CONTINUOUS LINEAR MAPPINGS AND DIFFERENTIALS - The Calculus of Variations - Advanced Calculus of Several Variables

Advanced Calculus of Several Variables (1973)

Part VI. The Calculus of Variations

Chapter 2. CONTINUOUS LINEAR MAPPINGS AND DIFFERENTIALS

In this section we discuss the concepts of linearity, continuity, and differentiability for mappings from one normed vector space to another. The definitions here will be simply repetitions of those (in Chapters I and II) for mappings from one Euclidean space to another.

Let E and F be vector spaces. Recall that the mapping φ : E → F is called linear if

Image

for all Image and Image.

Example 1 The real-valued function Image, defined on the vector space of all continuous functions on [a, b] by

Image

is clearly a linear mapping from Image.

If the vector spaces E and F are normed, then we can talk about limits (of mappings from E to F). Given a mapping f : E → F and Image, we say that

Image

if, given Image > 0, there exists δ > 0 such that

Image

The mapping f is continuous at Image if

Image

We saw in Section I.7 (Example 8) that every linear mapping between Euclidean spaces is continuous (everywhere). However, for mappings between infinite-dimensional normed vector spaces, linearity does not, in general, imply continuity.

Example 2 Let V be the vector space of all continuous real-valued functions on [0, 1], as in Example 2 of Section 1. Let E1 denote V with the 1-norm Image Image1, and let E0 denote V with the sup norm Image Image0. Consider the identity mapping λ : V → V as a mapping from E1 to E0,

Image

We inquire as to whether λ is continuous at Image (the constant zero function on [0, 1 ]). If it were, then, given Image > 0, there would exist δ > 0 such that

Image

However we saw in Section 1 that, given δ > 0, there exists a function φ : [0, 1 ] → Image such that

Image

It follows that λ : E1E0 is not continuous at Image.

The following theorem provides a useful criterion for continuity of linear mappings.

Theorem 2.1 Let L : E → F be a linear mapping where E and F are normed vector spaces. Then the following three conditions on L are equivalent:

(a)There exists a number c > 0 such that

Image

for all Image.

(b)L is continuous (everywhere).

(c)L is continuous at Image.

PROOF Suppose first that (a) holds. Then, given Image and Image > 0,

Image

so it follows that L is continuous at x0. Thus (a) implies (b).

To see that (c) implies (a), assume that L is continuous at 0, and choose δ > 0 such that Image implies ImageL(x)Image < 1. Then, given Image, it follows that

Image

so we may take c = 1/δ.

Image

Example 3 Let E0 and E1 be the normed vector spaces of Example 2, but this time consider the identity mapping on their common underlying vector space V as a linear mapping from E0 to E1,

Image

Since

Image

we see from Theorem 2.1 (with c = 1) that μ is continuous.

Thus the inverse of a one-to-one continuous linear mapping of one normed vector space onto another need not be continuous. Let L : E → F be a linear mapping which is both one-to-one and surjective (that is, L(E) = F). Then we will call L an isomorphism (of normed vector spaces) if and only if both L and L−1 are continuous.

Example 4 Let A, B : [a, b] → Imagen be continuous paths in Imagen, and consider the linear function

Image

defined by

Image

where the dot denotes the usual inner product in Imagen. Then the Cauchy-Schwarz inequality yields

Image

so an application of Theorem 2.1, with c = (b − a)(ImageAImage0 + ImageBImage0), shows that L is continuous.

We are now prepared to discuss differentials of mappings of normed vector spaces. The definition of differentiability, for mappings of normed vector spaces, is the same as its definition for mappings of Euclidean spaces, except that we must explicitly require the approximating linear mapping to be continuous. The mapping f : E → F is differentiable at Image if and only if there exists a continuous linear mapping L : E → F such that

Image

The continuous linear mapping L, if it exists, is unique (Exercise 2.3), and it is easily verified that a linear mapping L satisfying (1) is continuous at 0 if and only if f is continuous at x (Exercise 2.4).

If f : E → F is differentiable at Image, then the continuous linear mapping which satisfies (1) is called the differential of f at x, and is denoted by

Image

In the finite-dimensional case E = Imagen, F = Imagem that we are already familiar with, the m × n matrix of the linear mapping dfx is the derivative f′(x). Here we will be mainly interested in mappings between infinite-dimensional spaces, whose differential linear mappings are not representable by matrices, so derivatives will not be available.

Example 5 If f : E → F is a continuous linear mapping, then

Image

for all Image with h ≠ 0, so f is differentiable at x, with dfx = f. Thus a continuous linear mapping is its own differential (at every point Image).

Example 6 We now give a less trivial computation of a differential. Let g : ImageImage be a Image function, and define

Image

by

Image

We want to show that f is differentiable, and to compute its differential.

If f is differentiable at Image, then dfφ(h) should be the linear (in Image part of f(φ + h) − f(φ). To investigate this difference, we write down the second degree Taylor expansion of g at φ(t):

Image

where

Image

for some ξ(t) between φ(t) and φ(t) + h(t). Then

Image

where Image is defined by

Image

It is clear that Image is a continuous linear mapping, so in order to prove that f is differentiable at φ with dfφ = L, it suffices to prove that

Image

Note that, since g is a Image function, there exists M > 0 such that Imageg″(ξ(t))Image < 2M when ImagehImage0 is sufficiently small (why?). It then follows from (2) that

Image

and this implies (3) as desired. Thus the differential Image of f at φ is defined by

Image

The chain rule for mappings of normed vector spaces takes the expected form.

Theorem 2.2 Let U and V be open subsets of the normed vector spaces E and F respectively. If the mappings f : U → F and g : V → G (a third normed vector space) are differentiable at Image and Image respectively, then their composition h = g Image f is differentiable at x, and

Image

The proof is precisely the same as that of the finite-dimensional chain rule (Theorem II.3.1), and will not be repeated.

There is one case in which derivatives (rather than differentials) are important. Let φ : ImageE be a path in the normed vector space E. Then the familiar limit

Image

if it exists, is the derivative or velocity vector of φ at t. It is easily verified that φ′(t) exists if and only if φ is differentiable at t, in which case t(h) = φ′(t)h (again, just as in the finite-dimensional case).

In the following sections we will be concerned with the problem of minimizing (or maximizing) a differentiable real-valued function f : EImage on a subset M of the normed vector space E. In order to state the result which will play the role that Lemma II.5.1 does in finite-dimensional maximum–minimum problems, we need the concept of tangent sets. Given a subset M of the normed vector space E, the tangent set TMx of M at the point Image is the set of all those vectors Image, for which there exists a differentiable path Image such that φ(0) = x and φ′(0) = v. Thus TMx is simply the set of all velocity vectors at x of differentiable paths in M which pass through x. Hence the tangent set of an arbitrary subset of a normed vector space is the natural generalization of the tangent space of a submanifold of Imagen.

The following theorem gives a necessary condition for local maxima or local minima (we state it for minima).

Theorem 2.3 Let the function f : EImage be differentiable at the point x of the subset M of the normed vector space E. If Image for all Image sufficiently close to x, then

Image

That is, dfx(v) = 0 for all Image.

PROOF Given Image, let φ : ImageM be a differentiable path in E such that φ(0) = x and φ′(0) = v. Then the composition g = f Image φ : ImageImage has a local minimum at 0, so g′(0) = 0. The chain rule therefore gives

Image

as desired.

Image

In Section 4 we will need the implicit mapping theorem for mappings of complete normed vector spaces. Both its statement and its proof, in this more general context, are essentially the same as those of the finite-dimensional implicit mapping theorem in Chapter III.

For the statement, we need to define the partial differentials of a differentiable mapping which is defined on the product of two normed vector spaces E and F. The product set E × F is made into a vector space by defining

Image

If Image ImageE and Image ImageF denote the norms on E and F, respectively, then

Image

defines a norm on E × F, and E × F is complete if both E and F are (see Exercise 1.4). For instance, if E = F = Image with the ordinary norm (absolute value), then this definition gives the sup norm on the plane Image × Image = Image2(Example 1 of Section 1).

Now let the mapping

Image

be differentiable at the point Image. It follows easily (Exercise 2.5) that the mappings

Image

defined by

Image

are differentiable at Image and Image, respectively. Then dxf(a, b) and dyf(a, b), the partial differentials of f at (a, b), with respect to Image and Image, respectively, are defined by

Image

Thus dxf(a, b) is the differential of the mapping E → G obtained from f : E × F → G by fixing y = b, and dyf(a, b) is obtained similarly by holding x fixed. This generalizes our definition in Chapter III of the partial differentials of a mapping from Imagem+n = Imagem × Imagen to Imagek.

With this notation and terminology, the statement of the implicit mapping theorem is as follows.

Implicit Mapping Theorem Let f : E × F → G be a Image mapping, where E, F, and G are complete normed vector spaces. Suppose that f(a, b) = 0, and that

Image

is an isomorphism. Then there exists a neighborhood U of a in E, a neighborhood W of (a, b) in E × F, and a Image mapping φ : U → F such that the following is true: If Image and Image, then f(x, y) = 0 if and only if y = φ(x).

This statement involves Image mappings from one normed vector space to another, whereas we have defined only differentiable ones. The mapping g : E → F of normed vector spaces is called continuously differentiable, or Image, if it is differentiable and dgx(v) is a continuous function of (x, v), that is, the mapping (x, y) → dgx(v) from E × E to F is continuous.

As previously remarked, the proof of the implicit mapping theorem in complete normed vector spaces is essentially the same as its proof in the finite-dimensional case. In particular, it follows from the inverse mapping theorem for complete normed vector spaces in exactly the same way that the finite-dimensional implicit mapping theorem follows from the finite-dimensional inverse mapping theorem (see the proof of Theorem III.3.4). The general inverse mapping theorem is identical to the finite-dimensional case (Theorem III.3.3), except that Euclidean space Imagen is replaced by a complete normed vector space E. Moreover the proof is essentially the same, making use of the contraction mapping theorem. Finally, the only property of Imagen that was used in the proof of contraction mapping theorem, is that it is complete. It would be instructive for the student to reread the proofs of these three theorems in Chapter III, verifying that they generalize to complete normed vector spaces.

Exercises

2.1Show that the function Image is continuous on Image.

2.2If L : ImagenE is a linear mapping of Imagen into a normed vector space, show that L is continuous.

2.3Let f : E → F be differentiable at Image, meaning that there exists a continuous linear mapping L : E → F satisfying Eq. (1) of this section. Show that L is unique.

2.4Let f : E → F be a mapping and L : E → F a linear mapping satisfying Eq. (1). Show that L is continuous if and only if f is continuous at x.

2.5Let f : E × F → G be a differentiable mapping. Show that the restrictions φ : E → G and ψ : F → G of Eq. (4) are differentiable.

2.6If the mapping f : E × F → G is differentiable at Image, show that dxfp(r) = dfp(r, 0) and dyfp(s) = dfp(0, s).

2.7Let Image be a translate of the closed subspace V of the normed vector space E. That is, given Image. Then prove that TMx = V.