Hahn–Banach theorem

The Hahn–Banach theorem is a central tool in functional analysis (a field of mathematics). It allows the extension of bounded linear functionals defined on a subspace of some vector space to the whole space, and it also shows that there are "enough" continuous linear functionals defined on every normed vector space to make the study of the dual space "interesting". Another version of the Hahn–Banach theorem is known as the Hahn–Banach separation theorem or the hyperplane separation theorem, and has numerous uses in convex geometry.

History

The theorem is named for the mathematicians Hans Hahn and Stefan Banach, who proved it independently in the late 1920s. The special case of the theorem for the space of continuous functions on an interval was proved earlier (in 1912) by Eduard Helly,[1] and a more general extension theorem, the M. Riesz extension theorem, from which the Hahn–Banach theorem can be derived, was proved in 1923 by Marcel Riesz.[2]

The first Hahn-Banach theorem was proved by Eduard Helly in 1921 who showed that certain linear functionals defined on a subspace of a certain type of normed space () had an extension of the same norm.[3] Helly did this through the technique of first proving that a one-dimensional extension exists (where the linear functional has its domain extended by one dimension) and then using induction. In 1927, Hahn defined general Banach spaces and used Helly's technique to prove a norm preserving version of Hahn-Banach theorem for Banach spaces (where a bounded linear functional on a subspace has a bounded linear extension of the same norm to the whole space).[3] In 1929, Banach, who was unaware of Hahn's result, generalized it by replacing the norm-preserving version with the dominated extension version that uses sublinear functionals.[3] Whereas Helly's proof used mathematical induction, Hahn and Banach both used transfinite induction.[3]

The Hahn-Banach theorem arose from attempts to solve infinite systems of linear equations. This is needed to solve problems such as the moment problem, whereby given all the potential moments of a function one must determine if a function having these moments exists and if so then find it. Another such problem is the Fourier cosine series problem, whereby given all the potential Fourier cosine coefficients one must determine if a function having those coefficients exists and if so then find it.[3] Riesz and Helly solved the problem for certain classes of spaces (such as Lp([0, 1]) and C([a, b])) where they discovered that the existence of a solution was equivalent to the existence and continuity of certain linear functionals.[3] In effect, they needed to solve the following problem:[3]

(The vector problem) Given a collection of bounded linear functionals on a normed space X and a collection of scalars , determine if there is an x X such that fi(x) = ci for all i I.

To solve this, if X is reflexive then it suffices to solve the following dual problem:[3]

(The functional problem) Given a collection of vectors in a normed space X and a collection of scalars , determine if there is a bounded linear functional f on X such that f(xi) = ci for all i I.[3]

Riesz went on to define Lp([0, 1]) (1 < p < ) in 1910 and the lp spaces in 1913.[3] While investing these spaces he proved a special case of the Hahn-Banach theorem. Hally also proved a special case of the Hahn-Banach theorem in 1912.[3] In 1910, Riesz solved the functional problem for some specific spaces and in 1912, Helly solved it for a more general class of spaces. It wasn't until 1932 that Banach, in one of the first important applications of the Hahn-Banach theorem, solved the general functional problem. The following theorem states the general functional problem and characterizes its solution.[3]

Theorem (The functional problem):[3] Let X be a real or complex normed space, I a non-empty set, a family of scalars, and a family of vectors. Then there exists a continuous linear functional f on X such that f(xi) = ci for all i I if and only if there exists a K > 0 such that for any choice of scalars where all but finitely many si are 0, we necessarily have

.

One can use the above theorem to deduce the Hahn-Banach theorem.[3] If X is reflexive, then this theorem solves the vector problem.

Prerequisites and definitions

The most general formulation of the theorem needs some preparation. Given a real vector space X, a function p : XR is called sublinear if

  • Positive homogeneity: p(γx) = γ p(x) for all γR+, xX,
  • Subadditivity: p(x + y) ≤ p(x) + p(y) for all x, yX.

Every seminorm on X (in particular, every norm on X) and every linear functional on X is sublinear. Other sublinear functions can be useful as well, especially Minkowski functionals of convex sets.

If f is a linear functional on a topological vector space (TVS) X (e.g. a normed space) and if p is a continuous sublinear functional on X then |f| p implies that f is continuous. If f is a real-valued linear functional and p is a seminorm (not just a sublinear functional) then f p implies that |f| p.[3] If f(x) = R(x) + i I(x) is a complex-valued linear functional on a complex vector space X then I(x) = - R(ix) for all x X, so that I is completely determined by f's real part R, which we will also denote by Re f, and this in turn implies that f is entirely determined by R.[3] In particular, it follows that f is bounded (resp. continuous) if and only if R is bounded (resp. continuous).

If S is a subset of X and if f : S R is a function then we say that p dominates f on S if f(s) p(s) for all s S. We call a function F : X R and extension of f to X if F(s) = f(s) for all s S; if in addition F is a linear map then we call F a linear extension of f.

Formulation

The general template for the various versions of the Hahn-Banach theorem presented in this article is as follows:

X is a vector space, p is a sublinear functional on X, M is a vector subspace of X, and f is a linear functional on M satisfying (for instance) |f| p on M (and possibly some other conditions). One then concludes (possibly among other things) that there exists a linear extension F of f to X such that |F| p on X.

Sometimes X is assumed to have additional structure, such as a topology or a norm, but many of the statements are purely algebraic. We may apply a purely algebraic version of this theorem to the case where X is a topological vector space (TVS) (e.g. a normed space) as follows: by choosing p to be continuous, the inequality |F| p allows us to conclude that F is necessarily continuous. (As a side note, if p is continuous then the assumption "|f| p on M" also allows us to conclude that f is continuous).

Some of the statements are given only for real vector spaces or only real-valued linear functionals while others are given for real or complex vector spaces. One may apply a result that applies only to real-valued linear functionals to the complex case by recalling that a complex-valued linear functional c(x) = R(x) + i I(x) is continuous if and only if its real part, R, is continuous and that furthermore, the real part R completely determines the imaginary part I and thus completely determines c. The sublinear functional p is always real-valued (although it could possibly take on negative values if it is not assumed to be positive). If X is assumed to be a real (resp. complex) vector space then all linear functionals are assumed to be real-valued (resp. complex-valued). If the linear functional f is real-valued then you'll often see the condition f p whereas if f is complex-valued then you're more likely to see |f| p or Re f p.

Sometimes a functional is assumed to be bounded and other times it is assumed to be continuous. If X a is pseudometrizable TVS (e.g. a metrizable TVS or a normed space) then a linear map from X into any other TVS is continuous if and only if it is a bounded map (i.e. it maps bounded subsets of X to bounded subsets of the codomain). In particular, a linear functional on a normed or Banach space is continuous if and only if it is bounded.

Hahn-Banach theorem

The following lemma is fundamental to proving the general Hahn-Banach theorem and its basic prove first appeared in a 1912 paper by Helly where it was proved for the space C([a, b]).[3]

Lemma (One-dimensional dominated extension theorem):[3] Let X be a real vector space, p : X → ℝ a sublinear function, f : M → ℝ a linear functional on a proper vector subspace MX such that f p on M (i.e. f(m) p(m) for all m M), and x X an vector not in M. There exists a linear extension F : M x → ℝ of f to M x = span { M, x } such that F p on M x.

To prove this lemma, one first shows that for all m, n M, -p(- x - n) - f(n) p(m + x) - f(m) which allows us to define:

a = [ -p(- x - n) - f(n)], and b = [p(m + x) - f(m)]

from which we conclude "the decisive inequality"[3] that for any c [a, b],

-p(- x - n) - f(n) c p(m + x) - f(m).

For any m + rx M x, one then defines F(m + rx) := f(m) + rc, which gives us the desired extension.

Hahn–Banach dominated extension theorem:[3](Rudin 1991, Th. 3.2). If p : X → ℝ is a sublinear function, and f : MR is a linear functional on a linear subspace MX which is dominated by p on M, then there exists a linear extension F : X → ℝ of f to the whole space X that is dominated by p, i.e., there exists a linear functional F such that

Hahn–Banach theorem (alternative version). Set K = ℝ or C and let X be a K-vector space with a seminorm p : X → ℝ. If f : MK is a K-linear functional on a K-linear subspace M of X which is dominated by p on M in absolute value,

then there exists a linear extension F : XK of f to the whole space X, i.e., there exists a K-linear functional F such that

In the complex case of the alternate version, the C-linearity assumptions demand, in addition to the assumptions for the real case, that for every vector xM, we have ixM and f(ix) = if(x).

The extension F is in general not uniquely specified by f and the proof gives no explicit method as to how to find F. The usual proof for the case of an infinite dimensional space X uses Zorn's lemma or, equivalently, the axiom of choice. It is now known (see below) that the ultrafilter lemma, which is slightly weaker than the axiom of choice, is actually strong enough.

It is possible to relax slightly the subadditivity condition on p, requiring only that (Reed and Simon, 1980):

It is further possible to relax the positive homogeneity and the subadditivity conditions on p, requiring only that p is convex (Schechter, 1996).

This reveals the intimate connection between the Hahn–Banach theorem and convexity.

The Mizar project has completely formalized and automatically checked the proof of the Hahn–Banach theorem in the HAHNBAN file.[4]

Important consequences

Theorem:[3] If p is a sublinear functional on a real vector space X then there exists a linear functional f on X such that f p on X.

Theorem:[3] Let p be a sublinear functional on a real vector space X and let z X. Then there exists a linear functional f on X such that

  1. f(z) = p(z);
  2. -p(-x) f(x) p(x) for all x X;
  3. if p is a seminorm then | f | p.

If X is a TVS and p is continuous at 0, then f is continuous.

The theorem has several important consequences, some of which are also sometimes called "Hahn–Banach theorem":

  • (Continuous extensions on locally convex spaces): Let X be a real or complex locally convex topological vector space (TVS), M a vector subspace of X, and f a continuous linear functional on M. There f has a continuous linear extension to all of X.[3]

For normed spaces we have the following results:

  • If X is a normed vector space with linear subspace M (not necessarily closed) and if f : MK is continuous and linear, then there exists an extension F : XK of f which is also continuous and linear and which has the same operator norm as f (see Banach space for a discussion of the norm of a linear map). In other words, in the category of normed vector spaces, the space K is an injective object.
  • If X is a normed vector space with linear subspace M (not necessarily closed) and if z is an element of X not in the closure of M, then there exists a continuous linear map f : XK with f(x) = 0 for all x in M, f(z) = 1, and ||f|| = dist(z, M)−1.
  • In particular, if X is a normed vector space and z is an element of X, then there exists a continuous linear map f : XK such that f(z) = ||z|| and ||f|| ≤ 1. This implies that the natural injection J from a normed space X into its double dual V′′ is isometric.
  • If X is a normed space over the real or complex numbers, M is a closed proper vector subspace of X, and f is a continuous linear functional on M, then there exists a continuous linear extension F of f to X such that .[3]

Hahn–Banach separation theorems

Hahn–Banach separation theorems are the geometrical versions of the Hahn–Banach Theorem.[5][6] They have numerous uses in convex geometry,[7] optimization theory, and economics. The separation theorem is derived from the original form of the theorem.

Let X be a real vector space, A and B non-empty subsets of X, f 0 a real linear functional on X, s a scalar, and let . We also define the lower (resp. upper) half space to be { x X : f(x) s } (resp. { x X : f(x) s }). We define the strict lower (resp. strict upper) half space to be { x X : f(x) < s } (resp. { x X : f(x) > s }).

We say that we say that H (or f) separates A and B if sup f(A) s inf f(B) or equivalently, if f(a) s f(b) for all a A and b B. The separation is:

  • proper if ;[5]
  • strict if A H = and B H = or equivalently, if f(a) < s < f(b) for all a A and b B[3] (note that some authors define "strict" to mean that A H = or B H = );[5]
  • strong and we say that A and B are strongly separated by H if there exists an r > 0 such that f(a) s - r < s + r f(b) for all a A and b B.[3]

Note that A and B are separated (resp. strictly separated, strongly separated) if and only if the same is true of { 0 } and B - A.[3] If A and B are convex then they are strongly separated by a hyperplane if and only if there exists an absorbing convex U such that (A + U) B = .[3] We say that A and B are united if they cannot be properly separated.[5]

If a0 A and H separates A and { a0 } then H is called a supporting hyperplane of A at a0, a0 is called a support point of A, and f is called a support functional.[5] If A is convex and a0 A, then we call a0 a smooth point of A if there exists a unique hyperplane H such that a0 A H.[3] We call a normed space X smooth if at each point x in its unit ball there exists a unique closed hyperplane to the unit ball at x.[3] Köthe showed in 1983 that a normed space is smooth at a point x if and only if the norm is Gateaux differentiable at that point.[3]

Subsets of a half space

Theorem:[3] Let X be a real TVS, f a non-0 continuous real-valued linear functional on X, s a real number, H = , and G a subset of X having non-empty interior. If G is a subset of a half space (i.e. either the lower half space or the lower half space) then the closure of G also lies in that half space and furthermore, the interior of G lies in the corresponding strict half space.

Separation of sets

Theorem: Set K = R or C and let X be a topological vector space over K. If A, B are convex non-empty disjoint subsets of X, then:

  • If A is open then A and B are separated by a closed hyperplane; explicitly, this means that there exists a continuous linear map f : XK and sR such that Re(f(a)) < s ≤ Re(f(b)) for all aA, bB.
  • If A and B are open then A and B are strictly separated by a closed hyperplane; explicitly, this means that there exists a continuous linear map f : XK and sR such that Re(f(a)) < s < Re(f(b)) for all aA, bB.[3]
  • If X is locally convex, A is compact, and B closed, then there exists a continuous linear map f : XK and s, tR such that Re(f(a)) < t < s < Re(f(b)) for all aA, bB.

The following theorem may be used if the sets are not necessarily disjoint.[5]

Theorem: Let X be a real locally convex topological vector space and let A and B be non-empty convex subsets. If and then there exists a continuous and such that and for all (such is f is necessarily non-zero).

Theorem (Separation of a subspace and an open convex set):[3] Let M be a vector subspace of a topological vector space X and let U be a non-empty open convex subset of X that does not intersect M. Then there exists a continuous linear functional f on X such that f(m) = 0 for all m M and Re f > 0 on U, where Re f is the real part of f (if f is real-valued then Re f = f).

Separation of a point and set

The following is the Hahn–Banach separation theorem for a point and a set.[5]

Theorem. Let X be a real topological vector space and be convex with . If then is a support point of A.

Corollary. Let X be a real topological vector space, A a non-empty convex open subset of X, and . Then there exists a continuous such that for every a A.

Separation of a closed and compact set

Theorem:[3] Let A and B be non-empty disjoint convex subsets of a locally convex topological vector space X over the real or compact numbers. If A is closed and B is compact then they are strongly separated by a closed hyperplane (i.e. there exists a continuous real-valued linear functional f on X such that sup f(A) < inf f(B)).

Separation of points and disked neighborhoods of 0

Theorem:[3] Let U be a convex balanced neighborhood of 0 in a real or complex locally convex TVS X and z X be a point not in U. There exists a continuous real-valued linear functional f on X such that sup f(U) < inf f(z)).

Reflexive Banach spaces

Theorem:[8] A real Banach space is reflexive if and only if every pair of non-empty disjoint closed convex subsets, one of which is bounded, can be strictly separated by a hyperplane.

Characterizations

The above separation theorems may be generalized to the following theorems:[5]

Theorem. Let A and B be non-empty convex subsets of a topological vector space X.

  • A and B are strongly separated by a closed hyperplane if and only if there exists a convex open neighborhood U of 0 in X such that U (B - A) = (or equivalently, if (A + U) B = ).[3]
  • If and X is a vector space over the reals then:
if and only if there exists some continuous such that and .[5]
  • If X is locally convex then:
if and only if A and B are strongly separated by a closed hyperplane (i.e. if there exists some continuous such that ). (Such an f will necessarily be non-).[3][5]

Hahn-Banach sandwich theorem

Hahn-Banach sandwich theorem:[3] Let S be any subset of a real vector space X, let p be a sublinear functional on X, and let f : S ℝ be any map. If there exist positive numbers a and b such that for all x, y S,

then there exists a linear functional F on X such that F p on X and f F on S.

Geometric Hahn–Banach theorem

One form of Hahn–Banach theorem is known as the Geometric Hahn–Banach theorem, or Mazur's theorem.[9]

Theorem. Let K be a convex set having a nonempty interior in a real normed linear vector space X. Suppose X is a linear variety in X containing no interior points of K. Then there is a closed hyperplane in X containing X but containing no interior points of K; i.e., there is an element x* ∈ X* and a constant c such that for all vX and for all kint(K).

This can be generalized to an arbitrary topological vector space, which need not be locally convex or even Hausdorff, as:[10]

Theorem. Let M be a vector subspace of the topological vector space X. Suppose K is a non-empty convex open subset of X with KM = ∅. Then there is a closed hyperplane N in X containing M with KN = ∅.

Mazur-Orlicz theorem

The following theorem of Mazur-Orlicz (1953) is equivalent to the Hahn-Banach theorem.

Mazur-Orlicz theorem:[3] Let T be any set, r : T ℝ be any real-valued map, X be a real or complex vector space, v : T X be any map, and p be a sublinear functional on X. Then the following are equivalent:

  1. there exists a real-valued linear functional F on X such that F p on X and r Fv on T;
  2. for any positive integer n, any sequence s1, ..., sn of non-negative real numbers, and any sequence t1, ..., tn of elements of T,

Generalizations

Due to its importance, the Hahn-Banach theorem has been generalized many times.

Theorem (Andenaes, 1970):[3] Let M be a vector subspace of a real vector space X, p be a sublinear functional on X, f be a linear functional on M such that f p on M, and let S be any subset of X. Then there exists a linear functional F on X that extends f, satisfies F p on X, and is (pointwise) maximal in the following sense: if G is a linear functional on X ending f and satisfying G p on X, then G F implies that G = F on S.

Theorem:[3] Let X and Y be vector spaces over the same field, M be a vector subspace of X, and f : M Y be a linear map. Then there exists a linear map F : X Y that extends f.

Theorem:[3] Let M be a vector subspace of a real or complex vector space X, let D be an absorbing disk in X, and let f be a linear functional on M such that |f| 1 on M D. Then there exists a linear functional F on X extending f such that |F| 1 on D.

Relation to axiom of choice

As mentioned earlier, the axiom of choice implies the Hahn–Banach theorem. The converse is not true. One way to see that is by noting that the ultrafilter lemma (or equivalently, the Boolean prime ideal theorem), which is strictly weaker than the axiom of choice, can be used to show the Hahn–Banach theorem, although the converse is not the case.

The Hahn–Banach theorem is equivalent to the following:[11]

(∗): On every Boolean algebra B there exists a "probability charge", that is: a nonconstant finitely additive map from B into [0, 1].

(The Boolean prime ideal theorem is easily seen to be equivalent to the statement that there are always nonconstant probability charges which take only the values 0 and 1.)

In Zermelo–Fraenkel set theory, one can show that the Hahn–Banach theorem is enough to derive the existence of a non-Lebesgue measurable set.[12] Moreover, the Hahn–Banach theorem implies the Banach–Tarski paradox.[13]

For separable Banach spaces, D. K. Brown and S. G. Simpson proved that the Hahn–Banach theorem follows from WKL0, a weak subsystem of second-order arithmetic that takes a form of Kőnig's lemma restricted to binary trees as an axiom. In fact, they prove that under a weak set of assumptions, the two are equivalent, an example of reverse mathematics.[14][15]

Consequences

Topological vector spaces

If X is a topological vector space, not necessarily Hausdorff or locally convex, then there exists a non-zero continuous linear form if and only if X contains a nonempty, proper, convex, open set M.[16] So if the continuous dual space of X, X*, is non-trivial then by considering X with the weak topology induced by X*, X becomes a locally convex topological vector space with a non-trivial topology that is weaker than original topology on X. If in addition, X* separates points on X (which means that for each xX there is a linear functional in X* that's non-zero on x) then X with this weak topology becomes Hausdorff. This sometimes allows some results from locally convex topological vector spaces to be applied to non-Hausdorff and non-locally convex spaces.

Theorem:[3] A topological vector space X has a non-trivial continuous dual space if and only if it has a proper convex neighborhood of 0.

Theorem:[3] A U be a convex balanced neighborhood of 0 in a locally convex topological vector space X and suppose x X is not an element of U. Then there exists a continuous linear functional f on X such that

sup |f(U)| |f(x)|.

The dual space C[a, b]*

We have the following consequence of the Hahn–Banach theorem.

Proposition. Let −∞ < a < b < ∞. Then, FC[a, b]* if and only if there exists a (complex) measure ρ : [a, b] → R of bounded variation such that

for all uC[a, b]. In addition, |F| = X(ρ), where X(ρ) denotes the total variation of ρ.

Partial differential equations

The Hahn–Banach theorem is often useful when one wishes to apply the method of a priori estimates. Suppose that we wish to solve the linear differential equation for u, with f given in some Banach space X. If we have control on the size of u in terms of and we can think of u as a bounded linear functional on some suitable space of test functions g, then we can view f as a linear functional by adjunction: . At first, this functional is only defined on the image of P, but using the Hahn–Banach theorem, we can try to extend it to the entire codomain X. We can then reasonably view this functional as a weak solution to the equation.

See also

Notes

  1. O'Connor, John J.; Robertson, Edmund F., "Hahn–Banach theorem", MacTutor History of Mathematics archive, University of St Andrews.
  2. See M. Riesz extension theorem. According to Gȧrding, L. (1970). "Marcel Riesz in memoriam". Acta Math. 124 (1): I–XI. doi:10.1007/bf02394565. MR 0256837.CS1 maint: ref=harv (link), the argument was known to Riesz already in 1918.
  3. Narici 2011, pp. 177-220.
  4. HAHNBAN file
  5. Zălinescu, C. (2002). Convex analysis in general vector spaces. River Edge, NJ: World Scientific Publishing Co., Inc. pp. 5–7. ISBN 981-238-067-1. MR 1921556.
  6. Gabriel Nagy, Real Analysis lecture notes
  7. Harvey, R.; Lawson, H. B. (1983). "An intrinsic characterisation of Kähler manifolds". Invent. Math. 74 (2): 169–198. doi:10.1007/BF01394312.CS1 maint: ref=harv (link)
  8. Narici 2011, pp. 212.
  9. Luenberger, David G. (1969). Optimization by Vector Space Methods. New York: John Wiley & Sons. pp. 131–134. ISBN 978-0-471-18117-0.
  10. Treves, p. 184
  11. Schechter, Eric. Handbook of Analysis and its Foundations. p. 620.
  12. Foreman, M.; Wehrung, F. (1991). "The Hahn–Banach theorem implies the existence of a non-Lebesgue measurable set" (PDF). Fundamenta Mathematicae. 138: 13–19.CS1 maint: ref=harv (link)
  13. Pawlikowski, Janusz (1991). "The Hahn–Banach theorem implies the Banach–Tarski paradox". Fundamenta Mathematicae. 138: 21–22.
  14. Brown, D. K.; Simpson, S. G. (1986). "Which set existence axioms are needed to prove the separable Hahn–Banach theorem?". Annals of Pure and Applied Logic. 31: 123–144. doi:10.1016/0168-0072(86)90066-7.CS1 maint: ref=harv (link) Source of citation.
  15. Simpson, Stephen G. (2009), Subsystems of second order arithmetic, Perspectives in Logic (2nd ed.), Cambridge University Press, ISBN 978-0-521-88439-6, MR2517689
  16. Schaefer 1999, p. 47

References

    ISBN 0-12-622760-8.

    This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.