Function
Function
BOURBAKI’S
Integration
Notes by S. K. Berberian
Notes on text through Ch.V, §5, No. 4 (10-19-08)
Sterling K. Berberian, Prof. Emer.
The University of Texas at Austin
(e-mail: [email protected])
Inequalities of convexity
I.2, `. 15–16.
“. . .therefore there exists a λ0 > 0 such that the relation λ > λ0
implies λt ∈ K ;”
For example, set λ0 = ϕ(t1 , t2 , . . . , tn )−1 and cite the positive homo-
geneity of ϕ : if λ > λ0 , then ϕ(λt1 , . . . , λtn ) = λλ−1
0 > 1 , whence λt ∈ K .
I.2, `. 17.
“. . .the relations ti > 0 (1 6 i 6 n) imply αι1 t1 + · · · + αιn tn > 0 ”
Suppose first that ti > 0 for all i , and choose λ0 as above. Then, for
all λ > λ0 , λt ∈ K and therefore, for each ι ,
that is,
αι1 t1 + · · · + αιn tn > λ−1 βι ,
and letting λ → +∞ yields αι1 t1 + · · · + αιn tn > 0 .
If now t = (ti ) with ti > 0 for i = 1, . . . , n , then, for each ι , the
foregoing yields αι1 (t1 + ε) + · · · + αιn (tn + ε) > 0 for every ε > 0 , whence
αι1 t1 + · · · + αιn tn > 0 .
I.2, `. 18–19.
“. . .it is then clear that K is also the intersection of the half-spaces
ti > 0 (1 6 i 6 n) and the Uι such that βι > 0 ;”
If βι < 0 , the inequality (2) is trivially satisfied when the t i are > 0 .
I.2, `. 20.
“. . .since the origin does not belong to K ”
Writing 0 = (0, . . . , 0) , ϕ(0) = ϕ(2 · 0) = 2ϕ(0) implies ϕ(0) = 0 by
finiteness, whence 0 ∈/ K.
INT I.x2 inequalities of convexity Chap. 1
I.2, `. 20–21.
“. . .there exists at least one index ι such that β ι > 0 .”
If one had βι = 0 for all ι , then 0 = (0, . . . , 0) would satisfy all of
the inequalities defining K , that is, it would belong to all of the half-spaces
whose intersection is K , contrary to 0 ∈ / K.
I.2, `. −11 to −9.
“Now let C be the convex cone in Rn+1 defined by the relations ti > 0
(1 6 i 6 n + 1) , tn+1 6 ϕ(t1 , t2 , . . . , tn ) (the closure of the convex cone
generated in Rn+1 by the convex set K × {1} ) ”
It is clear from the continuity of ϕ that C is closed. Let D be the
convex cone generated by the convex set K × {1} , namely,
[
D= α(K × {1})
α>0
= {(αt1 , . . . , αtn , α) : α > 0 & (t1 , . . . , tn ) ∈ K } .
γ0 ϕ(t1 , . . . , tn ) = 1 ,
whence γ ϕ(t1 , . . . , tn ) > 1 for all γ > γ0 ; thus, for all γ > γ0 ,
we have ϕ(γt1 , . . . , γtn ) > 1 , so that (γt1 , . . . , γtn ) ∈ K and therefore
(γtn , . . . , γtn , 1) ∈ K × 1 , whence (t1 , . . . , tn , γ −1 ) ∈ γ −1 (K × {1}) ⊂ D ;
letting γ → +∞ one sees that (t1 , . . . , tn , 0) ∈ D .
Finally, suppose ϕ(t1 , . . . , tn ) = 0 . Let t0i > ti for i = 1, . . . , n . In
particular, t0i > 0 for i = 1, . . . , n , hence ϕ(t01 , . . . , t0n ) > 0 = tn+1 ,
thus (t01 , . . . , t0n , 0) ∈ D by the preceding sub-case; letting t 0i → ti yields
(t1 , . . . , tn , 0) ∈ D as desired.
Chap. 1 inequalities of convexity INT I.x3
I.2, `. −9 to −7.
“. . .it is immediate that C is also defined by the relations t i > 0
(1 6 i 6 n + 1) and
1 1
βι 6 αι1 · t1 + · · · + αιn · tn for all ι ∈ I ,
tn+1 tn+1
1
thus (t1 , . . . , tn ) ∈ K , so that (t1 , . . . , tn ) ∈ tn+1 K , whence
tn+1
for all ι ∈ I .”
Set ti = fi (x) for i = 1, . . . , n and set tn+1 = ϕ(t1 , . . . , tn ) . Obviously
on `. −10, −9, −8), thus (t1 , . . . , tn , tn+1 ) satisfies (3), which is precisely the
assertion (4).
I.2, `. −3, −2.
“ M (ϕ(f1 , f2 , . . . , fn )) is finite and
βι M ϕ(f1 , f2 , . . . , fn ) 6 αι1 M (f1 ) + αι2 M (f2 ) + · · · + αιn M (fn ) ”
βι f 6 αι1 f1 + · · · + αιn fn ,
whence M (f ) < +∞ .
I.3, Proposition 2.
Note that the condition α + β = 1 assures that the function ϕ(t 1 , t2 ) =
α β
t1 t2 is positively homogeneous.
I.4, `. 10, 11.
“. . .if g is a function belonging to F p (X, M ) and if |f | 6 |g| , then
f also belongs to F p (X, M ) ;”
Since |f |p 6 |g|p , one has M (|f |p ) 6 M (|g|p ) , therefore M (|f |p )1/p 6
M (|g|p )1/p < +∞ , thus f ∈ F (X, M ) .
I.4, `. 12.
“. . .the sum of two functions in F p (X, M ) also belongs to this set;”
1/p
For any finite numerical function f on X , write N p (f ) = M (|f |p ) .
By Minkowski’s inequality (Prop. 3), for any numerical functions f, g on X
one has Np (|f | + |g|) 6 Np (f ) + Np (g) , and it follows from |f + g| 6 |f | + |g|
that
Np (f + g) 6 Np (|f | + |g|) 6 Np (f ) + Np (g) .
Therefore if f and g belong to F p (X, M ) , that is, if Np (f ) and Np (g)
are finite, then Np (f + g) is also finite, thus f + g ∈ F p (X, M ) .
I.5, `. 7.
“. . .the corollary is proved.”
Explicitly, if 0 < r < p one sees that N r (f ) 6 Np (f ) by choosing q so
1 1 1
that = − and citing (9) with g = 1 .
q r p
Chap. 1 inequalities of convexity INT I.x5
I.5, `. 8–12.
Proposition 5. — For every finite numerical function f defined on X ,
the set I of values of 1/p (p > 0) such that N p (f ) is finite is either
empty or is an interval; if I is not reduced to a point, then the mapping
α 7→ log N1/α (f ) is either convex on I or is equal to −∞ on the interior
of I .
Rearrangement of the proof. Fix a finite numerical function f on X and
define J = {p > 0 : Np (f ) < +∞ }; thus, in the notation of the statement
of the proposition, I = {1/p : p ∈ J } .
claim 1: Either J = ∅ or J is an interval (possibly degenerate, i.e.,
reduced to a point).
In other words, the assertion is that J is a convex subset of ]0, +∞[ .
Suppose r, s ∈ J and 0 < α < 1 ; writing p = αr + (1 − α)s , we are
to show that p ∈ J , i.e., that M (|f |p ) < +∞ . Since r, s ∈ J we have
M (|f |r ) < +∞ and M (|f |s ) < +∞ ; by Hölder’s inequality (Prop. 2),
α 1−α
M |f |r )α (|f |s )1−α 6 M (|f |r ) M (|f |s ) ,
that is,
α 1−α
M (|f |p ) 6 M (|f |r ) M (|f |s ) ,
which can also be written
p rα s(1−α)
(∗) Np (f ) 6 Nr (f ) Ns (f ) ,
claim 3: ϕ is convex.
◦ ◦
Suppose r, s ∈ J ; thus 1/r, 1/s are typical elements of I . Given
0 < t < 1 , define p > 0 by the formula
1 1 1
(†) = t · + (1 − t) · ,
p r s
◦
so that 1 = tp/r + (1 − t)p/s ; thus 1/p ∈ I and the problem is to show that
1 1 1
ϕ 6 tϕ + (1 − t)ϕ ,
p r s
i.e., that
log Np (f ) 6 t log Nr (f ) + (1 − t) log Ns (f ) ,
equivalenty, that
t 1−t
Np (f ) 6 Nr (f ) Ns (f ) ,
i.e., that
1/p t/r (1−t)/s
M (|f |p 6 M (|f |r M |f |s ,
tp/r (1−t)p/s
i.e., that M (|f |p ) 6 M (|f |r M (|f |s .
Setting λ = tp/r , we have, by (†),
tp (1 − t)p
1−λ=1− = > 0,
r s
thus 0 < λ < 1 and our problem is to show that
λ 1−λ
(∗∗) M (|f |p ) 6 M (|f |r M (|f |s .
Note that
rλ + s(1 − λ) = tp + (1 − t)p = p ;
since M (|f |r ) < +∞ and M (|f |s ) < +∞ , it follows from Hölder’s inequal-
ity (Prop. 2) that
M (|f |p ) = M (|f |rλ+s(1−λ) ) = M (|f |r )λ · (|f |s )1−λ
λ 1−λ
6 M (|f |r ) M (|f |s ,
Riesz spaces
1 1
(y − x)+ = (|y − x| + y − x) = (|x − y| + y − x) ,
2 2
1
thus x + (y − x)+ = (2x + |x − y| + y − x) , whence (6).
2
II.3, `. 6–8.
“If A and B are two subsets of E each of which has a supremum, then
A + B also has a supremum and
II.8, `. −7.
“Corollary. — Let a be an element of a fully lattice-ordered space E ,
Ba the band generated by a , B0a the band of elements alien to a . For every
element x > 0 of E , the component of x in B a (for the decomposition
of E as the ordered direct sum of Ba and B0a ) is equal to sup inf(n|a|, x) .
n∈N
This follows from Proposition 6, applied to M = {a} , and Proposi-
tion 5.”
Adopt the notations of Prop. 6 and its proof. Thus (when M = {a} )
M1 = {t : 0 6 t 6 n|a| for some n ∈ N } ;
then M1 has the properties of the set A of Prop. 5, and M 2 is the set M
of Prop. 5. Let x ∈ E , x > 0 . By Prop. 5, one can write x = y + z with
y > 0 , z > 0 and
y = sup{v : v ∈ M1 and v 6 x }
= sup{v : 0 6 v 6 n|a| for some n ∈ N and v 6 x }
= sup{v : 0 6 v 6 inf(n|a|, x) for some n ∈ N }
= sup{v : v = inf(n|a|, x) for some n ∈ N }
= sup inf(n|a|, x) ,
n∈N
and where z ∈ M02 = M0 = M000 = B0a . Thus, as noted in the proof of Prop. 6,
x = y + z is the representation of x in the direct sum E = B a ⊕ B0a . In
particular, the component of x in Ba (namely y ) is given by the desired
formula.
§2. LINEAR FORMS ON A RIESZ SPACE
II.10, `. −5.
“Theorem 1. — 1◦ In order that a linear form L on a Riesz space E
be relatively bounded, it is necessary and sufficient that it be the difference
of two positive linear forms.
2◦ The ordered vector space Ω of relatively bounded linear forms on E
is a Riesz space that is fully lattice-ordered.”
The proof can perhaps be clarified by separating out a part of it as a
lemma:
Lemma 0. — Let L be a relatively bounded linear form on the Riesz
space E . Let P = {x ∈ E : x > 0 } and define M : P → [0, +∞[ by the
formula
M (x) = sup L(y) for all x ∈ P .
06y6x
γ = M (x + x0 ) = sup L(z) ;
06z6x+x0
We assert that in the ordered vector space Ω (with the positive linear forms
serving as the convex cone of positive elements), the elements L and 0 have
a supremum, namely M . For, on the one hand, M > 0 and M > L . On
the other hand, if N ∈ Ω satisfies N > 0 and N > L , then for every x > 0
one has
0 6 y 6 x ⇒ N (x) > N (y) > L(y) ,
whence
N (x) > sup L(y) = M (x) ,
06y6x
N = (L − M )+ + M ,
Since inf(L, M ) = L+M −sup(L, M ) (formula (8) of §1, No. 1), for all x > 0
one has
inf(L, M ) (x) = L(x) + M (x) − sup [L(y) + M (z)]
y>0, z>0, y+z=x
II.12, `. −4.
Let x > 0 and let s = sup L(y) . If |u| 6 x then L(u) 6 |L|(x)
|y|6x
by `. −6, hence s 6 |L|(x) .
On the other hand, if x = y + z with y > 0 and z > 0 , then (as noted
in connection with `. −7) the element u = y − z satisfies |u| 6 x , hence
L(y −z) = L(u) 6 s ; varying y and z (subject to y > 0 , z > 0 , y +z = x )
it follows from the formula displayed in `. −8 that |L|(x) 6 s .
II.14, `. 5.
“Then every continuous linear form x 0 ∈ E0 is relatively bounded”
Recall that the convex cone P is proper (TVS, II, §2, No. 5, Prop. 13).
One has
{y : |y| 6 x } = {y : − x 6 y 6 x }
= {y : x − y ∈ P and x + y ∈ P }
= {y : y − x ∈ −P and x + y ∈ P }
= (x − P) ∩ (−x + P) ;
A ∪ B = A ∪ (A ∩ X --- A) = (A ∪ A) ∩ (A ∪ X --- A) = A ∩ X = A .
◦ ◦
Incidentally, X --- A = { A , where A is the interior of A (GT, I, §1, No. 6,
formulas (2)). And when A is a closed set, A contains its boundary, since
A = A = A ∪ B.
Suppose now that A is a closed subset of X and f : A → Y is a con-
tinuous mapping of A into a topological space Y , such that the restriction
of f to its boundary B is constant, say f (B) = {y 0 } for some y0 ∈ Y .
Define g : X → Y to be the function equal to f on A and to y 0 on X --- A .
Since B is also the boundary of X --- A , g has the constant value y 0 on
the set X --- A = (X --- A) ∪ B . Thus, g is continuous on each of the closed
sets A and X --- A , whose union is X , consequently g is continuous on X
(GT, I, §3, No. 2, Prop. 4).
In particular, if K is a compact subset of the locally compact space X ,
E is a topological vector space over R or C , and f : K → E is a continuous
function that is zero on the boundary of K , then the extension g of f
to X defined by setting g(x) = 0 for all x ∈ X --- K is continuous, and
its support, being a closed subset of K , is compact. Conversely, if K is a
compact subset of X and g ∈ K (X, K; E) , that is, g is continuous on X
with support contained in K , i.e., {x : g(x) 6= 0 } ⊂ K , equivalently g is
equal to 0 on X --- K , then g is also equal to 0 on X --- K , hence on the
§1 measures on a locally compact space INT III.x2
boundary K ∩ X --- K of K , thus the restriction f = g K is a continuous
function on K equal to 0 on the boundary of K .
III.1, `. −5, −4.
“When C (K; E) is equipped with the topology of uniform convergence
in K , K (X, K; E) is a closed subspace of C (K; E) .”
More generally, suppose (gα ) is a directed family in K (X, K; E) that
converges pointwise to a function g ∈ C (K; E) . Then (assuming E is Haus-
dorff) g is equal to 0 on the boundary of K , hence belongs to K (X, K; E) .
For, if there existed a point x of the boundary of K such that g(x) 6= 0 ,
one could choose a neighborhood V of 0 in E such that g(x) ∈ / V ; but
there exists an α0 such that α > α0 ⇒ g(x) − gα (x) ∈ V , a contradiction
since gα (x) = 0 .
III.1, `. −2.
“. . .if the topology of E is defined by the semi-norms p n ”
Recall that a Fréchet space is a complete, metrizable locally convex
space, and that the topology of a metrizable locally convex space can be
defined by a countable set of semi-norms (TVS, II, §4, No. 1, comments
following the Corollary of Prop. 1).
III.2, `. 7, 8.
“If E is locally convex, one can therefore define on K (X; E) the direct
limit of the locally convex topologies of the K (X, K; E) ”
Recall (TVS, II, §4, No. 4, Prop. 5) that this topology (call it T ) is
the finest locally convex topology on K (X; E) that renders continuous the
canonical injections
uK : K (X, K; E) → K (X; E) ,
where, for each compact subset K of X , the space K (X, K; E) bears the
topology of uniform convergence in K ; and, for a linear mapping
u : K (X; E) → G ( G a locally convex space), the following conditions
are equivalent:
(a) u is continuous for the topology T ;
(b) for every compact subset K of X , the linear mapping
u ◦ uK : K (X, K; E) → G is continuous; that is,
(b0 ) for every compact subset K of X and for every directed family (f α )
in K (X, K; E) that converges uniformly to a function f ∈ K (X, K; E) , one
has u(fα ) → u(f ) in G .
The mappings uK are continuous when K (X, K; E) is equipped with
the topology of uniform convergence in K and K (X; E) is equipped with
the topology Tu of uniform convergence in X , consequently T is finer
INT III.x3 measures §1
than Tu (TVS, loc. cit., Example. Caution: In the cited Example, TVS
p. II.29, in lines 4 and 5 read “Denote by T K the topology induced on EK
by the topology Tu of uniform convergence on X ”; the sentence is correct
in EVT, p. II.31).
Scholium: The direct limit topology on K (X; E) is finer than the
topology of uniform convergence.
III.2, `. −5.
“. . .hence belongs to K (X, K; E) .”
Here f ∈ K (X, K0 ; E) and there exists a family (fα ) in K (X, K; E)
such that fα → f uniformly on K0 , hence uniformly on K0 --- K , hence
pointwise on K0 − K . Since the fα are 0 on X --- K , it follows that
f = 0 on K0 --- K ; but also f = 0 on X --- K0 , hence f = 0 on
X --- K = (X --- K0 ) ∪ (K0 --- K) , that is, f ∈ K (X, K; E).
III.2, `. −4, −3.
“(ii) The criterion for continuity in a direct limit (TVS, II, §4, No. 4,
Prop. 5) shows at once that the mapping f 7→ (pri ◦ f ) is continuous”
It suffices to show that for each i , f 7→ pr i ◦ f is a continuous mapping
K (X; E) → K (X; Ei ) . Fix a compact set K ⊂ X ; by the cited result in
TVS, it is enough to show that the composite mapping
III.3, `. 8–9.
“. . .it is immediate that the mapping f λ 7→ fλ00 of K (Xλ ; E) into
K (X; E) is continuous.”
For each compact Kλ ⊂ Xλ , the composite mapping
L
defined by f 7→ f 7→ (f Xλ ) . By the preceding discussion, K inter-
λ∈L
sects Xλ only for λ in some finite subset
L of L , thus the the above composite
has range contained in the subspace K (Xλ ; E) for some finite subset H
λ∈H
of L , and it defines a bicontinuous isomorphism of K (X, K; E) onto that
subspace; in particular, ϕK is continuous for every compact K ⊂ X , whence
the continuity of ϕ .
To see that ϕ−1 is continuous, it is enough, by the definition of topo-
logical direct sum, to note that for each λ ∈ L , the composite
M
K (Xλ ; E) → K (Xµ ; E) → K (X; E)
µ∈L
INT III.x5 measures §1
where W runs over any fundamental system of entourages for the uniform
structure on E (GT, X, §1, No. 1, Def. 1). Thus a fundamental system of
entourages for the uniform structure on C (K; E) is formed by the sets
where K runs over the set of all compact subsets of X . The deduced topol-
ogy is called the topology of compact convergence and will be denoted τ cc ;
it has as fundamental system of neighborhoods of 0 the sets
NK,α = {f ∈ C : f (K) ⊂ Nα }
thus hf ∈ VK,α (f ).
III.6, `. −7.
. . . the second assertion is an obvious consequence of the first ”
For definiteness, let us consider the case of real scalars (the same argu-
ment works in the complex case). The identification in question associates
Pn
with f = ϕk ⊗ yk (where ϕk ∈ K (X, K; R) and yk ∈ E ) the function
k=1
n
X
x 7→ ϕk (x)yk (x ∈ X) ,
k=1
where K runs over the set of all compact subsets of X ; citing (i) for E = R ,
we infer that
[
(ii) K (X; R) ⊗ E = K (X, K; R) ⊗ E .
K
INT III.x9 measures §1
Ψ : K (X; R2 ) → K (X; C)
namely
(Ψ ◦ Φ−1 )(f1 , f2 ) = Ψ Φ−1 (f1 , f2 )
= Ψ(g) (where g(x) = (f1 (x), f2 (x) ))
= θ◦g,
where (θ ◦ g)(x) = θ(g(x)) = θ (f1 (x), f2 (x) = f1 (x) + if2 (x) , thus
(Ψ ◦ Φ−1 )(f1 , f2 ) = f1 + if2 is the desired isomorphism.
III.14, `. 1–3.
“Moreover, we can suppose ζ so chosen that
By the cited proposition (see TVS II.21), the property of V just derived
assures that every positive linear form µ 0 on V (for the relative order) may
be extended to a positive linear form on K (X; R) , that is, to a positive
measure µ on X (No. 5, Th. 1); if, in addition, V is dense (for the direct
limit topology), then the extension is unique by the remarks at the beginning
of the present subsection.
III.16, `. 6–8.
“Definition 3. — A measure on a locally compact space X is said
to be bounded if it is continuous on K (X; C) for the topology of uniform
convergence.”
Note that a linear form on K (X; C) continuous for the topology of uni-
form convergence Tu (i.e., the norm topology) is automatically continuous
for the direct limit topology T (because T u ⊂ T ), hence is a measure.
III.16, `. 13–15.
“To say that µ is a bounded measure thus signifies that µ belongs to
the dual of the space K (X; C) normed by kf k ; we shall denote this dual
by M 1 (X; C) (or simply M 1 (X) when no confusion can result).”
The case X = N . The direct limit topology on K (X; C) remains a
slippery concept to grasp; it is instructive to trace through the concepts on
an atypical but simple example: X = N = {0, 1, 2, 3, . . .} with the discrete
topology.
(1) F (X) , the set of all functions f : N → C , may be identified with
the vector space (s) (in the notation of Banach’s book) of all sequences
(ck )k>0 of complex numbers with the termwise linear operations (or, when
viewed as a product space Cℵ0 , with the coordinatewise linear operations).
(2) C (X) = F (X) , as all functions from a discrete space to a topological
space ( C in this instance) are continuous.
(3) C b (X) : This is the linear subspace of F (X) consisting of all
bounded complex functions on X ; it may be identified with the space (m)
(in the notation of Banach’s book) of bounded sequences of complex num-
bers.
(4) K (X) : As the compact subsets K of X = N are its finite sub-
sets, K (X) is the set of all complex functions on X with finite support;
it may be identified with the linear subspace of (m) consisting of all com-
plex sequences that terminate in zeros. In particular, K (X, K) is finite-
dimensional, K (X) being the union of its linear subspaces K (X, K n ) ,
where Kn = {0, 1, 2, . . . , n} (n = 0, 1, 2, . . .) .
(5) The topology on K (X, K) , K compact (i.e., finite).
The topology of uniform convergence on K (X, K) is simply the topol-
ogy of pointwise convergence; since K (X, K) is finite-dimensional, this is
§1 measures on a locally compact space INT III.x14
uK : K (X, K) → K (X) , uK (f ) = f
X∞
1 |µ(ϕn ) − ν(ϕn )|
D(µ, ν) = n
·
n=0
2 1 + |µ(ϕn ) − ν(ϕn )|
∞
X
an µ(ϕn ) = |µ(f )| 6 kµ k · kf k = kµk · sup |an | .
n∈N
n=0
for all complex sequences (an ) that terminate in zeros. For every n ∈ N
choose a complex number an such that |an | = 1 and an cn = |cn | . By the
foregoing, applied to the sequence a 0 , a1 , a2 , . . . , an , 0, 0, 0, . . . ,
n X
X n
|ck | = ak ck 6 kµk · 1
k=0 k=0
P
∞
for every n ∈ N , thus the series ck is absolutely convergent with
n=0
P
∞
|cn | 6 kµk . On the other hand, for every finite sequence a 0 , a1 , . . . , an
n=0
of complex numbers, we have
X X X X
n n n ∞
ak ck 6 |ak | · |ck | 6 |ck | · max |ak | 6 |ck | · max |ak |;
06k6n 06k6n
k=0 k=0 k=0 k=0
P
∞
whence kµk 6 |cn | .
n=0
The set of all such measures, denoted M 1 (X) , is a linear subspace
of M (X) = K (X)∗ ; equipped with the norm just defined, it is the dual
space of the normed space K (X), k · k and is a Banach space (TVS, III,
§3, No. 8, Cor. 2 of Prop. 12).
The foregoing associates, to each bounded measure µ on X , the se-
P
∞
quence µ(ϕn ) satisfying |µ(ϕn )| < +∞ . The set of all complex se-
n=0
P
∞
quences c = (cn )n>0 satisfying |cn | < +∞ is a linear subspace of (s)
n=0
§1 measures on a locally compact space INT III.x18
P
∞
and is a Banach space for the norm kck = |cn | ; it is denoted (`(1) )
n=0
(Banach, p. 12). The correspondence µ 7→ µ(ϕn ) is an isometric linear
mapping of M 1 (X) onto (`(1) ) . The space (`(1) ) may be regarded as the
space of functions integrable with respect to the measure on X = N that
assigns measure 1 to every set {n} , namely, the linear form λ : K (X) → C
P
∞
defined by λ(f ) = f (n) (a finite sum) for every f ∈ K (X) .
n=0
Identifying K (X) with the space (c00 ) of sequences that terminate
in 0 ’s, its completion for the sup-norm k · k may be identified with the
space (c0 ) of sequences that converge to 0 ; the Banach space (c 0 ) has the
same dual as its dense linear subspace (c 00 ) , thus
0
(c0 )0 ∼
= (c00 )0 ∼
= K (X), k · k = M 1 (X) ∼
= (`(1) )
(cf. Hewitt–Stromberg (7.13) and (14.25), or Banach pp. 66, 67). Banach
spaces whose dual is isometrically isomorphic to the space of functions in-
tegrable for some measure are called L 1 -predual spaces (cf. the book of
H.E. Lacey, The isometric theory of classical Banach spaces, Grundlehren
math. Wiss. Bd. 208, Springer–Verlag, New York, 1974).
III.17, `. 9–13.
“Proposition 10. — For every measure µ on X ,
“For, taking into account the formula (12) that defines the absolute
value of a measure, the second member of (22) may be written
!
sup sup |µ(g)| = sup |µ(g)| .”
06f 61, f ∈K (X;R) |g|6f, g∈K (X;C) kgk61, g∈K (X;C)
Write α for the expression on the left (where, in the right side of (22),
|µ|(f ) has been replaced by its formula given by (12)), β for the expression
on the right (which is the definition of kµk ).
Proof that α 6 β : Suppose f ∈ K (X; R) , 0 6 f 6 1 . If g ∈ K (X; C)
and |g| 6 f , then kgk 6 kf k 6 1 , hence |µ(g)| 6 β ; varying g —and
then f —yields α 6 β .
Proof that β 6 α : Suppose g ∈ K (X; C) , kgk 6 1 ; we are to show that
|µ(g)| 6 α . Say g ∈ K (X, K; C) . Choose f ∈ K + (X) with 0 6 f 6 1 and
f = 1 on K . Evidently |g| 6 f , whence, citing (12), |µ(g)| 6 |µ|(f ) 6 α .
INT III.x19 measures §1
III.18, `. 1–4.
“Corollary 3. — For every real measure µ on a locally compact
space X,
“It suffices to make use of the formula (22) and the expression (9) for
|µ|(f ) when µ is a real measure and f ∈ K + (X) .”
Citing (22), then (9),
!
kµk = sup |µ|(f ) = sup sup µ(g) ;
f ∈K (X;R), 06f 61 f ∈K (X;R), 06f 61 g∈K (X;R), |g|6f
the g ’s involved on the right side of this equality are precisely the functions
g ∈ K (X; R) such that kgk 6 1 , thus the right side may be rewritten as
sup µ(g) , which equals the right side of (24) (where the absolute
g∈K (X;R), kgk61
value signs are superfluous since −f satisfies the same conditions as f ).
III.18, `. 7, 8.
“The canonical injection M 1 (X, R) → M 1 (X; C) is an isometry by
virtue of (24).”
Let µ be a bounded real measure on X . The left side of (24) is kµk
as calculated in M 1 (X; C) ; the right side is the norm of µ regarded as an
element of M 1 (X, R) .
III.18, `. −3 to −1.
“Let X be a locally compact space. On the space M (X; C) , one can
consider the topology of pointwise convergence in K (X; C) , which we shall
call the vague topology on M (X; C) .”
Since M (X; C) is by definition the dual space of the locally convex
space K (X; C) (equipped with the direct limit topology), the vague topol-
ogy is the “weak topology” σ M (X; C), K (X; C) (TVS, II, §6, No. 2,
Def. 2).
The space M (X; C) , equipped with the vague topology, may also be
described as the locally convex space L S K (X; C); C , where K (X; C)
bears the direct limit topology and S is the set of all 1-element subsets {f }
of K (X; C) , or, equivalently, the set of all finite subsets of K (X; C) (TVS,
III, §3, No. 1, p. TVS III.13, `. −3 to −1; recall that every finite subset of a
topological vector space is bounded).
§1 measures on a locally compact space INT III.x20
III.19, `. 3–6.
“To say that a filter F on M (X; C) converges vaguely to a measure µ 0
signifies that
µ0 (f ) = limµ, F µ(f )
for every function f ∈ K (X; R).”
This is a property of initial topologies (GT, I, §7, No. 6, Prop. 10).
Incidentally, a filter F on M (X; C) is vaguely Cauchy (i.e., Cauchy for
the uniformity associated with the vague topology) if and only if for every
f ∈ K (X; C) (or every f ∈ K (X; R) ), the filter base
F(f ) = {V(f ) : V ∈ F }
V = {µ ∈ M (X; C) : |µ(f )| 6 1 }
Si = {µ(fi ) : µ ∈ H }
by (13), the definition of H , and the relation (*), thus the criterion d) is
satisfied.
Proof of (ii). Let (µj ) be a directed family in H with µj → µ in
M (X; C) vaguely; we are to show that µ ∈ H . Thus, assuming f ∈ K + (X) ,
it is to be shown that |µ|(f ) 6 ν(f ) . Given any g ∈ K (X; C) such that
|g| 6 f , it will suffice by (12) to show that |µ(g)| 6 ν(f ) . For each index j ,
we have
|µj (g)| 6 |µj |(|g|) 6 ν(|g|) 6 ν(f )
§1 measures on a locally compact space INT III.x22
The first factor of the intersection is vaguely closed as noted in Prop. 14,
and the second factor, being the inverse image of {1} under the vaguely
continuous mapping µ 7→ µ(1) , is also vaguely closed, thus H is vaguely
INT III.x23 measures §1
III.23, `. 11–13.
“. . .every continuous function with values in a topological vector space E ,
defined on Y and with compact support, may be extended by continuity to
all of X , by giving it the value 0 on { Y ”
Suppose g ∈ K (Y; E) and let f be the extension by 0 of g to X ,
that is, (
g(x) for x ∈ Y
f (x) =
0 for x ∈ X --- Y .
§2 support of a measure INT III.x28
sup |ν(g)| ,
|g|6f, g∈K (Y;C)
which is the formula for |ν|(f ) ; thus it follows from (*) that |µ| Y = |ν| on
K+ (Y) , hence on K (Y; C) .
III.23, `. 20–22.
“If µ is real then the restrictions of µ + and µ− to Y are, respectively,
(µ|Y)+ and (µ|Y)− , by virtue of formula (8) of § 1, No. 5.”
From the formulas (INT II.2)
1 1
µ+ = (|µ| + µ) , µ− = (|µ| − µ) ,
2 2
restriction to K (Y; C) yields (with ν = µ Y as in the preceding Note)
1 1
µ+ Y = (|ν| + ν) = ν + , µ− Y = (|ν| − ν) = ν − .
2 2
More generally, if µ 1 , µ 2 are real measures on X , and if ν 1 = µ 1
Y , ν 2 =
µ2 Y , then the formulas (INT II.2)
1
sup(µ1 , µ2 ) = (µ1 + µ2 + |µ1 − µ2 |)
2
1
inf(µ1 , µ2 ) = (µ1 + µ2 − |µ1 − µ2 |)
2
yield, on restriction to K (Y; C) , the formulas
1
sup(µ1 , µ2 )Y = (ν1 + ν2 + |ν1 − ν2 |) = sup(ν1 , ν2 )
2
1
inf(µ1 , µ2 )Y = (ν1 + ν2 − |ν1 − ν2 |) = inf(ν1 , ν2 ) .
2
INT III.x31 measures §2
III.23, `. 23–26.
“One sees immediately
that if Y and Z are two open sets in X such
that Y ⊃ Z , and if µY and µZ are the restrictions of µ to Y and Z ,
then µZ is also the restriction of µY to the open subspace Z of the locally
compact space Y .”
If h ∈ K (Z; C) and h* ∈ K (Y, Z; C) is the extension by 0 of h
0
to Y , and if h* ∈ K (X, Y; C) is the extension by 0 of h* to X , then in
0
fact h* ∈ K (X, Z; C) is the extension by 0 of h to X , and so
0
(µZ)(h) = µ(h* ) = (µY)(h*) = (µY)Z (h) ,
for all f ∈ K (Y, K; C). Thus f 7→ ν(f ), obviously a linear form on K (Y; C),
is a measure on Y (§1, No. 3, (4)).
§2 support of a measure INT III.x32
(§1, No. 3, (4)); note that we are concerned here only with the norm topology
on K (Yαx , Kx ; C) —the direct limit topology on K (Y αx ; C) that induces
it is not at issue here. If f ∈ K (X, Kx ; C) , then Supp f ⊂ Kx ⊂ Yαx ,
so that f can be regarded as an element of K (Y αx , Kx ; C) , and hence
satisfies (*). Thus (*) may be written
(µα )Yα ∩Yβ = (µYα )Yα ∩Yβ = µYα ∩Yβ = (µYβ )Yα ∩Yβ = (µβ )Yα ∩Yβ ;
it follows from Prop. 1 that µ is the unique measure on X such that µ Yα =
µα for all α . Similarly, ν is the unique measure on X such that ν Yα = να
for all α . But
µα = µ Yα = ν Yα = ν α
§2 support of a measure INT III.x34
µU = 0 ⇔ |µU | = 0 ⇔ |µ|U = 0 ,
therefore, by posititivy,
thus
U ⊂ { Supp(µ) ⇔ U ⊂ { Supp(µ+ ) ∩ { Supp(µ− )
⇔ U ⊂ { Supp(µ+ ) ∪ Supp(µ−) ;
U -- Supp(µU) ⊂ U − U ∩ Supp(µ) ,
whence equality.
III.26, `. −7 to −5.
“For, it is the intersection of the vaguely closed hyperplanes with equa-
tion µ(f ) = 0 , where f runs over the set of functions in K (X; C) whose
support does not intersect F .”
§2 support of a measure INT III.x36
Supp(µ) ⊂ F ⇔ { F ⊂ { Supp(µ)
⇔ µ{ F = 0
⇔ µ { F (g) = 0 for all g ∈ K ( { F; C)
⇔ µ(f ) = 0 for all f ∈ K (X, { F; C)
\
⇔ µ∈ {ν ∈ M (X; C) : ν(f ) = 0 } ;
f ∈K (X, { F;C)
for each f , the set {ν ∈ M (X; C) : ν(f ) = 0 } is the kernel of the vaguely
continuous linear form ν 7→ ν(f ) .
III.26, `. −4 to −1.
“Suppose X is not compact: given a filter Φ on the space M (X; C)
of measures on X , we shall say that the support of a measure µ recedes
indefinitely along Φ if, for every compact subset K of X , there exists a set
M ∈ Φ such that Supp(µ) ∩ K = ∅ for every measure µ ∈ M .”
Some thoughts on the definition: Let X 0 = X ∪ {ω} be the one-point
compactification of X (GT, I, §9, No. 8), with X viewed as an open subspace
of X0 . The open sets in X0 are the open sets U in X together with the
sets {ω} ∪ (X --- K) with K compact in X (the latter are the only open sets
in X0 that contain ω ). Thus the open neighborhoods of ω in X 0 are the
sets of the form U0 = {ω} ∪ (X --- K) for some compact subset K of X .
Thus, the concept defined above says that given any open neighborhood
V0 of ω in X0 (equivalently, given any compact subset K of X ), there
exists a set M ∈ Φ such that Supp(µ) ⊂ V 0 (resp. Supp(µ) ∩ K = ∅ ) for all
µ ∈ M . It is the same to say that for every neighborhood
S V 0 of ω in X0 ,
such an M exists. Thus, writing Supp(M) = Supp(µ) , the concept
µ∈M
says that for every neighborhood V 0 of ω in X0 , there exists a set M ∈ Φ
such that Supp(M) ⊂ V0 ; this might be expressed suggestively as
lim Supp(M) = ω
M∈Φ
III.27, `. 6, 7.
“. . .which proves the proposition.”
To say that µ converges vaguely to a measure µ 0 with respect to a fil-
ter Φ on M (X; C) means that each vague neighborhood of µ 0 in M (X; C)
contains some set M ∈ Φ , that is, Φ ⊃ V0 , where V0 is the filter
of neighborhoods of µ0 in M (X; C) for the vague topology. Concisely,
Φ → µ0 vaguely.
An equivalent condition (see the note for III.19, `. 3–6) is that for each
f ∈ K (X; C) , one has
µ0 (f ) = lim µ,Φ µ(f ) ;
the meaning of this is that for each f ∈ K (X; C) one considers the linear
form Lf : M (X; C) → C defined by Lf (µ) = µ(f ) , and that
lim µ,Φ Lf (µ) = Lf (µ0 ) ;
that is, the filter on C with base Lf (Φ) converges to Lf (µ0 ) in C ; here,
Lf (Φ) = {Lf (M) : M ∈ Φ } , where
Lf (M) = {Lf (µ) : µ ∈ L } = {µ(f ) : µ ∈ M },
concisely and suggestively denoted M(f ) .
Equivalently, for every f ∈ K (X; C) and every ε > 0 there exists
an M ∈ Φ such that |µ(f ) − µ0 (f )| < ε for all µ ∈ M .
In particular, Φ converges to the measure 0 if and only if, for every
f ∈ K (X; C) and every ε > 0, there exists a set M ∈ Φ such that |µ(f )| < ε
for all µ ∈ M ; concisely, Φ → 0 vaguely.
The assertion of Prop. 7 is that if Φ is a filter on M (X; C) such
that “the support of µ recedes indefinitely along Φ ” then Φ → 0 vaguely.
To prove this, suppose f ∈ K (X; C) and ε > 0 ( ε will play no role in
the argument). Let K be the support of f . By hypothesis, there exists
an M ∈ Φ such that Supp(µ) ∩ K = ∅ for every µ ∈ M . In view of the
preceding paragraph, it will obviously suffice to show that µ(f ) = 0 for
every µ ∈ M . Fix µ ∈ M . We know that Supp(µ) ∩ K = ∅ , that is,
K ⊂ { Supp(µ) . Writing U = { Supp(µ) , we have µU = 0 , that is, µ = 0
on K (X, U; C) ; since Supp f = K ⊂ U , we have f ∈ K (X, U; C) , whence
µ(f ) = 0 , which completes the proof.
III.27, `. 9, 10.
“By definition, if the support of a function f ∈ K (X; C) does not
intersect the support of a measure µ , then µ(f ) = 0 ”
To say that Supp(f )∩Supp(µ) = ∅ is to say that Supp(f ) ⊂ { Supp(µ) .
Writing U = { Supp(µ) , we have µU = 0 , that is, µ = 0 on K (X, U; C)
and in particular µ(f ) = 0 .
§2 support of a measure INT III.x38
it is stronger than (*), in that it yields the same conclusion with a weaker
hypothesis. It is more precise in the sense that it enlarges the set of functions
for which the conclusion is true.
III.27, `. 15, 16.
“. . . V is an open set containing S by hypothesis”
For, S = Supp(µ) ⊂ {x : f (x) = 0 } ⊂ V .
III.27, `. 19.
“Since the support of f h does not intersect S . . .”
Supp(f h) ⊂ Supp(h) ⊂ { S .
III.27, `. 20, 21.
“. . .therefore |f − f h| 6 2ε on X ”
On V , |f − f h| 6 |f | + |f h| 6 |f | + |f | < 2ε , whereas on { V ,
|f − f h| = |f − f · 1| = 0 .
III.27, `. −3.
“. . . |f (x)| 6 ah(x) on Supp(µ) ”
On Supp(f ) ∩ Supp(µ) , one has |f (x)| 6 a = ah(x) , whereas on
{ Supp(f ) ∩ Supp(µ) one has |f (x)| = 0 6 ah(x) .
III.27, `. −3, −2.
“. . .therefore
|µ|(|f |) 6 a|µ|(h) 6 akµk ”
The measure |µ| is positive, and on Supp(|µ|) = Supp(µ) one has
ah − |f | > 0 , therefore |µ|(ah − |f |) > 0 by Cor. 2; thus
III.28, `. 7.
“. . .then g 6 bf /a , whence µ(g) 6 bµ(f )/a = 0 .”
f (y)
If y ∈ V then f (y) > a , whence g(y) 6 b = b · 1 6 b · ; whereas
a
if y ∈ { V then g(y) = 0 . Therefore 0 6 µ(g) 6 bµ(f )/a = 0 , and so
µ(g) = 0 .
The argument shows that for every g ∈ K + (X) with Supp g ⊂ V ,
one has µ(g) = 0 . If g ∈ K (X; C) with Supp g ⊂ V , then |g| ∈ K + (X)
and Supp |g| = Supp g ⊂ V , therefore µ(|g|) = 0 by the foregoing, and so
|µ(g)| 6 µ(|g|) = 0 . Thus µ(g) = 0 for all g ∈ K (X; C) with Supp g ⊂ V .
◦
A fortiori, µ(g) = 0 for all g ∈ K (X; C) with Supp g ⊂ V , in other
◦ ◦
words µ = 0 on K (X, V; C) , that is, V ⊂ { Supp(µ) , and in particular
x ∈ { Supp(µ) . Thus x ∈
/ Supp(µ) , as we wished to show.
III.28, `. 11-13.
“. . .there exists an open neighborhood V of x 0 such that at every point
of V ∩ S , g is zero”
Write N(g) = {x : g(x) 6= 0 } , Z(g) = { N(g) = {x : g(x) = 0 } ;
in particular, T = N(g) ∩ S , and the assertion of the proposition is that
T = Supp(g · µ) . By assumption, x0 is not in the closure of N(g) ∩ S , hence
there exists an open neighborhood V of x 0 such that V ∩ N(g) ∩ S = ∅ ,
whence V ∩ S ⊂ { N(g) = Z(g) —in other words, g = 0 on V ∩ S .
III.28, `. 13.
“. . .then f g is zero on S ”
We know that g = 0 on V ∩ S , and that f = 0 on { V hence on
{ V ∩ S , therefore f g = 0 on (V ∩ S) ∪ ( { V ∩ S) = S .
III.28, `. 14, 15.
“. . .in other words, the restriction of g · µ to V is zero.”
The argument shows that µ(f g) = 0 for all f ∈ K (X, V; C) , in other
words g ·µ = 0 on K (X, V; C) , whence V ⊂ { Supp(g ·µ) and in particular
x0 ∈ { Supp(g · µ) . We have shown that { T ⊂ { Supp(g · µ) , that is,
T ⊃ Supp(g · µ) . That’s half the battle.
III.28, `. 16–18.
“Conversely, assuming that the restriction of g · µ to an open neighbor-
hood W of a point x0 ∈ X is zero, let us show that there does not exist a
point of W ∩ S at which g is 6= 0 .”
One is assuming that x0 ∈ W , where W is open and W ⊂ { Supp(g·µ) ;
to put it another way, we are assuming that x 0 ∈
/ Supp(g · µ) (a closed set)
and that W is any open neighborhood of x 0 disjoint from Supp(g · µ) . If
§2 support of a measure INT III.x40
g · µ = 0 ⇔ Supp(g · µ) = ∅
and since
g = 0 on Supp(µ) ⇔ A = ∅ ,
the assertion of the corollary reduces to the triviality A = ∅ ⇔ A = ∅ .
III.29, `. 2–4.
“Proposition 12. — Let ai (1 6 i 6 n) be distinct points in a locally
compact space X . Every measure on X whose support is contained in the
set of the ai is a linear combination of the measures ε ai (1 6 i 6 n) .”
Pn
Conversely, if µ = ci εai then (No. 2, Prop. 4)
i=1
n
[ n
[
Supp(µ) ⊂ Supp(ci εai ) ⊂ Supp(εai ) = {a1 , . . . , an } ,
i=1 i=1
where the fact that Supp(εa ) = {a} is shown by the following argument:
if U = X --- {a} then for all f ∈ K (X, U; C) one has ε a (f ) = f (a) = 0 ,
thus U ⊂ { Supp(εa ) , that is, Supp(εa ) ⊂ { U = {a} ; since Supp(εa ) 6= ∅
(because εa 6= 0 ) necessarily Supp(εa ) = {a} (alternatively—cf. the note
for III.25, `. −7 to −5 —for every neighborhood V of a , there exists a
function f ∈ K (X, V; C) with εa (f ) = f (a) 6= 0 , whence a ∈ Supp(εa ) ).
III.29, `. 16–18.
“It suffices to prove that µ is orthogonal to the subspace V ◦ of K (X; C)
orthogonal to V (TVS, II, §6, No. 3, Cor. 2 of Th. 1), that is, that the rela-
tions hf, εa i = 0 , where a runs over the support of µ , imply hf, µi = 0 ”
In view of Prop. 12, V is the linear subspace of M (X; C) generated
by the measures εa with a ∈ Supp(µ) ; regarding K (X; C) and M (X; C)
as being in a (separating) duality via the bilinear form
(f, ν) 7→ hf, νi = ν(f ) f ∈ K (X; C), ν ∈ M (X; C) ,
(TVS, II, §6, No. 3, Prop. 4, (iii) for the case of real scalars, loc. cit., §8,
No. 4 for the case of complex scalars). By the theorem on bipolars (TVS, II,
§6, No. 3, Th. 1 and §8, No. 4), the polar V ◦◦ of V◦ in M (X; C) is equal
to the closure of V in M (X; C) for the topology σ M (X; C), K (X; C) ,
that is,
V = V◦◦ ,
where V is the vague closure of V in M (X; C) . Thus the assertion that
µ is vaguely adherent to V is equivalent to the assertion that µ ∈ V ◦◦ .
III.29, `. 19.
“. . .but this is just Prop. 8 of No. 3.”
The cited proposition states that
f = 0 on Supp(µ) ⇒ µ(f ) = 0 ,
that is,
hf, εa i = 0 for all a ∈ Supp(µ) ⇒ hf, µi = 0 ,
equivalently
f ∈ V◦ ⇒ hf, µi = 0 ,
in other words µ ∈ V◦◦ .
Preparation for Corollary 1 of Theorem 1:
(I). Let A = {a1 , . . . , an } be a finite set of pairwise distinct points of a
locally compact space X . In order that a measure µ on X have support A ,
Pn
it is necessary and sufficient that µ = ck εak with the ck nonzero scalars,
k=1
in which case
n
X
(i) kµk = |ck | .
k=1
P
n
Proof. If µ has support A , we know from Prop. 12 that µ = ck εk
k=1
for suitable scalars ck . If a coefficient, say c1 , were equal to 0 , then one
would have Supp(µ) ⊂ {a2 , . . . , an } by the note for III.29, `. 2–4, contrary
to hypothesis.
P
n
Conversely, if µ = ck εk for suitable nonzero scalars ck , then
k=1
Supp(µ) ⊂ {a1 , . . . , an } = A
INT III.x43 measures §2
by the note for III.29, `. 2–4; we assert that Supp(µ) = A . Assume to the
contrary that some ak is absent from Supp(µ) . Then there exists an open
neighborhood U of ak such that U∩Supp(µ) = ∅ , whence U ⊂ { Supp(µ) ,
and so µ(f ) = 0 for all f ∈ K (X, U; C) . Choose f ∈ K (X, U; C) such
that f (ak ) 6= 0 . For all indices j 6= k we have aj ∈ { U and so f (aj ) = 0 ,
thus
n
X n
X X
0 = µ(f ) = cj εaj (f ) = cj f (aj ) = ck f (ak ) + cj · 0 = ck f (ak ),
j=1 j=1 j6=k
P
n
hence kµk 6 |ck | .
k=1
On the other hand, let V1 , . . . , Vn be pairwise disjoint neighborhoods
of a1 , . . . , an , respectively, and choose functions f k ∈ K (X; R) such that
0 6 fk 6 1 , fk (ak ) = 1 and Supp fk ⊂ Vk ; in particular,
[
fk = 0 on Vj for k = 1, . . . , n .
j6=k
n
X
g= θk fk ;
k=1
n
X n
X n
X n
X
µ(g) = ck εak (g) = ck g(ak ) = ck θk = |ck | ,
k=1 k=1 k=1 k=1
P
n
thus kµk > |µ(g)| = |ck | , which completes the proof of (i).
k=1
§2 support of a measure INT III.x44
where bal conv {rεa1 , . . . , rεan } denotes the balanced convex envelope of the
set {rεa1 , . . . , rεan } in M (X; C) .
Proof. Recall that the balanced convex envelope of a subset S of a
vector space (real or complex) is the set of all finite linear
P P combinations
cj uj , where uj ∈ S and the cj are scalars such that |cj | 6 1 (TVS,
II, §8, No. 2).
P
n
By (I), µ = ck εak , where ck is 0 when the point ak does not
k=1
belong to Supp(µ) , and
n
X |ck |
kµk 6 r ⇔ 6 1.
r
k=1
The following general lemma prepares the way for item (III) below.
Lemma. — Let F, G be complex vector spaces in duality with respect
to a bilinear form (x, y) 7→ hx, yi (x ∈ F , y ∈ G ), and let M be a “circled”
( x ∈ M and |c| = 1 imply cx ∈ M ) subset of F . Then:
(1) The polar M◦ of M in G is given by the formula
R hcx, yi > −1
for all scalars c with |c| = 1 ; if c is chosen so that chx, yi = −|hx, yi| , then
and let B ◦ be the polar set of B in K (X; C) for the canonical duality
hf, νi = ν(f ) . Then
1
(iii) B ◦ = {f ∈ K (X; C) : |hf, εa i| 6 for all a ∈ S} .
r
Proof. Let P = {rεa : a ∈ S } , and let Pbc be the balancedPconvex
envelope of P , thatPis, the set of all finite linear combinations ck νk ,
where νk ∈ P and |ck | 6 1 .
We assert that B = Pbc . For, if ν ∈ B , so that kνk 6 r and
Supp(ν) = {a1 , . . . , an } with the ak pairwise distinct elements of S , then
by (II),
ν ∈ bal conv{rεa1 , . . . , rεan } ⊂ Pbc .
P
n
Conversely, suppose ν ∈ Pbc , so that ν = ck (rεak ) for suitable points
k=1
P
n
ak ∈ S and scalars ck such that |ck | 6 1 ; we can suppose (after
k=1
gathering together coefficients of a same ε ak ) that the points ak are pair-
wise distinct (at work here is the triangle inequality for the partial sums
of the ck ), so that the εak are linearly independent; then, omitting terms
with zero coefficients, we can further suppose that the c k are nonzero. Then,
Supp(ν) = {a1 , . . . , an } ⊂ S by (I), and kνk 6 r by (II), thus ν ∈ B .
Now let {M = cν : |c| = 1 and ν ∈ P } (the “circled envelope”
of P ). Clearly M and P have the same balanced convex envelope, thus
M◦ = (M◦◦ )◦ = (B ◦◦ )◦ = B ◦ ,
INT III.x47 measures §2
thus
1
f ∈ K (X; C) & |hf, εa i| 6 for all a ∈ S ⇒ |hf, µi| 6 1 ,
r
that is,
1
f ∈ K (X; C) & kf Sk 6 ⇒ |µ(f )| 6 1 .
r
III.29, `. 24–26.
“To prove the first assertion. . .”
This is covered by the preceding remarks, with Supp(µ) in the role
of S , r = kµk , and A in the role of B .
III.29, `. −8 to −5.
“. . .we note that
lim inf kνk > kµk
ν→µ, ν∈A
since the function ν 7→ kνk is lower semi-continuous for the vague topology
(§1, No. 9, Cor. 4 of Prop. 15)”
One must adapt the cited Prop. 15 to the situation of a limit with respect
to the relative topology induced on A by the vague topology. Let V be
the set of neighborhoods of µ for the vague topology on M (X; C) . By
definition,
lim inf kνk = sup inf kνk .
ν→µ, ν∈A V∈V ν∈V∩A
Given 0 < h < kµk , by the cited lower semi-continuity there exists a V ∈ V
such that ν ∈ V ⇒ kνk > h ; in particular, ν ∈ V ∩ A ⇒ kνk > h , hence
whence equality throughout (GT, IV, §5, No. 6, formula (11)); thus
by GT, IV, §5, No. 6, Cor. 1 of Th. 3 (the filter G in question being the
trace V ∩ A of V on A ).
Note: It follows that the mapping A ∪ {µ} → [0, +∞[ defined by
ν 7→ kνk is continuous at µ for the topology on A ∪ {µ} induced by the
vague topology.
III.30, `. 2–4.
“. . .for every ε such that 0 < ε < 1 , there exists, by virtue of Cor. 1,
a measure ν0 whose support is finite and contained in Supp(µ) and for
which ν0 − µ ∈ V and kµk > kν0 k > (1 − ε)kµk .”
With A as in Corollary 1, we know that
ν − µ = (ν − ν0 ) + (ν0 − µ) ∈ V + V ,
and given an ε > 0 such that εkµk · kfk k < 1 for all k , choose ν 0 ∈ B so
kµk
that ν0 −µ ∈ V and kµk > kν0 k > (1−ε)kµk . Setting ν = ·ν0 , we have
kν0 k
ν > 0 (because ν0 > 0 ), kνk = kµk (hence ν ∈ D ) and kν − ν0 k 6 kµk ,
whence ν − ν0 ∈ V and ν − µ ∈ V + V . This shows that µ ∈ D .
III.30, `. 13, 14.
“Again, it suffices to establish that µ belongs to the polar set of B ◦ ,
the polar set of B in the space K (X; R) ”
The point is that to exploit the argument of the preceding note, we
must show that µ ∈ B . It is clear that B is convex and that 0 ∈ B , thus
B = B◦◦ by the theorem on bipolars for the duality between M (X; R) and
K (X; R) (TVS, II, §6, No. 3, Th. 1).
III.30, `. 15–17.
“. . .but this means that for f ∈ K (X; R) the relations hf, ε a i >
−1/kµk for all a ∈ Supp(µ) imply hf, µi > −1 , which is a consequence
of No. 3, Cor. 2 of Prop. 8.”
We know, from item (I) of the Note for III.29, `. 19, that B consists
of 0 together with the set of all measures ν on X of the form
n
X
(∗) ν= c k ε ak ,
k=1
INT III.x51 measures §2
1
f ∈ B◦ ⇔ hf, εa i > − for all a ∈ Supp(µ) .
kµk
X 1 X
hf, νi = ck hf, εak i > − ck > −1 ,
kµk
whence f ∈ B◦ .
Thus, in order that µ belong to B = B◦◦ , it is necessary and sufficient
that
f ∈ B◦ ⇒ hf, µi > −1 ,
that is,
1
hf, εa i > − for all a ∈ Supp(µ) ⇒ hf, µi > −1 ,
kµk
that is,
1
(∗∗) f >− on Supp(µ) ⇒ µ(f ) > −1 .
kµk
thus x0 ∈ Supp(µ) .
III.30, `. −10 to −6.
“. . .we are to prove that there exists a function g ∈ K (X; C) such
that kg · µk 6 1 and such that
for 1 6 k 6 n .”
One has fk (x0 ) − µ(gfk ) = εx0 (fk ) − (g · µ)(fk ) = (εx0 − g · µ)(fk ) , thus
the indicated inequalities state that the vague neighborhood of ε x0 defined
by
{ν ∈ M (X; C) : |(εx0 − ν)(fk )| 6 δ (1 6 k 6 n) }
contains g · µ .
§2 support of a measure INT III.x54
kgk = kg 0 k = kf f0 k 6 kf k · kf0 k 6 kf k ,
therefore
III.31, `. 6–8.
“. . .the hypothesis kνk = 1 then implies that there exists a function
h ∈ K (X; C) , with support contained in U , such that khk 6 1 and such
that |αk (1 − ν(h))| 6 δ/2 for 1 6 k 6 n .”
It is here that we need Prop. C of the preceding note: Since kνk = 1 ,
there exists a function h ∈ K (X, U; C) such that khk 6 1 and |ν(h)| is as
close to kνk = 1 as we like, in particular, close enough so that
For,
III.31, `. 13.
“This proves the proposition, since kg ·µk = k(g 0 h)·µk 6 kg0 ·µk = 1 .”
INT III.x57 measures §2
For,
K (X; C) · µ = {g · µ : g ∈ K (X; C) }
V = {g · µ : g ∈ K (X; C) }
defined by (g, ρ) 7→ g·ρ continuous (jointly in the variables) for the indicated
topologies? I suspect that the answer is usually “no” and that the case of
X = N with the discrete topology may provide a counter-example (cf. TVS,
III, §5, Exer. 3, TVS, IV, §1, Exer. 7, and the note for III.16, `. 13–15).
A counterexample with X compact is proposed in §1, Exer. 12 d) (possible
candidate: the 1-point compactification of the discrete space N ?). The
following is a more elementary observation:
Proposition. The above mapping (g, ρ) 7→ g · ρ is always separately
continuous in the variables.
Proof. If ρj → ρ vaguely then g·ρj → g·ρ vaguely; for, if f ∈ K (X; C)
then
(g · ρj )(f ) = ρj (gf ) → ρ(gf ) = (g · ρ)(f ) .
On the other hand, if gj → g for the direct limit topology, the claim is
that gj · ρ → g · ρ vaguely for every measure ρ . This entails showing that
for every f ∈ K (X; C) and every measure ρ ,
ρ(gj f ) → ρ(gf ) ;
=0 (because g = 0 on { W ⊃ N ).
To summarize: Assuming b ∈
/ N , the existence of such a neighbor-
/ Supp(µ) (No. 2). Thus { N ⊂ { Supp(µ) ,
hood W of b assures that b ∈
i.e., Supp(µ) ⊂ N .
III.31, `. −1 to III.32, `. 1.
“. . .the restriction of µ to Va is therefore a point measure with support
{a} (No. 2, Prop. 5)”
Supp(µVa ) = Va ∩ Supp(µ) = Va ∩ N = {a} ; quote Prop. 12.
§3 integrals of continuous vectorial functions INT III.x60
III.32, `. 2.
“. . .is of the form h(a)εa . . .”
The scalar coefficient yielded by the cited Prop. 12 is here written h(a)
in anticipation of the function h being defined.
III.32, `. 2–4.
“Setting h(x) = 0 at the points of { N , and denoting by ν the measure
defined by the masses h(x) , the principle of localization shows that ν = µ .”
Let us first establish that the function h : X → C so defined is eligible
to define a discrete measure. Indeed, for every compact subset K of X ,
N ∩ K is finite; for, since N is closed, N ∩ K is a compact subset of the
discrete space N , hence is covered by finitely many of the open subsets {x}
(x ∈ N) of N . It follows from the discussion in §1, No. 3, Example I that
a discrete measure ν on X can be defined, for every f ∈ K (X; C) , by the
formula
X
ν(f ) = h(x)f (x)
x∈X
X
= h(a)f (a) (h = 0 on { N )
a∈N
X
= h(a)εa (f ) ( f = 0 on { Supp(f ) )
a∈N∩Supp(f )
From (i) and (iii), we see that (Vx )x∈X is an open covering of X such that
ν Vx = µ Vx for all x ∈ X ,
INT III.x61 measures §3
III.32, `. 14–16.
“. . . E0 * , equipped with the weak topology σ(E 0 *, E0 ) , may be canoni-
cally identified with the completion of E equipped with the weakened topology
σ(E, E0 ) .”
In TVS, II, §6, No. 7, Prop. 9, set F = E and G = E 0 , placed in duality
via hz, z0 i = z0 (z) .
{In particular, regarding E ⊂ E b (the completion of E for the weakened
topology σ(E, E ) ), if z ∈ E then the corresponding element of E 0 * is
0
and similarly v(cy*) = c v(y*) for every scalar c. If v(y*) = 0 then y*(e ι ) = 0
for all ι ∈ I , whence y* = 0 by linearity; thus v is injective. Given
any (cι )ι∈I ∈ CI , the linear form y* ∈ G* for which y*(e ι ) = cι for
all ι ∈ I satisfies v(y*) = (cι )ι∈I ; thus v is surjective, hence is a vector
space isomorphism.
§3 integrals of continuous vectorial functions INT III.x62
(GT, I, §7, No. 6, Prop. 10). Translated in terms of nets, this says that for
a directed family (yα∗ ) in G* ,
thus
n
[
{x : f (x) 6= 0 } = {x : (z0k ◦ f )(x) 6= 0 } ,
k=1
whence
n
[
{x : f (x) 6= 0 } = {x : (z0k ◦ f )(x) 6= 0 } ,
k=1
that is,
n
[
Supp(f ) = Supp(z0k ◦ f ) ;
k=1
Conditions under which this occurs are discussed in No. 3, Cor. 2 of Prop. 7
(assuming E to be a quasi-complete Hausdorff locally Rconvex space), but
only for functions f in K (X; E) . Criteria assuring that f dµ belongs to E
for functions f ∈ K f (X; E) are given in Ch. VI, §1, No. 2, Cor. of Prop. 7
(with extra conditions on E and f ) and Prop. 8 (with extra conditions on µ
and f ).
When E is a Banach space, so Rthat the quasi-completeness condition
cited above is trivially satisfied and f dµ ∈ E for every f ∈ K (X; E) , the
notation µ(f ) is authorized by Ch. IV, §4, No. 1, Def. 1, as well as by Ch. V,
§1, No. 3, Def. 3; one would suppose that it would also be tolerated in the
case that E is a quasi-complete Hausdorff locally convex space.
The issue does not come up in the case of a vector-valued measure m
(Ch. VI, §2, No. 1, Def. 1), where m is given at the outset R to be a continuous
linear mapping of K (X) into a Banach space E and f dm is defined to
be m(f ) for f ∈ K (X) (loc. cit., No. 2, Def. 2); similarly for the case
of integration of a vector-valued function f with respect to a vector-valued
measure m (loc. cit., No. 7, Prop. 11 and the remark following its proof).
A final instance of the use of µ(f ) occurs in the extension of the the-
ory to Hausdorff topological spaces X , with E a Banach space (remarks
following Prop. 16 in Ch. IX, §1, No. 10).
Summary.
R I see no clear pattern in the foregoing, but sense
R a tendency:
One
R writes f dµ = µ(f ) either when this is the definition of f dµ or when
f dµ turns out to be an element of the ambient locally convex space RE ; in
the two cases cited above where the notation µ(f ) is not proposed, f dµ
is an element of an algebraic dual E0 ∗ .
At any rate, the issue seems unimportant; but the above tour of defini-
tions may serve as a useful preview of forthcoming concepts.
III.33, `. −8 to −6.
P
n
“ . . . we then have f (x) = fi (x)ei for all x ∈ X , and
i=1
Z n
X
f dµ = µ(fi )ei . ”
i=1
INT III.x65 measures §3
P
n
Thus f = fi ei . To prove the second equality, it suffices to show that for
i=1
each i , the function gi = fi ei has integral µ(fi )ei . Indeed,
(z0 ◦ gi )(x) = z0 gi (x) = z0 fi (x)ei
= fi (x)z0 (ei ) = hei , z0 ifi (x) ,
Z
= hei , z i fi dµ = µ(fi )hei , z0 i
0
= hµ(fi )ei , z0 i ,
R
therefore gi dµ = µ(fi )ei .
III.34, `. 8–10.
“ . . . we have Z
f dεy = f (y)
R
because hf , z0 i dεy = hf (y), z0 i by definition.”
For all z0 ∈ E0 ,
Z Z Z
f dεy , z = hf , z i dεy = (z0 ◦ f ) dεy
0 0
= (z0 ◦ f )(y) = z0 f (y) = hf (y), z0 i .
f (X; E)
f ∈K ⇔ f (X; E1 ) ,
f1 ∈ K
and in this case, for all such pairings z 0 ↔ z01 , f ↔ f1 , one has
µ(z0 ◦ f ) = µ(z01 ◦ f1 ) ,
that is, Z Z
f dµ, z0 = f1 dµ, z01 ;
R
thus,
R when E0 and E01 are identified via the pairing z0 ↔ z01 , f dµ and
f1 dµ define the same element of the algebraic dual (E 0 )∗ = (E01 )∗ .
III.34, `. −14 to −11.
“. . .the mapping z0 7→ Rz0 which, to every continuous (complex) linear
form z0 on E , makes correspond the continuous (real) linear form z 7→
Rhz, z0 i on E0 , is an R-isomorphism of the dual E0 onto the dual E00
of E0 (TVS, II, §8, No. 1).”
To establish the notation for the next note, we review the construction
of the isomorphism (TVS II.61). Most of what is going on is pure vector
space theory; to keep the notation simple, it is best to leave the topology to
the end.
To set the stage: by E00 the author means (E0 )0 , the dual of the topo-
logical vector space E0 , namely, the real vector space E0 underlying E ,
equipped with the given topology on E . The assertion is that (E 0 )0 is
isomorphic, as a real vector space, to (E 0 )0 , that is, the real vector space
INT III.x67 measures §3
underlying the complex vector space E 0 ; identifying (E0 )0 and (E0 )0 via
this isomorphism then renders the notation E 00 unambiguous (alternatively,
it allows E00 to denote either of the two spaces according to the context).
Suppose first that E is any complex vector space (no topology). If f
is a (complex)
linear form on E (that is, f ∈ E ∗ ) then the formula g(z) =
R f (z) defines an R-linear form on E (concisely denoted Rf —or R ◦ f ,
where R : C → R is the R-linear mapping R(a + ib) = a ), that is, a linear
form on E0 , and so g ∈ (E0 )∗ (the algebraic dual of E0 ). Since
g(iz) = R f (iz) = R if (z) = −I f (z) ,
one has
(∗) f (z) = R f (z) + iI f (z) = g(z) − ig(iz) .
of the algebraic duals (A, II, §2, No. 5), defined by t v(h) = h ◦ v for all
linear forms h : (E0 )0 → R .
On the other hand, by the second paragraph above (with F replaced by
the complex vector space E0 ) there exists an isomorphism w of real vector
spaces,
∗
w : (E0 ∗ )0 → (E0 )0
INT III.x69 measures §3
(E0 ∗ )0 w - (E0 ) ∗ t
v - (E )0 ∗
0 0
ϕ : (E0 ∗ )0 → (E0 )0 ∗ .
The foregoing proof is efficient, but one yearns for a more direct con-
struction that, given an element of E 0 ∗ , exhibits the corresponding element
of (E0 )0 ∗ . Such a construction will now be given; the following (commuta-
tive) diagram will be helpful in keeping track of the argument.
f -C
v(z00 ) = z0 ∈ E0
6
v = u−1 R
?
z00 ∈ (E0 )0 -R
g
Given f ∈ E0 ∗ , the formula
g(z00 ) = Rf v(z00 ) z00 ∈ (E0 )0
defines a linear form on the real vector space (E 0 )0 , that is, g ∈ (E0 )0 ∗ .
We define ψ : E0 ∗ → (E0 )0 ∗ by ψ(f ) = g . It is routine to verify that ψ is
R-linear.
ψ is injective: If g = 0 then Rf v(z00 ) = 0 for all z00 ∈ (E0 )0 , that
is, (Rf )(z0 ) = 0 for all z0 ∈ E0 ( v is surjective); thus the real part of the
complex linear form f is 0 , whence f = 0 by a now-familiar argument.
ψ is surjective: Given g ∈ (E0 )0 ∗ , that is, a linear form g : (E0 )0 → R ,
we seek a linear form f : E0 → C such that ψ(f ) = g . For all z0 ∈ E0
define
f (z0 ) = g(Rz0 ) − i g R(iz0 )
(where R(iz0 ) denotes the R-linear form z → R iz0 (z) = −I z0 (z)
on E ) It is routine to show that f is R-linear and that f (iz 0 ) = i f (z0 ) ,
hence that f is C-linear, that is, f ∈ E 0 ∗ . Now, R f (z0 ) = g(Rz0 ) for all
z0 ∈ E0 ; in other words, for all z00 ∈ (E0 )0 ,
R f (v(z00 )) = g R(v(z00 ) = g u(v(z00 )) = g(z00 ) ,
§3 integrals of continuous vectorial functions INT III.x70
that is, Rf (v(z00 )) = g(z00 ) , which says that ψ(f ) = g .
Inasmuch as f ◦ v = t v(f ) figures in both constructions, one suspects
that ϕ = ψ . Indeed, for f ∈ E0 ∗ and z00 ∈ (E0 )0 , one has
ϕ(f ) (z00 ) = (t v ◦ w)(f ) (z00 )
= t v w(f ) (z00 ) = t v(Rf ) (z00 )
= (Rf ) ◦ v (z00 ) = Rf v(z00 ) = ψ(f ) (z00 ) ,
ψ(z0 *) = z00 * ,
Z Z
0 0
w 7→ (w ◦ g)(x) dν(x) = hg(x), w0 i dν(x) (w0 ∈ F0 )
DZ E Z
0
0
(1 ) g dν, w = hg, w0 i dν for all w0 ∈ F0 .
DZ E Z
f0 dµ0 , z00 = hf0 , z00 i dµ0 for all z00 ∈ (E0 )0 .
In a sense, the preceding formula is all that the assertion in the text says:
when µ ∈ M (X; C) is a real measure, the formula (1) remains true R (by
definition!) with subscripts 0 installed. The proof that ψ(z 0 *) = f0 dµ0
will show the relevance of the formula to the identifications of the preceding
note.
Before proceeding, we note that
Z Z
(*) R f dµ = (Rf ) dµ for all f ∈ K (X; C)
R R
(because (Rf ) dµ and (I f ) dµ are real).
R
Let z00 * = ψ(z0 *) , where z0 * = f dµ and ψ : E0 * → (E0 )0 * is
the R-linear
R bijection of the preceding note. Our problem is to show that
z00 * = f0 dµ0 . By the definition of ψ , z00 * is the linear form on (E0 )0
such that
z00 *(z00 ) = Rz0 * v(z00 ) for all z00 ∈ (E0 )0 ,
briefly,
(**) hz00 *, z00 i = R v(z00 ) ◦ f , µ .
But v(z00 ) ◦ f (x) = v(z00 ) f (x) for all x ∈ X , and, by the definition
of v ,
R v(z00 ) f (x) = z00 f (x) ,
thus R v(z00 ) ◦ f = z00 ◦ f ; substituting this in (**), we have, for all
z00 ∈ (E0 )0 ,
R R
thus f0 dµ0 = z00 * = ψ(z0 *) = ψ f dµ .
III.35, `. 6.
“. . . tt u extends the mapping u .”
b ∈ E0 * defined by
To each y ∈ E corresponds the element y
z ∈ F0 * by
Similarly, each z ∈ F defines b
This observation is not needed for the Proposition 2 that follows it.
III.36, `. 5, 6.
“If E is complex, we equip E with its underlying real vector space
structure, which, as we have seen, does not modify the formula (1).”
Working through the identifications entailed in deriving the case for
complex E from the case for the real space E 0 underlying E promises to
be daunting. Instead, a more direct argument, patterned after the proof of
the real case, is given in the Note for III.36, `. 9–11.
III.36, `. 7.
“(i) We know . . .”
We know it from TVS, II, §5, No. 3, Cor. 5 of Prop. 4, p. TVS II.39.
Strictly speaking, by f (S) is here meant the canonical image
T = { fd
(x) : x ∈ S }
of f (S) in E0 * .
III.36, `. 9–11.
“ . . . it therefore suffices to prove that, for z 0 ∈ E0 , the relation
hf (x), z0 i> 0 for all x ∈ S implies
Z
0
f dµ, z > 0 ”
T = { fd
(x) : x ∈ S } ⊂ E0 * .
§3 integrals of continuous vectorial functions INT III.x74
Since E0 * is equipped with the weak topology σ(E 0 *, E0 ) for the duality
defined by the bilinear form (f, z0 ) 7→ hf, z0 i = f (z0 ) ( f ∈ E0 *, z0 ∈ E0 ), its
dual for this topology is E0 , i.e., consists of the linear forms
f 7→ f (z0 ) = hf, z0 i (f ∈ E0 *) ,
in other words, Z
T ⊂ H z0 ⇒ f dµ ∈ Hz0 ,
i.e., that
DZ E
0
z ◦ f > 0 on S ⇒ f dµ, z0 > 0 .
The complex case. The structure of the proof is the same as that of the
real case. Assuming E is a locally convex space over C , equip E 0 * with
the topology σ(E0 *, E0 ) relative to the duality defined by the bilinear form
We again write
T = { fd
(x) : x ∈ S } ⊂ E0 * ,
where fd (x)(z0 ) = z0 f (x) for z0 ∈ E0 , and define C to be the closure
in E0 * of the convex cone generated by T . The definition of C takes place
in the context of the real vector space (E 0 *)0 equipped with the topology
σ(E0 *, E0 ) .
In the duality between E0 and E0 * , the continuous linear forms on E0 *
are the forms
f 7→ f (z0 ) (f ∈ E0 *) ,
in other words R(z0 ◦ f ) > 0 on S . Writing A for the set of all such z 0 ,
we have \
C= H z0
z0 ∈A
§3 integrals of continuous vectorial functions INT III.x76
R
and, as argued in the real case, to prove that f dµ ∈ C we need only prove
the implication
Z
0
R(z ◦ f ) > 0 on S ⇒ f dµ ∈ Hz0 ,
that is,
DZ E
0
R(z ◦ f ) > 0 on S ⇒ R f dµ, z0 > 0 ;
since
DZ E Z Z Z
0 0 0
R f dµ, z =R hf , z i dµ = R (z ◦ f ) dµ = R(z0 ◦ f ) dµ ,
T = { fd
(x) : x ∈ S } ⊂ E0 *
INT III.x77 measures §3
Hz0 ,a = {f ∈ E0 * : f (z0 ) 6 a } ,
that is,
(a) z0 ◦ f 6 a on S .
z0 ◦ f = (z0 ◦ f )g 6 ag on S ,
R
that is, ag − z0 ◦ f > 0 on S , therefore (ag − z0 ◦ f ) dµ > 0 by §2, No. 3,
Cor. 2 of Prop. 8, thus
Z
(∗∗) (z0 ◦ f ) dµ 6 aµ(g) .
§3 integrals of continuous vectorial functions INT III.x78
If we show that aµ(g) 6 akµk , the implication (a) ⇒ (b) will be proved. At
any rate, 0 6 µ(g) = |µ(g)| 6 kµk kgk = kµk .
case 1: If a > 0 then aµ(g) 6 akµk by the preceding inequality, there-
fore (b) follows from (**).
case 2: If a < 0 , then z0 ◦ f 6 a < 0 on S , therefore S ⊂ Supp(z0 ◦ f )
(in particular, µ has compact support), and since g = 1 on Supp(z 0 ◦ f ) ,
we infer that g = 1 on S . If we show that µ(g) = kµk then (**) will again
yield (b). Since g > 0 , for every h ∈ K (X; R) we have
−khkg 6 hg 6 khkg ,
therefore
−khkg 6 h 6 khkg on S ,
whence (since µ > 0 )
thus |µ(h)| 6 khkµ(g) for all h ∈ K (X; R) , consequently kµk 6 µ(g) . But
where z0 varies over E0 , thus the closed half-spaces of (E0 *)0 are the sets
Hz0 ,a = {f ∈ E0 * : R f (z0 ) 6 a } ,
0
R says0 that R(z ◦ f ) 6 a on S , and the condition
The condition on the left
on the right says that R(z ◦ f ) dµ 6 akµk , thus the problem is to show
that
Z
0
(†) R(z ◦ f ) 6 a on S ⇒ R(z0 ◦ f ) dµ 6 akµk .
whence
R(z0 ◦ f ) = R(z0 ◦ f )g 6 ag on S ,
therefore (since µ is positive)
Z
(††) R(z0 ◦ f ) dµ 6 aµ(g) .
case 1: If a > 0 then aµ(g) 6 akµk as in the real case, and (†) then
follows from (††) .
case 2: If a < 0 then S ⊂ Supp R(z0 ◦ f ) as in the real case, therefore
g = 1 on S . Then, for every h ∈ K (X; C) , we have |h| 6 khkg on S ,
hence µ(|h|) 6 khkµ(g) ; therefore
which shows that kµk 6 µ(g) . But µ(g) 6 kµk kgk = kµk , thus µ(g) = kµk
and (†) again follows from (††).
III.36, `. −9.
“ . . . then ν is bounded”
Let U be an open set in X such that K ⊂ U and U is compact. By
Prop. 5 of §2, No. 2,
thus ν has compact support, hence is bounded (§2, No. 3, Prop. 11).
§3 integrals of continuous vectorial functions INT III.x80
III.36, `. −9. R R
“ . . . and f dµ = f dν ∈ kνk · D by Prop. 4, (ii).”
Since Supp(f ) = K ⊂ U , we haveR f ∈ K (X, U; E) , hence f U ∈
R (U;
K E) (see RIII.23,
`. 14–16).
Thus by f dν (abuse of notation)
is meant
(f U) dν = (f U) d(µU) which, by the definition of µU , is equal to
R
f dµ since, for all z0 ∈ E0 ,
DZ E D Z E
(f U) dν, z0 = ν z0 ◦ (f U) = ν (z0 ◦ f )U = µ(z0 ◦ f ) = f dµ, z0 .
R
By the cited Prop. 4, (ii), (f U) dν ∈ kνk · D1, where D1 is the closed
convex envelope of (the canonical image of) f U Supp(ν) in E0 * . Now,
f U Supp(ν) = f U U∩Supp(µ) = f U∩Supp(µ) = f (U∩S) ⊂ f (S) ,
thus D1 ⊂ D , and so
Z Z
f dµ = (f U) dν ∈ kνk · D1 ⊂ a · D
with a = kνk .
III.36, `. −8, −7.
“The second result follows from this, since any complex measure may
be written as µ1 − µ2 + iµ3 − iµ4 , where the µj are positive.”
Attention must be paid to the supports of the positive measures µ j .
As noted in §2, No. 2, Prop. 2, if ν is a real measure (so that ν = ν + − ν − ),
then
Supp(ν) = Supp(|ν|) = Supp(ν +) ∪ Supp(ν −) ;
and if ν is a complex measure then |Rν| 6 |ν| and |I ν| 6 |ν| (§1, No. 6,
(17)) show that the supports of Rν and I ν —hence of their positive and
negative parts—are contained in Supp(|ν|) = Supp(ν) . Thus ν has a rep-
resentation of the indicated sort with Supp(µ j ) ⊂ Supp(ν) for 1 6 j 6 4 .
Then, by the previous result, for each j there exists a number a j > 0 such
that Z
f dµj ∈ aj · Dj ,
where Dj is the closed convex envelope of f Supp(µj ) ; but Supp(µj ) ⊂
Supp(ν) implies that Dj ⊂ D , thus
Z
f dµj ∈ aj · D for 1 6 j 6 4 ,
INT III.x81 measures §3
(Prop. 1).
III.36, `. −1.
“ . . . E0 * is complete”
For the uniform structure on E0 * derived from the weak topology
σ(E *, E0 ) (TVS, II, §6, No. 7, Prop. 9).
0
III.37, `. R1, 2.
“ . . . f dµ ∈ C for every measure µ belonging to the convex set H of
positive measures on X of total mass equal to 1 .”
(As in previous notes, I regard C to be the closed convex envelope of
the canonical image {fd (x) : x ∈ X} of f (X) in E0 * .) The point is that for
µ ∈ H one has R kµk = µ(1) = 1 (§1, No. 8, Cor. 2 of Prop. 10). By
Prop. 4, (ii), f dµ belongs to the closed convex envelope D of (the canon-
ical image of) f (S) in E0 * , where S = Supp(µ) , and since f (S) ⊂ f (X)
implies D ⊂ C , one has
Z
f dµ ∈ kµk · D = D ⊂ C
as claimed.
III.37, `. 5–7. R
“ . . . the image of H0 under the mapping µ 7→ f dµ is the convex
envelope C0 of f (X) in E0 * .”
P
n
If µ is a measure with finite support, say µ = ci εxi , where the xi
i=1
are distinct points of X (§2, No. 4, Prop. 12), then µ is bounded, kµk =
Pn
|ci | (see item (12) in the note for III.16, `. 13–15) and µ > 0 ⇔ c i > 0
i=1
for all i , thus
n
X
µ ∈ H0 ⇔ ci > 0 for all i and ci = 1 ;
i=1
III.37, `. 17.
“ . . . therefore D = D◦◦ (TVS, II, §6, No. 3, Cor. 3 of Th. 1).”
Moreover, since D is balanced—hence symmetric or circled, according
as E is real or complex—one has
z ∈ D ⇒ |hz, z0 i| 6 1 ,
that is,
It follows that
that is, |z0 ◦f | 6 q ◦f ; since both sides belong to K (X; R) and since |µ| > 0 ,
we infer that
|µ|(|z0 ◦ f |) 6 |µ|(q ◦ f ) ;
but |µ(z0 ◦ f )| 6 |µ|(|z0 ◦ f |) by the inequality (13) of §1, No. 6, therefore
|µ(z0 ◦ f )| 6 |µ|(q ◦ f ) ,
that is,
D Z E Z
0
f dµ, z 6 (q ◦ f ) d|µ| ,
R
Write a = (q ◦ f ) d|µ| ; the assertion of the proposition is that
(δ) q(z0 ) 6 a .
a−1 z0 ∈ D◦◦ = D , that is, q(a−1 z0 ) 6 1 ; thus q(z0 ) 6 a and (δ) is again
verified.
R
We remark that if E is a Banach space, then the condition f dµ ∈ E
is automatically verified (by Cor. 1 of Prop. 7 in No. 3 below), therefore
Z
Z
f dµ
6 kf (x)k d|µ|(x)
R
by the foregoing result; and if X is compact,
f dµ
6 kf k∞ kµk .
III.38, `. 3–5.
“ . . . the closed convex envelope C of f (X) in E is then precompact (for
the uniform structure induced by that of E ) (TVS, II, §4, No. 1, Prop. 3).”
By the cited Prop. 3, the balanced convex envelope B 0 of f (X) in E is
precompact, therefore so is its closure B = B 0 (GT, II, §4, No. 2, Prop. 1);
since B is closed and convex (it is the closed balanced convex envelope
of f (X) by TVS, I, §1, No. 5, Prop. 2 and II, §2, No. 6, Prop. 14) it con-
tains C , therefore C is also precompact (GT, II, §4, No. 2, Prop. 1).
In fact, C is compact, as noted in the Corollary of the cited Prop. 3.
III.38, `. 8, 9.
“ . . . C is the closed convex envelope of f (X) in E 0 * for the topology
σ(E *, E0 ) ”
0
are continuous for σ(E, E0 ) . The assertion to be proved is that θ(C) is the
closed convex envelope of θ f (X) in E0 * .
Let C0 be the convex envelope of f (X) in E . The closure of C 0 in E
is the same for σ(E, E0 ) and the original topology of E (TVS, IV, §1, No. 2,
Prop. 2), therefore C is also the closed convex envelope of f (X) in E for
σ(E, E0 ) , that is, C = C0 (closure for either topology).
Let D0 be the convex envelope of θ f (X) in E0 * and let D = D0 be
its closure. Since θ is linear, θ(C0 ) = D0 . We know that C is compact
INT III.x85 measures §3
where D is defined as in the precedingR note, and that note R shows that
ak D ⊂ θ(E) for 1 6 k 6 4 , therefore f dµ ∈ θ(E) ; that is, f dµ = b z0 for
some z0 ∈ ER . R
When f dµ = b z0 , z0 ∈ E , it is useful
R to redefine f dµ to be the
unique element z0 ∈ E such that hz0 , z0 i = hf , z0 i dµ for all z0 ∈ E0 .
III.38, `. 13–14.
“Since the duals of E and E b are identical, it suffices to apply Prop. 7
while regarding f as taking its values in E b .”
Strictly speaking, if E is regarded as a (dense) linear subspace of E b,
and if f is regarded as taking its values in E b , that is, as an element of
R
b
K (X; E) , then f dµ becomes (by Prop. 7) the unique element w 0 ∈ E b
R
0 0
such that hw0 , w i = hf , w i dµ for all w ∈ (E) 0 b . The fact that w0
0
b ; since such neighborhoods are fundamental (GT, II, §1, No. 2, Cor. 3
in E
of Prop. 2 and III, §3, No. 1), it will suffice to find a set V ∈ N such that
V ⊂ W (closure in E b ). Since E has the topology induced by E b , W ∩ E is
a neighborhood of 0 in E ; by assumption, there exists a set V ∈ N such
that V ⊂ W ∩ E , whence V ⊂ W = W .
III.38, `. 18.
“ . . . hence bounded . . . ”
Every precompact subset of a locally convex space is bounded (TVS,
III, §1, No. 2, Prop. 2).
An argument that does not appeal to the concept of total boundedness:
since f (X) is compact, it is bounded (by a simplification of the proof of
the Prop. 2 just cited, avoiding the concept of the completion of a uniform
space), therefore its closed convex envelope is also bounded (TVS, III, §1,
No. 2, Prop. 1).
III.38, `. −11 to −9.
“ 2◦ E is the dual of a barreled Hausdorff locally convex space G , and
E is equipped with an S-topology, where S is a covering of G by bounded
subsets (TVS, III, §4, No. 2, Cor. 4 of Th. 1).”
The adaptation of the cited Cor. 4 to the present notation is as follows:
Let G and F be two locally convex spaces, S a covering of G consisting
of bounded subsets. If G is barreled and F is Hausdorff and quasi-complete,
then the space LS (G; F) is Hausdorff and quasi-complete.
One takes F = K , where K = R or C according as the spaces are
real or complex, and one sets E = LS (G; K) . Recall that L (G; K) is the
vector space G0 of continuous linear forms on G , and E = L S (G; K) is
the vector space G0 equipped with the topology of uniform convergence in
the subsets M ∈ S of G .
III.38, `. −8, −7.
“For example, Cor. 2 of Prop. 7 can be applied when E is the weak
dual of a Banach space”
In the preceding note, one takes G to be a Banach space and S to
be the set of its 1-element subsets {y} (y ∈ G) (equivalently, the set of its
finite subsets).
III.38, `. −7, −6.
“ . . . or a space of measures M (Y; C) equipped with the vague topol-
ogy.”
Here one takes (in the foregoing) G = K (Y; C) equipped with the
direct limit topology (a barreled space, by §1, No. 1, Prop. 2) and S to be
the set of its 1-element subsets {f } (f ∈ K (Y; C) ) (§1, No. 9).
INT III.x87 measures §3
III.39, `. 2–5.
“Proposition 8. . . . ”
R
When f dµ ∈ E for all f ∈ K (X; E) (for example, when E is quasi-
complete, cf. No. 3, Cor. 2 of Prop. 7), it is a corollaryR of Prop. 8 that
the mapping Φ0 : K (X; E) → E defined by Φ0 (f ) = f dµ is the unique
continuous linear mapping K (X; E) → E such that g ·a 7→ µ(g)·a for every
vector a ∈ E and every function g ∈ K (X; C) . For, since E is a topological
subspace of Eb , equivalently, the topology of E is the initial topology for the
canonical injection ι : E → E b , the continuity of Φ0 is equivalent to that of
ι ◦ Φ0 = Φ (GT, I, §2, No. 3, Prop. 4).
III.39, `. 6–8. R
“To prove the continuity of the mapping f 7→ f dµ , it suffices to show
that its restriction to K (X, K; E) is continuous for every compact subset
K of X (TVS, II, §4, No. 4, Prop. 5).”
The issue has essentially been addressed in §1, No. 1 (see the note for
III.2, `. 7, 8); as some aspects are employed here for the first time, a review
is in order. R
In line with No. 3, Cor. 1 of Prop. 7, we regard f dµ as the unique
R
element w0 ∈ E b such that hw0 , w0 i = hf , w0 i dµ for all w0 ∈ (E)
b 0 , so that
R
f 7→ f dµ is a linear mapping of K (X; E) into the locally convex space E b.
Let Tu be the topology on K (X; E) of uniform convergence in X .
For every compact subset K of X , the topology on K (X, K; E) is by def-
inition the topology TK induced by Tu , that is, the topology of uniform
convergence (in X or in K , it comes to the same). On the other hand, the
(direct limit) topology T on K (X; E) (§1, No. 1) is by definition the “final
locally convex topology” for the family of mappings u K , where K runs over
the set of all compact subsets of X , and u K : K (X, K; E) → K (X; E) is
the canonical injection; that is, T is the finest locally convex topology on
K (X; E) for which all the uK are continuous (TVS, II, §4, No. 4, Prop. 5
and Example).
One knows that if G is a locally convex space and u : K (X; E) → G
is a linear mapping, then u is continuous (for T ) if and only if u ◦ u K is
continuous for every K (TVS, loc. cit.). The assertion at hand then follows
from the observation that u ◦ uK is the restriction of u to K (X, K; E) .
{In particular, when E = C , a linear form on K (X; E) is a measure
on X if and only if its restriction to every K (X, K; C) is continuous—as
noted following the definition of a measure (§1, No. 3)}.
To conclude the review, since Tu obviously renders the uK continuous,
T is finer than Tu , and the inclusion T ⊃ Tu may be proper (for example,
when X = N equipped with the discrete topology; see item (9) in the note
§3 integrals of continuous vectorial functions INT III.x88
TK = Tu ∩ K (X, K; E) ⊂ T ∩ K (X, K; E) ,
A = {x : p(x) 6 1 } ;
§3 integrals of continuous vectorial functions INT III.x90
|λ| z = λγ z ∈ ρA ⇔ λz ∈ ρA ,
p 6 q ⇔ {x : p(x) 6 1 } ⊃ {x : q(x) 6 1 };
for, assuming that the inclusion on the right holds, given any vector x and,
−1
for any ε > 0 , setting yε = q(x) + ε x , one has
−1
q(yε ) = q(x) + ε q(x) < 1 ,
−1
therefore p(yε ) 6 1 by hypothesis, and so q(x) + ε p(x) 6 1 , that is,
p(x) 6 q(x) + ε .
(iv) Given any ε > 0 , one has
For, writing U for the set on the left, U is convex and it is open by the
◦
continuity of p , whence U ⊂ A ; and U ⊂ A = A . In fact, U = A ; for,
if x ∈ A and 0 < α < 1 then p(αx) = α p(x) 6 α < 1 , thus αx ∈ U ,
and, since αx → x as α → 1 , every neighborhood of x contains such a
vector αx , whence x ∈ U . Since U is a convex open set, it is equal to
the interior of its closure (TVS, II, §2, No. 7, Cor. 1 of Prop. 16), that is,
◦
U = A.
(7) Let (Nα ) be a fundamental system of neighborhoods of 0 in E , so
that, by (1), the sets
Vα = {(x, y) ∈ E × E : x − y ∈ Nα }
Nβ + Nβ ⊂ Nα and Nγ ⊂ −Nα
§3 integrals of continuous vectorial functions INT III.x92
Nα = {x ∈ E : qα (x) 6 1 } ;
(∗) Uα = {f ∈ K (X, K; E) : pα (f ) 6 1 } .
III.39, `. 13–16.
“ . . . by No. 2, Prop. 6 we have, for every function f ∈ K (X, K; E) ,
Z Z Z
qα f dµ = qα hf dµ 6 h(x)qα f (x) d|µ|(x) 6 |µ|(h) · pα (f )
(2) With the above notations, the displayed inequalities in the text
b ,
become, for all f * ∈ K (X, K; E)
Z Z Z
∗ ∗ ∗
qα f dµ = qα hf dµ 6 h(x)qα∗ f ∗ (x) d|µ|(x) 6 |µ|(h)·p∗α (f ∗ ) ,
∗
where
(i) p∗α (f ∗ ) = sup qα∗ f ∗ (x) ;
x∈K
the first inequality follows from the cited Prop. 6 and the fact that
qα∗ ◦ (h · f ∗ ) = h · (qα∗ ◦ f ∗ ) as functions on X , whereas the second inequality
follows from (i). From this, we need only retain
Z
(ii) qα∗ f dµ 6 |µ|(h) · p∗α (f ∗ ) .
∗
§3 integrals of continuous vectorial functions INT III.x94
thus the right side of (ii) becomes |µ|(h) · p α (f ) . We wish to prove that
Z
(iii) q∗α f dµ 6 |µ|(h) · pα (f ) ,
DZ E Z
(ι ◦ f ) dµ, u = h(ι ◦ f )(x), ui dµ(x)
Z DZ E
= hf (x), ui dµ(x) = f dµ, u .
Z
= g(x)ha, z0 i dµ(x)
Z
0
= ha, z i g(x) dµ(x)
= hµ(g) · a, z0 i ;
R
b , hence may
thus (g · a) dµ is equal to the canonical image of µ(g) · a in E
be identified with the element µ(g) · a .
INT III.x95 measures §3
is continuous
for σ(E0 , E) , the relative compactness (i.e., boundedness) of
(v(H) (x0 ) = ϕ(H) follows from ϕ(H) ⊂ ϕ(H) and the compactness of H .
To summarize, we have shown that the closure of v(H) in K is con-
tained in K (X, K; C) and is compact, hence is a strictly compact subset
of K .
III.40, `. 11, 12.
“ . . . u is none other than the restriction to M (X; C) of the trans-
pose t v (in the algebraic sense)
The conventions of the preceding note are in force; in particular, E is
complete, and we abbreviate K = K (X; C) , M = M (X; C) , writing K *
and M * for their algebraic duals. The linear mappings
u : M → E , v : E0 → K
R
are defined by u(µ) = f dµ and v(z0 ) = z0 ◦ f ( µ ∈ M , z0 ∈ E0 ).
Equip E0 with the topology σ(E0 , E0 *) , and K with the topology
σ(K , K *) ; then v is continuous. For, if f ∈ K * then f ◦ v ∈ E 0 * is
continuous for σ(E0 , E0 *)
v-K f-
E0 C
and since σ(K , K *) is the initial topology for the linear forms f ∈ K *
(TVS, II, §6, No. 2, Def. 2), it follows that v is continuous (GT, I, §2, No. 3,
Prop. 4). The transposed linear mapping
t
v : K * → E0 *
not appear until 1981. Unable to find a suitable substitute reference in the
first edition at the time (1970’s) that I first studied Intégration, I improvised
a “direct proof” based on ingredients available in the first edition of Ch. III
of Esp. vect. top. When I was preparing (1998) the published translation,
Professor Jacques Dixmier guided me to the correct reference in TVS and
sketched the indicated proof of the continuity of u ; an exposition of his
proof (Proof #2) is given immediately after the following “direct proof”.
Proof #1. The conventions of the preceding two notes are in force.
−1
Assuming V is a neighborhood of 0 in E , let us show that u (V) is a
neighborhood of 0 in M for the topology τ scc of strictly compact conver-
gence on M . We can suppose that V is convex, balanced and closed in E .
For the canonical duality hz, z0 i = z0 (z) between E and E0 , the polar of V
in E0 is the set
(TVS, II, §8, No. 4); moreover, V ◦ is convex, balanced, and closed for the
weak topology σ(E0 , E) , and, by the theorem on bipolars (loc. cit., §6, No. 3,
Cor. 3 of Th. 1), V = V◦◦ (the polar of V◦ in E ).
Let H = V◦ ⊂ E0 ; since H◦ = V is a neighborhood of 0 in E , H is
equicontinuous on E (TVS, III, §3, No. 5, Prop. 7), therefore the closure
v(H) of v(H) in K is strictly compact (see (ii) of the Note for III.40,
`. 4–7).
Thus v(H) ∈ S , the set of all strictly compact subsets of K . The
topology τscc on M is the topology of uniform convergence in the sets
S ∈ S . It follows (GT, X, §1, No. 2) that if S ∈ S and D = {c ∈ C :
|c| 6 1 } is the closed unit disc in C , so that the set
D = {(c, d) ∈ C × C : c − d ∈ D }
v : E0 → K
it follows (TVS, II, §6, No. 4, Prop. 5) that, in the notations of (*), u and v
are continuous for the topologies σ in question, and each is the transpose of
the other (in the sense of ibid., Remark ), concisely t u = v (and t v = u ).
Our problem is to show that the linear mapping v : M → E remains
continuous when M is equipped with the topology of strictly compact con-
vergence (in general stronger than σ(E 02 , E2 ) = σ(M , K ) , the vague topol-
ogy on M ), and E with its original topology (in general stronger than
σ(E01 , E1 ) = σ(E, E0 ) , the weakened topology on E ); this is fertile terrain
for the S-topologies associated with dualities (TVS, III, §3, No. 5). Indeed,
if S2 is the set of all strictly compact subsets of E 2 = K , then the topol-
ogy of (E02 )S2 is the topology of strictly compact convergence in E 02 = M ;
and if S1 is the set of all subsets of E1 = E0 that are equicontinuous with
INT III.x101 measures §4
III.41, `. 6, 7.
“It is immediate that the image under ω of K (X × Y, K × L; C) is
contained in K (X, K; K (Y, L; C)) .”
If f : X × Y → C then ω(f ) : X → F (Y; C) is defined as follows: for
each x ∈ X , ω(f ) (x) : Y → C is defined by
ω(f ) (x) (y) = f (x, y) for all y ∈ Y .
If f is continuous then,
for each x ∈ X , the mapping y 7→ f (x, y) is con-
tinuous, hence ω(f ) (x) is continuous on Y , and so ω(f ) : X → C (Y; C) .
§4 products of measures INT III.x102
therefore ω(f ) (x) ∈ K (Y, L; C) for all x ∈ X . Moreover, if x ∈ X --- K ,
then
ω(f ) (x) (y) = f (x, y) = 0 for all y ∈ Y ,
therefore ω(f ) (x) = 0 , the zero element of K (Y, L; C) ; thus
ω(f ) ∈ F X, K; K (Y, L; C) .
To show that ω(f ) ∈ K X, K; K (Y, L; C) , we need only show that ω(f )
is continuous for the given topology on X and the norm topology on
K (Y, L; C) . Since the norm topology on K (Y, L; C) coincides with the
topology of uniform convergence in the compact subsets of Y , K (Y, L; C)
is a topological subspace of Cc (Y; C) ; the desired continuity of ω(f ) is
therefore a consequence of GT, X, §3, No. 4, Th. 3 (with Z = C and
fe = ω(f ) ).
III.41, `. 7–10.
“ . . . if u is a continuous mapping of X into K (Y, L; C) , with support
contained in K , then the mapping (x, y) 7→ u(x)(y) of X × Y into C is
continuous and has support contained in K × L ”
Define f : X × Y → C by f (x, y) = u(x) (y) . For every x ∈ X ,
ω(f ) (x) (y) = f (x, y) = u(x) (y) for all y ∈ Y ,
hence ω(f ) (x) = u(x) for all x ∈ X ; thus ω(f ) = u . The continuity of
ω(f ) : X → K (Y, L; C) ⊂ Cc (Y; C) then implies that of f , by GT, X, §3,
No. 4, Th. 3 (the “conversely” part, with Z = C and fe = ω(f ) ), and it is
clear that f (x, y) 6= 0 implies (x, y) ∈ K × L .
III.41, `. 14–18.
“ . . . the image under ω , of
K (X, K; C) ⊗ C K (Y, L; C)
P
n
An element f = gi ⊗hi of the displayed tensor product space, viewed
i=1
as a function on X × Y (the first identification referred to), is the function
n
X
f (x, y) = gi (x)hi (y) .
i=1
thus n
X
ω(f ) (x) = gi (x) · hi ∈ K (Y, L; C) for all x ∈ X .
i=1
K (X, K; C) ⊗ C K (Y, L; C) = K (X × Y, K × L; C) .
K (X, K) → K (X) → K (X × Y)
uK : K (X, K) → K (X × Y)
K (X, K) → K (X × Y, K × L) → K (X × Y)
K (X, K) → K (X × Y, K × L)
INT III.x105 measures §4
f ⊗g (f ∈ A, g ∈ B)
is total in K (X × Y) .
Proof. Let ν be a measure on X × Y such that ν(f ⊗ g) = 0 for all
f ∈ A and g ∈ B ; we are to show that ν = 0 . Fix g ∈ B . It follows
from the Lemma that the function f 7→ ν(f ⊗ g) is a continuous linear form
on K (X) ; since it vanishes on the total set A , it is identically zero. Thus
ν(f ⊗ g) = 0 for all f ∈ K (X) and g ∈ B . Similarly, fixing f ∈ K (X) ,
the validity of ν(f ⊗ g) = 0 for all g ∈ B implies that ν(f ⊗ g) = 0 for all
g ∈ K (Y) . Thus ν = 0 .
Corollary. — With notations as in the Proposition, if ν is a measure
on X × Y such that ν(f ⊗ g) = 0 for all f ∈ A and g ∈ B , then ν = 0 .
III.42, `. 3–5.
“ . . . for u = ω(f ) and for every y ∈ Y ,
Z Z Z
u(x) dλ(x), εy = u(x)(y) dλ(x) = f (x, y) dλ(x) ,
but (ηy ◦u)(x) = ηy u(x) = εy u(x) = u(x) (y) = f (x, y) , thus ηy ◦u =
f (·, y) and so
DZ E Z
u dλ, ηy = f (·, y) dλ
Z
= f (x, y) dλ(x) = h(y) ,
§4 products of measures INT III.x106
briefly
DZ E
(*) h(y) = u dλ, ηy for all y ∈ Y .
and it satisfies
Z Z
(ii) ν(f ) = f (x, y) dλ(x) dµ(y) for all f ∈ K (X × Y; C) .
and it satisfies
Z Z
(iv) ρ(k) = k(y, x) dµ(y) dλ(x) for all k ∈ K (Y × X; C) .
= ρ(f 0 )
Z
= f 0 (y, x)dρ(y, x)
Z
= f (x, y)dρ(y, x)
Z
= h(y)g(x)dρ(y, x)
= µ(h)λ(g) by (iii)
Z
= g(x)h(y) dν(x, y) by (i) .
§4 products of measures INT III.x108
= ρ(f 0 )
Z Z
= f 0 (y, x) dµ(y) dλ(x) by (iv)
Z Z
= f (x, y) dµ(y) dλ(x)
as claimed.
III.44, `. 1–4.
“Indeed, for every z0 ∈ E0 we have
Z Z ZZ Z Z
0
f dλ dµ, z = hf , z i dλ dµ = dµ hf , z0 i dλ
0
Z Z Z Z
0 0
= f dλ, z dµ = dµ f dλ, z
whence Z Z Z
f dν = f (·, y) dλ dµ(y)
Z Z
= f (x, y) dλ(x) dµ(y)
Z Z
= dµ(y) f (x, y) dλ(x) ;
INT III.x109 measures §4
R 0
Rthe argument shows that h dµ , a priori an element of E * , is equal to
f dν hence belongs to E . Similarly,
DZ E Z Z
f dν, z0 = hf (x, y), z0 i dµ(y) dλ(x)
DZ Z E
= f (x, ·) dµ dλ(x), z0 ,
whence Z Z Z
f dν = f (x, ·) dµ dλ(x)
Z Z
= dλ(x) f (x, y) dµ(y) .
III.44, `. −5.
“ . . . which proves formula (6).”
For all f ∈ K (X × Y; C) ,
Z
hf, (g ⊗ h) · (λ ⊗ µ)i = f · (g ⊗ h) d(λ ⊗ µ)
Z
= f (x, y)g(x)h(y) d(λ ⊗ µ)(x, y)
Z Z
= f (x, y)g(x)h(y) dµ(y) dλ(x) by (3)
Z Z
= f (x, y)h(y) dµ(y) g(x) dλ(x)
Z Z
= f (x, ·) d(h · µ) d(g · λ)(x)
Z
= f d (g · λ) ⊗ (h · µ) by (3)
= hf, (g · λ) ⊗ (h · µ)i ,
whence (g ⊗ h) · (λ ⊗ µ) = (g · λ) ⊗ (h · µ) .
III.45, `. 1–3.
“ . . . if U (resp. V ) is an open set in X (resp. Y ), then the restriction
of λ ⊗ µ to the product U × V is the product of the restrictions of λ to U
and of µ to V ”
We know from the first paragraph of §2, No. 1, that K (X, U; C) may
be identified with K (U; C) via the mapping f 7→ f U , and the measure
λU on U is the result of transporting the restriction λ K (X, U; C) to a
linear form on K (U; C) via this identification; that is,
(λU)(f U) = λ(f ) for all f ∈ K (X, U; C) .
§4 products of measures INT III.x110
(see
the note for III.23, `. 16–17). Similarly
for the relation between µ and
µV , and between
λ ⊗
µ and (λ ⊗ µ)
U × V .
Both (λU) ⊗ (µV) and (λ ⊗ µ)U × V are measures on U × V ; if g ∈
K (U; C) and h ∈ K (V; C) , then g ⊗ h ∈ K (U × V; C) , g 0 ⊗ h0 = (g ⊗ h)0
on X × Y , and
[(λU) ⊗ (µV)](g ⊗ h) = (λU)(g) · (µV)(h)
= λ(g 0 )µ(h0 )
= (λ ⊗ µ)(g 0 ⊗ h0 )
= λ ⊗ µ (g ⊗ h)0
= [(λ ⊗ µ)U × V)](g ⊗ h) ,
therefore (λU) ⊗ (µV) = (λ ⊗ µ)U × V by the uniqueness part of No. 1,
Th. 1.
III.45, `. 6–8.
“ . . . which proves the proposition, on taking into account the definition
of the support of a measure (§2, No. 2).”
Let U = X --- Supp λ . Then λU = 0 and, writing ν = λ ⊗ µ , we have
ν U × Y = (λU) ⊗ (µY) = 0 ⊗ µ = 0 ,
Supp ν ⊂ { (U × Y) = ( { U) × Y = (Supp λ) × Y .
(x, y) ∈ U × V ⊂ { Supp ν .
Then ν U × V = 0 , thus (λU) ⊗ (µV) = (λ ⊗ µ)U × V = 0 , therefore
λU = 0 or µV = 0 , whence U ⊂ { Supp λ or V ⊂ { Supp µ , and so
x ∈ { Supp λ or y ∈ { Supp µ .
INT III.x111 measures §4
III.47, `. 3.
“There exist . . . ”
By §1, No. 8, Prop. 10.
III.47, `. 13–17.
“Proposition 6. — When M (X; C) , M (Y; C) and M (X × Y; C)
are equipped with the topology of strictly compact convergence (§1, No. 10),
the bilinear mapping (λ, µ) 7→ λ ⊗ µ of M (X; C) × M (Y; C) into
M (X × Y; C) is hypocontinuous for the set of vaguely bounded subsets of
M (X; C) and M (Y; C) (TVS, III, §5, No. 3).”
The assertion is that if S and T are the sets of vaguely bounded
subsets of M (X; C) and M (Y; C) , respectively, then the bilinear mapping
u : (λ, µ) 7→ λ ⊗ µ is (S, T)-hypocontinuous (in the sense of TVS, III, §5,
No. 3), that is, (i) it is separately continuous (for the topologies of strictly
compact convergence), (ii) for every M ∈ S and every neighborhood W
of 0 in M (X × Y; C) , there exists a neighborhood V of 0 in M (Y; C)
such that u(M × V) ⊂ W , and (ii0 ) for every N ∈ T and every neighbor-
hood W of 0 in M (X × Y; C) , there exists a neighborhood U of 0 in
M (X; C) such that u(U × N) ⊂ W .
(The conditions (i) and (ii) express that ϕ is S-hypocontinuous, whereas
(i) and (ii0 ) express that it is T-hypocontinuous.)
Remarks. — Suppose E, F, G are locally convex spaces, ϕ : E × F → G
is bilinear, and S (resp. T ) is a set of bounded subsets of E (resp. F ).
If ϕ is continuous (“jointly”) then it is (S, T)-hypocontinuous for all such
S and T (TVS, III, §5, No. 3, discussion following Def. 2). Thus (S, T)-
hypocontinuity is in general stronger than separate continuity and weaker
than continuity. According to Exercise 3 for §4, the bilinear mapping
u : (λ, µ) 7→ λ ⊗ µ is continuous (not just separately) for the topologies
of strictly compact convergence. It then follows from the foregoing that
u is (S, T)-hypocontinuous when S (resp. T ) is any set of sets that
are bounded for the topology τscc of strictly compact convergence; but
the sets bounded for τscc are the same as the vaguely bounded sets (§1,
No. 10, Prop. 17), thus one recovers Prop. 6, enhanced with joint continuity
(for τscc ).
III.47, `. −14 to −11.
“the mapping ϕ of K (X, K; K (Y, L; C)) × M (X; C) Rinto K (Y, L; C),
such that ϕ(g, λ) is the function h defined by h(y) = g(x, y) dλ(x) , is
separately continuous by virtue of §3, No. 4, Props. 8 and 9.”
There is here an “abuse of notation”: a function
g ∈ K (X × Y, K × L; C)
§4 products of measures INT III.x112
ϕ : K (X × Y, K × L; C) × M (X; C) → E
R R
by ϕ(g, λ) = ω(g) dλ = g(x, ·) dλ(x) ∈ E b = E (§3, No. 3, Cor. 2 of
Prop. 7).
For fixed g ∈ K (X × Y, K × L; C) , hence fixed ω(g) ∈ K (X, K; E) ,
the mapping
Z
λ 7→ ϕ(g, λ) = ω(g) dλ λ ∈ M (X; C)
is continuous for the norm topology (§3, No. 4, Prop. 8), and since g 7→ ω(g)
is an isometry K (X × Y, K × L; C) → K (X, K; E) (No. 1, Lemma 1), the
mapping
Z
g 7→ ω(g) dλ = ϕ(g, λ) g ∈ K (X × Y, K × L; C)
= hϕ(f, λ), µi .
the validity of (*) for all f ∈ A says, by (ii), that λ ⊗ µ ∈ A ◦ . This proves
the assertion
(λ, µ) ∈ B × C◦ ⇒ λ ⊗ µ ∈ A◦ .
So to speak, B ⊗ C◦ ⊂ A◦ .
One notes that if µ ∈ M (Y; C) is such that λ ⊗ µ ∈ A ◦ for all λ ∈ B ,
then the validity of (*) for all f ∈ A and λ ∈ B shows, since ϕ(f, λ) then
runs over ϕ(A × B) = C , that µ ∈ C◦ . Thus if µ ∈ M (Y; C) , then
λ ⊗ µ ∈ A◦ for all λ ∈ B ⇒ µ ∈ C◦ ,
C◦ = {µ ∈ M (Y; C) : B ⊗ µ ⊂ A◦ } .
u : M (X; C) × M (Y; C) → M (X × Y; C) .
Let us take on faith (details later) that the sets of the form A ◦ (resp. C◦ )
are a fundamental system of neighborhoods of 0 in M (X × Y; C) (resp.
M (Y; C) ) for the topology of strictly compact convergence. As in the note
for III.47, `. 13–17, let us write S (resp. T ) for the set of vaguely bounded
subsets of M (X; C) (resp. M (Y; C) ).
Given a neighborhood W = A◦ of 0 in M (X × Y; C) and a set
B ∈ S that is vaguely closed (not a restriction, since, in any real or complex
topological vector space, the closure of a bounded set is bounded—because
the closed neighborhoods of 0 are fundamental), we have produced a neigh-
borhood C◦ of 0 in M (Y; C) such that u(B × C◦ ) ⊂ A◦ , informally
INT III.x115 measures §4
Vε = {(c, d) ∈ C × C : |c − d| 6 ε } (ε > 0) ;
W ε1 , S 1 ∩ · · · ∩ W εn , S n
whence Wε, A = W1, ε−1 A ; since the ε−1 A run over the closed, balanced
convex elements of A as A does, the sets
(*) U A = A◦
INT III.x117 measures §4
(TVS, II, §8, No. 4); thus the sets (*), as A runs over the set of balanced,
convex, strictly compact subsets of K , form a fundamental system of neigh-
borhoods of 0 ∈ M for the topology of strictly compact convergence.
Incidentally, by the theorem on bipolars (TVS, II, §6, No. 3, Th. 1),
◦◦
A is the closure of A in K for the weakened topology σ(K , M ) . But
the strongly closed convex set A is also closed for σ(K , M ) (loc. cit., Cor. 3
of Th. 1); thus A◦◦ = A .
III.48, `. 1, 2.
“The conclusion of Prop. 6 is no longer valid when the topology of
strictly compact convergence is replaced by the vague topology (Exer. 2 c)).”
A surprising result, since u is separately continuous for the vague topol-
ogy (Prop. 5).
In the cited exercise, X = Y = [0, 1] , and M (X; C) , M (X × X; C)
are equipped with the vague topology. Write u(λ, µ) = λ ⊗ µ as in the
preceding note. Let S be the set of vaguely bounded subsets of M (X; C) .
The set B = {µ ∈ M (X; C) : kµk 6 1 } is vaguely compact (§1, No. 9,
Cor. 2 of Prop. 15), hence vaguely bounded. If u were S-hypocontinuous
for the vague topology, then the restriction of u to B × M (X; C) would be
continuous (TVS, III, §5, No. 3, Prop. 4), contrary to part c) of the exercise.
III.48, `. 2–6.
“ . . . if B (resp. C ) is a vaguely bounded subset of M (X; C) (resp.
M (Y; C) ), then the image of B × C under the mapping (λ, µ) 7→ λ ⊗ µ
is vaguely bounded in M (X × Y; C) and therefore the restriction of this
mapping to B × C is vaguely continuous . . . ”
Let u : M (X)×M (Y) → M (X×Y) be the bilinear mapping u(λ, µ) =
λ ⊗ µ , and let S (resp. T ) be the set of vaguely bounded subsets of M (X)
(resp. M (Y) ); we know from Prop. 6 that u is (S, T)-hypcontinuous when
each of M (X) , M (Y) and M (X × Y) is equipped with the topology τ scc
of strictly compact convergence.
From §1, No. 10, Prop. 17, we know that S is also the set of subsets
of X that are bounded for τscc , and similarly for T and Y . By TVS, III,
§5, No. 3, Props. 4, 5, u(B × C) is bounded for τ scc —hence for the vague
topology—and the restriction of u to B × C is continuous when each of
B , C is equipped with the topology induced by τ scc , and B × C with the
product topology. Since B , C and u(B × C) are vaguely bounded, the
topologies induced on them by τscc and the vague topology are identical by
the Prop. 17 cited above; and the product topology on B × C coincides with
the topology induced by the product of the vague topologies on M (X) and
M (Y) (GT, I, §4, No. 1, Cor. of Prop. 3); therefore the restriction u B × C
is continuous when all spaces in sight are equipped with the vague topology.
§4 products of measures INT III.x118
III.48, `. 10–14.
“The set of linear combinations of complex functions of the form
(x1 , x2 , . . . , xn ) 7→ f1 (x1 )f2 (x2 ) · · · fn (xn ) ,
where fi ∈ K (Xi ; C) (1 6 i 6 n) , may be identified canonically with the
N
n
tensor product K (Xi ; C) , and it follows from Lemma 1 of No. 1, by
i=1
induction on n , that this tensor product is dense in K (X; C) .”
The case n = 2 is proved in the remarks following the cited Lemma 1.
Assume inductively that all is well for for n − 1 and consider the
case of n . Write Y = X1 × · · · × Xn−1 . The canonical homeomorphism of
X1 ×· · ·×Xn with Y×Xn permits the identification of K (X1 ×· · ·×Xn ) with
K (Y×Xn ) . By the induction hypothesis, the set of functions f 1 ⊗· · ·⊗fn−1
is total in K (Y) , therefore the set of functions
f1 ⊗ · · · ⊗ fn = (f1 ⊗ · · · ⊗ fn−1 ) ⊗ fn ( fi ∈ K (Xi ) for i = 1, . . . , n)
is total in K (Y × Xn ) = K (X1 × · · · × Xn ) (see the Proposition in the note
for III.41, `. −9, −8).
III.49, `. −13 to −11.
“The integral notation and formula (14) may be extended in an obvious
way to functions f ∈ K (X; E) with values in a Hausdorff locally convex
space E , such that f (X) is contained in a complete convex subset of E .”
Preparatory to computing the case n = 3 , let us review the Remark in
No. 1 (which is the case n = 2 ). In the context of a measure ν = λ ⊗ µ
on X × Y , if f ∈ K (X × Y; E)R is such that f (X × Y) ⊂ C ⊂ E with C
complete and convex, so that f dν ∈ E by §3, No. 3, Prop. 7, then
Z Z Z Z Z
f dν = dµ(y) f (x, y) dλ(x) = dλ(x) f (x, y) dµ(y)
and similarly
DZ E DZ Z E
0
f dν, z = dλ(x) f (x, y) dµ(y), z0
Z DZ E
= f (x, y) dµ(y), z0 dλ(x)
Z Z
= hf (x, y), z0 i dµ(y) dλ(x) .
as well as
Z Z Z Z
f dν = dµ1 (x1 ) f (x1 , x2 , x3 ) dµ2 (x2 )
dµ3 (x3 )
Z Z Z
= dµ3 (x3 ) dµ1 (x1 ) f (x1 , x2 , x3 ) dµ2 (x2 ) .
This accounts for the permutations (3,2,1) and (3,1,2). On the other hand,
Z Z Z Z
f dν = f d (µ1 ⊗ µ2 ) ⊗ µ3 = d(µ1 ⊗ µ2 )(x1 , x2 ) f (x1 , x2 , x3 ) dµ3 (x3 )
Z Z Z
= dµ1 (x1 ) dµ2 (x2 ) f (x1 , x2 , x3 ) dµ3 (x3 )
Z Z Z
= dµ1 (x1 ) dµ2 (x2 ) f (x1 , x2 , x3 ) dµ3 (x3 ) ,
as well as
Z Z Z Z
f dν = dµ1 (x1 ) f (x1 , x2 , x3 ) dµ3 (x3 )
dµ2 (x2 )
Z Z Z
= dµ2 (x2 ) dµ1 (x1 ) f (x1 , x2 , x3 ) dµ3 (x3 ) ,
§4 products of measures INT III.x120
Thus, assuming
M (Xα ; C) equipped with the vague topology, the family
M (Xα ; C) of topological vector spaces and the family (pαβ )∗ of contin-
uous linear mappings form an inverse system
(†) M (Xα ; C), (pαβ )∗
X = lim Xα
←−
Q
is a closed subspace of the compact space Xα (GT, I, §8, No. 2, Cor. 2 of
α
Prop. 7), the canonical mapping
pα : X → X α
INT III.x121 measures §4
The sense in which the term “inverse system” is applied to the family of
mappings
(pα )∗ : M (X; C) → M (Xα ; C)
is explained in S, III, §7, No. 2, Remark 2 (reviewed at the end of this note),
but what the property (*) is called is not important; what is important is
the role that it plays in Definition 2 below, as follows.
Consider the inverse limit of the system (†), namely
lim M (Xα ; C) ;
←−
Q
its elements are the families (µα ) ∈ M (Xα ; C) such that µα = (pαβ )∗ (µβ )
α
when α 6 β .
For example, suppose µ is a measure on the compact space X = lim Xα .
←−
For each α , define a linear form µα on C (Xα ; C) by
µα (f ) = µ(f ◦ pα ) f ∈ C (Xα ; C) ;
whence (pα )∗ (µ) = µα for all α . It then follows from (*) that, for α 6 β ,
(pαβ )∗ (µβ ) = (pαβ )∗ (pβ )∗ (µ)
(pαβ )∗ ◦ (pβ )∗ (µ) = (pα )∗ (µ) = µα ,
§4 products of measures INT III.x122
thus the family (µα ) is (in the sense of Def. 2) an inverse system of measures
admitting µ as inverse limit. One can therefore define a (linear) mapping
by the formula
Φ(µ) = (pα )∗ (µ) µ ∈ M (X; C) .
By Def. 2 the range of Φ consists of the set of all inverse systems of measures
that admit an inverse limit. A preview of Prop. 8:
By (i) of Prop. 8, Φ is injective.
By (ii) of Prop. 8, if (µα ) is an inverse system admitting an inverse
limit µ , then the family (kµα k) is bounded (indeed, (µα ) = Φ(µ) satisfies
kµα k 6 kµk for all α as noted above).
By (iii) of Prop. 8, if the pαβ are surjective, then every inverse sys-
tem (µα ) for which (kµα k) is bounded has an inverse limit; in this case,
the range of Φ is precisely the set of elements (µ α ) of lim M (Xα ; C) for
←−
which (kµα k) is bounded, and Φ is a bijection of M (X; C) onto this set
(a linear subspace of lim M (Xα ; C) ).
←−
By (iv) of Prop. 8, if the pαβ are surjective, then Φ is a bijection
of M+ (X; C) onto the set of elements (µα ) of lim M (Xα ; C) such that
←−
µα > 0 for all α , and, when µ ∈ M+ (X; C) , kµα k= kµk for all α .
To summarize, the family of mappings (pα )∗ and their property (*)
are the link between inverse systems of measures on the X α and possible
measures µ on the space X = lim Xα that may correspond to them.
←−
We now review the justification for calling the family (pα )∗ an “inverse
system” of mappings.
Let (Eα , fαβ ) be an inverse system, E = lim Eα , and fα : E → Eα the
←−
canonical mapping.
{In the application at hand, Eα = M (Xα ) (we drop the “ ;C” for
brevity), fαβ = (pαβ )∗ , E = lim M (Xα ) and fα the restriction of
Q ←−
prα : M (Xβ ) → M (Xα ) to lim M (Xβ ) . Thus fα (µβ ) = µα for every
β ←−
inverse system of measures µβ ∈ M (Xβ ) .}
Let (Fα , gαβ ) be a second inverse system, F = lim Fα , gα : F → Fα
←−
the canonical mapping.
{In the application at hand, Fα = M (X) for all α , gαβ is the identity
mapping on M (X) ; the elements (µα ) of lim Fα are then the “constant
←−
families” (µα ) where, for some µ ∈ M (X) , µα = µ for all α ; and the
INT III.x123 measures §4
gαβ fαβ
? ?
Fα u - Eα
α
id (pαβ )∗
? ?
M (X) - M (Xα )
(pα )∗
whose commutativity reduces to the relation
whenever α 6 β .}
By S, III, §7, No. 2, Cor. 1 of Prop. 1, there exists a unique mapping
u : F → E such that the diagram
F u- E
gα fα
? ?
Fα u - Eα
α
is commutative for all α . The mapping u is then called the inverse limit
of the family (uα ) , written u = lim uα (S, III, §7, No. 2).
←−
{In the application at hand, the diagram takes the form
§4 products of measures INT III.x124
M (X) u - lim M (X )
β
←−
id fα
? ?
M (X) - M (Xα )
(pα )∗
Q
where fα is the restriction of prα : M (Xβ ) → M (Xα ) to lim M (Xβ ) .
←−
The property required of u is that fα u(µ) = (pα )∗ (µ) for all α ; since
Φ(µ) = (pβ )∗ (µ) has this property, Φ = u by the uniqueness of u . Thus
(pα )∗ is an inverse system of mappings and Φ = lim (pα )∗ .}
←−
belongs to F . . . ”
In slow motion,
g + h = g β ◦ pβ + hγ ◦ pγ
(*) = gβ ◦ (pβα ◦ pα ) + hγ ◦ (gγα ◦ pα )
= (gβ ◦ pβα ) ◦ pα + (hγ ◦ pγα ) ◦ pα .
In particular,
A + B pα (x) = A pα (x) + B pα (x) for all x ∈ X ,
g + h = A ◦ pα + B ◦ pα = (A + B) ◦ pα = fα ◦ pα ,
III.52, `. 10.
“ . . . (pα )∗ (µ) = µα for all α ∈ I . . . ”
For every fα ∈ C (Xα ; C) ,
µα (fα ) = λ(fα ◦ pα ) = µ(fα ◦ pα ) = (pα )∗ (µ) (fα ) ,
whence µα = (pα )∗ (µ) .
III.52, `. 13.
“ . . . the relation µα = (pαβ )∗ (µβ ) implies that µα (1) = µβ (1) . . . ”
Write 1α ∈ C (Xα ; C) for the constant function equal to 1 . The point
is that 1α ◦ pαβ = 1β , thus
µβ (1β ) = µβ (1α ◦ pαβ ) = (pαβ )∗ (µβ ) (1α ) = µα (1α ) .
III.52, `. 15, 16.
“ . . . the subspace F obviously satisfies the property (P) of §1, No. 7,
Prop. 9 . . . ”
The constant function 1 belongs to F because 1 = 1 α ◦ pα , where
1α ∈ C (Xα ; C) is the constant function equal to 1 , and f = 1 trivially
satisfies the requirements of (P) in the cited Prop. 9.
III.52, `. −13 to −10.
“We know that (XJ , prJK ) is an inverse system of compact spaces, and
that the inverse limit of the system of continuous mappings (pr J ) is a hom-
eomorphism of X onto the inverse limit space lim XJ , permitting one to
←−
identify these two spaces (GT, I, §4, No. 4 and S, III, §7, No. 2, Remark 3).”
I had great difficulty applying the cited references, making many mis-
takes along the way; be suspicious of every move in the following exposition.
To recapitulate the notations, we are given a family of nonempty com-
pact spaces, and Y
X= Xλ
λ∈L
is the product topological space, the topology being defined as the initial
topology for the family (prλ )λ∈L of projection mappings
prλ : X → Xλ .
No order relation on the index set L is assumed.
Let F be the set of all nonempty finite subsets J of L ; F is an in-
creasing directed set for the order relation J ⊂ K on F . For each J ∈ F
one writes Y
XJ = Xλ ,
λ∈J
§4 products of measures INT III.x126
To summarize,
(XJ , prJK )
is an inverse system
Q of sets relative to the index set F (S, III, §7, No. 1).
Writing Z = XJ , the inverse limit
J∈F
lim XJ
←−
as functions X → XJ . We write
Y = lim XJ
←−
INT III.x127 measures §4
and, for J ∈ F ,
pJ : Y → X J
for the restriction to Y of the J ’th coordinate projection mapping Z → X J .
By GT, I, §9, No. 6, Prop. 8, Y = lim XJ is compact and nonempty;
←−
our problem is to show that X is homeomorphic to Y . At any rate, there
is a natural mapping h : X → Y , defined by
Q
Q
prλ Q pK
prK Q
Q
? sQ ?
Xλ pr XK
K,λ
where prK,λ is the λ’th coordinate projection on the product space X K . The
equality pK ◦ h = prK expresses the definition of h , and pr K,λ ◦ prK = prλ
is clear from the definition of prK,λ .
We prove that h is a homeomorphism in three steps:
(i) h is injective;
(ii) h is continuous;
(iii) h is surjective.
Proof of (i): If h(x) = h(x0 ) then, for every J ∈ F , prJ x = prJ x0 ;
thus if λ ∈ L and if J is an element of F containing λ , then
on the XJ for which the norms kµJ k are bounded (by (iii) of Prop. 8; see
also the note for III. 50, `. −14, −13); and, by (iv) of Prop. 8, Φ M+ (X)
is the set of all inverse systems (µJ )J∈F for which µJ > 0 for all J , in
which case kµJ k is constant, specifically, if Φ(µ) = (µ J )J∈F then µ > 0
and kµJ k = kµk for all J ∈ F .
Combining the mappings (i) and (ii), we have a mapping
that is,
J = {1, . . . , j} , K = {1, . . . , j, j + 1, . . . , k} .
INT III.x131 measures §4
III.53, `. 1, 2.
“For (µJ ) to be an inverse system of measures, it is therefore necessary
and sufficient that either µλ = 0 for all λ ∈ L or µλ (1) = 1 for all
λ ∈ L .”
As in the preceding notes on the Example, F denotes the set of all
nonempty finite subsets J of L . The condition (17) may be written
Y
(prJ,K )∗ (µK ) (fJ ) = µJ (fJ ) · µλ (1)
λ∈K -- J
§4 products of measures INT III.x132
XK = XJ × XK -- J = XK -- J × XJ
are a consequence of the associative law for products (GT, I, §4, No. 1,
Prop. 2), the equalities expressing canonical homeomorphisms. Thus if
y ∈ XJ and z ∈ XK -- J then (y, z) and (z, y) are representations of a
same element x of XK .
INT III.x133 measures §4
N
(ii) If fλ ∈ C (Xλ ; C) for all λ ∈ J , what does fλ mean? It is the
λ∈J
function f on XJ defined by
Y
f (xλ )λ∈J = fλ (xλ ) ,
λ∈J
µK = µJ ⊗ µK -- J = µK -- J ⊗ µJ
(see the formula (v) in the note for III.52, `. −13 to −10).
§4 products of measures INT III.x134
III.53, `. −2 to III.54, `. 3.
“The positive measure
!
Y
µ= µλ (1) µ0
λ∈L
Q
The factorization of µλ (1) in the preceding line is a special case
λ∈L
of an associativity theorem for multipliable families (GT, IV, §7, No. 5,
Remark ).
For all J ∈ F we have
O O −1
µ0J = µ0λ = µλ (1) µλ
λ∈J λ∈J
Y −1 O Y −1
= µλ (1) · µλ = µλ (1) · µJ ,
λ∈J λ∈J λ∈J
whence
Y
(*) µJ = µλ (1) · µ0J ;
λ∈J
whence the formula (18). And, since µ 0 (1) = 1 , the defining formula
Y
µ= µλ (1) µ0
λ∈L
Y Y Y
(1) µλ (1) = µλ (1) .
ρ∈R λ∈Lρ λ∈L
satisfying
Y
(3) νρ (fJ ◦ prJ ) = µJ (fJ ) µλ (1)
λ∈Lρ -- J
moreover, since the family νρ (1) ρ∈R
is multipliable, Prop. 9 yields a pos-
itive measure
O Y
(5) ν= νρ on Yρ = X
ρ∈R ρ∈R
§4 products of measures INT III.x136
one has
Y
(7) ν(gS ◦ prS ) = νS (gS ) νρ (1)
ρ∈R -- S
satisfying
Y
(11) µ(fJ ◦ prJ ) = µJ (fJ ) µλ (1)
λ∈L -- J
N
since µ = µλ on X , the problem is to show that ν = µ . So to speak,
Q λ∈L
Q
Xλ and Yρ are two representations of the same space X ; we are to
λ∈L ρ∈R
show that µ and ν are two representations of the same measure on X .
INT III.x137 measures §4
S
To this end, let J be a finite subset of L = Lρ and let fJ ∈
ρ∈R
C (XJ ; C) . The set J can intersect only finitely many of the mutually
S disjoint
sets Lρ ; let S be a nonempty finite subset of R such that J ⊂ Lρ , which
ρ∈S
we can suppose to be minimal, so that J ∩ L ρ 6= ∅ for every ρ ∈ S . Then
[ [
(14) J= J ∩ Lρ = Jρ ,
ρ∈S ρ∈S
(15) gS = fJ ◦ prJ ,
thus
S = {ρ1 , . . . , ρs } ( s = card S ) .
J = J ρ1 ∪ · · · ∪ J ρs ,
as follows: Y
fJ (x) = fJρ (xλ )λ∈Jρ .
ρ∈S
Q
Define gS ∈ C (YS ; C) = C ( Yρ ; C) by the formula
ρ∈S
O
(22) gS = gρ .
ρ∈S
Y
(23) prJ y = (xλ )λ∈J = (xλ )λ∈Jρ ρ∈S
∈ X Jρ = X J ,
ρ∈S
therefore
O Y
(24) fJ (prJ y) = f Jρ (xλ )λ∈Jρ ρ∈S
= fJρ (xλ )λ∈Jρ ;
ρ∈S ρ∈S
whence, by (24),
Y O
fJ (prJ y) = gρ (yρ ) = gρ (yρ )ρ∈S = gS (y) ,
ρ∈S ρ∈S
and so
(that is, (15) holds), in other words, taking into account (20) and (22),
O O
(26) gρ = fJρ ◦ prJ .
ρ∈S ρ∈S
N Q
We now consider νS = νρ ∈ M (
Yρ ; C) = M (YS ; C) (see (6))
ρ∈S Q ρ∈S
and apply it to the function gS ∈ C (YS ; C) = C ( Yρ ; C) :
ρ∈S
O O
νS (gS ) = νρ gρ by (6) and (22)
ρ∈S ρ∈S
Y
= νρ (gρ ) by (12) of No. 4
ρ∈S
Y
= νρ fJρ ◦ prJρ by (21)
ρ∈S
Yh Y i
= µ Jρ f Jρ · µλ (1) by (3)
ρ∈S λ∈Lρ -- Jρ
hY i h Y Y i
= µ Jρ f Jρ · µλ (1)
ρ∈S ρ∈S λ∈Lρ -- Jρ
h O O i h Y Y i
= µ Jρ f Jρ · µλ (1)
ρ∈S ρ∈S ρ∈S λ∈Lρ -- Jρ
Y
= µJ (fJ ) µλ (1) ,
λ∈M -- J
N N N
where µ Jρ = µλ = µJ by No. 4, Prop. 7, and where M =
ρ∈S ρ∈S λ∈Jρ
S
Lρ , so that by disjointness one has
ρ∈S
[ [ [
(27) M --- J = Lρ --- Jρ = (Lρ --- Jρ ) ,
ρ∈S ρ∈S ρ∈S
Noting that (15) has been verified, and substituting (28) into (18), we then
INT III.x141 measures §4
the second and third equalities by (28) and (4), respectively; the indices λ
for which µλ (1) occurs as a factor in (29) are those in
[ [ [ [
(M --- J) ∪ Lρ = Lρ --- Jρ ∪ Lρ (by (27))
ρ∈R -- S ρ∈S ρ∈S ρ∈R -- S
[ [ [
= Lρ ∪ Lρ --- Jρ
ρ∈S ρ∈R -- S ρ∈R
[
= Lρ --- J
ρ∈R
= L --- J ,
whence (29) becomes
Y
(30) ν(fJ ◦ prJ ) = µJ (fJ ) µλ (1) = µ(fJ ◦ prJ ) (by (11)).
λ∈L -- J
We have thus shown that for every finite J ⊂ L , and, behind the scenes, the
finite S ⊂ R chosen preparatory to (14), one has
ν(fJ ◦ prJ ) = µ(fJ ◦ prJ )
for all fJ running over a suitable set of “elementary functions” total in
C (XJ ; C) , hence
(*) ν(f ◦ prJ ) = µ(f ◦ prJ )
for all finite linear combinations of such f J ’s. If now f ∈ C (XJ ; C) is
arbitrary, it is the uniform limit of a sequence f n of such linear combinations.
whence
kfn ◦ prJ − f ◦ prJ k = k(fn − f ) ◦ prJ k 6 kfn − f k → 0 ,
consequently (*) holds for all f ∈ C (X J ; C) by the continuity of ν and µ .
Such functions
f ◦ prJ J ⊂ L finite, f ∈ C (XJ ; C)
form a dense linear subspace of C (X; C) by Lemma 3 of No. 5, therefore
ν = µ on C (X; C) by continuity.
Ouf!
CHAPTER IV
p
Extension of a measure. L spaces
ϕU = sup{g ∈ K+ : g 6 ϕU , Supp g ⊂ U } .
INT IV.x2 extension of a measure. lp spaces §1
IV.2, `. −13. P
“ . . . which again proves the relation µ*(f ) = α(x)f (x) .”
x∈X
Recall that µ ∈ M+ (X) , X a discrete space, K+ is equal to the
linear span of the characteristic functions ϕ {x} (x ∈ X) with finite real
coefficients > 0 , and α(x) = µ(ϕ{x} ) (x ∈ X) ; thus every g ∈ K+ has the
form X
g= g(x)ϕ{x}
x∈M
thus x ∈ D implies α(x)f (x) = 0 , and the right side S of (*) may be
written X
S= α(x)f (x) .
x∈X -- D
Let
A = {x ∈ X : f (x) = +∞ } .
where, for every x ∈ X --- A , both α(x) and f (x) are finite.
Let usPfirst show that S 6 µ*(f ) . Let M be any finite subset of X --- A ;
then g = f (x)ϕ{x} belongs to K+ and g 6 f , therefore µ(g) 6 µ*(f ) ,
x∈M
that is, X
α(x)f (x) 6 µ*(f ) ;
x∈M
the validity of this inequality for every finite M ⊂ X --- A shows that
S 6 µ*(f ) .
To prove the reverse inequality, let g ∈ K + be such that g 6 f and let
us show that µ(g) 6 S . Let M be the support of g , so that
X X
g= g(x)ϕ{x} and µ(g) = α(x)g(x) .
x∈M x∈M
sup g = sup ϕ ,
g∈H ϕ∈Φ
by the definitions.
§2 negligible functions and sets INT IV.x5
IV.11, `. 5, 6.
“These propositions are the translations of Props. 10 and 13 and of
Th. 3 of No. 3 for characteristic functions of sets.”
S P
Let A = An . Then ϕA 6 ϕAn , hence
n n
X X
µ*(ϕA ) 6 µ* ϕ An 6 µ*(ϕAn )
n n
by Props. 10 and 13, whence Prop. 18. If, moreover, (A n ) is increasing, then
(ϕAn ) is increasing and ϕA = sup ϕAn , therefore µ*(ϕA ) = sup µ*(ϕAn )
n n
by Th. 3.
x 7→ fn (x) →7 ϕ fn (x) ;
let us abbreviate it u = ϕ (fn ) , thus
u(x) = ϕ fn (x) (x ∈ X) .
Similarly, let v = ϕ (gn ) be the function
v(x) = ϕ gn (x) (x ∈ X) .
§3. Lp SPACES
IV.19, `. 20–24.
“We extend Def. 1 to the case of numerical functions, finite or not,
defined on X , by again setting
Z 1/p
∗ p
Np (f ) = |f | d|µ|
for such a function f . One sees immediately that the relations (3) and (4)
also hold for these functions when f + g is defined on X and α 6= 0 .”
It is convenient to simultaneously treat the proof of Prop. 2.
If f, g : X → R and |f | 6 |g| almost everywhere in X , then |f | p 6 |g|p
almost everywhere, therefore |µ|*(|f | p ) 6 |µ|*(|g|p ) by §2, No. 3, Prop. 6
1/p 1/p
and §1, No. 3, Prop. 10, hence |µ|*(|f |p ) 6 |µ|*(|g|p ) ; thus
(i) |f | 6 |g| a.e. ⇒ Np (f ) 6 Np (g) .
INT IV.x8 extension of a measure. lp spaces §3
IV.19, `. −1 to IV.20, `. 1.
“ . . . the definition of Np (f ) , and Th. 3 of §1, No. 3, show that N p (f ) =
sup Np (gn ) .”
n
By the definition of f , gn ↑ f , hence (gn )p ↑ f p , hence |µ|* (gn )p ↑
1/p 1/p
|µ|*(f p ) by Th.3 of §1, No. 3, hence |µ|* (gn )p ↑ |µ|*(f p ) as
claimed.
IV.20, `. 6.
“The proposition follows at once from Th. 1 of §2, No. 3.”
that is, f and g are equivalent if and only if N p (f − g) = 0 for some (hence
for all) p > 1 .
IV.20, `. 16–19.
“One can then define Np (f ) for a function with values in F (resp.
in R ) defined almost everywhere in X , by setting N p (f ) = Np (e f ) ; it is
then clear that the relations (3) and (4) again hold (assuming α 6= 0 and
f + g defined almost everywhere, in the case of numerical functions, finite
or not).”
Recall (§2, No. 5) that if f is any function defined almost everywhere
in X , with values in a set G (not necessarily vectorial or numerical), the
class fe is defined as follows: let f 0 be any extension of f to all of X , with
values in G , and let fe be the set of all functions g : X → G such that
g = f 0 almost everywhere in X , that is, fe is the equivalence class of f 0 in
F (X; G) , for the relation of equality almost everywhere; if f 00 is another
extension of f to X , then f 00 is equivalent to f 0 , whence it is clear that
fe is independent of the particular extension f 0 of f to X .
When G is equal to the Banach space F or to R , and f 0 is any
extension of f to all of X , then Np (f 0 ) is defined (Def. 1 and the remarks
following Prop. 2) and is independent of the particular extension f 0 (§2,
No. 3, Prop. 6), thus the definition N p (f) e = Np (fe0 ) ( = Np (f 0 ) ) depends
only on f . The author defines Np (f ) to be equal to Np (fe) ; but what it
all comes down to is that Np (f ) may unambiguously be defined to be equal
to Np (g) , where g is any function equivalent to any extension of f to X .
If f , g are functions defined almost everywhere in X with values in F ,
then f + g is defined almost everywhere and if f 0 and g0 are extension of f
and g to X , then f 0 + g0 extends f + g and the relation (4) for f 0 and g0
says precisely Np (f + g) 6 Np (f ) + Np (g) . Similarly for (3).
If f, g are defined almost everywhere in X and if f +g is defined almost
everywhere, then there exists a negligible set N such that f , g and f+ g
are defined 0
on { N ; if 0f (resp. g 0 ) is a function on X that extends f { N
(resp. g { N ), then f , g 0 and f 0 + g 0 clearly represent the equivalence
classes fe, ge and f] + g , respectively, consequently the relation (4) for f 0
0
and g says precisely Np (f + g) 6 Np (f ) + Np (g) . Similarly for (3).
IV.21, `. 4–6.
“This terminology extends at once to the case that the functions f n and
the function f are only defined almost everywhere (or have values in R ,
and are defined and finite almost everywhere).”
When the functions fn , f have values in the Banach space F , there
exists a negligible set N on whose complement the f n and f are defined;
in particular (for a pair of such functions) linear operations on the classes of
INT IV.x10 extension of a measure. l p spaces §3
such functions are readily defined, and the notation N p (fn − f ) is available
(see the preceding Note).
When the functions take their values in R and are, moreover, finite
almost everywhere, then there exists a negligible set N on whose comple-
ment the functions are defined and finite-valued; one then proceeds as in
the preceding paragraph. Note in this connection that if f is a function de-
fined almost everywhere in X and taking values in R , and if N p (f ) < +∞
(where Np (f ) is defined as in the preceding Note), then f is finite-valued
almost everywhere (§2, No. 3, Prop. 7).
IV.21, `. 7–9.
“ . . . the closure of 0 in this space is the linear subspace N F of negligible
mappings of X into F (No. 2, Prop. 3).”
The closure in question is the set {f ∈ F Fp : Np (f ) = 0 } (TVS, II,
§1, No. 2, Prop. 2), and Np (f ) = 0 if and only if f is negligible (§2, No. 1,
Def. 1). It follows that NF is contained in each of the spaces LFp to be
defined in No. 4 (Def. 2).
IV.21, `. −13, −12.
“From this, it follows that Np (f − f0 ) 6 ε Np (h) , whence the proposi-
tion.”
Since, for every ε > 0 , there exists an M ∈ B such that
whence
If f has the indicated property, then for any such g one clearly has
f − g ∈ FFp , hence f = (f − g) + g ∈ FFp ; thus f belongs to the closure of
K (X; F) in FFp , that is, f ∈ LFp . The converse is immediate from Def. 2.
IV.24, `. 15, 16.
“For, f is finite almost everywhere and N p (f − g) 6 Np (h − g) 6 ε ;
Prop. 7 therefore shows that f is p -th power integrable.”
Since the functions g, h of the statement are finite almost everywhere,
so is f ; redefining f on a negligible set, we can suppose that it is everywhere
defined and finite. Then, given ε > 0 , choose g and h as in the statement
of the proposition; replacing them by equivalent functions, we can suppose
that g, h ∈ L p . Then all three functions are everywhere defined, finite, and
satisfy g 6 f 6 h almost everywhere in X ; and since 0 6 f − g 6 h − g
almost everywhere, Np (f − g) 6 Np (h − g) 6 ε . Choose g0 ∈ K (X; R)
with Np (g − g0 ) 6 ε (Def. 2); then
Np (f − g0 ) = Np (f − g) + (g − g0 ) 6 Np (f − g) + Np (g − g0 ) 6 ε + ε ,
By 1◦ and the cited Prop. 6, the series with general term f nk+1 (x) −
fnk (x) is absolutely convergent for almost every x in X , and the function
g : X → F defined by
X
g(x) = [fnk+1 (x) − fnk (x)] = lim fnk (x) − fn1 (x)
k
k
at the points x where the series converges, and by g(x) = 0 at the remaining
points x , belongs to FFp , and, as k → ∞ ,
k−1
X
[fnj+1 − fnj ] = fnk − fn1 → g in FFp ,
j=1
that is, in mean of order p ; g ∈ LFp because the fnk − fn1 belong to LFp
and LFp is closed in FFp . Thus, setting f = g + fn1 , one has f ∈ LFp ,
fnk (x) → f (x) for almost every x in X , and f nk → fn1 + g = f in LFp ,
that is, in mean of order p . Since (fn ) is Cauchy in LFp and fnk → f
in LFp , it follows that fn → f in LFp (an easy application of the triangle
INT IV.x12 extension of a measure. l p spaces §3
inequality; for the abstract principle, see GT, II, §3, No. 2, Cor. 3 of Prop. 5).
Whence 3◦ .
IV.24, `. −2, −1.
“ . . . there exists a lower semi-continuous function g > h + |f n1 | such
that Np (g) < +∞ , which completes the proof.”
P
By definition, h(x) = |fnk+1 (x) − fnk (x)| for all x ∈ X , thus h =
P k
|fnk+1 − fnk | is the sum of the positive numerical functions |f nk+1 − fnk | .
k
By Th. 1 of No. 2,
X X
Np (h) 6 Np (|fnk+1 − fnk |) = Np (fnk+1 − fnk ) < +∞ ,
k k
k−1
X
fnj+1 (x) − fnj (x) = fnk (x) − fn1 (x) ,
j=1
P
k−1
thus fnk (x) = fn1 (x) + fnj+1 (x) − fnj (x) , therefore
j=1
k−1
X
|fnk (x)| 6 |fn1 (x)| + fnj+1 (x) − fnj (x) 6 |fn1 (x)| + h(x) ,
j=1
Np (|fn1 | + h) < +∞ ,
whence |µ|* (|fn1 | + h)p < +∞ . By the definition of |µ|* (§1, No. 3,
Def. 3) there exists a lower semi-continuous function u > 0 such that
Since t 7→ t1/p is continuous and increasing (GT, IV, §3, No. 3, with the
convention (+∞)1/p = +∞ ) it is easy to see that u1/p is also lower semi-
continuous (more generally, see GT, IV, §6, Exer. 4). Writing g = u 1/p , we
have |fn1 | + h 6 g and
1/p 1/p
Np (g) = |µ|*(g p ) = |µ|*(u) < +∞ ,
§3 lp spaces INT IV.x13
IV.27, `. −8 to −6.
“ . . . the image of the section filter F of H under this mapping is there-
fore a base of a Cauchy filter on R .”
The set of all ke f kp ( fe ∈ H ) has a finite supremum λ and is directed
upward by the relation 6 on H ; that is, the mapping Φ : H → R defined
by Φ(e f ) = ke f kp is increasing and has a finite supremum λ . The sets
g) = {e
H(e f ∈H: e
f >g
e} (e
g ∈ H)
are a base for the section filter F of the directed set H . The sets
Φ H(eg) = {ke f kp : e
f ∈ H, e
f >ge } (e g ∈ H)
form a base of a filter G on R (GT, I, §6, No. 6).
g k > λ − ε ; then
Given any ε > 0 , choose ge ∈ H so that ke
e
f ∈ H, e
f >g
e ⇒ gk 6 ke
λ − ε 6 ke fk 6 λ ,
thus Φ H(eg ) ⊂ [λ − ε, λ] ⊂ [λ − ε, λ + ε] . This shows that the filter G is
convergent (to λ ), hence is Cauchy.
IV.28, `. 10–12.
“ . . . every function f > 0 in L p is the limit (for convergence in mean
of order p ) of a sequence of continuous functions > 0 with compact support,
by the continuity of the mapping g 7→ |g| on L p (Prop. 11).”
By the definition of L p , there exists a sequence (fn ) in K such that
fn → f (in mean of order p ). Since |fn | → |f | = f by the cited Prop. 11,
one can suppose without loss of generality that f n > 0 for all n . Similarly,
since g − f ∈ L p and g − f > 0 , there exists a sequence (h n ) in K with
hn > 0 and hn → g − f . Then
fn + hn → f + (g − f ) = g ;
setting gn = fn + hn , we have 0 6 fn 6 gn , where fn , gn ∈ K+ and
fn → f , gn → g (in mean of order p ). By the special case considered
earlier,
p p p
Np (gn − fn ) 6 Np (gn ) − Np (fn ) for all n ,
and passage to the limit yields (9).
IV.31, `. 11, 12.
“ . . . limF Np (f − fα ) exists and is equal to the common limit 0 of all
of the sequences (Np (f − fαn )) .”
Explicitly: Given ε > 0 , we seek an index n such that N p (f − fα ) < ε
for all α ∈ An . Assume to the contrary that for every n there exists an
αn ∈ An such that Np (f − fαn ) > ε . This contradicts the fact, already
shown, that lim Np (f − fαn ) = 0 .
n→∞
§4 integrable functions and sets INT IV.x15
IV.33, `. 13.
P
“ . . . we have |µ|*(|f |) = |α(x)| · |f (x)| < +∞ (§1, No. 3, Example)”
x∈X
R∗
by the definition of |µ|(f ) , thus |µ|(f ) = lim N1 (fn ) = N1 (f ) = f d |µ| .
n
IV.34, `. 15–17.
“Corollary 1.”
The positive function f = |f | belongs to L 1 by §3, No. 5, Prop. 11,
and N1 (f ) = N1 (f ) .
INT IV.x16 extension of a measure. l p spaces §4
IV.35, `. 3, 4.
“This follows at once from the inequality (1) by passage to the limit, on
taking into account (3) and the continuity of N 1 (f ) on LF1 .”
By the definition of “ f integrable”, there exists a sequence f n ∈ KF
such that
(*) N1 (fn − f ) → 0 ,
By (1),
Z Z
(iii) fn dµ 6 |fn | d|µ| = N1 (fn ) ;
are continuous mappings of LF1 into the Hausdorff space G ; since they agree
on the dense subspace KF (cf. Ch. III, §3, No. 2, Prop. 2 and No. 3, Cor. 2
of Prop. 7), they are identical by the “Principle of extension of identities”
(GT, I, §8, No. 1, Cor. 1 of Prop. 2).
IV.35, `. −1.
Z n
! n Z
X X
(8) ak fk dµ = ak fk dµ .
k=1 k=1
example, if f1 (x) = −∞ then f1− (x) = sup(−f1 (x), 0) = +∞ and the sum
f1 (x)+f1− (x) is not defined. (This problem did not arise in §3, No. 6, Cor. 2
of Th. 5, because the functions fn were assumed to belong to L p .)
The cure: If the fn are replaced by equivalent functions f n0 in L 1 then
the upper envelope f 0 of the fn0 is equivalent to f ; f 0 may admit +∞ as
a value, but there is no problem with g 0 = f 0 + f10 − since f10 is finite-valued.
To assure that the sequence (fn0 ) is increasing, let An be a negligible
S set on
whose complement fn0 = fn , and redefine the fn0 to be 0 on An . Finally,
n
gn0 = fn0 + f10 − > f10 + f10 − = f10 + , thus the functions gn0 belong to L 1 and
are > 0 . The proof of Prop. 4 carries through for the f n0 as indicated (see
the next note), and the validity of Prop. 4 for the f n follows at once from
the function equivalences.
IV.36, `. 15, 16.
“ . . . the proposition follows from Th. 5 of §3, No. 6.”
As observed in the preceding note, we can suppose that the f n are
finite-valued, that is, belong to L 1 . Since (recall that fn + f1− > 0 ) the
norms
Z Z Z
− −
N1 (fn + f1 ) = (fn + f1 ) d|µ| = fn d|µ| + f1− d|µ|
have a finite upper bound, it follows from the cited Th. 5 that the function
g = f + f1− is integrable, therefore f = g − f1− is integrable. Th. 5 also
yields
N1 (f + f1− ) = sup N1 (fn + f1− ) ,
n
that is, Z Z
(f + f1− ) d|µ| = sup (fn + f1− ) d|µ| ,
n
whence Z Z
f d|µ| = sup fn d|µ| ,
n
R R R
thus fn d|µ| → f d|µ| , therefore (f − fn ) d|µ| → 0 ; but f − fn > 0
and, citing Prop. 2 of No. 2, one has
Z Z Z
(f − fn ) dµ 6 |f − fn | d|µ| = (f − fn ) d|µ| → 0 ,
R R
whence fn dµ → f dµ .
§4 integrable functions and sets INT IV.x19
IV.36, `. −8.
“The theorem follows from Lebesgue’s theorem (§3, No. 7, Cor. of
Th. 6) . . .”
Here the functions fα belong to LF1 ( F a Banach space), and f : X → F;
one applies the cited Cor. with p = 1 , to conclude
R that Rf ∈ L F1 and that
fα → f in mean with respect to F , whence fα dµ → f dµ in F with
respect to F by the continuity of the integral.
IV.36, `. −5 to IV.37, `. 5.
“Corollary 1.”
Here Ω will play the role of A in Th. 2, and the role of F will be
played by the neighborhood filter of t 0 in Ω (which, by hypothesis, has a
countable base). For each t ∈ Ω define f t : X → F by the formula
ft (x) = f (x, t) (x ∈ X) .
in other words, the family (ft )t∈Ω converges pointwise in X to ft0 with
respect to the filter F .
By the hypothesis c), for each t ∈ U we have
in other words
Z Z
f (x, t) dµ(x) → f (x, t0 ) dµ(x) in F as t → t0 ,
IV.37, `. 6–13.
“Corollary 2.”
The points x ∈ X for which the series with general term f n (x) does not
converge form a negligible set; redefining all of the f n to be 0 at such x , we
can suppose that f is defined everywhere in X . For each positive integer n ,
define
Xn
gn = fk ;
k=1
as claimed.
IV.38, `. 8.
“We may limit ourselves to the case of lower semi-continuous functions”
For, if H is a set of upper semi-continuous functions directed for > ,
then −H = {−f : f ∈ H } is a set of lower semi-continuous directed for 6 .
However, the case of lower semi-continuous functions of arbitrary sign,
directed for 6 , will be reduced to the case of Cor. 2 for sets of functions
> 0 , thus will entail the consideration of sets H of both types: lower semi-
continuous and directed for 6 as well as upper semi-continuous and directed
for > .
IV.38, `. 10–12.
“ . . . the upper (resp. lower) envelope of the f + (resp. f − ), for f ∈ H ,
is equal to g + (resp. g − ).”
Let h = sup f + . Since f 6 g implies f + = sup(f, 0) 6 sup(g, 0) = g + ,
f ∈H
it is obvious that h 6 g + . To prove that g + 6 h , let x ∈ X . Since g(x) =
sup f (x) , there exists a sequence fn ∈ H such that g(x) = sup fn (x) ,
f ∈H n
and since H is directed for 6 we can suppose that the sequence (f n )
is increasing. If g(x) = +∞ then fn (x) > 0 from some index onward,
therefore
g + (x) = +∞ = sup fn+ (x) 6 sup f + (x) = h(x) ;
n f ∈H
§4 integrable functions and sets INT IV.x21
if g(x) < +∞ then fn (x) → g(x) in R , whence fn+ (x) → g + (x) , so that
g + (x) = sup fn+ (x) 6 h(x) .
n
IV.38, `. 14. R R R
“ . . . then f + d|µ| 6 f d|µ| + f0− d|µ| ”
If f > f0 then f − = sup(−f, 0) 6 sup(−f0 , 0) = f0− ; since f =
f − f − , and f and f0 are finite almost everywhere, one has, for almost
+
every x ,
f + (x) = f (x) + f − (x) 6 f (x) + f0− (x) ,
R
whence the asserted inequality (which shows that sup f + d|µ| < +∞ ).
f ∈H
H+ = {f + : f ∈ H } (resp. H− = {f − : f ∈ H } )
consists of lower (resp. upper) semi-continuous functions (GT, IV, §6, No. 2,
Prop. 2 and its analog for upper semi-continuous functions) and is directed
for 6 (resp. > ). Moreover, from f > f0 one infers that, for almost every x ,
whence Z Z Z
f d|µ| > −
f0+ d|µ| − f d|µ| ,
R
which shows that inf f − d|µ| > −∞ .
f ∈H
Thus, the validity of both parts of Cor. 2 for functions > 0 will imply
that Z Z Z Z
+ +
g dµ = lim f dµ and g dµ = lim f − dµ ,
−
f ∈H f ∈H
as well as
Z Z Z Z
g + d|µ| = sup f + d|µ| and g − d|µ| = inf f − d|µ| ,
f ∈H f ∈H
whence
Z Z Z Z Z
+ − + −
g dµ = g dµ − g dµ = lim (f − f ) dµ = lim f dµ ,
f ∈H f ∈H
INT IV.x22 extension of a measure. l p spaces §4
as well as
Z Z Z Z
g d|µ| = (g + − g − ) d|µ| = g + d|µ| − g − d|µ|
Z Z
+
= sup f d|µ| − inf f − d|µ|
f ∈H f ∈H
Z + Z
+
= sup f d|µ| + sup (−f − ) d|µ|
f ∈H f ∈H
Z + Z
= lim +
f d|µ| + lim (−f − ) d|µ|
f ∈H f ∈H
Z Z
= lim f + d|µ| + (−f − ) d|µ|
f ∈H
Z Z
= lim f d|µ| = sup f d|µ| .
f ∈H f ∈H
To summarize: The validity of Cor. 2 for functions > 0 implies the validity
of the ‘lower semi-continuous half’ of Cor. 2 for functions of arbitrary sign
(hence the validity of the corollary as stated).
IV.38, `. −9 to −7.
“If H is directed for > and consists of upper semi-continuous integrable
functions f such that 0 6 f 6 f1 with f1 ∈ H , then there exists a lower
semi-continuous integrable function h such that f 1 6 h ”
Given a set H Rof upper semi-continuous functions > 0 , directed for >
and such that inf f d|µ| > −∞ , one chooses any f1 ∈ H ; then the lower
f ∈H
envelope g of H is equal to the lower envelope of the cofinal set
H1 = {f ∈ H : 0 6 f 6 f1 } ;
also, Z Z
inf f d|µ| = inf f d|µ| > −∞ ,
f ∈H1 f ∈H
briefly f 0 > k on V .
IV.38, `. −2.
“ . . . we can apply to them what has been proved above . . . ”
“them” is the set H0 = {f 0 : f ∈ H } , consisting of lower semi-
continuous integrable functions > 0 , and H 0 is directed for the relation 6 .
Recall that the lower envelope of H is denoted g in the statement of Cor. 2.
Let us write gb for the upper envelope of H 0 . {The text uses the notation g 0 ,
but this is confusing since g 0 is not derived from g in the same way that
f 0 is derived from f ; more R about this later.}
As just shown, sup f 0 d|µ| < −∞ , thus, by the case of Cor. 2 for
f ∈H
positive lower semi-continuous functions, we know that gb is integrable and
that Z Z Z Z
0
gb dµ = lim f dµ and gb d|µ| = sup f 0 d|µ| .
f ∈H f ∈H
0 6 f0 6 h for all f ∈ H ;
A = {x ∈ X : h(x) = +∞ }
R∗ R R∗
but h d|µ| = h d|µ| < +∞ (No. 2, Prop. 1), thus g d|µ| < +∞ ,
hence g is integrable by Cor. 1 Rof Prop. 5. By hypothesis, given any ε > 0
there exist such g and h with (h − g) d|µ| 6 ε , that is,
Z ∗ Z
N1 (h − g) = |h − g| d|µ| = (h − g) d|µ| 6 ε ,
0 6 h − g = (u + v) − (u − v)+ 6 (u + v) − (u − v) = 2v ,
whence
ε
N1 (h − g) 6 2 N1 (v) 6 2 · = ε,
2
R
that is, (h − g) d|µ| 6 ε as desired.
IV.40, `. 2, 3.
“ . . . since g 6 f , g is integrable by Prop. 4 of No. 3”
The
R gn are integrable
R (§3, No. 5, Cor. of Prop. 12) and g n ↑ g 6 f ;
since gn d|µ| 6 f d|µ| < +∞ R for all n , itR follows from Prop. 4 of No. 3
that g is integrable and that g d|µ| = lim gn d|µ| .
n
IV.40, `. −15.
“step function”
The French original is “fonction en escalier ” (literally “function in
steps”), signifying a (finite) linear combination of characteristic functions
of subintervals of R (FVR, II, §1, No. 3). Later in the section (No. 9,
Def. 4) a more general concept is introduced, signifying a function X → F
§4 integrable functions and sets INT IV.x27
P
( F a Banach space) that is a finite sum ai ϕMi , where the ai belong to F
i
and the Mi belong to a given clan Φ of subsets of X —so to speak, a linear
combination of characteristic functions of sets of Φ , with ‘coefficients’ a i
in F ; this concept is translated as “step function with respect to Φ ”, or
“ Φ-step function” (in the original, “fonction Φ-étagée” or “fonction étagée
sur les ensembles de Φ ”). A fuller discussion of how this double usage of
“step function” came about is given in the note for IV.66, `. −15, −14; for
the present, suffice it to note that “step function” in the sense of FRV is
intended also at the following places in Integration: the footnote on p. xi of
Vol. I, and, in the present chapter, the hints for Exers. 29 d) and 30 for §5,
and Exer. 17 for §6.
IV.40, `. −12 to −8.
“It follows that if f is a regulated function on R with compact support
(FRV, II, §1, No. 3), then f is integrable, because it is the uniform limit of
a sequence of step functions gn Rwith support Rcontained in a fixed compact
set (No. 3, Prop. 3); moreover, f dµ = lim gn dµ .”
n→∞
From the boldface notation and the references, we can suppose that
f : R → F , F a Banach space.
Let [a, b] be a compact interval in R such that Supp f ⊂ [a, b] and let
gn : R → F be a sequence of step functions such that g n → f uniformly
in [a, b] (FRV, II, §1, No. 3, Def. 3). Multiplying g n by the characteristic
function of [a, b] , one can suppose that
R Supp g n ⊂ [a,
R b] for all n ; then by
Prop. 3 of No. 3, f is integrable and f dµ = lim gn dµ .
n→∞
IV.40, `. −6, −5. R R +∞
“ . . . the integral f dµ is equal to the integral −∞ f (x) dx defined
in FRV, II, §2, No. 1.”
The following remark will be useful in the proof: R +∞ If f : R → F is a
regulated function ( F a Banach space) such that −∞ f (x) dx exists, then,
R +∞
for every closed interval [a, b] of R , −∞ (ϕ[a,b] f )(x) dx exists and
Z +∞ Z b
(ϕ[a,b] f )(x) dx = f (x) dx .
−∞ a
and
F00 (x) = f0 (x) = f (x) ,
therefore G0 and F0 differ by a constant vector (FRV, II, §1, Prop. 1) and
Z b
(ϕ[a,b] f )(x) dx = G(b) − G(a) = G0 (b) − G0 (a)
a
Z b
= F0 (b) − F0 (a) = F(b) − F(a) = f (x) dx .
a
the first and third terms of the right member are 0 because ϕ [a,b] f is equal
to 0 on ]r, a[ and on ]b, s[ (then consider its restrictions to [r, a] and
[b, s] as in the foregoing discussion to argue that the zero function serves
as primitives for these restrictions), whereas the middle term is equal to
Rb
a
f (x) dx , thus (†) is verified.
Now let K be the set of all compact intervals [r, s] of R , which is
directed for ⊂ . For every [r, s] ∈ K satisfying [r, s] ⊃ [a, b] , the equa-
lity (†) holds; since such intervals [r, s] are cofinal in K , it follows that the
limit Z s
lim [r,s]∈K (ϕ[a,b] f )(x) dx
r
§4 integrable functions and sets INT IV.x29
Rb
exists and is equal to a
f (x) dx , consequently
Z +∞ Z b
(ϕ[a,b] f )(x) dx = f (x) dx
−∞ a
by the definition of the left member (FRV, II, §2, No. 1); if, moreover,
Supp f ⊂ [a, b] , then ϕ[a,b] f = f and we have the following:
Z +∞ Z b
(††) f (x) dx = f (x) dx when Supp f ⊂ [a, b] .
−∞ a
R +∞
But also −∞ f (x) dx = (d − c)a . For, if F : R → R is the continuous
piecewise linear function
0 for x 6 c
F (x) = x−c for c < x < d
d−c for x > d
and so
Z Z +∞
(*) f dµ = f (x) dx
−∞
INT IV.x30 extension of a measure. l p spaces §4
for all such f (FRV, II, §2, No. 1, formula (1)). It follows by linearity that
(*) holds for every step function with compact support.
Now let f : R → F be a regulated function with compact support and
let (gn ) be a sequence of step functions with support contained in a fixed
compact interval [c, d] , such that g n → f uniformly in R (see the preceding
note). By (*),
Z Z +∞
(**) gn dµ = gn (x) dx for all n ;
−∞
R R
we know that f dµ = lim gn dµ , and it follows from FRV, II, §3, No. 1,
n→∞
Cor. 1 of Prop. 1 that
Z +∞ Z d Z d Z +∞
f (x) dx = f (x) dx = lim gn (x) dx = lim gn (x) dx
−∞ c n→∞ c n→∞ −∞
(the first and last equalities by (††)), so passage to the limit in (**) shows
that (*) holds for every regulated function f with compact support.
IV.40, `. −1 to IV.41, R `. 1. Rn
“ . . . therefore |f | dµ = lim −n |f (x)| dx by Th. 2 of No. 3”
R n→∞ R
By the cited Th. 2, |f | dµ = lim ϕIn |f | dµ . Since ϕIn |f | is regu-
n→∞
lated and has compact support,
Z Z +∞
(*) ϕIn |f | dµ = (ϕIn |f |)(x) dx
−∞
next that
Z s Z n
(ϕIn |f |)(x) dx = |f |(x) dx when [r, s] ⊃ [ − n, n] ,
r −n
exists, and Z Z
+∞ n
(ϕIn |f |(x) dx = |f |(x) dx ;
−∞ −n
§4 integrable functions and sets INT IV.x31
therefore by (*), Z Z n
ϕIn |f | dµ = |f |(x) dx ,
−n
by the theorem of the mean (FRV, I, §1, No. 5, Prop. 6), and it follows that
Rb Rs
for intervals [a, b] ⊂ [r, s] one has a f (x) dx 6 r f (x) dx (FRV, II, §1,
No. 5, formula (6)).
In particular, if F is a primitive for the regulated function |f | then F
is increasing. Thus if K is the set of compact intervals [r, s] of R , directed
by ⊂ , then the function
Z s
[r, s] 7→ |f |(x) dx
r
in other words,
Z s Z
lim |f |(x) dx = |f | dµ < +∞ .
r→−∞, s→+∞ r
Thus, in the terminology following FRV, II, §2, No. 1, Def. 1, the integral
of |f | over the interval (−∞, +∞) is convergent and
Z +∞ Z
|f |(x) dx = |f | dµ ;
−∞
INT IV.x32 extension of a measure. l p spaces §4
that is, in the language of FRV, II, §2, No. 3, Def. 2, “the integral of f is ab-
R +∞
solutely convergent”. It then follows from loc. cit., Prop. 4, that −∞ f (x) dx
exists; more precisely, see the next note.
IV.41, `. 2. R R +∞
“Moreover, f dµ = −∞ f (x) dx by Th. 2 of No. 3.”
R +∞
As remarked at the end of the preceding note, −∞ f (x) dx exists (by
FRV, II, §2, No. 3, Prop. 4). Thus, writing K for the set of compact intervals
[a, b] of R , directed by ⊂ ,
Z +∞ Z b
f (x) dx = lim f (x) dx .
−∞ [a,b]∈K a
On the other hand, by the cited Th. 2 (with |f | playing the role of g ; the
intervals [ − n, n] form a countable base for K ),
Z Z
f dµ = lim ϕ[a,b] f dµ .
[a,b]∈K
by the preceding Example, and, as shown at the beginning of the note for
IV.40, `. −6, −5,
Z +∞ Z b
(ϕ[a,b] f )(x) dx = f (x) dx ,
−∞ a
therefore
Z Z Z b Z +∞
f dµ = lim ϕ[a,b] f dµ = lim f (x) dx = f (x) dx .
[a,b]∈K [a,b]∈K a −∞
IV.41, `. 2–4. R +∞
“Conversely, suppose that −∞ f (x) dx is absolutely convergent; again,
R R +∞
by Th. 2 of No. 3, f dµ = −∞ f (x) dx .”
R +∞
Assuming f is a regulated function such that −∞ |f |(x) dx is conver-
Rn
gent (i.e., exists; equivalently, the integrals −n |f |(x) dx are bounded), the
problem is to show that f is µ-integrable. (We know from FRV, II, §2, No. 3,
R +∞
Prop. 4 that −∞ f (x) dx exists, but no use of this fact will be made.)
§4 integrable functions and sets INT IV.x33
R∗
We will show that |f | dµ < +∞ , so that the function |f | can play
the role of g in the cited Th. 2. The numerical function g = |f | is regulated
and > 0 . Let In = [ − n, n] (n = 1, 2, 3, . . .). For every n , the function
ϕIn g is regulated, positive, and has compact support; by the above Example,
ϕIn g is µ-integrable and
Z Z +∞
ϕIn g dµ = (ϕIn g)(x) dx .
−∞
R +∞
Since −∞
g(x) dx exists by hypothesis, we have
Z +∞ Z n
(ϕIn g)(x) dx = g(x) dx
−∞ −n
by
R the remarkR n at the beginning of the note for IV.40, `. −6, −5; thus
ϕIn g dµ = −n g(x) dx , and
Z Z n Z +∞
µ*(ϕIn g) = ϕIn g dµ = g(x) dx 6 g(x) dx < +∞ for all n
−n −∞
(the first equality by No. 2, Prop. 1); since ϕ In g ↑ g pointwise, by §1, No. 3,
Th. 3 one has
Z +∞
µ*(g) = sup µ*(ϕIn g) 6 g(x) dx < +∞ ,
n −∞
qualifying g for its role in the cited Th. 2: each ϕ In f is µ-integrable (it is reg-
ulated and has compact support), |ϕ In f | = ϕIn gR 6 g for all Rn, and ϕIn f → f
pointwise in R , hence f is µ-integrable (and f dµ = lim ϕIn f dµ ). The
n
R R +∞
formula f dµ = −∞ f (x) dx then follows from the preceding note.
IV.41, `. −10 to −8.
“ . . . for a set to be negligible, it is necessary and sufficient that it be of
measure zero with respect to |µ| .”
Every negligible function is integrable; for, the set N of negligible
functions (the functions f such that N 1 (f ) = 0 , equivalently Np (f ) = 0 for
1 6 p < +∞ ) is the closure of {0} in the topological vector space F p (§3,
No. 3), consequently N ⊂ L p for 1 6 p < +∞ (loc. cit., No. 4, Def. 2).
IV.42, `. 4, 5.
“1◦ If A and B are two integrable sets such that B ⊂ A , then the set
C = A --- B is integrable . . . ”
INT IV.x34 extension of a measure. l p spaces §4
IV.42, `. 10.
“ ϕA = inf ϕAn , therefore A is integrable (No. 3, Prop. 4).”
n
IV.44, `. −11.
“ f 6 ϕK + δϕB ”
For f (x) = 0 , the inequality is trivial. Suppose f (x) > 0 . Then
ϕA > f yields ϕA (x) > 0 , so ϕA (x) = 1 , x ∈ A ; if x ∈ / K (that is,
f (x) < δ ) then x ∈ A --- K = B and the desired inequality reduces to
f (x) 6 0 + δ · 1 , whereas if x ∈ K then ϕK (x) = 1 and the desired
inequality results from f (x) 6 ϕA (x) = 1 = ϕK (x) .
IV.45, `. 1–3.
“The condition is sufficient, because it says that, for the topology of
convergence in mean, ϕA is in the closure of the set of integrable func-
tions ϕK ( K an arbitrary compact subset of A ).”
Observe first that ϕA ∈ F 1 : for ε = 1 choose a compact set K1 ⊂ A
with |µ|*(A --- K1 ) 6 1 ; then
|µ|*(A) = |µ|* (A --- K1 ) ∪ K1
6 |µ|*(A --- K1 ) + |µ|*(K1 ) 6 1 + |µ|*(K1 ) < +∞ ,
thus ϕA ∈ F 1 . The sufficiency of the condition then follows from the fact
that L 1 is closed in F 1 (§3, No. 4, Def. 2).
The set F of all compact sets K ⊂ A is directed for ⊂ , and from
|µ|(A) − |µ|(K) = |µ|(A --- K) and the monotonicity of |µ| , one infers that
and then (No. 5, Prop. 7) |µ(A) − µ(K)| = |µ(A --- K)| 6 |µ|(A --- K) yields
µ(A) = lim µ(K) .
K,F
f = ϕ Kf f 6 ϕ Kf · 1 6 ϕ U ;
whence
|µ|*(U) = sup |µ|(Kf ) .
f ∈H
IV.46, `. 7, 8.
“For, we have seen that |µ|*(X) = kµk (§1, No. 2); the proposition
therefore follows from Prop. 10 of No. 6.”
By the remarks following §1, No. 2, Def. 2, |µ|*(X) = k |µ| k , and
k |µ| k = kµk (Ch. III, §1, No. 8, Cor. 1 of Prop. 10); µ is bounded if
and only if kµk < +∞ (loc. cit., III.16, `. −11, −10), hence if and only if
|µ|*(X) < +∞ , that is (No. 6, Prop. 10), X is µ-integrable. This Rmeans, in
turn, that ϕX = 1 is µ-integrableR (No. 5, Def. 2), in which case 1 d|µ| =
|µ|*(1) (No. 2, Prop. 1), that is, d|µ| = kµk , and one writes
Z Z
|µ|(X) = ϕX d|µ| = d|µ|
IV.46, `. −1 to IV.47, `. 3.
“It is clear that the sets MK form a filter base B on LFp , that the
functions belonging to MK are uniformly bounded, and that B converges
uniformly to f on every compact subset of X , whence the corollary.”
The functions hf belong to KF , hence to LFp (§3, No. 4, Def. 2), thus
MK ⊂ LFp for every compact set K ⊂ X . If K1 , K2 are compact and K is
a compact set such that K ⊃ K1 ∪ K2 , then MK ⊂ MK1 ∩ MK2 (because
h = 1 on K ⇒ h = 1 on K1 and h = 1 on K2 ), so that the MK form
a filter base on LFp . The uniform boundedness of the hf is clear from the
boundedness of f and the fact that khk = 1 .
Given any compact K0 ⊂ X , we are to show that B → f uniformly
on K0 (it will then follow from Prop. 13 that f ∈ L Fp and that B → f in
mean of order p ). Given any ε > 0 , suppose K ⊃ K 0 , K compact, and let
hf ∈ MK , where h ∈ K , 0 6 h 6 1 and h = 1 on K ; then
thus sup |(hf − f )(x)| = 0 < ε and the claimed uniform convergence on K 0
x∈K0
holds trivially.
IV.47,R `. 4, 5. R
“ f dµ is the limit with respect to B of the integrals hf dµ .”
The integral is (by definition) a continuous linear mapping L F1 → F
(No. 1, Def. 1).
IV.47, `. 13–17. R
“ the mapping f 7→ f dµ is continuous on the Banach space C b (X; F) ;
its restriction to the closure C 0 (X; F) of K (X; F) in C b (X; F) , that is,
to the space of continuous functions tending to 0 at the point at infinity
§4 integrable functions and sets INT IV.x39
(Ch. III, §1, No. 2, Prop. 3), is therefore the extension by continuity of the
integral to C 0 (X; F) .”
The notations C b (X; F) and C 0 (X; F) are established in Ch. III, §1,
No. 2. Measures are defined in the following subsection (loc. cit., No. 3), the
set M (X; C) of all (complex) measures µ being defined to be the vector
space dual to the topological vector space K (X; C) (equipped with the
direct limit topology).
Bounded measures µ are defined loc. cit., No. 8, the definition of µ(f )
being limited to f ∈ K (X; C) ; it is observed that the linear subspace
M 1 (X; C) of M (X; C) consisting of the bounded measures µ may be re-
garded as the dual of the vector space K (X; C) equipped with the topol-
ogy defined by the norm kf k = sup |f (x)| , but the domain of µ remains
x∈X
K (X; C) . The possibility of extending µ to C b (X; C) , and in particu-
lar to C 0 (X; C) , is not taken up, but it is is latent in the observation
that if g ∈ C b (X; C) then g · µ is a bounded measure (ibid., Prop. 12),
whence ϕX = 1 is (g · µ)-integrable
R (Prop. 12 of the present No. 7), invit-
ing the definition µ(g) = 1 d(g · µ) (No. 1, Def. 1). This is accomplished
in the present No., more generally for f ∈ C b (X; F) (Cor. of Prop. 13):
C b (X; F) ⊂ LFp (X, µ) and, by the inequality (20), the mapping f 7→ µ(f ) so
defined is a linear mapping C b (X; F) → F continuous for the norm topology
on C b (X; F) . R
Write L(f ) = f dµ ( µ a bounded measure, f ∈ C b (X; F) ), which is
a linear mapping C b (X; F) → F continuous for the norm topology, and let
L0 = LC 0 (X; F) be the restriction of L to C 0 (X; F) . One has
thus X is integrable (No. 6, Prop. 10) and so µ is bounded (No. 7, Prop. 12),
as already observed in Ch. III, §2, No. 3, Prop. 11.
INT IV.x40 extension of a measure. l p spaces §4
where the first inequality holds by the inequality (5) of No. 2, Prop. 5 (cited
also in the proof of (20), which, strictly speaking, is stated only for contin-
uous bounded functions).
IV.47, `. −3 to −1.
“In particular, if f is continuous on X then f is µ-integrable, since
f h ∈ K (X; F) for every function h ∈ K (X; R) equal to 1 on S (Ch. III,
§1, No. 2, Lemma 1).”
Let h be any such function (which exists by the cited lemma). Then
f −f h = 0 on S ; since { S is µ-negligible (§2, No. 2, Prop. 5), it follows that
|f − f h| = 0 almost everywhere, consequently N 1 (f − f h) = 0 (§2, No. 3,
Prop. 6). Thus, f −f h is negligible for µ (§2, No. 1, Def. 1) hence belongs to
LF1 (X, µ) (see the note for IV.21, `. 7–9); but f h ∈ K (X; F) ⊂ L F1 (X, µ)
(§3, No. 4, Def. 2), therefore f = (f − f h) + f h ∈ L F1 (X, µ) . (The same
argument shows that f ∈ LFp (X, µ) for 1 6 p < +∞ .)
Thus, when µ has compact support, every f ∈ C (X; F) is µ-integrable;
since such a function is bounded on every compact subset of X , and in
particular on S = Supp(µ) , the inequality (22) holds for every f ∈ C (X; F) .
IV.48, `. 9–11. R
“. . .the mapping f 7→ f dµ of C (X; F) into F is continuous for the
topology of compact convergence.”
The locally convex topology in question is defined by the family of semi-
norms pK (f ) = sup |f (x)| , where K ⊂ X is compact (GT, X, §1, No. 3,
x∈K
Example III and No. 6; TVS, II, §1, No. 2 and §4, No. 1, Cor. of Prop. 1;
and the note forR III.39, `. 8–11). The inequality (22) shows that the linear
mapping f 7→ f dµ ( f ∈ C (X; F) ) is continuous for the topology defined
by the semi-norm pS ( S the support of µ ), hence for the (finer) topology
defined by all the pK .
§4 integrable functions and sets INT IV.x41
IV.48, `. 12–14.
“Then, there is a compact set K ⊂ X and a number a > 0 such that
|µ(f )| 6 a · sup |f (x)| for every function f ∈ K (X; F) ”
x∈K
R
It follows from the hypothesis that the formula q(f ) = f dµ defines a
semi-norm on K (X; F) continuous for the topology of compact convergence;
as that topology is defined by the family of semi-norms p K (f ) = sup |f (x)|
x∈K
with K ⊂ X compact (see the preceding note), at issue is the character-
ization of the continuous semi-norms on K (X; F) (and C (X; F) ) for the
topology. As the topology of a topological vector space can be defined by
a set of semi-norms if and only if it is locally convex (TVS, II, §4, No. 1,
Cor. of Prop. 1), it is of interest (and notationally simpler) to consider the
general case.
Let E be a vector space (over R or C ) and let p be a semi-norm
on E . The sets
V(p, α) = {z ∈ E : p(z) 6 α } (α > 0)
form a fundamental system of neighborhoods of 0 for a locally convex topol-
ogy τp on E , and τp is the coarsest topology on E that makes p continu-
ous and is compatible with the additive group structure of E (TVS, II, §1,
No. 2).
Lemma 1. For a semi-norm q on E , the following conditions are equiv-
alent:
(a) q is continuous for τp ;
(b) τq ⊂ τp ;
(c) there exists a constant M > 0 such that q 6 Mp on E .
Proof. (b) ⇒ (a): q is continuous for τ q , hence a fortiori for the finer
topology τp .
(a) ⇒ (b): The topology τp is compatible with the additive structure
of E and makes q continuous, hence τq is coarser than τp .
(c) ⇒ (a): For every α > 0 , clearly p(z) 6 α/M ⇒ q(z) 6 α , thus
{z : q(z) 6 α } ⊃ {z : p(z) 6 α/M } ;
since the right side is a neighborhood of 0 for τ p (by the continuity of p
for τp ) so is the left side, consequently q is continuous at 0 for τ p . It
then follows from |q(z) − q(z0 )| 6 q(z − z0 ) that q is (uniformly) continuous
for τp .
(a) ⇒ (c): Let D = {λ : |λ| 6 1 } , a neighborhood of 0 in the field of
scalars. By hypothesis, q is continuous at 0 for τ p , hence the set
−1
V(q, 1) = {z : q(z) 6 1 } = q (D)
INT IV.x42 extension of a measure. l p spaces §4
0
−1
) 6 1 , that is, α p(z) + ε
therefore q(z q(z) 6 1 , in other words q(z) 6
M p(z) + ε ; since ε is arbitrary, q(z) 6 Mp(z) .
Lemma 2. Let Γ be a set of semi-norms on E and let τ be the (locally
convex ) topology on E defined by Γ , that is, the coarsest topology τ on E
that is compatible with the additive structure of E and is such that every
p ∈ Γ is continuous (TVS, II, §1, No. 2).
Assume, moreover, that if p1 , . . . , pn belong to Γ then there exists a
p ∈ Γ such that p > sup pi (the upper envelope of p1 , . . . , pn ). Then, for
16i6n
a semi-norm q on E , the following conditions are equivalent:
(i) q is continuous for τ ;
(ii) there exist a p ∈ Γ and a constant M > 0 such that q 6 Mp .
Proof. Clearly τ is the coarsest topology, compatible with the additive
structure of E , for which τ ⊃ τp for all p ∈ Γ .
(ii) ⇒ (i): By Lemma 1, q is continuous for τ p , hence a fortiori for
the finer topology τ .
(i) ⇒ (ii): The sets Dα = {λ : |λ| 6 α } (α > 0) form a base for
the neighborhoods of 0 in the field of scalars. By the definition of τ ,
a neighborhood base at 0 ∈ E for τ is given by the finite intersections
−1 −1
p 1 (Dα1 ) ∩ · · · ∩ p n (Dαn ) ,
−1 −1
p 1 (Dα ) ∩ · · · ∩ p n (Dα ) ( α > 0 , pi ∈ Γ for 1 6 i 6 n )
−1 −1 −1
p (Dα ) ⊂ p 1 (Dα ) ∩ · · · ∩ p n (Dα ) ,
{z : q(z) 6 1 } ⊃ {z : p(z) 6 α } ,
S
n
If K1 , . . . , Kn are compact and K = Ki , then pK > pKi for i = 1, . . . , n ,
i=1
therefore pK > sup pKi ; in view of Lemma 2, if q is a semi-norm on the
16i6n
space, then q is continuous for the topology of compact convergence if and
only if q 6 a · pK for some a > 0 and compact set K , whence the original
assertion of the text.
IV.48, `. 16, 17.
“ µ(h) = 0 for every function h ∈ K (X; C) whose support does not
intersect K ”
Recall that µ(ha) = µ(h)a (Ch. III, §3, No. 4, Prop. 8).
IV.48, `. 17, 18.
“ . . . which proves that Supp(µ) ⊂ K .”
By the foregoing argument, µ(h) = 0 for every h ∈ K (X; C) such
that Supp(h) ⊂ { K ; this means that the measure induced by µ on the
open set { K is 0 (Ch. III, §2, No. 1), therefore { K ⊂ { Supp(µ) by the
definition of Supp(µ) (loc. cit., No. 2, Def. 1).
INT IV.x44 extension of a measure. l p spaces §4
IV.48, `. 22–24.
“The set of measures on X with compact support may therefore be
identified with the dual C 0 (X; C) of the Hausdorff locally convex space
C (X; C) .”
Write Mc for the set of all measures on X with compact support. Since
the support of the sum of two measures is the union of their supports (Ch. III,
§2, No. 2, Prop. 4), it is clear that M c is a linear subspace of the space
M (X; C) of all measures on X . According to Prop. 14, for each measure µ
with Rcompact support, every f ∈ C (X; C) is µ-integrable, the linear form
f 7→ f dµ is continuous for the topology τ cc of compact convergence on
C (X; C) (hence is an element of the dual space C 0 (X; C) ), and it is the
only element of C 0 (X; C) that extends the linear form µ : K (X; C) → C .
For each µ ∈ Mc let us write µ0 ∈ C 0 (X; C) for its unique τcc -
continuous extension to C (X; C) . The mapping M c → C 0 (X; C) defined
by µ 7→ µ0 is linear; for instance, if µ1 , µ2 ∈ Mc then µ01 + µ02 extends
µ1 + µ2 , therefore (µ1 + µ2 )0 = µ01 + µ02 . The mapping is injective; for, the
restriction of µ0 to K (X; C) is µ , so if µ0 = 0 then µ = 0 . Thus µ 7→ µ0
is an injective linear mapping Mc → C 0 (X; C) ; our problem is to show that
it is also surjective.
Thus, given ν ∈ C 0 (X; C) , we seek a measure µ on X with compact
support such that µ0 = ν . Let µ = ν K (X; C) and let us show that µ
is a measure on X with compact support. It suffices to show that µ is
a measure, for, by Prop. 14, compactness of its support is assured by the
existence of the extension ν ; and then µ 0 = ν , because ν extends µ .
Given any compact subset K of X , it suffices to show that there exists
a constant M > 0 such that
where kf k = sup |f (x)| (Ch. III, §1, No. 3). Now, the topology τ cc on
x∈X
C (X; C) is defined by the set of semi-norms p L (f ) = sup |f (x)| ( L a com-
x∈L
pact subset of X ), and the semi-norm q(f ) = |ν(f )| ( f ∈ C (X; C) ) is
continuous for τcc ; as shown in the note for IV.48, `. 9–10, there exist a
compact set L and a constant M > 0 such that q 6 M · p L on C (X; C) ,
hence on K (X; C) . Thus, for f ∈ K (X; C) , we have
{f : pK (f ) 6 ε } ( K ⊂ X compact, ε > 0 }
{f : pn (f ) 6 ε } ⊂ {f : pK (f ) 6 ε } ,
{f : pn (f ) 6 ε } ( ε > 0 , n = 1, 2, 3, . . .)
◦ ◦ ◦ ◦
thus the sets Kn form an open covering of K ; since K1 ⊂ K2 ⊂ K3 ⊂ · · ·
◦ ◦
and since K is covered by finitely many of the Kn , one has K ⊂ Kn for
some n , whence K ⊂ Kn and so pK 6 pKn = pn .
It follows that the sets
{f : pn (f ) 6 1/m } (m, n = 1, 2, 3, . . .)
L0 = {µ0 : µ ∈ L } ⊂ C 0 (X; C) .
−1
thus Φ U(S, ε) ∩ Mc is a neighborhood of 0 in C 0 (X; C) for T , whence
the asserted continuity of Φ .
IV.49, `. 19, 20.
“It is clear that the set H0 of functions f h , where f runs over H , is
strictly compact in K (X; C) ”
The formula Ψ(f ) = f h defines a linear mapping Ψ : C (X; C) →
C (X; C) , and Ψ(H) = H0 . Moreover, if G is the support of h , then Ψ
takes all of its values in K (X, G; C) and in particular H 0 ⊂ K (X, G; C) .
It will suffice to prove that Ψ is continuous (for the topology τ cc of compact
convergence); for, this will imply that H 0 = Ψ(H) is compact in C (X; C)
for τcc , whence H0 is compact in K (X, G; C) for the induced topology
§4 integrable functions and sets INT IV.x49
V(J, ε) = {f ∈ C (X; C) : pJ (f ) 6 ε } ;
in other words
−1
V(J, ε/khk) ⊂ {f ∈ C (X; C) : Ψ(f ) ∈ V(J, ε) } = Ψ V(J, ε) ;
−1
this shows that, for every J and ε , Ψ V(J, ε) is a neighborhood of 0 in
C (X; C) for τcc , whence the continuity of Ψ .
IV.49, `. 21.
“ for every measure µ ∈ L , µ(f ) = µ(f h) for every function f ∈ H ”
As f ∈ C (X; C) and f h ∈ K (X; C) , a review of the notations is in
order. In the expression µ(f h) , µ is the original linear form on K (X; C) .
As noted at the beginning of the No., since µ has compact support, f is
µ-integrable in the sense of No. 1, Def. 1 ( f − f h = 0 on K ⊃ Supp(µ) and
{ Supp(µ) is negligible (§2, No. 2, Prop.R 5), therefore f = f h µ-almost
everywhere) and, by Prop. 14, f 7→ f dµ ( f ∈ C (X; C) ) is the unique
linear form on C (X; C) that is continuous for the topology of compact
convergence and extends theR original linear form f 7→ µ(f ) ( f ∈ K (X; C) ).
One also writes µ(f ) for f dµ ( f ∈ C (X; C) ), so that µ(f ) = µ(f h)
(because f = f h µ-almost everywhere).
In the note for IV.48, `. 22–24, we have introduced the ephemeral nota-
tion µ0 for the linear form in C (X; C) so defined, as an aid to understanding
the proofs; thus µ0 (f ) = µ(f ) ( f ∈ C (X; C) ) by the definition of µ 0 , and
µ0 (f ) = µ(f ) = µ(f h) = µ0 (f h) .
IV.49, `. 21, 22.
“ . . . whence the conclusion.”
The linear bijection Φ : C 0 (X; C) → Mc , Φ(µ0 ) = µ ( µ ∈ Mc ) was
introduced in the note for IV.49, `. 17–18, and shown to be continuous
for the topologies T and T 0 ∩ Mc , whence the continuity of the linear
INT IV.x50 extension of a measure. l p spaces §4
bijection ΦL0 : L0 → L for the topologies T ∩ L0 and T 0 ∩ L ; the desired
conclusion is that ΦL0 is a homeomorphism, thus we must show that the
inverse mapping (ΦL0 )−1 : L → L0 is continuous for T 0 ∩ L and T ∩ L0 .
Let us abbreviate Θ = (ΦL0 )−1 .
For every compact subset H of C (X; C) and every ε > 0 , write
(TVS, II, §8, No. 4), H◦ is also balanced, and H ⊂ (H◦ )◦ (loc. cit., § 6,
No. 3).
Now, for every µ ∈ Mc and every f ∈ C (X; C) , f is µ-integrableR (that
is, f ∈ LC1 (X, µ) ), the inequality (22) holds, and µ 0 (f ) = µ(f ) = f dµ
defines, via Prop. 14, the element µ0 of C 0 (X; C) ; thus |µ0 (f )| = |µ(f )| 6
kµk · pS (f ) , where S is the support of µ . When µ ∈ B and f ∈ H , so that
S ⊂ K , kµk 6 a and pK (f ) 6 c , we have pS 6 pK and
on M (X; C) . Let
C = {ν ∈ B : kνk = kµk }
C+ = {ν ∈ C : ν > 0 } .
By Cor. 2 of Th. 1 of Ch. III, §2, No. 4, µ belongs to the closure of C in
M (X; C) for the vague topology, hence for T 0 ; and when µ > 0 , µ belongs
to the closure of C+ in M (X; C) for the vague topology (loc. cit., Cor. 3),
hence for T 0 .
As observed in the preceding note, the mapping ν 7→ ν 0 defines a
homeomorphism B → B0 for the topologies T 0 ∩ B and T ∩ B0 , there-
fore µ0 belongs to the closure of C0 in C 0 (X; C) for T ; and when µ > 0 ,
µ0 belongs to the closure of (C+ )0 in C 0 (X; C) for T .
IV.53, `. −8, −7.
“ . . . for every compact subset S of X , µ and ν take on the same values
in K (X, S; C) .”
Let f ∈ K (X, S; C) . By Cor. 2, if ε > 0 there exists a finite linear
combination X
g= λ i ϕ Ki
i
|f − g| = |f − g|ϕS 6 ε ϕS ,
1
|µ(f ) − µ(gn )| 6 |µ|(S) ,
n
thus µ(gn ) → µ(f ) . Similarly ν(gn ) → ν(f ) . But µ(gn ) = ν(gn ) for all n ,
whence µ(f ) = ν(f ) .
IV.55, `. 10.
“ . . . the I(K, U) form a base for the topology T ”
INT IV.x54 extension of a measure. l p spaces §4
The formula displayed in `. 7 shows that the set of all I(K, U) is closed
under finite intersections, therefore the set of unions of arbitrary families of
sets I(K, U) is the set of all open sets for T . In particular, for each M ⊂ X ,
the set of all I(K, U) for which K ⊂ M ⊂ U forms a fundamental system of
neighborhoods of M . Note that T exists independently of any function α .
IV.55, `. 13.
“The condition (PCII ) expresses that Φ is dense in P(X) ”
If O is any nonempty open subset of P(X) then O contains some set
I(K, U) , and I(K, U) ∩ Φ 6= ∅ by (PCII ), whence O ∩ Φ 6= ∅ .
IV.55, `. 15–17.
“Th. 4 of No. 6 expresses that the function M 7→ µ(M) is continuous
on the clan of µ-integrable sets, for the topology induced by T .”
Here, µ can be any measure on X (not necessarily positive). Write Ψ
for the set of all µ-integrable subsets of X ; it follows from Props. 6 and 7
of No. 5 that Ψ is a clan, hence satisfies (PC I ). We are to show that the
function Ψ → C defined by A 7→ µ(A) (A ∈ Ψ) is continuous for the
topology T ∩ Ψ induced on Ψ by the topology T on P(X) .
Given M ∈ Ψ , let us show continuity at M . Given any ε > 0 , choose
K and G as in the cited Th. 4, so that M ∈ I(K, G) and |µ|(G --- K) 6 ε .
Then
N ∈ I(K, G) ∩ Ψ ⇒ |µ|(G --- N) 6 |µ|(G --- K) 6 ε ,
§4 integrable functions and sets INT IV.x55
and we have shown that the filter base α(V ∩ Φ) converges to µ(M) ∈ R .
{Indeed, µ(M) ∈ R+ since α > 0 . Thus µ(M) > 0 for every µ-integrable
INT IV.x56 extension of a measure. l p spaces §4
lim α(N)
N∈Φ, N→M
lim α(N)
N∈Φ, N→M
By 3◦ , one has
α(K) = inf α(M) ;
M∈Φ, M⊃K
K* ⊂ K ∪ K0 ⊂ U* ;
K ∪ K0 ⊂ M ⊂ U* .
INT IV.x58 extension of a measure. l p spaces §4
|α(M) − α(U)| 6 ε ,
|α(M) − α(K ∪ K0 )| 6 δ ,
therefore
But, by the note for `. 8–10, given any ε > 0 there exists a compact sub-
set K0 of U such that the difference between α(K 0 ) and α(U) is as small
as we like, therefore the preceding displayed inequality is in fact an equality.
IV.56, `. 17–19.
“ . . . by (PM00IV ), for every set M ∈ Φ such that K ⊂ M ⊂ U , there
exists a compact set K0 ⊂ M such that
α(M) 6 α(K0 ) + ε 6 b + ε ”
therefore U ∈ Φ , and α(U) = b (as one also knows from the proof of
“necessity” of the condition b < +∞ ).
IV.56, `. −8 to −3.
“The definition of Φ and α can now be transformed as follows (taking
into account (PCII )): in order that M ∈ Φ , it is necessary and sufficient
that, for every ε > 0 , there exist a compact set K and an open set U ∈ Φ
such that K ⊂ M ⊂ U and α(U) − α(K) 6 ε ; α(M) is, moreover, the
infimum of the α(U) for the open sets U ∈ Φ containing M , and the
supremum of the α(K) for the compact sets K ⊂ M .”
One is assuming that the hypotheses of Th. 5, including (PM IV ), are
fulfilled, and that, as shown in 3◦ :
(i) Every compact set K ⊂ X belongs to Φ , with α(K) ∈ R + defined
by the formula
α(K) = inf α(P) ;
P∈Φ, P⊃K
Let us write Φ e for the set of all M ⊂ X that satisfy the condition
in (A); we are to show that Φe = Φ and that the formulas (B) hold for every
M ∈ Φ . The proof is organized into three parts: Φ ⊂ Φ e, Φe ⊂ Φ , and the
formulas (B).
Proof of Φ ⊂ Φe . Let M ∈ Φ and let ε > 0 . Since α(N) → α(M) as
N → M while remaining in Φ , there exists a neighborhood I(K, U) of M
in P(X) such that
e , and so Φ ⊂ Φ
concisely α(U) − α(K) 6 ε , which proves that M ∈ Φ e.
§4 integrable functions and sets INT IV.x61
K ⊂ K0 ⊂ N ⊂ U 0 ⊂ U ,
K ⊂ K 0 ⊂ K0 ⊂ N0 ⊂ U0 ,
|α(N) − α(N0 )| 6 ε ;
r = inf{α(U) : M ⊂ U ∈ Φ , U open} ,
s = sup{α(K) : K ⊂ M , K compact } ;
K00 = K ∩ { U0 ⊂ M ∩ { N ⊂ U ∩ { K0 = U00 .
whence α(U00 --- K00 ) 6 α(U --- K) + α(U0 --- K0 ) by the subadditivity of α
for open sets belonging to Φ .
IV.57, `. −12, −11.
“ . . . M ∪ N belongs to Φ .”
This completes the proof that Φ is a clan (No. 9, Prop. 17). Conse-
quently α is subadditive on Φ by the ‘conditional subadditivity’ (p. IV.56,
`. −2, −1) proved earlier.
IV.57, `. −9, −8.
“ . . . since ε is arbitrary, we have α(M ∪ N) = α(M) + α(N) .”
Since ε is arbitrary, we have α(M ∪ N) > α(M) + α(N) , and the reverse
inequality holds by the subadditivity of α .
IV.57, `. −3 to −1.
“Since β is positive, |β(f )| 6 α(K) · kf k for every function f ∈ E (Φ)
whose support is contained in K ”
For such a function f , −kf k ϕK 6 f 6 kf k ϕK , whence
E (Φ) ∩ F(X, K; R)
IV.58, `. 8, 9.
“ . . . the restriction to K of the positive linear form β is therefore a
positive measure µ .”
Every positive linear form on K (X; R) is a positive measure (Ch. III,
§1, No. 5, Th. 1).
IV.58, `. 10.
“ there exists an open set U ∈ Φ such that K ⊂ U , µ(U) 6 µ(K) + ε ”
There exists an open set V with K ⊂ V ⊂ V ⊂ U and V compact;
replacing U by V , we can suppose that U is relatively compact, thereby
assuring that every function whose support is contained in U has compact
support.
IV.58, `. 13.
“Then µ(K) 6 µ(f ) 6 µ(U) 6 µ(K) + ε ”
To justify the notation µ(f ) at this stage, we must assume that f has
compact support; this is assured by arranging that U be relatively compact
(or by constructing f to have compact support). Moreover, the equality
µ(f ) = β(f ) will shortly be needed.
{Absent this precaution, we do not know that f is µ-integrable; of
course, after it has been shown that every open set belonging to Φ is
µ-integrable, a bounded continuous function with support contained in such
a set will be integrable (No. 4, Prop. 5).}
IV.58, `. 15.
“ . . . we see that |µ(K) − α(K)| 6 ε ”
From the above inequalities, we extract (noting that µ(f ) = β(f ) by
the definition of µ )
µ(K) 6 µ(f ) 6 µ(K) + ε
and α(K) 6 β(f ) = µ(f ) 6 α(K) + ε , that is,
By 4◦ ,
U∈Φ ⇔ b < +∞ ,
in which case α(U) = b . On the other hand, by Prop. 10 of No. 6,
U is µ-integrable ⇔ b < +∞ ⇔ U ∈ Φ,
M∈Φ ⇔ M is µ-integrable.
as observed in No. 6, Cor. 1 of Th. 4 (see the note for IV.45, `. 1–3). Since
µ = α on K , one concludes that µ(M) = α(M) .
IV.58, `. 22.
“ µ*(U) = sup α(M) for every open set U ”
M∈Φ, M⊂U
Every M ∈ Φ is µ-integrable and α(M) = α(M) = µ(M) = µ*(M) ;
if, moreover, M ⊂ U , then α(M) = µ*(M) 6 µ*(U) , so the supremum in
question is 6 µ*(U) .
Write K for the set of all compact subsets of X . By Cor. 4 of Th. 4
of No. 6,
d) β is subadditive. For,
by b) and c).
S
∞
e) If (Bk )k>1 is an increasing sequence in Ψ and B = Bk , then
k=1
For, setting C1 = B1 and Cn = Bn --- Bn−1 for n > 2 , the Cn are pairwise
S
∞
disjoint and Cn = B , therefore
n=1
∞
X n
X [
n
β(B) = β(Cn ) = sup β(Ck ) = sup β Ck
n n
n=1 k=1 k=1
= sup β(Bn ) = lim β(Bn )
n n
by the theorem on monotone limits (GT, IV, §5, No. 2, Th. 2).
S
∞ P∞
f ) For any sequence (Bk )k>1 in Ψ , β Bk 6 β(Bk ) (‘complete
k=1 k=1
S
n
subadditivity’). For, writing Cn = Bk for n = 1, 2, 3, . . . and C =
k=1
S
∞
Bk , by e) and d) one has
k=1
n
X ∞
X
β(C) = sup β(Cn ) 6 sup β(Bk ) = β(Bk ) .
n n
k=1 k=1
the first equality by §1, No. 2, Prop. 7, and the last by item e) of the note
for IV.58, `. −6, −5.
IV.59, `. 6, 7.
“ . . . if B is an element of Ψ contained in L , then B is µ-integrable”
Let S be the tribe of µ-integrable subsets of L . The set of all A ⊂ X
such that A∩L ∈ S is easily seen to be a tribe that contains every compact
subset of X , hence contains the tribe Ψ of all Borel sets in X ; that is,
B ∩ L is µ-integrable for every B ∈ Ψ . In particular, if B ∈ Ψ is contained
in L , then B = B ∩ L is µ-integrable.
IV.59, `. 8, 9.
“Since β(U) = µ*(U) and β(K) = µ(K) , we see that |µ*(B)−β(B)| 6
2ε .”
For,
where
IV.60, `. 6, 7.
“ . . . therefore the restriction of f to H is continuous.”
GT, I, §3, No. 2, Prop. 4.
IV.60, `. 17.
P
∞
“ |µ|(K --- K0 ) 6 ε/2n ”
n=1
Write An = K --- Kn and A = K --- K0 , and cite §4, No. 5, Cor. of
Prop. 8.
IV.60, `. −14 to −11.
“ . . . it comes to the same to say that a measurable set A is a set such
that, for every compact set K , there exist a negligible set N ⊂ K and a
partition (Kn ) of K --- N formed by a sequence of compact sets each of
which is contained either in K ∩ A or in K ∩ { A .”
Suppose A is measurable, that is, ϕA is measurable, and let K be
any compact set in X . By Def. 1, there exist a negligible set N ⊂ K and
a partition
(Kn ) of K --- N into a sequence of compact sets K n such that
ϕA Kn is continuous for each n . But 0, 1 are the only possible values
of ϕA ; let
−1 −1
K0n = Kn ∩ ϕA (1) , K00n = Kn ∩ ϕA (0) .
Then K0n , K00n partition Kn into subsets that are closed (by the continuity
of ϕA Kn ) hence compact, with K0n ⊂ A and K00n ⊂ { A . Thus N , together
with the compact sets (K0n ), (K00n ) satisfy the condition in the assertion.
The converse is immediate from Def. 1, since ϕ A is constant (hence
continuous) on each of the sets K ∩ A and K ∩ { A .
IV.60, `. −7.
“The condition is necessary . . . ”
Suppose A is measurable and let K be any compact set in X . As
noted following Def. 2, there exists a partition
∞
[ ∞
[
K=N∪ K0n ∪ K00n ,
n=1 n=1
§5 measurable functions and sets INT IV.x73
where N is negligible, the K0n and K00n are compact, and K0n ⊂ A ,
K00n ⊂ { A for all n . Then A ∩ K00n = ∅ for all n , so
∞
[
(∗) A ∩ K = (A ∩ N) ∪ K0n ;
n=1
S
∞
therefore K0n is integrable (§4, No. 5, Cor. of Prop. 8). Thus each of the
n=1
two terms on the right side of (∗) is integrable, hence so is A ∩ K .
IV.60, `. −5.
“The condition is sufficient . . . ”
Let K be any compact set in X . By hypothesis A ∩ K is integrable,
hence there exists a partition
∞
[
0
A∩K=N ∪ K0n ,
n=1
where N0 is negligible and the K0n are compact (§4, No. 6, Cor. 2 of Th. 4).
But ( { A) ∩ K = K --- A ∩ K is also integrable (§4, No. 5, Prop. 7), so there
is also a partition
∞
[
00
{A∩K = N ∪ K00n
n=1
00
with N negligible and the K00n compact. Then N = N0 ∪ N00 is negligible
and
∞
[ ∞
[
K = (A ∩ K) ∪ ({ A ∩ K) = N ∪ K0n ∪ K00n ,
n=1 n=1
A measurable ⇒ { A measurable.
‘capacity’ is different in GT and TG, but that with either definition, |µ|* has
the desired properties. In both cases, the property asserted for A is
established in GT, IX, §6, No. 9, Th. 5 and TG, IX, §6, No. 10, Th. 6,
respectively.
For the purposes of the proof at hand, one need read no further; the rest
of this note is devoted to the comparative analysis of the two definitions.
In GT, IX, §6, No. 9, Def. 8, a capacity on (a Hausdorff space) X is
defined to be a function f : P(X) → R satisfying the following axioms:
(CAI ) If A ⊂ B , then f (A) 6 f (B) .
(CAII ) If (An ) is any increasing sequence of subsets of X , then
[
f An = sup f (An ) .
n
n
For, let α be the infimum on the right side; if U is any open set containing K
(at least U = X qualifies) then f (K) 6 f (U) , whence f (K) 6 α .
(CA0III ) ⇒ (CA00III ): If f (K) = +∞ then also α = +∞ . If f (K) < +∞ ,
let a be any real number such that f (K) < a < +∞ ; since a > f (K) ,
by (CA0III ) there exists an open set U such that K ⊂ U and f (U) < a ,
INT IV.x76 extension of a measure. l p spaces §5
(GT, IX, §6, No. 9, Def. 9) or (TG, IX, §6, No. 10, Def. 10). For f = |µ|*
the condition reads
(∗) V i = Ni ∪ Ai (1 6 i 6 r) ,
Vε = V1ε1 ∩ · · · ∩ Vrεr ,
Vε0 = { V1 ∩ · · · ∩ { Vr = { (V1 ∪ · · · ∪ Vr ) ,
S
r
Since K ⊂ Vi , one has
i=1
r
[ [ [
(†) K=K∩ Vi = K ∩ Vε , K ∩ Vi = K ∩ Vε .
i=1 ε6=ε0 εi =1
INT IV.x78 extension of a measure. l p spaces §5
V ε1 = N ε1 ∪ A ε1 ,
where Nε1 is negligible, Aε1 is integrable, and Aε1 ⊂ Ai for at least one
value of i (in this instance, for every i ).
Consider now any ε such that ε 6= ε0 and ε 6= ε1 (that is, εi is equal
to 1 for at least one value of i and to −1 for at least one value). Then
\ \ \ [
Vε = Vi ∩ { Vi = Vi ∩ { Vi = R ∩ { S ,
εi =1 εi =−1 εi =1 εi =−1
T S
where R = Vε and S = Vε . By the argument used for Vε1 , one
εi =1 εi =−1
can write R as a disjoint union R = N0 ∪A
T with N0 negligible, A integrable
and A ⊂ Ai for some i (indeed, A = Ai ). Then
εi =1
Vε = R ∩ { S = (N0 ∪ A) ∩ { S
= (N0 ∩ { S) ∪ (A ∩ { S) = (N0 ∩ { S) ∪ (A --- A ∩ S) ,
where N0 ∩ { S is negligible,
T A --- A ∩ S is integrable and is contained in
some Ai (indeed, in Ai ). We write Nε = N0 ∩ { S and Aε = A --- A∩S .
εi =1
To summarize: for every ε 6= ε0 we may write
Vε = N ε ∪ A ε
(††) (∀ ε 6= ε0 ) K ∩ Vε = (K ∩ Nε ) ∪ (K ∩ Aε ) ,
§5 measurable functions and sets INT IV.x79
[
N= K ∩ Nε ,
ε6=ε0
where N is negligible, the Mj are integrable and, for each j , there exists
an index i such that f = gi on Mj .
{Though it is not needed for the rest of the proof of Prop. 4, we verify
that for each i (1 6 i 6 r), K ∩ Vi = K ∩ Vxi has the form
[
K ∩ Vi = N0i ∪ Mj ,
j∈Ji
with Pnj negligible and (Kmnj )m∈N a sequence of pairwise disjoint compact
sets such that the restriction to Kmnj of gi —hence also of f —is continuous
(No. 1, Def. 1). The partition (†††) doubly refined now takes the form
[ [
K=N∪ Nj ∪ Knj
j n∈N
[ [ [ [
= N∪ Nj ∪ Pnj ∪ Kmnj
j j n∈N m∈N
[ [ [
= N∪ Nj ∪ Pnj ∪ Kmnj
j j,n j,n,m
[
= N0 ∪ Kmnj
j,n,m
S S
with N0 = N ∪ Nj ∪ Pnj negligible and the Kmnj compact sets such
j j,n
that f Kmnj is continuous; since K is an arbitrary compact subset of X ,
f is measurable (No. 1, Def. 1).
IV.61, `. −4.
“By the principle of localization, every locally negligible set is measur-
able.”
Let A ⊂ X be locally measurable. Given any compact set K in X , it
suffices to show that A ∩ K is integrable (No. 1, Prop. 3). Indeed, A ∩ K is
negligible by Prop. 5 below (which is derived directly from Def. 3 without the
1 The N ’s here are different from the N ’s occurring in the preceding note (an
j i
ephemeral notation on the way to the partition (†††)).
§5 measurable functions and sets INT IV.x81
K1 = K ∩ { U = K --- U = K --- K ∩ U ;
IV.63, `. 11–13.
“Proposition 6.”
It is convenient to change the notation: assuming g : X → F is measur-
able and f : X → F is a function equal locally almost everywhere to g , let us
show that f is also measurable. Let N = {x ∈ X : f (x) 6= g(x) } , a locally
negligible set. The plan is to cite the principle of localization (Prop. 4).
For each x ∈ X let Vx be any compact (hence integrable) neighborhood
of x and let gx = g . Then Vx ∩ N is negligible (Prop. 5) and, on the subset
Vx --- Vx ∩ N of X --- N , f = g = gx , thus f = gx almost everywhere
in Vx . Quote Prop. 4.
IV.64, `. 6–9.
“Corollary 4.”
One will observe that the same proof works with R replaced by C . It
is tacit that F has its unique compatible Hausdorff topology (TVS, I, §2,
No. 3, Th. 2).
Sufficiency. Assume each fk : X → F is measurable. For each k ,
ek fk : X → F is the composite X → R → F , where
x 7→ fk (x) 7→ fk (x) · ek (x ∈ X) ;
X
n
u f (x) = u ek fk (x) = (f1 (x), . . . , fn (x)) .
k=1
If prk : Rn → R is the k ’th coordinate projection, then pr k u(f (x) =
fk (x) for all x , thus
fk = (prk ◦ u) ◦ f ;
since f is measurable and prk ◦ u is continuous, fk is measurable by Th. 1.
IV.64, `. 10–13.
“Corollary 5.”
The mapping [f · g] : F × G → H is the composite X → F × G → H
defined by
x 7→ f (x), g(x) 7→ [f (x) · g(x)] ,
where f and g are measurable and (u, v) → [u · v] is continuous, so the
composite is measurable by Th. 1.
The notation [f · g] is suggestive of a bilinear operation, but in fact
bilinearity plays no role in the proof and can be omitted from the hypothesis.
IV.64, `. −5.
“The first assertion obviously follows from the second ”
With notations as in 2◦ , f is continuous on K1 ; cite No. 1, Prop. 1.
IV.65, `. 4, 5.
“ Bn,r is a countable union of compact sets contained in K 0 , hence is
integrable ”
For any pair (α, β) ∈ A × A , write Fα,β : K0 → R for the function
Fα,β (x) = d fα (x), fβ (x) (x ∈ K0 ) .
−1
{x ∈ K0 : d fα (x), fβ (x) > 1/r } = Fα,β ([1/r, +∞[)
INT IV.x84 extension of a measure. l p spaces §5
S
∞
therefore Bn,r = Tm is integrable (§4, No. 5, Prop. 8).
m=1
Note that since An is a decreasing function of n , it is clear from (∗)
that, for fixed r , Bn,r is also a decreasing function of n ; and for fixed n ,
Bn,r is an increasing function of r .
IV.65, `. 6–8.
“If r is fixed, the intersection of the decreasing sequence of sets B n,r
(n = 1, 2, . . .) has measure zero, since f α (x) tends to f (x) almost every-
where in K0 with respect to the filter F ”
T
∞
Set Br = Bn,r . Since (Bn,r )n>1 is a decreasing sequence of in-
n=1
tegrable sets, we know that Br is integrable and |µ|(Bn,r ) → |µ|(Br )
as n → ∞ (§4, No. 5, Cor. of Prop. 7).
We are to show that |µ|(Br ) = 0 ; since K0 ∩ N is negligible (No. 2,
Prop. 5) and Br ⊂ K0 , it will suffice to show that Br ⊂ N .
Let x ∈ Br and assume to the contrary that x ∈ X --- N . Then,
fα (x) → f (x) with respect to F , hence the family fα (x) α∈A is Cauchy
with respect to F ; since the An form a base for F , given any ε > 0 there
exists an index n such that
(α, β) ∈ An × An ⇒ d fα (x), fβ (x) < ε .
the existence of such an n means that x does not belong to B n,r , contrary
to x ∈ Br ⊂ Bn,r .
Thus lim |µ|(Bn,r ) = |µ|(Br ) = 0 .
n→∞
§5 measurable functions and sets INT IV.x85
IV.65, `. 8.
“ . . . thus lim |µ|(Bn,r ) = 0 ”
n→∞
See the preceding note.
IV.65, `. 13–16.
“Let C be the complement of B in K0 ; by construction, fα (x) con-
verges uniformly to f (x) in C with respect to the filter F , and since the
restrictions of the fα to C are continuous, so is the restriction of f to C .”
Of course B ⊂ K0 . Recall that K0 ∩ N is negligible; for reasons that
will be clear in the following argument, it is convenient to take C to be the
complement of B ∪ (K0 ∩ N) in K0 . Writing N0 = K0 ∩ N , an elementary
calculation yields
Thus, for all x ∈ C and all r , we have x ∈ K 0 --- Bnr ,r , therefore (by the
definition of Bnr ,r )
d fα (x), fβ (x) < 1/r for all α, β ∈ Anr ,r .
In other words: given any integer r > 1 , for every (α, β) ∈ A nr ,r × Anr ,r
we have
d fα (x), fβ (x) < 1/r for all x ∈ C ,
whence sup d fα (x), fβ (x) 6 1/r ; this shows that the family of restric-
x∈C
tions fα C is uniformly Cauchy with respect
to F , and since f α → f
pointwise in C it follows that fα C → f C uniformly with respect to F
(GT, X, §1, No. 5, Prop. 5).
And since C ⊂ K 0 and the fα K0 are contin-
uous, we conclude that f C is continuous.
INT IV.x86 extension of a measure. l p spaces §5
therefore
n > m ⇒ α n ∈ An ⊂ Am .
then f is measurable.
2◦ For every compact subset K of X and every ε > 0 , there exists
a compact set K1 ⊂ K such that |µ|(K --- K1 ) 6 ε and such that the
restrictions of the fn to K1 are continuous and converge uniformly to f
on K1 .
In the language of Th. 2 and its proof, A = N and F is the filter on N
generated by the sets An = {k ∈ N : k > n } . Another way of packaging
the result:
∞
[
Ai ∩ K = N i ∪ Kin (i = 1, . . . , m) .
n=1
S
m
The Ai ∩ K are pairwise disjoint, with union K (because Ai = X ),
i=1
therefore
m
[ [
m [ ∞
m [
K= Ai ∩ K = Ni ∪ Kin ,
i=1 i=1 i=1 n=1
where the first union is negligible and the K in are pairwise disjoint compact
sets with f = ai on Kin , thus f is measurable by No. 1, Def. 1.
INT IV.x88 extension of a measure. l p spaces §5
If not “step function”, what were the alternatives for “fonction étagée”?
Harrap’s Unabridged gives “tiered” or “terraced” as possibilities for “étagée”.
The frequent translation of “ét. . .” as “st. . .” suggests “staged”. In the ter-
minology of P.R. Halmos’ Measure theory, a “fonction Φ-étagée” (Φ a clan)
is a “simple function” with respect to a ring Φ of subsets of a set (op. cit.,
p. 84), whereas T.H. Hildebrandt’s Introduction to the theory of integration
employs “finite-step function”, with “step function” reserved for functions
with a countable number of values (op. cit., p. 208).
After consulting twenty or so textbooks on Integration and Real Anal-
ysis, I can report the following usage:
(1) “step function” (or “step map”), for “fonction en escalier”: Asplund
& Bungart, Burrill & Knudsen, Dieudonné, Hewitt & Stromberg, K. Hoff-
man, Lang, McShane & Botts, K.A. Ross, Royden, Stromberg, Zaanen. The
overwhelming choice. {Not all books treated the topic, and Hildebrandt em-
ploys the interesting term “staircase function”, with “interval step function”
as possible alternative (op. cit., p. 218).}
(2) “step function” (or “step map”) with respect to a class of sets,
for “fonction étagée”: Dinculeanu, L.M. Graves, Lang, McShane, Segal &
Kunze, Zaanen.
(20 ) “simple function” for “fonction étagée”: Berberian, Burrill & Knud-
sen, Halmos, Hewitt & Stromberg, Munroe, McShane & Botts, Royden,
Rudin, Saks, Stromberg. Somewhat in the lead, but it translates “étagée”
poorly.
Two possible solutions seem to me to stand out for the usage in Inte-
gration:
(A) “step function” for “étagée”, and “interval step function” for “en
escalier”;
(B) “simple function” for “étagée” and “step function” for “en escalier”.
whereas if b = a then
−1 −1 −1
hn (b) = hn (a) = {y ∈ A : gn (y) = b }∪(X --- A) = A∩ g n (a) ∪(X --- A) .
By the Lemma of §4, No. 9 there exists in the clan generated by the V xk ∩K0
(hence in the clan of integrable sets and in the tribe of measurable sets)
a finite family A of pairwise disjoint sets such that each V xk ∩ K0 may be
expressed as the union of a subfamily of A . Let A 1 , A2 , . . . , Aq be a listing
(without repetitions) of those nonempty members of A that occur in the
expressions of Vx1 ∩ K0 , . . . , Vxr ∩ K0 ; then
q r
[ [
Aj = V xk ∩ K 0 = K 0 ,
j=1 k=1
−1 [
g n (a) = (X --- K0 ) ∪ Ak
ak =a
−1
(that is, the union of X --- K0 and those Ak for which ak = a ), and g n (aj )
is equal to [ [
Ak or (X --- K0 ) ∪ Ak
ak =aj ak =aj
according as aj 6= a or aj = a , respectively.
Given any y ∈ K0 , say y ∈ Aj , then
S
∞
(iii) Finally, suppose K = N ∪ Ki is a partition of K with N negli-
i=1
gible and
(Ki ) an infinite sequence of non-negligible compact sets such that
every f Ki is continuous.
Sn
For every n > 1 let Jn = Ki , so that (Jn ) is an increasing sequence
i=1
S
∞
of compact sets such that Jn = K --- N and f Jn is continuous for
n=1
every n .
For each n , the technique of the preceding case shows that there exists
a measurable step function gn such that
(∗) d gn (y), f (y) 6 1/n for all y ∈ Jn ,
S
∞
and, say, gn (y) = a for all y ∈ Ki = (K --- N) --- Jn . Given any
i=n+1
S
∞
y ∈ K --- N = Jn , we assert that gn (y) → f (y) . Say y ∈ Jn0 . Given
n=1
any ε > 0 , choose an integer n1 such that 1/n1 < ε ; increasing n1 if
necessary, we can suppose that n1 > n0 . Then, for all n > n1 we have
y ∈ J n0 ⊂ J n1 ⊂ J n ,
therefore by (∗), d gn (y), f (y) 6 1/n 6 1/n1 < ε . Thus gn → f pointwise
on K --- N , hence almost everywhere in K .
IV.67, `. 5, 6.
“It is clear that the sequence gn (x) converges to f (x) at every point
of K not belonging to N .”
See the preceding note.
IV.67, `. 12.
“With notations as in the proof of Th. 3 . . . ”
The notations in the note before the last are readily adapted to the
present corollary, following the analysis into cases (i)–(iii).
(i) If K is negligible, define gn = 0 for all n .
0 (ii) If K = N∪K0 is a partition of K with N negligible, K 0 compact and
f K continuous, given any integer n > 1 construct a function g n : X → F
as follows. As in the previous note, let K 0 = A1 ∪ · · · ∪ Aq be a partition
of K0 such that the Aj are nonempty measurable sets on each of which f
has oscillation 6 1/n . For each j = 1, . . . , q choose a point x j ∈ Aj and
let aj = f (xj ) . If j is an index for which |aj | > 1/n , that is, n|aj | − 1 > 0 ,
define
1
a0j = 1 − aj ∈ F ,
n|aj |
INT IV.x94 extension of a measure. l p spaces §5
1
and note that |a0j | = |aj | − 1/n , a0j − aj = − · aj and |a0j − aj | = 1/n .
n|aj |
Define
0
if x ∈ X --- K0 = (X --- K) ∪ N
gn (x) = 0 if x ∈ Aj and |aj | 6 1/n
0
aj if x ∈ Aj and |aj | > 1/n .
Each of the finitely many values of gn is assumed on a measurable subset
of X : [
−1
gn (0) = (X --- K0 ) ∪ Aj ,
|aj |61/n
|gn (x) − f (x)| = |f (x)| 6 |f (x) − f (xj )| + |f (xj )| 6 1/n + |aj | 6 1/n + 1/n ,
but
|f (xj )| − |f (x)| 6 |f (xj ) − f (x)| 6 1/n ,
§5 measurable functions and sets INT IV.x95
IV.68, `. 8–10.
“ . . . in the notations of Th. 3, condition b) is satisfied by taking H to
be the countable set formed by the values of all of the functions g n .”
A proof can be based directly on the definition of ‘measurable function’.
Given a compact set K ⊂ X , we seek a negligible set N ⊂ K and a countable
set H ⊂ F such that f (K --- N) ⊂ H . Since f is measurable, there exist
a negligibleS set N ⊂ K and a sequence (K n ) of compact sets such that
K --- N = Kn and f Kn is continuous for every n (No. 1, Def. 1; that
n
the Kn can be taken to be pairwise disjoint is not needed here). For each n ,
f (Kn ) is a compact metrizable subspace of F , hence has a countable base
(GT, IX, §2, No. 9, Prop. 16) and therefore a countable subset H n such that
§5 measurable functions and sets INT IV.x97
−1
∞
\ ∞
\
−1 −1
f (a) = f B1/p (a) = f B1/p (a)
p=1 p=1
IV.68, `. 15.
“It follows from condition a) that A n,p is measurable.”
We continue with the notations established in the preceding note. Fix
an integer p > 1 . For every integer n > 1 ,
An,p = {x ∈ K --- N : d f (x), an 6 1/p}
−1
= {x ∈ K --- N : f (x) ∈ B1/p (an ) } = (K --- N) ∩ f B1/p (an ) ,
n+1
[ n
[ n
[
Bn+1,p = Ak,p --- Ak,p = An+1,p --- Ak,p .
k=1 k=1 k=1
Define gm,p : X → F by
∞
[
b if x ∈ X --- Cm,p = (X --- K) ∪ N ∪ Bi,p
gm,p (x) = i=m+1
ai if x ∈ Bi,p for some (unique) i ∈ {1, , . . . , m} .
IV.68, `. 21–23.
“ . . . as m tends to infinity, gm,p converges pointwise to the function f p
equal to an on Bn,p (n > 1) and to b on N ∪ { K ”
Continuing the notations of the preceding note, the sets C m,p (m =
1, 2, 3, . . .) form an increasing sequence with union K --- N .
As Bn,p may be empty for some n , the definition of f p must be
adjusted. Let’s just calculate lim gm,p (x) . If x ∈ (X --- K) ∪ N then
m→∞
gm,p (x) = b for every m , so the limit is b . On the other hand, if x ∈ K --- N ,
then x ∈ Bn,p for a unique n ; clearly gm,p (x) = gn,p (x) = an for all m > n
(by the uniqueness mentioned in the definition of g m,p ), thus the limit is an .
So the pointwise limit of the gm,p as m → ∞ is the function fp : X → F
defined by
(
b if x ∈ (X --- K) ∪ N
fp (x) =
an if x ∈ Bn,p for some (unique) n .
Thus only the nonempty Bn,p figure in the definition of fp (but their union
is still K --- N ); if Bn,p = ∅ then an does not get used in the definition
(but, being an element of f (K --- N) , it will necessarily show up as a limit
of values of the fp ).
At any rate, fp is measurable, being the pointwise limit of a sequence
of measurable step functions (No. 4, Th. 2).
IV.68, `. 23–25.
“As p tends to infinity, fp (x) tends to f (x) for every x ∈ K --- N ,
and to b for x ∈ N ∪ { K ”
Let x ∈ K --- N . Given any integer p > 1 , x belongs to B n,p for
some n . We know that for all m > n , gm,p (x) = gn,p (x) = an and so
d f (x), gm,p (x) = d f (x), an 6 1/p ;
§5 measurable functions and sets INT IV.x101
letting m → ∞ , we have d f (x), fp (x) 6 1/p , whence fp (x) → f (x) as
p → ∞ . On the other hand, if x ∈ (X --- K) ∪ N then f p (x) = b for all p ,
therefore fp (x) → b trivially. Thus the function g K : X → F defined by
(
b if x ∈ (X --- K) ∪ N
gK (x) =
f (x) if x ∈ K --- N
is closed (GT, IV, §6, No. 2, Prop. 1), therefore −f is measurable (cite
Prop. 8 with D = Q ) hence so is f (No. 3, Th. 1 and the continuity of
c 7→ −c in R ).
IV.69, `. −9, −8.
“It is clear that d f (x), gn (x) 6 2/n for all x ∈ X .”
Say x ∈ Ak , so that f (x) ∈ Bk . Say x ∈ Ci ⊂ Ak , so that
a) S is measurable;
b) K ∩ S is measurable for every compact set K ⊂ X ;
c) K ∩ S is integrable for every compact set K ⊂ X .
Proof. a) ⇔ c): No. 1, Prop. 3.
c) ⇒ b): Suppose S satisfies c). Given K ⊂ X compact. To prove that
K ∩ S is measurable, it suffices by c) ⇒ a) to show that for every compact
K0 ⊂ X , the set K0 ∩(K∩S) is integrable; indeed, K0 ∩(K∩S) = (K0 ∩K)∩S
is integrable by c).
b) ⇒ c): Suppose S satisfies b). Given K ⊂ X compact. We know that
K ∩ S is measurable, hence K ∩ S = K ∩ (K ∩ S) is integrable by a) ⇒ c).
The following corollary is sharpened in No. 6, Th. 5 below.
Corollary. For a subset S ⊂ X ,
S integrable ⇒ S measurable.
Proof. Suppose S is integrable. Given any compact set K ⊂ X , one
knows that K is integrable (§4, No. 6, Cor. 1 of Prop. 10), hence so is K ∩ S
(§4, No. 5, Prop. 7, 2◦ ), thus S is measurable by the above proposition.
IV.70, `. 3, 4.
“ . . . for every z ∈ F ,
(loc. cit., Prop. 8, (i)); to show that |z| 6 ρ(z) , it suffices to show that
|hz, a0 i| 6 ρ(z) for every a0 ∈ B0 . Indeed, if a0 ∈ B0 then a0 belongs
to the weak closure of D , hence there exists a sequence in D , say a 0nk
(k = 1, 2, 3, . . . ) such that a0nk → a0 weakly as k → ∞ . In particular
hz, a0nk i → hz, a0 i , hence |hz, a0nk i| → |hz, a0 i| ; but
|hz, a0nk i| = (|hz, a0nk i|/|a0nk |) · |a0nk | 6 |hz, a0nk i|/|a0nk | 6 ρ(z) ,
and passage to the limit yields |hz, a 0 i| 6 ρ(z) .
Thus, writing b0n = a0n /|a0n | (a unit vector in F0 ), one has
|z| = sup |hz, b0n i|
n
for all z ∈ F .
IV.70, `. 5, 6.
“ A is thus the intersection of K and the sets defined by
|hf (x), a0n i − ha, a0n i| 6 r |a0n | ”
−1 −1
thus K ∩ g (B) is measurable, hence so is K ∩ f (B) as argued in the
preceding note.
IV.70, `. 21, 22.
−1
“ . . . the same reasoning as in Prop. 10 then shows that K ∩ f (B) is
measurable.”
See the preceding note.
IV.70, `. −10, −9. Q
“We may regard F as a subspace of a countable product En of
n
Banach spaces (TVS, II, §4, No. 3)”
As the proof of ‘necessity’ is immediate (see the note for IV.69, `. −2),
one is concerned here with the proof of ‘sufficiency’.
The cited reference refers back to an earlier section (TVS, II, §1, No. 3);
let us review the argument, which entails some subtleties.
If F is any locally convex space (but not necessarily Hausdorff), its
topology can be generated by a family of continuous semi-norms (TVS, II,
§4, No. 1, Cor. of Prop. 1), say (pι )ι∈I . For each ι , the set
Nι = {a ∈ F : pι (a) = 0 }
Assuming F Hausdorff,
Q let Eι be the Banach space
Q obtained by com-
pleting Fι . Since Fι is a topological subspace of Eι (GT, I, §4, No. 1,
ι ι Q
Cor. of Prop. 3), M is also a topological subspace of Eι (transitivity of
ι
induced topologies); thus, every Hausdorff locally convex space is isomor-
phic as a topological vector space to a (topological) subspace of a product
of Banach spaces (TVS, II, §1, No. 3, Prop. 3). {Since a continuous linear
mapping of topological vector spaces is uniformly continuous (GT, III, §3,
No. 1, Prop. 3) it follows that if F is Hausdorff then the uniform structures
of F and M are isomorphic; if, moreover, FQis complete, then so is M , con-
sequently M is a closed linear subspace of Fι (GT, II, §3, No. 4, Prop. 8)
Q ι
hence also of Eι (TVS, II, §4, last paragraph of No. 3).
ι
Consider now the space F of the present corollary. Since F is metriz-
able, it has a countable fundamental system (V n ) of neighborhoods of 0
(TVS, I, §3, No. 1), and since F is locally convex, the V n may be taken
to be closed, balanced and convex (TVS, II, §4, No. 1). If p n is the gauge
of Vn , the topology of F is generated by the family of semi-norms (p n ) (loc.
cit., Cor. of Prop. 1; see also the note for III.39,
Q `. 8–11). With notations
as in the foregoing discussion, if u : F → Fn is the mapping defined by
n
−1
u(a) = un (a) , where Fn = F/ pn (0) (a normed space) and un : F → Fn is
the quotient mapping, then u defines a topological vector spaceQisomorphism
of F onto M = u(F) regarded as a topological subspace of Fn . If En
n
is the Banach space
Q completion of Fn , then u may be regarded as linear
mapping F → En that defines a topological vector space isomorphism
n Q
of F onto the topological subspace M = u(F) of En . Since F has a
n
countable dense subset, so does the subspace pr n M of En (n = 1, 2, 3, . . .);
replacing En by the closure of prn M in En , we can suppose that the En
are separable Banach spaces. Identifying F with M , we have shown that
F is a topological linear subspace of the product of a sequence of separable
Banach spaces.
is measurable. Now,
h(un ◦ f )(x), a0n i = hun (f (x)), a0n i = hf (x), t un a0n i = hf (x), a0n ◦ un i ,
it follows from the cited Th. 1 that, taking u to be the canonical injection
f (X) → E , the mapping u ◦ f : X → E is measurable. But f takes its values
in F , and we wish to show that f : X → F is measurable.
Let K ⊂ X be compact. By the measurability of u ◦ f , there exist
a negligible set N ⊂ K and a partition (K i )i∈I of K --- N into compact
sets Ki on each of which u ◦ f is continuous, that is, the function
x 7→ (u ◦ f )(x) = u f (x) = f (x) ∈ F (x ∈ Ki )
IV.71, `. 9.
“ . . . each of the Sy is measurable ”
Sy is the inverse image, under the measurable mapping x 7→ hy, f (x)i ,
of the closed set {c : |c| 6 1 } in the field of scalars, hence is measurable by
No. 5, Prop. 7.
IV.71, `. 9, 10.
“ Xn is the intersection of the countable family of S y for y ∈ D ∩ Vn .”
Since D = F and Vn is open in F , one has Vn = Vn ∩ D ⊂ Vn ∩ D
(GT, I, §1, No. 6, Prop. 5), whence V n = Vn ∩ D . The following computa-
tion then depends on a formula for the polar of a balanced set (cf. TVS, II,
§8, No. 4) and the continuity of the linear forms f (x) :
x ∈ Xn ⇔ f (x) ∈ Vn◦
⇔ |hy, f (x)i| 6 1 for all y ∈ Vn
⇔ |hy, f (x)i| 6 1 for all y ∈ V n
⇔ |hy, f (x)i| 6 1 for all y ∈ Vn ∩ D
⇔ |hy, f (x)i| 6 1 for all y ∈ Vn ∩ D
⇔ x ∈ Sy for all y ∈ Vn ∩ D ,
T
that is, Xn = Sy .
y∈Vn ∩D
IV.71, `. 10–13.
“ . . . for every compact subset K of X and every ε > 0 , there exists an
integer n such that |µ| K --- (K ∩ Xn ) 6 ε/4 , and then a compact subset
K1 of K ∩ Xn such that |µ| (K ∩ Xn ) --- K1 6 ε/4 ”
Since the Xk are measurable, the K ∩ Xk form an increasing sequence
of integrable (No. 1, Prop. 3) sets with union K , therefore
|µ| K --- (K ∩ Xk ) → 0
(§4, No. 5, Cor. of Prop. 7), whence the existence of n . The existence of K 1
then follows from the ‘inner regularity’ of |µ| (§4, No. 6, Cor. 1 of Th. 4).
IV.71, `. 18, 19.
“ . . . the restriction of f to K2 is therefore continuous ”
Let (xj ) be a directed family in K2 such that xj → x ∈ K2 ; we are
to show that f (xj ) → f (x) in F0 for σ(F0 , F) , that is, for the topology
of pointwise convergence in F . We know that hy, f (x j )i → hy, f (x)i for
each y ∈ D , by the choice of K2 ; thus f (xj ) → f (x) for the topology of
INT IV.x110 extension of a measure. l p spaces §5
thus MK = |µ|(K) meets the requirements (§4, No. 6, Cor. 1 of Prop. 10) .
Addendum. Pursuing the addendum to the preceding note, suppose Y
is an arbitrary subspace of X , and let g ∈ K + (Y) . As observed there, g 0 is
lower semi-continuous on X ; it is also bounded and has compact support.
The argument in the preceding note shows that for every k ∈ R , the set
{x ∈ X : g 0 (x) > k } is either equal to X or is a compact subset of X ,
hence is measurable for µ (No. 1, Cor. 1 of Prop. 3); therefore g 0 is measur-
able (No. 5, Prop. 8). Since every g ∈ K (Y; C) is a linear combination of
four elements of K+ (Y) , it follows from the linearity of g 7→ g 0 that g 0 is
§5 measurable functions and sets INT IV.x113
for every Borel set B in K , in other words, for every Borel set B in X such
that B ⊂ K (GT, IX, §6, No. 3, Remark 2).
For, writing B(T) for the set of Borel sets in a topological space T
(the tribe generated by its open sets, equivalently by its closed sets), we have
for all A, B ∈ B(K) (loc. cit., Prop. 7). The same is true of the set function
B 7→ µ(B) ( B ∈ B(K) ). Let C be the set of all compact (i.e., closed)
subsets of K . By (i), (∗) holds for every set in C ; since C is closed under
finite unions and intersections, and since the tribe of subsets of K generated
by C is B(K) , it follows that (∗) holds for every set in B(K) (Th. 51.F in
Halmos’ book, p. 223; or S.K. Berberian, Measure and integration, Th. 2 on
p. 185, Macmillan, New York, 1965, reprinted by Chelsea). The heart of the
matter is that the clan generated by C is the set of all finite disjoint unions
of sets C --- D , where C, D ∈ C and D ⊂ C .
IV.74, `. −9.
“ (i) We can restrict ourselves to the case that H is compact. ”
Suppose the assertion proved for compact sets, so that, in particular,
µK (K) = µ(K) , and let J be an open set in K (that is, a set of the form
U ∩ K , where U is an open set in X ); then the set H = K --- J is compact,
therefore by assumption
Z = K ∪ U ∩ Y --- K ;
since K = K , we have
Z = K ∪ U ∩ Y --- K = K ∪ (U ∩ Y --- K) = U ∩ Y ⊂ U = U ,
K ∩ U ∩ Y --- K = (K ∩ Y) ∩ U ∩ Y --- K
= K ∩ Y ∩ U ∩ Y --- K
= K ∩ ClY (U ∩ Y --- K) ;
Z --- K ⊂ U ∩ Y --- K ,
h1 ϕU -- K = h1 ϕU − h1 ϕK = h1 − h1 ϕK ,
f1 ϕU -- K = f1 ϕU − f1 ϕK = f1 − f1 ϕK ,
IV.75, `. −2.
“ |µ|Y (f ) 6 |µY (h)| + ε(1 + 2kf k) 6 |µY |(f ) + ε(1 + 2kf k) ”
In slow motion: the inequalities
|µ|(f1 ) − |µ|Y (f ) 6 εkf k ,
|µ|(f1 ) 6 |µ(h1 )| + ε ,
|µ(h1 )| 6 |µ(h1 ) − µY (h)| + |µY (h)| 6 ε kf k + |µY (h)| ,
|µY (h)| 6 |µY |(|h|) 6 |µY |(f )
yield, respectively,
IV.76, `. 18.
“It is immediate (No. 2, Prop. 5) that d) implies a) ”
Assuming d), we are to show that for a set B ⊂ A ,
where the B∩Kn are negligible by the preceding paragraph, therefore B∩K
is negligible.
IV.76, `. −19, −18.
“ . . . one defines
S recursively a sequence
S (H p ) of sets of K such that
Hn+1 ⊂ B --- Hp and |µ|(B --- Hp ) 6 1/n (§4, No. 6, Th. 4). ”
p6n p6n
By b), choose H1 ∈ K with H1 ⊂ B and |µ|(B --- H1 ) 6 1 . Since
B --- H1 is integrable,
there exists a compact set B 1 ⊂ B --- H1 such that
|µ| (B --- H1 ) --- B1 6 1/4 by Cor. 1 of the cited Th. 4; if, by b), H 2 ∈ K is
chosen so that H2 ⊂ B1 and |µ|(B1 --- H2 ) 6 1/4 , then since H2 ⊂ B1 ⊂
B --- H1 , so that
(B --- H1 ) --- H2 = (B --- H1 ) --- B1 ∪ (B1 --- H2 ) ,
it follows that
|µ| (B --- H1 ) --- H2 6 1/4 + 1/4 ,
that is, |µ| B --- (H1 ∪ H2 ) 6 1/2 .
Suppose H1 , H2 , . . . , Hn in K already constructed satisfying the rele-
vant inclusions and inequalities, in particular
[ [ 1
Hn ⊂ B --- Hp and |µ| B --- Hp 6 .
n
p6n−1 p6n
§5 measurable functions and sets INT IV.x121
S S
Then B --- Hp is integrable; choose a compact set B n ⊂ B --- Hp
p6n p6n
such that [
1
|µ| B --- Hp --- Bn 6 ,
2(n + 1)
p6n
1
|µ|(Bn --- Hn+1 ) 6 .
2(n + 1)
S
Then Hn+1 ⊂ Bn ⊂ B --- Hp , thus
p6n
[ [
B --- Hp --- Hn+1 = B --- Hp --- Bn ∪ Bn --- Hn+1 ,
p6n p6n
therefore
[ 1 1
|µ| B --- Hp --- Hn+1 6 + ,
2(n + 1) 2(n + 1)
p6n
S 1
that is, |µ| B --- Hp 6
, which completes the induction.
p6n+1 n + 1
S
∞ S
Set N = B --- Hp ; then N ⊂ B --- Hp for all n , therefore
p=1 p6n
|µ|*(N) 6 1/n for all n , thus N is negligible.
IV.76, `. −17.
“It remains to prove that a) implies b).”
In view of |µ|(K0 --- K) = |µ|(K0 ) − |µ|(K) (§4, No. 5, Prop. 7), the
meaning of b) is that, for every compact subset K 0 of A ,
The proof by contradiction ultimately rests on the fact that the measure
(for |µ| ) of an integrable set is the supremum of the measures of its compact
subsets (§4, No. 6, Cor. 1 of Th. 4).
IV.76, `. −13, −12.
“ B is integrable and |µ|(B) = α ”
Because (Ln ) is an increasing sequence of integrable sets such that the
sequence |µ|(Ln ) has a finite supremum α (§4, No. 5, Prop. 8).
INT IV.x122 extension of a measure. l p spaces §5
B = (B --- A) ∪ (B ∩ A) ;
we show that both terms of the union are locally negligible. Since X --- A
is locally negligible, so is its subset B --- A . On the other hand, for every
K ∈ K , the set
(B ∩ A) ∩ K = B ∩ (A ∩ K) = B ∩ K
is by assumption negligible, therefore B ∩ A is locally negligible by the
µ-density of K in A and its property a) in Prop. 12.
IV.77, `. 10–12.
“If K is a compact subset of X , it comes to the same to say that a set
of compact subsets of K is µ-dense in K or that it is µ K -dense in K ; this
follows from Lemmas 2 and 3 of No. 7 and condition b) of Prop. 12.”
§5 measurable functions and sets INT IV.x123
IV.78, `. 13.
“ |µ|(K ∩ V) > 0 for every K ∈ HV ”
The set K ∩ V is a nonempty open set in the subspace K of X , and
Supp(µK) = K , therefore |µK |(K ∩ V) > 0 by §2, No. 2, Prop. 5; and
|µK |(K ∩ V) = |µ|K (K ∩ V) = |µ|(K ∩ V) by Lemmas 3 and 2 of No. 7.
IV.78, `. 15.
“ N is µ-measurable ”
S
The set K is measurable (see the note for IV.77, `. −8 to −6), and
K∈H
the difference of two measurable sets is measurable.
IV.78, `. 15, 16.
“If N were not locally negligible, it would contain a non-negligible com-
pact set L0 ”
If N is not locally negligible then there exists a compact set K in X
such that K∩N is not negligible (No. 2, Prop. 5). Since K∩N is measurable
and has finite exterior measure ( 6 |µ|(K) ), it follows that K∩N is integrable
(No. 6, Cor. 1 of Th. 5); thus |µ|(K ∩ N) = |µ|*(K ∩ N) > 0 , therefore there
exists a compact set L0 ⊂ K ∩ N such that |µ|(L0 ) > 0 by ‘inner regularity’
of |µ| (§4, No. 6, Cor. 1 of Th. 4).
IV.79, `. 3.
“ d) implies c) ”
§5 measurable functions and sets INT IV.x125
f-
A F
iA j
? ?
X g-G
INT IV.x126 extension of a measure. l p spaces §5
uK- f-
K A F
Q
Q
Q iA j
iK Q
Q
sQ ? ?
X g-G
is commutative; in particular,
(j ◦ f ) ◦ uK = (g ◦ iA ) ◦ uK = g ◦ (iA ◦ uK ) = g ◦ iK .
where
(Hn ) is a sequence of pairwise disjoint compact sets in S
B such that
f Hn = (f K)Hn is continuous for every n , and N = B --- Hn is µK -
n
negligible; in particular, Hn ∈ HK , thus the desired property c) is verified.
INT IV.x128 extension of a measure. l p spaces §5
IV.79, `. 7.
“ a) implies d). ”
Assuming that the set H = {K ⊂ A : K compact, f K continuous }
is µ-dense in A , we are to show that, given any y ∈ F , the extension of f
to X by y , that is, the function g : X → F defined by
(
f (x) for x ∈ A
g(x) =
y for x ∈ X --- A ,
whence |µ|(L --- K) 6 ε/4 + ε/2 + ε/4 by the subadditivity of |µ| (§4, No. 5,
Prop. 6).
An outline of the proof:
a) - b)
Q
k
Q
Q
Q
Q
? Q
d) - c)
§5 measurable functions and sets INT IV.x129
H is µ-dense in K ⇔ H is µK -dense in K
f is µ-measurable ⇔ f is µK -measurable
(the statement on the left, in the sense of Def. 8; the statement on the right,
in the sense of No. 1, Def. 1 applied in the compact space K ).
Proof of ⇒ : Suppose f is µ-measurable. By criterion a) of Prop. 15,
we know that H is µ-dense in K , hence is µ K -dense in K ; from the charac-
terization of density in b) of Prop. 12 (applied to the measurable subset K
of the compact space K ), it follows that f is µ K -measurable (No. 1, Prop. 1,
applied in the space K ).
Proof of ⇐ : Suppose f is µK -measurable. By No. 1, Prop. 1 and b) of
Prop. 12 (both applied in the space K ) we see that H is µ K -dense in K ,
hence µ-dense in K , therefore f is µ-measurable by a) of Prop. 15.
IV.80, `. 7, 8.
“The condition being obviously necessary . . . ”
Suppose f : B → F is µ-measurable
(in the sense of Def. 8). Fix
A ∈ A ; we are to show that f A is µ-measurable. Applying to B and f
the criterion c) of Prop. 15, there exist a topological space G , an injective
mapping j : F → G such that F has the initial topology for j , and a
µ-measurable mapping g : X → G such that g B = j ◦ f (see the note for
IV.79, `. 3,4); thus the diagram
f-
B F
iB j
? ?
X g-G
§5 measurable functions and sets INT IV.x131
uA- f-
A B F
Q
Q
Q iB j
iA Q
Q
sQ ? ?
X g-G
S
where all unions are ‘disjoint’, therefore K --- N = Lmn .
m,n
Thus it has been shown that, given any compact subset K of B , there
exists a partition of K formed of a µ-negligible
set N and a countable family
(Lmn ) of compact sets such that the f Lmn are continuous.
This shows that
the set H of all compact sets H ⊂ B such that f H is continuous satisfies
condition c) of Prop. 12 (as well, of course, the conditions (PL I ) and (PLII )),
therefore H is µ-dense in B (No. 8, Def. 6); thus condition a) of Prop. 15
is satisfied, consequently f : B → F is indeed µ-measurable (Def. 8).
IV.80, `. −15, −14.
“ . . . these generalizations are left to the reader.”
The proofs are sketched in the note after the next.
IV.80, `. −14 to −11.
“ . . . the principle of localization (No. 2, Prop. 4) remains valid when it
is assumed that each of the functions g x is only defined in Vx (or almost
everywhere in Vx ) and is measurable.”
The key to interpreting this statement is the phrase “and is measurable”.
Two interpretations of ‘measurable function’ are available: Def. 1 or Def. 8;
it is clearly Def. 8 that is intended here.
Let us call Prop. 40 the proposed generalization of Prop. 4. One is given:
(i) a topological space F and a mapping f : X → F ;
(ii) for each x ∈ X , a neighborhood Vx of x in X , a negligible sub-
set Nx of Vx , and a function gx , measurable in the sense of Def. 8, whose
(measurable) domain of definition B x contains
the (integrable, hence mea-
surable) set Ax = Vx --- Nx , such that gx Ax = f Ax .
§5 measurable functions and sets INT IV.x133
N = {x ∈ A : f is not continuous at x }
are measurable by the original Cor. 1, it follows from Def. 8 that sup(f, g)
and inf(f, g) are measurable functions A → R .
No. 3, Cor. 2 0 of Th. 1. Given a measurable subset A of X and a
function f : A → R , we are to show that f is measurable if and only
if f + and f − are measurable. Writing 0 for the function on A (or on X )
§5 measurable functions and sets INT IV.x135
in 2) above, HK is µ-dense
in K . By the argument in 3) above, the
set H = {H ∈ K : f H is continuous } is µ-dense in A , therefore f is
µ-measurable by the argument in 4) above.
Remark. A simpler result in this vein: If A , B are measurable sets
in X such that B ⊂ A , and if f : A → F is measurable ( F any topological
space), then f B is also measurable. For, f has a measurable extension
to X , therefore so does f B .
if B ⊂ F then
−1 −1
A ∩ h (B) = {x ∈ A : h(x) ∈ B } = {x ∈ A : f (x) ∈ B } = f (B) ,
−1
thus if B is a Borel set in F then f (B) is the intersection of two measurable
sets, hence is measurable.
−1
In particular, if B is any open set or closed set in F , then f (B) is a
measurable set of X that is contained in A , that is, a µ-measurable subset
of A .
Remark. The term ‘measurable subset of A ’ was avoided, as it begs the
question of whether A is a locally compact subspace of X equipped with a
measure of its own.
Suppose, indeed, that A is a locally compact subspace of X , hence is
a measurable
subset of X (No. 7, first paragraph). One has the R measure
µA = µA induced on A (No. 7, Def. 4), defined by µ A (g) = g 0 dµ for
g ∈ K (A; C) , where g 0 is the extension of g to X by 0 . Borrowing from
the future (Ch. V, §7, No. 1, Cor. of Prop. 2), we know that a set C ⊂ A
is µA -measurable if and only if it is µ-measurable. And (loc. cit., Prop. 2),
a mapping f : A → F is µA -measurable if and only if it is µ-measurable in
the sense of Def. 8, hence has a µ-measurable extension h : X → F ; in this
−1 −1
case, if C is a Borel set in F , then f (C) = A ∩ h (C) is a µ-measurable
set contained in A by the earlier discussion, hence (by the foregoing) is a
µA -measurable subset of A .
The bottom line: when A is a locally compact subspace of X , the term
‘measurable subset of A ’ is unambiguous.
No. 5, Th. 4 0 . As in the original Th. 4, let X be a locally compact
space equipped with a measure µ , F a metrizable topological space, and d
a metric on F compatible with its topology.
The original Th. 4 asserts that a mapping h : X → F is measurable if
and only if it satisfies the following two conditions (slightly reworded):
−1
a) For every closed ball B in F , h (B) is a measurable set in X ;
b) for every compact set K ⊂ X , there exist a countable set H ⊂ F
and a negligible set N ⊂ K such that h(K --- N) ⊂ H .
An equivalent formulation of b) is that for every compact set K ⊂
X , there exists a negligible set N ⊂ K such that the topological subspace
h(K --- N) of F has a countable dense subset, equivalently (GT, IX, §2,
No. 8, Prop. 12), has a countable base for open sets, in other words (loc.
cit., Def. 4) is a separable metrizable space. For, writing S = h(K --- N) , if
there exists a countable set H ⊂ F such that S ⊂ H , then the subspace H
of F has a countable base for open sets, therefore so does the subspace S ,
§5 measurable functions and sets INT IV.x141
−1 −1
A ∩ h (B) = {x ∈ A : h(x) ∈ B } = {x ∈ A : f (x) ∈ B } = f (B) ,
−1
thus f (B) is the intersection of two measurable sets, whence a 0 ) is satisfied.
And
f (K --- N) = (hA)(K --- N) ⊂ h(K --- N) ;
thus f (K --- N) is a subspace of the separable subspace h(K --- N) , whence b 0 )
is satisfied.
Conversely, suppose f satisfies a 0 ) and b 0 ). Fix a point y0 ∈ F and let
h : X → F be the extension of f to X by y0 ; to prove that f is measurable,
it will suffice by Def. 8 to show that h is measurable, and by the original
Th. 4 we need only verify that h satisfies the conditions a) and b).
a) If B is a closed ball in F , then
−1
−1 f (B) if y0 ∈
/B
h (B) =
−1
f (B) ∪ (X --- A) if y0 ∈ B ;
−1 −1
since f (B) is measurable by a 0 ), we see that h (B) is also measurable.
b) Let K ⊂ X be compact; we seek a negligible set N ⊂ K such that
h(K --- N) is separable. Since A is measurable, K ∩ A is integrable (No. 2,
Prop. 3), hence there exists a sequence (K n ) of pairwise disjoint compact
INT IV.x142 extension of a measure. l p spaces §5
S
subsets of K ∩ A such that the set N0 = K ∩ A --- Kn is negligible (§4,
n
No. 6, Cor. 2, 2◦ of Th. 4). For each n , since Kn ⊂ A is compact, by b 0 )
there exists a negligible
S set Nn ⊂ Kn such that f (Kn --- Nn ) is separable.
0
The set N = N ∪ Nn is negligible, N ⊂ K ∩ A ⊂ K , and
n
[ [ [
K ∩ A = N0 ∪ Kn = N 0 ∪ Nn ∪ (Kn --- Nn ) = N ∪ (Kn --- Nn ) ,
n n n
S
therefore K ∩ A --- N = (Kn --- Nn ) ; then
n
[
f (K ∩ A --- N) = f (Kn --- Nn ) ,
n
K --- N = [(K --- N) ∩ (X --- A)] ∪ [(K --- N) ∩ A] ⊂ (X --- A) ∪ [(K ∩ A --- N)] ,
whence
−1
{x ∈ A : f (x) > a } = {x ∈ A : h(x) > a } = A ∩ h ([a, +∞])
is measurable.
2) For every a ∈ D , the set
(
{x ∈ A : f (x) > a } if a > 0
{x ∈ X : h(x) > a } =
{x ∈ A : f (x) > a } ∪ (X --- A) if a 6 0
−1 T −1
therefore {x ∈ X : h(x) 6 b = h ([ − ∞, b]) = { h ([an , +∞]) is the
n
intersection of a sequence of measurable sets.
INT IV.x144 extension of a measure. l p spaces §5
Then:
f is measurable ⇔ h is measurable
⇔ hn is measurable for all n
⇔ fn is measurable for all n
(the first equivalence by Def. 8; the second by (∗) and the original Prop. 10;
the third by Def. 8), thus a) ⇔ b).
§5 measurable functions and sets INT IV.x145
(1) x 7→ hh(x), a0 i (x ∈ X)
(2) x 7→ hf (x), a0 i (x ∈ A) .
Then:
f is measurable ⇔ h is measurable
⇔ the functions (1) are all measurable
⇔ the functions (2) are all measurable
INT IV.x146 extension of a measure. l p spaces §5
(the first equivalence, by Def. 8; the second, by the original Cor. 2; the third,
by Def. 8).
No. 5, Prop. 11 0 . Statement and proof as in the preceding, with “Cor. 2”
replaced by Prop. 11” (but note that the conditions on F are different in
the two statements).
IV.81, `. 2, 3.
“ . . . if V0 ⊂ V , B0 ⊃ B and δ 0 6 δ , then
W(V0 , B0 , δ 0 ) ⊂ W(V, B, δ) ”
|µ|*(M) 6 |µ|*(M0 ) 6 δ 0 6 δ ,
−1
(f, g) ∈ W(V, B, δ) ⇔ (g, f ) ∈ W( V, B, δ) .
−1 −1
Writing W = W(V, B, δ) and W 0 = W( V, B, δ) , we then have W0 = W .
§5 measurable functions and sets INT IV.x147
IV.81, `. 4–6.
2
“Now, if V0 is an entourage such that V0 ⊂ V , then
W(V0 , B, δ/2) ◦ W(V0 , B, δ/2) ⊂ W(V, B, δ) .”
If δ1 , δ2 are > 0 , then
W(V0 , B, δ1 ) ◦ W(V0 , B, δ2 ) ⊂ W(V, B, δ1 + δ2 ) .
For, let (f, g) ∈ left side; then, for a suitable function h , one has (f, h) ∈
second factor and (h, g) ∈ first factor (GT, II, §1, No. 1, footnote). Since
f (x), h(x) ∈ V0 & h(x), g(x) ∈ V0 ⇒ f (x), g(x) ∈ V0 ◦ V0 ⊂ V ,
one has
f (x), g(x) ∈/V ⇒ h(x), g(x) ∈/ V0 or f (x), h(x) ∈/ V0 ;
thus, writing
M = {x ∈ B : f (x), g(x) ∈/ V}
M1 = {x ∈ B : h(x), g(x) ∈ / V0 }
M2 = {x ∈ B : f (x), h(x) ∈/ V0 } ,
we see that M ⊂ M1 ∪ M2 , therefore
|µ|*(M) 6 |µ|*(M1 ) + |µ|*(M2 ) 6 δ1 + δ2 ,
whence (f, g) ∈ right side.
IV.81, `. 9–11.
“ . . . there exists a compact set K ∈ K contained in B such that
|µ|(B --- K) 6 δ , and therefore W(V, K, δ) ⊂ W(V, B, 2δ) .”
Since B is integrable, there exists a compact set K 1 ⊂ B such that
|µ|(B --- K1 ) 6 δ/2 (§4, No. 6, Cor. 1 of Th. 4). In turn, since K is µ-dense
in A (such a K exists—cf. IV.77, `. 1), K 1 has a subset K ∈ K such that
|µ|(K1 --- K) 6 δ/2 (by b) of Prop. 12), therefore |µ|(B --- K) 6 δ/2 + δ/2 .
Suppose (f, g) ∈ W(V, K, δ) . Write
M1 = {x ∈ K : f (x), g(x) ∈ / V}
M2 = {x ∈ B : f (x), g(x) ∈ / V}.
By assumption |µ|*(M1 ) 6 δ ; we wish to show that |µ|*(M2 ) 6 2δ .
From K ⊂ B we see that M1 ⊂ M2 . Moreover, M2 --- M1 ⊂ B --- K :
for, if x ∈ M2 --- M1 then x ∈ M2 ⊂ B , therefore f (x), g(x) ∈ / V ; if
one had x ∈ K , then from x ∈ / M1 we would infer that f (x), g(x) ∈ V ,
a contradiction. From M2 --- M1 ⊂ B --- K we infer that
M2 ⊂ M1 ∪ (B --- K) ,
whence |µ|*(M2 ) 6 |µ|*(M1 ) + |µ|*(B --- K) 6 δ + δ .
INT IV.x148 extension of a measure. l p spaces §5
IV.81, `. −8.
“ . . . consequently (f, g) ∈ / W(V0 , B, α/2) .”
Let M0 = {x ∈ B : f (x), g(x) ∈ / V0 } ; we are to show that
|µ|*(M0 ) > α/2 .
Clearly K ⊂ M0 by the construction of V0 , and M0 ⊂ M because V0
contains the diagonal ∆ of F × F (if x ∈ M 0 then f (x), g(x) ∈ / V0 , hence
f (x), g(x) ∈ / ∆ , that is, f (x) 6= g(x) , and so x ∈ M ); thus K ⊂ M 0 ⊂ M .
Then
α/2 > |µ|(M --- K) = |µ|(M) − |µ|(K) = α − |µ|(K) ,
whence |µ|(K) > α − α/2 = α/2 and finally |µ|*(M 0 ) > |µ|(K) > α/2 .
To summarize: Write S = S (A, µ; F) ( F a Hausdorff uniform space)
and let B be a fundamental system of entourages for F . For every integrable
set B ⊂ A ,
\
{(f, g) ∈ S × S : f = g a.e. in B } = W(V, B, δ) .
V∈B, δ>0
N = {x ∈ B : f (x) 6= g(x) }
for all δ > 0 ; thus (f, g) ∈ W(V, B, δ) for every entourage V (in particular
for every V ∈ B ) and every δ > 0 , whence (f, g) ∈ right side.
Proof of ⊃ : Assuming (f, g) ∈ / left side, we are to show that
(f, g) ∈
/ right side.
Writing M = {x ∈ B : f (x) 6= g(x) } (an integrable set), by as-
sumption |µ|(M) = α > 0 , and the earlier argument shows that there
exists a closed entourage V0 such that (f, g) ∈ / W(V0 , B, α/2) . Choose
V ∈ B so that V ⊂ V0 ; then W(V, B, α/2) ⊂ W(V0 , B, α/2) , therefore
(f, g) ∈
/ W(V, B, α/2) and so (f, g) ∈ / right side.
IV.81, `. −7 to −5.
“ . . . if F is Hausdorff, then the intersection of all the entourages of
S (A, µ; F) is the set of pairs (f, g) such that f (x) = g(x) locally almost
everywhere in A .”
(It is then a triviality that the same is true of the intersection of any
fundamental system of entourages of S (A, µ; F) .)
INT IV.x150 extension of a measure. l p spaces §5
Recall that the entourages of S (A, µ; F) are the supersets of the sets
W(V, B, δ) , where δ > 0 , V runs over the set of entourages for the uni-
formity of F , and B runs over the set of integrable subsets of A . A fun-
damental system of entourages is also given by the sets W(V, K, δ) , where
V runs over a fundamental system of entourages for the uniformity of F ,
K runs over any set K of compact subsets of A that is µ-dense in A , and δ
runs over, say, the values 1/n ( n = 1, 2, 3, . . . ).
Write S = S (A, µ; F) and let N be the intersection of all the en-
tourages for S . To be specific, let K be the set of all compact subsets
of A , and let B be a fundamental system of entourages for the uniformity
of F , so that
\
N = W(V, K, δ)
K∈K, V∈B, δ>0
\ \
= W(V, K, δ)
K∈K V∈B, δ>0
\
= {(f, g) ∈ S × S : f = g a.e. in K }
K∈K
= {(f, g) ∈ S × S : for every compact set K ⊂ A , f = g a.e. in K }
(the third equality, by the formula proved in the preceding note). For each
pair (f, g) ∈ S × S , write
M(f, g) is a measurable set (by the argument in the note for IV.81, `. −14
to −11). The assertion to be proved is that
IV.82, `. 7.
“It follows from No. 10, Prop. 16 that ψ is bijective.”
S
where N0 = Nλ ; since the Aλ are locally countable, so are the Nλ ,
λ
therefore N0 is also locally negligible (loc. cit.).
Since N∪N0 is locally negligible and A --- (N∪N 0 ) = (A --- N) --- N0 =
A0 --- N0 , it follows that f = g locally almost everywhere (on A ), that is,
f˙ = ġ .
Q
ψ is surjective. Let u = (uλ ) ∈ S(Aλ , µ; F) ; we seek a function
λ
f ∈ S (A, µ; F) such that ψ(f˙) = u , that is, f˙λ = uλ for all λ .
For each λ , let g λ be any mapping in S (Aλ , µ; F) whose equivalence
class is uλ (it is not assumed that g λ is the restriction to Aλ of a mapping g
on A ). Let f0 = A0 → F be the mapping such that f0 Aλ = g λ for all λ
(disjointness); since g λ is µ-measurable in Aλ for every λ , it follows that f0
is µ-measurable in A0 (No. 10, Prop. 16). Let f1 : X → F be a measurable
extension of f0 to X (No. 10, Def. 8), and define f = f1 A ; then f is
µ-measurable (Def. 8), that is, f ∈ S (A, µ; F) , and, for all λ ,
fλ = f Aλ = (f1 A)Aλ = f1 Aλ = (f1 A0 )Aλ = f0 Aλ = g λ
IV.82, `. 9, 10.
PJ of λ ∈ L such that B ∩ Aλ 6= ∅ is countable (No. 9),
“ . . . the set
and |µ|(B) = |µ|(B ∩ Aλ ) ”
λ∈J
Since the family (Aλ )λ∈J is locally countable, J is countable (para-
graph following No. 9, Def. 7). With notations for A 0 and N as in the note
for `. 7, one then has
[
B = B ∩ A = (B ∩ A0 ) ∪ (B ∩ N) = B ∩ Aλ ∪ (B ∩ N) ,
λ∈J
where B ∩ N is negligible (No. 2, Prop. 5), whence the formula for |µ|(B)
(§4, No. 5, Prop. 9).
IV.82, `. 13–16.
“The image of T under ψ × ψ is . . . which proves the proposition.”
Let us pause to review product
Q uniform spaces. Let (X λ )λ∈L be a
family of uniform spaces, X = Xλ the product set, and prλ : X → Xλ
λ∈L
(λ ∈ L) the family of projection mappings. The product uniformity of X is
the initial uniform structure for the family (pr λ )λ∈L , that is, the coarsest
uniform structure on X that renders uniformly continuous every pr λ (GT,
II, §2, No. 6, Def. 4). If λ0 ∈ L and Wλ0 ⊂ Xλ0 × Xλ0 , then
Y
(prλ0 × prλ0 )−1 (Wλ0 ) = Wλ0 × (Xλ × Xλ ) .
λ6=λ0
If H runs over the set of finite subsets of L and if, for each λ ∈ H , W λ runs
over a fundamental system of entourages for X λ , then the sets (∗) form a
fundamental system of entourages for the product uniformity on X (GT, loc.
cit., No. 3, Prop. 4). {See also J.L. Kelley, General topology (Van Nostrand,
New York, 1955), p. 182.}
ψ −1 is uniformly continuous. Given a basic entourage W for S =
S (A, µ; F) , so that the set
is a basic entourage for S = S(A, µ; F) (see the note for IV.81, `. −5 to −1),
itQwill suffice to show that (ψ × ψ)(Ẇ) ⊃ V for some entourage V for
Sλ , where Sλ = S(Aλ , µ; F) for all λ ∈ L ; for then, writing
λ∈L
Y Y
Φ = ψ −1 × ψ −1 = (ψ × ψ)−1 : Sλ × Sλ → S × S ,
λ∈L λ∈L
Q
is an entourage for the product uniformity on Sλ . Let us show that
λ∈L
V ⊂ (ψ × ψ)(Ẇ) .
Let (u, v) ∈ V . Say u = (uλ )λ∈L , v = (vλ )λ∈L , so that (uλ , vλ ) ∈ Ẇλ
for all λ ∈ H (but Q no restriction on the u λ , vλ ∈ Sλ for λ ∈ L --- H ).
Since ψ : S → Sλ is bijective, there exist functions f, g ∈ S such
λ∈L
that ψ(f˙) = u and ψ(ġ) = v , and so (ψ × ψ) (f, ˙ ġ) = (u, v) . Thus, for
every λ ∈ L , uλ (resp. vλ ) is the equivalence class of fλ = f Aλ (resp.
gλ = g Aλ ). By the definition of V , for every λ ∈ H one has (u λ , vλ ) ∈ Ẇλ
and so
|µ|*({x ∈ Aλ : f (x), g(x) ∈ / V } 6 δ/2m .
Write
Mλ = {x ∈ Aλ : f (x), g(x) ∈ / V } for λ ∈ H ,
M = {x ∈ B : f (x), g(x) ∈ / V};
§5 measurable functions and sets INT IV.x155
as we wish to show. S
Now, M ⊂ B ⊂ A , and we have a partition A = N ∪ Aλ with N
λ∈L
locally negligible. Then
[
M=M∩B∩A=M∩B∩ N∪ Aλ
λ∈L
h [ i
= (M ∩ B ∩ N) ∪ M ∩ B ∩ Aλ
λ∈L
h [ i
= (M ∩ N) ∪ M ∩ B ∩ Aλ
λ∈J
h[ i
= (M ∩ N) ∪ M ∩ Aλ
λ∈J
h [ i h [ i
= (M ∩ N) ∪ M ∩ Aλ ∪ M ∩ Aλ ;
λ∈H λ∈J -- H
The first term on the right side is negligible; in the middle term, |µ|*(M λ ) 6
δ/2m for all λ ∈ H ; and the third term has outer measure 6 δ/2 by the
choice of H ; therefore
as we wished to show.
ψ is uniformly Qcontinuous. Given a basic entourage V for the prod-
uct uniformity on Sλ , it will suffice to find an entourage W for S =
λ∈L
S (A, µ; F) such that (ψ × ψ)(Ẇ) ⊂ V (GT, loc. cit.).
We can suppose that
Y Y
V = Ẇλ × Sλ × S λ ,
λ∈H λ∈L -- H
INT IV.x156 extension of a measure. l p spaces §5
since the outer measure of the right side is 6 δ (because (f, g) ∈ W ), the
outer measure of the left side is also 6 δ , in other words, (f λ , gλ ) ∈ Wλ ,
and so (uλ , vλ ) ∈ Ẇλ .
IV.82, `. −16 to −14.
“Since each An is the union of a negligible set and a sequence of compact
sets, we can suppose that the An are already compact and pairwise disjoint.”
S S
Say A = N∪ An with N locally negligible; replacing N by N --- An
n S Sn
we can suppose that N ∩ An = ∅ . Then, replacing An by An --- Ak
n k<n
(§4, No. 5) we can suppose that theSAn are pairwise disjoint. For each n ,
there exists a partition An = Nn ∪ Ank with Nn negligible and (Ank ) a
k
sequence of pairwise disjoint compact sets (§4, No. 6, Cor. 2 of Th. 4). Then
[ [
A = N ∪ Nn ∪ Ank ,
n n,k
S S
where N ∪ Nn is locally negligible, disjoint from Ank , and the Ank
n n,k
are pairwise disjoint compact sets.
IV.82, `. −14.
“Prop. 17 then allows us to suppose that A is compact.”
Writing S = S(A, µ; F) Q
and Sn = S(An , µ; F) , we know from Prop. 17
that the mapping ψ : S → Sn is an isomorphism of uniform spaces. If
n
§5 measurable functions and sets INT IV.x157
the compact
Q case is established, so that every S n is metrizable, it will follow
that Sn , hence S , is metrizable (GT, IX, §2, No. 4, Cor. 2 of Th. 1).
n
For, set
Vk = {(y, y 0 ) ∈ F × F : d(y, y 0 ) 6 δk } (k = 0, 1, 2, . . .)
Wk = W(Vk , B, δk ) ;
INT IV.x158 extension of a measure. l p spaces §5
that is,
|µ| {x ∈ B : d f (x), g(x) > δk } 6 δk .
(GT, II, §3, No. 1, second paragraph following Def. 2). Having chosen
n0 < . . . < nr−1 satisfying (∗) for k = 0, . . . , r − 1 , choose an index n r
such that (∗) is satisfied for k = r ; since (∗) continues to hold if n r is re-
placed by a larger integer, we can suppose that n r > nr−1 , which completes
the induction. In particular, for every k one has n k+1 > nk > nk and so
(fnk , fnk+1 ) ∈ Wk , thus
(∗∗) |µ| {x ∈ B : d fnk (x), fnk+1 (x) > δk } 6 δk
by the definition of Wk .
For each m > 0 , the sequence (fmn )n>0 is constructed inductively, as
follows.
m = 0 : Already defined is f0n = fn (n = 0, 1, 2, . . . ).
m = 1 : Set δ1,k = 1/2m+k+1 = 1/21+k+1 . Since (fn ) is Cauchy
in S , the argument for (∗∗) shows that (f n ) has a subsequence (f1n )n>0
(of course Cauchy) such that, writing
M1n = {x ∈ B : d f1n (x), f1,n+1 (x) > 1/21+n+1 }
(a measurable set, by No. 3, Th. 1 and No. 5, Prop. 7), one has |µ|(M 1n ) 6
1/21+n+1 . Note that
X X 1 1X 1 1
|µ|(M1n ) 6 = = · 1 = 1/21 .
21+n+1 2 2n+1 2
n>0 n>0 n>0
§5 measurable functions and sets INT IV.x159
∞
X
|µ|(Mm ) 6 |µ|(Mmn ) 6 1/2m ”
n=0
It is intended here that m > 1 ; the inequality then follows from (∗∗∗)
of the preceding note.
By the convention established at the end of theTpreceding note,
M0 = B ; the above inequalities assure that the set N = Mm is negligi-
m>0
ble, irrespective of the value of |µ|(M 0 ) .
IV.83, `. 4, 5.
“ . . . and, for every x ∈ B --- Mm , we have d(fmn (x), fm,n+p (x)) 6
1/2m+n for all n > 0 and all p > 0 ”
It is assumed here that m > 1 . Thus,
[ \
x ∈ B --- Mm = B --- Mmn = (B --- Mmn )
n>0 n>0
means that d fmn (x), fm,n+1 (x) 6 1/2m+n+1 for all n > 0 . It then follows
from the triangle inequality that
p−1 p−1
X X 1
d fmn (x), fm,n+p (x) 6 d fm,n+k (x), fm,n+k+1 (x) 6
2m+n+k+1
k=0 k=0
p−1
X
1 1 1
= < · 1.
2m+n 2k+1 2m+n
k=0
INT IV.x160 extension of a measure. l p spaces §5
IV.83, `. 8–10.
“ . . . if x ∈ B --- N , there exists an index m such that x ∈
/ M m , which
proves that the sequence (gn (x)) is a Cauchy sequence in F .”
/ N we know that there exists an index m > 0 such that
Since x ∈
x∈ / Mm . But x ∈ B = M0 , so m > 1 ; by the preceding note, (fmn (x))n>0
is Cauchy, hence so its its subsequence (g n (x))n>m .
IV.83, `. 11–15.
“If now B is the union of a sequence (B m ) . . . where Pm is negligi-
ble.”
The notations for m = 0 can be handled as follows. We can suppose
that B0 = ∅ ; setting P0 = ∅ and g0n = fn for all n > 0 , the sequence
(gn0 )n>0 is vacuously pointwise Cauchy in B0 --- P0 . The recursive defini-
tion then proceeds as in the text.
IV.83, `. −11, −10.
“ . . . we are thus reduced to proving the proposition when A is integrable ”
A product of nonempty uniform spaces is complete if and only if every
factor is complete (GT, II, §3, No. 5, Prop. 10).
IV.83, `. −9, −8.
“ . . . for every Cauchy sequence (fn ) in S (A, µ; F) there exists a sub-
sequence (fnk ) that is convergent in A --- N , where N is negligible ”
This is immediate from Lemma 4 and the completeness of F .
But . . . where does the Cauchy sequence (f n ) come from? We are to
show that S = S(A, µ; F) is complete, thus one begins with a Cauchy se-
quence (un ) in S . Say fn ∈ S = S (A, µ; F) with un = f˙n (in the
notation of Prop. 17); we need to know that (f n ) is Cauchy in S .
Lemma. Let A be any measurable subset of X , and W = W(V, B, δ) ,
where V is an entourage for F , B is an integrable subset of A , and δ > 0 ,
and let
Ẇ = {(f˙, ġ) : (f, g) ∈ W } .
Then, for a pair of functions f, g in S , one has
˙ ġ) ∈ Ẇ .
(f, g) ∈ W ⇔ (f,
⇒ : By the definition of Ẇ .
⇐ : Suppose (f˙, ġ) ∈ Ẇ. This means that there exists a pair (h, k) ∈ W
such that (f˙, ġ) = (ḣ, k̇), that is, locally almost everywhere, f = h and g = k.
Let M be a locally negligible subset of A such that f = h and g = k
everywhere in A --- M . Since B is integrable, B ∩ M is negligible (No. 2,
§5 measurable functions and sets INT IV.x161
(um , un ) ∈ Ẇ ⇔ (fm , fn ) ∈ W ,
thus the assertion follows from GT, II, §3, No. 1, second paragraph following
Def. 2.
IV.83, `. −7, −6.
“ . . . is then µ-measurable ”
By 1◦ of the generalization of Egoroff’s Theorem proved in the note for
IV.80, `. −17 to −14 (item “No. 4, Th. 2 0 in that note).
IV.83, `. −6, −5.
“ . . . it follows from the extension of Egoroff’s theorem mentioned in
No. 10 that the sequence (fnk ) converges in measure to f in A .”
By 2◦ of the generalization cited in the preceding note, if K ⊂ A
is compact and δ > 0 , there exists a compact set K 1 ⊂ K such that
|µ|(K --- K1 ) 6 δ and fnk → f uniformly on K1 (we do not need the
continuity of the restrictions to K 1 established there). We are to show that
(fnk ) converges to f in the uniform space S = S (A, µ; F) .
Given a basic entourage W for S , it will suffice to find an index k 0
such that
k > k0 ⇒ (fnk , f ) ∈ W
(GT, II, §3, No. 1, second paragraph following Def. 2). We can suppose that
W = W(V, K, δ) ,
INT IV.x162 extension of a measure. l p spaces §5
whence
|µ| {x ∈ K : d fnk (x), f (x) > ε } 6 |µ|(K --- K1 ) 6 δ
IV.84, `. 1.
“Corollary.”
This is a corollary of Lemma 4, using parts of the proof of Prop. 19.
IV.84, `. 10, 11.
“The assertion follows at once from the extension of Egoroff’s theorem
mentioned in No. 10.”
One is assuming that there exists a locally negligible set N ⊂ A such
that fn (x) → f (x) for all x ∈ A --- N . The argument that f is measurable
is given in the note for IV.83, `. −7, −6 (completeness of F is not needed
since the limits f (x) for x ∈ A --- N are assumed to exist); and the argument
that fn → f in the uniform space S (A, µ; F) (that is, in measure) is given
in the note for IV.83, `. −6, −5 (valid for A measurable but not necessarily
integrable; the role of integrability arises in the proof of Prop. 19 because of
the citation of Lemma 4 in order to obtain a limit function f , whereas f
is given here in advance).
IV.84, `. 15.
“ . . . f 0 is a µ-measurable mapping of B --- N into Fb”
Write gk = fnk B --- N ; we know that gk : B --- N → F is µ-measurable
( fnk has a measurable extension to X , therefore so does g k ; see item c00 )
in the note for IV.79, `. 3,4). Let i : F → F b be the canonical injection,
b
and gk = i ◦ gk : B --- N → F ; since i is continuous, gk0 is µ-measurable by
0
the generalization of No. 3, Th. 1 (item No. 3, Th. 1 0 in the note for IV.80,
`. −17 to −14). One has gk0 → f 0 pointwise (on all of B --- N ), therefore f 0
is µ-measurable by 1◦ of the generalization of Egoroff’s theorem (loc. cit.,
item No. 4, Th. 2 0 ).
IV.84, `. 15, 16.
“ . . . the sequence (fn ) converges in measure to f 0 in B --- N by (i) ”
Lemma. If C is a measurable subset of A , then the mapping f 7→ f C
of S (A, µ; F) into S (C, µ; F) is uniformly continuous.
For, f C is measurable (since f has a measurable extension to X , so
does f C ). If K is a compact subset of C then it is also a compact subset
of A ; let V be a basic entourage for F , let δ > 0 , and write
whence obviously
(f, g) ∈ WA ⇔ f C, g C ∈ WC .
b .
by (i) of the present Corollary, fn0 k → f 0 (in measure) in SC0 = S (C, µ; F)
On the other hand, since fn → f in measure, hence fnk → f in
measure, we know that fnk C → f C in measure by the Lemma; and since
i:F→F b is uniformly continuous, it follows that i ◦ (f n C) → i ◦ (f C) in
k
uk → u0 in S0C ; since
S0C is Hausdorff (IV.81, `. −7 to −1), it follows that
u = u0 , thus i◦(f C) = f 0 locally almost everywhere
in C , and since C is a
countable union of integrable sets, in fact i ◦ (f C) = f 0 almost everywhere
in C . But fnk (x) → f (x) for every x ∈ C , and f 0 (x) = f (x) for almost
every x ∈ C , therefore fnk (x) → f (x) for almost every x ∈ C . Since
B = C ∪ N and N is negligible, we conclude that f nk (x) → f (x) for almost
every x ∈ B ; this proves the assertion (ii) of the Corollary.
As for the original assertion, since f n → f in
measure, the
argument in
the paragraph before the last shows that i ◦ (f n C) → i ◦ (f C) in measure;
but we now know that i ◦ (f C) = f 0 almost everywhere in C , therefore
i ◦ (fn C) → f 0 in measure.
This is the sense in which the original assertion holds, avoiding the abuse of
notation where the fn take values in F whereas f 0 takes its values in F b.
§5 measurable functions and sets INT IV.x165
IV.84, `. −10.
“ T(B, δ) + T(B, δ) ⊂ T(B, 2δ) ”
Let f , g ∈ S (A, µ; F) . One has
in other words
(∗∗) |µ| {x ∈ B : |f (x)| > |α|δ } 6 δ ;
Since the sets W(Vδ , B, δ) form a fundamental system of entourages for the
uniformity of S , the sets
{f ∈ S : (f , g) ∈ W(Vδ , B, δ) }
for some integrable set B ⊂ A and some number δ 0 > 0 . Say T(B, δ) ⊂ U .
Then
2T(B, δ/2) ⊂ T(B, δ/2) + T(B, δ/2) ⊂ T(B, δ) ;
thus 2T(B, δ/2) ⊂ T(B, δ) for every δ > 0 , whence, by induction,
Choose n so large that 2n |λ| > 1 ; since T(B, δ/2n ) is balanced, one has
(2n |λ|)−1 T(B, δ/2n ) ⊂ T(B, δ/2n ) , whence
(∗) {f ∈ L p : Np (f ) 6 ρ } ⊂ T(B, δ) .
and we may write |µ|(C) . To assure that f ∈ T(B, δ) , in other words that
|µ|(B ∩ C) 6 δ , it suffices to assure that |µ|(C) 6 δ . But δ p |µ|(C) 6 ρp ,
that is, |µ|(C) 6 (ρ/δ)p ; so f ∈ T(B, δ) is assured if (ρ/δ)p = δ , that is,
ρp = δ p+1 , ρ = δ 1+1/p .
To recapitulate: given a neighborhood T(B, δ) of 0 in S , where
B ⊂ X is integrable and δ > 0 , setting ρ = δ 1+1/p assures that (∗) is
satisfied.
IV.85, `. 5–7.
“In view of (iii), it suffices to show for example that L F1 is dense in SF ,
since by definition K (X; F) is dense in L F1 for the topology of convergence
in mean.”
The argument is easily modified to cover the case of L Fp ; borrowing
from the next note, let us assume already established that L Fp is dense
in SF .
Let U be a nonempty open set in SF ; we are to show that
U ∩ K (X; F) 6= ∅ . By the foregoing assumption, U ∩ L Fp is nonempty,
and by (iii) it is an open set in LFp , therefore
U ∩ K (X; F) = U ∩ LFp ∩ K (X; F) = (U ∩ LFp ) ∩ K (X; F) 6= ∅
by the density of K (X; F) in LFp (§3, No. 4, Def. 2).
IV.85, `. 7–12.
“Now, let f be any element of SF . . . and obviously f −g ∈ T(B, δ) . ”
The objective is to prove that LFp is dense in SF . Given f ∈ SF and
a set T(B, δ) , so that f + T(B,
δ) is a basic neighborhood of f in S F , we
are to show that f + T(B, δ) ∩ LFp 6= ∅ , establishing that f belongs to the
closure of LFp in SF ; thus we seek a function g ∈ LFp such g ∈ f +T(B, δ) .
As shown in the proof of (i), there exists a number n > 0 such that
the (integrable) set C = {x ∈ B : |f (x)| > n } satisfies |µ|(C) 6 δ . Then
|f | 6 1/n on B --- C , thus the function g : X → F defined by g = ϕ B -- C f is
measurable (No. 3, remarks following Cor. 5 of Th. 1) and |g| 6 (1/n)ϕ B -- C .
It follows that g ∈ LFp , since
1/p
Np (g) = Np (|g|) 6 (1/n)Np (ϕB -- C ) = (1/n) |µ|(B --- C) < +∞
(No. 6, Th. 5).
INT IV.x170 extension of a measure. l p spaces §5
(ii) for every ε > 0 there exists a compact set K ⊂ X such that
Z
ϕX -- K f dµ 6 ε .
whence (ii). {We will see later that this number is ρ(X) .} ♦
Definition of the measure ν = f · µ .
For each h ∈ K = K (X; C) the function hf is µ-measurable (No. 3,
remark following Cor. 5 of Th. 1) and |hf | 6 khk f , whence
defines a linear form on K such that h > 0 ⇒ ν(h) > 0 , that is, a positive
measure ν on X (Ch. III, §1, No. 5, Th. 1). We write ν = f · µ , a notation
consistent with Ch. III, §1, No. 4 when f is continuous. Moreover,
Z
|ν(h)| 6 ν(|h|) 6 khk f dµ ,
thus ν is a bounded measure on X (loc. cit., No. 8, remark following Def. 3).
It follows that X is ν-integrable (§4, No. 7, Prop. 12) and that kνk = ν(X) =
sup{ν(K) : K ⊂ X compact } (§4, No. 6, Cor. 1 of Th. 4).
Note that a compact set K in X is a Gδ if and only if there exists a
decreasing sequence (hn ) in K+ (X)T such that hn ↓ ϕK : for, if (Un ) is a
sequence of open sets such that K = Un , let kn ∈ K+ (X) with kn = 1 on K
n
and Supp(kn ) ⊂ Un (III, §1, No. 2, Lemma 1) and let hn = inf(k1 , . . . , kn ) ;
conversely, if hn ↓ ϕK then for each n the set Kn = {x : hn (x) > 1 } =
T
∞ T
∞
{x : hn (x) > 1 − 1/m } is a compact Gδ , and Kn = K .
m=1 n=1
DeferringR any hypothesis on X , we observe that if K is a compact G δ
then ν(K) = ϕK f dµ ; for, ifR(hn ) is a decreasing sequence in K+ (X) such
that hn ↓ ϕK , thenR ν(hn ) ↓ RϕK dν = ν(K) (§4, No. 3, Prop. 4) whereas
hn f ↓ ϕK f implies hn f dµ ↓ ϕK f dµ , thus
Z Z
ν(K) = lim ν(hn ) = lim hn f dµ = ϕK f dµ .
n n
R
whence
R ρ(A) 6 f dµ < +∞ for all such A . In particular,
R ρ(K) =
ϕK f dµ for every compact set K ⊂ X , whence ρ(K) 6 f dµ . For every
open set U in X we know that
R
Thus the condition (ii) is clearly equivalent to ρ(X) = f dµ . I do not see
how to derive the condition (ii) from (2); but we established (ii) at the outset
(Prop. A), therefore
Z Z
(3) ϕX dρ = ρ(X) = f dµ .
On the other hand, the condition (i) does follow from the properties
of ρ :
Proposition B. — The condition (i) is satisfied without restrictions on
the locally compact space X .
Proof. Every µ-negligible set A in X is ν-negligible; for, such a set
A is µ-integrable, hence it is ρ-integrable and ρ(A) = 0 by (1). Since ρ
is a finite measure, it follows that, given any ε > 0 , there exists a δ > 0
such thatR if A is a µ-integrable set with µ(A) 6 δ then ρ(A) 6 ε ; since
ρ(A) = ϕA f dµ , this will prove (i).
The argument for set-measures (cf. Halmos, op. cit., p. 125, §30, Th. B)
can be adapted as follows. Assume to the contrary that for some ε > 0 ,
no such δ exists; that is, for every δ > 0 there exists a µ-integrable set A
such that µ(A) 6 δ but ρ(A) > ε . For every positive integer n , choose a
µ-integrable set An such that µ(An ) 6 1/2n and ρ(An ) > ε , and let A be
the set of all x ∈ X that belong to An for infinitely many n , that is,
∞
\ [
A= Ak ,
n=1 k>n
§5 measurable functions and sets INT IV.x175
S
∞
which is a µ-measurable set. Writing B n = Ak , Bn is µ-measurable
k=n
and
∞
X ∞
X 1 1
µ*(Bn ) 6 µ(Ak ) 6 k
= n−1 ,
2 2
k=n k=n
therefore Bn is µ-integrable (No. 6, Cor. 1 of Th. 5); since B n ⊃ A , it
follows that A is µ-integrable and µ(A) 6 µ(B n ) → 0 , whence A is
µ-negligible, so by assumption ρ(A) = 0 . But B n ↓ A and ρ is finite,
therefore ρ(Bn ) → ρ(A) ; since Bn ⊃ An , one has ρ(Bn ) > ε for all n ,
whence ρ(A) > ε > 0 , a contradiction. ♦
Finally,
Proposition C. If every compact set in X is a G δ , then ν = ρ .
Proof. As observed at the end of the discussion of ν = f · µ ,
Z
(4) ν(K) = ϕK f dµ for all compact sets K ⊂ X .
λ(A) = µ*(A) (A ∈ M) .
∞
X
λ(A) 6 λ(An )
n=1
S
n
For, writing Bn = Ak , one has
k=1
n
X ∞
X
λ(A) = sup λ(Bn ) = sup λ(Ak ) = λ(An )
n n
k=1 n=1
R = {A ∈ M : λ(A) < +∞ } ;
in which case λ(A) = sup λ(An ) (§4, No. 5, Prop. 8). The restriction of λ
n
to R is a finite set-measure on the ring R [MT, §5, pp. 30, 31].
Borel sets and H-Borel sets.
Let B be the tribe (σ-algebra) of Borel sets in X , that is, the tribe
generated by the closed subsets of X (GT, IX, §6, No. 3,
Def. 4). Of course
B ⊂ M (No. 4, Cor. 3 of Th. 2), and the restriction λ B is a set-measure
on B . However, the set of ‘Borel sets’ in the sense of Halmos [MT, §51,
INT IV.x178 extension of a measure. l p spaces §5
and this follows from the µ-integrability of U (§4, No. 6, Cor. 1 of Th. 4).
{Incidentally, if U ⊂ K1 with U open and K1 compact, then necessarily
U ∈ B1 ; indeed, it suffices that U can be covered by a sequence of compact
sets [MT, §51, Th. A, p. 219].}
To summarize:
Proposition A. If µ is a positive measure on the locally compact space X ,
then the restriction of µ* to the σ-ring generated by the compact subsets
of X is a regular Borel measure on X in the sense of Halmos [MT].
Conversely:
Proposition B. If λ is a regular Borel measure in the sense of [MT]
(defined on the σ-ring generated by the compact subsets of X ), then there
exists a unique measure µ on X (in the sense of Integration) such that
λ(K) = µ(K) for every compact set K in X .
§5 measurable functions and sets INT IV.x179
for every µ-integrable set A ⊂ X (see the first note for `. 19-27). Write
M for the tribe of all µ-measurable sets A ⊂ X , and K for the set of all
compact sets K ⊂ X ; since µ is assumed to be bounded, every A ∈ M
is µ-integrable, hence (∗) holds for all A ∈ M . Moreover, since |f | 6 a
µ-almost everywhere for each f ∈ H , one has
and Z
|f |p ϕX -- K d|µ| = ρf (X --- K) 6 ap |µ|(X --- K) 6 ε,
Note first of all that only compact sets K of measure > 0 are in play:
assuming Np (f ) > 0 , for ε sufficiently small the K in property (ii) of Def. 10
will have to be non-negligible.
Where are we going? In the notations of Prop. 20, we have
U ∩ (LFp × LFp ) ⊂ Up ,
(∗) W ∩ (H × H) ⊂ U ∩ (H × H) ,
W = W(V, K, δ) ,
we thus have
W = {(f , g) ∈ S × S : |µ|* M(f , g) 6 δ } .
by the definition of M in `. 4.
IV.86, `. −12 to −6.
“Lemma 5.”
One notes that the condition (i) reveals the Lemma as an analog, in lo-
cally compact spaces, of Lebesgue’s criterion for Riemann-integrability (see,
for example, §11.4, pp. 211–214 of my book A first course in real analysis,
Springer). Specialized to the case of a bounded function f : [a, b] → R
and Lebesgue measure µ on [a, b] , the lemma says (assuming Lebesgue’s
criterion) that f is Riemann-integrable if and only if, for every ε > 0 , there
exist continuous functions g, h : [a, b] → R , with h > 0 , such that
Z
|f − g| 6 h 6 2 sup |f | and h dµ 6 ε .
|g1 a1 + · · · + gn an | 6 M = sup |f |
INT IV.x182 extension of a measure. l p spaces §5
on X , whence k 6 2M .”
I am indebted to Professor Dixmier for crucial help with the proof of
Lemma 5.
For technical reasons, let us assume the g i and R ai so chosen that,
on writing k = |f − g1 a1 − · · · − gn an | , one has k dµ 6 ε/4 (instead
of ε/2 ). When the gi are replaced by Rsuitable functions g i0 , the function
k 0 = |f − gi0 a1 − · · · − gn0 an | will satisfy k 0 dµ 6 ε/2 .
R Write p = g1 a1 + · · · + gn an . Thus p : X → F , k = |f − p| and
k dµ 6 ε/4 . We seek a continuous function ϕ : X → R , 0 6 ϕ 6 1 , such
that, on defining
gi0 = ϕgi , p0 = ϕp = g10 a1 + · · · + gn0 an , k 0 = |f − p0 | ,
R
one has |p0 | 6 M and k 0 dµ 6 ε/2 . We define
M if |p(x)| > M
ϕ(x) = |p(x)|
1 if |p(x)| 6 M .
The continuity of ϕ is a special case of the following proposition (with
g = |p| ):
Proposition. Let X be any topological space, g : X → R a continuous
function > 0 , and M a real number > 0 . Let
U = {x ∈ X : g(x) > M }
and define ϕ : X → [0, 1] by
M if x ∈ U
ϕ(x) = g(x)
1 if x ∈ X --- U .
Then ϕ is continuous.
Proof. By the continuity of g , U is open and g > M on U . On the
other hand, X --- U = {x ∈ X : g(x) 6 M } , therefore
(∗) g(x) = M on U ∩ (X --- U) = U --- U .
(Incidentally, U --- U = U ∩ X --- U = ∂U (the boundary of U .)
Now, U , X --- U is a closed covering of X , so it will suffice to show
that the restrictions ϕU and ϕX --- U are continuous (GT, I, §3, No. 2,
Prop. 4). Indeed, ϕX --- U is the constant function 1 . On the other hand,
M
ϕ(x) = on U = U ∪ (U --- U) ;
g(x)
§5 measurable functions and sets INT IV.x183
M M
ϕ(x) = < 1 and p0 (x) = ϕ(x)p(x) = · p(x) ,
|p(x)| |p(x)|
thus
M
|p(x) − p0 (x)| = p(x) − p(x)
|p(x)|
M M
= 1− p(x) = 1 − |p(x)|
|p(x)| |p(x)|
= |p(x)| − M = |[p(x) − f (x)] + f (x)| − M
6 |p(x) − f (x)| + |f (x)| − M
6 |p(x) − f (x)| + M − M = |p(x) − f (x)| ,
therefore
for all x ∈ X .
IV.87, `. 6.
“Then 2M > l > k on X , and l = k on X --- N0 ”
We know that k : X → [0, 2M] . By definition (GT, IV, §5, No. 6)
l(x) = inf sup k(y) ,
V y∈V
where V runs over the filter of neighborhoods of x (or any base thereof),
therefore l : X → [0, 2M] , thus l 6 2M ; and l(x) > k(x) because x ∈ V
for all V .
If x ∈ X --- N0 then k is continuous at x , hence is upper semi-continuous
at x (GT, IV, §6, No. 2, comment following Def. 1), hence
(loc. cit., ‘dual’ of Prop. 3). The function l is upper semi-continuous (loc.
cit., the ‘dual’ of Prop. 4), and is called the upper semi-continuous regular-
ization of k .
IV.87, `. 7. R
“ . . . therefore l dµ 6 ε/2 .”
Since l = k almost everywhere, so that µ*(l) = µ*(k) < +∞ , it follows
1 1
that l ∈ FR R(§3, No.R3), hence l ∈ LR (§3, No. 4, third paragraph following
Def. 2) and l dµ = k dµ 6 ε/2 .
§5 measurable functions and sets INT IV.x185
IV.87, `. 7–9.
“ l is bounded and upper semi-continuous, hence is the lower envelope
of the set of bounded continuous functions > l .”
Of course 0 6 l 6 2M , and the upper semi-continuity of l was observed
in the note before the last. Then 2M − l is > 0 and lower semi-continuous,
that is, in the notation of §1, No. 1, 2M − l ∈ I + , therefore 2M − l is the
upper envelope of the functions g ∈ K + such that g 6 2M − l (loc. cit.,
Lemma); consequently l − 2M is the lower envelope of the functions −g ,
and so
l = (l − 2M) + 2M
is the lower envelope of the functions −g + 2M = 2M − g , which are indeed
bounded and continuous. Write
l 6 inf H 6 inf H0 = l ,
moreover,
|f − g1 a1 − · · · − gn an | = |f − p| = k 6 l 6 h 6 2M ,
IV.87, `. 14–15.
“For every x ∈ X , ω(x) is the oscillation of f − g 1 a1 − · · · − gn an
at x , therefore ω(x) 6 2h(x) .”
Let us write ωf (x) = ω(x) for the oscillation of f at x , that is (GT,
IX, §2, No. 3),
ωf (x) = inf δ f (V) ,
V
is the diameter of f (V) . Note that V 7→ δ f (V) is a decreasing function
of V , and, since |f | 6 M , one has 0 6 ωf 6 2M .
Why are we considering ω ?; because f is continuous at x if and only if
ω(x) = 0 (GT, loc. cit., comment following the proof of Prop. 4), therefore
{x ∈ X : ω(x) > 0 } = N
whence
(∗) δ (f + g)(V) 6 δ f (V) + δ g(V) .
and since they are decreasing functions of V , passage to the limit yields the
desired inequality.
(2) If, moreover, g is continuous, then ω g = 0 and so ωf +g 6 ωf ;
similarly ωf = ω(f +g)+(−g) 6 ωf +g .
(3) See GT, loc. cit., Prop. 4. ♦
Returning to the proof of (ii) ⇒ (i), writing g = −g 1 a1 − · · · − gn an
and w = f + g , we have |w| 6 h . It follows that ω w 6 2h . For, if V is a
neighborhood of x ∈ X then, for all y, y 0 ∈ V one has
but
h(y) = [h(y) − h(x)] + h(x) 6 |h(y) − h(x)| + h(x) 6 δ h(V) + h(x)
and similarly h(y 0 ) 6 δ h(V) + h(x) . Thus
|w(y) − w(y 0 )| 6 δ h(V) + h(x) + δ h(V) + h(x)
thus ωw 6 2h .
Finally, ωf = ωf +g = ωw 6 2h ; since ωf is upper semi-continuous,
hence measurable (No. 5, Cor. of Prop. 8), and h is integrable, it follows
that ωf is integrable and
Z Z
ωf dµ 6 2 h dµ 6 2ε .
0 6 k 6 2M , k = 0 on K , k = 2M on X --- K0 .
0 6 h 6 h0 6 2M , 0 6 h0 − h 6 2M − h 6 2M ,
h0 = h on K , and h0 = 2M on X --- K0 .
IV.87, `. −3,
R −2.
“Then (h0 − h) dµ 6 2Mµ*(X --- K) 6 2Mε .”
With notations as in the preceding note,
0 6 h0 − h = (h0 − h)ϕX -- K 6 2M · ϕX -- K ,
R
whence 0 6 (h0 − h) dµ 6 2Mµ(X --- K) 6 2Mε .
IV.87, `. −2, −1.
“ . . . h0 = h1 + 2M , where h1 ∈ K (X) .”
With the foregoing notations, set h 1 = h0 − 2M ; since h0 − 2M = 0 on
X --- K0 , one has h1 ∈ K (X) (and h1 6 0 ).
IV.88,R`. 1, 2. R R R
“ h0 dν = h1 dν +2Mkνk tends to h1 dµ+2Mkµk = h0 dµ with
respect to B .”
R R
h1 dν → h1 dµ because h1 ∈ K (X) and ν → µ vaguely, whereas
kνk → kµk by hypothesis, and
Z Z Z Z
0
h dν = h1 dν + 2M 1 dν = h1 dν + 2Mkνk
6 2(M + 2)ε ”
R R
where |f −p| 6 h and p dν − p dµ 6 ε by the preceding note, therefore
Z Z Z Z
f dν − f dµ 6 h dν + ε + h dµ 6 2(M + 1)ε + ε + ε .
IV.89, `. 3.
“ µn (A) tends to µ(A) .”
One proposes to apply Prop. 22 to the function f = ϕ A . The space X
is the disjoint union of the interior, exterior and boundary (‘frontier’) of A ,
◦
X = A ∪ ( { A)◦ ∪ ∂A
◦
(GT, I, §1, No. 6). Since f is constant on each of the open sets A and
( { A)◦ , it is continuous at those points, whereas ∂A = A ∩ { A shows that
it is discontinuous at the points of ∂A , therefore the set of discontinuities
of f is precisely ∂A , which is by assumption R negligible.
R Thus f satisfies
the conditions of Prop. 22, whence µ n (A) = f dµn → f dµ = µ(A) .
IV.89, `. 4, 5.
“ if pn denotes the number of integers k ∈ [0, n] such that e 2iπkθ ∈ A ,
then n−1 pn tends to µ(A ) as n tends to +∞ .”
For every integer k > 0 ,
Z (
1 if e2iπkθ ∈ A
νk (A) = ϕA dνk =
0 otherwise,
therefore
1 pn
µn (A) = ν0 (A) + · · · + νn (A) = ,
n+1 n+1
thus
pn n+1 pn n+1
= · = · µn (A) → µ(A) .
n n n+1 n
A stunning application of Prop. 22.
IV.89, `. −4 to −1.
“Corollary.”
A variation on the argument shifts the burden from the measure to the
function:
Proposition. Let µ be a positive measure, f : X → F an integrable
function that is equal to zero outside some integrable
R subset A of X . If C
is the closed convex envelope of f (X) , then f dµ ∈ µ(A) · C .
§6 convexity inequalities INT IV.x191
R
Proof. If µ(A) = 0 then f is negligible and f dµ = 0 ∈ 0 · C .
If µ(A) > 0 then g = ϕA meets the requirements of Th. 1, f g = f , and
R R
f dµ f g dµ
= R ∈ C. ♦
µ(A) g dµ
(X --- A) ∩ (X --- B) = ∅ ,
α 6 β & γ 6 δ ⇒ αγ 6 βδ .
IV.92, `. 5–7.
“ . . . for every integer m , there exist a locally negligible set H m and an
integer n0 such that |f (x) − fn (x)| 6 1/m for every integer n > n0 and
every x ∈ / Hm ”
For each m , let n0 (m) be an index such that
thus, for each n > n0 (m) , there exists a locally negligible set H mn such
that
IV.92, `. 10.
“ . . . the converse is immediate.”
That is, if there exists a locally negligible set H such that f n → f
uniformly on X --- H , then N∞ (fn − f ) → 0 . For,
Suppose now that x ∈ X --- A and |gk1 (x)| 6 N∞ (gk1 ) . Then, for all
n > k1 , one has
and passage to the limit yields |g(x)| 6 0+1+N ∞ (gk1 ) . Thus g is bounded
by 1 + N∞ (gk1 ) on the indicated set.
IV.92, `. −7.
“ . . . g belongs to LF∞ . ”
The set N = {x ∈ X : |gk1 (x)| > N∞ (gk1 ) } is locally negligible, so
g = ϕ { N g locally almost everywhere, and the measurable function ϕ { N g
is bounded by the preceding note.
IV.93, `. 2–4.
“ . . . if there exists a continuous function f with negligible compact
support and not identically zero ”
Such a function exists if and only if Supp µ 6= X :
If : Suppose Supp µ 6= X . Then U = X --- Supp µ is a nonempty
negligible open set (§2, No. 2, Prop. 5). Choose a nonzero function f ∈
K+ (X) such that Supp f ⊂ U , and a nonzero vector a ∈ F , and define
f = f a ; then f ∈ K (X; F) , f 6= 0 , and the set Supp f = Supp f ⊂ U is
negligible. Moreover, |f | = |f | · |a| 6 kf k · |a| < +∞ , thus f ∈ C b (X; F)
(Ch. III, §1, No. 2).
Only if : Suppose f ∈ K (X; F) with f 6= 0 and Supp f negligible.
Then the open set V = {x : f (x) 6= 0 } ⊂ Supp f is nonempty and negligible,
hence V ⊂ X --- Supp µ by the cited Prop. 5, thus Supp µ 6= X . ♦
For such a function f , one has |f | = 0 almost everywhere, hence locally
almost everywhere, therefore M∞ (|f |) = 0 ; thus N∞ (f ) = 0 < kf k . This
proves the implication
IV.93, `. 7.
“ . . . which shows that N∞ (f ) = kf k . ”
We are assuming that Supp µ = X and f ∈ C b (X; F) . At any rate,
N∞ (f ) 6 kf k . Supposing to the contrary that M ∞ (|f |) = N∞ (f ) < kf k ,
choose α so that M∞ (|f |) < α < kf k and let U = {x : |f (x)| > α } ,
a nonempty open set. We know from the definition of M ∞ that |f (x)| 6 α
locally almost everywhere, therefore U is locally negligible. Let V be a
nonempty open set with V compact and V ⊂ U ; then V = V ∩ U is
INT IV.x196 extension of a measure. l p spaces §6
negligible (§5, No. 2, Prop. 5), thus V is a nonempty negligible open set,
contradicting X --- Supp µ = ∅ (§2, No. 2, Prop. 5).
IV.93, `. 11–13.
“ . . . but its canonical image in L∞ ∞
F is a closed subspace of LF (which
can moreover be identified with C b (X; F) in the case contemplated).”
For any measure µ (with no asssumption about its support) every f ∈
C b (X; F) belongs to LF∞ and N∞ (f ) 6 kf k , thus the canonical injection
C b (X; F) → LF∞ is continuous for the norm topology on C b (X; F) and the
topology on LF∞ defined by the semi-norm N∞ . In turn, the quotient
mapping LF∞ → L∞ ∞
F defined by f 7→ ḟ is continuous, where LF is a
Banach space equipped with the norm k ḟ k∞ = N∞ (f ) (also written N∞ (ḟ ) )
(Prop. 2). The composite mapping C b (X; F) → L∞ F , defined by f 7→ ḟ ,
satisfies kḟ k∞ 6 kf k hence is continuous but in general not injective. Its
image in L∞ b
F may not be complete: if fn ∈ C (X; F) is a sequence such that
∞
(ḟn ) is Cauchy in LF , there exists an f ∈ LF∞ with kḟn − ḟ k∞ → 0 , but
it is not assured that there exists a g ∈ C b (X; F) such that f = g locally
almost everywhere.
However, when Supp µ = X (the case contemplated), one has kf k =
N∞ (f ) = kḟ k∞ , therefore the image of C b (X; F) in L∞ F is a complete,
∞
hence closed, linear subspace of LF . Thus, the mapping C b (X; F) → L∞ F
is a linear injection such that the norm (and topology) induced by L ∞ F on the
image coincide with those of C b (X; F) , whence the proposed identification
C b (X; F) ⊂ L∞ F .
IV.93, `. 15–17.
“This implies that the space K (X; F) of mappings of X into F , con-
tinuous with compact support, is in general not dense in L ∞ F ”
If the space K (X; F) ⊂ C b (X; F) were dense in L∞F then, since C b (X; F)
is closed in L∞ b ∞
F , it would follow that C (X; F) = LF .
W = W(V, K, δ) ,
V = {(a, b) ∈ F × F : |a − b| 6 ε }
therefore
W = (f , g) ∈ LF∞ × LF∞ : |µ|* {x ∈ K : |f (x) − g(x)| > ε} 6 δ .
Let
V = {(f , g) ∈ LF∞ × LF∞ : N∞ (f − g) < ε }
(the same ε ); it suffices to show that V ⊂ W , that is, for f , g ∈ L F∞ ,
N∞ (f − g) < ε ⇒ |µ|* {x ∈ K : |f (x) − g(x)| > ε} 6 δ .
is locally negligible (see the note for IV.90, `. 13, 14), therefore the set
Th. 1 or TVS, I, §1, No. 6, Prop. 5). This observation is useful for the proof
of Cor. 3 given below.
IV.94, `. −13.
“For, |hz, z0 i| 6 |z| · |z0 | . ”
The inequality, noted in the proof of TVS, IV, §1, No. 3, Prop. 8, assures
the continuity of the bilinear form (z, z 0 ) 7→ hz, z0 i (GT, IX, §3, No. 5, Th. 1).
For the definition of |z0 | , see GT, X, §3, No. 2, discussion preceding
Prop. 6; TVS, IV, §2, No. 4; or the final section of TVS (“Summary of some
important properties of Banach spaces”, TVS p. 355).
IV.94, `. −9 to −5.
“Corollary 3. — Let µ be a positive measure on X , F a real (resp.
complex ) Hilbert space. On the space L 2F , the symmetric (resp. Hermitian)
form Z
e e) 7→ hf , gi dµ
(f , g
hence Ψ is continuous (GT, IX, §3, No. 5, Th. 1), therefore so is Φ , and
|Φ(a, b)| 6 |a| · |b| . Since the proof of Cor. 1 does not require that Φ be
bilinear, we conclude that if f , g ∈ L F2 then Φ(f , g) ∈ LC1 and
Z
Φ(f , g) dµ 6 N2 (f )N2 (g) .
§6 convexity inequalities INT IV.x199
Now,
Φ(f , g) (x) = hf , gi(x) = hf (x), g(x)i ;
and if ef ∈ L2F is the class of f for equality almost everywhere (§3, No. 4,
Def. 2), then
Z 1/2
N2 (f ) = N2 (e
f) = |f (x)|2 dµ = ke
f k2 .
The integral in (∗) depends only on the classes of f and g , so we may define
a function L2F × L2F → C by
Z
(e e) 7→
f, g hf , gi dµ
one has
Z Z
2 2
he
f,e
fi = hf , f i dµ = |f |2 dµ = N2 (f ) = ke
f k2 ,
thus he
f,e
f i > 0 when e
f 6= 0 . The norm e
f 7→ ke
f k derived from this form is
ke
f k = he
f,e
f i1/2 = ke
f k2 ;
as L2F is complete for this norm, we conclude that it is a Hilbert space for
the inner product he ei .
f, g
A shortcut, based on the characterization of a Hilbert space as a Banach
space whose norm satisfies the “parallelogram law ”:
By the parallelogram law in F , |f + g| 2 + |f − g|2 = 2|f |2 + 2|g|2 ;
integration term-by-term yields
thus the norm of the Banach space L2F satisfies the parallelogram law, hence
L2F is a Hilbert space, with inner product
1 e
he
f ,e
gi = ek2 )2 − (ke
(kf + g f −gek2 )2 + i(ke
f + iegk2 )2 − i(ke g k 2 )2
f − ie
4Z Z Z Z
1n o
= |f + g|2 dµ − |f − g|2 dµ + i |f + ig|2 dµ − i |f − ig|2 dµ
4
Z
1
= |f + g|2 − |f − g|2 + i|f + ig|2 − i|f − ig|2 dµ
4
Z Z
= hf (x), g(x)i dµ(x) = hf , gi dµ .
IV.94, `. −4 to −1.
“Corollary 4.”
Immediate from Cor. 1, with Φ(λ, a) = λa ( λ scalar, a ∈ F ).
IV.95, `. 7, 8.
“ . . . since the inequality (8) is true for upper integrals (Ch. I, No. 2,
Cor. of Prop. 2) ”
One can suppose that the fi are everywhere finite and > 0 (§5, No. 6,
Cor. 3 of Th. 5), as required by the cited Cor. from Ch. I.
IV.95, `. −5 to −3.
P
n
“Suppose first that f is an integrable step function, f = a k ϕ Ak ,
k=1
where the Ak are pairwise disjoint (§4, No. 9, Lemma).”
The term ‘measurable step function’ is defined in the first paragraph
of §5, No. 5, but the term ‘integrable step function’ is nowhere explicitly
defined; we can infer its meaning from an earlier approximation theorem
(§4, No. 10, Cor. 1 of Prop. 19) cited later in the present proof. Proposed
definition: An integrable step function is an element of E F (Φ) (§4, No. 9,
Def. 4), where Φ is the clan (loc. cit., Prop. 17) of integrable subsets of X
(§4, No. 5, Props. 6 and 7).
Proposition. For a function f : X → F , the following conditions are
equivalent:
a) f is an integrable step function (in the above sense);
b) f is integrable and has only finitely many values.
Proof. a) ⇒ b): Obvious.
b) ⇒ a): Let a1 , . . . , an be the distinct nonzero values of f and, for
−1
each index k , let Ak = f (ak ) . Then every Ak is measurable (§5, No. 5,
§6 convexity inequalities INT IV.x201
Prop. 7) and
n
X
f= a k ϕ Ak .
k=1
P
n
The function |f | = |ak |ϕAk is also integrable (§4, No. 2, Cor. 1 of Prop. 1)
k=1
and, for each k , ϕAk 6 |ak |−1 |f | , whence Ak is integrable (§5, No. 6, Th. 5);
thus f is a step function with respect to the clan Φ of integrable sets. ♦
IV.95, `. −3 to IV.96, `. 1.
“For every ε > 0 , there exists (for every index k ) a vector a 0k ∈ F0
such that |a0k |q = |ak |p if p > 1 (resp. |a0k | = 1 if p = 1 ) and hak , a0k i >
(1 − ε)|ak | · |a0k | (TVS, IV, §1, No. 3, Prop. 8).”
Fix k . By the cited Prop. 8,
|ak | = sup |hak , a0 i| ,
a0 ∈F0 , |a0 |=1
IV.96, `. 12.
Z
hf1 , gi dµ > Np (f1 ) − ε > 1 − 2ε .
IV.96, `. −6.
“ . . . the set of x ∈ X such that |f (x)| > α is measurable ”
For, f is measurable (§5, No. 6, Th. 5), therefore so is |f | (§5, No. 3,
Cor. 6 of Th. 1), therefore so is the set in question (§5, No. 5, Prop. 8).
IV.96, `. −5.
“ . . . therefore it contains a compact set K of measure > 0 . ”
Let B = {x ∈ X : |f (x)| > α } . Since B is not locally negligible, there
exists a compact set H such that B∩H is not negligible (§5, No. 2, Prop. 5).
Since B is measurable, B ∩ H is integrable (§5, No. 6, Cor. 3 of Th. 5), and
since B ∩ H is not negligible it contains a compact set K of measure > 0
(§4, No. 6, Cor. 1 of Th. 4).
IV.96, `. −3 to −1.
“ . . . for every ε > 0 , there exists a partition of K 1 into a finite number
of integrable sets, in each of which the oscillation of f is 6 ε ”
Since f K1 is continuous, for every x ∈ K1 there exists an open neigh-
borhood Ux of x in X such that |f (y)−f (x)| 6 ε/2 on U x ∩K1 . Cover K1
§6 convexity inequalities INT IV.x203
with a finite number of open sets U1 , . . . , Un in X such that f K1 has os-
cillation 6 ε on each Ui ∩ K1 . The sets Ui ∩ K1 are integrable, and since
the integrable sets form a clan, the sets
[
U1 ∩ K1 , Ui ∩ K1 --- Uj ∩ K1 (i = 2, . . . , n)
j<i
and so
Z
hf , gi dµ = |a| > |ha, a0 i| − ε > (|a| − ε) − ε > α − 2ε .
µ(A)
◦ ◦ ◦
B ∩E ⊃ B∩E = B,
◦ ◦
therefore B ∩ E ⊃ B = B (TVS, II, §2, No. 6, Cor. 1 of Prop. 16), whereas
◦ ◦
B ∩ E ⊂ B = B , thus B ∩ E = B , whence obviously B ∩ E = B . Let
f ∈ LFp and let
s = sup |hf , gi| ;
g∈B∩E
Inspection of the proof shows that B can be replaced by its subset consisting
of all integrable step functions g such that |g| 6 1 . Of course |g| 6 1 ⇒
N∞ (g) 6 1 . In the reverse direction, if g : X → F 0 is a measurable function
such that N∞ (g) 6 1 then |g(x)| 6 1 on the complement of a locally
negligible set N ; the function g1 = ϕ { N g is then measurable, |g1 | 6 1
and g = g1 locally almost everywhere. Since the set A = {x : f (x) 6= 0 }
is integrable, its intersection with N is negligible; for, A ∩ N is integrable
(§5, No. 6, Cor. 3 of Th. 5), and every compact set K ⊂ A ∩ N is negligible
(§5, No. 2, Prop. 5), therefore A∩N is negligible (§4, No. 6, Cor. 1 of Th. 4).
It follows that hf , gi = hf , g1 i almost everywhere; thus, writing B 1 for the
set of measurable functions g : X → F0 such that |g| 6 1 , that is, in the
notation of §5, No. 11,
B1 = {g ∈ S (X, µ; F0 ) : |g| 6 1 } ,
one has
Z
(∗) sup hf , gi dµ = N1 (f ) .
g∈B 1
In fact, defining
B3 = {g ∈ K (X; F0 ) : |g| 6 1 }
(a subset of B1 ∩ B2 ) the argument will show that
Z
(***) sup hf , gi dµ = N1 (f ) .
g∈B 3
Z n
X n
X
|f |ϕ { K dµ = |ai |µ(Ai --- Ki ) 6 ε/n = ε .
i=1 i=1
IV.97, `. −3. P 0
“Setting h = ai hi ”
We require that h , hence the hi , have compact support: Ki ⊂ Vi ⊂
Hi ⊂ Ui for suitable open Vi and compact Hi ; a suitable hi then exists
by Lemma 1 of Ch. III, §1, No. 2.
IV.97, `. −1.
Z
|hf , hi|ϕ { K dµ 6 ε
§6 convexity inequalities INT IV.x207
IV.98, `. 1. R
“ . . . consequently | hf , hi dµ| > 1 − 3ε , which proves our assertion.”
Abbreviate the equation
Z Z Z
hf , hi dµ = hf , hiϕK dµ + hf , hiϕ { K dµ
by α = β + γ . Then
Z
|α| − |β| 6 |α − β| = |γ| 6 |hf , hi|ϕ { K dµ 6 ε ,
thus Z Z
β= hf , gi dµ − hf , giϕ { K dµ ;
and finally |α| > |β| − ε > (1 − 2ε) − ε , which verifies the asserted inequality.
As in the note for IV.97, `. −12, −11, writing B 3 = {k ∈ K (X; F0 ) :
|k| 6 1 } , the foregoing shows that h ∈ B 3 and
Z Z
sup hf , ki dµ > hf , hi dµ = |α| > 1 − 3ε .
k∈B3
Since ε is arbitrary,
Z
sup hf , ki dµ > 1 = N1 (f ) ,
k∈B3
IV.98, `. 4.
“ whose support is contained in a countable union of compact sets K n .”
Here, by “support” is meant the set {x : f (x) 6= 0 } ; it is not necessary
that its closure be contained in the union.
IV.98, `. 8.
“ . . . a special case of (9) ”
That is, (9) as generalized in Remark 1.
IV.98, `. 8, 9.
“ f is then equivalent to a function in L p (§5, No. 6, Th. 5).”
Since Np (f ) < +∞ , the set N = {x : f (x) = +∞ } is negligible (§2,
No. 3, Prop. 7). Then f ϕ { N is a finite-valued measurable function such
that f ϕ { N = f almost everywhere, whence Np (f ϕ { N ) = Np (f ) < +∞ ,
thus f ϕ { N ∈ L p by the cited Th. 5. In other words, f is p-th power
integrable (§3, No. 4, second paragraph after Def. 2).
Incidentally, the set {x : (f ϕ { N )(x) 6= 0 } is contained in the union
of a negligible set and a sequence of compact sets (§5, No. 6, Lemma 1),
hence the same is true of {x : f (x) 6= 0 } ; thus the assumption about the
‘support’ of f , not used in the case N p (f ) < +∞ , is in a sense redundant.
IV.98, `. 10.
“ . . . set fn = inf(n, f ϕKn ) . ”
It is understood that if f (x) = +∞ and x ∈ / K n then fn (x) =
inf(n, 0) = 0 by the convention +∞ · 0 = 0 . One can suppose that the
sequence (Kn ) is increasing.
Note that fn is measurable and 0 6 fn 6 nϕKn , hence fn is integrable.
If p < +∞ then Np (fn ) 6 n Np (ϕKn ) < +∞ , whereas N∞ (fn ) 6 n < +∞ ,
therefore (11) holds for fn for 1 6 p 6 +∞ .
The sequence (fn ) is increasing, with upper envelope equal to f , that
is, fn (x) ↑ f (x) for all x ∈ X . For, since ϕKn 6 ϕKn+1 , fn 6 fn+1 is
assured by f (x)ϕKn (x) 6 f (x)ϕn+1 (x) (even if f (x) = +∞ and x ∈ / Kn ).
If f (x) = +∞ then x ∈ Km for some index m ; then, for all n > m ,
f (x)ϕKn (x) = +∞ , therefore fn (x) = n , thus fn (x) ↑ +∞ = f (x) . If
0 < f (x) < +∞ then x ∈ Km for some m , and one can suppose that
m > f (x) ; then fn (x) = f (x) for all n > m , whence fn (x) ↑ f (x) . And if
f (x) = 0 then fn (x) = 0 for all n , fn (x) ↑ 0 = f (x) .
IV.98, `. 12, 13.
“ . . . passing to the limit (assuming, as we may, that the sequence (K n )
R∗
is increasing), we have sup |f g| dµ = +∞ (§1, No. 3, Th. 3).”
By assumption, Np (f ) = +∞ . We know (see the preceding note)
that the sequence (fn ) is increasing, with upper envelope f , and that
§6 convexity inequalities INT IV.x209
f = 0 a.e. ⇒ hf , gi = 0 a.e.
g = 0 l.a.e. ⇒ hf , gi = 0 a.e.
INT IV.x210 extension of a measure. l p spaces §6
Z ∗ 1/p
Np (f ) = |f |p d|µ| (0 < p < +∞)
thus J is the image of S ∩ ]0, 1] under the mapping s 7→ 1/s , that is,
J = {s−1 : s ∈ S ∩ ]0, 1] ,
0<α61 ⇒ α∈
/ S ⇒ N1/α (f ) = +∞ ,
(***) p 7→ log Np (f ) (p ∈ J)
(****) p 7→ Np (f ) (p ∈ J)
is continuous.
IV.98, `. −3 to −1.
“If s ∈ J then, for every finite number p > s , |f | p = |f |s |f |p−s , and
the inequality of the mean shows that
s/p (p−s)/p
(12) Np (f ) 6 Ns (f ) N∞ (f ) .”
s/p
Since s ∈ J and f is not negligible, 0 < Ns (f ) < +∞ . If N∞ (f ) =
+∞ the right side of (12) is equal to +∞ and the inequalilty is trivial.
Suppose N∞ (f ) < +∞ , that is, +∞ ∈ I , thus f is bounded in measure.
Note that N∞ (f ) > 0 . {For, N∞ (f ) = 0 would imply that the set A = {x :
S
∞
f (x) 6= 0 } is locally negligible; but |f | p is integrable, so A ⊂ N ∪ Kn
n=1
for a suitable negligible set N and compact sets K n (§5, No. 6, Lemma 1),
S
∞
whence A = (A ∩ N) ∪ A ∩ Kn is negligible (§5, No. 2, Prop. 5), contrary
n=1
to the assumption that f is not negligible.} Thus 0 < N ∞ (f ) < +∞ . Since
p − s > 0 , one has
p−s p−s
N∞ (|f |p−s ) = M∞ (|f |p−s ) = M∞ (|f |) = N∞ (f ) < +∞
p p−s s
that is, Np (f ) 6 N∞ (f ) Ns (f ) , whence the asserted inequality.
The argument shows that p ∈ J for all p ∈ (s, +∞) , therefore J must be
of the form ]a, +∞[ or [a, +∞[ , and then I = ]a, +∞] or [a, +∞] .
INT IV.x214 extension of a measure. l p spaces §6
The presumption is that J is of the form ]a, +∞[ or [a, +∞[ , so that
p → +∞ makes sense. If N∞ (f ) = +∞ , the inequality holds trivially.
If N∞ (f ) < +∞ then, applying Prop. 11 of GT, IV, §5, No. 6 to the in-
equality (12), one has
h s/p (p−s)/p i
lim sup Np (f ) 6 lim sup Ns (f ) N∞ (f )
p→+∞ p→+∞
h s/p (p−s)/p i
= lim Ns (f ) N∞ (f ) = 1 · N∞ (f )
p→+∞
to be convex on J−1 (whose endpoints are 1/s < 1/r , where 1/s = 0 if
s = +∞ ); with the convention that log(+∞) = +∞ , the addition of the
point (1/r, +∞) to its graph creates no new chords between pairs of points
to challenge convexity.
(ii) If s < +∞ and s ∈ / J , then Ns (f ) = +∞ and the problem is
to show that Np (f ) → +∞ as p ↑ s , with the same conclusion concerning
convexity.
(iii) When s = +∞ , of course s ∈ / J ; and J −1 has endpoints 0 < 1/r .
Whether +∞ ∈ I or +∞ ∈ / I , that is, N∞ (f ) < +∞ or N∞ (f ) = +∞ , the
problem is to show that Np (f ) → N∞ (f ) as p ↑ +∞ . It has already been
shown that
lim sup Np (f ) 6 N∞ (f ) ;
p→+∞
showing that
lim inf Np (f ) > N∞ (f )
p→+∞
IV.99,R`. 14. R r
“ |f |p ϕA d|µ| tends to |f | ϕA d|µ| (§4, No. 3, Prop. 4).”
When p > r , |f | ϕA 6 |f | ϕA ∈ L 1 assures that |f |r ϕA ∈ L 1 .
r p
IV.99, `. 15.
R p R∗ r
“ Therefore |f | d|µ| tends to |f | d|µ| ”
R∗ r
We need only consider the case that r ∈
/ J , so that |f | d|µ| = +∞ .
Then
Z Z Z Z Z ∗
|f | d|µ| = |f | ϕA d|µ|+ |f | ϕ { A d|µ| → |f | ϕA d|µ|+ |f |r ϕ { A d|µ|
p p p r
1 1
log ap → · (+∞) = +∞
p r
(GT, IV, §4, No. 3, Prop. 8), whence (a p )1/p → +∞ , that is, Np (f ) → +∞ .
IV.99, `. 17.
“The same reasoning may be applied at the point s if s < +∞ . ”
In particular, the roles of A and { A are reversed: when p < s ,
|f |s ϕ { A 6 |f |p ϕ { A ∈ L 1 assures that |f |s ϕ { A ∈ L 1 .
IV.99, `. −12.
“ . . . let a be a number such that 0 < a < N ∞ (f ) . ”
Since J 6= ∅ we know that N∞ (f ) > 0 (see the note for IV.98, `. −3
to −1). {It is not enough to note that f is not negligible; it might still be
locally negligible—that is, N∞ (f ) = 0 —but not when J 6= ∅ .}
IV.99, `. −10.
“ . . . non-negligible ”
Since N∞ (f ) is the infimum of the numbers M > 0 such that |f | 6 M
locally almost everywhere, it is not the case that |f | < a locally almost
§6 convexity inequalities INT IV.x217
everywhere. That is, the set A = {f : |f (x)| > a } is not locally negligible,
hence is not negligible.
IV.99, `. −7.
“ . . . which completes the proof.”
Letting a → N∞ (f ) , one has lim inf Np (f ) > N∞ (f ) . This completes
p→+∞
the proof of lim Np (f ) = N∞ (f ) and hence of the Proposition (see the
p→+∞
note for `. 4–6).
IV.99, `. −6 to −4.
“Corollary.”
By assumption, 1 6 r < p < s 6 +∞ and f ∈ LFr ∩ LFs ; we are to
show that f ∈ LFp , and can suppose that f is not negligible.
From f ∈ LFr we know that Nr (f ) < +∞ , so that r belongs to the
interval J of the proof of Prop. 4; if s < +∞ then similarly s ∈ J , therefore
Np (f ) < +∞ for all p ∈ [r, s] . On the other hand, if s = +∞ then f ∈ L F∞
is bounded in measure, N∞ (f ) < +∞ ; since J is nonempty, the proof of
Prop. 4 shows that J is an interval with right endpoint +∞ , consequently
Np (f ) < +∞ for every p ∈ [r, +∞[ .
IV.100, `. 9, 10.
“This is an immediate consequence of Prop. 4 above and of the Cor. of
Prop. 4 of Ch. I, No. 3.”
R
We may write simply kµk = |µ|(X) = d|µ| (§4, No. 7, Prop. 12). Since
µ is bounded, every measurable set in X is integrable (§5, No. 6, Cor. 1 of
Th. 5). It follows that every locally negligible set A in X is negligible; for,
A is measurable (§5, No. 2, sentence after Def. 3), hence integrable, therefore
S
∞
A = N∪ Kn with N negligible and the Kn compact (§4, No. 6, Cor. 2
n=1
of Th. 4), and the Kn = A ∩ Kn are also negligible (§5, No. 2, Prop. 5).
If N∞ (f ) < +∞ , then N∞ (f ) may be described as the infimum of all real
numbers α > 0 such that |f | 6 α almost everywhere, and N ∞ (f ) = 0
means that f = 0 almost everywherep (Nos. 2, 3). If N ∞ (f ) < +∞ then, for
p
all p ∈ [1, +∞[ , |f | 6 N∞ (f ) almost everywhere,
Z Z
p
p
|f | d|µ| 6 N∞ (f ) d|µ| ,
M (f ) = kµk−1 |µ|*(f ) (f ∈ P) ;
INT IV.x218 extension of a measure. l p spaces §6
then M satisfies the conditions 1◦ , 2◦ , 3◦ of Ch. I, No. 1 (§1, No. 3, Props. 10,
11, 12), and M (1) = kµk−1 |µ|*(1) = 1 . For 0 < p < +∞ and any function
f : X → F , define
1/p
N0p (f ) = M (|f |p
(cf. the note for IV.98, `. −10 to −7); since M (αf ) = αM (f ) for f ∈ P
and scalars α > 0 , we have
1/p
(∗) N0p (f ) = kµk−1 |µ|*(|f |p ) = kµk−1/p Np (f ) ,
IV.100, `. 11–13.
“Corollary.”
The corollary is valid for 1 6 r < s 6 +∞ , except that when s = +∞
the topology on LF∞ defined by the semi-norm N∞ is the topology of
‘uniform convergence locally almost everywhere’ (No. 3).
For a measurable function f : X → F let us write I f for the set of all
p ∈ [1, +∞] such that Np (f ) < +∞ . Suppose f ∈ LFs , that is, s ∈ If . By
Prop. 5, If is an interval with left endpoint 1 , therefore [1, s] ⊂ I f and in
particular r ∈ If , that is, f ∈ LFr . Thus LFs ⊂ LFr . Moreover, for a fixed
measurable function f , if If 6= ∅ then, with notation as in the preceding
note, the function
p 7→ N0p (f ) = kµk−1/p Np (f ) ( p ∈ If )
we see that the canonical injection L Fs → LFr is continuous for the respec-
tive semi-norm topologies; in other words, the topology on L Fs induced by
the Nr -topology on LFr is coarser than the Ns -topology.
Incidentally, KF ⊂ LFp for all p ∈ [1, +∞] , and the normed space L pF
derived from the semi-norm Np is a Banach space (No. 3, Prop. 2 and §3,
No. 4, Th. 2).
IV.100, `. −13.
“ N∞ (f ) = kf k = sup |f (x)| ”
x∈X
Every function f : X → F is continuous, hence measurable, and the
empty set is the only negligible set, hence the only locally negligible set.
IV.100, `. −13 to −11.
“ . . . if there exists a number α > 0 such that |f (x)| > α for infinitely
many values of x ∈ X , then Np (f ) = +∞ for every finite p ”
In this case I = ∅ or {+∞} according as |f | is unbounded or bounded.
INT IV.x220 extension of a measure. l p spaces §6
that is, g(0) 6 g(α) for all α ∈ ]0, 1/r] ; thus g takes its smallest value
at the left end-point 0 . Let us show that g is an increasing function. For
0 6 α, β 6 1/r , α 6= β , write
g(β) − g(α)
Mαβ = = Mβα
β −α
for the slope
of the chord of the graph of g joining the points α, g(α) and
β, g(β) . Since g is convex, for every α ∈ [0, 1/r] the function
with the discrete topology, µ is the atomic (hence discrete) measure defined
by a mass +1 at every point of X (Ch. III, §1, No. 3, Example I), and a
function on X is displayed as a sequence.
1 1
Let cn = . If 0 < p < +∞ , then cpn > for all sufficiently
log n n
P∞
large n , therefore cpn = +∞ . Thus I = {+∞} for this sequence. {This
i=1
example is given in Konrad Knopp’s Infinite sequences and series (Dover,
New York, 1956), p. 60, Example 7 of 3.2.1; and Theory and application of
infinite series (Hafner, New York, 1951), p. 119, Example g).}
A sequence for which I = ∅ : Interlace (c n ) with any unbounded se-
quence.
A sequence for which I = [1, +∞] : Any absolutely summable sequence.
IV.100, `. −4 to −1.
“Corollary.” (of Prop. 6)
The corollary is valid for 1 6 r < s 6 +∞ , except that when s = +∞
the topology on LF∞ defined by the semi-norm N∞ is the topology of
uniform convergence.
Let 1 6 r < s 6 +∞ and let f ∈ LFr . In the notation of Prop. 6,
r ∈ I . By Prop. 6, [r, +∞] ⊂ I , hence also s ∈ I , that is, f ∈ L Fs ;
thus LFr ⊂ LFs . Moreover, r < s implies by Prop. 6 that N r (f ) > Ns (f ) ,
therefore the canonical injection L Fr → LFs is continuous for the respective
semi-norm topologies; in other words, the topology on L Fr induced by the
Ns -topology of LFs is coarser than the Nr -topology.
§7. BARYCENTERS
IV.101, `. 7, 8. R
“ . . . the integral x dµ(x) is therefore defined and is an element of E 0 *
(Ch. III, §3, No. 1).”
Write iK : K → E for the canonical injection; then i K ∈ K (K; E) ⊂
R R R
f
K (K; E) and, by definition, iK dµ = iK (x) dµ(x) = x dµ(x) is the
unique element z ∈ E0 * such that, for all x0 ∈ E0 ,
Z Z
hz, x i = hiK (x), x i dµ(x) = hx, x0 i dµ(x)
0 0
Z
= (x0 K) dµ = µ(x0 K) = µ(x0 ◦ iK ) .
See also the note for III.33, `. 6–11.
§7 barycenters INT IV.x223
IV.101, `. 8, 9.
“Moreover, on K , the topology induced by the weak topology σ(E 0 *, E0 )
is identical with the original topology.”
As K ⊂ E , the assertion entails the identification of E as a linear
subspace of E0 * : for x ∈ E write x* for the linear form on E 0 defined by
x*(x0 ) = x0 (x) = hx, x0 i for all x0 ∈ E0 , and define the mapping (linear and
injective) θ : E → E0 * by θ(x) = x* (x ∈ E) .
The topology σ(E0 *, E0 ) on E0 * is the initial topology for the family
of ‘evaluation mappings’
(i) f 7→ f (x0 ) (f ∈ E0 *)
σ(E0 *, E0 ) ∩ E = σ(E, E0 ) .
IV.101, `. 9–12.
“ . . . if C is the closed convex envelope of K in E 0 * equipped with
σ(E *, E0 ) , then C ∩ E is the closed convex envelope of K in E for the
0
N = {x ∈ K : α(x) 6= 0 }
P
n
thus µ = λi εxi , where λi = α(xi ) .
i=1
There exist functions fi ∈ C (K) , 0 6 fi 6 1 , such that fi (xj ) = δij
for all i, j = 1, . . . , n , whence µ(fi ) = α(xi ) . The assumption that µ > 0
§7 barycenters INT IV.x225
is clearly equivalent to the condition λ i > 0 for all i (i.e., α(x) > 0 for all
x ∈ K ). Moreover, from
X
n
|µ(f )| 6 λi kf k f ∈ C (K)
i=1
P
n
we know that µ is bounded, with kµk 6 λi ; since K is compact, we
i=1
have in fact
n
X n
X
kµk = |µ|(1) = µ(1) = λi εxi (1) = λi
i=1 i=1
(Ch. III, §1, No. 8, Cor. 2 of Prop. 10), and to say that µ has total mass
P
n
equal to 1 means that λi = 1 , so that µ is a convex combination of
i=1
Dirac measures.
IV.101, `. −12.
R P
“ bµ = x dµ(x) = λi xi .”
i
0 0
For all x ∈ E ,
Z
R
hbµ , x0 i = x dµ(x), x0 = hx, x0 i dµ(x) = µ(x0 K)
X X
= λi εxi (x0 K) = λi (x0 K)(xi )
i i
X X
X
0
= λi x (xi ) = λi hxi , x0 i = λi xi , x 0 .
i i i
IV.101, `. −11.
“ x is the barycenter of the measure ε x .”
The discrete measure defined by α = ϕ{x} is µ = 1 · εx .
IV.101, `. −6, −5.
“This is nothing more than Prop. 5 of Ch. III, §3, No. 2 applied to the
canonical injection of K into E . ”
As in the note for `. 9–12, let us instead write B for the closed convex
envelope of K in E for the original topology T on E , reserving the letter C
for the closed convex envelope of K in E 0 * for topology T * = σ(E0 *, E0 ) as
in the sentence preceding Def. 1, where it is shown that C ∩ E = B (with E
canonically identified as a linear subspace of E 0 * , as in the note for `. 8,9).
INT IV.x226 extension of a measure. l p spaces §7
Write M1 for the set positive measures on K of total mass 1 , and set
S = {bµ : µ ∈ M1 } ,
IV.102, `. 16. R
“Therefore sup hα (x) dµ(x) = sup hα (bµ ) = f (bµ ) ”
α α
R
In the expression hα (x) dµ(x) , by hα is meant the restriction
of hα to K ; and the second equality holds because 1) sup (hα K) = f ,
α
and 2) bµ ∈ K by the preceding Remark.
IV.102, `. 17, 18.
“When µ is a discrete positive measure on K of total mass 1 , Prop. 2
yields anew the inequality that defines the convex functions on K . ”
R+ is convex means that ifPx1 , . . . , xn are distinct
To say that f : K → P
elements of K , and x = λi xi with λi > 0 and λi = 1 , then
i i
X X
f (x) 6 λi f (xi ) = λi εxi (f ) ;
i i
P
one knows from the Example following Def. 1 that λi εxi is the general
i
discrete positive measure on K of total mass 1 , and that its barycenter
is x . To formulate this result in terms of Prop. 2 we will make use of the
following:
Remark. If K is a compact space and µ is a positive discrete measure
on K , then every lower semi-continuous function f : K → R + is integrable;
Pn R
and if µ = λi εxi as in the note for IV.101, `. −15 to −13, then f dµ =
i=1
P
n
λi f (xi ) .
i=1
Since f is measurable for any measure (§5, No. 5, Cor. of Prop. 8)
Proof. P
and µ∗ = λi (εxi )* (§1, No. 3, Prop. 15), one is reduced to the case that
i
µ = εx (§5, No. 6, Th. 5). Let H be the set of all functions g ∈ C + (K)
such that g 6 f ; by the Lemma of §1, No. 1,
thus f (x) = (εx )*(f ) in the sense of loc. cit., Def. 1, hence also in the sense
R∗ R
of §1, No. 3, Def. 3. Thus f dεx = f (x) < +∞ , whence f dεx exists
(§4, No. 4, Cor. 1 of Prop. 5) and is equal to f (x) . ♦
Finally, with K as in Prop. 2, for a lower semi-continuous function
f > 0 to be convex, it is necessary and sufficient that it satisfy the inequality
of Prop. 2 for every positive discrete measure µ on K of total mass 1 .
Proof. Necessity: A special case of Prop. 2.
INT IV.x228 extension of a measure. l p spaces §7
R∗
Sufficiency: The assumption is that f (b µ ) 6 f dµ for every positive
discrete measure P µ on K of total mass 1 . If P x 1 , . . . , xn are distinct points
of K and if xP= λi xi with λi > 0 and λi = 1 , then x = bµ for the
measure µ = λi εxi by the Example following Def. 1, so by assumption
Z ∗
f (x) = f (bµ ) 6 f dµ ,
R P
whence, citing the above Remark, f (x) 6 f dµ = λi f (xi ) .
IV.102, `. −13.
“It suffices to note that inf g(x) = a is finite and apply Prop. 2 to
x∈K
g − a .”
The values of g are finite, and a = g(y) for some y ∈ K (GT, IV, §6,
No. 2, Th. 3).
The function f = g − a satisfies the hypotheses of Prop. 2, so
R∗
(g − a)(bµ ) 6 (g − a) dµ ; moreover, g − a is bounded and measurable,
and µ is bounded, therefore g − a is integrable (§5, No. 6, Th. 5) and
Z Z
g(bµ ) − a = (g − a)(bµ ) 6 (g − a) dµ = g dµ − a .
IV.102, `. −3 to IV.103, `. 1.
“For every point a ∈ K , there exists a closed convex neighborhood V a
of 0 in E such that
p
\ p
\ p
\
Wa = Wi = K ∩ (a + Vi ) = K ∩ a + Vi = K ∩ (a + Va ) ,
i=1 i=1 i=1
IV.103, `. 5–7.
“Each of the measures µj is positive, of total mass 1 , and its support
is contained in the compact convex set W aj ”
6 0 then µj (1) = α−1
If αj = −1
j µ(gj · 1) = αj αj = 1 , and, citing Ch. III,
§2, No. 3, Prop. 10,
IV.103, `. 7–9.
“ . . . by definition,
r
X
(3) µ= αj µj
j=1
since gj · µ = 0 if µ(gj ) = 0 ”
If αj = 0 , that is, if µ(gj ) = 0 , then, for every f ∈ C+ (K) with
0 6 f 6 1 , one has 0 6 gj f 6 gj , therefore 0 6 µ(gj f ) 6 µ(gj ) = 0 , thus
gj · µ = 0 , whence αj µj = 0 · εaj = 0 = gj · µ ; whereas if αj > 0 then
αj µj = gj · µ by the definition of µj . Thus
gj · µ = α j µj (j = 1, . . . , r) .
P
r P
r
Since gj = 1 , for every f ∈ C (K; C) one has f = gj f , therefore
j=1 j=1
r
X r
X r
X
µ(f ) = µ(gj f ) = (gj · µ)(f ) = (αj · µj )(f ) ,
j=1 j=1 j=1
whence (3).
IV.103, `. 11.
“Let xj be the barycenter of µj , which belongs to Waj (No. 1, Prop. 1) ”
At any rate, µj is a positive measure on K of total mass 1 , its barycen-
ter xj is defined, and xj ∈ K by the corollary of the cited Prop. 1 (since
the closed convex envelope of K in E is K itself).
INT IV.x230 extension of a measure. l p spaces §7
(the first three integrations are over K j , the last two over K ), thus (∗) is
verified.
§7 barycenters INT IV.x231
r
X r
X Z
0
= αj µj (z K) = αj hy, z0 i dµj (y)
j=1 j=1
Xr DX
r E
= αj hbµj , z0 i = α j b µj , z 0 ,
j=1 j=1
P
r
whence bµ = αj bµj . But x = bµ and xj = bµj by definition, thus
j=1
P
r P
r
x= αj xj . On the other hand bν = αj xj by the Example following
j=1 j=1
Def. 1 of No. 1, thus bν = x = bµ .
IV.103, `. 15, 16.
“ . . . since Supp(µj ) ⊂ Waj , |µj (fi ) − fi (aj )| 6 δ/2 for 1 6 i 6 p . ”
Writing Kj = Supp(µj ) , we have Kj ⊂ Waj ⊂ K , and K --- Kj is a
µj -negligible open set in K . Let ϕKj be the characteristic function of K j
in K . Since Kj ⊂ Waj , we know from (2) that the function f i − fi (aj ) · 1
in C (K; C) = K (K; C) satisfies
Xr r
X
|µ(fi ) − ν(fi )| = αj µj (fi ) − αj εxj (fi )
j=1 j=1
Xr
= αj [µj (fi ) − εxj (fi )]
j=1
r
X r
X
6 αj |µj (fi ) − εxj (fi )| 6 αj δ = δ .
j=1 j=1
IV.104, `. 1, 2.
“ . . . the hypothesis that x is the barycenter of ν may be written
Pr
x= λi xi . ”
i=1
Recalling that the xi belong to K0 , the computation in the note for
P
r P
r
IV.103, `. 13, 14 yields bν = λi bεxi , that is, x = λi xi .
i=1 i=1
IV.104, `. 9.
“Then x is the barycenter of λµ0 + (1 − λ)µ00 . ”
Arguing as in the preceding note, bλµ0 +(1−λ)µ00 = λbµ0 + (1 − λ)bµ00 =
λx0 + (1 − λ)x00 = x .
IV.104, `. 9, 10.
“Therefore λµ0 + (1 − λ)µ00 = εx . ”
By the assumption on εx .
IV.104, `. 10.
“Therefore µ0 and µ00 are proportional to εx ”
Note first that Supp(εx ) = {x} . {At any rate, Supp(εx ) 6= ∅ since
εx 6= 0 , so it suffices to show that Supp(εx ) ⊂ {x} , that is, K0 − {x} ⊂
K0 − Supp(εx ) , in other words, the restriction of ε x to the open subset Y =
K0 --- {x} of K0 is zero (Ch. III, §2, No. 2, Def. 1). Indeed, if g ∈ K (Y; C)
and g 0 is the extension by 0 of g to K0 (i.e., by setting g 0 (x) = 0 ), then
(εx )Y (g) = εx (g 0 ) = g 0 (x) = 0 .}
From 0 6 λµ0 6 εx one then infers (loc. cit., Prop. 3)
Let z ∈ K . Then
that is, z ∈
/ M ⇔ z ∈ q(U) .
IV.104, `. −6 to −4.
“By the Hahn-Banach theorem, for (a, b) to belong to S , it is necessary
and sufficient that h(a, b) > 0 for every continuous affine linear function h
on E × R such that h(x, u(x)) > 0 for x ∈ K .”
Review. A function f : E → R is affine linear if there exist a linear
form f0 on E and a real number α such that f (x) = f 0 (x) + α for all
x ∈ E (A, II, §9, No. 4, Prop. 6); f is continuous when f 0 ∈ E0 .
Thus, a function h : E × R → R is a continuous affine linear function
if there exist a continuous linear form h 0 ∈ (E × R)0 and an α ∈ R such
that h(x, t) = h0 (x, t) + α for all (x, t) ∈ E × R . But we can make the
identifications (E × R)0 = E0 ⊕ R0 = E0 ⊕ R (TVS, IV, §1, No. 5), so there
exist f0 ∈ E0 and β ∈ R such that h0 (x, t) = f0 (x) + βt , whence
h(x, t) = f0 (x) + βt + α = f0 (x) + α + βt ;
implies
Now, the closed half-spaces in E are the sets {x : f (x) > 0 } , where f is
a continuous affine linear function on E . Therefore the validity of (i) says
that a belongs to every closed half-space that contains K , so a belongs to
their intersection, which is equal to K (TVS, II, §5, No. 3, Cor. 1 of Prop. 4).
Conversely, if a ∈ K , then a belongs to every closed half-space con-
taining K , thus (i) is satisfied and so (∗) holds for every h with λ = 0 .
Conclusion: Assuming (a, b) ∈ S ,
case 2. λ = −1 .
For such h , the meaning of (∗) is
that is,
f (x) + u(x) > 0 on K ⇒ f (a) + b > 0 ,
that is,
−f (x) 6 u(x) on K ⇒ − f (a) 6 b ;
since f 7→ −f is a permutation of the set of all continuous affine linear
functions on E , we may write this implication as
(ii) f 6 u on K ⇒ f (a) 6 b .
Now, u is the supremum of the f K as f runs over the set of all continuous
affine linear functions such that f K 6 u (TVS, II, §5, No. 4, Prop. 5; see
the Remark at the end of this note); by (ii), f (a) 6 b for every such f ,
therefore u(a) 6 b .
Conversely, if a ∈ K (as is the case when (a, b) ∈ S , as noted above)
and u(a) 6 b , then f 6 u on K implies that f (a) 6 u(a) 6 b , so the
implication (ii) holds.
Conclusion: Assuming (a, b) ∈ S ,
case 3. λ = 1 .
For such h , the meaning of (∗) is
that is,
f (x) − u(x) > 0 on K ⇒ f (a) − b > 0 ,
that is,
IV.105, `. −14.
“ u(x) > u(y) + u(z) /2 ”
Proof #1. If f is a continuous affine linear function on E such that
f > u on K , then f > u on K by the definition of u , therefore
1 1 1 1
u(y) + u(z) 6 f (y) + f (z) = f (x) ,
2 2 2 2
and the assertion follows on taking the infimum of f (x) over all such f .
Proof #2. Since uK is the lower envelope of a family of concave (even
affine-linear) functions
f K minorized by the finite-valued function u , it
follows that uK is concave (TVS, II, §2, No. 9). That is, if y, z ∈ K and
x = αy + (1 − α)z with 0 6 α 6 1 , then u(x) > αu(y) + (1 − α)u(z) .
IV.105, `. −11, −10.
“ . . . there exists, by Prop. 1 of No. 1, a positive measure ν on G , of
total mass 1 , having (a, u(a)) as barycenter.”
Recall that S is the closed convex envelope of the compact set G ; note
that Prop. 1 does not require that the closed convex envelope be compact.
IV.105, `. −6.
Z Z
(6) a= x dµ(x) and u(a) = u(x) dµ(x) .
for, if w0 ∈ F0 then
DZ E Z Z Z
0 0 0
g dµ, w = hg, w i dµ = (w ◦ g) dµ = (w0 ◦ g) ◦ p dν
Z Z DZ E
0
0
= w ◦ (g ◦ p) dν = hg ◦ p, w i dν = (g ◦ p) dν, w0 .
Z
hx, z0 i + λu(x) dµ(x)
Z Z
0
= hx, z i dµ(x) + λ u(x) dµ(x)
DZ E Z
= x dµ(x), z0 + λ u(x) dµ(x)
D Z Z E
= x dµ(x), u(x) dµ(x) , z0 + λ ,
thus
Z Z Z
(††) (x, u(x)) dµ(x) = x dµ(x), u(x) dµ(x) ;
R R
from (†) and (††) we have x dµ(x), u(x) dµ(x) = (a, u(a)) , whence (6).
IV.106, `. 10.
“ . . . subspace A ”
Clearly a linear subspace of C (K; R) .
IV.106, `. 18.
“ . . . for this it suffices that hn (x) 6= hn (x0 ) ”
Let h be an affine linear function on K , and let x, x 0 be points of K
such that h(x) 6= h(x0 ) ; we are to show that h2 is strictly convex on the
segment [x, x0 ] = {λx + (1 − λ)x0 : 0 6 λ 6 1 } . That h2 is convex has
already been noted (TVS, II, §2, No. 8, Examples), and the computation in
the cited Examples show that if 0 < λ < 1 then
h2 λx + (1 − λ)x0 < λh2 (x) + (1 − λ)h2 (x0 ) .
Suppose y, y0 ∈ [x, 0 0
x ] 2with y 6= y , 2and0 let 0 < λ < 1 ; to prove that
2 0
h λy + (1 − λ)y < λh (y) + (1 − λ)h (y ) , it suffices by the foregoing to
show that h(y) 6= h(y0 ) . Say
y = ρx + (1 − ρ)x0 , y0 = σx + (1 − σ)x0 (0 6 ρ, σ 6 1) .
is continuous, and σ(H 0 , H ) is the initial topology for the family of linear
forms
α 7→ α(f ) = hf, αi (α ∈ H 0 )
indexed by the functions f ∈ H (TVS, II, §6, No. 2, Def. 2), whence the
assertion (GT, I, §2, No. 3, Prop. 4).
In more direct terms, given any x0 ∈ X , the continuity of iH at x0
may be seen as follows. A basic neighborhood V of i H (x0 ) in H 0 is given
by
V = {α ∈ H 0 : |α(fk ) − iH (x0 ) (fk )| < ε for k = 1, . . . , n } ,
that is, µ has barycenter iH (x) if and only if λ satisfies (10) for every
h∈H .
The setting for discussing the barycenter of µ is as follows. We have
a Hausdorff locally convex space E = H 0 , where H ⊂ C (X) is a normed
space, and E is equipped with the weak topology σ(H 0 , H ) ; then E0 =
(H 0 )0 = H (TVS, II, §6, No. 2, Prop. 3). We have a compact subspace
K = iH (X) of E , whose closed convex envelope C in E is compact by
part (i) of the present Proposition. Thus the measure µ on K is eligible to
have a barycenter bµ ∈ E0 * = (H 0 )0 * = H * (No. 1, Def. 1), and in fact
bµ ∈ C ⊂ E = H 0 (No. 1, Prop. 1 and its Corollary), characterized by the
property
DZ E Z Z
bµ (h) = hbµ , hi = α dµ(α), h = hα, hi dµ(α) = α(h) dµ(α)
K K K
Z Z
hbµ , hi = α(h) dµ(α) = ĥ(α) dµ(α)
K K
Z Z Z
0
= ĥ (z) dλ(z) = h(z) dλ(z) = h dλ .
X X
Therefore
bµ = iH (x) ⇔ hbµ , hi = hiH (x), hi for all h ∈ H
Z
⇔ h dλ = iH (x) (h) = h(x) for all h ∈ H ,
whence (∗).
While the notations are at hand, let us complete the proof of (ii). Con-
sider the statements
(a) iH (x) is an extremal point of of C ;
(b) εx isthe only positive measure λ on X satisfying the condition (10)
λ(h) = h(x) for all h ∈ H .
We are to show that (a) ⇔ (b).
Proof of (a) ⇒ (b). The measure λ = εx trivially satisfies (10) for
every h ∈ H . Assuming (a), suppose λ is any positive measure on X
satisfying (10) for every h ∈ H (in particular, 1 X ∈ H , and λ(1X ) =
1X (x) = 1 , so λ has total mass 1 ). Let µ be the corresponding measure on
iH (X) , which is also positive and of total mass 1 . We know from (∗) that
bµ = iH (x) ; but iH (x) is extremal in C by the assumption (a), therefore
µ = εiH (x) by the Corollary of Prop. 3 of No. 2. Since the measure on X
corresponding to εiH (x) is εx , we conclude that λ = εx , whence (b).
Proof of (b) ⇒ (a). Assume (b). To prove (a), it suffices to verify that
iH (x) satisfies the conditions of the above-cited Corollary. At any rate,
εiH (x) is a positive measure on iH (X) , of total mass 1 , whose barycenter
is iH (x) (No. 1, Example). On the other hand, suppose µ is a positive
measure on iH (X) , of total mass 1 , such that bµ = iH (x) , and let λ
be the corresponding measure on X . By (∗), λ satisfies (10) for every
h ∈ H , therefore, by the assumption (b), λ = ε x . As the measure on
iH (X) corresponding to εx is εiH (x) , we conclude that µ = εiH (x) . Thus
the conditions of the Corollary are satisfied, whence (a).
Remarks. 1. If λ is any measure on the compact space X , its
restriction
to the normed space H is a continuous linear form, and kλ H k 6 kλk .
The positive measures on X are the positive linear forms on K (X) = C (X)
(Ch. III, §1, No. 5, Th. 1), and kλk = λ(1) for all such measures λ (§4,
No. 7, Prop. 12); since 1 ∈ H and k1k = 1 , one has
kλH k > |(λH )(1)| = λ(1) = kλk ,
thus kλH k = kλk = λ(1) when λ > 0 .
§7 barycenters INT IV.x247
0
Let us write Φ : M (X) → H for the mapping defined by Φ(λ) =
λH . From kΦ(λ)k 6 kλk we see that Φ is a continuous linear mapping
between Banach spaces, and kΦk 6 1 ; in fact kΦk = 1 since, if λ is
any nonzero positive measure on X of norm 1 (for example, any Dirac
measure εx ) then kΦk > kΦ(λ)k = kλH k = kλk = 1 .
2. If x ∈ X , λ is a positive measure on X of total mass 1 , and µ is
the corresponding measure on iH (X) , the equivalence (∗) can be expressed
as
λH = εx H ⇔ bµ = iH (x) .
3. Writing M+ (X) for the set of all positive measures on X , the
assertion (ii) of Prop. 4, for a point x ∈ X , can be expressed as
(ii) {λ ∈ M+ (X) : Φ(λ) = εx H } = {εx } ⇔ iH (x) is extremal in C .
If Φ is injective then the condition on the left in (ii) holds for every x ∈ X ,
and so iH (X) is precisely the set of all extremal points of C ; this is the case
if H is dense in C (X) —for example, if f ∈ H ⇒ |f | ∈ H (M.H. Stone’s
theorem, GT, X, §4, No. 1, Th. 2) or if f, g ∈ H ⇒ f g ∈ H (loc. cit.,
No. 4, Prop. 6).
The case that H = C (X) has been taken up in §4, No. 8, Prop. 15
and is continued in Ch. VI, §1, No. 6, Remark 1). A (real) character of
a commutative algebra A over R with unity is an algebra epimorphism
A → R ; the characters of C (X) are precisely the ε x (Gillman and Jerison,
Rings of continuous functions, p. 57, item 4.9, Van Nostrand, Princeton,
N.J., 1960).
5. In a vague sense, iH (X) is a ‘linearization’ of the compact space X ,
‘tailored’ to the linear subspace H of C (X) , placing X in a structurally
richer context (topological vector spaces) than that of topological spaces.
Prop. 4 reformulates a property of a point x ∈ X with respect to H in
terms of measures on X ; measures on i H (X) play only an auxiliary role
in the proof. The property is reformulated in Prop. 6 in topological terms,
with measures playing a role only in the proof; this theme culminates in the
theorem of Errett Bishop (No. 5, Th. 2), where measures are nowhere in
sight. In Prop. 8 and in Choquet’s theorem (No. 6, Th. 3), it is measures
that are in the forefront.
All in all, Prop. 4 is a subtle, far-reaching result whose cunning remains
a mystery to me.
IV.107, `. −3 to −1.
“ . . . the assertion (ii) is just the translation of the criterion of No. 2,
Cor. of Prop. 3 for iH (x) to be an extremal point of C .”
See the preceding note.
INT IV.x248 extension of a measure. l p spaces §7
IV.108, `. 8, 9.
“ the weakly closed hyperplane of H 0 with equation hh, t0 i = hh, iH (x)i
is a support hyperplane of iH (X) .”
Order of events: Fix h ∈ H and let x be a point of X where h
attains its supremum. The linear form t 0 7→ hh, t0 i (t0 ∈ H 0 ) is continuous
for σ(H 0 , H ) ; the set H = {t0 ∈ H 0 : hh, t0 i = hh, iH (x)i } is a closed
hyperplane in H 0 , the relation
expresses that the points of iH (X) lie on the same side of H , and since
H contains at least the point iH (x) of iH (X) , it is a support hyperplane
of iH (X) (TVS, II, §5, No. 2, Def. 3). It follows that H is also a support
hyperplane of the closed convex envelope C of i H (X) in H 0 , therefore H
contains some extremal point t00 of C (loc. cit., §7, No. 1, Cor. of Prop. 1).
Necessarily t00 ∈ iH (X) (loc. cit., Cor. of Prop. 2), say t 00 = iH (y) ; then y
is by definition H -extremal, and
The assertion of the cited Prop. 1 in EVT (p. EVT II.22) is that
whence
α0 = inf [α0 , α00 ] = inf{h(a) : h ∈ Sf } .
Translated to the present context, this says that
sup h(x) = inf{λ(f ) : λ ∈ M+ (X) and λH = εx H } ,
h∈H , h6f
as asserted.
Incidentally, the infimum in the assertion is attained: there exists a
λ0 ∈ S such that λ0 (f ) = sup h(x) , but λ0 depends on f (be-
εx H
h∈H , h6f
cause α0 does) as well as on x . Similarly, there exists a λ 00 ∈ S such
εx H
00
that λ (f ) = inf h(x) .
h∈H , h>f
{This argument was perfectly clear to me when I first studied it in 1974;
it is lifted verbatim from my notes at the time. On rereading the argument
when preparing these notes, with TVS at my elbow, I found that I no longer
understood it. After several days of struggle, it occurred to me to look at
the French original, and all was clear again.
The problem: In TVS, instead of the conclusion (∗), one finds the con-
clusion
{h(a) : h ∈ Sf } ⊂ [α0 , α00 ] ,
which (i) follows at once from the fact that if y, z ∈ V with z 6 a 6 y , and
h ∈ Sf , then
f (z) = h(z) 6 h(a) 6 h(y) = f (y) ,
whence α0 6 h(a) 6 α00 , and (ii) is of no help in proving (∗). The problem
did not arise in 1974, as I was working from the 2nd French edition of Chs. I
and II of EVT.}
IV.108, `. 24–26.
“Suppose that x is H -extremal; it then follows from Prop. 4, (ii) that
for every function f ∈ C (X; R) ,
Since x is H -extremal, by
the cited Prop. 4 the only positive mea-
sure λ on X that extends εx H is λ = εx ; thus, in the notation of the
INT IV.x250 extension of a measure. l p spaces §7
{z0 ∈ H 0 : hh, z0 i = α }
and since hh, iH (x)i 6 hh, iH (y)i for all x ∈ X , the set
then there exists a point y ∈ F such that hh, i H (y)i = α ; that is, every
h ∈ H attains its supremum on iH (X) at some point of iH (F) = G .
This means: If H is any closed support hyperplane of i H (X) , then
there exists a point y ∈ F such that iH (y) ∈ H , in other words G∩H 6= ∅ .
Since G ⊂ iH (X) , G ∩ H = G ∩ iH (X) ∩ H , so it is equivalent to say (as
in the text) that G intersects iH (X) ∩ H .
IV.109, `. 15, 16.
“ . . . the condition c) signifies that every point of i H (X) is the barycen-
ter of a measure with support contained in G ”
Reviewing the notations employed in the proof of Prop. 4, let us write
λ for a measure on the compact space X , and µ for the corresponding
measure on the subspace iH (X) of H 0 homeomorphic to X , defined by
condition c) holds ⇔ C = D .
§7 barycenters INT IV.x253
whereas Z
hbν , hi = hz0 , hi dν(z0 ) = ν(f G) ,
G
whence hbµ , hi = µ(f ) = ν(f G) = hbν , hi .
Proof of ⇒ : The foregoing computations show that if ν is a positive
measure on G of total mass 1 , then the measure µ on i H (X) defined
by (†) is a positive measure of total mass 1 such that b µ = bν ; in particular,
bν = bµ ∈ C (No. 1, Prop. 1), and since D is equal to the set of all such b ν
(same Prop. 1), one has D ⊂ C .
INT IV.x254 extension of a measure. l p spaces §7
But (ϕG f )(iH (y)) is equal to f (iH (y)) when iH (y) ∈ G , and to 0 when
iH (y) ∈
/ G , thus ϕG f is the extension by 0 of the function g ∈ C (G)
defined by
g(iH (y)) = hiH (y), hi for y ∈ F ,
that is, ϕG f = g 0 . Thus
§7 barycenters INT IV.x255
Z Z Z
0
hbµ , hi = ϕG f dµ = g dµ = g dν
iH (X) iH (X) G
Z
= hz0 , hi dν(z0 ) = hbν , hi .
G
The dictionary between Choquet’s theorem (No. 2, Th. 1) and its ap-
plication here is as follows:
0
E ↔ H (equipped with σ(H 0 , H ) )
K ↔ C (the closed convex envelope of i H (X) )
M ↔ Cep (the set of extremal points of C ) .
One knows that iH ChH (X) = Cep ⊂ iH (X) ⊂ C (see the note for
IV.109, `. 12, 13).
To apply Choquet’s theorem, we must verify that C is a metrizable
subspace of H 0 . As noted in the proof of Prop. 4, (i), i H (X) is contained
in the set
B = {z0 ∈ H 0 : kz0 k 6 1 } ,
which is compact for σ(H 0 , H ) (TVS, III, §3, No. 4, Cor. 3 of Prop. 4),
whence C is contained in B and is compact. Moreover, since the com-
pact space X is assumed here to be metrizable, the Banach space C (X)
is separable, i.e., of countable type (GT, X, §3, No. 3, Th. 1), therefore its
linear subspace H is a normed space of countable type; it follows that B
is metrizable for σ(H 0 , H ) (TVS, loc. cit., Cor. 2 of Prop. 6, read “second
axiom of countability”), therefore so is C .
For consistency with the notations in the proof of Prop. 4, it is useful
to replace µ by λ in the statement of Prop. 8: Given any x ∈ X , we seek
a positive measure
R λ on X of total mass 1 such that λ X --- ChH (X) = 0
and h(x) = h dλ for all h ∈ H .
By Choquet’s theorem, Cep is the intersection of a sequence of open
sets in C (i.e., is a Gδ in C ), hence is a Borel set in C (§5, No. 4, Cor. 3
of Th. 2); and, since iH (x) ∈ C , there exists a positive measure ρ on C of
total mass 1 such that bρ = iH (x) and ρ(C --- Cep ) = 0 . From
Cep ⊂ iH (X) ⊂ C
we have ρ C --- iH (X) 6 ρ(C --- Cep ) = 0 , thus the open set C --- iH (X)
in C is ρ-negligible, whence C --- iH (X) ⊂ C --- Supp ρ (Ch. III, §2, No. 2,
Def. 1), that is, Supp ρ ⊂ iH (X) .
Let µ = ρiH (X) be the restriction of ρ to the compact subset i H (X)
of C ; thus,
µ(g) = ρ(g 0 ) for all g ∈ C (iH (X)) ,
where g 0 is the extension by 0 of g to C (§5, No. 7, Def. 4). In particular,
if g = 1 then g 0 = ϕiH (X) (the characteristic function of the subset i H (X)
of C ), and since g 0 = 1 ρ-almost everywhere, one has
Z
µ(1) = ϕiH (X) dρ = ρ(1) = 1 ,
C
§7 barycenters INT IV.x257
Thus bµ = bρ = iH (x) .
Let λ be the positive measure on X of total mass 1 that is paired
with µ as in the proof of Prop. 4, (ii), so that µ is derived from λ via the
homoemorphism iH of X onto iH (X) . Since bµ = iH (x) , we know that
Z
h(x) = h dλ for all h ∈ H
X
E ↔ R3
X ↔ X = {x ∈ R3 : kxk 6 1 }
F ↔ S2 = {x ∈ R3 : kxk = 1 }
H ↔ H ⊂ C (X) as described here.
by Prop. 7 (specifically, b) ⇒ a) ).
On the other hand, if x ∈ S2 then there exists a support (i.e., ‘tangent’)
hyperplane H of X that contains x (TVS, §5, No. 2, Prop. 3). Say
H = {y ∈ R3 : f (y) = c } ,
whence Mh ∩ ChH (X) = {x} and so x ∈ ChH (X) . Thus S2 ⊂ ChH (X) ,
and finally
S2 ⊂ ChH (X) ⊂ ŠH (X) ⊂ S2 ,
whence equality throughout.
IV.110, `. 13–15.
“The Hr -extremal points in X are again called H -extremal, the set
of them is denoted ChH (X) , and the closure of the latter set is denoted
ŠH (X) .”
Thus ChH (X) = ChHr (X) and
Note that the concept of H -extremal point does not depend on the existence
of measures, but is characterized in terms of measures in Prop. 4. If x ∈ X
and if λ is a positive measure on X of total mass 1 , the conditions
Z
(10) h(x) = h dλ for all h ∈ Hr
and
Z
0
(10) f (x) = f dλ for all f ∈ H
and
x ∈ ChH (X) ⇔ x ∈ ChHr (X) (by definition)
⇔ (10) holds only for λ = εx (Prop. 4)
⇔ (10)0 holds only for λ = εx .
INT IV.x260 extension of a measure. l p spaces §7
Then F ⊃ ŠHr (X) = ŠH (X) by “b) ⇒ a)” of Prop. 7, thus a) of the present
proposition holds.
IV.111, `. 14, 15.
“Since g − b ∈ H , the hypothesis on F implies that |g − b| 6 b ”
By the hypothesis on F , |g − b| attains its supremum at some point
a ∈ F ; but |g − b| is 6 b at every point of F , therefore for every x ∈ X
one has
|g(x) − b| 6 sup |g − b| = |g(a) − b| 6 b ,
that is, |g − b| 6 b .
Indeed, since g − b ∈ Hr , one can replace b) by the weaker condition
b)r For every g ∈ Hr , F intersects the set of points of X where |g|
attains its supremum.
IV.111, `. −13 to −11.
“ . . . a point where |f | attains its supremum is a point where one of the
functions f, −f attains its supremum.”
§7 barycenters INT IV.x261
IV.111, `. −7 to −1.
“Lemma 4.”
So to speak, if a ∈ X is an “approximate peak point” and has a count-
able neighborhood base, then a is a “peak point” (relative to H ).
{For the terminology, cf. the book of G.M. Leibowitz (Lectures on com-
plex function algebras, p. 54, Scott, Foresman, Glenview, IL, 1970).}
As indicated by the “resp. C (X; R) ” in the statement of the lemma,
the proof works for either the real or the complex case.
IV.112, `. 2, 3.
“ . . . let λ, µ, ε be numbers such that
The order of events: Fix λ with 0 < λ < 1 . Fix µ with 1 < µ < 1+λ .
Then choose any ε > 0 (which will also remain fixed) such that µ+ε 6 1+λ ,
for example ε = (1 + λ) − µ .
IV.112, `. 4–6.
“We are going to define, by induction on n (n > 1), a decreasing
sequence (Un ) of open neighborhoods of a such that U n ⊂ Vn for all n ,
and a sequence (hn ) of functions in H ”
To get a feeling for the intricate argument, I found it necessary to look
at n = 1, 2, 3 .
One is assuming that
whence
λ 1
(†) 0< < < 1.
µ µ
◦
The argument for n = 1. Define U1 = V1 , set U = U1 , and apply
the hypothesis to the inequalities (†), that is, with c = λ/µ , d = 1/µ and
U = U1 : there exists a function f ∈ H such that
1 λ
(121 ) |f | 6 1 , |f (a)| > , |f (x)| 6 for all x ∈ X --- U1 .
µ µ
1
Define h1 = f . Then
f (a)
|f | 1
(131 ) |h1 | = 6 6 µ;
|f (a)| |f (a)|
§7 barycenters INT IV.x263
f (a)
(141 ) h1 (a) = = 1;
f (a)
|f (x)| λ 1 λ
(151 ) for all x ∈ X --- U1 , |h1 (x)| = 6 · 6 · µ = λ;
|f (a)| µ |f (a)| µ
2
X
1
(161 ) for all y ∈ X , |λ h1 (y)| = λ|h1 (y)| 6 λµ < λ(1 + λ) = λj .
j=1
we have |g| 6 λ|h1 | + λ2 |h2 | , and to prove (162 ) we must show that |g(y)| <
P3
λ3 for all y ∈ X . At any rate, by (161 ) we know that
j=1
2
X
|g| < λj + λ2 |h2 | on X ;
j=1
whereas if y ∈ U2 then |h2 (y)| 6 µ by (132 ) and, by (172 ), |λh1 (y)| <
λ + ελ2 , therefore
P
4
To prove (163 ) we must show that |g(y)| < λj for all y ∈ X . At any
j=1
rate, by (162 ) we know that
3
X
|g| < λj + λ3 |h3 | on X ;
j=1
X2
j
λ h j (y) < λ + λ2 + ελ3 ,
j=1
therefore
|g(y)| = (λ + λ2 + ελ3 ) + λ3 µ
= λ + λ2 + (µ + ε)λ3
4
X
< λ + λ2 + (1 + λ)λ3 = λj (by (∗))
j=1
1
the series is normally convergent since kgk < +∞ , | log c| = | − log | =
c
1
log and
c
n 1 n
X∞
| log c| kgk
∞
X log kgk 1 1 kgk
= c = e(log c ) kgk = < +∞,
n=0
n! n=0
n! c
Pm [(log c)g]n
and f ∈ A because the partial sums belong to A and A is
n=0 n!
norm-closed.
IV.114, `. 3.
“ |f | 6 1 , |f (a)| > cε = d , |f (x)| 6 c for x ∈ X --- U . ”
Write g = h + ik , where h, k ∈ Ar ; thus h = Rg > 0 . For every
x ∈ X,
the first factor is positive and 6 1 since (log c)h(x) 6 0 , and the second
factor has absolute value 1 since (log c)k(x) is real, therefore |f (x)| 6 1 .
Since Rg > 0 and h(a) = Rg(a) 6 ε , one has 0 6 h(a) 6 ε , whereas
log c < 0 , therefore (log c)h(a) > (log c)ε ; citing the formula for f (x) as a
product,
log d
|f (a)| = exp[(log c)h(a)] > exp[(log c)ε] = exp (log c) · = d.
log c
Finally, if x ∈ X --- U then h(x) = Rg(x) > 1 , whereas log c < 0 ,
therefore (log c)h(x) 6 log c and
IV.114, `. 9.
“ c) Let M be the set of subsets M of X such that . . . ”
If M ∈ M and M0 ⊃ M then obviously M0 ∈ M ; thus minimal sets M
are of interest. By No. 4, Prop. 9, Ch A (X) ∈ M . The implication a)T⇒ c)
says that ChA (X) ⊂ M for every M ∈ M , therefore ChA (X) ⊂ M,
M∈M
and
T the reverse inclusion is immediate from Ch A (X) ∈ M ; thus ChA (X) =
M , ChA (X) is the smallest element of M , and
M∈M
M = {A ⊂ X : A ⊃ ChA (X) } .
In particular, ŠA (X) ∈ M .
IV.114, `. 12.
“ d) Let N be the set of subsets N of X such that . . . ”
Such a set N is called a boundary for the algebra A (cf. G.M. Leib-
owitz, op. cit., p. 53, Exer. 4). If N ∈ N and N 0 ⊃ N , then N0 ∈ N . By
Prop. 5 of No. 3 applied to Ar , ChA (X) = ChAr (X) ∈ N ; arguing as in
the preceding note, Tone sees that the meaning of the implication a) ⇒ d) is
that ChA (X) = N , ChA (X) is the smallest element of N , and
N∈N
N = {A ⊂ X : A ⊃ ChA (X) } ,
which is also equal to M . In particular, ŠA (X) ∈ N .
One calls ChA (X) the Choquet boundary, and ŠA (X) the Shilov bound-
ary, for A (Leibowitz, op. cit., p. 49). Another valuable reference for §7 is
the book of Robert R. Phelps, Lectures on Choquet’s theorem, Van Nostrand,
Princeton, NJ, 1966.
IV.114, `. −15, −14.
“ . . . we can restrict ourselves to the case that X does not reduce to the
single point a ”
When X = {a} , each of a), c), d) (resp. b)) is trivially (resp. vacuously)
true.
IV.115, `. 8, 9.
“ . . . the spaces X1 and X2 are homeomorphic, both being bounded
convex sets in R4 with nonempty interior.”
A convex body in a topological vector space over R or C is a closed
convex set with nonempty interior. The theorem “Any two compact convex
bodies in Rn are homeomorphic” is cited in the prerequisites of the book
of E.H. Spanier (Algebraic topology, p. 10, McGraw-Hill, New York, 1966;
reprinted by Springer-Verlag, New York); presumably a proof can be found
in one of the 29 books listed there, but I have not tracked it down.
§7 barycenters INT IV.x269
IV.115, `. 12.
“ . . . pointed convex cone in E .”
With vertex 0 (TVS, II, §2, No. 4).
IV.115, `. 12.
“One knows . . . ”
TVS, II, §2, No. 5, Prop. 13.
IV.115, `. −9.
“The Ck are disjoint convex cones with union C .”
The proof is delicate. Note that, since the f λ are affine, hence convex,
it follows that f is convex (FRV, I, §4, No. 2, Prop. 3). Since the f λ are
positively homogeneous, so is f : f (tx) = tf (x) for x ∈ C , t > 0 .
For every y ∈ C there is at least one index j such that f (y) = f j (y) ,
and y ∈ Ck if and only if k is the first such index; thus C is the union of
the Ck , and the Ck to which y belongs is unique, whence disjointness.
Consider C1 = {y ∈ C : f1 (y) = f (y) } . If y, y 0 ∈ C1 and 0 < t < 1 ,
we are to show that ty + (1 − t)y 0 ∈ C1 . At any rate, ty + (1 − t)y 0 ∈ C ,
and
f1 ty + (1 − t)y 0 = tf1 (y) + (1 − t)f1 (y 0 ) ( f1 is affine)
= tf (y) + (1 − t)f (y 0 ) (because y, y 0 ∈ C1 ) .
Assume to the contrary that ty + (1 − t)y 0 ∈ Ck for some k > 1 . Then, by
the definition of Ck ,
f ty + (1 − t)y 0 > f1 ty + (1 − t)y 0 = tf (y) + (1 − t)f (y 0 ) ,
which contradicts the convexity of f . Thus C 1 is convex; moreover, f C1 =
f1 C1 is affine. Since tC1 ⊂ C1 for t > 0 by the positive homogeneity of f 1
and f , C1 is a convex cone.
“At the other end”, consider
Cp = {y ∈ C : fk (y) < f (y) for 1 6 k < p } .
Of course y ∈ Cp ⇒ fp (y) = f (y) . Suppose y, y 0 ∈ Cp and 0 < t < 1 .
If ty + (1 − t)y 0 did not belong to Cp , then ty + (1 − t)y 0 ∈ Ck for some
k < p , therefore
f ty+(1 − t)y 0 = fk ty + (1 − t)y 0 (because ty + (1 − t)y 0 ∈ Ck )
= tfk (y) + (1 − t)fk (y 0 )
< tfp (y) + (1 − t)fp (y 0 ) (because y, y 0 ∈ Cp and k < p )
= fp ty + (1 − t)y 0
6 f ty + (1 − t)y 0 (by the definition of f ) ,
whence the absurdity f ty + (1 − t)y 0 < f ty + (1 − t)y 0 .
INT IV.x270 extension of a measure. l p spaces §7
by the associativity theorem. Resist writing f (x) for this sum; f is convex
on C but need not be additive.
For use below, we note that the convexity and positive homogeneity
of f implies that f is subadditive, that is, f (x + y) 6 f (x) + f (y) for
x, y ∈ C ; for,
1 1 h1 1 i
f (x + y) = 2f x + y 6 2 f (x) + f (y) = f (x) + f (y) .
2 2 2 2
IV.115, `. −5.
“ (19) f (x) = sup f (y1 ) + · · · + f (yp ) , ”
IV.116, `. 11.
“ . . . weakly complete”
Regarding the product space R × E as a weak locally convex space.
INT IV.x272 extension of a measure. l p spaces §7
x = x1 + · · · + xp , y = y1 + · · · + yp , zi = xi + yi for i = 1, . . . , p .
Since f (zi ) 6 f (xi ) + f (yi ) (see the note for IV.115, `. −7, −6) one has
p p p
X X X
f(x + y) = f (zi ) 6 f (xi ) + f (yi ) 6 f(x) + f(y) ;
i=1 i=1 i=1
p
X p
X p
X
f (y) = f (yi ) = f (ri y) = ri f (y) = 1 · f (y) .
i=1 i=1 i=1
INT IV.x274 extension of a measure. l p spaces §7
IV.117, `. 5, 6.
“Since λ*(K --- (K ∩ G)) = 0 , we have λ(f K) = λ(f K) .”
We have just shown that f = f on G . Since K = (K∩G)∪(K--- K∩G) ,
where f = f on K ∩ G and λ*(K --- K ∩ G) = 0 , we have f K = f K
λ-almost everywhere, therefore λ*(f K) = λ*(f K) (§2, No. 3, Prop. 6).
The function f (resp. f ) is lower (resp. upper) semi-continuous, by
hypothesis (resp. Lemma 6), hence so is its restriction to K (GT, IV, §6,
No. 2). It follows that f K and f K are measurable with respect to any
measure on K (§5, No. 5, Cor. of Prop. 8); being bounded, they are inte-
grable with respect to the (bounded)
measure λ (§5, No. 6, Th. 5), so one
can drop the asterisks: λ(f K) = λ(f K) .
IV.117, `. 9.
“Let x ∈ K be the barycenter of λ .”
Since K is compact and convex, the barycenter of λ —a priori an ele-
ment of E0 * —may be viewed as an element of K (No. 1, Cor. of Prop. 1).
IV.117, `. 9, 10.
“If g ∈ A then λ(g K) = g(x) .”
R
Recall that x = bλ = K y dλ(y) ∈ E0 * . If g ∈ A then g K = z 0 K for
some z 0 ∈ E0 , thus
Z Z
g(x) = hx, z i = hy, z i dλ(y) = (g K)(y) dλ(y) = λ(g K) .
0 0
IV.117, `. 10.
“Therefore λ(f K) = f(x) (§4, No. 4, Cor. 2 of Prop. 5).”
Since λ(g K) = g(x) for every g ∈ A , and since λ(1) = 1 , it is
immediate that the equality holds for every g ∈ A 0 . The functions in A 0
are continuous, f(K) is the lower envelope of a decreasing
directed set D
of functions g K with g ∈ A , and the numbers λ(g K) ( = g(x) ) are
0
by the cited Cor. 2, that is, λ(f K) = inf g(x) = f(x) .
g∈D
IV.117, `. 13.
“ . . . admitting a compact sole M ”
This means (TVS, II, §7, No. 3) that M = C ∩ H , where H is a closed
hyperplane in E that does not pass through the vertex 0 of C , the convex
§7 barycenters INT IV.x275
IV.117, `. −12.
“By Th. 3, λ(f ) = λ0 (f ) for every f ∈ S .”
Given f ∈ S , let f 0 : C → R be a continuous, positively homoge-
neous convex function such that f 0 M = f . Write M+1 (X) for the positive
measures of total mass 1 on a compact space X .
The dictionary between Th. 3 and its Corollary is as follows:
E ↔ E
C ↔ C
f ↔ f0
K ↔ M
f K ↔ f 0 M = f
λ, λ0 ∈ M1+ (K) ↔ λ, λ0 ∈ M1+ (M) .
Regard C (M) as a Riesz space in the usual way, with f 6 g the point-
wise order relation, and with f ∪ g and f ∩ g as the pointwise supremum
and infimum (Ch. III, §1, No. 5). As observed in the note for `. 21, 22, S is
closed under finite sups and infs. The first displayed relation is an applica-
tion, in the Riesz space C (M) , of the invariance of order under translation
by the element f2 + f4 of S (Ch. II, §1, No. 1, formula (5)), and shows
that S − S is closed under finite sups; and the second displayed relation is
a consequence of the first, showing that S − S is closed under finite infs.
(It follows that if f ∈ S − S then |f | = sup(f, −f ) ∈ S − S .)
Conclusion: S −S is itself a Riesz space for the order relation induced
by that of C (M) .
§7 barycenters INT IV.x277
IV.117, `. −8.
“Since hM ∈ S , S − S contains the constant functions.”
The half-space H0 = {y ∈ E : h(y) > 0 } is a pointed cone (albeit
degenerate) with vertex 0 , and M ⊂ H0 (since h(y) = 1 for y ∈ M ),
therefore C ⊂ H0 (see the note for `. 13), that is, h > 0 on C . Thus h M
qualifies for membership in S , that is, the constant function 1 M belongs
to S , whence the assertion.
IV.117, `. −6, −5.
“ . . . this form is the difference of two continuous linear forms that are
positive on C (TVS, II, §6, No. 8, Lemma 1).”
The weak completeness of C plays a role here.
IV.117, `. −5 to −3.
“It follows from the foregoing that for α, β real, there exists f ∈
S − S such that f (x) = α , f (y) = β .”
Say k ∈ E0 with k(x) 6= k(y) (Hahn–Banach), and write k = k 1 − k2
with
ki ∈ E0 and ki > 0 on C for i = 1, 2 (see the preceding note). Then
k M = k1 M − k2 M ∈ S − S , thus S − S does contain a function that
distinguishes between x and y ; the passage to a function in the vector space
S − S (containing 1 ) taking on specified values at x and y is carried out
in the proof of Stone’s theorem (GT, X, §4, No. 1, Th. 2).
CHAPTER V
Integration of measures
case 1. µ• (f ) = 0 .
Then f is locally negligible, and the convention (+∞)
· 0 in R assures
that (+∞)f is also locally negligible, whence µ • (+∞)f = 0 = (+∞)·0 =
(+∞) · µ• (f ) .
case 2. µ• (f ) > 0 .
Then (+∞) · µ• (f ) = +∞ , and f (t) > 0 ⇔ (+∞) · f (t) = +∞ (GT,
IV, §4, No. 3, Prop. 8). Write A = {t : f (t) > 0 } . By hypothesis, A is
not locally negligible, hence there exists a compact set K in T such that
A ∩ K is not negligible. But (+∞)f = +∞ on A ∩ K , thus (+∞)f · ϕ K is
not almost everywhere finite, therefore
µ* (+∞)f · ϕK = +∞ (Ch. IV, §2,
•
No. 3, Prop. 7), whence µ (+∞)f = +∞ and the equality in c) reduces
to +∞ = +∞ .
V.2, `. −2, −1.
The left member γ is obviously 6 the right member δ ; on the other hand,
for all K1 , K2 ∈ K one has
is negligible, and A ∪ B = C .
Let N be a negligible set such that f g , f h , and f (g + h) are finite-
valued on T --- N . Replacement of f by f ϕ T -- N does not change the
numbers α, β, γ , so we can suppose that f g , gh , and f (g + h) are finite
at every point of T . This is useful for the next note.
Caution: The functions f and g + h are permitted to have infinite
values but, at a point where one of them is infinite, the other must be 0 .
V.3, `. 12, 13.
“ . . . then v > f and u > v(g + h) ”
We can suppose, after excluding special cases and modifying f on a
negligible set, that f (g + h) is finite at every point of T (see the preceding
§1 essential upper integral INT V.x4
B = {t : (g + h)(t) = 0 } , C = {t : (g + h)(t) = +∞ }
INT V.x5 integration of measures §1
are also measurable (Ch. IV, §5, No. 5, Prop. 7), therefore the step function
(+∞) · ϕB on T is measurable (loc. cit.), so the function
v = (v A)0 + (+∞) · ϕB + 0 · ϕC = uw + (+∞) · ϕB
where the second equality holds by the cited Prop. 15; the third, by the
theorem on monotone limits (GT, IV, §5, No. 2, Th. 2) and the continuity
of addition in R+ (GT, IV, §4, No. 3); and the other three, by definitions.
V.4, `. 6, 7.
“It follows, by the definition of upper integral, that µ*(f ) 6 µ • (f ) ”
In the notation of Ch. IV, §1, No. 1, f ∈ I + (T) , therefore µ*(f ) =
sup µ(g) as g varies over the set of all functions in K + (T) such that g 6 f
(loc. cit., Def. 1). For any such g , write K(g) for its (compact) support;
then g = gϕK(g) 6 f ϕK(g) and
V.4, `. 20.
“Ch. IV, §1, No. 4, Prop. 19 ”
And Ch. IV, §4, No. 6, Prop. 10.
V.5, `. 14.
“ . . . f also has it.”
P
By assumption, f = fn . What does the notation mean? For
Pn∈N
every t ∈ T , the sum fn (t) exists in R+ as the supremum of the
n∈N
finite
P subsums (GT, IV, §7, No. 5, Prop. 2) and one defines a function
fn ∈ F+ (T) by
n∈N X X
fn (t) = fn (t) .
n∈N n∈N
for, all of the sums exist in R+ , and the equality holds by associativity (GT,
IV, §7, No. 5, Remark ). In other words,
XX X
hnk = hnk ,
n∈N k∈N (n,k)∈N×N
that is, X
f= hnk .
(n,k)∈N×N
V.7, `. 8.
“ inf µ*(hϕK ) > inf µ• (h) − a .”
h∈H h∈H
For all h ∈ H with h 6 h0 , one has µ• (hϕ { K ) 6 µ• (h0 ϕ { K ) 6 a and
so
µ• (h) = µ• (hϕK ) + µ• (hϕ { K ) 6 µ• (hϕK ) + a = µ*(hϕK ) + a ,
thus µ*(hϕK ) > µ• (h) − a , whence the asserted inequality.
V.7, `. −8, −7.
“The closure of 0 for this topology is the space N F∞ ”
p
The set in question is {f ∈ F F : Np (f ) = 0 } (TVS, II, §1, No. 2,
Prop. 2); to say that f : T → F is locally negligible means that µ • (|f |) = 0
(V.2, `. 11, 12). Thus the assertion is that
Np (f ) = 0 ⇔ f = 0 locally almost everywhere.
p
Given f ∈ F F let g ∈ FFp with f = g locally almost everywhere. By the
1/p
definitions, Np (f ) = Np (g) = µ*(|g|p ) .
§1 essential upper integral INT V.x10
Np (f ) = Np (g) = Np (e
g)
(cf. Ch. IV, §3, No. 2), thus the mapping is an isometry. Since F Fp is
complete (loc. cit., No. 3, Prop. 5), F Fp /NF is a Banach space (given a
Cauchy sequence (e gn ) in FFp /NF , the sequence (gn ) is Cauchy in FFp ,
p p
etc.), therefore so is F F /NF∞ , consequently F F is complete.
Caution: Note the author’s avoidance of the notation e f for the class
f + NF∞ of f for equality locally almost everywhere, as this notation has
been pre-empted for f + NF in Ch. IV, §3, No. 2. {In another context, the
notation ḟ = f + NF∞ has been employed (Ch. IV, §6, No. 3).}
V.7, `. −3, −2.
p p p
“We shall similarly denote by L F (T, µ) (or L F (µ) , or L F ) the sub-
p
space LFp + NF∞ of F F ”
p p
As for the notation L F , note that LFp is dense in L F , that is, LFp =
p p p p
L F . For, {0} ⊂ L F ⊂ F F and the closure of {0} in F F is NF∞ ,
p p
therefore the closure of {0} in L F is NF∞ ∩ L F = NF∞ (GT, I, §3, No. 1,
p
Prop. 1); since LFp contains {0} , its closure in L F is a linear subspace
p
containing LFp and NF∞ , hence containing LFp + NF∞ = L F .
INT V.x11 integration of measures §1
p p
Better yet, KF is dense in L F . For, if f ∈ L F , say f = g locally
almost everywhere, where g ∈ LFp , there exists a directed family (gj )
in KF such that Np (gj − g) → 0 (Ch. IV, §3, No. 4, Def. 2); since g j − g =
gj − f locally almost everywhere, one has
Np (gj − f ) = Np (gj − g) → 0 ,
p
that is, gj → f in L F .
V.7, `. −2, −1.
p p
“ . . . one can also characterize L F as the subspace of F F constituted
by the measurable mappings (Ch. IV, §5, No. 6, Th. 5).”
Note that every function in NF∞ is measurable, since it is equal locally
almost everywhere to the measurable function 0 (Ch. IV, §5, No. 2, Prop. 6).
p
Let f ∈ F F , say f = g locally almost everywhere with g ∈ F Fp ; then
f − g ∈ NF∞ , so f − g is measurable.
If f is measurable then so is g = f − (f − g) ; but g ∈ F Fp , therefore
p
g ∈ LFp by the cited Th. 5, thus f = g + (f − g) ∈ L Fp + NF∞ = L F .
Conversely, if f ∈ LFp + NF∞ then f is the sum of two measurable
mappings, hence is measurable.
V.8, `. −15, −14.
“Corollary.”
p
Necessity: f ∈ LFp ⊂ L F , and |f |p (hence also f ) is moderated by
Prop. 7, 1).
p
Sufficiency: If f ∈ L F is moderated then µ*(|f |p ) = µ• (|f |p ) < +∞
(Prop. 7, 2) and Prop. 9); moreover, f is measurable (Prop. 9) therefore
f ∈ LFp (Ch. IV, §5, No. 6, Th. 5).
V.8, `. −13 to −7.
“Definition 3.”
1
In the present context, f denotes elements of L F ; thus the composition
indicated in Def. 3,
1
L F → L1F → F
e = g + NF ∈ L1F (where g ∈ LF1 and f = g locally
is the mapping f 7→ g
almost everywhere), followed by the mapping g e 7→ µ(g) . The continuity of
the composite is immediate from the definitions:
N1 (f ) = N1 (g) = N1 (e
g) = ke
gk1 = µ(|g|)
(cf. Ch. IV, §3, No. 4) and |µ(g)| 6 µ(|g|) (Ch. IV, §4, No. 2, Prop. 2).
§1 essential upper integral INT V.x12
the equalities holding, respectively, by No. 1, Prop. 1, a), No. 2, Prop. 7, 2),
and Def. 3.
V.8, `. −4 to −2.
“If A is a set whose characteristic function is essentially integrable,
then A is said to be an essentially µ-integrable set ”
For a subset A of T , the following conditions are equivalent:
a) A is essentially integrable;
b) there exists an integrable set C such that ϕ A = ϕC locally almost
everywhere;
c) there exists an integrable set C such that the ‘symmetric difference’
of A and C , that is, the set (A --- C) ∪ (C --- A) , is locally negligible.
Proof. b) ⇔ c): For, (A --- C) ∪ (C --- A) = {t : ϕ A (t) 6= ϕC (t) } .
b) ⇒ a): Obvious from the definitions.
INT V.x13 integration of measures §1
which is a measurable set (Ch. IV, §5, No. 5, Prop. 7). If t ∈ C then
g(t) = 1 = ϕC (t) ; that is, g = ϕC on C . To show that ϕA = ϕC locally
almost everywhere, it will suffice to show that
{t : g(t) 6= ϕC (t) } ⊂ { B
V.9, `. −9 to −7.
“To establish a), it suffices to show that for every compact subset L
R∗ R∗
of T , f ϕL dµ = sup f ϕK dµ , where K runs over the set of subsets
K
of L belonging to K .”
For, assuming this to be shown, it would follow that
Z ∗ Z ∗
f ϕL dµ 6 sup f ϕK dµ ,
K∈K
R∗
whence, varying L , µ• (f ) 6 sup f ϕK dµ , whereas the reverse inequality
K∈K
follows trivially from the definition of µ • .
V.10, `. 2.
Z Z Z
“ f dµ − f ϕH dµ 6 |f |ϕ { H dµ ”
valid for f ∈ LF1 (θ) = LF1 (|θ|) (Ch. IV, §4, No. 2, Prop. 2), generalizes at
1
once for f ∈ L F (θ) .
1 1
Let f ∈ L F (θ) . Apply b) of Prop. 10 to µ = |θ| and |f | ∈ L R (µ) to
obtain Z Z
|f | d|θ| = lim |f |ϕK d|θ| ,
K
that is, Z
lim |f |ϕ { K d|θ| = 0 ;
K
R
whence lim f ϕ { K dθ = 0 , that is,
K
Z Z
f dθ = lim f ϕK dθ . ♦
K
N01 (f − gn ) 6 N1 (f − gn ) → 0 ,
θ 0 (f ) = lim θ 0 (gn ) in F .
n
µk (f ) = lim µk (gn ) in F ,
n
therefore
θ(f ) = lim θ(gn ) = lim µ1 (gn ) − µ2 (gn ) + iµ3 (gn ) − iµ4 (gn )
n n
= lim µ1 (gn ) − lim µ2 (gn ) + i lim µ3 (gn ) − i lim µ4 (gn )
n n n n
= µ1 (f ) − µ2 (f ) + iµ3 (f ) − iµ4 (f ) .
Applying b) to each of the four terms on the right, one obtains the desired
formula: for f ∈ LF1 (θ) ,
1
its extension to f ∈ L F (θ) is then immediate. ♦
INT V.x17 integration of measures §1
is bilinear.
Proof. It is clear that Mµ is a vector space. If f ∈ LF1 (µ) and (gn ) is
a sequence in KF such that µ(f − gn ) → 0 then, for every θ ∈ Mµ ,
|θ|*(f − gn ) → 0 ,
by the cited Th. 5, to show that f is µ-integrable it will suffice to show that
it is µ-measurable.
To this end, we employ the criterion of Ch. IV, §5, No. 1, Prop. 1. Let
K be any compact subset of T , and let ε > 0 . Since f is |θ|-measurable,
there exists a compact set H ⊂ K such that
|θ|*(K --- H) 6 ε/2 and f H is continuous ;
§1 essential upper integral INT V.x18
therefore
µ* K --- (H ∪ H0 ) = |θ|* K --- (H ∪ H0 ) + |θ 0 |* K --- (H ∪ H0 )
6 |θ|*(K --- H) + |θ 0 |*(K --- H0 ) 6 ε/2 + ε/2 ;
the coup de grace: f (H ∪ H0 ) is continuous (GT, Ch. I, §3, No. 2, Prop. 4),
thus f is µ-integrable.
Since |θ + θ 0 | 6 |θ| + |θ 0 | = µ (Ch. III, §1, No. 6) it follows from the
Lemma that f is (θ + θ 0 )-integrable and (θ + θ 0 )(f ) = θ(f ) + θ 0 (f ) . ♦
Remark. Recall that, for every measure ρ on T , the ρ-integrable sets
form a clan (Ch. IV, §4, No. 9, Example) that contains every compact set
(loc. cit., No. 6, Cor. 1 of Prop. 10), hence contains the clan R generated by
the compact sets; thus every set in R is ‘universally integrable’ in the sense
that it is ρ-integrable for every measure ρ on T . Thus, in hindsight, one
can omit the asterisks in the above computation (loc. cit., No. 5, Def. 2).
However, with the asterisks in place, the equality in the above display is
justified by the known property (|θ| + |θ 0 |)* = |θ|* + |θ 0 |* of outer measures
(Ch. IV, §1, No. 3, Prop. 15); to justify removing the asterisks would require
verifying the special case of the Theorem being proved, for the measures
|θ|, |θ 0 | and the (universally integrable) numerical function f = ϕ K -- (H∪H0 ) ,
which verification would bring back the asterisks.
The above Theorem extends, by induction, to finitely many complex
1 1
measures; but the analogous proposition for f ∈ L F (θ) ∩ L F (θ 0 ) , a special
case of §2, No. 2, Prop. 3, is more complicated (see (i) of Cor. 3 below for a
simpler proof).
Corollary 1. Let θ be any complex measure on T and suppose θ is
decomposed as in (∗∗) (for example, in the canonical way). Then
4
\
f∈ LF1 (θ) ⇔ f∈ LF1 (µK ) ,
k=1
|θ|(f ) = θ + (f ) + θ − (f ) , θ(f ) = θ + (f ) − θ − (f ) .
Proof. By Cor. 1, LF1 (θ) = LF1 (θ + ) ∩ LF1 (θ − ) and the second formula
holds for f ∈ LF1 (θ) ; and, since LF1 (θ) = LF1 (|θ|) and |θ| = θ + + θ − , the
first formula is immediate from the Theorem. ♦
n
\ 1 1
L F (θk ) = L F (µ) .
k=1
T
n 1 1
(ii) If θ = θ1 + · · · + θn then L F (θk ) ⊂ L F (θ) and, for every
k=1
T
n 1 R R R
f∈ L F (θk ) , one has f dθ = f dθ1 + · · · + f dθn .
k=1
T
n T
n
(iii) LF1 (θk ) = LF1 (µ) ⊂ LF1 (θ) and, for f ∈ LF1 (θk ) , one has
k=1 k=1
θ(f ) = θ1 (f ) + · · · + θn (f ) .
1 1
Proof. {Recall that for any complex measure θ , L F (θ) = L F (|θ|) by
definition.}
(i) For a function f : T → F , µ• (|f |) = |θ1 |• (|f ) + · · · + |θn |• (|f |) by
No. 1, Prop. 3, therefore µ• (|f |) < +∞ if and only if |θk |• (|f |) < +∞ for
1
every k ; in view of No. 3, Prop. 9, to show that f ∈ L F (µ) if and only if
1
f ∈ L F (θk ) for all k , it will suffice to show that
To this end, recall that µ* = |θ1 |* + · · · + |θn |* (Ch. IV, §1, No. 3, Prop. 15).
⇒: If f is µ-measurable, it is clear from |θ k |* 6 µ* and the definition
of measurability that f is θk -measurable.
⇐: Assuming f is θk -measurable for all k , we are to show that f is
µ-measurable. As in the proof of the Theorem, we employ the criterion of
Ch. IV, §5, No. 1, Prop. 1: if K ⊂ T is compact and ε > 0 , there exist
compact sets Hk ⊂ K (k = 1, . . . , n) such that f Hk is continuous and
§1 essential upper integral INT V.x20
|θk |*(K --- Hk ) 6 ε/n ; then f (H1 ∪ · · · ∪ Hn ) is continuous and
n
[ n
X n
[
µ∗ K --- Hj = |θk |* K --- Hj
j=1 k=1 j=1
Xn n
\
= |θk |* (K --- Hj )
k=1 j=1
Xn
6 |θk |*(K --- Hk ) 6 ε ,
k=1
S
∞ T
n
that is, A --- Kki ⊂ Nk ; then the set N = Nk is θk -negligible for
i=1 k=1
every k , hence is µ-negligible, and
n [
[ ∞ n
\ ∞
[ n
\
A --- Kki = A --- Kki ⊂ Nk = N ,
k=1 i=1 k=1 i=1 k=1
n S
S ∞
thus A ⊂ N ∪ Kki and so A is µ-moderated.
k=1 i=1
Application of (†) to (iii): the equivalences
n
\ 1
f∈ LF1 (θk ) ⇔ f ∈ L F (θk ) and f is θk -moderated (k = 1, . . . , n)
k=1
1
⇔ f ∈ L F (µ) and f is µ-moderated
⇔ f ∈ LF1 (µ)
hold, respectively, by the Cor. of Prop. 9; by (i) and (†); and the Cor. of
Prop. 9. ♦
V.10, `. −7, −6.
“ ν*(f ) 6 ν(g) 6 ε , or λ∗α (f ) > λ*(f ) − ε (Ch. IV, §1, No. 3,
Prop. 15).”
Since 0 6 f 6 g , one has ν*(f ) 6 ν*(g) = ν(g) = λ(g) − λ α (g) 6 ε by
the choice of g . By the cited Prop. 15, λ* = (λ α + ν)* = λ∗α + ν* , therefore
λ*(f ) − λ∗α (f ) = ν*(f ) 6 ε , whence the second asserted inequality.
V.10, `. −1.
“ λ• (f ) = sup λ• (fn ) = sup sup λ•α (fn ) = sup sup λ•α (fn ) = sup λ•α (f ) . ”
n∈N n∈N α∈A α∈A n∈N α∈A
V.11, `. 3, 4.
λ• (f ) = sup λ• (f ϕK ) = sup sup λ•α (f ϕK )
K∈K K∈K α∈A
= sup sup λα (f ϕK ) = sup λ•α (f ) .
•
α∈A K∈K α∈A
g λ-negligible ⇒ λα -negligible ,
whence g λ-measurable ⇒ g λα -measurable (Ch. IV, §5, No. 1, comment
following Def. 1).
V.11, `. 13.
“ . . . let L be a compact set such that . . . ”
The assumption that L be compact is unnecessary and confusing.
Assume g is λα -measurable for all α , and let K be the set of all
compact subsets K of T such that g K is continuous. To show that g
is λ-measurable, it suffices to show that K is λ-dense (Ch. IV, §5, No. 10,
Prop. 15, criterion a), with A = X = T ); to this end, it will suffice to show
that K satisfies criterion a) of Ch. IV, §5, No. 8, Prop. 12. Thus, given an
arbitrary subset L of T , we are to show that
V.12, `. −7 to −5.
“Corollary 1”
For every subset M of T ,
X
ν • (M) = λ•α (M)
α∈A
for every ν-integrable set M , and in particular for every compact set, or
relatively compact open set (Ch. IV, §4, No. 6, Cor. 1 of Prop. 10).
§2 summable families of positive measures INT V.x24
where (Kn ) is a sequence of compact sets and ν*(N) = 0 (No. 2, Def. 2);
then, for every α , λα * 6 ν* , therefore λ∗α (N) = 0 and so f is
λα -moderated. By the cited Prop. 7, ν • (f ) = ν*(f ) and λ•α (f ) = λ∗α (f )
for all α , so it follows from (3) that (4) holds with equality.
V.13, `. 11.
“This follows at once from Cor. 2 of Prop. 11 of §1.”
It does, provided one has verified the proposition for finite sums of
the λα . The crux of the matter: If µ, µ0 are positive measures on T , and
f : T → G , then
f is (µ + µ0 )-measurable ⇔ f is measurable for µ and for µ 0 .
From (µ + µ0 )* = µ* + µ0 * one sees that a subset N of T is negligible
for µ + µ0 if and only if it is negligible for each of µ, µ 0 .
Proof of ⇒: Immediate from the preceding remark and the definition
of measurability (Ch. IV, §5, No. 1, Def. 1).
Proof of ⇐: This is shown in the proof of the Theorem in the Note for
p. V.10, `. 13, 14 (see also Cor. 3 of the cited Theorem).
V.13, `. −10, −9.
“The first part of the statement therefore follows at once from Props. 2
and 1.”
At any rate,
X
(∗) λ•α (|f |) = ν • (|f |)
α∈A
by Prop. 1.
INT V.x25 integration of measures §2
both of which are defined and are elements of F ; Φ and Ψ define linear
mappings LF1 (ν) → F .
Φ is continuous for convergence in mean: for, by the definition of inte-
gral, it is the extension to LF1 (ν) of the mapping g 7→ ν(g) (g ∈ KF (T) )
by continuity with respect to the semi-norm N 1 = ν* (Ch. IV, §4, No. 1,
Def. 1).
Ψ is likewise continuous for convergence in mean, by virtue of the com-
putation
X Z XZ
X
|Ψ(g)| 6 g dλα 6 |g| dλα = λα (|g|)
α∈A α∈A α∈A
X
= λ•α (|g|) •
= ν (|g|) = ν(|g|) .
α∈A
V.13, `. −4 to −2.
“ . . . it contains the functions of the form f · a , where a ∈ F and f
denotes a finite integrable positive function (Prop. 1).”
Assuming f ∈ LR1 (ν) , f > 0 , the assertion is that the function f = f ·a
satisfies (6). R R
One has f ·a ∈ LF1 (ν) and (f ·a) dν = f dν ·a (Ch. IV, §4, No. 2,
Cor.
R 2 of Th. 1);R from λα 6 ν it follows that f ∈ LR1 (λα ) , consequently
(f · a) dλα = f dλα · a . By Prop. 1,
X
ν • (f ) = λ•α (f ) ,
α∈A
and since f is integrable (hence moderated) for ν and the λ α , this may be
written Z XZ
f dν = f dλα .
α∈A
Then Z Z XZ
(f · a) dν = f dν · a = f dλα · a
α∈A
X Z XZ
= f dλα · a = (f · a) dλα
α∈A α∈A
V.14, `. 8–11.
“Conversely, if A is finite and if f is λ α -integrable for all α ∈ A ,
then f is essentially ν-integrable by Prop. 3, and it sufficesPto verify that
ν*(|f |) < +∞ ; this follows at once from the relation ν* = λ∗α (Ch. IV,
α∈A
§1, No. 3, Prop. 15).”
The proof cites Prop. 3, which cites Prop. 2, which cites Cor. 2 of
Prop. 11 of §1, whose proof employs the concept of a λ-dense set K of
compact sets.
For A finite, there is a simpler proof: in the Note for p. V.10, `. 13, 14,
see Cor. 3, (iii) of the Theorem proved there.
P •
But for arbitrary A , the additivity property ν • = λα and the
α∈A
concept of λ-dense sets K prove their worth.
V.14, `. 12–17.
“Corollary 2”
See also the note for p. V.10, `. 13, 14 (Cor. 1 of the Theorem proved
there).
V.15, `. 3, 4.
“ . . . the linear form µα on K (T) is positive, therefore is a positive
measure, with support contained in K α .”
That µα is a measure is shown in Ch. III, §1, No. 5, Th. 1.
To show that Supp(µα ) ⊂ Kα , it suffices to show that µα { Kα = 0
(Ch. III, §2, No. 2, Def. 1 and the sentence before it). Indeed, if f ∈ K (T)
and Supp(f ) ⊂ { Kα then f ϕKα = 0 and so µα (f ) = µ(f ϕKα ) = 0 ; thus
µα { Kα = 0 (loc. cit., No. 1; for every g ∈ K ( { Kα ), the extension by 0
of g to T is such a function f ).
{The earlier-cited Prop. 14 (Ch. IV, §5, No. 9) also shows that one can
suppose that Supp(µKα ) = Kα ; an alternate proof of Prop. 4 based on the
measures µKα is given below in the Note for V.15, `. 11, 12.}
V.15, `. 8, 9.
“ . . . let A0 be the countable set formed by the α ∈ A such that
S ∩ Kα 6= ∅ .”
Remark following Ch. IV, §5, No. 9, Def. 7.
V.15, `. 9, 10.
“ . . . the set N ∩ S is µ-negligible ”
Ch. IV, §5, No. 2, Prop. 5.
§2 summable families of positive measures INT V.x28
(Ch. IV, §1, No. 3, Th. 3 or Ch. IV, §4, No. 3, Cor. 2 of Th. 2).
3rd equality: f ϕS∩Kα = (f ϕS )ϕKα = f ϕKα .
4th equality: For α ∈ A --- A0 one has S ∩ Kα = ∅ , whence f ϕKα =
(f ϕS )ϕKα = f ϕS∩Kα = f · 0 = 0 .
Proof of Prop. 4 based on the measures µ Kα . For each α ∈ A , the
correspondence
f 7→ µKα (f Kα ) (f ∈ K (T))
defines a positive linear form µ0α on K (T) , that is, a positive measure
on T . Then Supp(µ0α ) ⊂ Kα ; for, if f ∈ K (T) and Supp(f ) ⊂ { Kα , then
f Kα = 0 and µ0α (f ) = µKα (f Kα ) = 0.
Fix f ∈ K+ (T) and define A0 and S as above. One argues, as above,
that X
µ(f ) = µ(f ϕKα ) ;
α∈A0
INT V.x29 integration of measures §3
since f ϕKα = (fKα )0 , where
(f Kα )0 is the extension by 0 of f Kα
to T , µ (f Kα )0 = µKα (f Kα ) by the definition of µKα (Ch. IV, §5,
No. 7, Def. 4), and f Kα = 0 for α ∈ A --- A0 , one has
X X
µ(f ) = µ0α (f ) = µ0α (f ) .
α∈A0 α∈A
V.15, `. 16.
/ A0 ”
“Then µα = 0 for α ∈
With A0 defined as in the preceding note,
S suppose α ∈ / SA 0 . Then
Kα ∩ Ln = ∅ for all n , thus Kα ⊂ T --- Ln ; but T --- Ln is by
n n
construction µ-negligible, therefore so is K α . Then, forall f ∈ K (T) , the
function f ϕKα is µ-negligible, and so µα (f ) = µ f ϕKα = 0 .
V.17, `. 1–4.
“For, verifying that Λ is scalarly essentially integrable for a positive
measure η on T comes down to verifying that t 7→ λ t (g) is η-measurable
§3 integration of positive measures INT V.x30
and admits a finite essential upper integral, with respect to η , for every
function g ∈ K+ (X) .”
Prop. 9 of §1, No. 3. In what follows, the role of η is played by µ and
the µi .
V.17, `. 4, 5.
“The proposition therefore follows at once from Prop. 11 of §1, No. 4
and its Corollary 2.”
Suppose Λ is scalarly essentially µ-integrable. Given g ∈ K + (X) , let
gb : T → R+ be the function defined by gb(t) = λt (g) ; by assumption,
1
gb ∈ L R (µ) , thus gb is µ-measurable and µ• (b
g ) < +∞ (§1, No. 3, Prop. 9).
Since µ = sup µi , it follows from the cited Prop. 11 that
i∈I
sup µ•i (b
g ) = µ• (b
g ) < +∞ .
i∈I
(a special case of §2, No. 2, Prop. 3; cf. Cor. 3 of the Theorem in the Note
for V.10, `. 13–14). Given a mapping Λ : t 7→ λ t ∈ M+ (X) , for g ∈ K+ (X)
define gb : T → R+ by gb(t) = λt (g) . The following conditions are equivalent:
a) Λ is scalarly essentially µ-integrable;
1
b) gb ∈ L (µ) for all g ∈ K+ (X) ;
1
c) gb ∈ L (µα ) for all α ∈ A and all g ∈ K+ (X) ;
d) Λ is scalarly essentially µα -integrable
R for all α ∈ A , R
in which case the positive measures ν = λt dµ(t) and να = λt dµα (t)
are defined, where, for all g ∈ K+ (X) ,
Z X XZ X
ν(g) = λt (g) dµ(t) = µ• (b g) = µ•α (b
g) = λt (g) dµα (t) = να (g),
α∈A α∈A α∈A
P
that is, ν = να . This proves the Corollary for the case that A is finite.
α∈A P
In the general case, for every finite set J ⊂ A write µ J = µα ;
α∈J
then (µJ ) is an increasing directed family of positive measures, and we
are assuming that the family admits µ as supremum. We know from the
foregoing that Λ is scalarly essentially µ α -integrable for all α ∈ A if and
only if it isP
scalarly essentially µ J -integrable for all finite J ⊂ A , in which
case νJ = να for every J , that is,
α∈J
Z XZ
λt dµJ (t) = λt dµα (t) .
α∈J
§3 integration of positive measures INT V.x32
(the third equality by §2, No.P2, Prop. 1), whence the family (ν α )α∈A is
summable in M+ (X) , with να = ν .
α∈A
(see the Note for V.10, `. 13, 14, proof of (i) in Cor. 3 of the Theorem).
V.17, `. 17–25.
“Definition 1.”
As in the preceding note, regard the scalarly essentially
R µ-integrable
mapping Λ as a mapping F+ (X) → F+ (T) , let ν = λt dµ(t) and, as
in Ch. IV, §1, No. 1, write I+ (X) for the set of all lower semi-continuous
functions f > 0 on X . If f ∈ I+ (X) then f is universally measurable
(i.e., measurable for every measure on X ) by Ch. IV, §5, No. 5, Cor. of
INT V.x33 integration of measures §3
Thus, to say that Λ is µ-pre-adequate means that for every f ∈ I + (X) the
function Λ(f ) is µ-measurable and
ν • (f ) = µ• Λ(f ) ,
or ν 0 *(f ) = µ0• Λ(f ) , where Λ(f ) (t) = λ•t (f ) = λ∗t (f ) for all t ∈ T .
V.17, `. −6 to −4.
R “It can be shown that if Λ is µ-pre-adequate and if the measure ν =
λt dµ(t) is moderated—in particular if X is countable at infinity—then
Λ is µ-adequate (Exer. 7) ”
The exercise (p. V.97):
7) a) Let Λ : t 7→ λt be a Rscalarly essentially integrable mapping
of T into M+ (X) , and let ν = λt dµ(t) . Show that for every lower
semi-continuous function f > 0 defined on X ,
Z • Z • Z •
f (x) dν(x) 6 dµ(t) f (x) dλt (x) .
§3 integration of positive measures INT V.x34
(Ch. IV, §1, No. 1, Def. 1 and the remark following Def. 3 of No. 3).
Let g ∈ K+ (X) , g 6 f . Then, for every t ∈ T, λ•t (g) 6 λ•t(f ) ; that
is, Λ(g) 6 Λ(f ) pointwise on T , therefore µ • Λ(g) 6 µ• Λ(f ) . On the
other hand, by the definition of ν ,
ν(g) = µ• (b
g ) = µ• Λ(g) 6 µ• Λ(f )
(when g ∈ K+ (X) we may use the notation gb in place of Λ(g) ) and finally
ν • (f ) = ν*(f ) = sup ν(g) 6 µ• Λ(f ) .
g∈K+ (X), g6f
(†) ν 0 + ν 00 = ν .
(see, e.g., the Note for p. 17, `. 6–11), therefore, citing the definition of ν, ν 0
and ν 00 ,
ν(g) = µ• (b
g ) = (µ0 + µ00 )• (b
g ) = (µ0• + µ00• )(b
g)
0• 00•
= µ (b g ) + µ (b g ) = ν (g) + ν (g) = (ν + ν 00 )(g) ,
0 00 0
therefore
ν • (f ) = µ• Λ(f ) = (µ0 + µ00 )• Λ(f ) = (µ0• + µ00• ) Λ(f )
= µ0• Λ(f ) + µ00• Λ(f )
> ν 0• (f ) + ν 00• (f ) = (ν 0• + ν 00• )(f )
= (ν 0 + ν 00 )• (f ) = ν • (f )
(the inequality, by (††), and the last equality by (†)); we thus have equality
throughout, and since ν • (f ) = ν*(f ) < +∞ the inequalities in (††) must in
fact be equalities. This proves the first assertion of b).
Suppose now that f is merely a Rν-moderated lower semi-continuous
function > 0 . With µ0 6 µ and ν 0 = λt dµ0 (t) as above, we are to show
that the equality ν 0• (f ) = µ0• Λ(f ) again holds. The key idea (§1, No. 2,
Prop. 5, a)): every ν-moderated set is contained in the union of a sequence
of µ-integrable open sets.
Let A = {x ∈ X : f (x) 6= 0 } . By assumption there exists a
ν-moderated set B ⊂ X such that f = 0 on { B , in other words A ⊂ B ;
since subsets of moderated sets are moderated, it is the same to say that A
is ν-moderated. LetS(Un ) be a sequence of ν-integrable open sets in X
such that A ⊂ U = Un ; we can suppose that the sequence is increasing.
n
The functions f ϕUn are lower semi-continuous (GT, IV, §7, No. 2, Cor.
of Prop. 1 and Prop. 2) and the sequence (f ϕ Un ) is increasing; if x ∈ A
§3 integration of positive measures INT V.x36
fn = inf(n, f ϕUn ) ,
but from fn ↑ f it is immediate that λ•t (fn ) ↑ λ•t (f ) for every t ∈ T (§1,
No. 1, Prop. 1), that is, Λ(fn ) ↑ Λ(f) pointwise on T , and passage to the
limit in (∗) yields ν 0• (f ) = µ0• Λ(f ) .
R
Finally, suppose Λ is µ-pre-adequate, ν = λt dµ(t) , and let µ0 be a
positive measure 6 Rµ ; one knows that Λ is scalarly essentially µ 0 -integrable
and, defining ν 0 = λt dµ0 (t) , obviously ν 0 6 ν . As observed in the Note
for V.17, `. 17–25, in order that Λ be µ 0 -pre-adequate, it is necessary and
sufficient that ν 0• (f ) = µ0• Λ(f ) for every f ∈ I+ (X) , a condition that is
0
satisfied (as shown above) when S every f ∈ I + (X) is ν -moderated. If ν is
moderated, so that X = N ∪ Kn with (Kn ) a sequence of compact sets
n
and N ν-negligible, then N is also ν 0 -negligible, thus ν 0 is moderated and
0
so every function on
R X is ν -moderated. Conclusion: When Λ is µ-pre-
adequate and ν = λt dµ(t) is moderated, then Λ is µ0 -pre-adequate for
every positive measure µ0 6 µ , that is, Λ is µ-adequate.
V.17, `. −4.
“ . . . it is not known if these concepts are in general equivalent.”
The question implicitly posed: Is every µ-pre-adequate mapping
Λ : t 7→ λt automatically µ-adequate? {Perhaps the matter has been settled
by now? (5-14-2007).}
Exercise 7 (see the preceding note) and Prop. B below give an affirmative
answer for some special cases, and Exercise 8 (worked out later in this Note)
effectively reduces the problem to the case that µ has compact support.
First, some simple preliminaries:
Proposition A.PLet Λ : t 7→ λt be scalarly essentially µ-integrable, and
suppose that µ = µi for some summable family (µi )i∈I in M+ (T) . If Λ
i∈I
is µi -pre-adequate for all i ∈ I , then Λ is µ-pre-adequate.
INT V.x37 integration of measures §3
thus Λ is µ-pre-adequate. ♦
§3 integration of positive measures INT V.x38
(the term “decomposition lemma” is introduced at the end of Ch. II, §1,
No. 1). For every i , Λ is µ0i -pre-adequate (No. 1, Def. 1), therefore Λ is
µ0 -pre-adequate by the above Prop. A.
Necessity. If Λ is µ-adequate (in particular, scalarly essentially
µ-integrable) then, for every positive measure µ 0 6 µ , it is obvious from
the definition of adequacy that Λ is also µ 0 -adequate. In particular, Λ is
µi -adequate for every i .
Proof of b): Sufficiency. Suppose that Λ is scalarly essentially
µ-integrable and µi -adequate for all i ∈ I . Given any positive measure
µ0 6 µ , we are to show that Λ is µ0 -pre-adequate.
Define µ0i = inf(µ0 , µi ) for all i ∈ I ; then (µ0i )i∈I is an increasing
directed family whose supremum in M + (T) is equal to inf(µ0 , µ) = µ0
(A, Ch. VI, §1, No. 12, Prop. 13). Since µ 0i 6 µi , Λ is µ0i -pre-adequate for
every i ∈ I , therefore Λ is µ0 -pre-adequate by Prop. B.
Necessity. Same argument as in Part a).
Proof of c): Sufficiency. Suppose Λ is scalarly essentially µ-integrable
and µj -adequate for all j ∈ J . Let A be the set of all finite subsets
INT V.x39 integration of measures §3
P
α of J and, for every α ∈ A , define µα = µj . By assumption,
j∈α
µ = sup µα . By Part a), Λ is µα -adequate for every α ∈ A , therefore
α∈A
it is µ-adequate by Part b).
Necessity. Same argument as in Part a).
Proof of d): Sufficiency. This is the Corollary of Prop. A.
Necessity. If Λ is µ-adequate (in particular, scalarly essentially
µ-integrable) then it is µ0 -pre-adequate for every µ0 6 µ regardless of sup-
port.
V.18, `. 3–11.
“Proposition 2.”
The ingenious argument in the text merges the proofs of Parts a) and b),
at some cost in clarity of the structure of the argument. The following
rearrangement highlights structure, at some cost in repetition.
Proof of a): We employ the notations Λ(f ) and gb as in the preceding
notes (e.g., the Note for V.17, `. −6 to −4); in particular, when f ∈ I + (X) ,
Λ(f ) (t) = λ•t (f ) = λ∗t (f ) for all t ∈ T (§1, No. 1, Prop. 4), and when
g ∈ K+ (X) , gb abbreviates Λ(g) , so that gb(t) = λ •t (g) = λt (g) for
R all t ∈ T .
Since Λ is scalarly essentially µ-integrable, the measure ν = λt dµ(t) is
defined, by the formula ν(g) = µ• (b g ) for g ∈ K+ (X) (No. 1).
The vague continuity of Λ : T → M+ (T) means that for every g ∈
K+ (X) , the function t 7→ λt (g) is continuous on T , that is, gb is continuous.
Let f ∈ I+ (X) and define F = {g ∈ K+ (X) : g 6 f } . The functions
g ∈ F form an increasing directed family of functions whose upper envelope
is f (Ch. IV, §1, No. 1, Lemma), concisely g ↑ f (pointwise on X ). By
definition,
ν*(f ) = sup ν(g)
g∈F
(Ch. IV, §5, No. 5, Cor. of Prop. 8), hence need only show that ν • (f ) =
µ• Λ(f ) . Indeed, for all g ∈ F one has
whence
ν • (f ) = sup ν • (g) = sup µ*(b
g ) = µ* Λ(f ) = µ• Λ(f )
g∈F g∈F
(the 1st and 3rd equalities, by §1, No. 1, Prop. 4 and Ch. IV, §1, No. 1,
Th. 1; the 2nd, by the preceding display; the 4th by §1, No. 1, Prop. 4).
Proof of b): Suppose Λ is scalarly essentially µ-integrable and vaguely
µ-measurable. As in the text, let K be
the set, µ-dense in T , of all compact
sets K ⊂ T such that the mapping ΛK : K → M+ (X) is vaguely continuous
(Ch. IV, §5, No. 10, Prop. 15), and let P(µ α )α∈A be a summable family of
positive measures on T such that µ = µα and Supp µα ∈ K (§2, No. 3,
α∈A
Prop. 4). In particular, the restriction of Λ to Supp µ α is vaguely continuous
for all α ∈ A . To show that Λ is µ-adequate, it will suffice to show that
it is µα -adequate for every α ∈ A (see assertion c) of Exer. 8, proved in
the preceding Note). We are thus reduced to proving the following: If Λ is
scalarly essentially µ-integrable and its restriction to S = Supp µ is vaguely
continuous, then Λ is µ-adequate; since, for every positive measure µ 0 6 µ ,
Λ is µ0 -scalarly essentially integrable and Supp µ 0 ⊂ Supp µ (Ch. III, §2,
No. 2, Prop. 3), it suffices to show that Λ is µ-pre-adequate. In the text,
the argument for this is given inR the proof of Part a), as follows (in slightly
different notation). Let ν = λt dµ(t) ; by definition, ν(g) = µ• (b g ) for
g ∈ K+ (X) (§1, No. 1).
Let f ∈ I+ (X) and let F = {g ∈ K+ (X) : g 6 f } . Given any
g ∈ K+ (X) , by the vague continuity assumption the restriction to S of the
mapping t 7→ gb(t) = λt (g) is continuous, that is, gbS is continuous; write g
(instead of hg ) for the function on T defined by
(
gb(t) for t ∈ S
g(t) =
+∞ for t ∈ T --- S .
(the 3rd equality, by §2, No. 2, Prop. 8; the 4th by the next-to-last display).
V.18, `. 15–17.
“Similarly, set
(the 1st equality, by §1, No. 1, Prop. 4; the 2nd, by Ch. IV, §1, No. 1, Def. 1
and No. 3, Def. 3; the 3rd and 4th repeat definitions of the notations gb
and hg ).
V.18, `. −13.
“Set hf = sup hg ; then hf = hf on S .”
g∈F
§3 integration of positive measures INT V.x42
whereas if t ∈ S then
kfbn − fk
b K = sup |λt (fn ) − λt (f )| .
t∈K
§3 integration of positive measures INT V.x44
therefore
kfbn − fk
b K 6 ε kϕk
b K for n > n0
we see that fbK is the uniform limit of the continuous functions fbn K ,
hence is continuous.
Since f ∈ K (X) is arbitrary, this shows that the restriction to K of
the mapping Λ : t 7→ λt is vaguely continuous; and since K ∈ K is arbitrary
and K is µ-dense in T , Λ is indeed vaguely µ-measurable (Ch. IV, §5,
No. 10, Prop. 15). A tour de force.
V.20, `. −4 to −2.
“The first of the inequalities (6) then follows from the definition of
R∗
f (x) dν(x) (Ch. IV, §1, No. 3, Def. 3), and the second follows immediately
from it.”
The new actor in Prop. 3 is the function
Z ∗
(∗) t 7→ f (x) dλt (x) = λ∗t (f ) (t ∈ T)
and the second inequality of (6) becomes obvious: λ ∗t (f ) > λ•t (f ) for all
(1)), thus Λ*(f ) > Λ(f ) pointwise on T , whence
t ∈ T (§1, No. 1, formula
µ• Λ*(f ) > µ• Λ(f ) (loc. cit., Prop. 1). As for the first inequality, since
(Ch. IV, §1, No. 3, Def. 3), it will suffice to show that
ν*(g) > µ• Λ*(f )
for every lower semi-continuous function g > f ; for such a function g , one
has (§1, No. 1, Prop. 4)
A = {t ∈ T : h(t) < +∞ } .
(the last equality due to the fact that λ t is bounded locally almost every-
where, and Prop. 7 of § 1).”
As in earlier notes, let us write Λ(f ) for the function t 7→ λ •t (f )
and Λ*(f ) for the function t 7→ λ∗t (f ) , so that Λ(f ) 6 Λ*(f ) pointwise
on T by (1) of §1, No. 1.
For locally µ-almost every t ∈ T one has λ ∗t (1) = λ•t (1) < +∞ and
so λt is bounded (Ch. IV, §4, No. 7, Prop. 12); but if λ t is bounded then
it is moderated (§1, No. 2, Remark 2), therefore f is λ t -moderated (loc.
cit., comment following Def. 2) and so λ ∗t (f ) = λ•t (f ) (loc. cit., Prop. 7).
It follows that Λ*(f ) = Λ(f ) locally µ-almost
everywhere,
therefore locally
µα -almost everywhere, and so µ•α Λ*(f ) = µ•α Λ(f ) .
§3 integration of positive measures INT V.x48
for every α and for every function f > 0 on X , and summing over α
yields (8):
ν • (f ) > µ• Λ*(f ) = µ• Λ(f )
(§2, No. 2, Prop. 1). Note that the first equality in (6) is sharpened in (8)
R∗ R•
(because > ), and the second inequality in (6) becomes an equality
in (8).
V.21, `. 18–24.
“Corollary 1”
{Cor. 1, Prop. 4 and Prop. 5 are variations on the same general pattern;
as the variations are numerous and subtle, it is helpful to have the pattern
in mind as one thrashes through the details.
One is given a mapping Λ : t 7→ R λt that is scalarly essentially
µ-integrable, so that the measure ν = λt dµ(t) is defined; certain con-
ditions are imposed on f and ν , and one shows that the same conditions
are satisfied for “most” of the λt . The progression from a) to c) tends to
be from general to special:
(i) In assertion a), Λ is assumed to be µ-pre-adequate, and “most”
means “for locally µ-almost every t ”.
(ii) To the hypotheses of a), assertion b) adds the assumption that Λ
is vaguely continuous, and “most” is sharpened to “for µ-almost every t ”.
(iii) In assertion c), Λ is assumed to be µ-adequate, and the function
t 7→ λ•t (1) = λ∗t (1) is assumed to be finite-valued locally µ-almost every-
where, in other words, the measure λ t is assumed to be bounded for locally
µ-almost every t ∈ T ; “most” means “for locally µ-almost every t ”.}
R
Let Λ : t 7→ λt be scalarly essentially µ-integrable and let ν = λt dµ(t) .
Assume in a) and b) that Λ is µ-pre-adequate, and in c) that Λ is
µ-adequate. Let f ∈ F+ (X) and set
H = {t ∈ T : λ∗t (f ) > 0 } = {t ∈ T : Λ*(f ) (t) > 0 } .
V.22, `. 9–11.
“Let S be the set of t ∈ T such that N is not λ t -negligible: S is
locally µ-negligible (resp. µ-negligible) by Cor. 1 of Prop. 3.”
Part a) (resp. Part b)) of Cor. 1 is being applied to the ν-negligible
function ϕN , the set S being denoted H there.
V.22, `. 11–13.
“The sets Kn , B, N are measurable for every measure on X , and the
restriction of f to each of them is λt -measurable for every t ∈
/ S .”
Since B and the Kn are closed sets, all sets in the formula
[
N = (X --- B) --- Kn
n
are Borel sets, hence are ‘universally measurable’ (Ch. IV, §5, No. 4, Cor. 3
of Th. 2).
By construction, f B is a constant function, say equal to y 0 ∈ G at
every point of B ; its extension to X by y 0 is constant on X , hence contin-
uous, hence ‘universally measurable’, in particular measurable for every λ t
(t ∈/ S) .
The function f Kn is continuous. Let K be the set of all compact
subsets of Kn . Obviously K is closed under finite unions, thus K satisfies
the conditions (PLI ) and (PLII ) of Ch. IV, §5, No. 8, Prop. 12; moreover,
condition b) of the cited Prop. 12 is satisfied with A = K n , K = K0 , and
any measure on X , therefore K is dense in K n for every measure on X
(loc. cit., Def. 6). Thus K and f Kn satisfy condition a) of Ch. IV, §5,
No. 10, Prop. 15 with A = Kn , for every measure on X , therefore f Kn is
measurable for every measure on X (loc. cit., Def. 8) and in particular for
every λt (t ∈ / S) .
Let t ∈/ S . Then N is λt -negligible (hence also locally λt -negligible),
therefore f N is λt -measurable (see the Note for IV.79, `. −17, −16).
Thus, for every t ∈/ S , the restriction of f to B , N and every K n is
λt -measurable.
V.22, `. 13–14.
“The function f is therefore λt -measurable for every t ∈
/ S (Ch. IV,
§5, No. 10, Prop. 16).”
Continuing the preceding note: since the set consisting of B , N and
the Kn is countable, a fortiori locally countable, and since
[
X = B∪N∪ Kn ,
n
§3 integration of positive measures INT V.x52
it follows that f = f X is λt -measurable for every t ∈
/ S (Ch. IV, §5, No. 10,
Prop. 16).
V.22, `. −4 to −2.
“The assertions concerning the set N have already been established
(Prop. 4, and Cor. 2 of Prop. 3).”
Let
L = {t ∈ T : f is not λt -moderated }
M = {t ∈ T : f is not λt -measurable } ;
/ { L ∩ { M , thus
then t ∈ N means that t ∈
N = L ∪ M.
(Ch. IV, §1, No. 3, Th. 3 and Ch. V, §1, No. 1, Prop. 1). If, moreover,
the hn are ρ-measurable, then so is h (Ch. IV, §5, No. 4, Cor. 1 of Th. 2).
If the hn are ρ-moderated, then so is h (Ch. V, §1, No. 2, Remark 3).
P
(β) If the hn are ρ-measurable and h = hn , then h is ρ-measurable:
n∈N
the measurability of h0 + h1 follows from Ch. IV, §5, No. 3, Th. 1, ap-
plied to the mapping x 7→ h0 (x), h1 (x) ( x in the given space), with
u(a, b) = a + b for (a, b) ∈ A0 × A0 , where A0 = R ∪ {+∞} (GT, IV,
§4, No. 3, Prop. 7); the measurability of the finite sums h 0 + · · · + hn follows
INT V.x53 integration of measures §3
by Ch. IV, §5, No. 6, Cor. 4 of Th. 5 and Ch. V, §1, No. 1, Cor. of Prop. 2).
If the hn are ρ-moderated, then so is h (Ch. V, §1, No. 3, Remark 3). ∗
By the cited Prop. 6, there exists a sequence f n (n = 0, 1, 2, . . .) of
numerical functions > 0 on X such that f 0 satisfies 1), and the fn for
n > 1 satisfy 2) and are finite-valued. Moreover, the f n for n > 1 are
ν-measurable and ν-moderated: for f 0 , this is obvious, and it is obvious
the fn for n > 1 are moderated (for any measure on X ). It remains to
check that if n > 1 then fn is measurable. Let Kn be a compact set in X
such that fn Kn is continuous and fn = 0 outside Kn , let a ∈ R and let
A = {x ∈ X : fn (x) > a } ;
and, when Λ is vaguely continuous, that the function Λ*(f n ) : t 7→ λ∗t (fn )
is µ-measurable and µ-moderated, and satisfies
(ii) ν*(fn ) = µ* Λ*(fn ) ,
P
n
by (β). Thus Λ(gn ) = Λ(fk ) locally µ-almost everywhere on T , conse-
k=0
quently
n
X n
X
•
•
(iii) µ Λ(gn ) = µ Λ(fk ) = ν • (fk ) = ν • (gn ) for all n
k=0 k=0
by (α), and similarly ν • (f ) = sup ν • (gn ) . From (iii) and (iv) we see that
n
ν • (f ) = sup ν • (gn ) = sup µ• Λ(gn ) = µ• Λ(f ) .
n n
P
n
Since, by assumption, the Λ(fn ) are µ-measurable, so are the Λ(fk )
k=0
by (β), therefore so are the Λ(gn ) (Ch. IV, §5, No. 2, Prop. 6), therefore so
is Λ(f ) by (α). And since every gn is µ-measurable and µ-moderated, so
is f = sup gn by (α).
n
P
n
by (β), thus Λ*(gn ) = Λ*(fk ) µ-almost everywhere for all n ; since the
k=0
Λ*(fn ) are by assumption µ-measurable and µ-moderated, it follows that
Pn
Λ*(fk ) is µ-measurable and µ-moderated by (β), therefore Λ*(g n ) is
k=0
µ-measurable by Ch. IV, §5, No. 2, Prop. 6, and is obviously µ-moderated.
Moreover,
n
X
µ* Λ*(gn ) = µ* Λ*(fk ) for all n
k=0
by (β); but
n
X n
X
µ* Λ*(fk ) = ν*(fk ) = ν*(gn )
k=0 k=0
Finally,
ν*(f ) = sup ν*(gn ) = sup µ* Λ*(gn ) = µ* Λ*(f )
n n
H = {t ∈ T : f is not λt -negligible };
§3 integration of positive measures INT V.x56
thus Λ(h) = Λ(f ) + Λ(g) locally µ-almost everywhere. Since Λ(g) and
Λ(h) are essentially µ-integrable, so is Λ(f ) ; for, if u, v ∈ L 1 (µ) are equal,
locally µ-almost everywhere, to Λ(g) and Λ(h) , respectively (§1, No. 3,
Def. 3), then, for locally µ-almost every t ∈ T ,
and the ν-integrability of g and h implies that that Λ*(g) and Λ*(h) are
µ-integrable (Ch. IV, §5, No. 6, Th. 5).
As shown at the outset, g is λt -measurable and λt -moderated for
µ-almost every t , and similarly for h , so the same is true of f = h − g , and
thus Λ*(h) = Λ*(f ) + Λ*(g) µ-almost everywhere. Since Λ*(g) and Λ*(h)
are µ-integrable, they are finite µ-almost everywhere (Ch. IV, §2, No. 3,
Prop. 7), hence there exist functions u, v ∈ L 1 (µ) such that Λ(g) and Λ(h)
are equal, µ-almost everywhere, to u and v , respectively. Then, for
µ-almost every t ∈ T ,
v(t) = λ∗t (f ) + u(t) ,
therefore Λ*(f ) is equal µ-almost everywhere to the µ-measurable function
v − u , hence is µ-measurable (Ch. IV, §5, No. 2, Prop. 6). It follows that
µ* Λ*(h) = µ* Λ*(f ) + µ* Λ*(g)
thus Λ*(f ) is µ-integrable (Ch. IV, §5, No. 6, Th. 5), hence is µ-moderated
(§1, No. 2, Prop. 7), and satisfies (10).
Thus parts a) and b) of Prop. 5 are proved in full generality.
V.23, `. −12.
“ f is να -measurable and να -moderated ”
Because f is ν-measurable and 0 6 να 6 ν ; and because να is bounded
(§1, No. 2, Remark 2).
V.23, `. −9.
“It remains only to sum on α , applying Props. 1 and 2 of §2, No. 2.”
By the cited Prop. 1, f satisfies (9).
By the cited Prop. 2, the function Λ(f ) : t 7→ λ •t (f ) is µ-measurable.
By assertion a), N is locally µα -negligible for all α , therefore N is
locally µ-negligible by Cor. 2 of the cited Prop. 1.
INT V.x59 integration of measures §3
therefore Λ(f ) is essentially µ-integrable (§1, No. 3, Prop. 9). It follows that
Λ(f ) is finite locally µ-almost everywhere, that is, the set
P = {t ∈ T : λ•t (f ) = +∞ }
have (in these notes) denoted numerical functions that are defined every-
where on T , either because f > 0 or because f ∈ K (X) (No. 1); even in
the Cor. of Prop. 5 of No. 2, it is Λ(|f |) that figuresRin the proof, not Λ(f ) .
From now on, we have to deal with functions t 7→ f dλt that are defined
1
only on the set of t ∈ T such that f ∈ L (λt ) (or L 1 (λt ) ), and which
may have values in a Banach space.
A review of the core results on numerical functions that are involved:
(i) f is measurable ⇔ f + and f − are measurable, in which case
|f | = f + + f − is measurable (Ch. IV, §5, No. 3, Cors. 2 and 3 of Th. 1).
INT V.x61 integration of measures §3
H+ = {t ∈ T : f + is not λt -integrable },
H− = {t ∈ T : f − is not λt -integrable };
(Ch. IV, §4, No. 2, Cor. 2 of Th. 1). Thus, forRsuch an f , H = ∅ and,
writing Λ(f ) for the vector-valued function t 7→ f dλt (t ∈ T), one has
Λ(f ) = aΛ(f )
INT V.x63 integration of measures §3
1
Being finite-valued, Λ(f ) belongs to L R (µ) (resp. LR1 (µ) ). Let
h ∈ LR1 (µ) with Λ(f ) = h locally µ-almost everywhere (resp. µ-almost
everywhere;
R Rwhen Λ is vaguely continuous, h = Λ(f ) will serve). Then
Λ(f ) dµ = h dµ , thus Z
ν(f ) = h dµ .
(loc. cit., Cor. 2 of Th. 1); since Λ(f ) = ah locally µ-almost everywhere
(resp. µ-almost everywhere), Λ(f ) is essentially µ-integrable (resp. µ-inte-
grable) and
Z Z
R
Λ(f ) dµ = ah dµ = h dµ · a = ν(f ) · a = ν(f ) ,
(see item (β) in the Note for p. V.22, `. −2 to V.23, `. 3), therefore g is
ν-integrable.
V.24, `. −11, −10.
“ . . . N1 is locally µ-negligible (resp. µ-negligible) by formula (6) (resp.
(7)).”
Proof #1. As observed in the preceding note, the function g is ‘univer-
sally measurable’ and ν-integrable. In particular, g is λ t -measurable for all
t ∈ T , thus g is λt -integrable if and only if λ∗t (g) < +∞ ; in other words,
N1 = {t ∈ T : g is not λt -integrable },
therefore h0 is finite µ-almost everywhere (Ch. IV, §2, No. 3, Prop. 7);
let B be a µ-negligible set such that h 0 is finite on { B . If t ∈ { (A ∪ B) =
{ A ∩ { B , then h(t) = h(t0 ) and h0 (t) is finite, thus h(t) is finite on the
complement of the locally µ-negligible set A ∪ B . ♦
Proof #2: By (6) (resp. (7)) of No. 2, Prop. 3,
µ• Λ*(g) 6 ν*(g) < +∞ (resp. µ* Λ*(g) 6 ν*(g) < +∞ ) ,
INT V.x65 integration of measures §3
it is to the sequence (fn0 )n∈N and the measure λt that the cited Prop. 6 will
be applied.
As observed in the Notes for `. 19, 20 and 21–26, for every measure on X
the fn are integrable—hence so are the fn0 —and g is measurable. Since
t∈/ N1 , λ∗t (g) < +∞ , therefore g is λt -integrable and, by the argument in
the latter Note, X
λ∗t (|fn0 |) = λ∗t (g) < +∞ ,
n∈N
that is, the series with general term λ ∗t (|fn0 |) is summable. Therefore, by
the cited Prop. 6, for λt -almost every x ∈ X the series with general term
fn0 (x) is absolutely convergent in F ; the n ’th partial sum of the series is
equal to fn (x) , thus, writing Nt for the set of all such x , we may define a
function X → F by
lim fn (x) for x ∈ { Nt
n→∞
ft (x) =
0 for x ∈ Nt .
V.24, `. −3 to −1.
“Suppose that t does not belong to N1 ∪ N2 ; the sequence (fn ) con-
verges in mean in LF1 (λRt ) , and converges
R λt -almost everywhere to f . There-
fore f ∈ LF1 (λt ) and f dλt = lim fn dλt (Ch. IV, §4, No. 1).”
n→∞
/ N1 we may construct ft ∈ LF1 (λt ) as in the preceding Note;
Since t ∈
in particular, fn → ft λt -almost everywhere in X , and fn → ft in mean in
LF1 (λt ) .
Since t ∈
/ N2 , M is λt -negligible, that is, fn (x) → f (x) for λt -almost ev-
ery x ∈ X ; but already fn (x) → ft (x) for λt -almost every x ∈ X , therefore
f = ft λt -almost everywhere. Since ft ∈ LRF1 (λt ) we have Ralso f ∈ LF1 (λt )
and fn → f in mean in LF1 (λt ) , whence f dλt = lim fn dλt (Ch. IV,
n→∞
§1, No. 1, Def. 1).
In particular, since f is λt -integrable we have t ∈
/ H , and we have shown
that H ⊂ N1 ∪ N2 ; therefore H is locally µ-negligible (resp. µ-negligible).
V.25, `. 2–4. R
“ . . . the function t 7→ f dλt is equal locally µ-almost everywhere to
the limit of a sequence of µ-measurable functions; it is therefore µ-measur-
able.”
Since the fn belong to H Rwe know (see the Note for V.24, `. 19, 20)
that the functions Λ(fn ) : t 7→ fn dλt are defined everywhereR in T , are
essentially µ-integrable (resp. µ-integrable), and satisfy ν(f n ) = Λ(fn ) dµ .
Moreover (see the R precedingR Note) for every t ∈ { (N1 ∪ N2 ) one has
f ∈ LF1 (λt ) and fn dλt → f dλt .
Define a function Λ(f ) : T → F by the formula
( R
lim fn dλt for t ∈ { (N1 ∪ N2 )
n→∞
(∗) Λ(f ) (t) =
0 for t ∈ N1 ∪ N2 .
The functions ϕ { (N1 ∪N2 ) Λ(fn ) are λt -measurable and converge pointwise
on T to Λ(f ) , therefore Λ(f ) is λt -measurable (Ch. IV, §5, No. 4, Th. 2).
Moreover, Λ(fn ) → Λ(f ) locally µ-almost everywhere (resp. µ-almost
everywhere) in T by (∗).
V.25, `. 8–11.
R∗
“Now, the function t 7→ g(x) dλt (x) is essentially µ-integrable (resp.
µ-integrable) by Prop. 5. We may therefore apply Lebesgue’s theorem, which
yields
Z Z Z Z Z
dµ(t) f (x) dλt (x) = lim dµ(t) fn (x) dλt (x) = lim fn (x) dν(x). ”
n→∞ n→∞
INT V.x67 integration of measures §3
b) With notations as in the preceding Note, consider first the case that
Λ is vaguely continuous, the Λ(fn ) are µ-integrable, and Λ(fn ) → Λ(f )
µ-almost everywhere in T . By the case of a positive numerical function, we
know that the function Λ*(g) (defined everywhere on T ) is µ-integrable;
moreover, as just noted in the text, |Λ(f n )| 6 Λ*(g) on { (N1 ∪ N2 ) , that is,
µ-almost everywhere in T . It follows from Lebesgue’s theorem (Ch. IV, §3,
No. 7, Th. 6) that RΛ(f ) is µ-integrable
R and Λ(f n ) converges to Λ(f ) in mean
for µ , therefore Λ(fn ) dµ → Λ(f ) dµ . Finally since, by construction,
fn → f in mean for ν , we have
Z Z Z Z
f dν = lim fn dν = lim Λ(fn ) dµ = Λ(f ) dµ ,
n→∞ n→∞
we are
R to show thatR Hg is locally µ-negligible, Λ(g) is essentially µ-integrable,
and Λ(g) dµ = g dν .
The letter f was reserved for a Rν-integrable
R function such that g = f
locally ν-almost everywhere, so that g dν = f dν (§1, No. 3, Def. 3), and
to which part a) of the theorem is to be applied; we can suppose that if g
is a numerical function then f ∈ LR1 (ν) , that is, all values of f are finite
(which facilitates subtraction). Writing
H = {t ∈ T : f is not λt -integrable} ,
we know from part a) that H is locally µ-negligible, that the function Λ(f )
defined by Z
f dλt
for t ∈
/H
Λ(f ) (t) =
0 for t ∈ H
R R
is essentially µ-integrable, and that Λ(f ) dµ = f dν .
INT V.x69 integration of measures §3
Λ(f ) : t 7→ λt (f ) , Λ0 (f ) : t 7→ λ0t (f )
thus Λ0 is µ-pre-adequate.
(iii) If Λ is µ-adequate then so is Λ0 . For, given any positive mea-
sure ρ 6 µ , Λ = Λ0 locally ρ-almost everywhere and Λ is ρ-pre-adequate,
therefore so is Λ0 .
§3 integration of positive measures INT V.x70
Λ(f ) : t 7→ λt (f ) (t ∈ T)
R
Summarizing: Λ is scalarly essentially
R µ-integrable;
λt dµ(t) = ν , because
the left side of (∗) is equal to λt dµ(t) (f ) ; and Λ has the contemplated
relation to H .
Will any of this prove to be useful? (See also §4, No. 4, Remark.)
V.26, `. −6, −5.
“ N is a universally measurable set”
We are to show that the (numerical) function ϕ N is universally measur-
able. For every t ∈ T let Vt be a compact (hence ‘universally integrable’)
neighborhood of t in T . By the Principle of Localization (Ch. IV, §5, No. 2,
Prop. 4) it suffices to show that for every t ∈ T there exists a universally
INT V.x71 integration of measures §3
measurable function gt : T → R such that ϕN Vt = gt Vt . Now, ϕN Vt is
equal to 1 on Vt ∩ N and to 0 on Vt --- Vt ∩ N , therefore
ϕN Vt = ϕVt ∩N Vt ;
argument that they are all universally measurable is given in the Note for
V.22, `. 11–13.
V.26, `. −2.
“Definition 3.”
We record here some more-or-less immediate consequences of the defi-
nition.
(i) If Λ : t 7→ λt is pre-adequate for every positive measure on T with
compact support, then Λ is a diffusion.
For, if µ is a positive measure with compact support then every positive
measure µ0 6 µ also has compact support (Ch. III, §2, No. 2, Prop. 3).
(ii) If µ is a positive measure on T and Λ : t 7→ λ t is a mapping
T → M+ (X) such that Λ is µ-pre-adequate, it is implicit that Λ is scalarly
essentially µ-integrable (No. 1, Def. 1).
(iii) Let µ be any positive measure on T . One knows that µ is the sum
of a family (µα )α∈A of positive measures on T with compact support (§2,
No. 3, Prop. 4). Thus, if Λ : t 7→ λt is a diffusion then Λ is µα -adequate
(hence scalarly essentially µα -integrable) for every α , but one does not know
whether Λ is scalarly essentially µ-integrable.
However, if Λ is scalarly essentially µ-integrable and if µ is the sum of
a family (µα )α∈A of positive measures such that R Λ is µ α -pre-adequate for
every α , then Λ is µ-pre-adequate and (with signifying essential integral)
Z XZ
λt dµ(t) = λt dµα (t)
α∈A
thus Λ(g) is essentially µ-integrable (§1, No. 3, Prop. 9). But µ is bounded
(Ch. III, §2, No. 3, Prop. 11), hence moderated (§1, No. 2, Remark 2),
therefore every function on T is µ-moderated (loc. cit., Def. 2), whence
Λ(g) is µ-integrable (§1, No. 3, Cor. of Prop. 9).
V.27, `. −14, −13.
P 1
“ . . . then u is not integrable for the measure µ = ε
2 tn
with
n>1 n
compact support ”
1
(i) The family of measures εt (n > 1) is summable: for, if
n2 n
f ∈ K+ (T) then for every finite set J of integers > 1 one has
X 1 X 1 X 1
2
εtn (f ) = 2
f (tn ) 6 kf k < +∞ ,
n n n2
n∈J n∈J n>1
For, h is the upper envelope of the set A of all f ∈ K + (T) such that
f 6 h , and, by definition,
{Since h is universally measurable (Ch. IV, §5, No. 5, Cor. of Prop. 8), it
follows that h is εt -integrable ⇔ h(t) < +∞ (loc. cit., No. 6, Th. 5).}
(iii) µ*(u) = +∞ (whence the assertion).
By definition, µ*(u) is the infimum of µ*(h) for all lower semi-continuous
functions h > u , so it will suffice to show that µ*(h) = +∞ for all such h .
Indeed, by §2, No. 2, Cor. 3 of Prop. 1, one has
X 1 ∗ X 1 X 1
µ*(h) > ε t (h) = h(t n ) > u(tn ) = +∞
n
n2 n n
n2 n
n2
1
because u(tn ) > 1 . ♦
n2
R example. The computation in (ii) throws light on the formula
An
µ = εx dµ(x) of Ch. III, §3, No. 1, Example 2, where it is restricted to
functions in K (X) and is merely a notation.
In the present context, consider the mapping Λ : x 7→ ε x of X into
M+ (X) , that is, T = X and λx = εx for all x ∈ X . Since εx is bounded,
ε•x = ε∗x (§1, No. 2, Cor. 2 of Prop. 7), thus, for every function f > 0
on X , we have Λ(f ) (x) = ε•x (f ) = ε∗x (f ) for all x ∈ X . In particular, if
f ∈ K+ (X) the function
H = {x ∈ X : f is not εx -integrable}
Λ(f ) : x 7→ ε∗x (f )
R R
is µ-integrable, and f dν = Λ(f ) dµ , that is
Z Z
(∗) f dµ = Λ(f ) dµ .
As in the proof of the cited Th. 1, one extends this result, in particular (∗),
to an arbitrary µ-integrable function f by considering |f | = f + + f − and
f = f + −f − . We shall show that the underlying reason for the equality (∗) is
that Λ(f ) = f µ-almost everywhere; this entails a close look at the meaning
of εx -integrability.
Fix a point a ∈ X .
(i) The key fact is that Supp(εa ) = {a} . (See the Note for III.29,
`. 2–4.)
(ii) The negligible sets for εa are the subsets of X --- {a} , that is,
the subsets of X that do not contain a . For, X --- {a} = { Supp(εx ) is
εa -negligible; whereas if a ∈ A ⊂ X then ε ∗a (A) > ε∗a ({a}) = 1 because
ϕ{a} = ϕX = 1 εa -almost everywhere, therefore A is not ε a -negligible.
For something to happen εa -almost everywhere in X , it is necessary
and sufficient that it happen at a ; for, the set A where it does not happen
is εa -negligible if and only if a ∈
/ A.
For example, if f, g are functions on X then f = g ε a -almost every-
where if and only if f (a) = g(a) .
(iii) Every function f on X is εa -measurable (obvious from Ch. IV, §5,
No. 1, Def. 1, with N = K --- {a} ). It follows that f is ε a -integrable if and
only if ε∗a (|f |) < +∞ (loc. cit., No. 6, Th. 5).
(iv)
R A numerical function f is εa -integrable ⇔ f (a) is finite, in which
case f dεa = f (a) .
⇒: If f is εa -integrable, then the set A = {x ∈ X : f (x) = ±∞ } is
negligible for εa (Ch. IV, §2, No. 3, Prop. 7), thus a ∈
/ A.
INT V.x75 integration of measures §3
V.28, `. 16–18.
“ . . . this amounts to saying (in view of Prop. 8) that Λ is scalarly
essentially µ-integrable and hµ0 Λ, f i = hµ0 , Λf i for every positive measure
µ0 6 µ and every lower semi-continuous positive function f . ”
(i) Suppose that the diffusion Λ is scalarly essentially µ-integrable and
satisfies the stated condition. Then, for every positive measure µ 0 6 µ ,
one knows that Λ is scalarly essentially µ 0 -integrable (see the Note for V.17,
`. 12, 13), so that the measure µ0 Λ = ν 0 is defined; and, for every lower semi-
continuous function f > 0 , Λf is universally measurable by 1) of Prop. 8,
so the assumption that
ν 0• (f ) = µ0• (Λf )
for all such f says that Λ is µ0 -pre-adequate, and this, for every positive
measure µ0 6 µ , thus Λ is µ-adequate (No. 1, Def. 1).
(ii) Conversely, if µ is a positive measure on T such that Λ is
µ-adequate, then it is clear from No. 1, Def. 1 that Λ is scalarly essen-
tially µ-integrable and satisfies the stated equality for every positive measure
µ0 6 µ and every lower semi-continuous function f > 0 (with the added
spice, via 1) of Prop. 8, that Λf is universally measurable).
It turns out (Prop. 11 below) that the stated condition is superfluous:
the diffusion Λ is µ-adequate for a measure µ > 0 if and only if it is scalarly
essentially µ-integrable. {See also item (v) in the Note for V.26, `. −2.}
V.28, `. 19–28.
“Proposition 10.”
See the next Note.
V.28, `. 28.
“See also the next proposition.”
An excellent suggestion. Prop. 11 has already been verified in an earlier
note (item (v) in the Note for V.26, `. −2): the proposition says that if Λ is
a diffusion and µ is a positive measure on T , then Λ is µ-adequate if and
only if it is scalarly essentially µ-integrable. The proof of Prop. 10 can then
be conducted as follows.
§3 integration of positive measures INT V.x78
thus
Λ(f + g) : t 7→ (Λf )(t) + (Λg)(t) = (Λf + Λg)(t)
Λ(af ) : t 7→ a(Λf )(t) = a(Λf ) (t) ,
whence Λ(f + g) = Λf + Λg and Λ(af ) = a(Λf ) .
b) Assuming that Λ is µ-adequate and ν-adequate, so that, in partic-
ular, Λ is scalarly essentially integrable for both µ and ν , we are to show
that Λ is (µ + ν)-adequate and that (µ + ν)Λ = µΛ + νΛ , that is,
Z Z Z
λt d(µ + ν)(t) = λt dµ(t) + λt dν(t) .
(in the Note for V.10, `. 13, 14, see item (i) of Cor. 3 of the Theorem there),
therefore Λf is essentially (µ + ν)-integrable and
V.29, `. 6, 7.
“Corollary 1.”
Note that kλt k = λ∗t (1) < +∞ and kµk = µ*(1) < +∞ (Ch. IV,
§4, No. 7, Prop. 12); moreover, λ•t = λ∗t and µ• = µ* because bounded
measures are moderated (§1, No. 2, Remark 2 and Prop. 7).
To show that Λ is µ-adequate, it suffices (Prop. 11) to show that Λg is
µ-integrable (hence essentially µ-integrable) for every g ∈ K + (X) . At any
rate, Λg is universally measurable (Prop. 8). For all t ∈ T one has
therefore Λg is µ-integrable (Ch. IV, §5, No. 6, Th. 5). Thus Λ is scalarly
essentially µ-integrable and
Hf : x 7→ ηx• (f ) = ηx∗ (f )
V.30, `. 4–6.
“ . . . we shall denote by Γ the mapping ΛH of T into M + (Y) , and
by Γf the function t 7→ hγt , f i (an abuse of notation, since we do not yet
know whether Γ is a diffusion).”
This No. is about the composition of diffusions, an operation that is
applied to mappings. How are we to regard a diffusion as a mapping? Let
me count the ways . . .
1◦ When mappings Λ : t 7→ λt were introduced in No. 1, clearly Λ was
being regarded as a mapping T → M+ (X) .
2◦ But Λ also induces a mapping F+ (X) → F+ (T) : if f ∈ F+ (X)
the corresponding element of F+ (T) is the function
t 7→ λ•t (f ) (t ∈ T) ,
DΛ = {µ ∈ M+ (T) : Λ is µ-adequate } .
DEΛ = {µ ∈ M+ (T) : µE = µ ∈ DΛ } = DΛ
hλt H, f i = hλt , Hf i ,
{The parentheses on the left side of (∗) are provisional, pending the
status of Γ ; on the right side, they group together the symbols that make
up the universally measurable function on which Λ is acting.}
(ii) Let f > 0 be a universally measurable function on Y . By (i),
Hf is universally measurable on X . Since Λ is a bounded diffusion of X
in Y , it follows from the Scholium that Λ(Hf ) is a universally measurable
function on T and, for every measure µ ∈ D Λ (the domain of Λ ),
λ∗t (Hg) 6 kHk kgk λ∗t (1) = kHk kgk kλt k 6 kHk kgk kΛk
for all t ∈ T , whence 0 6 Λ(Hg) 6 kHk kgk kΛk . Thus the universally
measurable function Γ(g) = Λ(Hg) is bounded, hence it is integrable for the
bounded measure µ .
Anticipating
R that Λ will prove to be a diffusion, we denote by µΓ the
integral γt dµ(t) of Γ (without relinquishing control to auto-pilot).
(iv) If µ is a bounded measure on T , then (µΛ)H = µΓ .
Here, µΓ is the measure on Y defined in (iii). Since µ is a bounded
measure on T , by the Scholium µ ∈ DΛ and µΛ is a bounded measure
on X , whence µΛ ∈ DH and (µΛ)H is a bounded measure on Y . For
every g ∈ K+ (Y) ,
by the definition of µΓ , the equality (∗∗), and the fact that µΛ ∈ D H ; thus
the measures µΓ and (µΛ)H on Y are equal.
§3 integration of positive measures INT V.x84
that is, hµΓ, hi = hµ, Γ(h)i . Indeed, since µ and µΛ are bounded and h
is universally measurable, one has
(the first equality by (iv), the second and third by the Scholium, and the
fourth by (∗∗). Thus Γ is a diffusion.
Finally, kγt k = kλt Hk 6 kλt k kHk 6 kΛk kHk < +∞ by the Scholium,
thus the diffusion Γ is bounded.
(vi) If µ ∈ DΛ and µΛ ∈ DH (so that µΛ and (µΛ)H are defined)
then µ ∈ DΓ and µΓ = (µΛ)H .
Since Γ is a diffusion, it suffices by Prop. 11 to show that Γ is scalarly
essentially µ-integrable. Given g ∈ K + (Y) we are to show that Γg is
essentially µ-integrable.
At any rate, Γg and Hg are universally measurable (Scholium) and
consequently µΓ = (µΛ)H .
(vii) By definition, ΛH is a notation for Γ ; thus, assuming that µ ∈ D Λ
and µΛ ∈ DH , we know from (vi) that µ ∈ DΛH ; we are to show that the
INT V.x85 integration of measures §3
In the first line of the display, the first equality holds by (∗∗), while
hµΓ, f i = hµ, Γf i holds by the Scholium.
In the second line, the first equality holds by (vi), the second by (∗).
(viii) Let Λ : t 7→ λt and H : x 7→ ηx be bounded diffusions of T in X
and of X in Y , respectively. If µ is any bounded positive measure on T
then, by the Scholium, µ ∈ DΛ and µΛ is bounded, hence µΛ ∈ DH . Thus,
viewing Λ and H as mappings
Λ : DΛ → M+ (X) , H : DH → M+ (Y) ,
Λ ◦ H ⊂ Γ,
V.31, `. −15.
R∗ R•
“ f (x) dλt (x) = f (x) dλt (x) = f π(t) g(t) ”
The assertion is that λ∗t (f ) = λ•t (f ) = f π(t) g(t) . Since λt is a
bounded measure, λ•t = λ∗t (§1, No. 2, Cor. 2 of Prop. 7).
If g(t) = 0 then λt = 0 , so λ∗t (f ) = 0 , and f π(t) g(t) = 0 (even if
f π(t) = +∞ ).
Otherwise, 0 < g(t) < +∞ and λt = g(t)επ(t) , whence
λ∗t (f ) = g(t)ε∗π(t) (f ) = g(t)f π(t)
by item (v) of the Example in the Note for V.27, `. −14, −13.
V.31, `. −13, −12.
“Every function (with values in a topological space) defined on X is
λt -measurable for every t ∈ T . ”
See item (iii) of the Example cited above.
V.31, `. −12, −11.
“Every Rmapping f of X into a Banach space F is λ t -integrable for all
t ∈ T , and f (x) dλt (x) = f π(t) g(t) . ”
If g(t) = 0 then λt = 0 and it is trivial that f is λt -integrable with
integral 0 . {The gory details are worked out at the end of this note.}
Otherwise 0 < g(t) < +∞ and λt = g(t)επ(t) . We know that f is
measurable for λt ; moreover,
λ∗t (|f |) = g(t)ε∗π(t) (|f |) = g(t) · |f | π(t) < +∞
By Th. 1 of Ch. IV, §4, No. 2, u ◦ f is λt -integrable (see also the next Note)
and Z Z
u f dλt = (u ◦ f ) dλt ;
but Z Z
(u ◦ f ) dλt = g(t) (u ◦ f ) dεπ(t) = g(t)(u ◦ f ) π(t)
(the second equality, by item (iv) of the above-cited Example), whence the
asserted equality.
{Write θ for the zero measure on X . One has θ*(|f |) = 0 ; for, θ(g) = 0
for every g ∈ K+ (X) , therefore θ*(h) = 0 for every lower semi-continuous
function h > 0 (Ch. IV, §1, No. 2, Def. 2), whence θ*(f ) = 0 for every
numerical function f > 0 (loc. cit., No. 3, Def. 3). In particular |f | is
θ-negligible for every function f with values in a Banach space F or in R ,
thus f = 0 almost everywhere for θ , whence f is θ-integrable with inte-
gral 0 (p. IV.33, `. −3).}
V.31, `. −10 to −8.
“Finally, if f is an arbitrary numerical function defined on X , for f
to be λt -integrable
R it is necessary andsufficient that f π(t) g(t) be finite,
in which case f (x) dλt (x) = f π(t) g(t) . ”
As shown in item (iv) of the above-cited Example, R f is integrable
for επ(t) if and only if f π(t) is finite, in which case f dεπ(t) = f π(t) .
Then,
in which case the asserted equality is evident from the preliminary remark.
V.31, `. −4, −3.
“ 2◦ For every function f ∈ K (X) , the mapping t 7→ f π(t) g(t) is
essentially µ-integrable. ”
Defining λt = g(t)επ(t) for t ∈ T , the validity of 2◦ for every
f ∈ K+ (X) says that the mapping Λ : t 7→ λt ∈ M+ (X) is scalarly es-
sentially µ-integrable (§3, No. 1).
Conversely, if the assertion in 2◦ holds for every f ∈ K+ (X) , then it
holds for every f ∈ K (X) ; for,
(f ◦ π)g = (f + − f − ) ◦ π g = (f + ◦ π)g − (f − ◦ π)g
INT V.x89 integration of measures §4
K = {K ∩ K0 : K ∈ Kπ and K0 ∈ Kg }
is also µ-dense. To this end, we verify that K satisfies the conditions (PL I ),
(PLII ) and b) of Ch. IV, §5, No. 8, Prop. 12.
(PLI ): If K ∩ K0 ∈ K and B is a closed subset of K ∩ K0 , then B ⊂ K
and B ⊂ K0 , whence B ∈ Kπ ∩ Kg and so B = B ∩ B ∈ K .
(PLII ): Suppose K1 ∩ K01 and K2 ∩ K02 belong to K . Then the set
B = (K1 ∩ K01 ) ∪ (K2 ∩ K02 ) is a closed subset of the set
are µ-dense in S . Obviously KΛ|S ∩ Kg|S = K (the set in the assertion). One
proves, as in the preceding Note, that the set
T π-X
g f
? ?
R R
−1 −1
gR : π ({x}) 7→ inf g π ({x}) (x ∈ π(T)) ;
V.32, `. −1.
“ . . . and nonempty ”
−1
Since f (x) 6 a < +∞ , π (x) is nonempty by the definition of f .
V.32, `. −1 to V.33, `. 1.
−1
“ . . . there exists a t ∈ π (x) such that . . . ”
Recall that the restriction of g to any subspace S of T is lower semi-
continuous, since {t ∈ S : g(t) 6 a } = S ∩ {t ∈ T : g(t) 6 a } .
V.33, `. 10, 11.
“For every numerical function f > 0 defined on X ,
Z • Z •
(1) f (x) dν(x) = f π(t) g(t) dµ(t) . ”
(the second equality by §1, No. 1, Prop. 4; or by the fact that λ t is bounded,
hence moderated, cf. §1, No. 2, Prop. 7). Thus (∗) may be written ν*(1) =
µ• (g) , or, since µ is bounded,
ν*(1) = µ*(g) .
R
thus ν is bounded and kνk = g dµ .
V.33, `. −12.
“In view of formula (6) of §3, No. 2 . . . ”
The context here is the µ-adequate mapping Λ : t 7→ λ t = g(t)επ(t) , for
which the cited formula (6) yields
Z ∗ Z • Z ∗
f (x) dν(x) > dµ(t) f (x) dλt (x) ,
R∗
where f (x) dλt (x) = f π(t) g(t) (V.31, `. −15) and µ• = µ* (preceding
Note), thus Z ∗ Z ∗
f (x) dν(x) > f π(t) g(t) dµ(t) .
Ka = {t ∈ K : u(t) 6 a }
h(t) + ε
Ka = {t ∈ K : g(t) > 0 and 6 a}.
g(t)
u|K v
? ?
R R
INT V.x95 integration of measures §4
−1
v(x) = inf{u(s) : s ∈ π ({x}) ∩ K } 6 u(t)
−1
because t ∈ π ({x}) ; then
v (π(t) g(t) = v(x)g(t) 6 u(t)g(t) = h(t) + ε
by the definition of u .
V.34, `. 7–9.
Z ∗ Z ∗
f (x) dν(x) 6 v(x) dν(x)
Z ∗ Z ∗
(4) = v π(t) g(t) dµ(t) 6 h(t) + ε dµ(t)
K
Z ∗
= h(t) dµ(t) + εµ(1) .
that is, ν • (v) = µ• (v ◦ π)g (p. V.31, `. −15). Since v is lower semi-
continuous and µ is bounded, this may be written ν*(v) = µ* (v ◦ π)g ,
which is the asserted equality.
V.34, `. 13, 14.
“ . . . the set K of compact subsets K of T such that the restrictions of
π and g to K are continuous is µ-dense (Ch. IV, §5, No. 10, Prop. 15).”
See also the Note for V.32, `. 9–11.
V.34, `. 16, 17.
“ . . . the pair (π, g) being µα -adapted for every α ∈ A ”
If (π, g) is µ-adapted, then it is µ0 -adapted for every positive measure
0
µ 6 µ:
1◦ π and g are µ0 -measurable (because µ-negligible sets are µ 0 -negligi-
ble);
2◦ every essentially µ-integrable function is essentially µ 0 -integrable; in-
1 1 1
deed, writing µ00 = µ − µ0 , one has L (µ) = L (µ0 ) ∩ L (µ00 ) (see Corol-
lary 3 in the Note for V.10, `. 13, 14).
V.34, `. −8 to −6.
“Corollary.”
We propose to apply Th. 1 to the function f = ϕ N . Since ϕN ◦ π =
ϕπ−1 (N) , by Th. 1 one has
ν • (N) = ν • (ϕN ) = µ• (ϕN ◦ π)g = µ• ϕπ−1 (N) g ;
−1
{t : (ψ ◦ π)(t) 6= 0 } ⊂ π (K) ;
−1
since π is proper, π (K) is compact (GT, I, §10, No. 2, Prop. 6), therefore
−1
Supp(ψ ◦ π) = {t : (ψ ◦ π)(t) 6= 0} ⊂ π (K) ,
whence ψ ◦ π ∈ K (T) .
V.35, `. 4, 5.
“ . . . the pair (π, g) is therefore µ-adapted ”
The functions π and g are universally measurable and, for every
ψ ∈ K (X) , the function (ψ ◦ π)g is continuous with compact support,
hence is integrable with respect to every measure on T . Thus (π, g) is, so
to speak, ‘universally adapted’, that is, µ-adapted for every positive measure
µ on T .
Remark. It then follows from No. 1, Prop. 1 that the mapping
Λ : t 7→ λt = g(t)επ(t) is µ-adequate for every positive measure µ on T ,
in particular for every positive measure with compact support, thus Λ is a
diffusion of T in X (§3, No. 5, Def. 3) whose domain is all of M + (T) .
See also the Example in the Note for V.32, `. 9–11.
V.35, `. 5, 6.
“ . . . the mapping t 7→ g(t)επ(t) is vaguely continuous.”
For every ψ ∈ K (X) , the mapping t 7→ g(t)ε π(t) (ψ) = ψ (π(t) g(t) is
continuous (with compact support), whence the assertion (§3, No. 1).
Thus, for the diffusion Λ : t 7→ λt = g(t)επ(t) (see the preceding Note),
Λψ = (ψ ◦ π)g ∈ K+ (T) for all ψ ∈ K+ (X) .
V.35, `. −10.
“ 1◦ f(x) > f (x) for all x ∈ X (since g(t) > 0 for all t ∈ T ).”
If x ∈
/ π(T) , then f (x) = +∞ and the inequality is trivial. If x ∈ π(T)
−1 −1
then f(x) is defined as an infimum over π (x) ; now, for every t ∈ π (x)
one has
h(t)
f (x) = f π(t) 6
g(t)
−1
by the choice of h , and taking the infimum over t ∈ π (x) yields f (x) 6
f(x) .
§4 integration of positive point measures INT V.x98
V.35, `. −9.
“ 2◦ f π(t) g(t) 6 h(t) for all t ∈ T .”
Fix t ∈ T and set x = π(t) . Thus f π(t) = f(x) is defined as an
−1 −1
infimum over π (x) . Now, π (x) is compact since π is proper (GT, I, §10,
−1
No. 2, Th. 1), and h/g is lower semi-continuous on T , hence on π (x) ,
−1
therefore h/g attains its infimum on π (x) (GT, IV, §6, No. 2, Th. 3).
−1
Thus there exists a point t0 ∈ π (x) such that
h(t0 ) h(s) −1
f(x) = 6 for all s ∈ π (x) ,
g(t0 ) g(s)
and in particular f(x) 6 h(t)/g(t) , whence f π(t) g(t) 6 h(t) .
V.35, `. −2 to V.36, `. 4.
“ Proposition 3. ”
The main players:
T π-X
g f
? ?
R+ G
V.36, `. 7.
“ . . . is µ-dense in S (Ch. IV, §5, No. 10, Prop. 15).”
Since π is µ-measurable, its restriction to the µ-measurable
set S is
µ-measurable in the sense of loc. cit., Def. 8 (because π S has an extension
to T that is µ-measurable, namely π itself; see the condition c 00 ) in the
Note for IV.79, `. 3, 4). Therefore K is µ-dense in S by the equivalent
condition a) of the cited Prop. 15.
V.36, `. 8–10.
“ . . . it therefore suffices to prove that for every K ∈ K , the set of com-
pact subsets H of K , such that the restriction of f ◦ π to H is continuous,
is µ-dense in K (Ch. IV, §5, No. 8, Prop. 13).”
Write H for the set of all compact subsets H of S such that (f ◦π) H is
continuous. One sees easily that H satisfies the conditions (PL I ) and (PLII )
of loc. cit., Prop. 12, that is: if H ∈ H then every closed subset of H belongs
to H ; and if H1 , H2 ∈ H then H1 ∪ H2 ∈ H (GT, I, §3, No. 2, Prop. 4).
Thus if, for every K ∈ K , the set of all H ∈ H such that H ⊂ K is
shown to be µ-dense in K , it will follow from the cited Prop. 13 that H
INT V.x99 integration of measures §4
is µ-dense in S , and this will establish that (f ◦ π) S is µ-measurable by
criterion a) of Ch. IV, §5, No. 10, Prop. 15.
Review of the cited Prop. 13: With the notations K and H as in that
Prop., given K ∈ K note that the set of H ∈ H such that H ⊂ K is the
trace H ∩ K of H on K , that is,
{H ∈ H : H ⊂ K } = {H ∩ K : H ∈ H } .
HK0 = H ∩ K0 = {H ∈ H : H ⊂ K0 } ,
§4 integration of positive point measures INT V.x100
is µ-dense in K0 (loc. cit., No. 8, Prop. 13, with T playing the role of X
in Prop. 13, and S the role of A ). Since H satisfies (PL I ) and (PLII ), so
does HK0 .
Let K0 ∈ K . To show that HK0 is µ-dense in K0 , it will suffice to show
that HK0 satisfies the criterion c) of loc. cit., Prop. 12 (with T playing the
role of X , K0 the role of A , and HK0 the role K ). Thus, let K be any
compact subset of K0 . Then K ∈ K . Apply to K the arguments of the first
part of the proof to produce a partition
−1 [ −1
K = K ∩ π (N) ∪ K ∩ π (Cn ) ,
n
−1 −1
where K ∩ π (N) is µ-negligible (`. 14, 15) and K ∩ π (Cn ) ∈ H for all n
−1 −1
(`. 16); since K ∩ π (Cn ) ⊂ K ⊂ K0 , one has K ∩ π (Cn ) ∈ HK0 , and the
criterion c) is verified.
V.36, `. −17.
“ . . . let us show that N is locally ν-negligible.”
It will then follow from criterion a) of Ch. IV, §5, No. 8, Prop. 12 that
L is µ-dense (in X ).
V.36, `. −14.
“ . . . is by hypothesis µ-dense in S (Ch. IV, §5, No. 10, Prop. 15).”
The argument for this is given in the Note for V.32, `. 9–11.
V.36, `. −13.
−1
“It therefore suffices to prove that π (N) ∩ H is µ-negligible for every
H ∈ H .”
By criterion a) of Ch. IV, §5, No. 8, Prop. 12.
V.36, `. −12 to −10.
“ π(H) is compact and may be identified with the quotient space of H
by the equivalence relation π(t) = π(t 0 ) , π being identified with the canon-
ical mapping of H onto this quotient space (GT, I, §5, No. 2, Prop. 3).”
Since π H is a continuous mapping of the compact space H into the
Hausdorff space X , it is a closed mapping hence is eligible for application
of the cited Prop. 3. One has the diagram
p- h - π(H) i - X.
H H/R
0
where R is the equivalence relation in H defined by π(t) = π(t ) . The
elements of H/R are the sets of constancy of π H . Let us write ṫ for the
equivalence class of t ∈ H , that is,
−1 −1
ṫ = H ∩ π π(t) = π H π(t) (t ∈ H) ;
INT V.x101 integration of measures §4
thus p(t) = ṫ for t ∈ H , h(ṫ) = π(t) for ṫ ∈ H/R , and i is the canonical
injection of π(H) into X .
By definition, H/R bears the quotient topology, that is, the final topol-
ogy for the mapping p , and the topology of the subspace π(H) is the initial
topology for the mapping i . By the cited Prop. 3, h is a homeomorphism,
therefore the topology on π(H) is also the final topology for h ; by transi-
tivity (GT, I, §2, No. 4, Prop. 7) the topology on π(H) is the
final topology
for the mapping h ◦ p : H → π(H) , which is the mapping π H : H → π(H) .
Consequence: a mapping
π(H) → G will be continuous if and only if
its composition with π H is continuous (loc. cit., Prop. 6).
V.36, `. −10 to −8.
“Since the restriction of f ◦ π to H is continuous, the restriction of f
to π(H) is therefore continuous ”
The continuity of (f ◦ π) H = f π(H) ◦ (π H) implies the continuity
of f π(H) by the preceding Note.
V.36, `. −2 to V.37, `. 2.
“Remark.”
One is contemplating Prop. 3 with G = F (a Banach space).
Suppose (f ◦π)S is µ-measurable. This is equivalent to the measurabil-
ity of the function h = (f ◦π)ϕS , because h is the extension by 0 of (f ◦π) S
to T (Ch. IV, §5, No. 10, Prop. 15, criterion d)). Then hg is also measur-
able on T (loc. cit., No. 3, Cor. 5 of Th. 1). Since S = {t ∈ T : g(t) 6= 0 }
one has g = ϕS g , therefore (f ◦ π)g = (f ◦ π)ϕS g = hg is measurable.
Conversely, assume that (f ◦ π)g is measurable. The function g S is
measurable (its extension by 0 to T is the measurable function g ) and takes
its values in ]0, +∞[ ; composing it with the continuous function u(a) = 1/a
on ]0, +∞[ , it follows that the function
1
: t 7→ 1/g(t) (t ∈ S)
g S
But |f | is also essentially ν-integrable, and (|f | ◦ π)g = |h| , therefore |h| is
µα -integrable and
Z • Z •
(ii• ) |f | dνα = |h| dµα .
The equality
Z • XZ •
•
(i ) |f | dν = |f | dνα
α∈A
by (ii• ), (i• ) and the essential ν-integrability of f ; it then follows from the
cited Prop. 3 that h is essentially µ-integrable and
Z XZ
(iii) h dµ = h dµα (absolutely summable) .
α∈A
(the 1st and 3rd equality, by §2, No. 2, Prop. 1, the 2nd equality by (ii • )),
therefore f is essentially ν-integrable, and the relation (9) holds by the first
part of the argument.
§4 integration of positive point measures INT V.x104
V.38, `. 4, 5.
“The second part of the statement therefore follows from the first part
and Proposition 2.”
By the cited Prop. 2 and
the inequality (1) of §1, No. 1, we have the
diagram where h = (f ◦ π)g :
the first equality by the definition of the left member, the second because
|g0 |ϕK is moderated. Thus
Z • Z ∗
|g|ϕK dµ < +∞ ⇔ |g0 |ϕK dµ < +∞ ;
thus Z • Z ∗
|gh| dµ < +∞ ⇔ |g0 h| dµ < +∞ ,
f 0 h = g0 h almost everywhere
N = {t : f 0 (t) 6= g0 (t) }
1
In particular, the locally convex topology on L loc generated by the
semi-norms pK is characterized by the validity of the equivalence a) ⇔ b)
of the Proposition.
V.39, `. −12 to −10.
1
“The associated Hausdorff space, the quotient of L loc (T, θ; F) by the
∞
subspace NF of mappings that are zero locally almost everywhere, is de-
noted L1loc (T, θ; F) .”
The space NF∞ wasR defined in Ch. IV, §6, No. 3. As in the preceding
1 1
Note, we write pK (g) = |g|ϕK d|θ| for every g ∈ Lloc = Lloc (T, θ; F) and
every compact set K ⊂ T .
Proof that NF∞ ⊂ Lloc 1
: Let g ∈ NF∞ . For every compact K ⊂ T , the
function gϕK is locally negligible and moderated, hence negligible (§1, No. 2,
1
Cor. 1 of Prop. 7), therefore integrable, whence g ∈ L loc and pK (g) = 0
for every K .
1
Conversely, if g ∈ Lloc and pK (g) = 0 for every compact K ⊂ T , then
the integrable function |g|ϕK is negligible, whence g is locally negligible
(Ch. IV, §5, No. 2, Prop. 5). Thus
NF∞ = {g ∈ Lloc
1
: pK (g) = 0 for every compact set K ⊂ T } .
1
Since the semi-norms pK define the topology of Lloc , it follows that
NF∞ 1
is the closure of {0} in Lloc (TVS, II, §1, No. 2, Prop. 2, (i)), there-
1
fore the quotient space Lloc /NF∞ is the Hausdorff topological vector space
associated with Lloc (loc. cit., No. 3). The space L1loc = Lloc
1 1
/NF∞ is also
locally convex (TVS, II, §4, No. 4, Example I), its topology being defined by
1
the family of semi-norms ġ 7→ pK (g) ( ġ the equivalence class of g in Lloc
for equality locally almost everywhere); the details are as follows.
As L1loc bears the final topology for the quotient mapping u : L loc 1
→
1 1
Lloc (loc. cit.), a semi-norm q on Lloc is continuous if and only if the semi-
1
norm q◦u : g 7→ q(ġ) on Lloc is continuous. Every continuous semi-norm p
1 ∞
on Lloc is zero on NF (because the pK are), so the formula q(ġ) = p(g)
defines a semi-norm q on L1loc such that p = q ◦ u , and q is continuous
by the foregoing. In particular, for every compact set K ⊂ T the formula
qK (ġ) = pK (g) defines a continuous semi-norm on L 1loc . On the other hand,
if q is any continuous semi-norm on L 1loc , then the semi-norm q ◦u on Lloc 1
In detail:
1
(i) Definition of u . Let g ∈ Lloc (T, µ; F) , and fix an index α . One
knows that g is µ-measurable and gϕKα is µα -integrable (Def. 1 and
Prop. 1). Since µα 6 µ , it follows that g is also µα -measurable (§1, No. 4,
Cor. 2 of Prop. 11) and gϕKα is µα -integrable. Since µ∗α (T --- Kα ) = 0
(Ch. IV, §2, No. 2, Prop. 5), it follows that ϕ Kα = 1 µα -almost everywhere,
whence gϕKα = g µα -almost everywhere; thus g ∈ LF1 (T, µα ) . Moreover,
if g = g0 locally µ-almost everywhere then g = g 0 µα -almost everywhere,
whence a well-defined mapping
where the (ġ) on the right side is the family, indexed by α , whose α-th
coordinate is ġ ∈ L1F (T, µα ) ; in other words, prα ◦ u = uα .
(ii) Continuity of u . It suffices to show that for each α , the mapping
prα ◦ u = uα is continuous. The topology of L1F (T, µα ) is generated by the
norm function q α ,
Z Z
q α (ġ) = |g| dµα = |g|ϕKα dµα ;
(q α ◦ uα )(ġ) = q α uα (ġ) ġ ∈ L1loc (T, µ; F)
= q α (ġ) ġ ∈ L1F (T, µα )
Z Z
= |g| dµα = |g|ϕKα dµα
Z
6 |g|ϕKα dµ = pKα (g) = qKα (ġ)
( qKα is defined in the Note for V.39, `. −12 to −10), that is, q α ◦ uα 6 qKα ;
since qKα is a continuous semi-norm on L1loc (T, µ; F) , so is q α ◦ uα .
(iii) Since u is linear and continuous, it is uniformly continuous (GT, III,
§3, No. 1, Prop. 3); therefore, if FQis a Cauchy filter on L 1loc (T, µ; F) , then
the filter base u(F) is Cauchy on L1F (T, µα ) (GT, II, §3, No. 1, Prop. 3).
Q 1 α
Now, LF (T, µα ) is locally convex for the product topology (TVS, II, §4,
α
No. 3, Prop. 4), and, for the product uniformity (GT, II, §2, No. 6, Def. 4),
it is complete (GT, II, §3, No. 5, Prop. 10), therefore the Cauchy filter base
u(F) is convergent; say u(F) → (ḟα ) , where fα ∈ LF1 (T, µα ) for all α .
We can suppose that the fα are universally measurable, i.e., measurable
for every measure on T (§3, No. 4, Prop. 7).
(iv) Definition of f . Let (fα ) be the family of universally measurable
functions on T constructed in (iii), with f α µα -integrable for all α . Define
f : T → F by
fα (t) for t ∈ Kα
f (t) = [
0 for t ∈ T --- Kα .
α
is µ-integrable, the pair (I, g 0 ) is µ-adapted (§4, No. 1, Def. 1); therefore Λ
is vaguely measurable (and µ-adequate) (loc. R cit., Prop. 1). R
The measure ν is defined by ν(f ) = λt (f ) dµ(t) = g 0 (t)f (t) dµ(t)
0
for all f ∈ K (T) (§3, No. 1); since gf = g 0 f µ-almost everywhere, one is
permitted to write
Z Z
ν(f ) = gf dµ = g(t)f (t) dµ(t)
where µ1 = (Rθ)+ , . . .”
Understood, that when u takes its values in R , g 3 = g4 = 0 and g1 , g2
are not simultaneously equal to +∞ at any point; when u takes its values
in C , the gi are all finite-valued; when θ is a real measure, µ 3 = µ4 = 0 .
One need not take µ1 = (Rθ)+ , etc., but it is essential that the g i be
θ-measurable, 0 6 gi 6 |g| and 0 6 µj 6 µ for all i and j (the argument
of §2, No. 2, Cor. 2 of Prop. 3 is then applicable); thus the local integrability
of u for θ implies that gi is locally µj -integrable for all i and j .
It is clear from the earlier discussion that one can suppose u to be
1
defined and finite at every point of T , i.e., that u ∈ L loc (T, θ; C) .
{Conversely, if gi is µj -integrable for all i and j , then each g i is
integrable for θ (§2, No. 2, Cor. 2 of Prop. 3), therefore so is u ; it follows that
if gi is locally µj -integrable for all i and j , then u is locally θ-integrable.}
INT V.x119 integration of measures §5
V.40, `. −9 to −6.
“ . . . the mapping Z
f 7→ f (t)u(t) dθ(t)
for all f ∈ K (T) . It follows from the assumptions on the µ j that every
θ-integrable function h is µj -integrable for all j , and that
Z Z Z Z Z
(ii) h dθ = h dµ1 − h dµ2 + i h dµ3 − i h dµ4
(see the Note for V.10, `. 13, 14). In particular, for every f ∈ K (T) one
has
Z Z Z Z Z
(iii) f u dθ = f u dµ1 − f u dµ2 + i f u dµ3 − i f u dµ4 .
f u = f g1 − f g2 + if g3 − if g4 ;
substituting this expression for f u on the right side Rof (iii), it follows from
the linearity of integration that the linear form f 7→ f u dθ on K (T) is a
linear combination of the (sixteen) linear forms
Z
(iv) f 7→ f gi dµj f ∈ K (T) ,
1
whence the assertion. {Alternatively, one can suppose that u i ∈ Lloc etc.}
V.41, `. 7–9.
“ . . . if θ1 and θ2 are two measures on T , and if u is a function locally
integrable for θ1 and θ2 , then u is locally integrable for θ1 + θ2 and one
has u · (θ1 + θ2 ) = u · θ1 + u · θ2 .”
Let A = {t ∈ T : u(t) is defined and finite} , and let N = { A .
Since N is locally negligible for θ1 and θ2 , it is locally negligible for θ1 + θ2
( |θ1 + θ2 |• 6 |θ1 |• + |θ2 |• ); we can therefore suppose that u is defined and
finite everywhere on T .
By assumption, for every f ∈ K (T) one has f u ∈ L 1 (θ1 ) ∩ L 1 (θ2 ) ,
therefore f u ∈ L 1 (θ1 + θ2 ) and
Z Z Z
f u d(θ1 + θ2 ) = f u dθ1 + f u dθ2
(the left members are ‘essential integrals’—merely a notation for the right
members, which are ‘honest’ integrals).
INT V.x121 integration of measures §5
V.41, `. −3 to V.42, `. 2.
“ . . . then cn ∈ K1 , g = lim cn f , therefore |hθ, gi| = lim |hθ, cn f i| ,
n→∞ n→∞
and finally
sup |hθ, gi| 6 sup |hθ, cf i| .
|g|6|f |, g∈K (T;C) c∈K1
One concludes by observing that the first member of this inequality is equal
to h|θ|, |f |i (Ch. III, §1, No. 6, formula (12)).”
In slow motion,
gn (t) for t ∈ U
cn (t) = f (t)
0 for t ∈ { U .
From Supp cn ⊂ Supp gn one knows that cn has compact support; at issue
is the continuity of cn . From Supp gn ⊂ U one knows that the open sets U
and { (Supp gn ) have union T , so it suffices to show that the restrictions
of cn to U and { (Supp gn ) are continuous. Indeed, from the definition
of cn , it is clear that cn U is continuous; whereas
if t ∈ { (Supp gn ) then
cn (t) = 0 whether t ∈ U or t ∈ { U , thus cn { (Supp gn ) = 0 .
Note that cn f = gn : the equality holds at the points of U by the
definition of cn , and both members are equal to 0 on the subset { U
of { (Supp gn ) . It follows that kg − cn f k = kg − gn k → 0 , thus cn f → g
uniformly. Let K = U = Supp f ; then Supp(c n f ) ⊂ Supp f = K and
Supp g ⊂ Supp f = K (because |g| 6 |f | ), thus c n f and g belong to
K (T, K; C) , consequently cn f → g in K (T, K; C) . Since the norm topol-
ogy on K (T, K; C) is equal to the topology induced by the (inductive limit)
topology of K (T; C) , it follows that c n f → g in K (T; C) ; by the conti-
nuity of θ , θ(cn f ) → θ(g) , whence
lim |hθ, cn f i| = |hθ, gi| .
n→∞
Passing to the limit in the inequality |hθ, c n f i| 6 sup |hθ, cf i| , one has
c∈K1
in other words
h|θ|, |f |i 6 sup |hθ, cf i|
c∈K1
by the cited formula (12). This implies that in the inequalities of `. −9,
equality holds throughout. Thus the equalities (4) are verified for the case
that f ∈ K (T; C) .
INT V.x123 integration of measures §5
V.42, `. 6.
“ . . . the dense subspace K (T; C) .”
1
Note that the semi-norm N 1 on L C coincides with N1 on LC1 (§1,
No. 2, Prop. 7, 2)).
1
The closure of K (T; C) in L C is a linear subspace that contains
LC1 (Ch. IV, §3, No. 4, Def. 2) and NC∞ (§1, No. 3) hence it contains
1
LC1 + NC∞ = L C .
V.42, `. 8, 9.
|h|θ|, |f |i − h|θ|, |f 0 |i| 6 h|θ|, |f − f 0 |i = N1 (f − f 0 )
|hθ, cf i − hθ, cf 0 i| 6 h|θ|, |c| |f − f 0 |i 6 N1 (f − f 0 )
by the definitions, and similarly for the other five bracket expressions in the
displayed relations; one can therefore suppose that f and f 0 belong to LC1 ,
whence also N1 (f − f 0 ) = N1 (f − f 0 ) .
The first
line of the displayed relations then follows from the inequality
|f | − |f 0 | 6 |f − f 0 | and the computation
h|θ|, |f |i − h|θ|, |f 0 |i = h|θ|, |f | − |f 0 |i
6 |θ|, |f | − |f 0 | 6 h|θ|, |f − f 0 |i = N1 (f − f 0 )
(Ch. IV, §4, No. 2, Prop. 2); similarly for the second line of the display,
taking into account that c is (universally) measurable and |c| 6 1 .
From the first line of the display, it is clear that the first member of (4)
1
depends continuously on f ∈ L C . As to the second member, let us abbre-
viate by defining
1
α(f ) = sup |hθ, cf i| (f ∈ L C ) .
c∈K1
1
For f, f 0 ∈ L C and c ∈ K1 , one has
|α(f ) − α(f 0 )| 6 N1 (f − f 0 ) ,
1
sup(g1 , g2 ) · µ = (g1 · µ + g2 · µ + |g1 − g2 | · µ)
2
1
= (g1 · µ + g2 · µ + |(g1 − g2 ) · µ|)
2
1
= (g1 · µ + g2 · µ + |g1 · µ − g2 · µ|)
2
= sup(g1 · µ, g2 · µ)
(the second equality, by Prop. 2; the fourth, by the cited formula), and
similarly
inf(g1 , g2 ) · µ = inf(g1 · µ, g2 · µ) .
In particular, setting g1 = g and g2 = 0 , one has
V.42, `. −3 to −1.
“In the statements of this subsection, g denotes a positive numerical
function, defined everywhere and locally µ-integrable, θ denotes a complex
measure, and u a locally θ-integrable complex function.”
INT V.x125 integration of measures §5
R
that is, g(t)εt dµ(t) = ν .
V.43, `. 6.
“This follows from Th. 1 of §4, No. 2.”
By assumption, g is defined and > 0 everywhere in T , but may have
values equal to +∞ . With notations as in the preceding Note, (I, g 0 ) is
µ-adapted;
R in particular, g 0 (t) is finite and > 0 for every t ∈ T , and one
0
has ν = g (t)εt dµ(t) .
§5 measures defined by numerical densities INT V.x126
When f is only defined and > 0 locally µ-almost everywhere, the same
is true of f g and f g 0 , and f gR= f g 0 locally
R • µ-almost everywhere. Imitating
• 0
p. V.9, `. 1–5, if one defines f dν
R• = f dν
R•, where f 0 isR the extension
• 0
by 0 of f to T , and similarly f g dµ = (f g)0 dµ = f g dµ (the
convention 0 · (+∞) = 0 is at work in the last equality), then
Z • Z •
0
f (t) dν(t) = f 0 (t)g(t) dµ(t)
R• R•
by (∗); in other words f (t) dν(t) = f (t)g(t) dµ(t) , that is, f satis-
fies (∗) formally.
V.43, `. 13.
“ . . . the statement then follows at once from Prop. 3.”
In Prop. 3, put µ = |θ| , g = |u| , f = |f | ; then ν = g·µ = |u|·|θ| = |u·θ|
(No. 2, Prop. 2) and Prop. 3 yields
Z • Z • Z •
|u · θ|• (|f |) = |f | d(|u · θ|) = |f | |u| d|θ| = |uf | d|θ| = |θ|• (|uf |) ;
(the 1st and 4th equivalence, by Ch. IV, §5, No. 1, Def. 1), the 2nd by No. 2,
Prop. 2, and the 3rd by the ‘positive case’ already considered). This settles
the case that u is defined everywhere on T .
Suppose more generally that the locally θ-integrable function u (with
values in C or in R ) is defined only locally θ-almost everywhere, say
with domain A ⊂ T , where T --- A is locally θ-negligible. Choose u 0
1
in Lloc (T, θ; C) such that u = u0 locally θ-almost everywhere, and let N
be a locally θ-negligible set such that u(t) is defined and equal to u 0 (t) on
T --- N , that is, T --- N ⊂ A and u = u0 on T --- N . By Cor. 2, u·θ = u0 ·θ .
Let
S = {t ∈ A : u(t) 6= 0 } , S0 = {t ∈ T : u0 (t) 6= 0 } .
By the foregoing, we know that f S0 is θ-measurable
if and only if f is
measurable for u0 · θ = u · θ ; to show that f S is θ-measurable if and only
if f is measurable for u · θ , we need only show that
(∗) f S is θ-measurable ⇔ f S0 is θ-measurable.
One knows that the set S0 is θ-measurable (Ch. IV, §5, No. 5, Prop. 7),
therefore so is S0 ∩ (T --- N) . Note that
S ∩ (T --- N) = S0 ∩ (T --- N) ;
§5 measures defined by numerical densities INT V.x128
(the 1st and 3rd equivalences, by Ch. IV, §5, No. 10, Prop. 15, more precisely,
criterion d 0 ) in the Note for IV.79, `. 7; the 2nd, by Ch. IV, §5, No. 2,
Prop. 6), that is, (∗) holds.
V.43, `. −3.
“For, uf is the extension by 0 of (uf ) S to T .”
By assumption, u : T → C and f : T → G , where G = F (a Banach
space) or G = R . What does uf mean? The interpretation uf : t 7→
u(t)f (t) (t ∈ T) has consequences:
When G = F is a real Banach space, then u : T → R .
When G = F is a Banach space over C , then u : T → C .
When G = R , let us assume that u : T → R ; for, if f (t) = ±∞ ,
permitting u(t) to be a non-real complex number opens the door to too
many ‘infinities’.
In any case, if u(t) = 0 then u(t)f (t) = 0 (even if G = R and f (t) =
±∞ ), consequently
(
u(t)f (t) for t ∈ S
(uf )(t) =
0 for t ∈ T --- S .
INT V.x129 integration of measures §5
This means that uf is the extension by 0 of (uf ) S to T . Since S is
θ-measurable, it follows that
(i) uf is θ-measurable ⇔ uf S is θ-measurable
(Ch. IV, §5, No. 10, Prop. 15, more precisely, criterion d 0 ) in the Note for
1
IV.79, `. 7). But uS and are θ-measurable functions on S (Ch. IV,
uS
§5, No. 3, Th. 1, as generalized in the Note for IV.80, `. −17 to −14), so
1
from f S = · uf S it follows that
uS
(ii) uf S is θ-measurable ⇔ f S is θ-measurable
(Ch. IV, §5, No. 3, Cor. 5 of Th. 1, as generalized in the Note just cited).
On the other hand, by Prop. 4,
(iii) f S is θ-measurable ⇔ f is (u · θ)-measurable .
From (i)–(iii) we infer that uf is θ-measurable if and only if f is (u · θ)-
measurable, as asserted.
Suppose, more generally, that the locally θ-integrable function u is
defined only locally θ-almost everywhere in T , and let u 0 be the extension
by 0 of u to T . By the case already considered, f is (u 0 · θ)-measurable
if and only if u0 f is θ-measurable; but u · θ = u0 · θ , and uf = u0 f locally
θ-almost everywhere, whence again f is (u · θ)-measurable if and only if
uf is θ-measurable.
V.44, `. 6, 7.
“The case that u and θ are positive follows at once from Th. 2 of §4,
No. 4.”
Since u is locally θ-integrable, there exists a function u 0 ∈ Lloc
1
(T, θ; C)
such that u = u locally θ-almost everywhere. Then u · θ = u 0 · θ and
0
by No. 2, Def. 2.
Since f is essentially integrable for η , and 0 6 η ij 6 |η| , it follows
that f is essentially integrable for every η ij ; for, η-negligible sets are ηij -
negligible and so measurability of f for η implies its measurability for
•
the ηij , and, since ηij 6 |η|• , the essential integrability of f for the η ij
follows from ηij (|f |) 6 |η|• (|f |) < +∞ (§1, No. 3, Prop. 9 and Def. 3).
•
V.44, `. 15.
“The formula (6) follows immediately from this.”
Let us consider the case of a Banach space F (of which the case for R
is a simplification).
It will be convenient to write the decompositions (2) in the form
4
X
(2a) u= ai gi
i=1
4
X
(2b) θ= aj µj
j=1
(by (2b) and the Theorem in the Note for V.10, `. 13, 14). Thus each g i is
locally integrable for θ , and
4
X
gi · θ = aj (gi · µj ) .
j=1
Similarly, since the ai gi are locally integrable for θ , it follows from (2a) and
the above-cited Theorem that
4
X 4
X
u·θ = ai (gi · θ) = ai aj (gi · µj ) ,
i=1 i,j=1
P
4
that is, η = ai aj ηij . Since f is essentially integrable for η and the η ij ,
i,j=1
it follows from Corollary 3 of the above-cited Theorem that
Z X4 Z
(∗) f dη = ai aj f dηij .
i,j=1
But, by the ‘positive case’ already treated, one has the 16 equalities
Z Z
(∗∗) f dηij = gi f dµj (1 6 i, j 6 4).
Finally,
Z Z X X Z
(uf )dθ = ai gi f dθ = ai gi f dθ
i i
X Z X X Z
= ai gi f dθ = ai aj gi f dµj
i i j
X Z X Z
= ai aj gi f dµj = ai aj f dηij
i,j i,j
Z
= f dη
(the 2nd equality, by the additivity of the essential integral; the 4th, by the
above-cited Corollary 3; the 6th, by (∗∗); and the last, by (∗).
§5 measures defined by numerical densities INT V.x132
The first
R two members are being defined to be equal to the third, which
is equal to f dν .
V.45, `. 10, 11.
R∗ R•
“ . . . one defines similarly A f dµ and A f dµ .”
Let f be any function > 0 on T such that f A = f A , and define
Z •
f dµ = µ• (f ϕA ) ;
A
if also f 0 is a function > 0 on T such that f 0 A = f A , we know that
f 0 = f locally almost everywhere R • for ν = Rϕ A · µ (see the Note
R • 0for V.44,
0 •
R`.• −3 to V.45, `. 1), therefore f dν = f dν , whence f ϕA dµ =
R• fϕ A dµ by the preceding argument in the Example. Thus the symbol
A
f dµ is well-defined, i.e., is independent of the choice of f .
R∗
The situation for A f dµ is more delicate. One would like to define it
to be µ∗ (f ϕA ) . Assuming that f and f 0 are any two functions > 0 on T
that coincide with f on A , one knows that fϕ A = f 0 ϕA locally µ-almost
R∗
everywhere; for A f dµ to be well-defined, one must arrange that fϕ A =
f 0 ϕA µ-almost everywhere (Ch. IV, §2, No. 3, Prop. 6). This will be the
case, for example, if A is µ-moderated (§1, No. 3, Lemma). Caution: One
knows that µ• (fϕA ) = ν • (f ) by Prop. 3, but the corresponding equality
for outer measure is not justified. {Problem: If A is µ-moderated, is the
measure ϕA · µ moderated? (answered in the Note for V.48, `. 2, 3)}
Remarks. (i) If A is a µ-measurable set such that µ • (A) < +∞ —
in other words if the function ϕA is essentially µ-integrable (§1, No. 3,
INT V.x135 integration of measures §5
is well-defined ( f a function
> 0 defined on B ⊃ A , and f any function
> 0 on T such that f A = f A ). But fϕA is also µ-moderated, therefore
(§1, No. 2, Prop. 7)
(∗∗∗) ν*(f ) = µ• (f ϕA )
by (i). From (∗∗) and (∗∗∗) we have µ*(f ϕ A ) = ν*(f ) and (∗) may then
be written Z ∗ Z ∗ Z ∗
f dµ = f ϕA dµ = f dν .
A
V.45, `. 17.
“ . . . admitting in M (T) a supremum λ .”
Review: The criterion for this is that, for every f ∈ K + (T) , the (in-
creasing directed) family of real numbers λ α (f ) > 0 admit a finite upper
bound (Ch. II, §2, No. 2, Lemma). That the correspondence f 7→ sup λα (f )
α∈A
defines an additive function on K+ (T) follows from the theorem on mono-
tone limits (GT, Ch. IV, §5, No. 2, Th. 2) and the continuity of addition
in R ; that it is extendible to a linear form λ on K (T) then results from
the fact that every f ∈ K (T) is the difference of two functions in K + (T)
(Ch. II, §2, No. 1, Prop. 2), and, being a positive linear form on K (T) ,
λ is a measure on T (Ch. III, §1, No. 5, Th. 1).
This construction—as a result of which, the space M (T; R) of real
measures on T is a fully lattice-ordered Riesz space (loc. cit., Th. 3)—is
fundamental for the concept of essential integral (cf. §1, No. 4, Prop. 11).
V.45, `. −10.
“It is clear that the condition is necessary.”
Suppose g is locally integrable for λ . Given any h ∈ K + (T) , one
knows that gh is λ-integrable; since λ is the sum of the positive measures
λα and λ − λα , it follows that gh is integrable for every λ α and λα (gh) 6
λ(gh) < +∞ (see, e.g., the Note for V.10, `. 13, 14, the remark preceding
the Theorem), consequently g is locally integrable for λ α and g · λα 6 g · λ ,
whence the assertion.
Thus, by the preceding Note, the measure sup g · λα exists, and it is
α∈A
equal to g · λ by the following computation: for every h ∈ K + (T) ,
sup g · λα (h) = sup (g · λα )(h) = sup λα (gh) = λ(gh) = (g · λ)(h) ;
α∈A α∈A α∈A
the first equality, by Ch. II, §2, No. 2, Lemma; the third, because gh is
(universally) moderated and
(loc. cit., Remark 1). It then follows from P Prop. 5 that g is locally
µ-integrable and that g · µ = sup g · µJ = ρ = g · µα .
J∈F α∈A
V.46, `. 8, 9.
“ . . . if the family (Sα ) is locally countable (Ch. IV, §5, No. 9)”
The concept of local countability (loc. cit., Def. 7) has been defined
only for a set of subsets (not a family of subsets) of a topological space T .
§5 measures defined by numerical densities INT V.x138
Guided by the definition of a ‘locally finite family’ (GT, I, §1, No. 5, Def. 8),
the appropriate definition is as follows: call a family (S α )α∈A of subsets
of T locally countable if, for every t ∈ T , there exists a neighborhood V
of t such that the set of indices {α ∈ A : V ∩ S α 6= ∅ } is countable.
It is obvious that if the family (Sα )α∈A is locally countable in the
foregoing sense, then the set A = {Sα : α ∈ A } is locally countable in the
sense of the cited Def. 7 (Ch. IV, § 5, No. 9). However, the converse is false.
{For example, if the index set A is uncountable and S α = T for all α ∈ A ,
then the set A = {Sα : α ∈ A } = {T} is trivially locally countable, but
the family (Sα )α∈A is not.}
V.46, `. 9, 10.
“ . . . this amounts to saying
that, for every compact set K in T , the
set of α ∈ A such that gα K is not zero is countable.”
Consider the conditions:
a) The family of functions (gα )α∈A is locally countable, that is, the
family (Sα )α∈A is locally countable;
b) for every compact set K in T , the set of α ∈ A such that K∩S α 6= ∅
is countable;
c) for every compact set K in T , the set of α ∈ A such that g Kα 6= 0
is countable.
Since K ∩ Sα 6= ∅ ⇔ gα K 6= 0 , the equivalence b) ⇔ c) is trivial.
a) ⇒ b): The finite covering argument for locally countable sets A ,
following Def. 7 of Ch. IV, §5, No. 9, is readily adapted to the case of
families.
b) ⇒ a): When T is locally compact, every point t ∈ T has a compact
neighborhood K .
Thus the assertion “amounts to saying” entails the local compactness
of T .
V.46, `. 16, 17.
“It is clear that g is µ-measurable (Ch. IV, §5, No. 2, Prop. 4 and No. 4,
Cor. 1 of Th. 2).”
Lemma. If f is a function on the locally compact space T such that
f ϕK is µ-measurable for every compact set K ⊂ T , then f is µ-measurable.
Proof. For every t ∈ T let Vt be a compact (hence µ-integrable) neigh-
borhood of T . Since f ϕVt is by assumption µ-measurable, and
f = f ϕVt on Vt , f is µ-measurable by the Principle of localization (Ch. IV,
§5, No. 2, Prop. 4). ♦
Consider now the given function g . Given any compact set K in T , it
will suffice to show that gϕK is µ-measurable. Let
AK = {α ∈ A : gK K 6= 0 }
INT V.x139 integration of measures §5
for, if t ∈
/ K then both sides are equal to 0 at t , whereas if t ∈ K then
X X X
g(t)ϕK (t) = g(t) = gα (t) = gα (t) = gα (t)ϕK (t) .
α∈A α∈AK α∈AK
in the Note for V.39, `. 22–24); their supremum g (which may take on
infinite values) is then altered at most on a locally µ-negligible set—and the
measures gn · µ and g · µ , not at all (No. 3, Cor. 2 of Prop. 3). The relations
gn · µ 6 gn+1 · µ then imply that
0 6 gn+1 · µ − gn · µ = (gn+1 − gn ) · µ
Since the sets Jn are cofinal in the set F ordered by ⊂ , one has
X
gk0 = sup sJ = sup sJn = sup (gn+1 − g0 )
J∈F n∈N n∈N
k∈N
= −g0 + g ,
P
thus g = g0 + gk0 .
k∈N
Necessity. Assuming the sequence (g n · µ) has an upper bound in
M (T) , we are to show that g is locally µ-integrable. Since M (T) is fully
lattice-ordered (Ch. III, §1, No. 5, Th. 3), the measure ρ = sup (gn · µ)
n∈N
exists, as does
sup (gn · µ − g0 · µ) = ρ − g0 · µ .
n∈N
§5 measures defined by numerical densities INT V.x142
Then
this means that the family (gk0 · µ)k∈N is summable, with P sum equal to
0
ρ − g0 · µ . It then follows from Prop. 6 that the function gk = g − g0 is
k∈N
locally µ-integrable, whence so is g , and
X
(g − g0 ) · µ = gk0 · µ = ρ − g0 · µ ;
k∈N
defined in Ch. III, §1, No. 4. On the other hand, for h ∈ K (T) , the function
t 7→ (g · λt )(h) = λt (gh)
h, (g·λt ) dµ(t) = (g·λt )(h) dµ(t) = λt (gh) dµ(t) = ν(gh) = (g·ν)(h) ,
whence (9).
The formula is more memorable when written as a ‘distributive law’
Z Z
0
(9 ) g · λt dµ(t) = (g · λt ) dµ(t) ,
but (9) has the merit that it forces one to check the stringent hypotheses
on X and the mapping t 7→ λt ∈ M+ (X) .
V.47, `. 7, 8.
“ . . . to say that g is locally η-integrable is equivalent to saying that
gϕKn is η-integrable for every n .”
If g is locally η-integrable, then gϕ K is η-integrable for every compact
set K in T by condition b) of No. 1, Prop. 1.
Conversely, suppose gϕKn is η-integrable for every n . If K is any com-
pact set in T then K ⊂ Kn for some n (GT, I, §9, No. 9, Cor. 1 of Prop. 15),
and the η-integrability of gϕKn implies that of gϕK = (gϕKn )ϕK ; in par-
ticular, gϕK is η-measurable for every K , hence g is η-measurable (Lemma
in the Note for V.46, `. 16, 17), therefore g satisfies the cited condition b),
hence is locally η-integrable.
V.47, `. 10. S
“ . . . let H = Hn ”
n
By the preceding Note, g is not λt -integrable if and only if gϕKn is
not λt -integrable for some n , therefore
[
{t ∈ T : g is not λt -integrable } = Hn = H .
n
t 7→ λt (gϕKn ) (t ∈
/ Hn )
S
The local µ-negligibility of H = Hn then follows from Ch. IV, §5, No. 2,
n
Def. 3, and the first assertion of the statement follows from the fact that
{t ∈ T : g is not λt -integrable } = H
(preceding Note).
V.47, `. 14–16.
“ . . . we have, by Prop. 3 and by Prop. 5 of §3, No. 2,
Z • Z • Z • Z • Z • Z •
h d(g·ν) = (gh) dν = dµ(t) (gh) dλt = dµ(t) h d(g·λt ) . ”
R
Here, ν = λt dµ(t) . As the cited Prop. 3 employs the notation ν
in another sense, it is helpful to restate it: If η is a positive measure on
a locally compact space X and if g is a locally η-integrable function > 0
on X , then Z Z
• •
f d(g · η) = (f g) dη
t 7→ λt ∈ M+ (X) (t ∈ T)
R
is µ-adequate, with integral ν = λt dµ(t) , and gh is ν-measurable (be-
cause g and h are; see the Lemma at the end of this Note) and ν-moderated
(because g is ν-integrable; in fact, since X is countable at infinity, all func-
tions on X are moderated for every measure on X ).
INT V.x145 integration of measures §5
R
thus (g · λt ) dµ(t) = g · ν , which is the formula (9).
The following minor point, touched on in the Note for V.46, `. 17, 18,
seems not to have been established explicitly (if it has, I have forgotten
where):
Lemma. Let µ be a measure on a locally compact space X , and let
f, g : X → R be functions that are (with respect to µ ) measurable and are
> 0 locally almost everywhere on X . Then f g is measurable.
Proof. We can suppose that µ > 0 (Ch. IV, §2, No. 1, Def. 1 and §5,
No. 1, Def. 1). Redefining f (x) and g(x) to be 0 on the locally negligible
set where either number is < 0 , we can suppose that f and g are > 0
everywhere on X (Ch. IV, §5, No. 2, Prop. 6).
A function h > 0 on X is measurable if and only if the set
S = {x ∈ X : h(x) > a }
is measurable for every real number a > 0 (loc. cit., No. 5, comment pre-
ceding the Cor. of Prop. 8, plus the fact that S = X when a < 0 ); and since
(for the case a = 0 )
∞
[
{x : h(x) > 0 } = {x : h(x) > 1/n } ,
n=1
§5 measures defined by numerical densities INT V.x146
A = {x : (f g)(x) > a }
is measurable for every real number a > 0 . Now, (f g)(x) > a if and only
if one of the following three conditions holds:
f (x) = +∞ and g(x) > 0 ;
f (x) > 0 and g(x) = +∞ ;
0 < f (x) < +∞ , 0 < g(x) < +∞ and f (x)g(x) > a .
Thus, writing
one has
A = B ∪ C ∪ (D ∩ E ∩ {x : f (x)g(x) > a}) ;
the sets B, C, D, E are known to be measurable, so it suffices to show that
the set
D ∩ E ∩ {x : f (x)g(x) > a}
is measurable. This set can be written
is measurable for every real number a > 0 . Now, the finite-valued functions
ϕD f and ϕE g are measurable; for example, the set
is measurable for every real number a > 0 . One is thus reduced to proving
that the product of measurable functions with values in R is measurable,
and this is known from the comments following Ch. IV, §5, No. 3, Cor. 5 of
Th. 1. ♦
INT V.x147 integration of measures §5
In slow motion:
Z • Z • Z •
|g2 | f d|θ1 | = |g2 | f d|g1 · θ| = |g2 | f d(|g1 | · |θ|)
Z • Z •
= |g2 | f |g1 | d|θ| = |g2 g1 | f d|θ| .
§5 measures defined by numerical densities INT V.x148
that is, ϕA · µn = µn . Since µ• = sup µ•n (§1, No. 4, Prop. 11) it follows
n
that, for all g ∈ K+ (T) ,
(ϕA · µ)(g) = µ• (ϕA g) = sup µ•n (ϕA g) = sup µ•n (g) = µ• (g) = µ(g) ,
n n
thus ϕA · µ = µ . ♦
The example. Let T and µ be as in Ch. V, §2, Exer. 4 (on p. V.95).
By part a) of the exercise, µ is not moderated but µ = sup µn , where
n
(µn ) is anSincreasing sequence of positive measures with finite support. The
set A = Supp µn is the union of a sequence of compact sets, hence is
n
universally moderated, but ϕA · µ = µ by the Lemma, thus ϕA · µ is not
moderated.
Coda. This is a good place to stop—the Notes are getting too fre-
quent and too long, the document a bloated 586 pages, and the complicated
arrangement of the proof of the Lebesgue-Nikodym theorem in No. 5 exacer-
bates the trend; the reader who has gotten this far will probably find filling
in the gaps easier than studying how I filled them in.
As epitome of the formal beauty of Bourbaki’s treatment of integration,
I nominate the formula
p
|µ + iν| = µ2 + ν 2