Processing math: 60%
Research article

Refined stability of the additive, quartic and sextic functional equations with counter-examples

  • Received: 04 February 2023 Revised: 10 April 2023 Accepted: 12 April 2023 Published: 19 April 2023
  • MSC : 39B52, 39B72, 39B82

  • In this study, we utilize the direct method (Hyers approach) to examine the refined stability of the additive, quartic, and sextic functional equations in modular spaces with and without the Δ2-condition. We also use the direct approach to discuss the Ulam stability in 2-Banach spaces. Ultimately, we ensure that stability of above equations does not hold in a particular scenario by utilizing appropriate counter-examples.

    Citation: Hasanen A. Hammad, Hassen Aydi, Manuel De la Sen. Refined stability of the additive, quartic and sextic functional equations with counter-examples[J]. AIMS Mathematics, 2023, 8(6): 14399-14425. doi: 10.3934/math.2023736

    Related Papers:

    [1] Murali Ramdoss, Divyakumari Pachaiyappan, Inho Hwang, Choonkil Park . Stability of an n-variable mixed type functional equation in probabilistic modular spaces. AIMS Mathematics, 2020, 5(6): 5903-5915. doi: 10.3934/math.2020378
    [2] K. Tamilvanan, Jung Rye Lee, Choonkil Park . Hyers-Ulam stability of a finite variable mixed type quadratic-additive functional equation in quasi-Banach spaces. AIMS Mathematics, 2020, 5(6): 5993-6005. doi: 10.3934/math.2020383
    [3] Vediyappan Govindan, Inho Hwang, Choonkil Park . Hyers-Ulam stability of an n-variable quartic functional equation. AIMS Mathematics, 2021, 6(2): 1452-1469. doi: 10.3934/math.2021089
    [4] Bei Wang, Songsong Li, Yi Ouyang, Honggang Hu . Ready-made short basis for GLV+GLS on high degree twisted curves. AIMS Mathematics, 2022, 7(1): 306-314. doi: 10.3934/math.2022021
    [5] Zhihua Wang, Choonkil Park, Dong Yun Shin . Additive ρ-functional inequalities in non-Archimedean 2-normed spaces. AIMS Mathematics, 2021, 6(2): 1905-1919. doi: 10.3934/math.2021116
    [6] Zhenyu Jin, Jianrong Wu . Ulam stability of two fuzzy number-valued functional equations. AIMS Mathematics, 2020, 5(5): 5055-5062. doi: 10.3934/math.2020324
    [7] Cure Arenas Jaffeth, Ferrer Sotelo Kandy, Ferrer Villar Osmin . Functions of bounded (2,k)-variation in 2-normed spaces. AIMS Mathematics, 2024, 9(9): 24166-24183. doi: 10.3934/math.20241175
    [8] Jae-Hyeong Bae, Won-Gil Park . Approximate solution of a multivariable Cauchy-Jensen functional equation. AIMS Mathematics, 2024, 9(3): 7084-7094. doi: 10.3934/math.2024345
    [9] Kandhasamy Tamilvanan, Jung Rye Lee, Choonkil Park . Ulam stability of a functional equation deriving from quadratic and additive mappings in random normed spaces. AIMS Mathematics, 2021, 6(1): 908-924. doi: 10.3934/math.2021054
    [10] Zhihua Wang . Stability of a mixed type additive-quadratic functional equation with a parameter in matrix intuitionistic fuzzy normed spaces. AIMS Mathematics, 2023, 8(11): 25422-25442. doi: 10.3934/math.20231297
  • In this study, we utilize the direct method (Hyers approach) to examine the refined stability of the additive, quartic, and sextic functional equations in modular spaces with and without the Δ2-condition. We also use the direct approach to discuss the Ulam stability in 2-Banach spaces. Ultimately, we ensure that stability of above equations does not hold in a particular scenario by utilizing appropriate counter-examples.



    In many different settings, functional equations are essential to the investigation of stability problems. The first in challenging the stability of group homomorphisms was Ulam [1]. His work laid the groundwork for subsequent research on stability phenomena. If an equation allows only one unique solution, we refer to that equation as being stable. Ulam [1] formulated the following Cauchy functional equation:

    Ξ(s1+s2)=Ξ(s1)+Ξ(s2).

    In the context of a Banach space, Hyers [2] addressed Cauchy's functional equation in order to resolve this problems. Aoki [3] improved the work of Hyers by taking an unbounded Cauchy difference. Rassias [4] discussed additive mappings in his study, and Găvruţa [5] has already given identical results. For more details about the stability results, see [2,6,7,8,9,10,11,12,13].

    In 1950, Nakano [14] investigated the idea of modular linear spaces. Numerous writers have now extensively verified these hypotheses, e.g., Luxemburg [15], Amemiya [16], Musielak [17], Koshi [18], Mazur [19], Turpin [20] and Orlicz [21]. Both Orlicz spaces [22] and the concept of interpolation [17,22] have several applications in the setting of modular spaces.

    Several researchers examined stability in modular spaces via a fixed point approach of quasi-contractions without utilizing the Δ2-condition, as suggested by Khamsi [23]. In recent years, Sadeghi [24] produced results on stability of some functional equations combining the Δ2-condition with the Fatou property.

    First, we review some terminology, notations, and common characteristics of the theory of given spaces.

    Definition 1.1. [24] Let Q be a linear space over k (R or C). A function ϱ:Q[0,) is said to be modular if the hypotheses below hold for all ϖ,ρQ:

    (m1) ϱ(ϖ)=0ϖ=0;

    (m2) ϱ(aϖ)=ϱ(ϖ) for any scalar a with |a|=1;

    (m3) ϱ(a1ϖ+a2ρ)ϱ(ϖ)+ϱ(ρ) for any scalar a1,a20 with a1+a2=1.

    Also, ϱ is said to be convex modular, if the hypothesis (m3) is replaced by

    (m3) ϱ(a1ϖ+a2ρ)a1ϱ(ϖ)+a2ϱ(ρ) for any scalar a1,a20 with a1+a2=1.

    Additionally, the vector space induced by a modular ϱ,

    Qϱ={ϖ:ϱ(cϖ)0, as c},

    is a modular space (MS, for short). Denote by N the set of positive integers.

    Definition 1.2. [24] Let {ϖμ} be a sequence in an MS Qϱ.

    (i) If ϱ(ϖjϖ)0 as j, then {ϖj} is called ϱ-convergent to a point ϖ and we write ϖjϖ as j.

    (ii) If ϱ(ϖjϖξ)<ϵ for any ϵ>0 and for sufficiently large j,ξN, then {ϖj} is called ϱ-Cauchy.

    (iii) KQϱ is called ρ-complete if any ϱ-Cauchy sequence is ϱ-convergent.

    (iv) A modular ϱ is said to satisfy the Δ2-condition if ϱ(2ϖj)0 as j, whenever ϱ(ϖj)0 as j.

    If ϱ(ϖ)liminfμϱ(ϖμ), the modular ϱ possesses the Fatou property, whereas the sequence {ϖμ} is ϱ-convergent to ϖ in the MS Qϱ and vice versa.

    Proposition 1.1. [25] In MSs,

    (1) if ϖαϖ and λ is a constant vector, then ϖα+λϖ+λ;

    (2) if ϖαϖ and ραρ, then a1ϖα+a2ραa1ϖ+a2ρ, where a1,a20 and a1+a21.

    Remark 1.1. Suppose that ρ is convex and justifies the Δ2-condition with Δ2-constant r>0. If r<2, then ϱ(ϖ)rϱ(ϖ2)r2ϱ(ϖ), which suggests ϱ=0. Therefore, if ρ is a convex modular, we ought to obtain the Δ2-constant r2.

    It is clear that if μ is chosen from the analogous scalar field with |μ|>1 in MSs, then the convergence of a sequence {ϖα} to ϖ does not imply that {μϖα} converges to μϖ. This is due to the fact that in MSs, the multiples of the convergent sequence {ϖα} are convergent naturally.

    In 1960, the idea of linear 2-normed spaces was created by Gahler [26] as follows:

    Definition 1.3. Assume that Λ over R is a linear space with dimΛ>1 and a function .,.:Λ×ΛR is such that for all ϖ,ρ,Λ and ϑR,

    (i) ϖ,ρ=0 iff ϖ and ρ are linearly dependent;

    (ii) ϖ,ρ=ρ,ϖ;

    (iii) ϑϖ,ρ=|ϑ|ϖ,ρ;

    (iv) ϖ,+ρϖ,+ϖ,ρ.

    Then the function .,. is called a 2-norm on Λ, and (Λ,.,.) is called a linear 2-normed space (2-NS, for short).

    For example of a 2-NS, consider R2 endowed with a 2-norm defined by |ϖρ|=the area of the triangle with vertices 0, ϖ and ρ.

    It should be noted that, the assertion (iv) implies that

    ϖ+,ρϖ,ρ+,ρ and |ϖ,ρ,ρ|ϖ,ρ.

    Hence, the mapping ϖϖ,ρ is continuous from Λ onto R, for any fixed ρR.

    Definition 1.4. Let Λ be a linear 2-NS and {ϖj}j1 be a sequence in Λ.

    (1) A sequence {ϖj}j1 is called convergent if there exists an element ϖΛ such that

    limjϖjϖ,=0 for every Λ.

    If {ϖj}j1 converges to ϖ, then we can write ϖjϖ as j or limjϖj=ϖ and we say that ϖ is a limit point of {ϖj}j1.

    (2) Assume that ,ρΛ such that and ρ are linearly independent. Then {ϖj}j1 is called a Cauchy sequence in Λ, if

    limj,vϖjϖv,=0,

    and

    limj,vϖjϖv,ρ=0.

    Definition 1.5. A linear 2-NS in which every Cauchy sequence is a convergent sequence is called a 2-Banach space (2-BS, for short).

    Lemma 1.1. [27] Assume that (Λ,.,.) is a 2-NS. If ϖΛ and ϖ,ρ=0 for each ρΛ, then ρ=0.

    Lemma 1.2. [27] Let {ϖj}j1 be a convergent sequence in a linear 2-NS Λ, then,

    limjϖj,=limjϖj, for all  Λ.

    Sadeghi [24] has confirmed the stability findings of functional equations utilizing the Fatou property and the Δ2-condition in modular spaces. Our paper is aimed to discuss the refined stability of additive, quartic and sextic functional equations

    Ω((s1s2)+(s3s2)m+s4)+Ω((s2s3)+(s4s3)m+s1)+Ω((s3s4)+(s1s4)m+s2)+Ω((s4s1)+(s2s1)m+s3)=Ω(s1+s2+s3+s4),
    Ω(3s1+s2)+Ω(3s1s2)=9Ω(s1+s2)+9Ω(s1s2)+144Ω(s1)16Ω(s2),

    and

    Ω(s1+3s2)6Ω(s1+2s2)+15Ω(s1+s2)20Ω(s1)+15Ω(s1s2)6Ω(s12s2)+Ω(s13s2)=720Ω(s2),

    respectively, in MSs with and without the Δ2-condition and by the direct technique. Additionally, the Ulam stability in 2-BSs is examined. Finally, we show that the stability of these equations does not hold in a particular scenario using appropriate counter-examples.

    Here, we apply the direct technique to examine the stability theorems of the additive, quartic, and sextic functional equation. These results are considered as an improvement of forms due to Wongkum [28] and Sadeghi [24]. We assume here Λ is a linear space and Qϱ is a complete convex MS.

    Kim [29] in 2013 studied the stability of the additive functional equation in fuzzy BSs. Inspired by the technique of Kim [29], we aim to study the stability of the additive functional equation:

    Ω((s1s2)+(s3s2)m+s4)+Ω((s2s3)+(s4s3)m+s1)+Ω((s3s4)+(s1s4)m+s2)+Ω((s4s1)+(s2s1)m+s3)=Ω(s1+s2+s3+s4) (2.1)

    for any m>0 in modular spaces by ignoring the conditions of Δ2. For the convenience of notation, define the mapping Ω:ΛQϱ as

    ΔΩ(s1,s2,s3,s4)=Ω(s1s2m+s3+s4)+Ω(s2s3m+s4+s1)+Ω(s3s4m+s1+s2)+Ω(s4s1m+s2+s3)Ω(s1+s2+s3+s4),

    where s1,s2,s3,s4Λ, and m is a fixed nonzero integer.

    Theorem 2.1. Assume that there is a function Ξ:Λ4[0,) defined by

    Ξ(s1,s2,s3,s4)=μ=114μE(4μ1s1,4μ1s2,4μ1s3,4μ1s4)<, (2.2)

    such that a mapping Ω:ΛQϱ satisfies Ω(0)=0 and for all s1,s2,s3,s4Λ,

    ϱ(ΔΩ(s1,s2,s3,s4))Ξ(s1,s2,s3,s4). (2.3)

    Then there exists a unique additive mapping (AM) W:ΛQϱ fulfilling

    ϱ(Ω(s1)W(s1))Ξ(s1,s1,s1,s1) for all s1Λ. (2.4)

    Proof. Putting s1=s2=s3=s4 in (2.3), and setting Ξ(s1,s1,s1,s1)=U(s1), we have

    ϱ(4Ω(s1)Ω(4s1))Ξ(s1,s1,s1,s1)=U(s1). (2.5)

    Hence,

    ϱ(Ω(s1)14Ω(4s1))14U(s1). (2.6)

    Based on a mathematical induction, one can deduce that

    ϱ(Ω(s1)Ω(4μs1)4μ)μj=114jU(4j1s1), (2.7)

    for all s1Λ and all natural numbers μ. Clearly, (2.6) follows immediately form (2.7) if we take μ=1. Assume that the inequality (2.7) is true for μN, then we get

    ϱ(Ω(s1)Ω(4μ+1s1)4μ+1)=ϱ(14(Ω(4s1)Ω(4μ4s1)4μ)+14(4Ω(s1)Ω(4s1)))14ϱ(Ω(4s1)Ω(4μ4s1)4μ)+14ϱ(4Ω(s1)Ω(4s1))14μj=114jU(4js1)+14U(s1)=μj=114j+1U(4js1)+14U(s1)=μ+1j=114jU(4j1s1).

    It follows that the inequality (2.7) is true for every μN. Suppose that θ and η are natural numbers with θ<η. Using (2.7), we can write

    ϱ(Ω(4ηs1)4ηΩ(4θs1)4θ)=ϱ(14θ(Ω(4ηθs1)4ηθΩ(4θs1)))14θηθj=1U(4j14θs1)4j=ηθj=1U(4θ+j1s1)4θ+j=ημ=θ+1U(4μ1s1)4μ. (2.8)

    Inequalities (2.2) and (2.8) illustrate that {Ω(4ηs1)4η} is ϱ-Cauchy sequence in Qϱ. Since Qϱ is ϱ-complete, one can say {Ω(4ηs1)4η} is ϱ-convergent. Now, describe the mapping W:ΛQϱ as

    W(s1)=limηΩ(4ηs1)4η, s1Λ. (2.9)

    Hence,

    ϱ(4W(s1)W(4s1)44)=ϱ(144(Ω(4η+1s1)4ηW(4s1))+142(14W(s1)14Ω(4η+1s1)4η+1))144ϱ(W(4s1)Ω(4η+1s1)4η)+143ϱ(Ω(4η+1s1)4η+1W(s1)), (2.10)

    for all s1Λ. Applying (2.9) in (2.10) after taking the limit as η, we find that the right-hand side of (2.10) tends to 0. Thus, one gets

    4W(s1)=W(4s1), for all s1Λ. (2.11)

    Also, for all ηN, by (2.11), we observe that

    ϱ(Ω(s1)W(s1))=ϱ(ημ=14Ω(4μ1s1)Ω(4μs1)4μ+(Ω(4ηs1)4ηW(s1)))=ϱ(ημ=14Ω(4μ1s1)Ω(4μs1)4μ+14(Ω(4η14s1)4η1W(4s1))). (2.12)

    Since ημ=114μ+14<1, by (2.5) and (2.12), one can write

    ϱ(Ω(s1)W(s1))ημ=114μϱ(4Ω(4μ1s1)Ω(4μs1))+14ϱ(Ω(4η14s1)4η1W(4s1))ημ=114μU(4μ1s1)+14ϱ(Ω(4η14s1)4η1W(4s1))=ημ=114μΞ(4μ1s1,4μ1s1,4μ1s1,4μ1s1)+14ϱ(Ω(4η14s1)4η1W(4s1)). (2.13)

    Passing to the limit as η in (2.13), we have

    ϱ(Ω(s1)W(s1))Ξ(s1,s1,s1,s1) for all s1Λ.

    Therefore, the inequality (2.4) is true. Now, we shall prove that W is an AM. It is easy to observe that

    ϱ(14jΔΩ(4js1,4js2,4js3,4js4))14jϱ(ΔΩ(4js1,4js2,4js3,4js4))14jΞ(4js1,4js2,4js3,4js4), (2.14)

    for all s1,s2,s3,s4Λ. When j in (2.14), we get ϱ(ΔW(s1,s2,s3,s4))0. Hence,

    ΔW(s1,s2,s3,s4)=0.

    This implies that W is an additive mapping. For the uniqueness, assume that W1 and W2 are two AMs that satisfy (2.4). Then,

    ϱ(W1(s1)W2(s1)2)=ϱ(12(W1(4μs1)4μΩ(4μs1)4μ)+12(Ω(4μs1)4μW2(4μs1)4μ))12ϱ(W1(4μs1)4μΩ(4μs1)4μ)+12ϱ(Ω(4μs1)4μW2(4μs1)4μ)1214μ[ϱ(W1(4μs1)Ω(4μs1))+ϱ(W2(4μs1)Ω(4μs1))]14μΞ(4μs1,4μs1,4μs1,4μs1)u=μ+114uE(4u1s1,4u1s1,4u1s1,4u1s1)0, as u,

    which yields that W1=W2. This finishes the proof.

    The corollaries below follow immediately from Theorem 2.1:

    Corollary 2.1. If there exists a mapping Ω:ΛQϱ such that Ω(0)=0 and

    ϱ(ΔΩ(s1,s2,s3,s4))ε,

    for all s1,s2,s3,s4Λ, then there exists a unique AM W:ΛQϱ satisfying

    ϱ(Ω(s1)W(s1))ε2 for all s1Λ.

    Corollary 2.2. If there exists a mapping Ω:ΛQϱ such that Ω(0)=0 and

    ϱ(ΔΩ(s1,s2,s3,s4))ξ(s1p+s2p+s3p+s4p)

    for all s1,s2,s3,s4Λ, ξ>0 and p(0,1), then there exists a unique AM W:ΛQϱ fulfilling

    ϱ(Ω(s1)W(s1))4ξ44ps1p for all s1Λ.

    In the context of MSs, we present another stability result as in Theorem 2.1 with condition Δ2 as follows:

    Theorem 2.2. Let Q be a linear space and Qϱ fulfill the Δ2-condition with the mapping Ω:ΛQϱ such that

    ϱ(ΔΩ(s1,s2,s3,s4))Ξ(s1,s2,s3,s4),

    and

    limμuμΞ(s14μ,s24μ,s34μ,s44μ)=0 and j=1(u24)jΞ(s14μ,s14μ,s14μ,s14μ)<,

    for all s1,s2,s3,s4Λ. Then there exists a unique AM W:ΛQϱ, described as

    W(s1)=limμ4μΩ(s14μ),

    and

    ϱ(Ω(s1)W(s1))α4uj=1(u24)jΞ(s12j,s12j,s12j,s12j),

    for all s1Λ.

    Proof. Since ϱ verifies the Δ2-condition with α, Eq (2.3) implies that

    ϱ(ΔΩ(s1,s2,s3,s4))αΞ(s1,s2,s3,s4) for all s1,s2,s3,s4Λ.

    So, the proof of Theorem 2.1 directly leads to the conclusion.

    In this part, without using the Fatou property, the refined and Ulam stability of the following quartic functional equation are investigated:

    Ω(3s1+s2)+Ω(3s1s2)=9Ω(s1+s2)+9Ω(s1s2)+144Ω(s1)16Ω(s2), (2.15)

    in modular spaces Qϱ. For ease of notations, we can define a mapping Ω:ΛQϱ as

    ΔΩ(s1,s2)=Ω(3s1+s2)+Ω(3s1s2)+16Ω(s2)144Ω(s1)9Ω(s1+s2)9Ω(s1s2),

    for all s1,s2Λ.

    Theorem 2.3. Let Q be a linear space and Qϱ fulfill the Δ2-condition with a mapping Ξ:Λ×Λ[0,). Suppose also there exists a mapping Ω:ΛQϱ such that

    ϱ(ΔΩ(s1,s2))Ξ(s1,s2),limμu4μΞ(s13μ,s23μ)=0 and j=1(u43)jΞ(s13j,0)<, (2.16)

    for all s1,s2Λ. Then there exists a unique quartic mapping (QM) W:ΛQϱ described as

    W(s1)=limμ34μΩ(s13μ),

    and

    ϱ(Ω(s1)W(s1))12uj=1(u43)jΞ(s13j,0), (2.17)

    for all s1Λ.

    Proof. Consider Ω(0)=0 in view of Ξ(0,0)=0 along the convergence of

    j=1(u43)jΞ(0,0)<.

    Letting s2=0 in (2.16), we have

    ϱ(2Ω(3s1)2×34Ω(s1))Ξ(s1,0) for all s1Λ.

    Since j=112j<1, based on Δ2-condition of ϱ, the subsequent functional inequality can be written as

    ϱ(Ω(s1)34μΩ(s13μ))=ϱ(μj=113j(34j3Ω(s13j1)35jΩ(s13j)))1u3μj=1(u43)jΞ(s13j,0) for all s1Λ. (2.18)

    Now, in (2.18), replacing s1 with s13μ, we conclude that the series in (2.16) converges, and

    ϱ(34θΩ(s13μ)34(θ+μ)Ω(s13θ+μ))u4θϱ(Ω(s13μ)34μΩ(s13θ+μ))u4θ3μj=1(u43)jΞ(s13j+θ,0)3θuθ+3μ+βj=θ+1(u43)jΞ(s13j,0),

    for all s1Λ. Since 3θuθ+31, the right-hand side of the above inequality tends to 0. This proves that {34μΩ(s13μ)} is a ϱ-Cauchy sequence in Qϱ. Since Qϱ is ϱ-complete, it is ϱ-convergent in Qϱ. Define the mapping W:ΛQϱ by

    W(s1)=ϱ(limη34μΩ(s13μ)) for all s1Λ,

    that is,

    limηϱ(34μΩ(s13μ)W(s1))=0 for all s1Λ.

    Now, consider

    ϱ(Ω(s1)W(s1))12ϱ(2Ω(s1)2(34μ)Ω(s13μ))+12ϱ(2(34μ)Ω(s13μ)2W(s1))u2ϱ(Ω(s1)(34μ)Ω(s13μ))+u2ϱ((34μ)Ω(s13μ)W(s1))12u2μj=1(u43)jΞ(s13j,0)+u2ϱ((34μ)Ω(s13μ)W(s1)),

    for all s1Λ and all μ>1. Thus, the inequality is founded without utilizing the Fatou property. Letting μ, we have estimate of (2.17) of Ω as W. Replacing (s1,s2) with (s13μ,s23μ) in (2.16), one gets

    ϱ(34μΔΩ(s13μ,s23μ))u4μΞ(s13μ,s23μ)0, as μ.

    It follows from the convexity of ϱ that

    ϱ(1235W(3s1+s2)+1235W(3s1s2)+16235W(s2)9235W(s1+s2)9235W(s1s2)144235W(s1))1235ϱ(W(3s1+s2)34μΩ(3s1+s23μ))+1235ϱ(W(3s1s2)34μΩ(3s1s23μ))+16235ϱ(W(s2)16(34μ)Ω(s23μ))+9235ϱ(W(s1+s2)9(34μ)Ω(s1+s23μ))+9235ϱ(W(s1s2)9(34μ)Ω(s1s23μ))+144235ϱ(W(s1)144(34μ)Ω(s13μ))+1235ϱ(34μΩ(3s1+s23μ)+34μΩ(3s1s23μ)+16(34μ)Ω(s23μ)+9(34μ)Ω(s1+s23μ)+9(34μ)Ω(s1s23μ)+144(34μ)Ω(s13μ)),

    for all s1,s2Λ. Then the function W is a quartic (it is enough to let μ).

    For the uniqueness, suppose that W:ΛQϱ is a QM such that

    ϱ(Ω(s1)W(s1))12uj=1(u43)jΞ(s13j,0) for all s1Λ.

    Then, from the equations W(3μs1)=34μW(s1) and W(3μs1)=34μW(s1), one can write

    ϱ(W(s1)W(s1))13ϱ(3(34μ)W(s13μ)3(34μ)Ω(s13μ))+13ϱ(3(34μ)Ω(s13μ)3(34μ)W(s13μ))u4μ+13ϱ(W(s13μ)Ω(s13μ))+u4μ+13ϱ(Ω(s13μ)W(s13μ))3μ12uμ+3μj=1(u43)jΞ(s13j,0), for all s1Λ,

    for all sufficiently large integers μ. Letting μ, we conclude that W(s1)=W(s1), for all s1Λ. This completes the proof.

    Corollary 2.3. Suppose that (Λ,.) is a normed space and Qϱ fulfills Δ2-condition. Assume also there are ξ>0, p>log3u43 and the mapping Ω:ΛQϱ such that

    ϱ(ΔΩ(s1,s2))ξ(s1p+s2p) for all s1,s2Λ.

    Then there exists a unique QM W:ΛQϱ fulfilling

    Ω(s1)W(s1)u3ξ3p+1u4 for all s1Λ.

    Without utilizing the Δ2-condition or the Fatou property, we provide another stability result in an MS.

    Theorem 2.4. Assume that there are a mapping Ω:ΛQϱ that fulfills (2.16) and a function Ξ:Λ×Λ[0,) such that

    limμΞ(3μs1,3μs2)34μ=0 and j=1Ξ(3js1,0)34j< for all s1,s2Λ.

    Then there exists a unique QM W:ΛQϱ fulfilling

    ϱ(Ω(s1)14Ω(0)W(s1))134j=1Ξ(3js1,0)34j for all s1Λ.

    Proof. Setting s2=0 in (2.16), one can write

    ϱ(2Ω(3s1)2×34Ω(s1))Ξ(s1,0), for all s1Λ. (2.19)

    Taking Ω(s1)=Ω(s1)14Ω(0), then by the convexity of ϱ and using the fact μ1j=0134(j+1)<1, we get

    ϱ(Ω(s1)Ω(3μs1)34μ)ϱ(μ1j=0(34Ω(3js1)Ω(3j+1s1)34(j+1)))μ1j=0ϱ(34Ω(3js1)Ω(3j+1s1))34(j+1)134μ1j=0Ξ(3js1,0)34j for all s1Λ, μN.

    Then one gets {34μΩ(s13μ)} is a ϱ-Cauchy sequence and the mapping W:ΛQϱ is defined as

    W(s1)=ϱ(limμΩ(3μs1)34μ),

    that is

    ϱ(limμΩ(3μs1)34μW(s1))=0 for all s1Λ,

    without utilizing the Δ2-condition and the Fatou property. Furthermore, it is clear from the proof used in Theorem 2.3 that the mapping W satisfies the quartic functional equation.

    Now, using the Fatou property and the Δ2-condition, we demonstrate that (2.19) is true. According to the convexity of ϱ and using the fact μ1j=0134(j+1)+134<1, we have

    ϱ(Ω(s1)W(s1))=ϱ(μ1j=1(34Ω(3js1)Ω(3j+1s1)34(j+1))+Ω(3μs1)34μW(3s1)34)μ1j=0134(j+1)ϱ(34Ω(3js1)Ω(3j+1s1))+134ϱ(Ω(3μ13s1)34(μ1)W(3s1))134μ1j=0134jΞ(3js1,0)+134ϱ(Ω(3μ13s1)34(μ1)W(3s1)),

    for all s1Λ and all natural number μ>1. Letting μ in the above inequality, we get our desired result.

    Corollary 2.4. Assume that there exists a function Ξ:Λ×Λ[0,) such that

    limμΞ(3μs1,3μs2)34μ=0 and Ξ(3s1,0)<34MΞ(s1,0) for all s1,s2Λ,

    where M(0,1). If there exists a mapping Ω:ΛQϱ fulfilling (2.16), then there exists a unique QM W:ΛQϱ fulfilling

    ϱ(Ω(s1)14Ω(0)W(s1))Ξ(s1,0)34(1M) for all s1Λ.

    Corollary 2.5. Let (Λ,.) be a normed linear space. If there are the real numbers ξ>0, ε>0 and a mapping Ω:ΛQϱ such that

    ϱ(ΔΩ(s1,s2))ξ(s1p+s2p)+ε,

    for all s1,s2Λ, then there exists a unique QM W:ΛQϱ fulfilling

    ϱ(Ω(s1)14Ω(0)W(s1))3ξ343ps1p+ε4 for all s1,s2Λ,

    where s10 if p<0.

    Here, without using the Fatou property, the refined Ulam stability of the following sextic functional equation is introduced:

    Ω(s1+3s2)6Ω(s1+2s2)+15Ω(s1+s2)20Ω(s1)+15Ω(s1s2)6Ω(s12s2)+Ω(s13s2)=720Ω(s2),

    in an MS Qϱ.

    Theorem 2.5. Let Q be a linear space and Qϱ fulfilling the Δ2-condition with a mapping Ξ:Λ×Λ[0,). Assume also there is a mapping Ω:ΛQϱ such that

    ϱ(Ω(s1+3s2)6Ω(s1+2s2)+15Ω(s1+s2)20Ω(s1)+15Ω(s1s2)6Ω(s12s2)+Ω(s13s2)720Ω(s2))}Ξ(s1,s2), (2.20)

    and

    limμu6μΞ(s12μ,s22μ)=0, and j=1(u72)jΞ(s12j,0)<,

    for all s1,s2Λ. Then there exists a unique sextic mapping (SM) W:ΛQϱ, described as

    W(s1)=limμ26μΩ(s12μ),

    and

    ϱ(Ω(s1)W(s1))12uj=1(u72)jΞ(s12j,s12j), (2.21)

    for all s1Λ.

    Proof. Firstly, let Ω(0)=0 in view of Ξ(0,0)=0 along the convergence of

    j=1(u72)jΞ(0,0)<.

    Setting s1=s2 in (2.20), we get

    ϱ(Ω(4s1)6Ω(3s1)+15Ω(2s1)20Ω(s1)6Ω(s1)+Ω(2s1)720Ω(s1))Ξ(s1,s1).

    Since j=112j<1, and using Δ2-condition of ϱ, the next functional inequality can be written as

    ϱ(Ω(s1)26μΩ(s13μ))=ϱ(μj=112j(37j6Ω(s13j1)26jΩ(s12j)))1u6μj=1(u62)jΞ(s12j,s12j), for all s1Λ. (2.22)

    Replacing s1 with s12μ in (2.22), we see that the series in (2.16) converges, and

    ϱ(26θΩ(s12μ)26(θ+μ)Ω(s16θ+μ))u6θϱ(Ω(s12μ)26μΩ(s12θ+μ))u6θ6μj=1(u72)jΞ(s12j+θ,s12j+θ)2θuθ+5μ+βj=θ+1(u72)jΞ(s12j,s12j),

    for all s1Λ, which goes to 0 as θ since 2u1, then the right-hand side of the above inequality tends to 0. This proves that {26μΩ(s12μ)} is a ϱ-Cauchy sequence for all s1Λ and it is ϱ-convergent in Qϱ. Hence, we can define the mapping W:ΛQϱ by

    W(s1)=ϱ(limη26μΩ(s12μ)), that is, limηϱ(26μΩ(s12μ)W(s1))=0,

    for all s1Λ. Consequently, without employing the Fatou property from the Δ2-condition, the following inequality

    ϱ(Ω(s1)W(s1))12ϱ(2Ω(s1)2(26μ)Ω(s12μ))+12ϱ(2(26μ)Ω(s12μ)2W(s1))u2ϱ(Ω(s1)(26μ)Ω(s12μ))+u2ϱ((26μ)Ω(s12μ)W(s1))12uμj=1(u72)jΞ(s12j,s12j)+u2ϱ(26μΩ(s12μ)W(s1)),

    is true for all s1Λ and a nature number μ>1. Letting μ, we have estimate of (2.17) in Ω by W. Replacing (s1,s2) with (s12μ,s22μ) in (2.21), we obtain that

    ϱ(26μΩ(s1+3s22μ)6(26μ)Ω(s1+2s22μ)+15(26μ)Ω(s1+s22μ)20(26μ)Ω(s12μ)+15(26μ)Ω(s1s22μ)6(26μ)Ω(s12s22μ)+26μΩ(s13s22μ)720Ω(s22μ))u6μΞ(s12μ,s22μ)0, as μ, for all s1,s2Λ.

    It follows from the convexity of ϱ that

    ϱ(1784W(s1+3s2)6784W(s1+2s2)+15784W(s1+s2)20784W(s1)+15784W(s1s2)6784W(s12s2)+1784W(s13s2)720784W(s2))1784ϱ(W(s1+3s2)26μΩ(s1+3s23μ))+6784ϱ(W(s1+2s2)6(26μ)Ω(s1+2s22μ))+15784ϱ(W(s1+s2)15(26μ)Ω(s1+s22μ))+20784ϱ(W(s1)20(26μ)Ω(s12μ))+15784ϱ(W(s1s2)15(26μ)Ω(s1s22μ))+6784ϱ(W(s12s2)6(26μ)Ω(s12s22μ))+1784ϱ(W(s13s2)126μΩ(s13s22μ))+720784ϱ(W(s2)720(26μ)Ω(s22μ))+1784ϱ(26μΩ(s1+3s23μ)+6(26μ)Ω(s1+2s22μ)+15(26μ)Ω(s1+s22μ)+20(26μ)Ω(s12μ)+15(26μ)Ω(s1s22μ)+6(26μ)Ω(s12s22μ)+126μΩ(s13s22μ)+720(26μ)Ω(s22μ)),

    for all s1,s2Λ. Therefore, the mapping W is sextic (it is enough to let μ).

    For the uniqueness, let W:ΛQϱ be another SM satisfying

    ϱ(Ω(s1)W(s1))12uj=1(u72)jΞ(s12j,s12j), for all s1Λ.

    From the equations W(2μs1)=26μW(s1) and W(2μs1)=26μW(s1), we have

    ϱ(W(s1)W(s1))12ϱ(2(26μ)W(s12μ)2(26μ)Ω(s12μ))+12ϱ(2(26μ)Ω(s12μ)2(26μ)W(s12μ))u6μ+12ϱ(W(s12μ)Ω(s12μ))+u6μ+12ϱ(Ω(s12μ)W(s12μ))u6μ2j=1(u72)jΞ(s12j+μ,s12j+μ)2μ1uμ+3μj=1(u72)jΞ(s12j,s12j),

    for all s1Λ and for all sufficiently large natural numbers μ. Letting μ, we obtain that W=W and this completes the proof.

    Corollary 2.6. Let (Λ,.) be a normed space and Qϱ fulfill the Δ2-condition. If there exist a real number ξ>0, p>log2u62 and a mapping Ω:ΛQϱ such that

    ϱ(Ω(s1+3s2)6Ω(s1+2s2)+15Ω(s1+s2)20Ω(s1)+15Ω(s1s2)6Ω(s12s2)+Ω(s13s2)720Ω(s2))}ξ(s1p+s2p),

    for all s1,s2Λ, then there exists a unique SM W:ΛQϱ fulfilling

    Ω(s1)W(s1)u7ξ2p+1u8 for all s1Λ.

    Now, without employing the Δ2-condition and the Fatou property, the following theorem provides an alternative stability result of Theorem 2.5 in an MS.

    Theorem 2.6. If there exist a mapping Ω:ΛQϱ, which fulfills (2.20) with the function Ξ:Λ×Λ[0,) satisfying

    limμΞ(2μs1,2μs2)26μ=0 and j=1Ξ(2js1,2js1)26j< for all s1,s2Λ,

    then there exists a unique SM W:ΛQϱ such that

    ϱ(Ω(s1)17Ω(0)W(s1))126j=1Ξ(2js1,2js1)26j for all s1Λ. (2.23)

    Proof. Putting s1=s2 in (2.20), we have

    ϱ(Ω(4s1)6Ω(3s1)+15Ω(2s1)20Ω(s1)6Ω(s1)+Ω(2s1)720Ω(s1))=ϱ(ˆΩ(4s1)6ˆΩ(3s1)+15ˆΩ(2s1)20ˆΩ(s1)6ˆΩ(s1)+ˆΩ(2s1)720ˆΩ(s1))Ξ(s1,s1),

    where ˆΩ(s1)=Ω(s1)17Ω(0). From the convexity of ϱ and using the fact μ1j=0126(j+1)<1, we get

    ϱ(ˆΩ(s1)ˆΩ(2μs1)26μ)ϱ(μ1j=0(26ˆΩ(2js1)ˆΩ(2j+1s1)26(j+1)))μ1j=0ϱ(34ˆΩ(3js1)ˆΩ(3j+1s1))26(j+1)126μ1j=0Ξ(2js1,2js1)26j for all s1Λ, μN.

    It follows that the sequence {ˆΩ(2μs1)26μ} is ϱ-Cauchy and the mapping W:ΛQϱ is defined by

    W(s1)=ϱ(limμˆΩ(2μs1)26μ),

    that is,

    ϱ(limμˆΩ(2μs1)26μW(s1))=0 for all s1Λ,

    without utilizing the Δ2-condition and the Fatou property. Clearly, from the proof of Theorem 2.5, we conclude that the mapping W satisfies the sextic functional equation.

    Now, using the Fatou property and the Δ2-condition, we show that (2.23) holds. From the convexity property of ϱ and since μ1j=0126(j+1)+126<1, we get

    ϱ(ˆΩ(s1)W(s1))=ϱ(μ1j=1(26ˆΩ(3js1)ˆΩ(2j+1s1)26(j+1))+ˆΩ(2μs1)26μW(2s1)26)μ1j=0126(j+1)ϱ(26ˆΩ(2js1)ˆΩ(2j+1s1))+126ϱ(ˆΩ(2μ12s1)26(μ1)W(2s1))126μ1j=0126jΞ(2js1,2js1)+126ϱ(ˆΩ(2μ12s1)26(μ1)W(2s1)),

    for all s_{1}\in \Lambda and all natural number \mu > 1. As \mu \rightarrow \infty in the above inequality, we obtain our needed result.

    Corollary 2.7. Assume that there exists a mapping \Xi :\Lambda \times \Lambda \rightarrow \lbrack 0, \infty) satisfying

    \begin{equation*} \lim\limits_{\mu \rightarrow \infty }\frac{\Xi \left( 2^{\mu }s_{1}, 2^{\mu }s_{2}\right) }{2^{6\mu }} = 0\ \mathit{\text{and}}\ \Xi \left( 2s_{1}, 2s_{2}\right) < 2^{6}M^{\ast }\Xi \left( s_{1}, s_{2}\right) , \end{equation*}

    for all s_{1}, s_{2}\in \Lambda and for some M^{\ast }\in (0, 1). If there exists a mapping \Omega :\Lambda \rightarrow Q_{\varrho } fulfilling (2.20), then there exists a unique SM W:\Lambda \rightarrow Q_{\varrho } verifying

    \begin{equation*} \varrho \left( \Omega \left( s_{1}\right) -\frac{1}{7}\Omega \left( 0\right) -W(s_{1})\right) \leq \frac{\Xi \left( s_{1}, s_{1}\right) }{2^{6}(1-M^{\ast })}\; \ \mathit{\text{for all}}\ s_{1}\in \Lambda . \end{equation*}

    Corollary 2.8. Let \left(\Lambda, \left\Vert.\right\Vert \right) be a normed space. Assume that there are \xi > 0, \varepsilon > 0, p\in (-\infty, 2) and a mapping \Omega :\Lambda \rightarrow Q_{\varrho } such that

    \begin{equation*} \left. \begin{array}{c} \varrho \left( \Omega \left( s_{1}+3s_{2}\right) -6\Omega \left( s_{1}+2s_{2}\right) +15\Omega \left( s_{1}+s_{2}\right) -20\Omega \left( s_{1}\right) \right. \\ \left. +15\Omega \left( s_{1}-s_{2}\right) -6\Omega \left( s_{1}-2s_{2}\right) +\Omega \left( s_{1}-3s_{2}\right) -720\Omega \left( s_{2}\right) \right) \end{array} \right\} \leq \xi \left( \left\Vert s_{1}\right\Vert ^{p}+\left\Vert s_{2}\right\Vert ^{p}\right) +\varepsilon , \end{equation*}

    for all s_{1}, s_{2}\in \Lambda. Then there exists a unique SM W:\Lambda \rightarrow Q_{\varrho } fulfilling

    \begin{equation*} \varrho \left( \Omega \left( s_{1}\right) -\frac{1}{7}\Omega \left( 0\right) -W(s_{1})\right) \leq \frac{2\xi }{2^{6}-2^{p}}\left\Vert s_{1}\right\Vert ^{p}+\frac{\varepsilon }{3}, \end{equation*}

    for all s_{1}\in \Lambda.

    In this section, we discuss the stability of the involved functional equations by considering \Lambda as a linear normed space and Q as a 2 -BS.

    For the convenience of notations, define the mapping \Omega :\Lambda \rightarrow Q_{\varrho } as

    \begin{eqnarray*} R\Omega \left( s_{1}, s_{2}, s_{3}, s_{4}\right) & = &\Omega \left( \frac{ s_{1}-s_{2}}{m}+s_{3}+s_{4}\right) +\Omega \left( \frac{s_{2}-s_{3}}{m} +s_{4}+s_{1}\right) +\Omega \left( \frac{s_{3}-s_{4}}{m}+s_{1}+s_{2}\right) \\ &&+\Omega \left( \frac{s_{4}-s_{1}}{m}+s_{2}+s_{3}\right) -\Omega \left( s_{1}+s_{2}+s_{3}+s_{4}\right) , \end{eqnarray*}

    for each s_{1}, s_{2}, s_{3}, s_{4}\in \Lambda.

    Theorem 3.1. Suppose that there exists a function \Xi :\Lambda ^{4}\times Q\rightarrow \lbrack 0, \infty) such that

    \begin{equation} \lim\limits_{j\rightarrow \infty }E\left( 4^{j}s_{1}, 4^{j}s_{2}, 4^{j}s_{3}, 4^{j}s_{4}, \ell \right) = 0, \ \mathit{\text{for all}}\ s_{1}, s_{2}, s_{3}, s_{4}\in \Lambda \ \mathit{\text{and}}\ \ell \in Q. \end{equation} (3.1)

    If there is a mapping \Omega :\Lambda \rightarrow Q with \Omega (0) = 0 such that

    \begin{equation} \left\Vert R\Omega \left( s_{1}, s_{2}, s_{3}, s_{4}\right) , \ell \right\Vert \leq \Xi \left( s_{1}, s_{2}, s_{3}, s_{4}, \ell \right) , \end{equation} (3.2)

    and

    \begin{equation*} \widehat{\Xi }\left( s_{1}, \ell \right) = \sum\limits_{j = 1}^{\infty }\frac{1}{ 4^{j}}E\left( 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, \ell \right) < \infty , \end{equation*}

    for all s_{1}, s_{2}, s_{3}, s_{4}\in \Lambda and \ell \in Q, then there is a unique AM V:\Lambda \rightarrow Q fulfilling

    \begin{equation} \left\Vert \Omega \left( s_{1}\right) -V(s_{1}), \ell \right\Vert \leq \widehat{\Xi }\left( s_{1}, \ell \right) \ \mathit{\text{for all}}\ s_{1}\in \Lambda \ \mathit{\text{and all}}\ \ell \in Q. \end{equation} (3.3)

    Proof. Setting s_{1} = s_{2} = s_{3} = s_{4} in (3.2), we get

    \begin{equation} \left\Vert 4\Omega \left( s_{1}\right) -\Omega \left( 4s_{1}\right) , \ell \right\Vert \leq \Xi \left( s_{1}, s_{2}, s_{3}, s_{4}, \ell \right) . \end{equation} (3.4)

    Replacing s_{1} with 4^{j}s_{1} in (3.4), and using

    \begin{equation*} \left\Vert \frac{1}{4^{j+1}}\Omega \left( 4^{j+1}s_{1}\right) -\frac{1}{4^{j} }\Omega \left( 4^{j}s_{1}\right) , \ell \right\Vert \leq \frac{1}{4^{j+1}}\Xi \left( 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, \ell \right) , \end{equation*}

    for all s_{1}\in \Lambda, \ell \in Q and all j > 0 , one writes

    \begin{eqnarray} \left\Vert \frac{1}{4^{j+1}}\Omega \left( 4^{j+1}s_{1}\right) -\frac{1}{4^{r} }\Omega \left( 4^{r}s_{1}\right) , \ell \right\Vert &\leq &\sum\limits_{t = r}^{j}\left\Vert \frac{1}{4^{t+1}}\Omega \left( 4^{t+1}s_{1}\right) -\frac{1}{4^{t}}\Omega \left( 4^{t}s_{1}\right) , \ell \right\Vert \\ &\leq &\frac{1}{4}\sum\limits_{t = r}^{j}\frac{1}{4^{t}}\Xi \left( 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, \ell \right) , \end{eqnarray} (3.5)

    for all s_{1}\in \Lambda, \ell \in Q and all integers j > 0 and r > 0 with r\leq j. It follows from (3.4) and (3.5) that the sequence \left\{ \frac{\Omega \left(4^{j}s_{1}\right) }{4^{j}}\right\} is a Cauchy sequence in Q . The completeness of Q implies that the sequence \left\{ \frac{\Omega \left(4^{j}s_{1}\right) }{4^{j}}\right\} converges in Q for all s_{1}\in \Lambda. Therefore, we can define that mapping V:\Lambda \rightarrow Q as

    \begin{equation} V(s_{1}) = \lim\limits_{j\rightarrow \infty }\frac{\Omega \left( 4^{j}s_{1}\right) }{ 4^{j}}, \text{ for all }s_{1}\in \Lambda . \end{equation} (3.6)

    Hence,

    \begin{equation*} \lim\limits_{j\rightarrow \infty }\left\Vert \frac{\Omega \left( 4^{j}s_{1}\right) }{4^{j}}-V(s_{1}), \ell \right\Vert = 0, \text{ for all }s_{1}\in \Lambda \text{ and }\ell \in Q. \end{equation*}

    Putting t = 0 and let j\rightarrow \infty in (3.5), we have (3.3).

    Now, we shall show that V is an AM. Using (3.1), (3.2), (3.6) and Lemma 1.2, one gets

    \begin{eqnarray*} \left\Vert R\Omega \left( s_{1}, s_{2}, s_{3}, s_{4}\right) , \ell \right\Vert & = &\lim\limits_{j\rightarrow \infty }\left\Vert R\Omega \left( 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}\right) , \ell \right\Vert \\ &\leq &\lim\limits_{j\rightarrow \infty }\frac{1}{4^{j}}\Xi \left( 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, 4^{j}s_{1}, \ell \right) = 0. \end{eqnarray*}

    By Lemma 1.1,

    \begin{equation*} \left\Vert RV\left( s_{1}, s_{2}, s_{3}, s_{4}\right) , \ell \right\Vert = 0. \end{equation*}

    Thus, V is an AM. For the uniqueness, consider another AM V^{\prime }:\Lambda \rightarrow Q fulfilling (3.3). Then,

    \begin{eqnarray*} \left\Vert V(s_{1})-V^{\prime }(s_{1}), \ell \right\Vert & = &\lim\limits_{j\rightarrow \infty }\left\Vert V(4^{j}s_{1})-\Omega (4^{j}s_{1})+\Omega (4^{j}s_{1})-V^{\prime }(4^{j}s_{1}), \ell \right\Vert \\ &\leq &\lim\limits_{j\rightarrow \infty }\frac{1}{4^{j}}\widehat{\Xi }\left( 4^{j}s_{1}, \ell \right) = 0\text{ for all }s_{1}\in \Lambda \text{ and all } \ell \in Q. \end{eqnarray*}

    Based on Lemma 1.1, we have V(s_{1})-V^{\prime }(s_{1}) = 0 for all s_{1}\in \Lambda, which implies that V = V^{\prime }.

    Corollary 3.1. Assume that \mu :[0, \infty)\rightarrow \lbrack 0, \infty) is a function such that \mu (0) = 0 and the following assertions hold:

    (a_{1}) \mu (\varkappa \omega)\leq \mu (\varkappa)\mu (\omega);

    (a_{2}) For all \varkappa > 1, \mu (\varkappa) < \varkappa.

    If there exists a mapping \Omega :\Lambda \rightarrow Q_{\varrho } such that \Omega (0) = 0 and

    \begin{equation*} \left\Vert R\Omega \left( s_{1}, s_{2}, s_{3}, s_{4}\right) , \ell \right\Vert \leq \mu \left( \left\Vert s_{1}\right\Vert +\left\Vert s_{2}\right\Vert +\left\Vert s_{3}\right\Vert +\left\Vert s_{4}\right\Vert \right) +\mu \left( \ell \right) , \end{equation*}

    for all s_{1}, s_{2}, s_{3}, s_{4}\in \Lambda and \ell \in Q, then there is a unique AM V:\Lambda \rightarrow Q fulfilling

    \begin{equation} \left\Vert \Omega \left( s_{1}\right) -V(s_{1}), \ell \right\Vert \leq \left( \frac{4\mu \left( \left\Vert s_{1}\right\Vert \right) }{4-\mu \left( 4\right) }+\mu \left( \ell \right) \right) , \ \mathit{\text{for all}}\ s_{1}\in \Lambda \ \mathit{\text{and all}}\ \ell \in Q. \end{equation} (3.7)

    Proof. Consider

    \begin{equation*} \Xi \left( s_{1}, s_{2}, s_{3}, s_{4}, \ell \right) = \mu \left( \left\Vert s_{1}\right\Vert +\left\Vert s_{2}\right\Vert +\left\Vert s_{3}\right\Vert +\left\Vert s_{4}\right\Vert \right) +\mu \left( \ell \right) , \end{equation*}

    for all s_{1}, s_{2}, s_{3}, s_{4}\in \Lambda and \ell \in Q. Based on condition (a_{1}), we have

    \begin{equation*} \mu \left( 4^{j}\right) = \left( \mu \left( 4\right) \right) ^{j}, \end{equation*}

    and

    \begin{equation*} \Xi \left( 4^{j}s_{1}, 4^{j}s_{2}, 4^{j}s_{3}, 4^{j}s_{4}, \ell \right) \leq \left( \mu \left( 4\right) \right) ^{j}\left[ \mu \left( \left\Vert s_{1}\right\Vert +\left\Vert s_{2}\right\Vert +\left\Vert s_{3}\right\Vert +\left\Vert s_{4}\right\Vert \right) \right] +\mu \left( \ell \right) . \end{equation*}

    Using Theorem 3.1, we obtain (3.7).

    Corollary 3.2. Assume that \Game :\left([0, \infty)\right) ^{4}\rightarrow \lbrack 0, \infty) is a homogeneous function with degree q and \Omega :\Lambda \rightarrow Q is a mapping satisfying \Omega (0) = 0 and

    \begin{equation*} \left\Vert R\Omega \left( s_{1}, s_{2}, s_{3}, s_{4}\right) , \ell \right\Vert \leq \Game \left( \left\Vert s_{1}\right\Vert , \left\Vert s_{2}\right\Vert , \left\Vert s_{3}\right\Vert , \left\Vert s_{4}\right\Vert \right) \left\Vert \ell \right\Vert , \end{equation*}

    for all s_{1}, s_{2}, s_{3}, s_{4}\in \Lambda and \ell \in Q. Then there is a unique AM V:\Lambda \rightarrow Q fulfilling

    \begin{equation*} \left\Vert \Omega \left( s_{1}\right) -V(s_{1}), \ell \right\Vert \leq \frac{ \Game \left( \left\Vert s_{1}\right\Vert , \left\Vert s_{1}\right\Vert , \left\Vert s_{1}\right\Vert , \left\Vert s_{1}\right\Vert \right) \left\Vert \ell \right\Vert }{4-4^{p}}, \end{equation*}

    for all s_{1}\in \Lambda and all \ell \in Q, where p\in \mathbb{R} ^{+} with p < 1.

    Corollary 3.3. Let c\in \mathbb{R} ^{+} with c < 1 and \Game :\left([0, \infty)\right) ^{4}\rightarrow \lbrack 0, \infty) be a homogeneous function with degree a. Assume that \Omega :\Lambda \rightarrow Q is a mapping satisfying \Omega (0) = 0 and

    \begin{equation*} \left\Vert R\Omega \left( s_{1}, s_{2}, s_{3}, s_{4}\right) , \ell \right\Vert \leq \Game \left( \left\Vert s_{1}\right\Vert , \left\Vert s_{2}\right\Vert , \left\Vert s_{3}\right\Vert , \left\Vert s_{4}\right\Vert \right) +\left\Vert \ell \right\Vert , \end{equation*}

    for all s_{1}, s_{2}, s_{3}, s_{4}\in \Lambda and \ell \in Q. Then there is a unique AM V:\Lambda \rightarrow Q fulfilling

    \begin{equation*} \left\Vert \Omega \left( s_{1}\right) -V(s_{1}), \ell \right\Vert \leq \frac{ \Game \left( \left\Vert s_{1}\right\Vert , \left\Vert s_{1}\right\Vert , \left\Vert s_{1}\right\Vert , \left\Vert s_{1}\right\Vert \right) +\left\Vert \ell \right\Vert }{4-c}, \end{equation*}

    for all s_{1}\in \Lambda and all \ell \in Q.

    Corollary 3.4. Assume that a mapping \Omega :\Lambda \rightarrow Q satisfies \Omega (0) = 0 and

    \begin{equation*} \left\Vert R\Omega \left( s_{1}, s_{2}, s_{3}, s_{4}\right) , \ell \right\Vert \leq \left\Vert s_{1}\right\Vert ^{b}+\left\Vert s_{2}\right\Vert ^{b}+\left\Vert s_{3}\right\Vert ^{b}+\left\Vert s_{4}\right\Vert ^{b}+\left\Vert \ell \right\Vert , \end{equation*}

    for all s_{1}, s_{2}, s_{3}, s_{4}\in \Lambda and \ell \in Q. Then there is a unique AM V:\Lambda \rightarrow Q fulfilling

    \begin{equation*} \left\Vert \Omega \left( s_{1}\right) -V(s_{1}), \ell \right\Vert \leq \frac{ 2\left\Vert s_{1}\right\Vert ^{b}+\left\Vert \ell \right\Vert }{4-b}, \end{equation*}

    for all s_{1}\in \Lambda and all \ell \in Q, where b\in \mathbb{R} ^{+} with b < 1.

    In this part, we assume that the mapping \Omega :\Lambda \rightarrow Q is described as

    \begin{eqnarray*} R\Omega \left( s_{1}, s_{2}\right) & = &\Omega \left( 3s_{1}+s_{2}\right) +\Omega \left( 3s_{1}-s_{2}\right) +16\Omega \left( s_{2}\right) -144\Omega \left( s_{1}\right) \\ &&-9\Omega \left( s_{1}+s_{2}\right) -9\Omega \left( s_{1}-s_{2}\right) , \end{eqnarray*}

    for all s_{1}, s_{2}\in \Lambda.

    Theorem 3.2. Assume that \Xi :\Lambda ^{4}\times Q\rightarrow \lbrack 0, \infty) is a function such that

    \begin{equation} \lim\limits_{j\rightarrow \infty }\frac{1}{3^{4j}}\Xi \left( 3^{j}s_{1}, 3^{j}s_{1}, \ell \right) = 0, \end{equation} (3.8)

    for all s_{1}, s_{2}\in \Lambda and \ell \in Q. If there exists \Omega :\Lambda \rightarrow Q with \Omega (0) = 0 such that

    \begin{equation} \left\Vert R\Omega \left( s_{1}, s_{2}\right) , \ell \right\Vert \leq 2\Xi \left( s_{1}, s_{2}, \ell \right) , \end{equation} (3.9)

    and

    \begin{equation*} \widehat{\Xi }\left( s_{1}, \ell \right) = \frac{1}{3}\sum\limits_{j = 1}^{ \infty }\frac{1}{3^{4j}}E\left( 3^{j}s_{1}, 0, \ell \right) < \infty , \end{equation*}

    for all s_{1}\in \Lambda and \ell \in Q, then there exists a unique QM V_{4}:\Lambda \rightarrow Q fulfilling

    \begin{equation} \left\Vert \Omega \left( s_{1}\right) -V_{4}(s_{1}), \ell \right\Vert \leq \widehat{\Xi }\left( s_{1}, \ell \right) \ \mathit{\text{for all}}\ s_{1}\in \Lambda \ \mathit{\text{and all}}\ \ell \in Q. \end{equation} (3.10)

    Proof. Consider s_{2} = 0 in (3.9). We have

    \begin{equation*} \left\Vert 2\Omega \left( 3s_{1}\right) -2\times 3^{4}\Omega \left( s_{1}\right) , \ell \right\Vert \leq 2\Xi \left( s_{1}, 0, \ell \right) , \end{equation*}

    which implies that

    \begin{equation} \left\Vert \frac{\Omega \left( 3s_{1}\right) }{3^{4}}-\Omega \left( s_{1}\right) , \ell \right\Vert \leq \frac{1}{3^{4}}\Xi \left( s_{1}, 0, \ell \right) , \end{equation} (3.11)

    for all s_{1}\in \Lambda and \ell \in Q. In (3.11), replace s_{1} with 3^{j}s_{1}, to get

    \begin{equation*} \left\Vert \frac{1}{3^{4(j+1)}}\Omega \left( 3^{j+1}s_{1}\right) -\frac{1}{ 3^{4j}}\Omega \left( 3^{j}s_{1}\right) , \ell \right\Vert \leq \frac{1}{ 3^{4(j+1)}}\Xi \left( s_{1}, 0, \ell \right), \end{equation*}

    for all s_{1}\in \Lambda , \ell \in Q and all integer j > 0 . Hence,

    \begin{eqnarray} \left\Vert \frac{1}{3^{4(j+1)}}\Omega \left( 3^{j+1}s_{1}\right) -\frac{1}{ 3^{4m}}\Omega \left( 3^{m}s_{1}\right) , \ell \right\Vert &\leq &\sum\limits_{r = m}^{j}\left\Vert \frac{1}{3^{4(r+1)}}\Omega \left( 3^{r+1}s_{1}\right) -\frac{1}{3^{4r}}\Omega \left( 3^{r}s_{1}\right) , \ell \right\Vert \\ &\leq &\frac{1}{3}\sum\limits_{r = m}^{j}\frac{1}{3^{4r}}E\left( 3^{j}s_{1}, 0, \ell \right) , \end{eqnarray} (3.12)

    for all s_{1}\in \Lambda , \ell \in Q and all integers j\geq m > 0 . Therefore, from (3.9) and (3.12), the sequence \left\{ \frac{\Omega \left(3^{j}s_{1}\right) }{3^{4j}}\right\} is a Cauchy sequence in Q . The completeness of Q implies that the sequence \left\{ \frac{\Omega \left(3^{j}s_{1}\right) }{3^{j}}\right\} converges in Q for all s_{1}\in \Lambda. Therefore, we can describe the mapping V_{4}:\Lambda \rightarrow Q as

    \begin{equation} V(s_{1}) = \lim\limits_{j\rightarrow \infty }\frac{\Omega \left( 3^{j}s_{1}\right) }{ 3^{j}}\text{ for all }s_{1}\in \Lambda . \end{equation} (3.13)

    Hence,

    \begin{equation*} \lim\limits_{j\rightarrow \infty }\left\Vert \frac{\Omega \left( 3^{j}s_{1}\right) }{3^{j}}-V_{4}(s_{1}), \ell \right\Vert = 0\text{ for all }s_{1}\in \Lambda \text{ and }\ell \in Q. \end{equation*}

    Letting m = 0 and j\rightarrow \infty in (3.12), we have (3.10).

    Now, we shall show that V_{4} is a QM. Using (3.8), (3.9), (3.13) and Lemma 1.2, one can write

    \begin{eqnarray*} \left\Vert R\Omega \left( s_{1}, s_{2}\right) , \ell \right\Vert = \lim\limits_{j\rightarrow \infty }\left\Vert R\Omega \left( 3^{j}s_{1}, 3^{j}s_{1}\right) , \ell \right\Vert \leq \lim\limits_{j\rightarrow \infty }\frac{1}{3^{4j}}E\left( 3^{j}s_{1}, 3^{j}s_{2}, \ell \right) = 0, \end{eqnarray*}

    for all s_{1}\in \Lambda , \ell \in Q. By Lemma 1.1, we get

    \begin{equation*} \left\Vert RV_{4}\left( s_{1}, s_{2}, s_{3}, s_{4}\right) , \ell \right\Vert = 0. \end{equation*}

    Thus, V is a QM. For the uniqueness, consider another QM V_{4}^{\prime }:\Lambda \rightarrow Q fulfilling (3.10). Then

    \begin{eqnarray*} \left\Vert V_{4}(s_{1})-V_{4}^{\prime }(s_{1}), \ell \right\Vert & = &\lim\limits_{j\rightarrow \infty }\frac{1}{3^{4j}}\left\Vert V_{4}(3^{j}s_{1})-\Omega (3^{j}s_{1})+\Omega (3^{j}s_{1})-V_{4}^{\prime }(3^{j}s_{1}), \ell \right\Vert \\ &\leq &\lim\limits_{j\rightarrow \infty }\frac{1}{3^{4j}}\widehat{\Xi }\left( 3^{j}s_{1}, \ell \right) = 0\text{ for all }s_{1}\in \Lambda \text{ and all } \ell \in Q. \end{eqnarray*}

    Based on Lemma 1.1, we have V_{4}(s_{1})-V_{4}^{\prime }(s_{1}) = 0 for all s_{1}\in \Lambda, which implies that V_{4} = V_{4}^{\prime } and this completes the proof.

    Corollary 3.5. Let \mu :[0, \infty)\rightarrow \lbrack 0, \infty) be a given function with \mu (0) = 0 and

    (i) \mu (\varkappa \omega)\leq \mu (\varkappa)\mu (\omega),

    (ii) for all \varkappa > 1, \mu (\varkappa) < \varkappa.

    If there exists a mapping \Omega :\Lambda \rightarrow Q_{\varrho } with \Omega (0) = 0 and

    \begin{equation*} \left\Vert R\Omega \left( s_{1}, s_{2}\right) , \ell \right\Vert \leq \mu \left( \left\Vert s_{1}\right\Vert +\left\Vert s_{2}\right\Vert \right) +\mu \left( \ell \right) , \end{equation*}

    for all s_{1}, s_{2}\in \Lambda and \ell \in Q, then there exists a unique QM V_{4}:\Lambda \rightarrow Q fulfilling

    \begin{equation} \left\Vert \Omega \left( s_{1}\right) -V_{4}(s_{1}), \ell \right\Vert \leq \left( \frac{\mu \left( \left\Vert s_{1}\right\Vert \right) }{3-\mu \left( 3\right) }+\mu \left( \ell \right) \right) , \ \mathit{\text{for all}}\ s_{1}\in \Lambda \ \mathit{\text{and all}}\ \ell \in Q. \end{equation} (3.14)

    Proof. Assume that

    \begin{equation*} \Xi \left( s_{1}, s_{2}, \ell \right) = \mu \left( \left\Vert s_{1}\right\Vert +\left\Vert s_{2}\right\Vert \right) +\mu \left( \ell \right) , \end{equation*}

    for all s_{1}, s_{2}\in \Lambda and \ell \in Q. From the condition (i), we get

    \begin{equation*} \mu \left( 3^{j}\right) = \left( \mu \left( 3\right) \right) ^{j}, \end{equation*}

    and

    \begin{equation*} \Xi \left( 3^{j}s_{1}, 3^{j}s_{2}, \ell \right) \leq \left( \mu \left( 3\right) \right) ^{j}\left[ \mu \left( \left\Vert s_{1}\right\Vert +\left\Vert s_{2}\right\Vert \right) \right] +\mu \left( \ell \right) . \end{equation*}

    By utilizing Theorem 3.1, we obtain (3.14).

    Corollary 3.6. Assume that \Game :\left([0, \infty)\right) ^{2}\rightarrow \lbrack 0, \infty) is a homogeneous function with degree q and \Omega :\Lambda \rightarrow Q is a mapping satisfying \Omega (0) = 0 and

    \begin{equation*} \left\Vert R\Omega \left( s_{1}, s_{2}\right) , \ell \right\Vert \leq \Game \left( \left\Vert s_{1}\right\Vert , \left\Vert s_{2}\right\Vert \right) \left\Vert \ell \right\Vert , \end{equation*}

    for all s_{1}, s_{2}\in \Lambda and \ell \in Q. Then there exists a unique QM V_{4}:\Lambda \rightarrow Q fulfilling

    \begin{equation*} \left\Vert \Omega \left( s_{1}\right) -V_{4}(s_{1}), \ell \right\Vert \leq \frac{\Game \left( \left\Vert s_{1}\right\Vert , \left\Vert s_{1}\right\Vert \right) \left\Vert \ell \right\Vert }{3-3^{p}}, \end{equation*}

    for all s_{1}\in \Lambda and all \ell \in Q, where p\in \mathbb{R} ^{+} with p < 1.

    Corollary 3.7. Let c\in \mathbb{R} ^{+} with c < 1 and \Game :\left([0, \infty)\right) ^{4}\rightarrow \lbrack 0, \infty) be a homogeneous function with degree a. Assume that \Omega :\Lambda \rightarrow Q is a mapping satisfying \Omega (0) = 0 and

    \begin{equation*} \left\Vert R\Omega \left( s_{1}, s_{2}\right) , \ell \right\Vert \leq \Game \left( \left\Vert s_{1}\right\Vert , \left\Vert s_{2}\right\Vert \right) +\left\Vert \ell \right\Vert , \end{equation*}

    for all s_{1}, s_{2}\in \Lambda and \ell \in Q. Then there exists a unique AM V_{4}:\Lambda \rightarrow Q fulfilling

    \begin{equation*} \left\Vert \Omega \left( s_{1}\right) -V_{4}(s_{1}), \ell \right\Vert \leq \frac{\Game \left( \left\Vert s_{1}\right\Vert , \left\Vert s_{1}\right\Vert \right) +\left\Vert \ell \right\Vert }{3-c}, \end{equation*}

    for all s_{1}\in \Lambda and all \ell \in Q.

    Proof. We obtain the proof immediately, if we take in Theorem 3.2,

    \begin{equation*} \Xi \left( s_{1}, s_{2}, \ell \right) = \Game \left( \left\Vert s_{1}\right\Vert +\left\Vert s_{2}\right\Vert \right) +\Game \left( \ell \right) , \end{equation*}

    for all s_{1}, s_{2}\in \Lambda and \ell \in Q .

    Corollary 3.8. Assume that a mapping \Omega :\Lambda \rightarrow Q satisfies \Omega (0) = 0 and

    \begin{equation*} \left\Vert R\Omega \left( s_{1}, s_{2}\right) , \ell \right\Vert \leq \left\Vert s_{1}\right\Vert ^{b}+\left\Vert s_{2}\right\Vert ^{b}+\left\Vert \ell \right\Vert , \end{equation*}

    for all s_{1}, s_{2}\in \Lambda and \ell \in Q. Then there exists a unique AM V:\Lambda \rightarrow Q fulfilling

    \begin{equation*} \left\Vert \Omega \left( s_{1}\right) -V_{4}(s_{1}), \ell \right\Vert \leq \frac{2\left\Vert s_{1}\right\Vert ^{b}+\left\Vert \ell \right\Vert }{3-b}, \end{equation*}

    for all s_{1}\in \Lambda and all \ell \in Q, where b\in \mathbb{R} ^{+} with b < 1.

    By the same methods used in sections 3.1 and 3.2, we can obtain the refined stability of the sextic functional equation by defining a mapping \Omega :\Lambda \rightarrow Q as

    \begin{eqnarray*} R\Omega \left( s_{1}, s_{2}\right) & = &\Omega \left( s_{1}+3s_{2}\right) -6\Omega \left( s_{1}+2s_{2}\right) +15\Omega \left( s_{1}+s_{2}\right) -20\Omega \left( s_{1}\right) +15\Omega \left( s_{1}-s_{2}\right) \\ &&-6\Omega \left( s_{1}-2s_{2}\right) +\Omega \left( s_{1}-3s_{2}\right) -720\Omega \left( s_{2}\right) , \end{eqnarray*}

    for all s_{1}, s_{2}\in \Lambda, where \Lambda is a linear normed spaces and Q is a 2 -BS.

    With the aid of a pertinent example, it is demonstrated that the functional equations (2.1) and (2.15) are unstable in the singular condition. To Gajda's outstanding example in [30], which demonstrates the instability in Corollaries 2 and 3 of equations (2.1) and (2.15), respectively, we propose the following examples as counter-examples via the assumptions p\neq 1 and p\neq \log _{3}\frac{ u^{4}}{3}, respectively.

    Here, \mathbb{R} stands for a real space, \mathbb{Z} and \mathbb{Q} refer to the sets of integer ad rational numbers. Our counter-examples can be demonstrated as in [31,32].

    Remark 4.1. If a mapping \Omega : \mathbb{R} \rightarrow \Lambda fulfills the functional equation (2.1), then the following assertions are true:

    (R_{1}) For all s_{1}\in \mathbb{R}, z\in \mathbb{Z} and m\in \mathbb{Q}, \Omega \left(m^{z}s_{1}\right) = m^{z}\Omega \left(s_{1}\right);

    (R_{2}) For all s_{1}\in \mathbb{R}, if the mapping \Omega is continuous, then \Omega \left(s_{1}\right) = s_{1}\Omega \left(1\right).

    Example 4.1. Assume that \Omega : \mathbb{R} \rightarrow \mathbb{R} is a function described as

    \begin{equation*} \Omega \left( s_{1}\right) = \sum\limits_{j = 0}^{\infty }\frac{\varpi \left( 4^{j}s_{1}\right) }{4^{j}}, \end{equation*}

    where

    \begin{equation*} \varpi \left( s_{1}\right) = \left\{ \begin{array}{cc} \delta s_{1}, & \text{if }s_{1}\in (-1, 1), \\ \delta, & \text{otherwise.} \end{array} \right. \end{equation*}

    If we define a function \Omega : \mathbb{R} \rightarrow \mathbb{R} as in (2.1) such that

    \begin{equation*} \left\vert \Delta \Omega \left( s_{1}, s_{2}, s_{3}, s_{4}\right) \right\vert \leq 8\delta \left( \left\vert s_{1}\right\vert +\left\vert s_{2}\right\vert +\left\vert s_{3}\right\vert +\left\vert s_{4}\right\vert \right) , \end{equation*}

    for all s_{1}, s_{2}, s_{3}, s_{4}\in \mathbb{R}, then we cannot found an AM W: \mathbb{R} \rightarrow \mathbb{R} which satisfies

    \begin{equation*} \left\vert \Omega \left( s_{1}\right) -W(s_{1})\right\vert \leq \varkappa \left\vert s_{1}\right\vert , \end{equation*}

    for all s_{1}\in \mathbb{R}, where \xi and \varkappa are constants.

    Remark 4.2. If a mapping \Omega : \mathbb{R} \rightarrow \Lambda fulfills the functional equation (2.15), then the following hypotheses hold:

    (R_{1}) For all s_{1}\in \mathbb{R}, z\in \mathbb{Z} and m\in \mathbb{Q}, \Omega \left(m^{\frac{z}{3}}s_{1}\right) = m^{z}\Omega \left(s_{1}\right);

    (R_{2}) For all s_{1}\in \mathbb{R}, if a mapping \Omega is continuous, then \Omega \left(s_{1}\right) = s_{1}^{3}\Omega \left(1\right).

    Example 4.2. Assume that \Omega : \mathbb{R} \rightarrow \mathbb{R} is a function described as

    \begin{equation*} \Omega \left( s_{1}\right) = \sum\limits_{j = 0}^{\infty }\frac{\varpi \left( 3^{j}s_{1}\right) }{3^{4j}}, \end{equation*}

    where

    \begin{equation*} \varpi \left( s_{1}\right) = \left\{ \begin{array}{cc} \delta s_{1}^{3}, & {if }\ s_{1}\in (-1, 1), \\ \delta, & {otherwise.} \end{array} \right. \end{equation*}

    If we define a function \Omega : \mathbb{R} \rightarrow \mathbb{R} as in (2.15) such that

    \begin{equation*} \left\vert \Delta \Omega \left( s_{1}, s_{2}\right) \right\vert \leq 234\delta \left( \left\vert s_{1}\right\vert ^{3}+\left\vert s_{2}\right\vert ^{3}\right) , \end{equation*}

    for all s_{1}, s_{2}\in \mathbb{R}, then we cannot found a QM W: \mathbb{R} \rightarrow \mathbb{R} which satisfies

    \begin{equation*} \left\vert \Omega \left( s_{1}\right) -W(s_{1})\right\vert \leq \varkappa \left\vert s_{1}\right\vert ^{3}, \end{equation*}

    for all s_{1}\in \mathbb{R}, where \xi and \varkappa are constants.

    The concept of stability of a functional equation arises if we replace this functional equation by an inequality acting as a perturbation on the equation itself. Stability of the functional equation has been become an interesting subject over the last seventy years. Several results appeared in this direction. In our work, the direct method of Hyers has been utilized to study the refined stability of the additive, quartic, and sextic functional equations in modular spaces with and without the \Delta _{2} -condition. Moreover, we used the direct approach to investigate the Ulam stability in 2 -Banach spaces. At the end, some counter examples have been presented in order to ensure that the stability of these equations does not hold in a particular case. As future works, we look forward to study the stability of generalized additive, generalized quartic and generalized sextic functional equations. We will also study the effect of the multivalued mappings of these equations on D -metric spaces and generalized metric spaces.

    This work was supported in part by the Basque Government under Grant IT1555-22.

    The authors declare that they have no competing interests.



    [1] S. M. Ulam, A collection of mathematical problems, New York: Interscience, 1960.
    [2] D. H. Hyers, On the stability of the linear functional equation, Proc. N. A. S., 27 (1941), 222–224. https://doi.org/10.1073/pnas.27.4.222 doi: 10.1073/pnas.27.4.222
    [3] T. Aoki, On the stability of the linear transformation in Banach spaces, J. Math. Soc. Japan, 2 (1950), 64–66. https://doi.org/10.2969/jmsj/00210064 doi: 10.2969/jmsj/00210064
    [4] T. M. Rassias, On the stability of the linear mapping in Banach spaces, Proc. Amer. Math. Soc., 72 (1978), 297–300.
    [5] P. Gavruta, A generalization of the Hyers-Ulam-Rassias stability of approximately additive mappings, J. Math. Anal. Appl., 184 (1994), 431–436. https://doi.org/10.1006/jmaa.1994.1211 doi: 10.1006/jmaa.1994.1211
    [6] A. Charifi, R. Lukasik, D. Zeglami, A special class of functional equations, Math. Slovaca, 68 (2018), 397–404. https://doi.org/10.1515/ms-2017-0110 doi: 10.1515/ms-2017-0110
    [7] H. A. Hammad, R. A. Rashwan, A. Nafea, M. E. Samei, S. Noeiaghdam, Stability analysis for a tripled system of fractional pantograph differential equations with nonlocal conditions, J. Vib. Control, 2023. https://doi.org/10.1177/10775463221149232 doi: 10.1177/10775463221149232
    [8] H. A. Hammad, H. Aydi, H. Isik, M. De la Sen, Existence and stability results for a coupled system of impulsive fractional differential equations with Hadamard fractional derivatives, AIMS Math., 8 (2023), 6913–6941. https://doi.org/10.3934/math.2023350 doi: 10.3934/math.2023350
    [9] H. A. Hammad, R. A. Rashwan, A. Nafea, M. E. Samei, M. De la Sen, Stability and existence of solutions for a tripled problem of fractional hybrid delay differential equations, Symmetry, 14 (2022), 2579. https://doi.org/10.3390/sym14122579 doi: 10.3390/sym14122579
    [10] H. A. Hammad, M. Zayed, Solving a system of differential equations with infinite delay by using tripled fixed point techniques on graphs, Symmetry, 14 (2022), 1388. https://doi.org/10.3390/sym14071388 doi: 10.3390/sym14071388
    [11] H. A. Hammad, P. Agarwal, S. Momani, F. Alsharari, Solving a fractional-order differential equation using rational symmetric contraction mappings, Fractal Fract., 5 (2021), 159. https://doi.org/10.3390/fractalfract5040159 doi: 10.3390/fractalfract5040159
    [12] R. A. Rashwan, H. A. Hammad, M. G. Mahmoud, Common fixed point results for weakly compatible mappings under implicit relations in complex valued G-metric spaces, Inform. Sci. Lett., 8 (2019), 111–119. https://doi.org/10.18576/isl/080305 doi: 10.18576/isl/080305
    [13] H. A. Hammad, M. De la Sen, Fixed-point results for a generalized almost (s, q)-Jaggi F-contraction-type on b-metric-like spaces, Mathematics, 8 (2020), 63. https://doi.org/10.3390/math8010063 doi: 10.3390/math8010063
    [14] H. Nakano, Modulared semi-ordered linear spaces, Tokyo: Maruzen Company, Ltd., 1950.
    [15] W. A. J. Luxemburg, Banach function spaces, Ph.D. Thesis, Delft, the Netherlands: Technische Hogeschool Delft, 1955.
    [16] I. Amemiya, On the representation of complemented modular lattices, J. Math. Soc. Japan, 9 (1957), 263–279. https://doi.org/10.2969/jmsj/00920263 doi: 10.2969/jmsj/00920263
    [17] J. Musielak, Orlicz spaces and nodular spaces, Berlin, Heidelberg: Springer, 1983.
    [18] S. Koshi, T. Shimogaki, On F-norms of quasi-modular spaces, J. Fac. Sci. Hokkaido Univ. Ser. 1 Math., 15 (1961), 202–218.
    [19] B. Mazur, Modular curves and the Eisenstein ideal, Publ. Math. l'IHÉS, 47 (1977), 33–186. https://doi.org/10.1007/BF02684339 doi: 10.1007/BF02684339
    [20] P. Turpin, Fubini inequalities and bounded multiplier property in generalized modular spaces, Comment. Math., 1 (1978), 331–353.
    [21] W. Orlicz, Collected papers. Part I, II, Warsaw, Poland: PWN-Polish Scientific Publishers, 1988.
    [22] L. Maligranda, Orlicz spaces and interpolation, Campinas: Universidade Estadual de Campinas, 1989.
    [23] M. A. Khamsi, Quasicontraction mappings in modular spaces without \Delta _{2}-condition, Fixed Point Theory Appl., 2008 (2008), 916187. https://doi.org/10.1155/2008/916187 doi: 10.1155/2008/916187
    [24] G. Sadeghi, A fixed point approach to stability of functional equations in modular spaces, Bull. Malays. Math. Sci. Soc., 37 (2014), 333–344.
    [25] H. M. Kim, H. Y. Shin, Refined stability of additive and quadratic functional equations in modular spaces, J. Inequal. Appl., 2017 (2017), 146. https://doi.org/10.1186/s13660-017-1422-z doi: 10.1186/s13660-017-1422-z
    [26] S. Gähler, 2-metrische Räume und ihre topologische struktur, Math. Nachr., 26 (1963), 115–148. https://doi.org/10.1002/mana.19630260109 doi: 10.1002/mana.19630260109
    [27] W. G. Park, Approximate additive mappings in 2-Banach spaces and related topics, J. Math. Anal. Appl., 376 (2011), 193–202. https://doi.org/10.1016/j.jmaa.2010.10.004 doi: 10.1016/j.jmaa.2010.10.004
    [28] K. Wongkum, P. Chaipunya, P. Kumam, On the generalized Ulam-Hyers-Rassias stability of quadratic mappings in modular spaces without \Delta _{2}-conditions, J. Funct. Spaces, 2015 (2015), 461719. https://doi.org/10.1155/2015/461719 doi: 10.1155/2015/461719
    [29] H. M. Kim, I. S. Chang, E. Son, Stability of Cauchy additive functional equation in fuzzy Banach spaces, Math. Inequal. Appl., 16 (2013), 1123–1136.
    [30] Z. Gajda, On stability of additive mappings, Int. J. Math. Math. Sci., 14 (1991), 431–434.
    [31] N. Uthirasamy, K. Tamilvanan, H. K. Nashine, R. George, Solution and stability of quartic functional equations in modular spaces by using Fatou property, J. Funct. Spaces, 2022 (2022), 5965628. https://doi.org/10.1155/2022/5965628 doi: 10.1155/2022/5965628
    [32] S. A. Mohiuddine, K. Tamilvanan, M. Mursaleen, T. Alotaibi, Stability of quartic functional equation in modular spaces via Hyers and fixed-point methods, Mathematics, 10 (2022), 1938. https://doi.org/10.3390/math10111938 doi: 10.3390/math10111938
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1305) PDF downloads(56) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog