3
Hilbert Spaces for Engineers

David Hilbert is considered to be one of the most influential mathematicians of the 20th Century. He was a German mathematician, educated at Konigsberg, where he completed his PhD thesis under the direction of Ferdinand Lindemann. At the same time, Hermann Minkowski was also a PhD student of Lindemann – they presented their PhD’s in the same year (1885) and thereafter became good friends. Hilbert was acquainted with some of the best mathematicians of his time and contributed significantly to important mathematical problems. He also supported the candidacy of Emmy Noether to teach at the University of Gottingen – at these times, this university was considered one of the best mathematical centers of the world and women had difficulty in finding academic positions. The history of mathematics teems with anecdotes and quotes of Hilbert. At the 2nd International Congress of Mathematics held in Paris, in 1900, Hilbert presented a speech where he evoked unsolved problems. His speech was transcribed in a long article, first published in German, but which was quickly translated into English [WIN 02] and is known globally as the 23 Hilbert problems or the Hilbert program – this text has widely influenced Mathematics along the 20th Century. The Hilbert program addressed fundamental difficulties of Mathematics – some of them remain unresolved to this day – and their study led to a large development of new ideas and methods. For instance, the research connected to the second problem (the consistency of arithmetic) have led to the denial of any divine status to Mathematics and brought it to the dimension of a human activity based on human choices, biases and beliefs. The particularity of Mathematics remains in its quest for consistency and its extremely codified way to build results by logical deductions and demonstrations based on previous results. Mathematicians make choices by establishing axioms, which are the foundations of their buildings. Then, they patiently construct the whole structure by exploring the consequences of the axioms and generating results which are new bricks that make the building higher. The question of the compatibility of all the bricks among themselves is difficult and leads to the analysis of quality of the foundations, which is connected to the coherence of the results obtained, which may be weakened by paradoxes found in the development of the theory. The works of Kurt Godel [GOD 31] and Alfred Tarski [TAR 36] tend to show that we cannot be sure of the global compatibility of all the bricks and, so, of the coherence of the structure. But this apparent weakness transforms into force, since it gives power to imagination and total freedom for the construction of alternate theories, based on new axioms, which makes the possible developments of Mathematics virtually inexhaustible. This has led, for instance, to new geometries which have found applications in Physics and new function theories which have found their application in Physics and Engineering.

Ironically, Hilbert spaces were not formalized by Hilbert himself, but by other people. The three major contributions came from Maurice Fréchet [FRE 06], Stefan Banach [BAN 22] and John Von Neumann [VON 32]. Fréchet established a formalism based on the idea of distance in his PhD Thesis, prepared under the direction of Jacques Hadamard and presented in Paris in 1905. Banach presented the formalism of normed spaces in his PhD (1920 at Lvov), under the direction of Hugo Steinhaus – a former student of Hilbert. Banach also published an article presenting the formalization of normed spaces. Von Neumann worked with Hilbert and stated the axiomatic presentation of Hilbert spaces in his book. The original paper by Hilbert (the fourth of the series of six works on integral equations published between 1904 and 1910 [HIL 04a, HIL 04b, HIL 05, HIL 06a, HIL 06b, HIL 10, HIL 12] contains all the papers) did not present a formal theory and was limited to square summable functions [WEY 44].

In order to take the full measure of the importance of this work and the extent of its applications, the reader must recall that one of the main reasons for the existence of Mathematics is the need for the determination of the solutions of equations and the prediction of the evolution of systems or the determination of some of their parameters (namely, for safety or obtaining a desired result). Human activities often involve the generation of data, its analysis and conclusions about the future from observations or the determination of some critical parameters from data. This work is based on models representing the behavior of the system under analysis. However, the models involve equations that must be solved, either to generate the model or to determine the evolution of the system. For instance, Engineering and Physics use such an approach. Many of the interesting situations involve the solution of various types of equations (algebraic, differential, partial differential, integral, etc.). When studying integral, differential or partial differential equations, a major problem has to be treated, connected to the infinity: while the equation ximg and f(x) = 0 concerns a single value of x, the equation x(t) ∈ img and f(x(t),t) = 0, t ∈ (a, b) concerns infinite values of x (one for each t). Analogously, integral, differential or partial differential equations concern the determination of infinitely many values. Since the determination of infinite actions takes an infinite time, procedures like limits and approximations have been introduced. Instead of achieving an infinite amount of solutions, it seems more convenient to determine close approximations of the real solution by achieving a finite number of solutions.

The practical implementation of this simple idea begs answering some questions. For instance: what is the sense of “close”, i.e. who are the neighbors at a given distance of an element? How do we ensure that the approximations are close to “the real solution”, i.e. that there are neighbors of the real solution at an arbitrarily small distance? How do we obtain a sequence of solutions which are “closer and closer (i.e. convergent) to the real solution”, i.e. belonging to arbitrarily small neighborhoods of the real solution? Moreover, what is a solution? Mathematically, these questions may be formulated as follows: what is the set (x; img) = {y: distance between y and x≤ ε}? How do we ensure that an approximation images verifies images How do we generate an element xε img V(x; ε) for an arbitrarily small ε > 0? By taking ε = 1/n and denoting xn = xε, this last question reads as: how to generate an element xn such that distance between xn and x ≤ 1/n, i.e. how do we generate a sequence {xn: n img img*} which converges to x? How do we define x into a convenient way?

In the mathematical construction, elements such as the neighborhoods, the convergent sequences and approximations are defined by the topology of the space.

It is quite intuitive that we are interested in the construction of a theory and methods leading to the solution of the largest possible number of equations: for practical reasons, we want to be able to solve as many equations as possible. This means that we want to be able to generate the largest number of converging sequences {xn: nimg*} or, equivalently, that our main interest is the construction of spaces tending to include in the neighborhoods of given functions as many elements as possible (i.e. V(x; ε) is as large as possible).

Increasing the members of a neighborhood, of solvable equations or of convergent sequences consists of weakening a topology: weaker topologies have more convergent sequences, more solvable equations and more elements in a neighborhood. But weaker topologies also introduce weaker regularity and a more complex behavior of the functions, so that a convenient topology establishes a balance between regularity and solvability.

In order to generate convenient topologies, we observe that Mathematics is constructed into a hierarchical way: the starting level is Set Theory, which defines objects that are mere collections and only involves operations such as intersection, union and difference.

In the next level, we consider operations involving the elements of the set itself: groups, rings, vectorial spaces. At this level, we have no concept of “neighbour”, “convergence” or “approximation”. In order to give a mathematical sense to these expressions, we need a topology.

A convenient way to define a topology consists of the introduction of a tool measuring the distance between two objects — i.e. a metric d(u, v)giving the distance between two elements u, v of the set (this was the work of Fréchet): then, we may define neighborhoods of u by considering the elements v such that d(u, v) ≤ ε, where ε is a small parameter: V(u; ε) = {v: d (u, v) ≤ ε}.

We may use a simple way to measure distances, which consists of using a norm imgimgimg measuring the “length” of the elements of V: d(u, v) = imgu − vimg, i.e. the distance between u and v is measured as the length of the difference between the elements u, v (this was the work of Banach).

The next level consists of introducing a measure of the angle between the elements u, v − this is made by using the scalar product (u, v) and was the work of Von Neumann. We will see that such a structure has more complex neighbourhoods and leads to a larger number of converging sequences. Thus, this last level leads to more converging sequences, more complete neighborhoods and gives at the same time measures of angles, lengths and distances. It corresponds to Hilbert Spaces and furnishes an interesting framework for the definition of convenient topologies.

3.1. Vector spaces

Vector spaces are a fundamental structure for variational methods. The definition of a vector space assumes that two sets were defined: a set Λ of scalars and a set V of vectors. The set of scalars is usually the set of real numbers img or the set of complex numbers img — even if, in general, we may consider any field, i.e. any algebraical structure having operations of addition and multiplication with their inverses (subtraction and division). In the sequel, we consider vector spaces on img, i.e. we assume that Λ = img. However, in a general approach, V may appear as arbitrary, but in our applications V will be a functional space, i.e. the elements of V will be considered as functions. The formal definition of a vector space on img (resp. Λ) assumes that V possesses two operations: addition u + v of two elements u, v ∈ and multiplication λu of an element u of V by a scalar λ∈ img (respectively, Λ). Then,

DEFINITION 3.1.– Let V be such that:

images

Then, V is a vector space on img.

img

We have:

THEOREM 3.1.— Let V be a vector space space on img and SV. S is a vector space on img if and only if 0 ∈ S and λu + vV, ∀λimg u, vS. ∈ In this case, S is said to be a vector subspace of V.

PROOF.– We have(−l)u + 0 = −u(see exercises). Thus, −uS, ∀ uS. Thus, all the conditions of definition 3.1 are satisfied by S.

EXAMPLE 3.1.— Let us consider Ω ⊂ imgn and V = {v: imgk}. V is a vector space on img (see exercises). Thus any subset of V such that 0 ∈ S and λu + vV, ∀λimg, u, vS is also a vector space on img. For instance:

images

are both vector subspaces of V.

img

DEFINITION 3.2.— Let F = {φλ: λ ∈ Λ} ⊂ V, Λ ≠ ø, F ≠ ø. F is linearly independent (or free) if and only if a finite linear combination images of its elements is null if and only if all the coefficients ai are null

images

We denote by [F ] the set of the finite linear combinations of elements of F:

images

We say that F is a generator (or that F spans V) if and only if [F] = V. We say that F is a basis of F if and only if F is free and Fspans V. The number of elements of the basis is the dimension of V: if the basis contains n elements, we say that V is n-dimensional (ordim (V) = n); if V contains infinitely many elements, we say that V is infinite dimensional (or dim(V) = +∞).

img

Any vector space containing at least one element non null has a basis (see [SOU 10]). In the following, we are particularly interested in countable families F, i.e. in the situations where Λ = img*. We will extend the notion of “basis” to the notion of “Hilbert basis”, where we have not if [F] = V, but V is the limit of elements of [F].

3.2. Distance, norm and scalar product

The structure of a vector space does not contain elements for the definition of neighbors, i.e. does not allow us to give a sense to the expressions “close to” and “convergent”. In order to give a practical content to these ideas, we must introduce new concepts: the distance between elements of V, the norm of an element of V, the scalar product of elements of V.

3.2.1. Distance

The notion of distance allows the definition of neighbors and quantifies as far or as close are two elements.

DEFINITION 3.3.– Let d: V × Vimg be a function such that

images

Then, d is a distance on V.

As previously observed, we are not interested in the numeric value of a distance, but in the definition of families of neighbors in order to approximate solutions and study convergence of sequences. For such a purpose, the numeric value of a distance is not useful as information. For instance, if d is a distance on V and η: imgimg is a strictly increasing application such that η(0) = 0 and η(a + b) ≤ η(a) + η(b), then dη,(u, v) = η(d(u, v)) is a distance on V (see exercises). Thus, we may give an arbitrary value to the distance between two elements of V. Moreover, distances may be pathological: for instance, the defination

images

corresponds to a distance. This distance is not useful, since the small neigborhoods of u contain a single element, which is u itself: d(u, v)≤ ε ≤1 ⇒ v = u. Thus, V(u; ε) = {u} and the associated topology contains just a few converging sequences and neighborhoods are poor in elements.

Distances allow the definition of limits and continuity. Since our presentation focuses on Hilbert spaces, these notions are introduced later, when considering norms.

3.2.2. Norm

The notion of norm allows the definition of the length of a vector and, at the same time, may be used in order to define a distance. In Mechanics, norms may also be interpreted in terms of internal energy of a system.

DEFINITION 3.4.– Let img img img : Vimg be a function such that

images

Then, img img img is a norm on V.

img

Classical examples of norms on V = imgn are (x = (x1, …, x„)):

images

THEOREM 3.2.– Let img img img be a norm on V. Then d(u, v) = img u − v imgis a distance on V.

img

PROOF.– Immediate.

img

When considering functional spaces, such as subspaces of V = {v: Ω ⊂ imgnimgk} continuity may be an essential assumption in order to define norms. For instance, let us consider pimg* and

images

Let S = {v ∈ V: v is continuous on Ω and imgvimg < ∞ }, then img img img is a norm on S. The assumption of continuity is used in order to show that

images

Indeed, for v continuous on Ω:

images

If x ∈ Ω is such that imgv(x)imgp > 0, the, by continuity, there is ε > 0 such that imgv(y)imgp>0 on Bε(x) = {y ∈ Ω: imgy − ximg ≤ ε} (Corollary 3.1). Then,

images

and we have 0 > 0, what is a contradiction. The classical counterexample establishing that continuity must be assumed is furnished by a function v such that v(x) = 0 everywhere on Ω, except on a finite or enumerable number of points. For instance,

images

Here, img v img = 0, but v≠ 0. In the following, we show how to we lift this restriction by using equivalence relations that generate the functional space Lp(Ω) − there will be a price to pay: it will be the loss of the concept of value at a point: the individual value v (x) of v in a point x has no meaning for elements of Lp(Ω).

As an alternative norm, we may also consider

images

In this case, the assumption of continuity is not necessary. Indeed,

images

3.2.3. Scalar product

The notion of scalar product allows the definition of angles between vectors and, at same time, may be used in order to define a norm. In Mechanics, scalar products may also be interpreted as the virtual work of internal efforts.

DEFINITION 3.5.– Let (img, img): V × Vimg be a function such that

images

Then, (img, img) is a scalar product on V.

img

A classical example of scalar product on V = imgn is (x = (x1, …, xn), y = (y1, …, yn))

images

THEOREM 3.3.– Let (img, img): V × Vimg be a scalar product on V. Then, ∀u, v, ∈ V, λ ∈ img,

  1. i) img
  2. ii) img
  3. iii) img
  4. iv) img
  5. v) img
  6. vi) img

img

PROOF— We have

images

so that,

images

and we have (i). Thus:

images

By adding these two equalities, we have (ii).

Let f(λ) = (u + λv, u + λv). We have f(λ) ≥ 0, ∀λ ≥ 0. Thus,

images

and we have (iii).

Since

images

We have

images

what implies (iv). Replacing v by − v, we obtain (v).

Finally, we observe that

images

so that (from(iii))

images

and we have (vi).

img

THEOREM 3.4.— Let (img, img): V × Vimg be a scalar product on V. Then, images is a norm on V.

img

PROOF — we have

images

Finally, 3.2.5. (v) implies that imgu + vimgimguimg + imgvimg.

Functional spaces, such as subspaces of images generally involve continuity assumptions in order to define scalar products. Analogously to norms, the assumption of continuity is necessary in order to verify that

images

For instance, if

images

and S = {vV: v is continuous on Ω and (y, v) < ∞}, then (img, img) defines a scalar product on S. Analagously to the case of norms (use p = 2),

images

In the following, continuity assumption is lifted by using functional space L2(Ω). As previously observed, the consequence is the loss of the meaning of point values v(x)

This limitation may be avoided in particular situations where the scalar product implies the continuity. For instance, let us consider,

images

where A: B = Σi, j Aij Bij. In this case, the continuity of the elements u, v yields from , the existence of the gradients ∇u, ∇v and we may give a meaning to the punctual values u(x) and (x) .

DEFINITION 3.6.— The angle θ(u, v) between the vectors u and v is

images

u and v are orthogonal if and only if (u, v) = 0 Notation: uv. For S ⊂ V, we denote by S the set of the elements of V which are orthogonal to all the elements of S:

images

img

DEFINITION 3.7.– Let F = {φn : n ∈ Λ ⊂ img*} be a countable family. We say that F is an orthogonal family if and only if

images

We say that F is an orthonormed family if and only if it is orthogonal and, moreover,

images

img

A countable free family G = {øn: n ∈ Λ ⊂ img*} may be transformed into an orthonomal family F = {φn: n ∈ Λ ⊂ img*} by the Gram–Schmidt procedure, which has been created for the solution of least-square linear problems. It has been proposed by Jorgen Pedersen Gram [GRA 83] and formalized by Ehrlich Schmidt [SCM 07] (although early uses by Pierre Simon de Laplace have been noticed [LEO 13]). The Gram–Schmidt procedure reads as:

  1. 1) Set
    images
  2. 2) For k > 0, set
    images

In this case,

images

The practical implementation requests the evaluation of scalar products and norms.

As an example, let us consider Ω = (−1, 1) and

images

The family øn(x) = xn is not orthonormal, since (mod(p, q) denotes the remainder of the integer division p/q)

images

If the Gram–Schmidt procedure is applied, we have:

images

3.2.4. Cartesian products of vector spaces

In practice, we often face situations where V = V1 × … Vd is the Cartesian product of vector spaces. In this case, elements u, images are given by:

images

In such a situation:

  • – If the distance on Vi is di, then the distance d„ on V may be defined as:
    images
  • Analogously, if the norm on Vt is img img img , then the norm imgimgimgv on V may be dfined as:
    images
  • – Finally, if the scalar product on Vi is (img, img)i, then the scalar product (img, img)v on V may be defined as:
    images

    When images, we have images and

    images

3.2.5. A Matlab® class for scalar products and norms

Below we give an example of a Matlab® class for the evaluation of scalar products and norms. The class contains two scalar products:

images

Evaluated by methods sp0 and sp1, respectively. The associated norms are evaluated by methods n0 and n1, respectively. For each method, the functions are furnished by structured data. In both the cases, the structure has as properties dim, dimx and component. dim and dimx are integer values defining the dimension of the function and the dimension of its argument x. component is a cell array of length dim containing the values of the components and their gradients: each element of component is a structure having properties value and grad. When using ‘subprogram’, value and grad are Matlab subprograms (i.e. anonymous functions). For ‘table’, value is an array of values and grad is a cell arrays containing tables of values: u.component{i}.value is a table of values of component ui; and u. component {i} .grad {j} is a table of values of ∂ui/∂xj.

For ‘subprogram’, parameter xl contains the limits of the region, analogous to those used in class subprogram_integration. For ‘table’, xl is a structure containing the points of integration, analogously to those used in class riemann.

image
image

Program 3.1. A class for the determination of partial solutions of linear systems

EXAMPLE 3.2.– Let us evaluate both the scalar products and norms for the functions u, v: (0, 1) → img given by u(x) = x, v(x) = x2. The functions may be defined as follows:

image

Program 3.2. Definition of u and v in example 3.2

The code:


xlim.lower.x = 0;
xlim.upper.x = 1;
xlim.dim = 1;
u = u1();
v = v1();
sp = scalar_product();
sp0uv = sp.sp0(u,v, xlim,‘subprogram’);
n0u = sp.n0(u,xlim,‘subprogram’);
n0v = sp.n0(v,xlim,‘subprogram’);
sp1uv = sp.sp1(u,v, xlim,‘subprogram’);
n1u = sp.n1(u,xlim,‘subprogram’);
n1v = sp.n1(v,xlim,‘subprogram’);

produces the results sp0uv = 0.25, n0u = 0.57735, n0v = 0.44721, sp1uv = 1.25, n1u = 1.1547, n1v = 1.2383.

The exact values are, respectively, images images

We generate cell arrays containing tables of values by using:


x = 0:0.01:1;
p.x = x;
p.dim = 1;
U.dim = 1;
U.dimx = 1;
U.component = cell(u.dim,1);
U.component{1}.value =
spam.partition(u.component{1}.value,p);
U.component{1}.grad = cell(u.dim,1);
U.component{1}.grad{1} =
spam.partition(u.component{1}.grad,p);
V.dim = 1;
V.dimx = 1;
V.component = cell(v.dim,1);
V.component{1}.value =
spam.partition(v.component{1}.value,p);
V.component{1}.grad = cell(v.dim,1);
V.component{1}.grad{1} =
spam.partition(v.component{1}.grad,p);

Then, the code:


sp0UV = sp.sp0(U,V,p,'table');
n0U = sp.n0(U,p,'table');
n0V = sp.n0(V,p,'table');
sp1UV = sp.sp1(U,V,p,'table');
n1U = sp.n1(U,p,'table');
n1V = sp.n1(V,p,'table');

produces the results sp0UV = 0.25003, n0U = 0.57736, n0V = 0.44725, sp1UV = 1.25, n1U = 1.1547, n1V = 1.2383.

img

EXAMPLE 3.3.— Let us evaluate both the scalar products and norms for the functions u, v: (0, 1) × (0, 2) → img given by images images. The functions may be defined as follows:

image

Program 3.3. Definition of u and v in example 3.3

The code:


xlim.lower.x = 0;
xlim.upper.x = 1;
xlim.lower.y = 0;
xlim.upper.y = 2;
xlim.dim = 2;
u = u2();
v = v2();
sp = scalar_product();
sp0uv = sp.sp0(u,v, xlim,'subprogram');
n0u = sp.n0(u,xlim,'subprogram');
n0v = sp.n0(v,xlim,'subprogram');
sp1uv = sp.sp1(u,v, xlim,'subprogram');
n1u = sp.n1(u,xlim,'subprogram');
n1v = sp.n1(v,xlim,'subprogram');

produces the results sp0uv = 1.8333, n0u = 0.94281, n0v = 2.0976, sp1uv = 4.8333, n1u = 2.0548, n1v = 3.0111.

The exact values are, respectively img

We use the function


function v = select_index(ind,u,x)
aux = u(x);
v = aux(ind);
return;
end

and we generate cell arrays containing tables of values by the code below:


x = 0:0.01:1;
y = 0:0.01:2;
p.x = x;
p.y = y;
p.dim = 2;
U.dim = 1;
U.dimx = 2;
U.component = cell(u.dim,1);
U.component{1}.value =
spam.partition(u.component{1}.value,p);
U.component{1}.grad = cell(u.dim,1);
V.dim = 1;
V.dimx = 2;
V.component = cell(u.dim,1);
V.component{1}.value =
spam.partition(v.component{1}.value,p);
V.component{1}.grad = cell(u.dim,1);
for i = 1: u.dim
  for j = 1: p.dim
    du = @(x) select_index(j,u.component{i}.grad,x);
    U.component{i}.grad{j} = spam.partition(du,p);
    dv = @(x) select_index(j,v.component{i}.grad,x);
    V.component{i}.grad{j} = spam.partition(dv,p);
  end;
end
sp0UV = sp.sp0(U,V,p,'table');
n0U = sp.n0(U,p,'table');
n0V = sp.n0(V,p,'table');
sp1UV = sp.sp1(U,V,p,'table');
n1U = sp.n1(U,p,'table');
n1V = sp.n1(V,p,'table');

It produces the results sp0UV = 1.8334, n0U = 0.94284, n0V = 2.0977, sp1UV = 4.8334, n1U = 2.0548, n1V = 3.0111.

img

3.2.6. A Matlab® class for Gram–Schmidt orthonormalization

The class defined in the preceding section may be used in order to transform a general family G = {øi: 1 ≤ i ≤ n} into an orthonomal family F = {φi: 1 ≤ i ≤ n} by the Gram–Schmidt procedure.

Let us assume that the family G is defined by a cell array basis such that basis{i} defines øi, according to the standard defined in section 3.2.5 : basis{i} is a structure having properties dim, dimx and component. component is a structure having fields value and grad. The class given in program 3.4 generates a cell array hilbert_basis defining the family F, resulting from the application _ the Gram–Schmidt’s procedure to G. The class uses subprograms sp_space and n_space, which are assumed to evaluate the scala product and the norm.

image
image
image

Program 3.4. A class for the Gram-Schmidt orthogonalization

EXAMPLE 3.4.— Let us consider the family øi(x) = xi−1

image

Program 3.5. Definition of the family G in example 3.4

The family is created by the code


basis = cell(n,1);
for i = 1: n
  basis{i} = u3(i);
end;

We apply the procedure on (−1, 1) with the scalar product sp0:


xlim.lower.x = -1;
xlim.upper.x = 1;
xlim.dim = 1;
sp_space = @(u,v)
scalar_product.sp0(u,v,xlim,'subprogram');
n_space = @(u)
scalar_product.n0(u,xlim,'subprogram');
hb =
gram_schmidt.subprograms(basis,sp_space,n_space);

The results are shown in Figure 3.1 below:

image

Figure 3.1. Results for example 3.4 (using sp0 and ‘subprograms’). For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

We generate discrete data associated with these functions by using the code:


x = -1:0.02:1;
p.x = x;
p.dim = 1;
tb = cell(n,1);
for i = 1: n
  u = u3(i);
  U.dim = 1;
U.dimx = 1;
  U.component = cell(u.dim,1);
  U.component{1}.value =
spam.partition(u.component{1}.value,p);
  U.component{1}.grad = cell(u.dim,1);
  U.component{1}.grad{1} =
spam.partition(u.component{1}.grad,p);
  tb{i} = U;
end;

Then, the commands


sp_space = @(u,v)
scalar_product.sp0(u,v,p,'table');
n_space = @(u) scalar_product.n0(u,p,'table');
hb = gram_schmidt.tables(tb,sp_space,n_space);

furnish the result shown in Figure 3.2 below:

image

Figure 3.2. Results for example 3.4 (using sp0 and ‘tables’). For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

image

Figure 3.3. Results for example 3.4 using sp1 and ‘subprograms’. For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

image

Figure 3.4. Results for example 3.4 using sp1 and ‘tables’. For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

When the procedure is applied with the scalar product sp1, the results differ. For instance, we obtain the results exhibited in Figures 3.3 and 3.4:

As an alternative, we may define the basis by using the classes previously introduced (section 2.2.6). For instance,


zero = 1e-3;
xmin = -1;
xmax = 1;
degree = 3;
bp = polynomial_basis(zero,xmin,xmax,degree);
%
basis = cell(n,1);
for i = 1: n
  basis{i}.dim = 1;
  basis{i}.dimx = 1;
  basis{i}.component = cell(1,1);
  basis{i}.component{1}.value = @(x) bp.value(i,x);
  basis{i}.component{1}.grad = @(x) bp.d1(i,x);
end;

This definition leads to the same results.

img

3.3. Continuous maps

As previously indicated, distances allow the introduction of the concept of continuity of a function. For a distance defined by a norm (as in theorem 3.2),

DEFINITION 3.8.– Let V, W be two vector spaces having norms img img imgv, img img imgw, respectively. Let T: VW be a function. We say that T is continuous at the point uV if and only if

images

We say that T is continuous on AV if and only if ∀uA : T is continuous at u.

We say that T is uniformly continuous on AV if and only if δ is independent of u on A: δ(u, ε) = δ(ε), ∀u ∈ A.

img

An example of uniformly continuous function is T: Vimg given by T(u) = imguimg: we have imgimguimgimgvimgimgimgu − vimg so that δ(u, ε) = ε.

Continuous functions have remarkable and useful properties. For instance,

PROPOSITION 3.1.— Let V be a vector space having normimgimgimg. Let T: Vimg be continuous at the point uV. If T(u) > 0 (resp. (u)< 0), then there exists δ > 0 such that

images

img

PROOF.– Take ε = imgT(u)img/2 and δ = δ(u, e). From definition 3.5,

images

and we have the result stated.

img

COROLLARY 3.1.— Let V, W be two vector spaces having nomsimgimgimgv,imgimgimgw, respectively. Let F:VW be continuous at the point uV. If F(u) 0 then there exists δ > 0 such that

images

img

PROOF.– Since F is continuous at u,

images

Let T(v) = imgF(v)imgw. We have

images

therefore, that T is continuous at u and the result follows from proposition 3.1.

img

3.4. Sequences and convergence

Our main objective is the determination of numerical approximations to equations (algebraic, differential, partial differential, etc.). Approximations are closely connected to converging sequences.

3.4.1. Sequences

DEFINITION 3.9.— A sequence of elements of V (or a sequence on V) is an application img*n → un ∈ V. Such a sequence is denoted {un: n ∈ W} ⊂ V .

Let img* ∋ knkimg* be a strictly increasing map. Then img* ∋ imagesV is a subsequence images ⊂ {un: nimg*}

img

We observe that nk → +∞ when k → +∞ : we have nk+1 > nk so that nk+1 ≥nk + l. Thus, nk+1 ≥ n1+ k≥k and nk +∞ .

In the context of engineering, the main interest of sequences the main interest lies in their use in the construction of approximations, whether for functions (for instance, signals) or solutions of equations. Approximations are connected to the notion of convergence. In the following, we present this concept and we distinguish strong and weak convergence.

3.4.2. Convergence (or strong convergence)

DEFINITION 3.10.– Let V be a vector space having norm img img img. {un: nimg*} ⊂ V converges (or strongly converges) to uV if and only if imgunuimg → 0, when n → ∞. Notation: un → u or u = limn→ +∞ un.

img

EXAMPLE 3.5.— Let us consider un: (0,1) → img given by

images

and

images

We have

images

Thus, un → u on V = {v: img: v is continuous on Ω, and imgvimgp < ∞ }. Notice that the limit uV, since it is discontinuous at x = 1/2, while unV is continuous on Ω, ∀n.

img

THEOREM 3.5.− un → u if and only if images, for any subsequence.

img

PROOF.– Assume that imgun − uimg → 0. Then:

images

Let k0(ε) be such that images. Then

images

Thus, images. However, assume that images, for any subsequence. Let us suppose that un img u. Then, there exists ε > 0 such that,

images

Let us generate a subsequence as follows: n1 = n(1) and nk+1 = n(nk), ∀ k ≥ 1. Then

images

Thus, 0 < ε ≤ 0, what is a contradiction. So, un → u.

img

We have also

THEOREM 3.6.– LetV, W be two vector spaces having norms img img imgv, imgimgimgw, respectively. F: VW is continuous at uV if and only if unuF(un) → F(u) .

img

PROOF.– See [SOU 10].

One of the main differences between vector spaces formed by functions (i.e. functional spaces) and standard finite dimensional spaces such as imgn is the fact that bounded sequences may do not have convengent subsequences (see examples 3.6 and 3.7 below).

EXAMPLE 3.6.– Let us consider Ω = (0, 2π),

images

and S = {vV: v is continuous on Ω and (v, v) < ∞}. Let us consider the sequence defined by

images

Let v0 = 1. We have

images

Let vk(x) = xk (k ≥ 1). We have

images

Thus, ∀k ≥ 1,

images

Thus, (un, vk) → 0, ∀k ≥ 0, so that

(un, P) → 0, ∀ polynomial function P: (0, 2n) → img.

Let v: (0, 2π) → img be a continuous function. From Stone–Weierstrass approximation theorem (see theorem 3.13), there is a polynomial Pε: (0, 2π) → img such that imgv − Pεimg ≤ ε. Considering that

images

we have

images

and

images

Since ε > 0 is arbitrary, we have (un, v) 0. Assume that images. images is a subsequence of {(un, v): nimg*}, so that images. We also have

images

so that images. Thus, (u, v) = 0, ∀v, ∈ S. Taking v = u, we have (u, u) = 0. This equality implies that imguimg = 0. But

images

so that images and we have images, which is a contradiction. We conclude that the sequence {un: nimg*} has no converging subsequence.

img

3.4.3. Weak convergence

DEFINITION 3.11.– Let V be a vector space having scalar product (img, img). {un: nimg*} ⊂ V weakly converges to uV if and only if ∀vV: (un − u, v) → 0, when n → +∞. Notation: un img u.

img

EXAMPLE 3.7.– Let us consider Ω = (0, 1)

images

and un: (0, 1) → img given by un(x) = cos (2nπx). Let v0 = 1. We have

images

Let vk(x) = xk (k≥1). We have:

images

So that

images

Thus,

images

Let v be continuous on [0, 1] and ε > 0. From Stone–Weierstrass approximation theorem 3.13, there is a polynomial Pε: (0, 1) → img such that imgvPεimg, ≤ ε. Thus,

images

and we have

images

Since ε > 0 is arbitrary, we have (un, v) → 0, so that un img 0 (weakly) on V = {v: (0, 1) → img: v is continuous on [0,1] and imgvimg2 < ∞} , with the scalar product under consideration. Notice that unimg0.

(strongly) since

images

img

We have

THEOREM 3.7.– unu strongly if and only if un img u weakly and imgunimgimguimg when n → +∞.

img

PROOF.– Assume that unu strongly. Then,

images

Thus, un img u weakly and imgunimg imguimgwhen n → +∞. Converse is obtained by using the equality

images

img

THEOREM 3.8.– Let dim(V) = n < ∞. Then un → u strongly if and only if unimgu weakly.

img

PROOF.– Assume that unimg u weakly. Let F = {φ1, … , φn} be an orthonormal basis of V. Then (un, φi) → (u, φi), 1 ≤ i ≤ n , so that

images

img

3.4.4. Compactness

DEFINITION 3.12.– Let SV. S is compact if and only if any sequence {un: nimg*} ⊂ S has a subsequence images such that images. S is weakly compact if and only if any sequence {un: nimg*} ⊂ S has a subsequence images such that images.

img

We have:

THEOREM 3.9.– Let V be a vector space having norm img img img. If SV is compact then S is closed and bounded (i.e. there exists Mimg such that imguimgM, ∀uS).

img

PROOF.– Assume that S is not closed. Then, there exists a sequence {un: nimg*} ⊂ S, such that unu (resp. un img u) and uS. Thus, for any subsequence, images. Since S is compact, u ∈ S . Then, uS and u ∉ S, what is a contradiction.

Assume that S is not bounded. Then, ∀n > 0: ∃unS : imgun img > n . Since S is compact, {un: nimg*} ⊂ S has a subsequence images images such that images. Thus, images images is bounded and ∃Mimg such thatimages. Then M > nk, ∀k > 0. Since nk → +∞, there is k(M) such that nk(M) > M, so that M > nk(M) > M, what is a contradiction.

img

In finite dimensional situations, we have the converse:

THEOREM 3.10.– Let V be a vector space having norm img img img and dim(V) = n < ∞. Then SV is compact if and only if S is closed and bounded.

img

PROOF.– From theorem 3.9, we have the direct implication: if S is compact, then S is closed and bounded. For the opposite, assume that S is closed and bounded. Let F = {φ1, …, φn} be an orthonormal basis of V and T: V → imgn be the linear map T(u) = x, with xi = (u, φ1), 1 ≤ i ≤ n. T is bijective, since, on the one hand, ∀x imgn: ∃u = x1φ1 + …+ xnφn such that T(u) = x and, on the other hand, T(u) = T(v) ⇔ (u, φi) = (v, φi), 1 ≤ i ≤ nu = v. Moreover,

images

so that T(S) ⊂ imgnis closed and bounded. As a consequence, T(S) ⊂ imgnis compact (Heine-Borel-Lebesgue theorem. See [HOF 07], for instance).

Let images has a subsequence images such thatimages. Then images images.

img

The preceding example, 3.7, shows that, contrarily to finite dimensional situations, a bounded sequence of functions does not form a compact set. In the following we will see that, for a particular class of Hibert spaces, a bounded sequence forms a weakly compact set.

3.5. Hilbert spaces and completeness

One of the main difficulties in the practical manipulation of sequences is the verification of their convergence. In order to apply definition 3.9, we must know the limit u — otherwise, we cannot evaluate imgun − uimg. An attempt to obtain a way to verify the convergence without knowing the limit is furnished by Cauchy sequences:

DEFINITION 3.13.– Let V be a vector space having norm img img img. {un: nimg*} ⊂ V is a Cauchy sequence if and only if: ∀ε > 0, ∃ n0(ε) such that m, n ≥ n0(ε) ⇒ imgum − unimg ≤ ε.

We have:

PROPOSITION 3.2.– Let V be a vector space having norm img img img and {un: nimg*] ⊂ V. If unu strongly then {un.nimg*} is a Cauchy sequence.

img

PROOF.– Assume that unu strongly. Then,

images

Thus

images

img

Unfortunately, the opposite is not true: example 3.5 shows that a Cauchy sequence may have as limit an element that does not belong to V. Only at particular spaces, the limits of their Cauchy sequences remain in the space:

DEFINITION 3.14.– Let V be a vector space having norm img img img. V is complete if and only if all its Cauchy sequences converge to elements of V: {un: nimg*] ⊂ V is a Cauchy sequence ⇒ ∃ uV such that unu strongly.

img

DEFINITION 3.15.– A space V that possesses a scalar product and is complete is called a Hilbert space.

img

3.5.1. Fixed points

A fixed point of a map, as it names explicitly mentions, is an invariant point – i.e. a point such that its image by the map is itself:

DEFINITION 3.16.– Let F: VV be a map. We say that uV is a fixed point of F if and only if u = F(u).

img

A large number of numerical methods is based in the construction of fixed points, such as, for instance, Newton’s method for algebraic equations.

The following result is known as Banach’s fixed point theorem:

THEOREM 3.11.– Let V be a Hilbert space. Let F: VV be continuous on V. F is a contraction, i.e. if there exists there exists M<1 such that

images

If F is a contraction, then there exists an unique u img V such that u = F(u). Moreover, the sequence {un: n img img*} ⊂ V,un+1 = F(un) converges to u, ∀u0 imgV.

img

PROOF.– We have

images

so that

images

and

images

Thus,

images

Let ε > 0. There exists n0(ε) such that

images

Thus,

images

and the sequence is a Cauchy sequence. As a consequence, there exists an element u img V such that unu. We obtain u = F(u) from un+1 = F(un). Moreover, let v; = F(v). Then

images

so that (1 − M)imgu − vimg ≤ 0 ⇒ imgu – vimg = 0 ⇒ u = v.

img

The reader will find in the literature a large number of Fixed Point Theorems, established by Luitzen Egbertus Jan Brouwer [BRO 11], Juliusz Paweł Schauder [SHA 30], Andrei Nikolaievich Tikhonov [TIK 35] and Shizuo Kakutani [KAK 41]. Other developments may be found in [FAN 52, RIL 66, EAR 70]. A complete panorama of fixed points is given in [SMA 74].

img

3.6. Open and closed sets

Open and closed sets are a fondamental concept in variational methods. In fact, an alternative way to define a topology consists of giving the families of open and closed sets. Naively, a closed set is a set that contains its boundary. In mathematical terms, we talk about cluster points, i.e. points which cannot be separated from the set, since all their neighborhoods contain points of the set.

3.6.1. Closure of a set

DEFINITION 3.17.− Let V be a vector space having scalar product (img, img). Let S ⊂ V. u img V is a cluster point (or adhering point) of S if and only if

images

Analogously, u img V is a weak cluster point (or weak adhering point) of S if and only if

images

img

DEFINITION 3.18.− Let V be a vector space having scalar product (img, img). Let SV. The closure of S is denoted by images and defined as:

images

The weak closure of S is

images

img

We have

PROPOSITION 3.3.− Let SU. Then Simages.

img

PROOF.− We have Simages: ∀u img S : un = u, ∀n verifies {un:n img img*} ⊂ S and un → u.

Consider {un: n img img*} ⊂ S such that unu. Then {un: n img img*} ⊂ img, so that u img images.

3.6.2. Open and closed sets

DEFINITION 3.19.− Let SV. S is closed (respectively, weakly closed) if and only if S contains all its cluster points (resp., weak cluster points), i.e. images S is open (respectively, weakly open) if and only if V − S is closed (resp. weakly closed).

img

We have:

THEOREM 3.12.− S is open if and only if:

images

img

PROOF. Assume that S is open. Let us consider u img S: ∀ ε > 0, ∃ vε img S such that imgvε — uimg ≤ ε and vεS. Let images un = vs. We have images, so that un → u. Moreover, un = vεS, ∀S. Thus, {un: n img img*} ⊂ V − S. Since S is open, V − S is closed. Thus, u img V − S: then, we have u img S and u img V − S, which is a contradiction.

However, assume that

images

but V − S is not closed. The, ∃{un: n img img*} ⊂ V − S such that unu ∉ V −S . Thus, on the one hand, u img S and ∃ ε > 0 such that imgv uimgε v img S . On the other hand, ∃n0(ε): n ≥ n0(s) ⇒ imgun uimg ≤ ε. Thus: un img S, ∀nn0(ε) and {un: n ε img*} ⊂ VS, so that ∀nn0(ε): un img S img V − S = ø, which is a contradiction. Thus, S is closed.

img

Moreover:

PROPOSITION 3.4.− If S is weakly closed, then S is closed.

img

PROOF.− Assume that S is weakly closed. Consider {un: n img img*} ⊂ S such that unu. From theorem 3.7, un img u, so that u img S.

img

PROPOSITION 3.5.− S is weakly closed.

img

PROOF.− Consider {un:n img img*} ⊂ S such that un img u,. Then:

images

Thus, ∀v img S : (u, v) = limn→+∞ (un, v) = 0, so that u img S.

img

COROLLARY 3.2.− S⊥⊥ = (S) is weakly closed and images.

img

PROOF.− S⊥⊥ is weakly closed, from proposition 3.5: then, images = S⊥⊥. Moreover, for u img S ,(u, v) = 0, ∀ vimgS, so that u img S⊥⊥ and we have SS⊥⊥. From proposition 3.3, SimagesS⊥⊥.

img

PROPOSITION 3.6.− Let SV. Then images is closed. If S is a vector subspace, then imagesV is a closed subspace.

img

PROOF.− See [SOU 10].

img

3.6.3. Dense subspaces

DEFINITION 3.20.− Let V be a vector space having scalar product (img, img). Let SV. S is dense on V if and only if images.

img

A first example of dense subspace is furnished by the theorem of Stone-Weierstrass:

THEOREM 3.13.− Let Ω ⊂ imgn and F: Ω → img be a continuous function. Then, ∀ε > 0, ∃ Pε, Pε polynomial, such that imgF − Pε img0 ≤ ε.

img

PROOF.− See [DUR 12].

img

3.7. Orthogonal projection

Orthogonal projections are a basic tool in variational methods, often invoked and used in practice. In the sequel, V is a Hilbert space having scalar product(img, img).

3.7.1. Orthogonal projection on a subspace

DEFINITION 3.21.− Let SV. S ≠ ø. Let uV. The orthogonal projection Pu of u onto S is PuS such that imgu − Puimg imgu — vimg, ∀v img S, i.e.

images

img

We have

THEOREM 3.14.− Let S be a closed vector subspace. Then, ∀uV, there exists an unique PuS such that that imgu - Puimg images

img

PROOF.– Let images. Let n > 0: there exists unS such that

Thus, for m, nk,

images

so that

Since

we have images, so that equation [3.2] implies that

and the sequence {un: nimg*} ⊂ S is Cauchy sequence. Thus, there exists PuS (since S is closed and V is a Hilbert space) such that unPu. Equation [3.1] shows thatimages images. Moreover, if Qu ∈ S and images, then, the equality

images

shows that imgPu – Quimg2 0, so that Pu = Qu (analogous to equations [3.2], [3.3], [3.4]).

THEOREM 3.15.– Let S be a closed vector subspace and uV. Then, Pu is the orthogonal projection of u onto S if and only if

img

PROOF.– Let vS and f(λ) = imgu – Pu – λvimg2. We have

images

Then

images

If Pu is the orthogonal projection of u onto S, the minimum of f is attained at λ = 0and f′ (0) = 0, so that (uPu, v) = 0. However, if (uPu, v) = 0, then f′(λ) = 0 for λ = 0, so that the minimum is attained at λ = 0 and we have imgu – Puimg = min{imguvimg: vS} .

Equation [3.5] is a variational equation. It may be interpreted in terms of orthogonality: vector uPu is orthogonal to all the elements of the subspace S. A simple geometrical interpretation is given in Figure 3.5: in dimension 2, a vector subspace is a line passing through the origin.

image

Figure 3.5. Geometrical interpretation of the orthogonal projection on a vector subspace. For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

One of the main consequences of the orthogonal projection is [RSZ 07]

THEOREM 3.16.– Let SV be a vector subspace. S is dense on V if and only if S = {0}.

img

PROOF.– Let S be dense on V and vS Then, on the one hand, there exists a sequence {vn: nimg*} ⊂ V such that vn → v; on the other hand, (vn, v) = 0, ∀nimg*. Thus, from theorem 3.7, (v, v) = 0 ⇒ v = 0. Thus, S = {0}.

Let S = {0}. Assume that images. Then, there exists uV, u ∉ images . Since images is closed, images is open: from theorem 3.12, there is ε > 0 such that images. Let Pu be the orthogonal projection of u onto images: equation [3.5] shows that u – images. Since Simages, we have uPuS Thus, u — Pu = 0, so that u = Puimages. Thus, uimages and uimages, what is a contradiction.

img

3.7.2. Orthogonal projection on a convex subset

DEFINITION 3.22.– Let SV. S ≠ Ø. S is convex if and only if

images

img

THEOREM 3.17.– Let S be a closed non-empty convex. Then, ∀uV, there exists a unique PuS such that that imgu – Puimgimgu – vimg, ∀v , ∈ S. Moreover, Pu is the orthogonal projection of u onto S if

PROOF.– The proof of existence and uniqueness is analogous to these given for theorem 3.14. Since S is convex, equations [3.2], [3.3], [3.4] are verified and the conclusion is obtained by the same way (the sequence is a Cauchy sequence, V is Hilbert, S is closed). The inequality is obtained as follows: let vS and f(λ) = imgu – Pu – λ{v-Pu)img2. Since

images

we have f(0) ≤ f(λ), ∀λ ∈ (0, 1), so that f′(0) = -2(u – Pu, v – Pu) ≥ 0 and we have the inequality.

img

image

Figure 3.6. Geometrical interpretation of the orthogonal projection on a convex subset. For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

Equation [3.6] is a variational inequality. It may be interpreted in terms of orthogonality: vector u – Pu is orthogonal to the tangent space of set S. A simple geometrical interpretation in dimension 2 is given in Figure 3.6: we see that the angle between uPu and vPu is superior to π/2.

3.7.3. Orthogonal projection on an affine subspace

DEFINITION 3.23.– Let SV. S ≠ Ø. S is an affine subspace if and only if

images

img

image

Figure 3.7. Geometrical interpretation of the orthogonal projection on an affine subspace. For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

THEOREM 3.18.– Let S be a closed non-empty affine subspace. Then, ∀u img V, there exists an unique PuS such that that imgu – Puimgimgu – vimg, ∀vS. Moreover, Pu is the orthogonal projection of u onto S if and only if

img

PROOF.– Notice that S is a closed convex set, so that theorem 3.17 applies. In order to prove the equality, notice that, for vS, 2Pu – v ∈ S, so that we have

images

Thus, (u – Pu, v – Pu) 0 and (u – Pu, v – Pu) 0, which establishes the equality.

Equation [3.7] is a variational equation. It may be interpreted in terms of orthogonality: vector u – Pu is orthogonal to the differences of elements of S. A simple geometrical interpretation in dimension 2 is given in Figure 3.7: we see that uPu is orthogonal to v – Pu.

Affine subspaces are generally defined by introducing a translation of a vector subspace:

images
image

Figure 3.8. A second interpretation of the orthogonal projection on an affine subspace. For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

In this case

THEOREM 3.19.– Let S = S0 + {u0}, where S0 is a closed vector subspace and u0V. Then, ∀uV, there exists an unique PuS such that that imgu — Puimgimgu — vimg, ∀vS. Moreover, Pu is the orthogonal projection of u onto S if and only if

img

PROOF.– Notice that S is a closed affine subspace, so that theorem 3.18 applies. For the equality, notice that, for any W ∈ S0: W + PuS. Indeed, U = Pu – u0 S0 and, as a consequence, w + US0, so that w + Pu = w + U + u0S. Thus, equation [3.7] is equivalent to equation [3.8].

img

Equation [3.8] is a variational equation. A simple geometrical interpretation in dimension 2 is given in Figure 3.8: we see that uPu is orthogonal to S0 parallel to S, which is generated by a translation of S0.

3.7.4. Matlab® determination of orthogonal projections

Orthogonal projections may be assimilated to linear equations and determined into a way analogous to the one introduced in section 2.2.4.

Let us consider the situation where S is a subspace: we may consider a basis images and look for an approximation

images

Then, the variational equation [3.5] yields that

images

Thus, the coefficients U = (u1, …, un)t satisfy a linear system AU = B, with images and images.

When using the scalar product sp0, defined in section 3.2.5, and the basis defined in section 2.2.6, the coefficients U may be determined by using the method approx_coeffs. Notice that matrix A corresponds to methods varmat_mean, varmat_int, while vector B corresponds to methods sm_mean and sm_int. The method of evaluation is selected by _ choosing ‘variational_mean’, ‘variational_int’ or ‘variational_sp’.

Otherwise, matrix A and vector B may be determined by using the class introduced in section 3.2.5. We introduce below a new class corresponding to this last situation. Analogously to the approach presented in 3.2.6, we assume that basis{i} defines images as a structure having properties dim, dimx, component, where component is a cell array such that each element is a structure having as properties value and grad. u is assumed to be analogously defined. In this case, an example of class for orthogional projection is given in program 3.6. The class contains two methods: coeffs, which evaluates the coefficients of the orthogonal projection and funct, which generates a subprogram (anonymous function) evaluating the projection.

image
image

Program 3.6. A class for orthogonal projection

EXAMPLE 3.8.– Let us consider the family images 4 and the function u(x) = x5 on (−1, 1). Using the scalar product sp0, the orthogonal projection is images. The code


xlim.lower.x = -1;
xlim.upper.x = 1;
xlim.dim = 1;
u.dim = 1;
u.dimx = 1;
u.component = cell(u.dim,1);
u.component{1}.value = @(x) x^5;
u.component{1}.grad = @(x) 5*x^4;
sp_space = @(u,v)
scalar_product.sp0(u,v,xlim,'subprogram');
c = orthogonal_projection.coeffs(u,basis,sp_space);
pu = orthogonal_projection.funcproj(c,basis);

generates a subprogram pu which evaluates the orthogonal projection. Analogously,

U.dim = 1;
U.dimx = 1;
U.component = cell(U.dim,1);
U.component{1}.value =
spam.partition(u.component{1}.value,p);
U.component{1}.grad = cell(U.dim,1);
U.component{1}.grad{1} =
spam.partition(u.component{1}.grad,p);
sp_space = @(u,v)
scalar_product.sp0(u,v,p,'table');
c_tab =
orthogonal_projection.coeffs(U,tb,sp_space);
pu_tab =
orthogonal_projection.funcproj(c_tab,basis);

generates a subprogram pu_tab which evaluates the orthogonal projection. The results are shown in Figure 3.9.

img

EXAMPLE 3.9.– Let us consider the family images and the function u(x) = sin(x) on (0, 2π:). Using the scalar product sp0, the orthogonal projection is images images. The code


xlim.lower.x = 0;
xlim.upper.x = 2*pi;
xlim.dim = 1;
u.dim = 1;
u.dimx = 1;
u.component = cell(u.dim,1);
u.component{1}.value = @(x) sin(x);
u.component{1}.grad = @(x) cos(x);

defines u. pu and pu_tab are generated analogously to example 3.3. The results are_ shown in Figure 3.10.

img

image

Figure 3.9. Orthogonal projection of x5 onto a polynomial subspace (degree <= 3)

image

Figure 3.10. Orthogonal projection of sin(x) onto a polynomial subspace (degree <= 3). For a color version of the figure, see www.iste.co.uk/souzadecursi/variational.zip

3.8. Series and separable spaces

A series on V is an infinite sum of elements of V. In the framework of variational methods, the main use of series is the representation of elements of V by infinite sums of a given family FV. We are particularly interested in countable families: images, so that an element uV may be represented as

images

Such an expansion is usually referred to as a generalized Fourier series. Indeed, such an expansion was introduced by Fourier [POI 07], which proposed a solution of heat equation by series of trigonometric functions. His paper was presented by Siméon Denis Poisson and immediately aroused a controversy with Joseph-Louis Lagrange and Pierre-Simon Laplace, due to its mathematical shortcomings. The main question was the sense of the equality and, so, what the functions u admitting such an expansion are. Previously to Fourier, Daniel Bernoulli faced a controversy with Leonhard Euler for analogous reasons: Euler considered that expansions by using trigonometrical function could not represent other functions [BER 53a, BER 53b, EUL 53]. Poisson worked with Jacques Charles François Sturm and Joseph Liouville in order to get results that are more rigorous. However, their work found some difficulty in being accepted by the French Academy of Sciences, due to the incompleteness of the theory – of course, we may blame their referees with a few centuries of additional mathematical work in our possession, but the subject was difficult and about one century has been needed for a more satisfactory solution. A new result arrived with Peter Gustav Lejeune Dirichlet [DIR 29a], but Paul David Gustave du Bois-Reymond produced an example which generates new discussions (a continuous function having a Fourier series that diverges at a point) [DUB 73]. In a further development, these reflections led to the theory of integrals by Riemann and Lebesgue, as introduced in Chapter 1.

3.8.1. Series

DEFINITION 3.24.– A series images of elements of V is a sequence {Sn: nimg*}, such that images. If SnvV, we say that the series is convergent and its sum is images. A series that is not convergent is said to be divergent.

img

Sn is usually referred to as a partial sum. The series

images

is the series of the residual of order n. Rn is defined by the sequence

images

We have

THEOREM 3.20.– The following assertions are equivalent

images
images
images

img

PROOF.– Assume (i). Let images. Then images. Theorem 3.5 shows that all the subsequences of images converge to the same limit. Thus, for n > 1: Rn, n+k = Sn+k images, so that we have (ii).

Assume (ii). Then (iii) is immediate.

Assume (iii). We have images. Taking the limit for k → +∞, we have Sn−1 + Rn = Sm−1 + Rm. Thus, s = Sn−1 + Rn is independent of n and we have ∀n, k > 1: Sn+k = Sn–1 + Rn, ks, for k → +∞. Thus

images

what shows that Sk+1 → s, so that images converges.

img

3.8.2. Separable spaces and Fourier series

DEFINITION 3.25.– Let images and [F] be the set of the finite linear combinations of elements of F:

images

F is a total family if and only if images, i.e. [F] is dense on V. We say that F is a Hilbert basis if and only if F is orthonormed and total.

img

DEFINITION 3.26.– V is separable if and only if there exists FV such that F is total.

img

We have:

THEOREM 3.21.– Let images be a Hilbert basis of V. Let u, vV. Then

  1. i) img is a closed subspace.
  2. ii) The orthogonal projection img
  3. iii) img strongly in V.
  4. iv) img
  5. v) img
  6. vi) img converges.

img

PROOF.– See [SOU 10].

images is the Fourier series of u associated with the Hilbert basis F. Equality 3.21. (iv) is known as Bessel-Parseval equality.

Separable spaces have the weak Bolzano–Weierstrass property:

THEOREM 3.22.– Let V be a separable Hilbert space. Let {un: nimg*} ⊂ V be a bounded sequence (∃Mimg: ∀nimg* : imgunimgM). Then {un:nimg*} has a subsequence images which is weakly convergent.

img

PROOF.– See [SOU 10].

img

3.9. Duality

Hilbert spaces are associated with their dual spaces. If we consider Hilbert spaces as spaces of displacements of a mechanical system, dual spaces may be interpreted as the spaces of virtual works associated with these displacements. Riesz’s theorem establishes that any virtual work corresponds to a force [RSZ 09].

3.9.1. Linear functionals

DEFINITION 3.27.– A linear functional is an application ℓ: Vimg such that

images

img

A linear functional verifies ℓ(0) = 0 : since 0 = 2 × 0, we have (0) = (2 × 0). Thus, (0) = 2 × (0) and ℓ(0) = 0.

DEFINITION 3.28.– The norm of a linear functional ℓ is

images

img

We have

PROPOSITION 3.7.–

images

img

and

THEOREM 3.23.– Let :Vimg be a linear functional. is continuous if and only if img img< ∞.

img

PROOF .-Assume that img img< ∞ Then

images

so that satisfies Lipschitz’s condition and, thus, is continuous (observe that, for un → u, we have images images.

Assume that is continuous. Let us establish that imgimg < ∞.

images

Then

images

But

images

Thus, 0 > +∞, what is a contradiction.

img

COROLLARY 3.3.– Let : Vimg be a linear functional. is continuous if and only if there exists Mimg such that images In this case, imgimg ≤ M .

img

PROOF.– Assume that ℓ is continuous. Then M = imgimg verifies

images

Assume that there exists Mimg such that images images

images

and ℓ is continuous.

img

EXAMPLE 3.10.– Let us give an example of a linear functional which is not continuous and has imgimg = ∞. Let us consider

images

Let ℓ:Vimg be given by img(v) = v(0). Let us consider the sequence

images

Then, for 0 < α < 1/2,

images

while

images

Assume that there exists Mimg such that images images

images

So that

images

what is a contradiction, since Mimg. Thus, imgMimg such that imgℓ(v)imgM imgvimg, ∀vV: ℓ is not continuous and imgimg = ∞. This functional is kwown as Dirac’s delta or Dirac’s mass (see Chapter 6).

img

An important property of continuous linear functional is the continuous extension:

THEOREM 3.24.– Let SV be a dense subspace. Let ℓ : Simg be such that there exists M ∈ img such that ∀vS: imgℓ(v)imgM imgvimg. In this case, ℓ has an extension ℓ Vimg such that imgimg ≤ M .

img

PROOF.– Let vV Since S is dense on V, there is a sequence {vn: nimg*} ⊂ S such that vnv. Let us consider the sequence of real numbers {(vn): nimg*} ⊂ img. This is a Cauchy sequence, since

images

Then, (vn) → λ ∈ img. Let {wn: nimg*} ⊂ S such that wnv and {ℓ(vn) → ηimg. Then.

images

So that λ = η. Thus, we may define

images

is linear: if SvnvV, S ∋ unuV, αimg, then

images

so that (i + av) = ℓ(u) + aℓ(v). Finally, by taking the limit for n → ∞,

images

img

DEFINITION 3.29.– The topological dual space V’ associated with V is

images

img

3.9.2. Kernel of a linear functional

DEFINITION 3.30.– The Kernel of a linear functional is

images

img

Notice that Ker () is a subspace, since

images

and, since (αu + i) = αℓ(u) + ℓ(v),

images

We have

PROPOSITION 3.8.– Let : Vimg be a linear functional. is continuous if and only if Ker() is closed.

img

PROOF.– Assume that is continuous. Then

images

so that

images

and

images

Thus Ker() is closed.

Assume that Ker() is closed. Let us establish that imgimg < ∞. This is immediate if = 0, since, in this case, imgimg = 0. Let us assume 0 and imgimg = ∞. Then, on the one hand,

images

and, on the other hand, there exists ū such that (ū) = 1. Let

images

Then

images

so that {un}n>0Ker(ℓ). Moreover,

images

so that unū. Since Ker() is closed, we have ū ∈ iver(0) ⇒ (ū) = 0. Thus, 1 = (ū) = 0, what is a contradiction.

img

3.9.3. Riesz’s theorem

The main result is the Riesz’s theorem [RSZ 09, FRE 07a]:

THEOREM 3.25.– Let ℓ: Vimg be a continuous linear functional. Then there exists one and only one uV such that:

images

Moreover

images

img

PROOF.– The result is immediate when = 0, since, in this case, imgimg = 0 and u = 0. Let us assume ℓ ≠ 0. Then, there exists w such that (w) = 1. Let us denote Pw its orthogonal projection onto Ker(). We have w – Pw ≠ 0, since (w) = 1 0 = (Pw). Let us consider,

images

From the properties of the orthogonal projection, we have uKer(), so that, (u, Pw) = 0. Moreover,

images

Let vV: we have,

images

Since (v − ℓ(v)w) = (v) − (v)(w) = (v) (v) = 0, we have v(v)w ∈ Ker(ℓ) and

images

Moreover,

images

and

images

img

Riesz’s theorem establishes a connection between linear functionals and scalar products. Since scalar products represent the virtual work of internal efforts and linear functional represent the virtual work of external forces, Riesz’s theorem may be interpreted as a connection between the work of external forces and the variation of internal energy, which corresponds to the work of internal efforts. More precisely, Riesz’s theorem connects the virtual work of external forces to the virtual work of internal efforts.

Moreover, Riesz’s theorem connects a linear functional and a field: it may also be interpreted as the connection between a virtual work and a force. This connection is formalized by Riesz canonical isometry:

images

Riesz’s theorem shows that is a bijection, so that −1 exists. Moreover, on the one hand, img∏(u)img = imguimg, while, on the other hand, img−1()img = imgimg.

3.10. Generating a Hilbert basis

There are two basic ways for the generation of a Hilbert basis:

  1. 1) If a total family {ϕn}nimg is given, we may use the Gram-Schmidt procedure;
  2. 2) Otherwise, we may use the eigenfunctions ϕn associated with a pair order differential operator T:ϕn 0, T:ϕn = λnϕn, with homogeneous boundary conditions. For instance, on Ω = (0, 1), you may use images.

The use of a differential operator of pair order is justified by the fact that the inverse of T is a linear bounded compact self-adjoint operator. Let us recall that:

  1. 1) A linear operator T on V is an application T: VV such that images.
  2. 2) A bounded linear operator T on V is a linear operator T on V such that images.
  3. 3) A compact bounded linear operator T on V is a bounded linear operator T on V which transforms weakly convergent sequences into strongly convergent sequences, i.e. images images
  4. 4) T is self-adjoint if and only if images.

Compact bounded linear operators may be considered as a generalization of matrices to infinite dimensional spaces. They may be generated by the inversion of differential or partial differential operators. We have

THEOREM 3.26.– Let T: VV be a linear, bounded, compact, self-adjoint operator. Then, there exists a hilbertian basis of V formed by eigenvectors of T.

img

3.10.1. 1D situations

For instance, let us consider Ω = (a, b) and the space V = L2(Ω). The operator defined by:

images

is linear, bounded, compact and self-adjoint. Let ϕ be an eigenvector of T, associated with the eigenvalue α:

images

We have:

images

Let us observe that αn ≠ 0, since,

images

which is a contradiction. Thus, denoting λ = 1/α, we have:

images

In this case, the solution is a trigonometric family:

images

3.10.2. 2D situations

Analogously, if Ω ⊂ img2 is a regular bounded domain, H = L2(Ω),

images

is also linear, bounded, compact and self-adjoint Let ϕ be an eigenvector of T, associated with the eigenvalue α:

images

We have:

images

Here again, αn ≠ 0, since,

images

what is a contradiction? Thus, denoting λ = 1/α, we have:

images

In this case, the solution is not explicitly known for general domains: it is explicitly known only for particular shapes of Ω. Otherwise, it must be numerically determined. For special geometries, the eigenfunctions are often determined by separation of variables.

3.10.2.1. The case of a rectangle

If images, we may consider (x, y) = X(x)Y(y).

Then,

images

Thus, we have,

images

So,

images

and:

images

Moreover,

images

Thus, we have:

images

So,

images

The solution is a product of trigonometric families: X = um and Y = un are solutions of the 1D case,

images

3.10.2.2. The case of a disk

If Ω= {(x, y): x2+y2 < α2}, we may consider (x, y) = R(r) Θ (θ), = rcosθ, y = rsin θ. Let us denote by μ′ the derivative of a function μ(ξ) with respect to its single real variable ξ ∈ img − for instance R′ = dR/dr, Θ′ = dΘ/dθ, etc. Since

images

we have:

images

and

images

So,

images

Thus,

images

Let us introduce images: we have (recall that y′ = dy/dp, while R′ = dR/dr)

images

and

images

For β = −m2, this is the classical Bessel’s equation and Θ is a trigonometric function. Thus, denoting by Jm the m-th Bessel’s function and by zmn its n-th zero:

images

3.10.3. 3D situations

If Ω ⊂ img3 is a regular bounded domain, H = L2 (Ω),

images

is again linear, bounded, compact and self-adjoint. Thus, an eigenvector ø, associated with the eigenvalue α verifies, with λ = 1/α,

images

Here, the solution is explicitly known only for particular shapes of Ω and it is often determined by separation of variables.

3.10.3.1. The case of a rectangular parallelepiped

If images, we may consider images Then, we obtain:

images

3.10.3.2. The case of a cylinder

images, we may consider images. Since,

images

we have:

images

Thus,

images

So,

images

For, images, we take images images and we obtain, analogously to the 2D-case,

images

Thus, y is again the solution of Bessel’s equation, while Z and Θ are trigonometric functions and we have:

images

3.10.3.3. The case of a ball

For Ω = {(x, y, z) : x2 + y2 + z2 < α2}, we may consider images rcosφ associated with the eigenvalue α. Since,

images

we have, for λ = 1/α,

images

and

images

For images, we have (recall that y′ = dy/ds, while images)

images

and

images

Thus

images

and

images

This equation is known as Legendre’s equation. The solutions are the associated Legendre functions images. Moreover, Θ is a trigonometric function and img verifies

images

Taking again images, we have:

images

and

images

The solutions of this equation are the spherical Bessel’s functions. Thus, the eigenfunctions are:

images

zm+1/2,n is the n-th zero of Jm+1/2. The angular term images is called a spherical harmonic.

3.10.4. Using a sequence of finite families

An alternative to the generation of a hilbertian basis consists of the use of a family of finite families, i.e. FV such that,

images

For instance, families of finite elements may be used: in such a case, Fn corresponds to a particular discretization (i.e. a particular mesh) and F is the family of all the possible discretizations (i.e. of all the possible meshes). Finite volume approximations or wavelets may be interpreted by the same way.

3.11. Exercises

EXERCISE 3.1.– Let V be a vector space on img.

  1. 1) Show that −1. u = − u, ∀uV. (Hint: 1. u = u)
  2. 2) Show that 0. u = 0, ∀uV. (Hint: 0 = 0 + 0).

img

EXERCISE 3.2.– Let Ω imgn and V = {v: Ωimgk}.

  1. 1) Show that V is a vector space on E.
  2. 2)Show that
images

is a vector subspace of V.

img

EXERCISE 3.3.– Let d be a distance on V and η: imgimg be a strictly increasing application such that η(0) = 0 and η(a + b) ≤ η(a) + η(b). Show that dη(u, v) = η(d(u, v)) is a distance on V. (Hint: images

img

EXERCISE 3.4.– Show that images

img

EXERCISE 3.5.– Condiser

images

Let θ(m, n) be the angle between tmand tn. Show that

images

img

EXERCISE 3.6.– Consider

images

Let θ be the angle between sin(t) and cos(t). Show that:

images

img

EXERCISE 3.7.– Let images A}. Determine images, where M = (x, y, z) ∈ img3 and D is the line D = {λ(a, b, c): λ ∈ img}.

img

EXERCISE 3.8.– Let A⊂–imgn, Bimgn, images A, bB}. Find d(A, B) when A is the hyperbole A = {(x, y) ∈ img2: x1x2 = 1} and B is one of its asymptotes.

img

EXERCISE 3.9.– Let images. Show that:

  1. 1) images.
  2. 2) images.
  3. 3) images.
  4. 4) Let images. Show that, for any (a, b)img2, a < b, there exists n ≥ 1 such that0 < un < b − a. Conclude that there exists mimg such thata < mun < b.
  5. 5) Show that: ∀ximg: ∀ε, > 0, ∃xεD, imgx – xεimg ε. Conclude that d(x, D) = 0.

img

EXERCISE 3.10.− Let d(x, y) = imgx − yimg, f: imgimg satisfy: ∀α > 0, f−1((a,+∞)) is finite. Let D = {xRimgf(x) = 0}. Show that images

img

EXERCISE 3.11.− Let E = C0(I; Rn); I = (0, 1). We denote, for p ≥ 1 and uE,

images

The aim is to establish that imgimgp is a norm E and that E is not a complete space for this norm.

  1. 1) As a first step, establish Hölder’s inequality: for p, q ≥ 1 such that images (conjugate exponents),
    images
    1. a) Let f(t) = log(t) (natural basis). Show that f" < 0 on images images. Conclude that f is concave on images and that:
      images
    2. b) Show that
      images
    3. c) Conclude that
      images.
    4. d) Extend the result to images.
    5. e) Take = imguimg/imguimgp, b = imgvimg/imgvimgq and conclude.
  2. 2) As a second step, establish Minkowski’s inequality:
    images
    1. a) Set r = (imguimg + imgvimg)p, s = (imguimg + imgvimg), t = (imguimg + imgvimg)p1 Verify that r = st = imguimgt + imgvimgt.
    2. b) Verify that
      images
    3. c) Conclude that
      images
    4. d) Verify that (p−1)q = p. Conclude that images.
    5. e) Show Minkowski’s inequality.
  3. 3) Establish th imgimgp atis a norm on E.
  4. 4) Consider the sequence {un: n ≥ 0} given by
    images
    1. a) Let
      images

      Show that images

    2. b) Conclude that {un: n ≥ 0} is a Cauchy sequence of elements of (E, imgimgp).
    3. c) Let uE verify images on (E, imgimgP). Show that imguūimgp = 0. Conclude that u ∉ E and that on (E, imgimgp) is not complete.

      img

  5. 5) Let images and images
    1. a) Let images. Verify that images and use Holder’s inequality to show that
      images
    2. b) Since I is bounded, show that there exists a constant Cb(a, l) such that
      images
  6. 6) Let
    images.
    1. a) Show that imgimg∞ is a norm on E.
    2. b) Verify that
      images.
    3. This inequality justifies the generalization where the conjugate exponent of p = ∞ is q = 1 and the convention images.
    4. c) Since I is bounded, show that there exists a constant C (a, 1) such that, for aimg, a ≥ 1,
      images.
  7. 7) Let p ≥ 1, q ≥ 1 be conjugate exponents, eventually taking the value ∞, and u regular enough such that u(0) = 0. By convention images
    1. a) Show that
      images.
    2. b) Conclude that
      images.
    3. c) Since I is bounded, show that there exists a constant M (a, I) such that
      images.
    4. Conclude that there exists a contant M (a, p, I) img img such that:
      images.
    5. d) Let u be a regular function such that u(0) = u′(0) = 0. Verify that:
      images.
    6. Show that there exists a constant N(a, p, l) ∈ img such that:
      images.

img

EXERCISE 3.12.– Let I = (0, T), images images. We consider the ordinary differential equation (ODE)

images.

Assume that f satisfies a Lipschitz condition:

images.
  1. 1) Show that:
    images.
  2. 2) Let F: EE be defined as follows :
    images.
    1. a) Let u, vE. Verify that:
      images
    2. b) Conclude that
      images
    3. c) Show that:
      images
    4. d) Let images on images Show that:
      images.
    5. e) Conclude that F is continuous.
  3. 3) Let images
    1. a) Show that F is a contraction.
    2. b) Conclude that there exists one and only one yE such that y = F(y).
    3. c) Conclude that the ODE has a unique solution yE.

EXERCISE 3.13.– Let Ω = (0, 1) and (u, v) = img uvdx, C = {v: Ω) → img: v ≥ 0 on Ω}.

  1. 1) Show that C is convex
  2. 2) Let images Verify that images
  3. 3) Show that (u − u+, v − u+) ≤ 0, ∀vC.
  4. 4) Conclude that u+ is the orthogonal projection of u onto C.

EXERCISE 3.14.– Let Ω = (0, 1) and (u, v) = img (uv + u′v′)dx. Let S = {v: Ωimg: v(0) = 1}.

  1. 1) Show that S affine subspace
  2. 2) Verify that images.
  3. 1) Let images. Verify that:
    images
  4. 2) Conclude that images.
  5. 3) Conclude that images.
  6. 4) Show that U is the orthogonal projection of 0 onto S.

EXERCISE 3.15.– Let Ω = (0, 1) and images images. Let images.

  1. 1) Show that S is an affine subspace.
  2. 2) Verify that= {1} + S0, S0 = {v: Ωimg: v(1) = 1}.
  3. 3) Let U : Ω → img satisfy images 0 on Ω. Verify that:
    images
  4. 4) Conclude that ∀w ∈ S0: (U,w) = 0.
  5. 5) Conclude that ∀vS: (U, v − U) = 0.
  6. 6) Show that U is the orthogonal projection of 0 onto S.

EXERCISE 3.16.– Let Ω = (0, 1) × (0, 1) and images images. Let images

  1. 1) Show that S is an affine subspace.
  2. 2) Verify that= {u0} + S0 , S0 = {v: Ωimg: v = 0 on ∂Ω}.
  3. 3) Let U: Ωimg satisfy U = u0 on ∂Ω, ΔU = 0 on Ω. Verify that:
    images
  4. 4) Conclude that ∀w ∈ S0: (U, w) = 0.
  5. 5) Conclude that ∀vS: (U,v − U) = 0.
  6. 6) Show that U is the orthogonal projection of 0 onto S.

EXERCISE 3.17.– Let Ω = (0, 1) × (0, 1) and images images Let S = {v: Ω → img: v = u0 on ∂Ω}.

  1. 1) Show that S is an affine subspace.
  2. 2) Verify that= images
  3. 3) Let U: Ω img satisfy U = u0 on ∂Ω, U − ΔU = 0 on Ω. Verify that:
    images
  4. 4) Conclude that ∀w ∈ S0: (U,w) = 0.
  5. 5) Conclude that ∀vS: (U, vU) = 0.
  6. 6) Show that U is the orthogonal projection of 0 onto S
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset