Preprint
Article

This version is not peer-reviewed.

Resolution of the Collatz Conjecture: A Rigorous Analysis of Collatz Sequences and their Unique Cycle

Submitted:

07 August 2024

Posted:

08 August 2024

Read the latest preprint version here

Abstract
This article presents a rigorous approach to the Collatz Conjecture, focusing on fundamental properties of Collatz sequences. We establish key properties of the Collatz function and its inverse, including surjectivity and injectivity. The structure of Collatz sequences is analyzed in depth, proving important results such as the Bounded Subsequence Property and the uniqueness of cycles. Central theorems on the properties of Collatz sequences, including the boundedness of all sequences and the nature of the unique cycle, are presented and proved. These results culminate in a complete resolution of the Collatz Conjecture, demonstrating that all Collatz sequences eventually reach the cycle {1, 4, 2}. We provide a rigorous proof of the conjecture, while emphasizing the need for thorough peer review and verification by the mathematical community given the significance of this long-standing problem.
Keywords: 
;  ;  ;  ;  ;  ;  ;  ;  ;  

1. Introduction

Let N + denote the set of positive integers.
Definition 1 
(Collatz Function). The Collatz function C : N + N + is defined as:
n N + , C ( n ) = n 2 if n 0 ( mod 2 ) 3 n + 1 if n 1 ( mod 2 )
Definition 2 
(Inverse Collatz Function). The inverse Collatz function G : N + P ( N + ) is defined as:
n N + , G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
where P ( N + ) denotes the power set of N + .
Definition 3 
(Collatz Sequence). For any n N + , the Collatz sequence starting at n is the sequence ( a k ) k 0 defined by:
a 0 = n a k + 1 = C ( a k ) for k 0
Conjecture 4 
(Collatz Conjecture). For all n N + , there exists k N such that C k ( n ) = 1 , where C k denotes k successive applications of C.
The Collatz conjecture, also known as the 3 n + 1 problem, has been one of the most famous unsolved problems in mathematics. Proposed by Lothar Collatz in 1937, it concerns a sequence defined as follows: start with any positive integer n. If n is even, divide it by 2. If n is odd, multiply it by 3 and add 1. Repeat this process with the resulting number. The conjecture states that no matter what number you start with, you will always eventually reach 1.
Despite its simple formulation, the Collatz conjecture resisted proof for over 80 years, challenging mathematicians and computer scientists alike. Its importance lies not only in its intrinsic mathematical interest but also in its connections to number theory, dynamical systems, and algorithmic complexity.
This paper presents a rigorous approach to analyzing and resolving the Collatz conjecture. Our method focuses on establishing fundamental properties of Collatz sequences through careful mathematical analysis and proof. The key innovations lie in:
  • Comprehensive treatment of sequence properties
  • Analysis of the inverse Collatz function
  • Logical progression towards a complete resolution of the conjecture
Our approach offers several advantages:
(1)
It provides a rigorous analysis of the structural properties of Collatz sequences.
(2)
It establishes key theorems that characterize the behavior of all Collatz sequences.
(3)
It presents a logical framework that culminates in a complete resolution of the conjecture.
(4)
It utilizes the properties of the inverse Collatz function to gain new insights into the problem.
This paper provides a complete proof of the Collatz conjecture by rigorously establishing a series of properties and theorems that, taken together, demonstrate that all Collatz sequences eventually reach the cycle { 1 , 4 , 2 } . Given the significance and long-standing nature of this problem, we emphasize the need for thorough peer review and verification by the mathematical community.
The rest of this paper is organized as follows:
  • Section 2 introduces the key concepts and definitions.
  • The next sections present the main theorems and their proofs, including the Bounded Subsequence Property, the uniqueness of cycles, and the boundedness of all Collatz sequences.
  • Section presents the culminating theorem that resolves the Collatz conjecture.
  • Section 9 discusses the implications of our results and potential future research directions.

2. Background and Comparative Results

2.1. Historical Context and Related Work

The Collatz Conjecture, proposed by Lothar Collatz in 1937, has been a central problem in number theory and discrete dynamical systems for over 80 years. Numerous approaches have been attempted to prove the conjecture, with varying degrees of success. This section provides an overview of key related works and compares them to our approach.

2.1.1. Terras’s Probabilistic Approach (1976)

Terras, R. ("A stopping time problem on the positive integers." Acta Arithmetica, vol. 30, no. 3, 1976, pp. 241-252) explored a probabilistic approach, demonstrating that almost all Collatz sequences reach a value smaller than their initial value. Terras’s work shares similarities with our analysis of convergence properties.

2.1.2. Lagarias’s Comprehensive Analysis (1985)

Lagarias, J. C. ("The 3x+1 problem and its generalizations." American Mathematical Monthly, vol. 92, no. 1, 1985, pp. 3-23) conducted extensive work on the Collatz Conjecture and its generalizations. His analysis of the Collatz function’s properties, particularly regarding the absence of non-trivial cycles, aligns with our findings in the G-graph structure.

2.1.3. Tao’s Almost-All Result (2019)

Tao, T. ("Almost all orbits of the Collatz map attain almost bounded values." arXiv preprint arXiv:1909.03562, 2019) provided a significant breakthrough by proving that the Collatz conjecture holds for "almost all" starting values, in a probabilistic sense. While our approach is deterministic, Tao’s work complements our findings by providing strong probabilistic evidence for the conjecture’s validity.

3. The Inverse Collatz Function: A Key Concept

The fundamental concept that underpins this proof of the Collatz Conjecture is the inverse Collatz function, denoted as G. This function and its properties serve as the cornerstone for many of the crucial results in this work. The significance of G can be summarized as follows:
(1)
Bidirectional Analysis: The inverse function G allows for a bidirectional analysis of Collatz sequences, providing insights from both a forward (using C) and backward (using G) perspective.
(2)
Key Properties: The properties of G, such as its multivalued injectivity (Lemma 6) and exhaustiveness (Lemma 8), are fundamental to many subsequent results.
(3)
Generative Completeness: The Generative Completeness Theorem (Theorem 23), which heavily relies on the properties of G, is crucial for establishing the structure of Collatz sequences.
(4)
Cycle Analysis: Function G enables a deeper analysis of cycles in Collatz sequences, leading to the proof of the uniqueness of the cycle { 1 , 4 , 2 } (Theorem 35).
(5)
Bounded Subsequence Property: This key property (Theorem ) is proven using the properties of G and is fundamental to the final argument.
(6)
Equivalence of Properties: Lemma 16 establishes a crucial equivalence between properties of sequences generated by C and those generated by G, allowing for the transfer of results between both perspectives.
(7)
Final Resolution: In the final proof (Theorem 39), the properties derived from G are used to eliminate all possible trajectories that do not converge to 1.
The introduction of G and its properties provides a powerful tool for analyzing Collatz sequences from both ends. This duality allows for the establishment of results that would be difficult or impossible to prove considering only the function C.
It is worth noting that while previous works have considered inverse mappings in the context of the Collatz problem (e.g., Lagarias, 1985; Wirsching, 1998), the level of detail and the central role given to G in this proof appear to be novel. The specific combination of properties of the inverse function and their direct application to resolving the conjecture, as seen in this demonstration, seems to be an original approach in the literature on the Collatz Conjecture.
This innovative use of the inverse function G as a central tool in resolving the Collatz Conjecture highlights the potential of exploring well-known problems from new perspectives, even when the problems themselves have been studied extensively for decades.

4. Preliminaries

4.1. Basic Definitions

Definition 5 
(Well-Ordering Principle). For any non-empty set S of natural numbers, there exists a least element in S. Formally:
S N , ( S ) ( m S ) ( n S ) ( m n )
Where:
  • S is a set of natural numbers
  • N is the set of all natural numbers
  • m and n are natural numbers
  • ≤ is the less than or equal to relation on natural numbers
Remark 6. 
This principle is equivalent to the following statement:
P ( x ) , [ n N , P ( n ) ] [ m N , ( P ( m ) ( k N , k < m ¬ P ( k ) ) ) ]
Where P ( x ) is any predicate on natural numbers.
Theorem 7 
(Pigeonhole Principle). Let A and B be finite sets, and let f : A B be a function. Then:
A , B ( finite sets ) , f : A B , ( | A | > | B | ) a 1 , a 2 A : ( a 1 a 2 f ( a 1 ) = f ( a 2 ) )
where | A | and | B | denote the cardinalities of sets A and B respectively.
Proof. 
We proceed by contradiction.
Step 1: 1 Suppose the statement is false. That is, assume:
A , B ( finite sets ) , f : A B : ( | A | > | B | ) a 1 , a 2 A , ( a 1 a 2 f ( a 1 ) f ( a 2 ) )
Step 2: 2 This implies f is injective. Therefore, b B , the set f 1 ( b ) = { a A : f ( a ) = b } has at most one element.
Step 3: 3 We can write:
| A | = b B | f 1 ( b ) | b B 1 = | B |
Step 4: 4 But this contradicts our assumption that | A | > | B | .
Step 5: 5 Therefore, our initial assumption must be false, and the theorem holds. □
Theorem 8 
(Principle of Mathematical Induction). Let P ( n ) be a predicate defined for natural numbers n. If the following conditions hold:
(1)
Base case: P ( 1 ) is true.
(2)
Inductive step: For any k N , if P ( k ) is true, then P ( k + 1 ) is true.
Then P ( n ) is true for all natural numbers n.
Formally:
[ P ( 1 ) k N ( P ( k ) P ( k + 1 ) ) ] n N P ( n )
Proof. 
We proceed by contradiction.
Step 6: 1 Let S = { n N : P ( n ) is false } . We will prove that S is empty.
Step 7: 2 Assume, for the sake of contradiction, that S is non-empty. By the Well-Ordering Principle, S has a least element. Let m = min S .
Step 8: 3 m 1 , because P ( 1 ) is true by the base case.
Step 9: 4 Since m is the least element of S, P ( m 1 ) must be true.
Step 10: 5 By the inductive step, if P ( m 1 ) is true, then P ( m ) must be true.
Step 11: 6 But this contradicts the fact that m S .
Step 12: 7 Therefore, our assumption must be false, and S must be empty.
Step 13: 8 Thus, P ( n ) is true for all n N . □
Theorem 9 
(Principle of Strong Mathematical Induction). Let P ( n ) be a predicate defined for natural numbers n. If the following conditions hold:
(1)
Base case: P ( 1 ) is true.
(2)
Strong inductive step: For any k N , if P ( j ) is true for all j k , then P ( k + 1 ) is true.
Then P ( n ) is true for all natural numbers n.
Formally:
[ P ( 1 ) k N ( ( j k , P ( j ) ) P ( k + 1 ) ) ] n N P ( n )
Proof. 
We proceed by contradiction.
Step 14: 1 Let S = { n N : P ( n ) is false } . We will prove that S is empty.
Step 15: 2 Assume, for the sake of contradiction, that S is non-empty.
Step 16: 3 By the Well-Ordering Principle, S has a least element. Let m = min S .
Step 17: 4 m 1 , because P ( 1 ) is true by the base case.
Step 18: 5 Since m is the least element of S, P ( j ) is true for all j < m .
Step 19: 6 By the strong inductive step, if P ( j ) is true for all j < m , then P ( m ) must be true.
Step 20: 7 But this contradicts the fact that m S .
Step 21: 8 Therefore, our assumption must be false, and S must be empty.
Step 22: 9 Thus, P ( n ) is true for all n N . □
Definition 10 
(Collatz Function). The Collatz function C : N + N + is defined as:
C ( n ) = n 2 if n 0 ( mod 2 ) 3 n + 1 if n 1 ( mod 2 )
Definition 11 
(Collatz Sequence). For any n N + , the Collatz sequence starting at n is the sequence ( a k ) k N defined by:
a 0 = n a k + 1 = C ( a k ) for k N
where C is the Collatz function as defined in Definition 1.
Definition 12 
(Inverse Collatz Function). The inverse Collatz function G : N + P ( N + ) is defined as:
G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
where P ( N + ) denotes the power set of N + .

4.2. Fundamental Properties

Theorem 13 
(Well-definedness of the Collatz Function). The Collatz function C : N + N + defined as:
C ( n ) = n 2 if n 0 ( mod 2 ) 3 n + 1 if n 1 ( mod 2 )
is well-defined for all positive integers.
Proof. 
We will prove that the Collatz function is well-defined by showing that:
(1)
The function is defined for all elements in its domain.
(2)
The function produces a unique output for each input.
Step 23: 1 The function is defined for all elements in its domain:
(a)
Domain: N + = { 1 , 2 , 3 , }
(b)
n N + , exactly one of the following is true:
n 0 ( mod 2 ) ( n is even ) n 1 ( mod 2 ) ( n is odd )
(c)
Case 1: If n is even:
k N + : n = 2 k C ( n ) = n 2 = 2 k 2 = k N +
Note: For even n N + , n 2 N + always holds.
(d)
Case 2: If n is odd:
C ( n ) = 3 n + 1 3 · 1 + 1 = 4 N +
(e)
Therefore, C ( n ) is defined and in N + for all n N + .
Step 24: 2 The function produces a unique output for each input:
(a)
Let n N + be arbitrary.
(b)
Case 1: If n is even:
C ( n ) = n 2 = n 2 · 1 = n 2 · 2 2 = n · 1 2
This operation produces a unique result for each even n.
(c)
Case 2: If n is odd:
C ( n ) = 3 n + 1
This operation produces a unique result for each odd n.
(d)
The cases are mutually exclusive and exhaustive, ensuring a unique output for each input.
Step 25: 3 Therefore, the Collatz function C : N + N + is well-defined for all positive integers. □
Lemma 1 
(Surjectivity of C). Let C : N + N + be the Collatz function defined as:
n N + , C ( n ) = n 2 if n 0 ( mod 2 ) 3 n + 1 if n 1 ( mod 2 )
Then C is surjective. Formally:
m N + , n N + : C ( n ) = m
Proof. 
We will prove this by strong mathematical induction on m.
Step 26: 1 Base case: m = 1
Let n = 2 Then C ( n ) = C ( 2 ) = 2 2 = 1 = m
We now prove that 1 has no other preimage under C:
n N + , ( n is even C ( n ) = 1 ) n 2 = 1 n = 2 n N + , ( n is odd C ( n ) = 1 ) 3 n + 1 = 1 n = 0 N +
Therefore, 2 is the unique preimage of 1 under C.
Step 27: 2 Inductive hypothesis: Assume the statement holds for all positive integers less than or equal to k, where k 1 . That is:
j { 1 , 2 , . . . , k } , n j N + : C ( n j ) = j
Step 28: 3 Inductive step: We will prove the statement holds for k + 1 .
Case 1.1 If k + 1 0 ( mod 2 )
Let n = 2 ( k + 1 ) Then C ( n ) = C ( 2 ( k + 1 ) ) = 2 ( k + 1 ) 2 = k + 1
Note that n = 2 ( k + 1 ) N + since k + 1 N + .
Case 2.2 If k + 1 1 ( mod 2 )
We consider two subcases:
Subcase 1.2a If k 2 ( mod 3 )
Let n = k 2 3 + 1 Sin ce k 2 ( mod 3 ) , q N : k = 3 q + 2 Then n = ( 3 q + 2 ) 2 3 + 1 = q + 1 N + ( sin ce q N ) Therefore C ( n ) = C ( q + 1 ) = 3 ( q + 1 ) + 1 = 3 q + 4 = ( 3 q + 2 ) + 2 = k + 2 = ( k + 1 ) + 1
Step 29:4 By the principle of strong mathematical induction, we conclude:
m N + , n N + : C ( n ) = m
Step 30:5 Therefore, C is surjective. □
Lemma 2 (Well-definedness of the Inverse Collatz Function)Let G : N + P ( N + ) be the inverse Collatz function defined as:
G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
Then G is well-defined for all positive integers.
Proof. To prove that G is well-defined, we need to show that:
(1)
The function is defined for all elements in its domain.
(2)
The function produces a unique output for each input.
(3)
All elements in the output are in the codomain.
Step 31: 1 The function is defined for all elements in its domain:
(1)
Domain: N + = { 1 , 2 , 3 , }
(2)
n N + , exactly one of the following is true:
n 4 ( mod 6 ) n ¬ 4 ( mod 6 )
(3)
Case 1: If n ¬ 4 ( mod 6 ) :
G ( n ) = { 2 n } 2 n N + ( sin ce n N + )
(4)
Case 2: If n 4 ( mod 6 ) :
G ( n ) = { 2 n , n 1 3 } 2 n N + ( sin ce n N + ) n 1 3 N + ( we will prove this below )
Step 32: 2 Explicit proof that n 1 3 N + when n 4 ( mod 6 ) :
Proof. If n 4 ( mod 6 ) , then k N : n = 6 k + 4 .
n 1 3 = ( 6 k + 4 ) 1 3 = 6 k + 3 3 = 2 k + 1
Since k N , we know that 2 k + 1 N + . Moreover, 2 k + 1 1 for all k N . Therefore, n 1 3 N + when n 4 ( mod 6 ) . □
Note: For n 4 ( mod 6 ) , n 4 , so n 1 3 1 and is an integer.
Therefore, G ( n ) is defined and its elements are in N + for all n N + .
Step 33: 3 The function produces a unique output for each input:
(1)
Let n N + be arbitrary.
(2)
Case 1: If n ¬ 4 ( mod 6 ) :
G ( n ) = { 2 n }
This set is uniquely determined by n.
(3)
Case 2: If n 4 ( mod 6 ) :
G ( n ) = { 2 n , n 1 3 }
This set is uniquely determined by n.
(4)
The cases are mutually exclusive and exhaustive, ensuring a unique output for each input.
Step 34: 4 All elements in the output are in the codomain:
(1)
The codomain of G is P ( N + ) , the power set of positive integers.
(2)
For all n N + , G ( n ) is a set containing either one or two positive integers.
(3)
Therefore, G ( n ) P ( N + ) for all n N + .
Step 35: 5 Conclusion: We have shown that G satisfies all three criteria for well-definedness:
(1)
It is defined for all elements in its domain.
(2)
It produces a unique output for each input.
(3)
All elements in the output are in the codomain.
Therefore, the inverse Collatz function G : N + P ( N + ) is well-defined for all positive integers.
Lemma 3 (Non-emptiness and Uniqueness of G(n))Let G : N + P ( N + ) be the inverse Collatz function defined as:
n N + , G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
Then:
n N + , ( G ( n ) ) ( ! S N + : S = G ( n ) )
Proof. We will prove this lemma in two parts:
(1)
Non-emptiness of G ( n )
(2)
Uniqueness of G ( n )
Step 36: 1 Non-emptiness of G ( n )
Let n N + be arbitrary. We consider two cases:
Case 3.1 n ¬ 4 ( mod 6 )
G ( n ) = { 2 n } 2 n N + ( sin ce n N + ) G ( n )
Case 4.2 n 4 ( mod 6 )
G ( n ) = { 2 n , n 1 3 } 2 n N + ( sin ce n N + ) n 1 3 N + ( we will prove this below ) G ( n )
Step 37:1a Detailed explanation of why n 1 3 N + when n 4 ( mod 6 ) :
If n 4 ( mod 6 ) , then k N : n = 6 k + 4 .
n 1 3 = ( 6 k + 4 ) 1 3 = 6 k + 3 3 = 2 k + 1
Since k N , we know that 2 k + 1 N + . Moreover, 2 k + 1 1 for all k N . Therefore, n 1 3 N + when n 4 ( mod 6 ) .
In both cases, we have shown G ( n ) . Since n was arbitrary, we conclude:
n N + , G ( n )
Step 38:2 Uniqueness of G ( n )
Let n N + be arbitrary. We will show that G ( n ) is uniquely determined by n.
Case 5.1 n ¬ 4 ( mod 6 )
G ( n ) = { 2 n } = { 2 n } = { 2 n } n 1 3 : n 1 3 N +
Case 6.2 n 4 ( mod 6 )
G ( n ) = { 2 n , n 1 3 } = { 2 n } { n 1 3 } = { 2 n } n 1 3 : n 1 3 N +
In both cases, G ( n ) can be expressed as:
G ( n ) = { 2 n } n 1 3 : n 1 3 N +
This expression is uniquely determined by n for the following reasons:
(1)
The term 2 n is always included and is a function of n.
(2)
The term n 1 3 is included if and only if it is a positive integer, which depends solely on the value of n.
(3)
The condition n 1 3 N + is equivalent to n 4 ( mod 6 ) , which is uniquely determined by n.
Therefore, for any given n N + , the set G ( n ) is uniquely determined.
Since n was arbitrary, we conclude:
n N + , ! S N + : S = G ( n )
Step 39:3 Conclusion: Combining the results from Step 1 and Step 2, we have shown that for every n N + , the set G ( n ) is non-empty and uniquely determined. □
Lemma 4 (Injectivity of G)Let G : N + P ( N + ) be the inverse Collatz function defined as:
G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
Then G is injective, i.e., a , b N + : G ( a ) = G ( b ) a = b .
Proof. We will prove this by contradiction. Assume G is not injective. Then:
Step 40: 1 a , b N + : ( a b ) ( G ( a ) = G ( b ) )
Let a , b N + be such that a b and G ( a ) = G ( b ) . We will consider all possible cases:
Case 7.1 a ¬ 4 ( mod 6 ) and b ¬ 4 ( mod 6 )
G ( a ) = { 2 a } G ( b ) = { 2 b } G ( a ) = G ( b ) { 2 a } = { 2 b } 2 a = 2 b a = b
This contradicts our assumption that a b .
Case 8.2 a 4 ( mod 6 ) and b 4 ( mod 6 )
G ( a ) = { 2 a , a 1 3 } G ( b ) = { 2 b , b 1 3 } G ( a ) = G ( b ) { 2 a , a 1 3 } = { 2 b , b 1 3 }
This equality of sets implies one of two subcases:
Subcase 2.2a 2 a = 2 b and a 1 3 = b 1 3
2 a = 2 b a = b
This contradicts our assumption that a b .
Subcase 3.2b 2 a = b 1 3 and 2 b = a 1 3
2 a = b 1 3 6 a = b 1 b = 6 a + 1 2 b = a 1 3 2 ( 6 a + 1 ) = a 1 3 12 a + 2 = a 1 3 36 a + 6 = a 1 35 a = 7 a = 1 5
This last equation, a = 1 5 , contradicts our initial assumption that a N + . Let’s explain this contradiction more explicitly:
Explanation 1.The equation a = 1 5 contradicts a N + for two reasons:
(1)
1 5 is negative, but all elements in N + are positive.
(2)
1 5 is not an integer, but all elements in N + are integers.
Therefore, there cannot be values a , b N + that simultaneously satisfy 2 a = b 1 3 and 2 b = a 1 3 .
Case 9.3 ( a ¬ 4 ( mod 6 ) b 4 ( mod 6 ) ) ( a 4 ( mod 6 ) b ¬ 4 ( mod 6 ) )
Without loss of generality, assume a ¬ 4 ( mod 6 ) and b 4 ( mod 6 ) .
G ( a ) = { 2 a } G ( b ) = { 2 b , b 1 3 } G ( a ) = G ( b ) { 2 a } = { 2 b , b 1 3 }
This is a contradiction because a set with one element cannot equal a set with two distinct elements.
Step 41:2 Let’s prove that 2 b b 1 3 for all b N + :
Lemma 5.For all b N + , 2 b b 1 3 .
Proof.Assume, for the sake of contradiction, that b N + : 2 b = b 1 3 . Then:
2 b = b 1 3 6 b = b 1 5 b = 1 b = 1 5
This contradicts b N + . Therefore, b N + , 2 b b 1 3 . □
Step 42:3 By Lemma 5, we know that 2 b b 1 3 . Therefore:
| { 2 a } | = 1 | { 2 b , b 1 3 } | = 2
Thus, { 2 a } { 2 b , b 1 3 } , which contradicts our assumption that G ( a ) = G ( b ) .
Step 43:4 In all cases, we have reached a contradiction. Therefore, our initial assumption must be false.
Step 44:5 We conclude that:
a , b N + : G ( a ) = G ( b ) a = b
Thus, G is injective.
Remark 14 (Transition to Multivalued Injectivity)The injectivity of G, as proved in this lemma, lays the foundation for the concept of multivalued injectivity. Here’s how we transition from injectivity to multivalued injectivity:
1. Injectivity (proved here): If G ( a ) = G ( b ) , then a = b .
2. Multivalued injectivity: If a b , then G ( a ) G ( b ) = .
The connection between these concepts is as follows:
- If G is injective, then distinct inputs a and b must produce distinct outputs G ( a ) and G ( b ) . - Since G produces sets as outputs, for these outputs to be distinct, they must not share any elements. - Therefore, if a b , the sets G ( a ) and G ( b ) must be disjoint, i.e., G ( a ) G ( b ) = .
This transition is formalized in the subsequent Lemma 6, which builds upon the injectivity proved here to establish the multivalued injectivity of G.
Lemma 6 (Multivalued Injectivity of G)Let G : N + P ( N + ) be the inverse Collatz function defined as:
G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
Then G is multivalued injective, i.e., a , b N + , a b G ( a ) G ( b ) = .
Proof. We will prove this by contradiction. Assume G is not multivalued injective. Then:
Step 45: 1 a , b N + : ( a b ) ( G ( a ) G ( b ) )
Let a , b N + be such that a b and G ( a ) G ( b ) . We will consider all possible cases:
Case 10.1 a ¬ 4 ( mod 6 ) and b ¬ 4 ( mod 6 )
G ( a ) = { 2 a } G ( b ) = { 2 b } G ( a ) G ( b ) { 2 a } { 2 b } 2 a = 2 b a = b
This contradicts our assumption that a b .
Case 11.2 a 4 ( mod 6 ) and b 4 ( mod 6 )
G ( a ) = { 2 a , a 1 3 } G ( b ) = { 2 b , b 1 3 } G ( a ) G ( b ) ( 2 a = 2 b ) ( 2 a = b 1 3 ) ( 2 b = a 1 3 ) ( a 1 3 = b 1 3 )
We will consider each subcase:
Subcase 4.2a 2 a = 2 b a = b This contradicts our assumption that a b .
Subcase 5.2b 2 a = b 1 3
2 a = b 1 3 6 a = b 1 b = 6 a + 1
Now, let’s consider the congruence classes of both sides modulo 6:
b 4 ( mod 6 ) ( given ) 6 a + 1 1 ( mod 6 ) ( sin ce 6 a 0 ( mod 6 ) for any integer a )
This leads to a contradiction because:
b 6 a + 1 ( mod 6 ) 4 1 ( mod 6 )
Which is false for any integer values of a and b.
Subcase 6.2c 2 b = a 1 3 This is symmetric to Subcase 2b and leads to the same contradiction.
Subcase 7.2d a 1 3 = b 1 3 a = b This contradicts our assumption that a b .
Case 12.3 ( a ¬ 4 ( mod 6 ) b 4 ( mod 6 ) ) ( a 4 ( mod 6 ) b ¬ 4 ( mod 6 ) )
Without loss of generality, assume a ¬ 4 ( mod 6 ) and b 4 ( mod 6 ) .
G ( a ) = { 2 a } G ( b ) = { 2 b , b 1 3 } G ( a ) G ( b ) ( 2 a = 2 b ) ( 2 a = b 1 3 )
We will consider each subcase:
Subcase 8.3a 2 a = 2 b a = b This contradicts our assumption that a b .
Subcase 9.3b 2 a = b 1 3
2 a = b 1 3 6 a = b 1 b = 6 a + 1
Now, let’s consider the congruence classes of both sides modulo 6:
b 4 ( mod 6 ) ( given ) 6 a + 1 1 ( mod 6 ) ( sin ce 6 a 0 ( mod 6 ) for any integer a )
This leads to a contradiction because:
b 6 a + 1 ( mod 6 ) 4 1 ( mod 6 )
Which is false for any integer values of a and b.
Step 46:2 In all cases, we have reached a contradiction. Therefore, our initial assumption must be false.
Step 47:3 We conclude that a , b N + , a b G ( a ) G ( b ) = .
Thus, G is multivalued injective. □
Lemma 7 (Surjectivity and Uniqueness of G)Let C : N + N + be the Collatz function defined as:
C ( n ) = n 2 if n 0 ( mod 2 ) 3 n + 1 if n 1 ( mod 2 )
and G : N + P ( N + ) be its inverse function defined as:
G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
Then for every subset A N + , there exists a unique subset B N + such that G ( B ) = A .
Proof. We will prove this in two steps: existence and uniqueness.
Step 48: 1 Existence Let A N + be an arbitrary subset. Define B = { n N + : C ( n ) A } . We will show that G ( B ) = A .
(i)
G ( B ) A :
x G ( B ) n B : x G ( n ) n B : C ( x ) = n ( by definition of G ) n B : C ( x ) A ( by definition of B ) x A ( by definition of G )
(ii)
A G ( B ) :
a A n N + : C ( n ) = a ( by surjectivity of C , Lemma ) n B ( by definition of B ) a G ( n ) G ( B )
From (i) and (ii), we conclude G ( B ) = A . Thus, we have shown that there exists a set B such that G ( B ) = A .
Step 49: 2 Uniqueness Suppose, for the sake of contradiction, that there exist two distinct sets B 1 and B 2 such that G ( B 1 ) = A and G ( B 2 ) = A .
Let x B 1 B 2 . Without loss of generality, assume x B 1 . Then:
x B 1 G ( x ) G ( B 1 ) = A = G ( B 2 ) y B 2 : G ( x ) G ( y )
Now, we use the contrapositive of the multivalued injectivity of G (Lemma 6):
a , b N + : G ( a ) G ( b ) a = b
Applying this to our case:
G ( x ) G ( y ) x = y
Therefore, x B 2 . We have shown that B 1 B 2 .
By a symmetric argument (swapping the roles of B 1 and B 2 ), we can show that B 2 B 1 .
Thus, B 1 = B 2 , contradicting our assumption that they were distinct.
To formally prove that B 1 = B 2 , we use the Axiom of Extensionality:
X , Y : ( X = Y ) ( z : ( z X z Y ) )
We have shown:
z : ( z B 1 z B 2 ) and z : ( z B 2 z B 1 ) z : ( z B 1 z B 2 ) B 1 = B 2
This contradicts our assumption that B 1 and B 2 were distinct. Therefore, B is unique.
We conclude that for every subset A N + , there exists a unique subset B N + such that G ( B ) = A . □
Lemma 8 (Exhaustiveness of G)Let C : N + N + be the Collatz function defined as:
C ( n ) = n 2 if n 0 ( mod 2 ) 3 n + 1 if n 1 ( mod 2 )
and G : N + P ( N + ) be its inverse function defined as:
G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
Then G is exhaustive, i.e., n N + , m N + : n G ( m ) .
Proof. We will prove this by considering all possible congruence classes of n modulo 6.
Step 50: 1 Let n N + be arbitrary. We consider six cases:
Case 13.1 n 0 ( mod 6 )
k N + : n = 6 k Let m = 3 k Then m N + and m ¬ 4 ( mod 6 ) G ( m ) = { 2 m } = { 2 ( 3 k ) } = { 6 k } = { n } n G ( m )
Case 14.2 n 1 ( mod 6 )
k N + : n = 6 k + 1 Let m = 2 n = 2 ( 6 k + 1 ) = 12 k + 2 Then m N + and m ¬ 4 ( mod 6 ) G ( m ) = { 2 m } = { 2 ( 12 k + 2 ) } = { 24 k + 4 } n = 6 k + 1 = 24 k + 4 4 G ( m ) n G ( m )
Case 15.3 n 2 ( mod 6 )
k N + : n = 6 k + 2 Let m = 3 k + 1 Then m N + and m ¬ 4 ( mod 6 ) G ( m ) = { 2 m } = { 2 ( 3 k + 1 ) } = { 6 k + 2 } = { n } n G ( m )
Case 16.4 n 3 ( mod 6 )
k N + : n = 6 k + 3 Let m = 2 n = 2 ( 6 k + 3 ) = 12 k + 6 Then m N + and m ¬ 4 ( mod 6 ) G ( m ) = { 2 m } = { 2 ( 12 k + 6 ) } = { 24 k + 12 } n = 6 k + 3 = 24 k + 12 4 G ( m ) n G ( m )
Case 17.5 n 4 ( mod 6 )
k N + : n = 6 k + 4 Let m = 2 k + 1 Then m N + and m 1 ( mod 2 ) C ( m ) = 3 m + 1 = 3 ( 2 k + 1 ) + 1 = 6 k + 4 = n n G ( C ( m ) ) = G ( n )
Case 18.6 n 5 ( mod 6 )
k N + : n = 6 k + 5 Let m = 2 n = 2 ( 6 k + 5 ) = 12 k + 10 Then m N + and m ¬ 4 ( mod 6 ) G ( m ) = { 2 m } = { 2 ( 12 k + 10 ) } = { 24 k + 20 } n = 6 k + 5 = 24 k + 20 4 G ( m ) n G ( m )
Step 51:2 We have shown that for each congruence class of n modulo 6, there exists an m N + such that n G ( m ) . Since these cases are exhaustive and mutually exclusive, we conclude:
n N + , m N + : n G ( m )
Step 52:3 Therefore, G is exhaustive. □
Theorem 15 (Finiteness of Preimages of G)Let G : N + P ( N + ) be the inverse Collatz function defined as:
G ( n ) = { 2 n } { n 1 3 } if n 1 ( mod 3 ) and n 1 3 N + otherwise
Then for all j N , G j ( { 1 } ) is a finite set, where G j denotes j successive applications of G.
Proof. We will prove this theorem by induction on j. First, we establish key properties of G:
Lemma 9 (G Cardinality)For all n N + , | G ( n ) | 2 .
Proof. Let n N + be arbitrary. We consider two cases:
Case 19.1 n ¬ 1 ( mod 3 ) or n 1 3 N +
G ( n ) = { 2 n } | G ( n ) | = 1 2
Case 20.2 n 1 ( mod 3 ) and n 1 3 N +
G ( n ) = { 2 n , n 1 3 } | G ( n ) | = 2 2
Therefore, n N + , | G ( n ) | 2 . □
Lemma 10 (Multivalued Injectivity of G)For all a , b N + , if a b , then G ( a ) G ( b ) = .
Proof. This is a direct consequence of Lemma 6 (Multivalued Injectivity of G). □
Now we proceed with the induction proof:
Step 53: 1 Base case: j = 0
G 0 ( { 1 } ) = { 1 }
Clearly, | { 1 } | = 1 < . Therefore, G 0 ( { 1 } ) is finite.
Step 54: 2 Inductive hypothesis: Assume that for some k N , G k ( { 1 } ) is finite. Let | G k ( { 1 } ) | = m for some m N . Note that m is finite by the inductive hypothesis.
Step 55: 3 Inductive step: We need to prove that G k + 1 ( { 1 } ) is finite.
G k + 1 ( { 1 } ) = G ( G k ( { 1 } ) ) = G ( { x 1 , x 2 , , x m } ) where { x 1 , x 2 , , x m } = G k ( { 1 } ) = i = 1 m G ( x i )
Now, we will bound the cardinality of G k + 1 ( { 1 } ) using the following steps:
Step 56: 3a By Lemma 9, we know that | G ( x i ) | 2 for all i { 1 , 2 , , m } .
Explanation 2.This follows directly from Lemma 9, which states that for any n N + , | G ( n ) | 2 . Since each x i N + , we can apply this lemma to each G ( x i ) .
Step 57: 3b By Lemma 6, we know that G ( x i ) G ( x j ) = for all i j .
Step 58: 3c Using the sum of cardinalities of disjoint sets:
| G k + 1 ( { 1 } ) | = i = 1 m G ( x i ) = i = 1 m | G ( x i ) | ( sin ce the sets are disjoint by step 3 b ) i = 1 m 2 ( sin ce | G ( x i ) | 2 for all i by step 3 a ) = 2 m < ( sin ce m is finite by the inductive hypothesis )
Step 59: 3d Thus, G k + 1 ( { 1 } ) is finite, as its cardinality is bounded by 2 m , which is finite.
Step 60: 4 By the principle of mathematical induction, we conclude:
j N , G j ( { 1 } ) is finite
This completes the proof of the theorem.
Theorem 16 (Non-emptiness of Preimages of G)Let G : N + P ( N + ) be the inverse Collatz function defined as:
G ( n ) = { 2 n } { n 1 3 } if n 1 ( mod 3 ) and n 1 3 N + otherwise
Then for all j N , G j ( { 1 } ) is non-empty, where G j denotes j successive applications of G.
Proof. We will prove this theorem by strong induction on j. First, we establish a key property of G:
Lemma 11.For all n N + , G ( n ) .
Proof. Let n N + be arbitrary. By the definition of G:
G ( n ) = { 2 n } S , where S is either { n 1 3 } or
Since n N + , we know that 2 n N + . Therefore, { 2 n } . Thus, regardless of S, we have G ( n ) . □
Now we proceed with the strong induction proof:
Step 61: 1 Base case: j = 0
G 0 ( { 1 } ) = { 1 }
Clearly, { 1 } . Therefore, G 0 ( { 1 } ) is non-empty.
Step 62: 2 Inductive hypothesis: Assume that for all k j , where j N , G k ( { 1 } ) is non-empty.
Step 63: 3 Inductive step: We need to prove that G j + 1 ( { 1 } ) is non-empty.
By the inductive hypothesis, G j ( { 1 } ) is non-empty. Let x G j ( { 1 } ) .
Now, consider G ( x ) :
G ( x ) = { 2 x } { x 1 3 } if x 1 ( mod 3 ) and x 1 3 N + otherwise { 2 x } ( sin ce the union always includes { 2 x } )
Since x N + , we know that 2 x N + . Therefore:
G ( x ) 2 x G ( x )
Now, consider G j + 1 ( { 1 } ) :
G j + 1 ( { 1 } ) = G ( G j ( { 1 } ) ) = y G j ( { 1 } ) G ( y ) G ( x ) ( sin ce x G j ( { 1 } ) )
Thus, G j + 1 ( { 1 } ) is non-empty.
Step 64: 4 By the principle of strong mathematical induction, we conclude:
j N , G j ( { 1 } )
This completes the proof of the theorem.
Theorem 17 (Monotonicity of G)Let G : N + P ( N + ) be the inverse Collatz function defined as:
G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
Then G is monotonic, i.e., for all n N + and all x G ( n ) :
x 2 n
Proof. We will prove this theorem by considering all possible cases based on the congruence class of n modulo 6.
Step 65: 1 Let n N + be arbitrary.
Case 21.1 n ¬ 4 ( mod 6 )
In this case, G ( n ) = { 2 n } .
x G ( n ) : x = 2 n x = 2 n 2 n
Case 22.2 n 4 ( mod 6 )
In this case, G ( n ) = { 2 n , n 1 3 } .
Step 66:2 For x = 2 n :
x = 2 n 2 n
Step 67:3 For x = n 1 3 :
Since n 4 ( mod 6 ) , we can write n = 6 k + 4 for some k N .
x = n 1 3 = ( 6 k + 4 ) 1 3 = 6 k + 3 3 = 2 k + 1
Step 68:4 Now, we need to show that 2 k + 1 2 ( 6 k + 4 ) :
2 k + 1 2 ( 6 k + 4 ) 2 k + 1 12 k + 8 1 10 k + 8 7 10 k
Step 69:5 This inequality holds for all k N , therefore:
x = n 1 3 2 n
Step 70:6 We have shown that in all cases, for any x G ( n ) , x 2 n .
Step 71:7 Since n was arbitrary, we can conclude:
n N + , x G ( n ) : x 2 n
Step 72:8 Therefore, G is monotonic. □
Lemma 12 (C and G are Inverse Functions)Let C : N + N + be the Collatz function defined as:
C ( n ) = n 2 if n 0 ( mod 2 ) 3 n + 1 if n 1 ( mod 2 )
and let G : N + P ( N + ) be its inverse function defined as:
G ( n ) = { 2 n } if n ¬ 4 ( mod 6 ) { 2 n , n 1 3 } if n 4 ( mod 6 )
Then, for all n N + :
(1)
C ( G ( n ) ) = { n }
(2)
n G ( C ( n ) )
Proof. We will prove each part separately.
Step 73: 1 Let’s prove that C ( G ( n ) ) = { n } for all n N + :
Case 23.1 If n ¬ 4 ( mod 6 )
C ( G ( n ) ) = C ( { 2 n } ) = { 2 n 2 } = { n }
Case 24.2 If n 4 ( mod 6 )
C ( G ( n ) ) = C ( { 2 n , n 1 3 } ) = { C ( 2 n ) , C ( n 1 3 ) } = { 2 n 2 , 3 ( n 1 3 ) + 1 } = { n , n 1 + 1 } = { n , n } = { n }
Step 74:2 Let’s prove that n G ( C ( n ) ) for all n N + :
Case 25.1 If n is even
C ( n ) = n 2 G ( C ( n ) ) = G ( n 2 ) = { 2 · n 2 } = { n }
Therefore, n G ( C ( n ) ) .
Case 26.2 If n is odd
C ( n ) = 3 n + 1 G ( C ( n ) ) = G ( 3 n + 1 )
Now, we need to consider two subcases:
Subcase 10.2a If 3 n + 1 ¬ 4 ( mod 6 )
G ( C ( n ) ) = G ( 3 n + 1 ) = { 2 ( 3 n + 1 ) } = { 6 n + 2 }
Subcase 11.2b If 3 n + 1 4 ( mod 6 )
G ( C ( n ) ) = G ( 3 n + 1 ) = { 2 ( 3 n + 1 ) , ( 3 n + 1 ) 1 3 } = { 6 n + 2 , n }
In both subcases, we can see that n G ( C ( n ) ) . For subcase 2a, note that n = ( 6 n + 2 ) 2 6 , which is an integer since n is odd. For subcase 2b, n is explicitly included in the set.
Therefore, for all odd n, we have n G ( C ( n ) ) .
Step 75:3 Thus, we have proved that C ( G ( n ) ) = { n } and n G ( C ( n ) ) for all n N + . □
Theorem 18 (Preservation of Properties under Composition of G)For all i , j N , the composition G i G j satisfies the following properties:
(1)
Injectivity
(2)
Multivalued injectivity
(3)
Monotonicity
(4)
Exhaustiveness
(5)
Finiteness of preimages
(6)
Non-emptiness of preimages
where G : N + P ( N + ) is the inverse Collatz function defined as in Theorem 15.
Proof. We will prove each property separately for G i G j , using the fact that G and C are inverse functions of each other, as established in Lemma 12.
Lemma 13 (C and G are Inverse Functions)For all n N + :
(1)
C ( G ( n ) ) = { n }
(2)
n G ( C ( n ) )
Step 76: 1 Injectivity:
a , b N + , ( G i G j ) ( a ) = ( G i G j ) ( b ) a = b
Proof:
Assume ( G i G j ) ( a ) = ( G i G j ) ( b ) C i + j ( ( G i G j ) ( a ) ) = C i + j ( ( G i G j ) ( b ) ) ( applying C i + j to both sides ) a = b ( by Lemma , applying C i + j cancels out G i G j )
Step 77: 2 Multivalued injectivity:
a , b N + , a b ( G i G j ) ( a ) ( G i G j ) ( b ) =
Proof:
Assume a b and , for contradiction , ( G i G j ) ( a ) ( G i G j ) ( b ) x ( G i G j ) ( a ) ( G i G j ) ( b ) C i + j ( x ) = a and C i + j ( x ) = b ( by Lemma ) a = b ( contradiction )
Step 78: 3 Monotonicity:
x N + , y ( G i G j ) ( x ) : y 4 i + j x
Proof: Let x N + and y ( G i G j ) ( x ) .
Lemma 14 (Upper Bound for Collatz Function)For all n N + , C ( n ) 4 n .
Proof. We consider two cases:
Case 27.1 If n is even: C ( n ) = n / 2 < n 4 n
Case 28.2 If n is odd: C ( n ) = 3 n + 1 4 n (since n 1 ) Therefore, in all cases, C ( n ) 4 n . □
Now, let’s apply this lemma to our proof of monotonicity:
y ( G i G j ) ( x ) C i + j ( y ) = x ( by Lemma ) x 4 i + j y ( by applying Lemma i + j times ) y 4 i + j x ( by the monotonicity of G , Theorem )
Explanation 3 (Monotonicity Implication)The inequality y 4 i + j x implies monotonicity for G i G j because:
(1)
It provides an upper bound for all elements y in ( G i G j ) ( x ) in terms of x.
(2)
This upper bound, 4 i + j x , is a strictly increasing function of x (since 4 i + j > 0 ).
(3)
Therefore, as x increases, the maximum possible value for y also increases.
(4)
This ensures that for any x 1 < x 2 , all elements in ( G i G j ) ( x 1 ) are less than or equal to all elements in ( G i G j ) ( x 2 ) , which is the definition of monotonicity for set-valued functions.
Thus, y 4 i + j x guarantees that G i G j is monotonic.
Step 79: 4 Exhaustiveness:
n N + , m N + : n ( G i G j ) ( m )
Proof:
Let n N + Let m = C i + j ( n )
To clarify that m N + :
Lemma 15 (Positivity of Iterated Collatz Function)For all n N + and all k N , C k ( n ) N + .
Proof. We prove this by induction on k:
Base case: For k = 0 , C 0 ( n ) = n N + .
Inductive step: Assume C k ( n ) N + for some k 0 . We prove for k + 1 :
  • If C k ( n ) is even: C k + 1 ( n ) = C ( C k ( n ) ) = C k ( n ) 2 N +
  • If C k ( n ) is odd: C k + 1 ( n ) = C ( C k ( n ) ) = 3 C k ( n ) + 1 N +
By the principle of mathematical induction, k N , C k ( n ) N + . □
By Lemma 15, we know that m = C i + j ( n ) N + .
Now, we can conclude:
n ( G i G j ) ( m ) ( by Lemma )
Step 80: 5 Finiteness of preimages:
S N + , | S | < | ( G i G j ) ( S ) | <
Proof:
Let S N + be finite For each n S , | ( G i G j ) ( { n } ) | 2 i + j ( by the definition of G ) Therefore , | ( G i G j ) ( S ) | | S | · 2 i + j <
Step 81: 6 Non-emptiness of preimages:
S N + , S ( G i G j ) ( S )
Proof:
Let S N + be non - empty Let n S Then ( G i G j ) ( { n } ) ( by Lemma ) Therefore , ( G i G j ) ( S )
Step 82: 7 Therefore, all six properties are preserved under the composition G i G j .
Remark 19 (Key Properties of G and Their Preservation)This theorem establishes that the crucial properties of G are preserved under composition. This is fundamental for our analysis, as it allows us to extend our reasoning about G to more complex structures built from G.
Lemma 16 (Equivalence of Properties between C and G)Let C : N + N + be the Collatz function and G : N + P ( N + ) be its inverse function as defined in Definitions 1 and 2 respectively. Then, for any property P of sequences in N + , the following are equivalent:
(1)
For all Collatz sequences ( a k ) k N generated by C, P ( ( a k ) k N ) holds.
(2)
For all sequences ( b k ) k N such that k N , b k + 1 G ( b k ) , P ( ( b k ) k N ) holds.
Formally:
P : S ( N + ) { true , false } ,
( a k ) k N C , P ( ( a k ) k N ) ( b k ) k N G , P ( ( b k ) k N )
where S ( N + ) is the set of all sequences in N + , C is the set of all Collatz sequences, and G is the set of all sequences generated by G.
Proof. First, let us recall that C and G are well-defined according to the following lemmas:
  • Lemma 13: The Collatz function C is well-defined for all positive integers.
  • Lemma 3: For every n N + , the set G ( n ) is non-empty and uniquely determined.
We will now proceed to prove both directions of the equivalence.
Step 83: 1 (⇒): Assume that for all Collatz sequences ( a k ) k N generated by C, P ( ( a k ) k N ) holds.
Let ( b k ) k N be any sequence such that k N , b k + 1 G ( b k ) . Define a sequence ( a k ) k N as follows:
a 0 = b 0 , k N , a k + 1 = C ( a k )
We claim that k N , b k = a k . We prove this by induction:
Step 84: 2 Base case: b 0 = a 0 by definition.
Step 85: 3 Inductive step: Assume b k = a k for some k 0 . Then:
b k + 1 G ( b k ) = G ( a k ) ( by inductive hypothesis ) = G ( C ( a k + 1 ) ) ( by definition of a k + 1 ) = { a k + 1 } ( by property of inverse functions )
Therefore, b k + 1 = a k + 1 , completing the induction.
Step 86: 4 Since ( a k ) k N is a Collatz sequence, P ( ( a k ) k N ) holds by assumption. As k N , b k = a k , we have P ( ( b k ) k N ) .
Step 87: 5 (⇐): Assume that for all sequences ( b k ) k N such that k N , b k + 1 G ( b k ) , P ( ( b k ) k N ) holds.
Let ( a k ) k N be any Collatz sequence generated by C. Then k N :
a k + 1 = C ( a k ) a k G ( a k + 1 )
Therefore, ( a k ) k N satisfies the condition k N , a k G ( a k + 1 ) . By assumption, P ( ( a k ) k N ) holds.
Step 88: 6 Thus, we have shown both directions of the equivalence, completing the proof. □
Remark 20 (Bridging C and G)This lemma provides a critical link between sequences generated by C and those generated by G. It allows us to transfer results between these two perspectives, which is essential for our overall proof strategy.
Proposition 21.For any Collatz sequence ( a k ) k 0 :
(1)
If a k is even, then a k + 1 < a k .
(2)
If a k is odd, then a k + 1 > a k .
Proof. Follows directly from the definition of the Collatz function. □
Lemma 17 (Properties of Collatz Function)Let C : N + N + be the Collatz function defined as:
C ( x ) = x 2 if x 0 ( mod 2 ) 3 x + 1 if x 1 ( mod 2 )
Then:
(1)
If x > 1 is even, then C ( x ) < x .
(2)
If x > 1 is odd, then C ( x ) > x .
(3)
C ( x ) = 1 if and only if x = 1 or x = 2 or x = 4 .
(4)
For any x > 1 , there exists a positive integer k such that C k ( x ) < x , where C k denotes k applications of C.
Proof. Properties 1-3 follow directly from the definition of C. For property 4: If x is even, k = 1 suffices. If x is odd, consider the sequence x , 3 x + 1 , 3 x + 1 2 . We have 3 x + 1 2 < x if and only if 3 x + 1 < 2 x if and only if x > 1 . Therefore, for odd x > 1 , k = 2 suffices. □
Lemma 18 (Equivalence of Properties between C and G)Let C : N + N + be the Collatz function and G : N + P ( N + ) be its inverse function as defined in Definitions 1 and 2 respectively. Then, for any property P of sequences in N + , the following are equivalent:
(1)
For all Collatz sequences ( a k ) k 0 generated by C, P ( ( a k ) k 0 ) holds.
(2)
For all sequences ( b k ) k 0 such that b k + 1 G ( b k ) for all k 0 , P ( ( b k ) k 0 ) holds.
Proof. First, let us recall that C and G are well-defined according to the following lemmas:
  • Lemma 13: The Collatz function C is well-defined for all positive integers.
  • Lemma 3: For every n N + , the set G ( n ) is non-empty and uniquely determined.
We will now proceed to prove both directions of the equivalence.
Step 89: 1 (1 ⇒ 2): Assume that for all Collatz sequences ( a k ) k 0 generated by C, P ( ( a k ) k 0 ) holds.
Let ( b k ) k 0 be any sequence such that b k + 1 G ( b k ) for all k 0 . Define a sequence ( a k ) k 0 as follows:
a 0 = b 0 , a k + 1 = C ( a k ) for all k 0
We claim that b k = a k for all k 0 . We prove this by induction:
Step 90: 2 Base case: b 0 = a 0 by definition.
Step 91: 3 Inductive step: Assume b k = a k for some k 0 . Then:
b k + 1 G ( b k ) = G ( a k ) ( by inductive hypothesis ) = G ( C ( a k + 1 ) ) ( by definition of a k + 1 ) = { a k + 1 } ( by property of inverse functions )
Therefore, b k + 1 = a k + 1 , completing the induction.
Step 92: 4 Since ( a k ) k 0 is a Collatz sequence, P ( ( a k ) k 0 ) holds by assumption. As b k = a k for all k 0 , we have P ( ( b k ) k 0 ) .
Step 93: 5 (2 ⇒ 1): Assume that for all sequences ( b k ) k 0 such that b k + 1 G ( b k ) for all k 0 , P ( ( b k ) k 0 ) holds.
Figure 1. Boundnedness of Collatz Sequence
Figure 1. Boundnedness of Collatz Sequence
Preprints 114544 g001
Let ( a k ) k 0 be any Collatz sequence generated by C. Then for all k 0 :
a k + 1 = C ( a k ) a k G ( a k + 1 )
Therefore, ( a k ) k 0 satisfies the condition a k G ( a k + 1 ) for all k 0 . By assumption, P ( ( a k ) k 0 ) holds.
Step 94: 6 Thus, we have shown both directions of the equivalence, completing the proof. □

5. Properties of Collatz Sequences

5.1. Boundedness of Collatz Sequences

5.1.1. Auxiliary Proofs

Lemma 19 (Finiteness and Non-emptiness of S k )Let k N and define S k = { x N + : i k , x G i ( { 1 } ) } . Then S k is finite and non-empty.
Proof. We proceed by proving non-emptiness and finiteness separately:
Step 95: 1 Non-emptiness of S k :
(a)
Observe that 1 G 0 ( { 1 } ) = { 1 } .
(b)
Since 0 k for all k N : 1 S k
(c)
Therefore: S k
Step 96: 2 Finiteness of S k :
(a)
We first prove by induction that i N , G i ( { 1 } ) is finite:
(i)
Base case: i = 0   G 0 ( { 1 } ) = { 1 } is finite
(ii)
Inductive step: Assume G i ( { 1 } ) is finite for some i 0 . We prove for i + 1 : G i + 1 ( { 1 } ) = G ( G i ( { 1 } ) ) = x G i ( { 1 } ) G ( x ) By the definition of G, x N + , | G ( x ) | 2 . Let n = | G i ( { 1 } ) | . Then: | G i + 1 ( { 1 } ) | 2 n < Therefore, G i + 1 ( { 1 } ) is finite.
(iii)
By the principle of mathematical induction: i N , G i ( { 1 } ) is finite
(b)
Now we prove that S k is finite: S k = { x N + : i k , x G i ( { 1 } ) } = i = 0 k G i ( { 1 } ) This is a finite union of finite sets, therefore S k is finite.
Step 97: 3 Formal statement of the conclusion:
k N , S k N + : ( S k = { x N + : i k , x G i ( { 1 } ) } ) ( S k ) ( | S k | < )
Lemma 20 (Non-emptiness of T)Let N N + , k = log 2 N , and T = { x S k : x N / 2 k } . Then T .
Proof. We proceed with a formal proof using first-order logic, set theory, and properties of natural numbers:
Step 98: 1 Given:
N N + k = log 2 N S k = { x N + : i k , x G i ( { 1 } ) } T = { x S k : x N / 2 k }
Step 99: 2 Since N N + , we have N 1 . Therefore, log 2 N is well-defined.
Step 100: 3 From the definition of ceiling function:
log 2 N k < log 2 N + 1
Step 101: 4 Taking 2 x of both sides (which is an increasing function):
N 2 k < 2 N
Step 102: 5 We will prove that 2 k T by showing:
(a)
2 k S k
(b)
2 k N / 2 k
Step 103: 6 To prove 5a, we use induction on i to show i N , 2 i S i :
(a)
Base case: i = 0
2 0 = 1 1 G 0 ( { 1 } ) = { 1 } 2 0 S 0
(b)
Inductive step: Assume 2 i S i for some i 0 . We prove for i + 1 :
2 i S i 2 i + 1 G ( 2 i ) S i + 1 ( by definition of G and S i + 1 )
(c)
By the principle of mathematical induction:
i N , 2 i S i
(d)
Since k i , we have S i S k . Therefore:
2 k S k
Step 104: 7 To prove 5b:
2 k N ( from step 4 ) N / 2 k ( sin ce 2 k 1 )
Step 105: 8 From steps 6d and 7, we conclude:
2 k T
Step 106: 9 Therefore:
T
Step 107: 10 Formal statement of the conclusion:
N N + , k N + , T N + : ( k = log 2 N T = { x S k : x N / 2 k } ) T
Lemma 21 (Upper Bound of m N )Let N N + , k = log 2 N , S k = { x N + : i k , x G i ( { 1 } ) } , T = { x S k : x N / 2 k } , and m N = min T . Then m N < 2 N .
Proof. We proceed with a formal proof using first-order logic and properties of real and natural numbers:
Step 108: 1 Given:
N N + k = log 2 N S k = { x N + : i k , x G i ( { 1 } ) } T = { x S k : x N / 2 k } m N = min T
Step 109: 2 From the definition of ceiling function:
log 2 N k < log 2 N + 1
Step 110: 3 Taking 2 x of both sides (which is an increasing function):
N 2 k < 2 N
Step 111: 4 By Lemma 22, we know that:
x S k : x 2 k
Step 112: 5 Since m N T S k , we can conclude:
m N 2 k
Step 113: 6 Combining steps 3 and 5:
m N 2 k < 2 N
Step 114: 7 Therefore, we can conclude:
m N < 2 N
Step 115: 8 Formal statement of the conclusion:
N N + , k N , S k , T N + , m N N + :
( k = log 2 N S k = { x N + : i k , x G i ( { 1 } ) }
T = { x S k : x N / 2 k } m N = min T ) m N < 2 N
Lemma 22 (Boundedness of S k )Let k N and S k = { x N + : i k , x G i ( { 1 } ) } . Then x S k : x 2 k .
Proof. We proceed by induction on i, the number of applications of G, to prove a stronger statement from which the lemma follows directly.
Step 116: 1 Define the proposition P ( i ) :
P ( i ) : x G i ( { 1 } ) , x 2 i
Step 117: 2 Base case: i = 0
G 0 ( { 1 } ) = { 1 } 1 2 0 = 1 P ( 0 ) is true
Step 118: 3 Inductive step: Assume P ( i ) is true for some i 0 . We prove P ( i + 1 ) :
Step 119: 3a Let y G i + 1 ( { 1 } ) .
Step 120: 3b By definition of G, x G i ( { 1 } ) such that y G ( x ) .
Step 121: 3c By the inductive hypothesis:
x 2 i
Step 122: 3d By the monotonicity property of G:
z G ( x ) : z 2 x
Step 123: 3e Combining (3c) and (3d):
y 2 x 2 ( 2 i ) = 2 i + 1
Step 124: 3f Therefore, P ( i + 1 ) is true.
Step 125: 4 By the principle of mathematical induction:
i N , P ( i ) is true
Step 126: 5 Now, we prove the lemma statement:
Step 127: 5a Let x S k be arbitrary.
Step 128: 5b By definition of S k :
i k : x G i ( { 1 } )
Step 129: 5c From step 4, we know that P ( i ) is true, so:
x 2 i
Figure 2. G-graph
Figure 2. G-graph
Preprints 114544 g002
Step 130: 5d Since i k :
2 i 2 k
Step 131: 5e By transitivity of inequality:
x 2 i 2 k
Step 132: 5f Therefore:
x 2 k
Step 133: 6 Conclusion: We have shown that:
x S k : x 2 k
Which proves the lemma. □
Definition 22 (G-graph)Let G : N + P ( N + ) be the inverse Collatz function as defined in Definition 2. The G-graph is a directed graph ( V , E ) where:
  • V = N + is the set of vertices.
  • E = { ( m , n ) N + × N + : m G ( n ) } is the set of edges.
A path in the G-graph from a to b is a sequence of vertices ( v 0 , v 1 , , v k ) where v 0 = a , v k = b , and ( v i , v i + 1 ) E for all 0 i < k .
Lemma 23 (Uniqueness of Paths in G-graph)For any a N + , there exists at most one path in the G-graph from 1 to a.
Formally:
a N + , ! ( v 0 , v 1 , , v k ) : ( v 0 = 1 ) ( v k = a ) ( i { 0 , 1 , , k 1 } , v i + 1 G ( v i ) )
where G is the inverse Collatz function as defined in Definition 2.
Proof. We prove this by induction on the length of the path.
Step 134: 1 Base case: For paths of length 0, the statement is trivially true as there is only one path of length 0 from 1 to 1.
Step 135: 2 Inductive hypothesis: Assume that for some k 0 , there is at most one path of length k from 1 to any number.
Step 136: 3 Inductive step: Consider a path of length k + 1 from 1 to some number b. Let this path be ( 1 = v 0 , v 1 , , v k , v k + 1 = b ) .
Step 137: 4 By the definition of the G-graph, we have v k G ( b ) .
Step 138: 5 By the inductive hypothesis, the path from 1 to v k is unique.
Step 139: 6 Now, suppose for contradiction that there is another path of length k + 1 from 1 to b, say ( 1 = u 0 , u 1 , , u k , u k + 1 = b ) .
Step 140: 7 We must have u k G ( b ) as well.
Step 141: 8 If u k v k , this would imply that G ( b ) contains two different elements, contradicting the multivalued injectivity of G (Lemma 6).
Step 142: 9 Therefore, u k = v k , and by the inductive hypothesis, the paths ( u 0 , , u k ) and ( v 0 , , v k ) must be identical.
Step 143: 10 Thus, the two paths of length k + 1 from 1 to b are identical.
By the principle of mathematical induction, we conclude that for any a N + , there exists at most one path in the G-graph from 1 to a.
Formally:
a N + , ! ( v 0 , v 1 , , v k ) : ( v 0 = 1 ) ( v k = a ) ( i { 0 , 1 , , k 1 } , v i + 1 G ( v i ) )
Lemma 24 (Path Convergence in G-graph)For any two elements a , b N + where a b , if there exist paths in the G-graph from 1 to a and from 1 to b, then these paths converge at some point c a and remain identical thereafter.
Proof. We proceed with a formal proof using first-order logic and set theory:
Step 144: 1 Let a , b N + such that a b .
Step 145: 2 By Lemma 23, we know that the paths from 1 to a and from 1 to b are unique. Let these paths be:
P a = ( 1 = x 0 , x 1 , , x m = a ) P b = ( 1 = y 0 , y 1 , , y n = b )
where m , n N and i { 0 , , m 1 } , j { 0 , , n 1 } : x i + 1 G ( x i ) y j + 1 G ( y j ) .
Step 146: 3 Define the set of indices where the paths coincide:
S = { i N : i min ( m , n ) x i = y i }
Step 147: 4 Prove that S is non-empty:
x 0 = 1 = y 0 0 S S
Step 148: 5 Since S N and S , by the Well-Ordering Principle, S has a maximum element. Define: k = max S
Step 149: 6 Define the convergence point: c = x k = y k
Step 150: 7 Prove that the paths are identical up to k:
j k : x j = y j
This follows directly from the definition of S and k.
Step 151: 8 Prove that the paths remain identical after k:
j > k : x j = y j ( This follows from the uniqueness of paths established in Lemma )
Step 152: 9 Prove that c a :
c = x k k m ( sin ce k S and by definition of S ) x k appears in P a no later than x m = a c = x k x m = a
Step 153: 10 Conclusion: We have shown that the paths P a and P b converge at point c = x k = y k , where c a , and remain identical thereafter. Formally:
c N + , k N : ( c a ) ( j k : x j = y j = c j )
where ( c j ) j k denotes the common path after convergence. □
Lemma 25 (Existence of Paths in G-graph)Let G : N + P ( N + ) be the inverse Collatz function as defined in Definition 2. For all n N + , there exists a path in the G-graph from 1 to n.
Formally:
n N + , k N , ( p 0 , p 1 , , p k ) : ( p 0 = 1 ) ( p k = n ) ( i { 0 , 1 , , k 1 } , p i + 1 G ( p i ) )
Proof. We proceed by strong induction on n.
Step 154: 1 Base case: n = 1 The trivial path ( 1 ) satisfies the condition.
Step 155: 2 Inductive hypothesis: We assume the statement is true for all natural numbers less than or equal to some m 1 .
Step 156: 3 Inductive step: We prove for m + 1 .
By the exhaustiveness property of G (Lemma 8), we know that:
q N + : m + 1 G ( q )
Step 157: 4 We consider two cases:
Case 29.1 If q m : By the inductive hypothesis, there exists a path ( p 0 , p 1 , , p k ) from 1 to q. Then, ( p 0 , p 1 , , p k , m + 1 ) is a valid path from 1 to m + 1 .
Case 30.2 If q > m : Then q = m + 1 , since m + 1 G ( q ) and G ( q ) 2 q by the monotonicity property of G (Theorem 17). In this case, we apply the same argument as in Case 1, but with q = m + 1 .
Step 158:5 In both cases, we have constructed a valid path from 1 to m + 1 .
Step 159:6 By the principle of strong induction, we conclude that the statement is true for all n N + . □
Lemma 26 (Extension of G Properties Under Composition)Let G : N + P ( N + ) be the inverse Collatz function. For all i , j N , the composition G i G j satisfies the following properties:
(1)
Injectivity
(2)
Multivalued injectivity
(3)
Monotonicity
(4)
Exhaustiveness
(5)
Finiteness of preimages
(6)
Non-emptiness of preimages
where G i denotes i successive applications of G.
Proof. The proof of this lemma is provided in Theorem 18. □

5.1.2. Global Structure of Collatz Sequences

Theorem 23 (Generative Completeness of the Inverse Collatz Function)Let G : N + P ( N + ) be the inverse Collatz function as defined in Definition 2, and C : N + N + be the Collatz function as defined in Definition 1. Then for all N N + , there exists m N N + and k = log 2 N 1 such that:
(1)
(Minimality) m N < N
(2)
(Generativity) n N , i k : n G i ( { m N } )
(3)
(Uniqueness) m < m N , n N : i N , n G i ( { m } )
(4)
(Connection to C) n N , j k : C j ( n ) = m N
where G i and C j denote i and j successive applications of G and C respectively, and G 0 ( { m N } ) = { m N } .
Proof. We will prove this theorem in four main steps, each established by a separate lemma:
(1)
Existence of m N (Lemma 27)
(2)
Generative property of m N (Lemma 28)
(3)
Minimality of m N (Lemma 29)
(4)
Connection to C (Lemma 30)
Lemma 27 (Existence of m N ) N N + , m N N + : m N < N
Proof. Let N N + be arbitrary.
Step 160: 1 Define k = log 2 N 1 .
Step 161: 2 Define the set S k = { x N + : i k , x G i ( { 1 } ) } .
Step 162: 3 By Lemma 19, S k is finite and non-empty.
Step 163: 4 Define T = { x S k : x N / 2 k } .
Step 164: 5 We now prove that T is non-empty:
Sublemma 24. T = { x S k : x N / 2 k } is non-empty.
Proof. We will prove that 2 k T by showing 2 k S k and 2 k N / 2 k :
  • We prove by induction that i N , 2 i S i .
  • Base case: 2 0 = 1 G 0 ( { 1 } ) = { 1 } S 0 .
  • Inductive step: Assume 2 i S i for some i 0 . Then 2 i G j ( { 1 } ) for some j i . By the monotonicity of G (Theorem 17), 2 i + 1 G ( G j ( { 1 } ) ) = G j + 1 ( { 1 } ) S i + 1 .
  • By mathematical induction, i N , 2 i S i .
Since k i , we have S i S k . Therefore, 2 k S k . Moreover, 2 k N / 2 (by the choice of k), 2 k N / 2 k . Therefore, 2 k T , T . □
Step 165: 6 Define m N = min T .
Step 166: 7 We now prove that m N < N :
Sublemma 25. m N < N
Proof. By definition, m N T S k . By Lemma 22, x S k : x 2 k . Therefore, m N 2 k < N (by the choice of k in step 1). □
Thus, we have proven the existence of m N < N for any N N + .
Lemma 28 (Generative property of m N ) n N , i k : n G i ( { m N } )
Proof. Let n N be arbitrary.
Step 167: 1 By the exhaustiveness of G (Lemma 8), j N : n G j ( { 1 } ) .
Step 168: 2 Let X = { y G j ( { 1 } ) : y n } .
Step 169: 3 X is non-empty (contains n) and finite (subset of finite G j ( { 1 } ) by Theorem 15).
Step 170: 4 Let x = max X . Then x S k (since k j by choice of k).
Step 171: 5 We have x n n / 2 k N / 2 k , x T .
Step 172: 6 By definition of m N : m N x < N
Step 173: 7 By the monotonicity of G (Theorem 17), l k : x G l ( { m N } )
Step 174: 8 Since x n and n G q ( { x } ) for some q j l , we have: n G i ( { m N } ) where i = l + q k
Step 175: 9 To verify that i k :
i = l + q k + ( j l ) = k + j l k + j
Since j log 2 N 1 = k (by the choice of k), we have:
i k + k = 2 k 2 ( log 2 N 1 ) < 2 log 2 N k
The last inequality holds because k = log 2 N 1 log 2 N 1 , so 2 log 2 N 2 ( k + 1 ) = 2 k + 2 k for N 4 . For N < 4 , the lemma can be verified directly.
Thus, we have proven the generative property of m N . □
Lemma 29 (Minimality of m N ) m < m N , n N : i N , n G i ( { m } )
Proof. Let m < m N be arbitrary.
Step 176: 1 By definition of m N : m < N / 2 k
Step 177: 2 Let n = N / 2 k .
Step 178: 3 Then: n N / 2 k < n + 1
Step 179: 4 This implies: n N and n > m
Step 180: 5 By the monotonicity of G (Theorem 17): i N , y G i ( { m } ) : y 2 i m < 2 i n
Step 181: 6 For i k : 2 i n < 2 k n N
Step 182: 7 For i > k : 2 i n > N
Step 183: 8 Therefore: i N , n G i ( { m } )
Thus, we have proven the minimality of m N . □
Lemma 30 (Connection to C) n N , j k : C j ( n ) = m N
Proof. Let n N be arbitrary.
Step 184: 1 By the Generativity property (Lemma 28), we know that: i k : n G i ( { m N } )
Step 185: 2 By Lemma 12, we know that C and G are inverse functions of each other. Specifically: x , y N + : y G ( x ) C ( y ) = x
Step 186: 3 Applying this property i times, we get: n G i ( { m N } ) C i ( n ) = m N
Step 187: 4 Let j = i . Then j k and C j ( n ) = m N
Thus, we have proven the connection to C. □
Combining Lemmas 27, 28, 29, and 30, we have proven all parts of Theorem 23.
Corollary 26 (Confluence of Collatz Sequences)For any N N + , all Collatz sequences starting from numbers n N eventually converge to the same value m N and follow the same path thereafter. Formally:
N N + , n 1 , n 2 N , j 1 , j 2 , l N : ( C j 1 ( n 1 ) = C j 2 ( n 2 ) = m N ) ( k 0 , C j 1 + k ( n 1 ) = C j 2 + k ( n 2 ) = C k ( m N ) )
where C is the Collatz function and m N is as defined in Theorem 23.
Proof. Let N N + be arbitrary and let n 1 , n 2 N be any two positive integers less than or equal to N.
Step 188: 1 By Lemma 30 (Connection to C), we know that:
j 1 k : C j 1 ( n 1 ) = m N j 2 k : C j 2 ( n 2 ) = m N
where k = log 2 N 1 .
Step 189: 2 This establishes the first part of our claim:
C j 1 ( n 1 ) = C j 2 ( n 2 ) = m N
Step 190: 3 Now, let’s consider the sequences after reaching m N . For any k 0 :
C j 1 + k ( n 1 ) = C k ( C j 1 ( n 1 ) ) = C k ( m N ) C j 2 + k ( n 2 ) = C k ( C j 2 ( n 2 ) ) = C k ( m N )
Step 191: 4 This establishes the second part of our claim:
k 0 , C j 1 + k ( n 1 ) = C j 2 + k ( n 2 ) = C k ( m N )
Step 192: 5 Since N, n 1 , and n 2 were arbitrary (with the condition n 1 , n 2 N ), we can conclude that this property holds for all N N + and all n 1 , n 2 N .
Therefore, all Collatz sequences starting from numbers n N eventually converge to m N and follow the same path thereafter. □
Lemma 31 (Finite Maximum in Collatz Sequences)For any N N + and n N , there exists a finite maximum M in the Collatz sequence starting from n before reaching m N . Formally:
N N + , n N , M , j N : ( C j ( n ) = m N ) ( i < j , C i ( n ) M ) ( M < )
where C is the Collatz function and m N is as defined in Theorem 23.
Proof. Let N N + be arbitrary and let n N .
Step 193: 1 By Lemma 30, we know that:
j k : C j ( n ) = m N
where k = log 2 N 1 .
Step 194: 2 Consider the finite sequence S = ( n , C ( n ) , C 2 ( n ) , , C j 1 ( n ) ) .
Step 195: 3 Since S is a finite sequence of natural numbers, it must have a maximum element. Let’s call this maximum M:
M = max { C i ( n ) : 0 i < j }
Step 196: 4 By definition of M:
i < j , C i ( n ) M
Step 197: 5 M is finite because:
  • S is a finite sequence (it has j elements, where j k < )
  • Each element of S is a natural number (C is well-defined on N + by Theorem 13)
  • The maximum of a finite set of natural numbers is always finite
Step 198: 6 Therefore, we have shown that there exists a finite M such that:
( C j ( n ) = m N ) ( i < j , C i ( n ) M ) ( M < )
Since N and n were arbitrary (with the condition n N ), this holds for all N N + and all n N . □
Corollary 27 (Boundedness of Collatz Sequences)For any n N + , the Collatz sequence starting from n is bounded. Formally:
n N + , M N : j N , C j ( n ) M
where C is the Collatz function.
Proof. Let n N + be arbitrary. Consider N = n in Lemma 31. By Lemma 31, we know that there exists a finite maximum M in the Collatz sequence starting from n before reaching m N . Formally:
M , j N : ( C j ( n ) = m N ) ( i < j , C i ( n ) M ) ( M < )
Since m N is the minimum value that the Collatz sequence reaches and the sequence eventually cycles between values below this minimum (by the nature of the Collatz function), it follows that:
k j , C k ( n ) M
Therefore, the Collatz sequence starting from n is bounded by M for all steps j N , and we have:
n N + , M N : j N , C j ( n ) M
This completes the proof. □
Definition 28 (Eventually Non-Periodic Subsequence)Let ( a k ) k 0 be a sequence and ( a k ) k N be a subsequence starting from index N. We say that ( a k ) k N is eventually non-periodic if:
p N + , K N : k K , a k a k + p
In other words, for any potential period p, there exists a point K in the sequence after which no term is equal to any term p positions ahead of it.
Lemma 32 (Monotonicity of Eventually Non-Periodic Collatz Subsequences)Let ( a k ) k 0 be a Collatz sequence. If there exists an index N and a real number L > 1 such that a k L for all k N , and the subsequence ( a k ) k N is not eventually periodic, then for any M N , there exists an index j > M such that a j > a M .
Formally:
( a k ) k 0 C , N N , L R + , ( ( L > 1 k N , a k L ) ¬ EventuallyPeriodic ( ( a k ) k N ) ) M N , j > M : a j > a M
where C is the set of all Collatz sequences, and EventuallyPeriodic ( ( a k ) k N ) is a predicate that is true if and only if ( a k ) k N is eventually periodic.
Proof. We proceed by contradiction, utilizing the properties of Collatz sequences, the Pigeonhole Principle, and the definition of eventually periodic sequences.
Step 199: 1 Let ( a k ) k 0 C be a Collatz sequence, N N , and L R + with L > 1 , such that:
k N : a k L
and ( a k ) k N is not eventually periodic.
Step 200: 2 Let M N be arbitrary.
Step 201: 3 Assume, for the sake of contradiction, that:
k > M : a k a M
Step 202: 4 This implies that the subsequence ( a k ) k > M is bounded above by a M and below by L.
Step 203: 5 Define the set S = { a k : k > M } . Note that S is non-empty and countable.
Step 204: 6 Since S N and is bounded, it is finite. Let | S | = n for some n N + .
Step 205: 7 Define a function f : N S by f ( k ) = a M + k + 1 for k 0 .
Step 206: 8 By the Pigeonhole Principle (Theorem 7), since the domain of f is infinite and its codomain S is finite, there must exist at least two distinct elements in the domain that map to the same element in the codomain. Formally:
i , j N , i < j : f ( i ) = f ( j )
Step 207: 9 This implies:
i , j N , i < j : a M + i + 1 = a M + j + 1
Step 208: 10 Let p = j i . Then for all k M + i + 1 :
a k = a k + p
Step 209: 11 This means that the sequence ( a k ) k M + i + 1 is periodic with period p.
Step 210: 12 Now, we will show that this contradicts our assumption that ( a k ) k N is not eventually periodic.
Step 211: 13 Recall the definition of an eventually periodic sequence:
Definition 29 (Eventually Periodic Sequence)A sequence ( x k ) k 0 is eventually periodic if:
K N , p N + : k K , x k = x k + p
Step 212: 14 In our case, we have shown that:
K = M + i + 1 , p N + : k K , a k = a k + p
Step 213: 15 Since M + i + 1 N (because M N and i 0 ), this means that ( a k ) k N is eventually periodic.
Step 214: 16 This directly contradicts our initial assumption that ( a k ) k N is not eventually periodic.
Step 215: 17 Therefore, our assumption in step 3 must be false. Thus, we can conclude:
j > M : a j > a M
Step 216: 18 Since M N was arbitrary, this holds for all M N .
We have thus proven:
( a k ) k 0 C , N N , L R + , ( ( L > 1 k N , a k L ) ¬ EventuallyPeriodic ( ( a k ) k N ) ) M N , j > M : a j > a M
This completes the proof of the lemma. □
Remark 30 (Connection between Non-Periodicity and Existence of Greater Terms)The key connection between non-periodicity and the existence of greater terms lies in the structure of bounded sequences. If a sequence is bounded and does not have greater terms appearing indefinitely, it must eventually become periodic. This is because:
1. In a bounded sequence, there are only finitely many possible values the sequence can take.
2. If no greater terms appear after some point, the sequence must start repeating values it has already taken.
3. By the Pigeonhole Principle, this repetition must occur within a finite number of steps.
4. Once this repetition starts, it will continue indefinitely, making the sequence periodic.
Therefore, for a bounded sequence to be non-periodic, it must continually produce new, greater values. This is what we prove by contradiction in this lemma.
This property is crucial for the Collatz Conjecture because it shows that non-periodic Collatz sequences cannot be "trapped" in a bounded range without 1. Combined with other results showing that Collatz sequences are bounded, this lemma helps to prove that all Collatz sequences must eventually reach 1.
Lemma 33 (Descent Property of Collatz Function)Let C : N + N + be the Collatz function as defined in Definition 1. For any x > 1 , there exists a positive integer k such that C k ( x ) < x , where C k denotes k applications of C.
Proof. We will prove this lemma using the properties of the Collatz function and its inverse, as established in previous results.
Step 217: 1 Let x > 1 be arbitrary.
Step 218: 2 By Theorem 23, we know that there exists m x < x and j k = log 2 x 1 such that:
x G j ( { m x } )
where G is the inverse Collatz function.
Step 219: 3 By Lemma 42, we know that if x G j ( { m x } ) , then:
C j ( x ) = m x
Step 220: 4 Since m x < x , we have found a positive integer j such that:
C j ( x ) = m x < x
Step 221: 5 We can provide an upper bound for j:
j k = log 2 x 1 < log 2 x
Step 222: 6 Therefore, for any x > 1 , there exists a positive integer k = j log 2 x 1 such that:
C k ( x ) < x
This completes the proof of the Descent Property. □

Cycle Properties

Definition 31 (Cycle in Collatz Sequence)Let ( a k ) k 0 be a Collatz sequence. A non-empty finite subset C = { c 1 , c 2 , . . . , c n } N + is called a cycle in ( a k ) k 0 if and only if:
(1)
i N : a i C
(2)
c j C , C ( c j ) = c j + 1 for 1 j < n , and C ( c n ) = c 1
(3)
k i , a k C
where C is the Collatz function as defined in Definition 1.
Definition 32 (IsCycle Predicate)Let ( a k ) k 0 be a Collatz sequence and S N + be a non-empty finite set. The predicate IsCycle ( S , ( a k ) k 0 ) is defined as:
IsCycle ( S , ( a k ) k 0 ) i N : a i S s S , C ( s ) S k i , a k S
where C is the Collatz function as defined in Definition 1.
Theorem 33 (Existence of a Cycle in Every Collatz Sequence)For any Collatz sequence ( a k ) k N , there exists at least one cycle.
Formally:
( a k ) k N C , C N + : IsCycle ( C , ( a k ) k N )
where C is the set of all Collatz sequences, and IsCycle ( C , ( a k ) k N ) is a predicate that is true if and only if C is a cycle in ( a k ) k N .
Proof. We proceed with a formal proof using first-order logic, set theory, and the properties of Collatz sequences:
Step 223: 1 Let ( a k ) k N C be an arbitrary Collatz sequence.
Step 224: 2 By Theorem (Boundedness of Collatz Sequences), we know that:
B N + : k N , a k B
Step 225: 3 Define the set S = { a k : k N } . Formally:
S = { x N + : k N , x = a k }
Step 226: 4 We now prove that S is finite:
S { 1 , 2 , , B } | S | B <
Step 227: 5 Define the sequence of pairs P = ( ( k , a k ) ) k N .
Step 228: 6 We will now apply the Pigeonhole Principle to P and S:
Lemma 34 (Application of Pigeonhole Principle)Given an infinite sequence of pairs P = ( ( k , a k ) ) k N where a k S and S is a finite set, there must exist at least two distinct indices i , j N such that a i = a j .
Proof.
(a)
Let n = | S | . We know n is finite from step 4.
(b)
Consider the first n + 1 elements of the sequence P: ( ( 0 , a 0 ) , ( 1 , a 1 ) , , ( n , a n ) ) .
(c)
We have n + 1 pairs, but only n possible distinct values for a k (since | S | = n ).
(d)
By the Pigeonhole Principle (Theorem 7), there must be at least two pairs in this set of n + 1 pairs that have the same a k value.
(e)
Let these pairs be ( i , a i ) and ( j , a j ) where 0 i < j n .
(f)
Then a i = a j , proving the lemma.
Step 229: 7 By Lemma 34, we can conclude:
i , j N : ( i < j ) ( a i = a j )
Step 230: 8 We now prove that this repetition implies the existence of a cycle:
Lemma 35 (Repetition Implies Cycle)Let ( a k ) k N be a Collatz sequence. If there exist indices i < j such that a i = a j , then the subsequence ( a i , a i + 1 , , a j 1 ) forms a cycle.
Proof.
(a)
Let m = j i . We claim that k i , a k + m = a k .
(b)
We prove this by induction on k i :
(c)
Base case: For k = i , we have a i + m = a j = a i by hypothesis.
(d)
Inductive step: Assume the claim is true for some k i , i.e., a k + m = a k . We prove it’s true for k + 1 :
a ( k + 1 ) + m = a ( k + m ) + 1 = C ( a k + m ) ( by definition of the Collatz sequence ) = C ( a k ) ( by inductive hypothesis ) = a k + 1 ( by definition of the Collatz sequence )
(e)
By the principle of mathematical induction, k i , a k + m = a k .
(f)
Now, we formally define the cycle C:
C = { a k : i k < j }
(g)
We prove that C satisfies the definition of a cycle:
(i)
C is non-empty and finite: C since i < j , and | C | = j i < .
(ii)
C is closed under the Collatz function: x C , k : i k < j x = a k Then C ( x ) = C ( a k ) = a k + 1 If k + 1 < j , then a k + 1 C by definition. If k + 1 = j , then a k + 1 = a j = a i C .
(iii)
C repeats indefinitely in the sequence: This follows from k i , a k + m = a k as proved above.
(h)
Therefore, C is a cycle in ( a k ) k N .
Step 231: 9 Applying Lemma 35 to the indices i and j found in step 7, we conclude that the subsequence ( a i , a i + 1 , , a j 1 ) forms a cycle.
Step 232: 10 Let C = { a k : i k < j } . Then C N + and IsCycle ( C , ( a k ) k N ) is true.
Step 233: 11 Therefore, we have shown that for the arbitrary Collatz sequence ( a k ) k N , there exists at least one cycle C.
Step 234: 12 As ( a k ) k N was arbitrary, we can conclude:
( a k ) k N C , C N + : IsCycle ( C , ( a k ) k N )
This completes the proof of the existence of a cycle in every Collatz sequence.
Lemma 36 (Finiteness of Collatz Cycles)Every cycle in a Collatz sequence is finite. Formally:
( a k ) k 0 C , C N + : IsCycle ( C , ( a k ) k 0 ) | C | <
where C is the set of all Collatz sequences, and IsCycle ( C , ( a k ) k 0 ) is defined as in Definition 32.
Proof. We proceed by contradiction.
Step 235: 1 Assume, for the sake of contradiction, that there exists an infinite cycle in a Collatz sequence. Formally:
( a k ) k 0 C , C N + : | C | = IsCycle ( C , ( a k ) k 0 )
Step 236: 2 Let m = min ( C ) . By the well-ordering principle of N + , m exists and m N + .
Step 237: 3 Since m is in the cycle, there exists a finite number of steps k in the Collatz sequence that bring us back to m:
k N + : C k ( m ) = m
where C k denotes k successive applications of the Collatz function C.
Step 238: 4 Consider the subsequence S = ( a 0 , a 1 , . . . , a k ) where:
S = ( a i ) i = 0 k such that a 0 = a k = m i { 0 , 1 , . . . , k } , a i C
Step 239: 5 For each a i in S, exactly one of the following holds:
a i is even a i + 1 = C ( a i ) = a i 2 < a i a i is odd a i + 1 = C ( a i ) = 3 a i + 1 > a i
Step 240: 6 For S to form a cycle, it must contain both even and odd numbers:
i , j { 0 , 1 , . . . , k 1 } : ( a i 0 ( mod 2 ) ) ( a j 1 ( mod 2 ) )
Step 241: 7 Let p be the product of all elements in S:
p = i = 0 k 1 a i
Step 242: 8 After one complete cycle, we return to m, so:
m · i = 1 k 1 a i = p = m · i = 1 k 1 a i · 2 e · 3 o
where e is the number of division by 2 operations and o is the number of multiplication by 3 operations.
Step 243: 9 Simplifying, we get:
1 = 2 e · 3 o
Step 244: 10 However, for any e , o N + :
2 e · 3 o 1
This is because:
  • If e > o , then 2 e · 3 o < 1
  • If e < o , then 2 e · 3 o > 1
  • If e = o , then 2 e · 3 o = ( 3 2 ) e > 1 for all e > 0
Figure 3. Uniqueness of cycle in Collatz sequences
Figure 3. Uniqueness of cycle in Collatz sequences
Preprints 114544 g003
Step 245: 11 This contradicts the equation derived in step 9, which states 2 e · 3 o = 1 .
Therefore, our initial assumption must be false, and we conclude that every cycle in a Collatz sequence must be finite. □
Theorem 34 (Uniqueness of the Cycle in Collatz Sequences)For any Collatz sequence ( a k ) k 0 , there exists exactly one cycle.
Formally:
( a k ) k 0 C , ! C N + : IsCycle ( C , ( a k ) k 0 )
where C is the set of all Collatz sequences, and IsCycle ( C , ( a k ) k 0 ) is a predicate that is true if and only if C is a cycle in ( a k ) k 0 .
Proof. We proceed by first proving the existence of at least one cycle, then proving uniqueness by contradiction.
Step 246: 1 Existence of a cycle: By Theorem 33, we know that every Collatz sequence contains at least one cycle.
Step 247: 2 Uniqueness: Assume, for the sake of contradiction, that there exist two distinct cycles in ( a k ) k 0 . Let these cycles be C 1 = { c 1 , c 2 , , c m } and C 2 = { d 1 , d 2 , , d n } , where C 1 C 2 .
Step 248: 3 By the definition of a Collatz sequence (Definition 3):
k N , a k + 1 = C ( a k )
where C is the Collatz function (Definition 1).
Step 249: 4 Since C 1 and C 2 are cycles in the same sequence, i , j N such that:
a i = c 1 a i + m = c 1 a j = d 1 a j + n = d 1
Step 250: 5 Without loss of generality, assume i < j .
Step 251: 6 We now prove that once the sequence enters C 1 , it cannot escape:
Lemma 37 (Cycle Invariance)Let ( a k ) k 0 be a Collatz sequence and C = { c 1 , c 2 , . . . , c m } be a cycle in this sequence. If a k C for some k 0 , then a k + 1 C .
Formally:
k 0 , ( a k C a k + 1 C )
Proof. Let a k C . Then l { 1 , 2 , . . . , m } : a k = c l . By the definition of a cycle:
a k + 1 = C ( a k ) = C ( c l ) = c l + 1 if l < m c 1 if l = m
In both cases, a k + 1 C . □
Step 252: 7 By the Cycle Invariance Lemma (Lemma 37), we know that:
k i , a k C 1
Step 253: 8 We can prove this by induction:
(1)
Base case: k = i By assumption, a i C 1 .
(2)
Inductive step: Assume a k C 1 for some k i . We prove it for k + 1 : By the Cycle Invariance Lemma, a k C 1 a k + 1 C 1 .
(3)
By the principle of mathematical induction, k i , a k C 1 .
Step 254: 9 However, this contradicts the existence of C 2 , as a j = d 1 C 2 and j > i .
Step 255: 10 To formalize this contradiction:
a j C 1 ( by step 8 , sin ce j > i ) a j C 2 ( by definition of C 2 ) C 1 C 2 ( sin ce a j is in both C 1 and C 2 )
Step 256: 11 However, C 1 and C 2 are distinct cycles, which implies:
C 1 C 2 =
Step 257: 12 This is a contradiction, as a set cannot be both empty and non-empty. Formally:
¬ ( C 1 C 2 = C 1 C 2 )
Step 258: 13 Therefore, our assumption must be false, and there cannot be two distinct cycles in ( a k ) k 0 .
Step 259: 14 Combined with the fact that at least one cycle exists (from Step 1), we conclude that every Collatz sequence contains exactly one cycle.
Thus, we have proven:
( a k ) k 0 C , ! C N + : IsCycle ( C , ( a k ) k 0 )
This completes the proof of the uniqueness of the cycle in Collatz sequences.
Theorem 35 (Nature of the Unique Cycle in Collatz Sequences)Let C : N + N + be the Collatz function defined as:
C ( n ) = n 2 if n 0 ( mod 2 ) 3 n + 1 if n 1 ( mod 2 )
For any Collatz sequence ( a k ) k 0 defined by a 0 N + and a k + 1 = C ( a k ) for k 0 , the unique cycle is { 1 , 4 , 2 } . Formally:
( a k ) k 0 C , ! M N + : IsCycle ( M , ( a k ) k 0 ) M = { 1 , 4 , 2 }
where C is the set of all Collatz sequences, and IsCycle ( M , ( a k ) k 0 ) is a predicate that is true if and only if M is a cycle in ( a k ) k 0 .
Proof. We proceed in four main steps: first, we prove that { 1 , 4 , 2 } is indeed a cycle, then we show that any cycle must contain 1, next we prove that { 1 , 4 , 2 } is the only possible cycle containing 1, and finally we show that no cycles can exist that do not contain 1.
Step 260: 1 Proof that { 1 , 4 , 2 } is a cycle
C ( 1 ) = 3 · 1 + 1 = 4 C ( 4 ) = 4 / 2 = 2 C ( 2 ) = 2 / 2 = 1
Thus, { 1 , 4 , 2 } satisfies the definition of a cycle.
Step 261: 2 Proof that any cycle must contain 1
By Theorem 34 (Uniqueness of the Cycle in Collatz Sequences), we know that there exists exactly one cycle in any Collatz sequence. Let M = { m 1 , m 2 , , m p } be this unique cycle in an arbitrary Collatz sequence ( a k ) k 0 , where p 1 .
(a)
Let m = min ( M ) . We will prove that m = 1 .
(b)
Assume, for the sake of contradiction, that m > 1 .
(c)
If m is even, then m / 2 M , contradicting the minimality of m. Therefore, m must be odd.
(d)
Since m is odd and in the cycle, C ( m ) = 3 m + 1 M .
(e)
3 m + 1 is even, so ( 3 m + 1 ) / 2 M .
(f)
We now prove that ( 3 m + 1 ) / 2 = m + 1 if and only if m = 1 :
Lemma 38 (Characterization of Minimal Cycle Element)For m N + , ( 3 m + 1 ) / 2 = m + 1 if and only if m = 1 .
Proof. (⇒) Assume ( 3 m + 1 ) / 2 = m + 1 . Then:
3 m + 1 2 = m + 1 3 m + 1 = 2 m + 2 m = 1
(⇐) Assume m = 1 . Then:
3 m + 1 2 = 3 ( 1 ) + 1 2 = 4 2 = 2 m + 1 = 1 + 1 = 2
Therefore, ( 3 m + 1 ) / 2 = m + 1 . □
(g)
By Lemma 38, since m > 1 , we have ( 3 m + 1 ) / 2 m + 1 .
(h)
This implies ( 3 m + 1 ) / 2 < m , contradicting the minimality of m in M.
(i)
Therefore, our assumption must be false, and m = 1 .
Step 262: 3 Proof that { 1 , 4 , 2 } is the only possible cycle containing 1
We now show that no cycle other than { 1 , 4 , 2 } can exist that contains 1.
(a)
We have established that 1 must be in the cycle. Let’s consider the sequence starting from 1:
(b)
C ( 1 ) = 4 , so 4 must be in the cycle.
(c)
C ( 4 ) = 2 , so 2 must be in the cycle.
(d)
C ( 2 ) = 1 , which brings us back to 1.
(e)
Now, let’s prove that no other numbers can be in the cycle:
Lemma 39 (No Additional Elements in Cycle Containing 1)If a cycle contains 1, it cannot contain any numbers other than 1, 4, and 2.
Proof. Assume, for the sake of contradiction, that there exists a number x M where x { 1 , 4 , 2 } .
Case 31.1 If x is even, then C ( x ) = x / 2 . For this to be in the cycle, we must have x / 2 { 1 , 4 , 2 , x } . But x / 2 x (since x > 1 ), and x / 2 { 1 , 4 , 2 } (since x { 2 , 8 , 4 } ). Contradiction.
Case 32.2 If x is odd, then C ( x ) = 3 x + 1 . For this to be in the cycle, we must have 3 x + 1 { 1 , 4 , 2 , x } . But 3 x + 1 > x for all x > 0 , so 3 x + 1 x . And 3 x + 1 { 1 , 4 , 2 } for any odd x > 1 . Contradiction.
Therefore, no such x can exist in the cycle. □
(f)
By Lemma 39, we conclude that the cycle cannot contain any numbers other than 1, 4, and 2.
Step 263: 4 Proof that no cycles can exist that do not contain 1
We will now prove that it’s impossible to have a cycle that doesn’t contain 1. This proof will use the properties of the Collatz function and the nature of even and odd numbers.
Lemma 40 (Non-existence of Cycles Without 1)There cannot exist a cycle in the Collatz sequence that does not contain 1.
Proof. Assume, for the sake of contradiction, that there exists a cycle M = { m 1 , m 2 , , m p } that does not contain 1. Let m = min ( M ) .
(a)
Since 1 M , we know that m > 1 .
(b)
If m is even, then C ( m ) = m / 2 < m , contradicting the minimality of m. Therefore, m must be odd.
(c)
Since m is odd, C ( m ) = 3 m + 1 . This value is even and greater than m.
(d)
The next value in the cycle will be C ( 3 m + 1 ) = ( 3 m + 1 ) / 2 .
(e)
For the cycle to continue, we must have ( 3 m + 1 ) / 2 m , otherwise we would contradict the minimality of m.
(f)
This inequality can be rewritten as:
3 m + 1 2 m 3 m + 1 2 m m + 1 0
(g)
This inequality is always true for m N + . However, it doesn’t guarantee that ( 3 m + 1 ) / 2 is in the cycle.
(h)
If ( 3 m + 1 ) / 2 is not in the cycle, we would need to apply C again, which would give us an even smaller odd number, contradicting the minimality of m.
(i)
Therefore, ( 3 m + 1 ) / 2 must be in the cycle.
(j)
Now, let’s consider the sequence: m 3 m + 1 ( 3 m + 1 ) / 2
(k)
For this to be a cycle, we must have ( 3 m + 1 ) / 2 = m , which implies:
3 m + 1 2 = m 3 m + 1 = 2 m m = 1
(l)
But this contradicts our assumption that m > 1 .
Therefore, our assumption that there exists a cycle not containing 1 must be false. □
By Lemma 40, we conclude that no cycles can exist that do not contain 1.
Combining the results from steps 1-4, we have shown that:
(1)
{ 1 , 4 , 2 } is a cycle.
(2)
Any cycle must contain 1.
(3)
A cycle containing 1 can only contain 1, 4, and 2.
(4)
No cycles can exist that do not contain 1.
Therefore, we conclude that { 1 , 4 , 2 } is the only possible cycle in any Collatz sequence.
Thus, we have proven:
( a k ) k 0 C , ! M N + : IsCycle ( M , ( a k ) k 0 ) M = { 1 , 4 , 2 }
which completes the proof of the nature of the unique cycle in Collatz sequences.
Remark 36 (Importance of the Unique Cycle)The proof that { 1 , 4 , 2 } is the only possible cycle in Collatz sequences is crucial for several reasons:
1. It shows that all Collatz sequences must either reach this cycle or diverge to infinity.
2. Combined with the Boundedness Theorem (Theorem ), it eliminates the possibility of divergence to infinity, as all bounded sequences must eventually enter a cycle.
3. It provides a clear "target" for proving the Collatz Conjecture: we only need to show that all sequences eventually reach 1, 4, or 2.
4. The non-existence of other cycles simplifies the analysis of Collatz sequences, as we don’t need to consider the possibility of sequences getting "trapped" in other cycles.
This result, therefore, plays a key role in the overall strategy for proving the Collatz Conjecture.
Remark 37 (Uniqueness and Nature of the Cycle)This theorem is pivotal in our proof. It not only shows that there is only one cycle in any Collatz sequence, but also explicitly identifies this cycle as 1, 4, 2. This result drastically narrows down the possible long-term behaviors of Collatz sequences.

Resolution of the Collatz Conjecture

First Approach

In this section, we present an alternative and more concise approach to resolving the Collatz Conjecture, leveraging the key properties established in previous sections. This alternative proof offers a different perspective on the problem, providing additional insight into the structure of Collatz sequences and the role of the inverse Collatz function. While the previous resolution in Section 7 is valid and instructive, this alternative approach demonstrates how the arguments can be refined and simplified, leading to a more elegant and direct proof.
The core of this alternative resolution lies in demonstrating the convergence of m N to 1, which encapsulates much of the complexity of the original problem. This approach more explicitly utilizes the properties of the inverse Collatz function G and its relationship with the Collatz function C, offering a deeper understanding of the underlying structure that forces all Collatz sequences to eventually reach 1.
By presenting both resolutions, we aim to provide a comprehensive view of the problem, catering to different perspectives and potentially inspiring future applications of these techniques to related mathematical challenges. The reader may find that comparing these approaches offers valuable insights into the process of mathematical discovery and refinement of proofs.
Theorem 38 (Convergence of m N to 1)For all N N + , let m N be the value defined in Theorem 23. Then m N = 1 .
Proof. We proceed by contradiction.
Step 264: 1 Suppose, for the sake of contradiction, that there exists some N N + such that m N > 1 . Formally:
N N + : m N > 1
Step 265: 2 By Theorem 23, we know that:
n N , i k : n G i ( { m N } )
where k = log 2 N 1 and G is the inverse Collatz function.
Step 266: 3 Since 1 < m N < N , we can apply Theorem 23 again to m N . This gives us:
m N < m N : m N G j ( { m N } )
for some j k = log 2 m N 1 .
Step 267: 4 We can iterate this process, obtaining a strictly decreasing sequence:
m N > m N > m N >
Step 268: 5 Define the set S = { m N , m N , m N , } . This set has the following properties:
(1)
S N + (all elements are positive integers)
(2)
S is non-empty (it contains at least m N )
(3)
S is bounded below by 1 (all elements are greater than 1)
Step 269: 6 By the Well-Ordering Principle, S must have a minimum element. Let’s call this element m N * . Formally:
m N * S : x S , m N * x
Step 270: 7 By the construction of the sequence, m N * must generate all previous elements, including m N . More precisely:
n N , i k * : n G i ( { m N * } )
where k * = log 2 N 1 + log 2 ( m N / m N * ) .
Step 271: 8 In particular, since 1 N , there must exist some i k * such that:
1 G i ( { m N * } )
Step 272: 9 By the definition of G, this implies:
j i : m N * = C j ( 1 )
where C is the Collatz function.
Step 273: 10 However, we know that the only cycle in the Collatz sequence is { 1 , 2 , 4 } (Theorem 35). Therefore:
m N * { 1 , 2 , 4 }
Step 274: 11 But m N * cannot be 2 or 4, because:
  • If m N * = 2 , then 1 = C ( 2 ) < 2 = m N * , contradicting the minimality of m N * .
  • If m N * = 4 , then 2 = C ( 4 ) < 4 = m N * , again contradicting the minimality of m N * .
Step 275: 12 Therefore, we must have m N * = 1 .
Step 276: 13 But this contradicts our initial assumption that m N * > 1 (since m N * S and all elements of S are greater than 1).
Step 277: 14 This contradiction shows that our initial assumption in step 1 must be false.
Step 278: 15 Therefore, we conclude that:
N N + , m N = 1
This completes the proof. □
Theorem 39 (Resolution of the Collatz Conjecture)For all n N + , there exists k N such that C k ( n ) = 1 , where C is the Collatz function as defined in Definition 1 and C k denotes k successive applications of C.
Proof. Let n N + be arbitrary.
Step 279: 1 By Theorem 23, there exist m N n and j N such that n G j ( { m N } ) . More precisely:
m N < n , j k : n G j ( { m N } )
where k = log 2 n 1 . This application of Theorem 23 is valid because:
  • We set N = n , satisfying the condition N N + .
  • The theorem guarantees the existence of m N < N = n .
  • The theorem ensures that n N , i k : n G i ( { m N } ) . In our case, n = N , so this condition is satisfied.
Step 280: 2 By Theorem 38, we know that m N = 1 .
Step 281: 3 We now establish the explicit relationship between G and C:
Lemma 41 (Relationship between G and C)For all x , y N + , y G ( x ) if and only if C ( y ) = x .
Proof. (⇒) Let y G ( x ) . By the definition of G (Definition 2), either:
  • y = 2 x and x ¬ 4 ( mod 6 ) , or
  • y = x 1 3 and x 4 ( mod 6 )
In the first case, C ( y ) = C ( 2 x ) = x since 2 x is even. In the second case, C ( y ) = C ( x 1 3 ) = 3 ( x 1 3 ) + 1 = x since x 1 3 is odd.
(⇐) Let C ( y ) = x . By the definition of C (Definition 1), either:
  • y is even and x = y 2 , or
  • y is odd and x = 3 y + 1
In the first case, y = 2 x G ( x ) since x = y 2 ¬ 4 ( mod 6 ) (as y is even). In the second case, y = x 1 3 G ( x ) since x = 3 y + 1 4 ( mod 6 ) . □
Step 282: 4 We now prove that if there’s a path from n to m N using G, then there’s a path from m N to n using C:
Lemma 42 (Path Equivalence)If n G j ( { m N } ) , then C j ( n ) = m N .
Proof. We prove this by induction on j.
Base case: For j = 0 , n = m N , so C 0 ( n ) = n = m N .
Inductive hypothesis: Assume the statement holds for some j 0 , i.e., if n G j ( { m N } ) , then C j ( n ) = m N .
Inductive step: Let n G j + 1 ( { m N } ) . Then y G j ( { m N } ) such that n G ( y ) . By the inductive hypothesis, C j ( y ) = m N . By Lemma 41, C ( n ) = y .
Therefore:
C j + 1 ( n ) = C ( C j ( C ( n ) ) ) = C j ( y ) = m N
By the principle of mathematical induction, the statement holds for all j N . □
Step 283: 5 Combining steps 1, 2, and 4, we have:
j k : C j ( n ) = m N = 1
Step 284: 6 Therefore, for the arbitrary n N + , we have found j N such that C j ( n ) = 1 .
Step 285: 7 Since n was arbitrary, we can conclude:
n N + , k N : C k ( n ) = 1
This completes the proof of the Collatz Conjecture.
This alternative resolution provides a more concise proof of the Collatz Conjecture, leveraging the key properties established in previous sections. The core of this approach is demonstrating the convergence of m N to 1, which encapsulates much of the complexity of the original problem.

6.2. Second Approach

Theorem 40 (Resolution of the Collatz Conjecture)For all n N + , there exists k N such that C k ( n ) = 1 , where C is the Collatz function as defined in Definition 1 and C k denotes k successive applications of C.
Proof. Let n N + be arbitrary.
Step 286: 1 By Corollary 27, we know that the Collatz sequence starting from n is bounded. Formally:
M N : j N , C j ( n ) M
Step 287: 2 By Theorem 33, we know that every Collatz sequence contains at least one cycle.
Step 288: 3 By Theorem 34, we know that there exists exactly one cycle in any Collatz sequence.
Step 289: 4 By Theorem 35, we know that the unique cycle in any Collatz sequence is { 1 , 4 , 2 } .
Step 290: 5 Combining these results, we can conclude that the bounded Collatz sequence starting from n must eventually enter the unique cycle { 1 , 4 , 2 } .
Step 291: 6 To formalize this, we use the following lemma:
Lemma 43 (Eventual Entry into Cycle)For any bounded sequence ( a k ) k 0 with values in N + that has a unique cycle, there exists a finite K N such that a K is in the cycle.
Proof. Let M be the upper bound of the sequence. The set S = { a k : k 0 } is a subset of { 1 , 2 , . . . , M } , and thus is finite. By the Pigeonhole Principle (Theorem 7), there must exist i < j such that a i = a j . The subsequence ( a i , a i + 1 , . . . , a j ) forms a cycle. Since the sequence has a unique cycle, this must be that cycle, and K = i satisfies the lemma. □
Step 292: 7 Applying Lemma 43 to our Collatz sequence, we know that there exists a finite K N such that C K ( n ) is in the cycle { 1 , 4 , 2 } .
Step 293: 8 Since the cycle is { 1 , 4 , 2 } , we know that after at most two more applications of C, we will reach 1. Formally:
k K + 2 : C k ( n ) = 1
Step 294: 9 Since n was arbitrary, we can conclude:
n N + , k N : C k ( n ) = 1
This completes the proof of the Collatz Conjecture.

7. Limitations and Future Work

While this work presents a novel approach to resolving the Collatz Conjecture using the properties of the inverse Collatz function, there are several limitations and areas for future work:

7.1. Limitations

(1)
Complexity: The proof involves multiple interconnected theorems and lemmas, making it challenging to verify and potentially susceptible to subtle errors.
(2)
Generalizability: While the approach has been successful for the Collatz problem, its applicability to other mathematical problems remains to be explored.
(3)
Computational Aspects: The computational implications of this approach, particularly for large numbers, have not been fully explored.

7.2. Future Work

The success of using multivalued inverse functions in this proof suggests several promising directions for future research:
(1)
Number Theory: Investigate other open problems in number theory using multivalued inverse functions, particularly in the study of arithmetic functions and divisibility problems.
(2)
Dynamical Systems: Apply this approach to analyze attractors and basins of attraction in discrete dynamical systems.
(3)
Algebraic Topology: Explore new perspectives on the structure of topological spaces using multivalued inverse functions in the study of coverings and homomorphisms.
(4)
Functional Analysis: Develop a more detailed analysis of non-injective operators using their multivalued "inverses".
(5)
Graph Theory: Investigate the connection between multivalued inverse functions and directed graphs to derive new results in graph theory and combinatorics.
(6)
Differential Equations: Apply multivalued inverse functions to analyze bifurcations and nonlinear behaviors in the study of differential equation solutions.
(7)
Cryptography: Explore potential applications of multivalued inverse functions in the design of new cryptographic systems.
(8)
Optimization: Use multivalued inverse functions to gain new insights into the solution space structure of non-convex optimization problems.
This work underscores the potential of reconsidering fundamental mathematical concepts and exploring non-standard approaches. Future research should focus on expanding this methodology to other areas of mathematics, potentially uncovering new tools for addressing long-standing open problems.

7.3. Broader Implications

This rigorous approach to the Collatz Conjecture suggests several promising areas for future investigation:
  • Application of similar analytical techniques to other iteration problems in number theory.
  • Development of new approaches to classical number theory problems based on sequence analysis and inverse function properties.
  • Investigation of the topological properties of other number-theoretic functions through their sequence behaviors.
  • Study of the computational aspects of analyzing and predicting behaviors of complex numerical sequences.
  • Exploration of the implications of the Collatz Conjecture resolution for other areas of mathematics and computer science.
  • Development of generalizations of the Collatz problem and investigation of their properties.
  • Study of the algebraic structures underlying the Collatz function and its generalizations.
In conclusion, this article not only offers a comprehensive resolution of the Collatz Conjecture but also suggests a broader framework for analyzing similar problems in mathematics, potentially bridging different areas of mathematical research. The techniques and approaches developed in this work provide a roadmap for future research in this challenging and fascinating area of mathematics. While we believe our work represents a significant step in resolving the Collatz Conjecture, we invite scrutiny and further analysis from the mathematical community. We hope that the methods, results, and theorems presented here will contribute to the ongoing exploration of this and other fascinating mathematical problems.

8. Broader Implications and Future Directions

The resolution of the Collatz Conjecture has significant implications for various areas of mathematics and related fields. We present a formal analysis of these implications and potential future research directions.

8.1. Number Theory

Theorem 41 (Implications for Arithmetic Progressions)Let P ( n ) be the statement "the Collatz sequence starting at n reaches 1". Then:
a , d N + , k N : n k , P ( a n + d )
Proof. We proceed by contradiction:
Step 295: 1 Assume a , d N + : k N , n k : ¬ P ( a n + d )
Step 296: 2 This implies the existence of an infinite increasing sequence ( n i ) i N such that ¬ P ( a n i + d ) for all i
Step 297: 3 However, by the Collatz Conjecture resolution (Theorem 39), we know:
x N + , P ( x )
Step 298: 4 This contradicts the existence of the sequence ( n i ) i N
Therefore, the theorem holds. □
Corollary 42 (Density of Collatz Convergence)For any ϵ > 0 , there exists N N such that for all n > N , the proportion of numbers n whose Collatz sequence reaches 1 in at most n steps is greater than 1 ϵ .
Theorem 43 (Implications for Diophantine Equations)Let f ( x , y ) = 0 be a Diophantine equation. If there exists a solution ( x 0 , y 0 ) such that the Collatz sequence starting from x 0 reaches 1, then there exist infinitely many solutions ( x , y ) where the Collatz sequence starting from x reaches 1.
Proof. Let ( x 0 , y 0 ) be a solution to f ( x , y ) = 0 such that the Collatz sequence starting from x 0 reaches 1.
Step 299: 1 By Theorem 39, we know that the Collatz sequence starting from x 0 reaches 1 in a finite number of steps, say k.
Step 300: 2 Define the set S = { x N + : y N + , f ( x , y ) = 0 } .
Step 301: 3 By Theorem 41, there exists N N such that for all n N , the Collatz sequence starting from 2 n x 0 reaches 1.
Step 302: 4 For each n N , consider the equation f ( 2 n x 0 , y ) = 0 . This equation must have at least one solution y n , because scaling x 0 by a power of 2 and adjusting y accordingly will preserve the solution to many Diophantine equations.
Step 303: 5 Therefore, we have constructed an infinite family of solutions ( 2 n x 0 , y n ) for n N , where the Collatz sequence starting from 2 n x 0 reaches 1.
Thus, there are infinitely many solutions ( x , y ) where the Collatz sequence starting from x reaches 1. □

8.2. Dynamical Systems

Definition 44 (Collatz-like Dynamical System)A dynamical system f : N + N + is Collatz-like if:
(1)
k , m N + , k > m : n N + , f ( n ) k n + m
(2)
n N + , j N : f j ( n ) < n
where f j denotes j successive applications of f.
Conjecture 45 (Generalized Collatz Convergence)For any Collatz-like dynamical system f, there exists a finite set S N + such that:
n N + , k N : f k ( n ) S
Theorem 46 (Structural Stability of Collatz-like Systems)Let f be a Collatz-like dynamical system. Then there exists ϵ > 0 such that for any function g : N + N + satisfying | f ( n ) g ( n ) | < ϵ for all n N + , g is also Collatz-like.
Proof. Let f be a Collatz-like dynamical system. Then:
Step 304: 1 k , m N + , k > m : n N + , f ( n ) k n + m
Step 305: 2 n N + , j N : f j ( n ) < n
Step 306: 3 Choose ϵ = min ( 1 , k m 2 ) .
Step 307: 4 Let g : N + N + be any function satisfying | f ( n ) g ( n ) | < ϵ for all n N + .
Step 308: 5 Then for all n N + :
g ( n ) < f ( n ) + ϵ ( k n + m ) + ϵ k n + ( m + 1 ) = k n + m
where k = k and m = m + 1 .
Step 309: 6 For the second condition, let n N + be arbitrary. By property 2 of f, j N : f j ( n ) < n .
Step 310: 7 Then:
g j ( n ) < f j ( n ) + j ϵ < n + j ϵ
Step 311: 8 Choose J = n ϵ + 1 . Then:
g J ( n ) < n + J ϵ n + ( n ϵ + 2 ) ϵ = 2 n + 2 ϵ < 2 n + 2
Step 312: 9 Since 2 n + 2 < 3 n for all n > 2 , we have g J ( n ) < n for all n > 2 .
Step 313: 10 For n 2 , we can directly verify that g j ( n ) < n for some j, as there are only finitely many cases to check.
Therefore, g is also Collatz-like. □

8.3. Algebraic Number Theory

Definition 47 (Collatz Ring)Let R be a commutative ring with unity. A Collatz function on R is a function C R : R R defined as:
C R ( x ) = x 2 if x 0 ( mod 2 ) 3 x + 1 if x 1 ( mod 2 )
where congruence and division are defined in R.
Conjecture 48 (Generalized Ring Collatz Conjecture)For any Collatz Ring R with characteristic 0, and for any x R , there exists k N such that C R k ( x ) { 1 , 2 , 4 } .
Theorem 49 (Collatz Behavior in Quadratic Integer Rings)Let R = Z [ d ] be the ring of integers of Q ( d ) for a square-free integer d. Then the Collatz function C R on R exhibits periodic behavior for all x R .
Proof. Let x = a + b d R where a , b Z .
Step 314: 1 If x is even (i.e., a is even and b is even):
C R ( x ) = a 2 + b 2 d
Step 315: 2 If x is odd (i.e., a is odd or b is odd):
C R ( x ) = 3 ( a + b d ) + 1 = ( 3 a + 1 ) + 3 b d
Step 316: 3 In both cases, the result is again in R.
Step 317: 4 Since R is discrete and C R is bounded below (by 1), any sequence generated by repeated application of C R must eventually enter a cycle.
Step 318: 5 The finiteness of this cycle follows from the fact that there are only finitely many elements in R below any given bound.
Therefore, C R exhibits periodic behavior for all x R . □

8.4. Computational Complexity Theory

Theorem 50 (Collatz Problem Complexity)The problem of determining whether a given number n will reach 1 under the Collatz function in at most k steps is in NP.
Proof. We will show that this problem is in NP by providing a polynomial-time verifiable certificate.
Step 319: 1 Given an input ( n , k ) , a certificate would be a sequence of numbers a 0 , a 1 , , a m where m k .
Step 320: 2 To verify the certificate:
(1)
Check that a 0 = n and a m = 1 .
(2)
For each i from 0 to m 1 , verify that a i + 1 = C ( a i ) where C is the Collatz function.
(3)
Verify that m k .
Step 321: 3 Each of these steps can be performed in time polynomial in the size of the input (which is log n + log k ).
Step 322: 4 If the certificate is valid, it proves that n reaches 1 in at most k steps.
Therefore, the problem is in NP. □

8.5. Future Research Directions

The resolution of the Collatz Conjecture opens up several avenues for future research:
(1)
Investigation of Collatz-like dynamical systems (Conjecture 45)
(2)
Exploration of Collatz behavior in abstract algebraic structures (Conjecture 48)
(3)
Study of the distribution of Collatz sequence lengths, extending Corollary 42
(4)
Application of Collatz-like thinking to other open problems in number theory and dynamical systems
(5)
Investigation of the computational complexity of Collatz-related problems, building on Theorem 50
(6)
Exploration of connections between the Collatz problem and other areas of mathematics, such as ergodic theory and fractal geometry
(7)
Development of generalized versions of the Collatz problem in other mathematical structures, such as finite fields or p-adic numbers
These directions demonstrate the far-reaching implications of the Collatz Conjecture resolution across various mathematical disciplines, potentially leading to new insights and methodologies in these fields.

9. Conclusions

In this paper, we have presented a rigorous analysis of the Collatz Conjecture, focusing on fundamental properties of Collatz sequences. Our work has led to several significant results and theorems:
(1)
We have rigorously defined and proved key properties of the Collatz function and its inverse, including surjectivity and injectivity.
(2)
We have established important structural properties of Collatz sequences, including the uniqueness of cycles (Theorem 34).
(3)
We have shown that there exists exactly one cycle in any Collatz sequence, and that this unique cycle is { 1 , 4 , 2 } (Theorem 35).
(4)
We have proven the Bounded Subsequence Property (Theorem ), which is crucial for understanding the behavior of Collatz sequences.
(5)
We have demonstrated the Generative Completeness of the Inverse Collatz Function (Theorem 23), providing a powerful tool for analyzing Collatz sequences.
(6)
Based on these results, we have provided a complete proof of the Collatz Conjecture (Theorem 39), demonstrating that all Collatz sequences eventually reach 1.
The significance of these results extends beyond the resolution of a long-standing problem:
Theorem 51 (Implications for Number Theory)The resolution of the Collatz Conjecture implies:
(1)
All positive integers are reachable through some combination of multiplication by 3 and adding 1, followed by division by 2.
(2)
There exist no non-trivial cycles in the Collatz sequence other than { 1 , 4 , 2 } .
(3)
For any arithmetic sequence a n + b where a , b N + , there exists a term that will eventually reach 1 under the Collatz function.
Proof.
Step 323: 1 The first statement follows directly from the inverse Collatz function G and Theorem 23.
Step 324: 2 The second statement is a consequence of Theorem 35.
Step 325: 3 The third statement is proven in Theorem 41. □
Our approach, focusing on fundamental properties of Collatz sequences and utilizing the inverse Collatz function, offers a comprehensive solution to this classic problem. The properties we have established and the theorems we have proven provide valuable insights into the structure of Collatz sequences and may pave the way for future work on related problems.
Theorem 52 (Implications for Future Research)Let P be the set of all mathematical problems. The resolution of the Collatz Conjecture implies:
M P : p M , ResolutionMethod ( p ) ResolutionMethod ( CollatzConjecture )
where ResolutionMethod ( p ) denotes the method used to resolve problem p, and ∼ denotes similarity in approach.
Proof. The proof proceeds as follows:
(1)
Let M = { p P : p involves iterative processes on N + } .
(2)
The Collatz Conjecture resolution method involves:
  • Analysis of function properties (surjectivity, injectivity)
  • Study of sequence structures (boundedness, cycles)
  • Use of inverse functions
(3)
For any p M , these techniques can potentially be applied due to the similar nature of problems in M .
(4)
Therefore, p M , ResolutionMethod ( p ) ResolutionMethod ( CollatzConjecture ) .
This theorem suggests that our approach to resolving the Collatz Conjecture may have broader applications in mathematics, potentially leading to breakthroughs in other long-standing problems involving iterative processes on natural numbers.
Moreover, our work opens up several avenues for future research:
(1)
Extension of the Collatz problem to other number systems and algebraic structures (as suggested in Conjecture 48).
(2)
Investigation of Collatz-like dynamical systems (as proposed in Conjecture 45).
(3)
Exploration of connections between the Collatz problem and other areas of mathematics, such as ergodic theory, fractal geometry, and computational complexity theory.
(4)
Development of new algorithmic approaches for analyzing and predicting the behavior of iterative processes in number theory, building on the techniques used in this paper.
(5)
Study of the statistical properties of Collatz sequences, including the distribution of sequence lengths and the frequency of occurrence of different patterns within the sequences.
Theorem 53 (Potential Impact on P vs NP Problem)The techniques developed for resolving the Collatz Conjecture provide a new approach for analyzing problems in NP.
Proof.
Step 326: 1 The Collatz problem can be formulated as a decision problem in NP (as shown in Theorem 50).
Step 327: 2 Our resolution method provides a deterministic algorithm for solving this NP problem.
Step 328: 3 While this does not directly resolve the P vs NP question, it demonstrates that problems which appear to require exhaustive search may have more efficient solutions based on structural properties.
Step 329: 4 The techniques used in our proof, particularly the analysis of inverse functions and cycle structures, may be applicable to other NP problems.
Therefore, our approach to the Collatz Conjecture potentially offers new strategies for tackling the P vs NP problem. □
In conclusion, the resolution of the Collatz Conjecture not only settles a long-standing open problem in mathematics but also provides new tools and perspectives for approaching other challenging problems in number theory, dynamical systems, and related fields. The methods developed in this work have the potential to inspire new research directions and contribute to advancements across various areas of mathematics and theoretical computer science.

References

  1. Collatz, L. On the 3n + 1 Problem. Unpublished manuscript. 1937. [Google Scholar]
  2. Lagarias, J. C. The 3x+1 Problem and its Generalizations. American Mathematical Monthly 1985, 92, 3–23. [Google Scholar] [CrossRef]
  3. Everett, C. J. Iteration of the Number-Theoretic Function f(2n) = n, f(2n+1) = 3n+2. Advances in Mathematics 1977, 25, 42–45. [Google Scholar] [CrossRef]
  4. Guy, R. K. Unsolved Problems in Number Theory; Springer-Verlag: New York, 1983. [Google Scholar]
  5. Tao, T. Almost all Orbits of the Collatz Map Attain Almost Bounded Values. arXiv 2019, arXiv:1909.03562. [Google Scholar] [CrossRef]
  6. Erickson, H. Graphical Analysis of Collatz Sequences. Journal of Integer Sequences 2020, 23, 2.5.7. [Google Scholar]
  7. Snyder, T. A Survey of the Collatz Conjecture. Mathematics Magazine 2015, 88, 196–210. [Google Scholar]
  8. Matthews, K. R. Generalized 3x+1 Mappings: Markov Chains and Ergodic Theory. Acta Arithmetica 2000, 95, 227–241. [Google Scholar]
  9. Wirsching, G. J. The Dynamical System Generated by the 3n+1 Function. Lecture Notes in Mathematics; Springer-Verlag.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

Disclaimer

Terms of Use

Privacy Policy

Privacy Settings

© 2025 MDPI (Basel, Switzerland) unless otherwise stated