ar X iv :2 30 8. 10 91 4v 1 [ m at h. FA ] 2 0 A ug 2 02 3 ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS S. CAMBRONERO 1, D. CAMPOS 2, C. A. FONSECA-MORA 3, AND D. MENA 4 Abstract. The concept of supremum of a family of set functions was introduced by M. Veraar and I. Yaroslavtsev (2016) for families of measures defined on a measurable space. We expand this concept to include families of set functions in a very general setting. The case of families of signed measures is widely discussed and exploited. Contents 1. Introduction 1 2. The set function supremum 2 3. Sub-additive set functions on rings 6 4. Supremum of signed measures 8 4.1. Signed measures with density 11 4.2. Application: Vector measures on a space of bilinear forms 13 Declarations 14 References 14 1. Introduction Given a family of sets A and a family {fj}j∈J of real extended set functions defined on A, the classical supremum defined as f(A) := sup{fj(A) : A ∈ A} gives the least upper bound of the family under the natural order (see Definition 2.8). In practical situations, one might need to work with a smaller class of set functions, as for instance the class of measures over a sigma-algebra, or the class of finitely additive set functions over an algebra. The classical definition of supremum given above might produce an object which is not in the class. 2020 Mathematics Subject Classification. 28A10, 28B05. Key words and phrases. set function supremum; admissibility; signed measures. 1 http://arxiv.org/abs/2308.10914v1 https://orcid.org/0000-0001-6758-4942 https://orcid.org/0000-0002-3608-1151 https://orcid.org/0000-0002-9280-8212 https://orcid.org/0000-0002-9443-391X ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 2 In [9], they show that, in the class of (non-negative) measures on a measurable space (Ω,F), a suitable definition for the supremum of a family of measures (νj)j∈J is µ(A) = sup Π∈P(A) ∑ C∈Π sup j∈J νj(C), where P(A) is the family of all finite partitions of A by elements of F . In this paper, we show that this definition works perfectly well in larger classes of set functions and explore some nice consequences of this concept. Although the literature provides plenty of concepts and results related to that of the supremum of measures, we could only find the first reference in [9]. Some works like [5] and [7] have mentioned the existence of a supremum in the case of a family of measures, but not in a constructive way. In [9], they do provide a constructive definition for this particular kind of families, developing a few tools they needed in their construction of the quadratic variation of a cylindrical local martingale. When trying to apply this concept to our recent work on cylindrical martingale-valued measures [1], we have found it necessary to extend it to families of a more general class of set functions, for instance to families of random measures, signed measures or even random sub-additive set functions. In Section 2 we give the definition of the supremum of a family of set functions in the most general possible setting. We establish three levels of admissibility, that will be exploited according to the kind of properties we need the supremum to inherit from the given family. The weak admissibility is the least we can ask in order to define the set function supremum, but we need admissibility if we want this function to inherit at least finitely sub-additivity from the given family. In Section 3 we work with set functions defined on rings. The main result on this section, Theorem 3.4, shows that the set function supremum of an admissible family of sub-additive set functions is a signed measure, whenever we can control it from below. We also introduce the concept of set function infimum. Results in Section 4 can be considered as the main contributions of the paper. We concentrate on families of signed measures, for which the concepts of admissibility and strong admissibility are equivalent. We use the set function supremum to express the positive and negative part of signed measure, as well as the total variation. The main result, Theorem 4.11, generalizes a result stated on [9] for families of (non-negative) measures. We finish by giving some very pleasant applications of this result. 2. The set function supremum Throughout this section we consider a non-empty set S and a family A of subsets of S. Unless otherwise stated, no particular structure is assumed on A. We denote by C the class of all set functions f : A → R. ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 3 Definition 2.1. Consider a family (fj)j∈J on C. We say that this family is: (i) weakly admissible if, for any pair of disjoint sets C,D ∈ A, sup j∈J fj(D) = ∞ ⇒ sup j∈J fj(C) > −∞. (ii) admissible if for each C ∈ A we have sup j∈J fj(C) > −∞. (iii) strongly admissible if there is a ∈ R such that ∀C ∈ A, sup j∈J fj(C) > a. It is evident that every strongly admissible family is admissible, and every admissible family is weakly admissible. The importance of these properties will become clear as we establish and discuss the definition of set function supremum. Before that, we have the following example. Example 2.2. For A ∈ B(R) let |A| denote its Lebesgue measure. (i) For every n ∈ N let αn : B(R) → R be defined by αn(A) := |A ∩ (0, n]| − |A ∩ (−n, 0]| . The family (αn)n≥1 is strongly admissible. (ii) Let g : B(R) → R be defined by g(A) = |A ∩ (0,∞)| − |A ∩ (−∞, 0)| , ∀A ∈ B(R), A 6= R, and g(R) = 0. The single element family (g) is not weakly admissible since for C = (−∞, 0) and D = (0,∞) we have g(D) = ∞ but g(C) = −∞. The following results state useful sufficient conditions for admissibility and for strong admissibility. Their proofs are simple so we omit them. Lemma 2.3. Any family (fj)j∈J of functions from A into (−∞,∞] is admissible. Lemma 2.4. Consider a family (fj)j∈J on C. Assume that there exists a ∈ R and j0 ∈ J such that fj0(A) ≥ a for every A ∈ A. Then the family (fj)j∈J is strongly admissible. In particular, any family on C for which at least one of its members is non-negative, is strongly admissible. Definition 2.5. Consider a weakly admissible family (fj)j∈J on C. We define its set function supremum as ( ssup j∈J fj ) (A) := sup Π∈P(A) ∑ C∈Π sup j∈J fj(C), ∀A ∈ A, ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 4 where P(A) is the family of all finite partitions of A by elements of A. Remark 2.6. The weak admissibility condition avoids having undetermined sums in the last definition. When the family is admissible, the supremum does not achieve the value −∞. Example 2.7. Consider S = {0, 1} and A = 2S. For n = 1, 2, . . . define αn = n · δ{0} − ∞ · δ{1} (here ∞ · 0 = 0). Each αn is clearly a signed measure on (S, 2S). But notice that supn≥1 αn({0}) = +∞ and supn≥1 αn({1}) = −∞. This family is not weakly admissible so we are not able to apply the previous definition for A = S. It is clear from the definition that ssupj∈J fj ∈ C. In order to explore more properties of the set function supremum we will need to be able to compare elements in C. Definition 2.8. Given f, g ∈ C, we write f ≤ g whenever f(A) ≤ g(A), ∀A ∈ A. If this is the case, we say that g dominates f. It is clear that the relation “≤” introduces a partial order on C. We will show (see Corollary 2.13 below) that the set function supremum of an admissible family is the smallest finitely super-additive set function that dominates each fj. Here, a function g ∈ C is called finitely super-additive if whenever A,B ∈ A are disjoint and A∪B ∈ A, then g(A) + g(B) is a well defined extended real number and g(A) + g(B) ≤ g(A ∪ B). Lemma 2.9. Let (fj)j∈J be weakly admissible and f = ssupj∈J fj. If g ∈ C is finitely super-additive and dominates each fj, then it also dominates f . Proof. Let A ∈ A. For any Π ∈ P(A) we have ∑ C∈Π sup j∈J fj(C) ≤ ∑ C∈Π g(C) ≤ g(A), the last inequality is a consequence of the finite super-additivity of g. Since f(A) is the supremum of the left-hand side, this implies f(A) ≤ g(A). � In order to talk about super-additivity of the set function supremum, weakly admissibility of the family is not enough. Example 2.10. Consider S = {1, 2, 3, ...} and A = 2S. Define f on A by f(∅) = 0, f(A) = 1 if 1 /∈ A 6= ∅, f(A) = −∞ if 1 ∈ A. Let A = {1} and B = {2, 3, . . .}. The unitary family {f} is weakly admissible but not admissible. The supremum f̌ = ssup{f}, though well defined, is not supper-additive. In fact, notice that f̌(A) = −∞ and f̌(B) = +∞. Lemma 2.11. If (fj)j∈J is admissible, then f = ssupj∈J fj is finitely super-additive and dominates each fj. ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 5 Proof. Let A,B ∈ A be disjoint and such that A ∪ B ∈ A. Because of the admissibility, f(A), f(B) ∈ (−∞,∞]. Let c ∈ R such that c < f(A) + f(B). Let a < f(A) and b < f(B) such that a + b = c. Choose partitions ΠA ∈ P(A) and ΠB ∈ P(B) such that a < ∑ C∈ΠA sup j∈J fj(C), b < ∑ C∈ΠB sup j∈J fj(C). Then for each C ∈ ΠA ∪ ΠB, there exists jC ∈ J such that a < ∑ C∈ΠA fjC (C), b < ∑ C∈ΠB fjC(C). Since ΠA ∪ ΠB ∈ P(A ∪ B), it follows that c = a+ b < ∑ C∈ΠA∪ΠB fjC (C) ≤ f(A ∪ B). The inequality f(A) + f(B) ≤ f(A ∪ B) follows from the fact that c < f(A) + f(B) was arbitrary. � The results of these lemmas can be applied to quite general situations. For instance, A could be so small that no partition other that Π = {A} exists for each set A. Example 2.12. Consider S = {0, 1, 2, 3} and A = {{0}, {0, 1}, S}. Any set function f defined on A is additive. In fact, there are no disjoint sets on A to test. In this case, ( ssupj∈J fj ) (A) = supj∈J fj(A). We get a similar situation with A = {{0}, {1, 2}, S}. In this case we have two disjoint sets, but their union is not in A. By combining the results of Lemmas 2.9 and 2.11 we conclude the following: Corollary 2.13. For an admissible family (fj)j∈J , the supremum ssupj∈J fj is the smallest finitely super-additive set function that dominates each fj. Remark 2.14. Notice the special case of a unitary family {f}, with f > −∞. In this case, ssup{f} is the smallest finitely super-additive set function defined on A that dominates f . In particular, f = ssup{f} if and only if f is finitely super-additive. Remark 2.15. Consider an admissible family of set functions (νj)j∈J , defined on A ⊆ 2S. (i) If for each C ∈ A there exists j ∈ J such that νj(C) ≥ 0, then ssup j∈J νj = ssup j∈J (νj)+ = ssup {νj : j ∈ J} ∪ {0}, where (νj)+(A) = max{νj(A), 0}. (ii) Given k ∈ J define µj(A) := max {νj(A), νk(A)} and λj := ssup {νj , νk}. Then ssup j∈J νj = ssup j∈J µj = ssup j∈J λj. ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 6 (iii) If for each C ∈ A sup j∈J νj(C) ≥ − inf j∈J νj(C), then, ssup j∈J νj = ssup j∈J |νj | . 3. Sub-additive set functions on rings In this section we assume that the family A is a ring. More precisely, A is closed under finite unions and difference of sets, and therefore it is also closed under finite intersections. We will try to prove that ssup fj inherits the sub-additivity from the fj ’s. Since negative values are allowed, in the definition of sub-additivity we must assume that the sets are disjoint. We say that a function f ∈ C is sub-additive if f ( ∞ ⋃ n=1 Bn ) ≤ ∞ ∑ n=1 f(Bn) (3.1) for any countable family (Bn) of pairwise disjoint sets on A such that ∪Bn ∈ A. As part of the definition, the sum on the right-hand side must make sense independently of any reordering. Finite sub-additivity is defined analogously. Lemma 3.1. Consider an admissible family (fj)j∈J of functions defined on a ring A. If each fj is sub-additive, then f := ssup fj is also sub-additive. Proof. Let (Bn) be a sequence of pairwise disjoint sets on A such that B = ∪Bn ∈ A. If Π ∈ P(B), it follows that C ∩ Bn ∈ A for each C ∈ Π. By applying the sub-additivity of fj , and the super-additivity of f one gets ∑ C∈Π sup j∈J fj(C) ≤ ∑ C∈Π sup j∈J ∞ ∑ n=1 fj(C ∩ Bn) ≤ ∞ ∑ n=1 ∑ C∈Π f(C ∩ Bn) ≤ ∞ ∑ n=1 f(Bn). Taking the supremum on the left we get the result. � As the next example illustrates, in Lemma 3.1 the admissibility assumption can not be replaced by weak admissibility. Example 3.2. Consider S = {1, 2, . . .}, A = 2S and define αn(A) = ∑ k∈A pnk, where pnk = (−1)k for k ≤ n and pnk = −1 for k > n. The sum is always well defined and gives αn(A) = −∞ for any infinite set A. Each αn is sub-additive, and better yet, a signed measure. The family (αn) is weakly admissible but not admissible. For µ = ssupαn we have µ({k}) = (−1)k, consequently µ is not sub-additive. ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 7 Lemma 3.3. Suppose the family (fj)j∈J is admissible and each fj is finitely sub-additive on a ring A. Then f := ssupj∈J fj is finitely additive. Proof. The argument of Lemma 3.1 clearly works for finite families of disjoint sets. This gives the finite sub-additivity, and Lemma 2.11 finishes the work. � Recall that, if a set function defined on a measurable space (S,F) is sub-additive, non- negative and finitely additive, then it is a measure. This fact is used in the proof of the following useful result. Theorem 3.4. Consider an admissible family of sub-additive set functions defined on a measurable space (S,F). Let µ be the set function supremum of that family, and suppose there is a finite measure ν such that µ ≥ −ν. Then µ is a signed measure. In particular, if µ ≥ 0, then it is a measure. Proof. By Lemmas 3.1 and 3.3, µ is finitely additive and sub-additive, then so is the non- negative set function µ + ν. We conclude that µ + ν is a measure, so µ = (µ + ν) − ν is a signed measure. � Corollary 3.5. Let (νj)j∈J be a family of sub-additive set functions defined on a measurable space (S,F). If for some j ∈ J we have νj ≥ 0, then ν := ssupj∈J νj is a measure. In particular, the set function supremum of a family of measures on (S,Σ), is itself a measure. Proof. Since the family (nj) has a non-negative element, then it is strongly admissible (Lemma 2.4) and clearly its supremum is non-negative. Hence by Theorem 3.4 its supremum is a measure. The second assertion is an easy consequence of the former. � Remark 3.6. The result in Corollary 3.5 generalizes Lemma 2.6 in [9]. We finish this section by introducing the concept of set function infimum. Consider a family (αj)j∈J of set functions defined on a family of setsA. If the family (−αj) is weakly admissible, we can define iinf j∈J αj := − ssup j∈J (−αj). We immediately get ( iinf j∈J αj ) (A) = inf Π∈P(A) ∑ C∈Π inf j∈J αj(C). If (−αj) is admissible, then iinfj∈J αj is the biggest finitely sub-additive set function that is dominated by each αj . If A is a ring and each αj is finitely super-additive, then iinfj∈J αj is finitely additive; if each αj is super-additive, then iinfj∈J αj is super-additive. ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 8 4. Supremum of signed measures We now consider the supremum µ := ssupj∈J αj of an admissible family of signed measures defined on a σ-algebra A. Example 4.1. Consider the (admissible) family of finite signed measures (αn)n≥1 defined on (R,B(R)) by αn(A) := |A ∩ (0, n]| − |A ∩ (−n, 0]| . Since µ̌ = ssupn≥1 αn is finitely additive we have µ̌(A) = µ̌(A ∩ R−) + µ̌(A ∩ R+) = |A ∩ R+| − |A ∩ (−1, 0]| . This is clearly a signed measure. Example 4.2. Consider now the (strongly admissible) family of finite signed measures (αn)n≥1 given by αn(A) := |A ∩ (n− 1, n]| − |A ∩ (−n,−n + 1]| . In this case µ̌ = ssupn≥1 αn is the measure given by µ̌(A) = |A ∩ R+|. Since signed measures are sub-additive and finitely additive, by Theorem 3.4 µ is a signed measure if (and only if) there exists a finite measure ν such that µ ≥ −ν. In particular, if some member of the family misses the value −∞, then µ is a signed measure on A. The following lemma shows that in the present case, admissibility does the whole work. For a signed measure α, we consider the Jordan decomposition α = α+ − α−. Lemma 4.3. Consider a family (νj)j∈J of signed measures defined on a measurable space (S,F). The following are equivalent: (i) The family (νj) is admissible (ii) The family (νj) is strongly admissible (iii) There exists k ∈ J such that νk > −∞. Proof. If the family is admissible, by definition there exists k ∈ J such that νk(S) > −∞. The additivity implies that νk misses the value −∞. This proves that (i) implies (iii). To see that (iii) implies (ii) it is enough to take a = −ν− k (S). Finally, it is clear that (ii) implies (i). � Last lemma, combined with Theorem 3.4, gives us the following result. Theorem 4.4. If (νj)j∈J is an admissible family of signed measures defined on a measurable space (S,F), then µ := ssup νj is a signed measure with µ− < ∞. ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 9 We now use the concept of set function supremum to provide alternative expressions for α+, α− and var(α) = α+ +α−. We start by noticing that, for a signed measure α, ssup {α, 0} is the measure defined by ssup {α, 0}(A) := sup Π ∑ C∈Π max{α(C), 0} = sup Π ∑ C∈Π α(C)+. Lemma 4.5. For a signed measure α we have α+ = ssup{α, 0}, α− = ssup {−α, 0}. Proof. In fact, let Ω+ and Ω− be respectively the positive and negative sets of a Hahn decomposition for α. By considering the partition Π = {A ∩ Ω+, A ∩ Ω−} we have ssup{α, 0}(A) ≥ α(A ∩ Ω+) = α+(A). On the other hand, since for every A ∈ A we have α+(A) ≥ 0 and also α(A) = α+(A)− α−(A) ≤ α+(A), we must have ssup{α, 0}(A) ≤ a+(A), and thus the desired equality. To get the second identity, we apply what we have just proven to −α. � Remark 4.6. If |α| denotes the absolute value of α, which is a sub-additive set function, then var(α) = α+ + α− coincides with ssup {|α|} (the supremum of a unitary family). This readily follows from the definition of supremum of measures and the general definition of variation for a signed measure. Lemma 4.7. For a signed measure α we have var(α) = ssup{α,−α} Proof. In fact, by definition we have ssup{α,−α}(A) = sup Π ∑ C∈Π max{α(C),−α(C)} = sup Π ∑ C∈Π |α(C)| = var(α). � Because of Theorem 3.4, when (−αj) is an admissible family of sub-additive set functions, then µ := iinfj∈J αj is a signed measure if and only if there exists a finite measure ν such that µ ≤ ν. In particular, if there exists j such that αj < ∞, then µ is a signed measure. More particularly, if (µj)j∈J is a family of measures and (−µj) is admissible, according to Theorem 4.4, iinfj∈J νj is also a measure. Theorem 4.8. Consider a family (νj)j∈J of finite signed measures on the measurable space (Ω,F). The signed measure µ := ssupj∈J νj satisfies µ+ = ssup j∈J ν+ j , µ− ≤ iinf j∈J ν− j . ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 10 If J is countable, we also have µ− = iinf j∈J ν− j . Proof. It is clear that µ ≤ ssup ν+ j . Let Ω+ and Ω− form a Hahn decomposition of µ. We have µ ≥ 0 and µ ≥ νj on Ω+ and consequently µ ≥ ssup{νj, 0} = ν+ j on Ω+. This shows that on Ω+ we have µ+ = µ = ssup ν+ j . On Ω− we have νj ≤ µ ≤ 0 and consequently ν+ j = 0 for each j. We conclude µ = ssup ν+ j on Ω+ and on Ω−, so all over Ω. Now, on Ω− we also have ν− j = ssup{−νj , 0} = −νj and then µ− = −µ = − ssup νj = iinf(−νj) = iinf ν− j while on Ω+ 0 = µ− ≤ iinf j∈J ν− j . It remains to show that, for the countable case, iinf νj = 0 on Ω+. In that case, we can choose a Hahn decomposition {Ω+ j ,Ω − j } for each νj, it is easy to check that Ω+ = ∪Ω+ j and Ω− = ∩Ω− j is a Hahn decomposition for µ. Since iinf ν− j ≤ ν− k = 0 on Ω+ k for each k ∈ J , the result follows. � Using an analogous argument, we can infer the next result. Corollary 4.9. Consider a family (νj)j∈J of finite signed measures on the measurable space (Ω,F). The signed measure µ := iinfj∈J νj satisfies µ− = ssup j∈J ν− j , µ+ ≤ iinf j∈J ν+ j . If J is countable, we also have µ+ = iinf j∈J ν+ j . Theorem 4.10. Consider a family (νj)j∈J of finite signed measures on the measurable space (Ω,F), and let µ = ssupj∈J νj. Then var(µ) ≤ ssup j∈J |νj| = ssup j∈J var(νj). Proof. Since for all A ∈ F we have µ(A) = (ssup j∈J νj)(A) ≤ (ssup j∈J |νj|)(A), −µ(A) = (iinf j∈J (−νj))(A) ≤ (iinf j∈J |νj |)(A) ≤ (ssup j∈J |νj|)(A), then |µ| ≤ ssupj∈J |νj |. It follows that var(µ) = ssup{|µ|} ≤ ssupj∈J |νj |. Now, since for every j ∈ J we have |νj | ≤ var(νj) then ssupj∈J |νj| ≤ ssupj∈J var(νj). On the other hand, ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 11 for every j ∈ J , |νj | ≤ ssupj∈J |νj |, so var(νj) = ssup{|νj |} ≤ ssupj∈J |νj|. By definition of supremum of measures we get ssup j∈J var(νj) ≤ ssup j∈J |νj| and therefore we have the desired equality. � 4.1. Signed measures with density. We extend Lemma 2.8 on [9] to the case of signed measures. Theorem 4.11. Let (S,A, ν) be a σ-finite measure space. Let F be a family of measurable functions from S into (−∞,∞], such that ˆ f−dν < ∞ for each f ∈ F. Let (fj)j∈N be a sequence in F . Define f̌ = supj≥1 fj and assume that supf∈F f = f̌ . For each f ∈ F let µf be the signed measure defined by µf(A) = ˆ B fdν, A ∈ A. If we define µ̌ := ssupf∈F µf , then µ̌ = ssupj≥1 µfj and µ̌(A) = ˆ A f̌dν. (4.1) Proof. Clearly f̌ is measurable, so A 7→ ´ A f̌dν defines a signed measure that dominates each µf . This gives the inequality “≤” in (4.1). For the other inequality, we adapt the proof of Lemma 2.8 in [9], which is established for non-negative functions and measures. Let A ∈ A, n ∈ N and 0 < ε < 1. Define A1 = {s ∈ A : f1(s) > (f̌(s) ∧ n)(1− ε) > 0} and Aj+1 = {s ∈ A : fj+1(s) > (f̌(s) ∧ n)(1− ε) > 0} \ j ⋃ i=1 Ai. The sets Aj , j ≥ 1 are disjoint and cover A ∩ {f̌ > 0}, so µ̌(A ∩ {f̌ > 0}) = ∑ j≥1 µ̌(Aj) ≥ ∑ j≥1 µfj(Aj) = ∑ j≥1 ˆ Aj fjdν. It follows that µ̌(A ∩ {f̌ > 0}) ≥ (1− ε) ∑ j≥1 ˆ Aj (f̌ ∧ n)dν = (1− ε) ˆ A∩{f̌>0} (f̌ ∧ n)dν. ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 12 Since ε and n are arbitrary, we obtain µ̌(A ∩ {f̌ > 0}) ≥ ˆ A∩{f̌>0} f̌dν. If we now use A1 = {s ∈ A : f̌(s) < 0, f1(s) > f̌(s)(1 + ε)} and Aj+1 = {s ∈ A : f̌(s) < 0, fj+1(s) > f̌(s)(1 + ε)} \ j ⋃ i=1 Ai we get µ̌(A ∩ {f̌ < 0}) ≥ ˆ A∩{f̌<0} f̌dν. Now with A1 = {s ∈ A : −ε < f1(s) ≤ 0} and Aj+1 = {s ∈ A : −ε < fj+1(s) ≤ 0} \ j ⋃ i=1 Ai we get µ̌(A ∩ {f̌ = 0}) ≥ −εν(A ∩ {f̌ = 0}). Since ε is arbitrary and (by localization) ν may be assumed finite, we get µ̌(A∩{f̌ = 0}) ≥ 0. Finally µ̌(A) = µ̌(A ∩ {f̌ > 0}) + µ̌(A ∩ {f̌ = 0}) + µ̌(A ∩ {f̌ < 0}) ≥ ˆ A f̌dν. We have proven (4.1). If we replace F with (fj), we obtain ssup j≥1 µfj = ˆ A f̌dν = µ̌ and this completes the proof. � Corollary 4.12. Consider a signed measure µ given by µ(A) := ˆ A f(s) dν(s), where ν is a σ-finite measure. Then µ+(A) = ˆ A f+(s) dν(s), µ−(A) = ˆ A f−(s) dν(s), var(µ)(A) = ˆ A |f(s)| dν(s). Proof. By Theorem 4.11 we have µ+(A) = ssup{µ, 0}(A) = ˆ A sup{f(s), 0} dν(s) = ˆ A f+(s) dν(s). ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 13 For µ− we proceed similarly. Finally var(µ)(A) = ssup{µ,−µ}(A) = ˆ A sup{f(s),−f(s)} dν(s) = ˆ A |f(s)| dν(s) and the proof is complete. � Example 4.13. Consider f : R → R integrable and define α on (R,B(R)) by α(A) = ˆ A f(t)dt. By Corollary 4.12, the total variation of α is given by var(α)(A) = ˆ A |f(t)| dt. When applying this with A = (a, x], we can use Riemann-type partitions in the definition of the supremum (see for instance Remark 2.10 in [9]). We obtain a very direct proof of the identity V [F ; a, x] = ˆ x a |f(t)| dt, for F (x) := ˆ x a f(t)dt. 4.2. Application: Vector measures on a space of bilinear forms. Consider a σ-finite measure space (S,Σ, ν) and a separable Banach space X . We denote by Bil(X) the space of bounded bilinear forms defined on X ×X , and by L(X,X∗) the space of bounded linear operators from X to X∗. In both spaces, we use the σ-algebra of Borel sets. Consider Q : S → L(X,X∗) Bochner-integrable and define a Bil(X)-valued measure α by α(A)(x, y) = ˆ A 〈Q(s)x, y〉 dν(s), A ∈ Σ, (x, y) ∈ X ×X. Theorem 4.14. The vector measure α is of bounded variation and its variation is given by var(α)(A) = ˆ A ‖Q(s)‖L(X,X∗) dν(s), A ∈ Σ. In other words, var(α) << ν and d dν var(α) = ‖Q‖ . Proof. Notice that ˆ A ‖Q(s)‖ dν(s) = ˆ A sup ‖x‖=‖y‖=1 〈Q(s)x, y〉 dν(s), where the supremum can be computed over (xn, xm), for a dense sequence (xn) on the unit sphere SX . We do not use absolute value here because, when 〈Q(s)x, y〉 is negative, we can replace x with −x. We apply Theorem 4.11 with f(x,y) = 〈Q(·)x, y〉. We obtain ˆ A ‖Q‖ dν = ( ssup ‖x‖=‖y‖=1 α(·)(x, y) ) (A) = sup Π∈P(A) ∑ C∈Π sup ‖x‖=‖y‖=1 α(C)(x, y). ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 14 We conclude that ˆ A ‖Q‖ dν = sup Π∈P(A) ∑ C∈Π ‖α(C)‖ Bil = var(α)(A) and, since the left-hand side is finite, α is of bounded variation. � Acknowledgements This work was partially supported by The University of Costa Rica through the grant “C3019- Cálculo Estocástico en Dimensión Infinita”. Data Availability. Data sharing not applicable to this article as no datasets were generated or analysed during the current study. Declarations Conflict of interest The authors have no conflicts of interest to declare that are relevant to the content of this article. References [1] Cambronero, S.; Campos, D.; Fonseca-Mora, C.A.; Mena, D: Quadratic Variation for Cylindrical Martingale-Valued Measures, Preprint. [2] Diestel, J.; Uhl, J. J., Jr: Vector measures. Mathematical Surveys, No. 15. American Mathematical Society, Providence, R.I. (1977). [3] Dinculeanu, N.: Vector integration and stochastic integration in Banach spaces. Pure and Applied Mathematics. Wiley-Interscience, New York, (2000). [4] Folland, G.Real analysis. Modern techniques and their applications. Second edition. Pure and Applied Mathematics (New York). A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, (1999). [5] Kallenberg, O. Foundations of modern probability. Third edition. Probability Theory and Stochastic Modelling, 99. Springer, Cham, (2021). [6] Kluvánek, I.: The extension and closure of vector measure. Vector and operator valued measures and applications (Proc. Sympos., Alta, Utah, 1972), pp. 175–190. Academic Press, New York (1973). [7] Métivier, M; Pellaumail, J.: Stochastic integration, Probability and Mathematical Statistics, Academic Press, New York (1980). [8] Nygaard, O; Põldvere, M.: Families of vector measures of uniformly bounded variation, Arch. Math. 88 (2007). [9] Veraar, M. ; Yaroslavtsev, I.: Cylindrical continuous martingales and stochastic integration in infinite dimensions. Electron. J. Probab. 21, Paper No. 59, 53 pp. (2016) [10] Walsh, John B.: An introduction to stochastic partial differential equations. École d’été de probabilités de Saint-Flour, XIV—1984, 265–439, Lecture Notes in Math., 1180, Springer, Berlin (1986). ON THE SUPREMUM OF A FAMILY OF SET FUNCTIONS 15 Centro de Investigación en Matemática Pura y Aplicada, Escuela de Matemática, Universidad de Costa Rica Email address : 1 santiago.cambronero@ucr.ac.cr Email address : 2 josedavid.campos@ucr.ac.cr Email address : 3 christianandres.fonseca@ucr.ac.cr Email address : 4 dario.menaarias@ucr.ac.cr 1. Introduction 2. The set function supremum 3. Sub-additive set functions on rings 4. Supremum of signed measures 4.1. Signed measures with density 4.2. Application: Vector measures on a space of bilinear forms Declarations References