Dual system

In the field of functional analysis, a subfield of mathematics, a dual system, dual pair, or a duality over a field K (K is either the real or the complex numbers) is a triple (X, Y, b) consisting of two vector spaces over K and a bilinear map b : X × YK satisfying the following two separation axioms:[1]

  1. (S1) if xX is such that b(x, y) = 0 for all yY then x = 0;
  2. (S2) if yY is such that b(x, y) = 0 for all xX then y = 0.

According to Helmut H. Schaefer, "the study of a locally convex space in terms of its dual is the central part of the modern theory of topological vector spaces, for it provides the deepest and most beautiful results of the subject."[2]

Definition

A pairing or a pair is a triple (X, Y, b) consisting of two vector spaces over 𝕂 and a bilinear map b : X × Y → 𝕂, which we call the bilinear map associated with the pairing.[3] We may denote a pairing (X, Y, b) by b(X, Y) or and we may also denote b by (x, y) ↦ .

Notation: It is common practice to write instead of b(x, y), when no ambiguity can arise.

A dual system is a pairing (X, Y, b) satisfying the following two separation axioms:

  1. Y separates/distinguishes points of X: for all non-zero xX, there exists yY such that b(x, y) 0, and
  2. X separates/distinguishes points of Y: for all non-zero yY, there exists xX such that b(x, y) 0.

In this case we say that b is non-degenerate, we say that b places X and Y in duality (or in separated duality), and we call b the duality pairing or the canonical bilinear form of the duality.[3][1]

A subset S of Y is called total if for all x X, b(x, s) = 0 for all s S implies x = 0. A total subset of X is defined analogously: a subset S of X is called total if for all y Y, b(s, y) = 0 for all s S implies y = 0.

We say that the elements xX and yY are orthogonal and write x y if b(x, y) = 0. We say that two sets RX and SY are orthogonal and write R S if r and s are orthogonal for all rR and sS and we say that R is orthogonal to an element y Y if R is orthogonal to { y }. For R X, we define the orthogonal of R to be R := { y Y : R y }.

Dual definitions and results

Given a pairing (X, Y, b) we can define a new pairing (Y, X, ) where (y, x) := b(x, y).[3]

There is a repeating theme in duality theory, which is that any definition for a pairing (X, Y, b) has a corresponding dual definition for the pairing (Y, X, ).

Convention and Definition: Given any definition for a pairing (X, Y, b), one obtains a dual definition by applying it to the pairing (Y, X, ). If the definition depends on the order of X and Y (e.g. the definition of "the weak topology 𝜎(X, Y) defined on X by Y") then if we switch the order of X and Y then we mean that definition applied to (Y, X, ) (e.g. this gives us the definition of "the weak topology 𝜎(Y, X) defined on Y by X").

For instance, if we define "X distinguishes points of Y" (resp, "S is a total subset of Y") as above, then we immediately obtain the dual definition of "Y distinguishes points of X" (resp, "S is a total subset of X"). Once we define, for instance, 𝜎(X, Y) then we will automatically assume that 𝜎(Y, X) has been defined without mentioning the analogous definition. The same applies to many theorems.

Convention: Adhering to common practice, unless clarity is needed, whenever we give a definition (or result) for a pairing (X, Y, b) then we will omit mention the corresponding dual definition (or result) but nevertheless use it.

Identification of (X, Y) with (Y, X)

Although it is technically incorrect and an abuse of notation, we will also adhere to the following.

Convention: This article will use the common practice of treating a pairing (X, Y, b) interchangeably with (Y, X, ) and also denoting (Y, X, ) by (Y, X, b).

Polar sets

Suppose that (X, Y, b) is a pairing. The absolute polar or polar of a subset A of X is the set:[4]

.

Dually, the absolute polar or polar of a subset B of Y is denoted by and defined by

.

In this case, the absolute polar of a subset B of Y is also called the absolute prepolar or prepolar of B and may be denoted by .

The polar is necessarily a convex set containing 0 Y where if B is balanced then so is and if B is a vector subspace of X then so too is a vector subspace of Y.[5]

If A X then the bipolar of A, denoted by A∘∘, is A∘∘ := (A). Similarly, if B Y then the bipolar of B is B∘∘ := .

Competing definitions

Some authors define "polar sets" differently. For instance, some authors do not include absolute values around b(x, y) and/or use the real part of b (i.e. Re b(x, y)) in place of b(x, y). We now briefly discuss how these various definitions relate to one another.

If b is real-valued and A is symmetric (i.e. A = -A) then , where some authors use the right hand side to define the polar set. If b is complex-valued and A A for all real r then ,[3] where some authors use one of the latter two sets on the right hand side to define the polar set. Thus for all the common definitions of the polar set A to agree, it suffices that sA A for all scalars s of unit length (i.e. all scalars s such that |s| = 1). In particular, all definitions of the polar of A agree when A is a balanced set.

Examples

Restriction of a pairing

Suppose that (X, Y, b) is a pairing, M is a vector subspace of X, and N is a vector subspace of Y. Then the restriction of (X, Y, b) to M × N is the pairing . Note that if (X, Y, b) is a duality then it's possible for a restrictions to fail to be a duality (e.g. if Y { 0 } and N = { 0 }).

Convention: We will use the common practice of denoting the restriction by (M, N, b).

Canonical duality on a vector space

Suppose that X is a vector space and let denote the algebraic dual space of X (i.e. the space of all linear functionals on X). There is a canonical duality where is evaluation, which is called the natural or canonical bilinear functional on X × . This duality is often denoted by and it is the canonical duality.

If N is a vector subspace of then X distinguishes points of N so (X, M, c) is a duality if and only if N distinguishes points of X, or equivalently if M is total (i.e. n(x) = 0 for all n N implies x = 0).[3]

Suppose X is a topological vector space (TVS) with continuous dual space . Then the restriction of to X × defines a pairing for which X separates points of . If separates points of X (which is true if, for instance, X is a Hausdorff locally convex space) then this pairing forms a duality.[1]

Other examples

  • If (H, ) is a complex Hilbert space then forms a dual system if and only if dim H = 0. If H is non-trivial then doesn't even form pairing since the inner product is sesquilinear rather than bilinear.[3]
  • If (H, ) is a real Hilbert space then forms a dual system.
  • Suppose X = ℝ2, Y = ℝ3, and for all X and Y. Then (X, Y, b) is a pairing such that X distinguishes points of Y, but Y does not distinguish points of X. Furthermore, X := { y Y : X y } = { (0, 0, z) : z ℝ }
  • Let 1 < p < ∞, X := Lp(μ), Y := Lq(μ) (where q is such that 1/p + 1/q = 1), and b(f, g) := . Then (X, Y, b) is a dual system.
  • Let X and Y be vector spaces over the same field 𝕂. Then the bilinear form places X × Y and in duality.[1]
  • A sequence space X and its beta dual with the bilinear map defined as for x X, y forms a dual system.

Identification with a subspace of the algebraic dual

Suppose that (X, Y, b) is a pairing with X and Y being vector spaces over the field K. Denote the algebraic dual space of X by . For every y in Y, define the linear form b( , y) : XK by xb(x, y).[1]

Proposition: The following are equivalent:

  1. X distinguishes points of Y;
  2. The map yb( , y) defines an injection from Y into the algebraic dual space of X;[3]
  3. 𝜎(Y, X) is Hausdorff.[3]

If X distinguishes points of Y and if Z denotes the range of the injection yb( , y) then Z is a vector subspace of and the duality (X, Y, b) becomes canonically identified with the duality where is the natural evaluation map. In particular, Y can be identified as a vector subspace of X's algebraic dual and b can be replaced by the natural bilinear map.

Convention: Often, whenever yb( , y) is injective (especially when (X, Y, b) forms a dual pair) then we will use the common practice of assuming without loss of generality that Y is a vector subspace of the algebraic dual space of X, that b is the natural evaluation map, and we may also denote Y by .

In a completely analogous manner, if Y distinguishes points of X then it is possible for X to be identified as a vector subspace of Y's algebraic dual space.[1]

Weak topology

Suppose that (X, Y, b) is a pairing of vector spaces over 𝕂 and for all x X, let b(x, ) : Y 𝕂 be defined by yb(x, y) and for all y Y, let b( , y) : X 𝕂 be defined by xb(x, y). The weak topology on X induced by Y (and b) is the weakest TVS topology on X, denoted by 𝜎(X, Y, b) or simply 𝜎(X, Y), making all maps b( , y) : X 𝕂 continuous, as y ranges over Y.[3] We use X𝜎(X, Y, b), X𝜎(X, Y), or simply X𝜎 to denote X endowed with the weak topology 𝜎(X, Y, b). Similarly, we have the dual definition of the weak topology on Y induced by X (and b), which is of course denoted by 𝜎(Y, X, b) or simply 𝜎(Y, X): it is the weakest TVS topology on Y making all maps b(x, ) : Y 𝕂 continuous, as x ranges over X.[3] We also use the dual notation of Y𝜎(Y, X, b), Y𝜎(Y, X), or simply Y𝜎 to denote Y endowed with the weak topology 𝜎(Y, X, b).

Definition and notation: If we attach "𝜎(X, Y, b)" to a topological definition (e.g. 𝜎(X, Y, b)-converges, 𝜎(X, Y, b)-bounded, cl𝜎(X, Y, b)(S), etc.) then we mean that definition when the first space (i.e. X) carries the 𝜎(X, Y, b) topology. We may also omit mentioning b or even X and Y if no confusion will arise. So for instance, if we say that a sequence in Y "𝜎-converges" or "weakly converges" then this means that it converges in (Y, 𝜎(Y, X, b)) whereas if it were a sequence in X then this would mean that it converges in (X, 𝜎(X, Y, b)).

The topology 𝜎(X, Y) is locally convex since it is determined by the family of seminorms py : X ℝ defined by py(x) := |b(x, y)|, as y ranges over Y.[3] If x X and is a net in X, then we say that 𝜎(X, Y, b)-converges to x if converges to X in (X, 𝜎(X, Y)).[3] A net 𝜎(X, Y, b)-converges to x if and only if for all y Y, b(xi, y) converges to b(x, y). Note that if is a sequence of orthonormal vectors in Hilbert space, then converges weakly to 0 but does not norm-converge to 0 (or any other vector).[3]

If (X, Y, b) is a pairing and N is a proper vector subspace of Y such that (X, N, b) is a dual pair, then 𝜎(X, N, b) is strictly coarser than 𝜎(X, Y, b).[3]

Bounded subsets

If (X, Y, b) is a pairing and S X, then say that S is bounded if it is 𝜎(X, Y, b)-bounded, or equivalently, if < for all y Y.

Weakly complete

We call X 𝜎(X, Y, b)-complete or (if no ambiguity can arise) weakly-complete if (X, 𝜎(X, Y, b)) is a complete vector space. If is the algebraic dual of X then under the duality , is complete.[3] Note that there exist Banach spaces that are not weakly-complete (despite being complete).[3]

Weak representation theorem

The following theorem is of fundamental importance to duality theory.

Theorem (Weak representation theorem):[3] Let (X, Y, b) be a pairing over the field 𝕂. Then the continuous dual space of (X, 𝜎(X, Y, b)) is , where b( , y) : X 𝕂 is defined by xb(x, y). Furthermore,

  1. If f is a continuous linear functional on (X, 𝜎(X, Y, b)) then there exists some y Y such that f = b( , y); y is unique if and only if X distinguishes points of Y.
  2. The continuous dual space of (X, 𝜎(X, Y, b)) may be identified with Y/X, where X := { y Y : b(x, y) = 0 for all x X }.
    • Note that this is true regardless of whether or not X distinguishes points of Y or Y distinguishes points of X.

Polars and the weak topology

The following results are important for defining polar topologies. The bipolar theorem in particular "is an indispensable tool in working with dualities."[5]

Theorem: Let (X, Y, b) be a pairing and A X. Then

  1. The polar of A, A, is a closed subset of (Y, 𝜎(Y, X, b)).[3]
  2. The polars of the following sets are identical: (a) A; (b) the convex hull of A; (c) the balanced hull of A; (d) the 𝜎(X, Y, b)-closure of A; (e) the 𝜎(X, Y, b)-closure of the convex balanced hull of A;[3]
  3. The bipolar theorem: The bipolar of A, A∘∘, is equal to the 𝜎(X, Y, b)-closure of the convex balanced hull of A.[3]
  4. A is 𝜎(X, Y, b)-bounded if and only if A is absorbing in Y.[3]
  5. If in addition Y distinguishes points of X then A is 𝜎(X, Y, b)-bounded if and only if it is 𝜎(X, Y, b)-totally bounded.[3]

On topological vector spaces

Given a dual system , since Y is identified as a subspace of the algebraic dual of X, we can define on X the weak topology induced by Y, denoted by 𝜎(X, Y), as being the weakest TVS topology making all of the linear functionals in Y continuous. X endowed with this topology is denoted by X𝜎(X, Y) or X𝜎. If ℬ is a Hamel basis of Y then the topology 𝜎(X, Y) is generated by the seminorms as b ranges over ℬ.[5] The continuous dual space of X𝜎 is Y, meaning that if f is a linear form on X then f is continuous as a map on X𝜎 if and only if there exists some y Y such that for every x X.[5] If X is an infinite dimensional TVS then every weakly open neighborhood of the origin contains an infinite-dimensional vector subspace;[3] indeed, this vector subspace even has finite codimension in X.

Similarly, since X is identified as a subspace of the algebraic dual of Y, we can define on Y the weak topology induced by X, denoted by 𝜎(Y, X), as being the weakest TVS topology making all of the linear functionals in X continuous. Y endowed with this topology is denoted by Y𝜎. The continuous dual space of Y𝜎 is X, meaning that if f is a linear form on Y then f is continuous as a map on Y𝜎 if and only if there exists some x X such that for every y Y.[5] If X is a TVS and if Y is the continuous dual space of a X, then the weak topology 𝜎(Y, X) is called the weak-* topology.

If Y1 Y then the topology 𝜎(X, Y1) is coarser than 𝜎(X, Y) and furthermore, 𝜎(X, Y1) is strictly coarser than 𝜎(X, Y) if and only if Y1 Y.[5] In addition, the canonical bilinear form of places X and Y1 in duality if and only if Y1 is 𝜎(Y, X)-dense in Y.[5] So for instance, if X is a vector space and Y1 is a subspace of the algebraic dual then induces a duality between X and Y1 if and only if Y1 is -dense in .[5]

Note that if X is a TVS, then since X always distinguishes points on , is always Hausdorff.[3]

Polar sets

If X is any locally convex TVS, then the family of all barrels in X and the family of all subsets of that are convex, balanced, closed, and bounded in , correspond to each other by polarity (with respect to ).[5] It follows that a locally convex TVS X is barreled if and only if each bounded subset of is equicontinuous.[5]

Theorem Suppose that X is a separable TVS. Then every closed equicontinuous subset of is a compact metrizable space (under the subspace topology). If in addition X is metrizable then is separable.[5]

Here are some corollaries to the bipolar theorem:

  • For any subset S of X, .[5]
  • If is a family of 𝜎(X, Y)-closed subsets of X containing 0 X, then the polar of is the closed convex hull of .[5]
  • If X is a locally convex TVS then the polars (taken with respect to ) of any 0-neighborhood base forms a fundamental family of equicontinuous subsets of (i.e. given any bounded subset of , there exists a neighborhood S of 0 in X such that ).[5]
    • Conversely, if X is a locally convex TVS then the polars (taken with respect to ) of any fundamental family of equicontinuous subsets of form a neighborhood base of the origin in X.[5]
  • Let X be a TVS with a topology 𝜏. Then 𝜏 is a locally convex TVS topology if and only if 𝜏 is the topology of uniform convergence on the equicontinuous subsets of .[5] The last two results explain why equicontinuous subsets of the continuous dual space play such a prominent role in the modern theory of functional analysis: because equicontinuous subsets encapsulate all information about the locally convex space X's original topology.

Algebraic adjoint

Suppose that and are dual systems. If is a linear operator then we can define a linear operator by called the transpose or algebraic adjoint of u.[6] We say that the map is weakly continuous if the map is continuous, which happens if and only if , in which case the restriction of is continuous and .[6]

If A1 (resp. A2) is a subset of X1 (resp. X2) then

  • ;[6]
  • implies ;[6]
  • if A1 and A2 are convex, weakly closed sets containing 0 then implies .[6]

The kernel of is the subspace of Y2 orthogonal to the image of u.[6] The linear map u is injective if and only if the its image is a weakly dense subset of X2 (i.e. the image of u is dense in X2 when X2 is given the weak topology induced by Y2).[6]

Mackey topology

Given a dual system , since Y is identified as a subspace of the algebraic dual of X, we can define on X the Mackey topology induced by Y, denoted by 𝜏(X, Y), as being the strongest locally convex TVS topology such that the continuous dual space of (X, 𝜏) is Y. X endowed with this topology is denoted by X𝜏, its continuous dual space is Y, and it is a Mackey space.

Similarly, since X is identified as a subspace of the algebraic dual of Y, we can define on Y the Mackey topology induced by X, denoted by 𝜏(Y, X), as being the strongest locally convex TVS topology 𝜏 such that the continuous dual space of (Y, 𝜏) is X. Y endowed with this topology is denoted by Y𝜏 and its continuous dual space is X.

Mackey's theorem

Mackey's theorem is one of the central theorem in the theory of dual systems.

Theorem (Mackey) Let X be a Hausdorff locally convex TVS. Then the bounded subsets of X are the same for all Hausdorff locally convex TVS topologies on X compatible with the duality between X and .[7]

Lemma Let X be a Hausdorff locally convex TVS, B a barrel in X (i.e. a closed absorbing disk in X), and S a complete bounded disk in X. Then there exists a real r > 0 such that S rB.[7]

Comment

Associated with a dual pair is an injective linear map from X to given by

There is an analogous injective map from Y to .

In particular, if either of X or Y is finite-dimensional, these maps are isomorphisms.

See also

References

  1. Schaefer 1999, pp. 122–128.
  2. Schaefer 1999, p. 122.
  3. Narici 2011, pp. 225-273.
  4. Treves 2006, p. 195.
  5. Schaefer 1999, pp. 123–128.
  6. Schaefer 1999, pp. 128–130.
  7. Treves 2006, pp. 371–372.
  • Narici, Lawrence (2011). Topological vector spaces. Boca Raton, FL: CRC Press. ISBN 1-58488-866-0. OCLC 144216834.
  • Michael Reed and Barry Simon, Methods of Modern Mathematical Physics, Vol. 1, Functional Analysis, Section III.3. Academic Press, San Diego, 1980. ISBN 0-12-585050-6.
  • Rudin, Walter (1991). Functional analysis. McGraw-Hill Science/Engineering/Math. ISBN 978-0-07-054236-5.CS1 maint: ref=harv (link)
  • Schaefer, Helmut H.; Wolff, Manfred P. (1999). Topological Vector Spaces. GTM. 8. New York, NY: Springer New York Imprint Springer. ISBN 978-1-4612-7155-0. OCLC 840278135.CS1 maint: ref=harv (link)
  • Schmitt, Lothar M (1992). "An Equivariant Version of the Hahn–Banach Theorem". Houston J. Of Math. 18: 429–447.
  • Treves, Francois (2006). Topological vector spaces, distributions and kernels. Mineola, N.Y: Dover Publications. ISBN 978-0-486-45352-1. OCLC 853623322.CS1 maint: ref=harv (link)
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.