1. Introduction
We study questions of infinite-dimensional differential calculus in the setting of Keller’s
-theory [
1] (going back to [
2]). Applications to infinite-dimensional vector bundles are given, and also applications in the theory of locally convex Poisson vector spaces.
Differentiability properties of operator-valued maps. Our results are centred around the following basic problem: Consider locally convex spaces
X,
E and
F, an open set
and a map
to the space of continuous linear maps, endowed with the topology of uniform convergence on bounded sets. How are the differentiability properties of the operator-valued map
f related to those of
We show that if
is smooth, then also
f is smooth (Proposition 1). Conversely, exploiting the hypocontinuity of the bilinear evaluation map
we find natural hypotheses on
E and
F ensuring that smoothness of
f entails smoothness of
(Proposition 2; likewise for compact sets in place of bounded sets). Without extra hypotheses on
E and
F, this conclusion becomes false, e.g., if
is a non-normable, real, locally convex space with dual space
. Then,
is continuous linear and thus smooth, but
is the bilinear evaluation map taking
to
, which is discontinuous for non-normable
X (see [
3] (p. 2)) and hence not smooth in the sense of Keller’s
-theory. We also obtain results concerning finite-order differentiability properties, as well as real and complex analyticity. Furthermore,
can be replaced with the space
of continuous
k-linear maps
, if
are locally convex spaces. (Related questions also play a role in the comparative study of differential calculi [
1].) As a very special case of our studies, the differential
is
, for each
with
, locally convex spaces
E and
F, and
-map
on an open set
(see Corollary 1).
Applications to infinite-dimensional vector bundles. Apparently, mappings of the specific form just described play a vital role in the theory of vector bundles: If F is a locally convex space, M a (not necessarily finite-dimensional) smooth manifold and an open cover of M, then the smooth vector bundles , with fibre F, which are trivial over the sets , can be described by cocycles such that , is smooth (Proposition 3, Remark 7). Then, is smooth as a mapping to the space (see Proposition 1). In various contexts—for example, when trying to construct dual bundles—we are in the opposite situation: we know that each is smooth, and would like to conclude that also the mappings are smooth. Although this is not possible in general (as examples show), our results provide additional conditions ensuring that the conclusion is correct in the specific situation at hand. Notably, we obtain conditions ensuring the existence of a canonical dual bundle (Proposition 13). Without extra conditions, a canonical dual bundle need not exist (Example 2).
Besides dual bundles, we discuss a variety of construction principles of new vector bundles from given ones, including topological tensor products, completions, and finite or infinite direct sums. More generally, given a (finite- or infinite-dimensional) Lie group acting on the base manifold M, we discuss the construction of new equivariant vector bundles from given ones. Most of the constructions require specific hypotheses on the base manifold, the fibre of the bundle, and the Lie group.
As to completions, complementary topics were considered in the literature: Given an infinite-dimensional smooth manifold
M, completions of the tangent bundle with respect to a weak Riemannian metric occur in [
4] (p. 549), in hypotheses for a so-called
robust Riemannian manifold.
We mention that multilinear algebra and vector bundle constructions can be performed much more easily in an inequivalent setting of infinite-dimensional calculus, the convenient differential calculus [
3]. However, a weak notion of vector bundles is used there, which need not be topological vector bundles. Our discussion of vector bundles intends to pinpoint additional conditions ensuring that the natural construction principles lead to vector bundles in a stronger sense (which are, in particular, topological vector bundles).
The work [
5] was particularly important for our studies. For an open subset
U of a Fréchet space
E, smoothness of
is deduced from smoothness of
in the proof of [
5] (Proposition IV.6). A typical hypocontinuity argument already appears in the proof of [
5] (Lemma IV.7). In contrast to the local calculations in charts, the global structure on a dual bundle (and bundles of
k-forms) asserted in the first remark of [
5] (p. 339) is problematic if Keller’s
-theory is used, without further hypotheses.
Applications related to locally convex Poisson vector spaces. In the wake of works by Odzijewicz and Ratiu on Banach–Poisson vector spaces and Banach–Poisson manifolds [
6,
7], certain locally convex Poisson vector spaces were introduced [
8], which generalise the Lie–Poisson structure on the dual space of a finite-dimensional Lie algebra going back to Kirillov, Kostant and Souriau. By now, the latter spaces can be embedded in a general theory of locally convex Poisson manifolds (see [
9]; for generalisations of finite-dimensional Poisson geometry with a different thrust, cf. [
10]). Recall that many important examples of bilinear mappings between locally convex topological vector spaces are not continuous, but at least hypocontinuous (cf. [
11] for this classical concept). In
Section 12 and
Section 13, we provide the proofs for two fundamental results in the theory of locally convex Poisson vector spaces which are related to hypocontinuity. (These proofs were stated in the preprint version of [
8], but not included in the actual publication.) We show that the Poisson bracket associated with a continuous Lie bracket is always continuous (Theorem 1) and that the linear map
taking a smooth function to the associated Hamiltonian vector field is continuous (Theorem 2). Ideas from [
8] and the current article were also taken further in [
12] (
Section 13).
2. Preliminaries and Notation
We describe our setting of differential calculus and compile useful facts. Either references to the literature are given or a proof; the proofs can be looked up in
Appendix A.
Infinite-dimensional calculus. We work in the framework of infinite-dimensional differential calculus known as Keller’s
-theory [
1]. Our main reference is [
13] (see also [
14,
15,
16,
17]). If
, we let
and
for
. We write
and
. All topological vector spaces considered in the article are assumed Hausdorff, unless the contrary is stated. For brevity, Hausdorff locally convex topological vector spaces will be called locally convex spaces. As usual, a subset
M of a
-vector space is called
balanced if
for all
and
. The subset
M is called
absolutely convex if it is both convex and balanced. If
is a seminorm on a
-vector space
E, we write
for
and
. We also write
in place of
. If
E is a locally convex
-vector space, we let
be the dual space of continuous
-linear functionals
. We write
for the polar of a subset
. If
is a continuous
-linear map between locally convex
-vector spaces, we let
,
be the dual linear map. We say that a mapping
between topological spaces is a
topological embedding if it is a homeomorphism onto its image. We recall:
Definition 1. Let E and F be locally convex -vector spaces over and be an open subset. A map is called if it is continuous, in which case we set . Given and , we defineif the limit exists (using such that ). Let . We say that a continuous map is a-map
if the iterated directional derivativeexists for all such that and all , and if the mappings so obtained are continuous. Thus, . If is understood, we write instead of . As usual, -maps are also called smooth.
Remark 1. For , it is known that a map as before is if and only if f is and is (cf. [13] (Proposition 1.3.10)). Remark 2. If , it is known that a map as before is if and only if it is complex analytic
in the sense of [18] (Definition 5.6): f is continuous and for each , there exists a 0-neighbourhood such that and for all as a pointwise limit, where is a continuous homogeneous polynomial over of degree n, for each [13] (Theorem 2.1.12). Furthermore, f is complex analytic if and only if f is and is complex linear for all (see [13] (Theorem 2.1.12)). Complex analytic maps will also be called -analytic
or . Definition 2. If , then a map as before is called real analytic (or -analytic, or ) if it extends to a complex analytic mapping on some open neighbourhood of U in the complexification of E.
In the following, , unless the contrary is stated. We use the conventions and , for each . Furthermore, we extend the order on to an order on by declaring for each .
Remark 3. Compositions of composable -mappings are -mappings (see Proposition 1.3.4, Remark 2.1.13, and Proposition 2.2.4 in [13]). Thus, -manifolds modelled on locally convex -vector spaces can be defined in the usual way (see [13] (Chapter 3) for a detailed exposition). In this article, the word “manifold” (resp., “Lie group”) always refers to a manifold (resp., Lie group) modelled on a locally convex space. The following basic fact will be used repeatedly.
Lemma 1. For , let X, , and F be locally convex -vector spaces, be an open subset andbe a -map such that is k-linear, for each . Let and q be a continuous seminorm on F. Then, there exists a continuous seminorm p on X with , and continuous seminorms on for such thatfor all and . We shall also use the following fact:
Lemma 2. Let E and F be locally convex -vector spaces, be an integer and be a mapping such that is k-linear and symmetric for each . Let . Ifis , then also f is . Notably, f is continuous if h is continuous. -spaces, -spaces, -spaces, and -spaces. Recall that a topological space
X is said to be
completely regular if it is Hausdorff and its topology is initial with respect to the set
of all continuous real-valued functions on
X. Every locally convex space is completely regular, as with every Hausdorff topological group (cf. [
19] (Theorem 8.2)). Compare [
20,
21] for the following.
A topological space X is called a k-space if it is Hausdorff and a subset is closed if and only if is closed in K for each compact subset . Every metrisable topological space is a k-space, and every locally compact Hausdorff space. A Hausdorff space X is a k-space if and only if, for each topological space Y, a map is continuous if and only if f is k-continuous in the sense that is continuous for each compact subset . If X is a k-space, then also every subset which is open or closed in X, when the induced topology is used on M.
A topological space
X is called a
-space if it is Hausdorff and a function
is continuous if and only if
f is
k-continuous. Then also a map
to a completely regular topological space
Y is continuous if and only if it is
k-continuous (as the latter condition implies continuity of
for each
). For more information, cf. [
22].
Every
k-space is a
-space. The converse is not true:
is known to be a
-space for each set
I (see [
22]). If
I has cardinality
, then
is not a
k-space. (If
was a
k-space, then a certain non-discrete subgroup
G of
constructed in [
23] would be discrete, which is a contradiction (see [
13] (Remark A.6.16 (a)) for more details). Compare also [
22].)
The following facts are well known (cf. [
22]):
Lemma 3. - (a)
If a -space X is a direct product of Hausdorff spaces and , then is a -space.
- (b)
Every open subset U of a completely regular -space X is a -space in the induced topology.
Notably, U is a -space for each open subset U of a locally convex space E which is a -space. If is a -space, then also E.
Following [
8], a topological space
X is called a
-space if the Cartesian power
is a
k-space for each
, using the product topology. A Hausdorff space
X is called
hemicompact if
for a sequence
of compact subsets
such that each compact subset of
X is a subset of some
. Hemicompact
k-spaces are also called
-spaces. If
X and
Y are
-spaces, then the product topology makes
a
-space. Notably, every
-space is a
-space. See [
24,
25] for further information. Finite products of metrisable spaces being metrisable, every metrisable topological space is a
-space. Recall that a locally convex space
E is said to be a
Silva space or (DFS)-space if it is the locally convex inductive limit of a sequence
of Banach spaces such that each inclusion map
is a compact operator. Every Silva space is a
-space (see, e.g., [
13] (Proposition B13.13(g))).
Spaces of multilinear maps. Given
, locally convex
-vector spaces
and
F, and a set
of bounded subsets of
, we write
or
for the space of continuous
k-linear maps
, endowed with the topology
of uniform convergence on the sets
. Recall that finite intersections of sets of the form
yield a basis of 0-neighbourhoods for this (not necessarily Hausdorff) locally convex vector topology, for
U ranging through the 0-neighbourhoods in
F and
B through
. If
, then
is Hausdorff. If
, we abbreviate
. If
and
, we abbreviate
,
and
. We write
for the group of all automorphisms of the locally convex
-vector space
E. If
is the set of all bounded, compact, and finite subsets of
, respectively, we shall usually write “
b,” “
c,” and “
p” in place of
. For example, we shall write
,
, and
.
Remark 4. Let and F be complex locally convex spaces and be a map, defined on an open subset U of a real locally convex space. Let or . Since is a closed real vector subspace of , the map f is as a map to if and only if f is as a map to (see [13] (Lemma 1.3.19 and Exercise 2.2.4)). Given a
-map
as in Definition 1, we define
and
for
such that
.
Hypocontinuous multilinear maps. Beyond normed spaces, typical multilinear maps are not continuous, but merely hypocontinuous. Hypocontinuous bilinear maps are discussed in many textbooks. An analogous notion of hypocontinuity for multilinear maps (to be described presently) is useful to us. It can be discussed similarly to the bilinear case.
Lemma 4. For an integer , let be a separately continuous k-linear mapping and such that, for each , the mapis continuous. Let be a set of bounded subsets of . Consider the conditions: - (a)
For each and each 0-neighbourhood , there exists a 0-neighbourhood such that .
- (b)
The -linear map is continuous.
- (c)
is continuous, for each .
Then (a)
and (b)
are equivalent, and (b)
implies (c)
. Ifthen (a)
, (b)
, and (c)
are equivalent. Definition 3. A k-linear map β which satisfies the hypotheses and Condition (a) of Lemma 4 is called -hypocontinuous in its arguments . If , we also say that β is -hypocontinuous in the k-th argument. Analogously, we define -hypocontinuity of β in the j-th argument, if and a set of bounded subsets of are given.
We are mainly interested in b-, c-, and p-hypocontinuity, viz., in -hypocontinuity with respect to the set of all bounded subsets of , the set of all compact subsets, and the set of all finite subsets, respectively. If and are sets of bounded subsets of such that and is -hypocontinuous in its variables , then is also -hypocontinuous in the latter. The following is obvious from Lemma 4 (c) (as the elements of a convergent sequence, together with its limit, form a compact set):
Lemma 5. If is c-hypocontinuous in some argument, or in its arguments for some , then β is sequentially continuous.
In many cases, separately continuous bilinear maps are automatically hypocontinuous. Recall that a subset
B of a locally convex space
E is a
barrel if it is closed, absolutely convex, and absorbing. The space
E is called
barrelled if every barrel is a 0-neighbourhood. See Proposition 6 in [
11] (Chapter III, §5, no. 3) for the following fact.
Lemma 6. If is a separately continuous bilinear map and is barrelled, then β is -hypocontinuous in its second argument, with respect to any set of bounded subsets of .
Evaluation maps are paradigmatic examples of hypocontinuous multilinear maps.
Lemma 7. Let and F be locally convex -vector spaces and be a set of bounded subsets of with . Then, the -linear mapis -hypocontinuous in its arguments . If and is barrelled, then is also hypocontinuous in the first argument, with respect to any locally convex topology on which is finer than the topology of pointwise convergence, and any set of bounded subsets of . Lemma 8. Consider locally convex spaces and F with and a k-linear map .
- (a)
If β is sequentially continuous, then the composition is continuous for each continuous function on a topological space X which is metrisable or satisfies the first axiom of countability.
- (b)
If β is c-hypocontinuous in its arguments for some and X is a -space, then is continuous for each continuous function .
Lipschitz differentiable maps. In
Section 7, it will be useful to work with certain Lipschitz differentiable maps, instead of
-maps. We briefly recall concepts and facts.
Definition 4. Let E and F be locally convex -vector spaces, be open and be a map. We say that f is locally Lipschitz continuous
or if it has the following property: For each and continuous seminorm q on F, there exists a continuous seminorm p on E such that and Given , we say that f is if f is and is for each such that .
Every
-map is
(see, for example, [
13] (Exercise 1.5.4)). As a consequence, for each
, every
-map is
. Notably, every smooth map is
. Moreover, a
-map with finite
r is
if and only if
is
. The following facts are known, or part of the folklore.
Lemma 9. For locally convex spaces over and , we have:
- (a)
A map to a direct product of locally convex spaces is if and only each component is ;
- (b)
Compositions of composable -maps are ;
- (c)
Let F be a locally convex space and be a vector subspace which is closed in F, or sequentially closed. Then, a map is if and only if it is as a map to F.
- (d)
A map to a projective limit of locally convex spaces is if and only if is for all , where is the limit map.
Our concept of local Lipschitz continuity is weaker than the one in [
13] (Definition 1.5.4).
The compact-open -topology. If
E and
F are locally convex
-vector spaces,
is an open set and
, then the vector space
of all
-maps
carries a natural topology (the “compact-open
-topology”), namely the initial topology with respect to the mappings
for
such that
, where the right-hand side is endowed with the compact-open topology. Then,
is a locally convex
-vector space. If
F is a complex locally convex space, then also
. See, e.g., [
13] (§1.7) for further information, or [
26].
3. Differentiability Properties of Operator-Valued Maps
Let , , and . In this section, we establish the following proposition.
Proposition 1. Let , , and F be locally convex -vector spaces, X be a locally convex -vector space, and be an open subset. Let be a map such thatis . Then, the following holds: - (a)
f is as a map to .
- (b)
If , then f is as a map to .
Furthermore,for all with (resp., , in (b))
, all , , and . Corollary 1. Let E and F be locally convex -vector spaces and be a -map on an open subset , where . Then, the following holds:
- (a)
The map , is , for each such that .
- (b)
The map is , for each such that .
Furthermore, , for all with (resp., , all , and .
Proof. For each
such that
, the map
is
(see [
13] (Remark 1.3.13 and Exercise 2.2.7)), and
is
k-linear for each
, by [
13] (Proposition 1.3.17). Moreover,
. Thus, Proposition 1 applies with
in place of
f and
in place of
r. □
Given a topological space X and locally convex space F, we endow the space of continuous F-valued functions on X with the compact-open topology. It is known that this topology coincides with the topology of uniform convergence on compact sets. The next lemma will be useful when we discuss mappings to .
Lemma 10. Let X, E, and F be locally convex -vector spaces, and be open subsets, and be a -map, with . Then, also the mapis . If and f admits a complex analytic extension for suitable open neighbourhoods of U in and of W in , then is real analytic. Proof. We first assume that
, and proceed by induction. For
, the assertion is well known (see, e.g., [
13] (Proposition A.6.17)). Now assume that
. Given
and
, there exists
such that
, where
. Consider
Then,
, by the Mean Value Theorem. The integrand being continuous, also
g is continuous (by the Theorem on Parameter-Dependent Integrals, [
13] (Lemma 1.1.11)). Hence,
is continuous, by induction, and hence
as
, where
with
Since k is , the map is , by the inductive hypothesis. Notably, is continuous and hence is . Now, being with a -map, is .
The case . If f is , then f is for each . Hence, is for each (by the case already treated), and thus is .
Final assertion. By the
-case already treated, the map
is
. The restriction map
being continuous
-linear and thus
, it follows that the composition
is
and thus complex analytic. Since
extends
, we see that
is real analytic. □
Proof of Proposition 1. (a) Abbreviate
. Because
is a closed
-vector subspace of
and carries the induced topology,
f will be
as a map to
if we can show that
f is
as a map to
(see [
13] (Lemma 1.3.19 and Exercise 2.2.4)). Since
is
and
, the latter follows from Lemma 10. This is obvious unless
and
. In this case, the map
admits a
-analytic extension
to an open neighbourhood
Q of
in
. For each
, there exists an open, connected neighbourhood
of
x in
and a balanced, open 0-neighbourhood
such that
and
. Let
. Then,
is a
-analytic map which vanishes on
. Hence,
, by the Identity Theorem (see [
13] (Theorem 2.1.16 (c))). Then,
for all
such that
, by continuity. This implies that the map
is well defined. Since
g is
-analytic, the final statement of Lemma 10 applies.
(b) We prove the assertion for
first; then, also the case
follows. If
, let
. Given an open 0-neighbourhood
and bounded subset
, let
q be a continuous seminorm on
F such that
. By Lemma 1, there exist continuous seminorms
p on
X and
on
for
such that
and
for all
and all
. Since
B is bounded, we have
Choose
such that
. For each
, we get
for each
and thus
. Hence,
entailing that
f is continuous.
Induction step: Now, assume that
. Given
and
, there exists
such that
, where
. Consider
Then,
g is
and hence
, as a consequence of [
27] (Propositions 7.4 and 7.7). Since
is
k-linear in
v, it follows that
is continuous, by induction. As a consequence,
as
, where
with
Since h is and is k-linear in v, the map is , by induction. Hence, is continuous and thus f is . Now, f being with a -map, f is .
The case , . By Remark 4, we may assume that (the case then follows). Given , let be as in the proof of (a). Identifying with , the mapping g is complex k-linear in the second variable. Hence is -analytic, by the -case already discussed. Because the map , is continuous -linear, the composition is -analytic. However, this mapping extends . Hence, is real analytic and hence so is f, using that the open sets form an open cover of U.
Formula for the differentials: Let
with
,
,
and
. Exploiting that
,
is continuous and linear, we deduce that
for
f as a map to
. If
, the same calculation applies to
f as a mapping to
. □
For the special case of (a) when
and
X as well as
are metrisable, see already [
1] (Lemma 0.1.2).
4. Compositions with Hypocontinuous -Linear Maps
We study the differentiability properties of compositions of the form , where is a k-linear map which need not be continuous.
Lemma 11. Let be an integer, , X, and F be locally convex -vector spaces, be a k-linear map, and be a -map on an open subset . Assume that
- (a)
β is sequentially continuous and X is metrisable; or
- (b)
For some , the k-linear map β is c-hypocontinous in its variables . Moreover, is a -space, or and X is a -space, or and X is a -space.
Then, is a -map.
Proof. The case was treated in Lemma 8. We first assume that .
(a) Assuming (a), let
,
, and
be a sequence in
such that
as
and
for all
. Using the components of
, we can write the difference quotient
as the telescopic sum
which converges to
as
, using the sequential continuity of
. By Lemma 8,
is continuous, whence
is
. If
, then
is a
-map and
is
by induction; thus
is
. If
, the preceding shows that
is
for each
, whence
is
.
(b) If
is a
-space, then
and
U are
-spaces. By Lemma 5,
is sequentially continuous. The argument from (a) shows that
exists for all
and is given by (
5). Thus
is continuous, by Lemma 8, and thus
is
. Let
f be
now and assume
is
with
rth differential of the form
for
and
, where
ranges through
k-tuples of (possibly empty) disjoint sets
with
, and the following notation is used: For
, we let
be the cardinality of
and define
if
are the elements of
, abbreviating
(if
is empty, the symbol
is to be ignored). Holding
fixed, we can apply the case
to the function
and find that, for each
and
, the directional derivative at
x in the direction
exists and is given by
Thus, also
is of the form (
6), with
in place of
r. Using Lemma 8, we deduce from the preceding formula that the map
is continuous. Thus,
is continuous, by Lemma 2, and thus
is
.
If , then is for each and hence (still assuming (b)).
If
and
X is only assumed
, then
is continuous by the case
. Moreover, the restriction
is
for each finite-dimensional vector subspace
, by case (a). Hence,
f is
(and thus
) as a mapping to a completion of
F (see [
18] (Theorem 6.2)). Then,
f is also
as a map to
F, as all of its iterated directional derivatives are in
F.
Both in (a) and (b), it remains to consider the case
. Then,
f admits a
-analytic extension
, defined on an open neighbourhood
of
U in
. The complex
k-linear extension
of
is given by
for
with
and
for
. By the latter formula,
is sequentially continuous in the situation of (a), and
c-hypocontinuous in its arguments
in the situation of (b). The case
shows that
is complex analytic. As this mapping extends
, the latter map is real analytic. In case (b), we used here that
is a
-space. □
Moreover, the following variant will be useful.
Lemma 12. Let , , , and F be locally convex -vector spaces, and , be open subsets. Let and be a -bilinear map. Assume that is finite-dimensional and β is c-hypocontinuous in its first variable. Then, for all -maps and , also the following map is : Proof. We first prove the assertion for (from which the case follows). If , we have to show that g is continuous. If , then has a compact neighbourhood in . Then, is compact, and thus is continuous, by c-hypocontinuity. Hence, is continuous, where is the projection onto the first factor. Since is an open cover of , the map g is continuous.
Since
is sequentially continuous by Lemma 5, we see as in the preceding proof that the directional derivative
exists for all
and
, and is given by
Note that
and
are
-mappings
. Moreover,
and
are
-maps
(cf. Remark 1). By induction, the right-hand side of (
7) is a
-map. Hence,
g is
.
The case follows from the case as in the preceding proof. □
Remark 5. In a setting of differential calculus in which continuity on products is replaced with k-continuity (as championed by E. G. F. Thomas), every bilinear map β which is c-hypocontinuous in the second factor is smooth (see [28] (Theorem 4.1)); smoothness of for a smooth map f then follows from the Chain Rule (cf. also [29]). Likewise, β is smooth in the sense of convenient differential calculus. 5. Differentiability Properties of
For , the following result is essential for our constructions of vector bundles.
Proposition 2. Let , , , , and F be locally convex -vector spaces, X be a locally convex -vector space, and be an open subset. Then, the following holds.
- (a)
If is a -space, or and is a -space, or and is a -space, or all of the vector spaces are finite dimensional, thenis for each -map . - (b)
If holds and, moreover, is a -space or and is a -space, or and is a -space, thenis for each -map such that is a symmetric k-linear map for each . - (c)
If X is finite-dimensional, , and is barrelled, then , is for each -map .
- (d)
If all of the spaces are normable, then is for each -map .
Proof. Let be the evaluation map, which is c-hypocontinuous in its arguments by Lemma 7.
(a) Assuming the respective -property, the map is , by Lemma 11 (b). If are finite-dimensional, then equals , whence the conclusion of (a) is a special case of (d).
(b) By Lemma 11 (b), the map
is
, as
with
,
, which is continuous
-linear. Then, also
is
, by Lemma 2.
(c) The bilinear map is c-hypocontinuous in its first argument, by Lemma 7. Hence, is , by Lemma 12.
(d) If
are normable, then the evaluation map
is continuous
-linear and hence
, whence also
is
. □
Remark 6. If X and all of are metrisable, then the topological space is metrisable and hence a k-space. If X and all of are -spaces, then also is a -space and hence a k-space. In either case, we are in the situation of (a).
6. Infinite-Dimensional Vector Bundles
In this section, we provide foundational material concerning vector bundles modelled on locally convex spaces (cf. also [
13] (Chapter 3)). Notably, we discuss the description of vector bundles via cocycles, and define equivariant vector bundles.
Let , , and . The word “manifold” always refers to a manifold modelled on a locally convex space. Likewise, the Lie groups that we consider need not have finite dimension.
Definition 5. Let M be a -manifold and F be a locally convex -vector space. An-vector bundle of class
over
M, with typical fibre
F, is a -manifold E, together with a surjective -map and endowed with an -vector space structure on each fibre , such that, for each , there exists an open neighbourhood of x and a -diffeomorphism(called a “local trivialisation”) such that for each and the map is -linear (and hence an isomorphism of topological vector spaces, if we give the topology induced by E), where is the projection. In the situation of Definition 5, let
be an atlas of local trivialisations for
E,
i.e., a family of local trivialisations
of
E whose domains
cover
M. Then, given
, we have
for
,
, for some function
Here,
is
, as
is
in
. By Proposition 1,
is a
-map, and as a map to
, it is at least
(if
). Note that the “transition maps”
satisfy the “cocycle conditions”
Proposition 3. Let , . Assume that
- (a)
M is a -manifold modelled on a locally convex -vector space Z;
- (b)
E is a set and a surjective map;
- (c)
F is a locally convex -vector space;
- (d)
is an open cover of M;
- (e)
is a family of bijections such that for all ;
- (f)
depends -linearly on , for all , ;
- (g)
, is a -map.
Then, there is a unique -vector bundle structure of class on E making a local trivialisation for each .
Proof. For
, let
be the projection onto the first component. As the maps
are
, there is a uniquely determined
-manifold structure on
E making
a
-diffeomorphism for each
. Given
, we pick
with
; we give
the unique
-vector space structure making the bijection
an isomorphism of vector spaces. It is easy to see that the vector space structure on
is independent of the choice of
, and it is easily verified that we have turned
E into an
-vector bundle of class
with the asserted properties. □
Remark 7. Let M be a -manifold, F be a locally convex -vector space, be an open cover of M, and be a family of maps satisfying the cocycle conditions and such thatis , for all . Using Proposition 3, the usual construction familiar from the finite-dimensional case provides an -vector bundle of class , with typical fibre F, and a family of local trivialisations , whose associated transition maps are the given ’s. The bundle E is unique up to canonical isomorphism. Combining Proposition 3 and Proposition 2, we obtain:
Corollary 2. Retaining the hypotheses(a)–(f) from Proposition 3 but omitting (g), consider the following conditions:
for all , , and is ;
for all , , and is ;
- (i)
is a -space, or and is a -space, or and is a -space;
- (ii)
and F is barrelled;
- (iii)
F is normable.
If holds as well as (i) or (ii), then the conclusions of Proposition 3 remain valid. They also remain valid if and (iii) hold.
Example 2 below shows that Conditions (a)–(f) and alone are not sufficient for the conclusion of Proposition 3, without extra conditions on Z and F. Note that (i) is satisfied if both Z and F are metrisable, or both Z and F are -spaces.
Equivariant vector bundles. Beyond vector bundles, we shall discuss equivariant vector bundles in the following, i.e., vector bundles together with an action of a (finite- or infinite-dimensional) Lie group G. Choosing as a trivial group, we obtain results about ordinary vector bundles (without a group action), as a special case.
For the remainder of this section, and also in
Section 7, let
,
,
, and
with
. Let
G be a
-Lie group (modelled on a locally convex
-vector space
Y) and
M be a
-manifold. We assume that a
-action
is given. Then,
is called a
G-manifold of class .Definition 6. An equivariant
-vector bundle of class
over a G-manifold of class is an -vector bundle of class , together with a -actionsuch that for all , and is -linear. In other words, takes fibres linearly to fibres and coincides with on the zero section. The mapping is then equivariant in the sense that .
Example 1. If M is a G-manifold of class , with , then the tangent bundle is an equivariant -vector bundle of class in a natural way, with . In fact, the action has a tangent map , which is . Let be the 0-section. Identifying with in the usual way, we obtain a -map via It is easy to see that for and , whence and . Clearly, β is an action making an equivariant -vector bundle of class over the G-manifold M.
Induced action on an invariant subbundle. Given an
-vector bundle
of class
, with typical fibre
F, we call a subset
a
subbundle if there exists a sequentially closed
-vector subspace
such that for each
there exists a local trivialisation
of
E such that
. It readily follows from [
13] (Lemma 1.3.19 and Exercise 2.2.4) that there is a unique
-vector bundle structure of class
on
making
a local trivialisation of
, for each local trivialisation
as before. Then, the inclusion map
is
, and a mapping
from a
-manifold
N to
E with image in
is
as a mapping to
E if and only if its co-restriction to
is
, by the facts just cited. In the preceding situation, suppose that a
-Lie group
G acts
on
M and
E is an equivariant vector bundle of class
with respect to the action
. If
is invariant under the
G-action, i.e., if
, as a special case of the preceding observations, we deduce from the
-property of
that
and thus also
is
. We can summarise as follows.
Proposition 4. If E is an equivariant -vector bundle of class over a G-manifold M, then the action induced on any G-invariant subbundle is and thus makes the latter an equivariant -vector bundle of class .
7. Completions of Vector Bundles
Let be an equivariant -vector bundle of class , as in Definition 6, with typical fibre F and G-actions and . Assume that . Our goal is to complete the fibre of the bundle, i.e., to find a G-equivariant vector bundle whose typical fibre is a completion of the locally convex space F, and which contains E as a dense subset.
Let
be a completion of
F such that
and, for each
, let
be a completion of
such that
. We may assume that the sets
are pairwise disjoint for
. Consider the (disjoint) union
We shall turn
into an equivariant vector bundle. Consider the map
, defined using the continuous extension
of the linear map
via
for
,
, and
. It is clear that
makes
a
G-set. Let
be the map taking elements from
to
x. Then,
is
G-equivariant. If
is a local trivialisation of
E and
,
, we define
Then, the following holds:
Proposition 5. can be made an equivariant -vector bundle of class over the G-manifold M, such that is a local trivialisation of for each local trivialisation ψ of E.
Remark 8. Omitting the hypothesis that , assume instead that E is an equivariant -vector bundle of class . That is, both E and M are -manifolds (each admitting an atlas with transition maps of class ), a family of local trivialisations can be chosen with -transition maps, and the G-actions on E and M are . Then, also is an equivariant vector bundle of class (and hence of class .
Extension of differentiable maps to subsets of the completions. To enable the proof of Proposition 5, we need to discuss conditions ensuring that a -map (with locally convex spaces E and F) can be extended to a -map on an open subset of the completion of E, or at least to a -map. Although this is not possible in general, it is possible if F is normed and r is finite. This will be sufficient for our purposes. The natural framework for the discussion of the problem is not -maps, but Lipschitz differentiable maps, as in Definition 4.
Proposition 6. Let E be a locally convex -vector space, be a Banach space over , be open and be an -map, where . Let be a completion of E such that . Then, f extends to an -map on an open subset which contains U as a dense subset.
The following lemma enables an inductive proof of Proposition 6.
Lemma 13. Let , X be a locally convex -vector space, and be locally convex -vector spaces, with completions , and , respectively. Let be open and be a map such that is k-linear over for each . Assume that there exists an -map which extends f, defined on an open set in which is dense. Then, there exists an -mapwhich extends f, for some open subset in which U is dense. The maps are k-linear over , for each . Proof. For each
, there exists an open neighbourhood
of
x in
and a balanced, open 0-neighbourhood
such that
. After shrinking
, we may assume that
, whence
is dense in
. Given
such that
, consider the map
This map vanishes, because it is continuous and vanishes on the dense subset
. As a consequence, we obtain a well-defined map
for
,
and
with
. As
is
in
and these sets form an open cover of
, we see that
is
. Given
, the set
is dense in the open set
. Since
,
, and
f coincide on the set
, it follows that the continuous maps
and
coincide on the set
in which the former set is dense. Hence, setting
, a well-defined map
as in (
11) is obtained if we set
The final assertion follows by continuity from the k-linearity of the mappings for . □
Proof of Proposition 6. We proceed by induction on .
The case . Given
, there exists a continuous seminorm
q on
E such that
and
Then,
is a closed vector subspace of
E and
for
defines a norm on
making the map
,
continuous linear. By (
12), we have
for all
such that
. Hence,
is a well-defined map. Note that
is the open ball
in
. Let
be the completion of the normed space
; the extended norm will again be denoted by
. Applying (
12) to representatives, we see that
Hence,
h satisfies a global Lipschitz condition (with Lipschitz constant 1), and hence
h is uniformly continuous, entailing that
h extends uniquely to a uniformly continuous map
on the corresponding open ball
in
. Then,
for all
, by continuity. Let
be the continuous extension of the continuous linear map
. Then,
is an open neighbourhood of
x in
such that
. Moreover,
is a continuous map extending
, which furthermore satisfies
where we use the continuous seminorm
extending
q. Then
is an open subset of
and
is dense in
. Given
, the set
is dense in the open set
. Since
it follows that
. Hence
is a well-defined map. Since
is
for each
(by (
13)), the map
is
. Furthermore,
extends
f by construction.
Induction step. If
f is
, then
f extends to an
-map
on an open subset
such that
, and
extends to an
-map
on an open subset
W of
, by induction. Using Lemma 13, we find an open neighbourhood
V of
U in
and an
-map
which extends
. After replacing
and
V with their intersection, we may assume that
. If
and
, there exist open neighbourhoods
Q of
and
P of
in
, and
such that
. Then, the map
is continuous, being given by a parameter-dependent weak integral with continuous integrand. For
in the dense subset
of the set
, the Mean Value Theorem implies that
Then,
for all
, by continuity. Thus,
as
. Hence,
. Since
g is
, it follows that
is
. □
The conclusion of Proposition 6 becomes false in general if the Banach space
F is replaced by a complete locally convex space. In fact, there exists a smooth map
from a proper, dense vector subspace
E of
to a suitable power of
, which has no continuous extension to
for any
(see
Appendix B). Nonetheless, we have the following result.
Proposition 7. Let , X be a locally convex -vector space, and be locally convex -vector spaces, with completions , and , respectively. Let be open and be a mapping such that is k-linear over for each . If f is for some (resp., for some , then there exists a unique mapwhich is (resp., and extends f. The maps are k-linear over , for each . Proof. Abbreviate and . Assume first that . Since -maps are continuous and is dense in , there is at most one map with the asserted properties. We may therefore assume that . We may also assume that F is complete. Then, for some projective system of Banach spaces and continuous linear maps , with limit maps . We claim that has an -extension , for each . If this is true, then for , by uniqueness of continuous extensions. Hence, by the universal property of the projective limit, there exists a unique map such that . Then, and hence . Furthermore, is , by Lemma 9 (d). To prove the claim, note that Proposition 6 yields an -extension of to an open subset , which contains as a dense subset. Now, Lemma 13 yields an open subset in which U is dense, and an -extension of . Then, is as desired.
We now consider the case
. If
, by the density of
in
, for any real analytic extension
and
, the map
will be
k-linear over
. We may therefore assume that
. Let
be a
-analytic extension of
f, defined on an open subset
such that
. For each
, there exist an open
x-neighbourhood
and balanced open 0-neighbourhoods
and
such that
. We claim that there exists a
-analytic map
such that
. For
, the intersection
is connected and meets
whenever it is non-empty. Hence, by the Identity Theorem,
and
coincide on the intersection of their domains. We therefore obtain a well-defined
-analytic map
such that
for each
, using the open subset
of
. For each
, the map
is
k-linear over
. Using the Identity Theorem, we see that
is
k-linear over
for each
, and hence for each
by the Identity Theorem. By the case
,
g has a
-analytic extension
. Since
and
is dense in
, we deduce that
; we therefore obtain a map
for
,
. Since
is a
-analytic extension for
, the function
is
-analytic. To prove the claim, consider for
and
the
-analytic map
If
and
, we have for all
whence
for all
and
, by the Identity Theorem. Thus,
,
if
is a well-defined
-analytic extension of
. □
Proof of Proposition 5. It suffices to prove the strengthening described in Remark 8. Let
be a family of local trivialisations
of an
-vector bundle
E such that each local trivialisation is some
. Let
be the corresponding cocycle and
be the
-map
, which is
-linear in the second argument. By Proposition 7, there is a unique
-map
which extends
, and
is
-linear in the second argument. Thus, we obtain a map
By continuity and density, for all
, we have
for all
. Thus,
for all
. For all
, we have
as both sides are continuous in
and equality holds for
y in the dense subset
F of
; thus,
. Notably,
for all
and
, entailing that
. By the preceding, the
satisfy the cocycle conditions. Let
and
be as in (
8) and (
9); define
as in (
10), replacing
with
. For all
and
, we then have that
holds for all
, as equality holds for all
. As an analogue of Proposition 3 holds with
-maps in place of
-maps, we get a unique
-vector bundle structure of class
on
making
a local trivialisation for each
.
It is apparent that
is an action, and
is taken
-linearly to
by
, for each
and
. It only remains to show that
is
. To this end, let
and
; we show that
is
on
for some open neighbourhood
U of
in
G and an open neighbourhood
V of
in
M. Indeed, there exists a local trivialisation
of
E over an open neighbourhood
W of
in
M. The action
being continuous, we find an open neighbourhood
U of
in
G and an open neighbourhood
V of
in
M over which
E is trivial, such that
. Let
be a local trivialisation of
E over
V. Then,
for an
-map
, which is
-linear in the third argument. By Proposition 7, there is a unique extension of
A to an
-map
and the latter is
-linear in its third argument. For all
and
, we then have
for all
, as equality holds for all
. Thus,
is
. □
8. Tensor Products of Vector Bundles
Throughout this section, let , , , and such that . Let G be a -Lie group modelled on a locally convex -vector space Y, M be a -manifold modelled on a locally convex -vector space Z, and be a -action. For , let be an equivariant -vector bundle of class over M, whose typical fibre is a locally convex -vector space . Let be the G-action of class . Consider the set of all pairs of local trivialisations of and trivialising these over the same open subset of M. Using an index set I, we have , where is a local trivialisation of for , for each . Apparently, is an open cover of M.
For our first result concerning tensor products, Proposition 8, we assume that
is finite-dimensional. Then, fixing a basis
for
, the map
,
is an isomorphism of
-vector spaces. We give
the topology
, making
a homeomorphism. This topology makes
a locally convex
-vector space and
an isomorphism of topological
-vector spaces. It is easy to check (and well known) that the topology
is independent of the chosen basis. Let
be the basis dual to
. Our goal is to make the union
an equivariant
-vector bundle of class
over
M, with typical fibre
; the tensor products
are chosen pairwise disjoint here for
. Let
be the mapping which takes
to
x.
We define
via
for
and
, where
is the projection.
Given
and
, we have
for all
and
, where
is
and
an
-linear mapping. Then,
,
is
, and
for
and
, where
As , is a continuous linear map (where is the projection onto the -component), in view of the preceding formula is . Thus, by Proposition 3, there is a unique -vector bundle structure of class on making each a local trivialisation.
Note that , for , , defines an action of G on by -linear mappings, which makes an equivariant mapping and such that is -linear on for all and .
To show that
is
, let
and
. We pick
such that
. The mapping
being continuous, we find open neighbourhoods
U of
in
G and
V of
in
M such that
. There is
such that
. For
,
,
and
, we have
for some
-map
, which is
-linear in the final argument. Define
,
; then,
is
. If
,
and
, then
equals
which is a
-function of
. As a consequence,
is
and thus
, being
locally, is
. We summarise as follows.
Proposition 8. Let G be a -Lie group and M be a G-manifold of class . Let and be equivariant -vector bundles of class over M. If the typical fibre of is finite-dimensional, then , as defined above, is an equivariant -vector bundle of class over M.
Instead of
(as before) assume that
and
are Fréchet spaces and the modelling spaces of
G and
M are metrisable. The completed projective tensor product
over
then is a Fréchet space (cf. [
30] (p. 438, lines after Definitions 43.4)). We define
where the
for
are chosen pairwise disjoint. Let
be the map taking
to
x. Define
via
for
and
, where
is the projection. Note that
,
for
,
,
defines an action of
G on
E which makes
an equivariant mapping. We show:
Proposition 9. admits a unique structure of equivariant -vector bundle of class over M such that is a local trivialisation for each .
Proof. The uniqueness for prescribed local trivialisations is clear. Let us show the existence of the structure. Given
and
, we have
for all
and
, where
is
and
an
-linear mapping. By Proposition 1 (a), the map
is
. Now,
being continuous
-bilinear (as recalled in Lemma 14), we deduce that
is
. Hence,
,
is
, by Proposition 2 (a). We easily check that
holds for
as just defined, for all
and
. Hence,
can be made an
-vector bundle of class
in such a way that each
is a local trivialisation, by Proposition 3. Note that
is
-linear on
for all
and
. To show that
is
, let
,
,
i,
U,
V,
j and the
-map
be as in the proof of Proposition 8. By Proposition 1 (a),
,
is
. Hence,
is
, by the Chain Rule and Lemma 14. Using Proposition 2 (a), we find that the map
,
is
. We easily verify that
for all
. Thus,
is
in
, which completes the proof. □
We used the following fact:
Lemma 14. Let , , , and be Fréchet spaces over . Then, the following bilinear map is continuous: Proof. Let
be compact,
q be a continuous seminorm on
, and
. After increasing
q, we may assume that
for continuous seminorms
on
for
. By [
30] (p. 465, Corollary 2 to Theorem 45.2),
K is contained in the closed, absolutely convex hull of
for certain compact subsets
for
. For all
such that
, we have
using [
30] (Proposition 43.1). The assertion follows. □
Remark 9. If and are Hilbert spaces over with Hilbert space tensor product , and also and are Hilbert spaces over , then the bilinear mapis continuous, as . Replace the hypotheses in Proposition 9 with the requirements that G and M are modelled on metrisable locally convex spaces, and , are Hilbert spaces. We now use Remark 9 instead of Lemma 14, replace with the Hilbert space , Proposition 1 (a) with Proposition 1 (b) (so that operator-valued maps are only ) and use Proposition 2 (b) with in place of r. Repeating the proof of Proposition 9, we get:
Proposition 10. On , there is a unique equivariant -vector bundle structure of class over M whose typical fibre is the Hilbert space , such that is a local trivialisation for each .
Remark 10. If , G and M are modelled on metrisable spaces and both and are pre-Hilbert spaces with Hilbert space completions and , we can use the non-completed tensor product with the induced topology as the fibre and get an equivariant -vector bundle structure over M of class over M on , exploiting that the -bilinear map , is continuous.
9. Locally Convex Direct Sums of Vector Bundles
Let , , , such that , G be a -Lie group modelled on a locally convex space Y, and M be a -manifold modelled on a locally convex -vector space Z, together with a -action .
Let
and
be an equivariant
-vector bundle of class
over
M for
, with typical fibre a locally convex
-vector space
; let
be the
G-action and
be the projection onto the second component. We easily check that there is a unique
-vector bundle structure of class
on the “Whitney sum”
with the apparent map
, such that
,
is a local trivialisation of
E, for all families
of local trivialisations
, which trivialise the
s over a joint open subset
U of
M. Then,
for
,
yields an action of
G on
E. It is straightforward that
is
. Thus,
Proposition 11. If are equivariant -vector bundles of class over a G-manifold M of class , then also is an equivariant -vector bundle of class over M.
The following lemma allows infinite direct sums to be tackled.
Lemma 15. Let and be families of locally convex spaces over , with locally convex direct sums and , respectively. Let V be an open subset of a locally convex -vector space Z. Let , and assume that is a map which is linear in the second argument, for each . Moreover, assume that (a) or (b) holds:
- (a)
Z is finite-dimensional; or
- (b)
Z and each is a -space and I is countable.
If is of class for each , then also the following map is : Proof. If (b) holds, we may assume that
I is countably infinite, excluding a trivial case. Thus, assume that
. For each
, identify
with a vector subspace of
E, identifying
with
. For each
, we then have
where
is a
-space in the product topology. The inclusion map
is continuous and
-linear. Moreover,
is a
-map and so is
, for each
. Hence,
f is
on the open subset
of
, considered as the locally convex direct limit
, by [
31] (Proposition 4.5 (a)). This locally convex space equals
with the product topology (see [
32] (Theorem 3.4)).
If (a) holds, it suffices to prove the assertion for
. We proceed by induction.
The case . Let
; we show that
f is continuous at
. To this end, let
Q be an absolutely convex, open 0-neighbourhood in
F. There is a finite subset
such that
for all
. Let
. For each
, the intersection
is an absolutely convex, open 0-neighbourhood in
. For the absolutely convex hull, we get
. Since
is continuous for each
and
J is finite, we find a compact neighbourhood
K of
x in
V such that
for all
and
. Since
, where
K is compact and
is continuous, for each
, there is an absolutely convex, open 0-neighbourhood
in
such that
. Then,
is an open neighbourhood of
v in
E. Let
and
be given, say
, where
and
such that
and
. Then, for each
, since
, we obtain
For
, on the other hand, we have
As a consequence, , using the convexity of Q. We have shown that f is continuous at .
Induction step. Let
and assume the assertion is true for all numbers
. Given
,
, and
, we have
for some finite subset
. The map
,
is
, whence
exists in
and thus in
F; its
ith component is
in terms of partial differentials. Note that the mappings
,
and
,
are
and linear in
. By induction, the mappings
are
, using that
. Hence, also
is
, as
. Since
exists and is
, the continuous map
f is
. □
Remark 11. The conclusion of Lemma 15 does not hold for in the example , , , , using that the Taylor series of around 0 has radius of convergence for all .
Assuming now , consider a family of equivariant -vector bundles of class with typical fibre and G-action . We assume that (a) or (b) is satisfied:
- (a)
G and M are finite-dimensional; or
- (b)
I is countable and each as well as the modelling spaces of G and M are -spaces.
Moreover, we assume:
- (c)
For each , there exists an open neighbourhood U of x in M, such that, for each , the vector bundle admits a local trivialisation .
Thus, the
-vector bundle
is trivialisable for each
. Define
with pairwise disjoint direct sums and
,
. Then
is a
G-action such that
is
-linear for all
, where
. We readily deduce from Proposition 3 and Proposition 15 that there is a unique
-vector bundle structure of class
on
E such that
is a local trivialisation for
E, for each family
of local trivialisations as above. The latter makes
E an equivariant
-vector bundle of class
. In fact, the
-property of
can be checked using pairs of local trivialisations, as in the proofs of Propositions 5, 8, and 9. Then, apply Proposition 15, with
in place of
and
in place of
Z. Thus,
Proposition 12. In the preceding situation, is an equivariant -vector bundle of class over M.
Remark 12. If M is a -manifold, then every has an open neighbourhood U which is -diffeomorphic to a convex open subset W in the modelling space Z of M. If W can be chosen -paracompact, then every -vector bundle over U is trivialisable (see [12] (Corollary 15.10)). The latter condition is satisfied, for example, if Z is finite-dimensional, a Hilbert space, or a countable direct limit of finite-dimensional vector spaces (and hence a nuclear Silva space), cf. [3] (Theorem 16.10 and Corollary 16.16). If and Z has finite dimension, then each finite-dimensional holomorphic vector bundle over a, say, polycylinder in Z is -trivialisable (cf. [33]). Under suitable hypotheses, holomorphic Banach vector bundles over contractible bases are -trivialisable as well [34]. 10. Dual Bundles and Cotangent Bundles
In this section, we discuss conditions ensuring that a vector bundle has a canonical dual bundle. Let , , , and M be a -manifold modeled on a locally convex space Z.
Definition 7. Let be an -vector bundle of class , with typical fibre F. Consider the disjoint unionlet be the map taking to x, for each . Given such that , we say thatE has a canonical dual bundle
of class with respect to if can be made an -vector bundle of class over M, with typical fibre and bundle projection p, such thatis a local trivialisation of , for each local trivialisation of E. To pinpoint situations where the dual bundle exists, we recall a fact concerning the formation of dual linear maps (see [
8] (Proposition 16.30)):
Lemma 16. Let E and F be locally convex spaces, and . If the evaluation homomorphism , is continuous, thenis a continuous linear map. Remark 13. Let F be a locally convex -vector space over . It is known that is continuous if and only if F is quasi-barrelled, i.e., every bornivorous barrel in F is a 0-neighbourhood [35] (Proposition 2 in Section 11). In particular, is continuous if F is bornological or barrelled. It is also known that is continuous (and actually a topological embedding) if F is a -space. If , this follows from [36] (Theorem 2.3) and [37] (Lemma 14.3) (cf. also [37] (Propositions 2.3 and 2.4)). If and F is a -space, then is a topological embedding for the real topological vector space underlying F. Now, as a real topological vector space, using that a continuous -linear functional is determined by its real part. Transporting the complex vector space structure from to , the latter becomes a complex locally convex space. Thus, can be identified with , and it is easy to verify that corresponds to if we make the latter identification. Proposition 13. Let be an -vector bundle of class , with typical fibre F. Let . If , let ; if , assume and set . Consider the following conditions:
- ()
The modelling space Z of M is finite-dimensional, is continuous, and is barrelled.
- ()
is continuous and, moreover, is a -space, or and is a -space, or and is a -space.
- ()
F is normable.
If or is satisfied with , then E has a canonical dual bundle of class with respect to . If , , or is satisfied with , then E has a canonical dual bundle of class with respect to .
For , condition () of Proposition 13 is satisfied, for example, if F is a reflexive locally convex space (then is continuous and is barrelled, being reflexive.)
Proof. Let
be the disjoint union
, and
be as in Definition 7. Let
be a family such that the
form the set of all local trivialisations of
E. Let
be the associated cocycle (as explained before Proposition 3). Then,
is
and hence
is
, by Proposition 1. Given
, we define
as in (
15), using
instead of
. Then,
for all
and
shows that
where
. If (
) or (
) holds, then
is continuous by hypothesis. If
and (
) holds, then
is an isometric embedding (as is well known) and hence continuous. Thus,
,
is a continuous
-linear map (Lemma 16). Since
is
, we deduce that
is
. Thus Condition
of Corollary 2 is satisfied, with
in place of
r. Conditions (a)–(f) being apparent, the cited corollary provides an
-vector bundle structure of class
on
. □
Without specific hypotheses, a canonical dual bundle need not exist.
Example 2. Let A be a unital, associative, locally convex topological -algebra whose group of units is open in A, and such that the inversion map is continuous. Then, ι is smooth (and indeed -analytic); see, e.g., [13] (Propositions 10.1.12 and 10.1.13). We assume that the locally convex space underlying A is a non-normable Fréchet–Schwartz space and hence Montel, ensuring that . For example, we might take , where K is a connected, compact, smooth manifold of positive dimension (cf. [13] (Lemma 10.2.2 (c))). Let with and . We consider the trivial vector bundle(Thus, , the tangent bundle). Then, E is a -vector bundle of class over the base , with typical fibre A. Both and , are global trivialisations of E. Identifying with the set , we consider the associated bijections for (cf. (15)). Thus, , and for , . The map , is for , where is the projection onto the second factor. Then, also , is , by Proposition 1 (a). Now, A being Fréchet and thus barrelled, the evaluation homomorphism is continuous; since A is metrisable and hence a k-space, also is continuous (see Remark 13). Since is , we deduce with Lemma 16 that also , is . Definefor . Then, is discontinuous. To see this, we compose with the map , , which evaluates functionals at the identity element , and recall that is continuous. Then, for and . However, A being a non-normable locally convex space, the bilinear, separately continuous evaluation map , is discontinuous, and hence so is its restriction to the non-empty open subset , as is readily verified. Now, being discontinuous, also is discontinuous (and therefore not ). As a consequence, also is discontinuous. Summing up: There is no canonical vector bundle structure of class on because the two vector bundle structures on making (resp., a global trivialisation do not coincide.
Remark 14. In the preceding situation, set , , , for , and . If we let play the role of E in Proposition 3 and the role of in Proposition 3 (e), then all of Conditions (a)–(f) of Proposition 3 and Condition of Corollary 2 are satisfied for (with ). However, there is no -vector bundle structure on making each a trivialisation, as just observed, i.e., the conclusion of Corollary 2 becomes false.
Remark 15. Let , , with and M be a -manifold modelled on a locally convex space Z. Then, the tangent bundle is a -vector bundle of class over M, with typical fibre Z. Pick a locally convex vector topology on . Let be the set of all maps as in (15), with in place of , for ψ ranging through the set of all local trivialisations of (alternatively, only those of the form for charts of M, using the bundle projection ). Let us say that M has a canonical cotangent bundle
of class with respect to if admits a -vector bundle structure of class over M with typical fibre , which makes each a local trivialisation (with , ). Then, the evaluation mapmust be continuous and hence Z normable. For , this is explained in [17] (Remark 1.3.9) (written after Example 2 was found) if . This implies the case . As the diffeomorphism f employed as a change of charts is real analytic, the case follows and also the complex case, using a -analytic extension of f. When is the compact-open topology, existence of a canonical cotangent bundle for M even implies that Z is finite-dimensional. (If ε is continuous on , then there exists a compact subset and a 0-neighbourhood such that . Hence, . Since is a 0-neighbourhood in and compact (by Ascoli’s Theorem), is locally compact and hence finite-dimensional. As separates points on Z, also Z must be finite-dimensional.) Cotangent bundles are not needed to define 1-forms on an infinite-dimensional manifold
M. Following [
38], these can be considered as smooth maps on
which are linear on the fibres (and a similar remark applies to differential forms of higher order).
Differentiability properties of the -action on the dual bundle. Let , , , with , and G be a -Lie group modelled on a locally convex -vector space Y. Let M be a -manifold modelled on a locally convex -vector space Z and be a G-action of class .
Proposition 14. Let be an equivariant -vector bundle of class , with typical fibre F and G-action of class . Let . If , set ; if , assume and set . Consider the following conditions:
- (a)
is continuous, and, moreover, is a -space, or and is a -space, or and is a -space;
- (b)
M and G are finite-dimensional, is continuous, and is barrelled; or
- (c)
F is normable.
If and(a)
or(b)
holds, then E has a canonical dual bundle of class with respect to , and the map , defined using adjoint linear maps viafor , , turns into an equivariant -vector bundle of class over the G-manifold M. If and(a), (b)
, or (c)
is satisfied, then the same conclusion holds. Proof. In view of Proposition 13, the hypotheses imply that
E has a canonical dual bundle
of class
. It is apparent that
is an action, and
is taken
-linearly to
by
, for each
and
. It therefore only remains to show that
is
. To this end, let
and
; we show that
is
on
, for some open neighbourhood
U of
in
G and an open neighbourhood
V of
in
M. Indeed, there exists a local trivialisation
of
E over an open neighbourhood
W of
in
M. The action
being continuous, we find an open neighbourhood
U of
in
G and an open neighbourhood
V of
in
M over which
E is trivial, such that
. Let
be a local trivialisation of
E over
V. Then
for a
-map
, which is
-linear in the third argument. By Corollary 1, the map
,
is
. In view of the hypotheses, Lemmas 16 and 13 entail that also
,
is
-map. Now, again using the specific hypotheses, Proposition 2 shows that also the mapping
,
is
. However, for
,
, and
, we calculate
using the notation as in (
15). We conclude that
is
. □
Example 3. For elementary examples, recall that the group of all smooth diffeomorphisms of a σ-compact, finite-dimensional smooth manifold M can be made a smooth Lie group, modelled on the (LF)-space of compactly supported smooth vector fields on M (see [13,15]). The natural action is smooth [13]. In view of Example 1, Proposition 14 (b), Proposition 8 and Proposition 4, we readily deduce that also the natural action of on is smooth, as well as the natural actions on , for all , and the natural action on the subbundles and of given by symmetric and exterior powers, respectively. 11. Locally Convex Poisson Vector Spaces
We discuss a slight generalisation of the concept of a locally convex Poisson vector space introduced in [
8]. Fix
.
A bounded set-functor
associates with each locally convex
-vector space
E a set
of bounded subsets of
E, such that
for each continuous
-linear map
between locally convex
-vector spaces (cf. [
8] (Definition 16.15)). Given locally convex
-vector spaces
E and
F, we shall write
as a shorthand for
. We write
.
Throughout this section, we let
be a bounded set-functor such that, for each locally convex space
E, we have
Then,
for each
, and we get a continuous linear point evaluation
Definition 8. A locally convex Poisson vector space
with respect to is a locally convex -vector space E such that is a -space anda topological embedding, together with a bilinear map , , which makes a Lie algebra, is -hypocontinuous in its second argument, and satisfies writing . Remark 16.
- (a)
Definition 16.35 in [8] was more restrictive; E was assumed to be a -space there. - (b)
In [8] (16.31 (b)), the following additional condition was imposed: For each and , the set is bounded in , where is the evaluation map. As we assume (16), the latter condition is automatically satisfied, by [8] (Proposition 16.11 (a) and Proposition 16.14). - (c)
Let us say that a locally convex space E is-reflexive if is an isomorphism of topological vector spaces.
- (d)
Of course, we are mostly interested in the case where is continuous, but only hypocontinuity is required for the basic theory.
Definition 9. Let be a locally convex Poisson vector space with respect to , and be open. Given , we define a function viawhere , is the evaluation map and .Condition (17) in Definition 8 enables us to define a map via Using Lemma 11 instead of [
8] (Theorem 16.26), we see as in the proof of [
8] (Theorem 16.40 (a)) that the function
is
. The
-function
is called the
Poisson bracket of
f and
g. Using Lemma 11 instead of [
8] (Theorem 16.26), we see as in the proof of [
8] (Theorem 16.40 (b)) that
is a
-map; it is called the
Hamiltonian vector field associated with
f. As in [
8] (Remark 16.43), we see that the Poisson bracket just defined makes
a Poisson algebra.
We shall write “
b” and “
c” in place of
if
is the bounded set functor, taking a locally convex space
E to the set
of all bounded subsets and compact subsets of
E, respectively. Both of these satisfy the hypothesis (
16).
In the following, we describe new results for locally convex Poisson vector spaces over . We mention that the embedding property of is automatic in this case, as is a -space in Definition 9; thus, E is a -space and Remark 13 applies.
Example 4. Let be a family of finite-dimensional real Lie algebras . Endow with the locally convex direct sum topology, which coincides with the finest locally convex vector topology. Then, is c-reflexive, as with every vector space with its finest locally convex vector topology (see [39] (Theorem 7.30 (a))). As a consequence, also is c-reflexive (cf. [39] (Proposition 7.9 (iii))). Using [40] (Proposition 7.1), we see that the component-wise Lie bracket is continuous on the locally convex space , which is naturally isomorphic to the locally convex direct sum . We set and give the continuous Lie bracket making an isomorphism of topological Lie algebras. Thenand are -spaces, being Cartesian products of locally compact spaces (see [22]). Thus, is a locally convex Poisson vector space over , in the sense of Definition 8. If J has cardinality and for all (e.g., if we take an abelian 1-dimensional Lie algebra for each ), then is not a k-space. Hence, E is not a -space, and hence it is not a Poisson vector space in the more restrictive sense of [8]. 12. Continuity Properties of the Poisson Bracket
If E and F are locally convex -vector spaces and an open subset, we endow with the compact-open -topology. Our goal is the following result:
Theorem 1. Let be a locally convex Poisson vector space with respect to . Let be open. Then, the Poisson bracketis c-hypocontinuous in its second variable. If is continuous, then also the Poisson bracket is continuous. Various auxiliary results are needed to prove Theorem 1. With little risk of confusion with subsets of spaces of operators, given a 0-neighbourhood and a compact set , we shall write .
Lemma 17. Let be locally convex spaces and be open. Then, the linear mapis continuous. Proof. By Corollary 1,
for each
. As
D is linear and also
,
is linear for each
,
is linear, whence it will be continuous if it is continuous at 0. We pick a typical 0-neighbourhood in
, say
with a compact subset
and a 0-neighbourhood
. After shrinking
V, we may assume that
for some compact set
and 0-neighbourhood
.
We now recall that for
, we have
for all
,
and
(cf. Corollary 1). Since
is an open 0-neighbourhood in
and the map
,
is continuous, we see that the set
of all
such that
is a 0-neighbourhood in
. In view of (
21), we have
for each
. Hence,
from (
20) is continuous at 0, as required. □
Lemma 18. Let X be a Hausdorff topological space, F be a locally convex space, be compact and be compact. Let , be the evaluation map. Then, is compact.
Proof. The map
,
is continuous by [
20] (§3.2 (2)). Thus,
is compact in
. The map
,
is continuous by [
20] (Theorem 3.4.2). Hence,
is compact. □
Lemma 19. Let E, , , and G be locally convex -vector spaces and be a bilinear map which is c-hypocontinuous in its second argument. Let be an open subset and . Assume that is a -space, or and E is a -space, or and E is a -space. Then, the following holds:
- (a)
We have for all . The mapis bilinear. For each compact subset and 0-neighbourhood , there is a 0-neighbourhood such that . - (b)
For each , the map , is continuous and linear.
- (c)
If β is also c-hypocontinuous in its first argument, then is c-hypocontinuous in its second argument and c-hypocontinuous in its first argument.
- (d)
If β is continuous, then is continuous.
Proof. (a) By Lemma 11, . The bilinearity of is clear. It suffices to prove the remaining assertion for each . To see this, let be a compact subset and be a 0-neighbourhood. Since the topology on is initial with respect to the family of inclusion maps for , there exists and a 0-neighbourhood Q in such that . If the assertion holds for r, we find a 0-neighbourhood such that . Then, is a 0-neighbourhood in and .
The case . Let be compact and be a 0-neighbourhood. Then, for some compact subset and some 0-neighbourhood . By Lemma 18, the set is compact, where is the evaluation map. Since is c-hypocontinuous in its second argument, there exists a 0-neighbourhood with . Then, .
Induction step. Let
be a compact subset and
be a 0-neighbourhood. The topology on
is initial with respect to the linear maps
,
and
,
(by [
26] (Lemma A.1 (d))). Note that the ordinary
-topology is used there, by [
26] (Proposition 4.19 (d) and Lemma A2). After shrinking
W, we may therefore assume that
with absolutely convex 0-neighbourhoods
and
. Applying the case
to
, we find a 0-neighbourhood
such that
. The map
,
is continuous linear and
,
is smooth, whence
,
is continuous linear (cf. [
26] (Lemma 4.4) or [
13] (Proposition 1.7.11)). By (
5),
The subsets
and
are compact. Using the case
(with
in place of
U), which holds as the inductive hypothesis, we find 0-neighbourhoods
such that
and
. Then,
is an open 0-neighbourhood in
. Since
, we deduce from (
22) that
Thus, . Now, is a 0-neighbourhood in such that .
(b) Since is bilinear, the map is linear. Its continuity follows from (a), applied with the singleton .
(c) By (a) just established, the condition in Lemma 4 (a) is satisfied. By (b), the map is continuous in its first argument. Interchanging the roles of and , we see that is also continuous in its second argument and hence c-hypocontinuous in its second argument. Likewise, is c-hypoocontinuous in its first argument.
(d) If
is continuous and hence smooth, then
is smooth and hence continuous, as a very special case of [
26] (Proposition 4.16) or [
13] (Corollary 1.7.13). □
Proof of Theorem 1. By Lemma 17, the mapping
,
is continuous and linear. By Lemma 19 (c), the bilinear map
is
c-hypocontinuous in its second argument; if
is continuous, then also
, by Lemma 19 (d). The evaluation map
,
is
c-hypocontinuous in its first argument, by Proposition 7. As a consequence,
,
is continuous linear by Lemma 19 (b). Since
by definition, we see that
is a composition of continuous maps if
is continuous, and hence continuous. In the general case,
is a composition of a bilinear map which is
c-hypocontinuous in its second argument and continuous linear maps, whence
is
c-hypocontinuous in its second arguemnt. □
13. Continuity of the Map Taking to the Hamiltonian Vector Field
In this section, we show the continuity of the mapping which takes a smooth function to the corresponding Hamiltonian vector field, in the case .
Theorem 2. Let be a locally convex Poisson vector space with respect to . Let be an open subset. Then, the mapis continuous and linear. Proof. Let
be the evaluation homomorphism and
. Then,
V is a vector subspace of
and
. The composition map
,
is hypocontinuous with respect to equicontinuous subsets of
, by Proposition 9 in [
11] (Chapter III, §5, no. 5). If
is compact, then the polar
is a 0-neighbourhood in
, entailing that
is equicontinuous. Hence,
takes compact subsets of
E to equicontinuous subsets of
, and hence
is
c-hypocontinuous in its first argument. By Lemma 19 (c),
,
is continuous linear. Moreover, the map
,
is continuous linear by Lemma 17. Furthermore,
is continuous linear since
is
c-hypocontinuous in its second argument (see Lemma 4 (b)), whence
is continuous linear (see, e.g., [
26] (Lemma 4.13), or [
13] (Corollary 1.7.13)). Hence,
is continuous and linear. □