Dual norm

In functional analysis, the dual norm is a measure of size for a continuous linear function defined on a normed vector space.

Definition

Let be a normed vector space with norm and let be the dual space. The dual norm of a continuous linear functional belonging to is the non-negative real number defined[1] by any of the following equivalent formulas:

where and denote the supremum and infimum, respectively. The constant 0 map always has norm equal to 0 and it is the origin of the vector space If then the only linear functional on is the constant 0 map and moreover, the sets in the last two rows will both be empty and consequently, their supremums will equal instead of the correct value of 0.

The map defines a norm on (See Theorems 1 and 2 below.)

The dual norm is a special case of the operator norm defined for each (bounded) linear map between normed vector spaces.

The topology on induced by turns out to be as strong as the weak-* topology on

If the ground field of is complete then is a Banach space.

The double dual of a normed linear space

The double dual (or second dual) of is the dual of the normed vector space . There is a natural map . Indeed, for each in define

The map is linear, injective, and distance preserving.[2] In particular, if is complete (i.e. a Banach space), then is an isometry onto a closed subspace of .[3]

In general, the map is not surjective. For example, if is the Banach space consisting of bounded functions on the real line with the supremum norm, then the map is not surjective. (See space). If is surjective, then is said to be a reflexive Banach space. If then the space is a reflexive Banach space.

Mathematical optimization

Let be a norm on The associated dual norm, denoted is defined as

(This can be shown to be a norm.) The dual norm can be interpreted as the operator norm of , interpreted as a matrix, with the norm on , and the absolute value on :

From the definition of dual norm we have the inequality

which holds for all x and z.[4] The dual of the dual norm is the original norm: we have for all x. (This need not hold in infinite-dimensional vector spaces.)

The dual of the Euclidean norm is the Euclidean norm, since

(This follows from the Cauchy–Schwarz inequality; for nonzero z, the value of x that maximises over is .)

The dual of the -norm is the -norm:

and the dual of the -norm is the -norm.

More generally, Hölder's inequality shows that the dual of the -norm is the -norm, where, q satisfies , i.e.,

As another example, consider the - or spectral norm on . The associated dual norm is

which turns out to be the sum of the singular values,

where This norm is sometimes called the nuclear norm.[5]

Examples

Dual norm for matrices

The Frobenius norm defined by

is self-dual, i.e., its dual norm is

The spectral norm, a special case of the induced norm when , is defined by the maximum singular values of a matrix, i.e.,

has the nuclear norm as its dual norm, which is defined by

for any matrix where denote the singular values.

Some basic results about the operator norm

More generally, let and be topological vector spaces and let [6] be the collection of all bounded linear mappings (or operators) of into . In the case where and are normed vector spaces, can be given a canonical norm.

Theorem 1  Let and be normed spaces. Assigning to each continuous linear operator the scalar:

defines a norm on that makes into a normed space. Moreover, if is a Banach space then so is [7]

Proof

A subset of a normed space is bounded if and only if it lies in some multiple of the unit sphere; thus for every if is a scalar, then so that

The triangle inequality in shows that

for every satisfying This fact together with the definition of implies the triangle inequality:

Since is a non-empty set of non-negative real numbers, is a non-negative real number. If then for some which implies that and consequently This shows that is a normed space.[8]

Assume now that is complete and we will show that is complete. Let be a Cauchy sequence in so by definition as This fact together with the relation

implies that is a Cauchy sequence in for every It follows that for every the limit exists in and so we will denote this (necessarily unique) limit by that is:

It can be shown that is linear. If , then for all sufficiently large integers n and m. It follows that

for sufficiently all large m. Hence so that and This shows that in the norm topology of This establishes the completeness of [9]

When is a scalar field (i.e. or ) so that is the dual space of .

Theorem 2  For every define:

where by definition is a scalar. Then

  1. This is a norm that makes a Banach space.[10]
  2. Let be the closed unit ball of . For every
    Consequently, is a bounded linear functional on with norm
  3. is weak*-compact.
Proof

Let denote the closed unit ball of a normed space When is the scalar field then so part (a) is a corollary of Theorem 1. Fix There exists[11] such that

but,

for every . (b) follows from the above. Since the open unit ball of is dense in , the definition of shows that if and only if for every . The proof for (c)[12] now follows directly.[13]

See also

Notes

  1. Rudin 1991, p. 87
  2. Rudin 1991, section 4.5, p. 95
  3. Rudin 1991, p. 95
  4. This inequality is tight, in the following sense: for any x there is a z for which the inequality holds with equality. (Similarly, for any z there is an x that gives equality.)
  5. Boyd & Vandenberghe 2004, p. 637
  6. Each is a vector space, with the usual definitions of addition and scalar multiplication of functions; this only depends on the vector space structure of , not .
  7. Rudin 1991, p. 92
  8. Rudin 1991, p. 93
  9. Rudin 1991, p. 93
  10. Aliprantis 2006, p. 230
  11. Rudin 1991, Theorem 3.3 Corollary, p. 59
  12. Rudin 1991, Theorem 3.15 The Banach–Alaoglu theorem algorithm, p. 68
  13. Rudin 1991, p. 94

References

  • Aliprantis, Charalambos D.; Border, Kim C. (2006). Infinite Dimensional Analysis: A Hitchhiker's Guide (3rd ed.). Springer. ISBN 9783540326960.
  • Boyd, Stephen; Vandenberghe, Lieven (2004). Convex Optimization. Cambridge University Press. ISBN 9780521833783.
  • Kolmogorov, A.N.; Fomin, S.V. (1957). Elements of the Theory of Functions and Functional Analysis, Volume 1: Metric and Normed Spaces. Rochester: Graylock Press.
  • Narici, Lawrence; Beckenstein, Edward (2011). Topological Vector Spaces. Pure and applied mathematics (Second ed.). Boca Raton, FL: CRC Press. ISBN 978-1584888666. OCLC 144216834.
  • Rudin, Walter (1991). Functional Analysis. International Series in Pure and Applied Mathematics. 8 (Second ed.). New York, NY: McGraw-Hill Science/Engineering/Math. ISBN 978-0-07-054236-5. OCLC 21163277.
  • Schaefer, Helmut H.; Wolff, Manfred P. (1999). Topological Vector Spaces. GTM. 8 (Second ed.). New York, NY: Springer New York Imprint Springer. ISBN 978-1-4612-7155-0. OCLC 840278135.
  • Trèves, François (2006) [1967]. Topological Vector Spaces, Distributions and Kernels. Mineola, N.Y.: Dover Publications. ISBN 978-0-486-45352-1. OCLC 853623322.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.