Cinematic Strawberry

Logo

Subjective Numbers: A Category-Theoretic Construction


Abstract

This paper introduces subjective numbers, a novel mathematical framework constructed within category theory that formalizes intrinsic perspective as a fundamental property of mathematical objects. Traditional mathematics assumes objective universality, with mathematical truths independent of observer. In contrast, subjective numbers possess inherent "viewpoints" dictating their relationships with other mathematical entities. These perspectives create a mathematics where relations are perspective-dependent, challenging fundamental principles of equality and equivalence relations. We present a comprehensive category-theoretic construction, demonstrating emergent properties such as path-dependent truth and perspective incompleteness (where certain equivalences are discoverable only through cross-perspective inference). Rigorous proofs establish consistency relative to ZFC and non-triviality. This paradigm shift offers a mathematically robust approach to modeling phenomena where viewpoint and context are intrinsic determinants of truth and relation.

Symbol Description
P Set of perspectives.
p, q, r, s, t Elements of P (specific perspectives).
V Value space (typically or ).
a, b, c Elements of V (values).
a(p) Subjective number with value a and perspective p.
=p Subjective equality relation from perspective p (i.e. a(p) =p b(q) means that from p's viewpoint the values are considered equal).
Rp Relation function for perspective p, mapping tuples (s, t, a, b) to {true, false}.
CPAp Cross-perspective adoption function for perspective p.
a(p → q) Subjective number with value a undergoing a perspective transition from p to q.
Rp→q Relation function for the perspective transition from p to q (i.e. it satisfies Rp→q(p → q, r, a, b) = Rq(q, r, a, b)).
𝓢𝓝 Category of subjective numbers.
Mor𝓢𝓝(p, q) Set of morphisms from perspective p to q in 𝓢𝓝.
f(a, p → q) Morphism representing the subjective number a(p) viewed from perspective q.
f(b, q → r) ∘ f(a, p → q) = f((a,b), p → r) Composition of morphisms in 𝓢𝓝, where the composite records the combined value history.
and : Equivalence on value histories (e.g. (a,0) ∼ a, ((x,y),z) ∼ (x,(y,z))).
: Equivalence on morphisms induced by .
π1 Projection function extracting the first component from an ordered pair (e.g. π1((a,b)) = a).
p ⊕ q Composite perspective formed by fusing perspectives p and q.
Rp⊕q Relation function for the composite perspective p ⊕ q.
+d, +p⊕q, +n Subjective addition operations:
+d: Dominant perspective addition (result inherits the first operand’s perspective).
+p⊕q: Perspective fusion addition (result’s perspective is the fused p ⊕ q).
+n: Novel perspective addition (result is assigned a new perspective).
×d, ×p⊕q, ×n Subjective multiplication under the corresponding propagation rules.
-r, ÷r, ^r Subjective subtraction, division, and exponentiation under propagation rule r.
d(p, q) Perspective distance function between perspectives p and q (often defined via an integral measuring differences between Rp and Rq).
μ A σ‑finite measure on V×V (used for integration in the definition of d(p, q)).
λ Lebesgue measure (typically when V = ).
𝟙{condition} Indicator function for a given condition (encodes true as 1 and false as 0).
f Generic fusion function used in defining composite relation functions (e.g. for fusing evaluations in p ⊕ q).
η Parameter in [0,1] (representing randomness or bias) used in novel perspective generation.
g Generation function for the relation function of a novel perspective.
Φ Translation function from subjective mathematical statements to modal logic.
L Set of truth values in a multi‑valued logic.
vi Elements of L (specific truth values).
s Modal necessity operator from perspective s.

1. Introduction

1.1 Motivation and Central Idea

Mathematics, in its conventional form, operates under an implicit assumption of objective universality: mathematical truths are considered independent of any observer or context. This foundational principle of objectivity, particularly powerful in fields prioritizing symmetry and invariance, introduces significant limitations in mathematics' ability to formally represent perspective-dependent phenomena. The elegance and efficacy of traditional mathematics are deeply rooted in symmetry, from the symmetric nature of equality itself to the commutative properties of fundamental operations and the symmetries underpinning geometric understanding. However, this focus on symmetry presents challenges when formalizing inherently asymmetrical, directional, or viewpoint-dependent patterns.

Consider a basic symmetry principle: in conventional mathematics, if a = b, then necessarily b = a. Yet, in many contexts, evaluations fundamentally defy this symmetry. Evaluations of quality, preference, or relevance are often directional and perspective-dependent. In social dynamics, one agent may perceive another as equal, while the reverse perception may not hold. In cognitive systems, information relevance is viewpoint-specific. This fundamental asymmetry of relations manifests across diverse domains, yet lacks precise mathematical formalization within standard frameworks.

This paper introduces subjective numbers as a rigorous mathematical solution to this representational gap. A subjective number is constructed within category theory to inherently embody not just a value, but an intrinsic perspective that fundamentally governs its relationships with other mathematical entities. Unlike approaches that treat perspective as an external parameter, we construct subjective numbers as mathematical objects where perspective is an intrinsic property, defined through category-theoretic morphisms that formalize how perspectives relate and interact. In our notation for the relation function Rp(s, t, a, b), the first perspective p is always the vantage perspective evaluating the relationship between subjective numbers with perspectives s and t and values a and b.

For instance, a subjective number 5(p) might consider itself equal to 5(q) from its perspective (denoted 5(p) =p 5(q)), while 5(q) might not reciprocate this evaluation from its perspective (denoted ¬(5(q) =q 5(p))). This inherent asymmetry represents a foundational departure from traditional mathematical assumptions, yet provides a consistent framework for modeling perspective-dependent phenomena. Subjective numbers offer a novel mathematical lens for examining systems where viewpoint, directionality, and intrinsic perspective are not merely contextual factors but are mathematically constitutive of the objects and relations themselves. This framework allows for the formalization of path-dependent truths, perspective incompleteness, and non-symmetric relations within a rigorous mathematical structure.

Symmetry Principle

Universe 00110000

1.2 Key Distinctions and Innovations

Subjective numbers are distinguished from related concepts by several key innovations:

Beyond Symmetric Mathematics

Classical Approach: Traditional mathematics implicitly assumes objective universality, where mathematical truths are independent of observer or context. This has led to powerful symmetry principles like "if a = b, then b = a".

Subjective Numbers' Perspective: In this framework, mathematical objects have intrinsic perspectives that govern their relationships. A subjective number 5(p) can consider itself equal to 5(q) while 5(q) does not reciprocate this evaluation, creating a rigorous foundation for modeling inherently asymmetric, directional phenomena.

2. Preliminaries and Metatheoretical Framework

2.1. Category Theory Background

We briefly recall essential category theory concepts used in our construction. For readers less familiar with category theory, we present these concepts intuitively before providing formal definitions.

Intuitively, a category consists of objects (which can be thought of as mathematical structures) and morphisms (which represent structure-preserving maps between objects). The key feature of category theory is that it focuses on relationships between objects rather than their internal structure. This relationship-centric approach makes category theory particularly suitable for our construction of subjective numbers, where perspective-dependent relationships are fundamental.

Definition 2.1 (Category): A category 𝒞 consists of:

  1. A collection of objects, denoted by Ob(𝒞).
  2. For each pair of objects A, B ∈ Ob(𝒞), a set Mor𝒞(A, B) of morphisms from A to B. If f ∈ Mor𝒞(A, B), we write f: A → B.
  3. For each triple of objects A, B, C ∈ Ob(𝒞), a composition law that assigns to each pair of morphisms f: A → B and g: B → C a morphism g ∘ f: A → C, called the composition of g and f.
  4. For each object A ∈ Ob(𝒞), an identity morphism idA: A → A.

These data are subject to the following axioms:

Definition 2.2 (Quotient Category): A quotient category is formed when we define an equivalence relation on the morphisms of a category that is compatible with composition. Two morphisms are considered equivalent if they belong to the same equivalence class. The objects of the quotient category remain the same as the original category, but the morphisms become equivalence classes of the original morphisms. For a quotient category to be well-defined, the equivalence relation must satisfy the following compatibility conditions:

  1. If f ≡ g, then dom(f) = dom(g) and cod(f) = cod(g) (i.e., equivalent morphisms must have the same source and target objects).
  2. If f₁ ≡ f₂ and g₁ ≡ g₂, and if f₁ and g₁ are composable, then f₁ ∘ g₁ ≡ f₂ ∘ g₂ (i.e., composition must be well-defined on equivalence classes).
  3. For each object A, idA ≡ idA (i.e., identity morphisms must be equivalent only to themselves).

These conditions ensure that the quotient category preserves the essential structural properties of the original category while grouping morphisms according to the desired equivalence relation.

Functors are structure-preserving maps between categories, capturing how entire categorical structures relate to each other:

Definition 2.3 (Functor): A functor F: 𝒞 → 𝒟 between categories 𝒞 and 𝒟 consists of:

  1. A mapping F: Ob(𝒞) → Ob(𝒟) that assigns to each object A in 𝒞 an object F(A) in 𝒟.
  2. For each pair of objects A, B in 𝒞, a mapping FA,B: Mor𝒞(A, B) → Mor𝒟(F(A), F(B)) that assigns to each morphism f: A → B in 𝒞 a morphism F(f): F(A) → F(B) in 𝒟.

These mappings must preserve composition and identity morphisms:

2.2. Metatheoretical Framework for Subjective Mathematics

Before proceeding to the formal construction of subjective numbers, we must establish the metatheoretical framework within which we reason about perspective-dependent mathematics. This addresses a fundamental question: How do we, as mathematicians, study a system where mathematical truth itself depends on perspective?

2.2.1 Metatheoretical Stance: In developing and analyzing subjective numbers, we adopt the standard metatheoretical stance of mathematics. This stance allows us to reason about the entire system of perspectives while remaining outside that system. Just as a set theorist reasons about sets without requiring an axiom positing the existence of a "meta-set theorist," we reason about perspective-dependent mathematics without requiring an axiomatized "meta-perspective" within the system.

This approach recognizes the distinction between:

  1. Object-level mathematics: The subjective numbers themselves, with their intrinsic perspectives and perspective-dependent relations
  2. Meta-level mathematics: Our reasoning about the system of subjective numbers, which follows standard mathematical methodology

When we make statements about "all perspectives" or system-wide properties, these are metatheoretical observations about the subjective numbers framework, not statements made from within any particular perspective in the system.

2.2.2 Philosophical Foundation of Perspective: Within our framework, a perspective is not merely a formal index or label but represents a mathematical vantage point with its own internally consistent logic and evaluation criteria. Each perspective represents:

  1. A coherent system of evaluation with its own standards of equality and relation
  2. An origination point for mathematical assertions that may differ from other perspectives
  3. A mathematical entity that can itself be studied, related to other perspectives, and combined through various operations

This philosophical foundation grounds our mathematical formalization while clarifying that perspectives are not arbitrary but represent meaningful mathematical viewpoints with their own internal structure.

2.2.3. Metatheoretical Principle of Static Continuity

Metatheoretical Principle (Static Continuity): Unless otherwise specified, we assume that once an operation on fixed perspectives produces a given result at one time, it must produce the same result at any other time.

Formal expression: ∀p, q, r ∈ P, ∀a, b, c ∈ V, ∀φ ∈ Operations: (a(p) φp b(q) = c(r) at time t1) → (a(p) φp b(q) = c(r) at time t2).

Motivation: This principle does not forbid the introduction of time. Rather, it specifies that if time is to be part of the framework, it must be introduced explicitly. It establishes a baseline of stability for operations in the subjective numbers framework, ensuring predictability and mathematical coherence in the absence of explicitly time-dependent elements. This principle allows us to initially focus on the core innovation of perspective-dependence, setting a foundation for controlled extensions into dynamic systems in future work.

2.3. Formal Definitions

2.3.1 Relation Function Formalization

The relation function Rp for each perspective p is a core component of our framework. Formally, we define:

Rp: P × P × V × V → {true, false}

For any perspectives s, t ∈ P and values a, b ∈ V, the expression Rp(s, t, a, b) indicates whether perspective p considers the subjective number a(s) to be equal to b(t).

Note on “Undefined” vs. “False”: Throughout examples, we may informally say the relation Rp(s, t, a, b) is undefined when perspective p has no information about a(s) and b(t). Formally, however, Rp always maps into {true, false}. Whenever Rp(s, t, a, b) is not explicitly declared true, it defaults to false. Hence, undefined is just an informal way of saying not recognized as equal in perspective p.

When a relation involves perspective p itself, we often write Rp(p, q, a, b) to denote how p relates its own value a to perspective q's value b.

Definition 2.4 (Well-Formed Relation Function):

A relation function Rp is considered well-formed if it satisfies the following properties:

  1. Internal Consistency: If Rp(s, t, a, b) = true and Rp(t, u, b, c) = true, then Rp(s, u, a, c) = true (transitivity within perspective p).
  2. Self‑Reflexivity: For all a ∈ V, Rp(p, p, a, a) = true (a perspective considers a value equal to itself).
  3. Value Invariance: If a = c in V, then Rp(s, t, a, b) = Rp(s, t, c, b); similarly, if b = d in V, then Rp(s, t, a, b) = Rp(s, t, a, d). (Numerical equality in V implies relation equality.)

This characterization ensures that while perspective p may differ from others in how it evaluates equality, it remains internally coherent and respects basic numeric identities.

2.3.2. Cross-Perspective Adoption Function

This function represents a form of "trust" or "acceptance" between perspectives. When CPAp(q, r) = true, perspective p is willing to incorporate q's judgments about r into its own reasoning. The Cross-Perspective Inference axiom (Axiom 4, see Section 4) references CPAp directly: if p trusts q's evaluation about r, then p extends transitivity across perspectives accordingly. Formally,

The Cross-Perspective Inference axiom (Axiom 4) references CPAp directly: if p trusts q’s evaluation about r, then p extends transitivity across perspectives accordingly. Formally,

∀p, q, r ∈ P, ∀a, b, c ∈ V: (Rp(p, q, a, b) ∧ Rq(q, r, b, c) ∧ CPAp(q, r)) → Rp(p, r, a, c).

We treat CPAp(q,r) as a primitive function in the system, so there is no circular definition: the axiom system introduces CPAp as a parameter, and Axiom 4 describes precisely how cross-perspective acceptance connects to multi-perspective transitivity. This ensures that each perspective retains control over which other viewpoints it “trusts,” preventing unrestricted chains of inference that might lead to contradictions.

2.3.3 Perspective Transition Notation

For any two perspectives p, qP, we introduce the notation a(p → q) to represent a subjective number with value a that temporarily adopts perspective q while originating from perspective p. This perspective transition construct is essential for defining how perspectives can shift and interact with the cross-perspective adoption function. Formally:

a(p → q) is a subjective number with relation function Rp → q such that for all rP and all x, yV:

Rp → q(p → q, r, x, y) = Rq(q, r, x, y)

This means that a(p → q) evaluates relations in exactly the same way that perspective q would, despite having originated from perspective p. This notation provides a formal bridge between our categorical construction and the axiom system, especially for Axiom 6 (Perspective Adoption).

Formal Framework and Definitions

Universe 00110000

3. Formal Framework and Definitions

3.1. Definition of the Category 𝓢𝓝

We define the category 𝓢𝓝 (the category of subjective numbers) as follows:

  1. Objects: The objects of 𝓢𝓝 are the perspectives, which form a non-empty set P = {p, q, r, ...}.
  2. Morphisms: For any two perspectives p, qP and for any value aV (where V is a value space such as or ), we define a morphism f(a, p → q): p → q. The set of morphisms from p to q is given by:
    Mor𝓢𝓝(p, q) = {f(a, p → q) | a ∈ V}
  3. Composition: For morphisms f(a, p → q): p → q and f(b, q → r): q → r, we define composition as:
    f(b, q → r) ∘ f(a, p → q) = f((a,b), p → r)
    This enhanced composition rule records both values in an ordered pair, preserving information about the intermediate perspective. This approach:
    • Records both the original value and intermediate value, creating a richer structure that captures the path of perspective transitions
    • Makes morphisms "path-aware," which is crucial for modeling domains where evaluation history matters
    • Forms a quotient category structure through explicit equivalence relations to satisfy categorical axioms

    Morphism Composition in Practice

    Let's illustrate the composition of morphisms with a concrete example. Consider three perspectives p, q, and r, with values in ℝ. Suppose we have:

    • A morphism f(3, p → q): p → q representing value 3 transitioning from perspective p to q
    • A morphism f(5, q → r): q → r representing value 5 transitioning from perspective q to r

    To compose these morphisms, we follow our composition rule:

    f(5, q → r) ∘ f(3, p → q) = f((3,5), p → r)

    The resulting morphism f((3,5), p → r): p → r records the history of values (first 3, then 5) as we transition from perspective p to r via q.

    To understand why this composition stores the value history, consider a case where values undergo transformations between perspectives. For example, if perspective q doubles incoming values and perspective r adds 1 to incoming values, then:

    • The value 3 from perspective p becomes 6 in perspective q
    • The value 6 from perspective q becomes 7 in perspective r

    By recording (3,6) in our morphism, we capture this transformation history, which would be lost if we only stored the final value 7. This is crucial for scenarios where the sequence of perspectives matters and not just the final result.

    To simplify notation in cases where the complete value history isn't needed, we define a projection function π1((a,b)) = a that extracts the first component.
  4. Identity Morphisms: For each perspective pP, the identity morphism is defined by idp = f(idV, p → p), where idVV denotes the designated identity element of the value space. For the standard case where V = ℝ, this identity element is 0, making idp = f(0, p → p). For the case where V is a Boolean algebra, the identity element is true0, making idp = f(true0, p → p). This choice of identity element ensures that when we later define operations in Section 6, the identity morphism composes appropriately with other morphisms while preserving the algebraic value semantics: if f(a, p → q) represents a subjective number with value a, then f(a, p → q) ∘ idp = f(a, p → q) both as a morphism and in terms of the represented value a.

3.2. Relationship Between Categorical Construction and Relation Functions

The connection between the categorical morphisms and the relation functions is established through the following equivalence:

For perspectives p, qP and values a, bV:

a(p) =q b(t) ⟺ Rq(p, t, a, b) = true

This means that subjective equality from perspective q between subjective numbers a(p) and b(t) corresponds precisely to the relation function Rq evaluating to true when comparing these subjective numbers. This equivalence provides the crucial link between our categorical construction and the axiom system presented in Section 4, allowing us to move seamlessly between the two formulations.

3.2.1. Quotient Category Construction

In order to define the category 𝓢𝓝 properly, we introduce an equivalence relation on the morphisms to identify certain morphisms as "the same" in the quotient. Specifically, for values a, b in the underlying space V, we let f(a, p→q) and f(b, p→q) be two morphisms with the same source p and target q. We say f(a, p→q) ≡ f(b, p→q) if and only if their respective "value histories" belong to the same class under our value-history equivalence (denoted ).

Definition 3.2.1 (Value-History Equivalence Relation): We define the equivalence relation on value histories as the smallest equivalence relation satisfying the following conditions:

  1. Identity Properties:
    • For any value a ∈ V, (idV, a) ∼ a and (a, idV) ∼ a, where idV is the identity element in V.
  2. Associativity: For any values a, b, c ∈ V, ((a, b), c) ∼ (a, (b, c)).
  3. Closure Properties: The relation is closed under:
    • Reflexivity: For all value histories h, h ∼ h.
    • Symmetry: If h1 ∼ h2, then h2 ∼ h1.
    • Transitivity: If h1 ∼ h2 and h2 ∼ h3, then h1 ∼ h3.

Definition 3.2.2 (Morphism Equivalence Relation): Two morphisms

f(h1, p q) and f(h2, p q)

are declared equivalent (written f(h1, p q) f(h2, p q)) if and only if:

  1. They have the same source and target objects (i.e., the same p and q), and
  2. Their value histories h1 and h2 are equivalent under the value-history equivalence relation .

Proposition 3.2.1 (Compatibility with Composition): The morphism equivalence relation is compatible with morphism composition. That is, if f1 ≡ f2 and g1 ≡ g2, and if the compositions g1 ∘ f1 and g2 ∘ f2 are defined, then g1 ∘ f1 ≡ g2 ∘ f2.

Proof: Let f1 = f(h1, p→q), f2 = f(h2, p→q), g1 = f(k1, q→r), and g2 = f(k2, q→r).

By definition of morphism composition: g1 ∘ f1 = f(k1, q→r) ∘ f(h1, p→q) = f((h1,k1), p→r)

Similarly: g2 ∘ f2 = f(k2, q→r) ∘ f(h2, p→q) = f((h2,k2), p→r)

Given that f1 ≡ f2, we know h1 ∼ h2. Similarly, from g1 ≡ g2, we have k1 ∼ k2.

We need to show that (h1,k1) ∼ (h2,k2), which would imply g1 ∘ f1 ≡ g2 ∘ f2.

To establish this, we define the value-history equivalence relation to satisfy the following additional property:

This property ensures that respects the tuple construction used in composition. Since h1 ∼ h2 and k1 ∼ k2, by the congruence property, (h1,k1) ∼ (h2,k2).

Therefore, g1 ∘ f1 ≡ g2 ∘ f2, demonstrating that the equivalence relation is compatible with composition. □

This quotient construction yields the category 𝓢𝓝 where morphisms are equivalence classes under , ensuring that the category structure is well-defined. In particular, this construction guarantees that:

When we work with the category 𝓢𝓝, we are actually working with these equivalence classes rather than with the individual morphisms themselves. This abstraction enables us to focus on the essential structure of perspective transitions without being distracted by irrelevant distinctions in representation.

3.3. Verification of Category Axioms for 𝓢𝓝

Theorem 3.1: The structure 𝓢𝓝 defined above forms a quotient category.

Proof: We verify that 𝓢𝓝 satisfies the category axioms after appropriate quotient construction.

(1) Associativity of Composition
Let f(a, p → q): p → q, f(b, q → r): q → r, and f(c, r → s): r → s be any three morphisms in 𝓢𝓝.
We need to show that (f(c, r → s) ∘ f(b, q → r)) ∘ f(a, p → q) = f(c, r → s) ∘ (f(b, q → r) ∘ f(a, p → q)).
By the definition of composition:

Computing the left-hand side:
(f(c, r → s) ∘ f(b, q → r)) ∘ f(a, p → q) = f((b,c), q → s) ∘ f(a, p → q) = f((a,(b,c)), p → s)

And computing the right-hand side:
f(c, r → s) ∘ (f(b, q → r) ∘ f(a, p → q)) = f(c, r → s) ∘ f((a,b), p → r) = f((a,b),c), p → s)

To establish associativity, we define an equivalence relation on value histories such that: ((a,b),c) ≡ (a,(b,c)) for all a, b, c ∈ V.

This equivalence relation is:

We now verify that this equivalence relation is compatible with the category structure. For any morphisms f: A → B, g: B → C, h₁: C → D, h₂: C → D, and k: D → E where h₁ ≡ h₂, we need to prove:
k ∘ h₁ ∘ g ∘ f ≡ k ∘ h₂ ∘ g ∘ f

This follows directly from our definition of the equivalence relation. If h₁ ≡ h₂, then the resulting composition with other morphisms preserves this equivalence. This is because our equivalence relation is defined on the value histories themselves, independent of the source and target perspectives. The composition with additional morphisms merely extends the value history in a consistent way, preserving the equivalence relationship between the resulting composite morphisms.

Therefore, the equivalence relation is compatible with composition, meaning that if h₁ ≡ h₂ and g₁ ≡ g₂, then (h₁,g₁) ≡ (h₂,g₂). This ensures that composition is well-defined on equivalence classes.

Under this equivalence relation, the morphisms f((a,(b,c)), p → s) and f((a,b),c), p → s) represent the same morphism in the quotient category. This equivalence relation is structurally coherent and ensures composites are well-defined through equivalence classes of value histories.

(2) Identity Laws
For any perspective pP, the identity morphism is given by idp = f(0, p → p).
Let f(a, p → q): p → q be an arbitrary morphism. Then:
f(a, p → q) ∘ idp = f(a, p → q) ∘ f(0, p → p) = f((0,a), p → q)

Similarly, for any morphism f(b, q → p): q → p:
idp ∘ f(b, q → p) = f(0, p → p) ∘ f(b, q → p) = f((b,0), q → p)

We extend our equivalence relation to include: (0,a) ≡ a and (a,0) ≡ a for all a ∈ V. In our quotient category, this means the morphisms f((0,a), p → q) and f(a, p → q) are identified as the same morphism, as are f((b,0), q → p) and f(b, q → p). This equivalence relation preserves categorical structure while ensuring that composing with identity morphisms acts as expected, satisfying the identity laws in the quotient category. □

3.4. Perspective Distance Function

We define the perspective distance d(p, q) = ∫V × V |Rp(p, q, a, b) − Rq(q, p, b, a)| dμ(a,b). For this integral to be well-defined, we fix a σ-finite measure μ on V×V (typically a product of Lebesgue measures on a bounded domain, or any measure for which the indicator functions arising from Rp remain integrable).

In particular, we require:

Under these assumptions, Theorem 3.2 holds: d(p, q) is finite for standard examples like comparison relations or ≤-based definitions. If V is unbounded, we either (1) impose that Rp differs from Rq only on a subset of bounded measure, or (2) define dK(p, q) restricted to a compact region K and then consider a limit as K expands, whenever the limit exists. Thus the definition can be adapted to preserve well-definedness in various practical settings.

Example 3.4.1: Computing Perspective Distance

Let's calculate the perspective distance between two simple perspectives p and q in a concrete case:

Let's restrict our calculation to the interval K = [0,1] × [0,1] with Lebesgue measure λ. To compute dK(p, q), we evaluate:

dK(p, q) = ∫K |Rp(p, q, a, b) - Rq(q, p, b, a)| dλ(a,b)

For any pair (a,b), note that Rp(p, q, a, b) = 𝟙a ≤ b and Rq(q, p, b, a) = 𝟙b ≥ a. Since a ≤ b if and only if b ≥ a, we have Rp(p, q, a, b) = Rq(q, p, b, a) for all (a,b).

Therefore, dK(p, q) = ∫K |Rp(p, q, a, b) - Rq(q, p, b, a)| dλ(a,b) = ∫K 0 dλ(a,b) = 0

This shows that perspectives p and q have zero distance—they agree on all evaluations despite using different relation functions. Conversely, if we had defined Rq(q, p, b, a) = true if and only if b > a (strict inequality), then we would find dK(p, q) = λ({(a,b) ∈ K : a = b}) = 0 (since the diagonal has measure zero), also yielding zero distance.

For a non-zero example, let's instead define Rq(q, p, b, a) = true if and only if b ≥ 2a. Now Rp(p, q, a, b) ≠ Rq(q, p, b, a) when a ≤ b < 2a. The distance becomes:

dK(p, q) = λ({(a,b) ∈ [0,1]² : a ≤ b < 2a})=∫01amin(2a,1) db da

Evaluating this integral gives us dK(p, q) = 0.25, indicating a positive distance between these perspectives.

3.4.1 Measure-Theoretic Specification for Perspective Distance

We formalize our definition of the perspective distance function to ensure its well-definedness on both bounded and unbounded domains under rigorous measure-theoretic principles.

Definition 3.3 (Measure for Perspective Distance): Let V be equipped with a σ-algebra ΣV (e.g., the Borel σ-algebra if V = ). We define the product measure space (V × V, ΣV ⊗ ΣV, μ) where:

Measurability Requirement: For the perspective distance function d(p, q) = ∫V×V |Rp(p, q, a, b) − Rq(q, p, b, a)| dμ(a,b) to be well-defined, we require that for all perspectives p, q in P, the function (a,b) ↦ |Rp(p,q,a,b) − Rq(q,p,b,a)| is measurable with respect to ΣV ⊗ ΣV.

Numerical Encoding for Integration: For integration purposes in the perspective distance function, we numerically encode true as 1 and false as 0, creating characteristic functions of the form:

This encoding ensures that the sets {(a,b) ∈ V × V : Rp(p,q,a,b) = true} are measurable subsets of V×V, allowing the perspective distance function to be well-defined. Under standard assumptions (e.g., when the relation functions are defined using Borel-measurable operations like comparisons, arithmetic functions, or continuous transformations of values), the measurability condition is naturally satisfied.

Localized Perspective Distance for Unbounded Domains: For practical computation and to ensure convergence for unbounded domains, we restrict the integral to a compact set K ⊂ V × V:

dK(p, q) = ∫K |Rp(p,q,a,b) − Rq(q,p,b,a)| dμ(a,b).

As K increases to cover V×V, we can define d(p, q) = limK→V×V dK(p,q) when this limit exists. This approach ensures d(p,q) remains finite when Rp and Rq satisfy appropriate boundedness conditions.

Convergence Conditions: The perspective distance function is guaranteed to converge in the following cases:

  1. When V is a bounded domain (e.g., a closed interval [a,b] ⊂ ℝ).
  2. When the difference between relation functions vanishes outside a bounded subset of V×V.
  3. When the relation functions exhibit "locality properties" where perspectives primarily differ in their evaluations of nearby values.
Theorem 3.2 (Well-Definedness for Standard Relation Functions): Let V = with the Lebesgue measure λ. The perspective distance function d(p, q) is well-defined and finite under any of the following conditions:

(a) Comparison relations: Rp(p, q, a, b) = 𝟙a ≤ b and Rq(q, p, b, a) = 𝟙b ≤ a
(b) ε-equality relations: Rp(p, q, a, b) = 𝟙|a-b| ≤ εp and Rq(q, p, b, a) = 𝟙|b-a| ≤ εq for some εp, εq > 0
(c) Continuous relations: Rp(p, q, a, b) = σ(fp(a,b)) and Rq(q, p, b, a) = σ(fq(b,a)) where σ is the sigmoid function and fp, fq are continuous functions

Proof: For case (a), we analyze the difference function D(a, b) = |𝟙a ≤ b - 𝟙b ≤ a|. This function takes the value 1 exactly when a < b or b < a (i.e., when a ≠ b), and 0 when a = b.

The integral becomes: d(p, q) = ∫ℝ² D(a, b) dλ(a, b)

Since the diagonal set {(a, b) ∈ ℝ² : a = b} has Lebesgue measure zero, this integral equals the measure of the set where a ≠ b. For practical computation, we typically restrict to a compact subset K ⊂ ℝ², yielding: d(p, q) = λ({(a, b) ∈ K : a ≠ b}) = λ(K) - λ({(a, b) ∈ K : a = b}) = λ(K)

For example, if K = [0,1]², then d(p, q) = 1, which is finite.

For case (b), we analyze the difference function: D(a, b) = |𝟙|a-b| ≤ εp - 𝟙|b-a| ≤ εq|

Since |a-b| = |b-a|, this simplifies to: D(a, b) = |𝟙|a-b| ≤ εp - 𝟙|a-b| ≤ εq|

This function equals 1 precisely on the set where either εq < |a-b| ≤ εp or εp < |a-b| ≤ εq . Assuming without loss of generality that εp > εq, this set becomes {(a,b) : εq < |a-b| ≤ εp }. When restricted to a compact set K ⊂ ℝ², this set has finite Lebesgue measure, making the integral finite.

For case (c), the difference function is: D(a, b) = |σ(fp(a,b)) - σ(fq(b,a))|

Since the sigmoid function σ is bounded (taking values in (0,1)), we have D(a, b) ≤ 1 for all (a, b). The composition of continuous functions (fp, fq) with the continuous sigmoid function σ yields continuous functions, making D(a,b) continuous and thus measurable. When restricted to a compact set K ⊂ ℝ², the integral is bounded by the measure of K, making it finite.

In all three cases, the difference function is measurable and bounded, and when restricted to a compact domain, the integral is finite, ensuring well-definedness of the perspective distance function. □

This distance function provides a concrete mathematical tool for quantifying differences between perspectives, which is valuable for both theoretical analysis and practical applications of subjective numbers.

4. Axioms for Subjective Mathematics

We present a formal axiom system for subjective numbers. These axioms provide the rigorous foundation for subjective mathematics by specifying the essential properties and operations of subjective numbers. This axiomatic system is directly connected to the categorical construction presented in Section 3, with each axiom corresponding to specific properties of the category 𝓢𝓝.

Axiom 1 (Subjective Reflexivity): For any subjective number a(p), a(p) =p a(p).

Formal expression: ∀p ∈ P, ∀a ∈ V: Rp(p, p, a, a) = true.

Motivation: This axiom guarantees that every subjective number is self-identical from its own perspective, reflecting the identity property in 𝓢𝓝. Categorically, this corresponds to the identity morphism property in the category 𝓢𝓝.

Axiom 2 (Non-Symmetric Equality): If a(p) =p b(q), this does not necessarily imply that b(q) =q a(p).

Formal expression: ∃p,q ∈ P, ∃a,b ∈ V: Rp(p, q, a, b) ≠ Rq(q, p, b, a).

Motivation: This axiom introduces perspective-dependence by allowing the evaluation of equality to differ between perspectives. In the categorical framework, this corresponds to the potential asymmetry in morphisms between perspectives. Example 5.1 demonstrates this axiom.

Axiom 3 (Subjective Transitivity): For any perspectives ℓ, p, q, r ∈ P and any values a, b, c ∈ V, if R_ℓ(p, q, a, b) and R_ℓ(q, r, b, c) are true, then R_ℓ(p, r, a, c) must be true.

Formal expression: ∀ℓ, p, q, r ∈ P, ∀a, b, c ∈ V: (R(p, q, a, b) ∧ R(q, r, b, c)) → R(p, r, a, c).

Motivation: This axiom ensures that the evaluation within a single perspective is consistent and transitive. In category theory terms, this corresponds to the associativity of composition in the category 𝓢𝓝.

Axiom 4 (Cross-Perspective Inference): If a(p) =p b(q) and b(q) =q c(r), then a(p) =p c(r) if and only if CPAp(q, r) = true.

Formal expression: ∀p, q, r ∈ P, ∀a, b, c ∈ V: (Rp(p, q, a, b) ∧ Rq(q, r, b, c) ∧ CPAp(q, r)) ↔ Rp(p, r, a, c).

The bidirectional implication (↔) in Axiom 4 is necessary for two key reasons:

  1. The forward direction (→) ensures that when perspective p trusts q's evaluation of r, transitivity can extend across perspectives.
  2. The reverse direction (←) ensures that transitivity across perspectives must be explicitly authorized through the CPA function, preventing uncontrolled inference chains.

Without this bidirectional constraint, the framework could either allow unauthorized transitive inferences or fail to properly utilize authorized trust relationships. In the category 𝓢𝓝, this axiom governs how morphisms can be composed across different perspectives while preserving the proper evaluation semantics.

Example: Combined Incompleteness and Cross-Perspective Adoption

Let P = {α, β, γ} be three distinct perspectives with subjective numbers 7(α), 4(β), and 10(γ). Suppose:

Meanwhile, we define a cross-perspective adoption function CPAα(β, γ) = true, meaning α trusts β's evaluation about γ.

By Axiom 4 (Cross-Perspective Inference), from α's viewpoint we have: 7(α) =α 4(β) and 4(β) =β 10(γ); since α adopts β's evaluation of γ, it follows that 7(α) =α 10(γ).

However, from γ's direct perspective, 10(γ)γ 7(α). This illustrates how a certain "equality" can be derived by one perspective using cross-perspective adoption but contradicted by another perspective's direct evaluation — a hallmark of perspective incompleteness in subjective numbers.

Motivation: This axiom governs the conditions under which transitivity extends across different perspectives.

Axiom 5 (Value Consistency): If a(p) =p b(q) and a = c (in V), then c(p) =p b(q).

Formal expression: ∀p, q ∈ P, ∀a, b, c ∈ V: (Rp(p, q, a, b) ∧ (a = c)) → Rp(p, q, c, b).

Motivation: This axiom ensures that the inherent numerical equality is preserved within any subjective evaluation, reflecting the category-theoretic property that morphisms with the same numerical value are equivalent when their source and target perspectives are the same.

Axiom 6 (Perspective Adoption): A subjective number a(p) may temporarily adopt another perspective q, denoted by a(p → q), to evaluate relations as if observed from q.

Formal expression: ∀p, q ∈ P, ∀a ∈ V: Define a(p → q) such that ∀r ∈ P, ∀b ∈ V: Rp → q(p → q, r, a, b) = Rq(q, r, a, b).

Motivation: This axiom provides a mechanism for shifting perspectives, which is crucial for dynamic interactions. In the category 𝓢𝓝, this corresponds to the perspective transition notation introduced in Section 2.3.3 and used in defining morphisms.

Axiom 7 (Perspective Distinctness): For any two distinct perspectives p and q in P, there must exist at least one pair of subjective numbers a(p) and b(q) such that Rp(p, q, a, b) ≠ Rq(q, p, b, a).

Formal expression: ∀p, q ∈ P: (p ≠ q) → ∃a, b ∈ V: Rp(p, q, a, b) ≠ Rq(q, p, b, a).

Motivation: This axiom ensures that distinct perspectives yield non-trivial differences in evaluations, which corresponds to the category-theoretic property that different objects in 𝓢𝓝 exhibit distinct morphism behaviors.

4.1 Local Consistency

Theorem 4.1 (Local Consistency): Axioms 1 and 3 ensure that for any fixed perspective p ∈ P, the subjective equality relation =p forms a preorder (i.e., it is reflexive and transitive).

Proof:

  1. Reflexivity: By Axiom 1, for every a ∈ V, a(p) =p a(p) since Rp(p, p, a, a) = true.
  2. Transitivity: By Axiom 3, if a(p) =p b(q) and b(q) =p c(r), then a(p) =p c(r).

Thus, =p is a reflexive and transitive relation, making it a preorder. □

4.2 Independence of Axiom 7

Theorem 4.2 (Independence of Axiom 7): Axiom 7 is independent of Axioms 1 through 6; that is, there exist models in which Axioms 1–6 hold while Axiom 7 fails, and another model in which all seven axioms hold.

Proof: We present two models:

Lemma 4.2.1 (Model M1):

Model M1 satisfies Axioms 1–6 but fails Axiom 7.

Let P = {p, q, r} be a set of three distinct perspectives, and let the value space V = ℝ. We define relation functions explicitly as follows:

Relation function for perspective p:

Relation function for perspective q:

Relation function for perspective r:

Cross-Perspective Adoption functions:

By construction, Model M1:

This ensures that Model M1 satisfies Axioms 1–6 but not Axiom 7, demonstrating the intended independence.

Lemma 4.2.2 (Model M2):

Model M2 satisfies all seven axioms.

To demonstrate that Axiom 7 is independent rather than contradictory to Axioms 1-6, we now construct Model M2 that satisfies all seven axioms. We modify Model M1 by changing the relation functions between perspectives p and q to introduce the required perspective distinctness while maintaining all other properties of Model M1.

We keep the same setup as in Model M1 but change the relation functions between p and q as follows:

With this modification, for a = 3 and b = 2, we have:
Rp(p, q, 3, 2) = false (since 3 ≰ 2)
Rq(q, p, 2, 3) = false (since 2 ≱ 3)

This modification satisfies Axiom 7 because for any two distinct perspectives, we can find some pair of values where their evaluations differ. In particular, for perspectives p and q, the pair a = 3 and b = 2 gives:
Rp(p, q, 3, 2) = false (since 3 ≰ 2)
Rq(q, p, 2, 3) = false (since 2 ≱ 3)

Thus, Model M2 satisfies all seven axioms, completing our proof of Theorem 4.2. □

4.3 Minimal Axiomatization

Theorem 4.3 (Minimal Axiomatization): The axiom system is minimal; that is, for each axiom i (with i ∈ {1, ..., 7}), there exists a model satisfying all axioms except axiom i in which axiom i is violated.

Proof: We construct, for each axiom, an explicit model that satisfies all axioms except the one under consideration:

  1. Excluding Axiom 1 (Subjective Reflexivity): Let P = {p, q}, and define the relation functions so that for a specific a ∈ V, Rp(p, p, a, a) = false, while for all other a ∈ V and all other pairs (s, t) in P × P, define Rs(s, t, a, b) as standard equality (true if a = b, false otherwise). In this model, Axioms 2–7 hold (since the altered evaluation is confined to self-reflexivity in perspective p), but Axiom 1 is violated.
  2. Excluding Axiom 2 (Non-Symmetric Equality): Define P = {p, q}, and set for every s, t ∈ P and all a, b ∈ V,
    Rs(s, t, a, b) = true, if a = b, false, if a ≠ b.
    This definition yields a symmetric relation; hence, for any a, b, we have Rp(p,q, a, b) = Rq(q, p, b, a). This contradicts the intended non-symmetry of Axiom 2, while Axioms 1 and 3–7 remain satisfied.
  3. Excluding Axiom 3 (Subjective Transitivity): Choose P = {p, q, r} and select three values a, b, c ∈ V. Define the relation functions so that:
    • Rp(p, q, a, b) = true,
    • Rp(q, r, b, c) = true,
    • but Rp(p, r, a, c) = false. All other evaluations follow the standard equality rule. This configuration violates transitivity as required by Axiom 3, while the other axioms are maintained.
  4. Excluding Axiom 4 (Cross-Perspective Inference): Let P = {p, q, r}, and pick some a, b, c ∈ V. Define the relation functions so that:
    • Rp(p, q, a, b) = true (so a(p) =p b(q))
    • Rq(q, r, b, c) = true (so b(q) =q c(r))
    • We also set CPAp(q, r) = true (meaning perspective p trusts q for evaluating r)
    • But define Rp(p, r, a, c) = false, so from p's viewpoint, a(p) and c(r) are not equal.

    This directly violates the "if and only if" condition of Axiom 4. Despite having a(p) =p b(q), b(q) =q c(r), and CPAp(q,r) = true, we do not obtain a(p) =p c(r). All other axioms remain satisfiable in this model, so only Axiom 4 fails.

  5. Excluding Axiom 5 (Value Consistency): Let P = {p, q}. For some a, b, c ∈ V with a = c (numerical equality), define: Rp(p, q, a, b) = true and Rp(p, q, c, b) = false. This violates the requirement that numerical equality in V should ensure consistent evaluations, thereby breaking Axiom 5, while the other axioms hold under the standard assignments.
  6. Excluding Axiom 6 (Perspective Adoption): Let P = {p, q, r} and V = ℝ. Define a relation function framework similar to Model M2 above, but with a critical difference:
    • Define the relation functions Rp, Rq, and Rr as in Model M2.
    • However, for the perspective transition p → q, define Rp → q(p → q, r, a, b) ≠ Rq(q, r, a, b) for some r ∈ P and a, b ∈ V.

    Specifically, we define: Rq(q, r, 3, 4) = true (since 3 < 4) but Rp → q(p → q, r, 3, 4) = false

    The verification that this model satisfies Axioms 1-5 and 7 proceeds as follows:

    • Axiom 1 (Subjective Reflexivity): The defined relation functions satisfy reflexivity for all perspectives.
    • Axiom 2 (Non-Symmetric Equality): The model includes non-symmetric equality between perspectives (e.g., Rp(p, r, 3, 5) = true while Rr(r, p, 5, 3) = false).
    • Axiom 3 (Subjective Transitivity): The relation functions are defined to ensure transitivity within each perspective.
    • Axiom 4 (Cross-Perspective Inference): The CPA functions are defined appropriately to satisfy this axiom.
    • Axiom 5 (Value Consistency): The relation functions preserve numerical equality as required.
    • Axiom 7 (Perspective Distinctness): Distinct perspectives have different evaluation patterns as demonstrated earlier.

    This directly violates Axiom 6, which requires that a perspective transition should adopt the exact evaluation behavior of the target perspective. Therefore, we have constructed a model that satisfies Axioms 1-5 and 7 while violating Axiom 6, completing this part of the minimality proof.

  7. Excluding Axiom 7 (Perspective Distinctness): As demonstrated in Model M1 of Theorem 4.2, take P = {p, q}, and define all relation functions by standard equality. In this setting, for every a, b ∈ V, Rp(p, q, a, b) = Rq(q, p, b, a), and thus no distinctiveness in evaluation between different perspectives arises, violating Axiom 7 while satisfying Axioms 1–6.

Since for each axiom there exists a model where that axiom is violated while the remaining axioms are satisfied, the axiom system is minimal. □

Properties of Subjective Numbers

Universe 00110000

5. Properties of Subjective Numbers

From our category-theoretic construction and axiomatization of subjective numbers, several important mathematical properties emerge naturally. This section examines these properties to develop a deeper understanding of the framework.

Property 5.1 (Subjective Reflexivity): For any subjective number a(p), it holds that a(p) =p a(p). Subjective equality is reflexive from within its own perspective. This property is derived from Axiom 1.

This property derives directly from Axiom 1 and is a fundamental characteristic of the framework. It establishes that every subjective number equals itself from its own perspective, providing a consistent starting point for evaluations.

Property 5.2 (Non-Symmetric Equality): If a(p) =p b(q), it is not necessarily the case that b(q) =q a(p). Subjective equality is inherently non-symmetric. This property is derived from Axiom 2.

This property, which follows from Axiom 2, represents a foundational departure from classical mathematics. It formalizes the notion that evaluations can be direction-dependent and perspective-specific, capturing asymmetric relations that occur in many complex systems.

Example 5.1 (Non-Symmetric Equality): Consider subjective numbers 4(p) and 5(q). In a particular model, we might define relation functions such that Rp(p, q, 4, 5) = true but Rq(q, p, 5, 4) = false. This would mean 4(p) =p 5(q) (perspective p considers its 4 equal to perspective q's 5), but 5(q)q 4(p) (perspective q does not reciprocate this evaluation). This asymmetry is perfectly consistent within the framework and demonstrates Axiom 2.

Property 5.3 (Subjective Transitivity): If a(p) =r b(q) and b(q) =r c(s), then a(p) =r c(s). Subjective equality is transitive from the perspective of the evaluator (r in this case). This property is derived from Axiom 3.

This property, derived from Axiom 3, ensures that reasoning within a single perspective remains consistent. While equality across perspectives may be asymmetric, the framework maintains logical coherence within each individual perspective.

Property 5.4 (Controlled Cross-Perspective Inference): If a(p) =p b(q) and b(q) =q c(r), then a(p) =p c(r) holds if and only if CPAp(q, r) = true. Transitivity across perspectives is controlled and conditional, not automatic. This property is derived from Axiom 4.

This property, which follows directly from Axiom 4, regulates how information and evaluations propagate across different perspectives. It formalizes the idea that accepting another perspective's evaluation is an explicit choice governed by the CPA function.

Property 5.5 (Value Consistency): If a(p) =p b(q) and a = c as numerical values (in the value space V), then c(p) =p b(q). If two subjective numbers share the same numerical value when viewed from a given perspective, their subjective equality relations from that perspective will be consistent with respect to value substitution. This property is derived from Axiom 5.

Derived from Axiom 5, this property ensures that subjective evaluations respect the underlying numerical equality in the value space V. This maintains mathematical consistency while allowing for the novel features of subjective evaluation.

Theorem 5.1 (Category-Theoretic Preorder Structure) The subjective equality relation =p for any fixed perspective p forms a preorder (reflexive and transitive).

Proof:

We need to establish that for any fixed perspective p ∈ P, the relation =p is both reflexive and transitive.

Reflexivity: By Axiom 1 (Subjective Reflexivity), for any subjective number a(p), we have a(p) =p a(p), which establishes the reflexivity of =p.

Transitivity: By Axiom 3 (Subjective Transitivity), if a(s) =p b(t) and b(t) =p c(u), then a(s) =p c(u), which establishes the transitivity of =p.

In our category 𝓢𝓝, this preorder structure emerges naturally from the morphism structure. The existence of a morphism f(a, p→q) corresponds to a subjective equality a(p) =q a(q). More precisely, "a(s) =p b(t)" means Rp(s,t,a,b) = true.

The reflexivity and transitivity of =p correspond precisely to the identity morphisms and composition in the category 𝓢𝓝.

Note that while =p forms a preorder, it does not generally form a partial order since antisymmetry is not guaranteed. This aligns with our framework's design, where asymmetric evaluations between perspectives are a fundamental feature. □

6. Operations and Algebraic Structure

To make subjective numbers a functional mathematical framework, we define operational rules for combining and manipulating them, with particular attention to perspective propagation through operations. This section establishes the algebraic structure of subjective numbers and proves key properties that distinguish it from classical number systems.

6.1 Perspective Propagation in Operations

When subjective numbers are combined through operations, determining the perspective of the resulting subjective number is crucial. We define three fundamental perspective propagation rules with formal definitions and illustrated examples.

Comparison of Perspective Propagation Rules

Property Dominant
Perspective
Perspective
Fusion
Novel
Perspective
Definition a(p) +d b(q) = (a + b)(p)
(First operand's perspective)
a(p) +p⊕q b(q) = (a + b)(p⊕q)
(Composite perspective)
a(p) +n b(q) = (a + b)(r)
(Brand new perspective r)
Associativity Holds when the same first perspective dominates Depends on fusion's associative properties (e.g. conjunctive fusion is associative) Generally fails (new perspective each time)
Distributivity Fully distributive over +, × from the same perspective Only if the fusion satisfies certain conditions (e.g. logical AND/OR structures) Fails (different operations create distinct novel perspectives)
Identity / Inverses Local identities and inverses exist per perspective Possible under specific fusion rules (e.g. idempotent fusion), not always global No stable identity or inverse (perspective changes each operation)

Definition 6.1 (Dominant Perspective Rule): In this rule, the perspective of the first operand in an operation becomes the dominant perspective, determining the perspective of the result.

Formal Definition: For any binary operation φ: V × VV on values in V and subjective numbers a(s) and b(t), the dominant perspective rule yields:

a(s) +d b(t) = (a + b)(s)

where the subscript d in +d indicates the dominant perspective. The resulting subjective number inherits perspective s from the first operand.

Example 6.1: Consider two subjective numbers 3(a) and 4(b) with addition under the dominant perspective rule, denoted +d. The result is: 3(a) +d 4(b) = (3 + 4)(a) = 7(a)

The resulting subjective number reflects perspective a's view of the value 7.

Definition 6.2 (Perspective Fusion Rule): In this rule, the perspectives of both operands are combined to create a new, composite perspective for the result.

Formal Definition: For a binary operation φ: V × VV and perspective fusion, the rule is:

a(s) φs⊕t b(t) = (a φ b)(s ⊕ t)

where s ⊕ t represents a new, composite perspective with a relation function defined as:

Rs ⊕ t(s ⊕ t, r, x, y) = f(Rs(s, r, x, y), Rt(t, r, x, y))

for any perspective r and values x, y ∈ V. Here, f: {true, false} × {true, false}{true, false} is a fusion function that combines evaluations from perspectives s and t. Common choices for f include:

The choice of fusion function depends on the specific application and determines how the composite perspective evaluates relations with other perspectives.

Example 6.2: Using conjunctive fusion for addition: 3(a) +a ∧ b 4(b) = 7(a ⊕ b)

The perspective a ⊕ b only recognizes a relation as true if both original perspectives would agree, making it more restrictive than either individual perspective.

Definition 6.3 (Novel Perspective Rule): In this rule, the operation generates a completely new perspective for the result, distinct from the perspectives of either operand.

Formal Definition: For operation φ: V × VV and novel perspective generation:

a(s) φn b(t) = (a φ b)(u)

where u is a new perspective object distinct from s and t. The relation function Ru for this novel perspective is defined as:

Ru(u, r, x, y) = g(Rs(s, r, x, y), Rt(t, r, x, y), η)

where g: {true, false} × {true, false} × [0,1]{true, false} is a generation function that creates new perspective relations, and η ∈ [0,1] is a source of randomness or creativity in the generation process.

The parameter η can be systematically defined in various ways:

This allows for both deterministic and non-deterministic novel perspective generation.

Example 6.3: Using the novel perspective rule for addition: 3(a) +n 4(b) = 7(z)

The resulting perspective z might evaluate relations in ways that neither a nor b would, representing an emergent viewpoint not reducible to either original perspective.

6.2 Formal Definitions of Basic Operations

Using the perspective propagation rules established in 6.1, we formally define the basic arithmetic operations for subjective numbers.

Definition 6.4 (Subjective Addition): For subjective numbers a(s) and b(t), their sum under perspective propagation rule r ∈ {dominant, fusion, novel} is defined as:

a(s) +r b(t) = (a + b)(u)

where the perspective u of the resulting subjective number is determined by rule r applied to perspectives s and t.

Definition 6.5 (Subjective Multiplication): For subjective numbers a(s) and b(t), their product under perspective propagation rule r is:

a(s) ×r b(t) = (a × b)(u)

where again, the perspective u is determined by rule r.

Definition 6.6 (Subjective Subtraction): For subjective numbers a(s) and b(t), their difference under perspective propagation rule r is:

a(s) -r b(t) = (a - b)(u)

Definition 6.7 (Subjective Division): For subjective numbers a(s) and b(t) with b ≠ 0, their quotient under rule r is:

a(s) ÷r b(t) = (a ÷ b)(u)

Definition 6.8 (Subjective Power): For subjective numbers a(s) and b(t), the power operation under rule r is:

a(s) ^r b(t) = (a^b)(u)

where a^b represents exponentiation in the value space V, and perspective u is determined by rule r.

6.3 Algebraic Properties

The introduction of perspective into mathematical operations creates algebraic structures with properties that differ significantly from classical number systems.

6.3.1 Non-Commutativity of Subjective Operations

Theorem 6.1 (Non-Commutativity of Subjective Operations): Subjective addition (+r) and multiplication (×r) are, in general, non-commutative.

Proof: Consider subjective numbers a(p) and b(q) under the dominant perspective rule. We have:
a(p) +d b(q) = (a + b)(p), while
b(q) +d a(p) = (b + a)(q).
Since the perspectives p and q are generally distinct, these two sums represent different subjective numbers (even though a + b = b + a in V), thereby demonstrating non-commutativity.

The same argument applies to multiplication. For subjective numbers a(p) and b(q):
a(p) ×d b(q) = (a × b)(p)
b(q) ×d a(p) = (b × a)(q)

Even though a × b = b × a in V, the resulting subjective numbers differ in their perspectives. Since the perspective is an intrinsic part of the subjective number, we have:
(a × b)(p) ≠ (b × a)(q) as subjective numbers whenever p ≠ q.

This non-commutativity extends to the other perspective propagation rules as well:

For the fusion rule, a(p) +p⊕q b(q) = (a + b)(p⊕q) while b(q) +q⊕p a(p) = (b + a)(q⊕p). If the fusion operation ⊕ is not commutative (which is often the case with perspective fusion operations), then p⊕q ≠ q⊕p, making the addition non-commutative.

For the novel perspective rule, each operation generates a new perspective, so a(p) +n b(q) = (a + b)(r) while b(q) +n a(p) = (b + a)(s) with r ≠ s, again yielding different subjective numbers.

Therefore, subjective addition and multiplication are generally non-commutative across all perspective propagation rules. □

6.3.2 Perspective-Dependent Associativity

Theorem 6.2 (Perspective-Dependent Associativity): Subjective addition (+r) and multiplication (×r) are associative under specific conditions that depend on the perspective propagation rule. Full associativity holds for the dominant perspective rule when consistently applied, but may fail under other propagation rules.

Proof: We prove this by examining different perspective propagation rules for both addition and multiplication:

(1) Dominant Perspective Rule
First, let's address addition. Consider subjective addition with the dominant perspective rule and three subjective numbers a(p), b(q), c(r).

Since (a + b + c)(p) = (a + b + c)(p), we have: ((a(p) +d b(q)) +d c(r)) = (a(p) +d (b(q) +d c(r)))

Now, for multiplication with the dominant perspective rule:

Since (a × b × c)(p) = (a × b × c)(p), we have: ((a(p) ×d b(q)) ×d c(r)) = (a(p) ×d (b(q) ×d c(r)))

Therefore, subjective addition and multiplication under the dominant perspective rule are associative when the first operand's perspective consistently dominates.

(2) Perspective Fusion Rule
For the perspective fusion rule with conjunctive fusion (f(u, v) = u ∧ v), associativity depends on the associativity of the fusion operation .

For addition with perspective fusion:

For these expressions to be equal, we need (p ⊕ q) ⊕ r = p ⊕ (q ⊕ r), which is precisely the associativity property of the fusion operation ⊕.

For conjunctive fusion, where p ⊕ q is defined such that Rp⊕q(p⊕q, x, y, z) = Rp(p, x, y, z) ∧ Rq(q, x, y, z), associativity follows from the associativity of logical conjunction (∧). Therefore, addition with conjunctive perspective fusion is associative.

Similarly for multiplication with conjunctive perspective fusion:

Again, the associativity depends on whether (p ⊕ q) ⊕ r = p ⊕ (q ⊕ r), which holds for conjunctive fusion.

(3) Novel Perspective Rule
For the novel perspective rule, each operation generates a completely new perspective. With addition:

Since v and z are distinct novel perspectives with independent relation functions, in general: (a + b + c)(v) ≠ (a + b + c)(z)

Similarly for multiplication with the novel perspective rule:

Again, since v ≠ z, multiplication under the novel perspective rule is generally not associative.

This demonstrates that associativity in subjective operations is perspective-dependent, with different propagation rules yielding different associativity properties. □

6.3.3 Local Transitivity and Preorder Structure

As established in Theorem 5.1, for any fixed perspective p ∈ P, the subjective equality relation =p is reflexive (Axiom 1) and transitive (Axiom 3), thus forming a preorder. It generally fails antisymmetry, so it does not constitute a partial order.

This preorder structure provides a foundation for algebraic operations within the subjective numbers framework by ensuring that equality comparisons remain consistent from a single perspective's viewpoint, even as we allow for perspective-dependent evaluations across different perspectives.

6.3.4 Distributivity Properties

Theorem 6.3 (Distributivity Properties): Subjective multiplication distributes over subjective addition under specific conditions determined by the perspective propagation rules.

Proof: We provide a detailed analysis of distributivity for each propagation rule:

(1) Dominant Perspective Rule - Full Analysis: For subjective numbers a(p), b(q), c(r), analyzing the left side of the distributive property:
a(p) ×d (b(q) +d c(r)) = a(p) ×d (b+c)(q) = (a × (b+c))(p)

For the right side:
(a(p) ×d b(q)) +d (a(p) ×d c(r)) = (a × b)(p) +d (a × c)(p) = ((a × b) + (a × c))(p)

The key insight is that under the dominant perspective rule, the perspective of the first operand (p in this case) propagates through all operations, ensuring both sides maintain perspective p. Further, since (a × (b+c)) = ((a × b) + (a × c)) in the value space V (by the distributive property of V), and both expressions have perspective p, they are identically the same subjective number.

To verify this formally, we check if these are the same subjective number from perspective p:
(a × (b+c))(p) =p ((a × b) + (a × c))(p) ⟺ Rp(p, p, a × (b+c), (a × b) + (a × c)) = true

Since a × (b+c) = (a × b) + (a × c) in V, and by Axiom 1 (reflexivity), we have:
Rp(p, p, a × (b+c), a × (b+c)) = true

Therefore, distributivity holds under the dominant perspective rule.

(2) Perspective Fusion Rule - Detailed Analysis: For perspective fusion with the fusion operator :

Left side:
a(p) ×p⊕q (b(q) +q⊕r c(r)) = a(p) ×p⊕(q⊕r) (b+c)(q⊕r) = (a × (b+c))(p⊕(q⊕r))

Right side:
(a(p) ×p⊕q b(q)) +(p⊕q)⊕(p⊕r) (a(p) ×p⊕r c(r)) = (a × b)(p⊕q) +(p⊕q)⊕(p⊕r) (a × c)(p⊕r) = ((a × b) + (a × c))((p⊕q)⊕(p⊕r))

For these to be equal, we need:
(a × (b+c))(p⊕(q⊕r)) =? ((a × b) + (a × c))((p⊕q)⊕(p⊕r))

This requires the perspective compositions to be equivalent: p⊕(q⊕r) = (p⊕q)⊕(p⊕r). Formally, for the relation functions associated with these composite perspectives, we need:
Rp⊕(q⊕r)(p⊕(q⊕r), s, x, y) = R(p⊕q)⊕(p⊕r)((p⊕q)⊕(p⊕r), s, x, y)

For conjunctive fusion with defined using logical AND:
Rp⊕(q⊕r)(p⊕(q⊕r), s, x, y) = Rp(p, s, x, y) ∧ Rq(q, s, x, y) ∧ Rr(r, s, x, y)
R(p⊕q)⊕r((p⊕q)⊕r, s, x, y) = (Rp⊕q(p⊕q, s, x, y)) ∧ Rr(r, s, x, y) = (Rp(p, s, x, y) ∧ Rq(q, s, x, y)) ∧ Rr(r, s, x, y)

Here, we've used associativity of logical AND. Therefore, distributivity holds for subjective operations under conjunctive fusion.

However, for more complex fusion operations, this property might not hold. For example, with weighted fusion:
Rp⊕q(p⊕q, s, x, y) = 𝟙w1Rp(p, s, x, y) + w2Rq(q, s, x, y) > θ

In this case, the expressions p⊕(q⊕r) and (p⊕q)⊕(p⊕r) may yield different perspectives unless the weights satisfy specific constraints.

(3) Novel Perspective Rule - Analysis: For the novel perspective rule, each operation generates a completely new perspective with unique identity. The left and right sides of the distributive property yield:

Left side: (a × (b+c))(t) with a new perspective t
Right side: ((a × b) + (a × c))(w) with a new perspective w

Since t ≠ w by construction (novel perspectives are distinct), these subjective numbers cannot be equal from all perspectives, and distributivity fails for the novel perspective rule.

This detailed analysis demonstrates that distributivity in subjective mathematics is perspective-dependent, with different propagation rules yielding different algebraic properties. These differences highlight how subjective numbers can model phenomena where algebraic properties depend on the context or perspective of evaluation. □

In Theorem 6.3, we showed that subjective multiplication ×r distributes over subjective addition +r under perspective rules that preserve the underlying algebraic properties of V. Specifically, when perspective fusion is defined by an associative and commutative operator (such as conjunctive fusion f(u,v) = u ∧ v), the usual distributive laws hold locally within the quotient category.

However, if the fusion operator is neither associative nor commutative, or if p ⊕ q is defined by a more complex weighting or non-idempotent rule, then distributivity may fail. In such cases, we only guarantee local distributive laws for operations within a single perspective or under fusion rules that satisfy the relevant algebraic identities. Consequently, the scope of Theorem 6.3 is limited to fusion functions f that respect associativity and identity behavior. If f does not meet these properties, further analysis is required to determine which distributive or associative laws continue to hold.

6.4 Subjective Algebraic Structures

The algebraic properties of subjective numbers lead to novel mathematical structures that extend and generalize classical algebraic structures like groups, rings, and fields.

Definition 6.9 (Subjective Group): A set of subjective numbers G equipped with a binary operation r using perspective propagation rule r forms a subjective group if the following conditions are met:

  1. Closure: For all a(p), b(q) ∈ G, the result a(p)r b(q) ∈ G.
  2. Subjective Associativity: For all a(p), b(q), c(r) ∈ G, the operation r is associative when evaluated from a fixed perspective.
  3. Subjective Identity: There exists a subjective identity element e(i) ∈ G such that for all a(p) ∈ G, either a(p)r e(i) =p a(p) or e(i)r a(p) =p a(p) or both, depending on the perspective propagation rule.
  4. Subjective Inverse: For each a(p) ∈ G, there exists a subjective inverse a^{-1}(q) ∈ G such that either a(p)r a^{-1}(q) =p e(i) or a^{-1}(q)r a(p) =p e(i) or both, depending on the perspective propagation rule.

Unlike a classical group, a subjective group allows for non-commutative operations and perspective-dependent identities and inverses. The specific characteristics of a subjective group depend on the chosen perspective propagation rule.

Definition 6.10 (Subjective Ring): A set of subjective numbers R equipped with two binary operations +r and ×r using perspective propagation rule r forms a subjective ring if:

  1. (R, +r) forms a subjective abelian group (with respect to a fixed evaluation perspective).
  2. (R, ×r) forms a subjective monoid (a subjective group without the inverse property).
  3. ×r distributes over +r from the left and right, when evaluated from a fixed perspective.

Definition 6.11 (Subjective Field): A subjective ring (F, +r, ×r) forms a subjective field if:

  1. (F \ {0(p)}, ×r) forms a subjective abelian group (with respect to a fixed evaluation perspective).
  2. The distributive property holds for perspective-consistent operations.
Theorem 6.4 (Subjective Group Structure): Under the dominant perspective rule, the set of subjective numbers with the same perspective p and values from an abelian group V forms a subjective group under subjective addition.

Proof: Let Gp = {a(p) | a ∈ V} be the set of subjective numbers with perspective p and values from an abelian group V. We verify the subjective group axioms:

  1. Closure: For a(p), b(p) ∈ Gp, we have a(p) +d b(p) = (a + b)(p) ∈ Gp since a + b ∈ V.
  2. Associativity: Since (a + b) + c = a + (b + c) in V, the operation +d is associative: (a(p) +d b(p)) +d c(p) = a(p) +d (b(p) +d c(p)).
  3. Identity: The subjective number 0(p) serves as the identity, since a(p) +d 0(p) = (a + 0)(p) = a(p) and 0(p) +d a(p) = (0 + a)(p) = a(p).
  4. Inverse: For each a(p) ∈ Gp, the inverse is (-a)(p), since a(p) +d (-a)(p) = (a + (-a))(p) = 0(p) and (-a)(p) +d a(p) = ((-a) + a)(p) = 0(p).

Therefore, Gp forms a subjective group under +d. Furthermore, since the underlying value set V is an abelian group, Gp has the additional property that a(p) +d b(p) = b(p) +d a(p) for all a(p), b(p) ∈ Gp. However, this commutativity property does not extend to subjective numbers with different perspectives. □

Theorem 6.5 (Algebraic Classification): The algebraic structure formed by subjective numbers under perspective-propagating operations can be classified as follows:

1. For fixed perspective p and dominant perspective rule:
(Vp, +d) is isomorphic to (V, +) as an algebraic structure.
• The isomorphism is given by the mapping a(p) ↦ a.

2. For mixed perspectives and dominant perspective rule:
• The structure is a disjoint union of classical algebraic structures, each corresponding to a single perspective.
• The global structure forms a categorically enriched algebraic system where operations reflect the category-theoretic construction.

3. For perspective fusion rule with conjunctive fusion:
• The resulting structure creates a hierarchy of increasingly restrictive evaluation criteria.
• This forms a lattice structure on the set of perspectives, where the meet operation corresponds to perspective fusion.

4. For novel perspective rule:
• The resulting structure forms a free algebra over the generating set of perspectives.
• Each operation creates new perspectives that are not reducible to the original perspectives.

This classification contextualizes subjective numbers within the broader landscape of algebraic structures, showing how they both extend and transform classical algebraic systems. □

6.5 Consistency Relative to ZFC

Theorem 6.6 (Consistency Relative to ZFC): The axiom system for subjective numbers is consistent relative to ZFC.

Proof: We construct a complete model 𝓜 of subjective numbers within ZFC as follows:

  1. Model Definition: Let P = {p1, p2, p3} be three perspectives, and let V = be the value space. Define the set of subjective numbers as SN = V × P, where each subjective number is represented as an ordered pair (a, p) with a ∈ V and p ∈ P.
  2. Relation Function Specification: For each perspective piP, we define:

    For perspective p1:

    • Rp1(p1, p1, a, a) = true for all a ∈ ℝ
    • Rp1(p1, p1, a, b) = false for all a ≠ b
    • Rp1(p1, p2, a, b) = (a = b) for all a, b ∈ ℝ
    • Rp1(p1, p3, a, b) = (a < b) for all a, b ∈ ℝ

    For perspective p2:

    • Rp2(p2, p2, a, a) = true for all a ∈ ℝ
    • Rp2(p2, p2, a, b) = false for all a ≠ b
    • Rp2(p2, p1, a, b) = (a ≥ b) for all a, b ∈ ℝ
    • Rp2(p2, p3, a, b) = (a = 2b) for all a, b ∈ ℝ

    For perspective p3:

    • Rp3(p3, p3, a, a) = true for all a ∈ ℝ
    • Rp3(p3, p3, a, b) = false for all a ≠ b
    • Rp3(p3, p1, a, b) = (a = b+1) for all a, b ∈ ℝ
    • Rp3(p3, p2, a, b) = (a = b/2) for all a, b ∈ ℝ

    Cross-Perspective Adoption functions:

    • CPAp1(p2, p3) = true
    • CPAp2(p3, p1) = true
    • CPAp3(p1, p2) = true
    • All other CPA values are false
  3. Verification of Axioms:
    • Axiom 1 (Subjective Reflexivity): We defined Rpi(pi, pi, a, a) = true for all pi ∈ P and a ∈ ℝ.
    • Axiom 2 (Non-Symmetric Equality): For a = 3 and b = 5, we have Rp1(p1, p3, 3, 5) = true but Rp3(p3, p1, 5, 3) = false.
    • Axiom 3 (Subjective Transitivity): The relation functions use standard numerical relations (=, <, ≥, etc.) that are transitive by construction.
    • Axiom 4 (Cross-Perspective Inference): The CPA functions are defined to make this axiom hold.
    • Axiom 5 (Value Consistency): The relation functions are constructed to preserve numerical equality.
    • Axiom 6 (Perspective Adoption): The model supports perspective transitions.
    • Axiom 7 (Perspective Distinctness): By construction, for any two perspectives, there exists a pair of values with asymmetric evaluations.
  4. Categorical Construction Verification: The model instantiates the category 𝓢𝓝 as defined in Section 3.1, with:
    • Objects: The perspectives P = {p1, p2, p3}.
    • Morphisms: For each pair of perspectives and value, morphisms are defined as required.
    • Composition: The composition follows the rules established in Section 3.1 and in 3.2.1
    • Identity Morphisms: For each perspective, identity morphisms use the zero value as defined.
  5. Absence of Contradiction: The model satisfies all axioms of subjective numbers within ZFC. To address how value history equivalence relations maintain consistency within ZFC, we note that:
    • The equivalence relation on value histories ((a,b) ≡ a for identity morphisms, ((a,b),c) ≡ (a,(b,c)) for associativity) is formally definable in ZFC as it operates on ordered pairs and tuples.
    • This equivalence relation is compatible with composition in 𝓢𝓝 since: (1) it preserves source and target objects, (2) it ensures composition is well-defined on equivalence classes, and (3) it respects identity morphisms.
    • The well-definedness of this equivalence relation ensures that our quotient category construction remains within the expressive power of ZFC, avoiding potential inconsistencies.

Since our model is rigorously constructible within ZFC using standard set-theoretic constructions, and since all operations on perspectives, morphisms, and relation functions can be formalized using standard mathematical machinery available in ZFC, the axiom system for subjective numbers must be consistent relative to ZFC.

Theorem 6.7 (Non-Triviality): The subjective numbers framework is non-trivial; it permits models with genuinely perspective-dependent behavior that cannot be reduced to classical mathematics.

Proof: We have already demonstrated in Theorem 4.2 the existence of Model M2, which exhibits perspective-dependent behavior not reducible to classical equality. Specifically, in this model, we have Rp(p, q, a, b) ≠ Rq(q, p, b, a) for certain values a, b ∈ V, meaning that from perspective p, the subjective number a(p) may equal b(q), but from perspective q, this equality does not hold.

This asymmetric evaluation cannot be modeled in classical mathematics, where equality is necessarily symmetric. Since our framework permits such models while remaining consistent, it is both non-trivial and expressively more powerful than classical mathematics for representing perspective-dependent phenomena. □

Perspective Relations and Equivalence Classes

Universe 00110000

7. Perspective Relations and Equivalence Classes

7.1 Subjective Equivalence Relations

A fundamental concept in subjective mathematics is equivalence from a specific perspective. We define subjective equivalence relations as follows:

Definition 7.1 (Subjective Equivalence Relation): A subjective equivalence relation =s from perspective s is characterized by:

  1. Reflexivity: a(s) =s a(s) for all a(s) (by Axiom 1)
  2. s-Transitivity: If a(p) =s b(q) and b(q) =s c(r), then a(p) =s c(r) (by Axiom 3)
  3. Potential Non-Symmetry: a(p) =s b(q) does not necessarily imply b(q) =s a(p) (by Axiom 2)

This differs from traditional equivalence relations by not requiring symmetry, making it a reflexive and transitive relation (a preorder) but not necessarily an equivalence relation in the standard sense.

Example 7.1 (Expert Evaluation): Three experts—Alice, Bob, and Carol—each have their own perspective on evaluating research papers. Alice (perspective a) might consider Bob's evaluation (perspective b) equivalent to her own when assessing paper P, written as P(a) =a P(b). However, Bob might not reciprocate this assessment, so P(b)b P(a). This asymmetry reflects the real-world scenario where experts may not mutually recognize each other's evaluations as equivalent, and demonstrates Axiom 2.

7.2 Subjective Equivalence Classes

From perspective s, we define a subjective equivalence class [a(p)]s as the set of all subjective numbers that are considered equal to a(p) from perspective s:

Definition 7.2 (Subjective Equivalence Class): [a(p)]s = {b(q) | a(p) =s b(q)}

These subjective equivalence classes possess properties distinct from classical equivalence classes:

Theorem 7.1 (Class Intersection): For distinct perspectives s and t, the intersection [a(s)]s ∩ [a(t)]t may be non-empty, empty, or may contain a(s) but not a(t) (or vice versa), depending on the relation functions Rs and Rt.

Proof: By Axiom 7 (Perspective Distinctness), distinct perspectives must differ in at least one relational evaluation. Consider the three possible cases for the intersection of subjective equivalence classes:

  1. Case 1 - Non-empty intersection with mutual recognition: If Rs(s, t, a, a) = true and Rt(t, s, a, a) = true, then a(s) ∈ [a(t)]t and a(t) ∈ [a(s)]s, making the intersection contain at least {a(s), a(t)}.
  2. Case 2 - Empty intersection: If Rs(s, t, a, a) = false and Rt(t, s, a, a) = false, then a(s) ∉ [a(t)]t and a(t) ∉ [a(s)]s. Further, it is possible to define the relation functions such that no subjective number appears in both equivalence classes, making the intersection empty.
  3. Case 3 - Asymmetric recognition: If Rs(s, t, a, a) = true but Rt(t, s, a, a) = false, then a(t) ∈ [a(s)]s but a(s) ∉ [a(t)]t, creating asymmetric class membership.

Thus, the structure of subjective equivalence classes can vary significantly depending on the relational evaluations between perspectives. □

Theorem 7.2 (Non-Partition Property): Subjective equivalence classes from a given perspective do not necessarily partition the set of all subjective numbers.

Proof: Classical equivalence classes form a partition because: (1) every element belongs to some class, (2) no element belongs to two different classes, and (3) no class is empty.

For subjective equivalence classes, condition (2) can be violated. Consider subjective numbers 5(p), 7(q), and 10(r), and define relation functions such that:

Note that this relation function definition for perspective s maintains internal transitivity as required by Axiom 3, since transitivity applies only when all evaluations are from the same fixed perspective. In this case, Rs defines a transitive relation when considering all evaluations from perspective s.

From these relations, 7(q) ∈ [5(p)]s and 7(q) ∈ [10(r)]s, so 7(q) belongs to both the equivalence class of 5(p) and the equivalence class of 10(r). However, the equivalence classes [5(p)]s and [10(r)]s are distinct because 5(p)s 10(r). This violates the partition property that each element belongs to exactly one equivalence class.

The key insight is that while transitivity holds for all evaluations from a single perspective (as ensured by Axiom 3), the resulting equivalence classes need not form a partition of the set of subjective numbers. Therefore, subjective equivalence classes from perspective s do not necessarily partition the set of all subjective numbers. □

Example 7.2 (Explicit Equivalence Classes): Consider perspectives P = {a, b, c} with relation functions defined for value 5 such that:

The resulting equivalence classes are:

Note the cyclical pattern of recognition, where each perspective recognizes equality with one other perspective but not with the third. This example demonstrates Theorem 7.2.

This non-partition property is a fundamental difference between subjective mathematics and classical mathematics, reflecting how perspective-dependent evaluation creates overlapping rather than cleanly separated categories.

8. Connections to Established Mathematics

8.1 Relation to Modal and Multi-valued Logics

Subjective mathematics exhibits profound connections to modal logic, multi-valued logic, and non-classical logical systems designed to handle context-dependent truth, modality, and non-standard truth values. These connections provide valuable insights into the nature of subjective mathematics while highlighting its unique contributions.

Modal Logic Correspondence

Modal logic extends classical logic by introducing operators that qualify statements according to different "modes" of truth—notably possibility (◇) and necessity (□). This aligns with subjective mathematics' treatment of perspective-dependent truth.

Theorem 8.1 (Preservation of Logical Consistency via Modal Logic Translation): There exists a truth-preserving translation Φ from subjective mathematical statements (within a single perspective) to statements in modal logic, such that logical consistency within each perspective is preserved under this translation.

Proof: We define a translation function Φ that maps statements in subjective mathematics to statements in modal propositional logic. For subjective mathematical statements evaluated from perspective s:

Now, we must address how cross-perspective adoption (CPA) is handled in this translation. For statements involving cross-perspective inference via Axiom 4, we introduce the additional translation:

Apply the translation Φ to each component of this inference:

In modal logic, the inference rule corresponding to Subjective Transitivity becomes:

s(a = b) ∧ □s(b = c) s(a = c)

This inference rule, s(a = b) ∧ □s(b = c) → □s(a = c), is a valid inference in standard modal logic systems, particularly in systems that include the K axiom (Distribution Axiom) and assume the underlying equality relation in the possible worlds is transitive. The validity in modal logic arises from the properties of the necessity operator s and the inherent transitivity of equality within each possible world (perspective s).

Since any valid inference within subjective mathematics from a single perspective s translates to a valid inference in modal logic under Φ, the translation Φ preserves logical consistency within each perspective. This means that if a set of statements is consistent within subjective mathematics from perspective s, their modal logic translations will also be consistent in modal logic within the context of modality s. This does not imply global consistency across all perspectives, but ensures that reasoning within each perspective remains logically sound. □

Theorem 8.2 (Modal Logic Incompleteness for Cross-Perspective Inference): Standard modal logic systems (such as K, T, S4, and S5) are insufficient to fully capture cross-perspective inference in subjective mathematics.

Proof: Consider the cross-perspective inference axiom (Axiom 4), which states: ∀p, q, r ∈ P, ∀a, b, c ∈ V: (Rp(p, q, a, b) ∧ Rq(q, r, b, c) ∧ CPAp(q, r)) ↔ Rp(p, r, a, c)

In subjective mathematics, the truth of a(p) =p c(r) from premises a(p) =p b(q) and b(q) =q c(r) explicitly depends on the cross-perspective adoption function CPAp(q, r).

Under the translation Φ, we would have:

Standard modal logic provides no direct mechanism to conditionally control inferential transitions between different modal operators p and q based on a function like CPAp(q, r). The inference: p(a = b) ∧ □q(b = c) → □p(a = c) is not generally valid in any standard modal logic system (K, T, S4, or S5) without additional machinery.

To fully capture cross-perspective inference, we would need to extend modal logic with:

  1. A formal representation of the CPA function
  2. Modified inference rules that explicitly incorporate this function
  3. A semantics that allows controlled transitivity across modalities

Since standard modal logic lacks these features, it cannot completely capture the cross-perspective inference capabilities of subjective mathematics. □

8.2 Beyond Modal Logic: Unique Aspects of Subjective Mathematics

While the modal logic translation demonstrates a deep connection, subjective mathematics extends beyond what modal logic typically addresses in several key ways:

  1. Intrinsic Perspective: In subjective mathematics, perspective is an intrinsic property of mathematical objects themselves, not just an operator applied to propositions. This differs from modal logic where modalities are typically external operators.
  2. Non-Symmetric Relations: Unlike the accessibility relations in Kripke frames, which can be symmetric or asymmetric depending on axioms chosen, subjective mathematics inherently supports non-symmetric equality as a foundational property.
  3. Object-Level vs. Proposition-Level: Modal logic operates primarily at the level of propositions, whereas subjective mathematics builds perspective into the very nature of mathematical objects.
  4. Algebraic Structure: Subjective mathematics inherits and extends algebraic properties through operations like addition and multiplication, which don't have direct parallels in standard modal logic.
Theorem 8.3 (Embedding of Multi-Valued Logic): Any finitely-valued logic with n truth values can be embedded within the subjective numbers framework using n perspectives.

Proof: Let L = {v1, v2, ..., vn} be the set of truth values in a multi-valued logic with valuation function V mapping propositions to elements of L.

We construct an embedding as follows:

  1. Define a set of n perspectives P = {p1, p2, ..., pn}, one corresponding to each truth value.
  2. For each proposition φ with valuation V(φ) = vi, represent it as a subjective number 1(pi).
  3. Define relation functions such that: Rpi(pi, pj, 1, 1) = f(vi, vj) where f is a function that captures the relationship between truth values in the multi-valued logic.

For example, in Łukasiewicz's three-valued logic with values {0, 1/2, 1}, we would define three perspectives {p0, p1/2, p1} and encode the logic's conjunction operation as subjective equality relations across these perspectives.

For a concrete implementation, let's encode Łukasiewicz's three-valued conjunction (∧) as follows:

0 1/2 1
0 0 0 0
1/2 0 1/2 1/2
1 0 1/2 1

We can define Rpi such that for propositions φ and ψ with values V(φ) = vi and V(ψ) = vj, the conjunction φ ∧ ψ corresponds to subjective equality evaluations between the perspectives pi and pj.

This construction demonstrates that subjective numbers can express the same distinctions as multi-valued logic, but with the added capability of embedding these distinctions directly into mathematical objects rather than just truth values of propositions. □

Theorem 8.4 (Strengthened Expressivity): The subjective numbers framework strictly extends the expressivity of both modal logic and multi-valued logic for representing perspective-dependent mathematical phenomena.

Proof: We have established in Theorem 8.1 that subjective mathematics can express modal logic reasoning within a single perspective, and in Theorem 8.3 that it can encode multi-valued logics.

To show strict extension, we need to demonstrate capabilities unique to subjective numbers. Consider the following:

  1. Perspective-dependent operations: In subjective mathematics, we can define operations where both the operands and the result have perspectives: a(p) +r b(q) = (a + b)(u) Neither modal logic nor multi-valued logic has direct mechanisms for operations on perspectival entities where the perspective propagates through the operation.
  2. Cross-perspective controlled inference: The CPA function in Axiom 4 provides fine-grained control over when inference can cross perspective boundaries, which neither modal logic nor multi-valued logic directly support.
  3. Equivalence class structure: As shown in Theorem 7.2, subjective equivalence classes can form non-partition structures that have no direct analogue in classical many-valued logics.

Since subjective mathematics can express everything in modal and multi-valued logic while also providing capabilities beyond them, it strictly extends their expressivity. □

Conclusion Visualization

Universe 00110000

9. Selective Fusion of Perspectives

In earlier sections, we introduced the notion of perspective fusion, denoted p ⊕ q, which formally merges the evaluations of two perspectives into a single, composite perspective. While fusion is a powerful way to combine knowledge or viewpoints, it can also lead to paradox if performed indiscriminately—especially when self-referential statements are involved. In this section, we show how restricting fusion to “compatible” perspectives (i.e., those that do not generate contradictions) preserves consistency, thereby avoiding liar-like paradoxes, yet still allows beneficial “wisdom of crowds” effects when contradictions do not arise.

9.1 Contradiction-Free (Selective) Fusion

To formalize the restriction, we define a compatibility condition that detects whether two (or more) perspectives can be safely fused without producing inconsistency or self-referential paradox.

Theorem 9.1 (Selective Fusion Consistency): Let p, q be two perspectives whose evaluation functions are Rp and Rq, respectively. Suppose a compatibility test Compatible(p, q) returns true if and only if merging p and q into a single vantage p ⊕ q does not introduce contradiction or self‐reference loops. Then:
  1. If Compatible(p, q) = true, define p ⊕ q (e.g., via conjunctive or disjunctive fusion). This new perspective’s relation function, Rp⊕q, is constructed by applying a fusion operator f to the individual evaluations, so that for all x, y ∈ V:
    Rp⊕q(p ⊕ q, x, y) = f(Rp(p, q, x, y), Rq(q, p, y, x))
    Since Compatible(p, q) = true ensures Rp(p, q, x, y) = Rq(q, p, y, x) for all x, y, the fused evaluation is well‐defined and free of internal contradiction.
  2. If Compatible(p, q) = false, then there exists at least one pair x, y ∈ V for which the evaluations conflict. Any attempt to define Rp⊕q in this case would yield a contradiction or induce a self‐reference loop (e.g., a sentence that simultaneously forces “I am true” and “I am false”). Therefore, the system prohibits forming p ⊕ q when Compatible(p, q) = false.

Proof: Assume that for perspectives p and q the compatibility test yields Compatible(p, q) = true. By definition, for every pair of values x, y ∈ V the relation functions satisfy Rp(p, q, x, y) = Rq(q, p, y, x). When constructing the fused perspective p ⊕ q, we define its relation function as

Rp⊕q(p ⊕ q, x, y) = f(Rp(p, q, x, y), Rq(q, p, y, x))

where f is a fusion operator (for example, logical conjunction) that outputs a consistent truth value when provided with equal inputs. Because the compatibility condition guarantees identical inputs from Rp and Rq, the fused evaluation is unambiguous and the new perspective p ⊕ q remains consistent.

Conversely, if Compatible(p, q) = false, then there exists some x, y ∈ V for which Rp(p, q, x, y) ≠ Rq(q, p, y, x). In such cases, any attempt to define a fused relation Rp⊕q via the operator f would produce conflicting outputs—yielding a self‐reference loop or direct contradiction (for example, generating a scenario where a statement is forced to be simultaneously true and false). Thus, the system disallows the formation of p ⊕ q under these conditions.

Therefore, selective fusion is only permitted when it does not introduce paradoxes, preserving local consistency within every fused perspective. □

Example 9.1: Perspective-Dependent Operations and the Compatibility Test

Consider two perspectives, P and Q, over the value space V = ℝ. Let us define two subjective numbers: 3(P) and 4(Q).

Under the Dominant Perspective Rule, if perspective P is taken as dominant, the addition is defined as:

3(P) + 4(Q) = 7(P)

Alternatively, if we wish to fuse the perspectives using a Fusion Rule, we first perform the compatibility test. For instance, if both perspectives determine equality solely by numerical identity, then for every x, y ∈ ℝ we have:

RP(P, Q, x, y) = RQ(Q, P, y, x)

Hence, Compatible(P, Q) = true, and we can form the fused perspective P ⊕ Q. The addition then becomes:

3(P) + 4(Q) = 7(P ⊕ Q)

The result carries the fused perspective, representing a coherent combination of the individual viewpoints.

9.1.1 Category-Theoretic Foundation of Selective Fusion

To provide a rigorous foundation for selective fusion within our category-theoretic framework, we now formalize the compatibility test as a functorial construction in 𝓢𝓝.

Definition 9.1.1 (Compatibility Functor): Let Comp: 𝓢𝓝 × 𝓢𝓝 → Set be a functor that maps each pair of perspectives (p, q) to the set:

Comp(p, q) = {(x, y) ∈ V × V | Rp(p, q, x, y) = Rq(q, p, y, x)}

This functor identifies all pairs of values (x, y) for which perspectives p and q agree on their mutual evaluations.

Definition 9.1.2 (Categorical Compatibility Test): Two perspectives p, q ∈ P are compatible, denoted Compatible(p, q) = true, if and only if:

∀x, y ∈ V: (x, y) ∈ Comp(p, q)

In other words, perspectives are compatible when they agree on all mutual evaluations.

Definition 9.1.3 (Categorical Selective Fusion): When Compatible(p, q) = true, the fused perspective p ⊕ q is constructed as a pullback in 𝓢𝓝, with canonical projections:

πp: p ⊕ q → p and πq: p ⊕ q → q

The relation function for this fused perspective is defined as:

Rp⊕q(p ⊕ q, r, x, y) = f(Rp(p, r, x, y), Rq(q, r, x, y))

where f: {true, false} × {true, false} → {true, false} is a fusion function that combines the evaluations from the constituent perspectives.

Theorem 9.1.1 (Functorial Properties of Selective Fusion): Selective fusion, when defined on compatible perspectives, gives rise to a partial functor ⊕: 𝓢𝓝_c × 𝓢𝓝_c → 𝓢𝓝, where 𝓢𝓝_c is the subcategory of 𝓢𝓝 restricted to compatible perspectives.

Proof: To establish that selective fusion forms a partial functor, we need to verify that:

  1. It maps objects (compatible perspectives) to objects (fused perspectives).
  2. It maps morphisms (between compatible perspectives) to morphisms (between fused perspectives).
  3. It preserves identity morphisms and composition.

For (1), we've defined how selective fusion maps compatible perspectives p and q to the fused perspective p ⊕ q.

For (2), given morphisms f(a, p → p') and f(b, q → q') where both Compatible(p, q) = true and Compatible(p', q') = true, the fusion functor induces a morphism f((a,b), p⊕q → p'⊕q').

For (3), the fusion functor maps identity morphisms idp and idq to idp⊕q, and preserves composition through the universal property of pullbacks.

When perspectives are not compatible, the fusion is undefined, making this a partial functor rather than a total one. This mathematical structure formalizes the selective nature of fusion—it exists only when it can be defined consistently. □

Corollary 9.1.1 (Paradox Avoidance): The category-theoretic construction of selective fusion ensures that no contradiction or paradox can arise from fusing compatible perspectives.

Proof: By Definition 9.1.2, perspectives p and q are compatible only when Rp(p, q, x, y) = Rq(q, p, y, x) for all values x, y. This ensures that no conflicting evaluations exist.

When constructing the fused perspective p ⊕ q as a pullback, the universal property ensures that any relation function defined on p ⊕ q is consistent with the original relation functions Rp and Rq.

Since the original perspectives have no conflicting evaluations (by the compatibility test), and the fusion preserves their consistent evaluations, the fused perspective cannot contain any contradiction or paradox. □

This category-theoretic formalization provides a rigorous foundation for selective fusion, showing how it integrates with the broader framework of subjective numbers and establishing its consistency-preserving properties. The partial functorial structure captures precisely the essential property that fusion should only be allowed when it preserves mathematical coherence.

Example 9.2: Cyclic Resolution of Self-Reference in the Liar Paradox

Consider the liar statement L that asserts “This statement is false.” Let denote the intrinsic perspective of L and let E denote an external evaluative perspective.

The evaluation of L is modeled by two morphisms:

The composite morphism is:

f¬ ∘ ftruth: ℓ → ℓ

Suppose initially the intrinsic evaluation in is true. Then:

  1. ftruth maps true in to true in E.
  2. f¬ maps true in E to false in .

Repeating the cycle with the new value:

  1. ftruth maps false in to false in E.
  2. f¬ maps false in E to true in .

This establishes a 2-cycle in which the truth value of L alternates between true and false with each complete cycle. At no point does any single perspective (either or E) register both true and false simultaneously. Consequently, local classical consistency is maintained while the cyclic behavior prevents a fixed-point paradox.

10. Conclusion and Future Directions

We have presented subjective numbers as a systematic way to embed perspectives into the very structure of mathematical objects. By formalizing perspective as an intrinsic property of each number, rather than an external parameter, we capture asymmetries and directional evaluations that classic frameworks overlook. This approach clarifies how truth and equality can depend on the specific vantage point from which they are assessed—extending or modifying standard algebraic structures without sacrificing local consistency or coherence.

Through a precise category-theoretic construction, we demonstrated that subjective numbers satisfy an axiomatic system (Axioms 1–7) ensuring internal consistency, path-dependent evaluations, and a rich algebraic structure. We showed how these constructions remain compatible with familiar Boolean operations and preserve classical logic within each individual perspective, with non-classical effects emerging only when multiple perspectives interact. Our measure-theoretic extension further quantifies disagreement between perspectives, illustrating how subtle perspective shifts can be captured in a rigorous manner.

10.1 Summary of Key Contributions

10.2 Philosophical and Mathematical Outlook

The framework invites new ways to model settings where knowledge, assessment, or meaning depends crucially on the observer’s standpoint—aligning well with domains like cognitive science, social choice theory, and logic of agency. Rather than treating perspective as a nuisance, our construction shows it can be a first-class concept in mathematics. By connecting category theory and perspective-laden valuations, we gain a flexible means to study asymmetry and context-dependence in myriad applications.

10.3 Future Directions

There remain significant frontiers for subjective numbers:

Overall, this foundation lays out a coherent vision in which mathematics of perspective evolves past strict symmetry, enabling further research to bring subjective numbers into broader logical, computational, and philosophical discussions.

Appendix: Mathematical Examples for Subjective Numbers

This appendix provides detailed examples that illustrate key concepts of the subjective numbers framework, helping to build intuition for this novel mathematical structure.

Example A1: Illustrative Model of Subjective Numbers with Explicit Relation Functions

Let us construct a concrete model of subjective numbers with three perspectives P = {p, q, r} and values in . We explicitly define relation functions that demonstrate key properties of the framework.

Perspectives:

Relation Functions: For perspective p, define Rp as follows:

For perspective q, define Rq as follows:

For perspective r, define Rr as follows:

Verification of Axioms: We can verify that this model satisfies all axioms:

  1. Subjective Reflexivity: Each perspective satisfies Rs(s, s, a, a) = true, demonstrating Axiom 1.
  2. Non-Symmetric Equality: Consider Rp(p, r, 3, 5) = true since 3 ≤ 5, but Rr(r, p, 5, 3) = false since 5 ≠ 3 + 1, demonstrating Axiom 2.
  3. Subjective Transitivity: For each perspective, if a(s) =p b(t) and b(t) =p c(u), then a(s) =p c(u), demonstrating Axiom 3.

Specific Evaluations: From perspective p:

From perspective q:

From perspective r:

This model demonstrates perspective-dependent evaluation, non-symmetric relations, and distinctive relation patterns for each perspective, validating the subjective numbers framework.

Example A2: Perspective Propagation in Operations

Let us examine perspective propagation through operations in detail using the three fundamental rules.

Consider the perspectives P = {a, b, c} and subjective numbers 3(a), 4(b), and 5(c). We define relation functions such that:

Dominant Perspective Rule: When performing addition with the dominant perspective rule:

  1. 3(a) +d 4(b) = 7(a)
  2. 4(b) +d 3(a) = 7(b)
  3. 5(c) +d 3(a) = 8(c)

For subjective numbers 7(a) and 7(b), consider their relationships:

Despite having the same numerical value, 7(a) and 7(b) are distinct subjective numbers with different perspectives.

Perspective Fusion Rule: For perspective fusion, we define the composite perspective relation function Rab such that:

Now, when performing addition with perspective fusion:

  1. 3(a) +a∨b 4(b) = 7(a⊕b)
  2. 4(b) +b∨c 5(c) = 9(b⊕c)

For evaluating relations:

Novel Perspective Rule: For the novel perspective rule, we generate entirely new perspectives with distinct relation functions:

  1. 3(a) +n 4(b) = 7(z) where z is a new perspective
  2. 4(b) +n 5(c) = 9(w) where w is a new perspective

For the novel perspective z created from a and b, we define the relation function Rz to include elements from both source perspectives:

This example illustrates how each perspective propagation rule creates distinct patterns of relations in the resulting subjective numbers, demonstrating the flexibility of the framework for modeling diverse perspective interactions.

Example A3: Path-Dependent Truth

This example demonstrates how truth in subjective mathematics can be path-dependent, where different evaluation paths yield different valid conclusions.

Consider perspectives P = {α, β, γ} with the following specific relation functions:

  1. For perspective α:
    • Reflexivity: Rα(α, α, a, a) = true for all a ∈ ℝ (reflexivity)
    • Cross-Perspective: Rα(α, β, 5, 5) = true (perspective α considers its 5 equal to perspective β's 5)
    • Rα(α, α, 5, 3) = false (perspective α considers 5 and 3 different in direct evaluation)
    • Rα(α, γ, a, b) is undefined for all a, b (no direct relation with γ)
  2. For perspective β:
    • Reflexivity: Rβ(β, β, a, a) = true for all a ∈ ℝ (reflexivity)
    • Cross-Perspective: Rβ(β, γ, 5, 5) = true (perspective β considers its 5 equal to perspective γ's 5)
    • Rβ(β, α, a, b) is undefined for all a, b (no direct relation with α)
  3. For perspective γ:
    • Reflexivity: Rγ(γ, γ, a, a) = true for all a ∈ ℝ (reflexivity)
    • Cross-Perspective: Rγ(γ, α, 5, 3) = true (perspective γ considers its 5 equal to perspective α's 3)
    • Rγ(γ, β, a, b) is undefined for all a, b (no direct relation with β)

We also define the cross-perspective adoption relation:

Now, let's analyze the subjective equality relation between 5(α) and 3(α) from perspective α using different evaluation paths:

Direct Evaluation from Perspective α: By direct evaluation from perspective α's relation function, we have: Rα(α, α, 5, 3) = false Therefore, from perspective α's direct viewpoint, 5(α)α 3(α).

Evaluation via Perspective Chain α → β → γ → α:

  1. Start with 5(α)
  2. From perspective α: 5(α) =α 5(β) (since Rα(α, β, 5, 5) = true)
  3. Perspective α adopts perspective β's evaluation of γ: 5(β) =α 5(γ) (since Rβ(β, γ, 5, 5) = true and CPAα(β, γ) = true)
  4. Perspective α adopts perspective γ's evaluation of α: 5(γ) =α 3(α) (since Rγ(γ, α, 5, 3) = true and CPAα(γ, α) = true)
  5. By transitivity (within perspective α's adopted viewpoints): 5(α) =α 3(α)

Thus, we have derived 5(α) =α 3(α) by following the perspective chain α → β → γ → α, even though direct evaluation gives 5(α)α 3(α). This example demonstrates path-dependent truth.

Example A4: Perspective Incompleteness

This example demonstrates the Perspective Incompleteness phenomenon with explicit relation functions showing that certain mathematical truths are inaccessible from any single perspective but become available through perspective fusion.

Consider a system with three perspectives P = {α, β, γ} focused on the subjective numbers 7(α), 4(β), and 10(γ).

Define relation functions with the following specific knowledge distribution:

  1. For perspective α:
    • Rα(α, α, 7, 7) = true (reflexivity)
    • Rα(α, β, 7, 4) = true (perspective α recognizes equivalence with β's 4)
    • Rα(α, γ, 7, 10) is undefined (perspective α has no knowledge about relation to γ's 10)
  2. For perspective β:
    • Rβ(β, β, 4, 4) = true (reflexivity)
    • Rβ(β, γ, 4, 10) = true (perspective β recognizes equivalence with γ's 10)
    • Rβ(β, α, 4, 7) is undefined (perspective β has no knowledge about relation to α's 7)
  3. For perspective γ:
    • Rγ(γ, γ, 10, 10) = true (reflexivity)
    • Rγ(γ, α, 10, 7) = false (perspective γ explicitly rejects equivalence with α's 7)
    • Rγ(γ, β, 10, 4) is undefined (perspective γ has no knowledge about relation to β's 4)

Now, we analyze what mathematical truths are accessible from each individual perspective:

From Perspective α:

From Perspective β:

From Perspective γ:

Now, create fused perspective α⊕β with relation function Rα⊕β defined as:

From this fused perspective α⊕β:

This demonstrates perspective incompleteness: the relation between 7(α) and 10(γ) is inaccessible from any single perspective (α, β, or γ), yet becomes accessible through perspective fusion (α⊕β). In fact, perspective γ explicitly contradicts this derived relation (10(γ)γ 7(α)), showing how the fusion of perspectives can reveal mathematical truths that are not just unknown but potentially contradicted by individual perspectives. This example demonstrates Perspective Incompleteness.

This realization 7(α) =α⊕β 10(γ) represents an emergent mathematical truth that exists only at the collective level, not derivable from any individual perspective in isolation.

Example A5: Constructing a Penrose Stairs Analogue with Subjective Numbers

To further solidify the intuition behind subjective numbers and their capacity to model paradoxical, viewpoint-dependent structures, we present a concrete construction that mirrors the famous Penrose stairs illusion. This example demonstrates how subjective numbers can create a mathematical analogue of an impossible object, exhibiting local consistency while forming a globally paradoxical loop due to perspective-dependent relations.

Construction of the Penrose Stairs Analogue:

  1. Perspectives as Steps: We define a set of four perspectives, P = {p1, p2, p3, p4}, each representing a "step" in our subjective staircase.
  2. Subjective Numbers as Levels: We consider subjective numbers representing different "levels" of ascent, using integer values for simplicity:
    • level0(p1): Subjective number with value 0 and perspective p1.
    • level1(p2): Subjective number with value 1 and perspective p2.
    • level2(p3): Subjective number with value 2 and perspective p3.
    • level3(p4): Subjective number with value 3 and perspective p4.
  3. Asymmetric Subjective Equality Relations for "Ascent": We define the relation functions Rp1, Rp2, Rp3, and Rp4 to establish an asymmetric subjective equality that mimics an ascending staircase:
    • From Perspective p1: "Level 0" is subjectively equal to "Level 1" (from perspective p2): Rp1(p1, p2, 0, 1) = true. In other words, level0(p1) =p1 level1(p2).
    • From Perspective p2: "Level 1" is subjectively equal to "Level 2" (from perspective p3): Rp2(p2, p3, 1, 2) = true. That is, level1(p2) =p2 level2(p3).
    • From Perspective p3: "Level 2" is subjectively equal to "Level 3" (from perspective p4): Rp3(p3, p4, 2, 3) = true. Meaning, level2(p3) =p3 level3(p4).
    • Closing the Loop: From Perspective p4: "Level 3" is subjectively equal to "Level 0" (from perspective p1): Rp4(p4, p1, 3, 0) = true. Concretely, level3(p4) =p4 level0(p1).
    • Non-Equality for Direct Comparison: To maintain the illusion of distinct levels from each perspective and to ensure asymmetry, we stipulate that direct self-comparisons of different levels within the same perspective are not subjectively equal. Formally: Rpi(pi, pi, n, m) = false if n ≠ m, and Rpi(pi, pj, n, m) ≠ Rpj(pj, pi, m, n) for i ≠ j, n ≠ m. Reflexivity still holds as Rpi(pi, pi, n, n) = true.
  4. The Penrose Stairs Effect: Path-Dependent "Ascent" and Paradoxical Loop: Starting from level0(p1), let us trace the subjective equality relations, simulating a "walk" on our subjective Penrose stairs:
    • Step 1 (Perspective p1): We have level0(p1) =p1 level1(p2) . We effectively "ascend" to level1(p2) .
    • Step 2 (Perspective p2): We have level1(p2) =p2 level2(p3) . We further "ascend" to level2(p3) .
    • Step 3 (Perspective p3): We have level2(p3) =p3 level3(p4) . We continue to "ascend" to level3(p4) .
    • Step 4 (Perspective p4 - Closing the Loop): We have level3(p4) =p4 level0(p1) . Paradoxically, we "ascend" back to something subjectively equal to our starting point, level0(p1) , completing a loop of continuous ascent.
    This cyclical chain of subjective equalities, level0(p1) → level1(p2) → level2(p3) → level3(p4) → level0(p1) , mirrors the impossible loop of the Penrose stairs. Each step appears to ascend locally (from one perspective to the next), yet globally, the chain returns to its starting level, creating a paradoxical closed loop of continuous ascent. This example demonstrates path-dependent truth.

This construction can be visualized as a directed graph where each node represents a subjective number (level0(p1), level1(p2), etc.), and a directed edge from node A to node B exists if A =p B from the perspective p associated with node A.