Farey sequence

Last updated
Farey diagram to F9 represented with circular arcs. In the SVG image, hover over a curve to highlight it and its terms. Farey diagram horizontal arc 9.svg
Farey diagram to F9 represented with circular arcs. In the SVG image, hover over a curve to highlight it and its terms.
Farey diagram to F9. Farey diagram square 9.svg
Farey diagram to F9.
Symmetrical pattern made by the denominators of the Farey sequence, F9. Farey sequence denominators 9.svg
Symmetrical pattern made by the denominators of the Farey sequence, F9.
Symmetrical pattern made by the denominators of the Farey sequence, F25. Farey sequence denominators 25.svg
Symmetrical pattern made by the denominators of the Farey sequence, F25.

In mathematics, the Farey sequence of order n is the sequence of completely reduced fractions, either between 0 and 1, or without this restriction, [lower-alpha 1] which when in lowest terms have denominators less than or equal to n, arranged in order of increasing size.

Contents

With the restricted definition, each Farey sequence starts with the value 0, denoted by the fraction 0/1, and ends with the value 1, denoted by the fraction 1/1 (although some authors omit these terms).

A Farey sequence is sometimes called a Farey series, which is not strictly correct, because the terms are not summed. [2]

Examples

The Farey sequences of orders 1 to 8 are :

F1 = { 0/1,1/1 }
F2 = { 0/1,1/2,1/1 }
F3 = { 0/1,1/3,1/2,2/3,1/1 }
F4 = { 0/1,1/4,1/3,1/2,2/3,3/4,1/1 }
F5 = { 0/1,1/5,1/4,1/3,2/5,1/2,3/5,2/3,3/4,4/5,1/1 }
F6 = { 0/1,1/6,1/5,1/4,1/3,2/5,1/2,3/5,2/3,3/4,4/5,5/6,1/1 }
F7 = { 0/1,1/7,1/6,1/5,1/4,2/7,1/3,2/5,3/7,1/2,4/7,3/5,2/3,5/7,3/4,4/5,5/6,6/7,1/1 }
F8 = { 0/1,1/8,1/7,1/6,1/5,1/4,2/7,1/3,3/8,2/5,3/7,1/2,4/7,3/5,5/8,2/3,5/7,3/4,4/5,5/6,6/7,7/8,1/1 }
Centered
F1 = { 0/1,1/1 }
F2 = { 0/1,1/2,1/1 }
F3 = { 0/1,1/3,1/2,2/3,1/1 }
F4 = { 0/1,1/4,1/3,1/2,2/3,3/4,1/1 }
F5 = { 0/1,1/5,1/4,1/3,2/5,1/2,3/5,2/3,3/4,4/5,1/1 }
F6 = { 0/1,1/6,1/5,1/4,1/3,2/5,1/2,3/5,2/3,3/4,4/5,5/6,1/1 }
F7 = { 0/1,1/7,1/6,1/5,1/4,2/7,1/3,2/5,3/7,1/2,4/7,3/5,2/3,5/7,3/4,4/5,5/6,6/7,1/1 }
F8 = { 0/1,1/8,1/7,1/6,1/5,1/4,2/7,1/3,3/8,2/5,3/7,1/2,4/7,3/5,5/8,2/3,5/7,3/4,4/5,5/6,6/7,7/8,1/1 }
Sorted
 F1 = {0/1,                                                                                                          1/1}  F2 = {0/1,                                                   1/2,                                                   1/1}  F3 = {0/1,                               1/3,                1/2,                2/3,                               1/1}  F4 = {0/1,                     1/4,      1/3,                1/2,                2/3,      3/4,                     1/1}  F5 = {0/1,                1/5, 1/4,      1/3,      2/5,      1/2,      3/5,      2/3,      3/4, 4/5,                1/1}  F6 = {0/1,           1/6, 1/5, 1/4,      1/3,      2/5,      1/2,      3/5,      2/3,      3/4, 4/5, 5/6,           1/1}  F7 = {0/1,      1/7, 1/6, 1/5, 1/4, 2/7, 1/3,      2/5, 3/7, 1/2, 4/7, 3/5,      2/3, 5/7, 3/4, 4/5, 5/6, 6/7,      1/1}  F8 = {0/1, 1/8, 1/7, 1/6, 1/5, 1/4, 2/7, 1/3, 3/8, 2/5, 3/7, 1/2, 4/7, 3/5, 5/8, 2/3, 5/7, 3/4, 4/5, 5/6, 6/7, 7/8, 1/1} 

Farey sunburst

Plotting F6 numerators vs denominators Sunburst 8.png
Plotting F6 numerators vs denominators
Starbursts of iterations 1-10 superimposed Farey sunbursts 1-10.svg
Starbursts of iterations 1–10 superimposed

Plotting the numerators versus the denominators of a Farey sequence gives a shape like the one to the right, shown for F6.

Reflecting this shape around the diagonal and main axes generates the Farey sunburst, shown below. The Farey sunburst of order n connects the visible integer grid points from the origin in the square of side 2n, centered at the origin. Using Pick's theorem, the area of the sunburst is 4(|Fn|1), where |Fn| is the number of fractions in Fn.

Farey sunburst of order 6, with 1 interior (red) and 96 boundary (green) points giving an area of
html.skin-theme-clientpref-night .mw-parser-output div:not(.notheme)>.tmp-color,html.skin-theme-clientpref-night .mw-parser-output p>.tmp-color,html.skin-theme-clientpref-night .mw-parser-output table:not(.notheme) .tmp-color{color:inherit!important}@media(prefers-color-scheme:dark){html.skin-theme-clientpref-os .mw-parser-output div:not(.notheme)>.tmp-color,html.skin-theme-clientpref-os .mw-parser-output p>.tmp-color,html.skin-theme-clientpref-os .mw-parser-output table:not(.notheme) .tmp-color{color:inherit!important}}
1 +
96/2 - 1 = 48, according to Pick's theorem Farey sunburst 6.svg
Farey sunburst of order 6, with 1 interior (red) and 96 boundary (green) points giving an area of 1 + 96/2 − 1 = 48, according to Pick's theorem

History

The history of 'Farey series' is very curious — Hardy & Wright (1979) [3]
... once again the man whose name was given to a mathematical relation was not the original discoverer so far as the records go. — Beiler (1964) [4]

Farey sequences are named after the British geologist John Farey, Sr., whose letter about these sequences was published in the Philosophical Magazine in 1816. Farey conjectured, without offering proof, that each new term in a Farey sequence expansion is the mediant of its neighbours. Farey's letter was read by Cauchy, who provided a proof in his Exercices de mathématique, and attributed this result to Farey. In fact, another mathematician, Charles Haros, had published similar results in 1802 which were not known either to Farey or to Cauchy. [4] Thus it was a historical accident that linked Farey's name with these sequences. This is an example of Stigler's law of eponymy.

Properties

Sequence length and index of a fraction

The Farey sequence of order n contains all of the members of the Farey sequences of lower orders. In particular Fn contains all of the members of Fn1 and also contains an additional fraction for each number that is less than n and coprime to n. Thus F6 consists of F5 together with the fractions 1/6 and 5/6.

The middle term of a Farey sequence Fn is always 1/2, for n > 1. From this, we can relate the lengths of Fn and Fn1 using Euler's totient function  :

Using the fact that |F1| = 2, we can derive an expression for the length of Fn: [5]

where is the summatory totient.

We also have :

and by a Möbius inversion formula  :

where μ(d) is the number-theoretic Möbius function, and is the floor function.

The asymptotic behaviour of |Fn| is :

The index of a fraction in the Farey sequence is simply the position that occupies in the sequence. This is of special relevance as it is used in an alternative formulation of the Riemann hypothesis, see below. Various useful properties follow:

The index of where and is the least common multiple of the first numbers, , is given by: [6]

Farey neighbours

Fractions which are neighbouring terms in any Farey sequence are known as a Farey pair and have the following properties.

If a/b and c/d are neighbours in a Farey sequence, with a/b < c/d, then their difference c/d  a/b is equal to 1/bd. Since

this is equivalent to saying that

.

Thus 1/3 and 2/5 are neighbours in F5, and their difference is 1/15.

The converse is also true. If

for positive integers a,b,c and d with a < b and c < d then a/b and c/d will be neighbours in the Farey sequence of order max(b,d).

If p/q has neighbours a/b and c/d in some Farey sequence, with

then p/q is the mediant of a/b and c/d in other words,

This follows easily from the previous property, since if bpaq = qcpd = 1, then bp + pd = qc + aq, p(b + d) = q(a + c), p/q = a + c/b + d.

It follows that if a/b and c/d are neighbours in a Farey sequence then the first term that appears between them as the order of the Farey sequence is incremented is

which first appears in the Farey sequence of order b + d.

Thus the first term to appear between 1/3 and 2/5 is 3/8, which appears in F8.

The total number of Farey neighbour pairs in Fn is 2|Fn|  3.

The Stern–Brocot tree is a data structure showing how the sequence is built up from 0 (= 0/1) and 1 (= 1/1), by taking successive mediants.

Equivalent-area interpretation

Every consecutive pair of Farey rationals have an equivalent area of 1. [7] See this by interpreting consecutive rationals r1 = p/q and r2 = p′/q′ as vectors (p, q) in the x–y plane. The area of A(p/q, p′/q′) is given by qp′ − qp. As any added fraction in between two previous consecutive Farey sequence fractions is calculated as the mediant (⊕), then A(r1, r1r2) = A(r1, r1) + A(r1, r2) = A(r1, r2) = 1 (since r1 = 1/0 and r2 = 0/1, its area must be 1).

Farey neighbours and continued fractions

Fractions that appear as neighbours in a Farey sequence have closely related continued fraction expansions. Every fraction has two continued fraction expansions in one the final term is 1; in the other the final term is greater by 1. If p/q, which first appears in Farey sequence Fq, has continued fraction expansions

[0; a1, a2, ..., an 1, an, 1]
[0; a1, a2, ..., an 1, an + 1]

then the nearest neighbour of p/q in Fq (which will be its neighbour with the larger denominator) has a continued fraction expansion

[0; a1, a2, ..., an]

and its other neighbour has a continued fraction expansion

[0; a1, a2, ..., an 1]

For example, 3/8 has the two continued fraction expansions [0; 2, 1, 1, 1] and [0; 2, 1, 2], and its neighbours in F8 are 2/5, which can be expanded as [0; 2, 1, 1]; and 1/3, which can be expanded as [0; 2, 1].

Farey fractions and the least common multiple

The lcm can be expressed as the products of Farey fractions as

where is the second Chebyshev function. [8] [9]

Farey fractions and the greatest common divisor

Since the Euler's totient function is directly connected to the gcd so is the number of elements in Fn,

For any 3 Farey fractions a/b, c/d and e/f the following identity between the gcd's of the 2x2 matrix determinants in absolute value holds: [10]

[6]

Applications

Farey sequences are very useful to find rational approximations of irrational numbers. [11] For example, the construction by Eliahou [12] of a lower bound on the length of non-trivial cycles in the 3x+1 process uses Farey sequences to calculate a continued fraction expansion of the number log2(3).

In physical systems with resonance phenomena, Farey sequences provide a very elegant and efficient method to compute resonance locations in 1D [13] and 2D. [14]

Farey sequences are prominent in studies of any-angle path planning on square-celled grids, for example in characterizing their computational complexity [15] or optimality. [16] The connection can be considered in terms of r-constrained paths, namely paths made up of line segments that each traverse at most rows and at most columns of cells. Let be the set of vectors such that , , and , are coprime. Let be the result of reflecting in the line . Let . Then any r-constrained path can be described as a sequence of vectors from . There is a bijection between and the Farey sequence of order given by mapping to .

Ford circles

Comparison of Ford circles and a Farey diagram with circular arcs for n from 1 to 9. Each arc intersects its corresponding circles at right angles. In the SVG image, hover over a circle or curve to highlight it and its terms. Comparison Ford circles Farey diagram.svg
Comparison of Ford circles and a Farey diagram with circular arcs for n from 1 to 9. Each arc intersects its corresponding circles at right angles. In the SVG image, hover over a circle or curve to highlight it and its terms.

There is a connection between Farey sequence and Ford circles.

For every fraction p/q (in its lowest terms) there is a Ford circle C[p/q], which is the circle with radius 1/(2q2) and centre at (p/q, 1/ 2q2 ). Two Ford circles for different fractions are either disjoint or they are tangent to one another—two Ford circles never intersect. If 0 < p/q < 1 then the Ford circles that are tangent to C[p/q] are precisely the Ford circles for fractions that are neighbours of p/q in some Farey sequence.

Thus C[2/5] is tangent to C[1/2], C[1/3], C[3/7], C[3/8], etc.

Ford circles appear also in the Apollonian gasket (0,0,1,1). The picture below illustrates this together with Farey resonance lines. [17]

Apollonian gasket (0,0,1,1) and the Farey resonance diagram. Apolloinan gasket Farey.png
Apollonian gasket (0,0,1,1) and the Farey resonance diagram.

Riemann hypothesis

Farey sequences are used in two equivalent formulations of the Riemann hypothesis. Suppose the terms of are . Define , in other words is the difference between the kth term of the nth Farey sequence, and the kth member of a set of the same number of points, distributed evenly on the unit interval. In 1924 Jérôme Franel [18] proved that the statement

is equivalent to the Riemann hypothesis, and then Edmund Landau [19] remarked (just after Franel's paper) that the statement

is also equivalent to the Riemann hypothesis.

Other sums involving Farey fractions

The sum of all Farey fractions of order n is half the number of elements:

The sum of the denominators in the Farey sequence is twice the sum of the numerators and relates to Euler's totient function:

which was conjectured by Harold L. Aaron in 1962 and demonstrated by Jean A. Blake in 1966. [20] A one line proof of the Harold L. Aaron conjecture is as follows. The sum of the numerators is . The sum of denominators is . The quotient of the first sum by the second sum is .

Let bj be the ordered denominators of Fn, then: [21]

and

Let aj/bj the jth Farey fraction in Fn, then

which is demonstrated in. [22] Also according to this reference the term inside the sum can be expressed in many different ways:

obtaining thus many different sums over the Farey elements with same result. Using the symmetry around 1/2 the former sum can be limited to half of the sequence as

The Mertens function can be expressed as a sum over Farey fractions as

  where    is the Farey sequence of order n.

This formula is used in the proof of the Franel–Landau theorem. [23]

Next term

A surprisingly simple algorithm exists to generate the terms of Fn in either traditional order (ascending) or non-traditional order (descending). The algorithm computes each successive entry in terms of the previous two entries using the mediant property given above. If a/b and c/d are the two given entries, and p/q is the unknown next entry, then c/d = a + p/b + q. Since c/d is in lowest terms, there must be an integer k such that kc = a + p and kd = b + q, giving p = kc  a and q = kd  b. If we consider p and q to be functions of k, then

so the larger k gets, the closer p/q gets to c/d.

To give the next term in the sequence k must be as large as possible, subject to kd  b  n (as we are only considering numbers with denominators not greater than n), so k is the greatest integer  n + b/d. Putting this value of k back into the equations for p and q gives

This is implemented in Python as follows:

deffarey_sequence(n:int,descending:bool=False)->None:"""Print the n'th Farey sequence. Allow for either ascending or descending."""a,b,c,d=0,1,1,nifdescending:a,c=1,n-1print(f"{a}/{b}")while0<=c<=n:k=(n+b)//da,b,c,d=c,d,k*c-a,k*d-bprint(f"{a}/{b}")

Brute-force searches for solutions to Diophantine equations in rationals can often take advantage of the Farey series (to search only reduced forms). While this code uses the first two terms of the sequence to initialize a, b, c, and d, one could substitute any pair of adjacent terms in order to exclude those less than (or greater than) a particular threshold. [24]

See also

Footnotes

  1. The sequence of all reduced fractions with denominators not exceeding n, listed in order of their size, is called the Farey sequence of order n.” With the comment: “This definition of the Farey sequences seems to be the most convenient. However, some authors prefer to restrict the fractions to the interval from 0 to 1.” — Niven & Zuckerman (1972) [1]

Related Research Articles

In number theory, an arithmetic, arithmetical, or number-theoretic function is generally any function f(n) whose domain is the positive integers and whose range is a subset of the complex numbers. Hardy & Wright include in their definition the requirement that an arithmetical function "expresses some arithmetical property of n". There is a larger class of number-theoretic functions that do not fit this definition, for example, the prime-counting functions. This article provides links to functions of both classes.

<span class="mw-page-title-main">Euclidean algorithm</span> Algorithm for computing greatest common divisors

In mathematics, the Euclidean algorithm, or Euclid's algorithm, is an efficient method for computing the greatest common divisor (GCD) of two integers (numbers), the largest number that divides them both without a remainder. It is named after the ancient Greek mathematician Euclid, who first described it in his Elements . It is an example of an algorithm, a step-by-step procedure for performing a calculation according to well-defined rules, and is one of the oldest algorithms in common use. It can be used to reduce fractions to their simplest form, and is a part of many other number-theoretic and cryptographic calculations.

<span class="mw-page-title-main">Fibonacci sequence</span> Numbers obtained by adding the two previous ones

In mathematics, the Fibonacci sequence is a sequence in which each number is the sum of the two preceding ones. Numbers that are part of the Fibonacci sequence are known as Fibonacci numbers, commonly denoted Fn. The sequence commonly starts from 0 and 1, although some authors start the sequence from 1 and 1 or sometimes from 1 and 2. Starting from 0 and 1, the sequence begins

The Möbius function μ(n) is a multiplicative function in number theory introduced by the German mathematician August Ferdinand Möbius (also transliterated Moebius) in 1832. It is ubiquitous in elementary and analytic number theory and most often appears as part of its namesake the Möbius inversion formula. Following work of Gian-Carlo Rota in the 1960s, generalizations of the Möbius function were introduced into combinatorics, and are similarly denoted μ(x).

In mathematics, the classic Möbius inversion formula is a relation between pairs of arithmetic functions, each defined from the other by sums over divisors. It was introduced into number theory in 1832 by August Ferdinand Möbius.

Shor's algorithm is a quantum algorithm for finding the prime factors of an integer. It was developed in 1994 by the American mathematician Peter Shor. It is one of the few known quantum algorithms with compelling potential applications and strong evidence of superpolynomial speedup compared to best known classical algorithms. On the other hand, factoring numbers of practical significance requires far more qubits than available in the near future. Another concern is that noise in quantum circuits may undermine results, requiring additional qubits for quantum error correction.

<span class="mw-page-title-main">Euler's totient function</span> Number of integers coprime to and not exceeding n

In number theory, Euler's totient function counts the positive integers up to a given integer n that are relatively prime to n. It is written using the Greek letter phi as or , and may also be called Euler's phi function. In other words, it is the number of integers k in the range 1 ≤ kn for which the greatest common divisor gcd(n, k) is equal to 1. The integers k of this form are sometimes referred to as totatives of n.

In arithmetic and computer programming, the extended Euclidean algorithm is an extension to the Euclidean algorithm, and computes, in addition to the greatest common divisor (gcd) of integers a and b, also the coefficients of Bézout's identity, which are integers x and y such that

In mathematics, a generating function is a representation of an infinite sequence of numbers as the coefficients of a formal power series. Unlike an ordinary series, the formal power series is not required to converge: in fact, the generating function is not actually regarded as a function, and the "variable" remains an indeterminate. Generating functions were first introduced by Abraham de Moivre in 1730, in order to solve the general linear recurrence problem. One can generalize to formal power series in more than one indeterminate, to encode information about infinite multi-dimensional arrays of numbers.

<span class="mw-page-title-main">Practical number</span> Number such that it and all smaller numbers may be represented as sums of its distinct divisors

In number theory, a practical number or panarithmic number is a positive integer such that all smaller positive integers can be represented as sums of distinct divisors of . For example, 12 is a practical number because all the numbers from 1 to 11 can be expressed as sums of its divisors 1, 2, 3, 4, and 6: as well as these divisors themselves, we have 5 = 3 + 2, 7 = 6 + 1, 8 = 6 + 2, 9 = 6 + 3, 10 = 6 + 3 + 1, and 11 = 6 + 3 + 2.

In mathematics, the mediant of two fractions, generally made up of four positive integers

<span class="mw-page-title-main">Minkowski's question-mark function</span> Function with unusual fractal properties

In mathematics, Minkowski's question-mark function, denoted ?(x), is a function with unusual fractal properties, defined by Hermann Minkowski in 1904. It maps quadratic irrational numbers to rational numbers on the unit interval, via an expression relating the continued fraction expansions of the quadratics to the binary expansions of the rationals, given by Arnaud Denjoy in 1938. It also maps rational numbers to dyadic rationals, as can be seen by a recursive definition closely related to the Stern–Brocot tree.

In mathematics, the resultant of two polynomials is a polynomial expression of their coefficients that is equal to zero if and only if the polynomials have a common root, or, equivalently, a common factor. In some older texts, the resultant is also called the eliminant.

<span class="mw-page-title-main">Stern–Brocot tree</span> Ordered binary tree of rational numbers

In number theory, the Stern–Brocot tree is an infinite complete binary tree in which the vertices correspond one-for-one to the positive rational numbers, whose values are ordered from the left to the right as in a search tree.

A hyperelliptic curve is a particular kind of algebraic curve. There exist hyperelliptic curves of every genus . If the genus of a hyperelliptic curve equals 1, we simply call the curve an elliptic curve. Hence we can see hyperelliptic curves as generalizations of elliptic curves. There is a well-known group structure on the set of points lying on an elliptic curve over some field , which we can describe geometrically with chords and tangents. Generalizing this group structure to the hyperelliptic case is not straightforward. We cannot define the same group law on the set of points lying on a hyperelliptic curve, instead a group structure can be defined on the so-called Jacobian of a hyperelliptic curve. The computations differ depending on the number of points at infinity. Imaginary hyperelliptic curves are hyperelliptic curves with exactly 1 point at infinity: real hyperelliptic curves have two points at infinity.

In mathematics, a Redheffer matrix, often denoted as studied by Redheffer (1977), is a square (0,1) matrix whose entries aij are 1 if i divides j or if j = 1; otherwise, aij = 0. It is useful in some contexts to express Dirichlet convolution, or convolved divisors sums, in terms of matrix products involving the transpose of the Redheffer matrix.

In mathematics, infinite compositions of analytic functions (ICAF) offer alternative formulations of analytic continued fractions, series, products and other infinite expansions, and the theory evolving from such compositions may shed light on the convergence/divergence of these expansions. Some functions can actually be expanded directly as infinite compositions. In addition, it is possible to use ICAF to evaluate solutions of fixed point equations involving infinite expansions. Complex dynamics offers another venue for iteration of systems of functions rather than a single function. For infinite compositions of a single function see Iterated function. For compositions of a finite number of functions, useful in fractal theory, see Iterated function system.

In mathematics, a transformation of a sequence's generating function provides a method of converting the generating function for one sequence into a generating function enumerating another. These transformations typically involve integral formulas applied to a sequence generating function or weighted sums over the higher-order derivatives of these functions.

In number theory, the prime omega functions and count the number of prime factors of a natural number Thereby counts each distinct prime factor, whereas the related function counts the total number of prime factors of honoring their multiplicity. That is, if we have a prime factorization of of the form for distinct primes , then the respective prime omega functions are given by and . These prime factor counting functions have many important number theoretic relations.

The purpose of this page is to catalog new, interesting, and useful identities related to number-theoretic divisor sums, i.e., sums of an arithmetic function over the divisors of a natural number , or equivalently the Dirichlet convolution of an arithmetic function with one:

References

  1. Niven, Ivan M.; Zuckerman, Herbert S. (1972). An Introduction to the Theory of Numbers (Third ed.). John Wiley and Sons. Definition 6.1.
  2. Guthery, Scott B. (2011). "1. The Mediant". A Motif of Mathematics: History and Application of the Mediant and the Farey Sequence. Boston: Docent Press. p. 7. ISBN   978-1-4538-1057-6. OCLC   1031694495 . Retrieved 28 September 2020.
  3. Hardy, G.H.; Wright, E.M. (1979). An Introduction to the Theory of Numbers (Fifth ed.). Oxford University Press. Chapter III. ISBN   0-19-853171-0.
  4. 1 2 Beiler, Albert H. (1964). Recreations in the Theory of Numbers (Second ed.). Dover. Chapter XVI. ISBN   0-486-21096-0. Cited in "Farey Series, A Story". Cut-the-Knot.
  5. Sloane, N. J. A. (ed.). "SequenceA005728". The On-Line Encyclopedia of Integer Sequences . OEIS Foundation.
  6. 1 2 Tomas, Rogelio (January 2022). "Partial Franel sums" (PDF). Journal of Integer Sequences. 25 (1).
  7. Austin, David (December 2008). "Trees, Teeth, and Time: The mathematics of clock making". American Mathematical Society . Rhode Island. Archived from the original on 4 February 2020. Retrieved 28 September 2020.
  8. Martin, Greg (2009). "A product of Gamma function values at fractions with the same denominator". arXiv: 0907.4384 [math.CA].
  9. Wehmeier, Stefan (2009). "The LCM(1,2,...,n) as a product of sine values sampled over the points in Farey sequences". arXiv: 0909.1838 [math.CA].
  10. Tomas Garcia, Rogelio (August 2020). "Equalities between greatest common divisors involving three coprime pairs" (PDF). Notes on Number Theory and Discrete Mathematics. 26 (3): 5–7. doi: 10.7546/nntdm.2020.26.3.5-7 . S2CID   225280271.
  11. "Farey Approximation". NRICH.maths.org. Archived from the original on 19 November 2018. Retrieved 18 November 2018.
  12. Eliahou, Shalom (August 1993). "The 3x+1 problem: new lower bounds on nontrivial cycle lengths". Discrete Mathematics. 118 (1–3): 45–56. doi: 10.1016/0012-365X(93)90052-U .
  13. Zhenhua Li, A.; Harter, W.G. (2015). "Quantum Revivals of Morse Oscillators and Farey–Ford Geometry". Chem. Phys. Lett. 633: 208–213. arXiv: 1308.4470 . Bibcode:2015CPL...633..208L. doi:10.1016/j.cplett.2015.05.035. S2CID   66213897.
  14. Tomas, R. (2014). "From Farey sequences to resonance diagrams" (PDF). Physical Review Special Topics - Accelerators and Beams. 17 (1): 014001. Bibcode:2014PhRvS..17a4001T. doi: 10.1103/PhysRevSTAB.17.014001 .
  15. Harabor, Daniel Damir; Grastien, Alban; Öz, Dindar; Aksakalli, Vural (26 May 2016). "Optimal Any-Angle Pathfinding In Practice". Journal of Artificial Intelligence Research. 56: 89–118. doi: 10.1613/jair.5007 .
  16. Hew, Patrick Chisan (19 August 2017). "The Length of Shortest Vertex Paths in Binary Occupancy Grids Compared to Shortest r-Constrained Ones". Journal of Artificial Intelligence Research. 59: 543–563. doi: 10.1613/jair.5442 .
  17. Tomas, Rogelio (2020). "Imperfections and corrections". arXiv: 2006.10661 [physics.acc-ph].
  18. Franel, Jérôme (1924). "Les suites de Farey et le problème des nombres premiers". Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen. Mathematisch-Physikalische Klasse (in French): 198–201.
  19. Landau, Edmund (1924). "Bemerkungen zu der vorstehenden Abhandlung von Herrn Franel". Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen. Mathematisch-Physikalische Klasse (in German): 202–206.
  20. Blake, Jean A. (1966). "Some Characteristic Properties of the Farey Series". The American Mathematical Monthly. 73 (1): 50–52. doi:10.2307/2313922. JSTOR   2313922.
  21. Kurt Girstmair; Girstmair, Kurt (2010). "Farey Sums and Dedekind Sums". The American Mathematical Monthly. 117 (1): 72–78. doi:10.4169/000298910X475005. JSTOR   10.4169/000298910X475005. S2CID   31933470.
  22. Hall, R. R.; Shiu, P. (2003). "The Index of a Farey Sequence". Michigan Math. J. 51 (1): 209–223. doi: 10.1307/mmj/1049832901 .
  23. Edwards, Harold M. (1974). "12.2 Miscellany. The Riemann Hypothesis and Farey Series" . In Smith, Paul A.; Ellenberg, Samuel (eds.). Riemann's Zeta Function. Pure and Applied Mathematics. New York: Academic Press. pp. 263–267. ISBN   978-0-08-087373-2. OCLC   316553016 . Retrieved 30 September 2020.
  24. Routledge, Norman (March 2008). "Computing Farey series". The Mathematical Gazette . Vol. 92, no. 523. pp. 55–62.

Further reading