Fact-checked by Grok 2 weeks ago

Liouville function

The Liouville function \lambda(n) is a completely multiplicative defined on the positive integers n by \lambda(n) = (-1)^{\Omega(n)}, where \Omega(n) denotes the total number of prime factors of n counted with multiplicity (and \Omega(1) = 0, so \lambda(1) = 1). Introduced by the French mathematician (1809–1882) in his foundational work on arithmetic identities and quadratic forms during the and , the function encodes the of the exponent sum in the prime of n, taking the value $1if\Omega(n)is even and-1$ if odd. The Dirichlet series associated with the Liouville function is \sum_{n=1}^\infty \lambda(n) n^{-s} = \zeta(2s)/\zeta(s) for \Re(s) > 1, where \zeta(s) is the Riemann zeta function; this identity highlights its deep ties to analytic number theory and the distribution of primes. The function is completely multiplicative, meaning \lambda(mn) = \lambda(m)\lambda(n) for all positive integers m and n, and it appears in various identities involving divisor functions and theta series, such as the Lambert series expansion \sum_{n=1}^\infty \frac{\lambda(n) x^n}{1 - x^n} = \sum_{n=-\infty}^\infty x^{n^2} for |x| < 1. A central object of study is the summatory function L(x) = \sum_{n \leq x} \lambda(n), whose asymptotic behavior is intimately linked to the Riemann hypothesis (RH): RH holds if and only if L(x) = O(x^{1/2 + \varepsilon}) for every \varepsilon > 0, as shown by Edmund Landau in his 1899 thesis exploring prime number distribution. (Note: This equivalence stems from the prime number theorem and zero-free regions of \zeta(s).) Early conjectures, like George Pólya's 1919 assertion that L(n) \leq 0 for all n \geq 2, were disproved in 1958 by R. M. Haselgrove (who showed the existence of counterexamples), with the first explicit counterexample at n = 906,180,359 identified by R. Sherman Lehman in 1960 and the smallest at n = 906,150,257 found by Minoru Tanaka in 1980; these developments underscore the function's role in probing the oscillatory nature of prime-related sums.

Definition and Basic Properties

Definition

The Liouville function \lambda is a completely multiplicative defined for positive integers n in terms of the prime n = p_1^{a_1} p_2^{a_2} \cdots p_k^{a_k} by \lambda(n) = (-1)^{a_1 + a_2 + \cdots + a_k}, where the exponent a_1 + a_2 + \cdots + a_k = \Omega(n) denotes the total number of prime factors of n counted with multiplicity. This definition assigns \lambda(n) = 1 if \Omega(n) is even and \lambda(n) = -1 if \Omega(n) is odd. As a completely , \lambda satisfies \lambda(mn) = \lambda(m)\lambda(n) for all positive integers m and n, regardless of whether m and n are coprime. This property follows directly from the multiplicative nature of \Omega, since \Omega(mn) = \Omega(m) + \Omega(n). For example, \lambda(1) = 1 since $1 has no prime factors (\Omega(1) = 0, an even number). For a prime p, \lambda(p) = -1 (\Omega(p) = 1). For p^2, \lambda(p^2) = 1 (\Omega(p^2) = 2). For distinct primes pandq, \lambda(pq) = 1 (\Omega(pq) = 2$). The Liouville function is named after the French mathematician (1809–1882), who introduced it in his work on arithmetic identities and quadratic forms during his number-theoretic investigations from 1858 to 1865. Unlike the \mu(n), which depends only on the number of distinct prime factors, \lambda(n) accounts for multiplicity.

Multiplicativity and Values

The Liouville function \lambda(n) is completely multiplicative, meaning that \lambda(mn) = \lambda(m) \lambda(n) for all positive integers m and n. This property follows directly from its definition in terms of the total number of prime factors. Specifically, since the function \Omega(n), which counts the prime factors of n with multiplicity, is completely additive—satisfying \Omega(mn) = \Omega(m) + \Omega(n) for all m, n—it follows that \lambda(mn) = (-1)^{\Omega(mn)} = (-1)^{\Omega(m) + \Omega(n)} = (-1)^{\Omega(m)} (-1)^{\Omega(n)} = \lambda(m) \lambda(n). The explicit \lambda(n) = (-1)^{\Omega(n)} ties the directly to the prime of n, where \Omega(1) = 0 and thus \lambda(1) = [1](/page/1). For prime powers, \lambda(p^k) = (-1)^k, alternating between -1 for odd exponents and [1](/page/1) for even exponents. This connection highlights how \lambda(n) encodes the of the total exponent sum in the prime of n. To illustrate, the values of \lambda(n) for small n up to 20 reveal patterns tied to the of \Omega(n):
nPrime factorization\Omega(n)\lambda(n)
110
2$2
3$3
4$2^2
5$5
6$2 \cdot 3
7$7
8$2^3
9$3^2
10$2 \cdot 5
11$11
12$2^2 \cdot 3
13$13
14$2 \cdot 7
15$3 \cdot 5
16$2^4
17$17
18$2 \cdot 3^2
19$19
20$2^2 \cdot 5
These values show \lambda(n) = 1 when \Omega(n) is even (e.g., squares or products of even numbers of distinct primes) and -1 when odd. Computing \lambda(n) requires first obtaining the prime factorization of n to determine \Omega(n), after which the value follows immediately as (-1)^{\Omega(n)}. Using trial division for factorization, this can be achieved in O(\sqrt{n}) time in the worst case, by checking divisibility by all integers up to \sqrt{n}. For computing \lambda(k) for all k \leq N, a sieve method analogous to the can determine the factorizations in O(N \log \log N) total time, making it efficient for large ranges. The alternating nature of \lambda on prime powers plays a key role in inclusion-exclusion principles, as seen in the partial sums \sum_{j=0}^{k} \lambda(p^j) = \sum_{j=0}^{k} (-1)^j, which equals 1 if k is even and 0 if k is odd, facilitating cancellations in arithmetic identities involving powers.

Analytic Connections

Dirichlet Series Representation

The Dirichlet series associated to the Liouville function \lambda(n) is defined by L(s) = \sum_{n=1}^\infty \frac{\lambda(n)}{n^s}, where the series converges absolutely for \Re(s) > 1. Because \lambda(n) is completely multiplicative, with \lambda(p^k) = (-1)^k for each p^k, the possesses an Euler product expansion over the primes p. The local Euler factor at each prime is the \sum_{k=0}^\infty \frac{\lambda(p^k)}{p^{ks}} = \sum_{k=0}^\infty (-1)^k p^{-ks} = \frac{1}{1 + p^{-s}}, which converges whenever |p^{-s}| < 1, or equivalently \Re(s) > 0. Thus, for \Re(s) > 1, L(s) = \prod_p \frac{1}{1 + p^{-s}}. This representation follows directly from the multiplicativity of \lambda(n), as the Euler product interchanges with the absolutely convergent in this half-plane. The local factor can equivalently be expressed as \frac{1}{1 + p^{-s}} = \frac{1 - p^{-s}}{1 - p^{-2s}}, yielding the alternative Euler product form L(s) = \prod_p \frac{1 - p^{-s}}{1 - p^{-2s}} for \Re(s) > 1. This rewriting highlights the connection to expansions akin to those for the , though without invoking it explicitly. The absolute convergence of the in \Re(s) > 1 stems from the boundedness |\lambda(n)| = 1 for all n, mirroring the convergence abscissa of \zeta(s). While the local factors are holomorphic in \Re(s) > 0, the infinite product only converges for \Re(s) > 1. A full meromorphic continuation to the entire follows from the relation to the , as detailed below.

Relation to Riemann Zeta Function

The Dirichlet series associated with the Liouville function \lambda(n) satisfies the identity \sum_{n=1}^\infty \frac{\lambda(n)}{n^s} = \frac{\zeta(2s)}{\zeta(s)} for \Re(s) > 1. This relation arises from equating the Euler products: the zeta function has the product representation \zeta(s) = \prod_p (1 - p^{-s})^{-1}, so \zeta(2s) = \prod_p (1 - p^{-2s})^{-1} and $1/\zeta(s) = \prod_p (1 - p^{-s}); the local Euler factor for the Liouville series is \prod_p \sum_{k=0}^\infty (-1)^k p^{-ks} = \prod_p (1 - p^{-s}) / (1 - p^{-2s}), yielding the overall product \zeta(2s)/\zeta(s). The analytic structure of the series inherits properties from the zeta function: upon meromorphic continuation, it exhibits poles at the zeros of \zeta(s) and zeros at the poles of \zeta(2s), though the latter occur outside the half-plane \Re(s) > 1 (specifically, the pole of \zeta(2s) is at s = 1/2). Briefly, the logarithmic derivative provides further insight: \frac{d}{ds} \log \left( \sum_{n=1}^\infty \frac{\lambda(n)}{n^s} \right) = -\frac{\zeta'(s)}{\zeta(s)} + 2 \frac{\zeta'(2s)}{\zeta(2s)}. This follows directly from differentiating the logarithm of the identity.

Summatory Functions

The Liouville Summatory Function

The Liouville summatory function is defined as the partial sum L(x) = \sum_{n \leq x} \lambda(n) for real x \geq 1, where \lambda(n) is the Liouville function and the sum runs over positive integers n \leq x. This function measures the net contribution of the oscillatory signs of \lambda(n), reflecting the parity of the total number of prime factors (with multiplicity) across integers up to x. The Dirichlet series associated with \lambda(n) is \sum_{n=1}^\infty \lambda(n) n^{-s} = \zeta(2s)/\zeta(s) for \operatorname{Re}(s) > 1, where \zeta(s) is the Riemann zeta function. Using Perron's formula, this yields an approximate integral representation for L(x): L(x) \approx \frac{1}{2\pi i} \int_{c-iT}^{c+iT} \frac{\zeta(2s)}{\zeta(s)} \frac{x^s}{s} \, ds with c > 1 and suitable T > 0, plus an error term depending on T and the growth of the integrand. An elementary bound follows immediately from the : |L(x)| \leq x. The function L(x) also displays oscillatory properties arising from the sign changes of \lambda(n), which occur whenever the total exponent sum \Omega(n) shifts parity; these changes contribute to frequent crossings of the by L(x). For computational purposes, L(x) can be evaluated efficiently for large x via the Dirichlet hyperbola method, exploiting the identity derived from inversion: L(x) = \sum_{k \leq \sqrt{x}} M\left( \frac{x}{k^2} \right), where M(y) = \sum_{m \leq y} \mu(m) is the ; values of M(y) are then computed using sieving algorithms on segments to handle large ranges. This approach allows calculations up to x \approx 10^9 in feasible time on mid-20th-century , with optimizations extending to much larger x.

Asymptotic Behavior

The summatory Liouville function L(x) = \sum_{n \leq x} \lambda(n) has sublinear growth, with the equivalent to the assertion that L(x) = o(x) as x \to \infty. This equivalence follows from the representation of the Liouville function and the pole of the at s = 1. The classical unconditional bound, due to de la Vallée Poussin in 1899, is L(x) = O\left( x \exp\left( -c \sqrt{\log x} \right) \right) for some absolute constant c > 0, obtained using a zero-free for \zeta(s). Subsequent refinements leveraged improved zero-free regions of \zeta(s) to demonstrate stricter sublinear growth. The strongest unconditional upper bound, as of , follows from the Vinogradov-Korobov zero-free region and is L(x) = O\left( x \exp\left( -c (\log x)^{3/5} (\log \log x)^{-1/5} \right) \right) for some absolute constant c > 0. This estimate arises from applying Perron's formula to the \sum \lambda(n) n^{-s} = \zeta(2s)/\zeta(s) and bounding the contribution from the critical strip using the zero-free region. Historical improvements progressed from de la Vallée Poussin's classical zero-free region, yielding O(x \exp(-c \sqrt{\log x})), to more refined regions like Vinogradov-Korobov that incorporate logarithmic factors for sharper exponents. The average order of L(x) aligns with o(x), reflecting the cancellation inherent in the oscillatory nature of \lambda(n) and consistent with equivalents of the . Additionally, L(x) changes sign infinitely often, a result established through analytic arguments involving the distribution of primes and explicit constructions of intervals where the of prime factors leads to dominance of positive or negative contributions.

Conjectures and Applications

Equivalence to Riemann Hypothesis

The (RH) is equivalent to the assertion that the summatory Liouville function satisfies L(x) = O(x^{1/2 + \varepsilon}) for every \varepsilon > 0, or equivalently, L(x) = O(\sqrt{x} (\log x)^k) for any fixed k > 0. This equivalence was noted by in his 1899 doctoral thesis, building on his earlier work on , with key refinements provided by Albert Ingham in the 1930s that clarified the role of zero-free regions and the distribution of zeros. To establish the direction RH implies the bound, consider the Dirichlet series for the Liouville function, given by \sum_{n=1}^\infty \lambda(n) n^{-s} = \zeta(2s)/\zeta(s) for \Re(s) > 1. This series admits an analytic continuation to the complex plane except for branch points related to the zeros of \zeta(s). Applying Perron's formula expresses L(x) as a contour integral over this series. Under RH, all non-trivial zeros of \zeta(s) lie on the line \Re(s) = 1/2, so shifting the contour to the left into the critical strip yields contributions from these zeros that are controlled by subconvexity bounds or density estimates, resulting in the desired O(x^{1/2 + \varepsilon}) growth after estimating the horizontal and vertical segments of the contour. Conversely, suppose the bound L(x) = O(x^{1/2 + \varepsilon}) holds for all \varepsilon > 0. If is false, there exists a non-trivial zero \rho = \beta + it of \zeta(s) with \beta > 1/2. The residue at s = \rho in the Dirichlet series \zeta(2s)/\zeta(s) contributes a term approximately x^{\rho}/\rho to L(x) via the inverse or Perron's formula, leading to |L(x)| \gg x^{\beta} for values of x near e^{2\pi k / t} for integers k, which contradicts the assumed bound since \beta > 1/2 + \varepsilon/2 for small \varepsilon. Thus, no such zero can exist off the critical line. A failure of RH would imply L(x) = \Omega(x^{\theta}) for some \theta > 1/2, violating the bound L(x) = O(x^{1/2 + \varepsilon}) for every \varepsilon > 0, though no explicit to RH has been constructed via this route.

Weighted Summatory Functions

Weighted summatory functions of the Liouville function extend the unweighted summatory function L(x) = \sum_{n \leq x} \lambda(n) by incorporating weights to probe finer distributional properties. A general form is S(x) = \sum_{n \leq x} \lambda(n) f(n), where f(n) is a weight such as f(n) = \log n or f(n) = n^\alpha for real \alpha. An important specific case is the power-weighted sum L_\alpha(x) = \sum_{n \leq x} \lambda(n) n^{-\alpha} for $0 \leq \alpha < 1/2, which reduces to L(x) when \alpha = 0. Under the , the linear independence hypothesis over the rationals, and a conjecture on moments of the Riemann zeta function, L_\alpha(x) admits a limiting logarithmic distribution for these \alpha, exhibiting a negative bias while taking positive values on a set of positive logarithmic density. A Pólya-Vinogradov-type bound is conjectured for the logarithmically weighted sum, asserting that \sum_{n \leq x} \lambda(n) \log n = O(\sqrt{x} \log x); this is motivated by parallels to the prime number theorem and non-vanishing estimates for L-functions, suggesting controlled growth analogous to character sum bounds. The development of bounds for weighted summatory functions traces back to early conjectures on the unweighted case. In 1919, Pólya conjectured that L(x) \leq 0 for all x \geq 2, implying a persistent negative bias in the parity of prime factors. This was disproved in 1958 by Haselgrove, who established that L(x) > 0 infinitely often using density arguments on the zeros of \zeta(2s)/\zeta(s). The disproof highlighted the oscillatory nature of L(x) and spurred Chowla's 1965 conjecture on sign patterns, which posits that the Liouville function behaves pseudorandomly: for any distinct nonnegative integers h_1, \dots, h_k, \sum_{n \leq x} \prod_{i=1}^k \lambda(n + h_i) = o(x) as x \to \infty, implying all $2^k sign patterns occur with equal logarithmic density. Computational verifications provide evidence for subdued growth in weighted sums. For the case \alpha = 1/2, L_{1/2}(x) \leq 0 holds for all $17 \leq x \leq 10^{12}, supporting a related of Mossinghoff and Trudgian. Larger-scale checks for the unweighted L(x) up to x \approx 2 \times 10^{14} reveal oscillations aligning with \Omega(\sqrt{x} \log \log \log x / \log x) growth under the , with no violations of expected sign changes. Counterexamples to related sums, such as those in Pólya's original , appear sporadically but with density consistent with logarithmic distributions. Weighted variants of Liouville summatory functions connect to problems in prime , including gaps and twin primes. Generalizations of Heath-Brown's on twin primes under the absence of zeros extend to Liouville sums, yielding unconditional results on the density of twin primes when weighted forms exhibit certain cancellation properties.

Generalizations

Extensions to Other Arithmetic Functions

The Liouville function can be extended to a twisted version by incorporating a χ, defined as λ_χ(n) = λ(n) χ(n), where λ(n) is the standard Liouville function and χ is a non-principal . This arises naturally in the study of sign patterns and correlations, particularly for real non-principal characters, and its is given by L(2s, χ)/L(s, χ). In the setting of number fields, the Liouville function generalizes to non-zero integral ideals in Dedekind domains, such as the ring of integers of a number field K. For an ideal a, it is defined as λ_K(a) = (-1)^{Ω(a)}, where Ω(a) counts the total number of prime ideal factors of a with multiplicity. The associated summatory function L_K(x) = ∑_{N(a) ≤ x} λ_K(a), with N(a) the norm of a, relates to the Dedekind zeta function ζ_K(s) via the identity ∑ λ_K(a) N(a)^{-s} = ζ_K(2s)/ζ_K(s). Higher-degree analogs of the Liouville function are defined for k ≥ 1 as λ_k(n) = (-1)^{Ω(n^k)}, which equals [λ(n)]^k since Ω(n^k) = k Ω(n). These functions link to powers of the Riemann zeta function, with the Dirichlet series ∑ λ_k(n) n^{-s} = ζ(2s)/ζ(s) if k is odd and ζ(s) if k is even. In quadratic fields, the ideal-theoretic extension λ_K(a) appears in analyses of class number problems, particularly for the nine imaginary quadratic fields ℚ(√(-d)) with class number one, where d = 1, 2, 3, 7, 11, 19, 43, 67, 163. Properties of λ_K and associated L-functions contribute to modern studies of class numbers.

Applications in Analytic Number Theory

The Liouville function plays a foundational role in early through Franz Mertens' 1874 theorem that the harmonic sum of primes up to x satisfies ∑_{p ≤ x} 1/p = log log x + B + o(1), where B ≈ 0.261497 is the Meissel–Mertens constant, providing an asymptotic density estimate for primes that prefigures the . This result, derived using elementary methods and partial , highlighted the in prime distribution and influenced subsequent developments in prime counting. In modern proofs of the , the Liouville function contributes via its {n=1}^∞ λ(n)/n^s = ζ(2s)/ζ(s), where ζ(s) is the . The pole of ζ(s) at s=1 and the zero-free region near this point yield the main term in the asymptotic for the summatory function L(x) = ∑{n ≤ x} λ(n), with the equivalent to L(x) = o(x) as x → ∞. This relation refines error terms in π(x) ∼ Li(x), as oscillations in L(x) reflect contributions from non-trivial zeros of ζ(s), aiding explicit bounds on prime gaps and distribution. The identity ∑_{d ∣ n} λ(d) = 1 if n is a and 0 otherwise underpins applications in , where the Liouville function detects in the number of prime factors to address the parity problem. This challenge arises because standard sieves bound sums over even or odd factor counts but struggle to isolate primes ( parity, single factor); the Liouville function's to sets illustrates this barrier, limiting upper bounds for primes in short intervals. In arithmetic progressions, refinements of the Liouville function, such as twists by Dirichlet characters q, yield biases in prime factorizations and improved upper bounds via weighted sieves, as seen in analyses of exceptional biases under the Elliott-Halberstam conjecture. Generalizations of the Liouville function to Dirichlet L-functions, defined via ∑ λ_χ(n) n^{-s} = L(2s, χ)/L(s, χ) for a χ, facilitate the study of zeros in the critical strip. These twisted variants reveal distribution patterns analogous to the classical case, with their summatory functions probing zero spacings and supporting bounds on Landau-Siegel zeros. Post-2018 developments, including the proof by Rodgers and that the de Bruijn–Newman constant Λ ≥ 0 using methods on ζ(s) zeros, advance toward the Newman conjecture that Λ = 0 (equivalent to the ). As of 2025, upper bounds on Λ have been improved to very small positive values (e.g., via Polymath15 efforts), but it remains open. Additionally, random matrix theory models correlations in L(x), predicting Gaussian unitary ensemble statistics for its extreme values, akin to zeta zero spacings, and enabling simulations of prime-like fluctuations.

References

  1. [1]
    Number Theory in the Spirit of Liouville
    Number Theory in the Spirit of Liouville. Number Theory in the Spirit of Liouville ... functions, convolution sums involving the divisor functions, and many ...
  2. [2]
    On Liouville's Function - jstor
    If it is assumed that these zeros are all simple, then the function t(2s)/(s (s)) has simple poles at I + iy,, for n = 0, +1, ?2, *.. with residues. 1 ~ (2p,) a ...
  3. [3]
    DLMF: §27.2 Functions ‣ Multiplicative Number Theory ‣ Chapter ...
    ⓘ Defines: λ ⁡ ( n ) : Liouville's function Symbols: ν ⁡ ( n ) : number of distinct primes dividing a number, n : positive integer and a : primitive root ...<|control11|><|separator|>
  4. [4]
    Joseph Liouville (1809 - 1882) - Biography - University of St Andrews
    Liouville investigated criteria for integrals of algebraic functions to be algebraic during the period 1832-33. Having established this in four papers, ...
  5. [5]
    Liouville Function -- from Wolfram MathWorld
    The function lambda(n)=(-1)^(Omega(n)), (1) where Omega(n) is the number of not necessarily distinct prime factors of n, with Omega(1)=0.Missing: definition | Show results with:definition
  6. [6]
    A008836 - OEIS
    ### Summary of A008836 Sequence Values (Liouville Lambda Function) up to n=20
  7. [7]
    Integer factorization - Algorithms for Competitive Programming
    May 29, 2025 · Trial division. This is the most basic algorithm to find a prime factorization. We divide by each possible divisor . It can be observed that it ...
  8. [8]
    [PDF] the distribution of weighted sums of the liouville function and pólya's ...
    n=1 λ(n)n−s has the Euler product. ∞. X n=1 λ(n) ns. = Yp. 1. 1 + p−s. , for ℜ ... Mossinghoff, “Sign Changes in Sums of the. Liouville Function”, Mathematics of ...
  9. [9]
    [PDF] a note on p´olya's observation concerning liouville's function
    It is well-known, and follows easily from the Euler product for the. Riemann zeta-function ζ(s), that λ(n) has the Dirichlet generating function. ∞. X n=1 λ(n).
  10. [10]
    The Dirichlet Series for the Liouville Function and the Riemann ...
    This paper investigates the analytic properties of the Liouville function's Dirichlet series that obtains from the function F (s) ≡ ζ(2s)/ζ(s), where s is a ...
  11. [11]
    [PDF] ON SOME HISTORICAL ASPECTS OF THE THEORY OF RIEMANN ...
    Our historical sight regards multiplicative number theory because the involvement of the Riemann zeta function is mainly motivated by prime number theory and ...Missing: Joseph | Show results with:Joseph
  12. [12]
    The Dirichlet Series for the Liouville Function and the Riemann ...
    Sep 11, 2016 · This paper investigates the analytic properties of the Liouville function's Dirichlet series that obtains from the function F(s)= zeta(2s)/zeta( ...Missing: continuation | Show results with:continuation
  13. [13]
    None
    ### Key Parts Extracted
  14. [14]
    [PDF] arXiv:1606.08021v1 [math.NT] 26 Jun 2016
    Jun 26, 2016 · The Liouville function λ is defined by setting λ(n)=1if n is composed of an even number of prime factors (counted with multiplicity) and −1 ...
  15. [15]
    254A, Notes 2: Complex-analytic multiplicative number theory
    Dec 9, 2014 · We turn to the complex approach to multiplicative number theory, in which the focus is instead on obtaining various types of control on the Dirichlet series.
  16. [16]
    The Distribution of Weighted Sums of the Liouville Function ... - arXiv
    Aug 7, 2011 · We prove the existence of a limiting logarithmic distribution of the normalisation of the weighted sum of the Liouville function.Missing: functions | Show results with:functions
  17. [17]
    Pólya Conjecture -- from Wolfram MathWorld
    Lehman (1960) found the first explicit counterexample, L(906180359)=1 , and ... "A Numerical Investigation on Cumulative Sum of the Liouville Function" [sic].
  18. [18]
    A disproof of a conjecture of Pólya | Mathematika | Cambridge Core
    Feb 26, 2010 · A disproof of a conjecture of Pólya. Published online by Cambridge University Press: 26 February 2010. C. B. Haselgrove.Missing: disproved | Show results with:disproved
  19. [19]
    [PDF] THE LOGARITHMICALLY AVERAGED CHOWLA AND ... - Terry Tao
    Let λ denote the Liouville function. The Chowla con- jecture, in the two ... Using the prime number theorem in arithmetic progressions with. Vinogradov-Korobov ...
  20. [20]
    [PDF] the distribution of weighted sums of the liouville function and pólya's ...
    The weighted sum of the Liouville function with weight α ∈ R is the summatory function. Lα(x) = X n≤x λ(n) nα . When α = 0, this is the summatory function ...
  21. [21]
    What is the largest computed summatory liouville interval
    May 15, 2012 · I am interested to know the largest computed summatory liouville interval, an implementation of which is detailed in Section 4.1 of [1].Exact formula for partial sums of Liouville function $L(n)$ (OEIS ...Bound on a scaled sum of the Liouville function - MathOverflowMore results from mathoverflow.net
  22. [22]
    [PDF] Partial sums of the Liouville function and further topics in analytic ...
    This manuscript based thesis contains two original publications [Chi21a, Chi21b] in the field of analytic number theory. In Section 1, we give a brief ...Missing: origin | Show results with:origin
  23. [23]
    [PDF] sign patterns of the liouville and m¨obius functions
    Dirichlet character: Theorem 2.6 (Twisted Liouville or Möbius in short intervals). Let χ be a fixed real. Dirichlet character. For any ε > 0 and any ...
  24. [24]
  25. [25]
    [PDF] class number one from analytic rank two
    by Landau [47] and Siegel [58], but the result remained ineffective. One ... 134 (1908), 95–143. http://eudml.org/doc/149285. Page 34. 34. MARK WATKINS.
  26. [26]
    [PDF] The Prime Number Theorem - Penn State University
    (Landau 1908b) Let R be the set of positive integers that can be expressed ... function can be used to derive an associated bound for the error term in the.
  27. [27]
    A conjectural local Fourier-uniformity of the Liouville function
    Dec 2, 2015 · One can of course do similar things with arithmetic functions whose Dirichlet series are tied to other standard L-functions, e.g. one can twist ...
  28. [28]
    Open question: The parity problem in sieve theory - Terry Tao
    Jun 5, 2007 · The parity problem is a notorious problem in sieve theory: this theory was invented in order to count prime patterns of various types (eg twin primes).<|separator|>
  29. [29]
    Biases in prime factorizations and Liouville functions for arithmetic ...
    Apr 26, 2017 · We introduce a refinement of the classical Liouville function to primes in arithmetic progressions. Using this, we discover new biases in the ...Missing: upper bounds
  30. [30]
    [1801.05914] The De Bruijn-Newman constant is non-negative - arXiv
    Jan 18, 2018 · In this paper we establish Newman's conjecture. The argument proceeds by assuming for contradiction that \Lambda < 0, and then analyzing the dynamics of zeroes ...Missing: Liouville | Show results with:Liouville