notion of equivalence, in practice, lies in the fact that it allows us to “translate” apparently disparate properties between categories that, at first glance, seem to have no connection.

Theorem 5.4 (Gelfand–Naimark (Categorical Version)).

The categories 𝒞op\mathcal{HC}^{\text{op}} and 𝒞𝒜𝒰\mathcal{CAU} are equivalent. In particular, for every commutative CC^{\ast}-algebra 𝒜\mathcal{A} with unit, there exists a compact Hausdorff space XX such that 𝒜\mathcal{A} is isometrically \ast-isomorphic to C(X)C(X).

Proof.

In light of the previous definitions and observations, the proof of this theorem will follow the following steps:

  • Step 1: We define FF, a covariant functor from 𝒞𝒜𝒰\mathcal{CAU} to the category 𝒞op\mathcal{HC}^{\text{op}}.

  • Step 2: We define GG, a covariant functor from the category 𝒞op\mathcal{HC}^{\text{op}} to 𝒞𝒜𝒰\mathcal{CAU}.

  • Step 3: We define a natural transformation τ:Id𝒞𝒜𝒰GF\tau:Id_{\mathcal{CAU}}\to G\circ F, which we must prove to be an isomorphism.

  • Step 4: We define a natural transformation μ:Id𝒞opFG\mu:Id_{\mathcal{HC}^{\text{op}}}\to F\circ G, which we must prove to be an isomorphism.

To solve Step 1, we consider the functor F:𝒞𝒜𝒰𝒞opF:\mathcal{CAU}\to\mathcal{HC}^{\text{op}} defined on objects by:

F:Ob(𝒞𝒜𝒰)\displaystyle F:\text{Ob}(\mathcal{CAU}) \displaystyle\to Ob(𝒞op)\displaystyle\text{Ob}(\mathcal{HC}^{\text{op}})
𝒜\displaystyle\mathcal{A} \displaystyle\mapsto F(𝒜):=𝒜^\displaystyle F(\mathcal{A}):=\widehat{\mathcal{A}}

and on morphisms by:

F:Mor𝒞𝒜𝒰(𝒜,)\displaystyle F:\text{Mor}_{\mathcal{CAU}}(\mathcal{A},\mathcal{B}) Mor𝒞op(𝒜^,^)\displaystyle\to\text{Mor}_{\mathcal{HC}^{op}}(\widehat{\mathcal{A}},\widehat{\mathcal{B}})
ϕ\displaystyle\phi F(ϕ):^𝒜^ψψϕ\displaystyle\mapsto\!\begin{aligned} F(\phi):\widehat{\mathcal{B}}&\to\widehat{\mathcal{A}}\\ \psi&\mapsto\psi\circ\phi\end{aligned}

For the second step, consider G:𝒞op𝒞𝒜𝒰G:\mathcal{HC}^{\text{op}}\to\mathcal{CAU} defined on objects by:

G:Ob(𝒞op)\displaystyle G:\text{Ob}(\mathcal{HC}^{\text{op}}) \displaystyle\to Ob(𝒞𝒜𝒰)\displaystyle\text{Ob}(\mathcal{CAU})
X\displaystyle X \displaystyle\mapsto G(X):=C(X)\displaystyle G(X):=C(X)

and on morphisms by:

G:Mor𝒞op(X,Y)\displaystyle G:\text{Mor}_{\mathcal{HC}^{\text{op}}}(X,Y) Mor𝒞𝒜𝒰(C(X),C(Y))\displaystyle\to\text{Mor}_{\mathcal{CAU}}(C(X),C(Y))
f\displaystyle f G(f):C(X)C(Y)ggf\displaystyle\mapsto\!\begin{aligned} G(f):C(X)&\to C(Y)\\ g&\mapsto g\circ f\end{aligned}

It can be easily verified that both FF and GG satisfy the conditions in Definition 5.2, thus they are covariant functors between the categories 𝒞\mathcal{HC} and 𝒞𝒜𝒰\mathcal{CAU}.

For the third step, we define τ:Id𝒞𝒜𝒰GF\tau:\text{Id}_{\mathcal{CAU}}\to G\circ F, supported by the Gelfand transform, as the collection of morphisms for each commutative unital CC^{\ast}-algebra 𝒜\mathcal{A}:

τ(𝒜):𝒜\displaystyle\tau(\mathcal{A}):\mathcal{A} C(𝒜^)\displaystyle\to C(\widehat{\mathcal{A}})
a\displaystyle a τ(𝒜)(a):=a^:𝒜^ϕa^(ϕ)=ϕ(a).\displaystyle\mapsto\!\begin{aligned} \tau(\mathcal{A})(a):=\widehat{a}:\widehat{\mathcal{A}}&\to\mathbb{C}\\ \phi&\mapsto\widehat{a}(\phi)=\phi(a)\end{aligned}.

The Gelfand-Naimark Theorem guarantees that τ(𝒜)\tau(\mathcal{A}) is an isometric \ast-isomorphism for every commutative unital CC^{\ast}-algebra 𝒜\mathcal{A}. Thus, to conclude that τ\tau is an isomorphism between the functors, we only need to prove that it is a natural transformation, i.e., for every ϕ:𝒜\phi:\mathcal{A}\to\mathcal{B} \ast-homomorphism that preserves the unit, the following diagram commutes:

𝒜\textstyle{\mathcal{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ(𝒜)\scriptstyle{\tau(\mathcal{A})}ϕ\scriptstyle{\phi}C(𝒜^)\textstyle{C(\widehat{\mathcal{A}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G(F(ϕ))\scriptstyle{G(F(\phi))}\textstyle{\mathcal{B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ()\scriptstyle{\tau(\mathcal{B})}C(^)\textstyle{C(\widehat{\mathcal{B}})}

Indeed, for each a𝒜a\in\mathcal{A}, we need to show that the functions G(F(ϕ))(a^)=τ()(ϕ(a))G(F(\phi))(\widehat{a})=\tau(\mathcal{B})(\phi(a)), i.e., for each ψ^\psi\in\widehat{\mathcal{B}}, G(F(ϕ))(a^)(ψ)=τ()(ϕ(a))(ψ)G(F(\phi))(\widehat{a})(\psi)=\tau(\mathcal{B})(\phi(a))(\psi). This follows from the following sequence of equalities:

G(F(ϕ))(a^)(ψ)\displaystyle G(F(\phi))(\widehat{a})(\psi) =\displaystyle= a^[(F(ϕ))(ψ)]=a^(ψϕ)\displaystyle\widehat{a}[(F(\phi))(\psi)]=\widehat{a}(\psi\circ\phi)
=\displaystyle= ψ(ϕ(a))=ϕ(a)^(ψ)\displaystyle\psi(\phi(a))=\widehat{\phi(a)}(\psi)
=\displaystyle= τ()(ϕ(a))(ψ)\displaystyle\tau(\mathcal{B})(\phi(a))(\psi)

Finally, for the fourth step, we define μ:Id𝒞FG\mu:\text{Id}_{\mathcal{HC}}\to F\circ G as the collection of morphisms for each Hausdorff compact space XX:

μ(X):X\displaystyle\mu(X):X C(X)^\displaystyle\to\widehat{C(X)}
x\displaystyle x μ(X)(x)=ex:C(X)fex(f)=f(x).\displaystyle\mapsto\!\begin{aligned} \mu(X)(x)=e_{x}:C(X)&\to\mathbb{C}\\ f&\mapsto e_{x}(f)=f(x)\end{aligned}.

Theorem 4.4 guarantees that μ(X)\mu(X) is an isomorphism in Top, i.e., a homeomorphism. We only need to prove that μ\mu is a natural transformation, i.e., the following diagram commutes:

X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(X)\scriptstyle{\mu(X)}f\scriptstyle{f}C(X)^\textstyle{\widehat{C(X)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F(G(f))\scriptstyle{F(G(f))}Y\textstyle{Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(Y)\scriptstyle{\mu(Y)}C(Y)^\textstyle{\widehat{C(Y)}}

Indeed, for xXx\in X and each gC(Y)g\in C(Y), we have:

[F(G(f))(ex)](g)\displaystyle[F(G(f))(e_{x})](g) =\displaystyle= (ex)[G(f)(g)]=g(f(x))=ef(x)(g)\displaystyle(e_{x})[G(f)(g)]=g(f(x))=e_{f(x)}(g)
=\displaystyle= μ(Y)(f(x))(g)\displaystyle\mu(Y)(f(x))(g)

This completes the proof of the theorem.

6 Applications

In this section, we present two applications as a consequence of the interpretation of the Gelfand–Naimark Theorem in categorical terms. The first of them, with a more algebraic flavor, is inspired by the Nullstellensatz or the Hilbert’s Nullstellensatz ([19, Section 3.8, p. 62]), which establishes a categorical equivalence between the category of finitely generated and reduced \mathbb{C}-algebras and the category of affine algebraic varieties over \mathbb{C}. In general terms, the Nullstellensatz is a result that, when translated into categorical terms, establishes connections between the space and its geometric information, and its \mathbb{C}-algebra of regular functions and its algebraic information. In a similar vein, it acts similarly to the Gelfand–Naimark theorem.

The second application, with a more analytical flavor, will demonstrate the simplicity brought by a global theorem of this kind in determining properties and characterizing elements of commutative CC^{\ast}-algebras with unit, starting from the knowledge and manipulation of these elements on the CC^{\ast}-algebra of continuous functions over a Hausdorff-Compact space.

Before stating and presenting our main applications, we need to introduce some notation. For each CC^{\ast}-algebra 𝒜\mathcal{A}, we define Max(𝒜)\text{Max}(\mathcal{A}) as the set of maximal ideals of 𝒜\mathcal{A}. We endow this set with a topology called the Zariski topology, which, as we will see, has analytical significance.

The Zariski topology associates to each ideal I𝒜I\subseteq\mathcal{A} the set

V(I)={𝔪Max(𝒜)|𝔪I}.V(I)=\left\{\mathfrak{m}\in\text{Max}(\mathcal{A})|\,\mathfrak{m}\supseteq I\right\}.

For these sets, we have the following properties:

V((0))=Max(𝒜);V(I)V(J)=V(IJ);\displaystyle V((0))=\text{Max}(\mathcal{A});\quad V(I)\cup V(J)=V(I\cap J);
V(𝒜)=ϕ;αΛV(Iα)=V(αΛIα).\displaystyle V(\mathcal{A})=\phi;\quad\bigcap_{\alpha\in\Lambda}V(I_{\alpha})=V(\sum_{\alpha\in\Lambda}I_{\alpha}).

In this way, the complements of each V(I)V(I) form a topology on Max(𝒜)\text{Max}(\mathcal{A}) by considering each V(I)V(I) as a closed set in this topology. Notice that, in particular, each point 𝔪Max(𝒜)\mathfrak{m}\in\text{Max}(\mathcal{A}) is closed since V(𝔪)=𝔪V(\mathfrak{m})={\mathfrak{m}}.

Theorem 6.1.

The topological spaces XX and Max(C(X))\text{Max}(C(X)) are homeomorphic. Specifically, the function

ζ:X\displaystyle\zeta:X \displaystyle\to Max(C(X))\displaystyle\text{Max}{(C(X))}
x\displaystyle x \displaystyle\mapsto kerex\displaystyle\ker{e_{x}}

is actually a homeomorphism. In particular, Max(C(X))\text{Max}(C(X)) and C(X)^\widehat{C(X)} are homeomorphic, where C(X)^\widehat{C(X)} is equipped with the weak topology.

Proof.

First, let’s observe that ζ\zeta is well-defined. In fact, since C(X)C(X) is an algebra that separates points, exe_{x} is a nonzero surjective \ast-homomorphism, and therefore ker(ex)\ker(e_{x}) is a maximal ideal, as the quotient C(X)/ker(ex)C(X)/\ker(e_{x})\simeq\mathbb{C} is a field. Now let’s show that ζ\zeta is continuous. For every ideal IC(X)I\subseteq C(X), we have

ζ1(V(I))\displaystyle\zeta^{-1}(V(I)) =\displaystyle= {xX:ζ(x)V(I)}\displaystyle\{x\in X:\zeta(x)\in V(I)\}
=\displaystyle= {xX:Iker(ex)}\displaystyle\{x\in X:I\subseteq\ker(e_{x})\}
=\displaystyle= {xX:f(x)=0 for all fI}\displaystyle\{x\in X:f(x)=0\text{ for all }f\in I\}
=\displaystyle= fIf1({0})\displaystyle\bigcap_{f\in I}f^{-1}(\{0\})

where the last expression is a closed set in XX as it is an arbitrary intersection of closed sets. Clearly, ζ\zeta is injective, since if xyx\neq y, by Urysohn’s Lemma ([18, Ch. 4, Theorem 33.1]), there exists fC(X)f\in C(X) such that f(x)=0f(x)=0 and f(y)0f(y)\neq 0, i.e., ker(ex)ker(ey)\ker(e_{x})\neq\ker(e_{y}).

For surjectivity, let ηMax(C(X))\eta\in\text{Max}(C(X)) be a maximal ideal, and let’s show that there exists xXx\in X such that η=ker(ex)\eta=\ker(e_{x}). In other words, we need to prove that every maximal ideal arises as the kernel of an evaluation \ast-homomorphism. Indeed, since η\eta is maximal, it is a closed ideal, and therefore C(X)/ηC(X)/\eta is a commutative unital CC^{\ast}-algebra. Moreover, the maximality of η\eta implies that C(X)/ηC(X)/\eta is a division algebra (in fact, a field), and hence C(X)/ηC(X)/\eta\cong\mathbb{C} by the Gelfand-Mazur theorem (2.17). Thus, the projection \ast-homomorphism (which is surjective and preserves the unit)

ϕ:C(X)\displaystyle\phi:C(X) \displaystyle\rightarrow C(X)/η\displaystyle C(X)/\eta\cong\mathbb{C}
f\displaystyle f \displaystyle\mapsto ϕ(f)=f^=f+η\displaystyle\phi(f)=\hat{f}=f+\eta

satisfies η=ker(ϕ)\eta=\ker(\phi), and by Theorem 4.4, there exists xXx\in X such that ϕ=ex\phi=e_{x}, i.e., η=ker(ex)\eta=\ker(e_{x}).

In conclusion, ζ\zeta is a bijective function that is continuous on a Hausdorff-compact space, and therefore it is a homeomorphism.

The particular case follows from Theorem 4.4, which tells us that the function

C(X)^\displaystyle\widehat{C(X)} \displaystyle\to Max(C(X))\displaystyle\text{Max}{(C(X))}
ϕ\displaystyle\phi \displaystyle\mapsto kerϕ.\displaystyle\ker{\phi}.

is a homeomorphism. ∎

Theorem 6.2.

Let XX be a compact Hausdorff space, and let C(X)C(X) be its corresponding CC^{\ast}-algebra of continuous functions on XX. There is a bijective correspondence

{YXYis closed}{closed ideals ofC(X)}\displaystyle\{Y\subseteq X\mid Y\,\text{is closed}\}\longleftrightarrow\{\text{closed ideals of}\,\,C(X)\}
Proof.

For each closed set YXY\subseteq X (which is compact), the inclusion map i:YXi:Y\to X is continuous and injective, and it induces, as a consequence of the Gelfand-Naimark Theorem 5.4, a surjective \ast-homomorphism i:C(X)C(Y)i^{\ast}:C(X)\rightarrow C(Y) between the CC^{\ast}-algebras. The kernel of this map is

keri:=IY={fC(X):f(y)=0,yY}.\ker{i^{*}}:=I_{Y}=\{f\in C(X):f(y)=0,\,\forall y\in Y\}.

Let’s show that IYI_{Y} is closed in the weak topology. Indeed, for {fn}nIY\{f_{n}\}_{n\in\mathbb{N}}\subset I_{Y} a sequence of continuous functions that converges to fC(X)f\in C(X), we have that for every ϵ>0\epsilon>0 and each yYy\in Y, |f(y)|=|fn(y)f(y)|<ϵ|f(y)|=|f_{n}(y)-f(y)|<\epsilon, which implies that fIYf\in I_{Y}.

Thus, the natural algebraic isomorphism between C(X)/IYC(X)/I_{Y} and C(Y)C(Y) induced by ii^{\ast} (see Proposition 3.4) is actually a \ast-algebra isomorphism.

Conversely, every closed ideal II of C(X)C(X) defines a surjective projection \ast-homomorphism between the CC^{\ast}-algebras

ϕ:C(X)C(X)/I\phi:C(X)\hookrightarrow C(X)/I

such that kerϕ=I\ker{\phi}=I. Now, by virtue of the Gelfand-Naimark Theorem 5.4, for the CC^{\ast}-algebra C(X)/IC(X)/I, there exists a compact set YY such that α:C(Y)C(X)/I\alpha:C(Y)\stackrel{{\scriptstyle\simeq}}{{\rightarrow}}C(X)/I is a \ast-isomorphism. Thus, the surjective \ast-homomorphism α1ϕ\alpha^{-1}\circ\phi corresponds, by the same theorem, to a continuous and injective function f:YXf:Y\rightarrow X whose image f(Y)f(Y) is homeomorphic to YY and closed in XX, i.e., compact. Identifying YY with its image, we conclude that the ideal II corresponds to the compact subset YY of XX, as desired. ∎

Theorem 6.3.

Let 𝒜\mathcal{A} be a commutative CC^{\ast}-algebra with unity. For every a𝒜a\in\mathcal{A}, we have r(a)=ar(a)=\|a\|.

Proof.

By the categorical Gelfand-Naimark Theorem (5.4), the theorem reduces to proving the equality for fC(X)f\in C(X) with XX a compact Hausdorff space. In this latter case, by example (2.19), we conclude that r(f)=|f|r(f)=|f|_{\infty}. ∎

Next, we present an application where the categorical theorem allows us to go back and forth with information about a specific property in the CC^{\ast}-algebra 𝒜\mathcal{A} to the set of functions C(𝒜^)C(\widehat{\mathcal{A}}). In this specific case, the property of interest is invertibility, i.e., an element aa is invertible in 𝒜\mathcal{A} if and only if a^\hat{a} is invertible in C(𝒜^)C(\widehat{\mathcal{A}}) and vice versa.

Definition 6.4.

Given a CC^{\ast}-algebra 𝒜\mathcal{A}, an element a𝒜a\in\mathcal{A} is called:

  1. 1.

    Self-adjoint if a=aa=a^{*}.

  2. 2.

    Unitary if aa=aa=eaa^{*}=a^{*}a=e.

  3. 3.

    Projection if a=aa^{*}=a and a2=aa^{2}=a.

  4. 4.

    Positive if a=bba=bb^{*} for some b𝒜b\in\mathcal{A}.

Theorem 6.5.

Let a𝒜a\in\mathcal{A}.

  1. 1.

    If aa is self-adjoint, then σ(a)\sigma(a)\subset\mathbb{R}.

  2. 2.

    If aa is unitary, then σ(a)S1\sigma(a)\subset S^{1}.

  3. 3.

    If aa is a projection, then σ(a){0,1}\sigma(a)\subset\{0,1\}.

  4. 4.

    If aa is positive, then σ(a)+0\sigma(a)\subset\mathbb{R}^{+}\cup{0}.

Proof.

It is clear that σ(a)=σ(a^)=Im(a^)\sigma(a)=\sigma(\hat{a})=\text{Im}(\hat{a}).

  1. 1.

    A function ff is self-adjoint if f(z)¯=f(z)\overline{f(z)}=f(z) for every element in the domain. Since the spectrum of ff is its image, we have σ(f)¯=σ(f)\overline{\sigma(f)}=\sigma(f) and therefore σ(f)\sigma(f)\subset\mathbb{R}. In particular, for the function a^\hat{a}, we have σ(a)=σ(a^)\sigma(a)=\sigma(\hat{a})\subset\mathbb{R}.

  2. 2.

    A function ff is unitary if f(z)¯f(z)=1\overline{f(z)}f(z)=1 for every element in the domain. Since the spectrum of ff is its image, the elements of the spectrum have norm 1 and therefore σ(f)S1\sigma(f)\subset S^{1}. In particular, for the function a^\hat{a}, we have σ(a)=σ(a^)S1\sigma(a)=\sigma(\hat{a})\subset S^{1}.

  3. 3.

    A function ff is a projection if it is self-adjoint, and therefore its spectrum is real; and furthermore, f2(z)=f(z)f^{2}(z)=f(z) for every element in the domain. This implies that f(z)=1f(z)=1 or f(z)=0f(z)=0 for every element zz in the domain. Since the spectrum of ff is its image, we have σ(f){0,1}\sigma(f)\subset\{0,1\}. In particular, for the function a^\hat{a}, we have σ(a)=σ(a^){0,1}\sigma(a)=\sigma(\hat{a})\subset\{0,1\}.

  4. 4.

    A function ff is positive if f(z)f(z) is real and f(z)0f(z)\geq 0 for every element in the domain. Since the spectrum of ff is its image, we have σ(f)+{0}\sigma(f)\subset\mathbb{R}^{+}\cup\{0\}. In particular, for the function a^\hat{a}, we have σ(a)=σ(a^)+{0}\sigma(a)=\sigma(\hat{a})\subset\mathbb{R}^{+}\cup\{0\}.

References

  • [1] Steve Awodey. Category Theory. Oxford University Press, may 2006.
  • [2] Frank F. Bonsall and John Duncan. Complete Normed Algebras. Springer Berlin Heidelberg, 1973.
  • [3] Haim Brezis. Functional Analysis, Sobolev Spaces and Partial Differential Equations. Springer International Publishing, 2011.
  • [4] Paul Civin and Bertram Yood. Involutions on banach algebras. Pacific Journal of Mathematics, 9(2):415–436, jun 1959.
  • [5] Alain Connes, Joachim Cuntz, Erik Guentner, Nigel Higson, Jerome Kaminker, and John E. Roberts. Noncommutative Geometry. Springer Berlin Heidelberg, 2004.
  • [6] John Conway. A Course in Abstract Analysis. American Mathematical Society, oct 2012.
  • [7] John B. Conway. Functions of One Complex Variable. Springer US, 1973.
  • [8] John B. Conway. A Course in Functional Analysis. Springer New York, 2007.
  • [9] Robert Doran. CC^{*}-Algebras: 1943-1993: A Fifty Year Celebration: AMS Special Session Commemorating the First Fifty Years of CC^{*}-algebra Theory, volume 167. American Mathematical Society, 1994.
  • [10] Robert Doran. Characterizations of C* Algebras: the Gelfand Naimark Theorems. CRC Press, 05 2018.
  • [11] Robert Doran and James M. Fel. Representations of *-Algebras, Locally Compact Groups, and Banach *-Algebraic Bundles: Basic Representation Theory of Groups and Algebras. Academic Press, 1988.
  • [12] Ronald G. Douglas. Banach Algebra Techniques in Operator Theory. Springer New York, 1998.
  • [13] I. Gelfand and M. Naimark. On the imbedding of normed rings into the ring of operators in hilbert space. Rec. Math. [Mat. Sbornik], 12:197–213, 1943.
  • [14] Thierry Giordano, Ian F. Putnam, and Christian F. Skau. Topological orbit equivalence and cc^{*}-crossed products. Journal für die reine und angewandte Mathematik, 469:51–111, 1995.
  • [15] Erwin Kreyszig. Introductory Functional Analysis with Applications, volume 1. Wiley, New York, 1978.
  • [16] Alexander Kumjian and Jean Renault. Kms states on cc^{*}-algebras associated to expansive maps. Proceedings of the American Mathematical Society, 134:2067–2078, 2006.
  • [17] Saunders Mac Lane. Categories for the Working Mathematician. Springer New York, 1978.
  • [18] James Munkres. Topology. Prentice Hall, Upper Saddle River, NJ, second edition edition, 2000.
  • [19] Miles Reid. Undergraduate Algebraic Geometry. Cambridge University Press, dec 1988.
  • [20] Jean Renault. A Groupoid Approach to CC^{*}-Algebras. Springer Berlin Heidelberg, 1980.
  • [21] Jean Renault. CC^{*}-Algebras and Dynamical Systems. Coloquio Brasileiro de Matemáticas, IMPA, 2009.
  • [22] Joseph C. Varilly. An Introduction to Noncommutative Geometry, volume 4. European Mathematical Society, 2006.