Recall that every closed subspace of a Hilbert space E splits in E ([P2], section 3.10.2(II), Theorem 3.147(2)). Corollary 1.33 shows that any immersion of class C p admits a local retraction r of class C p . Immersions are therefore local sections (see [P1], section 1.1.1 (III)).
([P2], section 3.2.2 (IV), Theorem 3.5(3)) implies the following result:
The composition of two immersions is an immersion, the composition of two submersions is a submersion and, given a subimmersion f, an immersion i, and a submersion s, the mapping i ∘ f ∘ s is a subimmersion (exercise). However, the composition of two subimmersions is not always a subimmersion ([DIE 93], Volume 3, section 16.8, Problem 1(b)).
In this section,
.
(I) Affine spaces
Given a vector space E, an affine space
attached to the space E is a homogeneous space of the additive group E ([P1], section 2.2.8(II)) such that the (transitive) action of E on
is faithful, i.e. such that the neutral element 0 is the only element of E that fixes every element of
. The action of x ∈ E on
is written as P + x. We say that E is the space of translations of
, its elements are the translations of
and, if dim (E) < ∞, this quantity is called the dimension of
. Given some origin O chosen from
, the elements of
are all of the form O + x (x ∈ E).
If Q = P + x, we write
(the bipoint of origin P and endpoint Q)
7
. If E is a locally convex space, the sets O + U = {O + x : x ∈ U}, where the U are the open sets of E, define a topology on
. When equipped with this topology,
is called a locally convex affine space. Every point of such a space admits a fundamental system of convex neighborhoods. We can similarly define the concepts of affine topological space, affine normed space, affine pre-Hilbert space, etc.
(II) Lagrange variations
Let
be a locally convex affine space, F a locally convex space, A some non-empty subset of
, a some point of A and f : A → F a mapping.
We say that f admits a Lagrange first variation δ f (a): E → F : h ↦ δ f (a) [h] at the point a if, for all h ∈ E,
If this condition is satisfied, we say that f admits a Lagrange second variation δ 2 f (a): E → F if
The Lagrange variation of order n, δ n f (a) : E n → F : h ↦ δ n f (a) [h], is defined inductively in the same way.
(III) Gateaux differentiability It is easy to show using the generalized Goursat theorem (section 1.2.5 (III)) that, if E and F are complex locally convex spaces and f admits a Lagrange first variation at the point a, then δ f (a) is linear ([HIL 57], Theorem 26.3.2). In general, we have the following result:
The set
of G-differentiable mappings at the point a is an affine space. A G-differentiable mapping is not necessarily continuous. If E is a normed vector space and f is differentiable at the point a, then it is also G-differentiable at this point and D
f (a) = D
G
f (a) (exercise). On product spaces, we may define the Gateaux partial differential D
1
G
f (a
1, a
2) in the first variable and, similarly, in the second variable, etc. In the conditions of Theorem 1.9, where
and
(and B denotes an open subset of F containing f (A)), we have
and (instead of [1.5])
(exercise *: see [ALE 87] section 2.2.2).
Let
be a normed affine space, F a locally convex space, A a non-empty open subset of
, [a, a + h] a segment contained in A and f : A → F a G-differentiable mapping. Then, the mean value theorem (Theorem 1.13(i)) remains valid (exercise
*: see [ALE 87], section 2.2.3) in the form
where |.|γ is a continuous semi-norm on F. The claims (ii) and (iii) of this theorem also hold after making analogous adjustments. From this, we deduce the following result:
Thus, we do not need to distinguish between Fréchet and Gateaux differentiation when talking about mappings of class C p (p > 0).
In this section,
.
(I) Euler condition
Let
be a locally convex affine space, A some non-empty open subset of
, a some point of A and
a mapping that admits a Lagrange first variation δJ (a) at the point a.
(II) Euler-Lagrange equation
The Euler–Lagrange equation is essentially the Euler condition applied to calculus of variations. A full treatment would require another volume
8
; we will content ourselves with briefly mentioning the simplest part, which is sufficient for our purposes. Let E be a Banach space and suppose that
, t
1 < t
2.
Let Ω1, Ω2 be non-empty open subsets of E and let
be a mapping (known as the Lagrangian in mechanics) of class C
1 with partial differential
. Let x
1, x
2 ∈ Ω1 and write
for the set of mappings x : [t
1, t
2] → Ω1 of class C
1 satisfying x (t
1) = x
1, x (t
2) = x
2 whose derivatives
take values in Ω2. The set
is an open subset of the normed affine space O + X, where X is equipped with the norm
; X is a Banach space whose elements satisfy the condition h (t
1) = h (t
2) = 0 (exercise). Let
The Lagrangian
is said to be regular at x = x
* if
is invertible in
for every t ∈ [t
1, t
2].
(III) Legendre condition
Suppose that
is of class C
2 and x
* satisfies the Euler–Lagrange equation. Expanding to second order, we have
where P and Q are evaluated at the point (t, x * (t)) and h is evaluated at the point t.
(I)
topology Let E be a real locally convex space and suppose that A is a non-empty subset of E. A smooth curve in A is defined as a mapping c : I → A of class C
∞, where I is a non-empty open interval of
, for example ]−1, 1[. Any such curve is said to be analytic if it is of class C
ω
.
If E is a complex locally convex space, suppose again that A is a non-empty subset of E. Then, A can be viewed as a subset A 0 of the real locally convex space E 0 obtained from E by restriction of the field of scalars ([P2], section 3.2.2 (II)), and any smooth curve in A 0 is said to be a smooth curve in A.
The topology of
is finer than the topology of E, so every open subset of E is open in
. If the space E is bornological ([P2], section 3.4.4
(I), Definition 3.61), it has the finest locally convex topology of all locally convex topologies coarser than the topology of
([KRI 97], Corollary 4.6). “Convenient” differential calculus is performed with mappings defined in open sets of
.
Recall that every Fréchet space and every Silva space is bornological and complete ([P2], sections 3.4.4 (I) and 3.8.2(II)). We have the following result ([KRI 97], Theorem 4.11):
(II)
spaces
The following definition will be useful:
Theorem 1.49 shows that Fréchet (and in particular Banach spaces) and Silva spaces are
spaces. Any quasi-complete locally convex space, and hence any
space, is convenient (exercise). Nonetheless,
spaces are sufficiently general for our purposes, so we will restrict attention to them to simplify the statements of results.
(II) Mappings of class
If E is a Banach space and f is of class C
r
(r ∈ {∞, ω}), then f is of class
. J. Boman showed the following result in 1967 ([KRI 97], Corollary 3.14):
In the real analytic case, the following result ([KRI 97], Chapter II, section 10.4) generalizes Boman’s theorem:
(III) Holomorphic mappings
Assume that
.
A holomorphic curve in a non-empty open subset A of a complex
space E is a holomorphic mapping from a disk in the complex plane, for example the open disk
of center 0 and radius 1, into A. A
-holomorphic mapping (or a mapping of class
) from A into F, where F is a complex
space, is a mapping that transforms the holomorphic curves in A into holomorphic curves in F. With this notation, the mapping f : A → F is holomorphic if and only if λ ∘ f ∘ c is a holomorphic function for every continuous linear form λ ∈ F
∨ and every holomorphic curve
. The following result ([KRI 97], Chapter II, section 7.9) generalizes the classical Hartogs theorem ([P2], section 4.3.2
(II), Corollary 4.80):
In this section,
.
Let E be a normed vector space. The topological notions of normal space and paracompact space ([P2], sections 2.3.10 and 2.3.11) inspire the following definitions:
Recall that (ψ i ) i ∈ I is a subordinate partition of unity of (U i ) i ∈ I if and only if supp (ψ i ) ⊂ U i , the family (supp (ψ i )) i ∈ I is locally finite and ∑ i ∈ I ψ i = 1 ([P2], section 2.3.12).
As a metrizable space, every Banach space is paracompact by Stone’s theorem ([P2], section 2.3.10, Theorem 2.57). The next result ([ABR 83], Propositions 5.5.18 and 5.5.19), whose proof was established by Bonic and Frampton in 1966, follows from the fact that any separable Banach space is a Lindelöf space ([P2], section 2.6.3) (exercise).
It was shown in [TOR 73] that every reflexive (not necessarily separable) Banach space is C 1-paracompact and that the separability condition is not required in Corollary 1.62:
This implies the result stated in ([P2], section 4.4.1, Theorem 4.88)
We can define
-regularity,
-normality and
-paracompactness of a locally convex space in the obvious ways; the statement of Theorem 1.59 still holds if C
∞ is replaced by
; moreover, if
is a
-regular Lindelöf space, then it is
-paracompact ([KRI 97], Proposition 16.2). We already know that any nuclear space E has a topology defined by a family of pre-Hilbert norms ([P2], section 3.11.3(III)). Each of these semi-norms is of class
in E − {0}. Nuclear Fréchet spaces (also known as
spaces), nuclear Silva spaces (also known as
spaces), and countable products of
spaces and
spaces are paracompact Lindelöf spaces (ibid.). The following result is analogous to Corollary 1.62 ([KRI 97], Theorem 16.10):
(I) Notion of the solution of a differential equation
Let I be an interval of
with non-empty interior
, Ω a non-empty open subset of
and f a mapping from I × Ω into E. We say that a mapping φ : I → E is a solution (or integral) of the differential equation
if the conditions (ODE 1,2,3) below are satisfied:
A Cauchy problem involves determining a function φ that is a solution of [1.22] and that satisfies the Cauchy condition:
If φ satisfies the conditions (ODE 1,2,3) and [1.23], then, for all t ∈ I,
Conversely, suppose that [1.24] holds, (ODE 1) is satisfied and t ↦ f (t, φ (t)) is locally λ-integrable. Then, [1.23] holds; furthermore, Lusin’s measurability criterion ([P2], section 4.1.6(II)) implies that the function t ↦ f (t, φ (t)) is λ-measurable for every locally absolutely continuous function φ : I → E if the conditions (Cat 1,2) below are satisfied:
Suppose further that:
(Cat
3) For all x
0 ∈ Ω and every r > 0 such that B
r
(x
0) ⊂ Ω, where B
r
(x
0) is the open ball of center x
0 and radius r in E, there exists a locally λ-integrable function m from I into
such that | f (t, x)| ≤ m (t) for all (t, x) ∈ I × B
r
(x
0).
Then, for every locally absolutely continuous function φ : I → E satisfying φ (I) ⊂ B r (x 0), the mapping t ↦ | f (t, φ (t))| is locally λ-integrable in I, so t ↦ f (t, φ (t)) is locally λ-integrable in I ([P2], section 4.1.2(I)).
(II) Existence theorem
(III) Uniqueness theorem The fourth Carathéodory condition can be stated as follows (reusing some of the notation of Cat 3):
(Cat
4) For all x
0 ∈ E and every real number r > 0 such that B
r
(x
0) ⊂ Ω, there exists a locally λ-integrable function k from I into
such that
Any locally Lipschitz function in the second variable clearly satisfies the conditions Cat 1,2,3,4, and we may therefore apply Theorem 1.70 to this function. We also have the following result:
(IV) Differential equations in implicit form
Let I be an open interval of
, t
0 a point of I, F a Banach space and Ω a non-empty open subset of F
n + 1, where n is a natural integer. Let g : I × Ω → F be a mapping of class C
1 and consider the differential equation:
Let η 0 = (η 0 0, η 0 1,…, η 0 n − 1) ∈ F n and ξ 0 ∈ F such that (η 0, ξ 0) ∈ Ω, g (t 0, η 0, ξ 0) = 0, and suppose that the following condition (Inv) is satisfied:
(Inv)
is invertible in
.
By the implicit function theorem (Theorem 1.30), there exists an open neighborhood J × U of (t 0, η 0) in I × F n , an open neighborhood V of ξ 0 in F, where U × V ⊂ Ω, and a mapping h : J × U → V of class C 1, such that, for all (t, η, ξ) ∈ J × U × V,
Therefore, any mapping ψ : J → F such that
and ψ
(n) (t) ∈ V for all t ∈ J is a solution of [1.26] in J if and only if ψ is a solution of the following differential equation on J, said to be in explicit form:
Let x i = y (i − 1) (i = 1,…, n) and x =(x 0, …, x n ). For (t, x) ∈ J × U × V, the differential equation [1.27] is equivalent to [1.22], where f = (f 1,…, f n ) and
The mapping f : J × U → F n is of class C 1 and hence locally Lipschitz in x by the mean value theorem (Theorem 1.13(i)). We can therefore apply Theorem 1.72 according to Remark 1.73. Let x 0 = (η 0 1, …, η 0 n − 1); any solution φ = (φ 0, …, φ n ) of [1.22] of class C 1 in J satisfies the Cauchy condition [1.23] if and only if φ = φ 1 is a solution of [1.26], of class C n in J and ψ (t 0) = η 0 1, …, ψ (n − 1) (t 0) = η 0 n − 1.
(I) Let I be an interval of
with non-empty interior
and let
. Consider the linear differential equation
where
and b : I → E are locally λ-integrable. The Carathéodory conditions Cat
1,2,3,4 are all satisfied. (In particular, setting f (t, x) = A (t).x + b (t), we have
which shows that Cat
4 is satisfied.) Hence, for all
and every x
0 ∈ E, there exists a unique solution φ (.; t
0, x
0) of [1.28] in I satisfying [1.23].
The linear equation [1.28] is said to be homogeneous if b = 0, in which case
(II) The set of solutions of [1.29] (where A is locally λ-integrable) is an
-vector space. Let φ
0 (.; t
0, x
0) be a solution of [1.29] in I satisfying [1.23]. The mapping x
0 ↦ φ
0 (t; t
0, x
0) is a bijective linear mapping Φ (t, t
0) from E onto E, and Φ(., t
0) is identical to the solution of the differential equation
for U (t 0) = 1 E . For all t 1, t 2, t 3 ∈ I, we have Φ(t 3, t 1) = Φ (t 3, t 2) ∘ Φ(t 2, t 1) and Φ(t 1, t 2) = Φ(t 2, t 1)− 1.
(III) The above shows that the general solution of [1.29] is of the form t ↦ Φ(t, t 0)ξ. The “variation of constants” method (exercise) allows us to obtain the following solution (defined in I) of [1.28] and the Cauchy condition [1.23]:
(IV) The integration of linear differential equations with constant coefficients is a classical problem and is performed using the Jordan normal form ([P1], section 3.4.3 (IV)): see, for example, [BOU 10], section 12.5.2.
Let I be an interval of
with non-empty interior, E a Banach space (see Remark 1.73), Ω a non-empty open subset of E, Λ a topological space and f a mapping from I × Ω × Λ into E. Write f
λ
(t, x) for the value of f at the point (t, x, λ) ∈ I × Ω × Λ. It is possible to show the following result ([BOU 76], Chapter 4, section 1.6, Theorem 3):
Suppose now that Λ is an open subset of a normed vector space F and that the mapping (t, x, λ) ↦ f
λ (t, x) is continuous with continuous partial differentials
13
and
. Then, f
λ
is locally Lipschitz in the second variable x (see section 1.5.1
(V)). Suppose further that the mappings x
0 : Λ → Ω: λ → x
0 (λ) and t
0 : Λ → I : λ ↦ t
0 (λ) are of class C
1 in Λ. For all λ
ο ∈ Λ, Theorem 1.80 implies that there exist an open neighborhood V of λ
0 in Λ and an open interval J ⊂ I such that t
0 (λ
0) ∈ J, where the sets V and J satisfy the following condition: for all λ ∈ V, t
0 (λ) ∈ J and there exists a solution φ
λ
= φ (.,t
0 (λ), x
0 (λ)) of [1.31] in J taking the value x
0 (λ) ∈ Ω at the point t
0 (λ). Then, we have the following result:
Note that, in the linear equation [1.32],
; moreover,
y
, A(t) ∘
y
, and
b
(t) belong to
. In the Cauchy condition, f
λ
(t
0
λ),
,
,
, and f
λ (t
0 (λ),
.