Keywords

1 Introduction

The Hyers–Ulam stability problems of functional equations were originated by S. M. Ulam in 1940 when he proposed the following question [36]:

Let \(f\) be a mapping from a group G 1 to a metric group G 2 with metric \(d(\cdot, \cdot)\) such that

$$d(f(x y),\, f(x) f(y))\leq \varepsilon.$$

Then does there exist a group homomorphism \(h\) and \(\delta_\epsilon>0\) such that

$$d(f(x),\, h(x))\leq\delta_\epsilon$$

for all \(x\in G_1\) ?

One of the first assertions to be obtained is the following result, essentially due to D. H. Hyers [20], that gives an answer for the question of Ulam.

Theorem 1

Suppose that S is a commutative semigroup, B is a Banach space, \(\epsilon\ge0\) , and \(f:S \to B\) satisfies the inequality

$$\|f(x+y)-f(x)-f(y)\|\le \epsilon$$
(1)

for all \(x,\,y\in S\) . Then there exists a unique function \(A:S \to B\) satisfying

$$A(x+y)=A(x)+A(y)$$
(2)

and

$$\|f(x)-A(x)\|\le \epsilon$$
(3)

for all \(x\in S\) .

In 1950, this result was generalized by T. Aoki [4] and D.G. Bourgin [9, 8]. In 1978 T.M. Rassias generalized the Hyers’ result to new approximately linear mappings [?]. Since then the stability problems have been investigated in various directions for many other functional equations. Among the results, the stability problem in a restricted domain was investigated by F. Skof, who proved the stability problem of the inequality (1) in a restricted domain [35]. Several papers have been published on the Hyers–Ulam stability in restricted domains for a large variety of functional equations including the Jensen functional equation [24], quadratic type functional equations [23], mixed type functional equations [30], and Jensen type functional equations [31]. The results can be summarized as follows: Let X and B be a real normed space and a real Banach space, respectively. For fixed \(d\ge0\), if \(f:X \to B\) satisfies the functional inequalities (such as that of Cauchy, quadratic, Jensen, and Jensen type, etc.) for all \(x, y\in X\) with \(\|x\|+\|y\|\ge d\), then the inequalities hold for all \(x, y\in X\).

In [14, 15], generalizing the restricted domains such as \(\|x\|+\|y\|\ge d\) in a normed space to some abstract domains in a group, we consider the stability problem of Pexider equation and Jensen-type equations in the restricted domains. In the present paper, we consider a weak version of Hyers–Ulam stability of the Pexider equation when the target space of the functions in given functional inequalities are not a normed space but a 2-divisible commutative group. Note that the existence of the approximate additive function A in Theorem 1 is due to the completeness of the target space B. For example, if Y is a noncomplete normed space and \(f:S \to Y\) satisfies (1), then we can only find a Cauchy sequence \(a_n: S\to Y\) such that

$$|a_n(x+y)-a_n(x)-a_n(y)|\le 2^{-n}\epsilon$$
(4)

for all \(x, y\in S\), \(n=1,2, 3, \ldots\), and

$$|f(x)-a_n(x)|\le \epsilon$$
(5)

for all \(x\in S\) and \(n=1,2, 3, \ldots\). Throughout this paper, we denote a commutative group by G and a 2-divisible commutative group by H respectively, \(0\in V\subset H\) and \(W\subset G\times G\). Also, we denote a Banach space and a real normed space by B and Y, respectively, and \(f, g, h:G\to H\)(or Y, B). In Sect. 2 of this chapter, we consider the behavior of \(f:G\to H\) satisfying

$$f(x+y)-f(x)-f(y)\in V$$
(6)

for all \(x, y\in G\). As a result we prove that there exists a Cauchy-type sequence \(a_n:G\to H\) (which is a Cauchy sequence when H = Y) such that

$$f(x)-a_n(x)\in 2^{-n}\big(V+2~V+\ldots +2^{n-1}V\big)$$
(7)

for all \(x\in G\). In Sect. 3, we consider

$$f(x+y)-g(x)-h(y)\in V$$
(8)

for all \((x, y)\in W\subset G\times G\). As the main result we prove that under some assumptions on W, if \(f,\, g,\, h\) satisfy (8) then there exist approximate Cauchy-type sequences \(a_n,\, b_n\), and c n for \(f,\, g\), and h respectively. From our result we obtain the Hyers–Ulam stability theorem for Pexider equation when \(f, g, h:G\to B\).

2 A Weak Stability of Pexider Equation

For subsets \(V, V_1, V_2\) of H, \(v\in V\), and \(n\in \mathbb{N}\), we define

$$nv= \underbrace{v+\cdots +v}_{n-\text{times}}, nV =\{nv:v\in V\}, 2^{-n}V=\{h\in H: 2^n h\in V\},$$

and

$$V_1+V_2=\{v_1 +v_2: v_1\in V_1, v_2\in V_2\}.$$

We call \(a_n:G \to H\) a V-Cauchy sequence if

$$a_{m+n}(x)-a_m(x)\in 2^{-m-n} \big(V+2~V+\ldots +2^{n-1}V\big)$$

for all \(m, n=1, 2, 3, \ldots,\) and \(x\in G\).

First we consider the weak version of the Hyers–Ulam stability theorem for the Cauchy equation.

Theorem 2

Suppose that \(f:G \to H\) satisfies

$$f(x+y)-f(x)-f(y)\in V$$
(9)

for all \(x,y\in G\) . Then there exists a V -Cauchy sequence \(a_n:G \to H\) satisfying

$$a_n(x+y)-a_n(x)-a_n(y)\in 2^{-n}V,$$
(10)

and

$$a_n(x)-f(x)\in 2^{-n}\big(V+2~V+\ldots +2^{n-1}V\big)$$
(11)

for all \(x, y\in G\) and \(n\in \mathbb{N}\) .

Proof

Note that since H is 2-divisible, for each \(n\in \mathbb{N}\) and \(x\in G\) we can choose an a n (x) such that

$$2^{n}a_n(x)= f(2^{n}x).$$
(12)

Replacing y by x in (9) and using induction argument we have

$$\begin{aligned} 2^{n-1}f(2x)-2^nf(x)&\in 2^{n-1}V\nonumber\\ 2^{n-2}f(4x)-2^{n-1}f(2x)&\in 2^{n-2}V\nonumber\\ \cdots \cdots\cdots\cdots\cdots \cdots \nonumber\\ 2~f(2^{n-1}x)-4~f(2^{n-2}x)&\in 2~V\nonumber\\ f(2^{n}x)-2~f(2^{n-1}x)&\in V\nonumber\end{aligned}$$

for all \(x\in G\). Thus it follows that

$$f(2^{n}x)-2^{n}f(x)\in V+2~V+\ldots +2^{n-1}V$$
(13)

for all \(x\in G\). Now it follows from (12) and (13) that

$$a_n(x)-f(x)\in 2^{-n}\big(V+2~V+\ldots +2^{n-1}V\big)$$
(14)

for all \(x\in G\). Replacing x by \(2^m x\) in (13) and using (12) we have

$$a_{m+n}(x)-a_m(x)\in 2^{-m-n} \big(V+2~V+\ldots +2^{n-1}V\big)$$
(15)

for all \(x\in G\), which implies that a n is V-Cauchy. Replacing x by \(2^n x\) and y by \(2^n y\) in (9) and using (12) we have

$$a_{n}(x+y)-a_n(x)-a_n(y)\in 2^{-n} V$$
(16)

for all \(n\in \mathbb{N}\) and \(x\in G\). This completes the proof.

Let \(\langle Y, \|\cdot\|\rangle\) be a normed space and \(V=\{x\in Y: \|x\|\le \epsilon\}\). Then we have

$$2^{-n}\big(V+2~V+\ldots +2^{n-1}V\big)\subset \{x\in Y:\|x\|\le \epsilon\}$$

for all \(n\in \mathbb{N}\), and

$$2^{-m-n}\big(V+2~V+\ldots +2^{n-1}V\big)\subset \{x\in Y:\|x\|\le 2^{-m}\epsilon\}$$

for all \(m, n\in \mathbb{N}\). Thus in this case, every V-Cauchy sequence is a Cauchy sequence. Now as a direct consequence of Theorem 2 we have the following.

Corollary 1

Let \(\epsilon>0\) . Suppose that \(f:G \to Y\) satisfies

$$\|f(x+y)-f(x)-f(y)\|\le \epsilon$$
(17)

for all \(x,y\in G\) . Then there exists a Cauchy sequence \(a_n:G \to Y\) satisfying

$$\|a_{n}(x+y)-a_n(x)-a_n(y)\|\le 2^{-n}\epsilon$$
(18)

for all \(n\in \mathbb{N}\) and \(x, y\in G\) , and

$$\|a_n(x)-f(x)\|\le \epsilon$$
(19)

for all \(x\in G\) .

In particular, if \(f:G \to B\), then there exists \(A:G \to B\) such that

$$\lim_{n\to \infty}a_n(x)=A(x).$$

Letting \(n\to \infty\) in (18) we have

$$A(x+y)-A(x)-A(y)=0$$
(20)

for all \(x, y\in G\). We call a function \(A:G \to B\) satisfying (20) an additive function. Thus as a direct consequence of Corollary 1 we have the well known Hyers–Ulam stability theorem.

Corollary 2

Let \(\epsilon>0\) . Suppose that \(f:G \to B\) satisfies

$$\|f(x+y)-f(x)-f(y)\|\le \epsilon$$
(21)

for all \(x,y\in G\) . Then there exists an additive function \(A:G \to B\) such that

$$\|f(x)-A(x)\|\le \epsilon$$
(22)

for all \(x\in G\) .

Throughout this chapter we denote

$$V^*=\{v_1+v_2-v_3-v_4: v_j\in V,\, j=1, 2, 3,4\}.$$

Theorem 3

Suppose that \(f, g, h:G \to H\) satisfy

$$f(x+y)-g(x)-h(y)\in V$$
(23)

for all \(x,y\in G\) . Then there exist \(V^*\) -Cauchy sequences \(a_n, b_n, c_n:G \to H\) satisfying

$$\begin{aligned} a_{n}(x+y)-a_n(x)-a_n(y)&\in 2^{-n}V^{*}\end{aligned}$$
(24)
$$\begin{aligned} b_{n}(x+y)-b_n(x)-b_n(y)&\in2^{-n}V^{*}\end{aligned}$$
(25)
$$\begin{aligned} c_{n}(x+y)-c_n(x)-c_n(y)&\in2^{-n}V^{*}\end{aligned}$$
(26)

for all \(n\in \mathbb{N}\) and \(x, y\in G\) , and

$$\begin{aligned} a_n(x)-f(x)+f(0)& \in V_n^{*},\end{aligned}$$
(27)
$$\begin{aligned} b_n(x)-g(x)+g(0)&\in V_n^{*}, \end{aligned}$$
(28)
$$\begin{aligned} c_n(x)-h(x)+h(0)&\in V_n^{*},\end{aligned}$$
(29)

and

$$a_n(x+y)-b_n(x)-c_n(y)\in V_n^{**}$$
(30)

for all \(n\in \mathbb{N}\) and \(x, y\in G\) , where

$$V_n^{*}= 2^{-n}(V^*+2~V^*+\ldots +2^{n-1}V^*),$$
$$V_n^{**}= V-V+V_n^{*}-V_n^{*}-V_n^{*}.$$

Proof

Let \(D(x, y)=f(x+y)-g(x)-h(y)\). Then we have

$$\begin{aligned} &f(x+y)-f(x)-f(y)+f(0)=D(x, y)+D(0, 0)-D(x,0)-D(y, 0)\in V^*\end{aligned}$$
(31)
$$\begin{aligned} &g(x+y)-g(x)-g(y)+g(0)=D(x, y)+D(y, 0)-D(x+y,0)-D(0, y)\in V^*\end{aligned}$$
(32)
$$\begin{aligned} &h(x+y)-h(x)-h(y)+h(0)=D(x, y)+D(0, x)-D(0,x+y)-D(x, 0)\in V^*\end{aligned}$$
(33)

for all \(x, y\in G\). Thus, in view of (31), (32), and (33), using Theorem 2 for \(f(x)-f(0),\, g(x)-g(0),\, h(x)-h(0)\), we obtain (24)–(29). Now, putting \(x=y=0\) in (23), we have

$$f(0)-g(0)-h(0)\in V.$$
(34)

Then, by (23), (27), (28), (29), and (34) we get (30).

This completes the proof.

In particular, let \(V=\{x\in Y: \|x\|\le \epsilon\}\). Then we have

$$V_n^*\subset \{x\in Y:\|x\|\le 4\epsilon\}, V_n^{**}\subset \{x\in Y:\|x\|\le 14\epsilon\}$$

for all \(n\in \mathbb{N}\). Thus as a direct consequence of Theorem 3 we have the following.

Corollary 3

Let \(\epsilon>0\) . Suppose that \(f, g, h:G \to Y\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(35)

for all \(x,y\in G\) . Then there exist Cauchy sequences \(a_n, b_n, c_n:G \to Y\) satisfying

$$\begin{aligned} \|a_{n}(x+y)-a_n(x)-a_n(y)\|&\le2^{-n+2}\epsilon\end{aligned}$$
(36)
$$\begin{aligned} \|b_{n}(x+y)-b_n(x)-b_n(y)\|&\le2^{-n+2}\epsilon\end{aligned}$$
(37)
$$\begin{aligned} \|c_{n}(x+y)-c_n(x)-c_n(y)\|&\le2^{-n+2}\epsilon\end{aligned}$$
(38)

for all \(n\in\mathbb{N}\) and \(x, y\in G\) , and

$$\begin{aligned} \|f(x)-a_n(x)-f(0)\|& \le 4\epsilon,\end{aligned}$$
(39)
$$\begin{aligned} \|g(x)-b_n(x)-g(0)\|&\le 4\epsilon,\end{aligned}$$
(40)
$$\begin{aligned} \|h(x)-c_n(x)-h(0)\|&\le 4\epsilon\end{aligned}$$
(41)

and

$$\|a_n(x+y)-b_n(x)-c_n(y)\|\le 14\epsilon$$
(42)

for all \(n\in \mathbb{N}\) and \(x, y\in G\) .

Corollary 4

Let \(\epsilon>0\) . Suppose that \(f, g, h:G \to B\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(43)

for all \(x,y\in G\) . Then there exists an additive function \(A:G \to B\) such that

$$\begin{aligned} \|f(x)-A(x)-f(0)\|& \le 4\epsilon,\nonumber\\ \|g(x)-A(x)-g(0)\|&\le 4\epsilon, \nonumber\\ \|h(x)-A(x)-h(0)\|&\le 4\epsilon\nonumber\end{aligned}$$

for all \(x\in G\) .

Proof

Let \(A_1(x)=\lim_{n\to \infty}a_n(x),\,\,A_2(x)=\lim_{n\to \infty}b_n(x),\,\,A_3(x)=\lim_{n\to \infty}c_n(x)\). Then it follows from (36)–(38) that for each \(j=1, 2, 3\), A j is an additive function. Letting \(n \to \infty\) in (39)–(41) we have

$$\begin{aligned} \|f(x)-A_1(x)-f(0)\|& \le 4\epsilon,\nonumber\\ \|g(x)-A_2(x)-g(0)\|&\le 4\epsilon, \nonumber\\ \|h(x)-A_3(x)-h(0)\|&\le 4\epsilon\nonumber\end{aligned}$$

for all \(x\in G\). Finally, letting \(n\to \infty\) in (42) we have

$$\|A_1(x+y)-A_2(x)-A_3(y)\|\le 14\epsilon$$
(44)

for all \(x,y\in G\). Putting y = 0 and x = 0 in (44) separately, we have

$$\begin{aligned} \|A_1(x)-A_2(x)\|&\le 14\epsilon\nonumber\\ \|A_1(y)-A_3(y)\|&\le 14\epsilon\nonumber\end{aligned}$$

for all \(x,y \in G\), which implies that \(A_1=A_2\) and \(A_1=A_3\). This completes the proof.

3 Weak Stability of Pexider Equation in Restricted Domains

It is a frequent situation to consider a functional equation satisfied in a restricted domain or satisfied under a restricted condition [3, 57, 1012, 15, 18, 28, 3235]. In this section we consider the weak version of the Hyers–Ulam stability theorem in some restricted domains in G. We use the following usual notations. Let \(G\times G=\{(a_1, a_2):a_1, a_2\in G\}\) be the product group. For a subset K of \(G\times G\) and \(a\in G\times G\), we define \(a+K=\{a+k: k \in K\}\). For given \(x, y\in G\) we denote the sets of points of the forms (not necessarily distinct) in \(G\times G\) by \(P_{x,y}, \,Q_{x,y},\, \text{and} R_{x, y}\), respectively as,

$$\begin{aligned} P_{x, y}&=\{(0, 0), (x,0), (0, y), (x, y)\},\nonumber\\ Q_{x, y}&=\{(y, 0), (0,y), (x, y), (x+y, 0)\},\nonumber\\ R_{x, y}&=\{(x, 0), (0,x), (x, y), (0, x+y)\},\nonumber\end{aligned}$$

where 0 is the identity element of G. The set \(P_{x, y}\) can be viewed as the vertices of a rectangle in \(G\times G\), and \(Q_{x, y}\) and \(R_{x, y}\) can be viewed as the vertices of parallelograms in \(G\times G\).

Definition 1

Let \(W\subset G\times G\). We introduce the following conditions \((C1),\, (C2),\) and \((C3)\) on W: For any \(x, y\in G\), there exist \(z_1, z_2, z_3\in G\) such that

$$\begin{aligned} (C1)& (-z_1, z_1)+P_{x, y}\subset W,\nonumber\\ (C2)& (0,z_2)+Q_{x, y} \subset W,\nonumber\\ (C3)& (z_3, 0)+R_{x, y} \subset W,\nonumber\end{aligned}$$

respectively.

Example

1 Let G be a real normed space. For \(\alpha, \beta, d \in \mathbb{R}\), let

$$\begin{aligned} U&=\{(x, y)\in G\times G:\,\ \alpha\|x\|+\beta \|y\|\ge d\},\end{aligned}$$
(45)
$$\begin{aligned} V&=\{(x, y)\in G\times G:\,\ \|\alpha x+\beta y\|\ge d\}.\end{aligned}$$
(46)

Then U satisfies \((C1)\) if \(\alpha + \beta>0\), \((C2)\) if \(\beta>0\) and \((C3)\) if \(\alpha> 0\), and V satisfies \((C1)\) if \(\alpha \ne \beta\), \((C2)\) if \(\beta \ne 0\) and \((C3)\) if \(\alpha \ne 0\).

Example

2 Let G be a real inner product space. For \(d\ge 0,\,x_0, y_0\in G\)

$$\begin{aligned} U=\{(x, y)\in G\times G:\,\langle x_0, x\rangle +\langle y_0, y\rangle \ge d\}.\end{aligned}$$
(47)

Then U satisfies \((C1)\), if \(x_0\ne y_0\), \((C2)\) if \(y_0\ne 0\) and \((C3)\) if \(x_0\ne 0\).

Example

3 Let G be the group of nonsingular square matrices with the operation of matrix multiplication. For \(\alpha, \beta \in \mathbb{R},\, \,\delta, d\ge0\), let

$$\begin{aligned} U&=\{(P_1, P_2)\in G\times G:|\det P_1|^\alpha|\det P_2|^\beta\le \delta\},\end{aligned}$$
(48)
$$\begin{aligned} U&=\{(P_1, P_2)\in G\times G:|\det P_1|^\alpha|\det P_2|^\beta\ge d\}.\end{aligned}$$
(49)

Then U satisfies \((C1)\) if \(\alpha \ne \beta\), \((C2)\) if \(\beta \ne 0\), and \((C3)\) if \(\alpha \ne 0\).

In the following one can see that if \(P_{x, y},Q_{x, y}\), and \(R_{x, y}\) are replaced by arbitrary subsets of four points (not necessarily distinct) in \(G\times G\), respectively, the conditions become stronger, that is, there are subsets \(U_j,\, j=1, 2, 3,\) which satisfy the conditions \((C1)\), \((C2)\), and \((C3)\), respectively, but \(U_j,\,j=1, 2, 3,\) fail to fulfill the following conditions (2.6), (2.7), and (2.8), respectively: For any subset \(\{p_1, p_2, p_3, p_4\}\) of points (not necessarily distinct) in \(G\times G\), there exists a \(z\in G\) such that

$$\begin{aligned} (e, z)\{p_1, p_2, p_3, p_4\}(z^{-1}, e)\subset U_1,\end{aligned}$$
(50)
$$\begin{aligned} \{p_1, p_2, p_3, p_4\}(e,z)\subset U_2,\end{aligned}$$
(51)
$$\begin{aligned} (z, e)\{p_1, p_2, p_3, p_4\}\subset U_3,\end{aligned}$$
(52)

respectively.

Now we give examples of \(U_1,\,U_2, \,U_3\) which satisfy \((C1), (C2)\), and \((C3)\), respectively, but not (50), (51), and (52), respectively.

Example

4 Let \(G=\mathbb{Z}\) be the group of integers. Enumerating

$$\mathbb{Z}\times \mathbb{Z} =\{(a_1, b_1), (a_2, b_2), \ldots, (a_n, b_n), \ldots\}$$

such that

$$|a_1|+|b_1| \le |a_2| +|b_2| \le \cdots \le |a_n|+|b_n|\le \cdots,$$

and let \(P_n =\{(0,0), (a_n,0),(0, b_n), (a_n, b_n)\},\,\,n=1, 2, \ldots\). Then it is easy to see that \(U_1=\bigcup_{n=1}^\infty(P_n+(-2^n, 2^n))\) satisfies the condition \((C1)\). Now let \(P=\{(x_1, y_1),(x_2,y_2)\}\subset\mathbb{Z}\times \mathbb{Z}\) with \(x_2> x_1,\, y_2>y_1, \,(x_1+y_1)(x_2+y_2)>0\). Then \(P+(-z, z)\) is not contained in U 1 for all \(z\in \mathbb{Z}\). Indeed, let \((a, b) \in P_n +(-2^n, 2^n), (c, d)\in P_{n+1} +(-2^{n+1}, 2^{n+1})\). Then we have \(a>c,\, b<d\) for all \(n=1, 2, \ldots\). Thus it follows from \(x_2> x_1,\, y_2>y_1\) that if \(P+(-z, z)\subset U_1\), then \(P+(-z, z)\subset P_n +(-2^n, 2^n)\) for some \(n\in \mathbb{N}\), which implies that the line segment joining the points of \(P+(-z, z)\) intersects the line y = -x in \(\mathbb{R}^2\), contradicting to the condition \((x_1+y_1)(x_2+y_2)>0\). Similarly, let \(Q_n =\{(b_n,0), (0, b_n),(a_n, b_n), (a_n +b_n, 0)\}\) and \(R_n =\{(a_n,0), (0, a_n),(a_n, b_n), (0, a_n +b_n)\},\,\,n=1, 2, \ldots\). Then it is easy to see that \(U_2=\bigcup_{n=1}^\infty(Q_n+(0, 2^n))\) satisfies the condition \((C2)\) but not (2.7) and \(U_3=\bigcup_{n=1}^\infty(R_n+(2^n,0))\) satisfies the condition \((C3)\) but not (52).

As in Sect. 2, we denote

$$V^*=\{v_1+v_2-v_3-v_4: v_j\in V,\, j=1, 2, 3,4\}.$$

Theorem 4

Let W satisfy the condition \((C1)\) . Suppose that \(f, g, h:G \to H\) satisfy

$$f(x+y)-g(x)-h(y)\in V$$
(53)

for all \((x,y)\in W\) . Then there exists a \(V^*\) -Cauchy sequence \(a_n:G \to H\) satisfying

$$a_n(x+y)-a_n(x)-a_n(y)\in 2^{-n}V^*$$
(54)

for all \(n\in\mathbb{N}\) and \(x, y\in G\) and

$$a_n(x)-f(x)+f(0)\in 2^{-n}\big(V^*+2~V^*+\ldots +2^{n-1}V^*\big)$$
(55)

for all \(x\in G\) .

Proof

For given \(x, y\in G\), choose \(z\in G\) such that \((-z, z)+P_{x, y}\subset W\). Then we have

$$\begin{aligned} f(x+y)-g(x-z)-h(z+y)&\in V, \nonumber\\ -f(x)+g(x-z)+h(z)&\in -V, \nonumber\\ -f(y)+g(-z)+h(z+y)&\in -V, \nonumber\\ + f(0)-g(-z)-h(z)&\in V.\nonumber\end{aligned}$$

Thus it follows that

$$f(x+y)-f(x)-f(y)+f(0)\in V+(-V)+(-V)+V=V^*$$
(56)

for all \(x, y\in G\).

Now by Theorem 2, there exists a \(V^*\)-Cauchy sequence \(a_n:G \to H\) satisfying (54) and (55). This completes the proof.

In particular, let \(V=\{x\in Y: \|x\|\le \epsilon\}\). Then we have

$$V^*\subset \{x\in Y:\|x\|\le 4\epsilon\}, 2^{-n}\big(V^*+2~V^*+\ldots +2^{n-1}V^*\big)\subset \{x\in Y:\|x\|\le 4\epsilon\}$$

for all \(n\in \mathbb{N}\), and

$$2^{-m-n}\big(V^*+2~V^*+\ldots +2^{n-1}V^*\big)\subset \{x\in Y:\|x\|\le 2^{-m+2}\epsilon\}$$

for all \(m, n\in \mathbb{N}\). Thus in this case, every \(V^*\)-Cauchy sequence is a Cauchy sequence. Now as a direct consequence of Theorem 4 we have the following.

Corollary 5

Let W satisfy the condition \((C1)\) and \(\epsilon\ge0\) . Suppose that \(f, g, h:G \to Y\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(57)

for all \((x,y)\in W\) . Then there exists a Cauchy sequence \(a_n:G \to Y\) satisfying

$$\|a_{n}(x+y)-a_n(x)-a_n(y)\|\le 2^{-n+2}\epsilon$$
(58)

for all \(n\in\mathbb{N}\) and \(x, y\in G\) , and

$$\|a_n(x)-f(x)+f(0)\|\le 4\epsilon$$
(59)

for all \(x\in G\) .

As a direct consequence of Corollary 5 we have the following.

Corollary 6

Let W satisfy the condition \((C1)\) and \(\epsilon\ge0\) . Suppose that \(f, g, h:G \to B\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(60)

for all \((x,y)\in W\) . Then there exists an additive function \(A_1:G \to B\) and

$$\|f(x)-A_1(x)-f(0)\|\le 4\epsilon$$
(61)

for all \(x\in G\) .

Theorem 5

Let W satisfy the condition \((C2)\) . Suppose that \(f, g, h:G \to H\) satisfy

$$f(x+y)-g(x)-h(y)\in V$$
(62)

for all \((x,y)\in W\) . Then there exists a \(V^*\) -Cauchy sequence \(b_n:G \to H\) satisfying

$$b_n(x+y)-b_n(x)-b_n(y)\in 2^{-n}V^*$$
(63)

for all \(n\in\mathbb{N}\) and \(x, y\in G\) , and

$$b_n(x)-g(x)+g(0)\in 2^{-n}\big(V^*+2~V^*+\ldots +2^{n-1}V^*\big)$$
(64)

for all \(x\in G\) .

Proof

For given \(x, y\in G\), choose \(z\in G\) such that \((0, z)+Q_{x, y}\subset W\). Then we have

$$\begin{aligned} -f(x+y+z)+g(x+y)+h(z)&\in -V, \nonumber\\ f(x+y+z)-g(x)-h(y+z)&\in V, \nonumber\\ f(y+z)-g(y)-h(z)&\in V, \nonumber\\ - f(y+z)+g(0)+h(y+z)&\in -V.\nonumber\end{aligned}$$

Thus it follows that

$$g(x+y)-g(x)-g(y)+g(0)\in -V+V+V-V=V^*$$
(65)

for all \(x, y\in G\). Now by Theorem 2, there exists a sequence \(b_n:G \to H\) satisfying (63) and (64). This completes the proof.

In particular, if \(f, g, h:G \to Y\) we have the following.

Corollary 7

Let W satisfy the condition \((C2)\) and \(\epsilon\ge0\) . Suppose that \(f, g, h:G \to Y\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(66)

for all \((x,y)\in W\) . Then there exists a Cauchy sequence \(b_n:G \to Y\) satisfying

$$\|b_{n}(x+y)-b_n(x)-b_n(y)\|\le 2^{-n+2}\epsilon$$
(67)

for all \(n\in\mathbb{N}\) and \(x, y\in G\) , and

$$\|b_n(x)-g(x)+g(0)\|\le 4\epsilon$$
(68)

for all \(x\in G\) .

In particular, if \(f, g, h:G \to B\) we have the following.

Corollary 8

Let W satisfy the condition \((C2)\) and \(\epsilon\ge0\) . Suppose that \(f, g, h:G \to B\) satisfy

$$\left\|f(x+y)-g(x)-h(y)\right\|\le \epsilon$$
(69)

for all \((x,y)\in W\) . Then there exists a unique additive function \(A_2:G \to B\) such that

$$\|g(x)-A_2(x)-g(0)\|\le 4\epsilon$$
(70)

for all \(x\in G\) .

Theorem 6

Let W satisfy the condition \((C3)\) . Suppose that \(f, g, h:G \to H\) satisfy

$$f(x+y)-g(x)-h(y)\in V$$
(71)

for all \((x,y)\in W\) . Then there exists a \(V^*\) -Cauchy sequence \(c_n:G \to H\) satisfying

$$c_n(x+y)-c_n(x)-c_n(y)\in 2^{-n}V^*$$
(72)

for all \(n\in\mathbb{N}\) and \(x, y\in G\) and

$$c_n(x)-h(x)+h(0)\in 2^{-n}\big(V^*+2~V^*+\ldots +2^{n-1}V^*\big)$$
(73)

for all \(x\in G\) .

Proof

For given \(x, y\in G\), choose \(z\in G\) such that \((0, z)+Q_{x, y}\subset W\). Then we have

$$\begin{aligned} -f(z+x+y)+g(z)+h(x+y)&\in -V, \nonumber\\ f(z+x+y)-g(z+x)-h(y)&\in V, \nonumber\\ f(z+x)-g(z)-h(x)&\in V, \nonumber\\ - f(z+x)+g(z+x)+h(0)&\in -V.\nonumber\end{aligned}$$

Thus it follows that

$$h(x+y)-h(x)-h(y)+h(0)\in -V+V+V-V=V^*$$
(74)

for all \(x, y\in G\). Now by Theorem 2, there exists a sequence \(c_n:G \to H\) satisfying (72) and (73). This completes the proof.

In particular, if \(f, g, h:G \to Y\) we have the following.

Corollary 9

Let W satisfy the condition \((C3)\) and \(\epsilon\ge0\) . Suppose that \(f, g, h:G \to Y\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(75)

for all \((x,y)\in W\) . Then there exists a Cauchy sequence \(c_n:G \to Y\) satisfying

$$\|c_{n}(x+y)-c_n(x)-c_n(y)\|\le 2^{-n+2}\epsilon$$
(76)

for all \(n\in\mathbb{N}\) and \(x, y\in G\) , and

$$\|c_n(x)-h(x)+h(0)\|\le 4\epsilon$$
(77)

for all \(x\in G\) .

In particular, if \(f, g, h:G \to B\) we have the following.

Corollary 10

Let W satisfy the condition \((C3)\) and \(\epsilon\ge0\) . Suppose that \(f, g, h:G \to B\) satisfy

$$\left\|f(x+y)-g(x)-h(y)\right\|\le \epsilon$$
(78)

for all \((x,y)\in W\) . Then there exists a unique additive function \(A_3:G \to B\) such that

$$\|h(x)-A_3(x)-h(0)\|\le 4\epsilon$$
(79)

for all \(x\in G\) .

Theorem 7

Let W satisfy all the conditions \((C1),\, (C2),\) and \((C3)\) . Suppose that \(f, g, h:G \to H\) satisfy

$$f(x+y)-g(x)-h(y)\in V$$
(80)

for all \((x,y)\in W\) . Then there exist \(V^*\) -Cauchy sequences \(a_n, b_n, c_n:G \to H\) satisfying

$$\begin{aligned} a_{n}(x+y)-a_n(x)-a_n(y)&\in 2^{-n}V^{*}\end{aligned}$$
(81)
$$\begin{aligned} b_{n}(x+y)-b_n(x)-b_n(y)&\in2^{-n}V^{*}\end{aligned}$$
(82)
$$\begin{aligned} c_{n}(x+y)-c_n(x)-c_n(y)&\in2^{-n}V^{*}\end{aligned}$$
(83)

for all \(n\in \mathbb{N}\) and \(x, y\in G\) , and

$$\begin{aligned} a_n(x)-f(x)+f(0)& \in V_n^{*},\end{aligned}$$
(84)
$$\begin{aligned} b_n(x)-g(x)+g(0)&\in V_n^{*}, \end{aligned}$$
(85)
$$\begin{aligned} c_n(x)-h(x)+h(0)&\in V_n^{*}\end{aligned}$$
(86)

for all \(n\in \mathbb{N}\) and \(x\in G\) , and

$$a_n(x+y)-b_n(x)-c_n(y)\in V_n^{**}.$$
(87)

for all \(n\in \mathbb{N}\) and \(x, y\in G\) , where

$$V_n^{*}= 2^{-n}(V^*+2~V^*+\ldots +2^{n-1}V^*),$$
$$V_n^{**}= V+V+V+V+V-V-V-V-V-V+V_n^{*}-V_n^{*}-V_n^{*}.$$

Proof

From Theorems 4, 5, and 6, it remains to show (87). By the condition \((C1)\), for given \(x, y\in G\), choose \(z\in G\) such that \((-z, z), (x-z,z+y) \in W\). Then from (80) we have

$$\begin{aligned} f(x+y)-g(x-z)-h(z+y)&\in V,\end{aligned}$$
(88)
$$\begin{aligned} -f(0)+g(-z)+h(z)&\in -V.\end{aligned}$$
(89)

Also, by (65) and (74) we have

$$\begin{aligned} g(x-z)-g(x)-g(-z)+g(0)&\in V+V-V-V,\end{aligned}$$
(90)
$$\begin{aligned} h(z+y)-h(z)-h(y)+h(0)&\in V+V-V-V.\end{aligned}$$
(91)

for all \(x, y, z\in G\). From (88)–(91), we have

$$f(x+y)-g(x)-h(y)-f(0)+g(0)+h(0)\in V+V+V+V+V-V-V-V-V-V$$
(92)

for all \(x,y \in G\). Using (84), (85), (86), and (92) we have

$$a_n(x+y)-b_n(x)-c_n(y)\in V+V+V+V+V-V-V-V-V-V+V_n^{*}-V_n^{*}-V_n^{*}.$$
(93)

This completes the proof.

In particular, let \(V=\{x\in Y: \|x\|\le \epsilon\}\). Then we have

$$V_n^*\subset \{x\in Y:\|x\|\le 4\epsilon\}, V_n^{**}\subset \{x\in Y:\|x\|\le 22\epsilon\}$$

for all \(n\in \mathbb{N}\). Thus as a direct consequence of Theorem 7 we have the following.

Corollary 11

Let W satisfy the conditions \((C1),\, (C2),\) and \((C3)\) and \(\epsilon\ge0\) . Suppose that \(f, g, h:G \to Y\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(94)

for all \((x,y)\in W\) . Then there exist Cauchy sequences \(a_n, b_n, c_n:G \to Y\) satisfying

$$\begin{aligned} \|a_{n}(x+y)-a_n(x)-a_n(y)\|&\le2^{-n+2}\epsilon\end{aligned}$$
(95)
$$\begin{aligned} \|b_{n}(x+y)-b_n(x)-b_n(y)\|&\le2^{-n+2}\epsilon\end{aligned}$$
(96)
$$\begin{aligned} \|c_{n}(x+y)-c_n(x)-c_n(y)\|&\le2^{-n+2}\epsilon\end{aligned}$$
(97)

for all \(n\in\mathbb{N}\) and \(x, y\in G\) ,

$$\begin{aligned} \|f(x)-a_n(x)-f(0)\|& \le 4\epsilon,\end{aligned}$$
(98)
$$\begin{aligned} \|g(x)-b_n(x)-g(0)\|&\le 4\epsilon,\end{aligned}$$
(99)
$$\begin{aligned} \|h(x)-c_n(x)-h(0)\|&\le 4\epsilon\end{aligned}$$
(100)

for all \(n\in\mathbb{N}\) and \(x\in G\) , and

$$\|a_n(x+y)-b_n(x)-c_n(y)\|\le 22\epsilon$$
(101)

for all \(n\in \mathbb{N}\) and \(x, y\in G\) .

Corollary 12

Let W satisfy the conditions \((C1),\, (C2),\) and \((C3)\) and \(\epsilon\ge0\) . Suppose that \(f, g, h:G \to B\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(102)

for all \((x,y)\in W\) . Then there exists an additive function \(A:G \to B\) such that

$$\begin{aligned} \|f(x)-A(x)-f(0)\|& \le 4\epsilon,\nonumber\\ \|g(x)-A(x)-g(0)\|&\le 4\epsilon, \nonumber\\ \|h(x)-A(x)-h(0)\|&\le 4\epsilon\nonumber\end{aligned}$$

for all \(x\in G\) .

Proof

Let \(A_1(x)=\lim_{n\to \infty}a_n(x),\,\,A_2(x)=\lim_{n\to \infty}b_n(x),\,\,A_3(x)=\lim_{n\to \infty}c_n(x)\). Then it follows from (95)–(97) that for each \(j=1, 2, 3\), A j is additive. Letting \(n \to \infty\) in (98)–(100) we have

$$\begin{aligned} \|f(x)-A_1(x)-f(0)\|& \le 4\epsilon,\nonumber\\ \|g(x)-A_2(x)-g(0)\|&\le 4\epsilon, \nonumber\\ \|h(x)-A_3(x)-h(0)\|&\le 4\epsilon\nonumber\end{aligned}$$

for all \(x\in G\). Finally letting \(n\to \infty\) in (101) we have

$$\|A_1(x+y)-A_2(x)-A_3(y)\|\le 22\epsilon$$
(103)

for all \(x,y\in G\). Putting y = 0 and x = 0 in (103) separately, we have

$$\begin{aligned} \|A_1(x)-A_2(x)\|&\le 22\epsilon\nonumber\\ \|A_1(y)-A_3(y)\|&\le22\epsilon\nonumber\end{aligned}$$

for all \(x,y \in G\), which implies that \(A_1=A_2\) and \(A_1=A_3\). This completes the proof.

In particular, if G is a normed vector space we have the following.

Corollary 13

Let \(d>0\) . Suppose that \(f, g, h:G \to B\) satisfy

$$\|f(x+y)-g(x)-h(y)\|\le \epsilon$$
(104)

for all \(\|x\|+\|y\|\ge d\) . Then there exists an additive function \(A:G \to B\) such that

$$\begin{aligned} \|f(x)-A(x)-f(0)\|& \le 4\epsilon,\nonumber\\ \|g(x)-A(x)-g(0)\|&\le 4\epsilon, \nonumber\\ \|h(x)-A(x)-h(0)\|&\le 4\epsilon\nonumber\end{aligned}$$

for all \(x\in G\) .

Finally we give another interesting example of the set \(W\subset \mathbb{R}^n \times \mathbb{R}^n\) with finite Lebesgue measure satisfying all the conditions \((C1)\).

Lemma 1

Let \(D:=\{(x_1, y_1), (x_2, y_2), (x_3, y_3), \ldots\}\) be a countable dense subset of \(\mathbb{R}^2\) . For each \(j=1, 2, 3, \ldots,\) we denote by

$$R_j=\{(x, y)\in \mathbb{R}^2: |x-x_j| <1,\, |y-y_j|<2^{-j}\epsilon\}$$

the rectangle in \(\mathbb{R}^2\) with center \((x_j, y_j)\) and let \(W=\bigcup_{j=1}^\infty R_j\) . It is easy to see that the Lebesgue measure m ( W ) of U satisfies \(m(W)\le \epsilon\) . Now for \(d>0\) , let

$$W_d=W\cap\{(x, y)\in \mathbb{R}^2: |x|+|y|>d\}.$$

Then W d satisfies \((C1)\).

Proof

For given \(x, y\in \mathbb{R}\) we choose a \(p\in \mathbb{R}\) such that

$$\begin{aligned} &|p|\ge d+|x|+|y|+1.\end{aligned}$$
(105)

We first choose \((x_{i_1}, y_{i_1})\in K\) such that

$$\begin{aligned} &|-p-x_{i_1}|+|p-y_{i_1}|<\frac 1 4,\end{aligned}$$
(106)

and then we choose \((x_{i_2}, y_{i_2})\in K\), \((x_{i_3}, y_{i_3})\in K\) and \((x_{i_4}, y_{i_4})\in K\) with \(1<i_1<i_2<i_3<i_4\), step by step, satisfying

$$\begin{aligned} &|x-y_{i_1}-x_{i_2}|+ |y_{i_1}-y_{i_2}|<2^{-i_1-1},\end{aligned}$$
(107)
$$\begin{aligned} &|x-y_{i_2}-x_{i_3}|+|y+y_{i_2}-y_{i_3}|<2^{-i_2 -1},\end{aligned}$$
(108)
$$\begin{aligned} &|y-y_{i_3}-x_{i_4}|+|y_{i_3}-y_{i_4}|<2^{-i_3 -1}.\end{aligned}$$
(109)

Let

$$\begin{aligned} z_1 &= y_{i_1}-p,\nonumber\\ z_2 &= y_{i_2}-y_{i_1},\nonumber\\ z_3 &= y_{i_3}-y_{i_2}-y,\nonumber\\ z_4 &=y_{i_4}-y_{i_3},\nonumber\end{aligned}$$

and

$$z =z_1 + z_2 + z_3 + z_4.$$

Then from (106)–(109) we have

$$\begin{aligned} |z_1|< \frac 1 4, \,\,|z_2|<2^{-i_1-1},\,\,|z_3|<2^{-i_2 -1},\,\,|z_4|<2^{-i_3 -1},\,\, |z|< \frac 1 2.\end{aligned}$$
(110)

Thus from (105), (106), and (110) we have

$$\begin{aligned} |-p-z|+|p+z|& \ge 2(|p|- |z|)\ge 2\big(|p|-\frac 1 2\big)\end{aligned}$$
(111)
$$\begin{aligned} &> 2d \ge d,\nonumber\end{aligned}$$
$$\begin{aligned} |-p-z-x_{i_1}|&\le|-p-x_{i_1}|+|z|\end{aligned}$$
(112)
$$\begin{aligned} &<\frac 1 4+\frac 1 2<1,\nonumber\end{aligned}$$

and

$$\begin{aligned} |p+z-y_{i_1}|=|z_2+z_3+z_4| &<2^{-i_1-1}+2^{-i_2 -1}+2^{-i_3 -1}<2^{-i_1}.\end{aligned}$$
(113)

The inequalities (111), (112), and (113) imply

$$(-p-z, p+z)\in W_d.$$
(114)

Also from the inequalities

$$\begin{aligned} |x-p-z|+|p+z|& \ge 2\big(|p|-|x|-|z|\big)>2\big(|p|-|x|-\frac 1 2 \big)> d,\nonumber\end{aligned}$$
$$\begin{aligned} |x-p-z-x_{i_2}|&\le|x-y_{i_1}-x_{i_2}|+|z_2|+|z_3|+|z_4|\nonumber\\ &<\frac 1 8+\frac 1 8+\frac 1 16+\frac 1 32<1,\nonumber\end{aligned}$$

and

$$\begin{aligned} |p+z-y_{i_2}|=|z_3 +z_4|&<2^{-i_2 -1}+2^{-i_3 -1}<2^{-i_2},\nonumber\end{aligned}$$

we have

$$(x-p-z, p+z)\in W_d.$$
(115)

Similarly, using the inequalities

$$\begin{aligned} |x-p-z-x_{i_3}|&\le|x-y_{i_2}-x_{i_3}|+|z_3|+|z_4|<1,\nonumber\\ |y+p+z-y_{i_3}|&=|z_4|<2^{-i_3},\nonumber\\ |-p-z-x_{i_4}|&\le |y-y_{i_3}-x_{i_4}|+ |z_4|<1,\nonumber\\ |y+p+z-y_{i_4}|&=0,\nonumber\end{aligned}$$

we have

$$(x-p-z, y+p+z),\, (-p-z, y+p+z) \in W_d.$$
(116)

Let \(\{(x_1, y_1), (x_2, y_2), (x_3, y_3), \ldots\}\) be defined as above. For each \(j=1, 2, 3, \ldots\), let

$$S_j=\{(x, y):x, y\in \mathbb{R}:\, |x+y-x_j-y_j| <1,\, |x-y-x_j+y_j|<2^{-j}\epsilon\}$$

and let \(V=\bigcup_{j=1}^\infty S_j\). Then V satisfies \(m(V)\le \epsilon\). For fixed \(d>0\), let

$$V_d=V\cap\{(x, y)\in \mathbb{R}^2: |x|+|y|>d\}.$$

Using the similar method as in the proof of Lemma 1 we can show that V d satisfies the conditions \((C1), (C2)\), and \((C3)\).

As a direct consequence of Lemma 1 we have the following.

Theorem 8

Let \(d>0\) . Suppose that \(f:\mathbb{R} \to \mathbb{R}\) satisfies

$$\left|f(x+y)-f(x)-f(y)\right|\le \epsilon$$
(117)

for all \((x,y)\in W_d\) . Then there exists a unique additive function \(A:\mathbb{R} \to \mathbb{R}\) such that

$$\begin{aligned} |f(x)-A(x)|&\le 3\epsilon\end{aligned}$$
(118)

for all \(x\in \mathbb{R}\) .

Proof

It follows from (115) and (116) that for given \(x, y\in \mathbb{R}\) there exist \(p, z\in \mathbb{R}\) satisfying

$$\begin{aligned} |f(x+y)-f(x)-f(y)| &\le|-f(x)+f(x-p-z)+f(p+z)| \nonumber\\ &+ |f(x+y)-f(x-p-z)-f(y+p+z)| \nonumber\\ &+ |-f(y)+f(-p-z)+f(y+p+z)|\nonumber\\ &\le 3\epsilon.\nonumber\end{aligned}$$

Using Theorem A we get the result.

As a consequence of Theorem 8 we obtain an asymptotic behavior of

$$C_d(f):=\sup_{(x, y)\in W_d}|f(x+y)- f(x)-f(y)|\to 0$$
(119)

as \(d \to \infty\).

Theorem 9

Suppose that \(f:\mathbb{R} \to \mathbb{R}\) satisfies the condition

$$C_d(f)\to 0$$
(120)

as \(d\to \infty\) . Then f is an additive function.

Proof

By the condition (120), for each \(j\in \mathbb{N}\), there exists \(d_j>0\) such that

$$|f(x+y)- f(x)-f(y)|\le \frac 1 j$$

for all \((x, y)\in W_{d_j}\). By Theorem 8, there exists a unique additive function \(A_j:\mathbb{R} \to \mathbb{R}\) such that

$$\begin{aligned} |f(x)-A_j(x)|\le \frac 3 j\end{aligned}$$
(121)

for all \(x\in \mathbb{R}\). From (121), using the triangle inequality we have

$$\begin{aligned} |A_j(x)-A_k(x)|\le \frac 3 j +\frac 3 k\le 6\end{aligned}$$
(122)

for all \(x\in \mathbb{R}\) and all positive integers \(j, k\). Now, the inequality (122) implies \(A_j=A_k\). Indeed, for all \(x\in \mathbb{R}\) and all rational numbers \(r>0\) we have

$$\begin{aligned} |A_j(x)-A_k(x)|=\frac 1 r|A_j(rx)-A_k(rx)|\le \frac 6 r.\end{aligned}$$
(123)

Letting \(r\to \infty\) in (123) we have \(A_j=A_k\). Thus, letting \(j\to \infty\) in (121) we get the result.