452
Views
0
CrossRef citations to date
0
Altmetric
Research Article

A foundational theory of an extended Cauchy integral for generalized regulated real-valued functions on

ORCID Icon &
Article: 2174363 | Received 14 Sep 2022, Accepted 09 Jan 2023, Published online: 20 Feb 2023

Abstract

In this manuscript, we provide a generalization of the Cauchy integral to appropriate families of real-valued functions defined in Rn. The results are motivated by a generalization of regulated functions. The foundational theory for this extended Cauchy integral is developed in detail in this self-contained work. We derive some properties for appropriate families of functions and introduce the Cauchy integral for appropriate step functions. Using these results, the integral for generalized regulated functions is presented next together with some basic properties of the generalized Cauchy integral, including the linearity, the non-negativity and the monotonicity properties. Various illustrative examples on the applicability of our theoretical results are provided.

2010 Mathematics Subject Classification:

1. Introduction

Nowadays, the Riemann–Darboux integral is still a popular tool in science and engineering in view of its simplicity and intuitiveness [Citation1]. Moreover, since the fundamental theorem of calculus holds, this integral has become a popular topic for developing operative mathematical skills at the undergraduate level [Citation2,Citation3]. On the other hand, however, the theory of the Riemann–Darboux integral presents a number of limitations from a mathematical point of view. As an example, the exchange of the derivative (or the limit) and the integral operators requires strong assumptions on the family of functions involved [Citation4]. Among other important results, the dominated convergence theorem and the monotone convergence theorem for the Riemann–Darboux integral need to suppose that the convergence of the functions is uniform. This is a very restrictive hypothesis, especially when only point-wise convergence is required for those results to be valid under Lebesgues' theory of integration.

Despite having great historical and theoretical importance, another important shortcoming in the Riemann–Darboux theory is that functions which are integrable cannot be completely characterized within this theory. Indeed, the characterization of these functions is carried out employing arguments from Lebesgues' theory of measure and integration [Citation5]. In fact, it is well known that a function is integrable exactly when its collection of discontinuities has a Lebuesgue measure equal to zero. On the other hand, the Cauchy integral provides a theory in which the integrable functions are completely characterized without a need to appeal to Lebegues' theory [Citation6]. This is obviously a theoretical advantage over the Riemann–Darboux integral, in spite of the fact that Cauchy's integral theory is less popular and known.

In the present manuscript, we will extend the definition of the Cauchy integral given in [Citation7]. In that work, the integral is introduced as the countable limit of integrals of step functions that converge uniformly to a regulated function. Our approach will hinge on generalizing the concept of regulated functions following ideas and results derived in [Citation8]. It is important to mention here that Cauchy's integral was defined in [Citation9] using primitives of regulated functions. However, in the present work, we believe that it is more natural and convenient to extend the definition given in [Citation7] and avoid the use of primitives. To that end, Section 2 of this work will study appropriate families, that is, families of functions on which generalized regulated functions are defined. We will establish then the appropriate family of interest in this work. In Section 3, we will define the generalized Cauchy integral for appropriate step functions. Then we will define the generalized Cauchy integral for generalized regulated functions in Section 4, and demonstrate some properties, like the linearity, the non-negativity and the monotonicity. Some illustrative examples are provided in Section 5. Finally, this work closes with a section of concluding remarks and discussions.

2. Preliminaries

Generalized regulated functions were introduced in 1979 by T. Davison [Citation8], by using the so-called ‘appropriate families’. The present section is devoted to recall those families and establish the generic appropriate family of R which will be used throughout this work.

Definition 2.1

In general, assume that X is some fixed topological space, and let F be a collection of sets in X. The class F is an appropriate family when the next three requirements hold:

  1. Under the operations of unions and intersections of sets, F is a lattice [Citation10].

  2. If C and D are sets in F and DC, it follows that CDF.

  3. A basis for the topology of X is the collection of all open sets of X in F.

Generalized regulated functions will be introduced through appropriate families using the following definition.

Definition 2.2

Let F be an appropriate family on the topological space X, and assume that xX. Then f:XR will be called a regulated function at x if for each ε>0 we can find NF, such that:

  1. N is neighbourhood of x, and

  2. there are sets A1,AkF such that N=A1Ak, with the property that |f(t)f(s)|ε, for all i=1,,k and s,tAi.

For the remainder of this manuscript, we will let X be Rn with the standard topology and nN. Additionally, we will let R represent the collection of all regulated functions.

Definition 2.3

As usual, intervals are sets of the following forms: (a,b):={xR:a<x<b},[a,b]:={xR:axb},[a,b):={xR:ax<b},(a,b]:={xR:a<xb}.Here, ab are real numbers. Intervals of the first type are called open, while those of the second type are called closed. A box in Rn is a set of the form B=I1××In, where the sets I1,,In are intervals in the real numbers. When the intervals I1,,In are open and have the same length, then we say that B is an open box.

In what follows, we let F be the collection of finite unions of boxes in Rn. We claim that F is an appropriate family.

Lemma 2.1

Under unions and intersections of sets, the family F is a lattice.

Proof.

Clearly, (F,) is a partial order. Let U,VF. The definition of F assures that (1) U=i=1m1Ui,V=j=1m2Vj,(1) with Ui and Vj boxes, for each i=1,,m1, j=1,m2. It is obvious that the supremum (or least upper bound) of U and V is the set sup{U,V}=(i=1m1Ui)(j=1m2Vj)=k=1m1+m2Rk,where Rk=Uk when k=1,,m1, and Rk=Vkm1 if k=m1+1,,m1+m2. On the other hand, the infimum (or greatest lower bound) of U and V is (2) inf{U,V}=(i=1m1Ui)(j=1m2Vj)=i=1m1[j=1m2(UiVj)]=k=1m1m2Sk,(2) where each Sk is a set of the form UiVj. Since each Ui and Vj are boxes, it follows that Rk is also a box. As a consequence, we conclude that F is a lattice.

Lemma 2.2

The condition UVF holds if VU and U,VF.

Proof.

Assume that U,VF are of the form (Equation1), and notice that UV=i=1m1[j=1m2(UiVj)].We claim that UiVj is a finite union of boxes. Indeed, the case when UiVj= is trivial, so let us suppose that UiVj. By definition, Ui and Vj have the forms Ui=I1××In,Vj=J1××Jn,where Ik,Jk are intervals, for each k{1,,n}. Notice that IkJk is equal to an interval Rk0 or the union of two disjoint intervals Rk1 and Rk2. Let us consider the cross-products of all the possible combinations of these intervals, that is, the boxes of the form (3) R1r××Rnr,(3) where kr is equal to k0 if IkJk is a single interval, or is equal to the values k1 and k2 if IkJk is the union of two intervals. Obviously, these boxes are disjoint. In this way, UiVj is equal to the finite disjoint union of the boxes (Equation3), and we can denote this fact as follows: UiVj=k=1mijRkij.As a consequence, UV=i=1m1[j=1m2(UiVj)]=i=1m1[j=1m2(k=1mijRkij)].Finally, using induction, we will check now that j=1m2(k=1mijRkij) is a finite union of boxes. If m2=2, then we can mimic the argument to derive (Equation2). Let us suppose that the claim is true for m2=r>2. Notice then that j=1r+1(k=1mijRkij)=[j=1r(k=1mijRkij)](k=1mi,r+1Rki,r+1)=(k=1pRk^)(k=1mi,r+1Rki,r+1)=k=1lRk.The last equality was obtained as in the proof of (Equation2). This result follows now by mathematical induction.

Lemma 2.3

A basis for the topology of X is the collection of sets in F which are open.

Proof.

Let B denote the set of open boxes in Rn (which is a basis for the standard topology on Rn), and denote by τ(B) the topology generated by B. Define then the collection B=Fτ(B). It will be shown firstly that B is the basis for a topology in X. More precisely, the following conditions will be established [Citation11]:

  1. There are basis elements that contain x, for each xX.

  2. Suppose that xB1B2 and B1,B2B. Then xBB1B2 for some BB.

To check that the first condition is satisfied, let x=(x1,,xn)Rn, and define the box RxF by Rx=(x11,x1+1)××(xn1,xn+1). Clearly, xRx and RxB, which show that (i) is true. To establish property (ii), let B1,B2B and suppose x=(x1,,xn)B1B2. Notice that B1=i=1m1R1i,B2=j=1m2R2j.Since xB1B2, then xR1i and xR2j, for some indexes i and j. On the other hand, each intersection R1iR2j is a box, which means that it is a cross-product of intervals Ik, each one of them with endpoints ak and bk, where k=1,,n. Let us define it now (4) δ=mink=1,,n{|bkxk|,|xkak|},(4) (5) B=(x1δ,x1+δ)××(xnδ,xn+δ)R1iR2j.(5) Clearly, xB and BB1B2 and BB. Conclude that B is a basis for a topology.

Assume that τ(B) represents the smallest topology containing B. The following discussion will establish that τ(B)=τ(B). To prove that τ(B)τ(B), let BB and xRn satisfy xB. Due to the fact that B is an open box, then BB and xBB. Conversely, to prove that τ(B)τ(B), let x=(x1,,xn)Rn and BB satisfy xB. By definition, B has the form B=i=1mRi,where Ri is a box, for each i=1,m. Since xB, then xRi for some i{1,,m}. We know that Ri is a cross-products of intervals Ik, each one with ends ak and bk, where k=1,,n. Let us define δ as (Equation4), and B as B in (Equation5). Evidently, BB and xBB, which implies that τ(B)τ(B). The claim that τ(B)=τ(B) has been proved now.

Theorem 2.1

The collection F is an appropriate family.

Proof.

The conclusion follows from Lemmas 2.1–2.3.

3. The integral for step functions

In this stage, we will extend Cauchy's integral for a suitable family of step functions. More precisely, we will define the integral for appropriate step functions. In a first step, we will recall some standard (though important) concepts.

Definition 3.1

Define the length of any of the intervals in Definition 2.3 as |I|=ba. If B=I1××In is a box, we define the volume of B through |B|=|I1|×|I2|××|In|. Given R=i=1mBiF where the sets Bi are disjoint, we define the volume of R as |R|=|B1|++|Bm|. In general, we let ||=0 for the sake of convenience.

By definition, it is clear that the length and volume are non-negative quantities.

Lemma 3.1

If RF, then

  1. R is a disjoint finite union of boxes.

  2. Whenever R can be represented as a finite disjoint union R1Rk of boxes, then |R| is independent of the partition. More precisely, if R1Rk is another partition of R then |R1|++|Rk|=|R1|++|Rk|.

Proof.

See [Citation12].

Corollary 3.1

If BR and R,BF are satisfied, then |B||R|.

Proof.

It is obvious that R=(RB)B. From 3.1 we obtain |R|=|RB|+|B|. The non-negativity of the volume assures now that |B||R|.

We are now in a position to extend Cauchy's integral to appropriate step functions. To that end, let χS denote the characteristic function on the set S, for each SF. It is easy to see that χS is a regular function [Citation8]. Assume that E represents the subspace of R spanned by the set {χS:SF} as a vector space over R. The appropriate step functions will be those which are elements of E (see [Citation8]).

Lemma 3.2

If φE, then there exist distinct a1,amR{0} and nonempty disjoint sets S1,,SmF, with the property that (6) φ=i=1maiχSi.(6) This representation is unique and is known as the standard representation of φ.

Proof.

Since φE, then there are b1,bpR and G1,,GpF such that φ=j=1pbjχGj.In a first step, we will show that there exist disjoint sets H1,,HqF and constants c1,,cqR, such that j=1pbjχGj=i=1qciχHi.To that end, we proceed by induction. If p = 2, then we have that j=12bjχGj=b1χG1G2+(b1+b2)χG1G2+b2χG2G1.In this way, we have represented φ as a sum of characteristic functions (each one multiplied by a constant) on disjoint sets in F. Let us suppose the claim is true when p = k. Then j=1k+1bjχGj=j=1kbjχGj+bk+1χGk+1=i=1lciχHi+bk+1χGk+1=i=1lciχHiGk+1+i=1l(ci+bk+1)χHiGk+1+bk+1χGk+1i=1lHi,where H1,,HlF are disjoint sets. This proves our claim. As a consequence, φ has a representation of the form φ=j=1pbjχGj=r=1qcrχHr.where H1,,HqF are disjoint sets, and c1,,cqR. Notice that we may suppose these sets are nonempty, otherwise we can omit the terms with characteristic functions on the empty set, and the equality would still hold.

Denote the different nonzero members of the collection {cr:r{1,,q}} by ai, for i=1,,m, and let Si be the union of the sets Hr satisfying ai=cr. Clearly φ=r=1qcrχHr=i=1maiχSi,where a1,amR are distinct nonzero numbers and S1,,SmF are nonempty disjoint sets.

We will prove next that the standard representation is unique. To that end, suppose that there exist different nonzero numbers d1,dlR and nonempty disjoint sets D1,,DlF, with the property that φ=j=1ldjχDj.Firstly, we claim that i=1mSi=j=1mDj. Indeed, if ti=1mSi then tSr for some r{1,,m}. Thus, 0arχSr(t)=i=1maiχSi(t)=j=1ldjχDj(t).As a consequence of this, there is some k{1,,l} with the property that χDk(t)0, so tDk. Therefore, i=1mSij=1mDj, while the other inclusion is demonstrated analogously. Let r{1,,m} and tSr. Then there exists some k{1,,l} such that tDk. This implies that ar=arχSr(t)=i=1maiχSi(t)=j=1ldjχDj(t)=dkχDk(t)=dk.Therefore, for each r{1,,m}, there exists some k{1,,l}, such that cr=dk. In the same way, it is possible to check that for each k{1,,l}, there is some r{1,,m} such that cr=dk. Thus m = l and the coefficients of the representations are the same. Finally, if r{1,,m} and tSr, then f(t)=i=1maiχSi(t)=ar=dk,which yields tGk. This means that SrGk, In the same way, it is possible to prove that GkSr. Thus the sets associated to the characteristic functions of both representations are the same, whence the uniqueness readily follows.

Definition 3.2

Let φE be written in the standard representation (Equation6). We define the integral of φ as φ(t)dt=i=1mai|Si|.

Clearly, the integral is well defined since the standard representation is unique.

Lemma 3.3

Suppose that φE is given in the standard representation (Equation6). Suppose that b1,bpR and G1,,GpF are such that (7) φ=j=1pbjχGj.(7) Then i=1mai|Si|=j=1pbj|Gj|.

Proof.

We may suppose that all the coefficients bj are nonzero since, otherwise, can omit the terms with zero coefficient from (Equation7). We consider two cases for the proof.

Case 1. The sets Gj are disjoint. Using arguments similar to those employed in Lemma 3.2, it is easy to prove that i=1mSi=j=1lGj. Then Si=j=1p(SiGj),Gj=i=1m(SiGj).Using now Lemma 3.1 we obtain that |Si|=j=1p|SiGj|,|Gj|=i=1m|SiGj|.Moreover, if r{1,,m} and k{1,,p} satisfy that SrGk is nonempty, then for each tSrGk we have that ar=i=1maiχSi(t)=j=1lbjχGj(t)=bk.As a consequence, ar=bk when SrGk. From this, i=1mai|Si|=i=1mj=1SiGjpai|SiGj|=i=1mj=1SiGjpbj|SiGj|=j=1pbj|Gj|.Case 2. The sets Gj are not disjoint. Proceeding by induction and letting p = 2, j=12bj|Gj|=b1|G1G2|+(b1+b2)|G1G2|+b2|G2G1|.The right end is a summation of volumes of disjoint sets (multiplied by some constant). Therefore, we can apply the first case and obtain j=12bj|Gj|=i=1mai|Si|.Suppose now that the conclusion is true when p = k. Therefore, j=1k+1bj|Gj|=j=1kbj|Gj|+bk+1|Gk+1|=i=1lci|Hi|+bk+1|Gk+1|where H1,,HlF are nonempty and pairwise disjoint and the constants c1,,clR are different and nonzero by the induction hypothesis. Thus, j=1k+1bj|Gj|=i=1lci(|HiGk+1|+|HiGk+1|)+bk+1(|Gk+1i=1lHi|+i=1l|Gk+1Hi)|)=i=1lci|HiGk+1|+i=1l(ci+bk+1)|HiGk+1|+bk+1|Gk+1i=1lHi|.The right-hand side is a sum of volumes of disjoint sets (multiplied by a constant), which means that we can apply the first case to reach i=1lci|HiGk+1|+i=1l(ci+bk+1)|HiGk+1|+bk+1|Gk+1i=1lHi|=j=1mai|Si|.As a consequence, j=1k+1bj|Gj|=j=1mai|Si|which is what we wanted to prove.

Lemma 3.4

If φ,ψE and cR, then

(i)

(t)dt=cφ(t)dt, and

(ii)

(φ+ψ)(t)dt=φ(t)dt+ψ(t)dt.

Proof.

Let φ and ψ be given in the standard representation, say, φ=i=1maiχSi,ψ=j=1pbjχGjTo prove (i), notice that (t)dt=i=1mcai|Si|=ci=1mai|Si|=cφ(t)dt.To establish (ii) now, observe that φ+ψ=i=1maiχSij=1pGj+i=1mj=1p(ai+bj)χSiGj+j=1pbjχGji=1mSi.Lemma 3.3 guarantees now that (φ+ψ)(t)dt=i=1mai|Sij=1pGj|+i=1mj=1p(ai+bj)|SiGj|+j=1pbj|Gji=1mSi|=i=1mai|Sij=1pGj|+i=1maij=1p|SiGj|+j=1pbji=1m|SiGj|+j=1pbj|Gji=1mSi|=i=1mai[|Sij=1pGj|+|Si(j=1pGj)|]+j=1pbj[|Gji=1mSi|+|Gj(i=1mSi)|]=i=1mai|Si|+j=1pbj|Gj|=φ(t)dt+ψ(t)dtwhich is what we wanted to prove.

4. The integral for regulated functions

In this section, we will define Cauchy's integral for generalized regulated functions. To that end, we will present some useful properties of Cauchy's integral for appropriate step functions.

Definition 4.1

Let BF and φE. The integral of φ over B is defined by Bφ(t)dt=φ(t)χB(t)dt.

Clearly, this concept is well defined. Indeed, if BF and φE have standard representation given by (Equation6), then φχB=(i=1maiχSi)χB=i=1maiχBχSi=i=1maiχBSi.This means that φχBE.

The next result is similar to the last theorem proved in [Citation8]. However, we needed to apply different hypotheses to reach the conclusion. The proof requires to be modified in the way.

Lemma 4.1

Let BF, and assume that f is some real-valued function defined in Rn. Then f is regulated on B¯ exactly in the case that there is some sequence (φn)E that converges uniformly on B¯ to f.

Proof.

Let us suppose f is regulated on B, and let ε>0. We will construct a function φεE satisfying |f(x)φε(x)|<ε for each xB. When xB¯, there is some neighbourhood Nx=A1xAkxx of x, satisfying |f(s)f(t)|ε, for each s,tAix. Here, A1x,,AkxxF. Since Nx is a neighbourhood of x, then the interior of Nx contains x. If we let Sx=intNx, then B=i=1mRii=1mRi¯xB¯Sx,where R1,,Rm are boxes from the definition of BF. Compactness of the set i=1mRi¯ implies that the open covering {Sx:xB¯} has a finite subcovering, say, {Sx1,,Sxr}. As a consequence, Bi=1rSxii=1rj=1kxiAjxi.Relabeling we have that Bi=1pVi, and |f(t)f(s)|<ε for each s,tVi. Using the second property of appropriate families, we can suppose that the sets Vi are disjoint. Let yiVi be arbitrary, and define φε=i=1pf(yi)χVi.If xB¯, then there is some j{1,,p} such that xVj. Hence |f(x)φε(x)|=|f(x)i=1pf(yi)χVi(x)|=|f(x)f(yj)|<ε.The uniform convergence of (φ1m)mNE to f on B readily follows.

Conversely, suppose that the sequence (φn)E converges uniformly on B¯ to f, and fix ε>0. By hypothesis, there exists some nN, with the property that, for each xB¯, it follows that |f(x)φn(x)|<ε3. Let xB¯ and recall that φnR. Then there exists a neighbourhood N=A1Ak of x with each AiF, such that |φn(s)φn(t)|ε3, for each s,tAi. This means that if s,tAi, then |f(s)f(t)||f(s)φn(s)|+|φn(s)φn(t)|+|φn(t)f(t)|<ε.As a consequence f is regulated on B¯, as desired.

Lemma 4.2

Let BF, assume f is regulated on B¯, and suppose (φm),(ψm)E are two sequences that converge uniformly on B¯ to f. Then

  1. (Bφm(t)dt) is a Cauchy sequence.

  2. limmBφm(t)dt=limmBψm(t)dt.

Proof.

Observe firstly that the sequences (χBφm),(χBψm)E converge uniformly to χBf. Let χBφm=i=1hmaimχSim,χBψm=j=1lmbjmχGjmbe the standard representations for χBφ and χBψ, respectively.

To prove (i), fix ε>0, and notice that the absolute convergence of (χBφm) to χBf assures the existence of some MN, such that if mM, then |χBφm(t)χBf(t)|<ε, for each tRn. Here, ε=ε2|B|. Let m,kM, and assume that i{1,,hm} and j{1,,hk}. Then the following inequalities are satisfied:

  1. If SimSjk then |aimajk|<2ε.

  2. If Sim(j=1hkSjk) then |aim|<ε.

  3. ] If Sjk(i=1hmSim) then |ajk|<ε.

Additionally, the facts that i=1hmSimB and j=1hkSjkB, from Corollary 3.1 yield the following inequality: |(i=1hmSim)(j=1hkSjk)||B|.As a consequence, BφmBφk=i=1hmaim|Sim|j=1hkajk|Sjk|=i=1hmaim|Simj=1hkSjk|+i=1hmaim|Sim[j=1hkSjk]|j=1hkajk|Sjki=1hmSim|j=1lmajk|Sjk[i=1hmSim]|=i=1hmaim|Simj=1hkSjk|+i=1hmaimj=1hk|SimSjk|j=1hkajk|Sjki=1hmSim|j=1hkajki=1hm|SimSjk|<2ε(i=1hm|Simj=1hkSjk|+i=1hmj=1hk|SimSjk|+j=1hk|Sjki=1hmSim|)=2ε|(i=1hmSim)(j=1hkSjk)|2ε|B|=ε.We conclude (Bφm(t)dt) is a Cauchy sequence.

To establish (ii) now, let ε>0. Uniform convergence of the sequences (χBφm) and (χBψm) to χBf imply that there exists MN such that, if mM, then

(c)

|χBφm(t)χBf(t)|<ε, for each tRn, and

(d)

|χBψm(t)χBf(t)|<ε, for each tRn.

Here, ε=ε2|B|. If we proceed now as in the proof of (i), we can readily obtain the conclusion of part (ii).

The following is our definition of the generalized Cauchy integral. Our definition will hinge on the definition for the one-dimensional case presented in [Citation7].

Definition 4.2

Let BF, and suppose that f is a regulated function on B¯. We define the integral of f over B as Bf(t)dt=limmBφm(t)dtwhere (φm)E is any sequence which converges uniformly to f on B¯.

Notice that this concept is well defined in light of Lemmas 4.1 and 4.2.

Theorem 4.1

Let BF and cR. If f and g are regulated functions on B¯, then

  1. Bcf(t)dt=cBf(t)dt.

  2. B(f+g)(t)dt=Bf(t)dt+Bg(t)dt.

Proof.

Let (φm),(ψm)E be sequences that converge uniformly to f and g in B¯, respectively. Property (i) follows from Lemma 3.4 and the following identities: Bcf(t)dt=limmcEφm(t)dt=climmEφm(t)dt=cBf(t)dt.On the other hand, also follows from Lemma 3.4 and the identities B(f+g)(t)dt=limm(Eφm(t)dt+Eψm(t)dt)=limmEφm(t)dt+limmEψm(t)dt=Bf(t)dt+Bg(t)dt.This proves the result.

Theorem 4.2

Let BF, and assume that f is a regulated function on B¯. If f is non-negative on B¯, then Bf(x)dx0.

Proof.

Let (φm)E be a sequence that converge uniformly to f on B¯. For each mN and xB¯, define φm(x)=max{0,φm(x)}. It is easy to verify that (φm)E converges uniformly to f. Fix mN, and let i=1kaiχSibe the standard representation of χBφm. Clearly ai0, so Bφm(x)dx=i=1kai|Si|0.This last inequality holds for all mN. Making m now, we obtain Bf(x)dx=limmBφm(x)dx0which is what we wanted to prove.

Theorem 4.3

Assume that BF, and suppose that f and g are regulated functions on B¯. If fg on B¯, then Bf(x)dxBg(x)dx.

Proof.

By hypothesis fg0 on B¯. The conclusion readily follows now from the linearity of the integral and Theorem 4.2.

5. Illustrative examples

The purpose of the present section is to present some concrete examples to illustrate the applicability of the theorems above. In the following, we will observe the nomenclature employed in those results. We just need to point out that B will represent a box in Rn, as defined at the beginning of Section 3.

  1. To start with, observe that if f is a constant function equal to cR in all B, then f is integrable on B and, moreover, Bf(x)dx=c|B|.In particular, it follows from this fact that the integral of the constant function 1 over B is equal to the volume of B.

  2. If f and g are continuous functions, then they are regulated and, in that case, the properties in Theorem 4.1 are satisfied. It is worth pointing out that the class of generalized regulated functions obviously includes the continuous functions. Moreover, regulated functions are obviously bounded.

  3. An easy application of the monotonicity property results if we consider the function f:B¯R given by f(x)=x12++xn2, where x=(x1,,xn). In this case, it is obvious (both from the definition and the monotonicity of the integral) that Bf(x)dx0.Moreover, if we let g:B¯R be given as g(x)=x12, then evidently Bf(x)dxBg(x)dx.These properties were known for functions defined on an interval of R. The advantage, in this case, is that they have been extended theoretically to functions defined in Rn. These and other trivial examples confirm the validity of the theoretical results demonstrated in the present manuscript.

The previous were some easy consequences from the analytical results derived in the present work. However, the following are more elaborate applications of the extended Cauchy integral for generalized regulated real-valued functions on Rn. For the sake of illustration, the functions considered will be defined on R2. Our examples will be numerically oriented rather than theoretical.

In the first of our examples, we provide a means to estimate the value of the generalized Cauchy integral of a function, and the approximate error associated with that estimation.

Example 5.1

Define the function f:R2R by f(x,y)=515(x2+y2).For the sake of illustration, let us fix the box B=[0,4]×[0,3]R2. Observe that the function f is continuous, so regulated. Thus, f is regulated in the generalized sense of this work and so, integrable in the sense of Cauchy. Let us consider uniform partitions of the intervals [0,4] and [0,3] of the forms P1:x0<x1<x2<x3<x4 and P2:y0<y1<y2<y3, respectively. Notice then that P1={0,1,2,3,4} and P2={0,1,2,3}, and these partitions induce a partition of B into twelve squares of equal area. These squares will be represented by Ri and, evidently, |Ri|=1, for each i=1,2,,12. For convenience, Figure  shows the graph of the function f on the rectangle B. It is worthwhile pointing out that the rectangles on the xy-plane represent the squares Ri used in this example.

Figure 1. Graph of the function z=515(x2+y2) as a function of x and y on the domain (x,y)[0,4]×[0,3]. The rectangles on the xy-plane represent a partition of B into equal squares of area equal to 1.

Figure 1. Graph of the function z=5−15(x2+y2) as a function of x and y on the domain (x,y)∈[0,4]×[0,3]. The rectangles on the xy-plane represent a partition of B into equal squares of area equal to 1.

Let ϕ,ψ:BR be step functions defined as follows: for each i{1,2,,12} and (x,y)Ri, agree that ϕ(x,y)=min{f(x,y):(x,y)Ri},ψ(x,y)=max{f(x,y):(x,y)Ri}.For simplification, let us convey that the values for ϕ and ψ on the squares Ri are given by ai and bi, respectively, for each i=1,2,,12. Obviously, the inequalities ϕfψ are satisfied. Using the definition of the integral for step functions, we obtain that Bϕ(t)dt=i=112ai|Ri|=f(1,1)+f(2,1)+f(3,1)+f(4,1)+f(1,2)+f(2,2)+f(3,2)+f(4,2)+f(1,3)+f(2,3)+f(3,3)+f(4,3)=235+205+155+85+205+175+125+55+155+125+75+0=30.8.In a similar fashion, observe that Bψ(t)dt=i=112bi|Ri|=f(0,0)+f(1,0)+f(2,0)+f(3,0)+f(0,1)+f(1,1)+f(2,1)+f(3,1)+f(0,2)+f(1,2)+f(2,2)+f(3,2)=47.6.Integrability of the functions ϕ, f and ψ together with Theorem 4.3 assure then that 30.8=Bϕ(t)d(t)Bf(t)dtBψ(t)dt=47.6.Moreover, elementary algebra establishes that |Bf(t)dt39.2|=|Bf(t)dt12[Bϕ(t)dt+Bψ(t)dt]|12[Bψ(t)dtBϕ(t)dt]=8.4.As a conclusion, the numerical value of the Cauchy integral of the function f on B is estimated as 39.2, with a numerical error equal to 8.4. As expected by the theory developed in this work, better estimates can be obtained by choosing step functions over finer mesh grids on the x and y directions.

In the following example, we obtain the value of the extended Cauchy integral for the function in the previous example. The approach provided in the solution may be extended to calculate the integral for any polynomial function.

Example 5.2

Consider the function f and the set B from Example 5.1. For each nN, let P1(n) and P2(n) be the partitions of [0,4] and [0,3], respectively, consisting of sub-intervals of lengths equal to 1n. Let P(n) be the partition of B induced by P1(n) and P2(n), for each nN. It is obvious that P consists of 12n2 identical squares of equal area. Denote those squares by Ri(n), for each i=1,2,,12n2, and note that |Ri|=1n2. Define now the function ψn:BR by ψn(x,y)=max{f(x,y):(x,y)Ri(n)},for each i{1,2,,12n2} and (x,y)Ri. Clearly, ψn is a step function, for each nN. We will check now that ψn converges uniformly on B to the function f.

Fix nN, and observe that each square Ri has the form [x,x+1n]×[y,y+1n], for suitable xP1(n) and yP2(n). The definition of the function ψn shows that ψn(x,y)=f(x,y), for each (x,y)Ri. Moreover, |ψn(x,y)f(x,y)||f(x,y)f(x+1n,y+1n)|=25[1n2+1nx+1ny]25[1n2+7n].From this fact, we can readily establish that the sequence of step functions (ψn) converges uniformly to f on B. It follows then that (8) Bf(t)dt=limnBψn(t)dt.(8) To calculate each of the integrals on the right-hand side, observe that the nodes of the partitions P1(n):x0<x1<<x4n and P2(n):y0<y1<y3n are defined, respectively, by xi=i/n and yj=j/n, for each i=0,1,,4n and j=0,1,,3n. Let Rij=[i/n,(i+1)/n]×[j/n,(j+1)/n], and observe that ψn(x,y)=f(i/n,j/n), for each (x,y)Rij, i=0,1,,4n and j=0,1,,3n.

For convenience, we allow aij=f(i/n,j/n) in what follows. In such way, Bψn(t)dt=i=04n1j=03n1aij|Rij|=1n2i=04n1j=03n1[5i2+j25n2]=6015n4i=04n1j=03n1(i2+j2)=6015n4[3ni=04n1i2+4nj=13n1j2]=602n(4n1)(8n1)5n32n(3n1)(6n1)5nnHere, we employed the well-known formula for the sum of squares in the last step. Substituting this last identity into Equation (Equation8) and taking then the limit when n tends to infinity, we easily obtain that Bf(t)dt=40.Alternatively, we could have applied the linearity of the generalized Cauchy integral to calculate separately the integrals of each summand in the expression of f. Adding the values of those integrals would lead us to the same result for the integral of f.

It is important to notice here that the estimate derived in Example 5.1 (which is equal to 39.2) is in qualitative agreement with the value of the Cauchy integral of f obtained in the last example (which is equal to 40). Moreover, observe that the error obtained in the former example constitutes an upper-bound estimate.

The last example establishes the existence of non-integrable functions.

Example 5.3

Let B=[0,1]×[0,1], and consider the function f:R2R given by f(x,y)={0,x,yQ,1,otherwise.Using contradiction, it is easy to observe that there is no sequence of simple functions defined on B which converges uniformly to f. It follows that the function is not integrable in the extended sense of Cauchy.

6. Conclusions

Using appropriate families of functions and generalized regulated functions, the present manuscript extended Cauchy's integral to real-valued functions defined in Rn. As it has been done for other theories [Citation13], the present work provides a self-contained and detailed generalized theory for the generalized Cauchy integral. This extension was reached by defining the integral as the limit of a sequence of integrals of appropriate step functions that converge to a generalized regulated function. In addition, some basic properties of this integral were proved, like the linearity, the preservation of non-negativity and the monotonicity. We consider that one of the advantages of this integral is that it is not necessary to study measure theory to characterize the integrable functions, unlike the Riemann integral. This could be more advantageous from a pedagogical point of view, especially as a topic for elementary analysis courses. On the other hand, the authors of this manuscript are convinced that there are many more properties associated with this integral. For instance, one could study its relationship with the classical Riemann–Darboux integral. Here, it is worth pointing out that the author of [Citation14] showed that every regulated function is Riemann–Darboux integrable, but the converse is not true in general. A natural direction of investigation is whether a similar situation prevails with generalized regulated functions. Obviously, this is an interesting topic for research that the authors intend to tackle in the near future.

As one of the anonymous reviewers of this manuscript pointed out, the examples provided in this work are still simple. This is due to the fact that the theory formulated herein can still be developed tremendously. With more theory, better examples can be provided to illustrate the applications of this integration theory. To that effect, counterparts of strong theorems (like the Lebesgue's dominated convergence theorem, the monotone convergence theorem, Fatou's lemma, Fubini's theorem, the fundamental theorem of calculus, etc.) are required. However, the derivation of those results (if possible) will entail much more efforts and time and would require additional manuscripts. In its present form, our work has tackled the difficulties to extend Cauchy's integral to Rn, and the arguments have been nontrivial (as it can be checked above). Knowing that there is still a lot more to be done in this topic, we have included the adjective ‘foundational’ in the title of this manuscript to emphasize the foundational aspect of this theory. As part of future works, we will strive to develop this theory even more towards stronger properties of the Cauchy integral in Rn which will lead to more complicated and interesting applications.

Acknowledgments

The authors would like to thank the anonymous reviewers and the associate editor in charge of handling this manuscript for all their suggestions. Their comments and criticisms helped us to improve the quality of this work. The authors confirm that all the research meets ethical guidelines and adheres to the legal requirements of the study country.

Data availability statement

Data sharing is not applicable to this article as no new data were created or analysed in this study.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

One of the authors (J.E.M.-D.) was funded by the National Council of Science and Technology of Mexico through grant A1-S-45928.

References

  • Bartle RG. Return to the Riemann integral. Am Math Mon. 1996;103(8):625–632.
  • Pfeffer WF. The multidimensional fundamental theorem of calculus. J Aust Math Soc. 1987;43(2):143–170.
  • Sobczyk G, Sánchez OLón. Fundamental theorem of calculus. Adv Appl Clifford Algebras. 2011;21(1):221–231.
  • Bartle RG. The elements of real analysis. Wiley; 1976. Bibliyografya ve İndeks, New York.
  • Bohner M, Guseinov G. Riemann and Lebesgue integration. In: Advances in dynamic equations on time scales. Springer; 2003. p. 117–163.
  • Mattila P, Melnikov MS, Verdera J. The Cauchy integral, analytic capacity, and uniform rectifiability. Ann Math. 1996;144(1):127–136.
  • Viloria N, Cadenas R. Integral de Cauchy: alternativa a la integral de Riemann. Divulgaciones Math. 2003;11(1):49–53.
  • Davison TMK. A generalization of regulated functions. Am Math Mon. 1979;86(3):202–204.
  • Dieudonné J. Foundations of modern analysis. New York: Academic Press; 1969.
  • Nation J. Notes on lattice theory; 1998 Feb.
  • Munkres JR. Topology. Prentice Hall; 2000. (Featured titles for topology).
  • Tao T. An introduction to measure theory. American Mathematical Society; 2011. (Graduate studies in mathematics).
  • Feng X, Sutton M. A new theory of fractional differential calculus. Anal Appl. 2021;19(04):715–750.
  • Berberian SK. Regulated functions: Bourbaki's alternative to the Riemann integral. Am Math Mon. 1979;86(3):208–211.