
In this work, the notion of function of bounded variation in 2-normed spaces was established (Definition 4.2), the set of functions of bounded (2,k)-variation was endowed with a norm (Theorem 4.5), and it was proved that such a set is a Banach space (Theorem 4.6). In addition, the fundamental properties of the functions of bounded (2,k)-variation in the formalism of 2-Hilbert and 2-normed spaces were studied (see Theorems 4.1, 4.3, 4.4). Also, it was shown how to endow a 2-normed space with a function of bounded (2,k)-variation from a classical Hilbert space (Proposition 4.1). A series of examples and counterexamples are presented that enrich the results obtained in this work (4.1 and 4.2).
Citation: Cure Arenas Jaffeth, Ferrer Sotelo Kandy, Ferrer Villar Osmin. Functions of bounded (2,k)-variation in 2-normed spaces[J]. AIMS Mathematics, 2024, 9(9): 24166-24183. doi: 10.3934/math.20241175
[1] | Kaifang Liu, Lunji Song . A family of interior-penalized weak Galerkin methods for second-order elliptic equations. AIMS Mathematics, 2021, 6(1): 500-517. doi: 10.3934/math.2021030 |
[2] | 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 |
[3] | Tingting Wen . On the number of integers which form perfect powers in the way of $ x(y_1^2+y_2^2+y_3^2+y_4^2) = z^k $. AIMS Mathematics, 2024, 9(4): 8732-8748. doi: 10.3934/math.2024423 |
[4] | Cencen Dou, Jiagui Luo . Complete solutions of the simultaneous Pell's equations $ (a^2+2)x^2-y^2 = 2 $ and $ x^2-bz^2 = 1 $. AIMS Mathematics, 2023, 8(8): 19353-19373. doi: 10.3934/math.2023987 |
[5] | Changsheng Luo, Jiagui Luo . Complete solutions of the simultaneous Pell equations $ (a^2+1)y^2-x^2 = y^2-bz^2 = 1 $. AIMS Mathematics, 2021, 6(9): 9919-9938. doi: 10.3934/math.2021577 |
[6] | Tong Wei . The distribution of ideals whose norm divides $ n $ in the Gaussian ring. AIMS Mathematics, 2024, 9(3): 5863-5876. doi: 10.3934/math.2024285 |
[7] | Zaffar Iqbal, Xiujun Zhang, Mobeen Munir, Ghina Mubashar . Hilbert series of mixed braid monoid $ MB_{2, 2} $. AIMS Mathematics, 2022, 7(9): 17080-17090. doi: 10.3934/math.2022939 |
[8] | Zhihua Wang, Choonkil Park, Dong Yun Shin . Additive $ \rho $-functional inequalities in non-Archimedean 2-normed spaces. AIMS Mathematics, 2021, 6(2): 1905-1919. doi: 10.3934/math.2021116 |
[9] | Xiaofan Xu, Yongchao Xu . Some results on deep holes of generalized projective Reed-Solomon codes. AIMS Mathematics, 2019, 4(2): 176-192. doi: 10.3934/math.2019.2.176 |
[10] | Keun Young Lee, Gwanghyun Jo . The dual of a space of compact operators. AIMS Mathematics, 2024, 9(4): 9682-9691. doi: 10.3934/math.2024473 |
In this work, the notion of function of bounded variation in 2-normed spaces was established (Definition 4.2), the set of functions of bounded (2,k)-variation was endowed with a norm (Theorem 4.5), and it was proved that such a set is a Banach space (Theorem 4.6). In addition, the fundamental properties of the functions of bounded (2,k)-variation in the formalism of 2-Hilbert and 2-normed spaces were studied (see Theorems 4.1, 4.3, 4.4). Also, it was shown how to endow a 2-normed space with a function of bounded (2,k)-variation from a classical Hilbert space (Proposition 4.1). A series of examples and counterexamples are presented that enrich the results obtained in this work (4.1 and 4.2).
In 1881, Jordan [12] discovered in the work of Dirichlet the notion of function of bounded variation and proved that, for this class of functions, the Fourier conjecture is valid. He also showed that the function f:[a,b]→R has bounded variation on [a,b] if and only if f is the difference of monotone functions (nowadays, this result is known as Jordan's Representation Theorem).
The notion of function of bounded variation has been studied and generalized in different contexts, studying different structures and properties in spaces with research interest. For example, Chistyakov in [4,5,6] studied a concept of a function of bounded generalized variation in the sense of Jordan-Riesz-Orlicz for functions f:[a,b]→X, where X is a normed or metric space. More recently, the notion of functions of bounded variation in spaces with indefinite metric was introduced by Ferrer, Guzmán, Naranjo [10]. Furthermore, functions of bounded variation have multiple applications in various fields, for example, in image processing, Brokman, Burger, Gilboa [1] presented an analysis of the total-variation (TV) on non-Euclidean parametrized surfaces, a natural representation of the shapes used in 3D graphics. Among the results achieved in this research is a new method to generalize the convexity of sets from the plane to surfaces, derived by characterizing the TV eigenfunctions on surfaces. Additionally, Bugajewska, Bugajewski, Hudzik [2] investiged solutions of nonlinear Hammerstein and Volterra-Hammerstein integral equations in the space of functions of bounded ϕ-variation in the sense of Young. Later, Bugajewska, Bugajewski, Lewicki [3] explored the superposition operator and solutions to non linear integral equations in spaces of functions of generalized bounded ϕ-variation. In the field of nonlinear analysis, Xie, Liu, Li, Huang [18] examined the bounded variation capacity (BV capacity) and characterized the Sobolev-type inequalities associated with BV functions within a general framework of strictly local Dirichlet spaces with a doubling measure, utilizing the BV capacity.
On the other hand, the notion of a 2-normed vector space was introduced in 1963 by Siegfried Gähler [11] as a generalization of normed spaces. It has continued to be studied by several authors, among which the following stand out: White [17], who introduced the concept of 2-Banach space; Diminnie [8,9], which generated numerous results in the context of these spaces; and Lewandowska [14], who studied some properties of bounded 2-linear operators in a 2-normed set, endowed the set of bounded 2-linear operators of a norm, and proved the Banach-Steinhaus theorem for a family of bounded 2-linear operators.
In this paper, we extend the notion of bounded variation function in normed spaces to 2-normed spaces, endow the set of functions of bounded variation in 2-normed spaces with a norm, and extend the main properties of such functions. Moreover, in 2-normed spaces, we construct several examples of functions of bounded variation and prove the most relevant results for such functions.
Spaces with 2- inner product and spaces with 2-norm are extensions of spaces with inner product and normed spaces, respectively. So far, the theory of functions of bounded variation has not been studied in this context. Considering the historical importance of classical bounded variation functions, we consider that the lack of a detailed study of bounded variation functions in these new spaces provides a significant gap in the mathematical literature, preventing that the classical results of bounded variation functions from being extended to these more general spaces. Knowing that classical bounded variation functions have multiple applications, we consider that the exploration of bounded variation functions in spaces with 2-inner product and 2-normed product is as an important opportunity for the development of functional analysis and other related areas. In our research work, we aim to address this gap and answer two key questions:
(ⅰ) How to give a definition of bounded variation that allows to successfully relate the classical theory of functions of bounded variation to functions of bounded variation in 2-normed spaces?
(ⅱ) How appropriate are the concepts of (2,k)-variation and functions of bounded (2,k)-variation to formally extend some classical results of functions of bounded variation to 2-normed spaces?
Definition 3.1. [11,15] Let X to be a complex vector space of dimension d, where 2≤d≤∞. A 2-norm on X is a function ‖⋅,⋅‖:X×X→R that satisfies the following four conditions:
(2N1) ‖x,y‖=0 if and only if x and y are linearly dependent;
(2N1) ‖x,y‖=‖y,x‖;
(2N1) ‖x,αy‖=|α|‖x,y‖, for all α∈C;
(2N1) ‖x,y+z‖≤‖x,y‖+‖x,z‖.
If ‖⋅,⋅‖ is a 2-norm for X, then the pair (X,‖⋅,⋅‖) is called a 2-norm space.
Next, we will see an introductory example of 2-normed spaces.
Example 3.1. (Example 2-normed space) The set of all ordered triples of real numbers, R3, equipped with the function ‖⋅,⋅‖Δ:R3×R3→R, defined by
‖x,y‖Δ= the area of the triangle with vertices 0,x,y, |
it is a 2-norm (see Figure 1).
Indeed, let us see that ‖⋅,⋅‖Δ satisfies the conditions of being a norm for R3.
Recall that the area of the parallelogram generated by points three points A,B,C∈R3 is given by
S(A,B,C)=‖→AB×→AC‖. |
Let x,y,z∈R3 and α∈R.
(2N1) (⇒) If ‖x,y‖Δ=0, then ‖x×y‖2=0, i.e., x×y=0, which implies that {x,y} is linearly dependent.
(⇐) If {x,y} is linearly dependent, then x×y=0, therefore 0=‖0‖2=‖x×y‖2=‖x,y‖Δ.
(2N2) ‖x,y‖Δ=‖x×y‖2=‖−(y×x)‖2=‖y×x‖2=‖y,x‖Δ.
(2N3) ‖x,αy‖Δ=‖x×αy‖2=‖α(x×y)‖2=|α|‖x×y‖2=|α|⋅‖x,y‖Δ.
(2N4) ‖x,(y+z)‖Δ=‖x×(y+z)‖2=‖(x×y)+(x×z))‖2≤‖(x×y)‖+‖(x×z))‖2=‖x,y‖Δ+‖x,z‖Δ.
Proposition 3.1. [11] Let (X,‖⋅,⋅‖) be 2-norm space. For all x,y,z∈X and α∈C, it follows that:
(1) ‖x,y‖≥0.
(2) ‖x,y+αx‖=‖x,y‖.
Definition 3.2. [7] Let X be a complex vector space of dimension d≥2. A 2-inner product is a function
(⋅,⋅|⋅):X×X×X→C |
that satisfies the following conditions:
(2I1) (x1+x2,y|z)=(x1,y|z)+(x2,y|z).
(2I1) (x,x|z)=(z,z|x);
(2I1) (y,x|z)=¯(x,y|z);
(2I1) (αx,y|z)=α(x,y|z) for all α∈C;
(2I1) (x,x|z)≥0 and (x,x|z)=0 if and only if x and z are linearly dependent.
The pair (X,(⋅,⋅|⋅)) is called a 2-inner product space (or pre-2-Hilbert space).
Remark 3.1. Note that in (2I3), (y,x|z)=¯(x,y|z). If y=x, we obtain (x,x|z)= ¯(x,x|z), which guarantees (x,x|z)∈R.
Proposition 3.2. [7] Let (X,⟨⋅,⋅⟩) be space with inner product. Then, the following expression defines a 2-inner product on X.
(x,y|z)=|⟨x,y⟩⟨x,z⟩⟨z,y⟩⟨z,z⟩|=⟨x,y⟩⟨z,z⟩−⟨x,z⟩⟨z,y⟩. |
This 2-inner product is called a standard 2-inner product and guarantees that every inner product space can be viewed as a 2-inner product space.
Proposition 3.3. [7] Let (X,(⋅,⋅|⋅)) be space with 2-inner product and x,y,z∈X. Then, it holds that:
|(x,y|z)|2≤(x,x|z)(y,y|z). |
The inequality above is the analogue of the Cauchy-Schwarz inequality in spaces with inner product.
Remark 3.2. Given a space with 2-inner product (H,(⋅,⋅|⋅)), we can define a 2-norm for H, called the induced 2-norm of the 2-inner product given by
‖h,ξ‖:=√(h,h|ξ)h,ξ∈H. |
Example 3.2. Let us consider C3 with the usual inner product. We can induce a 2-inner product from this inner product by the function
(⋅,⋅|⋅):C3×C3×C3→C(a,b,k)↦(a,b|k)=⟨a,b⟩⟨k,k⟩−⟨a,k⟩⟨k,b⟩, a,b,k∈C3. |
According to Remark 3.2, the function
‖⋅,⋅‖:C3×C3→C(a,k)↦‖a,k‖=√(a,a|k), a,k∈C3. |
is a 2-norm. Then,
‖x,k‖:=√⟨x,x⟩⟨k,k⟩−⟨x,k⟩⟨k,x⟩=√⟨x,x⟩⟨k,k⟩−|⟨x,k⟩|2,x,k∈C3. |
Remark 3.3. According to [17], given (X,‖⋅,⋅‖) is a 2-normed space, the application ‖⋅,⋅‖:X×X→R is continuous. In particular, given k∈X, the function ‖⋅,k‖:X→R is continuous.
Definition 3.3. Let (X,‖⋅,⋅‖) be a 2-normed space, k∈X, [a,b] a closed interval. A function f:[a,b]→X is said to be (2,k)-bounded, if there exists α∈R+ such that for all x∈[a,b], it satisfies ‖f(x),k‖≤α.
Definition 3.4. [17] A sequence {xn}n∈N in a 2-normed space X is called a convergent sequence if there exists a x∈X such that the limn→∞‖xn−x,y‖=0 for all y∈X. If {xn}n∈N converges to x, we write xn→x and we call x the limit of {xn}n∈N.
Definition 3.5. [13] Let (X,‖⋅,⋅‖) be a 2-normed space, k∈X. A sequence {xn}n∈N in X is called k-Cauchy if for all ϵ>0, there exists N∈N, such that if m,n>N, then ‖xn−xm,k‖<ϵ.
Definition 3.6. [13] A 2-normed vector space in which every k-Cauchy sequence is a convergent sequence is called a (2,k)-Banach space.
Definition 3.7. [14] A space with 2-inner product (H,(⋅,⋅|⋅)) is said to be a 2-Hilbert space if it is complete with respect to the 2-norm induced by the 2-inner product. We reserve the letter H for 2-Hilbert spaces.
We now introduce the notion of (2,k)-variation of a function in 2-normed spaces.
Definition 4.1. ((2,k)-Variation of a function) Let (X,‖⋅,⋅‖) be a 2-normed space, k∈X, [a,b] a closed interval, and f:[a,b]→X a function. The (2,k)-variation of f over [a,b] denoted by Vba(f,X,k), is as:
Vba(f,X,k)=sup{n∑i=1‖f(ti)−f(ti−1),k‖:P={t0,t1,⋯,tn}∈P([a,b])}, |
Clearly, Vba(f,X,k)≥0.
Theorem 4.1. Let (X,‖⋅,⋅‖) be a 2-normed space, k,l∈X, α∈C and f:[a,b]→X a function. Then, we have that:
(i) Vba(αf,X,k)=|α|Vba(f,X,k).
(ii) Vba(f+g,X,k)≤Vba(f,X,k)+Vba(g,X,k).
(iii) Vba(f,X,k+l)≤Vba(f,X,k)+Vba(f,X,l).
Proof. Let P={t0,t1,⋯,tn} be a partition for the interval [a,b]. Using Definition 4.1, we obtain that
(i) Vba(αf,X,k)=sup{n∑i=1‖(αf)(ti)−(αf)(ti−1),k‖}=sup{n∑i=1‖α(f(ti)−f(ti−1)),k‖}=|α|sup{n∑i=1‖f(ti)−f(ti−1),k‖}=|α|Vba(f,X,k). |
Therefore, Vba(αf,X,k)=|α|Vba(f,X,k).
(ii) Vba(f,X,k)+Vba(g,X,k)=sup{n∑i=1‖f(ti)−f(ti−1),k‖}+sup{n∑i=1‖g(ti)−g(ti−1),k‖}=sup{n∑i=1‖f(ti)−f(ti−1),k‖+‖g(ti)−g(ti−1),k‖}≥sup{n∑i=1‖(f(ti)−f(ti−1))+(g(ti)−g(ti−1)),k‖}=sup{n∑i=1‖(f(ti)+g(ti))−(f(ti−1)+g(ti−1)),k‖}=Vba(f+g,X,k). |
Therefore, Vba(f+g,X,k)≤Vba(f,X,k)+Vba(g,X,k).
Remark 4.1. Note that, in particular, for the constant function g(t)=λk, where λ∈C, we have that
Vba(f+λk,X,k)=sup{n∑i=1‖(f+λk)(ti)−(f+λk)(ti−1),k‖}=sup{n∑i=1‖f(ti)+λk−f(ti−1)−λk,k‖}=sup{n∑i=1‖f(ti)−f(ti−1),k‖}=Vba(f,X,k). |
(iii) Vba(f,X,k)+Vba(f,X,l)=sup{n∑i=1(‖(f(ti)−f(ti−1)),k‖)}+sup{n∑i=1(‖(f(ti)−f(ti−1)),l‖)}=sup{n∑i=1(‖(f(ti)−f(ti−1)),k‖+‖(f(ti)−f(ti−1)),l‖)}≥sup{n∑i=1‖f(ti)−f(ti−1),k+l‖}=Vba(f,X,k+l). |
This proves that Vba(f,X,k+l)≤Vba(f,X,k)+Vba(f,X,l).
Next, we introduce the notion of a function of bounded (2,k)-variation in 2-normed spaces.
Definition 4.2. Let (X,‖⋅,⋅‖) be a 2-normed space and k∈X. A function f:[a,b]→X is said to be of bounded (2,k)-variation on [a,b] if Vba(f,X,k) is finite; in other words, if there exists a constant c∈R+ such that Vba(f,X,k)≤c.
The set of all functions of bounded (2,k)-variation on [a,b] are denoted by BV([a,b],2,k); in other words,
BV([a,b],2,k)={f:[a,b]→X:Vba(f,X,k)<∞}. |
Example 4.1. Consider the 2-normed space (C3,‖⋅,⋅‖) given in Example 3.2. Let us see that f:[a,b]→C3 defined by f(x)=(xi,2xi,3xi) is of bounded (2,k)-variation.
Consider P={a,t1,⋯,tn−1,b} a partition of [a,b].
On one side
f(tj)−f(tj−1)=(tji,2tji,3tji)−(tj−1i,2tj−1i,3tj−1i)=((tj−tj−1)i,(2tj−2tj−1)i,(3tj−3tj−1)i). |
Thus,
Vba(f,C3,k)=sup{n∑j=1√⟨f(tj)−f(tj−1),f(tj)−f(tj−1)⟩⟨k,k⟩−|⟨f(tj)−f(tj−1),k⟩|2}. |
Then, ⟨f(tj)−f(tj−1),f(tj)−f(tj−1)⟩ is equal to
⟨((tj−tj−1)i,(2tj−2tj−1)i,(3tj−3tj−1)i),((tj−tj−1)i,(2tj−2tj−1)i,(3tj−3tj−1)i)⟩=(tj−tj−1)(i)¯(tj−tj−1)(i)+(2tj−2tj−1)(i)¯(2tj−2tj−1)(i)+(3tj−3tj−1)(i)¯(3tj−3tj−1)(i)=|(tj−tj−1)|2+|(2tj−2tj−1)|2+|(3tj−3tj−1)|2=|(tj−tj−1)|2+|2|2|(tj−tj−1)|2+|3|2|(tj−tj−1)|2=14|(tj−tj−1)|2. |
On the other hand,
⟨k,k⟩=k1¯k1+k2¯k2+k3¯k3=|k1|2+|k2|2+|k3|2. |
In addition,
|⟨f(tj)−f(tj−1),k⟩|2=|⟨((tj−tj−1)i,(2tj−2tj−1)i,(3tj−3tj−1)i),(k1,k2,k3)⟩|2=|(tj−tj−1)i¯(k1)+(2tj−2tj−1)i¯(k2)+(3tj−3tj−1)i¯(k3)|2=|(tj−tj−1)(¯(k1)+2¯(k2)+3¯(k3))i|2=|(tj−tj−1)(¯(k1)+2¯(k2)+3¯(k3))|2. |
Thus, we obtain that
Vba(f,C3,k)=sup{n∑j=1√(14|(tj−tj−1)|2)(|k1|2+|k2|2+|k3|2)−|(tj−tj−1)(¯k1+2¯k2+3¯k3)|2}=sup{n∑j=1|(tj−tj−1)|√14(|k1|2+|k2|2+|k3|2)−|¯k1+2¯k2+3¯k3|2}. |
Taking √14(|k1|2+|k2|2+|k3|2)−|¯k1+2¯k2+3¯k3|2=M, we have that
Vba(f,C3,k)=Msup{n∑j=1|(tj−tj−1)|}=Msup{n∑j=1(tj−tj−1)}=M(b−a). |
Taking 2M(b−a)=ξ, we have that
Vba(f,C3,k)≤ξ |
Therefore f:[a,b]→C3 is of bounded (2,k)-variation.
In the following result, it is shown that a bounded variation function in an inner product space can be considered as a function of bounded (2,k)-variation with respect to the induced standard 2-inner product space.
Proposition 4.1. Let (X,⟨⋅,⋅⟩) be a space with an inner product, k∈X and f:[a,b]→X a function. If f is of bounded variation on (X,⟨⋅,⋅⟩), then f is of bounded (2,k)-variation under the 2-norm induced by the standard 2-inner product.
Proof. Let P={t0,t1,⋯,tn}∈P([a,b]). Consider the 2-norm defined in Example 3.2. If f is of bounded variation in (X,⟨⋅,⋅⟩), then there exists α∈R+ such that
Vba(f,X)=sup{n∑i=1‖f(ti)−f(ti−1)‖}≤α. |
Then,
‖f(ti)−f(ti−1),k‖=√⟨f(ti)−f(ti−1),f(ti)−f(ti−1)⟩⟨k,k⟩−|⟨f(ti)−f(ti−1),k⟩|2≤√⟨f(ti)−f(ti−1),f(ti)−f(ti−1)⟩⟨k,k⟩=√⟨f(ti)−f(ti−1),f(ti)−f(ti−1)⟩√⟨k,k⟩=‖f(ti)−f(ti−1)‖⋅‖k‖. |
Thus,
Vba(f,X,k)=sup{n∑i=1‖f(ti)−f(ti−1),k‖}≤sup{n∑i=1‖f(ti)−f(ti−1)‖⋅‖k‖}=‖k‖sup{n∑i=1‖f(ti)−f(ti−1)‖}≤‖k‖α. |
Therefore, f is of bounded (2,k)-variation with the standard 2-inner product.
Proposition 4.2. Let (X,‖⋅,⋅‖) be a 2-normed space, k∈X and f:[a,b]→X a function of bounded (2,k)-variation on [a,b]. If Vba(f,X,k)=0, then f(x)=mk+12(f(a)+f(b)) for all x∈(a,b).
Proof. Suppose that Vba(f,X,k)=0, i.e.,
Vba(f,X,k)=sup{n∑i=1‖f(ti)−f(ti−1),k‖:P∈P[a,b]}=0. |
Hence, for any participation of [a,b], we have
n∑i=1‖f(ti)−f(ti−1),k‖=0. | (4.1) |
Consider x∈(a,b). Then, in particular, for the partition P={a,x,b} of [a,b]. By (4.1), we have that
‖f(x)−f(a),k‖+‖f(b)−f(x),k‖=0. |
Thus,
‖f(x)−f(a),k‖=0 and ‖f(b)−f(x),k‖=0. |
Therefore,
{f(x)−f(a),k} is linearly dependent and {f(b)−f(x),k} is linearly dependent. |
That is, there exist α,β∈C non-null, such that
f(x)−f(a)=αk and f(b)−f(x)=βk. |
Then,
f(x)−f(a)+f(x)−f(b)=αk−βk, |
Thus,
f(x)=α−β2k+12(f(a)+f(b)) for all x∈(a,b). |
Therefore, f(x)=mk+12(f(a)+f(b)) for all x∈(a,b).
Theorem 4.2. Let (X,‖⋅,⋅‖) be a 2-normed space and k∈X. If f:[a,b]→X is a function of bounded (2,k)-variation, then f is (2,k)-bounded.
Proof. Since f is of bounded (2,k)-variation, then there exists M>0 such that
Vba(f,X,k)=sup{n∑i=1‖f(ti)−f(ti−1),k‖:P∈P[a,b]}≤M, |
Consider x∈(a,b) and also consider the partition P={a,x,b} of [a,b]. Then, in particular,
sup{‖f(x)−f(a),k‖+‖f(b)−f(x),k‖}≤M |
Then,
‖f(x),k‖=12‖2f(x),k‖=12‖f(x)+0+f(x),k‖=12‖(f(x)−f(a))+(f(x)−f(b))+(f(a)+f(b)),k‖≤12(‖f(x)−f(a),k‖+‖f(x)−f(b),k‖+‖f(a),k‖+‖f(b),k‖)≤‖f(x)−f(a),k‖+‖f(b)−f(x),k‖+‖f(a),k‖+‖f(b),k‖≤M+‖f(a),k‖+‖f(b),k‖. |
Taking M+‖f(a),k‖+‖f(b),k‖=α, we have that
‖f(a),k‖,‖f(b),k‖≤α. |
Thus, there exists α∈R+ such that for all x∈[a,b], we have that ‖f(x),k‖≤α. Therefore, f is (2,k)-bounded.
The reciprocal of the above theorem is not always true.
Example 4.2. Let us consider the space (C3,‖⋅,⋅‖) and the 2-norm defined in Example 3.2 by
‖x,k‖:=√⟨x,x⟩⟨k,k⟩−|⟨x,k⟩|2, x∈C3. |
Let us see that f:[√11,4]→C3 defined by
f(t)={(i,i,i), if t isrational, t∈[√11,4], (0,0,0), if t isirrational, t∈[√11,4]. |
and k=(0,−i,−i)∈C3.
Lets see that f is (2,k)-bounded, but it is not of bounded (2,k)-variation.
Indeed, if t is rational,
‖f(t),(0,−i,−i)‖=‖(i,i,i),(0,−i,−i)‖=√⟨(i,i,i),(i,i,i)⟩⟨(0,−i,−i),(0,−i,−i)⟩−|⟨(i,i,i),(0,−i,−i)⟩|2 |
Now,
⟨(i,i,i),(i,i,i)⟩=i¯i+i¯i+i¯i=3i¯i=3i(−i)=3 |
In addition,
⟨(0,−i,−i),(0,−i,−i)⟩=0−i¯(−i)−i¯(−i)=1+1=2 |
Also,
|⟨(i,i,i),(0,−i,−i)⟩|2=|i¯0−1−1|2=|−2|2=4 |
Therefore,
‖f(t),(0,−i,−i)‖=√3(2)−4=√2 |
Hence,
‖f(t),(0,−i,−i)‖={√2, if t isrational, t∈[√11,4], 0, if t isirrational, t∈[√11,4]. |
Therefore, ‖f(t),(0,−i,−i)‖≤√2 for all t∈[√11,4]. Thus, f is (2,k)-bounded in [√11,4].
Now, let us see that f is not of (2,k)-bounded variation.
Consider t0=√11, since between any two reals there exists a rational and an irrational number, we can choose t1 as a rational number between √11 and 4, t2 as an irrational number between t1 and 4, t3 as a rational number between t2 and 4, and so on; t2j would be an irrational number between t2j−1 and 4, t2j+1 would be a rational number between t2j and 4. Finally we choose tn=4. Then,
V4√11(f,X,(0,−i,−i))=sup{n∑j=1‖f(tj)−f(tj−1),k‖:P∈P[a,b]}≥n∑j=1‖f(tj)−f(tj−1),(0,−i,−i)‖=‖f(t1)−f(t0),(0,−i,−i)‖+⋯+‖f(tn)−f(tn−1),(0,−i,−i)‖=‖−(i,i,i),(0,−i,−i)‖+⋯+‖(i,i,i),(0,−i,−i)‖=‖(i,i,i),(0,−i,−i)‖+⋯+‖(i,i,i),(0,−i,−i)‖=√2+⋯+√2=√2n. |
Note that a partition of the interval [√11,4] was constructed starting with √11 and then alternating between rational and irrational numbers until it ends at 4, for which V4√11(f,C3,(0,−i,−i)) is not finite. Therefore, f is not of bounded (2,k)-variation in (C3,‖⋅,⋅‖).
Theorem 4.3. Let (X,‖⋅,⋅‖) be a 2-normed space, k∈X and f,g:[a,b]→X functions of bounded (2,k)-variation on [a,b]. Then, αf and f+g are also functions of bounded (2,k)-variation over [a,b].
Proof. Since f and g are of bounded (2,k)-variation, then there exist λ,β∈R+ such that
Vba(f,X,k)=sup{n∑i=1‖f(ti)−f(ti−1),k‖:P∈P[a,b]}≤λ |
and
Vba(g,X,k)=sup{n∑i=1‖g(ti)−g(ti−1),k‖:P∈P[a,b]}≤β |
(ⅰ) From Theorem 4.1 we have that
Vba(αf,X,k)=|α|Vba(f,X,k). |
Also, |α|∈R+, thus
Vba(αf,X,k)=|α|Vba(f,X,k)=|α|sup{n∑i=1‖f(ti)−f(ti−1),k‖:P∈P[a,b]}≤|α|λ. |
Taking |α|λ=σ,
Vba(αf,X,k)≤σ. |
Therefore, αf is of bounded (2,k)-variation on [a,b].
Remark 4.2. Note that if we take α=−1, then −f is also of bounded (2,k)-variation.
(ⅱ) Consider P∈P[a,b]. Taking λ+β=σ, we have that
σ=λ+β≥Vba(f,X,k)+Vba(g,X,k)=sup{n∑i=1‖f(ti)−f(ti−1),k‖}+sup{n∑i=1‖g(ti)−g(ti−1,k)‖}=sup{n∑i=1‖f(ti)−f(ti−1),k‖+‖g(ti)−g(ti−1),k‖}≥sup{n∑i=1‖(f(ti)−f(ti−1))+(g(ti)−g(ti−1)),k‖}=sup{n∑i=1‖(f(ti)+g(ti))−(f(ti−1)+g(ti−1)),k‖}=Vba(f+g,X,k). |
Therefore, f+g is of bounded (2,k)-variation on [a,b].
Remark 4.3. From the theorem above, it follows that the set BV([a,b],X,k) is a vector space.
Theorem 4.4. Let (X,‖⋅,⋅‖) be a 2-normed space, k∈X, f:[a,b]→X a function of bounded (2,k)-variation on [a,b] and c∈(a,b). Then, it is satisfied that f is of bounded (2,k)-variation in [a,c] and in [c,b]. Furthermore, Vca(f,X,k)+Vbc(f,X,k)≤Vba(f,X,k).
Proof. If f is of bounded (2,k)-variation, then there exists M>0 such that for any partition P of [a,b] it follows that
Vba(f,X,k)=sup{n∑i=1‖f(ti)−f(ti−1),k‖:P∈P[a,b]}≤M. |
Let us consider the partition P={a,t1,t2,t3,⋯,tn−1,b}. Since c∈(a,b), then
P={a,t1,t2,t3,⋯,tp−1,c,tp+1,⋯,tn−1,b} |
Then,
p∑i=1‖f(ti)−f(ti−1),k‖≤p∑i=1‖f(ti)−f(ti−1),k‖+n∑i=p‖f(ti)−f(ti−1),k‖=n∑i=1‖f(ti)−f(ti−1),k‖≤sup{n∑i=1‖f(ti)−f(ti−1),k‖:P∈P[a,b]}≤M. |
Therefore, M is an upper bound of the set
{p∑i=1‖f(ti)−f(ti−1),k‖:P∈P[a,c]}. |
Then,
Vca(f,X,k)=sup{p∑i=1‖f(ti)−f(ti−1),k‖:P∈P[a,c]}≤M. |
Therefore, f is of bounded (2,k)-variation on the interval [a,c]. Similarly, it is proved that f is of bounded (2,k)-variation on the interval [c,b].
Let us consider t∈(a,b] and the partitions Π[a,c],Π[c,b], given by:
Π[a,c]={a,t1,⋯,tk−1,c} |
and
Π[c,b]={c,s1,⋯,sm−1,b}. |
As Π[a,c]⊂Π[a,b] and Π[c,b]⊂Π[a,b], it follows that
supΠ[a,c]{n∑i=1‖f(ti)−f(ti−1),k‖}+supΠ[b,c]{n∑i=1‖f(ti)−f(ti−1),k‖}≤supΠ[a,b]{n∑i=1‖f(ti)−f(ti−1),k‖} |
Where the supremum is chosen over the set of all partitions Π[a,b].
Consequently, by Definition 4.1., we conclude that
Vca(f,X,k)+Vbc(f,X,k)≤Vba(f,X,k). |
Proposition 4.3. The following relation ∼ defined in BV([a,b],2,k) is an equivalence relation.
f∼g⟺f−g=αk,forsomeα∈C |
Proof. ⅰ) There exists 0∈C such that f−f=0k, so f∼f. Therefore, ∼ is reflexive.
ⅱ) If f∼g, then f−g=αk for some α∈C, then, there exists β=−α∈C such that g−f=βk, so g∼f. Therefore, ∼ is symmetric.
ⅲ) If f∼g and g∼h, then, there exists λ,γ∈C such that f−g=λk and g−h=γk, then f−h=γk+λk, so f−h=(γ+λ)k, where γ+λ∈C so f∼h. Therefore, ∼ is transitive.
By virtue of the proposition above, the equivalence class of f∈BV([a,b],2,k) under the relation ∼ is given by:
[f]={g∈BV([a,b],2,k):f∼g}. |
In addition,
λ[f]={λg∈BV([a,b],2,k):f∼g}. |
Remark 4.4. [f+g]=[f]+[g] and λ[f]=[λf].
Remark 4.5. The set of all equivalence classes under the relation ∼ given in Proposition 4.3 will be noted by BV∼k[a,b]. This is
BV∼k[a,b]={[f]:f∈BV([a,b],2,k)} |
Next, we will endow the set BV∼k[a,b] of functions of (2,k)-bounded variation of a norm.
Theorem 4.5. Let (X,‖⋅,⋅‖) be a 2-normed space, k∈X. The functional ‖⋅‖BV∼k[a,b]:BV∼k[a,b]→R defined by
‖[f]‖BV∼k[a,b]=‖f(a),k‖+‖f(b),k‖+Vba(f,X,k),f∈BV([a,b],2,k), |
is a norm for the set BV∼k[a,b].
Proof. Let [f],[g]∈BV∼k[a,b] and λ∈C, then
(N1) From the definition of 2-norm and (2,k)-variation, we have that ‖f(a),k‖≥0, ‖f(b),k‖≥0 and Vba(f,X,k)≥0, obtaining that
‖[f]‖BV∼k[a,b]=‖f(a),k‖+‖f(b),k‖+Vba(f,X,k)≥0. |
(N2) (⟹) Let us suppose that ‖[f]‖BV∼k[a,b]=0, then
‖[f]‖BV∼k[a,b]=‖f(a),k‖+‖f(b),k‖+Vba(f,X,k)=0. |
Then, ‖f(a),k‖=0, ‖f(b),k‖=0 and Vba(f,X,k)=0. Thus
f(a)=αk and f(b)=βk |
for some α,β∈C. From Vba(f,X,k)=0, using Proposition 4.2 and taking into account that f(a)=αk and f(b)=βk, we have that for all t∈[a,b], there exists γ∈C such that f(t)=γk.
Therefore, f∼0. Then, [f]=[0].
(⟸) Lets suppose that [f]=[0], since f∈[f]=[0], then f∈[0]; in other words, there exists γ∈C such that f(x)−0=γk for all x∈[a,b], so
‖[f]‖BV∼k[a,b]=‖f(a),k‖+‖f(b),k‖+Vba(f,X,k)=‖γk,k‖+‖γk,k‖+Vba(γk,X,k)=0+0+0=0. |
Therefore, ‖[f]‖BV∼k[a,b]=0.
(N3)
‖[λf]‖BV∼k[a,b]=‖λf(a),k‖+‖λf(b),k‖+Vba(λf,X,k)=|λ|‖f(a),k‖+|λ|‖f(b),k‖+|λ|Vba(f,X,k)=|λ|(‖f(a),k‖+‖f(b),k‖+Vba(f,X,k))=|λ|⋅‖[f]‖BV∼k[a,b]. |
(N4) As BV∼k[a,b] is a vector space, then, [f]+[g]∈BV∼k[a,b]. Also,
‖[f]‖BV∼k[a,b]+‖[g]‖BV∼k[a,b]=(‖f(a),k‖+‖f(b),k‖+Vba(f,X,k))+(‖g(a),k‖+‖g(b),k‖+Vba(g,X,k))=‖f(a),k‖+‖g(a),k‖+‖f(b),k‖+‖g(b),k‖+Vba(f,X,k)+Vba(g,X,k)≥‖f(a)+g(a),k‖+‖f(b)+g(b),k‖+Vba(f+g,X,k)=‖[f]+[g]‖BV∼k[a,b]. |
Therefore, ‖[f]+[g]‖BV∼k[a,b]≤‖[f]‖BV∼k[a,b]+‖[g]‖BV∼k[a,b]. Thus, it is proved that ‖⋅‖BV∼k[a,b] is a norm for BV∼k[a,b].
Proposition 4.4. Let f,g∈BV∼k[a,b]. If f∼g, then ‖[f]‖BV∼k[a,b]=‖[g]‖BV∼k[a,b].
Proof. If f∼g, then there exists α∈C such that f−g=αk. Thus,
‖[f]‖BV∼k[a,b]=‖g(a)+αk,k‖+‖g(b)+αk,k‖+Vba(g+αk,X,k)=‖g(a),k‖+‖g(b),k‖+Vba(g,X,k)=‖[g]‖BV∼k[a,b]. |
Theorem 4.6. If (X,‖⋅,⋅‖) is a (2,k)-Banach space, then the space (BV∼k[a,b],‖⋅‖BV∼k[a,b]) of all functions of bounded (2,k)-variation on [a,b], is a Banach space.
Proof. Let {[fn]}n∈N a Cauchy sequence in (BV∼k[a,b],‖⋅‖BV∼k[a,b]) and ϵ>0. Then, there exists N∈N such that if m,n≥N, we have that
‖[fn−fm]‖BV∼k[a,b]=‖[fn]−[fm]‖BV∼k[a,b]<ϵ. | (4.2) |
Since {[fn]}n∈N is of Cauchy, then it is bounded, so there exists M∈R+ such that for all n∈N
‖[fn]‖BV∼k[a,b]<M. | (4.3) |
From (4.2) we have that
‖(fn−fm)(a),k‖+‖(fn−fm)(b),k‖+Vba(fn−fm,X,k)=‖[fn−fm]‖BV∼k[a,b]<ϵ. | (4.4) |
Thus,
‖fn(a)−fm(a),k‖=‖(fn−fm)(a),k‖<ϵ and ‖fn(b)−fm(b),k‖=‖(fn−fm)(b),k‖<ϵ; |
which ensures that {fn(a)}n∈N and {fn(b)}n∈N are sequences k-Cauchy at (X,‖⋅,⋅‖).
On the other hand, from the inequality (4.4), we have that
Vba(fn−fm,X,k)<ϵ. | (4.5) |
Suppose t∈(a,b). Let us consider the partition P={a,t,b}. Using the Definition 4.1 and the inequality (4.5), we have
‖(fn−fm)(t)−(fn−fm)(a),k‖+‖(fn−fm)(b)−(fn−fm)(t),k‖≤Vba(fn−fm,X,k)<ϵ. |
From this, we have that
‖(fn−fm)(t),k‖−‖(fn−fm)(a),k‖≤‖(fn−fm)(t)−(fn−fm)(a),k‖<ϵ |
Then,
‖(fn−fm)(t),k‖<ϵ+‖(fn−fm)(a),k‖ |
Thus,
‖(fn−fm)(t),k‖<2ϵ. |
As a result, the sequence {fn(t)}n∈N is a uniform sequence k-Cauchy in X.
Since X is (2,k)-Banach, there exists a function f:[a,b]→X such that:
f(t)=limn→∞fn(t),t∈[a,b]. |
Consequently, for all ϵ>0 and t∈[a,b], there exists N∈N such that if n≥N we have that
‖fn(t)−f(t),k‖<ϵ. | (4.6) |
Similarly, from (4.5), we obtain
limm→∞Vba(fn−fm,X,k)<ϵ. |
In other words,
Vba(fn−f,X,k)<ϵ. | (4.7) |
Let us see that [f]∈BV∼k[a,b].
n∑j=1‖f(tj)−f(tj−1),k‖=n∑j=1‖f(tj)−f(tj−1)+(fn(tj)−fn(tj−1))−(fn(tj)−fn(tj−1)),k‖≤n∑j=1‖(f−fn)(tj)−(f−fn)(tj−1),k‖+n∑j=1‖fn(tj)−fn(tj−1),k‖≤n∑j=1‖(f−fn)(tj),k‖+n∑j=1‖(f−fn)(tj−1),k‖+n∑j=1‖fn(tj)−fn(tj−1),k‖. |
Using the inequalities (4.7) and (4.3), we have that for all ϵ>0, n∑j=1‖f(tj)−f(tj−1),k‖<ϵ+M. So, n∑j=1‖f(tj)−f(tj−1),k‖≤N. Hence [f]∈BV∼k[a,b].
Now, of the inequalities (4.6) and (4.7) we have ‖(fn−f)(a),k‖<ϵ, ‖(fn−f)(b),k‖<ϵ and Vba(fn−f,X,k)<ϵ. Therefore,
‖[fn−f]‖BV∼k[a,b]=‖(fn−f)(a),k‖+‖(fn−f)(b),k‖+Vba(fn−f,X,k)<ϵ |
Thus, {[fn]}n∈N converges to [f] with the norm ‖⋅‖BV∼k[a,b]. Therefore, the set BV∼k[a,b] of all functions of bounded (2,k)-variation in [a,b] is a Banach space.
The notion of a function of bounded (2,k)-variation in 2-normed spaces presented in Definition 4.2 is a generalization of the functions of bounded variation in normed spaces. Moreover, any function of bounded variation in a space with inner product is a function of bounded (2,k)-variation in the 2-normed space induced by the standard 2-inner product (Proposition 4.1). In this context, we envision that the functions of generalized bounded ϕ-variation on spaces with inner product can be viewed as functions of generalized bounded ϕ-variation in the context of the theory of spaces with 2-inner product. Therefore, we believe that the results obtained by Bugajewska et al. in [2,3] can be extended to the context of the theory of 2-normed spaces. On the other hand, if f a function of (2,k)-variation null on [a,b], then the restriction of f to the interval (a,b), f|(a,b), is of the form mk+12(f(a)+f(b)) (Proposition 4.2). Also, every function of bounded (2,k)-variation is (2,k)-bounded, but the reciprocal is not true in general, as can be seen in the counterexample (Example 4.2). The set of functions of bounded (2,k)-variation over a (2,k)-Banach space is a Banach space (Theorem 4.6).
Yazdi, Zarei, Adumene, Abbassi and Rahnamayiezekavat [19] characterized and addressed uncertainty in digitized process systems, which are essential and unavoidable today. These must be integrated with the design, control, and modeling supports of feasible and optimal engineering systems. The classical theory of functions of bounded variation represents an effective option to address these problems. For example, Weerasinghe [16] considered an infinite horizon discounted cost minimization problem for a one-dimensional stochastic differential equation model. Although the operating cost function is not necessarily convex, sufficient conditions are obtained to guarantee zero-control optimality due to the inclusion of a functions of bounded variation process. Considering that classical functions of bounded variation can be viewed as functions of bounded (2,k)-variation, our study represents a novel option for conducting research such as [16,19].
Cure Jaffeth, Ferrer Kandy and Ferrer Osmin: Conceptualization, methodology, validation, software, writing-original draft, writing-review & editing. All authors contributed equally to the manuscript. All authors have read and approved the final version of the manuscript for publication.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors declare that they have no conflict of interest in this work.
[1] |
J. Brokman, M. Burger, G. Gilboa, Spectral total-variation processing of shapes-theory and applications, ACM T. Graphic., 43 (2024), 1–20. https://doi.org/10.1145/3641845 doi: 10.1145/3641845
![]() |
[2] |
D. Bugajewska, D. Bugajewski, H. Hudzik, BVϕ-solutions of nonlinear integral equations, J. Math. Anal. Appl., 287 (2003), 265–278. https://doi.org/10.1016/S0022-247X(03)00550-X doi: 10.1016/S0022-247X(03)00550-X
![]() |
[3] |
D. Bugajewska, D. Bugajewski, G. Lewicki, On nonlinear integral equations in the space of functions of bounded generalized ϕ-variation, J. Integral Equ. Appl., 21 (2009), 1–20. https://doi.org/10.1216/JIE-2009-21-1-1 doi: 10.1216/JIE-2009-21-1-1
![]() |
[4] |
V. V. Chistyakov, On mappings of bounded variation, J. Dyn. Control Syst., 3 (1997), 261–289. https://doi.org/10.1007/BF02465896 doi: 10.1007/BF02465896
![]() |
[5] |
V. V. Chistyakov, On the theory of multivalued mappings of bounded variation of one real variable, Sb. Math., 189 (1998), 797–819. https://doi.org/10.1070/SM1998v189n05ABEH000321 doi: 10.1070/SM1998v189n05ABEH000321
![]() |
[6] |
V. V. Chistyakov, Metric-valued mappings of bounded variation, J. Math. Sci., 111 (2002), 3387–3429. https://doi.org/10.1023/A:1016054010760 doi: 10.1023/A:1016054010760
![]() |
[7] | Y. J. Cho, Theory of 2-inner product spaces, Nova Publishers, 2001. |
[8] |
C. Diminnie, S. Gähler, A. White, Strictly convex linear 2-normed spaces, Math. Nachr., 59 (1974), 319–324. https://doi.org/10.1002/mana.19740590127 doi: 10.1002/mana.19740590127
![]() |
[9] |
C. Diminnie, S. Gähler, A. White, Remarks on strictly convex and strictly 2-convex 2-normed spaces, Math. Nachr., 88 (1979), 363–372. https://doi.org/10.1002/mana.19790880128 doi: 10.1002/mana.19790880128
![]() |
[10] |
O. Ferrer, C. Guzmán, J. Naranjo, Strongly bounded variation functions in Krein spaces, J. Math. Comput. Sci., 36 (2025), 237–250. https://doi.org/10.22436/jmcs.036.02.08 doi: 10.22436/jmcs.036.02.08
![]() |
[11] |
S. Gähler, Lineare 2-normierte Räume, Math. Nachr., 28 (1964), 1–43. https://doi.org/10.1002/mana.19640280102 doi: 10.1002/mana.19640280102
![]() |
[12] | C. Jordan, Sur la serie de Fourier, CR Acad. Sci. Paris, 92 (1881), 228–230. |
[13] | H. Mazaheri, R. Kazemi, Some results on 2-inner product spaces, Novi Sad J. Math., 37 (2007), 35–40. |
[14] |
Z. Lewandowska, Bounded 2-linear operators on 2-normed sets, Galsnik Mathematicki, 39 (2004), 303–314. https://doi.org/10.3336/gm.39.2.11 doi: 10.3336/gm.39.2.11
![]() |
[15] |
S. N. Lal, S. Bhattacharya, C. Sreedhar, Complex 2-normed linear spaces and extension of linear 2-functionals, Z. Anal. Anwend., 20 (2001), 35–53. https://doi.org/10.4171/ZAA/1003 doi: 10.4171/ZAA/1003
![]() |
[16] |
A. Weerasinghe, A bounded variation control problem for diffusion processes, SIAM J. Control Optim., 44 (2005), 389–417. https://doi.org/10.1137/S0363012903436119 doi: 10.1137/S0363012903436119
![]() |
[17] |
A. G. White, 2-Banach spaces, Math. Nachr., 42 (1967), 43–60. https://doi.org/10.1002/mana.19690420104 doi: 10.1002/mana.19690420104
![]() |
[18] |
X. Xie, Y. Liu, P. Li, J. Huang, The bounded variation capacity and Sobolev-type inequalities on Dirichlet spaces, Adv. Nonlinear Anal., 13 (2024), 20230119. https://doi.org/10.1515/anona-2023-0119 doi: 10.1515/anona-2023-0119
![]() |
[19] |
M. Yazdi, E. Zarei, S. Adumene, R. Abbassi, P. Rahnamayiezekavat, Uncertainty modeling in risk assessment of digitalized process systems, Method. Chem. Process Saf., 6 (2022), 389–416. https://doi.org/10.1016/bs.mcps.2022.04.005 doi: 10.1016/bs.mcps.2022.04.005
![]() |