1 Introduction and preliminaries

In 1940, Ulam [23] asked the first question on the stability problem. In 1941, Hyers [9] solved the problem of Ulam. This result was generalized by Aoki [1] for additive mappings and by Rassias [20] for linear mappings by considering an unbounded Cauchy difference. In 1994, a further generalization was obtained by Găvruta [8]. Rassias [16,17,18,19] generalized Hyers result. During the last two decades, a number of papers and research monographs have been published on various generalizations and applications of the generalized Hyers-Ulam stability to a number of functional equations and mappings (see [2, 5,6,7, 11, 15, 16, 21]). We also refer the readers to the books: Czerwik [3] and Hyers, Isac and Rassias [10].

The theory of modulars on linear spaces and the corresponding theory of modular linear spaces were founded by Nakano [12] and were intensively developed by Amemiya, Koshi, Shimogaki, Yamamuro [14, 25] and others.

Definition 1.1

Let X be a vector space over a field K (\(\mathbb {R}\) or \(\mathbb {C}\)). A generalized functional \(\rho : X\longrightarrow [0,\infty ]\) is called a modular if for arbitrary \(x, y \in X,\) \(\rho \) satisfies:

  1. (a)

    \(\rho (x) = 0\) if and only if \( x = 0\),

  2. (b)

    \( \rho (a x) = \rho (x)\) for every scalar a with \(|a| = 1,\)

  3. (c)

    \(\rho (a x+b y)\le \rho (x) + \rho (y)\), whenever \(a,b \ge 0\) and \(a+b =1\).

If we replace (c) by

(\(c'\)) \(\rho (a x+b y)\le a\rho (x) +b\rho (y)\), whenever \(a,b \ge 0\) and \(a+b =1,\) then the modular \(\rho \) is called convex. A modular \(\rho \) defines a corresponding modular space, i.e., the vector space \( X_{\rho }\) given by:

$$\begin{aligned} X_{\rho }=\{x\in X|\,\rho (\lambda x)\longrightarrow 0\, as \,\lambda \longrightarrow 0\}. \end{aligned}$$

Definition 1.2

Let \(\{x_{n}\}\) and x be in \( X_{\rho }\). Then

  1. (i)

    The sequence \(\{x_{n}\}\), with \(x_{n} \in X_{\rho }\), is \(\rho \)-convergent to x if \(\rho (x_{n}-x)\rightarrow 0\) as \(n\rightarrow \infty \).

  2. (ii)

    The sequence \(\{x_{n}\}\), with \(x_{n} \in X_{\rho }\), is called \(\rho \)-Cauchy if \(\rho (x_{n}-x_{m})\rightarrow 0\) as \(n,m\rightarrow \infty \).

  3. (iii)

    A subset S of \( X_{\rho }\) is called \(\rho \)-complete if and only if any \(\rho \)-Cauchy sequence is \(\rho \)-convergent to an element of S.

Fatou property The modular \(\rho \) has the Fatou property if and only if \(\rho (x)\le \displaystyle \liminf _{n\rightarrow \infty }\rho (x_{n})\) whenever the sequence \(\{x_{n}\}\) is \(\rho \)-convergent to x. A function modular is said to satisfy the \(\Delta _{\alpha }\)-condition \((\alpha \in \mathbb {N}\),\(\alpha \ge 2)\) if there exists \(\kappa >0\) such that \(\rho (\alpha x)\le \kappa \rho (x),\) for all \(x \in X_{\rho }\).

Remark

\(\Delta _{\alpha }\)-condition implies \(\Delta _{2}\)-condition.

Definition 1.3

Let \( X_{\rho }\) be a modular space and C be a nonempty subset of \( X_{\rho }\). The self-map \(T:C\rightarrow C\) is said to be quasicontraction if there exists \(k < 1\) such that

$$\begin{aligned} \rho \big (Tx-Ty\big )\le k\max \Big \{\rho (x-y),\rho (x-Ty), \rho (y-Tx),\rho (x-Tx),\rho (y-Ty)\Big \}, \end{aligned}$$

for any \(x, y\in C.\)

Definition 1.4

Given a modular space \( X_{\rho }\), a nonempty subset \(C\subseteq X_{\rho }\), and a mapping \(T:C\rightarrow C\), the orbit of T around a point x is the set

$$\begin{aligned} O(T) := \{x, Tx, T^{2}x,\ldots \}, \end{aligned}$$

the quantity

$$\begin{aligned} \delta _{\rho }(T) := \sup \{\rho (u-v) | u,v\in O(T)\}, \end{aligned}$$

is then associated to T and is called the orbital diameter of T at x. In particular, if \(\delta _{\rho }(T)<\infty ,\) one says that T has a bounded orbit at x.

Theorem 1.5

([13]) Let \( X_{\rho }\) be a modular space such that \(\rho \) satisfies the Fatou property and let \(C\subseteq X_{\rho }\) be a \(\rho \)-complete subset. If \(T:C\rightarrow C\) is a quasicontraction and T has a bounded orbit at \(x_{0}\), then the sequence \(\{T^{n}x_{0}\}\) is \(\rho \)-convergent to a point \(\omega \in C.\)

Stability of quadratic and generalized Jensen functional equation in modular spaces have been investigated in [22] and [24].

In this paper, we investigate the generalized Hyers–Ulam stability of the \(\alpha \)- cubic functional equation

$$\begin{aligned}&f(\alpha x+y)+f(\alpha x-y)+f(x+\alpha y)-f(x-\alpha y)\nonumber \\&\quad =2\alpha f(x+y)+2\alpha (\alpha ^2-1)[f(x)+f(y)], \end{aligned}$$
(1.1)

with \(\alpha \in {\mathbb {N}}, \alpha \ne 1\) via the extensive studies of fixed point theory in modular spaces.

2 Stability of \(\alpha \)-cubic functional equation (1.1)

Throughout this section, we assume that \(\rho \) is a convex modular on \(\rho \)-complete modular space \(X_{\rho }\) with the Fatou property such that satisfies the \(\Delta _{\alpha }\)-condition with \(0<\kappa \le \alpha \). In addition, let V be a linear space. For convenience, we use the following abbreviation for a given function \(f:V\longrightarrow X_{\rho }\):

$$\begin{aligned} D_{\alpha }f(x,y):= & {} f(\alpha x+y)+f(\alpha x-y)+f(x+\alpha y)-f(x-\alpha y)\\&-2\alpha f(x+y)-2\alpha (\alpha ^2-1)[f(x)+f(y)] \end{aligned}$$

with \(\alpha \in {\mathbb {N}}, \alpha \ne 1\) and for all \(x,y\in V.\) We shall need the following lemmas:

Lemma 2.1

If a mapping \(f:X\rightarrow Y\) satisfies the functional equation

$$\begin{aligned} f(x+\alpha y)-f(x-\alpha y)=\alpha [f(x+y)-f(x-y)]+2\alpha (\alpha ^{2}-1)f(y), \end{aligned}$$
(2.1)

with \(\alpha \in {\mathbb {N}}, \alpha \ne 1\) and for all \(x, y\in X,\) then f is cubic.

Proof

Replacing (xy) with (0, 0) in (2.1), we get \(2\alpha (\alpha ^{2}-1)f(0)=0\) with \(\alpha \in {\mathbb {N}}, \alpha \ne 1\). Therefore \(f(0)=0\). Replacing (xy) with (0, x) and \((0,-x)\) in (2.1), we get, respectively, equations:

$$\begin{aligned} f(\alpha x)-f(-\alpha x)= & {} \alpha [(2\alpha ^{2}-1)f(x)-f(-x)],\nonumber \\ f(-\alpha x)-f(\alpha x)= & {} \alpha [(2\alpha ^{2}-1)f(-x)-f(x)]. \end{aligned}$$
(2.2)

By adding these two equations, one can obtain \(f(-x)=-f(x)\). By using (2.2) and \(f(-x)=-f(x)\), we get \(f(\alpha x)=\alpha ^{3}f(x)\) with \(\alpha \in {\mathbb {N}}, \alpha \ne 1\) and for all \(x\in X\) (See [4]). \(\square \)

Lemma 2.2

If a mapping \(f:X\rightarrow Y\) satisfies (1.1) for all \(x, y\in X,\) then f is cubic.

Proof

Replacing (xy) with (0, 0) in (1.1), we get \(f(0)=0\). Replacing (xy) with (x, 0) in (1.1), we get,

$$\begin{aligned} f(\alpha x)=\alpha ^{3} f(x), \end{aligned}$$
(2.3)

for all \(x\in X.\) By setting \(x=0\) and using (2.3), we get \(f(-y)=-f(y)\) for all \(y\in X,\) that is f is odd. Replacing (xy) with \((x,-y)\) in (1.1) and using oddness of f, we get,

$$\begin{aligned}&f(\alpha x-y)+f(\alpha x+y)+f(x-\alpha y)-f(x+\alpha y)\nonumber \\&\quad =2\alpha f(x-y)+2\alpha (\alpha ^2-1)[f(x)-f(y)] \end{aligned}$$
(2.4)

for all \(x,y\in X.\) It follows from (1.1) and (2.4) that

$$\begin{aligned} f(x+\alpha y)-f(x-\alpha y)=\alpha [f(x+y)-f(x-y)]+2\alpha (\alpha ^2-1)f(y), \end{aligned}$$
(2.5)

for all \(x,y\in X.\) It follows from Lemma 2.1 that f is cubic. \(\square \)

Theorem 2.3

Let \(\varphi : V^{2}\longrightarrow [0, +\infty )\) be a function such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{1}{\alpha ^{3n}}\varphi (\alpha ^{n}x,\alpha ^{n}y)=0, \end{aligned}$$
(2.6)

and

$$\begin{aligned} \varphi (\alpha x,\alpha y)\le \alpha ^{3} L\varphi (x,y), \end{aligned}$$
(2.7)

for all \(x,y\in V\) with \(L<1.\) Suppose that \(f:V\longrightarrow X_{\rho }\) satisfies the condition

$$\begin{aligned} \rho \big (D_{\alpha }f(x,y)\big )\le \varphi (x,y), \end{aligned}$$
(2.8)

for all \(x,y\in V\) and \(f(0)=0\). Then there exists a unique cubic mapping \(C_{\alpha }: V\longrightarrow X_{\rho }\) such that

$$\begin{aligned} \rho \big (C_{\alpha }(x)-f(x)\big )\le \frac{1}{\alpha ^{3}(1- L)}\varphi (x,0), \end{aligned}$$
(2.9)

for all \(x\in V\).

Proof

We consider the set

$$\begin{aligned} M= \{g:V\rightarrow X_{\rho }\} \end{aligned}$$

and define the function \(\overline{\rho }\) on M as follows,

$$\begin{aligned} \overline{\rho }(g)=:\inf \{c>0:\, \rho (g(x))\le c\varphi (x,0),\;\, \forall x\in V\} . \end{aligned}$$

We show that \(\overline{\rho }\) is a convex modular on M. It is also easy to verify that \(\overline{\rho }\) satisfies the axioms (a) and (b) of a modular. We will next show that \(\overline{\rho }\) is convex, and hence \((c')\) is satisfied. Let \(\epsilon >0\) be given. Then there exist real constants \(c_{1}>0\) and \(c_{2}>0\) such that

$$\begin{aligned} \overline{\rho }(g)\le c_{1}\le \overline{\rho }(g)+\epsilon ,\,\; \overline{\rho }(h)\le c_{2}\le \overline{\rho }(h)+\epsilon . \end{aligned}$$

Also

$$\begin{aligned} \rho (g(x))\le c_{1}\varphi (x,0),\,\;\rho (h(x))\le c_{2}\varphi (x,0). \end{aligned}$$

for all \(x\in V.\) If \(a+b=1\) and \(a,b\ge 0\), then we get

$$\begin{aligned} \rho (a g(x) +b h(x))\le & {} a\rho (g(x)) +b \rho (h(x)) \\\le & {} (c_{1}a+c_{2}b )\varphi (x,0), \end{aligned}$$

so we get

$$\begin{aligned} \overline{\rho }(a g+b h)\le a\overline{\rho }(g)+b \overline{\rho }(h)+(a+b)\epsilon , \end{aligned}$$

This concludes that \(\overline{\rho }\) is a convex modular on M. Now we show that \(M_{\overline{\rho }}\) is \(\overline{\rho }\)-complete. Let \(\{g_{n}\}\) be a \(\overline{\rho }\)-Cauchy sequence in \(M_{\overline{\rho }}\) and let \(\epsilon > 0\) be given. There exists a positive integer \(n_{0}\in \mathbb {N}\) such that

$$\begin{aligned} \overline{\rho }(g_{n}-g_{m})<\epsilon , \end{aligned}$$
(2.10)

for all \(n,m\ge n_{0}\). We have

$$\begin{aligned} \rho (g_{n}(x)-g_{m}(x))\le \epsilon \varphi (x,0) \end{aligned}$$
(2.11)

for all \(x\in V\) and \(n,m\ge n_{0}\). Therefore if x is any given point of V, \(\{g_{n}(x)\}\) is a \(\rho \)-Cauchy sequence in \(X_{\rho }\). Since \(X_{\rho }\) is \(\rho \)-complete, so \(\{g_{n}(x)\}\) is convergent in \(X_{\rho }\), for each \(x\in V \). Hence, we can define a function \(g:V\rightarrow X_{\rho }\) by:

$$\begin{aligned} g(x):=\lim _{n\rightarrow \infty }\displaystyle g_{n}(x), \end{aligned}$$
(2.12)

for all \(x\in V.\) Since \(\rho \) satisfies the Fatou property, it follows from (2.11) that

$$\begin{aligned} \rho (g_{n}(x)-g(x))\le \liminf _{m\rightarrow \infty }\displaystyle \rho (g_{n}(x)-g_{m}(x))\le \epsilon \varphi (x,0) , \end{aligned}$$
(2.13)

so

$$\begin{aligned} \overline{\rho }(g_{n}-g)\le \epsilon , \end{aligned}$$
(2.14)

for all \(n\ge n_{0}\). Thus, \(\{g_{n}\}\) is \(\overline{\rho }\)-converges, so that \(M_{\overline{\rho }}\) is \(\overline{\rho }\)-complete.

Now we show that \(\overline{\rho }\) satisfies the Fatou property. Suppose that \(\{g_{n}\}\) is a sequence in \(M_{\overline{\rho }}\) which is \(\overline{\rho }\)- convergent to an element \(g\in M_{\overline{\rho }}\). Let \( \epsilon > 0\) be given. For each \(n\in \mathbb {N}\), let \(c_{n}\) be a constant such that

$$\begin{aligned} \overline{\rho }(g_{n})\le c_{n}\le \overline{\rho }(g_{n})+\epsilon . \end{aligned}$$
(2.15)

so

$$\begin{aligned} \rho (g_{n}(x))\le c_{n}\varphi (x,0), \end{aligned}$$
(2.16)

for all \(x\in V\). Since \(\rho \) satisfies the Fatou property, we have

$$\begin{aligned} \rho (g(x))\le & {} \liminf _{n\rightarrow \infty }\displaystyle \rho (g_{n}(x))\\\le & {} \liminf _{n\rightarrow \infty }\displaystyle c_{n}\varphi (x,0)\\\le & {} \left[ \liminf _{n\rightarrow \infty } \overline{\rho }(g_{n})+\epsilon \right] \varphi (x,0) \end{aligned}$$

Thus, we have

$$\begin{aligned} \overline{\rho }(g)\le \liminf _{n\rightarrow \infty }\displaystyle \overline{\rho } (g_{n})+\epsilon . \end{aligned}$$

So \(\overline{\rho }\) satisfies the Fatou property. We consider the function \(\tau :M_{\overline{\rho }}\rightarrow M_{\overline{\rho }}\) defined by:

$$\begin{aligned} \tau g(x)=\frac{1}{\alpha ^{3}}g(\alpha x), \end{aligned}$$

for all \(x\in V\) and \(g\in M_{\overline{\rho }}\). Let \(g,h\in M_{\overline{\rho }}\) and let \(c \in [0,1]\) be an arbitrary constant with \(\overline{\rho }(g-h)<c\). From the definition of \(\overline{\rho }\), we have \(\rho (g(x)-h(x))\le c\varphi (x,0)\) for all \(x\in V\). By (2.7) and the last inequality, we get

$$\begin{aligned} \rho \big (\frac{g(\alpha x)}{\alpha ^{3}}-\frac{h(\alpha x)}{\alpha ^{3}}\big )\le & {} \frac{1}{\alpha ^{3}}\rho (g(\alpha x)-h(\alpha x))\\\le & {} \frac{1}{\alpha ^{3}}c\varphi (\alpha x,0)\\\le & {} cL\varphi (x,0), \end{aligned}$$

for all \(x\in V\). Hence, \(\overline{\rho }(\tau g-\tau h)\le L\overline{\rho }(g-h),\) for all \(g,h\in M_{\overline{\rho }}\), that is, \(\tau \) is a \(\overline{\rho }\)-contraction. Next, we show that \(\tau \) has a bounded orbit at f. Letting \(y =0\) in (2.8), we get

$$\begin{aligned} \rho \left( \frac{f(\alpha x)}{\alpha ^{3}}-f(x)\right)\le & {} \frac{1}{\alpha ^{3}}\varphi (x,0), \end{aligned}$$
(2.17)

for all \(x\in V\). Replacing x with \(\alpha x\) in (2.17), we get

$$\begin{aligned} \rho \left( \frac{f(\alpha ^{2} x)}{\alpha ^{3}}-f(\alpha x)\right)\le & {} \frac{1}{\alpha ^{3}}\varphi (\alpha x,0), \end{aligned}$$
(2.18)

By using (2.17) and (2.18), we get

$$\begin{aligned} \rho \left( \frac{f(\alpha ^{2} x)}{\alpha ^{6}}-f(x)\right)\le & {} \rho (\frac{f(\alpha ^{2}x)}{\alpha ^{6}}-\frac{f(\alpha x)}{\alpha ^{3}})+\rho (\frac{f(\alpha x)}{\alpha ^{3}}-f(x))\nonumber \\\le & {} \frac{1}{\alpha ^{6}}\varphi (\alpha x,0)+\frac{1}{\alpha ^{3}}\varphi (x,0), \qquad \qquad \end{aligned}$$
(2.19)

for all \(x\in V\). By induction, we can easily see that

$$\begin{aligned} \rho (\frac{f(\alpha ^{n}x)}{\alpha ^{3n}}-f(x))\le & {} \sum _{i=1}^{n}\frac{1}{\alpha ^{3i}}\varphi (\alpha ^{i-1}x,0) \nonumber \\\le & {} \frac{1}{L\alpha ^{3}}\varphi (x,0)\sum \limits _{i=1}^{n}L^{i}\nonumber \\\le & {} \frac{1}{\alpha ^{3}(1-L)}\varphi (x,0), \end{aligned}$$
(2.20)

for all \(x\in V\). It follows from inequality (2.20) that

$$\begin{aligned} \rho \left( \frac{f(\alpha ^{n}x)}{\alpha ^{3n}}-\frac{f(\alpha ^{k}x)}{\alpha ^{3k}}\right)\le & {} \frac{1}{2}\rho \left( 2\frac{f(\alpha ^{n}x)}{\alpha ^{3n}}-2f(x)\right) +\frac{1}{2}\rho \left( 2\frac{f(\alpha ^{k}x)}{\alpha ^{3k}}-2f(x)\right) \\\le & {} \frac{\kappa }{2}\rho \left( \frac{f(\alpha ^{n}x)}{\alpha ^{3n}}-f(x)\right) +\frac{\kappa }{2}\rho \left( \frac{f(\alpha ^{k}x)}{\alpha ^{3k}}-f(x)\right) \\\le & {} \frac{\kappa }{\alpha ^{3}(1-L)}\varphi (x,0), \end{aligned}$$

for every \(x\in V\) and \(n,k \in \mathbb {N}\), By the definition of \(\overline{\rho }\), we conclude that

$$\begin{aligned} \overline{\rho }(\tau ^{n} f-\tau ^{k} f)\le \frac{\kappa }{\alpha ^{3}(1-L)}, \end{aligned}$$

which implies the boundedness of an orbit of \(\tau \) at f. It follows from Theorem 1.5 that, the sequence \(\{\tau ^{n}f\}\) \(\overline{\rho }\)-converges to \(C_{\alpha }\in M_{\overline{\rho }}\). Now, by the \(\overline{\rho }\)-contractivity of \(\tau \), we have

$$\begin{aligned} \overline{\rho }(\tau ^{n} f-\tau C_{\alpha })\le L\overline{\rho }(\tau ^{n-1} f-C_{\alpha }). \end{aligned}$$

Passing to the limit \(n\rightarrow \infty \) and applying the Fatou property of \(\overline{\rho }\), we obtain that

$$\begin{aligned} \overline{\rho }(\tau C_{\alpha }-C_{\alpha })\le & {} \liminf _{n\rightarrow \infty }\overline{\rho }(\tau C_{\alpha }-\tau ^{n} f)\\\le & {} L\liminf _{n\rightarrow \infty }\overline{\rho }(C_{\alpha }-\tau ^{n-1} f)=0. \end{aligned}$$

Therefore, \(C_{\alpha }\) is a fixed point of \(\tau \). Letting \(x = \alpha ^{n}x\) and \(y = \alpha ^{n}y\) in (2.8), we get

$$\begin{aligned} \rho \big (D_{\alpha }f(\alpha ^{n}x,\alpha ^{n}y)\big )\le \varphi (\alpha ^{n}x_,\alpha ^{n}y), \end{aligned}$$

for all \(x,y\in V\). Therefore

$$\begin{aligned} \rho \big (\frac{1}{\alpha ^{3n}}D_{\alpha }f(\alpha ^{n}x,\alpha ^{n}y)\big )\le \frac{1}{\alpha ^{3n}}\varphi (\alpha ^{n}x_,\alpha ^{n}y), \end{aligned}$$
(2.21)

Employing the limit \(n\rightarrow \infty \), we get

$$\begin{aligned} D_{\alpha }C_{\alpha }(x,y)=0, \end{aligned}$$

for all \(x,y\in V\). It follows from Lemma 2.2, that \(C_{\alpha }\) is cubic. By using (2.20), we get (2.9).

To prove the uniqueness of \(C_{\alpha },\) let \(C:V\rightarrow X_{\rho }\) be another cubic mapping satisfying (2.9). Then, C is a fixed point of \(\tau \).

$$\begin{aligned} \overline{\rho }\big (C_{\alpha }-C\big )= \overline{\rho }\big (\tau C_{\alpha }-\tau C\big )\le L \overline{\rho }\big (C_{\alpha }-C\big ), \end{aligned}$$

which implies that \(\overline{\rho }\big (C_{\alpha }-C\big )=0\) or \(C_{\alpha }=C\). This completes the proof.

Corollary 2.4

Let X be a Banach space, \(\varphi : V^{2}\longrightarrow [0, +\infty )\) be a function such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{1}{\alpha ^{3n}}\varphi (\alpha ^{n}x,\alpha ^{n}y)=0, \end{aligned}$$

and

$$\begin{aligned} \varphi (\alpha x,\alpha y)\le L\alpha ^{3}\varphi (x,y), \end{aligned}$$

for all \(x,y\in V\) with \(L<1.\) Suppose that \(f:V\longrightarrow X\) satisfies the following condition

$$\begin{aligned} \big \Vert D_{\alpha }f(x,y)\big \Vert \le \varphi (x,y), \end{aligned}$$

\(x,y\in V\) and \(f(x)=0.\) Then there exists a unique cubic mapping \(C_{\alpha }: V\longrightarrow X\) such that

$$\begin{aligned} \big \Vert C_{\alpha }(x)-f(x)\big \Vert \le \frac{1}{\alpha ^{3}(1-L)}\varphi (x,0), \end{aligned}$$

for all \(x\in V\).

Proof

It is known that every normed space is modular space with the modular \(\rho (x) = \Vert x\Vert \) and satisfies the \(\Delta _{\alpha }\)-condition with \( \kappa = \alpha \). \(\square \)

Theorem 2.5

Let \(\varphi : V^{2}\longrightarrow [0, +\infty )\) be a function such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\kappa ^{3n}\varphi \left( \frac{x}{\alpha ^{n}},\frac{y}{\alpha ^{n}}\right) =0, \end{aligned}$$
(2.22)

and

$$\begin{aligned} \varphi \left( \frac{x}{\alpha },\frac{y}{\alpha }\right) \le \frac{L}{\alpha ^{3}}\varphi (x,y), \end{aligned}$$
(2.23)

for all \(x,y\in V\) with \(L<1.\) Suppose that \(f:V\longrightarrow X_{\rho }\) satisfies the condition

$$\begin{aligned} \rho \big (D_{\alpha }f(x,y)\big )\le \varphi (x,y), \end{aligned}$$
(2.24)

for all \(x,y\in V\) and \(f(0)=0.\) Then there exists a unique mapping \(C_{\alpha }: V\longrightarrow X_{\rho }\) such that

$$\begin{aligned} \rho \big (C_{\alpha }(x)-f(x)\big )\le \frac{L}{\alpha ^{3}(1- L)}\varphi (x,0), \end{aligned}$$
(2.25)

for all \(x\in V\).

Proof

We consider the set

$$\begin{aligned} M=\{g:V\rightarrow X_{\rho }\} \end{aligned}$$

and define the function \(\overline{\rho }\) on M as follows,

$$\begin{aligned} \overline{\rho }(g)=:\inf \{c>0:\, \rho (g(x))\le c\varphi (x,0),\;\, \forall x\in V\} . \end{aligned}$$

Similar to the proof of Theorem 2.3, we have:

  1. 1.

    \(\overline{\rho }\) is a convex modular on M

  2. 2.

    \(M_{\overline{\rho }}\) is \(\overline{\rho }\)-complete.

  3. 3.

    \(\overline{\rho }\) satisfies the Fatou property.

Now, we consider the function \(\tau :M_{\overline{\rho }}\rightarrow M_{\overline{\rho }}\) defined by:

$$\begin{aligned} \tau g(x)=\alpha ^{3}g\left( \frac{x}{\alpha }\right) , \end{aligned}$$

for all \(x\in V\) and \(g\in M_{\overline{\rho }}.\) Let \(g,h\in M_{\overline{\rho }}\) and let \(c \in [0,1]\) be an arbitrary constant with \(\overline{\rho }(g-h)<c\). From the definition of \(\overline{\rho }\), we have \(\rho (g(x)-h(x))\le c\varphi (x,0)\) for all \(x\in V\). By the assumption and the last inequality, we get

$$\begin{aligned} \rho \left( \alpha ^{3}g\left( \frac{x}{\alpha }\right) -\alpha ^{3}h\left( \frac{x}{\alpha }\right) \right)\le & {} \kappa ^{3}\rho \left( g\left( \frac{x}{\alpha }\right) -g\left( \frac{x}{\alpha }\right) \right) \\\le & {} \kappa ^{3} c\varphi \left( \frac{x}{\alpha },0\right) \\\le & {} cL\varphi (x,0), \end{aligned}$$

for all \(x\in V\). Hence, \(\overline{\rho }(\tau g-\tau h)\le L\overline{\rho }(g-h),\) for all \(g,h\in \mathfrak {M_{\overline{\rho }}}\) that is, \(\tau \) is a \(\overline{\rho }\)-contraction. Next, we show that \(\tau \) has a bounded orbit at f. Letting \(y=0\) in (2.24), we get

$$\begin{aligned} \rho (\alpha ^{3}f(x)-f(\alpha x))&\le \varphi (x,0), \end{aligned}$$
(2.26)

for all \(x\in V\). Replacing x with \(\frac{x}{\alpha }\) in (2.26), we get

$$\begin{aligned} \rho \left( \alpha ^{3}f\left( \frac{x}{\alpha }\right) -f(x)\right)\le & {} \varphi (\frac{x}{\alpha },0), \end{aligned}$$
(2.27)

for all \(x\in V\). Replacing x with \(\frac{x}{\alpha }\) in (2.27), we get

$$\begin{aligned} \rho \left( \alpha ^{3}f\left( \frac{x}{\alpha ^{2}}\right) -f\left( \frac{x}{\alpha }\right) \right)\le & {} \varphi (\frac{x}{\alpha ^{2}},0), \end{aligned}$$
(2.28)

for all \(x\in V\). By using (2.26), (2.27) and (2.28), we get

$$\begin{aligned} \rho (\alpha ^{6}f(\frac{x}{\alpha ^{2}})-f(x))\le & {} \rho (\alpha ^{6}f(\frac{x}{\alpha ^{2}})- \alpha ^{3}f(\frac{x}{\alpha }))+\rho (\alpha ^{3}f(\frac{x}{\alpha })- f(x))\nonumber \\\le & {} \kappa ^{3}\rho (\alpha ^{3}f(\frac{x}{\alpha ^{2}})- f(\frac{x}{\alpha }))+\rho (\alpha ^{3}f(\frac{x}{\alpha })-f(x))\nonumber \\\le & {} \alpha ^{3}\varphi (\frac{x}{\alpha ^{2}},0)+\varphi (\frac{x}{\alpha },0), \end{aligned}$$
(2.29)

for all \(x\in V\). By induction, we can easily see that

$$\begin{aligned} \rho \big (\alpha ^{3n}f(\frac{x}{\alpha ^{n}})-f(x)\big )\le & {} \frac{1}{\alpha ^{3}}\sum _{i=1}^{n}\alpha ^{3i}\varphi (\frac{x}{\alpha ^{i}},0)\nonumber \\\le & {} \frac{1}{\alpha ^{3}}\varphi (x,0)\sum _{i=1}^{n}L^{i}\nonumber \\\le & {} \frac{L}{\alpha ^{3}(1-L)}\varphi (x,0), \end{aligned}$$
(2.30)

for all \(x\in V\). It follows from inequality (2.30) that

$$\begin{aligned} \rho \big (\alpha ^{3n}f(\frac{x}{\alpha ^{n}})-\alpha ^{3k}f(\frac{x}{\alpha ^{k}})\big )\le & {} \frac{1}{2}\rho (2\alpha ^{3n}f(\frac{x}{\alpha ^{n}})-2f(x))+\frac{1}{2}\rho (2\alpha ^{3k}f(\frac{x}{\alpha ^{k}})-2f(x))\nonumber \\\le & {} \frac{kL}{\alpha ^{3}(1-L)}\varphi (x,0), \end{aligned}$$
(2.31)

for every \(x\in V\) and \(n,k \in \mathbb {N}\), By the definition of \(\overline{\rho }\), we conclude that

$$\begin{aligned} \overline{\rho }(\tau ^{n} f-\tau ^{k} f)\le \frac{kL}{\alpha ^{3}(1-L)}, \end{aligned}$$

which implies the boundedness of an orbit of \(\tau \) at f. It follows from Theorem 1.5 that, the sequence \(\{\tau ^{n}f\}\) \(\overline{\rho }\)-converges to \(C_{\alpha }\in M_{\overline{\rho }}\). Now, by the \(\overline{\rho }\)-contractivity of \(\tau \), we have

$$\begin{aligned} \overline{\rho }(\tau ^{n} f-\tau C_{\alpha })\le L\overline{\rho }(\tau ^{n-1} f-C_{\alpha }). \end{aligned}$$

Employing the limit \(n\rightarrow \infty \) and applying the Fatou property of \(\overline{\rho }\), we obtain that

$$\begin{aligned} \overline{\rho }(\tau C_{\alpha }-C_{\alpha })\le & {} \liminf _{n\rightarrow \infty }\overline{\rho }(\tau C_{\alpha }-\tau ^{n} f)\\\le & {} L\liminf _{n\rightarrow \infty }\overline{\rho }(C_{\alpha }-\tau ^{n-1} f)=0. \end{aligned}$$

Therefore, \(C_{\alpha }\) is a fixed point of \(\tau \). Letting \(x = \displaystyle \frac{x}{\alpha ^{n}}\) and \(y=\displaystyle \frac{y}{\alpha ^{n}}\) in (2.24), we get

$$\begin{aligned} \rho \big (D_{\alpha }f(\frac{x}{\alpha ^{n}},\frac{y}{\alpha ^{n}})\big )\le \varphi (\frac{x}{\alpha ^{n}},\frac{y}{\alpha ^{n}}), \end{aligned}$$

for all \(x,y\in V\). Therefore

$$\begin{aligned} \rho \left( \alpha ^{3n}D_{\alpha }f\left( \frac{x}{\alpha ^{n}},\frac{y}{\alpha ^{n}}\right) \right) \le \kappa ^{3n}\varphi \left( \frac{x}{\alpha ^{n}},\frac{y}{\alpha ^{n}}\right) , \end{aligned}$$

Passing to the limit \(n\rightarrow \infty \), we get

$$\begin{aligned} D_{\alpha }C_{\alpha }(x,y)=0 \end{aligned}$$

for all \(x,y\in V\). It follows from Lemma 2.2 that \(C_{\alpha }\) is cubic. By using (2.30), we get (2.25). \(\square \)

Corollary 2.6

Let X be a Banach space, \(\varphi : V^{2}\longrightarrow [0, +\infty )\) be a function such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\alpha ^{3n}\varphi \left( \frac{x}{\alpha ^{n}},\frac{y}{\alpha ^{n}}\right) =0, \end{aligned}$$

and

$$\begin{aligned} \varphi \left( \frac{x}{\alpha },\frac{y}{\alpha }\right) \le \frac{L}{\alpha ^{3}}\varphi (x,y), \end{aligned}$$

for all \(x,y\in V\) with \(L<1.\) Suppose that \(f:V\longrightarrow X\) satisfies the condition

$$\begin{aligned} \big \Vert D_{\alpha }f(x,y)\big \Vert \le \varphi (x,y), \end{aligned}$$

for all \(x,y\in V\) and \(f(0)=0.\) Then there exists a unique cubic mapping \(C_{\alpha }: V\longrightarrow X\) such that

$$\begin{aligned} \big \Vert C_{\alpha }(x)-f(x)\big \Vert \le \frac{L}{\alpha ^{3}(1-L)}\varphi (x,0), \end{aligned}$$

for all \(x\in V\).

Proof

It is known that every normed space is modular space with the modular \(\rho (x) = \Vert x\Vert \) and satisfies the \(\Delta _{\alpha }\)-condition with \( \kappa = \alpha \). \(\square \)

Remark 2.7

In Corollaries 2.4 and 2.6, by replacing \(\varphi \) with:

$$\begin{aligned} \varphi (x,y)= & {} \Vert x\Vert ^{p}+\Vert y\Vert ^{p},\\ \varphi (x,y)= & {} \Vert x\Vert ^{p}\Vert y\Vert ^{q},\\ \varphi (x,y)= & {} \Vert x\Vert ^{p}+\Vert y\Vert ^{p}+\Vert x\Vert ^{r}\Vert y\Vert ^{s}, \end{aligned}$$

under suitable conditions, it is possible to obtain some corollaries.