This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Naïve Infinitesimal Analysis
Its Construction and Its Properties

Anggha Nugraha
[email protected]
Maarten McKubre-Jordens Hannes Diener
Abstract

This paper aims to build a new understanding of the nonstandard mathematical analysis. The main contribution of this paper is the construction of a new set of numbers, <\mathbb{R^{Z_{<}}}, which includes infinities and infinitesimals. The construction of this new set is done naïvely in the sense that it does not require any heavy mathematical machinery, and so it will be much less problematic in a long term. Despite its naïvety character, the set <\mathbb{R^{Z_{<}}} is still a robust and rewarding set to work in. We further develop some analysis and topological properties of it, where not only we recover most of the basic theories that we have classically, but we also introduce some new enthralling notions in them. The computability issue of this set is also explored. The works presented here can be seen as a contribution to bridge constructive analysis and nonstandard analysis, which has been extensively (and intensively) discussed in the past few years.

Acknowledgements

The first author received financial support from Indonesia Endowment Fund for Education that enables the research of this article. We would also like to thank Professor Elemér Rosinger and Dr. Josef Berger for their invaluable inputs.

1 Background and Aim

There have been many attempts to rule out the existence of inconsistencies in mathematical and scientific theories. Since the 1930s, we have known (from Gödel’s results) that it is impossible to prove the consistency of any interesting system (in our case, this is a system capable of dealing with arithmetic and analysis). One of the famous examples of inconsistency is as follows. Suppose we have a function f(x)=ax2+bx+cf\left(x\right)=ax^{2}+bx+c and want to find its first derivative. By using Newton’s ‘definition of the derivative’:

f(x)\displaystyle f^{\prime}\left(x\right) =f(x+h)f(x)h\displaystyle=\frac{f\left(x+h\right)-f\left(x\right)}{h}
=a(x+h)2+b(x+h)+cax2bxch\displaystyle=\frac{a\left(x+h\right)^{2}+b\left(x+h\right)+c-ax^{2}-bx-c}{h}
=ax2+2axh+2h2+bx+bh+cax2bxch\displaystyle=\frac{ax^{2}+2axh+2h^{2}+bx+bh+c-ax^{2}-bx-c}{h}
=2axh+2h2+bhh\displaystyle=\frac{2axh+2h^{2}+bh}{h}
=2ax+h+b\displaystyle=2ax+h+b (1)
=2ax+b\displaystyle=2ax+b (2)

In the example above, the inconsistency is located in treating the variable hh (some researchers speak of it as an infinitesimal). It is known from the definition that hh is a small but non-trivial neighbourhood around xx and, because it is used as a divisor, cannot be zero. However, the fact that it is simply omitted at the end of the process (from Equation 1 to Equation 2) indicates that it was, essentially, zero after all. Hence, we have an inconsistency.

This problem of inconsistency has been ‘resolved’ in the 19th19^{\textnormal{th}} century111If we look historically, the debates of the use of infinitesimals have a long and vivid history. Their early appearance in mathematics was from the Greek atomist philosopher Democritus (around 450 B.C.E.), only to be dispelled by Eudoxus (a mathematician around 350 B.C.E.) in what was to become “Euclidean” mathematics. by the concept of limit, but its (intuitive) naïve use is still common nowadays, e.g. in physics [48]. In spite of that, interesting and correct results are still obtained. This outlines how firmly inconsistent infinitesimal reasoning (which is a reasoning with prima facie inconsistent infinitesimals) is entrenched in our scientific community and it means that inconsistency is something that, if unavoidable, should be handled appropriately. Actually, inconsistency would not have been such a problem if the logic used was not explosive [50]. The problem is that our mathematical theory is mostly based on classical logic, which is explosive. Thus, one promising solution is to change the logic into a non-explosive one and this is the main reason for the birth of paraconsistent mathematics which uses paraconsistent logic as its base.

Recent advances in paraconsistent mathematics have been built on developments in set theory [51], geometry [29], arithmetic [28], and also the elementary research at calculus [28] and [9]. A first thorough study to apply paraconsistent logic in real analysis was based on the early work, such as [11] and [27]. While Rosinger in [39] and [38] tried to elaborate the basic structure and use of inconsistent mathematics, McKubre-Jordens and Weber in [26] analysed an axiomatic approach to the real line using paraconsistent logic. They succeed to show that basic field and also compactness theorems hold in that approach. They can also specify where the consistency requirement is necessary. These preliminary works in [26] and [52] show how successful a paraconsistent setting to analysis can be. On the side of the non-standard analysis, it can be seen for example in [1] and [15] that it is still well-studied and still used in many areas.

The underlying ideas of the research described on this paper are as follows. We have two languages: 𝔏\mathfrak{L}, the language of real numbers \mathbb{R}, and 𝔏\mathfrak{L^{*}}, the language of the set of hyperreal numbers \mathbb{{}^{*}R}. The language 𝔏\mathfrak{L^{*}} is an extension of the language 𝔏\mathfrak{L}. It can be shown that each of those two sets forms a model for the formulas in its respective language.

Speaking about the hyperreals, the basic idea of this system is to extend the set \mathbb{R} to include infinitesimal and infinite numbers without changing any of the elementary axioms of algebra. Transfer principle holds an important role in the formation of the set \mathbb{{}^{*}R} in showing what are still preserved in spite of this extension. However, there are some problems with the transfer principle, notably its non-computability (see Subsection 2.2). To avoid its use, one can logically think to simply collapsing the two languages into one language 𝔏^\mathfrak{\widehat{L}} which corresponds to the set ^\mathbb{\widehat{R}}, a new set of numbers constructed by combining the two set of axioms of \mathbb{R} and \mathbb{{}^{*}R}. This is what we do here (see Section 3). Nevertheless, there is at least one big problem from this idea: a contradiction.

We can at present consider two possible ways of resolving this contradiction. The first way is to change the base logic into paraconsistent logic. There are many paraconsistent logics that are available at the moment. This could be a good thing, or from another perspective, be an additional difficulty as we need to choose wisely which paraconsistent logic we want to use at first, i.e. which one is the most appropriate or the best for our purpose. But then, to be able to do this, we need to know beforehand which criteria to use and this, in itself, is still an open question.

The second way we could consider is to have a subsystem in our theory. This idea arose from a specific reasoning strategy, Chunk & Permeate, which was introduced by Brown and Priest in 2003 [9]. Using this strategy, we divide our set ^\mathbb{\widehat{R}} into some consistent chunks and build some permeability relations between them. This process leads us to the creation of the sets \mathbb{R^{Z}} and <\mathbb{R^{Z_{<}}}. In our view, this idea is more sensible and promises to be more useful than the first. Moreover, after further analysing this idea, we produce some new interesting and useful notions that will be worth to explore even further (see Sections 4-6).

The aim of this paper is to build a new model of the nonstandard analysis. By having the new set produces in this paper, we would have real numbers, infinities, and infinitesimals in one set and would still be able to do our “usual” analysis in, and with this set. Moreover, in terms of Gödel’s second incompleteness theorem, if we can build a new structure for nonstandard mathematical analysis which is resilient to contradiction, we would open the door to having not just a sound, but a complete mathematical theory. To put it simply, like Weber said, ‘In light of Gödel’s result, an inconsistent foundation for mathematics is the only remaining candidate for completeness’ [50].

2 Some Preliminaries

To understand and carry our investigation, it is essential to have an accurate grasp of the received view about formal language, the reals and hyperreals, and the paraconsistent logic. In turn, to analyse these matters, it will be useful to fix some terminologies.

2.1 Reals, Hyperreals, and Their Respective Languages

Formal language is built by its syntax and semantics. Here are the symbols that are used in our language:

variables : aa bb cc \dots x1x_{1} x2x_{2} \dots
grammatical signs : ( ) ,
connectives : \land \lor ¬\neg \rightarrow
quantifiers : \forall \exists
constant symbols : 11 2.5-2.5 π\pi 5\sqrt{5} \dots
function symbols : ++ - sin\sin tan\tan \dots
relation symbols : == << >> \leq \geq \dots

Like in natural language, a sentence is built by its term. Terms and sentences are defined as usual. Example 2.1 uses the simple language \mathfrak{I} to build some statements about integer numbers, \mathbb{Z}.

Example 2.1.

In addition to our usual connectives, variables, quantifiers and grammatical symbols in \mathbb{Z}, \mathfrak{I} also contains:

constant symbols : ,2,1,0,1,2,\dots,-2,-1,0,1,2,\dots
function symbols : q(x)=x2q(x)=x^{2}
add(x,y)=x+y\text{add}(x,y)=x+y
mul(x,y)=x×y\text{mul}(x,y)=x\times y
relation symbols : P(x) for “x is positive”P(x)\textnormal{ for ``}x\textnormal{ is positive''}
E(x,y) for “x and y are equal”E(x,y)\textnormal{ for ``}x\textnormal{ and }y\textnormal{ are equal''}

In this language \mathfrak{I}, one can translate an English statement ‘squaring any integer number will give a positive number’ as x\forall x P(s(x))P(s(x)).

Now we can define a language for the set hyperreals. We define a language 𝔏\mathfrak{L} whose every sentence, if true in reals, is also true in hyperreals.

Definition 2.2 (Language 𝔏\mathfrak{L}).

The language 𝔏\mathfrak{L} consists of the usual defined variables, connectives, and grammatical signs in \mathbb{R}, and the following:

constant symbols : one symbol for every real number
function symbols : one symbol for every real-valued function of any finite
number real variables
relation symbols : one symbol for every relation on real numbers of any finite
number real variables

Semantics in our language is described by its model. This model gives an interpretation of the sentences of the language, such that we may know whether they are true or false in that model.

Definition 2.3 (Model of a Language).

Suppose that we have a language 𝔄\mathfrak{A}. A model for 𝔄\mathfrak{A} consists of:

  1. 1.

    a set AA so that each constant symbol in 𝔄\mathfrak{A} corresponds to an element of AA,

  2. 2.

    a set FF of functions on AA so that each function symbol in 𝔄\mathfrak{A} corresponds to a function in FF,

  3. 3.

    a set RR of relations on AA so that each relation symbol in 𝔄\mathfrak{A} corresponds to a relation in RR.

Example 2.4.

For our language \mathfrak{I} over integer numbers, its model is the set A=A=\mathbb{Z} with several functions and relations already well-defined in \mathbb{Z}.

Theorem 2.5 (Reals as a Model).

The real number system \mathbb{R} is a model for the language 𝔏\mathfrak{L}.

Proof.

Take A=A=\mathbb{R} and FF and RR as set of all functions and relations, respectively, which are already well-defined in \mathbb{R}. ∎

Then, by using the definition of a model, we defined what hyperreal number system is.

Definition 2.6.

A hypereal number system is a model for the language 𝔏\mathfrak{L} that, in addition to all real numbers, contains infinitesimal and infinite numbers.

Now suppose that \mathbb{{}^{*}R} is the set of all hyperreal numbers. Our goal now is to show that \mathbb{{}^{*}R} is a model for the language 𝔏\mathfrak{L}. To show this, we need to extend the definition of relations and functions on \mathbb{R} into \mathbb{{}^{*}R}. This extension can also be seen in [18].

Definition 2.7 (Extended Relation).

Let RR be a kk-variable relation on \mathbb{R}, i.e. for every x1,x2,,xkx_{1},x_{2},\dots,x_{k}, R(x1,x2,,xk)R(x_{1},x_{2},\dots,x_{k}) is a sentence that is either true or false. The extension of RR to \mathbb{{}^{*}R} is denoted by R{}^{*}R. Suppose that 𝐱1,𝐱2,,𝐱k\mathbf{x}_{1},\mathbf{x}_{2},\dots,\mathbf{x}_{k} are any hyperreal numbers whose form is {x1n},{x2n},,{xkn}\{x_{1n}\},\{x_{2n}\},\dots,\{x_{kn}\}, respectively. We define R(𝐱𝟏,𝐱𝟐,,𝐱𝐤){}^{*}R(\mathbf{x_{1}},\mathbf{x_{2}},\dots,\mathbf{x_{k}}) as true iff

{n|R(x1n,x2n,,xkn)\{n|R(x_{1n},x_{2n},\dots,x_{kn}) is true in }\mathbb{R}\}

is big.222A ‘big set’ is a set of natural numbers so large that it includes all natural numbers with the possible exception of finitely many [18]. Otherwise, R(𝐱𝟏,𝐱𝟐,,𝐱𝐤)R(\mathbf{x_{1}},\mathbf{x_{2}},\dots,\mathbf{x_{k}}) is false.

Example 2.8.

Suppose that

={1,2,3,4,5,}ⓩ=\{1,2,3,4,5,...\}

By taking k=1k=1 in Definition 2.7, we are able to have a relation I(x)=I(x)=xx is an integer”. The relation I()I(ⓩ) is true. This is because the set of indexes where relation I(x)I(x) is tru, is a big set. Thus, we conclude that ⓩ is actually a hyperinteger.

Definition 2.9 (Extended Function).

Let ff be an kk-variables function on \mathbb{R}. The extension of ff to \mathbb{{}^{*}R} is denoted by f{}^{*}f. Suppose that 𝐱1,𝐱2,,𝐱k\mathbf{x}_{1},\mathbf{x}_{2},\dots,\mathbf{x}_{k} are any hyperreal numbers whose form is {x1n},{x2n},,{xkn}\{x_{1n}\},\{x_{2n}\},\dots,\{x_{kn}\}, respectively. We define f(𝐱1,𝐱2,,𝐱k){}^{*}f(\mathbf{x}_{1},\mathbf{x}_{2},\dots,\mathbf{x}_{k}) by

f(𝐱1,𝐱2,,𝐱k)={f(x11,x21,,xk1),f(x12,x22,,xk2),f(x13,x23,,xk3),}{}^{*}f(\mathbf{x}_{1},\mathbf{x}_{2},\dots,\mathbf{x}_{k})=\{f(x_{11},x_{21},\dots,x_{k1}),f(x_{12},x_{22},\dots,x_{k2}),f(x_{13},x_{23},\dots,x_{k3}),\dots\}

Example 2.10.

Suppose that

={2,4,6,8,}ⓔ=\{2,4,6,8,...\}

By taking k=1k=1 in Definition 2.9, we might have, for example, a well-defined hypersinus function:

sin()={sin(2),sin(4),sin(6),sin(8),}\sin(ⓔ)=\{\sin(2),\sin(4),\sin(6),\sin(8),...\}

Theorem 2.11 (Hyperreals as a Model).

The set \mathbb{{}^{*}R} is a model for the language 𝔏\mathfrak{L} that contains infinitesimals and infinities.

Proof.

Take A=A=\mathbb{{}^{*}R} in Definition 2.3 with all of the functions f{}^{*}f defined in Definition 2.9 and relations R{}^{*}R defined in Definition 2.7. ∎

2.2 Transfer Principle and Its Problems

Definition 2.12 (Transfer Principle).

Let SS be a sentence in 𝔏\mathfrak{L}. The transfer principle says that:

SS is true in the model \mathbb{R} for 𝔏\mathfrak{L} iff SS is true in the model \mathbb{{}^{*}R} for 𝔏\mathfrak{L}.

As Goldblatt said in [18], ‘The strength of nonstandard analysis lies in the ability to transfer properties between \mathbb{R} and \mathbb{{}^{*}R}.’ But, there are some serious problems with the transfer principle. Some of them are: it is non-computable in the sense of there is no good computable representation of the hyperreals to start with; it really depends intrinsically on the mathematical model or language we use; we are prone to get things wrong when not handled correctly (especially because of human error).

Furthermore, there is an all too often overlooked, yet major deficiency with the transfer principle: it performs particularly poorly upon a rather simple “cost-return” analysis. Namely, on the one hand, the mathematical machinery which must be set up in advance in order to use the transfer principle is of such a considerable technical complication and strangeness with a lack of step-by-step intuitive insight that, ever since 1966, when Abraham Robinson published the first major book on the subject — that is, for more than half a century by now — none of the more major mathematicians ever chose to switch to the effective daily use of nonstandard analysis, except for very few among those who have dealt with time continuous stochastic processes, and decided to use the “Loeb Integral” introduced in 1975. On the other hand, relatively few properties of importance can ever be transferred, since they are not — and cannot be, within usual nonstandard analysis — formulated in terms of first order logic.

One possible solution for overcoming (some of) these problems is simply by not using the transfer principle, by throwing together the two sets \mathbb{R} and {}^{*}\mathbb{R}, i,e. combining their languages and axioms. However, Example 2.13 shows that if we just simply combine the two languages, it will pose one big problem of contradictions which can lead to absurdity if we use classical logic. That is the reason why we use paraconsistent logic as it is resilient against local contradiction.

Example 2.13.

Take the well-ordering principle for our example. This principle says that: “every non-empty set of natural numbers contains a least element”. Call a set S={xN:x is infinite}S=\{x\in N^{*}:x\textnormal{ is infinite}\}. Let ss be its least element. Note that ss is infinite and so s1s-1. Thus, s1Ss-1\in S and it makes ss is not the least element. Therefore, there exists ll such that ll is the least element of SS and there is no ll such that ll is the east element of SS.

In addition to the problems with the transfer principle, there are also some downsides of the construction of the set \mathbb{{}^{*}R} itself. Indeed, the usual construction of the hyperreal set \mathbb{R} involves an ultrafilter on \mathbb{N}, the existence of which is justified by appealing to the full Axiom of Choice whose validity is still a great deal to discuss [49, 22]. Moreover, it also relies on some heavy and non-constructive mathematical machineries such as Zorn’s lemma, the Hahn-Banach theorem, Tychonoff’s theorem, the Stone-Cech compactification, or the boolean prime ideal theorem. On the side of the nonstandard analysis itself, there are some critiques as can be seen in [7, 13, 14, 19, 45, 47]. Most of them are related with its non-constructivism and its difficulties to be used in class teaching. This problem can be solved by building a naïve constructive non-standard set and making sure that it is still a useful set by redefining some well-known notions in there.

2.3 Paraconsistent Logics in Mathematics

Generally, paraconsistent logics are logics which permit inference from inconsistent information in a non-trivial fashion [36]. Paraconsistent logics are characterized by rejecting the universal validity of the principle ex contradictione quodlibet (ECQ) which is defined below.

Definition 2.14 (ECQ Principle).

The principle of explosion, ECQ, is the law which states that any statement can be proven from a contradiction.

By admitting the ECQ principle in one theory, if that theory contains a single inconsistency, it becomes absurd or trivial. This is something that, in paraconsistent logics, does not follow necessarily.

Paraconsistent logicians believe that some contradictions does not necessarily make the theory absurd. It just means that one has to be very careful when doing deductions so as to avoid falling from contradiction into an absurdity. In other words, classical and paraconsistent logic treat contradiction in different ways. The former treats contradiction as a global contradiction (making the theory absurd), while the latter treats some contradictions as a local contradiction. In other words, classical logic cannot recognise if there is an interesting structure in the event of a contradiction.

Definition 2.15 (Paraconsistent Logic).

Suppose that AA is a logical statement. A logic is called paraconsistent logic iff

A,B\exists A,B such that A¬ABA\land\lnot A\nvdash B.

The symbol ΓA\Gamma\vdash A simply means that there exists a proof of AA from set of formulas Γ\Gamma, in a certain logic.

There are at least two different approaches to paraconsistent logics. The first is by adding another possible value, both true and false, to classical truth values while the second one is called the relevant-approach. The idea of the relevant-approach is simply to make sure that the conclusion of an implication must be relevant to its premise(s). Those two paraconsistent logics, respectively, are Priest’s Paraconsistent Logic LP and Relevant Logic R.333Note that in general, relevant logic differs from paraconsistent logic. When someone claims that they use relevant logic, it implies that they use paraconsistent logic, but not vice versa. Using paraconsistent logic does not necessarily mean using relevant logic, e.g. the logic LPLP^{\supset} below is not relevant logic. More explanations on each of them can be seen, for example, in [34], [2], [35], [24], and [12].

When we are applying paraconsistent logic to a certain theory, there will be at least two terms that we have to be aware of: inconsistency and incoherence. The first term, inconsistency, is applicable if there occurs a contradiction in a system. Meanwhile, the second term, incoherence, is intended for a system which proves anything (desired or not). In classical logic, there is no difference between these two terms because of the ECQ principle. Thus, if a contradiction arises inside a theory, anything that the author would like to say can be proved or inferred within that theory. This is something that likewise does not have to happen in paraconsistent logic.

In mathematical theory, foundation of mathematics is the study of the basic mathematical concepts and how they form more complex structures and concepts. This study is especially important for learning the structures that form the language of mathematics (formulas, theories, definitions, etc.), structures that often called metamathematical concepts. A philosophical dimension is hence central to this study. One of the most interesting topics in the foundation of mathematics is the foundation of real structure, or analysis.

Generally, it is known that the construction of real numbers is categorised in classical logic — while there is an advancement in paraconsistent logic such as in [4], this has not yet been extensively explored. However, it seems viable to make a further study of real structure by developing paraconsistent foundations of analysis.

3 The Creation of The New Sets

As noted before, the transfer principle is useful as well as fairly problematic at the same time. One way to avoid the unnecessary complications of the transfer principle is by collapsing the two languages involved into one language 𝔏^\mathfrak{\widehat{L}}. Simply collapsing the two, however, causes additional problems. One of the problems that can be expected to appear is contradiction, but we can use a paraconsistent logic to handle this when it arises.

In this section, we construct a new number system ^\mathbb{\widehat{R}} through its axiomatisation by ‘throwing’ the axioms of \mathbb{R} and \mathbb{{}^{*}R} together. The set ^\mathbb{\widehat{R}} is the number system on which the language 𝔏^\mathfrak{\widehat{L}} will be based and it contains positive and negative infinities, and also infinitesimals.444We are not necessarily expecting the resulting system to have contradictions, but we will make sure that we maintain coherency by not allowing contradictions to become an absurdity. It does make sense to insert infinities (and their reciprocals, infinitesimals) into ^\mathbb{\widehat{R}} as some of the contradictions in mathematics come from their existence and also because they are still used in today’s theory as can be seen in [48].

3.1 The New Set ^\mathbb{\widehat{R}}

For the sake of clarity, Axioms 3.13.7 give the axiomatisation of the number system ^\mathbb{\widehat{R}}.

Axiom 3.1 (Additive Property of ^\mathbb{\widehat{R}}).

In the set ^\mathbb{\widehat{R}}, there is an operator ++ that satisfies:
A1: For any x,y^,x+y^x,y\in\mathbb{\widehat{R}},x+y\in\mathbb{\widehat{R}}. A2: For any x,y^,x+y=y+xx,y\in\mathbb{\widehat{R}},x+y=y+x. A3: For any x,y,z^,(x+y)+z=x+(y+z)x,y,z\in\mathbb{\widehat{R}},(x+y)+z=x+(y+z). A4: There is 0^0\in\mathbb{\widehat{R}} such that x+0=xx+0=x for all x^x\in\mathbb{\widehat{R}}. A5: For each x^x\in\mathbb{\widehat{R}}, there is x^-x\in\mathbb{\widehat{R}} such that x+(x)=0x+(-x)=0.

Axiom 3.2 (Multiplicative Property of ^\mathbb{\widehat{R}}).

In the set ^\mathbb{\widehat{R}}, there is an operator \cdot that satisfies:
M1: For any x,y^,xy^x,y\in\mathbb{\widehat{R}},x\cdot y\in\mathbb{\widehat{R}}. M2: For any x,y^,xy=yxx,y\in\mathbb{\widehat{R}},x\cdot y=y\cdot x. M3: For any x,y,z^,(xy)z=x(yz)x,y,z\in\mathbb{\widehat{R}},(x\cdot y)\cdot z=x\cdot(y\cdot z). M4: There is 1^1\in\mathbb{\widehat{R}} such that (1=0)(1=0)\rightarrow\bot and x^\forall x\in\mathbb{\widehat{R}} x1=xx\cdot 1=x. M5: For each x^x\in\mathbb{\widehat{R}}, if (x=0)(x=0)\rightarrow\bot, then there is y^y\in\mathbb{\widehat{R}} such that xy=1x\cdot y=1.

Axiom 3.3 (Distributive Property of ^\mathbb{\widehat{R}}).

For all x,y,z^x,y,z\in\mathbb{\widehat{R}}, x(y+z)=(xy)+(xz)x\cdot(y+z)=(x\cdot y)+(x\cdot z).

Axiom 3.4 (Total Partial Order Property of ^\mathbb{\widehat{R}}).

There is a relation \leq in ^\mathbb{\widehat{R}}, such that for each x,y,z^x,y,z\in\mathbb{\widehat{R}}:
O1: Reflexivity: xxx\leq x, O2: Transitivity: (xyyz)xz(x\leq y\land y\leq z)\rightarrow x\leq z, O3: Antisymmetry: (xyyx)x=y(x\leq y\land y\leq x)\leftrightarrow x=y, O4: Totality: (xy)yx(x\leq y\rightarrow\bot)\rightarrow y\leq x. O5: Addition order: xyx+zy+zx\leq y\rightarrow x+z\leq y+z. O6: Multiplication order: (xyz0)xzyz(x\leq y\land z\geq 0)\rightarrow xz\leq yz.

Axiom 3.5 (Completeness Property of ^\mathbb{\widehat{R}}).

Every non-empty bounded above subset of ^\mathbb{\widehat{R}} has a least upper bound (see Definition LABEL:def:ClassicalLeastUpperBound).

Axiom 3.6 (Infinitesimal Property of ^\mathbb{\widehat{R}}).

The set ^\mathbb{\widehat{R}} has an infinitesimal (see Definition 3.8 for what infinitesimal is).

Axiom 3.7 (Archimedean Property of ^\mathbb{\widehat{R}}).

For all x,y>0x,y>0, n\exists n such that nx>ynx>y (see Definition 3.9 for the operator >>).

Notice what Axiom 3.5 and Axioms 3.6 and 3.7 cause. The first axiom, which states the completeness property of ^\mathbb{\widehat{R}}, causes computability issues in our set. The last two axioms, they cause the consistency trouble. Infinity and infinitesimals are formally defined in Definition 3.8.

Definition 3.8.

An element x^x\in\mathbb{\widehat{R}} is

  • infinitesimal iff n\forall n\in\mathbb{N} |x|<1n|x|<\frac{1}{n};

  • finite iff r\exists r\in\mathbb{R} |x|<r|x|<r;

  • infinite iff r\forall r\in\mathbb{R} |x|>r|x|>r;

  • appreciable iff xx is finite but not an infinitesimal;

By using notation ϵ\epsilon as an infinitesimal, an infinity ω\omega is defined as a reciprocal of ϵ\epsilon, i.e. 1ϵ\frac{1}{\epsilon}.

Definition 3.9.

For any numbers x,y^x,y\in\mathbb{\widehat{R}},

  1. 1.

    xy:=x<yx\geq y:=x<y\rightarrow\bot

  2. 2.

    x<y:=xy(x=y)x<y:=x\leq y\land(x=y\rightarrow\bot)

  3. 3.

    x>y:=xy(x=y)x>y:=x\geq y\land(x=y\rightarrow\bot)

If we look further, the set ^\mathbb{\widehat{R}} is actually an inconsistent set. Example 3.10 gives one of these contradictions.

Example 3.10.

Suppose that we have a set S={x^:|x|<1n for all n}S=\{x\in\mathbb{\widehat{R}}:\left\lvert x\right\rvert<\frac{1}{n}\textnormal{ for all }n\in\mathbb{N}\}. In other words, the set SS consists of all infinitesimals in ^\mathbb{\widehat{R}}. It is easily proven that SS is not empty and bounded above. So by Completeness Axiom, SS has a least upper bound. Suppose that zz is its least upper bound (which also means that zz must be an infinitesimal). Because zz is an infinitesimal, 2z2z is also an infinitesimal and this means 2z2z is also in ^\mathbb{\widehat{R}}. By using Definition 3.9, z<2zz<2z and so, zz is not the least upper bound of SS. Now suppose that sup S=2zS=2z. The same argument can be used to show that 2z2z is not supremum of SS but 3z3z. We can build this same argument infinitely to show that there does not exist ss such that sup S=sS=s. Thus, we have s:sup S=s\exists s:\textnormal{sup }S=s and s:sup S=s\nexists s:\textnormal{sup }S=s.

This kind of contradiction forces us to use a non-explosive logic such as paraconsistent logic instead of classical logic to do our reasoning in ^\mathbb{\widehat{R}}. Furthermore, we choose a particular paraconsistent reasoning strategy, Chunk and Permeate (C&P), to resolve our dilemma. The detail explanations of this strategy can be seen in [9].

3.2 Chunks in 𝔏^\mathfrak{\widehat{L}} and The Creation of The Set <\mathbb{R^{Z_{<}}}

Using the C&P strategy, we divided 𝔏^\mathfrak{\widehat{L}} into some consistent chunks — naturally, there might be several ways to do it (e.g. one can have an idea to divide the original set into two, three, or even more chunks). Nevertheless, we found out that one particular way to have just two different chunks, as provided in here, is the most interesting one as it leads to the creation of a new model. One chunk is a set which contains Axioms (3.13.4,3.6), while the other chunk is a set which contains Axioms (3.13.5,3.7). The consistency of each chunk was proved by providing a model for each of them.

3.2.1 Model for The First Chunk

One of the possible — and interesting — chunks is a set consists of Axioms 3.1-3.4 and Axiom 3.6. Here we proved the consistency of this chunk and also some corollaries that we have.

It is well-established that the set of hyperreals, \mathbb{{}^{*}R}, clearly satisfied those axioms [6]. Nevertheless, the construction of hyperreals \mathbb{{}^{*}R} depends on highly non-constructive arguments. In particular, it requires an axiom of set theory, the well-ordering principle, which assumes into existence something that cannot be constructed [21]. Here we proposed to take a look at a simpler set. Remember that our set has to contain not just \mathbb{R}, but also infinitesimals and infinities (and the combinations of the two). We take \mathbb{R^{Z}}, functions from integers to real numbers, as our base set. The member of \mathbb{R^{Z}} consists of standard and non-standard parts. The standard part of a certain number simply shows its finite element (the real part), while the non-standard part shows its infinite or infinitesimal part (see Definition 3.11).

Definition 3.11 (Member of \mathbb{R^{Z}}).

A typical member of \mathbb{R^{Z}} has the form 𝐱=ϵi,x^,ϵj\mathbf{x}=\langle\epsilon_{-i},\widehat{x},{\epsilon_{j}}\rangle where x^\widehat{x}\in\mathbb{R} and ϵn\epsilon_{n} denotes the sequence of the constant part of infininitesimals if n>0n>0, and infinities if n<0n<0.

Notice that the symbol x^\widehat{x} in Definition 3.11 signs the standard part of a number in \mathbb{R^{Z}}. Thus, the member of \mathbb{R^{Z}} can be seen as a sequence of infinite numbers. Example 3.12 gives an overview of how to write a number as a member of \mathbb{R^{Z}}.

Example 3.12 (Numbers in \mathbb{R^{Z}}).
  1. 1.

    The number 1 is written as 1=0,0,1^,0,0,\textbf{1}=\langle\dots 0,0,\widehat{1},0,0,\dots\rangle.

  2. 2.

    The number ϵ\epsilon is written as ϵ=0,0^,1,0,\mathbf{\epsilon}=\langle\dots 0,\widehat{0},1,0,\dots\rangle.

  3. 3.

    The number ω\omega (one of the infinities) is written as ω=0,1,0^,0,\mathbf{\omega}=\langle\dots 0,1,\widehat{0},0,\dots\rangle.

  4. 4.

    The number 2+2ϵω22+2\epsilon-\omega^{2} is written as 𝟐+𝟐ϵω𝟐=,0,1,0,2^,2,0,\mathbf{2+2\epsilon-\omega^{2}}=\langle\dots,0,-1,0,\widehat{2},2,0,\dots\rangle.

By using this form, all of the possible numbers can be written in \mathbb{R^{Z}}. However, this infinite form is problematic in a number of ways. For example, multiplication cannot be easily defined and there might exist multiple inverses if the set \mathbb{R^{Z}} was going to be used (see Example 3.16 in [32, p. 47]). Because of this, the semi-infinite form is motivated and the modified set is denoted by <\mathbb{R^{Z_{<}}}. The only difference between <\mathbb{R^{Z_{<}}} and \mathbb{R} is that, for any number 𝐱\mathbf{x}, we will not have an infinite sequence on the left side of its standard part. See example 3.13 and compare to Example 3.12.

Example 3.13 (Numbers in <\mathbb{R^{Z_{<}}}).
  1. 1.

    A number 1 is written as 1=1^,0,0,\textbf{1}=\langle\widehat{1},0,0,\dots\rangle.

  2. 2.

    A number ϵ\epsilon is written as ϵ=0^,1,0,\mathbf{\epsilon}=\langle\widehat{0},1,0,\dots\rangle.

  3. 3.

    A number ω\omega (one of the infinities) is written as ω=1,0^,0,\mathbf{\omega}=\langle 1,\widehat{0},0,\dots\rangle.

  4. 4.

    A number 2+2ϵω22+2\epsilon-\omega^{2} is written as 𝟐+𝟐ϵω𝟐=1,0,2^,2,0,\mathbf{2+2\epsilon-\omega^{2}}=\langle-1,0,\widehat{2},2,0,\cdots\rangle.

Some of the properties of the set <\mathbb{R^{Z_{<}}} are as follows, while their proof (when needed) and some examples of them can be seen in [32, p. 51–57].

Definition 3.14 (Addition and Multiplication in <\mathbb{R^{Z_{<}}}).

For any number 𝐱=xz\mathbf{x}=\langle x_{z}\rangle and 𝐲=yz\mathbf{y}=\langle y_{z}\rangle in <\mathbb{R^{Z_{<}}}, define:

𝐱+𝐲=xz+yz:z\displaystyle\mathbf{x}+\mathbf{y}=\langle x_{z}+y_{z}:z\in\mathbb{Z}\rangle

and 𝐱×𝐲\mathbf{x}\times\mathbf{y} is calculated by:

𝐱×𝐲=(i=maiϵi)×^(j=nbjϵj)=(kckϵk),\displaystyle\mathbf{x}\times\mathbf{y}=\left(\sum_{i=-m}a_{i}\epsilon^{i}\right)\widehat{\times}\left(\sum_{j=-n}b_{j}\epsilon^{j}\right)=\left(\sum_{k\in\mathbb{Z}}c_{k}\epsilon^{k}\right),

where ck=i+j=kaibjc_{k}=\sum_{i+j=k}a_{i}b_{j}.

Definition 3.15 (Order in <\mathbb{R^{Z_{<}}}).

The set <\mathbb{R^{Z_{<}}} is endowed with ^\widehat{\leq}, the lexicographical ordering.

Proposition 3.16.

The set <\mathbb{R^{Z_{<}}} satisfies the additive property in Axiom 3.1.

Proposition 3.17.

The set <\mathbb{R^{Z_{<}}} satisfies the multiplicative property in Axiom 3.2.

Proposition 3.18.

The set <\mathbb{R^{Z_{<}}} satisfies the distributive property in Axiom 3.3.

Proposition 3.19.

The set <\mathbb{R^{Z_{<}}} satisfies the total order property in Axiom 3.4.

The following results show how to find an inverse of any members of <\mathbb{R^{Z_{<}}} and its uniqueness property.

Proposition 3.20.

The number 𝟏+ω\mathbf{1+\omega} has a unique inverse.

Proposition 3.21.

The number ω+^ϵ\mathbf{\omega\widehat{+}\epsilon} has a unique inverse.

Lemma 3.22.

For any r,sr,s\in\mathbb{R}, a number rϵ+^sωr\mathbf{\epsilon}\widehat{+}s\mathbf{\omega} has a unique inverse.

Lemma 3.23.

For any rr\in\mathbb{R}, a number r+^ωr\widehat{+}\mathbf{\omega} (or r+^ϵ)r\widehat{+}\mathbf{\epsilon}) has a unique inverse.

Theorem 3.24.

For any number 𝐱<\mathbf{x}\in\mathbb{R^{Z_{<}}}, 𝐱\mathbf{x} has a unique inverse.

3.2.2 Model for The Second Chunk

The most evident model for this second chunk is the set of real numbers, \mathbb{R}. Thus, so far, we have already had two chunks in 𝔏^\mathfrak{\widehat{L}} and we proved their consistencies by providing a model for each of them.

3.3 Grossone Theory and The Set <\mathbb{R^{Z_{<}}}

Theories that contain infinities have always been an issue and have attracted much research, for example [10, 17, 20, 25, 37]. Note that the arithmetic developed for infinite numbers was quite different with respect to the finite arithmetic that we are used to dealing with. For example, Sergeyev in [44] created the Grossone theory. The basic idea of this theory is to treat infinity as an ‘normal’ number, so that our usual arithmetic rules apply. He named this infinite number Grossone and denoted it with . The four axioms that form this Grossone theory and its details can be seen [44].

It is very important to emphasis that is a number, and so it works as a usual number. For example, there exist numbers such as 100,3+16,ln\text{\char 172}-100,\text{\char 172}^{3}+16,\ln\text{\char 172}, and etc. Also for instance, 1<\text{\char 172}-1<\text{\char 172}.555This is unlike the way the usual infinity, \infty, behaves where for example, 1=\infty-1=\infty. It also differs from how Cantor’s cardinal numbers behave. The introduction for this new number makes us able to rewrite the set of natural numbers \mathbb{N} as:

={1,2,3,4,5,6,,2,1,}\mathbb{N}=\{1,2,3,4,5,6,\dots,\text{\char 172}-2,\text{\char 172}-1,\text{\char 172}\}.

Furthermore, adding the Infinite Unit Axiom (IUA) to the axioms of natural numbers will define the set of extended natural numbers \mathbb{{}^{*}N}:

={1,2,3,4,5,6,,1,+1,,21,2,}\mathbb{{}^{*}N}=\{1,2,3,4,5,6,\dots,\text{\char 172}-1,\text{\char 172}\,\text{\char 172}+1,\dots,\text{\char 172}^{2}-1,\text{\char 172}^{2},\dots\},

and the set \mathbb{{}^{*}Z}, extended integer numbers, can be defined from there.

Here we argued that our new set <\mathbb{R^{Z_{<}}} provides the model of Grossone theory and therefore proves its consistency rather in a deftly way.

Definition 3.25.

In our system <\mathbb{R^{Z_{<}}}, the number is written as:

=1,0^,0,\text{\char 172}=\langle 1,\widehat{0},0,\dots\rangle.

Proposition 3.26.

For every finite number 𝐫<\mathbf{r}\in\mathbb{R^{Z_{<}}}, 𝐫<\mathbf{r}<\text{\char 172}.

Proof.

The order in set <\mathbb{R^{Z_{<}}} is defined lexicographically. Now suppose that 𝐫=0,r^,0,0,\mathbf{r}=\langle 0,\widehat{r},0,0,\dots\rangle where r^\widehat{r}\in\mathbb{R}. Then it is clear that 𝐫<\mathbf{r}<\text{\char 172}. ∎

Proposition 3.27.

All of the equations in the Identity Axiom of Grossone theory are also hold in <\mathbb{R^{Z_{<}}}.

The fractional form of can also be defined in <\mathbb{R^{Z_{<}}} as:

for any nn\in\mathbb{N}, n=1n,0^,0,\frac{\text{\char 172}}{n}=\langle\frac{1}{n},\widehat{0},0,\dots\rangle.

Speaking about the inverse, one of the advantages of having the set <\mathbb{R^{Z_{<}}} is to be able to see what the inverse of a number looks like, not like in the Grossone theory. See Example 3.28 for more details.

Example 3.28.

In Grossone theory, the inverse of 1+\frac{1}{\text{\char 172}}+\text{\char 172} is just 11+\frac{1}{\frac{1}{\text{\char 172}}+\text{\char 172}}. While in our set <\mathbb{R^{Z_{<}}}, 1+\frac{1}{\text{\char 172}}+\text{\char 172} is written as ϵ+ω\epsilon+\omega and its inverse is

0^,1,0,1,0,1,0,1,0,=ϵϵ3+ϵ5ϵ7+\langle\widehat{0},1,0,-1,0,1,0,-1,0,\dots\rangle=\epsilon-\epsilon^{3}+\epsilon^{5}-\epsilon^{7}+\dots.

In other words, the more explicit form of 11+\frac{1}{\frac{1}{\text{\char 172}}+\text{\char 172}} is a series (1)n+1(2n1)(-1)^{n+1}\text{\char 172}^{-(2n-1)} for n=1,2,3,n=1,2,3,\dots\in\mathbb{N}.

In [23], Gabriele Lolli analysed and built a formal foundation of the Grossone theory based on Peano’s second order arithmetic. He also gave a slightly different notion of some axioms that Sergeyev used. One of the important theorems in Lolli’s paper is the proof that Grossone theory – or at least his version of it – is consistent. However, as he said also in [23], ‘The statement of the theorem is of course conditional, as apparent from the proof, upon the consistency of PAμ2\text{PA}_{\mu}^{2}’ while ‘its model theoretic proof is technically rather demanding’.

Thus, through what presented in this subsection, we have proposed a new way to prove the consistency of Grossone theory by providing a straightforward model of it. There is no need for complicated model-theoretic proofs. The set <\mathbb{R^{Z_{<}}} is enough to establish the consistency of Grossone theory in general. Moreover, the development in the next sessions can also be seen, at least in part, as a contribution to the development of Grossone theory.

4 Topology on The Set <\mathbb{R^{Z_{<}}}

Some topological properties of the set <\mathbb{R^{Z_{<}}} are discussed in this section. However, there are a number of definitions and issues that should be addressed first in order to understand how those properties will be applied to our set properly.

4.1 Metrics in <\mathbb{R^{Z_{<}}}

We defined what is meant by a distance (metric) between each pair of elements of <\mathbb{R^{Z_{<}}}.

Definition 4.1.

A metric ρ\rho in a set XX is a function

ρ:X×X[0,)\displaystyle\rho:X\times X\rightarrow\left[0,\infty\right)

where for all x,y,zXx,y,z\in X, these four conditions are satisfied:

  1. 1.

    ρ(x,y)0\rho(x,y)\geq 0,

  2. 2.

    ρ(x,y)=0\rho(x,y)=0 if and only if x=yx=y,

  3. 3.

    ρ(x,y)=d(y,x)\rho(x,y)=d(y,x),

  4. 4.

    ρ(x,z)ρ(x,y)+ρ(y,z)\rho(x,z)\leq\rho(x,y)+\rho(y,z).

When that function ρ\rho satisfies all of the four conditions above except the second one, ρ\rho is called a pseudo-metric666It is not without reason that we introduced the concept of the pseudo-metric here. This kind of metric will make sense when we are in \mathbb{R}. For example, the distance between 0 and ϵ\epsilon is 0 as our lens is not strong enough to distinguish those two numbers in \mathbb{R}. on XX.

Now we define two functions d\operatorname{\texttt{d}} and dψ\operatorname{\texttt{d}_{\texttt{$\psi$}}} in <\mathbb{R^{Z_{<}}} as follows:

Definition 4.2.

For all 𝐱,𝐲<\mathbf{x},\mathbf{y}\in\mathbb{R^{Z_{<}}},

d:<×<< and dψ:<×<\displaystyle\operatorname{\texttt{d}}:\mathbb{R^{Z_{<}}}\times\mathbb{R^{Z_{<}}}\rightarrow\mathbb{R^{Z_{<}}}\textnormal{ and }\operatorname{\texttt{d}_{\texttt{$\psi$}}}:\mathbb{R^{Z_{<}}}\times\mathbb{R^{Z_{<}}}\rightarrow\mathbb{R}

where d(𝐱,𝐲)=|𝐲𝐱|\operatorname{\texttt{d}}(\mathbf{x},\mathbf{y})=\left\lvert\mathbf{y}-\mathbf{x}\right\rvert and dψ(𝐱,𝐲)=St(|𝐲𝐱|)\operatorname{\texttt{d}_{\texttt{$\psi$}}}(\mathbf{x},\mathbf{y})=\texttt{St}(\left\lvert\mathbf{y}-\mathbf{x}\right\rvert).

It can be easily verified that d\operatorname{\texttt{d}} is a metric in <\mathbb{R^{Z_{<}}} (and so (<,d)\left(\mathbb{R^{Z_{<}}},\operatorname{\texttt{d}}\right) forms a metric space) and dψ\operatorname{\texttt{d}_{\texttt{$\psi$}}} is a pseudo-metric in <\mathbb{R^{Z_{<}}} (and so (<,dψ)\left(\mathbb{R^{Z_{<}}},\operatorname{\texttt{d}_{\texttt{$\psi$}}}\right) forms a pseudo-metric space).

4.2 Balls and Open Sets in <\mathbb{R^{Z_{<}}}

Now that we have the notion of distance in <\mathbb{R^{Z_{<}}}, we can define what it means to be an open set in <\mathbb{R^{Z_{<}}} by first defining what a ball is in <\mathbb{R^{Z_{<}}}.

Definition 4.3.

A ball of radius 𝐲\mathbf{y} around the point 𝐱<\mathbf{x}\in\mathbb{R^{Z_{<}}} is

B𝐱(𝐲)={𝐳<d1(𝐱,𝐳)<𝐲},\displaystyle B_{\mathbf{x}}(\mathbf{y})=\{\mathbf{z}\in\mathbb{R^{Z_{<}}}\mid d_{1}(\mathbf{x},\mathbf{z})<\mathbf{y}\},

where d1(𝐱,𝐲)d_{1}(\mathbf{x},\mathbf{y}) is either d(𝐱,𝐲)\operatorname{\texttt{d}}(\mathbf{x},\mathbf{y}) or dψ(𝐱,𝐲)\operatorname{\texttt{d}_{\texttt{$\psi$}}}(\mathbf{x},\mathbf{y}).

We require this additional definition in order to set forth our explanation about balls properly:

Definition 4.4.

The sets Δm\Delta^{m} and Δm\Delta^{\underset{\downarrow}{m}} are defined as follows:

Δm={𝐱:𝐱=amϵm}\Delta^{m}=\{\mathbf{x}:\mathbf{x}=a_{m}\mathbf{\epsilon}^{m}\} and Δm=nmΔn\Delta^{\underset{\downarrow}{m}}=\bigcup\limits_{n\geq m}\Delta^{n},

where m{0}m\in\mathbb{N}\cup\{0\}, ama_{m}\in\mathbb{R} and am0a_{m}\neq 0 whenever m1m\geq 1.

We have to be careful here as unlike in classical topology, there are different notions of balls that can be described as follows. The first possible notion of balls is when we use d\operatorname{\texttt{d}} as our metric and having 𝐲>0\mathbf{y}>0 as our radius. In this case, in <\mathbb{R^{Z_{<}}}, the ball around a point 𝐱\mathbf{x} with 𝐲\mathbf{y} radius is an interval (𝐱𝐲,𝐱+𝐲)(\mathbf{x}-\mathbf{y},\mathbf{x}+\mathbf{y}). Note that by using 𝐲\mathbf{y} as the radius, beside having the usual balls with “real” radius (St-balls) (that is when 𝐲=r\mathbf{y}=r\in\mathbb{R}), we also have some infinitesimally small balls (e-balls) when 𝐲=𝔢Δm\mathbf{y}=\mathfrak{e}\in\Delta^{\underset{\downarrow}{m}} for any given mm. The second possible notion is while we use the same metric d\operatorname{\texttt{d}}, we have 1/n\nicefrac{{1}}{{n}} for some nn\in\mathbb{N} as its radius. This produces balls (rat-balls) in the form of (𝐱1/n,𝐱+1/n)\left(\mathbf{x}-\nicefrac{{1}}{{n}},\mathbf{x}+\nicefrac{{1}}{{n}}\right). The third possibility is by using dψ\operatorname{\texttt{d}_{\texttt{$\psi$}}} as our metric. In this case, interestingly, the balls around a point 𝐱\mathbf{x} with 1/n\nicefrac{{1}}{{n}} radius will be in the form of the following set:

{𝐲St(𝐲)(St(𝐱)1/n,St(𝐱)+1/n)}\{\mathbf{y}\mid\texttt{St}(\mathbf{y})\in(\texttt{St}(\mathbf{x})-\nicefrac{{1}}{{n}},\texttt{St}(\mathbf{x})+\nicefrac{{1}}{{n}})\}.

We call this kind of balls as psi-balls. See Table 1 for the summary of these possibilities of balls in our sets.

Table 1: Some types of balls both in \mathbb{R} and <\mathbb{R^{Z_{<}}}
Balls in \mathbb{R} and <\mathbb{R^{Z_{<}}}
The Set The Metric The Form of The Balls
\mathbb{R} ρ(x,y)\rho(x,y)   =|xy|=\left\lvert x-y\right\rvert Bx(r)B_{x}(r)   =(xr,x+r)=(x-r,x+r)
<\mathbb{R^{Z_{<}}} d(𝐱,𝐲)\operatorname{\texttt{d}}(\mathbf{x},\mathbf{y})   =|𝐱𝐲|=\left\lvert\mathbf{x}-\mathbf{y}\right\rvert B𝐱(𝐫)B_{\mathbf{x}}(\mathbf{r})   =(𝐱𝐫,𝐱+𝐫)=(\mathbf{x}-\mathbf{r},\mathbf{x}+\mathbf{r})
<\mathbb{R^{Z_{<}}} d(𝐱,𝐲)\operatorname{\texttt{d}}(\mathbf{x},\mathbf{y})   =|𝐱𝐲|=\left\lvert\mathbf{x}-\mathbf{y}\right\rvert B𝐱(1/n)=(𝐱1/n,𝐱+1/n)B_{\mathbf{x}}(\nicefrac{{1}}{{n}})=(\mathbf{x}-\nicefrac{{1}}{{n}},\mathbf{x}+\nicefrac{{1}}{{n}})
<\mathbb{R^{Z_{<}}} dψ(𝐱,𝐲)=St(|𝐱𝐲|)\operatorname{\texttt{d}_{\texttt{$\psi$}}}(\mathbf{x},\mathbf{y})=\texttt{St}(\left\lvert\mathbf{x}-\mathbf{y}\right\rvert) B𝐱(1/n)=BSt(𝐱)(1/n)={𝐲St(𝐲)B_{\mathbf{x}}(\nicefrac{{1}}{{n}})=B_{\texttt{St}(\mathbf{x})}(\nicefrac{{1}}{{n}})=\{\mathbf{y}\mid\texttt{St}(\mathbf{y})\in
            (St(𝐱)1/n,St(𝐱)+1/n)}\quad(\texttt{St}(\mathbf{x})-\nicefrac{{1}}{{n}},\texttt{St}(\mathbf{x})+\nicefrac{{1}}{{n}})\}
Remark 4.5.

In \mathbb{R}, the 𝔢\mathfrak{e}-ball does not exist, whereas in <\mathbb{R^{Z_{<}}} there are infinitely many 𝔢\mathfrak{e}-balls around every point there.

Finally, we defined what it means to be an open set in <\mathbb{R^{Z_{<}}}. Notice that because we had two notions of ball in our set, i.e. St-balls and 𝔢\mathfrak{e}-ball, it led us to two different notions of openness as follows.

Definition 4.6 (St-open).

A subset O<O\subseteq\mathbb{R^{Z_{<}}} is St-open iff

xO\forall x\in O n s.t. Bx(1n)O\exists n\in\mathbb{N}\textnormal{ s.t. }B_{x}\left(\frac{1}{n}\right)\subseteq O.

Definition 4.7 (𝔢\mathfrak{e}-open).

A subset O<O\subseteq\mathbb{R^{Z_{<}}} is e-open iff

xO\forall x\in O 𝔢Δm s.t. Bx(𝔢)O\exists\mathfrak{e}\in\Delta^{\underset{\downarrow}{m}}\textnormal{ s.t. }B_{x}(\mathfrak{e})\subseteq O.

Remember that the set Δm\Delta^{\underset{\downarrow}{m}} is defined in Definition 4.4.

Example 4.8.

The interval (2,3)(\textbf{2},\textbf{3}) in <\mathbb{R^{Z_{<}}} is St-open and also 𝔢\mathfrak{e}-open.

Example 4.9.

The interval (0,𝔢)(\textbf{0},\mathbf{\mathfrak{e}}) in <\mathbb{R^{Z_{<}}} is 𝔢\mathfrak{e}-open, but not St-open.

Example 4.9 gives us the theorem below:

Theorem 4.10.

For any set U<U\subseteq\mathbb{R^{Z_{<}}},

<⊧̸If U is St-open, then U is 𝔢-open.\displaystyle\mathbb{R^{Z_{<}}}\not\models\textit{If }U\textit{ is {St}-open, then }U\textit{ is }\mathfrak{e}\textit{-open}.

Using the two definition of openness given in Definitions 4.6 and 4.7, we defined what it means by two points are topologically distinguishable. There are also two different notions of distinguishable points as can be seen in Definitions 4.11 and 4.12.

Definition 4.11 (St-distinguishable).

Any two points in <\mathbb{R^{Z_{<}}} are St-distinguishable if and only if there is a St-open set containing precisely one of the two points.

Definition 4.12 (𝔢\mathfrak{e}-distinguishable).

Any two points in <\mathbb{R^{Z_{<}}} are e-distinguishable if and only if there is an 𝔢\mathfrak{e}-open set containing precisely one of the two points.

4.3 Topological Spaces

Definition 4.13.

Let XX be a non-empty set and τ\uptau a collection of subsets of XX such that:

  1. 1.

    XτX\in\uptau,

  2. 2.

    τ\emptyset\in\uptau,

  3. 3.

    If O1,O2,,OnτO_{1},O_{2},\dots,O_{n}\in\uptau, then k=1nOkτ\bigcap_{k=1}^{n}O_{k}\in\uptau,

  4. 4.

    If OατO_{\alpha}\in\uptau for all αA\alpha\in A, then αAOατ\bigcup_{\alpha\in A}O_{\alpha}\in\uptau.

The pair of objects (X,τ)(X,\uptau) is called a topological space where XX is called the underlying set, the collection τ\uptau is called the topology in XX, and the members of τ\uptau are called open sets.

Note that if τ\uptau is the collection of open sets of a metric space (𝒳,ρ)(\mathcal{X},\rho), then (𝒳,τ)(\mathcal{X},\uptau) is a topological metric space, i.e. a topological space associated with the metric space (X,ρ)(X,\rho). There are at least three interesting topologies in <\mathbb{R^{Z_{<}}} as can be seen in Definition 4.14 below.

Definition 4.14.

The standard topology τSt\uptau_{\textnormal{St}} on the set <\mathbb{R^{Z_{<}}} is the topology generated by all unions of St-balls. The 𝔢\mathfrak{e}-topology in <\mathbb{R^{Z_{<}}}, τ𝔢\uptau_{\mathfrak{e}}, is the topology generated by all unions of 𝔢\mathfrak{e}-balls and the third topology in <\mathbb{R^{Z_{<}}} is pseudo-topology, τψ\uptau_{\psi}, when it is induced by dψ\operatorname{\texttt{d}_{\texttt{$\psi$}}}.

Axiom 4.15.

(<,τn),(\mathbb{R^{Z_{<}}},\uptau_{n}), (<,τ𝔢),(\mathbb{R^{Z_{<}}},\uptau_{\mathfrak{e}}), and (<,τψ)(\mathbb{R^{Z_{<}}},\uptau_{\psi}) form topological metric space with d\operatorname{\texttt{d}} as their metrics (for the first two) and dψ\operatorname{\texttt{d}_{\texttt{$\psi$}}} for the third one.

Theorem 4.16.

(<,τn)(\mathbb{R^{Z_{<}}},\uptau_{n}) is not a Hausdorff space but it is a preregular space.777Hausdorff space and preregular space are defined as usual.

Proof.

<\mathbb{R^{Z_{<}}} does not form a Hausdorff space because under the topology τSt\uptau_{\textnormal{St}}, there are two distinct points, ϵ=0^,1\mathbf{\epsilon}=\langle\widehat{0},1\rangle and ϵ+1=1^,1\mathbf{\epsilon}+\textbf{1}=\langle\widehat{1},1\rangle for example, which are not neighbourhood-separable. It is impossible to separate those two points with St-ballss as 1/n>𝔢\nicefrac{{1}}{{n}}>\mathfrak{e} for every nn\in\mathbb{N} and 𝔢Δm\mathfrak{e}\in\Delta^{\underset{\downarrow}{m}}. However, it is a preregular space as every pair of two St-distinguishable points in <\mathbb{R^{Z_{<}}} can be separated by two disjoint neighbourhoods. This follows directly from Definition 4.11. ∎

Theorem 4.17.

(<,τ𝔢)(\mathbb{R^{Z_{<}}},\uptau_{\mathfrak{e}}) is a non-connected space and it forms a Hausdorff space.

Proof.

We observe that for all 𝐱𝟎<\mathbf{x_{0}}\in\mathbb{R^{Z_{<}}} and 𝔢Δm\mathfrak{e}\in\Delta^{\underset{\downarrow}{m}}, the balls B𝔢(𝐱𝟎)B_{\mathfrak{e}}(\mathbf{x_{0}}) are 𝔢\mathfrak{e}-open and so is the whole space. To show that <\mathbb{R^{Z_{<}}} is not connected, let

S1\displaystyle S_{1} ={𝐱<(𝐱0) or (𝐱>0 and 𝐱Δm)} and\displaystyle=\{\mathbf{x}\in\mathbb{R^{Z_{<}}}\mid(\mathbf{x}\leq\textbf{0})\textnormal{ or }(\mathbf{x}>\textbf{0}\textnormal{ and }\mathbf{x}\in\Delta^{\underset{\downarrow}{m}})\}\textnormal{ and }
S2\displaystyle S_{2} ={𝐱<(𝐱>0) and 𝐱Δm}.\displaystyle=\{\mathbf{x}\in\mathbb{R^{Z_{<}}}\mid(\mathbf{x}>\textbf{0})\textnormal{ and }\mathbf{x}\notin\Delta^{\underset{\downarrow}{m}}\}.

The sets S1S_{1} and S2S_{2} are 𝔢\mathfrak{e}-open, disjoint and moreover, we have that <=S1S2\mathbb{R^{Z_{<}}}=S_{1}\cup S_{2} (and so <\mathbb{R^{Z_{<}}} is not connected). For any 𝐱,𝐲<\mathbf{x},\mathbf{y}\in\mathbb{R^{Z_{<}}}, Bx(d(x,y)/2)B_{x}(\nicefrac{{\operatorname{\texttt{d}}(x,y)}}{{2}}) and By(d(x,y)/2)B_{y}(\nicefrac{{\operatorname{\texttt{d}}(x,y)}}{{2}}) are open and disjoint. Thus, <\mathbb{R^{Z_{<}}} forms a Hausdorff space. ∎

We will now state the usual definition of the basis of a topology τ\uptau.

Definition 4.18.

Let (X,τ)(X,\uptau) be a topological space. A basis for the topology τ\uptau is a collection \mathcal{B} of subsets from τ\uptau such that every UτU\in\uptau is the union of some collections of sets in \mathcal{B}, i.e.

Uτ\forall U\in\uptau, \exists\mathcal{B^{*}}\subseteq\mathcal{B} s.t. U=BBU=\bigcup\limits_{B\in\mathcal{B^{*}}}B

Example 4.19.

On \mathbb{R} with its usual topology, the set ={(a,b):a<b}\mathcal{B}=\{(a,b):a<b\} is a topological basis.

Definition 4.20.

Let (X,τ)(X,\uptau) be a topological space and let xXx\in X. A local basis of xx is a collection of open neighbourhoods of xx, x\mathcal{B}_{x}, such that for all UτU\in\uptau with xUx\in U, Bx\exists B\in\mathcal{B}_{x} such that xBUx\in B\subset U.

Definition 4.21.

Let (X,τ)(X,\uptau) be a topological space. Then (X,τ)(X,\uptau) is first-countable if every point xXx\in X has a countable local basis.

Definition 4.22.

Let (X,τ)(X,\uptau) be a topological space. Then (X,τ)(X,\uptau) is second-countable if there exists a basis \mathcal{B} of τ\uptau that is countable.

Theorem 4.23.

(<,τ𝔢)(\mathbb{R^{Z_{<}}},\tau_{\mathfrak{e}}) is first countable but not second-countable.888Note that the space (<,τn)(\mathbb{R^{Z_{<}}},\tau_{n}) is still second-countable.

Proof.

From Axiom 4.15 and because every metric space is first-countable, it follows that (<,τ𝔢)(\mathbb{R^{Z_{<}}},\tau_{\mathfrak{e}}) is first-countable. However, there cannot be any countable bases in τ𝔢\tau_{\mathfrak{e}} as the uncountably many open sets Ox=(x𝔢,x+𝔢)O_{x}=(x-\mathfrak{e},x+\mathfrak{e}) are disjoint. ∎

5 Calculus on <\mathbb{R^{Z_{<}}}

It has been proved previously that the set <\mathbb{R^{Z_{<}}} forms a field. Remember that for any 𝐱<\mathbf{x}\in\mathbb{R^{Z_{<}}},

𝐱=xn,x(n1),,x2,x1,x^,x1,x2,x3,\mathbf{x}=\langle x_{-n},x_{-(n-1)},\dots,x_{-2},x_{-1},\widehat{x},x_{1},x_{2},x_{3},\dots\rangle

where

St(𝐱)=x^(\mathbf{x})=\widehat{x},

Nst(𝐱)ϵ={x1,x2,x3,}{}_{\epsilon}(\mathbf{x})=\{x_{1},x_{2},x_{3},\dots\}, and

Nst(𝐱)ω={xn,x(n1),x1}{}_{\omega}(\mathbf{x})=\{x_{-n},x_{-(n-1)},\dots x_{-1}\}.

In other words, for every 𝐱<\mathbf{x}\in\mathbb{R^{Z_{<}}},

𝐱=\mathbf{x}= Nstω(𝐱)(\mathbf{x}) ++ St(𝐱)(\mathbf{x}) ++ Nstϵ(𝐱)(\mathbf{x}).

Note that we can think of St(), Nstϵ(), and Nstω() as linear functions – that is for any 𝐱,𝐲<\mathbf{x},\mathbf{y}\in\mathbb{R^{Z_{<}}} and a constant cc\in\mathbb{R},

St(𝐱+𝐲)=(\mathbf{x}+\mathbf{y})=St(𝐱)+(\mathbf{x})+St(𝐲)(\mathbf{y}), St(c𝐱)=c(c\mathbf{x})=cSt(𝐱)(\mathbf{x}),

Nstϵ(𝐱+𝐲)=(\mathbf{x}+\mathbf{y})=Nstϵ(𝐱)+(\mathbf{x})+Nstϵ(𝐲)(\mathbf{y}), Nstϵ(c𝐱)=c(c\mathbf{x})=cNStϵ(𝐱)(\mathbf{x}),

Nstω(𝐱+𝐲)=(\mathbf{x}+\mathbf{y})=Nstω(𝐱)+(\mathbf{x})+Nstω(𝐲)(\mathbf{y}), and Nstω(c𝐱)=c(c\mathbf{x})=cNstω(𝐱)(\mathbf{x}).

Definition 5.1.

Suppose that niϵ(𝐱)\texttt{ni}_{\epsilon}(\mathbf{x}) denotes the non-infinitesimal part of 𝐱<\mathbf{x}\in\mathbb{R^{Z_{<}}}, i.e. niϵ(𝐱)=Nstω(𝐱)+St(𝐱)\texttt{ni}_{\epsilon}(\mathbf{x})=\texttt{Nst}_{\omega}(\mathbf{x})+\texttt{St}(\mathbf{x}) and function in <\mathbb{R^{Z_{<}}} be defined in the usual way. Then a function ff in <\mathbb{R^{Z_{<}}} is microstable if and only if

niϵ(f(x+ϵ))=niϵ(f(x))\texttt{ni}_{\epsilon}(f(x+\epsilon))=\texttt{ni}_{\epsilon}(f(x)),

Example 5.2.

Suppose that a function ff in <\mathbb{R^{Z_{<}}} is defined as follows:

f(𝐱)={1,if St(𝐱)>00,else.f(\mathbf{x})=\begin{cases}1,&\text{if }\texttt{St}(\mathbf{x})>0\\ 0,&\text{else.}\end{cases}

Then f(𝐱)f(\mathbf{x}) is a microstable function.

Theorem 5.3.

Microstability is closed under addition, multiplication, and composition.999The proof of this theorem can be seen in [32, p. 73].

Now for every function ff defined in <\mathbb{R^{Z_{<}}}, we are going to have the operator Derf\texttt{Der}_{f} which takes a 2-tuple in (×<)\left(\mathbb{R}\times\mathbb{R^{Z_{<}}}\right) as its input and returns a member of <\mathbb{R^{Z_{<}}} as the output, i.e.:

Derf:×<<\texttt{Der}_{f}:\mathbb{R}\times\mathbb{R^{Z_{<}}}\rightarrow\mathbb{R^{Z_{<}}}.

Eventually, this operator will be called a derivative of ff.

Using Newton’s original definition (and a slight change of notation), if a function f(x)f(x) is differentiable, then its derivative is given by:

Derf(𝐱,ϵ)=f(𝐱+ϵ)f(𝐱)ϵ.\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon})=\cfrac{f(\mathbf{x}+\mathbf{\epsilon})-f(\mathbf{x})}{\mathbf{\epsilon}}. (3)

Now suppose that we want to find a derivative of ff where ff is a function defined in <\mathbb{R^{Z_{<}}}. We can certainly use Equation 3 to calculate it as that equation holds for any function ff. But how is this calculation related to the calculus practised in classical mathematics? Note that using Newton’s definition to calculate the derivative will necessarily involve an inconsistent step. This inconsistency is located in the treatment given to the infinitesimal number. Thus it makes sense that in order to explore the problem posed above, we will use a paraconsistent reasoning strategy which is called Chunk and Permeate.

5.1 Chunk and Permeate for Derivative in <\mathbb{R^{Z_{<}}}

Details on the Chunk & Permeate reasoning strategy can be seen in [9]. Before applying this strategy for the derivative in <\mathbb{R^{Z_{<}}}, define a set EE which consists of any algebraic terms such that they satisfy:

St(Derf(𝐱,ϵ))=f (x)\texttt{St}(\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon}))=f\textnormal{ }^{\prime}(x),

where f (x)f\textnormal{ }^{\prime}(x) denotes the usual derivative of ff in \mathbb{R}. We will need this set EE when we try to define the permeability relation between chunks.

Proposition 5.4.

The set EE as defined above is inhabited.

Proof.

We want to show that the set EE has at least one element in it. It is clear that the identity function id(x)=x\texttt{id}(x)=x is in EE because for all ϵ\epsilon:

St(Derx(𝐱,ϵ))\displaystyle\texttt{St}(\texttt{Der}_{x}(\mathbf{x},\mathbf{\epsilon})) =St(𝐱+ϵ𝐱ϵ)\displaystyle=\texttt{St}\left(\cfrac{\mathbf{x}+\mathbf{\epsilon}-\mathbf{x}}{\mathbf{\epsilon}}\right)
=St(ϵϵ)\displaystyle=\texttt{St}\left(\cfrac{\mathbf{\epsilon}}{\mathbf{\epsilon}}\right)
=St(1)=1=f (x).\displaystyle=\texttt{St}(\textbf{1})=1=f\textnormal{ }^{\prime}(x).

Theorem 5.5.

If ff and gg are microstable functions in EE and cc is any real constant, then

  1. 1.

    f±gf\pm g are in EE,

  2. 2.

    cfcf is in EE,

  3. 3.

    fgfg is in EE,

  4. 4.

    fg\cfrac{f}{g} is in EE, and

  5. 5.

    fgf\circ g is in EE.

The proof of the above theorem is rather long and so can be seen in [32, p. 76].

Now we are ready to construct the chunk and permeate structure, called ^\mathfrak{\widehat{R}}, which is formally written as ^={ΣS,ΣT},ρ,T\mathfrak{\widehat{R}}=\langle\{\Sigma_{S},\Sigma_{T}\},\rho,T\rangle where the source chunk ΣS\Sigma_{S} is the language of <\mathbb{R^{Z_{<}}}, the target chunk ΣT\Sigma_{T} is the language of \mathbb{R}, and ρ\rho is the permeability relation between SS and TT.

The source chunk ΣS\Sigma_{S}

As stated before, this chunk is actually the language of the set <\mathbb{{R^{Z_{<}}}} and therefore, it consists of all six of its axioms. The source chunk requires one additional axiom to define what it means by derivative. This additional axiom can be stated as:

S1: Df=Derf(𝐱,ϵ)Df=\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon})

where Derf(𝐱,ϵ)\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon}) is defined in Equation 3.

The target chunk ΣT\Sigma_{T}

Again, the target chunk contains the usual axiom for the set or real numbers, \mathbb{R}. There is only one additional axiom needed for this chunk:

T1: 𝐱\forall\mathbf{x} 𝐱=St(𝐱)\mathbf{x}=\texttt{St}(\mathbf{x}).

Note that the axiom T1 above is actually equivalent to saying that 𝐱\forall\mathbf{x} Nst(𝐱)=0\texttt{Nst}(\mathbf{x})=0.

The permeability relation

The permeability relation ρ(S,T)\rho(S,T) is the set of equations of the form

Df=gDf=g

where fEf\in E. The function gg which is permeated by this permeability relation will be the first derivative of ff in \mathbb{R}. This permeability relation shows that the derivative notion is permeable to the set \mathbb{R}.

Example 5.6.

Suppose that f(𝐱)=3𝐱f(\mathbf{x})=\textbf{3}\mathbf{x} for all 𝐱\mathbf{x}. First, working within ΣS\Sigma_{S}, the operator DD is applied to ff such that:

Df\displaystyle Df =Derf(𝐱,ϵ)\displaystyle=\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon})
=3(𝐱+ϵ)3𝐱ϵ\displaystyle=\cfrac{\textbf{3}(\mathbf{x}+\mathbf{\epsilon})-\textbf{3}\mathbf{x}}{\mathbf{\epsilon}}
=3ϵϵ=3.\displaystyle=\cfrac{\textbf{3}\mathbf{\epsilon}}{\mathbf{\epsilon}}=\textbf{3}.

Note that St(Derf(𝐱,ϵ))=St(3)=3=f (x)\texttt{St}(\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon}))=\texttt{St}(\textbf{3})=3=f\textnormal{ }^{\prime}(x), and so f(x)Ef(x)\in E. Permeating the last equation of DfDf above to ΣT\Sigma_{T} gives us:

Df=3\displaystyle Df=3

and so the derivative of f(x)=3xf(x)=3x is 33.

Example 5.7.

Suppose that f(𝐱)=𝐱2+2𝐱+3f(\mathbf{x})=\mathbf{x}^{\textbf{2}}+\textbf{2}\mathbf{x}+\textbf{3} for all 𝐱\mathbf{x}. First, working within ΣS\Sigma_{S}, the operator DD is applied to ff such that:

Df\displaystyle Df =Derf(𝐱,ϵ)\displaystyle=\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon})
=(𝐱+ϵ)2+2(𝐱+ϵ)+3𝐱22𝐱3ϵ\displaystyle=\cfrac{(\mathbf{x}+\mathbf{\epsilon})^{\textbf{2}}+\textbf{2}(\mathbf{x}+\mathbf{\epsilon})+\textbf{3}-\mathbf{x}^{\textbf{2}}-\textbf{2}\mathbf{x}-\textbf{3}}{\mathbf{\epsilon}}
=2𝐱ϵ+ϵ2+2ϵϵ\displaystyle=\cfrac{\textbf{2}\mathbf{x}\mathbf{\epsilon}+\mathbf{\epsilon}^{\textbf{2}}+\textbf{2}\mathbf{\epsilon}}{\mathbf{\epsilon}}
=2𝐱+ϵ+2.\displaystyle=\textbf{2}\mathbf{x}+\mathbf{\epsilon}+\textbf{2}.

Note that the standard part of 2𝐱+ϵ+2\textbf{2}\mathbf{x}+\mathbf{\epsilon}+\textbf{2} will depend on the domain of 𝐱\mathbf{x}. That is:

St(2𝐱+ϵ+2)={2x+2,if x2,else.\texttt{St}(\textbf{2}\mathbf{x}+\mathbf{\epsilon}+\textbf{2})=\begin{cases}2x+2,&\text{if }x\in\mathbb{R}\\ 2,&\text{else.}\end{cases}

In other words, if (and only if) Nst(𝐱)=0\texttt{Nst}(\mathbf{x})=0, i.e. xx\in\mathbb{R}, DfDf can be permeated into ΣT\Sigma_{T}. Thus, if xx is a real number, then we have the derivative of f(x)=x2+2x+3=2x+2f(x)=x^{2}+2x+3=2x+2.

Example 5.8.

Suppose that f(𝐱)=sign(𝐱)f(\mathbf{x})=\operatorname{sign}(\mathbf{x}) is defined as:

sign(𝐱)={𝟏,if St(𝐱)>0𝟎,if St(𝐱)=0𝟏,if St(𝐱)<0\operatorname{sign}(\mathbf{x})=\begin{cases}\mathbf{1},&\text{if }\texttt{St}(\mathbf{x})>0\\ \mathbf{0},&\text{if }\texttt{St}(\mathbf{x})=0\\ \mathbf{-1},&\text{if }\texttt{St}(\mathbf{x})<0\end{cases}

First, working within ΣS\Sigma_{S}, the operator DD is applied to ff so that:

Df\displaystyle Df =Derf(𝐱,ϵ)\displaystyle=\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon})
=sign(𝐱+ϵ)sign(𝐱)ϵ\displaystyle=\cfrac{\operatorname{sign}(\mathbf{x}+\mathbf{\epsilon})-\operatorname{sign}(\mathbf{x})}{\mathbf{\epsilon}}
=0 (because x St(𝐱)=St(𝐱+ϵ))\displaystyle=0\text{ (because $\forall x$ $\texttt{St}(\mathbf{x})=\texttt{St}(\mathbf{x}+\epsilon)$)}

Note that St(Derf(𝐱,ϵ))=St(𝟎)=0=f (x)\texttt{St}(\texttt{Der}_{f}(\mathbf{x},\mathbf{\epsilon}))=\texttt{St}(\mathbf{0})=0=f\textnormal{ }^{\prime}(x), and so f(x)Ef(x)\in E. Permeating the last equation of DfDf above to ΣT\Sigma_{T} gives us:

Df=0\displaystyle Df=0

and so the derivative of f(𝐱)=sign(𝐱)f(\mathbf{x})=\operatorname{sign}(\mathbf{x}) is 0 for all 𝐱\mathbf{x}. Notice that this is not the case in \mathbb{R}, where the derivative of the sign\operatorname{sign} function at x=0x=0 is not defined because of its discontinuity. However, this is not really a bizarre behaviour because if we look very closely at the infinitesimal neighbourhood of 𝐱\mathbf{x} when St(𝐱)=0\texttt{St}(\mathbf{x})=0, the function sign(𝐱)\operatorname{sign}(\mathbf{x}) will look like a straight horizontal line and so it makes a perfect sense to have 0 as the slope of the tangent line there. Moreover, this phenomenon also happens in distribution theory where sign\operatorname{sign} function has its derivative everywhere.

5.2 Transcendental Functions in <\mathbb{R^{Z_{<}}}

As we know, there are some special functions defined in real numbers and two of them are the trigonometric and the exponential functions. How then are these functions defined in <\mathbb{R^{Z_{<}}}? Here we propose to define them using power series.

The first two trigonometric functions that we are going to discuss are the sin\sin and cos\cos functions. Using the MacLaurin power series, these two functions are defined as follows:

sin(𝐱)=n=0(1)n(2n+1)! 𝐱2n+1\displaystyle\sin(\mathbf{x})=\sum\limits_{n=0}\cfrac{(-1)^{n}}{(2n+1)!}\text{ }\mathbf{x}^{2n+1} (4)

and

cos(𝐱)=n=0(1)n(2n)! 𝐱2n.\displaystyle\cos(\mathbf{x})=\sum\limits_{n=0}\cfrac{(-1)^{n}}{(2n)!}\text{ }\mathbf{x}^{2n}. (5)

The exponential function is defined as:

exp(𝐱)=n=01n!𝐱n.\displaystyle\exp(\mathbf{x})=\sum\limits_{n=0}\cfrac{1}{n!}\mathbf{x}^{n}. (6)

Note that the MacLaurin polynomial is just a special case of Taylor polynomial with regards to how the function is approximated at 𝐱=0\mathbf{x}=\textbf{0}.

Example 5.9.

Suppose that we have 𝐱=x+aϵ=x^,a,0,0,\mathbf{x}=x+a\mathbf{\epsilon}=\langle\widehat{x},a,0,0,\dots\rangle where x,ax,a\in\mathbb{R}. We want to know what sin(𝐱)\sin(\mathbf{x}) is. Based on Equation 4,

sin(𝐱)=sin(x+ϵ)=(x+ϵ)13!(x+ϵ)3+15!(x+ϵ)517!(x+ϵ)7+\displaystyle\sin(\mathbf{x})=\sin(x+\mathbf{\epsilon})=(x+\mathbf{\epsilon})-\tfrac{\textbf{1}}{\textbf{3}!}(x+\mathbf{\epsilon})^{\textbf{3}}+\tfrac{\textbf{1}}{\textbf{5}!}(x+\mathbf{\epsilon})^{\textbf{5}}-\tfrac{\textbf{1}}{\textbf{7}!}(x+\mathbf{\epsilon})^{\textbf{7}}+\dots

Our task now is to find all the members of Nstϵ(sin(𝐱))\texttt{Nst}_{\mathbf{\epsilon}}(\mathbf{\sin(x)}) and also St(sin(𝐱))\texttt{St}(\mathbf{\sin(x)}). These are shown in Table 2. Note that from the way the sin\sin function is defined, xi=0x_{i}=0 xiNstω(sin(𝐱))\forall x_{i}\in\texttt{Nst}_{\omega}(\sin(\mathbf{x})). Thus from Table 2, we get:

sin(𝐱)\displaystyle\sin(\mathbf{x}) =sin(x+aϵ)\displaystyle=\sin(x+a\mathbf{\epsilon})
=sin(x)^,acos(x),a22!sin(x),a33!cos(x),a44!sin(x),a55!cos(x),,\displaystyle=\langle\widehat{\sin(x)},a\cos(x),-\tfrac{a^{2}}{2!}\sin(x),-\tfrac{a^{\textbf{3}}}{\textbf{3}!}\cos(x),\tfrac{a^{4}}{4!}\sin(x),\tfrac{a^{\textbf{5}}}{\textbf{5}!}\cos(x),\dots\rangle,

and we also get

sin(ϵ)=0^,1,0,13!,0,15!,=ϵ13!ϵ3+15!ϵ5\displaystyle\sin(\mathbf{\epsilon})=\langle\widehat{0},\textbf{1},0,-\tfrac{\textbf{1}}{\textbf{3}!},0,\tfrac{\textbf{1}}{\textbf{5}!},\dots\rangle=\mathbf{\epsilon}-\tfrac{\textbf{1}}{\textbf{3}!}\mathbf{\epsilon}^{\textbf{3}}+\tfrac{\textbf{1}}{\textbf{5}!}\mathbf{\epsilon}^{\textbf{5}}-\dots

for an infinitesimal angle ϵ\mathbf{\epsilon}.

Table 2: St(sin(𝐱))\texttt{St}(\sin(\mathbf{x)}) and the first four members of Nstϵ(sin(𝐱))\texttt{Nst}_{\epsilon}(\sin(\mathbf{x}))
Expanded Form Simplified Form
real-part =𝐱13!𝐱3+15!𝐱517!𝐱7+=n=01n(2n+1)! 𝐱2n+1\begin{array}[]{ll}=&\mathbf{x}-\tfrac{1}{3!}\mathbf{x}^{3}+\tfrac{1}{\textbf{5}!}\mathbf{x}^{\textbf{5}}-\tfrac{1}{\textbf{7}!}\mathbf{x}^{\textbf{7}}+\dots\\ =&\sum_{n=0}\tfrac{-1^{n}}{(2n+1)!}\text{ }\mathbf{x}^{2n+1}\end{array} =sin(𝐱)=\sin(\mathbf{x})
ϵ\mathbf{\epsilon}-part =aϵ12!𝐱2aϵ+14!aϵ𝐱416!aϵ𝐱6+=ϵ(a12!a𝐱2+14!a𝐱416!a𝐱6+)=ϵn=01n(2n)! a𝐱2n\begin{array}[]{ll}=&a\mathbf{\epsilon}-\tfrac{1}{2!}\mathbf{x}^{2}a\mathbf{\epsilon}+\tfrac{1}{4!}a\mathbf{\epsilon}\mathbf{x}^{4}-\tfrac{1}{6!}a\mathbf{\epsilon}\mathbf{x}^{6}+\dots\\ =&\mathbf{\epsilon}(a-\tfrac{1}{2!}a\mathbf{x}^{2}+\tfrac{1}{4!}a\mathbf{x}^{4}-\tfrac{1}{6!}a\mathbf{x}^{6}+\dots)\\ =&\mathbf{\epsilon}\sum_{n=0}\tfrac{-1^{n}}{(2n)!}\text{ }a\mathbf{x}^{2n}\end{array} =ϵ(acos(𝐱))=\mathbf{\epsilon}(a\cos(\mathbf{x}))
ϵ2\mathbf{\epsilon}^{2}-part =33!a2ϵ2𝐱+105!a2ϵ2𝐱3217!a2ϵ2𝐱5+=12!a2ϵ2𝐱+24!a2ϵ2𝐱336!a2ϵ2𝐱5+=ϵ2(12!𝐱+24!a2𝐱336!a2𝐱5+)=ϵ2n=01n+1(n+1)(2n+2)! a2𝐱2n+1\begin{array}[]{ll}=&-\tfrac{3}{3!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}+\tfrac{\textbf{10}}{\textbf{5}!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{3}-\tfrac{21}{\textbf{7}!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{\textbf{5}}+\dots\\ =&-\tfrac{1}{2!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}+\tfrac{2}{4!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{3}-\tfrac{3}{6!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{\textbf{5}}+\dots\\ =&\mathbf{\epsilon}^{2}(-\tfrac{1}{2!}\mathbf{x}+\tfrac{2}{4!}a^{2}\mathbf{x}^{3}-\tfrac{3}{6!}a^{2}\mathbf{x}^{\textbf{5}}+\dots)\\ =&\mathbf{\epsilon}^{2}\sum_{n=0}\tfrac{-1^{n+1}(n+1)}{(2n+2)!}\text{ }a^{2}\mathbf{x}^{2n+1}\end{array} =ϵ2(a22sin(𝐱))=\mathbf{\epsilon}^{2}(-\tfrac{a^{2}}{2}\sin(\mathbf{x}))
ϵ3\mathbf{\epsilon}^{3}-part =13!a3ϵ3+105!a3ϵ3𝐱2357!a3ϵ3𝐱4+849!a3ϵ3𝐱6=ϵ3(13!a3𝐱0+105!a3𝐱2357!a3𝐱4+849!a3𝐱6)=ϵ3n=01n+16(2n)! a3𝐱2n\begin{array}[]{ll}=&-\tfrac{1}{3!}a^{3}\mathbf{\epsilon}^{3}+\tfrac{\textbf{10}}{\textbf{5}!}a^{3}\mathbf{\epsilon}^{3}\mathbf{x}^{2}-\tfrac{\textbf{3{5}}}{\textbf{7}!}a^{3}\mathbf{\epsilon}^{3}\mathbf{x}^{4}+\tfrac{\textbf{84}}{\textbf{9}!}a^{3}\mathbf{\epsilon}^{3}\mathbf{x}^{6}-\dots\\ =&\mathbf{\epsilon}^{3}(-\tfrac{1}{3!}a^{3}\mathbf{x}^{0}+\tfrac{\textbf{10}}{\textbf{5}!}a^{3}\mathbf{x}^{2}-\tfrac{\textbf{3{5}}}{\textbf{7}!}a^{3}\mathbf{x}^{4}+\tfrac{\textbf{84}}{\textbf{9}!}a^{3}\mathbf{x}^{6}-\dots)\\ =&\mathbf{\epsilon}^{3}\sum_{n=0}\tfrac{-1^{n+1}}{6(2n)!}\text{ }a^{3}\mathbf{x}^{2n}\end{array} =ϵ3(a36cos(𝐱))=\mathbf{\epsilon}^{3}(-\tfrac{a^{3}}{6}\cos(\mathbf{x}))
ϵ4\mathbf{\epsilon}^{4}-part =55!a4ϵ4𝐱357!a4ϵ4𝐱3+1269!a4ϵ4𝐱533011!a4ϵ4𝐱7+=ϵ4(14!𝐱56!a4𝐱3+148!a4𝐱53010!a4𝐱7+)=ϵ4n=01n24(2n+1)! a4𝐱2n+1\begin{array}[]{ll}=&\tfrac{\textbf{5}}{\textbf{5}!}a^{4}\mathbf{\epsilon}^{4}\mathbf{x}-\tfrac{\textbf{3{5}}}{\textbf{7}!}a^{4}\mathbf{\epsilon}^{4}\mathbf{x}^{3}+\tfrac{\textbf{126}}{\textbf{9}!}a^{4}\mathbf{\epsilon}^{4}\mathbf{x}^{\textbf{5}}-\tfrac{\textbf{3{30}}}{11!}a^{4}\mathbf{\epsilon}^{4}\mathbf{x}^{\textbf{7}}+\dots\\ =&\mathbf{\epsilon}^{4}(\tfrac{1}{4!}\mathbf{x}-\tfrac{\textbf{5}}{6!}a^{4}\mathbf{x}^{3}+\tfrac{\textbf{14}}{8!}a^{4}\mathbf{x}^{\textbf{5}}-\tfrac{\textbf{30}}{\textbf{10}!}a^{4}\mathbf{x}^{\textbf{7}}+\dots)\\ =&\mathbf{\epsilon}^{4}\sum_{n=0}\tfrac{-1^{n}}{24(2n+1)!}\text{ }a^{4}\mathbf{x}^{2n+1}\end{array} =ϵ4(a424sin(𝐱))=\mathbf{\epsilon}^{4}(\tfrac{a^{4}}{24}\sin(\mathbf{x}))
Example 5.10.

Suppose that we have 𝐱=x+aϵ\mathbf{x}=x+a\epsilon where x,ax,a\in\mathbb{R}. Here we try to find what cos𝐱\cos\mathbf{x} is. With a similar method to the one used in Example 5.9, we have a calculation like what is shown in Table 3.

Table 3: St(cos(𝐱))\texttt{St}(\cos(\mathbf{x)}) and the first two members of Nstϵ(cos(𝐱))\texttt{Nst}_{\epsilon}(\cos(\mathbf{x}))
real-part =112!𝐱2+14!𝐱416!𝐱6+=n=01n(2n)! 𝐱2n\begin{array}[]{ll}=&1-\tfrac{1}{2!}\mathbf{x}^{2}+\tfrac{1}{\textbf{4}!}\mathbf{x}^{\textbf{4}}-\tfrac{1}{\textbf{6}!}\mathbf{x}^{\textbf{6}}+\dots\\ =&\sum_{n=0}\tfrac{-1^{n}}{(2n)!}\text{ }\mathbf{x}^{2n}\end{array} =cos(𝐱)=\cos(\mathbf{x})
ϵ\mathbf{\epsilon}-part =22!aϵ𝐱+44!aϵ𝐱366!aϵ𝐱5+=ϵ(a𝐱+13!a𝐱315!a𝐱5+)=ϵn=01n+1(2n+1)! a𝐱2n+1\begin{array}[]{ll}=&-\tfrac{2}{2!}a\mathbf{\epsilon}\mathbf{x}+\tfrac{\textbf{4}}{\textbf{4}!}a\mathbf{\epsilon}\mathbf{x}^{3}-\tfrac{\textbf{6}}{\textbf{6}!}a\mathbf{\epsilon}\mathbf{x}^{5}+\dots\\ =&\mathbf{\epsilon}(-a\mathbf{x}+\tfrac{1}{3!}a\mathbf{x}^{3}-\tfrac{1}{5!}a\mathbf{x}^{5}+\dots)\\ =&\mathbf{\epsilon}\sum_{n=0}\tfrac{-1^{n+1}}{(2n+1)!}\text{ }a\mathbf{x}^{2n+1}\end{array} =ϵ(asin(𝐱))=\mathbf{\epsilon}(-a\sin(\mathbf{x}))
ϵ2\mathbf{\epsilon}^{2}-part =12!a2ϵ2+64!a2ϵ2𝐱2156!a2ϵ2𝐱4+288!a2ϵ2𝐱6=ϵ2n=01n+1(n+1)(2n+1)(2n+2)! a2𝐱2n=ϵ2n=0121n+1(2n)! 𝐱2n\begin{array}[]{ll}=&-\tfrac{1}{2!}a^{2}\mathbf{\epsilon}^{2}+\tfrac{\textbf{6}}{\textbf{4}!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{2}-\tfrac{\textbf{15}}{\textbf{6}!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{\textbf{4}}+\tfrac{\textbf{28}}{8!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{\textbf{6}}-\dots\\ =&\mathbf{\epsilon}^{2}\sum_{n=0}\tfrac{-1^{n+1}(n+1)(2n+1)}{(2n+2)!}\text{ }a^{2}\mathbf{x}^{2n}\\ =&\mathbf{\epsilon}^{2}\sum_{n=0}\tfrac{1}{2}\tfrac{-1^{n+1}}{(2n)!}\text{ }\mathbf{x}^{2n}\\ \end{array} =ϵ2(a22cos(𝐱))=\mathbf{\epsilon}^{2}(-\tfrac{a^{2}}{2}\cos(\mathbf{x}))

Thus from Table 3, we get:

cos(𝐱)=cos(x+aϵ)=cos(x)^,asin(x),a22!cos(x),a33!sin(x),a44!cos(x),,\displaystyle\cos(\mathbf{x})=\cos(x+a\mathbf{\epsilon})=\langle\widehat{\cos(x)},-a\sin(x),-\tfrac{a^{2}}{2!}\cos(x),\tfrac{a^{3}}{3!}\sin(x),\tfrac{a^{4}}{4!}\cos(x),\dots\rangle,

and we also get

cos(ϵ)=1^,0,12,0,=112!ϵ2+\displaystyle\cos(\mathbf{\epsilon})=\langle\widehat{1},0,-\tfrac{1}{2},0,\dots\rangle=1-\tfrac{\textbf{1}}{\textbf{2}!}\mathbf{\epsilon}^{\textbf{2}}+\dots

for an infinitesimal angle ϵ\mathbf{\epsilon}.

Example 5.11.

With the same 𝐱\mathbf{x} as in Examples 5.9 and 5.10, we try to know what exp(𝐱)\exp(\mathbf{x}) is. Based on Equation 6,

exp(𝐱)=exp(x+aϵ)=1+(x+aϵ)12!(x+aϵ)2+13!(x+aϵ)3+\displaystyle\exp(\mathbf{x})=\exp(x+a\mathbf{\epsilon})=\textbf{1}+(x+a\mathbf{\epsilon})-\tfrac{\textbf{1}}{\textbf{2}!}(x+a\mathbf{\epsilon})^{\textbf{2}}+\tfrac{\textbf{1}}{\textbf{3}!}(x+a\mathbf{\epsilon})^{\textbf{3}}+\dots

Our task now is to find all the members of Nstϵ(exp(𝐱))\texttt{Nst}_{\epsilon}(\exp(\mathbf{x})) and also St(exp(𝐱))\texttt{St}(\exp(\mathbf{x})), which are shown in Table 4. Note that from the way we define the function exp\exp, xiNstω(exp(𝐱))\forall x_{i}\in\texttt{Nst}_{\omega}(\exp(\mathbf{x})) xi=0x_{i}=0. Thus from Table 4, we get:

exp(𝐱)=exp(x+aϵ)=exp(x)^,aexp(x),a22!exp(x),a33!exp(x),a44!exp(x),,\displaystyle\exp(\mathbf{x})=\exp(x+a\mathbf{\epsilon})=\langle\widehat{\exp(x)},a\exp(x),\tfrac{a^{2}}{2!}\exp(x),\tfrac{a^{3}}{3!}\exp(x),\tfrac{a^{4}}{4!}\exp(x),\dots\rangle,

and we also get

exp(ϵ)=1^,1,12!,13!,14!,=1+ϵ+12!ϵ2+13!ϵ3\displaystyle\exp(\mathbf{\epsilon})=\langle\widehat{1},1,\tfrac{1}{2!},\tfrac{1}{3!},\tfrac{1}{4!},\dots\rangle=1+\mathbf{\epsilon}+\tfrac{\textbf{1}}{\textbf{2}!}\mathbf{\epsilon}^{\textbf{2}}+\tfrac{\textbf{1}}{\textbf{3}!}\mathbf{\epsilon}^{\textbf{3}}-\dots

for an infinitesimal angle ϵ\mathbf{\epsilon}.

Table 4: St(exp(𝐱))\texttt{St}(\exp(\mathbf{x)}) and The First Three Members of Nstϵ(exp(𝐱))\texttt{Nst}_{\epsilon}(\exp(\mathbf{x}))
Expanded Form Simplified Form
real-part =1+𝐱+12!𝐱2+13!𝐱3+=n=01n! 𝐱n\begin{array}[]{ll}=&1+\mathbf{x}+\tfrac{1}{2!}\mathbf{x}^{2}+\tfrac{1}{3!}\mathbf{x}^{3}+\dots\\ =&\sum_{n=0}\tfrac{1}{n!}\text{ }\mathbf{x}^{n}\end{array} =exp(𝐱)=\exp(\mathbf{x})
ϵ\mathbf{\epsilon}-part =aϵ+22!aϵ𝐱+33!aϵ𝐱2+44!aϵ𝐱3ϵ+=ϵ(a+a𝐱+12!a𝐱2+13!a𝐱314!a𝐱4+)=ϵn=01n! a𝐱n\begin{array}[]{ll}=&a\mathbf{\epsilon}+\tfrac{2}{2!}a\mathbf{\epsilon}\mathbf{x}+\tfrac{3}{3!}a\mathbf{\epsilon}\mathbf{x}^{2}+\tfrac{\textbf{4}}{\textbf{4}!}a\mathbf{\epsilon}\mathbf{x}^{3}\mathbf{\epsilon}+\dots\\ =&\mathbf{\epsilon}(a+a\mathbf{x}+\tfrac{1}{2!}a\mathbf{x}^{2}+\tfrac{1}{3!}a\mathbf{x}^{3}-\tfrac{1}{\textbf{4}!}a\mathbf{x}^{\textbf{4}}+\dots)\\ =&\mathbf{\epsilon}\sum_{n=0}\tfrac{1}{n!}\text{ }a\mathbf{x}^{n}\end{array} =ϵ(aexp(𝐱))=\mathbf{\epsilon}(a\exp(\mathbf{x}))
ϵ2\mathbf{\epsilon}^{2}-part =12!a2ϵ2𝐱+33!a2ϵ2𝐱+64!a2ϵ2𝐱2+105!a2ϵ2𝐱3+=ϵ2(12!a2+33!a2𝐱+64!a2𝐱2+105!a2𝐱3+)=ϵ2n=0(n+1)(n+2)2(n+2)! a2𝐱n=ϵ2n=012(n!) a2𝐱n\begin{array}[]{ll}=&\tfrac{1}{2!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}+\tfrac{3}{3!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}+\tfrac{6}{\textbf{4}!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{2}+\tfrac{\textbf{10}}{5!}a^{2}\mathbf{\epsilon}^{2}\mathbf{x}^{3}+\dots\\ =&\mathbf{\epsilon}^{2}(\tfrac{1}{2!}a^{2}+\tfrac{3}{3!}a^{2}\mathbf{x}+\tfrac{6}{\textbf{4}!}a^{2}\mathbf{x}^{2}+\tfrac{\textbf{10}}{5!}a^{2}\mathbf{x}^{3}+\dots)\\ =&\mathbf{\epsilon}^{2}\sum_{n=0}\tfrac{(n+1)(n+2)}{2(n+2)!}\text{ }a^{2}\mathbf{x}^{n}\\ =&\mathbf{\epsilon}^{2}\sum_{n=0}\tfrac{1}{2(n!)}\text{ }a^{2}\mathbf{x}^{n}\end{array} =ϵ2(a22exp(𝐱))=\mathbf{\epsilon}^{2}(\tfrac{a^{2}}{2}\exp(\mathbf{x}))
ϵ3\mathbf{\epsilon}^{3}-part =13!a3ϵ3+44!a3ϵ3𝐱+105!a3ϵ3𝐱2+206!a3ϵ3𝐱3+357!a3ϵ3𝐱4+=ϵ3(13!a3+44!a3𝐱105!a3𝐱2+206!a3𝐱3+357!a3𝐱4+)=ϵ3n=0(n+1)(n+2)(n+3)6(n+3)! a3𝐱n=ϵ3n=016(n!) a3𝐱n\begin{array}[]{ll}=&\tfrac{1}{3!}a^{3}\mathbf{\epsilon}^{3}+\tfrac{\textbf{4}}{\textbf{4}!}a^{3}\mathbf{\epsilon}^{3}\mathbf{x}+\tfrac{\textbf{10}}{5!}a^{3}\mathbf{\epsilon}^{3}\mathbf{x}^{2}+\tfrac{\textbf{20}}{6!}a^{3}\mathbf{\epsilon}^{3}\mathbf{x}^{3}+\tfrac{\textbf{35}}{7!}a^{3}\mathbf{\epsilon}^{3}\mathbf{x}^{\textbf{4}}+\dots\\ =&\mathbf{\epsilon}^{3}(\tfrac{1}{3!}a^{3}+\tfrac{\textbf{4}}{\textbf{4}!}a^{3}\mathbf{x}-\tfrac{\textbf{10}}{5!}a^{3}\mathbf{x}^{2}+\tfrac{\textbf{20}}{6!}a^{3}\mathbf{x}^{3}+\tfrac{\textbf{35}}{7!}a^{3}\mathbf{x}^{\textbf{4}}+\dots)\\ =&\mathbf{\epsilon}^{3}\sum_{n=0}\tfrac{(n+1)(n+2)(n+3)}{6(n+3)!}\text{ }a^{3}\mathbf{x}^{n}\\ =&\mathbf{\epsilon}^{3}\sum_{n=0}\tfrac{1}{6(n!)}\text{ }a^{3}\mathbf{x}^{n}\end{array} =ϵ3(a36exp(𝐱))=\mathbf{\epsilon}^{3}(\tfrac{a^{3}}{6}\exp(\mathbf{x}))

From the preceding discussion, we have the following proposition.

Proposition 5.12.

For the sin\sin, cos\cos, and exp\exp functions:

  1. 1.

    Dersin𝐱(𝐱,ϵ)=sin(𝐱+ϵ)sin(𝐱)ϵ=cos(x)^,12!sin(x),13!cos(x),\textnormal{{Der}}_{\sin{\mathbf{x}}}(\mathbf{x},\mathbf{\epsilon})=\tfrac{\sin(\mathbf{x}+\mathbf{\epsilon})-\sin(\mathbf{x})}{\mathbf{\epsilon}}=\langle\widehat{\cos(x)},-\tfrac{1}{2!}\sin(x),-\tfrac{1}{3!}\cos(x),\dots\rangle, and so we have:

    St(Dersin(𝐱)(𝐱,ϵ))=cos(𝐱)\textnormal{{St}}(\textnormal{{Der}}_{\sin(\mathbf{x})}(\mathbf{x},\mathbf{\epsilon}))=\cos(\mathbf{x})

  2. 2.

    Dercos𝐱(𝐱,ϵ)=cos(𝐱+ϵ)cos(𝐱)ϵ=sin(x)^,12!cos(x),13!sin(x),\textnormal{{Der}}_{\cos{\mathbf{x}}}(\mathbf{x},\mathbf{\epsilon})=\tfrac{\cos(\mathbf{x}+\mathbf{\epsilon})-\cos(\mathbf{x})}{\mathbf{\epsilon}}=\langle\widehat{-\sin(x)},-\tfrac{1}{2!}\cos(x),\tfrac{1}{3!}\sin(x),\dots\rangle, and so

    St(Dercos(𝐱)(𝐱,ϵ))=sin(𝐱)\textnormal{{St}}(\textnormal{{Der}}_{\cos(\mathbf{x})}(\mathbf{x},\mathbf{\epsilon}))=-\sin(\mathbf{x})

  3. 3.

    Derexp𝐱(𝐱,ϵ)=exp(𝐱+ϵ)exp(𝐱)ϵ=exp(x)^,12!exp(x),13!exp(x),\textnormal{{Der}}_{\exp{\mathbf{x}}}(\mathbf{x},\mathbf{\epsilon})=\tfrac{\exp(\mathbf{x}+\mathbf{\epsilon})-\exp(\mathbf{x})}{\mathbf{\epsilon}}=\langle\widehat{\exp(x)},\tfrac{1}{2!}\exp(x),\tfrac{1}{3!}\exp(x),\dots\rangle, and so

    St(Derexp(𝐱)(𝐱,ϵ))=exp(𝐱)\textnormal{{St}}(\textnormal{{Der}}_{\exp(\mathbf{x})}(\mathbf{x},\mathbf{\epsilon}))=\exp(\mathbf{x})

5.3 Continuity

In this subsection, we try to pinpoint what the good definition for continuous functions is. We also decide whether we can permeate it between \mathbb{R} and <\mathbb{R^{Z_{<}}}. Note that if the domain and codomain of a function is not explicitly stated, they will be determined from the specified model.

Definition 5.13 (EDCLASS\text{ED}_{\text{CLASS}}).

A function f:f:\mathbb{R}\rightarrow\mathbb{R} is continuous at a point aa\in\mathbb{R} if, given nn\in\mathbb{N}, there exists a mm\in\mathbb{N} such that

|f(x)f(a)|<1n|f(x)-f(a)|<\frac{1}{n} whenever |xa|<1m|x-a|<\frac{1}{m}.

The function ff is called continuous on an interval II iff ff is continuous at every point in II.

Definition 5.14 (ED).

A function ff is continuous at a point 𝐜<\mathbf{c}\in\mathbb{R^{Z_{<}}} if, given 𝔢𝟏<>0\mathbf{\mathfrak{e}_{1}}\in\mathbb{R^{Z_{<}}}>\textbf{0}, there exists a 𝔢2<>0\mathfrak{e}_{2}\in\mathbb{R^{Z_{<}}}>\textbf{0} such that

|𝐟(𝐱)𝐟(𝐜)|<𝔢𝟏|\mathbf{f(x)}-\mathbf{f(c)}|<\mathbf{\mathfrak{e}_{1}} whenever |𝐱𝐜|<𝔢𝟐|\mathbf{x}-\mathbf{c}|<\mathbf{\mathfrak{e}_{2}}.

That function 𝐟\mathbf{f} is called continuous function over an interval II iff 𝐟\mathbf{f} is continuous at every point in II.

Proposition 5.15.

There exists a function ff in <\mathbb{R^{Z_{<}}} which is continuous under Definition 5.14, but discontinuous under Definition 5.13, i.e.

<f\mathbb{R^{Z_{<}}}\models\exists f s.t. (EDCLASS(f)¬ED(f))(\textnormal{ED}_{\textnormal{CLASS}}(f)\land\neg\textnormal{ED}(f)).

Proof.

Suppose that Δ={xn,|x|<1n}\Delta=\{x\mid\forall n\in\mathbb{N},\left\lvert x\right\rvert<\frac{1}{n}\} – in other words, Δ\Delta is a set of all infinitesimals – and consider the indicator function around Δ\Delta, that is

𝟙Δ(x)={1,xΔ0,otherwise.\displaystyle\mathds{1}_{\Delta}(x)=\begin{cases}1,&x\in\Delta\\ 0,&\textnormal{otherwise.}\end{cases}

Then EDCLASS(𝟙Δ)\textnormal{ED}_{\textnormal{CLASS}}(\mathds{1}_{\Delta}) but ¬ED(𝟙Δ)\neg\textnormal{ED}(\mathds{1}_{\Delta}). ∎

Remark 5.16.

Note that:

  1. 1.

    The set Δ\Delta in \mathbb{R} only has 0 as its member. That is Δ={0}\mathbb{R}\models\Delta=\{0\}.

  2. 2.

    In \mathbb{R}, both Definitions 5.13 and 5.14 are equivalent, that is for any function ff, EDCLASS(f)ED(f)\mathbb{R}\models\textnormal{ED}_{\textnormal{CLASS}}(f)\leftrightarrow\textnormal{ED}(f).

Property 5.17 (EVP).

If II is an interval and f:IJf:I\rightarrow J, we say that ff has the extreme value property iff ff has its maximum value on II. That is,

abI\forall a\leq b\in I, x[a,b]\exists x\in[a,b] s.t. y[a,b](f(y)f(x))\forall y\in[a,b](f(y)\leq f(x)).

Property 5.18 (IVP).

If II is an interval, and f:I=[a,b]Jf:I=[a,b]\rightarrow J, we say that ff has the intermediate value property iff

c(f(a),f(b))\forall c^{\prime}\in(f(a),f(b)), c(a,b)\exists c\in(a,b) s.t. f(c)=cf(c)=c^{\prime}.

Theorem 5.19.

EDEVP\mathbb{R}\models\textnormal{ED}\rightarrow\textnormal{EVP}

Proof.

The proof of this theorem can be found in any standard book for Analysis course (in [3] for example). ∎

Theorem 5.20.

There is a function ff such that

<ED(f)¬EVP(f)\mathbb{R^{Z_{<}}}\models\textnormal{ED}(f)\land\lnot\textnormal{EVP}(f).

Proof.

Take the function ff on [1,2][1,2] as defined below:

f(x)={1nxmn (reduced fraction)0otherwise.\displaystyle f(x)=\begin{cases}\frac{1}{n}&x\sim\frac{m}{n}\textnormal{ (reduced fraction)}\\ 0&\text{otherwise}.\end{cases}

Remark 5.21.

This research now reaches an especially engrossing object. The function f(x)f(x) in Theorem 5.20 can be used to construct a fractal-like object. Fractals are classically defined as geometric objects that exhibit some form of self-similarity. Figure 1 shows what the function f(x)f(x) in Theorem 5.20 looks like, and also what occurs when we zoom in on a particular point. In this sense, the function from Theorem 5.20 is an infinitesimal fractal. Formally speaking, suppose that we have a function f:<f:\mathbb{R^{Z_{<}}}\rightarrow\mathbb{R} and let us define another function

F:<<\displaystyle F:\mathbb{R^{Z_{<}}}\rightarrow\mathbb{R^{Z_{<}}}

by

F(x)=f(x)+ϵf(x)+ϵ2f(x)+=f(x)^,f(x),f(x),.\displaystyle\begin{aligned} F(x)&=f(x)+\epsilon f(x)+\epsilon^{2}f(x)+\dots\\ &=\langle\widehat{f(x)},f(x),f(x),\dots\rangle.\end{aligned}

Then, that function F(x)F(x) will define an infinite fractal (if niω(1ϵ×^F(x))=F(x)\texttt{ni}_{\omega}(\frac{1}{\epsilon}\widehat{\times}F(x))=F(x)) or infinitesimal fractals (if niω(ϵ×^F(x))=F(x)\texttt{ni}_{\omega}(\epsilon\widehat{\times}F(x))=F(x)), where niω(𝐱)\texttt{ni}_{\omega}(\mathbf{x}) denotes the non-infinity part of 𝐱\mathbf{x}.

Refer to caption
Figure 1: Illustration of the infinitesimal fractal from the function defined in Theorem 5.20

Figure 2 shows the relationship among the three definitions of continuity in both \mathbb{R} and <\mathbb{R^{Z_{<}}}. The proof of each of the relations there can be seen in [32, p. 94–98].

EVPIVPED
(a) In \mathbb{R}
EVPIVPED
(b) In <\mathbb{R^{Z_{<}}}
Figure 2: Relationship among the three definitions of continuity

The obvious question worth asking is how do we define continuity in our set <\mathbb{R^{Z_{<}}}. As seen before, there are three possible ways to define it, namely: with the ϵ\epsilon-δ\delta definition (ED), with extreme value property (EVP), or with the intermediate value property (IVP). We will now discuss them one by one. Firstly, through IVP. The IVP basically says that for every value within the range of the given function, we can find a point in the domain corresponding to that value. Will this work in our set <\mathbb{R^{Z_{<}}}? Let us consider the <\mathbb{R^{Z_{<}}}-valued function f(x)=x2f(x)=x^{2} on [a,b][a,b] for any a,b<a,b\in\mathbb{R^{Z_{<}}} and let us assume that IVP holds. It follows that for every cc^{\prime} between f(a)=a2f(a)=a^{2} and f(b)=b2f(b)=b^{2}, c(a,b)\exists c\in(a,b) such that f(c)=c2=cf(c)=c^{2}=c^{\prime}. The only cc which satisfies that last equation is c=cc=\sqrt{c^{\prime}}, which cannot be defined in our set <\mathbb{R^{Z_{<}}}. Thus, IVP, even though it is somehow intuitively “obvious”, it does not work in <\mathbb{R^{Z_{<}}}. This phenomenon is actually not uncommon if we want to have a world with infinitesimals (or infinities) in it. See [5, p. 107] for example.

However, note that in \mathbb{R}, the function x2x^{2} still satisfies IVP. Now, is there a function in <\mathbb{R^{Z_{<}}} that satisfies IVP? Consider the identity function f(x)=xf(x)=x. This function clearly satisfies IVP in both domains, and so we have the following theorem.

Theorem 5.22.

There exists a function ff such that (IVP(f)<¬IVP(f))(\mathbb{R}\models\textnormal{IVP}(f)\land\mathbb{R^{Z_{<}}}\models\lnot\textnormal{IVP}(f)), and there exists a function gg s.t. ,<IVP(g)\mathbb{R},\mathbb{R^{Z_{<}}}\models\textnormal{IVP}(g).

Hence, from the argument above we also argued that defining continuity in our set with IVP is not really useful.

Secondly, in regards to EVP. This is clearly not a good way to define continuity in our set because even in the set of real numbers, there are some continuous functions which do not satisfy EVP themselves. So the last available option now is the third one, which is the ϵ\epsilon-δ\delta (ED) definition. We argued that this definition is the best way to define continuity in <\mathbb{R^{Z_{<}}}. Moreover, in this way, it preserves much of the spirit of classical analysis on \mathbb{R} while retaining the intuition of infinitesimals.

It is important to note that in the ED definition of continuity (Definition 5.14), there are two variables which are in play, i.e. 𝔢1\mathfrak{e}_{1} and 𝔢2\mathfrak{e}_{2}. When we applied this definition on our set, these two variables hold important (or rather, very interesting) roles where we will have different levels of continuity from the same function. What we mean is that these two variables can greatly vary depending on how far (‘deep’) we want to push (observe) them, e.g. 𝔢2\mathfrak{e}_{2} can be a real number (𝔢2Δ0\mathfrak{e}_{2}\in\Delta^{0}), or it can be in Δ4\Delta^{4}, Δ8\Delta^{8} and so on. Remember that these two numbers, 𝔢1\mathfrak{e}_{1} and 𝔢2\mathfrak{e}_{2}, will determine how subtle we want our intervals to be (see Figure 3 for illustration).

Refer to caption
Refer to caption
(a) An 𝔢1\mathfrak{e}_{1} bound & its 𝔢2\mathfrak{e}_{2} neighbourhood fulfilling Definition 5.14
Refer to caption
(b) A smaller bound & its neighbourhood
Figure 3: Illustration of 𝔢1\mathfrak{e}_{1} and 𝔢2\mathfrak{e}_{2} intervals

Thus this definition of continuity works as follows. Suppose that we have a function ff and we want to decide whether it is continuous or not. With this concept of two variables, we will have what we call as (k,n)(k,n)-continuity where k,n{0}k,n\in\mathbb{N}\cup\{0\}.

Definition 5.23 ((k,n)(k,n)-Continuity).

A function ff is (k,n)(k,n)-continuous at a point 𝐜\mathbf{c} iff 𝔢𝟏𝐤>0\forall\mathbf{\mathfrak{e}_{1_{k}}}>0, 𝔢2n>0\exists\mathfrak{e}_{2_{n}}>0 such that

if |𝐱^𝐜|<𝔢𝟐𝐧|\mathbf{x}\widehat{-}\mathbf{c}|<\mathbf{\mathfrak{e}_{2_{n}}}, then |𝐟(𝐱)^𝐟(𝐜)|<𝔢𝟏𝐤|\mathbf{f(x)}\widehat{-}\mathbf{f(c)}|<\mathbf{\mathfrak{e}_{1_{k}}}

where 𝔢1p,𝔢2pΔp\mathfrak{e}_{1_{p}},\mathfrak{e}_{2_{p}}\in\Delta^{p}.

Definition 5.24.

A function ff is said to be (k,n)(k,n)-continuous iff it is (k,n)(k,n)-continuous at every point in the given domain.

Remark 5.25.

From the definition of the set Δm\Delta^{m}, note that for any r<,r\in\mathbb{R^{Z_{<}}}, dΔp,d\in\Delta^{p}, and eΔp+1,e\in\Delta^{p+1},

(re,r+e)(rd,r+d).\displaystyle(r-e,r+e)\subseteq(r-d,r+d).

To be able to grasp a better understanding of Definition 5.23, see the examples below.

Example 5.26.

Consider the <\mathbb{R^{Z_{<}}}-valued function f(x)f(x) defined as follows:

f(𝐱)={𝐱St(𝐱)1𝐱+^𝟏otherwise.f(\mathbf{x})=\begin{cases}\mathbf{x}&\texttt{St}(\mathbf{x})\leq 1\\ \mathbf{x}\widehat{+}\mathbf{1}&\textnormal{otherwise.}\end{cases}

First we need to understand clearly how this function actually works. Figure 4, where ii denotes an arbitrary infinitesimal number, illustrates to us what the function f(𝐱)f(\mathbf{x}) looks like. Notice that at 𝐱=𝟏\mathbf{x}=\mathbf{1}, what looks like a point in real numbers is actually a (constant) line when we zoom in deep enough into <\mathbb{R^{Z_{<}}}101010We have to be really careful here because if the first condition there was 𝐱1\mathbf{x}\leq 1 (instead of St(𝐱)1\texttt{St}(\mathbf{x})\leq 1), then there would be no line there — it will be exactly one point.. So how about the continuity of this function? It is obvious that f(𝐱)f(\mathbf{x}) is not 0,00,0-continuous (by taking, for example, 𝔢10=12\mathfrak{e}_{1_{0}}=\frac{1}{2} and 𝐱=1.5\mathbf{x}=\mathbf{1.5}). However, interestingly enough, it is (0,1)\mathit{(0,1)}-continuous by taking 𝔢21Δ1\mathfrak{e}_{2_{1}}\in\Delta^{1}. Why was that? The fact that 𝔢21Δ1\mathfrak{e}_{2_{1}}\in\Delta^{1} and that it has to depend on 𝔢10\mathfrak{e}_{1_{0}} means that 𝔢10\mathfrak{e}_{1_{0}} has to be in Δ1\Delta^{1} as well. Now, assigning 𝔢21=𝔢10\mathfrak{e}_{2_{1}}=\mathfrak{e}_{1_{0}} is sufficient to prove its 0,1\mathit{0,1}-continuity.

Refer to caption
Figure 4: Illustration of Function f(x)f(x) in Example 5.26
Example 5.27.

The identity function f(𝐱)=𝐱f(\mathbf{x})=\mathbf{x} for all 𝐱<\mathbf{x}\in\mathbb{R^{Z_{<}}} is (0,0)\mathit{(0,0)}-continuous, just like in reals. However, it is not (1,0)\mathit{(1,0)}-continuous because for any point 𝐜\mathbf{c}, there is an 𝔢11=ϵ\mathfrak{e}_{1_{1}}=\mathbf{\epsilon} such that for every 𝔢20=r\mathfrak{e}_{2_{0}}=r where rr\in\mathbb{R}, |𝐱^𝐜|<r|\mathbf{x}\widehat{-}\mathbf{c}|<r but f(𝐱)f(c)ϵf(\mathbf{x})-f(c)\geq\mathbf{\epsilon}. In fact, identity function is (k,n)(k,n)-continuous only when knk\leq n, but not otherwise.

The next theorem below is very interesting in as much as it enables us to classify whether a function is a constant function or not by using (k,n)(k,n)-continuity.

Theorem 5.28.

For any function ff, if ff is (k,n)(k,n)-continuous for any k,n{0}k,n\in\mathbb{N}\cup\{0\}, then ff is a constant function.

Proof.

Here we want to prove its contrapositive, in other words, if ff is not constant, then there exist (k,n)(k,n) such that ff is not (k,n)(k,n)-continuous. Because ff is not constant, there will be a,ba,b in the domain such that f(a)f(b)f(a)\neq f(b) and suppose that |f(a)f(b)|Δm|f(a)-f(b)|\in\Delta^{m} such that |ab|Δl|a-b|\in\Delta^{l}. By this construction, ff will not be (m,l1)(m,l-1)-continuous and so we can take k=mk=m and n=l1n=l-1. ∎

The theorem below is a generalisation of Theorem 5.28.

Theorem 5.29.

If there exists mm for all kk such that a function ff is (k,m)(k,m)-continuous, then ff will be constant in Δm\Delta^{m}-neighbourhood.

The next interesting question is: what is the relation between, for example, (0,1)\mathit{(0,1)}-continuity and (0,2)\mathit{(0,2)}-continuity? In general, what is the relation between (k,n)(k,n)-continuity and (k,(n+1))(k,(n+1))-continuity? And also between (k,n)(k,n)-continuity and ((k+1),n)((k+1),n)-continuity? See these two theorems below.

Theorem 5.30.

For any function ff, if ff is (k,n)(k,n)-continuous, then ff is also (k,(n+1))(k,(n+1))-continuous.

Proof.

Suppose that a function ff is (k,n)(k,n)-continuous at point 𝐜\mathbf{c}. This would mean that 𝔢1k\forall\mathfrak{e}_{1_{k}}, 𝔢2n\exists\mathfrak{e}_{2_{n}} such that if |𝐱^𝐜|<𝔢2n|\mathbf{x}\widehat{-}\mathbf{c}|<\mathfrak{e}_{2_{n}}, then |f(𝐱)^f(𝐜)|<𝔢1k|f(\mathbf{x})\widehat{-}f(\mathbf{c})|<\mathfrak{e}_{1_{k}}. By using the same 𝔢1k\mathfrak{e}_{1_{k}} and from Remark 5.25, we can surely find 𝔢2(n+1)=𝔢2n×^ϵ\mathfrak{e}_{2_{(n+1)}}=\mathfrak{e}_{2_{n}}\widehat{\times}\mathbf{\epsilon} such that for all x(𝐜^𝔢2(n+1),𝐜+^𝔢2(n+1))x\in(\mathbf{c}\widehat{-}\mathfrak{e}_{2_{(n+1)}},\mathbf{c}\widehat{+}\mathfrak{e}_{2_{(n+1)}}), f(x)(f(𝐜)^𝔢1k,f(𝐜)+^𝔢1k)f(x)\in(f(\mathbf{c})\widehat{-}\mathfrak{e}_{1_{k}},f(\mathbf{c})\widehat{+}\mathfrak{e}_{1_{k}}). ∎

Example 5.31.

By Theorem 5.30, the function f(x)f(x) in Example 5.26 is also (0,2)\mathit{(0,2)}-continuous, and also (0,3)\mathit{(0,3)}-continuous, and so on.

Theorem 5.32.

For any function ff, if ff is ((k+1),n)((k+1),n)-continuous, then ff is also (k,n)(k,n)-continuous.

Proof.

Suppose that a function ff is ((k+1),n)((k+1),n)-continuous at point 𝐜\mathbf{c} and the set Δm\Delta^{m} defined as in Theorem 5.30. The fact that ff is ((k+1),n)((k+1),n)-continuous means that 𝔢1(k+1)\forall\mathfrak{e}_{1_{(k+1)}}, 𝔢2n\exists\mathfrak{e}_{2_{n}} such that if |𝐱^𝐜|<𝔢2n|\mathbf{x}\widehat{-}\mathbf{c}|<\mathfrak{e}_{2_{n}}, then |f(𝐱)^f(𝐜)|<𝔢1(k+1)|f(\mathbf{x})\widehat{-}f(\mathbf{c})|<\mathfrak{e}_{1_{(k+1)}} is hold. Here we want to prove that 𝔢1k\forall\mathfrak{e}_{1_{k}}, 𝔢2n\exists\mathfrak{e}_{2_{n}} such that if |𝐱^𝐜|<𝔢2n|\mathbf{x}\widehat{-}\mathbf{c}|<\mathfrak{e}_{2_{n}}, then |f(𝐱)^f(𝐜)|<𝔢1k|f(\mathbf{x})\widehat{-}f(\mathbf{c})|<\mathfrak{e}_{1_{k}}. This actually follows directly from Remark 5.25 as (f(𝐜)^𝔢1(k+1),f(𝐜)+^𝔢1(k+1))(f(𝐜)^𝔢1k,f(𝐜)+^𝔢1k)(f(\mathbf{c})\widehat{-}\mathfrak{e}_{1_{(k+1)}},f(\mathbf{c})\widehat{+}\mathfrak{e}_{1_{(k+1)}})\subseteq(f(\mathbf{c})\widehat{-}\mathfrak{e}_{1_{k}},f(\mathbf{c})\widehat{+}\mathfrak{e}_{1_{k}}). ∎

Now suppose that ff and gg are two (k,n)(k,n)-continuous functions in <\mathbb{R^{Z_{<}}}. We will examine how the arithmetic of those two continuous functions works. It is clear that (k,n)(k,n)-continuity is closed under addition and subtraction, i.e. f+gf+g and fgf-g are both (k,n)(k,n)-continuous. The composition and multiplication of two continuous functions are particularly interesting as can be seen in Theorem 5.33 and Theorem 5.34, respectively.

Theorem 5.33.

If ff is a (k,n)(k,n)-continuous function and gg is an (n,q)(n,q)-continuous function, then fgf\circ g will be (k,q)(k,q)-continuous.

Proof.

Since ff is (k,n)(k,n)-continuous at g(c)g(c), our definition of continuity tells us that for all 𝔢1k>0\mathfrak{e}_{1_{k}}>0, there exists 𝔢2n\mathfrak{e}_{2_{n}} such that

if |g(x)g(a)|<𝔢2n\left\lvert g(x)-g(a)\right\rvert<\mathfrak{e}_{2_{n}}, then |f(g(x))f(g(a))|<𝔢1k\left\lvert f(g(x))-f(g(a))\right\rvert<\mathfrak{e}_{1_{k}}.

Also since gg is (n,q)(n,q)-continuous at cc, there exists 𝔢2q\mathfrak{e}_{2_{q}} such that

if |xa|<𝔢2q\left\lvert x-a\right\rvert<\mathfrak{e}_{2_{q}}, then |g(x)g(a)|<𝔢2n\left\lvert g(x)-g(a)\right\rvert<\mathfrak{e}_{2_{n}}.

I have taken 𝔢1n=𝔢2n\mathfrak{e}_{1_{n}}=\mathfrak{e}_{2_{n}} here. Now this tells us that for all 𝔢1k>0\mathfrak{e}_{1_{k}}>0, there exists 𝔢2q\mathfrak{e}_{2_{q}} (and an 𝔢2n\mathfrak{e}_{2_{n}}) such that

if |xa|<𝔢2q\left\lvert x-a\right\rvert<\mathfrak{e}_{2_{q}}, then |g(x)g(a)|<𝔢2n\left\lvert g(x)-g(a)\right\rvert<\mathfrak{e}_{2_{n}} which implies that |f(g(x))f(g(c))|<𝔢1k\left\lvert f(g(x))-f(g(c))\right\rvert<\mathfrak{e}_{1_{k}},

which is what we wanted to show. ∎

Theorem 5.34.

Suppose that f,gf,g are finite-valued functions. If ff is a (k,n)(k,n)-continuous function and gg is an (l,o)(l,o)-continuous function, then the function H=fgH=f\cdot g will be (max{k,l},min{n,o})(\max\{k,l\},\min\{n,o\})-continuous.

Proof.

Let f,gf,g be given such that ff is (k,n)(k,n)-continuous and gg is (l,o)(l,o)-continuous. Now let HH be defined by H(x)=f(x)g(x)H(x)=f(x)g(x) and so we want to show that HH is (max{k,l},min{n,o})(\max\{k,l\},\min\{n,o\})-continuous, that is, for all c<c\in\mathbb{R^{Z_{<}}}, for every 𝔢1max{k,l}>0\mathfrak{e}_{1_{\max\{k,l\}}}>0, there exists 𝔢2min{n,o}>0\mathfrak{e}_{2_{\min\{n,o\}}}>0 such that for all x<x\in\mathbb{R^{Z_{<}}} with |xc|<𝔢2min{n,o}\left\lvert x-c\right\rvert<\mathfrak{e}_{2_{\min\{n,o\}}}, |H(x)H(c)|<𝔢1max{k,l}\left\lvert H(x)-H(c)\right\rvert<\mathfrak{e}_{1_{\max\{k,l\}}} holds.
Now let cc and 𝔢1max{k,l}\mathfrak{e}_{1_{\max\{k,l\}}} be given and we choose 𝔢2\mathfrak{e}_{2} such that 𝔢2Δmin{n,o}\mathfrak{e}_{2}\in\Delta^{\min\{n,o\}}, i.e. 𝔢2=𝔢2min{n,o}\mathfrak{e}_{2}=\mathfrak{e}_{2_{\min\{n,o\}}}. Then for all x<x\in\mathbb{R^{Z_{<}}} with |xc|<𝔢2min{n,o}\left\lvert x-c\right\rvert<\mathfrak{e}_{2_{\min\{n,o\}}},

|H(x)H(c)|\displaystyle\left\lvert H(x)-H(c)\right\rvert =|f(x)g(x)f(c)g(c)|\displaystyle=\left\lvert f(x)g(x)-f(c)g(c)\right\rvert
=|f(x)g(x)f(x)g(a)+f(x)g(a)f(c)g(c)|\displaystyle=\left\lvert f(x)g(x)-f(x)g(a)+f(x)g(a)-f(c)g(c)\right\rvert
|f(x)g(x)f(c)g(c)|+|f(x)g(a)f(c)g(c)|\displaystyle\leq\left\lvert f(x)g(x)-f(c)g(c)\right\rvert+\left\lvert f(x)g(a)-f(c)g(c)\right\rvert
=|f(x)(g(x)g(a))|+|g(a)(f(x)f(c))|\displaystyle=\left\lvert f(x)(g(x)-g(a))\right\rvert+\left\lvert g(a)(f(x)f(c))\right\rvert
<|f(x)|𝔢1l+|g(a)|𝔢1k\displaystyle<\left\lvert f(x)\right\rvert\mathfrak{e}_{1_{l}}+\left\lvert g(a)\right\rvert\mathfrak{e}_{1_{k}} (7)

Note that because ff and gg are limited-valued function, then |f(x)|𝔢1l\left\lvert f(x)\right\rvert\mathfrak{e}_{1_{l}} and |g(a)|𝔢1k\left\lvert g(a)\right\rvert\mathfrak{e}_{1_{k}} are still in Δl\Delta^{l} and Δk\Delta^{k}, respectively. This means that the right side of Inequality 7 will be in Δmax{k,l}\Delta^{\max\{k,l\}} and so H(x)H(c)<𝔢1max{k,l}H(x)-H(c)<\mathfrak{e}_{1_{\max\{k,l\}}} is hold. ∎

It is worth pointing out here that the definition of (k,n)(k,n)-continuity is a much more fine-grained notion than the classical continuity. This is self-explanatory by the use of those two variables kk and nn which makes us able to take much more infinitesimals — in other words, we will be able to examine a far greater depth — than in the classical definition. Furthermore, there might be some possible connections to one of the quantum phenomenons in physics: ‘action at a distance’. This concept is typically characterized in terms of some cause producing a spatially separated effect in the absence of any medium by which the causal interaction is transmitted [16] and closely connected to the question of what the deepest level of physical reality is [30, pg. 168]. Note that research on this phenomenon is still being conducted up until now, as can be seen for example [33], [46] and [53].

5.4 Topological Continuity

Definition 5.35.

A function ff from a topological space (X,τ1)(X,\uptau_{1}) to a topological space (Y,τ2)(Y,\uptau_{2}) is a function f:XYf:X\rightarrow Y.

From now on, we will abbreviate this function notation by f:XYf:X\rightarrow\ Y or simply ff every time the topologies in XX and YY need not be explicitly mentioned. Also, f1f^{-1} denotes the inverse image of ff as usual.

Definition 5.36 (St-continuous).

A function f:XYf:X\rightarrow Y between topological spaces is standard topologically continuous, denoted by St-continuous, if

f1(U)Xf^{-1}(U)\subseteq X is St-open whenever UYU\subseteq Y is St-open.

Definition 5.37 (𝔢\mathfrak{e}-continuous).

A function f:XYf:X\rightarrow\ Y between topological spaces is infinitesimally topologically continuous, denoted by 𝔢\mathfrak{e}-continuous, if

f1(U)Xf^{-1}(U)\subseteq X is 𝔢\mathfrak{e}-open whenever UYU\subseteq Y is 𝔢\mathfrak{e}-open.

Theorem 5.38.

Suppose that X,Y<X,Y\subseteq\mathbb{R^{Z_{<}}}. Under the metric d\operatorname{\texttt{d}}, a function f:XYf:X\rightarrow Y is St-continuous if and only if ff satisfies EDCLASS\textnormal{ED}_{\textnormal{CLASS}} definition (Definition 5.13).

Proof.

We need to prove the implication both ways.

  1. 1.

    We want to prove that if ff is St-continuous, then ff satisfies EDCLASS{}_{\textnormal{CLASS}}. Suppose that ff is St-continuous and let 𝐱𝐨X\mathbf{x_{o}}\in X and n>0n\in\mathbb{N}>0. Then, the ball

    Bf(𝐱𝟎)(1/n)={𝐲Yd(𝐲,f(𝐱𝟎))<1/n}\displaystyle B_{f(\mathbf{x_{0}})}(\nicefrac{{1}}{{n}})=\{\mathbf{y}\in Y\mid\operatorname{\texttt{d}}(\mathbf{y},f(\mathbf{x_{0}}))<\nicefrac{{1}}{{n}}\}

    is open in YY, and hence f1(Bf(𝐱𝟎))f^{-1}(B_{f(\mathbf{x_{0}})}) is open in XX. Since 𝐱𝟎f1(Bf(𝐱𝟎))\mathbf{x_{0}}\in f^{-1}(B_{f(\mathbf{x_{0}})}), there exists some balls of radius 1/m\nicefrac{{1}}{{m}} for some mm\in\mathbb{N} such that

    B𝐱𝟎(1/m)f1(Bf(𝐱𝟎)).\displaystyle B_{\mathbf{x_{0}}}(\nicefrac{{1}}{{m}})\subseteq f^{-1}(B_{f(\mathbf{x_{0}})}).

    This is exactly what the EDCLASS{}_{\textnormal{CLASS}} says.

  2. 2.

    We want to prove that if ff satisfies EDCLASS{}_{\textnormal{CLASS}}, then ff is St-continuous. Suppose that ff satisfies EDCLASS{}_{\textnormal{CLASS}} and let UYU\subseteq Y is open. By Definition 4.6, for all 𝐲U\mathbf{y}\in U there exists some dy=1/nyd_{y}=\nicefrac{{1}}{{n_{y}}} where nyn_{y}\in\mathbb{N} such that

    B𝐲(dy)U\displaystyle B_{\mathbf{y}}(d_{y})\subseteq U

    and in fact,

    U=yUBdy(𝐲).\displaystyle U=\bigcup_{y\in U}B_{d_{y}}(\mathbf{y}). (8)

    Now we claim that f1(U)f^{-1}(U) is open in XX and suppose that 𝐱𝟎f1(U)\mathbf{x_{0}}\in f^{-1}(U). Then f(𝐱𝟎)Uf(\mathbf{x_{0}})\in U and so from Equation 8, f(𝐱𝟎)Bdy0(𝐲𝟎)f(\mathbf{x_{0}})\in B_{d_{y_{0}}}(\mathbf{y_{0}}) for some 𝐲𝟎U\mathbf{y_{0}}\in U and dy0=1/ny0d_{y_{0}}=\nicefrac{{1}}{{n_{y_{0}}}} for some ny0n_{y_{0}}\in\mathbb{N}, i.e. d(f(𝐱𝟎),𝐲𝟎)<dy0\operatorname{\texttt{d}}(f(\mathbf{x_{0}}),\mathbf{y_{0}})<d_{y_{0}}. Now define

    e=dy0d(f(𝐱𝟎),𝐲𝟎)>0.\displaystyle e=d_{y_{0}}-\operatorname{\texttt{d}}(f(\mathbf{x_{0}}),\mathbf{y_{0}})>0. (9)

    By Definition 5.13, there exists some mm\in\mathbb{N} such that

    if 𝐱X and d(𝐱,𝐱𝟎)<1/m, then d(f(𝐱),f(𝐱𝟎))<e.\displaystyle\textnormal{if }\mathbf{x}\in X\textnormal{ and }\operatorname{\texttt{d}}(\mathbf{x},\mathbf{x_{0}})<\nicefrac{{1}}{{m}},\textnormal{ then }\operatorname{\texttt{d}}(f(\mathbf{x}),f(\mathbf{x_{0}}))<e. (10)

    Now we claim that

    B𝐱𝟎(1/m)f1(U),\displaystyle B_{\mathbf{x_{0}}}(\nicefrac{{1}}{{m}})\subseteq f^{-1}(U), (11)

    which will actually show that f1(U)f^{-1}(U) is indeed open. To this end, let xB𝐱𝟎(1/m)x\in B_{\mathbf{x_{0}}}(\nicefrac{{1}}{{m}}), i.e. d(𝐱,𝐱𝟎)<1/m\operatorname{\texttt{d}}(\mathbf{x},\mathbf{x_{0}})<\nicefrac{{1}}{{m}}. Then from (10), we have d(f(𝐱),f(𝐱𝟎))<e\operatorname{\texttt{d}}(f(\mathbf{x}),f(\mathbf{x_{0}}))<e. Then, the triangle inequality and (9) imply that

    d(f(𝐱),𝐲𝟎)d(f(𝐱),f(𝐱𝟎))+d(f(𝐱𝟎),𝐲𝟎)<e+dψ(f(𝐱𝟎),𝐲𝟎)=dy0.\displaystyle\operatorname{\texttt{d}}(f(\mathbf{x}),\mathbf{y_{0}})\leq\operatorname{\texttt{d}}(f(\mathbf{x}),f(\mathbf{x_{0}}))+\operatorname{\texttt{d}}(f(\mathbf{x_{0}}),\mathbf{y_{0}})<e+\operatorname{\texttt{d}_{\texttt{$\psi$}}}(f(\mathbf{x_{0}}),\mathbf{y_{0}})=d_{y_{0}}.

    This means that f(x)B𝐲𝟎(dy0)Uf(x)\in B_{\mathbf{y_{0}}}(d_{y_{0}})\subseteq U, so that xf1(U)x\in f^{-1}(U). Therefore, (11) holds, as claimed.

And so from those two points above, we have proved what we want. ∎

Theorem 5.39.

Suppose that X,Y<X,Y\subseteq\mathbb{R^{Z_{<}}}. Under the metric d\operatorname{\texttt{d}}, a function f:XYf:X\rightarrow Y is 𝔢\mathfrak{e}-continuous if and only if ff satisfies ED definition.

Proof.

The proof of this theorem is similar with the one in Theorem 5.38 with some slight modifications in the distances (from 1/n\nicefrac{{1}}{{n}} for some nn\in\mathbb{N} into 𝔢<\mathbf{\mathfrak{e}}\in\mathbb{R^{Z_{<}}}). ∎

5.5 Convergence

When we are talking about sequences, it is necessary to talk also about what it means when we say that a sequence is convergent to a particular number. This subsection presents not only some possible definitions that can be used to define convergence in <\mathbb{R^{Z_{<}}}, but also the problems which occur when we apply them in <\mathbb{R^{Z_{<}}}.

Definition 5.40 (Classical Convergence).

A sequence sns_{n} converges to ss iff,

m\forall m\in\mathbb{N}, N\exists N such that n>N\forall n>N, |sns|<1m\left\lvert s_{n}-s\right\rvert<\frac{1}{m}.

We write CC(sn,s)\textnormal{CC}(s_{n},s) to denote that a sequence sns_{n} is classically convergent to ss.

Definition 5.40 above is the standard definition of how we define the notion of convergent classically.

Definition 5.41 (Hyperconvergence).

A sequence sns_{n} converges to ss iff,

r>0\forall r>0, N\exists N such that n>N\forall n>N, |sns|<r\left\lvert s_{n}-s\right\rvert<r.

We write HC(sn,s)\textnormal{HC}(s_{n},s) to denote that a sequence sns_{n} is hyperconvergent to ss. The interpretation of rr can be either in \mathbb{R} or in <\mathbb{R^{Z_{<}}}.

Example 5.42.

Suppose that we have a sequence sn=ϵns_{n}=\epsilon^{n} as follows:

S1=ϵ=0^,1,0,S_{1}=\epsilon=\langle\widehat{0},1,0,\dots\rangle

S2=ϵ2=0^,0,1,0,S_{2}=\epsilon^{2}=\langle\widehat{0},0,1,0,\dots\rangle

S3=ϵ3=0^,0,0,1,0,S_{3}=\epsilon^{3}=\langle\widehat{0},0,0,1,0,\dots\rangle

This sequence sns_{n} will hyperconverge to 0^,0,0,\langle\widehat{0},0,0,\dots\rangle, i.e. sns_{n} satisfies HC(sn,0)(s_{n},\textbf{0}).

Theorem 5.43.

For any sequence sns_{n}, CC(sn,s)HC(sn,s)\mathbb{R}\models\textnormal{CC}(s_{n},s)\leftrightarrow\textnormal{HC}(s_{n},s).

Proof.

The proof from HC to CC is obvious. Now suppose that a sequence sns_{n} satisfies CC(sn)(s_{n}) and w.l.o.g. we assume that the rr in HC definition is between 0 and 1. From the Archimedean property of reals we know that for every 0<r<10<r<1, we can find an mm\in\mathbb{N} such that 1m<r\frac{1}{m}<r, and so because of CC(sn)(s_{n}), we have |sns|<1m<r\left\lvert s_{n}-s\right\rvert<\frac{1}{m}<r. ∎

Theorem 5.44.

For any sequence sns_{n} in <\mathbb{R^{Z_{<}}}, HC(sn,ss_{n},s) always implies CC(sn,ss_{n},s). However, there exists a sequence (tn)(t_{n}) such that

<CC(tn,s)↛HC(tn,s)\mathbb{R^{Z_{<}}}\models\textnormal{CC}(t_{n},s)\not\rightarrow\textnormal{HC}(t_{n},s).

Proof.
  1. 1.

    To prove the first clause, suppose that a sequence sns_{n} satisfies HC(sn,s)(s_{n},s). This means that we are able to find a number NN such that nN\forall n\geq N, |sns|<r\left\lvert s_{n}-s\right\rvert<r for any r<r\in\mathbb{R^{Z_{<}}} which includes infinitesimals. By using the same NN, sns_{n} will satisfy CC(sn,s)(s_{n},s) .

  2. 2.

    To prove the second clause, take the sequence tn=1nt_{n}=\frac{1}{n} where nn\in\mathbb{N}. This sequence satisfies CC(tn,0)(t_{n},0), but it does not satisfy HC(tn,s)(t_{n},s) for any ss (as any rΔr\in\Delta will satisfy the negation of Definition 5.41).

Lemma 5.45.

Let (sn)(s_{n}) be a sequence in <\mathbb{R^{Z_{<}}} such that HC(sn,ss_{n},s) is hold. Then, HC(|sn|,|s|\left\lvert s_{n}\right\rvert,\left\lvert s\right\rvert) is hold.

Proof.

Let r>0<r>0\in\mathbb{R^{Z_{<}}} be given. Then this means that there exists NN\in\mathbb{N} such that m>N\forall m>N, |sms|<r\left\lvert s_{m}-s\right\rvert<r. Therefore, we also have

m>N||sm||s|||sms|<r.\displaystyle\forall m>N\textnormal{, }\left\lvert\left\lvert s_{m}\right\rvert-\left\lvert s\right\rvert\right\rvert\leq\left\lvert s_{m}-s\right\rvert<r.

Hence, HC(|sn|,|s|\left\lvert s_{n}\right\rvert,\left\lvert s\right\rvert) is true. ∎

Theorem 5.46.

Let X<X\subset\mathbb{R^{Z_{<}}} and f:X<f:X\rightarrow\mathbb{R^{Z_{<}}}. Then ff is 𝔢\mathfrak{e}-continuous at x0Xx_{0}\in X iff for any sequence xnx_{n} in XX that satisfies HC(xn,x0)(x_{n},x_{0}), the sequence f(xn)f(x_{n}) satisfies HC(f(xn),f(x0))(f(x_{n}),f(x_{0})).

Proof.

Suppose that ff is 𝔢\mathfrak{e}-continuous at x0x_{0} and let the sequence xnx_{n} be defined in XX and that xnx_{n} hyper converges to x0x_{0}. Now let 𝔢>0\mathfrak{e}>0 be given. Then from Theorem 5.39, there exists 𝔢2>0<\mathfrak{e}_{2}>0\in\mathbb{R^{Z_{<}}} such that

if xXx\in X and |xx0|<𝔢2\left\lvert x-x_{0}\right\rvert<\mathfrak{e}_{2}, then |f(x)f(x0)|<𝔢\left\lvert f(x)-f(x_{0})\right\rvert<\mathfrak{e}.

Now since xnx_{n} hyper converges to x0x_{0}, then there exists NN\in\mathbb{N} such that nN\forall n\geq N |xnx0|<𝔢2\left\lvert x_{n}-x_{0}\right\rvert<\mathfrak{e}_{2}. Thus we have

nN |f(xn)f(x0)|<𝔢.\displaystyle\forall n\geq N\textnormal{ }\left\lvert f(x_{n})-f(x_{0})\right\rvert<\mathfrak{e}.

and so the sequence f(xn)f(x_{n}) hyper converges to f(x0)f(x_{0}).
For the converse, we will prove the contrapositive. Suppose that ff is not 𝔢\mathfrak{e}-continuous at x0x_{0}. Then it means that there exists 𝔢0>0<\mathfrak{e}_{0}>0\in\mathbb{R^{Z_{<}}} such that for all 𝔢2>0<\mathfrak{e}_{2}>0\in\mathbb{R^{Z_{<}}}, there exists xXx\in X such that |xx0|<𝔢2\left\lvert x-x_{0}\right\rvert<\mathfrak{e}_{2} but |f(x)f(x0)|>𝔢0\left\lvert f(x)-f(x_{0})\right\rvert>\mathfrak{e}_{0}. In particular, for all nn\in\mathbb{N}, there exists xnXx_{n}\in X such that |xnx0|<𝔢2\left\lvert x_{n}-x_{0}\right\rvert<\mathfrak{e}_{2} and |f(xn)f(x0)|>𝔢0\left\lvert f(x_{n})-f(x_{0})\right\rvert>\mathfrak{e}_{0}. Thus xnx_{n} is a sequence in XX that hyper converges to x0x_{0}, but the sequence f(xn)f(x_{n}) does not hyper converge to f(x0)f(x_{0}). ∎

Definition 5.47.

Let sns_{n} be a sequence in <\mathbb{R^{Z_{<}}}. Then we say that sns_{n} is a hyper-Cauchy sequence iff 𝔢<\forall\mathfrak{e}\in\mathbb{R^{Z_{<}}}, N\exists N\in\mathbb{N} such that

l,mN |slsm|<𝔢.\displaystyle\forall l,m\geq N\textnormal{ }\left\lvert s_{l}-s_{m}\right\rvert<\mathfrak{e}.
Theorem 5.48.

Every hyper convergent sequence in <\mathbb{R^{Z_{<}}} is a hyper-Cauchy sequence.

Proof.

Let sns_{n} be a sequence in <\mathbb{R^{Z_{<}}} that satisfies HC(sn,s)(s_{n},s). We want to show that sns_{n} is hyper-Cauchy. Let 𝔢<\mathfrak{e}\in\mathbb{R^{Z_{<}}} be given. Then there exists NN\in\mathbb{N} such that n>N\forall n>N, |sns|<𝔢2\left\lvert s_{n}-s\right\rvert<\frac{\mathfrak{e}}{2}. Then for all l,m>Nl,m>N, we have

|slsm|=|sls(sms)||sls|+|sms|<𝔢2+𝔢2=𝔢\displaystyle\left\lvert s_{l}-s_{m}\right\rvert=\left\lvert s_{l}-s-(s_{m}-s)\right\rvert\leq\left\lvert s_{l}-s\right\rvert+\left\lvert s_{m}-s\right\rvert<\frac{\mathfrak{e}}{2}+\frac{\mathfrak{e}}{2}=\mathfrak{e}

and so sns_{n} is hyper-Cauchy. ∎

Conjecture 5.49.

The set <\mathbb{R^{Z_{<}}} is hyper-Cauchy complete with respect to the 𝔢\mathfrak{e}-topology.

Lemma 5.50.

Let sns_{n} be a sequence in <\mathbb{R^{Z_{<}}} whose members are just real numbers – that is, for all ssns\in s_{n}, Nstϵ(s)=Nstω(s)=\texttt{Nst}_{\epsilon}(s)=\texttt{Nst}_{\omega}(s)=\emptyset. Then sns_{n} is hyper-Cauchy if and only if there exists NN\in\mathbb{N} such that sm=sNs_{m}=s_{N} for all mNm\geq N.

Proof.

Let sns_{n} be a hyper-Cauchy sequence in <\mathbb{R^{Z_{<}}} whose members are real numbers. Then there exists NN\in\mathbb{N} such that

|smsl|<ϵ for all m,lN.\displaystyle\left\lvert s_{m}-s_{l}\right\rvert<\epsilon\textnormal{ for all }m,l\geq N. (12)

Since sns_{n} is a sequence of real numbers, we obtain from Inequality 12 that for all m,lNm,l\geq N, |smsl|=0\left\lvert s_{m}-s_{l}\right\rvert=0 and so sm=sNs_{m}=s_{N} for all mNm\geq N.
Conversely, let sns_{n} be a sequence in <\mathbb{R^{Z_{<}}} whose members are real numbers and assume that there exists NN\in\mathbb{N} such that sm=sNs_{m}=s_{N} for all mNm\geq N. Now let 𝔢>0\mathfrak{e}>0 be given. We have that for all l,mNl,m\geq N, |smsl|=0<𝔢\left\lvert s_{m}-s_{l}\right\rvert=0<\mathfrak{e} and so sns_{n} is hyper-Cauchy. ∎

Another possible way to define convergence in our set is through the concept of \ell^{\infty} as follows:

Definition 5.51 (<\mathbb{R^{Z_{<}}}-Convergence).

Suppose that sns_{n} is a sequence where every member of it is another sequence itself, i.e.

sn=(sn)1,(sn)2,(sn)3,,(sn)i,s_{n}=(s_{n})_{1},(s_{n})_{2},(s_{n})_{3},\dots,(s_{n})_{i},\dots.

Then, sns_{n} converges to 𝐬\mathbf{s} iff m\forall m\in\mathbb{N}, N\exists N such that

nN\forall n\geq N, i\forall i |(sn)isi|<1m\left\lvert(s_{n})_{i}-s_{i}\right\rvert<\frac{1}{m}.

We write RC(sn,s)(s_{n},s) to denote that a sequence sns_{n} is <\mathbb{R^{Z_{<}}}-convergent to ss.

Example 5.52.

The sequence sn=1n^,0,0,s_{n}=\langle\widehat{\frac{1}{n}},0,0,\dots\rangle is <\mathbb{R^{Z_{<}}}-convergent to 0.

The next interesting question is which of the three definitions above can be used to define convergence in <\mathbb{R^{Z_{<}}}? Unfortunately, neither of them is adequate to serve as the definition of convergence in our set. The three examples below demonstrate the reason. The first example shows that when Classical Convergence is adopted in <\mathbb{R^{Z_{<}}}, convergence is no longer unique. While the second one shows how adopting Definition 5.41 gave something unexpected occurs in our set, the last example shows why <\mathbb{R^{Z_{<}}}-convergence is not adequate.

Example 5.53.

Suppose that sns_{n} is a sequence defined by:

sn=0^,n,0,s_{n}=\langle\widehat{0},n,0,\dots\rangle.

Then by using Definition 5.40 above and the fact that any infinitesimals are less than any rational numbers, sns_{n} classically converges to 100ϵ100\epsilon, 200ϵ200\epsilon, 300ϵ300\epsilon, and so on. In other words, the sequence sns_{n} satisfies (CC(sn,100ϵ)(s_{n},100\epsilon)), (CC(sn,200ϵ)(s_{n},200\epsilon)), (CC(sn,300ϵ)(s_{n},300\epsilon)), and so on.

Example 5.54.

Using Definition 5.41, the sequence sn=1n^,0,0,s_{n}=\langle\widehat{\frac{1}{n}},0,0,\dots\rangle does not converge in the usual sense to 0, i.e. sns_{n} does not satisfy HC(sn,0)(s_{n},\textbf{0}). Taking r=ϵ=0^,1,0,r=\epsilon=\langle\widehat{0},1,0,\dots\rangle and n=N+1n=N+1 will show this.

Example 5.55.

The sequence sn=ϵns_{n}=\epsilon^{n} does not <\mathbb{R^{Z_{<}}}-converge to 0, as it should do intuitively.

Thus, this leaves us with the three definitions of convergence used in <\mathbb{R^{Z_{<}}}. There is no one definition of convergence in our set. This is not necessarily a bad thing, it simply means that our notion of convergence will differ from that of classical analysis.

Note that our attempts to have a proper notion of continuity and convergence in <\mathbb{R^{Z_{<}}} can be used in the area of reverse mathematics. From what we have done here, it can help us to gain a better understanding about some necessary condition, for example, for a function ff to be continuous or for a sequence to be convergent.

6 Some Notes on The Computability in <\mathbb{R^{Z_{<}}}

A computable function is a function ff which could, in principle, be calculated using a mechanical calculation tool and given a finite amount of time. In the language of computer science, we would say that there is an algorithm computing the function. A computable real number is, in essence, a number whose approximations are given by a computable function.

The notion of a function \mathbb{N}\to\mathbb{N} being computable is well understood. In fact, all definitions that so far capturing this idea (such as Turing Machines, Markov Algorithms, Lambda Calculus, the (partial) recursive functions, and many more) have all led to the same class of functions. This, in turn, has led to the so called Church-Markov-Turing thesis, which says that this class is exactly what computable intuitively means. Given computable pairing functions also, immediately, lead to a notion of computability for other function types such as km\mathbb{N}^{k}\to\mathbb{N}^{m}, \mathbb{N}\to\mathbb{Z} or \mathbb{N}\to\mathbb{Q}. If we see a real number as a sequence of rational approximation, we also get a definition of a computable real number.

However, we have to be a bit careful. There are many equivalent formulations for when a real number rr is computable, that work well in practice. This happens such as when

  • there is a finite machine that computes a quickly converging111111That is with a fixed modulus of Cauchyness. Cauchy sequence that converges to rr, or

  • it can be approximated by some computable function f:f:\mathbb{N}\rightarrow\mathbb{Z} such that: given any positive integer nn, the function produces an integer f(n)f(n) such that

    f(n)1naf(n)+1n\frac{f(n)-1}{n}\leq a\leq\frac{f(n)+1}{n}.

We denote the set of all computable real numbers by c\mathbb{R}_{c}. It is well known (and also well studied) that many real numbers, such as π\pi or ee, are computable. However, not every real number is computable.

One possibility that does not turn out to be useful is to write down a real number by using its decimal representation.121212Consider a number rr such that there is an algorithm whose input is nn, and it will give the nthn^{\textnormal{th}}-digit of rr’s decimal representation. The set of all real numbers that have a computable decimal representation is denoted by d\mathbb{R}_{d}.

Remark 6.1.

Although the set d\mathbb{R}_{d} is closed under the usual arithmetic operations, we have to be careful of what it really means. Take, for example, addition. We know that if xx and yy are in d\mathbb{R}_{d}, then x+yx+y is also in d\mathbb{R}_{d}. However, it does not mean that the addition operator itself is computable.

These ideas of computability can be extended to infinitesimals. In <\mathbb{R^{Z_{<}}}, we define its member to be computable if it satisfies the condition as stated in Definition 6.2.

Definition 6.2.

A number 𝐳<\mathbf{z}\in\mathbb{R^{Z_{<}}} is computable iff there is a computable function ff such that f(n,)f(n,\cdot) are computable numbers and

𝐳=f(1,),f(2,),,f(l,)^,\mathbf{z}=\langle f(1,\cdot),f(2,\cdot),\dots,\widehat{f(l,\cdot)},\dots\rangle

where l=f(0,0)l=f(0,0) denotes the index where the St(𝐳)\texttt{St}(\mathbf{z}) is. We denote the set of all computable members of <\mathbb{R^{Z_{<}}} by cZ<\mathbb{R}_{c}^{Z_{<}}.

In this section, we showed that the standard arithmetic operations (functions) in <\mathbb{R^{Z_{<}}} are computable (provided that the domain and codomain of those functions are (in) cZ<\mathbb{R}_{c}^{Z_{<}}). This was done by explicitly showing the program for each one of them. We actually uses a concrete implementation of these ideas in the programming language Python, whose syntax should be intuitively understandable even by those not familiar with it. There is also no need to show that our programs are correct, since they are so short that such a proof would be trivial.

Assuming that we already had a working implementation of c\mathbb{R}_{c}, our class cZ<\mathbb{R}_{c}^{Z_{<}} could be implemented as in Listing LABEL:mycode. There we defined the members of our set cZ<\mathbb{R}_{c}^{Z_{<}} (basically just a container for the index ll as in Definition 6.2 and the sequence of digits) and how their string representation would look like.

class infreal:
def __init__(self, digits, k=0):
self.k = k
self.digits = digits
def __repr__(self):
if self.k == 0:
return ”^” + ”, .join([str(self.digits(i)) for i in range(self.k,self.k+7)]) + ”, …”
else:
return ”, .join([str(self.digits(i)) for i in range(self.k)]) + ”, ^” + ”, .join([str(self.digits(i)) for i in range(self.k,self.k+7)]) + ”, …”
def __getitem__(self, key): return self.digits(key)
Listing 1: How to define the members of cZ<\mathbb{R}_{c}^{Z_{<}}.
Example 6.3.

Suppose that we want to write the number 𝟏=1^,0,0,\mathbf{1}=\langle\widehat{1},0,0,\dots\rangle. Then by writing

One=infreal(lambda n:one if n==0 else zero, 0)

where zero and one are the real numbers 0 and 11, respectively, we just created the number 𝟏\mathbf{1} in our system. The second argument of the function infreal is just to give how many digits we want to have before the real part of our number (the number with a hat). Its input and output will look like as follows:

>>> zero = real(0)
>>> one = real(1)
>>> One = infreal(lambda n: one if n==0 else zero, 0)
>>> One
^1, 0, 0, 0, 0, 0, 0, …\end{lstlisting}
Furthermore, we will also be able to know what is its $n^{\textnormal{th}}$ digit for any $n\in\mathbb{N}$. See the code below:
\begin{lstlisting}[language=Python, numbers=none]
>>> One
^1, 0, 0, 0, 0, 0, 0,
>>> One[0]
1
>>> One[-56]
0
>>> One[2454]
0\end{lstlisting}
\end{example}
\begin{example}
Similar to \Cref{ex:6.4}, the numbers $\mathbf{\epsilon}$ and $\mathbf{\omega}$ could also be defined in our system.
\begin{lstlisting}[language=Python, numbers=none]
>>> Epsilon = infreal(lambda n: one if n == 1 else zero, 0)
>>> Epsilon
^0, 1, 0, 0, 0, 0, 0,
>>> Omega = infreal(lambda n: one if n == 0 else zero, 1)
>>> Omega
1, ^0, 0, 0, 0, 0, 0, 0, \end{lstlisting}
Also, we will be able to have exotic numbers such as $\mathbf{\me}+\mathbf{2\me\epsilon}+\mathbf{3\me\epsilon}^2+\mathbf{4\me\epsilon}^3+\dots$ and its code will be as follows:
\begin{lstlisting}[language=Python, numbers=none]
>>> e = exp(rational(1,1))
>>> Funny = infreal(lambda n: real(rational(n + 1, 1)) * e if n > -1 else zero, 0)
>>> Funny
^2.71828, 5.43656, 8.15485, 10.8731, 13.5914, 16.3097, 19.028,
>>> Funny[43532]
118335
>>> Funny[-12964]
0\end{lstlisting}
\end{example}
\Cref{thm:addIsComputable}-\ref{thm:mulIsComputable} show that addition, subtraction, and multiplication in $\Rzlc$ are computable.
\begin{theorem}
\label{thm:addIsComputable}
Suppose that we have $x,y\in\Rzlc$. Then the function $\widehat{+}_c$ defined by
\begin{eqnarray*}
\widehat{+}_c: & \Rzlc & \rightarrow\Rzlc\\
& (\mathbf{x},\mathbf{y})& \mapsto \mathbf{x}\widehat{+}\mathbf{y}
\end{eqnarray*}
is computable.
\end{theorem}
\begin{proof}
The following code shows that the function $\widehat{+}_c$ defined above is computable.
\begin{lstlisting}[language=Python, numbers=none]
def __add__(self, other):
k = max(self.k, other.k)
return infreal(lambda n: self.digits(n - (k-self.k)) + other.digits(n - (k-other.k)), k)\end{lstlisting}
\end{proof}
\begin{example}
Suppose that we want to add $\mathbf{\epsilon}$, $\mathbf{\omega}$, and $\mathbf{1}$. Then we will have:
\begin{lstlisting}[language=Python, numbers=none]
>>> Omega
1, ^0, 0, 0, 0, 0, 0, 0,
>>> Epsilon
^0, 1, 0, 0, 0, 0, 0,
>>> Epsilon + Omega + One
1, ^1, 1, 0, 0, 0, 0, 0, …\end{lstlisting}
\end{example}
\begin{theorem}
\label{thm:minusIsComputable}
Suppose that we have $x,y\in\Rzlc$. Then the function $\widehat{-}_c$ defined by
\begin{eqnarray*}
\widehat{-}_c: & \Rzlc & \rightarrow\Rzlc\\
& (\mathbf{x},\mathbf{y})& \mapsto \mathbf{x}\widehat{-}\mathbf{y}
\end{eqnarray*}
is computable.
\end{theorem}
\begin{proof}
The following code shows that the function $\widehat{-}_c$ defined above is computable.
\begin{lstlisting}[language=Python,escapeinside={(*}{*)}]
def __neg__(self): return infreal(lambda n: -self[n], self.k) (*\label{line8}*)
def __sub__(self, other): return (self + (-other))\end{lstlisting}
The definition in line \ref{line8} shows that the additive inverse function is computable.
\end{proof}
\begin{example}
Suppose that we want to add $\mathbf{\epsilon}\widehat{-}\mathbf{\omega}$ to $\mathbf{1}$. Then we will have:
\begin{lstlisting}[language=Python, numbers=none]
>>> Epsilon - Omega + One
-1, ^1, 1, 0, 0, 0, 0, 0, …\end{lstlisting}
\end{example}
\begin{theorem}
\label{thm:mulIsComputable}
Suppose that we have $x,y\in\Rzlc$. Then the function $\widehat{\times}_c$ defined by
\begin{eqnarray*}
\widehat{\times}_c: & \Rzlc & \rightarrow\Rzlc\\
& (\mathbf{x},\mathbf{y})& \mapsto \mathbf{x}\widehat{\times}\mathbf{y}
\end{eqnarray*}
is computable.
\end{theorem}
\begin{proof}
The following code shows that the function $\widehat{\times}_c$ defined above is computable.
\begin{lstlisting}[language=Python,escapeinside={(*}{*)}, numbers=none]
def __mul__(self, other):
k = self.k + other.k
def digits(n):
if n < 0:
return zero
else:
return reduce((lambda x,y:x+y), [self.digits(i) * other.digits(n - i) for i in range(n + 1)])
return infreal(digits, k) \end{lstlisting}
\end{proof}
\begin{example}
Suppose that we want to add $\mathbf{\epsilon}\widehat{\times}\mathbf{\omega}$ to $\mathbf{1}$ and also $-\mathbf{\epsilon}^2$ to $\mathbf{1}$. Then we will have:
\begin{lstlisting}[language=Python, numbers=none]
>>> Epsilon * Omega + One
0, ^2, 0, 0, 0, 0, 0, 0,
>>> Epsilon * -Epsilon + One
^1, 0, -1, 0, 0, 0, 0, …\end{lstlisting}
\end{example}
\subsection{Some Remarks on Non-Computability in $\Rzl$}
\begin{remark}
Even though division on $\R_c$ is computable (assuming the input does not equal $0$), the same can not be said of $\Rzlc$.
\end{remark}
\begin{remark}
Suppose that we have a number $\mathbf{x}\in\Rzl$. Then without any further information, the process of finding $\mathbf{x}^{-1}$ (the multiplicative inverse of $\mathbf{x}$) is not computable. One extra information needed to make it computable in $\Rzl$ is how many digits we want to have in $\mathbf{x}^{-1}$, which of course will affect the accuracy of our result. More precisely, it is known that it is not possible to give an algorithm that, a given number $a \in \R_c$, decides whether $a = 0$ or $\lnot a =0$. So let $a \in \R_c$ and consider $\mathbf{z} = a + \mathbf{\epsilon}$. We have $ \mathbf{\epsilon} \neq 0$. If $a = 0$ then $\mathbf{z}^{-1} = \mathbf{\omega}$. If $a \neq 0$ then $z^{-1} < \mathbf{\omega}$. Thus by checking whether the $\omega$-part of $z^{-1}$ is less than $1$ or greater than $0$, we would be able to decide whether $a = 0$ or $\lnot a =0$.
\end{remark}
\begin{remark}
Similarly surprising, we can show that the absolute value function, which is computable for $\R_c$, is not computable for $\Rzlc$. Here the absolute value function is the function
\[ \left| \mathbf{z} \right| = \begin{cases}
\mathbf{z} & \text{if } \mathbf{z} \geq 0 \\ -\mathbf{z} & \text{if } \mathbf{z} < 0
\end{cases} \]
Similar to the above, take $a \in \R_c$ and consider $z = |a| - \mathbf{\epsilon}$. If $a =0$ then $|z| =- |a| + \mathbf{\epsilon}$. If $a \neq 0$ then $|z| = |a| - \mathbf{\epsilon}$. Thus by checking the $\epsilon$-part of $|z|$, we would be able to decide whether $a = 0$ or $\lnot a =0$.
\end{remark}
This also leads to comparison between numbers not being computable. Now this is also the case in $\R_c$. However, for numbers $x,y \in \R_c $ such that $x \neq y$, we can decide whether $x<y$ or $x > y$. This does not extend to $\Rzlc$:
\begin{remark}
Comparison among the members in $\Rzlc$ is not computable.
Again, let $a \in \R_c$ and consider $\mathbf{x} = |a| + \mathbf{\epsilon}$ and $\mathbf{y} = \textbf{2}|a|$. If $a = 0$ then $\mathbf{x} > \mathbf{y}$, and if $a \neq 0$ then $\mathbf{x} <\mathbf{y}$. Thus, once again we would be able to decide whether $a = 0$ or $\lnot a =0$.
\end{remark}

7 Conclusion and Suggestions for Further Research

In this article we treated the nonstandard real numbers in the spirit of the ‘Chunk and Permeate’ approach. The sets \mathbb{R} and \mathbb{{}^{*}R} were thrown together (that means: combining their languages and axioms) and arising inconsistency issues were dealt with using paraconsistent reasoning strategy. In this case — and this is one of the interesting novelties — the procedure was made explicit by introducing new sets ^\mathbb{\widehat{R}} and in turn <\mathbb{R^{Z_{<}}}, the separate ‘chunks’. In a sense, we transferred the ‘Chunk and Permeate’ approach from the theoretical level to the (explicit) model level. Whereas the impact on infinitesimal mathematics was only sketched in [9], it was worked out in detail here, using the sets ^\mathbb{\widehat{R}} and <\mathbb{R^{Z_{<}}}. After introducing the theoretical background, we constructed the new model of nonstandard analysis in detail. The remaining part of this paper lies in an extensive discussion of topological, applied (in the sense of calculus), and computability issues of the obtained model. A side result of the constructed set <\mathbb{R^{Z_{<}}} was a direct consistency proof of the Grossone theory, see [23].

On the topological aspect in Section 4, we introduced some new notions of metrics, balls, open sets, and etc in <\mathbb{R^{Z_{<}}} together with their properties. On the applied aspect in Section 5, some new concepts on the calculus in <\mathbb{R^{Z_{<}}} were discussed, e.g. derivative (we successfully developed a permeability relation such that the derivative function in <\mathbb{R^{Z_{<}}} can be permeated to \mathbb{R}), continuity, and convergence. For the two last issues, some new notions were introduced in this article. First of all, we discussed the three possible notions of continuity that can be applied to either \mathbb{R} or <\mathbb{R^{Z_{<}}}. We also determined how they relate to each other in their respective model. While doing that, we discovered a new kind of fractals — infinitesimal fractals. After analysing three possible notions of continuity, we decided that the best notion that can be used in our setting <\mathbb{R^{Z_{<}}} is the ϵ\epsilon-δ\delta definition and by doing that, we do not only preserve much of the spirit of classical analysis but also retain the intuition of infinitesimals. After establishing our position, we introduced a more detailed notion of continuity which is called (k,nk,n)-continuity (as can be seen in Definition 5.23). We explored how this new notion of continuity behaves, e.g. what happens with the composition of two continuous functions and also how this notion behaves under multiplication. It is worth pointing out here that this new notion of continuity is a much more fine-grained notion than the classical continuity. Last but not least, we showed that the set <\mathbb{R^{Z_{<}}} has nice computability features. We succeeded in building a program, in Python, to show that we can have a computable number <\mathbb{R^{Z_{<}}}. The set of all these computable numbers is denoted by cZ<\mathbb{R}_{c}^{Z_{<}}. We also showed some interesting remarks regarding this computability issue.

In term of further research, we indicated some possible areas of further development as follows. First, one could try to do infinitesimal analysis using the relevant logic R. The comparison between the results (perhaps) gotten in R and the one described here might be interesting, especially in term of usefulness and simplicity. Secondly, regarding the ‘transfer principle’, our intuition says that it is equivalent to the notion of permeability in the Cchunk & Permeate strategy. One could try to formally prove it, or disprove it. Thirdly, in term of computability issue, using the calculus on <\mathbb{R^{Z_{<}}}, one could try to formulate the necessary and sufficient conditions for the derivatives of functions, for example, on a computer to exist. And perhaps, showing also how to find these derivatives whenever they exist. This, of course, can also be applied to the other notions. Fourthly, as been said in the previous sections, some results described in this article could help us to gain a better understanding in another area of research (the two that were mentioned in Section 5 are reverse mathematics and quantum physics). One could try to work out the details on this.

In general, with the new consistent sets created in this work, new opportunities awaits mathematicians. One of the joys of mathematics is to explore a world which has no physical substance, and yet is everywhere in every aspect of our lives. Infinities and infinitesimals offer ways to explore hitherto unseen aspects of our world and our universe, by giving us the vision to see the greatest and smallest aspects of life. Even a naïve set, when it demonstrates harmony, offer another dimension of even clearer precision. In a wide sense, the work on this article can also be seen as a contribution to bridge (the antipodes) constructive analysis and nonstandard analysis. This problem has been extensively (and intensively) discussed in the past few years (see for example [40, 41, 42, 8, 43, 31]).

References

  • [1] Arkeryd, L.O., Cutland, N.J., Henson, C.W.: Nonstandard analysis: Theory and applications, vol. 493. Springer Science & Business Media (2012)
  • [2] Avron, A.: Natural 3-valued logics-characterization and proof theory. The Journal of Symbolic Logic 56(01), 276–294 (1991)
  • [3] Bartle, R.G., Sherbert, D.R.: Introduction to Real Analysis, vol. 2. Wiley (1992)
  • [4] Batens, D., Mortensen, C., Priest, G., Bendegem, J.P.V.: Frontiers in Paraconsistent Logic. Research Studies Press (2000)
  • [5] Bell, J.L.: A primer of infinitesimal analysis. Cambridge University Press (1998)
  • [6] Benci, V., Di Nasso, M.: Alpha-theory: an elementary axiomatics for nonstandard analysis. Expositiones Mathematicae 21(4), 355–386 (2003)
  • [7] Bishop, E., et al.: H. jerome keisler, elementary calculus. Bulletin of the American Mathematical Society 83(2), 205–208 (1977)
  • [8] Bournez, O., Ouazzani, S.: Cheap non-standard analysis and computability. arXiv preprint arXiv:1804.09746 (2018)
  • [9] Brown, B., Priest, G.: Chunk and permeate, a paraconsistent inference strategy. part i: The infinitesimal calculus. Journal of Philosophical Logic 33(4), 379–388 (2004)
  • [10] Cantor, G.: Contributions to the Founding of the Theory of Transfinite Numbers. 1. Open Court Publishing Company (1915)
  • [11] Da Costa, N.C., et al.: On the theory of inconsistent formal systems. Notre Dame Journal of Formal Logic 15(4), 497–510 (1974)
  • [12] Dunn, J.M., Restall, G.: Relevance logic. In: Handbook of philosophical logic, pp. 1–128. Springer (2002)
  • [13] Earman, J.: Infinities, infinitesimals, and indivisibles: the leibnizian labyrinth. Studia Leibnitiana pp. 236–251 (1975)
  • [14] Edwards, H.: Euler’s definition of the derivative. Bulletin of the American Mathematical Society 44(4), 575–580 (2007)
  • [15] Fletcher, P., Hrbacek, K., Kanovei, V., Katz, M.G., Lobry, C., Sanders, S.: Approaches to analysis with infinitesimals following robinson, nelson, and others. Real Analysis Exchange 42(2), 193–252 (2017)
  • [16] French, S.: Action at a distance. In: The Routledge Encyclopedia of Philosophy. Taylor and Francis (2005). www.rep.routledge.com/articles/thematic/action-at-a-distance/v-1. doi:10.4324/9780415249126-N113-1
  • [17] Gödel, K.: Consistency of the Continuum Hypothesis.(AM-3), vol. 3. Princeton University Press (2016)
  • [18] Goldblatt, R.: Lectures on The Hyperreals. Springer (1998)
  • [19] Gray, J.: The real and the complex: a history of analysis in the 19th century. Springer (2015)
  • [20] Hardy, G.H.: Orders of infinity. Cambridge University Press (2015)
  • [21] Henle, M.: Which Numbers are Real? The Mathematical Association of America (2012)
  • [22] Kanovei, V., Reeken, M.: Nonstandard analysis, axiomatically. Springer Science & Business Media (2013)
  • [23] Lolli, G.: Metamathematical investigations on the theory of grossone. Applied Mathematics and Computation 255, 3–14 (2015)
  • [24] Mares, E.: Relevance logic. In: E.N. Zalta (ed.) The Stanford Encyclopedia of Philosophy, spring 2014 edn. (2014). Available online at http://plato.stanford.edu/archives/spr2014/entries/logic-relevance/
  • [25] Mayberry, J.P.: The foundations of mathematics in the theory of sets, vol. 82. Cambridge University Press (2000)
  • [26] McKubre-Jordens, M., Weber, Z.: Real analysis in paraconsistent logic. Journal of Philosophical Logic 41(5), 901–922 (2012)
  • [27] Mortensen, C.E.: Models for inconsistent and incomplete differential calculus. Notre Dame Journal of Formal Logic 31(2), 274–285 (1990)
  • [28] Mortensen, C.E.: Inconsistent Mathematics, vol. 312. Springer (1995)
  • [29] Mortensen, C.E.: Inconsistent Geometry. College Publications (2010)
  • [30] Musser, G.: Spooky Action at a Distance: The Phenomenon that Reimagines Space and Time–and what it Means for Black Holes, the Big Bang, and Theories of Everything. Macmillan (2015)
  • [31] Normann, D., Sanders, S.: Computability theory, nonstandard analysis, and their connections. The Journal of Symbolic Logic 84(4), 1422–1465 (2019)
  • [32] Nugraha, A.: Naïve infinitesimal analysis. Ph.D. thesis, School of Mathematics and Statistics University of Canterbury, New Zealand (2018). https://ir.canterbury.ac.nz/handle/10092/16627
  • [33] Popkin, G.: Einstein’s ‘spooky action at a distance’ spotted in objects almost big enough to see (2018). Available online at http://www.sciencemag.org/news/2018/04/einstein-s-spooky-action-distance-spotted-objects-almost-big-enough-see. doi:10.1126/science.aat9920
  • [34] Priest, G.: The logic of paradox. Journal of Philosophical Logic 8(1), 219–241 (1979)
  • [35] Priest, G.: Minimally inconsistent lp. Studia Logica 50(2), 321–331 (1991)
  • [36] Priest, G.: Paraconsistent logic. In: Handbook of Philosophical Logic, pp. 287–393. Springer (2002)
  • [37] Robinson, A.: Non-standard analysis. Princeton University Press (1974)
  • [38] Rosinger, E.E.: On the safe use of inconsistent mathematics. arXiv preprint arXiv:0811.2405 (2008)
  • [39] Rosinger, E.E.: Basic Structure of Inconsistent Mathematics (2011). URL https://hal.archives-ouvertes.fr/hal-00552058. MSC 00A05,00A69,00A71,00A99,03Bxx,03B60,03B99
  • [40] Sanders, S.: On the connection between nonstandard analysis and constructive analysis. Logique et Analyse pp. 183–210 (2013)
  • [41] Sanders, S.: The effective content of reverse nonstandard mathematics and the nonstandard content of effective reverse mathematics. arXiv preprint arXiv:1511.04679 (2015)
  • [42] Sanders, S.: From nonstandard analysis to various flavours of computability theory. In: International Conference on Theory and Applications of Models of Computation, pp. 556–570. Springer (2017)
  • [43] Sanders, S.: The gandy–hyland functional and a computational aspect of nonstandard analysis. Computability 7(1), 7–43 (2018)
  • [44] Sergeyev, Y.D.: Arithmetic of infinity. Edizioni Orizzonti Meridionali (2003)
  • [45] Sergeyev, Y.D.: The olympic medals ranks, lexicographic ordering, and numerical infinities. The Mathematical Intelligencer 37(2), 4–8 (2015)
  • [46] Shalm, L.K., Meyer-Scott, E., Christensen, B.G., Bierhorst, P., Wayne, M.A., Stevens, M.J., Gerrits, T., Glancy, S., Hamel, D.R., Allman, M.S., et al.: Strong loophole-free test of local realism. Physical review letters 115(25), 250402 (2015)
  • [47] Spalt, D.: Cauchy’s continuum-a historiographic approach via cauchy’s sum theorem. Archive for History of Exact Sciences 56(4), 285–338 (2002)
  • [48] Susskind, L., Hrabovsky, G.: The Theoretical Minimum: What You Need To Know To Start Doing Physics. Basic Books (2014)
  • [49] Tao, T.: A cheap version of nonstandard analysis. https://terrytao.wordpress.com/2012/04/02/a-cheap-version-of-nonstandard-analysis/ (2012)
  • [50] Weber, Z.: Inconsistent mathematics. In: J. Fieser & B. Dowden (Eds.). The Internet Encyclopedia of Philosophy (2009). Available online at http://www.iep.utm.edu/math-inc/
  • [51] Weber, Z.: Extensionality and restriction in naive set theory. Studia Logica 94(1), 87–104 (2010)
  • [52] Weber, Z., McKubre-Jordens, M.: Paraconsistent measurement of the circle. Australasian Journal of Logic 14(1) (2017)
  • [53] Weston, M.M., Slussarenko, S., Chrzanowski, H.M., Wollmann, S., Shalm, L.K., Verma, V.B., Allman, M.S., Nam, S.W., Pryde, G.J.: Heralded quantum steering over a high-loss channel. Science advances 4(1), e1701230 (2018)