Hyperbolic Functions of Bounded Variation and Riemann-Stieltjes Integral involving Strong Partitions of Hyperbolic Intervals
E-mail: [email protected]
(⋆) Escuela Superior de Ingeniería Mecánica y Eléctrica. Instituto Politécnico Nacional. Edif. 5, 3er piso, U.P. Adolfo López Mateos. 07338, Mexico City, MEXICO
Email: [email protected])
Abstract
In this paper, we define two types of partitions of an hyperbolic interval: weak and strong. Strong partitions enables us to define, in a natural way, a notion of hyperbolic valued functions of bounded variation and hyperbolic analogue of Riemann-Stieltjes integral. We prove a deep relation between both concepts like it occurs in the context of real analysis.
Math Subject Classification (2020):: 30G35, 28B15, 26B15.
Keywords: Hyperbolic numbers, partitions, bounded variation functions, Riemann-Stieltjes integral.
1 Introduction
Literature on hyperbolic numbers like, for instance, [1, 2] highlights that these numbers themselves are not so complicated and have hybrid behavior between real numbers and complex numbers. The hyperbolic numbers born as a real sub-algebra of Tessarine numbers introduced by J. Cockle in [3] who focused the study in an hypercomplex analysis context with a Cauchy-Riemann type system. However, there exists an isomorphism between the hyperbolic numbers plane and the Cartesian product of the real numbers set with itself that changes the hypercomplex direction to more similar real analysis framework.
With this aim, a partial order was introduced in hyperbolic numbers, see the classical reference [4], which provides conditions of holomorphicity and continuity for functions of hyperbolic variable early presented in [1], see also [5]. Partial order is used to prove that natural domains are rectangles in the hyperbolic plane.
In recent years, we have seen the great success of hyperbolic intervals partitioning. For example, in [6, 7, 8] the concept was applied to certain classification of Cantor type sets in the hyperbolic plane. An special example was considered in [2, 9], where the focus is to consider sets of probabilities defined like a division of the unit hyperbolic interval .
To understand how an hyperbolic interval can be divided in Section 3, two types of partitions are defined, where the prominent type: strong partition, allows us to concern in the later sections with the notion of natural hyperbolic functions.
Sections 4 and 5 are devoted to important applications of strong partitions. First, we introduce a notion of hyperbolic functions of bounded variation and indicate how the set of discontinuities may be defined like in real numbers occurs. Second, we proceed with the definition of Riemann-Stieltjes integral of a hyperbolic valued functions and its relation with the derived Riemann integral presented in [1, 10].
2 Hyperbolic Numbers
Hyperbolic numbers are a generalization of complex numbers, which are classical extension of the real numbers by the inclusion of an imaginary unit, whereas the hyperbolic numbers also do it by a new square root such that . These are introduced by J. Cockle in [3] like a sub-algebra from the nowadays well know bicomplex numbers, see [1, 4]. Whereas each nonzero complex number has a multiplicative inverse, this is no longer true for all nonzero hyperbolic numbers.
The real ring of hyperbolic numbers is the commutative ring denoted usually by
Because there exists a bijection between Euclidean plane and hyperbolic numbers, set is also known as hyperbolic numbers plane.
There are two very special zero divisors and mutually complementary elements in which are
Given we have
Therefore, and a ring isomorphism maps into the direct product . In this way we obtain what will be referred to as idempotent representation.
Real line is endowed into the hyperbolic plane by the function . On the other side the idempotent projections of a subset are the real sets
2.1 Partial order
Hyperbolic numbers are a partially ordered set with a binary relation given by
for , where and .
The strict order is defined in a similar way, indicating the strict order in the real line
The strict order is not partial order with the constraint that , since it implies that one of the following two cases could occur: nor and , or and . These cases have been deleted in the strict order.
Partial order give the possibility to define in a natural way different types of intervals. If and are related number such that , then an hyperbolic closed interval is defined by
Likewise, an open hyperbolic interval setting up
We will consider others type of intervals, for instance, and to be defined in much the same way.
By the length of an hyperbolic interval (equally valid for all types of intervals) we mean
2.2 Natural hyperbolic functions
Let be a domain. A function may also be viewed as functions of two real variables on the Euclidean plane. Into hyperbolic numbers plane there exists a well defined Cauchy-Riemann equations theory (see [1, 12, 13]). If is differentiable in , it fulfill
Moreover, if satisfies the Cauchy-Riemann equations, the idempotent components of are functions of one real variable, see [1, 14] for more details. This implies that the derivative of , denoted by , is then computed by the partial derivatives on every component, which are taken as the total derivatives at the point .
It is easy to check that every domain can be extended to the minimum open interval that contains it, for a fuller treatment [1].
Hence, will be defined on and if it is regarded in the idempotent representation , then and are real valued functions over and respectively.
By this reason we realize that functions are the natural subject of study in the hyperbolic plane, where , , are real valued functions of one variable and can be either an open or a close hyperbolic interval. Therefore, under previous features, we shall call a natural hyperbolic function.
2.3 Continuous hyperbolic functions
With the partial order in the hyperbolic numbers, several concepts can be extended to objects defined in the hyperbolic plane. For example in [5, 15] the concept of hyperbolic metric spaces was studied.
The duple is an hyperbolic metric space when is a no-empty set and is a function with the next requirements for all ,
-
1)-
.
-
2)-
.
-
3)-
.
Hyperbolic numbers form an hyperbolic metric space with the usual hyperbolic metric such that for every and , , we have
Continuity of functions between two hyperbolic metric spaces is a concept already treated in the literature, see [5]. A function between hyperbolic metric spaces is say to be continuous, if for every hyperbolic positive number there exists a such that for every , with we have that .
A natural hyperbolic function in idempotent representation, with usual hyperbolic metric on , is continuous if and only if every component is real continuous.
3 Partitions Involving Hyperbolic Intervals
Let us start with a brief discussion of the possibility to recover all information about an hyperbolic interval after a division in sub-intervals. For instance, the length of the original hyperbolic interval. This can be found in [6].
Example 3.1.
Considering the interval and dividing it by nine sub-intervals
Therefore
The Example 3.1 shows a regular partition of an square in the Euclidean plane. It fulfill, in Lebesgue sense, that the sum of the areas of every sub-squares is equal to the total area of the big square.
By this reason, an interval may be considered as a square in the Euclidean plane. A partition in sub-rectangles such that
with denotes the Lebesgue measure, will be called a regular partition.
Remark 3.2.
Previous definition is not restricted to partitions of rectangles generated by real intervals. For every , could be a measurable set and if , then has measure zero.
In order to guarantee a positive answer we make the following natural definition of hyperbolic interval partition.
Definition 3.3.
A collection of sub-intervals from is a weak partition when
This type of partitions has the disadvantage of being able to be constituted by disjoint sub-intervals as Fig. 1 shows.


Figure 1(a) has three sub-intervals from . All of them have length equal to . Therefore
In the other hand Fig. 1(b) has four intervals where and have length equal to , but for the two remaining intervals we have
So, the sum of lengths of the four intervals being equal to .
In [6], to avoid the disjoint intervals issue, a condition under which a collection of points into an hyperbolic interval provides a collection of sub-intervals whose lengths add up to the length of the biggest interval and do not have empty intersection was established there.
Definition 3.4.
Let be a finite collection of points in the interval such that when . We say that is a strong partition, if both conditions are fulfill
-
1)-
is a chain on .
-
2)-
, and
There are two differences between Def. 3.4 and that given in [6]. The first is that equality in (3.4-2) is avowed, meanwhile in [6] an strict relation is required. As a consequence a third condition relative to the absence of zero divisor in the lengths among sub-intervals of the kind is established. Inclusion of equality into the second requirement do not alter the proof of the next theorem.
Theorem 3.5.
If is a strong partition of , then
As Fig. 2(a) shows, Definition 3.4 enables degenerate sub-intervals to be built, where and are of this kind of interval. While Fig. 2(b) is the extension of the uniform real partition with norm equal to .


Strong partitions can generate partitions on real interval, only from the points that define it. So, if is an strong partition, let the projection sets
(3-E1) |
where for every .
It is possible to build real partitions with regular and weak partitions by projections of the endpoints of every intervals. Our interest is in strong partitions because they are in spirit similar to that of the real intervals context.
3.1 Real partitions define a strong partition
Two real intervals and define the hyperbolic interval with and . So, a natural question arise, How can we create a strong partition from two real partitions and ?.
By definition of strong partition, we need that initial and final points match with and . Therefore we have and .
The general process to get an hyperbolic point is taking points and with and , but if , then , in a similar way if , then . We define .
Previous step only can be repeated in a maximum of times. And it finishes when .
This procedure generates a strong partition . Figure 3 shows some examples.


4 Hyperbolic Functions of Bounded Variation
In this section, the concept of hyperbolic valued functions of bounded variation is introduced. To do this, we give a brief exposition of the notion of supremum of a set in the hyperbolic plane.
Definition 4.1.
Let be a no empty subset of . The supremum of is defined to be the number
Definition 4.1 agrees with the idea that the supremum of is the least element in that is greater than or equal to all elements of . Even when there are elements into the set that are no related, all of them are related with the supremum by the partial order. For a fuller treatment we refer the reader to [7, 8, 17, 18].
Let and an hyperbolic valued function with idempotent representation .
If is a strong partition of , then we can consider the quantity
(4-E2) |
Let denote the family of all strong partitions for and we define the set
Definition 4.2.
We say that an hyperbolic valued function is of bounded variation, when
Previous statement is equivalent to say that
are bounded sets in the real line.
Definition 4.3.
The total variation of a function of bounded variation, is the quantity
According to Def. 4.1, we have
Due the bijection from to , a function may be viewed as a function of two real variables in the Euclidean plane. Therefore, Def. 4.2 implies that is a function of bounded Vitali variation (see [19, 20, 21, 22]).
Because every component of a natural hyperbolic function relies from the respective component in the point, the sum in Eq. 4-E2 is simplified. So, if is a natural hyperbolic function, then
(4-E3) |
Remark 4.4.
Let us denote by the collection of all partitions of the real interval for every and introduce the sets
Theorem 4.5.
If is a natural hyperbolic function, then
Proof.
Taking an element in the set , by Remark 4.4, the projections and are partitions over and respectively and it has the form in the Eq. 4-E3.
Reciprocally, two partitions and define a strong partition (see Section 3.1). Partition fulfill with and , therefore even if the process in Sec. 3.1 generates points, where and denote the cardinality of and , no additional elements in the sum are added, since in degenerated intervals or , implying that
∎
Corollary 4.6.
Let be a natural hyperbolic function. The function is of hyperbolic bounded variation if and only if the idempotent component functions and are functions of real bounded variation.
On account of this result the set of discontinuities for an natural hyperbolic function of bounded variation is well defined, which is due to the fact that a real function of bounded variation only has jump discontinuities and therefore the set of discontinuities is numerable, see [11, Sec. 6.8].
Lemma 4.7.
If is a natural hyperbolic functions of bounded variation, then the set of discontinuities is the numerable union of perpendicular line segments to idempotent axes.
Proof.
The components and from are real functions of bounded variation so, there exist two set and of with all discontinuities for and respectively.
For every point and , the point is a point of discontinuity for (See Sec. 2.3). Thus, the set of discontinuities contains the union .
Similarly, the set of discontinuities of contains the union .
The union contains all discontinuities of , because if there exist a discontinuity of , then is a discontinuity of or is a discontinuity of , but this implies that or . ∎
Theorem 4.8.
The set of discontinuities from a natural hyperbolic function of bounded variation is of zero measure with the Lebesgue measure in the Euclidean plane.
Proof.
It is a consequence that every line in the Euclidean plane has zero measure and numerable union of these set again has zero measure. ∎
This result can not be translated to hyperbolic Lebesgue measure defined in [10]. Since, the Lebesgue measure is defined as , where is the Lebesgue measure in the real line, implies that the set have not zero measure for all , because is not a real set of zero measure.
5 Hyperbolic Valued Riemann-Stieltjes Integral
Strong partitions can be applied to define a Riemann-Stieltjes type integral over hyperbolic valued functions.
The diameter of a real partition is defined as the maximum into the set of all lengths of successive intervals generated by ,
Although the diameter of a partition can be extended to strong partitions in the hyperbolic numbers plane in a direct way, it is not convenient for an extension of Riemann-Stieltjes integral because even if the maximum length is taken, its projections could not coincide with the diameter of the projections of the strong partition.
Definition 5.1.
Let be a strong partition of . The diameter of is defined to be the hyperbolic number
Although, Riemann-Stieltjes integral can be defined on general hyperbolic valued functions over an interval, our focus will be on the case of natural hyperbolic functions.
Definition 5.2.
Let and be two hyperbolic functions. An hyperbolic number is called the Riemann-Stieltjes integral of respect to , if for every there exists a such that
for any strong partition that fulfill the property and whatever selection , with
The quantity is called the Riemann-Stieltjes sum. In addition, when such exists, it will be denoted by .
When and in Def. 5.2 are assumed to be natural hyperbolic functions, the Riemann-Stieltjes sum is analogue to Eq. 4-E3 and hence
But and are real valued functions defined on the respective projections of , likewise and are real valued functions defined on the whole real line. Thus, the Riemann-Stieltjes over the hyperbolic plane is the sum of classic Riemann-Stieltjes sum on the partition generated by projections of (see [11, Sec. 7.3]).
By definition of usual metric on (for more details Sec. 2.3) and when natural hyperbolic functions are assumed, we have
(5-E4) |
where the integrals in the left side in Eq. 5-E4 are classical real valued Riemann-Stieltjes integrals.
Theorem 5.3.
An hyperbolic natural function is hyperbolic Riemann-Stieltjes integrable with respect to a natural hyperbolic function if and only if every component and are real Riemann-Stieltjes integrable functions respect to and .
Proof.
The proof is followed by previous comments. ∎
From now on, we make the assumption that and are natural hyperbolic functions.
Introduce the identity function , which is a natural hyperbolic function
There is a very close connection between the Riemann–Stieltjes integral and the Riemann integral we are aiming to classify. Indeed, the hyperbolic Riemann-Stieltjes integral of a natural hyperbolic function with respect to can be viewed as the hyperbolic Riemann integral introduced in [1, Ch. IV] or a particular case of Lebesgue integral following [10, Sec. 3].
Results in [1] requires non-self-intersecting continuous loop (Jordan curve). Taking the straight line that joins the two extreme points of an hyperbolic interval we get a loop of this kind. Therefore, for every strong partition the union of lines that join every sub-interval , where , is a Jordan loop and
Another way to consider the hyperbolic Riemann integral is through the use of an hyperbolic valued measure . Taking into account the definition of Lebesgue integral in [10] and since the real Lebesgue integral restricted to a closed interval reduces to the Riemann integral, the relation
holds.
Let us mention an important property of the hyperbolic Riemann-Stieltjes integral when the integrator is an holomorphic function.
Theorem 5.4.
Let be an holomorphic and continuously differentiable function, a natural hyperbolic Riemann-Stieltjes integrable function with respect to . Then
Proof.
By holomorphic assumption on , it is a natural hyperbolic functions with derivative , for . Also, since is continuously differentiable its idempotent component have continuous derivatives of any order. Therefore and are functions of bounded variation and [11, Thm. 7.8] makes easy to see that
Combining these equalities the result is obtained. ∎
Remark 5.5.
Although Theorem 5.4 establish a direct relation between Riemann-Stieltjes and Riemann integral, the integrability of respect to is required. So, it should be convenient to see under what conditions the integrability holds.
Theorem 5.6.
Suppose that is a continuous natural hyperbolic function and is a natural hyperbolic function of bounded variation. Then is Riemann-Stieltjes integrable with respect to .
Proof.
Since components and of a continuous natural hyperbolic function are real continuous functions (see Sec. 2.3), Corollary 4.6 shows that and are real functions of bounded variation. Therefore, the integrals
exist, which is clear from [11, Thm. 7.27]. Finally, by Theorem 5.3, the result is obtained. ∎
Declarations
Funding
Instituto Politécnico Nacional (grant number SIP20211188) and Postgraduate Study Fellowship of the Consejo Nacional de Ciencia y Tecnología (CONACYT) (grant number 774598).
Conflict of interest
No conflict of interest regarding the work is reported in this paper.
Author contributions
G. Y. Téllez-Sanchez: Visualization, Investigation, Writing- Original draft preparation, J. Bory-Reyes: Conceptualization, Supervision, Writing- Reviewing and Editing. Both authors approved the final form of the manuscript.
Availability of data and material
No data were used to support this study.
Code availability
Not applicable.
References
- [1] J. C. Vignaux, A. Durañona y Vieda. On the theory of functions of an hyperbolic complex variable. Univ Nac. La Plata. Publ. Fuc. Ci. Fisicomat. Contrib. 104 (1935) 139-183.
- [2] G. Y. Téllez-Sánchez, J. Bory Reyes. Extension of the Shannon Entropy and the Chaos Game Algortihm to Hyperbolic Numbers Plane. Fractals. 29(1) (2021) 2150013.
- [3] J. Cockle. On Certain Functions Resembling Quaternions and on a New Imaginary in Algebra. London-Dublin-Edinburgh Philosophical Magazine and Science Journal. 32(3) (1848) 435-439.
- [4] G. Sobczyk. The Hyperbolic Number Plane. The College Mathematics Journal. 26(4) (1995) 268-280.
- [5] S. Kumar Datta, C. Ghosh, J. Saha. On the Ring of Hyperbolic Valued Functions. Ganita. 70(2) (2020) 175-184.
- [6] A. S. Balankin, J. Bory-Reyes, M. E. Luna-Elizarrarás and M. Shapiro. Cantor-Type Sets in Hyperbolic Numbers. Fractals. 24(4) (2016) 1650051.
- [7] G. Y Tellez-Sanchez, J. Bory-Reyes. More about Cantoy Like Sets in Hyperbolic Numbers. Fractals. 25(5) (2017) 1750046.
- [8] G. Y. Tellez Sanchez. Aspectos Básicos de una geometría Fractal sobre Números Hiperbólicos (Thesis). (SEPI ESFM-IPN, Mexico City, MX, 2018).
- [9] D. Alpay, M. Luna-Elizarraráz, M. Shapiro. Kolmogorov’s Axioms for Probabilities with Values in Hyperbolic Numbers. Advances in Applied Clifford Algebras, 27 (2017) 913-929.
- [10] R. Kumar, K. Sharma. Hyperbolic valued measures and Fundamental law of probability. Global Journal of Pure and Applied Mathematics. 13(10) (2017). 7163-7177.
- [11] T. Apostol. Mathematical Analysis. 2nd ed. (Addison Wesley, Michigan USA, 1974) p. 506.
- [12] A. E. Motter, M. A. F. Rosa. Hyperbolic Calculus. Advances in Applied Clifford Algebras. 8(1) (1998) 109-128.
- [13] A. Khrennikov, G. Segre. An Introduction to Hyperbolic Analysis. ArXiv. math-ph/0507053 (2005).
- [14] G. Y. Téllez-Sánchez, J Bory-Reyes. Generalized Iterated Function Systems on Hyperbolic Number Plane. Fractals. 27(4) (2019) 1950045.
- [15] R. Kumar, H. Saini. Topological Bicomplex Modules. Advances in Applied Clifford Algebras. 26 (2016) 1249–1270.
- [16] M. E. Luna-Elizarrarás, M. Shapiro, D. C. Struppa, A. Vajiac. Bicomplex Holomorphic Functions: The Algebra, Geometry and Analysis of Bicomplex Numbers. 1st ed. (Birkhäuser/Springer, Switzerland, 2015) p. 231.
- [17] M. E. Luna-Elizarrarás, C. O. Pérez-Regalado, M. Shapiro. On Linear Functionals and Hahn-Banach Theorems for Hyperbolic and Bicomplex Modules. Advances in Applied Clifford Algebras. 24 (2014) 1105-1129.
- [18] H. Saini, A. Sharma, R. Kumar. Some Fundamental Theorems of Functional Analysis with Bicomplex and Hyperbolic Scalars. Advances in Applied Clifford Algebras. 30(66) (2020).
- [19] G. Vitali. Sui gruppi di punti e sulle funzioni di variabili reali. Atti della Accademia delle Scienze di Torino. 43 (1908) 75-92.
- [20] M. Fréchet. Extension au cas d’intégrales multiples d’une définition de l’intégrale due à Stieltjes. Nouvelles Annales de Mathématiques. 4(10) (1910) 241-256.
- [21] J. A. Clarkson, C. R. Adams. On Definitions of Bounded Variation for Functions of Two Variables. Transactions of the American Mathematical Society. 35(4) (1933) 824-854.
- [22] C. Aistleitner, J. Dick. Functions of bounded variation, signed measures, and a general Koksma-Hlawka inequality. Acta Arithmetica. 167(2) (2015) 143-171.