Fact-checked by Grok 2 weeks ago

Euler's totient function

Euler's totient function, commonly denoted as φ(n), is an that counts the positive up to a given positive integer n which are relatively prime to n. This count includes 1 and excludes any multiples of the prime factors of n, making φ(n) a measure of the "density" of integers coprime to n in the range from 1 to n. The explicit formula for φ(n) is φ(n) = n ∏_{p | n} (1 - 1/p), where the product runs over the distinct prime factors p of n. For a prime power p^k, φ(p^k) = p^k - p^{k-1}, and the function is multiplicative: if m and n are coprime, then φ(mn) = φ(m) φ(n). A key property is that the sum of φ(d) over all positive divisors d of n equals n, i.e., ∑_{d | n} φ(d) = n. Additionally, φ(n) is even for all n ≥ 3, and it satisfies inequalities such as φ(n) > c n / \log \log n for some constant c and sufficiently large n. Leonhard Euler first introduced the totient function in 1763 in his paper "Demonstration of a new method in the theory of arithmetic" (E271), where he used it to prove a generalization of known as : if a and n are coprime, then a^{φ(n)} ≡ 1 \pmod{n}. The modern notation φ(n) was adopted by Carl Friedrich Gauss in 1801 in his Disquisitiones Arithmeticae, while the term "totient" was coined by J. J. Sylvester in 1879. The function plays a central role in analytic number theory, appearing in the Dirichlet series ∑ φ(n)/n^s = ζ(s-1)/ζ(s) for Re(s) > 2, where ζ(s) is the Riemann zeta function. Beyond , Euler's totient function is essential in , particularly in the algorithm, where φ(n) for a n = pq (with p and q distinct primes) determines the exponent for to ensure secure key generation and encryption. It also underlies pseudorandom number generation and primality testing in .

Definition, History, and Notation

Definition

Euler's totient function, denoted \phi(n), counts the number of positive integers k such that $1 \leq k \leq n and k is coprime to n, formally defined as \phi(n) = \left| \{ k : 1 \leq k \leq n, \gcd(k,n)=1 \} \right|. Two positive integers are coprime if their greatest common divisor is 1, meaning they share no prime factors in common. For example, with n=6, the integers from 1 to 6 are checked for coprimality: only 1 and 5 satisfy \gcd(k,6)=1, so \phi(6)=2. This counting of coprimes to n relies on the inclusion-exclusion principle, which systematically subtracts and adds back multiples of the prime factors of n to determine the size of the set of integers up to n avoiding those factors. By convention, \phi(1)=1, as the single integer 1 is coprime to itself. The function is named after the mathematician Leonhard Euler, who introduced it in his 1763 paper Theoremata arithmetica nova methodo demonstrata.

Historical Development

Leonhard Euler introduced the totient function in his 1763 paper "Theoremata arithmetica nova methodo demonstrata" (E271), where he defined it as the number of positive integers up to n that are relatively prime to n. Euler provided initial computations for small values and established key properties, including the formula for prime powers \phi(p^m) = p^{m-1}(p-1) and the multiplicative property \phi(ab) = \phi(a)\phi(b) when a and b are coprime, laying the groundwork for its product formula \phi(n) = n \prod_{p \mid n} (1 - 1/p) developed around the same period. These contributions appeared in the context of his broader investigations into arithmetic progressions and divisor counts. In 1801, referenced Euler's function in his seminal work , introducing the notation \phi(n) and crediting Euler while expanding on its properties, such as the sum-of-divisors identity \sum_{d \mid n} \phi(d) = n. Gauss's adoption and notation helped standardize the function within , emphasizing its role in understanding the structure of integers and residues modulo n. This publication marked a pivotal milestone, integrating the totient into systematic treatments of congruences and primitive roots. The term "totient" was coined in 1879 by James Joseph Sylvester in his paper "On certain ternary cubic-form equations," where he formalized the function's study in the context of ternary cubic forms and proposed alternative notations like \tau(n), though \phi(n) prevailed. Later in the 19th century, Camille Jordan generalized the totient to higher powers with Jordan's totient function J_k(n) = n^k \prod_{p \mid n} (1 - 1/p^k) for k \geq 1, extending its applications to counting k-tuples coprime to n. By the late 19th and early 20th centuries, the totient function evolved into a cornerstone of modern , influencing Dirichlet's proofs on primes in arithmetic progressions (1837) through its appearance in L-functions and density estimates, and later in via works on the and distribution of primes. Its multiplicative nature and connections to group orders in facilitated advancements in and in the 20th century.

Notation and Terminology

The standard notation for Euler's totient function is \phi(n), where n is a positive , introduced by in his 1801 work to denote the count of integers up to n that are relatively prime to n. Earlier, Leonhard Euler described the concept without a dedicated symbol in his 1758 paper, referring to it descriptively as the "multitude of numbers less than N prime to N", and later employed \pi N in a 1775 publication to represent the same quantity. A variant form, \varphi(n) or \phi(n) with the variant phi, is also commonly used interchangeably in modern texts, while informal abbreviations like \mathrm{tot}(n) occasionally appear in computational contexts. The term "totient" itself was coined in by , who proposed the symbol \tau(n) but favored the name derived from the Latin totiens ("so many times"), evoking a sense of akin to "". Despite this, the function is most widely known as "Euler's totient function" in recognition of Euler's foundational contributions. To distinguish it from related concepts, Euler's totient function \phi(n) specifically counts coprime residues modulo n, whereas Jordan's totient functions J_k(n) for k > 1, introduced by Camille Jordan, generalize this to count k-tuples of integers up to n that are coprime to n in a componentwise manner, with J_1(n) = \phi(n).

Computation

Multiplicative Property

One of the fundamental properties of Euler's totient function \phi is its multiplicativity. Specifically, if m and n are positive integers such that \gcd(m, n) = 1, then \phi(mn) = \phi(m) \phi(n). This property arises because the integers between 1 and mn that are coprime to mn are exactly those coprime to both m and n, allowing the count of such integers to factor naturally under coprimality. A standard proof of this multiplicativity relies on the . When \gcd(m, n) = 1, the theorem provides a ring \mathbb{Z}/mn\mathbb{Z} \cong \mathbb{Z}/m\mathbb{Z} \times \mathbb{Z}/n\mathbb{Z}. This restricts to a between the multiplicative groups of units (\mathbb{Z}/mn\mathbb{Z})^\times and (\mathbb{Z}/m\mathbb{Z})^\times \times (\mathbb{Z}/n\mathbb{Z})^\times, since an element is invertible mn it is invertible both m and n. The orders of these groups are \phi(mn), \phi(m), and \phi(n), respectively, so the implies \phi(mn) = \phi(m) \phi(n). As a consequence, the value of \phi(n) for any positive n > 1 is fully determined by the of n, since n can be uniquely decomposed into coprime factors and \phi multiplies over them.

Product Formula

The explicit product for Euler's totient function \phi(n) expresses it in terms of the distinct prime factors of n. Suppose n = p_1^{k_1} p_2^{k_2} \cdots p_r^{k_r}, where the p_i are distinct primes and each k_i \geq 1. Then, \phi(n) = n \prod_{i=1}^r \left(1 - \frac{1}{p_i}\right). This closed-form expression was introduced by Leonhard Euler in his 1763 paper on arithmetic theory. It provides a direct computational tool based on the prime factorization of n, leveraging the fundamental theorem of arithmetic for its general applicability. The formula derives from the known values of \phi at prime powers, combined with the multiplicative property of the function. For a prime power p^k, the totient is \phi(p^k) = p^k - p^{k-1} = p^k \left(1 - \frac{1}{p}\right), which counts the integers from 1 to p^k that are not multiples of p. Since \phi is multiplicative—meaning \phi(mn) = \phi(m) \phi(n) when \gcd(m,n) = 1—it follows that \phi(n) = \prod_{i=1}^r \phi(p_i^{k_i}) for the prime factorization of n. Substituting the prime power expression yields the product formula. The term \left(1 - \frac{1}{p_i}\right) in the product arises from the inclusion-exclusion principle applied to the definition of \phi(n), which counts integers up to n coprime to n by excluding those sharing a with n. For the full set of distinct primes dividing n, the proportion of such is precisely \prod_{p \mid n} \left(1 - \frac{1}{p}\right), as the inclusion-exclusion expansion over the primes p_i simplifies to this product form. This connection underscores the formula's roots in methods for counting residue classes n.

Divisor Sum Formula

One alternative expression for Euler's totient function \phi(n) is the divisor sum formula, which involves a over the divisors of n weighted by the \mu: \phi(n) = \sum_{d \mid n} \mu(d) \frac{n}{d}. This representation, first derived by Mertens in , complements the product formula by providing a direct arithmetic series form. The derivation proceeds via Möbius inversion applied to the known \sum_{d \mid n} \phi(d) = n, which counts the total number of integers from 1 to n partitioned by their gcd with n. Here, the left side \phi(d) over divisors d of n, where each \phi(d) counts the integers n that are multiples of n/d and coprime to n/d. By the Möbius inversion theorem for arithmetic functions, if g(n) = \sum_{d \mid n} f(d) with g(n) = n and f = \phi, then \phi(n) = \sum_{d \mid n} \mu(d) g(n/d) = \sum_{d \mid n} \mu(d) (n/d). This inversion directly yields the formula and establishes its equivalence to the count of integers coprime to n. The appearance of the Möbius function \mu(d) in the sum encodes the inclusion-exclusion corrections for overcounting multiples of prime factors in the coprimality condition. For instance, positive terms (\mu(d) = 1) arise for square-free d with an even number of prime factors, subtracting intersections, while negative terms (\mu(d) = -1) add back triple intersections, and \mu(d) = 0 for non-square-free d eliminates higher powers. This structure mirrors the classical inclusion-exclusion principle but generalizes it via the divisor poset. Computationally, the divisor sum formula is advantageous when the prime factorization of n is unknown but the list of divisors can be enumerated efficiently, as it requires only evaluating \mu(d) for each d \mid n (by checking square-freeness and parity of distinct primes) rather than identifying the primes explicitly. For example, for n = 30, the divisors are 1, 2, 3, 5, 6, 10, 15, 30; computing \mu(1) = 1, \mu(2) = -1, \mu(3) = -1, \mu(5) = -1, \mu(6) = 1, \mu(10) = 1, \mu(15) = 1, \mu(30) = -1 yields \phi(30) = 30(1 - 1/2 - 1/3 - 1/5 + 1/6 + 1/10 + 1/15 - 1/30) = 8, verifiable against the product formula. This approach proves useful in scenarios like sieve methods or when n has many small divisors.

Other Methods

One specialized computational technique for expressing Euler's totient function \phi(n) leverages on the \mathbb{Z}/n\mathbb{Z}, particularly through transforms involving the function and Ramanujan sums. The Ramanujan sum c_n(m) is defined as c_n(m) = \sum_{\substack{1 \leq k \leq n \\ \gcd(k,n)=1}} \exp\left(2\pi i \frac{km}{n}\right), and notably, c_n(0) = \phi(n), providing a direct link to the totient via this Fourier-like sum over primitive roots of unity. A key formula arises from the of the gcd-dependent function f(k) = \gcd(k,n): \sum_{k=1}^n \gcd(k,n) \exp\left(-2\pi i \frac{k}{n}\right) = \phi(n). This equality follows from the general property that the Fourier transform F_f(m,n) = \sum_{k=1}^n f(\gcd(k,n)) \exp(-2\pi i k m / n) equals the Dirichlet convolution (f * c_{\bullet}(m))(n), where for f the identity function \mathrm{id}(k) = k and m=1, the convolution yields \phi(n) via the relation \phi = \mathrm{id} * \mu and properties of Ramanujan sums. These approaches are applied in , particularly for sieving and probabilistic evaluations of arithmetic sums. For instance, character sums decompose coprimality conditions orthogonally, facilitating estimates in sieving methods like those for the distribution of primes or lattice points free of certain factors, where \phi(n) emerges in densities or error terms. In probabilistic contexts, such expansions aid in analyzing random models over residue classes, such as estimating the probability that two integers are coprime via character averages. For large n, these methods are theoretically elegant but inefficient for direct computation compared to factorization-based techniques. The gcd-exponential sum requires O(n \log n) time to evaluate all terms, while the character sum involves \phi(n) \sim n / \log \log n terms, each needing character evaluation (potentially costly without precomputation). In contrast, the product or divisor sum formulas achieve O(\sqrt{n}) time assuming efficient factorization, though the latter remains the primary bottleneck for unfactored large n without special structure; thus, Fourier/character methods excel in aggregate analytic settings rather than isolated evaluations.

Specific Values and Examples

Values for Small Integers

The values of Euler's totient function φ(n) for small positive integers n provide a direct illustration of how many numbers up to n are coprime to n, revealing initial patterns in the distribution of totatives. The following table lists φ(n) for n from to 30, computed as the count of integers k where 1 ≤ k ≤ n and gcd(k, n) = 1; this sequence corresponds to OEIS A000010.
nφ(n)nφ(n)nφ(n)
1111102112
211242210
3213122322
42146248
541582520
621682612
7617162718
841862812
9619182928
104208308
These values highlight patterns across primes, composites, and prime powers: for a prime p, φ(p) = p - 1, as seen with φ(3) = 2, φ(5) = 4, and φ(7) = 6. For powers of 2, φ(2^k) = 2^{k-1}, evident in φ(2) = 1, φ(4) = 2, φ(8) = 4, and φ(16) = 8. Additionally, φ(n) is even for all n ≥ 3, a consistent observation starting from φ(3) = 2 onward in the table. Empirically, the of coprimes φ(n)/n decreases as n incorporates more distinct prime factors, reflecting fewer totatives relative to n; for instance, among the first values, this ratio is nearly 1 for primes like 29 (28/29 ≈ 0.965) but drops to 8/ ≈ 0.267 for the product of the first three primes. This pattern extends to s (products of the first k primes), where φ(n) grows but the φ(n)/n diminishes further due to the multiplicative accumulation of prime factors; examples include φ(6) = 2 for the 3# and φ() = 8 for the 5# .

Illustrative Computations

To illustrate the computation of Euler's totient function \phi(n) using the product formula \phi(n) = n \prod_{p \mid n} \left(1 - \frac{1}{p}\right), where the product runs over the distinct prime factors p of n, consider the following examples. This formula accounts for the prime of n and correctly handles repeated prime factors by applying the factor \left(1 - \frac{1}{p}\right) only once per distinct prime p. For n = 12, first factorize as $12 = 2^2 \times 3. Then, \phi(12) = 12 \left(1 - \frac{1}{2}\right) \left(1 - \frac{1}{3}\right) = 12 \times \frac{1}{2} \times \frac{2}{3} = 4. The integers up to 12 coprime to 12 are 1, 5, 7, and 11, confirming the count of 4. (Hardy and Wright, 1979) For n = 30 = 2 \times 3 \times 5, \phi(30) = 30 \left(1 - \frac{1}{2}\right) \left(1 - \frac{1}{3}\right) \left(1 - \frac{1}{5}\right) = 30 \times \frac{1}{2} \times \frac{2}{3} \times \frac{4}{5} = 8. The coprime integers up to 30 are 1, 7, 11, 13, 17, 19, 23, and 29. (Hardy and Wright, 1979) For a larger composite n = 100 = 2^2 \times 5^2, \phi(100) = 100 \left(1 - \frac{1}{2}\right) \left(1 - \frac{1}{5}\right) = 100 \times \frac{1}{2} \times \frac{4}{5} = 40. Note that the exponents in the prime powers do not affect the subtraction term beyond the initial factorization, as the formula subtracts the proportion divisible by each distinct prime only once. A common pitfall is mistakenly applying \left(1 - \frac{1}{p}\right) multiple times for higher powers of p, which would undercount; the correct approach uses the distinct primes as shown. (Hardy and Wright, 1979)

Core Theorems and Identities

Euler's Theorem

Euler's theorem states that if a and n > 1 are coprime positive integers, then a^{\phi(n)} \equiv 1 \pmod{n}, where \phi(n) is Euler's totient function. This result, proved by Leonhard Euler in his 1763 paper "Demonstration of a new method in the theory of arithmetic" (E271), generalizes Fermat's Little Theorem, which corresponds to the special case where n = p is prime and \phi(p) = p-1. The modern proof relies on the structure of the of integers modulo n, denoted (\mathbb{Z}/n\mathbb{Z})^\times, which consists of the residue classes coprime to n under multiplication modulo n. This group has \phi(n), as it contains exactly \phi(n) elements. Since a is coprime to n, the residue class of a belongs to this group, and the of this element divides the group \phi(n). By , raising the element to the power \phi(n) yields the , so a^{\phi(n)} \equiv 1 \pmod{n}. A key corollary is that if \gcd(a, n) = 1, then for any integer k \geq 0, a^k \equiv a^{k \mod \phi(n)} \pmod{n}; this allows efficient computation of large powers modulo n by reducing the exponent modulo \phi(n).

Menon's Identity

Menon's identity, named after P. K. Menon who proved it in 1965, relates the Euler totient function \varphi(n) to a sum involving greatest common divisors over the units modulo n. Specifically, for every positive integer n \geq 2, \sum_{\substack{1 \leq a \leq n \\ \gcd(a,n)=1}} \gcd(a-1, n) = \varphi(n) \cdot \tau(n), where \tau(n) denotes the number of positive divisors of n. This identity counts, for each possible divisor d of n, the number of reduced residues a modulo n such that \gcd(a-1, n) = d, weighted by d, and equates it to the product of the density of units and the divisor count. For example, when n = 6, the reduced residues are a = 1,5; \gcd(1-1,6) = 6 and \gcd(5-1,6) = 2, so the sum is $6 + 2 = 8, and \varphi(6) \cdot \tau(6) = 2 \cdot 4 = 8. A proof proceeds by verifying multiplicativity of both sides for coprime arguments, reducing the problem to s p^\nu. For a , the left side sums \gcd(a-1, p^\nu) over a coprime to p. The residues coprime to p^\nu are those not divisible by p, and \gcd(a-1, p^\nu) takes values p^k for k = 0 to \nu, with the number of a yielding each p^k being \varphi(p^{\nu-k}). Summing gives \sum_{k=0}^\nu p^k \cdot \varphi(p^{\nu-k}) = (\nu + 1)(p^\nu - p^{\nu-1}), which equals \varphi(p^\nu) \cdot \tau(p^\nu) = p^{\nu-1}(p-1) \cdot (\nu + 1). The general case follows by the . This approach highlights the identity's arithmetic nature without relying on group actions, though combinatorial proofs using permutations of residue classes also exist. Generalizations extend the identity to prime powers and beyond. For instance, in the of k-free residues n^k, a variant states \sum \gcd_k(a - s, n^k) = \phi_k(n) \cdot \tau(n), where \gcd_k and \phi_k are generalized to avoid multiples of the k-th power of primes dividing n. Further analogs appear in Dedekind domains and with Dirichlet characters, preserving the structure \sum \gcd(a-1, n) = \varphi(n) [\tau(n)](/page/Tau) but adapted to algebraic settings. These extensions underscore the 's role in and combinatorial .

Divisibility Properties

The divisibility properties of Euler's totient function \phi(n) concern conditions under which a fixed positive k divides \phi(n) for certain n. A fundamental result is that for any fixed positive k, there exist infinitely many positive integers n such that k \mid \phi(n). This follows from Dirichlet's theorem on primes in arithmetic progressions, which guarantees infinitely many primes q in the progression q \equiv 1 \pmod{k}. For any such prime q, \phi(q) = q - 1 \equiv 0 \pmod{k}, so k \mid \phi(q). This infinitude implies that every natural number divides \phi(n) for some n. The proof leverages the same arithmetic progression argument, ensuring at least one such prime q exists for each k. Specific examples illustrate these properties. For k = 2, \phi(n) is even for all n > 2, since if n > 2, then either n is even (so \phi(n) is even by the formula \phi(n) = n \prod_{p \mid n} (1 - 1/p)) or n has an odd prime factor p (so p-1 is even and divides \phi(n)). For an odd prime p, Dirichlet's theorem ensures infinitely many primes q \equiv 1 \pmod{p} such that p \mid q - 1 = \phi(q). These cases highlight how the structure of \phi(n) accommodates divisibility by small integers through prime choices aligned with modular conditions.

Advanced Formulas and Generating Functions

Sum Over Divisors

One of the fundamental properties of Euler's totient function \phi(n) is the \sum_{d \mid n} \phi(d) = n, where the is taken over all positive s d of n. This states that the of the totient values over the divisors of n equals n itself. To prove this , consider the set \{1, 2, \dots, n\}. Partition this set into subsets S_d = \{k \mid 1 \leq k \leq n, \gcd(k, n) = d\} for each d of n. For a fixed d \mid n, the elements k in S_d satisfy \gcd(k/d, n/d) = 1 and $1 \leq k/d \leq n/d, so the size of S_d is \phi(n/d). over all such subsets gives n = \sum_{d \mid n} |S_d| = \sum_{d \mid n} \phi(n/d). Substituting e = n/d (which also runs over all divisors of n) yields \sum_{e \mid n} \phi(e) = n. This partitioning shows that every from 1 to n belongs to exactly one subset, confirming the . The function h(n) = \sum_{d \mid n} \phi(d) is multiplicative, meaning that if \gcd(m, n) = 1, then h(mn) = h(m) h(n). This follows from the multiplicativity of \phi, as the identity holds for prime powers and extends by the Chinese Remainder Theorem. A natural generalization arises from Dirichlet convolution: for any multiplicative arithmetic function f, the function g(n) = \sum_{d \mid n} \phi(d) f(n/d) is also multiplicative. This is because the Dirichlet convolution of two multiplicative functions is multiplicative, and here \phi and f both satisfy the property. When f is the constant function f(k) = 1, this recovers the original identity g(n) = n. This identity plays a key role in Möbius inversion, which provides a reciprocal relation between sums over divisors. Specifically, since n = \sum_{d \mid n} \phi(d), applying Möbius inversion yields the explicit formula \phi(n) = \sum_{d \mid n} \mu(d) (n/d), where \mu is the . This inversion expresses \phi(n) directly in terms of the and establishes a deep connection between the totient and inclusion-exclusion principles in .

Generating Functions

The Dirichlet series for Euler's totient function \phi(n) is given by F(s) = \sum_{n=1}^\infty \frac{\phi(n)}{n^s}, which converges absolutely for \Re(s) > 2 and equals \frac{\zeta(s-1)}{\zeta(s)}, where \zeta(s) denotes the . This identity follows from the fundamental relation \sum_{d \mid n} \phi(d) = n. Applying the transform to both sides yields \sum_{n=1}^\infty \frac{1}{n^{s-1}} = \sum_{n=1}^\infty \frac{\sum_{d \mid n} \phi(d)}{n^s} = \zeta(s-1) on the left, while the right-hand side becomes \sum_{d=1}^\infty \frac{\phi(d)}{d^s} \sum_{k=1}^\infty \frac{1}{(dk)^s} = F(s) \zeta(s), since n = dk with k ranging over positive integers. Equating the two expressions gives F(s) \zeta(s) = \zeta(s-1), so F(s) = \frac{\zeta(s-1)}{\zeta(s)}. The multiplicativity of \phi(n) implies that F(s) admits an Euler product representation F(s) = \prod_p \left( \sum_{k=0}^\infty \frac{\phi(p^k)}{p^{ks}} \right), where the sum starts with the k=0 term equal to 1 (as \phi(1) = 1) and \phi(p^k) = p^k - p^{k-1} = p^{k-1}(p-1) for k \geq 1. The local factor at each prime p simplifies to \sum_{k=0}^\infty \frac{\phi(p^k)}{p^{ks}} = \frac{1 - p^{1-s}}{1 - p^{-s}}, which aligns with the ratio of zeta factors in the closed form, confirming the product structure via the Euler products for \zeta(s-1) and \zeta(s). A related generating function arises from differentiating F(s) with respect to s: \frac{d}{ds} F(s) = -\sum_{n=1}^\infty \frac{\phi(n) \log n}{n^s}. Substituting the expression for F(s) and applying the yields F'(s) = \frac{\zeta'(s-1) \zeta(s) - \zeta(s-1) \zeta'(s)}{\zeta(s)^2}, so \sum_{n=1}^\infty \frac{\phi(n) \log n}{n^s} = \frac{\zeta(s-1) \zeta'(s) - \zeta'(s-1) \zeta(s)}{\zeta(s)^2}. This series converges for \Re(s) > 2 and provides a logarithmic variant tied to the derivatives of the function. The function F(s) admits to the half-plane \Re(s) > 1 except for a simple at s=2. The originates from the simple of \zeta(s-1) at s=2, with residue \frac{1}{\zeta(2)} = \frac{6}{\pi^2}, while \zeta(s) is holomorphic and nonzero in \Re(s) \geq 1 except at s=1. Zeros of F(s) in this region occur precisely at the zeros of \zeta(s), though none exist for \Re(s) > 1. The relies on the known meromorphic of \zeta(s) to the , excluding its at s=1.

Additional Identities

One notable derived identity for Euler's totient function involves doubling the argument. Specifically, if n is odd, then \phi(2n) = \phi(n); if n is even, then \phi(2n) = 2 \phi(n). This relation arises from the behavior of the function on powers of 2 combined with its multiplicativity for coprime factors. A related identity holds for powers: for any positive integer k \geq 1, \phi(n^k) = n^{k-1} \phi(n). This provides a building on the value at the base n. For the specific form of the factorial n!, the totient function admits an exact closed form using the standard product formula over its prime factors, which are all primes p \leq n: \phi(n!) = n! \prod_{\substack{p \leq n \\ p \text{ prime}}} \left(1 - \frac{1}{p}\right). This expression counts the integers up to n! that are coprime to all primes at most n. Bijection-based identities also emerge in group theory contexts. When primitive roots exist modulo n (i.e., when (\mathbb{Z}/n\mathbb{Z})^\times is cyclic), the number of such primitive roots is exactly \phi(\phi(n)). This counts the generators of the via a with integers coprime to the group order \phi(n).

Asymptotic Behavior

Growth Rate

The asymptotic growth of Euler's totient function \phi(n) relative to n is captured by the relation \phi(n) \sim \frac{6}{\pi^2} n in the sense of average order, meaning that \lim_{x \to \infty} \frac{1}{x} \sum_{n \leq x} \frac{\phi(n)}{n} = \frac{6}{\pi^2} \approx 0.607927. This \frac{6}{\pi^2} equals \prod_p (1 - 1/p), the infinite Euler product over all primes p, which converges to the of the at 2, since \zeta(2) = \pi^2/6. The result reflects the fact that, on average, about 60.79% of the positive integers up to n are coprime to n. The proof of this asymptotic follows from evaluating the summatory function \sum_{n \leq x} \phi(n) = \frac{3}{\pi^2} x^2 + O(x \log x), which implies the average order via partial summation. One approach uses the identity \sum_{d \mid m} \phi(d) = m for each m \leq x, inverting it with the to express the sum in terms of arithmetic progressions, and then applying estimates from the distribution of primes. Alternatively, the \sum_{n=1}^\infty \phi(n) n^{-s} = \zeta(s-1)/\zeta(s) for \Re(s) > 2 can be analyzed near the pole at s=2 to derive the main term. Mertens' theorems on prime products provide the necessary control over error terms in the Euler product expansion. For individual values, \phi(n) exhibits fluctuations, but a uniform lower bound holds: there exists a constant c > 0 such that \phi(n) > c \frac{n}{\log \log n} for all integers n \geq 3. This bound arises from bounding the product \phi(n)/n = \prod_{p \mid n} (1 - 1/p) below by considering the contribution of small primes, using Mertens' estimate \prod_{p \leq y} (1 - 1/p) \sim e^{-\gamma}/\log y where \gamma is the Euler-Mascheroni constant, and optimizing over the prime factors of n. The density interpretation ties back to the constant \frac{6}{\pi^2}, as it represents the natural density of the set of integers coprime to a "typical" n, or equivalently, the probability that two randomly chosen positive integers are coprime.

Ratios of Consecutive Values

The ratios of consecutive values of Euler's totient function, \phi(n+1)/\phi(n), exhibit substantial fluctuations arising from the distinct prime factorizations of n and n+1, which are always coprime. When n is prime, \phi(n) = n-1 \approx n, but \phi(n+1) can be much smaller if n+1 has several small prime factors, yielding ratios less than 1. In the opposite case, if n is the product of the first few small primes and n+1 is prime, \phi(n) is relatively small while \phi(n+1) = n \approx n+1, producing ratios greater than 1 that grow with the number of prime factors in n. Examples illustrate this behavior. For n=7 (prime), \phi(7) = 6, and \phi(8) = \phi(2^3) = 4, so \phi(8)/\phi(7) = 4/6 = 2/3. For n=8, \phi(9) = \phi(3^2) = 6, so \phi(9)/\phi(8) = 6/4 = 1.5. Larger ratios occur for n=30 = 2 \times 3 \times 5, \phi(30) = 8, and n+1=31 prime, \phi(31) = 30, giving \phi(31)/\phi(30) = 30/8 = 3.75. Similarly, for n=210 = 2 \times 3 \times 5 \times 7, \phi(210) = 48, and n+1=211 prime, \phi(211) = 210, yielding \phi(211)/\phi(210) = 210/48 \approx 4.375. For n=2310 = 2 \times 3 \times 5 \times 7 \times 11, \phi(2310) = 480, and n+1=2311 prime, \phi(2311) = 2310, so \phi(2311)/\phi(2310) = 2310/480 = 4.8125. These examples demonstrate how the ratio can deviate from 1 based on . More generally, the extreme values of \phi(m)/m for m \approx n imply that \limsup_{n \to \infty} \phi(n+1)/\phi(n) = \infty and \liminf_{n \to \infty} \phi(n+1)/\phi(n) = 0, as the relative totient can approach 1 or its minimal order e^{-\gamma} / \log \log n independently for coprime arguments. The average value of the is approximately 1, consistent with the asymptotic \sum_{k=1}^n \phi(k) \sim 3n^2 / \pi^2, which implies the typical size of \phi(n) is (6/\pi^2) n \approx 0.607 n, so consecutive ratios fluctuate around 1 on average.

Totient Numbers

The totient summatory function, denoted T(n) or \Phi(n), is defined as T(n) = \sum_{k=1}^n \varphi(k), where \varphi denotes Euler's totient function. This cumulative sum aggregates the number of positive integers up to each k that are coprime to k, providing a measure of the total "coprime counts" up to n. A key interpretive role of T(n) arises in the enumeration of rational numbers: it precisely counts the number of reduced fractions a/b with $1 \leq b \leq n, $1 \leq a \leq b, and \gcd(a, b) = 1. For each fixed denominator b, there are exactly \varphi(b) such numerators a, yielding the total via the summation. This connection highlights the function's utility in problems involving farey sequences and rational approximations, where the distinct fractions in (0, 1] with bounded denominators are cataloged. Asymptotically, T(n) \sim \frac{3}{\pi^2} n^2 as n \to \infty, reflecting the average order of \varphi(k) \approx \frac{6}{\pi^2} k and the quadratic growth from . This equivalence underscores the of coprime pairs near the of n^2. Regarding basic properties, T(n) is strictly increasing because \varphi(k) \geq 1 for all k \geq 1, ensuring T(n+1) = T(n) + \varphi(n+1) > T(n). Simple bounds follow from the asymptotic, such as c_1 n^2 < T(n) < c_2 n^2 for constants $0 < c_1 < \frac{3}{\pi^2} < c_2 and sufficiently large n, though tighter error terms involve the Riemann hypothesis.

Ford's Theorem

Ford's theorem, published in 1998, establishes a fundamental result on the multiplicities of values in the image of Euler's totient function φ. Specifically, for every integer k ≥ 2, there exists a positive integer m such that the equation φ(n) = m has exactly k positive integer solutions n. This theorem resolves several open questions about the structure of the set of totient numbers, demonstrating that the multiplicity function A(m) = |{n : φ(n) = m}| attains every value k ≥ 2 infinitely often. The proof is largely constructive, relying on the selection of suitable primes in specific arithmetic progressions (residue classes) to build preimages n with controlled totient values and to ensure the precise number of such preimages. By leveraging sieve methods and estimates on the distribution of primes in residue classes, Ford constructs examples for each k, showing that the image of φ is "flexible" in terms of how many times each value is attained. A key aspect of the structure of totient values highlighted by Ford's work is the distinction between odd and even totients. Odd values of φ(n) are exceptional and occur only when n = 1 or n = 2 or n is a product of distinct (primes of the form 2^{2^j} + 1 for j ≥ 0) or twice such a product. Since only five Fermat primes are known (3, 5, 17, 257, and 65537), the possible odd totients are limited to finitely many forms, such as 1, 2, 4, 6, 10, 12, 16, 18, 20, 24, 28, 30, 32, 34, 36, 40, 42, 48, 54, 60, 64, 72, 80, 84, 90, 96, 108, 120, 128, 144, 160, 168, 180, 192, 216, 240, 256, 288, 320, 336, 360, 384, 432, 480, 512, 576, 640, 672, 720, 768, 864, 960, 1024, and so on up to products involving the known Fermat primes. This characterization severely restricts the odd part of the image of φ. Ford's results have significant implications for the surjectivity of φ onto the even positive integers. While it remains an open problem whether every sufficiently large even integer is a totient (i.e., whether φ is surjective onto the even numbers above some bound), Ford's analysis shows that the set of even totients is asymptotically dense among the even integers in a precise sense: the number of distinct even totients up to x is asymptotic to c \frac{x}{\log x} \exp\left( C (\log_3 x - \log_4 x)^2 + C' \log_3 x \right) for explicit constants C ≈ 0.8178 and C' ≈ 2.177, where the exponential factor grows slowly but ensures that nearly all even integers with a "typical" prime factorization structure (around (\log \log \log x)^2 distinct prime factors in their preimages) are attained. This supports the conjecture that non-totient even integers, if they exist beyond the known finite list (such as 14, 26, 34, 38, 50), must be sparse and possess unusual prime factorizations, often avoiding certain residue classes or having too many small prime factors. The constructive nature of the proof further implies that for even m with bounded number of prime factors (up to 6 or so, with finite exceptions), m is a totient, advancing understanding toward potential surjectivity.

Perfect Totient Numbers

A perfect totient number is a positive integer n for which the sum of the iterated applications of , starting from \phi(n) and including the terminal 1, equals n. Formally, if \phi^{(k)}(n) denotes the k-th iterate of \phi, then n is perfect totient if n = 1 + \sum_{k=1}^{c} \phi^{(k)}(n), where c is the smallest integer such that \phi^{(c)}(n) = 2 (with \phi(2) = 1). This sum terminates before the cycle at 1 to avoid divergence. The powers of 3 form an infinite family of perfect totient numbers. For n = 3, \phi(3) = 2, so the sum is $2 + 1 = 3. For n = 9, \phi(9) = 6, \phi(6) = 2, so the sum is $6 + 2 + 1 = 9. For n = 27, \phi(27) = 18, \phi(18) = 6, \phi(6) = 2, so the sum is $18 + 6 + 2 + 1 = 27. This pattern holds for all $3^k with k \geq 1, as the iteration chain scales accordingly. Other small perfect totient numbers include 15, 39, 111, 183, 255, and 327. For n=15: \phi(15) = 8, \phi(8) = 4, \phi(4) = 2, so the sum is $8 + 4 + 2 + 1 = 15. For n=39, \phi(39) = 24, \phi(24) = 8, \phi(8) = 4, \phi(4) = 2, so the sum is $24 + 8 + 4 + 2 + 1 = 39. For n=255, the iteration chain is \phi(255) = 128, \phi(128) = 64, \phi(64) = 32, \phi(32) = 16, \phi(16) = 8, \phi(8) = 4, \phi(4) = 2, so the sum is $128 + 64 + 32 + 16 + 8 + 4 + 2 + 1 = 255. All perfect totient numbers are odd, as even n lead to sums less than n due to \phi(n) \leq n/2 for even n > 2, and the iteration decreases rapidly. Constructions for additional perfect totient numbers include numbers of the form 3p, where p is a prime of the form 4 \times 3^N + 1 for some integer N. Examples include p = (giving ), p = 37 (giving 111), and p = (giving 327). Further families exist, such as those from Venkataraman's result involving certain large primes times 3. Up to 5 \times 10^9, there are 30 perfect totient numbers not powers of 3, plus 9 more from the construction. While there are infinitely many perfect totient numbers due to the powers of 3, it remains an open question whether there are infinitely many perfect totient numbers that are not powers of 3, as known constructions rely on the existence of primes in specific arithmetic progressions whose infinitude is unresolved. No perfect totient numbers of the form 3^k p with k \geq 4 exist under certain conditions on p, but the general case is open.

Applications

In Cyclotomy and Geometry of Numbers

In the theory of cyclotomic fields, Euler's totient function \phi(n) gives the degree of the extension \mathbb{Q}(\zeta_n)/\mathbb{Q}, where \zeta_n is a primitive nth root of unity. This field is the nth cyclotomic field, and its minimal polynomial over \mathbb{Q} is the nth cyclotomic polynomial, which has degree \phi(n). The Galois group \mathrm{Gal}(\mathbb{Q}(\zeta_n)/\mathbb{Q}) is isomorphic to the multiplicative group of units modulo n, (\mathbb{Z}/n\mathbb{Z})^\times, whose order is precisely \phi(n). For a prime p, this yields \phi(p) = p-1, representing the order of the Galois group of the pth cyclotomic extension. The totient function also features prominently in the constructibility of regular polygons using straightedge and compass. According to the Gauss-Wantzel , a regular n-gon is constructible n = 2^k \prod p_i for distinct Fermat primes p_i and nonnegative k, a condition equivalent to \phi(n) being a power of 2. This equivalence arises because the vertices of the n-gon lie in the real subfield of \mathbb{[Q](/page/Q)}(\zeta_n), which has degree \phi(n)/2 over \mathbb{[Q](/page/Q)}, and constructible numbers generate extensions of degree a power of 2. Gauss established the sufficiency, while Wantzel proved the necessity in 1837. In the geometry of numbers, \phi(n) appears in the study of primitive lattice points—those in \mathbb{Z}^d whose coordinates are coprime, meaning \gcd(x_1, \dots, x_d) = 1. The enumeration of such points in bounded regions, such as polygons or balls, often involves sums of \phi(m) over divisors m, reflecting the proportion of coprime residue classes modulo m. Minkowski's theorems on convex bodies guarantee the existence of nonzero lattice points in symmetric convex sets of sufficient volume, and adaptations to primitive points incorporate densities \phi(n)/n to account for coprimality conditions relative to a lattice basis. For instance, in \mathbb{Z}^2, the number of primitive points (with \gcd(x,y)=1) with coordinates up to N in absolute value is asymptotically \frac{6}{\pi^2} (2N+1)^2 \approx \frac{24 N^2}{\pi^2}, linking the totient to the distribution of visible points from the origin.

In Analytic Number Theory

In analytic number theory, Euler's totient function φ(q) plays a central role in the distribution of prime numbers within arithmetic progressions. The prime number theorem for arithmetic progressions states that for integers q > 1 and a with 1 ≤ a ≤ q and gcd(a, q) = 1, the number of primes π(x; q, a) up to x that are congruent to a q satisfies \pi(x; q, a) \sim \frac{\mathrm{Li}(x)}{\phi(q)}, where Li(x) is the logarithmic integral, asymptotically equivalent to x / log x. This result, established by de la Vallée Poussin in 1896, extends the classical and quantifies the equal distribution of primes among the φ(q) residue classes coprime to q. The proof relies on Dirichlet L-functions L(s, χ) associated to characters χ modulo q. For the principal character χ_0, L(s, χ_0) behaves like the Riemann zeta function ζ(s), yielding the main term, while the non-vanishing of L(1, χ) for all non-principal characters χ modulo q ensures no exceptional biases that would disrupt the equidistribution. Orthogonality of characters implies that the total density of primes across these φ(q) classes sums to the overall prime density, so each class receives a proportion 1/φ(q). This non-vanishing at s=1 was first proved by Dirichlet for the infinitude of primes in progressions and extended analytically by de la Vallée Poussin using contour integration and growth estimates on L(s, χ). A potential obstruction arises from real non-principal characters, where an exceptional zero β close to 1 could cause larger errors in the asymptotic. Siegel's theorem addresses this by providing an ineffective bound: if L(β, χ) = 0 for a real primitive character χ modulo q with β > 1 - c / log(q φ(q)), then c > 0 depends on ε in an ineffective way, ensuring that such zeros, if they exist, do not significantly affect the for most q. This bound implies controlled exceptions in the error term for π(x; q, a) - Li(x)/φ(q), with the relative error o(1) holding uniformly for q up to exp(c sqrt(log x)) under stronger assumptions, as refined in the Siegel-Walfisz theorem. Generalizations extend this framework to higher moments of the error term or twisted sums involving primes in progressions, such as moments of ψ(x; q, a) - x/φ(q), where effective non-vanishing results from zero-density estimates improve the φ(q)-weighted bounds. These developments, building on work by and others, connect φ(n) to subconvexity bounds for L-functions and applications in .

In Cryptography

Euler's totient function plays a central role in the Rivest–Shamir–Adleman () cryptosystem, a foundational public-key scheme. In , two large distinct prime numbers p and q are selected, and the modulus n is computed as their product, n = pq. The totient \phi(n) = (p-1)(q-1) is then calculated privately by the key generator. A public exponent e is chosen such that $1 < e < \phi(n) and \gcd(e, \phi(n)) = 1, ensuring e is coprime to \phi(n). The private exponent d is derived as the modular multiplicative inverse of e modulo \phi(n), satisfying ed \equiv 1 \pmod{\phi(n)}. The public key consists of the pair (n, e), which is shared openly, while the private key (n, d) is kept secret. For encryption, a message m (with $0 \leq m < n) is transformed into ciphertext c = m^e \mod n. Decryption recovers the original message via m = c^d \mod n. This process relies on Euler's theorem, which states that if \gcd(m, n) = 1, then m^{\phi(n)} \equiv 1 \pmod{n}, implying m^{k\phi(n) + 1} \equiv m \pmod{n} for any integer k. Since d \equiv e^{-1} \pmod{\phi(n)}, it follows that c^d = (m^e)^d = m^{ed} = m^{k\phi(n) + 1} \equiv m \pmod{n}. The security of RSA fundamentally depends on the computational difficulty of factoring the modulus n back into its prime factors p and q. Without knowledge of p and q, an adversary cannot efficiently compute \phi(n), which is required to determine the private exponent d from the public e and n. This hardness of the integer factorization problem underpins RSA's resistance to attacks, as no efficient classical algorithm is known for factoring sufficiently large semiprimes. Variants of RSA, such as multi-prime RSA, extend the scheme by using more than two primes to form n, accelerating decryption for certain applications while maintaining reliance on totient computation for key generation. In multi-prime RSA, \phi(n) is computed as the product \prod (p_i - 1) over all distinct primes p_i dividing n, with the Chinese Remainder Theorem often applied to speed up operations using partial totients. Padding schemes like Optimal Asymmetric Encryption Padding (OAEP) are incorporated in modern implementations to enhance security against chosen-ciphertext attacks, but the core totient-based key setup remains unchanged.

Unsolved Problems

Lehmer's Conjecture

Lehmer's conjecture, proposed by in 1932, posits that if the φ(n) divides n − 1 for a positive integer n, then n must be prime. This unsolved problem, often called Lehmer's totient problem, seeks to determine whether there exist any composite "Lehmer numbers" satisfying φ(n) | (n − 1). Lehmer himself proved that any such composite n, if it exists, must be odd and square-free with at least seven distinct . Subsequent work has strengthened these conditions significantly. Any composite Lehmer number must be a Carmichael number (satisfying λ(n) | (n − 1), where λ is the Carmichael function) with at least 15 distinct prime factors, and n > 10^{30}. Computational searches have confirmed no such composite n exists up to 10^{30}. The existence of a Lehmer number would have important implications for the theory of primitive roots, as φ(n) | (n − 1) would imply that the exponent of the modulo n divides n − 1, precluding the existence of a primitive root for composite n in this class (consistent with known non-cyclic structure for s greater than 2). It would also represent a particularly strong form of , where the full totient value divides n − 1 rather than just the Carmichael exponent.

Carmichael's Conjecture

Carmichael's conjecture, proposed by Robert D. Carmichael in 1922, asserts that the Euler totient function \phi has no unique values in its range. Specifically, for any positive integer m, the preimage \phi^{-1}(m) is either empty or contains at least two distinct positive integers n; in other words, |\phi^{-1}(m)| \neq 1 for all m. This implies that if \phi(n) = m for some n, then there exists at least one other k \neq n such that \phi(k) = m. The conjecture remains unsolved, but extensive evidence supports its validity. Computational checks have confirmed it for all m up to extraordinarily large bounds: Schlafly and verified in 1994 that no counterexamples exist for m < 10^{10{,}000{,}000}. On the theoretical front, Kevin Ford's 1999 analysis established that any counterexample m must exceed \exp(\exp(36.62)), an immense threshold far beyond current computational reach. Ongoing searches for counterexamples continue, driven by efficient algorithms to compute totient preimages. The multiplicativity of \phi plays a key role in understanding why even totient values (all except m=1) tend to have multiple preimages. For instance, if m is even and \phi(n)=m with n odd, then \phi(2n)=m as well, since \phi(2n)=\phi(2)\phi(n)=m when n is odd. This property ensures that many even m automatically satisfy the conjecture. Ford's theorem complements this by quantifying the abundance of totients, proving that the normal order of the number of solutions |\phi^{-1}(m)| grows like e^\gamma \log\log\log m + O(1/\log\log\log m), implying that singletons, if they exist, are exceptionally rare. The irregularities in the values of Euler's totient function φ(n) for certain highly composite numbers, particularly primorials, are conjecturally linked to the distribution of zeros of the Riemann zeta function. Specifically, Jean-Louis Nicolas proved that the Riemann hypothesis is equivalent to the inequality φ(N_k) < N_k / (e^γ log log N_k) holding for all sufficiently large primorials N_k = ∏{i=1}^k p_i (the product of the first k primes), particularly for k ≥ 30, where γ is the Euler-Mascheroni constant. This criterion arises from the asymptotic behavior of the product ∏{p | N_k} (1 - 1/p)^{-1} ≈ e^γ log log N_k, derived from Mertens' theorems, and the inequality ensures that the totient ratio does not exceed the expected value based on prime distribution under the hypothesis. If the Riemann hypothesis is false, the inequality fails for infinitely many k, leading to φ(N_k) > N_k / (e^γ log log N_k) for those primorials. Unconditionally, explicit lower bounds for φ(n) provide partial insight into these irregularities. Rosser and Schoenfeld established that φ(n) > n / (e^γ log log n + 3 / log log n) for all n ≥ 3. This implies φ(n)/n > 1 / (e^γ log log n + 3 / log log n) ≈ 0.56 / log log n for large n. Under the generalized (GRH), stronger bounds are available, such as φ(n) > c n (log log log n) / log log n for some c > 0 and sufficiently large n, reflecting improved control over prime gaps and distributions in arithmetic progressions. The connection to the zeta function stems from the fact that errors in prime-counting analogs for φ(n) depend on zero-free regions of ζ(s). The summatory function ∑_{n ≤ x} φ(n) = 3x^2 / π^2 + E(x), where the error E(x) satisfies E(x) = O(x log x) unconditionally, but under the , E(x) = O(x^{1/2} log^2 x). Deviations in E(x) are tied to nontrivial zeros off the critical line, and analogous irregularities would propagate to individual φ(n) values through Dirichlet series representations like ∑ φ(n)/n^s = ζ(s-1)/ζ(s). If the Riemann hypothesis is false, a nontrivial zero with real part β > 1/2 would imply large deviations in prime distributions, causing φ(n)/n to exceed its asymptotic expectation for infinitely many primorial n, resulting in unexpectedly high counts of integers coprime to n relative to n. This would manifest as significant fluctuations in coprime proportions, disrupting applications relying on uniform bounds for φ(n).

References

  1. [1]
    Totient Function -- from Wolfram MathWorld
    The totient function phi(n), also called Euler's totient function, is defined as the number of positive integers <=n that are relatively prime to.
  2. [2]
    [PDF] Number Theory: Introduction to Euler's Totient Function
    Jun 13, 2023 · Number theory is the study of relationships and properties of numbers. We learned about and worked with prime numbers and congruence ...
  3. [3]
    Math Origins: The Totient Function
    Leonhard Euler's totient function, φ(n), is an important object in number theory, counting the number of positive integers less than or equal to n which are ...
  4. [4]
    [PDF] The Role of Number Theory in Modern Cryptography
    2.8 Euler's Totient ϕ function. The Totient function is defined as the number of positive integers less than or equal to n that are relatively prime to . Two ...
  5. [5]
    [PDF] Class 17 Principle of Inclusion-Exclusion Euler's Function
    Principle of Inclusion-Exclusion. Euler's Function φ(n). Let φ(n) be the number of positive integers x ≤ n which are mutually prime to n i.e. have no common ...Missing: source | Show results with:source
  6. [6]
    [PDF] theoremata arithmetica - nova methodo demonstrata.
    THEOREMATA ARITHMETICA. NOVA METHODO DEMONSTRATA. Practe. Auctore. L. EVLER'Ò. ▻. + racter varias computandi operationes, quae vulgo in Arithmetica tradi ...
  7. [7]
  8. [8]
    Disquisitiones arithmeticae : Gauss, Carl Friedrich, 1777-1855
    Aug 11, 2018 · Publication date: 1801. Topics: Number theory. Publisher: Lipsiae : In commiss. apud Gerh. Fleischer, jun. Collection: smithsonian.Missing: totient function
  9. [9]
    A000010 - OEIS
    A000010 - OEIS. Euler totient function phi(n): count numbers <= n and prime to n. Number of elements in a reduced residue system modulo n. a(n) equals the ...<|control11|><|separator|>
  10. [10]
    [PDF] Dr. Z.'s Number Theory Lecture 15 Handout: Euler's Totient Function
    Very important property: φ(n) is multiplicative, in other words, φ(mn) = φ(m)φ(n) , if gcd(m, n)=1 . Proof: By the Chinese Remainder Theorem, the mapping ...<|control11|><|separator|>
  11. [11]
    [PDF] 4 Euler's Totient Function
    Euler's totient function φ : N → N is defined by2 φ(n) = \. \{0 < a ≤ n ... Here are the first few values of Euler's function; we also list the units.
  12. [12]
  13. [13]
    [PDF] 6.6. The Inclusion-Exclusion Principle and Euler's Function
    Feb 27, 2022 · In this section, we state (without a general proof) the Inclusion-Exclusion. Principle (in Corollary 6.57) concerning the cardinality of the ...Missing: source | Show results with:source
  14. [14]
    [PDF] The Möbius Function and Möbius Inversion - TigerWeb
    Mar 11, 2021 · Today, Möbius inversion concerns a different kind of sum: a divisor sum of an arithmetic function. ... Euler's totient function ϕ(n).<|control11|><|separator|>
  15. [15]
    [PDF] Möbius Inversion Formula. Multiplicative Functions - UC Berkeley math
    Taking the sum-functions of these, we obtain the relations: Sµ = e, Se = I, SI = τ, and Sid = σ. These examples suggest that the sum-function is multiplicative, ...
  16. [16]
    [PDF] notes on m¨obius inversion and inclusion-exclusion - garsia at york
    d|n φ(n/d)kd . One way to give a formula for the Euler phi function is to use the principle of inclusion- exclusion to show that.
  17. [17]
    [PDF] THE FOURIER TRANSFORM OF FUNCTIONS OF THE GREATEST ...
    The Ramanujan sum cn(m) is multiplicative in n and the Dirichlet convolution preserves the multiplicativity of functions. ! Here are some more examples ...
  18. [18]
    [PDF] Section 3, Dirichlet's theorem 1 Introduction. - NYU Courant
    The Euler product formulas in this section apply to Dirichlet series called L ... The Euler totient function is φ(n) = |Gn|. Figure 3 lists. Gn and φ(n) for ...
  19. [19]
    [PDF] Dirichlet characters and L-functions Dirichl
    Here ϕ is the Euler phi (“totient”) function, ϕ(q) = |(Z/qZ)∗|. We have just proved the cases (q, a) = (4, ±1) of Dirichlet's theorem. The same method.
  20. [20]
    Euler's Totient Theorem -- from Wolfram MathWorld
    A generalization of Fermat's little theorem. Euler published a proof of the following more general theorem in 1736. Let phi(n) denote the totient function.
  21. [21]
    [PDF] Euler's Theorem - Trinity University
    We can now state and prove our main result. Theorem 2 For any n ∈ N, if (a,n)=1, then a&(n) ≡ 1 (mod n). Proof. If (a,n) = 1, then a + nZ ∈ (Z/nZ)×.
  22. [22]
    IAAWA Euler's Theorem - UTK Math
    Let Z n × be the group of units modulo . n . Then . | Z n × | = ϕ ( n ) . 🔗. The following theorem is an important result in number theory, due to Leonhard ...
  23. [23]
    Proofs, generalizations and analogs of Menon's identity: a survey
    Oct 14, 2021 · Menon's identity states that for every positive integer n one has \sum (a-1,n) = \varphi(n) \tau(n), where a runs through a reduced residue ...
  24. [24]
    [PDF] Euler's ϕ function
    Idea: Most numbers have many different prime factors, so most values of ϕ are divisible by a large power of 2. But most numbers are not divisible by a large ...
  25. [25]
    [PDF] Chapter 3 Dirichlet series and arithmetic functions
    (ii) We first have to verify that the convolution product of two multiplicative functions is again multiplicative. Here we use that if m, n are two coprime ...
  26. [26]
    [PDF] The Mobius Function and Mobius Inversion - Ursinus Digital Commons
    Jan 17, 2021 · F. Mertens. Ueber einige asymptotische Gesetze der Zahlentheorie. Journal für die Reine und. Angewandte Mathematik. (Crelle's Journal), 77:289 – ...
  27. [27]
    [PDF] An Lntroduction To The Theory Of Numbers Third Edition
    Our purpose is to present a reasonably complete introduction to the theory of numbers within the compass of a single volume The basic concepts are.
  28. [28]
    Introduction to Analytic Number Theory - SpringerLink
    This book is the first volume of a two-volume textbook for undergraduates and is indeed the crystallization of a course offered by the author at the California ...
  29. [29]
    [PDF] Fixed points for discrete logarithms - Dartmouth Mathematics
    where τ0(n) denotes the number of squarefree divisors of n. Since ϕ(n) > cn/log log n for all n > 2 and some positive constant c, and since τ0(n) = Oǫ(nǫ) ...
  30. [30]
    [PDF] Euler Totient Function Inequality - arXiv
    Mar 24, 2018 · Let d(n) = Pd|n1. An integer n ∈ N is called highly composite if and only if d(m) < d(n) for all integers m<n. Definition 6.2. Let σ(n) = Pd|nd.
  31. [31]
    [PDF] Arithmetic and Asymptotic Properties of Restricted Totient Sums - arXiv
    Sep 10, 2025 · This paper studies restricted sums, Υ(x, p) and ∆(x, p), related to Euler's totient function, and their arithmetic and asymptotic properties, ...<|control11|><|separator|>
  32. [32]
    The Distribution of Totients | The Ramanujan Journal
    Ford, K. The Distribution of Totients. The Ramanujan Journal 2, 67–151 (1998). https://doi.org/10.1023/A:1009761909132. Download citation. Issue Date: March ...
  33. [33]
    [PDF] Some remarks on Euler's totient function - HAL
    Jul 18, 2012 · If p is an odd prime such that p − 1 divides 2k, then p must be of the form 2i + 1. The only primes of this form are Fermat numbers, hence if n ...
  34. [34]
    [PDF] On Perfect Totient Numbers
    Define the arithmetic function S by S(n) = φ(n) + φ2(n) + ··· + φc(n) + 1, where φc(n) = 2. We say n is a perfect totient number if S(n) = n. We give a list of ...Missing: properties | Show results with:properties
  35. [35]
    A082897 - OEIS
    327 is a perfect totient number because 327 = 216 + 72 + 24 + 8 + 4 + 2 + 1. Note that 216 = phi(327), 72 = phi(216), 24 = phi(72) and so on.
  36. [36]
    Sequence A082897: Perfect Totient Numbers - Robert Munafo
    Mar 28, 2022 · If a prime P is of the form P=4×3N+1 for some integer N, then 3P is a perfect totient number[1]. The primes P of this form are: 5, 13, 37, 109, ...Missing: properties | Show results with:properties
  37. [37]
    A Note On Perfect Totient Numbers
    In this note we prove that there are no perfect totient numbers of the form 3kp, k ≥ 4, where s = 2a 3b + 1, r = 2c 3d s + 1, q = 2e 3f r + 1, ...
  38. [38]
    [PDF] Section V.8. Cyclotomic Extensions
    Jan 2, 2016 · The Euler phi function is defined on N as ϕ(n) equals the number of elements in the set {1,2,...,n} which are relatively prime to n. For example ...Missing: totient | Show results with:totient
  39. [39]
    [PDF] 7 Cyclotomic Extensions - Columbia Math Department
    If w is any root of unity, then the field extension F(w)j F is called a cyclotomic extension. We point out two facts about roots of unity.
  40. [40]
    Why was Wantzel overlooked for a century? The changing ...
    Aug 10, 2025 · Gauss went on to demonstrate that a regular n-gon is constructible if Euler's totient ϕ(n) is a power of 2. He stopped short of proving that ...<|control11|><|separator|>
  41. [41]
    [PDF] arXiv:1509.02201v2 [math.NT] 31 Jul 2019
    Jul 31, 2019 · Key words and phrases. Primitive points in polygons, visible points, Euler's Totient function, Error term, rational polygons. 1 ...
  42. [42]
    [PDF] The Prime Number Theorem
    His paper treats not only the case of the arithmetic progression but also that of a binary quadratic form. de la Vallée Poussin concludes that ((1+it) 0 from ...
  43. [43]
    [PDF] First proof of L(1,χ) 6= 0 1. First proof of non-vanishing on Re(s)=1
    Sep 24, 2013 · The subtle element in Dirichlet's theorem about primes in arithmetic progressions is non-vanishing of L- functions L(s, χ) at s = 1. Here ...<|control11|><|separator|>
  44. [44]
    Chapter 11 Error bounds for primes in arithmetic progressions
    A zero of L ( s , χ ) arising as an exception in Theorem 11.6 (necessarily for χ a real character) is called an exceptional zero , or Siegel zero , of .
  45. [45]
    [PDF] SIEGEL ZERO Out line: 1. Introducing the problem of existence of ...
    Abstract. We talk about the effect of the positions of the zeros of Dirichlet L-function in the distribution of prime numbers in arithmetic progressions.
  46. [46]
    Primes in arithmetic progressions - Kiran S. Kedlaya
    For any positive integers m , N with , gcd ( m , N ) = 1 , the set of primes congruent to m modulo N has Dirichlet density 1 / ϕ ( N ) in the set of all primes ...Missing: phi( q)
  47. [47]
    [PDF] Method for Obtaining Digital Signatures & Public-Key Cryptosystems
    R.L. Rivest, A. Shamir, and L. Adleman. ∗. Abstract. An encryption method is presented with the novel property that publicly re- vealing an encryption key ...
  48. [48]
    [PDF] History of integer factorization - Purdue Computer Science
    The integer factorization problem is well known as the mathematical founda- tion of the public-key cryptosystem RSA. When RSA was introduced forty.
  49. [49]
    [PDF] On the security of multi-prime RSA
    Multi-prime RSA is a variant of RSA in which the modulus is the product of more than two distinct primes. In this work we collect the strongest known algebraic ...
  50. [50]
    On Euler's totient function - Project Euclid
    October 1932 On Euler's totient function. D. H. Lehmer · DOWNLOAD PDF + SAVE TO MY LIBRARY. Bull. Amer. Math. Soc. 38(10): 745-751 (October 1932).
  51. [51]
    Richard Pinch: Publications
    Lehmer's Totient Problem asks whether there is an integer n such that φ ... Back to Richard Pinch's personal home page or mathematics page. Updated 20 ...
  52. [52]
    A new perspective on Lehmer's totient problem - MathOverflow
    Jan 27, 2021 · Lehmer's totient problem asks if there are any composite integers n with φ(n) | n−1. It is known that any such n must be odd. It must also be a charmichael ...nt.number theory - Lehmer's totient problemGeneralized Lehmer Euler ConjectureMore results from mathoverflow.net
  53. [53]
    Lehmer's totient problem over Fq[x] - ScienceDirect
    Lehmer's totient problem is to determine the set of Lehmer numbers. To the best of our knowledge, the current best result is due to Richard G.E. Pinch (see [9]) ...
  54. [54]
    carmichael's conjecture on the euler function
    Let the multiplicity of an integer be the number of times it occurs as a value of <j)(x), where </> is the Euler function. Table 1 shows the multiplicities ...
  55. [55]
    Solving φ(x)=n, where φ(x) is Euler's totient function - testing ...
    Sep 9, 2010 · Carmichael's Totient Function Conjecture (1922, first stated in reference 6 below) states that Euler's function takes each value at least twice.
  56. [56]
  57. [57]