Fact-checked by Grok 2 weeks ago

Krein–Milman theorem

The Krein–Milman theorem is a cornerstone result in functional analysis, stating that every nonempty compact convex subset of a Hausdorff locally convex topological vector space is equal to the closed convex hull of its extreme points. This theorem provides a geometric characterization of such sets, emphasizing the role of extreme points—those that cannot be expressed as nontrivial convex combinations of other points in the set—in determining the entire structure. Proved by Ukrainian mathematicians Mark Krein and David Milman in 1940, the theorem was originally published in Studia Mathematica under the title "On Extreme Points of Regular Convex Sets." Their work built on earlier ideas, such as Hermann Minkowski's 1911 results on finite-dimensional convex bodies, but extended the principle to infinite-dimensional spaces where and convexity interact in more subtle ways. The proof relies on the Hahn–Banach separation theorem to ensure the existence of extreme points and their density in the , assuming the separates points. The theorem's significance lies in its applications across and beyond, enabling the representation of points in compact convex sets as barycenters of measures supported on extreme points—a key tool in Choquet theory. Notable uses include Louis de Branges' proof of the via extreme points of function algebras, Joram Lindenstrauss' demonstration of Lyapunov's convexity theorem for vector measures, and representations of completely monotone functions as Laplace transforms of measures on [0, ∞). It also underpins generalizations in non-commutative settings, such as operator convexity and matrix convex sets, highlighting its enduring influence in modern mathematics.

Background Concepts

Extreme Points

In a convex set K subset of a real vector space, a point x \in K is called an extreme point if it cannot be expressed as a nontrivial convex combination of distinct points in K; that is, whenever x = t y + (1-t) z with y, z \in K and $0 < t < 1, it follows that y = z = x. This definition captures points that lie at the "corners" of the set, resisting decomposition into averages of other elements within K. Equivalently, x is extreme if it does not belong to the relative interior of any line segment contained entirely in K. In finite-dimensional spaces, extreme points correspond intuitively to the vertices of polytopes. For example, consider the standard simplex in \mathbb{R}^n, where the extreme points are the standard basis vectors e_i (with one coordinate 1 and the rest 0). A more intricate illustration arises in the of n \times n doubly stochastic matrices, whose extreme points are precisely the . These examples highlight how extreme points form the skeletal structure supporting the entire convex set. In infinite-dimensional settings, extreme points exhibit richer behavior. For the closed unit ball in the \ell^1 space over the natural numbers, the extreme points consist of sequences with exactly one entry equal to \pm 1 and all others zero. This contrasts with finite dimensions, where such sets are polytopal, underscoring the potential for uncountably many extreme points or even their absence in non-compact cases. The collection of extreme points of K, denoted \operatorname{ext}(K), possesses notable properties, including closure under certain affine restrictions and mappings that preserve convexity. In finite dimensions, \mathbb{R}^n, every nonempty compact convex set equals the convex hull of its extreme points—a result serving as a precursor to broader theorems in convex analysis. This finite-dimensional fact motivates investigations into when \operatorname{ext}(K) is nonempty, particularly for compact convex sets in more general spaces, though algebraic considerations alone do not guarantee existence without additional structure.

Convex Sets in Topological Vector Spaces

A topological vector space (TVS) is a vector space over the real or complex numbers equipped with a topology such that the maps (x, y) \mapsto x + y and (\lambda, x) \mapsto \lambda x (for scalars \lambda) are continuous. This topological structure ensures that convergence in the space aligns with algebraic operations, forming the foundation for studying continuity of linear functionals and operators. In such spaces, neighborhoods of the zero vector play a crucial role in defining the topology. A TVS is locally convex if its topology admits a base of neighborhoods of the origin consisting of convex, absorbing, and balanced sets. Equivalently, the topology can be generated by a family of seminorms \{p_i\}_{i \in I}, where each p_i: X \to [0, \infty)&#36; is a sublinear functional satisfying p_i(\lambda x) = |\lambda| p_i(x)for scalars\lambda$, and local convexity follows from the Minkowski functional associated to convex absorbing sets. Absorbing sets contain scalar multiples of every vector for sufficiently small scalars, balanced sets are symmetric under multiplication by scalars of modulus at most 1, and these properties ensure the space supports a rich theory of convex analysis. In Hausdorff locally convex TVS, disjoint convex sets—one compact and the other closed—can be strictly separated by a continuous linear functional, leveraging the . Convex sets in a TVS are subsets closed under convex combinations, i.e., for x_1, \dots, x_n \in C and \lambda_i \geq 0 with \sum \lambda_i = 1, the point \sum \lambda_i x_i \in C. Compact convex sets in these spaces are particularly significant, as they exhibit strong topological stability. The closed convex hull of a set A, denoted \overline{\mathrm{co}}(A), is the smallest closed convex set containing A, obtained as the closure of the algebraic convex hull \mathrm{co}(A). In infinite-dimensional locally convex spaces, \mathrm{co}(A) may fail to be closed, necessitating the topological closure for applications involving extremal structures. In a Hausdorff locally convex TVS, every compact set is closed and bounded, where boundedness means the set is contained in \lambda U for some neighborhood U of zero and scalar \lambda > 0; however, the converse fails in infinite dimensions, unlike in finite-dimensional spaces.

Statement of the Theorem

Classical Krein–Milman Theorem

The classical Krein–Milman theorem asserts that if X is a locally Hausdorff and K \subset X is a nonempty compact set, then K = \overline{\mathrm{co}}(\mathrm{ext} K), where \mathrm{ext} K denotes the set of extreme points of K and \overline{\mathrm{co}}(\mathrm{ext} K) is the closed hull of \mathrm{ext} K. This formulation relies on the local convexity of the topology, which ensures the existence of sufficiently many continuous linear functionals to separate points, and the Hausdorff property, which guarantees that compact sets are closed. The compactness of K plays a crucial role: it implies that \mathrm{ext} K is nonempty (by the axiom of choice, via Zorn's lemma applied to the partially ordered set of subsets without extreme points) and that the closed hull of \mathrm{ext} K is compact, hence closed within K. A key immediate implication is the existence of extreme points in any nonempty compact convex set in such spaces, resolving a foundational question in convex analysis. Furthermore, since the closed convex hull consists of limits of convex combinations of finitely many points from \mathrm{ext} K, every element of K can be approximated arbitrarily closely by such finite convex combinations of extreme points. This finite-dimensional approximation property underscores the theorem's utility in reducing infinite-dimensional problems to extremal structures.

Strong Krein–Milman Theorem

The strong Krein–Milman theorem provides a version of the classical result that extends its applicability by considering the in locally Hausdorff topological vector spaces. Specifically, if K is a nonempty subset of such a space that is compact in the (i.e., weakly compact), then K equals the weak closure of the of its extreme points: K = \overline{\mathrm{co}(\mathrm{ext}(K))}^w, where \mathrm{ext}(K) denotes the set of extreme points of K, \mathrm{co} the , and the overline with superscript w the closure in the . This formulation ensures that the theorem holds without requiring compactness in a stronger , such as the . A key distinction from the classical Krein–Milman theorem, which assumes in the original (often the in Banach spaces), is that weak is a milder condition. In Banach spaces, the is coarser than the , so sets that are not norm-compact can still be weakly compact, broadening the theorem's scope to important examples like bounded closed sets in infinite-dimensional spaces. This makes the strong version particularly relevant for studying duality and reflexivity, where plays a central role. In reflexive Banach spaces, the Eberlein–Šmulian theorem guarantees that every bounded closed convex set is weakly compact, directly linking such sets to the strong Krein–Milman theorem. For instance, consider the real Hilbert space \ell^2, which is reflexive. Its closed unit ball B = \{ x \in \ell^2 : \|x\|_2 \leq 1 \} is weakly compact, and the extreme points of B are exactly the unit vectors \{ x \in \ell^2 : \|x\|_2 = 1 \}. The weak closure of the convex hull of these extreme points recovers the entire ball B. In the complex case, the extreme points include unit vectors up to multiplication by a complex phase e^{i\theta}, but the result holds analogously.

Proof and Analysis

Outline of the Proof

The proof of the classical Krein–Milman theorem, which states that a nonempty compact K of a Hausdorff locally is the closed of its extreme points, divides into two primary steps: demonstrating the nonemptiness of the set of extreme points \operatorname{ext}(K), and verifying that K = \overline{\operatorname{co}}(\operatorname{ext}(K)). The argument relies on the via , separation properties from the , and the topological features of compactness and local convexity. To establish that \operatorname{ext}(K) is nonempty, consider the family of all nonempty closed proper faces of K, where a face F \subseteq K is a subset such that whenever a of points from K lies in F, both endpoints also lie in F. This family is nonempty because, for distinct x, y \in K, the guarantees a continuous linear functional \ell \in X^* separating x and y (i.e., \ell(x) > \ell(y)), and the set F_\ell = \{z \in K : \ell(z) = \max_{w \in K} \ell(w)\} forms a proper closed face. Order these faces by reverse inclusion (larger faces smaller in the order). By , there exists a minimal element F in this poset. ensures F is compact, and minimality implies F cannot contain a nontrivial , so F is a singleton \{p\} with p \in \operatorname{ext}(K). The Hausdorff local convexity of the space ensures that such separating hyperplanes exist robustly, preventing pathological separations. The second step shows \overline{\operatorname{co}}(\operatorname{ext}(K)) = K by proving both inclusions, with the nontrivial direction K \subseteq \overline{\operatorname{co}}(\operatorname{ext}(K)) via contradiction. Assume there exists x \in K \setminus \overline{\operatorname{co}}(\operatorname{ext}(K)). The Hahn–Banach separation theorem, applicable due to local convexity, yields a continuous linear functional f \in X^* such that f(x) > \sup \{f(y) : y \in \overline{\operatorname{co}}(\operatorname{ext}(K))\}. Define the supporting face F = \{z \in K : f(z) = f(x)\}, which is nonempty, closed, convex, and compact. By the first step, F contains an extreme point p \in \operatorname{ext}(K), but then f(p) = f(x) > \sup f(\operatorname{ext}(K)), a contradiction. Thus, every x \in K lies in \overline{\operatorname{co}}(\operatorname{ext}(K)). This closure argument leverages the weak* topology in dual space contexts, where for x \in K, x belongs to the weak* closure of finite convex combinations of \operatorname{ext}(K), ensuring density. A refinement via the Choquet representation theorem extends this by expressing each x \in K as a barycenter of a supported on \operatorname{ext}(K), with finite combinations dense in the sense of the theorem's hull. This follows from the above structure and uses measures on compact sets, but the core proof remains the separation-based . The Hausdorff assumption on the guarantees that points can be separated from closed sets, underpinning the functional constructions throughout.

Consequences and Examples

One important corollary of the Krein–Milman theorem is that every compact convex subset of a separable metrizable locally convex topological vector space is the closed convex hull of a countable set of its extreme points. This follows from the fact that the set of extreme points is a G_\delta subset in the metrizable topology, allowing selection of a countable dense subset whose convex combinations approximate the set under closure. In finite-dimensional spaces, the Krein–Milman theorem recovers the classical Minkowski–Steinitz theorem, which states that every compact is the of its extreme points (its vertices). For example, in \mathbb{R}^n is precisely the of its finitely many vertices, which are the extreme points. This finite-dimensional case was established by Minkowski for three dimensions and generalized by Steinitz to arbitrary finite dimensions. A concrete illustration in infinite dimensions is the closed unit ball in L^1[0,1], which is convex and bounded but not compact in the norm topology. Unlike compact cases, it contains no extreme points at all: for any f \in L^1[0,1] with \|f\|_1 = 1, one can partition [0,1] into two sets of positive measure and split f into two distinct functions f_1, f_2 with \|f_1\|_1 = \|f_2\|_1 = 1/2 such that f = f_1 + f_2, showing f lies in the interior of a line segment within the ball. Thus, the Krein–Milman theorem does not apply directly, as compactness fails. The Krein–Milman theorem also connects to reflexivity in via James' theorem and related characterizations. Specifically, a is reflexive if and only if its closed unit ball is weakly compact, in which case the theorem implies that the ball equals the closed of its extreme points. In non-reflexive spaces like C[0,1], the unit ball contains extreme points (the constant functions \pm 1) but is not the closed of them, as the weak closure requires more structure. As a highlighting the role of , consider the open unit in any infinite-dimensional normed space. It is but lacks points entirely, since every point is interior and can be expressed as a strict of nearby points within the . Without , the theorem provides no guarantee of points or their hull reconstructing the set.

Generalizations

Extensions to Non-Compact Sets

The classical Krein–Milman theorem requires of the to ensure that it equals the closed of its extreme points in a Hausdorff locally . Extensions to non-compact sets have been developed under additional structural assumptions on the or the set, such as sigma-compactness or metrizability, often involving variants where the is generated by a countable of extreme points. In sigma-compact locally , a non-compact analog holds for certain closed bounded sets, such as norm-closed balls in of continuous linear operators to C(S) where S is a Stonean ; these sets equal the closure of the of their extreme points in the strong . In metrizable locally convex spaces, particularly for compact metrizable subsets, the theorem strengthens to state that the set is the closed of its exposed extreme points—a subset of extreme points defined by supporting hyperplanes—rather than all extreme points. This variant leverages the separability induced by metrizability to ensure the exposed points suffice for the hull. For broader non-compact cases in metrizable spaces, countable collections of extreme points can dense the hull when the set is sequentially compact. A related development concerns absorbing sets, which contain a neighborhood of the origin and thus "absorb" the space. The Bishop–Phelps theorem, stating that norm-attaining functionals are dense in the of a , connects to Krein–Milman by implying that spaces with the Bishop–Phelps property equate the Krein–Milman property (every bounded closed set has an ) with the strong version (every such set is the closed of its s), particularly for absorbing sets like unit balls. However, the focus remains on Krein–Milman variants ensuring s generate absorbing sets under relaxed compactness. Counterexamples illustrate the necessity of compactness or equivalent conditions. In the Banach space c_0 of real sequences converging to zero under the supremum norm, the closed unit ball B = \{ x \in c_0 : \|x\|_\infty \leq 1 \} is non-compact, convex, and bounded, with extreme points precisely the vectors \pm e_n ( vectors scaled by \pm 1). The closed convex hull of these extreme points is the set of all x \in c_0 such that \sum_n |x_n| \leq 1, which misses interior points of B, for example, x = (1/2, 1/2, 1/2, 0, \dots), which has \ell^1-norm $3/2 > 1 while the supremum norm is $1/2 < 1 (an interior point). This demonstrates that without , the hull may fail to recover the full set. In complete metrizable topological spaces (Fréchet spaces), and sequential compactness coincide, allowing the Krein–Milman theorem to apply directly to sequentially compact sets without additional changes. This equivalence facilitates proofs using sequences rather than nets, extending the theorem's utility to non-compact but sequentially compact subsets in such spaces.

Versions in Non-Locally Convex Spaces

In topological vector spaces lacking local convexity, the Krein–Milman theorem fails to hold in its classical form, primarily because the Hahn-Banach separation theorem, essential for identifying and isolating extreme points, requires the presence of a separating family of continuous linear functionals, which is guaranteed only in locally convex settings. Without this separation property, compact convex sets may lack extreme points altogether, rendering the closed of such points insufficient to recover the original set. A seminal counterexample was constructed by J. W. Roberts in 1977, demonstrating a compact subset of a non-locally complete (an F-space) that contains no extreme points. This example resides within a space modeled on L^p(\mu) for $0 < p < 1, where the p-norm induces a non-locally topology; here, the unit ball is compact in a suitable , yet the set has empty extreme boundary, violating the theorem outright. Such constructions highlight that in p-normed spaces with $0 < p < 1, compact sets can exist without the structure needed for the theorem to apply, necessitating adjustments like restricting to subsets embeddable in locally spaces. Subsequent analyses, including re-examinations by N. J. Kalton and N. T. Peck, confirmed the robustness of Roberts' example and explored its implications for fixed-point properties, showing that while the full Krein–Milman assertion breaks down, certain sets in these spaces retain partial extremal structure if they admit affine embeddings into Hausdorff locally topological vector spaces. In broader non-locally settings, such as certain Fréchet spaces (complete metrizable topological vector spaces) or uniformizable topologies without convexity bases, the theorem's validity depends on supplementary conditions like quasi-completeness or barrelledness to ensure the existence of extreme points or approximate representations via combinations, though these remain active areas of investigation with no universal extension established.

Applications

In Functional Analysis

The Krein–Milman theorem plays a central role in Choquet theory, which provides an integral representation for elements of a metrizable compact K in a locally topological vector space as barycenters of s supported on the extreme points \operatorname{ext}(K). Specifically, for any point x \in K, there exists a \mu_x on K such that \mu_x(\operatorname{ext}(K)) = 1 and x is the integral of the with respect to \mu_x, enabling the of sets into measures concentrated on their boundaries. This representation sharpens the topological closure asserted by Krein–Milman, particularly in metrizable cases where the support can be restricted to extreme points without closure. In the context of C*-algebras, the theorem applies to the state space, the set of positive linear functionals of norm 1 equipped with the , which is compact and . The extreme points of this state space are the pure states, and Krein–Milman implies that every state is a weak* of combinations of pure states. The Gelfand–Naimark–Segal (GNS) links each state \phi to a \pi_\phi on a , where pure states yield irreducible representations, facilitating the study of the algebra's structure through its boundary points. For algebras, the set of normal —those continuous with respect to the ultraweak topology—forms a compact of the , and its correspond to the normal pure . By the Krein–Milman theorem, every normal arises as a weak* of combinations of these , and each such induces an of the algebra, connecting the facial structure of the space to the algebra's . An illustrative example arises in duality theory for measure spaces: the unit ball of the space of signed measures M(K) on a compact K (the dual of C(K)), has extreme points consisting of the Dirac delta measures at points of K. This identification, via Krein–Milman, underscores the theorem's role in characterizing the structure of dual spaces and supports integral representations in broader duality frameworks.

In Optimization and Game Theory

In convex optimization, the Krein–Milman theorem underpins the principle that optimal solutions to linear objectives over compact convex feasible sets are attained at extreme points, reducing the search space significantly. This is particularly evident in , where the feasible region forms a in finite dimensions, and the theorem's finite-dimensional analog guarantees that the polytope is the of its vertices, corresponding to basic feasible solutions. The simplex method exploits this structure by pivoting between these extreme points to find an optimum, ensuring efficiency in solving problems like or . In , the theorem illuminates the structure of equilibria in finite normal-form games. The mixed space for each player is the probability simplex over pure strategies, a compact whose extreme points are precisely the pure strategies. Any mixed strategy is thus a of these extreme points, per the theorem, allowing equilibria to be expressed as distributions over pure profiles. Pure strategy equilibria occupy the extreme points of the joint strategy space, while mixed equilibria arise as convex combinations thereof, facilitating computational analysis and purification results. For infinite strategy games, such as zero-sum games with compact convex payoff sets, the Krein–Milman theorem enables equilibrium existence by decomposing the strategy spaces into their extreme points, often linking to Kakutani's for upper hemicontinuous best-response correspondences. In these settings, the value of the game is achieved through strategies that can be approximated via extreme points, ensuring the lies in the closed . This framework extends von Neumann's finite to continuous action spaces, like differential games or economic models with infinite types. A example arises in models, where the feasible set of allocations—such as distributing limited goods under linear constraints—forms a compact . The theorem implies that Pareto-optimal allocations occur at extreme points, simplifying the identification of efficient outcomes in applications like , without needing to evaluate interior points.

Foundational and Historical Aspects

Relation to

The proof of the Krein–Milman theorem employs , which is equivalent to the , to demonstrate the existence of extreme points. Specifically, it applies to the of faces of the compact convex set, ordered by inclusion, to obtain a maximal face, from which an extreme point is derived. In (ZF) without the , the Krein–Milman theorem does not hold in general, as there exist models where compact convex sets in locally convex topological vector spaces lack extreme points. Counterexamples exist in models of ZF with dependent choice, such as those where all sets of reals are Lebesgue measurable, particularly for non-separable spaces. The Krein–Milman theorem, together with the , implies the over ZF, as shown by constructing choice functions from extreme points in appropriately defined dual spaces. However, weaker versions hold without full choice: the suffices to prove the theorem for separable locally convex spaces, where the set can be approximated by countable dense subsets. This dependence highlights that while the theorem implies the in conjunction with principles like the Boolean prime ideal theorem, choice-free alternatives exist for restricted settings, such as those relying solely on dependent choice.

Historical Development

The concept of extreme points in convex sets originated in the early with foundational work in finite dimensions. In 1911, demonstrated that every compact body in finite-dimensional is the of its extreme points, providing the first systematic study of this relation in and laying the groundwork for later generalizations. This result highlighted the structural role of extreme points in describing sets geometrically. Building on this, Eduard Steinitz in 1916 extended insights into polytopes, showing that bounded polyhedral sets in finite dimensions can be represented as the of their vertices, which are precisely the extreme points, further emphasizing the finite approximation of structures. The Krein–Milman theorem emerged in 1940 as a significant advancement to infinite-dimensional settings, proved by Mark Krein and David Milman in their seminal paper "On extreme points of regular sets." Published in Studia Mathematica, the theorem states that a compact subset of a Hausdorff locally topological vector space equals the closed of its extreme points, extending Minkowski's finite-dimensional result to abstract linear topological spaces. This development was deeply influenced by the post-Hilbert era of , particularly the theory of Banach spaces established in the and by , which provided the topological framework for studying infinite-dimensional convexity beyond Hilbert spaces. In the 1940s, V. L. Šmulian strengthened aspects of the theorem, particularly in relation to topologies on Banach spaces, with early contributions like his 1939 note on differing topologies, which informed subsequent refinements of extremal properties in normed spaces. The 1950s saw further evolution through generalizations by Gustave Choquet, whose 1953–1954 theory of capacities introduced integral representations over extreme points, enabling broader applications in measure theory and for non-locally convex spaces. While the core theorem has remained stable without major revisions since the mid-20th century, its applications have expanded into modern fields.

References

  1. [1]
    Krein-Milman Theorem -- from Wolfram MathWorld
    In the field of functional analysis, the Krein-Milman theorem is a result which characterizes all (nonempty) compact convex subsets K of "sufficiently nice" ...
  2. [2]
    The Krein–Milman Theorem | SpringerLink
    Mar 7, 2020 · The Krein–Milman theorem asserts that in a Hausdorff locally convex space all points of a compact convex set can be approximated by convex ...
  3. [3]
    The Krein–Milman Theorem and Its Applications - SpringerLink
    Jul 10, 2018 · We demonstrate Krein–Milman theorem on existence of extreme points in convex compact sets and give a number of applications.
  4. [4]
  5. [5]
    Krein–Milman Theorem - SpringerLink
    A theorem stating that a compact closed set can be represented as the convex hull of its extreme points. First shown by H. Minkowski [4] and studied by some ...Missing: original | Show results with:original
  6. [6]
    The Krein-Milman theorem in operator convexity
    The Krein-Milman theorem is without doubt one of the cornerstones of functional analysis. With the rise of non-commutative functional analysis and related ...
  7. [7]
    [PDF] Extreme points and the Krein–Milman theorem - Caltech
    This proves the lack of compactness directly. Theorem 8.14 (The Krein–Milman Theorem) Let A be a compact convex subset of a locally convex vector space, X. Then.Missing: original paper
  8. [8]
    [PDF] Some Aspects of Convexity - Indian Academy of Sciences
    We shall try to indicate how this problem is solved as a result of three results. - namely, the Krein-Milman theorem, the Birkhoff-von. Neumann theorem, and the ...
  9. [9]
    [PDF] Questions about Extreme Points - arXiv
    Nov 2, 2022 · We discuss the geometry of the unit ball—specifically, the structure of its extreme points (if any)—in subspaces of L1 and L∞ on the circle ...
  10. [10]
    [PDF] Extreme points of convex matrix sets - Chi-Kwong Li
    An element x in a convex set S is an extreme point if x 6= (x1 + x2)/2 for any x1, x2 ∈ S. Krein-Milman Theorem / Caratheodory Theorem. If S ⊂ RN is a compact ( ...
  11. [11]
    [PDF] A short account of topological vector spaces Normed spaces, and ...
    In this chapter, we give a minimum introduction to topological vector spaces. The interested reader is referred to standard text-books of Functional Analysis, ...Missing: textbook | Show results with:textbook
  12. [12]
    [PDF] Functional Analysis
    locally convex (and, actually, ∅ and X are the only convex open subsets of X). As mentioned earlier, a main goal of Functional Analysis is the representation of.Missing: textbook | Show results with:textbook
  13. [13]
    [PDF] TOPOLOGICAL VECTOR SPACES1 1. Definitions and basic facts.
    Locally convex first countable spaces are metrizable via a countable family of seminorms. Example. Let X = C(Rn) be the space of continuous functions, with the ...
  14. [14]
    [PDF] Bounded subsets of topological vector spaces
    This together with Corollary 2.1.7 gives that in any Hausdorff t.v.s. a compact subset is always bounded and closed. In finite dimensional Hausdorff. t.v.s. we ...
  15. [15]
    [PDF] 1. Topological Vector Spaces
    Such sets are convex, balanced and absorbing. Therefore, by Theorem 1.4, (X,J ) is locally convex. Suppose f is a linear functional on X which is J -continuous.
  16. [16]
    On extreme points of regular convex sets - EuDML
    Krein, M., and Milman, D.. "On extreme points of regular convex sets." Studia Mathematica 9.1 (1940): 133-138. <http://eudml.org/doc/219061>.Missing: Banach spaces
  17. [17]
    [PDF] Functional Analysis Lecture Notes - Michigan State University
    point of K can be written as a convex combination of (at most) N + 1 extreme points. ... constant functions. By. Theorem 3.1 of these notes, we may extend L ...
  18. [18]
    [PDF] FUNCTIONAL ANALYSIS | Second Edition Walter Rudin
    Rudin, Walter, (date). Functional analysis/Walter Rudin.-2nd ed. p. em. -(international series in pure and applied mathematics).
  19. [19]
    [PDF] Banach Spaces V: A Closer Look at the w- and the w -Topologies
    In this section we discuss two important, but highly non-trivial, results concerning the weak topology (w) on Banach spaces and the weak dual topology (w∗) on ...
  20. [20]
    [PDF] Minimal concave functions over an obstacle: the ... - MSU Math
    Set of extreme points of the unit ball Lp (1 < p < ∞) coincides with {f : ||f ||Lp = 1}. Paata Ivanisvili. Concave Envelope. Page 116. Introduction. Example.
  21. [21]
    [PDF] 1 Krein-Milman theorem
    We are going to prove a following wonderful theorem. Theorem 1.1. Let X be a locally convex linear toplogical vector space. Let.
  22. [22]
    [PDF] The Krein–Milman Theorem - A Project in Functional Analysis
    Nov 29, 2016 · Proof of Krein–Milman. For conv(extK) ⊆ K,. K compact, convex, and K ⊇ extK. =⇒ K closed, convex, and K ⊇ extK. Page 53. Proof of Krein–Milman.
  23. [23]
    [PDF] Convex sets and their integral representations
    Jun 8, 2012 · 2 THE KREIN-MILMAN THEOREM. 2 The Krein-Milman Theorem. By ... subset of a locally convex Hausdorff topological vector space E, and that x0.
  24. [24]
    [PDF] Geometry of polynomial spaces and polynomial inequalities - TEMat
    According to the Krein-Milman theorem (or its finite dimensional version proved by Minkowski in 3-dimensional spaces and by Steinitz for any dimension), the set ...
  25. [25]
    A Banach space where the closed unit ball is the convex hull of its ...
    Sep 26, 2019 · Let X be a Banach space where the closed unit ball equals the convex hull of its extreme points. Is it true that this implies X is reflexive?Unit ball as norm closure of the convex hull of extreme points in ℓ∞(X)If the closed unit ball of Banach space has at least one extreme point ...More results from mathoverflow.net
  26. [26]
    [PDF] A Convexity Primer
    Sep 7, 2019 · The unit ball of C[0,1] does contain extreme points, but the entire thing is not a closed convex hull of its extreme points. I assert that the.
  27. [27]
    A NON-COMPACT KREIN-MILMAN THEOREM - Project Euclid
    This paper describes a class of closed bounded convex sets which are the closed convex hulls of their extreme points. It includes all compact ones and those ...
  28. [28]
    [PDF] On the Krein-Milman theorem for convex compact metrizable sets
    May 1, 2016 · Abstract. The Krein-Milman theorem states that a convex compact subset of a Haus- dorff locally convex topological space, is the closed ...Missing: original | Show results with:original
  29. [29]
  30. [30]
    On Bishop–Phelps and Krein–Milman Properties - MDPI
    A real topological vector space is said to have the Krein–Milman property if every bounded, closed, convex subset has an extreme point.
  31. [31]
    Elementary applications of Krein-Milman - MathOverflow
    May 19, 2012 · The Krein-Milman Theorem is used in the proof of Birkhoff's Theorem that the set of bistochastic matrices is the convex envelop of permutations ...Generalizations of the Birkhoff-von Neumann TheoremKrein Milman theorem without the axiom of choiceMore results from mathoverflow.netMissing: Neumann | Show results with:Neumann
  32. [32]
    [PDF] 3. Topological vector spaces
    aU. Since A is the intersection of convex sets it is convex. It is balanced because for every b ≤ 1, bA.
  33. [33]
    [PDF] Quasi-Banach Spaces - Nigel Kalton Memorial
    Krein-Milman Theorem fails in general quasi-Banach spaces by developing powerful new techniques. Quasi-Banach spaces (Hp-spaces when p < 1) were also used ...
  34. [34]
  35. [35]
    [PDF] Continuous version of the Choquet Integral Representation Theorem
    Apr 5, 2005 · µx(extK)=1. Re all that if the set extK is losed then the Choquet theorem is equivalent to the Krein Milman theorem. losed bounded onvex subset ...
  36. [36]
    [PDF] extreme Points of the Generalized State Space of a Commutative C
    In contrast, Farenick and Morenz have shown that these generalized state spaces contain many C∗-extreme points. In fact, they prove a Krein-Milman type theorem.
  37. [37]
    Dirac measures are extreme points of unit ball of $C(K)
    May 27, 2017 · The idea of the proof is that the variation measure |μ| is equal to δx, so that |μ| is an extreme point of the convex set P of probability ...Extreme points of $B_{C(K)^{*}}$ - Mathematics Stack ExchangeKrein-Milman and dual spaces - Mathematics Stack ExchangeMore results from math.stackexchange.com
  38. [38]
    [PDF] TMA 4180 Optimeringsteori Linear Programming Basics
    Task 13: How does the Krein-Milman Theorem work for a cube? We are going to identify the extreme points in as so-called basic points, or basic feasible ...
  39. [39]
    [PDF] Nash equilibrium and generalized integration for infinite normal form ...
    Dec 10, 2004 · By the Krein–Milman theorem (e.g. Dunford and Schwartz, 1957, Theorem V.8.4), M(µ) is the closed convex envelope of its necessarily non ...
  40. [40]
    Nash equilibria in oo-dimensional spaces: an approximation theorem
    Nash equilibria in oo-dimensional spaces: an approximation theorem. 745 through the Krein-Milman theorem, giving conditions under which the convex closure of ...
  41. [41]
    [PDF] Zornian Functional Analysis - arXiv
    Oct 29, 2020 · If the axiom of choice fails but BPI holds, then the Krein–Milman theorem fails. Proof. If the axiom of choice fails then there is a normed ...
  42. [42]
    The Axiom of Choice - Stanford Encyclopedia of Philosophy
    Jan 8, 2008 · Krein-Milman Theorem: the unit ball \(B\) of the dual of a real normed linear space has an extreme point, that is, one which is not an ...
  43. [43]
    Minkowski's development of the concept of convex bodies - jstor
    Oct 2, 2007 · Clearly, Minkowski at this point conceived of the nowhere concave bodies with middle point as a tool in number theory. The connection to ...Missing: extreme | Show results with:extreme
  44. [44]
    History of Banach Spaces and Linear Operators
    Mar 21, 1985 · ... Minkowski ... Furthermore, it may serve as a reference and guide for beginners who want to learn Banach space theory with some historical flavor.<|control11|><|separator|>