1. Introduction
For centuries, mathematicians have been captivated by the enigmatic allure of perfect numbers, defined as positive integers whose proper divisors sum precisely to the number itself [
1]. This fascination traces back to ancient Greece, where Euclid devised an elegant formula for generating even perfect numbers through Mersenne primes, numbers of the form
where
p is prime [
1]. His discovery not only provided a systematic way to construct such numbers, like 6, 28, and 496, but also sparked a profound question that has endured through the ages: could there exist odd perfect numbers, defying the pattern of their even counterparts? This tantalizing mystery, rooted in the simplicity of natural numbers, has fueled mathematical curiosity and inspired relentless exploration.
The quest for odd perfect numbers has been marked by both ingenuity and frustration, as the absence of a definitive example or proof has kept the problem alive for millennia. Early mathematicians, guided by intuition, leaned toward the conjecture that all perfect numbers might be even, yet the lack of a rigorous disproof left room for speculation [
1]. Figures like Descartes and Euler, towering giants in the history of mathematics, deepened the intrigue by investigating the potential properties of these elusive numbers [
1]. Euler, in particular, highlighted the challenge, noting, “Whether
… there are any odd perfect numbers is a most difficult question”. Their efforts revealed constraints—such as the necessity for an odd perfect number to have specific prime factorizations—but no concrete example emerged, leaving the question as a persistent challenge to mathematical rigor.
Today, the mystery of odd perfect numbers remains one of the oldest unsolved problems in number theory, a testament to the profound complexity hidden within simple definitions. Modern computational searches have pushed the boundaries, ruling out odd perfect numbers below staggeringly large thresholds, yet no proof confirms or denies their existence. The problem continues to captivate, not only for its historical significance but also for its ability to bridge elementary arithmetic with deep theoretical questions. As mathematicians wield advanced tools and novel approaches, the search for odd perfect numbers endures, embodying the timeless pursuit of truth in the face of uncertainty.
Despite extensive research, no odd perfect numbers have been discovered, and numerous constraints have been established. This paper resolves the conjecture by proving that odd perfect numbers do not exist. Using a proof by contradiction, we assume the existence of an odd perfect number N, which satisfies , where is the sum of its divisors. By leveraging the inequality for odd N, and the bound for odd perfect numbers with at least 10 distinct prime factors, we derive a contradiction by showing that cannot simultaneously be less than and at least 3. This result confirms that all perfect numbers are even, resolving a longstanding open problem in number theory.
2. Background and Ancillary Results
In 1734, Leonhard Euler solved the celebrated Basel problem, determining the exact value of the Riemann zeta function at
. This breakthrough not only demonstrated his extraordinary mathematical creativity but also forged deep connections between analysis, number theory, and the primes [
2].
Proposition 1.
The Riemann zeta function evaluated at satisfies:
where:
is the n-th prime number,
n ranges over the natural numbers, and
is the fundamental constant arising in diverse mathematical contexts, from geometry to number theory.
Euler’s proof ingeniously bridges the infinite series and an infinite product over primes, revealing the surprising appearance of π in the limit.
Definition 1. In number theory, the p-adic order of a positive integer n, denoted , is the highest exponent of a prime number p that divides n. For example, if , then and .
The divisor sum function, denoted , is a fundamental arithmetic function that computes the sum of all positive divisors of a positive integer n, including 1 and n itself. For instance, the divisors of 12 are , yielding . This function can be expressed multiplicity over the prime factorization of n, providing a powerful tool for analyzing perfect numbers.
Proposition 2.
For a positive integer with prime factorization [3]:
where indicates that p is a prime divisor of n.
Proposition 3. Similarly, Euler’s totient function, which counts the integers up to n that are coprime to n, is given by [4].
The abundancy index, defined as , maps positive integers to rational numbers and quantifies how the divisor sum compares to the number itself. The following Proposition provides a precise formula for based on the prime factorization.
Proposition 4.
Let be the prime factorization of n, where are distinct primes and are positive integers. Then [5]:
In our proof, we utilize the following propositions:
Proposition 5. A positive integer n is a perfect number if and only if , meaning .
Proposition 6. Any odd perfect number N must have at least 10 distinct prime factors [6,7].
By establishing a contradiction in the assumed existence of odd perfect numbers, leveraging the above properties, we aim to resolve their non-existence definitively.
3. Main Result
This is a key finding.
Lemma 1.
Let n be an odd positive integer, be Euler’s totient function, which counts the number of integers up to n that are coprime to n, and be the divisor sum function, which sums all positive divisors of n. Then:
Proof. Let
n be an odd positive integer with prime factorization
where
are distinct odd primes (i.e.,
),
are their multiplicities, and
(allowing
when
).
The Euler totient function is multiplicative and given by:
since
.
Similarly, the divisor sum function is multiplicative with:
since
.
We analyze the ratio:
Substituting the expressions for
and
:
Multiplying these yields:
Since
and
, the term
increases with
. Thus:
Moreover, since
, we have:
where the right-hand product starts at
.
Using the identity for the Euler product of the Riemann zeta function:
we derive:
By transitivity, we obtain:
completing the proof. □
This is a main insight.
Lemma 2.
Let N be an odd perfect number, i.e., a positive odd integer such that , where denotes the sum of the divisors of N. Then, the ratio of N to its Euler totient function , which counts the number of integers up to N that are coprime to N, satisfies:
Proof. Assume
N is an odd perfect number, so
. Since
N is odd, its prime factorization involves only odd primes. Let:
where
are distinct odd primes (i.e.,
), and
are their multiplicities, with
being the number of distinct prime factors.
The Euler totient function
is given by:
since for a prime power
, we have
, and
is multiplicative across distinct primes. Thus, the ratio is:
To find a lower bound for , we need to maximize the product , which decreases as m increases or as the primes grow larger. The maximum value of the product occurs when m is minimized and the primes are as small as possible.
The smallest possible
m for an odd perfect number is conjectured to be large due to known constraints (e.g.,
N must have many prime factors), but we proceed by considering the smallest odd primes to establish a lower bound. Suppose
N has exactly
distinct prime factors (a conservative estimate, as odd perfect numbers, if they exist, likely have more). Take the first 10 odd primes:
,
,
,
,
,
,
,
,
,
. Compute the product
:
If
N has more than 10 distinct prime factors (
), include the next prime, e.g.,
, so
. This reduces the product further:
yielding:
which is greater than 3.264. As
m increases, the product
continues to decrease, making
larger.
If
, use the first
m odd primes. For example, if
(primes 3 to 29):
so:
However, odd perfect numbers are conjectured to have significantly more than 9 distinct prime factors due to the condition
requiring a large divisor sum, which typically necessitates many prime factors (e.g., modern bounds suggest at least 10 distinct primes). Thus,
is a reasonable assumption for the minimal case.
Since
and the product is maximized (i.e., largest denominator, smallest ratio) when using the smallest
m and smallest primes, the approximate value of 3.264 corresponds to
with the first 10 odd primes. For
, the ratio is at least 3, and typically larger. Therefore, for any odd perfect number
N:
□
This is the main theorem.
Theorem 1. Odd perfect numbers do not exist.
Proof. Suppose, for the sake of contradiction, that an odd perfect number
N exists. A perfect number satisfies
, where
is the sum of all positive divisors of
N (including 1 and
N itself). Since
N is odd, its prime factorization consists solely of odd primes:
where
are distinct odd primes (
), and
are their multiplicities.
Since
N is perfect, the abundancy index is:
Consider the ratio involving the Euler totient function
, which counts the number of integers up to
N coprime to
N. For an odd positive integer
N, it is known that:
Rewrite this inequality:
Since
, we have:
Thus:
Given
(since
N is perfect), compute the left-hand side:
Since
, we have
, so:
Thus, the inequality becomes:
For an odd perfect number
N, constraints imply it has at least 10 distinct prime factors (a known lower bound in number theory). The ratio
is given by:
where
. To find a lower bound, maximize the product
by choosing the smallest
and the smallest odd primes:
. Compute:
Thus:
For
, the product decreases (e.g., multiply by
for
), making
larger. Hence:
From Lemma 1, we have:
However, from Lemma 2:
This leads to a contradiction, since:
Therefore, our assumption that an odd perfect number
N exists must be false.
Since the assumption of an odd perfect number leads to a contradiction, we conclude that no such number exists. □
Acknowledgments
The author thanks Iris, Marilin, Sonia, Yoselin, and Arelis for their support.
References
- Dickson, L.E. History of the Theory of Numbers; Number 256, Carnegie Institution of Washington: Washington, D.C, United States, 1919. [Google Scholar]
- Ayoub, R. Euler and the Zeta Function. The American Mathematical Monthly 1974, 81, 1067–1086. [Google Scholar] [CrossRef]
- Lagarias, J.C. An Elementary Problem Equivalent to the Riemann Hypothesis. The American Mathematical Monthly 2002, 109, 534–543. [Google Scholar] [CrossRef]
- Dimitrov, S. Inequalities involving arithmetic functions. Lithuanian Mathematical Journal 2024, 64, 421–452. [Google Scholar] [CrossRef]
- Hertlein, A. Robin’s Inequality for New Families of Integers. Integers 2018, 18. [Google Scholar] [CrossRef]
- Ochem, P.; Rao, M. Odd perfect numbers are greater than 101500. Mathematics of Computation 2012, 81, 1869–1877. [Google Scholar] [CrossRef]
- Nielsen, P. Odd perfect numbers, Diophantine equations, and upper bounds. Mathematics of Computation 2015, 84, 2549–2567. [Google Scholar] [CrossRef]
|
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. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).