1 Introduction

Let X be an abelian metric group with an invariant metric and Y be a real topological vector space. Assume that K is a subsemigroup of Y (i.e. \(K+K\subset K\)). Denote by n(Y) the family of all nonempty subsets of Y, and by \({\mathcal {B}}(Y)\) and \(\mathcal{C}\mathcal{C}(Y)\) its subfamilies of all bounded subsets of Y and all convex compact subsets of Y, respectively.

Definition 1

A set-valued map (s.v. map for short) \(F:X\rightarrow n(Y)\) is called K-subadditive if

$$\begin{aligned} F(x_1)+F(x_2) \subset F(x_1+x_2)+K \end{aligned}$$
(1)

for all \(x_1,x_2 \in X\). If F satisfies

$$\begin{aligned} F(x_1+x_2)\subset F(x_1)+F(x_2)+K \end{aligned}$$
(2)

for all \(x_1,x_2 \in X\) then it is called K-superadditive.

The concepts of K-subadditivity and K-superadditivity were introduced in [11], following the notions of K-midconvexity and K-midconcavity from [13]. Clearly, if F is a single-valued function, \(Y={\mathbb {R}}\) and \(K=[0,\infty )\), then K-subadditivity means the classical subadditivity, i.e. \(f:X\rightarrow {\mathbb {R}}\) satisfies

$$\begin{aligned} f(x_1+x_2)\le f(x_1)+f(x_2)\;\;\;\text{ for } \text{ every }\;x_1,x_2\in X, \end{aligned}$$

as well as K-superadditivity means the classical superadditivity, i.e. \(-f\) is subadditive.

Let us recall also the notion of K-lower (K-upper) semicontinuity from [13] which generalizes the classical notion of upper (lower, resp.) semicontinuity of a single-valued real function.

Definition 2

Let \(x_0\in X\). The s.v. map \(F:X\rightarrow n(Y)\) is called:

  • K-lower semicontinuous at \(x_0\), if for every neighborhood W of 0 in Y there exists a neighborhood U of 0 in X such that

    $$\begin{aligned} F(x_0)\subset F(x)+W+K\;\;\;\text{ for }\;x\in x_0+U, \end{aligned}$$
  • K-upper semicontinuous at \(x_0\), if for every neighborhood W of 0 in Y there exists a neighborhood U of 0 in X such that

    $$\begin{aligned} F(x)\subset F(x_0)+W+K\;\;\;\text{ for }\;x\in x_0+U. \end{aligned}$$

Moreover, F is K-lower (K-upper) semicontinuous on X, if it is K-lower (K-upper, resp.) semicontinuous at each point \(x\in X\).

Definition 3

A s.v. map \(F:X\rightarrow {\mathcal {B}}(Y)\) is called:

  • weakly K-upper bounded on a set \(A\subset X\), if there is \(B\in {\mathcal {B}}(Y)\) such that

    $$\begin{aligned} F(x)\cap (B-K)\ne \emptyset \;\;\;\text{ for } \text{ all }\;x\in A, \end{aligned}$$
  • K-upper bounded on a set \(A\subset X\), if there is \(B\in {\mathcal {B}}(Y)\) such that

    $$\begin{aligned} F(x)\subset B-K\;\;\;\text{ for } \text{ all }\;x\in A, \end{aligned}$$
  • weakly K-lower bounded on a set \(A\subset X\), if there is \(B\in {\mathcal {B}}(Y)\) such that

    $$\begin{aligned} F(x)\cap (B+K)\ne \emptyset \;\;\;\text{ for } \text{ all }\;x\in A, \end{aligned}$$
  • K-lower bounded on a set \(A\subset X\), if there is \(B\in {\mathcal {B}}(Y)\) such that

    $$\begin{aligned} F(x)\subset B+K\;\;\;\text{ for } \text{ all }\;x\in A. \end{aligned}$$

The above idea of (weak) K-upper/K-lower boundedness was introduced in [13] as a generalization of the notion of boundedness from above/below of single-valued real functions. More precisely, weak \([0,\infty )\)-upper boundedness, as well as \([0,\infty )\)-upper boundedness of a single-valued real function means its classical boundedness from above. The same holds for (weak) \([0,\infty )\)-lower boundedness of a single-valued real function.

Definition 4

A s.v. map \(F:X\rightarrow {\mathcal {B}}(Y)\) is called:

  • locally (weakly) K-upper (K-lower) bounded at \(x\in X\), if it is (weakly) K-upper (K-lower, resp.) bounded on some neighborhood of x,

  • locally K-bounded at \(x\in X\), if it is both locally weakly K-upper bounded and locally K-lower bounded at x,

  • locally K-bounded on X, if it is locally K-bounded at each point of X.

It is well known that for every subadditive real function defined on an abelian metric group boundedness from above on a “large” (in the sense of category or measure) set implies its local boundedness on the whole domain (see e.g. [12, Theorem 16.2.3]). Some recent results on subadditive functions can be found e.g. in [3]–[7].

In [2] the following generalization of the mentioned classical result was proved.

Theorem 1

[2, Theorem 2.2] Let X be an abelian metric group with an invariant metric and \(f:X\rightarrow {\mathbb {R}}\) be a subadditive function. If f is bounded from above on a non-null-finite set \(A\subset X\) then f is locally bounded on X.

The concept of null-finite sets was introduced in [1].Footnote 1

Definition 5

A subset A of an abelian metric group X is called null-finite if there exists a sequence \((x_n)_{n\in {\mathbb {N}}}\) tending to zero in X such that the set \(\{n\in {\mathbb {N}}:x+x_n\in A\}\) is finite for every \(x\in X\).

In a complete abelian metric group X with an invariant metric the following sets are not null-finite: open sets, non-meager sets with the Baire property, sets of positive Haar measure provided X is locally compact, universally measurable sets which are not Haar-null, Borel sets which are not Haar-meager (see [1, Theorems 5.1 and 6.1]).

The notions of a Haar-null set and a Haar-meager set have been introduced by Christensen [8] and Darji [9], respectively. A subset B of an abelian Polish group X is called:

  • Haar-meager if there exist a Borel set \(A\supset B\), a compact metric space K and a continuous function \(f:K\rightarrow X\) such that \(f^{-1}(A+x)\) is meager in K for every \(x\in X\);

  • Haar-null if there exist a universally measurable set \(A\supset B\) and a \(\sigma \)-additive probability Borel measure \(\mu \) on X such that \(\mu (A+x)=0\) for every \(x\in X\).

It was proved in [8] and [9] that every Haar-meager set is meager and, moreover, in every locally compact abelian Polish group the notions of a Haar-meager set and a Haar-null set are equivalent to the notions of a meager set and a set of Haar measure zero, respectively.

In [11] we proved the following generalization of Theorem 1.

Theorem 2

[11, Theorems 2 and 3] Let X be an abelian metric group with an invariant metric and Y be a locally convex real topological vector space. Assume that \(A\subset X\) is a non-null-finite set, K is a subsemigroup of Y, and \(F:X\rightarrow {\mathcal {B}}(Y)\) is a s.v. map. If F satisfies one of the following conditions:

  1. (i)

    F is K-subadditive and weakly K-upper bounded on A,

  2. (ii)

    F is K-superadditive and K-lower bounded on A,

then F is locally K-bounded on X.Footnote 2

Here we use the above theorem to show some relationships between a K-subadditivie or K-superadditive s.v. map \(F:X\rightarrow \mathcal{C}\mathcal{C}(Y)\) and the functionals \(f_{y^*}:X\rightarrow {\mathbb {R}}\) defined by

$$\begin{aligned} f_{y^*}(x) = \inf y^*(F(x))\;\;\;\text{ for }\; x\in X, \end{aligned}$$
(3)

where \(y^*\in K^*\) and \(K^*\) means the set of all real continuous functionals on a real topological vector space Y which are non-negative on K, i.e.

$$\begin{aligned} K^*=\{y^*\in Y^*:y^*(y)\ge 0 \ \text{ for } \text{ every } \ y\in K \}. \end{aligned}$$

The results obtained refer to [10, Theorem 5] and [14, Theorem 1], where the continuity of a K-midconvex (K-midconcave) s.v. map was proved under the assumption that the functionals (3) are lower (upper, resp.) semicontinuous. This paper relates also to [1, Theorem 9.1].

2 Main results

Theorem 3

Let X be an abelian metric group with an invariant metric, and \(A\subset X\) be a non-null-finite set. Assume that Y is a locally convex real topological vector space, K is a subsemigroup of Y, and \(F:X\rightarrow {\mathcal {B}}(Y)\) is a s.v. map. If one of the following conditions holds:

  1. (i)

    F is K-subadditive and weakly K-upper bounded on A,

  2. (ii)

    F is K-superadditive and weakly K-lower bounded on A,

then for every \(y^*\in K^*\) the functional \(f_{y^*}:X\rightarrow {\mathbb {R}}\) defined by (3) is locally bounded on X.

Proof

Assume that F satisfies (i) (if F satisfies (ii), the proof runs in the same way). Then there exists a bounded set \(B\subset Y\) such that

$$\begin{aligned} F(x)\cap (B-K)\ne \emptyset , \ \ x\in A. \end{aligned}$$
(4)

Fix any \( y^*\in K^*\) and take the functional \(f_{y^*}\) defined by (3). Since F is K-subadditive and \(y^*\in K^*\), we have

$$\begin{aligned} y^*(F(x_1))+ y^*(F(x_2)){} & {} = y^*(F(x_1)+F(x_2))\subset y^*(F(x_1+x_2)+K ) \\{} & {} \subset y^*(F(x_1+x_2))+[0,\infty ) \end{aligned}$$

for all \(x_1,x_2\in X\). Hence

$$\begin{aligned} f_{y^*}(x_1)+f_{y^*}(x_2){} & {} = \inf y^*(F(x_1))+\inf y^*(F(x_2))\\{} & {} =\inf \left( y^*(F(x_1))+ y^*(F(x_2))\right) \ge \inf y^*(F(x_1+x_2) \\{} & {} =f_{y^*}(x_1+x_2), \end{aligned}$$

which means that \(f_{y^*}\) is subadditive. By (4), for every \(x\in A\) we have

$$\begin{aligned} y^*(F(x))\cap y^*(B-K)\ne \emptyset . \end{aligned}$$

Hence

$$\begin{aligned} y^*(F(x))\cap \big ( y^*(B)+(-\infty , 0]\big ) \ne \emptyset . \end{aligned}$$
(5)

But the set \(y^*(B)\) is bounded, i.e. \(y^*(B)\subset [m, M]\) for some \(m<M\). Then, by (5),

$$\begin{aligned} y^*(F(x))\cap (-\infty , M] \ne \emptyset , \end{aligned}$$

which means that

$$\begin{aligned} f_{y^*}(x)\le M, \ \ x\in A. \end{aligned}$$

Consequently, in view of Theorem 1, \(f_{y^*}\) is locally bounded on X. \(\square \)

Now, which assumptions on the functionals \(f_{y^*}\) defined by (3) for \(y^*\in K^*\) imply the local K-boundedness on X of a K-subadditive s.v. map F? The next theorem gives an answer.

Let us recall that in a real vector space by a convex cone we mean the set K satisfying \(K+K\subset K\) and \(tK\subset K\) for every \(t\in [0,\infty )\).

Theorem 4

Let X be a complete metric space with an invariant metric, Y be a real normed space, and K be a convex cone in Y. If a s.v. map \(F:X\rightarrow \mathcal{C}\mathcal{C}(Y)\) is K-subadditive and for every \(y^*\in K^*\) the functional \(f_{y^*}:X\rightarrow {\mathbb {R}}\) defined by (3) is lower semicontinuous on X, then F is locally K-bounded on X.

Proof

Let B be the unit ball in Y and \(B_n:=nB\), \(n\in {\mathbb {N}}\). Denote

$$\begin{aligned} A_n:=\{x\in X:F(x)\cap \textrm{ cl}\,({B}_n - K)\ne \emptyset \},\;\; n\in {\mathbb {N}}. \end{aligned}$$

Clearly, \(\bigcup _{n\in {\mathbb {N}}} A_n = X.\) We will prove that the sets \(A_n\) are closed.

If \(A_n=X\), it is obvious. So, fix \(n\in {\mathbb {N}}\) such that \(A_n\ne X\) and fix \(x_0\in X\setminus A_n\). Then

$$\begin{aligned} F(x_0)\cap \textrm{ cl}\,(B_n-K)=\emptyset . \end{aligned}$$

Since \(F(x_0)\) is compact convex and \(B_n\) is convex, by the separation theorem (see [14, Lemma 1]) there exists a functional \(y^* \in K^*\) such that

$$\begin{aligned} \inf y^*(F(x_0)) > \sup y^*(B_n). \end{aligned}$$

Let

$$\begin{aligned} \varepsilon := \inf y^*(F(x_0))-\sup y^*(B_n). \end{aligned}$$

Since \(y^*\) is continuous at zero, we can find a neighborhood V of zero in Y with \(y^*(V)\subset \left( -\frac{\varepsilon }{2},\frac{\varepsilon }{2}\right) \). By the lower semicontinuity of \(f_{y^*}\) at \(x_0\) there exists a neighborhood U of \(x_0\) such that

$$\begin{aligned} f_{y^*}(x) > f_{y^*}(x_0)-\frac{\varepsilon }{2}\;\;\;\text{ for } \text{ every }\;x\in U. \end{aligned}$$

Now, we prove that

$$\begin{aligned} F(x)\cap \textrm{ cl}\,(B_n -K) =\emptyset \;\;\;\text{ for } \text{ every }\; x\in U. \end{aligned}$$
(6)

Indeed, if for some \(x\in U\) it was not true, there would exist a point

$$\begin{aligned} z\in F(x)\cap \textrm{ cl}\,(B_n -K)\subset F(x)\cap (B_n-K-V). \end{aligned}$$

Let \(z=b-k-v\) with some \(b\in B_n\), \(k\in K\) and \(v\in V\). Then

$$\begin{aligned} y^*(b)= & {} y^*(z)+y^*(k)+y^*(v)> y^*(z)-\frac{\varepsilon }{2} \\\ge & {} f_{y^*}(x)-\frac{\varepsilon }{2}>f_{y^*}(x_0)-\varepsilon = \sup y^*(B_n), \end{aligned}$$

a contradiction.

Thus (6) holds and, consequently, \(U\subset X{\setminus } A_n\) which shows that \(X{\setminus } A_n\) is open and hence \(A_n\) is closed.

Since X is a complete metric space and \(X=\bigcup _{n\in {\mathbb {N}}} A_n\), by the Baire category theorem there exists \(n_0 \in {\mathbb {N}}\) such that \(\textrm{ int}\,A_{n_0} \ne \emptyset .\) By the definition of \(A_{n_0}\) we have

$$\begin{aligned} F(x)\cap \big ((B_{n_0}+B)-K\big )\ne \emptyset \;\;\;\text{ for }\; x\in A_{n_0}. \end{aligned}$$

Since the set \(B_{n_0}+B\) is bounded, this means that F is weakly K-upper bounded on the set \(A_{n_0}\) with non-empty interior. Consequently, by Theorem 2, F is locally K-bounded on X. This finishes the proof. \(\square \)

The next example shows that it is not possible to get the same result in Theorem 4 if we weaken the assumption on the functionals \(f_{y^*}\), i.e. replace lower semicontinuity on X of \(f_{y^*}\) by local boundedness from below on X.

Example 1

Let \(K=[0,\infty )\) and \(F:{\mathbb {R}}\rightarrow \mathcal{C}\mathcal{C}({\mathbb {R}})\) be defined by

$$\begin{aligned} F(x)=\big [|a(x)|, |a(x)|+1 \big ]\;\;\;\text{ for } \text{ every } \;x\in {\mathbb {R}}, \end{aligned}$$

where \(a:{\mathbb {R}}\rightarrow {\mathbb {R}}\) is a discontinuous additive function. Clearly, then F is K-subadditive, even locally K-lower bounded on X, and every \(y^* \in K^*\) is given by \(y^*(x)=cx\) with some \(c\ge 0\). Hence

$$\begin{aligned} f_{y^*}(x)= \inf y^*\big (F(x)\big ) = \inf c\big [|a(x)|, |a(x)|+1\big ]=c|a(x)|\ge 0 \end{aligned}$$

which means that \(f_{y^*}\) is bounded from below on the whole \({\mathbb {R}}\), but F is not locally weakly K-upper bounded at any point.

Indeed, if for some open set \(U\subset {\mathbb {R}}\) and some \([m,M]\subset {\mathbb {R}}\) we had \(F(x)\cap \big ([m,M]-K\big )\ne \emptyset \) for \(x\in U\), then \(a(x) \le |a(x)| \le M\) for \(x\in U\), which would be impossible because of the discontinuity of a.

Problem 1

Let X, Y, K and F be as in Theorem 4. Can we obtain the same result replacing the lower semicontinuity of \(f_{y^*}\) on X by upper semicontinuity on X for every \(y^*\in K^*\), or by local boundedness on X for every \(y^*\in K^*\)?

We know that the answer to Problem 1 is positive in the case when \(Y={\mathbb {R}}\) and \(K=[0,\infty )\) (even under the weaker assumptions that for some nonzero \( y^* \in K^*\) the functional \(f_{y^*}\) is upper semicontinuous at a point of X or it is locally bounded at a point of X).

Indeed, let \(F(x)=[m(x),M(x)]\) for \(x\in X\). Then the K-subadditivity of F means the subadditivity of m. Moreover, \(f_{y*}(x)=c\,m(x)\), \(x\in X\), with \(c>0\) and the upper semicontinuity (local boundedness) of \(f_{y*}\) at some \(x_0\in X\) implies the upper semicontinuity (local boundedness, resp.) of m at \(x_0\), and hence m is locally bounded on X in view of Theorem 1. Thus, for every \(x\in X\) we can find a neighborhood U of x such that \(m(U)\subset [a,A]\) with some \(a<A\). For every \(t\in U\) we get

$$\begin{aligned} F(t)\cap ([a,A]-K){} & {} =[m(t),M(t)]\cap (-\infty , A]\ne \emptyset ;\\ F(t){} & {} = [m(t),M(t)]\subset [a,\infty )=[a,A]+K, \end{aligned}$$

which means that F is locally K-bounded on X.

Let us recall that the subadditivity of a single-valued real function \(f:X\rightarrow {\mathbb {R}}\) means the superadditivity of the function \(-f\). Unfortunately there is no analogous property for K-subadditive s.v. maps, i.e. the K-superadditivity of F does not mean the \(\pm K\)-subadditivity of \(-F\) (see [11, Examples 2 and 3]). That is why we have to prove an analogous result for K-superadditive s.v. maps independently.

Theorem 5

Let X be a complete metric space with an invariant metric, Y be a real normed space and K be a convex cone in Y. If a s.v. map \(F:X\rightarrow \mathcal{C}\mathcal{C}(Y)\) is K-superadditive and for every \(y^*\in K^*\) the functional \(f_{y^*}:X\rightarrow {\mathbb {R}}\) defined by (3) is upper semicontinuous on X, then F is locally K-bounded on X.

Proof

Let B be the unit ball in Y and \(B_n:=nB\) for \(n\in {\mathbb {N}}\). Define

$$\begin{aligned} A_n:=\{x\in X:F(x)\subset \textrm{ cl}\,({B}_n + K)\},\;\; n\in {\mathbb {N}}. \end{aligned}$$
(7)

Then \(\bigcup _{n\in {\mathbb {N}}} A_n = X.\) We will show that \(F(x)\subset \textrm{ cl}\,(B_n+K)\) for every \(x\in \textrm{ cl}\,A_n\) and \(n\in {\mathbb {N}}\).

For a proof by contradiction suppose that there are \(n_0\in {\mathbb {N}}\), \(x_0\in \textrm{ cl}\,A_{n_0}\) and \(z\in F(x_0) {\setminus } \textrm{ cl}\,(B_{n_0}+K).\) Since the set \(\textrm{ cl}\,(B_{n_0}+K)\) is convex and closed, by the separation theorem (see e.g [15, Theorem 3.4]) there exists a continuous linear functional \(y^* \in Y^*\) such that

$$\begin{aligned} y^*(z) < \inf y^*\big (\textrm{ cl}\,(B_{n_0}+ K) \big ). \end{aligned}$$
(8)

Since

$$\begin{aligned} y^*(k)\ge y^*(z)-y^*(b_0)=:M\;\;\;\text{ for } \text{ all }\;k\in K \end{aligned}$$

with arbitrarily fixed \(b_0 \in B_{n_0},\) and, moreover,

$$\begin{aligned} y^*(k) = \frac{1}{m}y^*(mk) \ge \frac{1}{m} M\;\;\;\text{ for } \text{ every }\; m\in {\mathbb {N}}, \end{aligned}$$

for \(m\rightarrow \infty \) we obtain \(y^*(k)\ge 0\) for all \(k\in K\) which means that \(y^* \in K^*.\) Now, put

$$\begin{aligned} \varepsilon :=\inf y^*\big (\textrm{ cl}\,(B_{n_0}+K)\big )-y^*(z). \end{aligned}$$

By the upper semicontinuity of \(f_{y^*}\) at \(x_0\) there exists a neighborhood U of \(x_0\) such that

$$\begin{aligned} f_{y^*}(x) < f_{y^*}(x_0) + \varepsilon \;\;\;\text{ for } \text{ every }\; x\in U. \end{aligned}$$
(9)

Since \(x_0\in \textrm{ cl}\,A_{n_0}\), there exists an \(x_1 \in A_{n_0} \cap U.\) Then, according to (9) and the definition of \(\varepsilon \), we obtain

$$\begin{aligned} \begin{array}{ll} f_{y^*}(x_1)\!\!\! &{}< f_{y^*}(x_0) + \varepsilon \le y^*(z) +\varepsilon =\inf y^* \big (\textrm{ cl}\,(B_{n_0}+K) \big ) \\ &{}\le \inf y^*\big (F(x_1) \big )= f_{y^*}(x_1). \end{array} \end{aligned}$$

This contradiction proves that F is K-lower bounded on \(\textrm{ cl}\,A_n\) for every \(n\in {\mathbb {N}}\), i.e.

$$\begin{aligned} F(x)\subset \textrm{ cl}\,(B_{n}+K)\;\;\;\text{ for } \text{ every }\; x\in \textrm{ cl}\,A_{n}. \end{aligned}$$

Since \(X=\bigcup _{n\in {\mathbb {N}}} \textrm{ cl}\,A_n\) and X is complete, in view of the Baire category theorem there exists \(n_0 \in {\mathbb {N}}\) such that \(\textrm{ int}\,\textrm{ cl}\,A_{n_0} \ne \emptyset .\) Moreover,

$$\begin{aligned} F(x)\subset \textrm{ cl}\,(B_{n_0}+K)\subset B_{n_0}+B+K\;\;\;\text{ for } \text{ every }\; x\in \textrm{ cl}\,A_{n_0}, \end{aligned}$$

so F is K-lower bounded on an open set. Now, to complete the proof it is enough to apply Theorem 2. \(\square \)

Modifying Example 1, it is easy to observe that it is impossible to get the same result in Theorem 5 assuming the local boundedness from above on X (instead of upper semicontinuity on X) of functionals \(f_{y^*}\).

Example 2

Let \(K=(-\infty ,0]\) and \(F:{\mathbb {R}}\rightarrow \mathcal{C}\mathcal{C}({\mathbb {R}})\) be defined by

$$\begin{aligned} F(x)=\big [|a(x)|, |a(x)|+1 \big ]\;\;\;\text{ for } \text{ every } \;x\in {\mathbb {R}}, \end{aligned}$$

where \(a:{\mathbb {R}}\rightarrow {\mathbb {R}}\) is a discontinuous additive function. Clearly, then F is K-superadditive, even locally K-upper bounded on X, and every \(y^* \in K^*\) is given by \(y^*(x)=cx\) with some \(c\le 0\). Hence

$$\begin{aligned} f_{y^*}(x)= \inf y^*\big (F(x)\big ) = \inf c\big [|a(x)|, |a(x)|+1\big ]=c\big (|a(x)|+1\big )\le 0 \end{aligned}$$

which means that \(f_{y^*}\) is bounded from above on the whole \({\mathbb {R}}\), but F is not locally K-lower bounded at any point.

The following question arises which is analogous to Problem 1.

Problem 2

Let X, Y, K and F be as in Theorem 5. Can we obtain the same result replacing the upper semicontinuity of \(f_{y^*}\) on X by lower semicontinuity on X for every \(y^*\in K^*\), or by local boundedness on X for every \(y^*\in K^*\)?

For now we can give a positive answer only in the case \(Y={\mathbb {R}}\) and \(K=[0,\infty )\) (the solution runs in the same way as the solution of Problem 1).

Remark 1

Notice that in the conclusion of Theorems 4 and 5 we are not able to get the local K-upper boundedness (instead of the local weak K-upper boundedness) of F on X.

For example, let \(K=[0,\infty )\) and \(F:{\mathbb {R}}\rightarrow \mathcal{C}\mathcal{C}({\mathbb {R}})\) be given by

$$\begin{aligned} F(x)=\left\{ \begin{array}{ll} \left[ 0,\frac{1}{|x|}\right] , &{}\quad x\ne 0,\\[1.5ex] \{0\},&{}\quad x=0. \end{array} \right. \end{aligned}$$

Clearly, F is K-subadditive and K-superadditive. Moreover, for every \(y^*\in K^*\) (i.e. \(y^*(x)=cx\) with \(c\ge 0\)) the functional \(f_{y^*}=0\), so it is continuous on \({\mathbb {R}}\). But F is not K-upper bounded at 0.

3 Applications and final remarks

One can easily observe that if \(F:X\rightarrow {\mathcal {B}}(Y)\) is a s.v. map, where X is a metric space and Y is a real normed space, then

  1. (i)

    the K-upper semicontinuity of F at a point \(x_0\in X\) implies local K-lower boundedness at this point,

  2. (ii)

    the K-lower semicontinuity of F at a point \(x_0\in X\) implies local weak K-upper boundedness at this point.

Indeed, to obtain (i) it is enough to put the bounded set \(B:=F(x_0)+W\) with a fixed neighbourhood \(W\subset Y\) of 0. The proof of (ii) runs by contradiction. Fix a neighborhood \(W\subset Y\) of 0. Then, by the K-lower semicontinuity of F, there is a neighbourhood U of \(x_0\) such that

$$\begin{aligned} F(x_0)\subset F(x)+W+K\;\;\;\text{ for } \text{ every }\;\;x\in U. \end{aligned}$$

If for the bounded set \(B:=F(x_0)-W\) we could find \(x_1\in U\) such that \(F(x_1)\cap (B-K)=\emptyset \), then

$$\begin{aligned} (F(x_1)+W+K)\cap F(x_0)=\emptyset . \end{aligned}$$

This contradiction proves that F is weakly K-upper bounded at \(x_0\).

Moreover, in view of Examples 1 and 2, a K-subadditive s.v. map which is locally K-lower bounded on X, as well as K-superadditive s.v. map which is locally K-upper bounded on X, needn’t be locally K-bounded on X.

However, as an immediate consequence of Theorems 4 and 5, we obtain the following result.

Corollary 6

Let X be a complete metric space with an invariant metric, Y be a real normed space, and K be a convex cone in Y. If a s.v. map \(F:X\rightarrow \mathcal{C}\mathcal{C}(Y)\) satisfies one of the following two conditions:

  1. (i)

    F is K-subadditive and K-upper semicontinuous on X,

  2. (ii)

    F is K-superadditive and K-lower semicontinuous on X,

then F is locally K-bounded on X.

Proof

Assume that F satisfies (i) (if F satisfies (ii) the proof runs in the same way). Let B be the unit ball in Y. Fix arbitrary \(x_0\in X\), \(y^* \in K^*\) and \(\varepsilon >0\). Since F is K-upper semicontinuous at \(x_0\), there exists a neighborhood U of 0 in X such that

$$\begin{aligned} F(x) \subset F(x_0) + \frac{\varepsilon }{\Vert y^*\Vert }B + K\;\;\;\text{ for } \text{ every }\;x\in x_0 +U. \end{aligned}$$

Hence

$$\begin{aligned} y^*\big (F(x)\big ) \subset y^*\big (F(x_0)\big ) + \frac{\varepsilon }{\Vert y^*\Vert }[-\Vert y^*\Vert , \Vert y^*\Vert ] + [0,\infty ), \ \ \ x\in x_0 +U. \end{aligned}$$

Consequently,

$$\begin{aligned} f_{y^*}(x)\ge f_{y^*}(x_0) -\varepsilon , \ \ \ x\in x_0 +U, \end{aligned}$$

which shows that \(f_{y^*}\) is lower semicontinuous at \(x_0\). To complete the proof it is enough to apply Theorem 4. \(\square \)

In Theorems 4 and 5 we assume that K is a convex cone in a real normed space, but considerations on K-subadditivity/K-superadditivity seem to be much more natural for a subsemigroup K.

It is easy to check that if K is a subsemigroup with 0, then \(\textrm{ conv}\,K\) is a convex cone. Clearly then

  • K-subadditivity implies \(\textrm{ conv}\,K\)-subadditivity,

  • K-superadditivity implies \(\textrm{ conv}\,K\)-superadditivity,

  • local K-boundedness at a point implies local \(\textrm{ conv}\,K\)-boundedness at the same point,

  • \(y^*\in K^*\) implies \(y^*\in (\textrm{ conv}\,K)^*\).

Consequently, by Theorems 4 and 5, we get the following corollary.

Corollary 7

Let X be a complete metric space with an invariant metric, Y be a real normed space, and K be a subsemigroup of Y with 0. Assume that \(F:X\rightarrow \mathcal{C}\mathcal{C}(Y)\) is a s.v. map and \(f_{y^*}:X\rightarrow {\mathbb {R}}\) is the functional defined by (3) for every \(y^*\in K^*\). If one of the following conditions holds:

  1. (i)

    F is K-subadditive and \(f_{y^*}\) is lower semicontinuous on X for every \(y^*\in K^*\),

  2. (ii)

    F is K-superadditive and \(f_{y^*}\) is upper semicontinuous on X for every \(y^*\in K^*\),

then F is locally \(\textrm{ conv}\,K\)-bounded on X.

Problem 3

Is it possible to get the stronger result that F is locally K-bounded on X in Corollary 7?

The next corollary gives an answer in the case \(Y={\mathbb {R}}.\)

Corollary 8

Let X be a complete metric space with an invariant metric, and K be a subsemigroup of \({\mathbb {R}}\) with 0. Assume that \(F:X\rightarrow \mathcal{C}\mathcal{C}({\mathbb {R}})\) is a s.v. map and \(f_{y^*}:X\rightarrow {\mathbb {R}}\) is the functional defined by (3) for every \(y^*\in K^*\). If one of the following conditions holds:

  1. (i)

    F is K-subadditive and \(f_{y^*}\) is lower semicontinuous on X for every \(y^*\in K^*\),

  2. (ii)

    F is K-superadditive and \(f_{y^*}\) is upper semicontinuous on X for every \(y^*\in K^*\),

then F is locally K-bounded on X.

Proof

Assume that F satisfies (i) (the proof for (ii) is similar). Let \(F(x)=[m(x),M(x)]\) for \(x\in X\).

The case when \(\textrm{ conv}\,K=\{0\}\) is trivial because then \(K=\{0\}\), so it is enough to use Corollary 7.

Now, consider the case when \(\textrm{ conv}\,K=[0,\infty )\). By Corollary 7F is locally \(\textrm{ conv}\,K\)-bounded on X; i.e., for every \(x\in X\) there are a neighborhood U of x and intervals \([m_1,M_1],[m_2,M_2]\) with \(m_i<M_i\) for \(i=1,2\) such that for any \(t\in U\)

$$\begin{aligned}{} & {} {[}m(t),M(t)]\cap (-\infty ,M_1]=F(t)\cap ([m_1,M_1]-\textrm{ conv}\,K)\ne \emptyset ,\\{} & {} {[}m(t),M(t)]=F(t)\subset [m_2,M_2]+\textrm{ conv}\,K=[m_2,\infty ). \end{aligned}$$

Hence \(m(U)\subset [m_2,M_1].\) Since \(0\in K\), for \(B=[m_2,M_1]\) and every \(t\in U\)

$$\begin{aligned} F(t)\cap (B-K)\supset [m(t),M(t)] \cap B=[m(t),M(t)]\cap [m_2,M_1]\ne \emptyset , \end{aligned}$$

which means that F is weakly K-upper bounded at \(x\in X\). Hence, in view of Theorem 2, F is locally K-bounded on X.

If \(\textrm{ conv}\,K=(-\infty ,0]\), then the proof runs in a similar way (then M is bounded at x).

Finally, let \(\textrm{ conv}\,K={\mathbb {R}}\). Then there are \(k_1,k_2\in K\) with \(k_1<0\) and \(k_2>0\). Moreover, for every \(x\in X\)

$$\begin{aligned}{} & {} F(x)\cap \big ([k_1,k_2]-K\big )=F(x)\cap {\mathbb {R}}\ne \emptyset ,\\{} & {} F(x)\subset {\mathbb {R}}=[k_1,k_2]+K, \end{aligned}$$

which means that F is locally K-bounded on X. \(\square \)