Fact-checked by Grok 2 weeks ago

Square-free integer

A square-free integer, also known as a squarefree integer, is a positive whose prime factorization contains no squared prime factors, meaning it is not divisible by any greater than 1. Equivalently, an n is square-free if for every prime p, p^2 does not divide n. By convention, 1 is considered square-free, as it has no prime factors at all. Square-free integers play a fundamental role in , appearing in contexts such as the distribution of primes, arithmetic functions, and Diophantine equations. Examples include 1, 2, 3, 5, 6, 7, 10, 11, 13, 14, 15, 17, 19, 21, 22, 23, 26, 29, 30, 31, and so on (OEIS A005117). A key characterization is provided by the \mu(n), where |\mu(n)| = 1 if n is square-free and \mu(n) = 0 otherwise; thus, the indicator function for square-freeness is \mu(n)^2. The square-free integers have an asymptotic of $6/\pi^2 \approx 0.607927, meaning that the proportion of positive integers up to x that are square-free approaches this value as x tends to ; more precisely, the Q(x) satisfies Q(x) = (6x/\pi^2) + [O](/page/O)(\sqrt{x}) . This arises from the Euler product \prod_p (1 - 1/p^2) = 1/\zeta(2), where \zeta is the and the product is over all primes p. Every positive integer can be uniquely expressed as the product of a square-free integer and a , highlighting their structural importance in the integers.

Definition and Basic Concepts

Definition

A square-free integer is a positive integer that is not divisible by any perfect square greater than 1. Equivalently, its prime factorization consists of distinct primes with no repetitions, expressed as n = p_1 p_2 \cdots p_k where the p_i are distinct primes. The term "square-free" was introduced in contexts around the early , with the equivalent term "quadratfrei" appearing in 1912 in Edmund Landau's work on the distribution of non-square-free numbers.

Examples and Small Square-Free Integers

A square-free integer is exemplified by , which is considered square-free vacuously, as it has no prime factors and thus no squared prime dividing it. All prime numbers are square-free, since their prime factorization consists of a single prime raised to the first power. Products of distinct primes, such as $6 = 2 \times 3 and $30 = 2 \times 3 \times 5, are also square-free, as no prime appears more than once in their factorization. In contrast, integers like $12 = 2^2 \times 3, $4 = 2^2, $9 = 3^2, and $8 = 2^3 are not square-free, because they are divisible by the square of a prime. The sequence of the first 20 positive square-free integers is 1, 2, 3, 5, 6, 7, 10, 11, , , , 17, 19, 21, , 23, , 29, , 31.

Basic Properties

A square-free integer is characterized by having no squared prime factors in its prime factorization, which immediately implies certain algebraic closure properties under multiplication. Specifically, if m and n are square-free positive integers with \gcd(m, n) = 1, then their product mn is also square-free, as the prime factors of m and n are distinct and each appears to the first power in mn. This follows directly from the unique factorization theorem in the integers, ensuring no exponent exceeds 1 in the factorization of mn. The set of square-free positive integers is thus closed under whenever the operands are coprime, forming a multiplicative subsemigroup of the positive integers under this condition. Two square-free integers are coprime they share no common prime factors, which is equivalent to the general definition of coprimality since the absence of shared primes prevents any common greater than 1. For example, 6 (primes 2 and 3) and 35 (primes 5 and 7) are both square-free and coprime, with their product 210 also square-free. The for square-freeness, defined as \chi(n) = 1 if n is square-free and \chi(n) = 0 otherwise, is a multiplicative . That is, for coprime positive integers m and n, \chi(mn) = \chi(m) \chi(n). This multiplicativity arises because \chi(n) = |\mu(n)|, where \mu is the , which is itself multiplicative, and the preserves this property for coprime arguments. In the \mathbb{Z}, square-free ideals correspond to products of distinct prime ideals. The nonzero principal ideals (n) that are radical—meaning equal to their own —are precisely those generated by square-free n > 1, as the of (n) consists of multiples of the square-free kernel of n. For instance, the ideal (6) = (2)(3) is radical, while (4) = (2)^2 is not.

Factorization and Decomposition

Square-Free Factorization

Every positive integer n > 1 admits a unique decomposition n = k \cdot m^2, where k is a square-free positive integer and m is a positive integer; this is known as the square-free factorization of n. This representation follows directly from the fundamental theorem of arithmetic, which guarantees the uniqueness of the prime factorization of n. Given the prime factorization n = \prod p_i^{a_i} with distinct primes p_i and positive integers a_i, the square-free part k is obtained by including each prime p_i to the first power if a_i is odd and excluding it if a_i is even, yielding k = \prod_{a_i \ odd} p_i. The corresponding m is then \prod p_i^{\lfloor a_i / 2 \rfloor}. This process extracts the square-free component by isolating the "odd residue" of each exponent modulo 2. Note that the radical \mathrm{rad}(n) = \prod_{a_i \geq 1} p_i, which is the product of all distinct primes dividing n, coincides with k precisely when all exponents a_i are odd but generally differs by including primes with even exponents. To compute the square-free part algorithmically, first determine the prime factorization of n using trial division or more advanced methods such as the elliptic curve method; then, for each prime factor p_i^{a_i}, divide out the highest square power p_i^{2 \lfloor a_i / 2 \rfloor} to leave the contribution p_i^{a_i \mod 2}, and multiply these remainders to obtain k. Equivalently, compute the largest square divisor s^2 = \prod p_i^{2 \lfloor a_i / 2 \rfloor} of n and set k = n / s^2. For example, consider n = [72](/page/72) = 2^3 \cdot [3](/page/3)^2. The exponent of 2 is 3 (odd), so include 2; the exponent of is 2 (even), so exclude . Thus, the square-free part is k = 2, and m = 2^{\lfloor 3/2 \rfloor} \cdot [3](/page/3)^{\lfloor 2/2 \rfloor} = 2^1 \cdot [3](/page/3)^1 = 6, verifying [72](/page/72) = 2 \cdot 6^2 = 2 \cdot [36](/page/36).

Square-Free Core

The square-free core of a positive n, denoted c(n), is the unique square-free positive such that n = c(n) \cdot k^2 for some positive k. This core is obtained by dividing n by its largest square , effectively removing all even powers from the prime while retaining a single factor for each prime with an odd exponent. Formally, c(n) = \frac{n}{\prod_{p \mid n} p^{2 \left\lfloor v_p(n)/2 \right\rfloor}} = \prod_{p \mid n} p^{v_p(n) \mod 2}, where the product is over the distinct primes p dividing n, and v_p(n) denotes the exponent of p in the prime factorization of n. By construction, c(n) is always square-free, as its prime factorization consists solely of distinct primes raised to the first power (or 1 if all exponents in n are even). The integer k in the decomposition n = c(n) \cdot k^2 is given by k = \prod_{p \mid n} p^{\left\lfloor v_p(n)/2 \right\rfloor}, representing the "square root" of the square component of n. For n = 1, c(1) = 1. Representative examples include: c(12) = c(2^2 \cdot 3) = 3, since $12 = 3 \cdot 2^2; c(18) = c(2 \cdot 3^2) = 2, since $18 = 2 \cdot 3^2; c(30) = c(2 \cdot 3 \cdot 5) = 30, since $30 = 30 \cdot 1^2; and c(36) = c(2^2 \cdot 3^2) = 1, since $36 = 1 \cdot 6^2. Notably, c(n) = 1 whenever n is a perfect square. The square-free core function c(n) is multiplicative: if \gcd(n, m) = 1, then c(nm) = c(n) \cdot c(m). This follows from the independence of prime factors in coprime arguments, preserving the mod-2 exponent structure.

Square-Free Factors of Integers

The square-free factors of an n > 1 are defined as its square-free divisors, which are the positive divisors of n that are not divisible by any greater than 1. These factors consist precisely of the products of distinct prime factors of n, and they are exactly the divisors of the of n, denoted \mathrm{rad}(n), which is the product of all distinct primes dividing n. The number of square-free factors of n is $2^{\omega(n)}, where \omega(n) denotes the number of distinct prime factors of n. This count arises because each square-free factor corresponds to a of the set of distinct primes dividing n; the empty subset yields the divisor , while nonempty subsets yield the products of those primes. For instance, if n = p_1^{a_1} p_2^{a_2} \cdots p_k^{a_k} with distinct primes p_i and exponents a_i \geq 1, then the square-free factors are generated by taking all possible products \prod_{i \in S} p_i over subsets S \subseteq \{1, 2, \dots, k\}. As an example, consider n = 30 = 2 \cdot 3 \cdot 5, which has \omega(30) = 3. The square-free factors are 1, 2, , 5, $2 \cdot 3 = 6, $2 \cdot 5 = 10, $3 \cdot 5 = 15, and $2 \cdot 3 \cdot 5 = 30, totaling $2^3 = 8. In sieve methods of , the square-free factors of n play a key role in counting square-free multiples of n up to a given bound, as they facilitate inclusion-exclusion arguments over the prime factors to exclude numbers divisible by higher powers of primes.

Equivalent Characterizations

Prime Factorization Characterization

A positive n > 1 is square-free its prime n = p_1^{e_1} p_2^{e_2} \cdots p_k^{e_k}, where the p_i are distinct primes and each e_i \geq 1, satisfies e_i = 1 for all i. This condition ensures that no prime square divides n, as any such square would require an exponent of at least 2. The 1 is conventionally considered square-free, having an empty prime . This characterization follows from the , which asserts that every integer greater than 1 factors uniquely into primes (up to ordering). To see the equivalence, suppose n has an exponent e_i \geq 2; then p_i^2 divides n, so n is not square-free. Conversely, if all e_i = 1, no prime square divides n, confirming square-freeness. The proof of the theorem proceeds by on n: primes are trivially factored, and for composites, repeated division by the smallest prime factor yields the full , with uniqueness ensured by the fact that primes are irreducible and any common factor in alternative factorizations must match. In , this notion extends to Dedekind domains, where every nonzero factors uniquely into s. An is square-free if it is a product of distinct s, mirroring the case by excluding higher powers of any . For example, in the of a number field, square-free ideals play a key role in ramification theory and class group computations. In contrast to square-free integers, cube-free integers permit exponents up to 2 in their , meaning they are divisible by no greater than 1. This characterization relies historically on Euclid's demonstration in the Elements of the infinitude of primes and the existence of prime factorizations (circa 300 BCE), with uniqueness rigorously established by in his (1801).

Möbius Function Characterization

A positive integer n is square-free if and only if the satisfies \mu(n) \neq 0. More precisely, when n is square-free, \mu(n) = (-1)^{\omega(n)}, where \omega(n) denotes the number of distinct prime factors of n; if n has a repeated prime factor, then \mu(n) = 0. This provides a direct arithmetic test for square-freeness using the , distinct from examining the exponents in the prime factorization. The Möbius function \mu(n) is defined such that \mu(1) = 1, \mu(n) = 0 if there exists a prime p with p^2 \mid n, and \mu(n) = (-1)^k if n is a product of exactly k distinct primes. Consequently, \mu(n) vanishes precisely when n is not square-free, alternating in sign based on the parity of the number of prime factors otherwise. The absolute value |\mu(n)| thus acts as the for square-free integers, equaling 1 if n is square-free and 0 otherwise. This characterization arises directly from the definition of \mu(n), which encodes the inclusion-exclusion principle to detect squared factors in the prime of n. For instance, if n = p_1 p_2 \cdots p_k with distinct primes p_i, then \mu(n) = (-1)^k \neq 0; but if n = p^2 m for some prime p and m, the presence of the square forces \mu(n) = 0. The function's link to square-freeness finds application in sieve methods for counting square-free s up to a bound x. Specifically, the count is given by \sum_{n \leq x} |\mu(n)| = \sum_{d \leq \sqrt{x}} \mu(d) \left\lfloor \frac{x}{d^2} \right\rfloor, extending the by sieving out multiples of squares via the inversion.

Radical and Other Equivalents

A fundamental arithmetic characterization of square-free s involves the function, defined as \mathrm{rad}(n) = \prod_{p \mid n} p, the product of the distinct prime factors of a positive n. Since \mathrm{rad}(n) divides n and \mathrm{rad}(n) = n if and only if no prime square divides n, an n > 1 is square-free precisely when \mathrm{rad}(n) = n. This equivalence highlights the square-free nature as the absence of higher prime powers in the . The concept of square-freeness extends naturally to over . A f(x) \in \mathbb{Q} is square-free if its irreducible contains no repeated factors, equivalent to \gcd(f, f') = 1 in characteristic zero, ensuring no multiple in the . This property can be verified using resultants, as \mathrm{Res}(f, f') = 0 if and only if f and f' share a common factor, indicating a repeated . Recent research has explored the of square-free values taken by at points. For instance, and Zarhin (2024) established explicit bounds on the square-free for certain multivariable , refining earlier asymptotic results and confirming that under mild conditions, the proportion of points k \in \mathbb{Z}^s for which f(k) is square-free approaches the expected of $6/\pi^2. Such studies bridge and square-freeness, with applications to Diophantine problems.

Analytic Number Theory

Dirichlet Series

The Dirichlet series generating function for the indicator of square-free positive integers is the sum over all n \geq 1 of |\mu(n)| n^{-s}, where \mu denotes the . This series equals \zeta(s)/\zeta(2s) for \Re(s) > 1, where \zeta is the . The equality follows from the Euler product formula for the zeta function, yielding \frac{\zeta(s)}{\zeta(2s)} = \prod_p \frac{1 - p^{-2s}}{1 - p^{-s}} = \prod_p (1 + p^{-s}), with the infinite product ranging over all primes p. The local factor $1 + p^{-s} at each prime corresponds to the contribution from the (1) and the single p (since higher powers p^k for k \geq 2 are excluded for square-free integers). This converges absolutely in the half-plane \Re(s) > 1 and admits a meromorphic continuation to the whole , determined by the meromorphic nature of \zeta(s) and \zeta(2s). It features a simple at s = 1, arising from the simple pole of \zeta(s) there (while \zeta(2) is finite and nonzero), and is holomorphic elsewhere except at points where \zeta(2s) vanishes but \zeta(s) does not. Such generating functions emerged as fundamental tools in 19th-century , extending Bernhard Riemann's 1859 investigation of the function to study arithmetic properties like square-freeness through multiplicative structure.

Asymptotic Density

The asymptotic density of square-free integers, defined as the natural density \lim_{x \to \infty} Q(x)/x where Q(x) counts the square-free positive integers up to x, equals $6/\pi^2 \approx 0.607927. This value indicates that square-free integers comprise roughly 60.79% of all positive integers. The derivation stems from the Dirichlet series for the characteristic function of square-free integers, \sum_{n=1}^\infty \mu^2(n)/n^s = \zeta(s)/\zeta(2s), evaluated at s=1, yielding $1/\zeta(2). Since \zeta(2) = \pi^2/6, the density simplifies to $6/\pi^2. An elementary proof employs the inclusion-exclusion principle: Q(x) = \sum_{n \leq x} \sum_{d^2 \mid n} \mu(d) = \sum_{d=1}^\infty \mu(d) \lfloor x/d^2 \rfloor. For large x, the sum approximates x \sum_{d=1}^\infty \mu(d)/d^2 = x/\zeta(2) = (6/\pi^2)x, with the main term arising from the Euler product \prod_p (1 - 1/p^2) = 1/\zeta(2). Classical estimates provide the error term Q(x) = (6/\pi^2)x + O(\sqrt{x}), where the O(\sqrt{x}) bound accounts for the finite sum up to d \approx \sqrt{x} and discrepancies from the floor function. This density extends to arithmetic progressions. For residues a modulo q with \gcd(a,q)=1, the number of square-free integers up to x congruent to a modulo q is asymptotically c_{a,q} x, where c_{a,q} > 0 is given by (1/\phi(q)) \sum_{\chi \pmod{q}} \overline{\chi}(a) L(1,\chi)/L(2,\chi) via orthogonality of Dirichlet characters and the series \sum \mu^2(n) \chi(n)/n^s = L(s,\chi)/L(2s,\chi); such densities are positive since L(2,\chi) \neq 0.

Gaps Between Square-Free Integers

A gap between square-free integers refers to the difference between consecutive such numbers in the natural number sequence. While the asymptotic density of square-free integers is \frac{6}{\pi^2} \approx 0.6079, implying an average gap of approximately \frac{\pi^2}{6} \approx 1.64493 between them, local fluctuations lead to larger maximal gaps. These gaps often occur in regions dense with multiples of small prime squares, such as near multiples of 4, 9, or 25, where sequences of non-square-free numbers arise due to divisibility by p^2 for small primes p. Upper bounds on the maximal gap between square-free integers up to x have been progressively improved since the mid-20th century. Early results by Fogels () established that every (x, x + x^{\theta}] with \theta > \frac{2}{5} contains a square-free integer for sufficiently large x, later refined by Roth () to \theta > \frac{3}{13} \approx 0.2308. A significant advancement came from Filaseta and Trifonov (1990), who proved using elementary methods that there exists a constant c > 0 such that every (x, x + c x^{1/5} \log x] contains a square-free integer for sufficiently large x. This exponent of \frac{1}{5} = 0.2 was improved in 2024 by , who showed that for some \eta > 0, every (x, x + x^{1/5 - \eta}] contains a square-free integer for sufficiently large x. Constructions for lower bounds on maximal gaps typically employ the to engineer intervals devoid of square-free numbers. For instance, to create a gap of at least g, solve the system of congruences m \equiv -k \pmod{p_k^2} for k = 0, 1, \dots, g-1, where p_k are the first r primes and r is chosen appropriately; this ensures each integer in [m, m+g-1] is divisible by some p^2. This construction shows that there exist infinitely many such intervals with h \gg \frac{\log x}{\log \log x}, demonstrating that maximal gaps grow at least like \log x / \log \log x. More refined constructions yield gaps of size \Omega\left( \frac{\log x \log \log \log x}{\log \log x} \right). Recent work has focused on making these upper bounds explicit. In 2023, McNew et al. provided the first explicit version of the Filaseta-Trifonov bound, proving that for any x \geq 2, every interval (x, x + 11 x^{1/5} \log x] contains a square-free integer, with the constant 11 improvable for very large x. This result builds on the original method while incorporating numerical optimizations for practical verification. Ongoing research continues to explore tighter exponents.

Distribution and Computation

Computation of the Counting Function Q(x)

The counting function Q(x), which enumerates the number of square-free positive integers not exceeding x, admits the exact formula Q(x) = \sum_{n \le x} |\mu(n)| = \sum_{k=1}^{\lfloor \sqrt{x} \rfloor} \mu(k) \left\lfloor \frac{x}{k^2} \right\rfloor, where \mu denotes the . This expression arises from applying Möbius inversion to the of square-free integers, which equals |\mu(n)|. For computational purposes, the summation is truncated at k = \lfloor \sqrt{x} \rfloor, as the floor function \left\lfloor x / k^2 \right\rfloor vanishes for larger k. The values of \mu(k) up to \sqrt{x} can be precomputed efficiently using a linear sieve variant of the , which identifies square-free k and counts their distinct prime factors in O(\sqrt{x}) time overall. A naive evaluation of the sum requires O(\sqrt{x}) arithmetic operations after sieving, making it suitable for moderate x. For larger x, optimizations such as input segmentation—dividing the sum into blocks—or adaptations of the Dirichlet hyperbola method can accelerate computation by exploiting symmetries in the floor terms, achieving sublinear time relative to x in practice. Tables of Q(x) for large x (up to $10^{10} or more) have been computed using these summation-based methods since the , leveraging early electronic computers to verify asymptotic behaviors and error estimates.

Tables of Q(x), 6/π² x, and Relative Error

The counting function Q(x), which enumerates the number of square-free integers up to x, is asymptotically approximated by \frac{6}{\pi^2} x, where \frac{6}{\pi^2} \approx 0.607927101854. To illustrate the accuracy of this approximation, the following presents values for x = 10^k with k = 1 to $10, including Q(x), the approximation \frac{6}{\pi^2} x, and the normalized error measure R(x) = \frac{|Q(x) - \frac{6}{\pi^2} x|}{\sqrt{x}}. These data are sourced from the (OEIS) A071172.
kx = 10^kQ(x)\frac{6}{\pi^2} xR(x)
11076.079271018540.291
21006160.79271018540.0207
3608607.9271018540.00231
460836079.271018540.0373
56079460792.71018540.00408
6607926607927.1018540.00110
760792916079271.018540.00632
86079269460792710.18540.00162
9607927124607927101.8540.000700
101000000000060792709426079271018.54030.000766
The values of Q(x) were computed using efficient algorithms for large n, enabling extensions up to x = 10^{36}. The normalized error R(x) oscillates but remains bounded as O(1), consistent with the error term in the asymptotic expansion Q(x) = \frac{6}{\pi^2} x + O(\sqrt{x}). This convergence is evident in plots of Q(x) against \frac{6}{\pi^2} x, where the ratio approaches the density \frac{6}{\pi^2} for large x.

Algorithms for Testing and Decomposition

To test whether an n is square-free, one standard deterministic approach is trial division: for each prime p \leq \sqrt{n}, check if p^2 divides n; if no such p exists, then n is square-free. This method runs in O(\sqrt{n}) time, which is feasible for moderate-sized n but impractical for large cryptographic integers. An alternative deterministic check computes \gcd(n, \prod_{p \leq \sqrt{n}} p^2); if the result is 1, then n is square-free, though constructing the product requires significant preprocessing and remains inefficient for large n. Probabilistic methods can accelerate testing by first applying compositeness tests like Miller-Rabin to identify potential prime factors, followed by verifying that no prime squares divide n. However, fully confirming square-freeness generally requires complete , as partial factorization may miss higher powers. There is no known polynomial-time for testing square-freeness or extracting the square-free part of an , and the problem is believed to be as hard as . Under the Generalized Riemann Hypothesis (GRH), an using the explicit formula for L-functions can certify square-freeness in subexponential time by bounding the number of small prime factors via analytic estimates. For square-free decomposition—expressing n = k \cdot m^2 where k is square-free (the square-free part of n)—standard methods rely on full using algorithms like the for small factors or the Number Field Sieve (NFS) for large semiprimes, followed by removing squared factors from the prime . is particularly effective for finding factors up to 50-60 digits, with expected running time O(\exp(c \sqrt{\log p \log \log p})) for a prime factor p, while NFS achieves subexponential time O(\exp(c (\log n)^{1/3} (\log \log n)^{2/3})) for general n. Recent advances leverage class groups of binary quadratic forms to compute the decomposition directly without full . In a 2023 algorithm, the square-free part and square part are found by solving representation problems in the class group of discriminant -4ab (where n = a b^2) using reduction and composition operations, achieving heuristic expected time O(\log^4 n) under GRH for class group computations. This class group method has applications in , particularly for generating square-free discriminants in protocols based on ideal class groups of quadratic fields, such as the CSIDH , where efficient decomposition ensures secure parameter selection without exposing full .

Representations and Conjectures

Encoding as Numbers

Square-free integers admit a natural encoding as strings, where the bits indicate the presence or absence of specific primes in their . The primes are ordered increasingly as p_1 = 2, p_2 = [3](/page/3), p_3 = 5, and so on. For a square-free integer n = \prod_{i \in S} p_i, where S is a of indices, the corresponding has a 1 in the i-th if i \in S and 0 otherwise. The number 1, being the , corresponds to the all-zero . This mapping establishes a between the positive square-free integers and the numbers with finite support (i.e., only finitely many 1-bits). By the , every finite subset of primes corresponds uniquely to a square-free integer via their product, and vice versa for square-free integers greater than 1. The facilitates systematic enumeration of square-free integers by generating all strings of increasing length and computing the product of primes at positions with 1-bits. For instance:
  • The string $0\ldots0 (all zeros) maps to 1.
  • \ldots001 (1 in ) maps to p_1 = 2.
  • \ldots010 (2 in ) maps to p_2 = 3.
  • \ldots011 (3 in ) maps to $2 \times 3 = 6.
  • \ldots100 (4 in ) maps to p_3 = 5.
Properties of this encoding highlight connections to subset enumeration: the total number of square-free integers using only the first k primes is exactly $2^k, corresponding to all s of those primes. The asymptotic of square-free integers, $6/\pi^2 \approx 0.6079, arises from the proportion of subsets whose products remain below a given bound, reflecting the thinning effect of prime growth in the representation. This encoding proves useful in computational enumeration and generating functions for square-free integers. For example, the summing over square-free n is \sum | \mu(n) | n^{-s} = \zeta(s)/\zeta(2s) = \prod_p (1 + p^{-s}) for \Re(s) > 1, where each factor $1 + p^{-s} encodes the choice of including or excluding the prime p. Similar product forms appear in ordinary generating functions like \prod_p (1 + x^p), enabling efficient computation via dynamic programming or recursive generation in algorithms for listing or counting square-frees up to a bound.

Erdős Square-Free Conjecture

The Erdős square-free conjecture posits that every positive greater than 1 can be expressed as the sum of a square-free and a power of 2. Formally, for every m > 1, there exist a square-free k and a non-negative l such that m = k + 2^l. This conjecture was posed by , with an early written formulation appearing in his 1977 paper on combinatorial problems. Erdős motivated the problem as part of broader inquiries into additive representations involving square-free numbers, suggesting that it might be approachable by replacing primes with square-frees in variants of Goldbach-type problems. Computational verifications have provided strong empirical support for the . Initial checks by Andrew Odlyzko extended up to $10^7. More recently, Kevin Hercher confirmed the representation holds for all integers up to $2^{50} \approx 1.126 \times 10^{15}, with no counterexamples found; in these cases, the required exponents l were at most 13. These exhaustive searches, leveraging efficient algorithms for square-free testing, have ruled out small counterexamples and motivated further theoretical progress. Partial results establish the for sufficiently large odd integers using arguments and connections to theory. Erdős himself proved that the representation holds for almost all odd integers, in the sense of , by exploiting the positive of square-free numbers, which is $6/\pi^2 \approx 0.607. More rigorously, Granville and Soundararajan demonstrated that the is equivalent to the infinitude of primes p for which the multiplicative order of 2 modulo p^2 equals p-1, implying it holds for all sufficiently large m if there are only finitely many Wieferich primes (primes p where p^2 divides $2^{p-1} - 1); only two such primes, 1093 and 3511, are known below $10^{32}. This links the problem to deeper questions in progressions and has inspired studies of square-free numbers as elements of additive bases, such as representations of integers as sums of distinct square-frees.

References

  1. [1]
    Squarefree -- from Wolfram MathWorld
    A number is said to be squarefree (or sometimes quadratfrei; Shanks 1993) if its prime decomposition contains no repeated factors.
  2. [2]
    square-free number in nLab
    Jan 22, 2023 · A square-free number is a natural number n n such that for all prime numbers p p , p 2 p^2 does not divide n n . 2. Properties. Given a square- ...Missing: definition | Show results with:definition
  3. [3]
    square-free number - PlanetMath.org
    Mar 22, 2013 · A square-free number is a natural number Mathworld Planetmath that contains no powers greater than 1 in its prime factorization.Missing: theory | Show results with:theory
  4. [4]
    Are negative numbers square-free? - Mathematics Stack Exchange
    Mar 25, 2021 · 2. Square free means not divisible by the square of a prime. · negative numbers are always square-free as long as you are in real field. A number ...
  5. [5]
    Simple (easy to understand) definition of squarefree integers.
    Jun 27, 2013 · With that definition, 0 certainly is not squarefree, and this is what I would expect: squarefree numbers should not be very divisible in a ...
  6. [6]
    Math Origins: The Totient Function
    Even though Euler's original definition appeared in 1763, it wasn't until 1801 that Gauss introduced the ϕ notation and 1879 that Sylvester used the word ...
  7. [7]
    Squarefree Part -- from Wolfram MathWorld
    That part of a positive integer left after all square factors are divided out. For example, the squarefree part of 24=2^3·3 is 6, since 6·2^2=24.
  8. [8]
    A005117 - OEIS
    Squarefree numbers: numbers that are not divisible by a square greater than 1. (Formerly M0617). 2033. 1, 2, 3, 5, 6, 7, 10, 11, ...
  9. [9]
    [PDF] Introduction to the Theory of Numbers
    Nov 21, 2014 · Page 1. AN INTRODUCTION. TO THE. THEORY. OF NUMBERS. Page 2. Page 3. AN INTRODUCTION. TO THE. THEORYOFNUMBERS. BY. G. H. HARDY. AND. E. M. ...
  10. [10]
    An Introduction to the Theory of Numbers - Google Books
    ... multiplicative my² natural density non-negative number theory obtain odd ... square-free summands suppose values write x₁ zero. Bibliographic ...
  11. [11]
    Characteristic Function of Square-Free Integers is Multiplicative
    Apr 25, 2019 · Characteristic Function of Square-Free Integers is Multiplicative ; :={n∈Z:∀k∈Z>1:k2∤n}. That is, let S be the set of all square-free positive ...
  12. [12]
    [PDF] MAS439 Lecture 8 Maximal, Prime, Radical
    ... radical. Page 17. Radical ideals in Z. Lemma. The ideal (n) is radical if and only if n is square-free – that is, n has no repeated prime factors. Proof. We ...
  13. [13]
    An Efficient Exact Quantum Algorithm for the Integer Square-free ...
    Feb 10, 2012 · Here we propose an efficient and exact quantum algorithm for finding the square-free part of a large integer - a problem for which no efficient classical ...
  14. [14]
    [PDF] On the radical of a perfect number - Florian Luca and Carl Pomerance
    For a positive integer n we put rad(n) = Y p|n p, where p runs over primes. The number rad(n) is called the radical of n, or the squarefree kernel of n. Let ...Missing: definition | Show results with:definition
  15. [15]
    [PDF] The three squares theorem and enchanted walks, with various ... - IJS
    Aug 4, 2012 · For any non-zero rational number r, the integer core(r) := sgn(r) p pep(r) mod 2 is called the square-free core of the number r; we also define ...
  16. [16]
    [PDF] Multiplicative partitions of numbers with a large squarefree divisor
    As usual, rad(n) denotes the radical of n, i.e., its largest squarefree divisor. f(n) = x/L(x)1+α+o(1). F(nk) = x/L(x)1+o(1).
  17. [17]
    [PDF] Sieve Methods - cs.wisc.edu
    We apply sieves to study the distribution of square-free numbers, smooth numbers, and prime numbers. The first chapter is a discussion of the basic sieve ...
  18. [18]
  19. [19]
    [PDF] Square-free ideals and SR condition
    May 23, 2023 · Abstract. In this article, we introduce the concept of a square-free ideal in any ring. We study equivalent definitions and some algebraic ...
  20. [20]
    [PDF] arXiv:1809.05432v2 [math.NT] 9 May 2019
    May 9, 2019 · Suppose that L is a finite separable extension over K of degree d, n is a square-free ideal of OK, and NK(n) is finite. Let OL be the ...
  21. [21]
    Cubefree -- from Wolfram MathWorld
    A number is said to be cubefree if its prime factorization contains no tripled factors. All primes are therefore trivially cubefree.
  22. [22]
    Prime numbers - MacTutor History of Mathematics
    Euclid also gives a proof of the Fundamental Theorem of Arithmetic: Every integer can be written as a product of primes in an essentially unique way. Euclid ...
  23. [23]
    [PDF] Introduction to analytic number theory
    Apostol, Tom M. Introduction to analytic number theory. (Undergraduate texts in mathematics). ” Evolved from a course (Mathematics 160) offered at the ...
  24. [24]
    [PDF] Introduction to Sieves - UC Berkeley math
    That's because a square-free number has no prime squares in its factorization. Since µ(1) = 1, the sum Pd2|n µ(d) will be 1. • If n is not square-free, it ...
  25. [25]
    [PDF] Introduction to the abc conjecture Arshay Sheth - University of Warwick
    Feb 28, 2023 · We have that rad(N) = N if and only if N is squarefree i.e. no square (other than 1) divides N. Examples of squarefree numbers: 10,35,any ...
  26. [26]
    [PDF] squarefree values of multivariable polynomials - MIT Mathematics
    An integer n is called squarefree if for all prime numbers p we have p2. ∤ n ... coprime positive integers satisfying. Date: September 28, 2003 (revised ...
  27. [27]
    [PDF] Chapter 1 Dirichlet Series — I
    By using Theorem 9 we can determine the Dirichlet series generating functions of λ(n) and of µ(n) in terms of the Riemann zeta function. Corollary 10. For σ > 1 ...
  28. [28]
    [PDF] DIRICHLET SERIES The Riemann zeta-function ζ(s ... - Keith Conrad
    If an = 1 for all n then f(s) = ζ(s), which converges for σ > 1. It does not converge at s = 1. Example 2. If an = χ(n) for a Dirichlet character χ, f ...
  29. [29]
    A simple proof that the square-free numbers have density 6/pi^2
    Dec 9, 2017 · Abstract. In this note we give a simple proof of the well-known result that the square-free numbers have density 6 ∕ π2.Missing: derivation | Show results with:derivation
  30. [30]
    On the distribution of squarefree numbers - ScienceDirect
    It is elementary to prove that Q ( x ) = ∑ n ≤ x | μ ( n ) | = 6 π − 2 x + Δ ( x ) , where Δ ( x ) is the error term, and Δ ( x ) = O ( x 1 / 2 ) . The problem ...
  31. [31]
    Remainder estimates for squarefree integers in arithmetic progression
    The author gives remainder estimates for squarefree integers in arithmetic progressions corresponding to those of Barban, Bombieri, Davenport, Halberstam, and ...
  32. [32]
    Large gap between two consecutive square-free numbers
    Apr 16, 2016 · How to prove the stronger bound? By estimating the index where the constructed gap between successive squarefree numbers occurs.The two consecutive square free numbers have arbitrarily large gaps.largest number of consecutive square-free positive integersMore results from math.stackexchange.com
  33. [33]
    A013928 - OEIS
    For n >= 1 define an n X n (0, 1) matrix A by A[i, j] = 1 if gcd(i, j) = 1, A[i, j] = 0 if gcd(i, j) > 1 for 1 <= i,j <= n . The rank of A is a(n + 1).
  34. [34]
    254A, Notes 1: Elementary multiplicative number theory - Terry Tao
    Nov 23, 2014 · We now present a basic method to improve upon the estimate (38), namely the Dirichlet hyperbola method. The point here is that the error ...
  35. [35]
    Computing the Summation of the M¨obius Function P - Project Euclid
    Our method is elementary, and was inspired by [Lehman 1960]. We give a table of certain values of M(x) for x up to 1016, and also some computation times.
  36. [36]
    A071172 - OEIS
    Number of squarefree integers <= 10^n. 20. 1, 7, 61, 608, 6083, 60794, 607926, 6079291, 60792694, 607927124, 6079270942 ...
  37. [37]
    [1107.4890] Counting Square-Free Numbers - arXiv
    Jul 25, 2011 · Abstract:The main topic of this contribution is the problem of counting square-free numbers not exceeding n. Before this work we were able ...
  38. [38]
    [PDF] Detecting square-free numbers via the explicit formula - OSU Math
    Let χ be a primitive Dirichlet character of conductor ∆. Explicit formula relates the zeros, the coefficients, and the conductor. g(y)eixy dx = sin(xY /2)2 (xY ...
  39. [39]
    Detecting squarefree numbers - Project Euclid
    We present an algorithm, based on the explicit formula for L-functions and con- ditional on the generalized Riemann hypothesis, for proving that a given integer ...
  40. [40]
    Fast square-free decomposition of integers using class groups - arXiv
    Aug 11, 2023 · In this paper we present an algorithm based on class groups of binary quadratic forms that finds the square-free decomposition of n, ie a and b, in heuristic ...
  41. [41]
    [PDF] (76.38) Problems and results on combinatorial number theory, III., r
    Is it true that every sufficiently large odd integer is of the form 2k + 0 where 0 is squarefree? Is there in fact an odd integer not of this form? The ...
  42. [42]
    Erdős Problem #11
    Is every odd n the sum of a squarefree number and a power of 2? #11: [Er77c] ... Hercher [He24b] has verified this is true for all odd integers up to 250≈1.12×1015 ...
  43. [43]
    [PDF] On the Sum of a Squarefree Integer and a Power of Two
    Apr 14, 2025 · Erd˝os conjectured that every odd number greater than one can be expressed as the sum of a squarefree number and a power of two. Subsequently, ...
  44. [44]
    A Binary Additive Problem of Erdös and the Order of 2 mod p 2
    We show that the problem of representing every odd positive integer as the sum of a squarefree number and a power of 2, is strongly related to the problem.