Fact-checked by Grok 2 weeks ago

Compact operator

In , a compact operator is a bounded linear T: X \to Y between normed linear spaces X and Y such that the image T(B) of every bounded subset B \subseteq X is relatively compact in Y. This means that the closure of T(B) is compact, providing a notion of "approximate finite-dimensionality" for infinite-dimensional spaces. Compact operators generalize finite-rank operators, as every finite-rank operator is compact, and conversely, every compact operator on a can be approximated in the by a sequence of finite-rank operators. A fundamental property of compact operators is their behavior under s and s: if T is compact, then so is its T^* (when defined on s), and the of two compact operators is compact. In , the spectrum of a compact operator on an infinite-dimensional consists solely of zero and a of eigenvalues that can accumulate only at zero, with corresponding eigenspaces of finite dimension except possibly for the zero eigenvalue. This spectral structure underpins the , which characterizes solvability of equations involving compact perturbations of the identity operator. Compact operators arise naturally in applications such as operators on function spaces and operators with compact resolvents, playing a central role in the study of partial differential equations and . They form a closed two-sided ideal in the of bounded operators on a , facilitating their analysis within operator algebras.

Definitions

In topological vector spaces

A is a over the real or complex numbers endowed with a such that the vector addition and operations are continuous. This framework generalizes normed spaces by not requiring a or , allowing for a broad class of spaces where and are defined topologically rather than metrically. In a X, a B \subseteq X is bounded if for every neighborhood V of the $0 \in X, there exists a t > 0 such that B \subseteq tV. Bounded sets capture the intuitive notion of "smallness" in the topological sense, absorbed by scalar multiples of neighborhoods of zero, without reliance on a . A A of a Y is relatively compact if its \overline{A} is compact in the topology of Y. Compactness here means every open cover of \overline{A} admits a finite subcover, providing a generalization of finite-dimensional behavior to infinite-dimensional settings. A linear T: X \to Y between topological vector spaces is compact if the image T(B) of every bounded subset B \subseteq X is relatively compact in Y. This definition extends the notion of compactness from metric-induced topologies (where relative compactness equates to total boundedness) to arbitrary topological vector spaces, accommodating non-metrizable and non-complete structures while preserving the essential property that compact operators "tame" infinite-dimensional spaces by mapping large sets to nearly finite ones.

In normed spaces

In normed linear spaces, a linear operator T: X \to Y between normed spaces X and Y is defined to be compact if the image under T of every bounded subset of X is relatively compact in Y, meaning its closure is compact in the norm topology of Y. Equivalently, T is compact if the image of the closed unit ball \{ x \in X : \|x\| \leq 1 \} is precompact in Y, i.e., for every \epsilon > 0, there exists a finite \epsilon-net covering the image. This definition leverages the norm to quantify boundedness, where a set S \subseteq X is bounded if \sup_{x \in S} \|x\| < \infty, enabling precise metric-based arguments for compactness that rely on the completeness or lack thereof in Y. An alternative characterization emphasizes sequential compactness: T is compact if and only if, for every bounded sequence \{x_n\} in X, there exists a subsequence \{x_{n_k}\} such that \{T x_{n_k}\} converges in the norm of Y. This equivalence holds because the norm-induced metric allows relatively compact sets to be sequentially compact in normed spaces, bridging set-theoretic compactness with sequence behavior essential for operator analysis. The norm plays a central role here by defining convergence (\|T x_{n_k} - T x_{n_\ell}\| \to 0 as k, \ell \to \infty) and boundedness uniformly across sequences, without requiring completeness of X or Y. Compact operators are necessarily continuous (bounded), as the image of the unit ball being precompact implies it is bounded in Y, but the converse does not hold: for instance, the identity operator on an infinite-dimensional normed space is continuous yet not compact, since the unit ball is not relatively compact. This distinction underscores that compactness imposes a stronger "finite-dimensional-like" constraint via the norm topology, distinguishing it from mere continuity which only requires \|T x\| \leq M \|x\| for some M > 0.

In Banach spaces

In Banach spaces, which are complete normed spaces, the notion of a compact operator refines the definition from general normed spaces by leveraging to ensure that the image of s has . Specifically, a T: X \to Y between Banach spaces X and Y is compact if the image T(B) of any B \subset X is precompact in Y, meaning its \overline{T(B)} is compact in the norm topology. This precompactness implies total boundedness of T(B), and the of Y guarantees that the closure is a complete totally bounded set, hence compact. A key consequence of this definition in the Banach setting is that compact operators map the closed unit ball B_X = \{ x \in X : \|x\| \leq 1 \} to a set whose has empty interior in Y whenever Y is infinite-dimensional. This follows from Riesz's lemma, which states that in any infinite-dimensional normed space, a closed can be approximated by a nearly orthogonal to it; iteratively applying this lemma to a supposed open ball in \overline{T(B_X)} yields a sequence of points with pairwise distances bounded below by a positive constant, contradicting the total boundedness of T(B_X). Thus, unless T has finite-dimensional range, \overline{T(B_X)} cannot contain any open ball, highlighting how completeness enables precise control over the geometry of images. The Arzelà–Ascoli theorem provides an important analogy for understanding in specific Banach spaces, such as continuous functions on compact sets. In C(K) for compact K, a subset is relatively compact if it is bounded and , mirroring how compact operators produce images that are "uniformly approximable" in the norm; this equicontinuity condition parallels the total boundedness ensured by in general Banach spaces. Compact operators between Banach spaces are completely continuous, meaning they map weakly convergent sequences in X to norm convergent sequences in Y. This property arises because weak convergence in X implies boundedness, and the compactness of T ensures the image sequence has a norm-convergent subsequence, with the full sequence converging due to the operator's linearity and continuity.

Characterizations and Properties

Equivalent formulations

A compact operator T: X \to Y between Banach spaces X and Y can be equivalently characterized as the norm limit of a sequence of finite-rank operators. This property highlights the "finite-dimensional" nature of compact operators in infinite-dimensional settings, where the of the finite-rank operators in the topology coincides with the compact operators. To see the equivalence to the standard definition that T maps the closed unit ball of X into a precompact subset of Y, consider the forward direction: finite-rank operators map bounded sets to relatively compact sets, and the set of compact operators is closed under norm limits. For the reverse, if T(B_X) is precompact, cover it with finitely many balls of radius \epsilon/2; project onto the span of corresponding centers to obtain a finite-rank approximation within \epsilon in operator norm. Another equivalent formulation is the sequential criterion: T is compact if and only if, for every bounded sequence \{x_n\} in X, the sequence \{T x_n\} in Y admits a convergent . This follows from the fact that precompact sets in Banach spaces are sequentially precompact, so any sequence in T(B_X) has a Cauchy (hence convergent, by ) subsequence; conversely, if T(B_X) is not precompact, it contains a sequence without convergent subsequences by the failure of total boundedness. In the specific context of operators into spaces of continuous functions, such as T: X \to C(K) where K is compact and X is a , compactness of T is equivalent to the image T(B_X) being pointwise bounded and on K, by the Arzelà–Ascoli theorem. The proof leverages the theorem's characterization of relatively compact subsets of C(K): ensures uniform control, and pointwise boundedness with the unit ball's boundedness yields relative compactness. These formulations unify across normed and Banach spaces, extending to topological vector spaces where total boundedness is replaced by appropriate precompactness notions, though is key for sequential .

Algebraic and topological properties

Compact operators on normed linear spaces are inherently linear maps that preserve the structure. Moreover, every compact operator is bounded, meaning it maps bounded sets to bounded sets and is continuous with respect to the norm topology. This boundedness follows directly from the definition, as the image of the closed unit under a compact operator is relatively compact and hence bounded. The collection of compact operators from a X to another Y, denoted K(X, Y), forms a of the space of s B(X, Y). Consequently, the sum of finitely many compact operators is compact, and scalar multiples of a compact operator remain compact. Additionally, K(X, Y) is a two-sided in B(X, Y): if T \in K(X, Y) and S \in B(Z, X), U \in B(Y, W) for s Z, W, then U T S \in K(Z, W). This property arises because the of a with a compact operator yields a compact operator, as the relatively compact image under the compact part remains relatively compact after bounded pre- and post-. In the context of dual spaces, the adjoint of a compact operator is also compact. Specifically, if T: X \to Y is compact between Banach spaces, then its T^*: Y^* \to X^* is compact; this is known as Schauder's theorem. The proof relies on the fact that the adjoint maps the unit ball of Y^* to a set whose weak* closure is compact in X^*, leveraging the Alaoglu theorem. For compact operators with separable ranges, a key topological factorization property holds: such an operator T: X \to Y factors through the space of continuous functions on a . That is, there exists a K, a A: X \to C(K), and a B: C(K) \to Y such that T = B A, where the range of A is dense in a separable of C(K). This factorization underscores the "finite-dimensional-like" topological behavior of compact operators, embedding their action into the structure of functions on compact sets.

Spectral properties

The spectral properties of compact operators are fundamentally , distinguishing them from more general bounded operators on infinite-dimensional s. For a compact operator T on a X, the \sigma(T) consists of 0 together with a set of eigenvalues, where each non-zero eigenvalue \lambda \neq 0 has finite geometric multiplicity, meaning \dim \ker(T - \lambda I) < \infty. Moreover, the non-zero part of the \sigma(T) \setminus \{0\} can only accumulate at 0, ensuring that for any \varepsilon > 0, there are only finitely many eigenvalues with |\lambda| \geq \varepsilon. This discreteness arises from the Riesz–Schauder theorem, which characterizes the of compact operators as \sigma(T) = \{0\} \cup \sigma_p(T), where \sigma_p(T) denotes the point (eigenvalues). A key consequence is the for compact operators: for any \lambda \neq 0, the operator T - \lambda I is either invertible or possesses a finite-dimensional , with the being closed and of finite if the is non-trivial. In the latter case, the \ind(T - \lambda I) = \dim \ker(T - \lambda I) - \codim \ran(T - \lambda I) is finite, reflecting the finite-dimensional nature of the eigenspaces. If the underlying space X is infinite-dimensional, then 0 belongs to the \sigma(T), serving as the sole possible of eigenvalues. Furthermore, if the of T is infinite-dimensional, the operator admits infinitely many distinct non-zero eigenvalues, which must accumulate solely at 0 due to compactness. The r(T) = \sup \{ |\lambda| : \lambda \in \sigma(T) \} of a compact operator T equals the \lim_{n \to \infty} \|T^n\|^{1/n}, a general formula for bounded operators that simplifies here to the supremum of the values of its eigenvalues, since the beyond 0 is purely point spectral. This radius is finite and attained as the largest eigenvalue modulus if the is non-trivial, or zero if T is quasinilpotent (i.e., all eigenvalues vanish except possibly at 0). These properties underpin the utility of compact operators in and approximation theory, where the eigenvalues provide a countable basis for understanding the operator's action.

Historical Development

Origins in integral equation theory

The origins of compact operators trace back to the early 20th-century study of integral equations, particularly those motivated by physical problems in potential theory and heat conduction. In 1903, Ivar Fredholm published a seminal paper introducing a determinant method to solve integral equations of the second kind, expressed as \phi(s) = f(s) + \lambda \int_a^b K(s,t) \phi(t) \, dt, where the kernel K(s,t) was approximated by finite-rank expansions to ensure solvability. This approach highlighted the role of finite-rank kernels, akin to those in Volterra equations, in establishing existence and uniqueness results, laying the groundwork for operators whose images could be approximated by finite-dimensional subspaces. David Hilbert significantly advanced this framework between 1904 and 1910 through a series of six papers, culminating in his 1912 treatise Grundzüge einer allgemeinen Theorie der linearen Integralgleichungen, where he focused on equations with symmetric continuous kernels. Hilbert demonstrated that solutions could be expanded in series of eigenfunctions, with eigenvalues forming a sequence converging to zero, for what he termed "completely continuous" transformations—now recognized as compact operators—enabling the resolution of Fredholm's alternative in infinite dimensions. These operators mapped bounded sets to precompact ones, mirroring finite-dimensional behavior while addressing the challenges of infinite-dimensional function spaces. Around 1906, Hilbert introduced the concept of compact operators in the functional analysis context, emphasizing their spectral properties for symmetric operators. This marked a pivotal shift from concrete formulations to abstract in L^2 spaces, where Hilbert employed orthogonal expansions to represent solutions, bridging equations with emerging structures. Such operators, like those with square-integrable kernels, exemplify early compact examples, though their detailed forms are explored elsewhere.

Evolution in functional analysis

The abstraction of compact operators from concrete integral formulations to general topological settings emerged in the early 20th century, building briefly on David Hilbert's foundational work with integral equations around 1904–1910. This shift emphasized properties like approximation by finite-rank operators and spectral behavior in infinite-dimensional spaces. In the 1910s and 1920s, Maurice Fréchet and Hugo Steinhaus laid early topological groundwork for these concepts. Fréchet developed notions of compactness in metric spaces through sequential criteria and finite ε-net coverings, providing characterizations essential for defining operator compactness in emerging functional-analytic frameworks. Steinhaus advanced related ideas via studies on dual spaces, such as identifying the dual of L^1[a,b] as L^\infty[a,b], which supported boundedness and continuity properties of linear operators in normed settings. A pivotal advancement came in 1918 with Frigyes Riesz's theorem on compact operators in Hilbert spaces, proving that such operators—termed "vollständig stetig" or completely continuous—can be uniformly approximated by finite-rank operators. Riesz's spectral theory further established that the non-zero part of the spectrum consists solely of eigenvalues with finite multiplicity, accumulating only at zero, thus generalizing Fredholm's earlier results on integral equations to abstract operator settings. Stefan Banach's 1932 monograph Théorie des opérations linéaires crystallized these developments by formally defining compact operators on Banach spaces as bounded linear maps sending the unit ball to a relatively compact set. Banach explicitly connected this definition to Riesz's results and Juliusz Schauder's (1930), which implied non-empty spectra for compact operators, thereby embedding compactness into the axiomatic structure of complete normed spaces. Developments in the 1930s, led by Marshall Stone, integrated compact operators more deeply into , extending Riesz's results to unbounded operators and axiomatic formulations that unified Hilbert and analyses. Stone's frameworks addressed resolvent properties and spectral measures, paving the way for modern generalizations like nuclear operators and applications in partial differential equations.

Compact Operators on Hilbert Spaces

Specific definitions and features

In Hilbert spaces, a bounded linear T: H \to H is compact if it maps the closed unit ball of H to a precompact subset of H. This property leverages the inner product structure, ensuring that the image has compact closure. Equivalently, in Hilbert spaces, compact operators are completely continuous, meaning they map weakly convergent sequences in H to strongly convergent sequences in H. For compact operators, the provides a with respect to an . Specifically, if T: H \to H is compact and , then H admits an \{e_i\}_{i=1}^\infty of eigenvectors such that T e_i = \lambda_i e_i for real eigenvalues \lambda_i satisfying |\lambda_1| \geq |\lambda_2| \geq \cdots and \lambda_i \to 0 as i \to \infty. Eigenspaces corresponding to distinct nonzero eigenvalues are orthogonal, and those for nonzero eigenvalues are finite-dimensional. Thus, T can be expressed as T u = \sum_{i=1}^\infty \lambda_i \langle u, e_i \rangle e_i for all u \in H. More generally, every compact operator T: H \to H on a admits a singular value decomposition T = U \Sigma V^*, where U and V are unitary s (or partial isometries extending to unitaries on the orthogonal complements of their kernels), and \Sigma is a diagonal operator with nonnegative singular values \sigma_1 \geq \sigma_2 \geq \cdots \geq 0 satisfying \sigma_n \to 0 as n \to \infty. The singular values are the square roots of the eigenvalues of T^* T, and there exist orthonormal bases \{v_n\} in H and \{u_n\} in H such that T v_n = \sigma_n u_n and T^* u_n = \sigma_n v_n. This decomposition allows representation as T x = \sum_{n=1}^\infty \sigma_n \langle x, v_n \rangle u_n for x \in H. The orthonormal bases in the singular value decomposition facilitate approximations by finite-rank operators that preserve orthogonality, as truncating the series to the first N terms yields a rank-N operator acting on the span of \{v_1, \dots, v_N\} and mapping to the orthogonal span of \{u_1, \dots, u_N\}. Such approximations converge in the to T, maintaining the inner product structure inherent to the .

Approximation theorems

In Hilbert spaces, a bounded linear operator T: H \to H is compact if and only if it is the uniform limit of a sequence of finite-rank operators. This characterization highlights the "finite-dimensional" nature of compact operators, allowing them to be approximated arbitrarily closely in the by operators of finite rank. For compact operators, this approximation can be achieved using finite-rank projections derived from the . Specifically, if T is compact and , the provides an \{e_n\} of eigenvectors with corresponding eigenvalues \lambda_n \to 0, and the partial sums T_n = \sum_{k=1}^n \lambda_k |e_k\rangle\langle e_k| are finite-rank orthogonal projections scaled by the eigenvalues, satisfying \|T - T_n\| \to 0 as n \to \infty. A key consequence of compactness in Hilbert spaces is the existence of an adapted to the operator. For any compact operator T: H \to H, there exists an \{e_n\} of H such that \|T e_n\| \to 0 as n \to \infty. This property follows from the fact that compactness implies \|T x_n\| \to 0 for every orthonormal \{x_n\} in H, and one can construct such a basis by selecting subsequences where the images converge appropriately. Conversely, if \|T e_n\| \to 0 for some \{e_n\}, then T maps the unit ball to a , confirming compactness. Picard's iteration provides a practical for solving equations of the form (I - T)x = y where T is a compact operator on a H and y \in H. Assuming \|T\| < 1, the iterates defined by x_0 = y and x_{n+1} = T x_n + y converge to the unique solution x = (I - T)^{-1} y in the norm topology. This successive approximation scheme is particularly useful for integral equations with compact kernels, as the contraction property ensures linear convergence near the solution, and the method extends to cases where \|T\| \geq 1 via analytic continuation or perturbation when the spectral radius of T is less than 1. Error estimates for finite-rank approximations of compact operators rely on the singular value decomposition (SVD), which, as discussed previously, expresses T as T = \sum_{n=1}^\infty \sigma_n |u_n\rangle \langle v_n| with singular values \sigma_n \downarrow 0. The optimal rank-n approximation is the partial sum T_n = \sum_{k=1}^n \sigma_k |u_k\rangle \langle v_k|, and the approximation error satisfies \|T - T_n\| = \sigma_{n+1}. This bound quantifies how rapidly finite-rank operators can approximate T, with the decay rate of \sigma_n determining the efficiency; for example, if \sigma_n = O(n^{-p}) for p > 0, the error decreases polynomially.

Completely continuous operators

A completely continuous operator is a bounded linear operator between normed spaces that maps every weakly convergent sequence to a norm convergent sequence. This property emphasizes the operator's ability to "strengthen" convergence from the weak topology to the norm topology. In Banach spaces, every compact operator is completely continuous, as the image of the unit ball under a compact operator is relatively compact, implying the desired convergence property for sequences. The converse holds in reflexive Banach spaces, including Hilbert spaces, where completely continuous operators coincide with compact operators; the concept of completely continuous operators was introduced by in 1918, building on Hilbert's work on integral equations, and extended to general Banach spaces by in 1932, though the full equivalence requires reflexivity. For example, in the non-reflexive Banach space \ell_1, the identity operator is completely continuous due to the Schur property (weak convergence implies norm convergence) but is not compact, as its image of the unit ball is not relatively compact. Historically, the term "completely continuous" (translating from the German vollständig stetig) was used in early 20th-century literature on integral operators, following Hilbert's foundational work on . later generalized the concept to operators between arbitrary Banach spaces in his 1932 monograph Théorie des opérations linéaires. In modern , "compact operator" has become the standard terminology, as it more directly captures the precompactness of images of bounded sets and avoids the historical association with specific sequence convergence properties. In non-Banach settings, such as incomplete normed spaces, the notions of completely continuous and compact operators do not necessarily coincide, with counterexamples arising from the lack of affecting relative compactness in the range space. For instance, certain inclusion operators into incomplete subspaces can satisfy the sequence convergence condition but fail to map bounded sets to relatively compact ones due to missing limit points in the space.

Connections to Fredholm operators

Fredholm operators are bounded linear operators between Banach spaces that possess finite-dimensional kernels and closed ranges with finite-dimensional cokernels. The index of a Fredholm operator T, denoted \operatorname{ind}(T), is defined as the difference between the dimensions of its kernel and cokernel, \operatorname{ind}(T) = \dim \ker(T) - \dim \coker(T). Compact operators form a significant subclass within this framework: on infinite-dimensional Banach spaces, every compact operator is Fredholm with index zero, meaning \dim \ker(K) = \dim \coker(K) < \infty for any compact K. This property arises because the image of the unit ball under a compact operator is precompact, ensuring the finite dimensionality of the relevant subspaces. A key consequence is the , which characterizes the solvability of equations involving the plus a compact operator. Specifically, for a compact operator K on a , the operator I + K is invertible if and only if -1 is not an eigenvalue of K, or equivalently, if \ker(I + K) = \{0\}; in this case, I + K is bijective with a bounded inverse. This extends the classical alternative from finite-dimensional linear algebra to infinite dimensions and underscores the invertibility stability near the for compact perturbations. More broadly, the Fredholm property is stable under compact perturbations: if T is Fredholm, then T + K is also Fredholm for any compact K, and \operatorname{ind}(T + K) = \operatorname{ind}(T). This index invariance is crucial for perturbation analyses in operator theory. In applications to partial differential equations, compact operators play a pivotal role through the resolvents of . For a second-order L on a bounded with suitable conditions, the resolvent (L - \lambda I)^{-1} is compact on L^2(\Omega) for \lambda in the , making L - \lambda I a compact of the and thus Fredholm of zero. This compactness ensures a for , with eigenvalues accumulating only at infinity, and enables the to determine solvability: solutions to Lu = f exist uniquely if f is orthogonal to the of the , or more generally, under conditions for nonhomogeneous terms. Such structures are foundational in proving existence and regularity results for elliptic value problems.

Examples

Finite-dimensional and finite-rank operators

All bounded linear operators between finite-dimensional normed spaces are compact, since the closed unit ball in a finite-dimensional space is compact, and the image of a bounded set under such an operator remains bounded and thus compact. This follows directly from the fact that in finite dimensions, boundedness implies compactness for subsets. A finite-rank operator T: X \to Y between normed spaces is a bounded linear operator whose range is finite-dimensional. Such operators are compact because the image of the closed unit ball under T is bounded in a finite-dimensional space, hence relatively compact. Equivalently, for any bounded sequence in the domain, the image sequence has a convergent subsequence in the finite-dimensional range. Representative examples include orthogonal projections onto finite-dimensional subspaces of a , which are finite-rank and thus compact. Another example is any acting on \mathbb{R}^n or \mathbb{C}^n, as these map between finite-dimensional spaces and are therefore compact. On separable , finite-rank are dense in the space of compact with respect to the ; that is, every compact is the norm limit of a sequence of finite-rank . This density result characterizes compact on such spaces and underpins their approximation properties.

Integral and differential operators

Integral operators provide fundamental examples of compact operators in infinite-dimensional spaces. A prominent class is the Hilbert-Schmidt operators, which act on the L^2(\Omega) for a \Omega via the formula (Tf)(x) = \int_{\Omega} k(x,y) f(y) \, d\mu(y), where the kernel k \in L^2(\Omega \times \Omega). Such operators are compact because their allows approximation by finite-rank operators, with the singular values forming a square-summable sequence that decays to zero. The Volterra operator offers another concrete example of compactness, defined on the space of continuous functions C[0,1] by (Vf)(x) = \int_0^x k(x,t) f(t) \, dt, where k is continuous on the $0 \leq t \leq x \leq 1. Despite the kernel being continuous (hence bounded), V is compact on C[0,1] equipped with the supremum norm, as the image of the unit ball is equicontinuous and uniformly bounded, hence precompact by the Arzelà-Ascoli theorem. Multiplication operators provide examples in discrete settings. In the sequence space \ell^2, the diagonal operator (Mf)_n = m_n f_n is compact if and only if m_n \to 0 as n \to \infty. A classic non-compact operator is the identity on \ell^2, whose image of the unit ball consists of all sequences with \ell^2-norm at most 1, which is not precompact since it contains the orthonormal basis vectors with no convergent subsequence. Differential operators, such as the d/dx realized as a from the H^1(\mathbb{R}) to L^2(\mathbb{R}), are typically non-compact on unbounded domains because they map bounded sets to sets lacking or uniform boundedness in higher norms, failing to produce precompact images.

References

  1. [1]
    [PDF] Chapter 9. Compact Operators
    May 16, 2017 · An operator T ∈ B(X, Y ) is a compact operator if, for all bounded sets B ⊆ X, the set T(B) is relatively compact in Y .
  2. [2]
    [PDF] Compact operators - Purdue Math
    Mar 13, 2025 · Definition. A bounded linear operator A on a Hilbert space H is compact if it has any of the following properties: (1) There exists a sequence ...
  3. [3]
    Lecture 20: Compact Operators and the Spectrum of a Bounded ...
    Description: We show that compact operators are precisely limits of finite-rank operators. Then, we define invertible linear operators and begin exploring ...
  4. [4]
    [PDF] On compact operators - Alen Alexanderian
    Aug 26, 2024 · T is called a compact operator if for all bounded sets E ⊆ X, T(E) is relatively compact in Y . By Definition 2.4, if E ⊂ X is a bounded set, ...
  5. [5]
    [PDF] Chapter 9. Compact Operators
    May 23, 2017 · Introduction and Basic Definitions. Relatively compact set, necessary and sufficient conditions for a set to be relatively compact (Lemma. 9.1.A) ...<|control11|><|separator|>
  6. [6]
    [PDF] Functional Analysis, Math 7320 Lecture Notes from March 09, 2017
    Mar 9, 2017 · To address this question we introduce the concept of a compact operator, which will serve as a natural generalization of a finite-rank operator ...
  7. [7]
    [PDF] 5.24. - compact operators 379 - LSU Math
    Every linear operator defined on a finite-dimensional normed linear space is compact. EXAMPLE 3. Consider the multiplication operator. F: x(t) →ƒ(t)x(t).
  8. [8]
  9. [9]
    Topology | SpringerLink
    Jan 26, 2011 · Many topological vector spaces, in particular ... compact operator if it is continuous and maps bounded sets to relatively compact sets.
  10. [10]
    None
    ### Summary of Compact Operators from http://www.mostefanadir.com/Integral%20Equations_files/2%20Compact%20Operators.pdf
  11. [11]
    [PDF] Lecture 08: Compact Sets. Compact Operators.
    Definition 3. Let X and Y be normed space over K (K = R or K = C). The operator A : X ж Y is called a compact operator iff. 1. A is continuous, and. 2. A ...
  12. [12]
    [PDF] Characterization of Compact Operators Abstract 1 Introduction
    Keywords: Compact Operator, Normed linear spaces, Strong ... ii) Every bounded sequence in a normed linear space of finite dimension has a subsequence.
  13. [13]
    [PDF] Fredholm-Riesz 1. Compact operators on Banach spaces
    A continuous linear operator T : X → Y on Banach spaces is compact when T maps bounded sets in X to pre-compact sets in Y , that is, sets with compact closure. ...
  14. [14]
    [PDF] VI.5 Compact operators 199 - LSU Math
    Definition Let X and Y be Banach spaces. An operator Te L(X, Y) is called compact (or completely continuous) if T takes bounded sets in X into precompact sets ...
  15. [15]
    [PDF] 18.102 S2021 Lecture 20. Compact Operators and the Spectrum of ...
    Last lecture, we introduced the concept of a compact operator: an operator A ∈ B(H) (recall that H always denotes a Hilbert space) is compact if K({||u|| ≤ 1}) ...
  16. [16]
    [PDF] Compact operators on Banach spaces - Jordan Bell
    Nov 12, 2017 · In this note I prove several things about compact linear operators from one. Banach space to another, especially from a Banach space to ...Missing: relatively empty interior
  17. [17]
    [PDF] Banach Spaces - UC Davis Math
    We leave the proof of the following properties of compact operators as an exercise. ... The converse is also true for compact operators on many Banach spaces, ...
  18. [18]
    [PDF] Essays in analysis Compact operators
    Apr 16, 2020 · All Hilbert spaces in this essay are assumed to be separable, which means they possess countable orthonormal bases. 1. The singular value ...
  19. [19]
    [PDF] Lecture 6 - Functional analysis
    4 Compact operators and their spectral properties. In this section, we generally consider Banach spaces over either real or complex numbers, however, there ...
  20. [20]
    [PDF] Appendix A Compact Operators
    Given two Banach spaces X, Y we denote by KpX, Y q the set of compact ... In this section we discuss the spectral properties of a compact operator. We ...<|control11|><|separator|>
  21. [21]
    [PDF] Bounded and compact operators. Spectral theorem.
    Nov 8, 2023 · Bounded and compact operators. Spectral theorem. Page 17. 2.3 Spectral Radius ... Therefore, the operator T is compact as the limit of compact ...
  22. [22]
    [PDF] FREDHOLM, HILBERT, SCHMIDT Three Fundamental Papers on ...
    Dec 15, 2011 · From this work emerged four general forms of integral equations now called Volterra and Fredholm equations of the first and second kinds (a ...
  23. [23]
    [PDF] On the origin and early history of functional analysis - DiVA portal
    ... Hilbert's faith and enthusiasm in the methods invented by Fredholm. During the years 1904–1906, Hilbert published six papers on integral equations which.
  24. [24]
  25. [25]
    None
    Below is a merged and comprehensive summary of the historical development of compact operators in functional analysis, integrating all information from the provided segments. To retain maximum detail, I will use a structured table format for key contributors, their contributions, theorems, definitions, references, and URLs, followed by a narrative summary for additional context and generalizations. This approach ensures all details are preserved while maintaining clarity and density.
  26. [26]
  27. [27]
    The 100th Jubilee of Riesz Theory - EMS Press
    From now on, we will employ the attributes 'compact' and 'completely continuous' in this way, which has become standard. Let L(X, Y) denote the Banach spaces of ...
  28. [28]
    [PDF] Théorie des opérations linéaires
    49. S. Banach. Théorie des opérations linéaires. 1. Page 6. 2. Introduction. Si la suite {x(t)} de fonctions à p-ième puissance sommable. (p1) converge presque ...
  29. [29]
    Mackey on Stone - Celebratio Mathematica
    Stone had worked on eigenfunction expansions (i.e., a form of spectral theory) for ordinary differential operators in his dissertation and in his early work up ...Missing: post- | Show results with:post-
  30. [30]
    [PDF] Compact Operators on Hilbert Space
    Feb 18, 2012 · We only need the Cauchy-. Schwarz-Bunyakowsky inequality and the definition of self-adjoint compact operator. ... topological vector space is ...
  31. [31]
    [PDF] 09b. Compact operators
    Apr 3, 2019 · of compact operators on Hilbert spaces are not shared by compact operators on Banach spaces. ... topological vector spaces, and is best delayed.
  32. [32]
    [PDF] Compact Operators in Hilbert Space - UW Math Department
    Compact Operators in Hilbert Space. Hart Smith ... Singular Value Decomposition Theorem. If T ∈ B(H) is compact with singular values {ρk }, there exists.
  33. [33]
    [PDF] Ordinary Differential Equations and Dynamical Systems
    ... Picard iteration. Un- fortunately, it is not suitable for actually finding ... compact operator is again compact (Problem 5.9). In combination with ...
  34. [34]
    [PDF] On the foundations of Completely Continuous, Compact and ...
    Dec 10, 2019 · compact operators. The Italian mathematician.Riesz (1918) who had ... Kothe, Topological vector spaces, I, II New York Heidelberg Berlin.
  35. [35]
    [PDF] Continuous operators on Hilbert spaces
    A continuous linear operator is of finite rank if its image is finite-dimensional. A finite-rank operator is compact, since all balls are pre-compact in a ...
  36. [36]
    [PDF] A Short History of Operator Theory - NYU Stern
    In 1916 Riesz created the theory of what he called "completely continuous" operators, now more familiarly compact operators. Since he wrote this in ...
  37. [37]
    Completely Continuous Operators on Banach Spaces - ResearchGate
    In this chapter we define the notion of a completely continuous operator from a Banach space to another Banach space and we present some simple properties ...
  38. [38]
    Compact operators and completely continuous operators
    Sep 4, 2012 · A compact operator between Banach spaces is an operator that maps bounded sets into relatively compact sets, while a completely continuous operator maps all ...Missing: normed | Show results with:normed
  39. [39]
    Completely Continuous Operators on Banach Spaces - SpringerLink
    In this chapter we define the notion of a completely continuous operator from a Banach space to another Banach space and we present some simple properties ...<|control11|><|separator|>
  40. [40]
    [PDF] Perturbation Theory
    Page 1. C L A S S I C S I N M A T H E M AT I C S. Tosio Kato. Perturbation Theory for Linear Operators ... There are a subject index, an author index and a ...
  41. [41]
    [PDF] Fredholm Theory - Yale Math
    Apr 25, 2018 · An alternate definition of compact operators is that the image of the unit ball is pre-compact. This implies that if fn is a bounded ...
  42. [42]
    [PDF] Chapter 4: Elliptic PDEs - UC Davis Math
    The operator (L − λI)−1 is called the resolvent of L, so this property is some- times expressed by saying that L has compact resolvent. As discussed in ...
  43. [43]
    [PDF] Compact Linear Operators - UW Math Department
    Definition. For X,Y Banach spaces, an operator T ∈ B(X,Y) is compact if the image of the unit ball of X has compact closure in Y. T compact implies that T(E) ...
  44. [44]
    [PDF] Compact Sets and Compact Operators 1 Compact and Precompact ...
    We start with the finite-rank operators. If the range of a bounded operator K is finite dimensional, then we say that K is a finite-rank operator. Proposition ...
  45. [45]
    [PDF] Hilbert spaces
    An operator K ∈ B(H), bounded on a separable Hilbert space, is compact if and only if it is the limit of a norm-convergent sequence of finite rank operators.
  46. [46]
    [PDF] 3 Compact operators on Hilbert space
    Let T be an integral operator on L2(E) with kernel K(x,y). Assume that K ∈ L2(E2). Then T is a bounded operator with ∥T ∥ ≤ ∥K∥L2(E2). Moreover, T is compact.
  47. [47]
    [PDF] THE VOLTERRA OPERATOR Let V be the indefinite integral ...
    By Arzela-Ascoli, V : Lp[0,1] → C[0,1] is compact. The preceeding argument does not go through when V acts on L1[0,1]. In this case equicontinuity fails, as is ...
  48. [48]
    [PDF] Operator theory on Hilbert spaces
    We start by considering multiplication operators on the Hilbert space L2(Rd). For any measurable complex function ϕ on Rd let us define the multiplication op-.
  49. [49]
    [PDF] 5 Compact linear operators - math.uzh.ch
    We note that every compact operator T is bounded. ... Let (Tn)n≥1 be a sequence of compact linear operators from a normed space X into a Banach spaces Y .
  50. [50]
    [PDF] Functional Analysis, Sobolev Spaces and Partial Differential Equations
    ... Compact Operator . . . . . . . . . . . . . . . . . . . . . . . . . . 162. 6.4 ... topological vector spaces. The second geometric form (Theorem 1.7) ...<|control11|><|separator|>