Unit fraction

Last updated

Slices of approximately 1/8 of a pizza Pizza-3007395.jpg
Slices of approximately 1/8 of a pizza

A unit fraction is a positive fraction with one as its numerator, 1/n. It is the multiplicative inverse (reciprocal) of the denominator of the fraction, which must be a positive natural number. Examples are 1/1, 1/2, 1/3, 1/4, 1/5, etc. When an object is divided into equal parts, each part is a unit fraction of the whole.

Contents

Multiplying two unit fractions produces another unit fraction, but other arithmetic operations do not preserve unit fractions. In modular arithmetic, unit fractions can be converted into equivalent whole numbers, allowing modular division to be transformed into multiplication. Every rational number can be represented as a sum of distinct unit fractions; these representations are called Egyptian fractions based on their use in ancient Egyptian mathematics. Many infinite sums of unit fractions are meaningful mathematically.

In geometry, unit fractions can be used to characterize the curvature of triangle groups and the tangencies of Ford circles. Unit fractions are commonly used in fair division, and this familiar application is used in mathematics education as an early step toward the understanding of other fractions. Unit fractions are common in probability theory due to the principle of indifference. They also have applications in combinatorial optimization and in analyzing the pattern of frequencies in the hydrogen spectral series.

Arithmetic

The unit fractions are the rational numbers that can be written in the form

where can be any positive natural number. They are thus the multiplicative inverses of the positive integers. When something is divided into equal parts, each part is a fraction of the whole. [1]

Elementary arithmetic

Multiplying any two unit fractions results in a product that is another unit fraction: [2]

However, adding, [3] subtracting, [3] or dividing two unit fractions produces a result that is generally not a unit fraction:

As the last of these formulas shows, every fraction can be expressed as a quotient of two unit fractions. [4]

Modular arithmetic

In modular arithmetic, any unit fraction can be converted into an equivalent whole number using the extended Euclidean algorithm. [5] [6] This conversion can be used to perform modular division: dividing by a number , modulo , can be performed by converting the unit fraction into an equivalent whole number modulo , and then multiplying by that number. [7]

In more detail, suppose that is relatively prime to (otherwise, division by is not defined modulo ). The extended Euclidean algorithm for the greatest common divisor can be used to find integers and such that Bézout's identity is satisfied:

In modulo- arithmetic, the term can be eliminated as it is zero modulo . This leaves

That is, is the modular inverse of , the number that when multiplied by produces one. Equivalently, [5] [6]

Thus division by (modulo ) can instead be performed by multiplying by the integer . [7]

Combinations

Several constructions in mathematics involve combining multiple unit fractions together, often by adding them.

Finite sums

Any positive rational number can be written as the sum of distinct unit fractions, in multiple ways. For example,

These sums are called Egyptian fractions, because the ancient Egyptian civilisations used them as notation for more general rational numbers. There is still interest today in analyzing the methods used by the ancients to choose among the possible representations for a fractional number, and to calculate with such representations. [8] The topic of Egyptian fractions has also seen interest in modern number theory; for instance, the Erdős–Graham conjecture [9] and the Erdős–Straus conjecture [10] concern sums of unit fractions, as does the definition of Ore's harmonic numbers. [11]

A pattern of spherical triangles with reflection symmetry across each triangle edge. Spherical reflection patterns like this with
2
x
{\displaystyle 2x}
,
2
y
{\displaystyle 2y}
, and
2
z
{\displaystyle 2z}
triangles at each vertex (here,
x
,
y
,
z
=
2
,
3
,
5
{\displaystyle x,y,z=2,3,5}
) only exist when
1
x
+
1
y
+
1
z
>
1
{\displaystyle {\tfrac {1}{x}}+{\tfrac {1}{y}}+{\tfrac {1}{z}}>1}
. Icosahedral reflection domains.png
A pattern of spherical triangles with reflection symmetry across each triangle edge. Spherical reflection patterns like this with , , and triangles at each vertex (here, ) only exist when .

In geometric group theory, triangle groups are classified into Euclidean, spherical, and hyperbolic cases according to whether an associated sum of unit fractions is equal to one, greater than one, or less than one respectively. [12]

Infinite series

Many well-known infinite series have terms that are unit fractions. These include:

Matrices

A Hilbert matrix is a square matrix in which the elements on the th antidiagonal all equal the unit fraction . That is, it has elements

For example, the matrix

is a Hilbert matrix. It has the unusual property that all elements in its inverse matrix are integers. [19] Similarly, Richardson (2001) defined a matrix whose elements are unit fractions whose denominators are Fibonacci numbers:

where denotes the th Fibonacci number. He calls this matrix the Filbert matrix and it has the same property of having an integer inverse. [20]

Adjacency and Ford circles

Fractions with tangent Ford circles differ by a unit fraction Ford circles colour.svg
Fractions with tangent Ford circles differ by a unit fraction

Two fractions and (in lowest terms) are called adjacent if

which implies that they differ from each other by a unit fraction:

For instance, and are adjacent: and . However, some pairs of fractions whose difference is a unit fraction are not adjacent in this sense: for instance, and differ by a unit fraction, but are not adjacent, because for them . [21]

This terminology comes from the study of Ford circles. These are a system of circles that are tangent to the number line at a given fraction and have the squared denominator of the fraction as their diameter. Fractions and are adjacent if and only if their Ford circles are tangent circles. [21]

Applications

Fair division and mathematics education

In mathematics education, unit fractions are often introduced earlier than other kinds of fractions, because of the ease of explaining them visually as equal parts of a whole. [22] [23] A common practical use of unit fractions is to divide food equally among a number of people, and exercises in performing this sort of fair division are a standard classroom example in teaching students to work with unit fractions. [24]

Probability and statistics

A six-sided die has probability 1/6 of landing on each side Dice 2005.jpg
A six-sided die has probability 1/6 of landing on each side

In a uniform distribution on a discrete space, all probabilities are equal unit fractions. Due to the principle of indifference, probabilities of this form arise frequently in statistical calculations. [25]

Unequal probabilities related to unit fractions arise in Zipf's law. This states that, for many observed phenomena involving the selection of items from an ordered sequence, the probability that the th item is selected is proportional to the unit fraction . [26]

Combinatorial optimization

In the study of combinatorial optimization problems, bin packing problems involve an input sequence of items with fractional sizes, which must be placed into bins whose capacity (the total size of items placed into each bin) is one. Research into these problems has included the study of restricted bin packing problems where the item sizes are unit fractions. [27] [28]

One motivation for this is as a test case for more general bin packing methods. Another involves a form of pinwheel scheduling, in which a collection of messages of equal length must each be repeatedly broadcast on a limited number of communication channels, with each message having a maximum delay between the start times of its repeated broadcasts. An item whose delay is times the length of a message must occupy a fraction of at least of the time slots on the channel it is assigned to, so a solution to the scheduling problem can only come from a solution to the unit fraction bin packing problem with the channels as bins and the fractions as item sizes. [27]

Even for bin packing problems with arbitrary item sizes, it can be helpful to round each item size up to the next larger unit fraction, and then apply a bin packing algorithm specialized for unit fraction sizes. In particular, the harmonic bin packing method does exactly this, and then packs each bin using items of only a single rounded unit fraction size. [28]

Physics

The hydrogen spectral series, on a logarithmic scale. The frequencies of the emission lines are proportional to differences of pairs of unit fractions. Hydrogen spectrum.svg
The hydrogen spectral series, on a logarithmic scale. The frequencies of the emission lines are proportional to differences of pairs of unit fractions.

The energy levels of photons that can be absorbed or emitted by a hydrogen atom are, according to the Rydberg formula, proportional to the differences of two unit fractions. An explanation for this phenomenon is provided by the Bohr model, according to which the energy levels of electron orbitals in a hydrogen atom are inversely proportional to square unit fractions, and the energy of a photon is quantized to the difference between two levels. [29]

Arthur Eddington argued that the fine-structure constant was a unit fraction. He initially thought it to be 1/136 and later changed his theory to 1/137. This contention has been falsified, given that current estimates of the fine structure constant are (to 6 significant digits) 1/137.036. [30]

See also

Related Research Articles

<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

<span class="mw-page-title-main">Pythagorean triple</span> Integer side lengths of a right triangle

A Pythagorean triple consists of three positive integers a, b, and c, such that a2 + b2 = c2. Such a triple is commonly written (a, b, c), and a well-known example is (3, 4, 5). If (a, b, c) is a Pythagorean triple, then so is (ka, kb, kc) for any positive integer k. A primitive Pythagorean triple is one in which a, b and c are coprime (that is, they have no common divisor larger than 1). For example, (3, 4, 5) is a primitive Pythagorean triple whereas (6, 8, 10) is not. A triangle whose sides form a Pythagorean triple is called a Pythagorean triangle and is a right triangle.

<span class="mw-page-title-main">Taylor series</span> Mathematical approximation of a function

In mathematics, the Taylor series or Taylor expansion of a function is an infinite sum of terms that are expressed in terms of the function's derivatives at a single point. For most common functions, the function and the sum of its Taylor series are equal near this point. Taylor series are named after Brook Taylor, who introduced them in 1715. A Taylor series is also called a Maclaurin series when 0 is the point where the derivatives are considered, after Colin Maclaurin, who made extensive use of this special case of Taylor series in the mid-18th century.

In mathematics, Pascal's triangle is a triangular array of the binomial coefficients arising in probability theory, combinatorics, and algebra. In much of the Western world, it is named after the French mathematician Blaise Pascal, although other mathematicians studied it centuries before him in Persia, India, China, Germany, and Italy.

<i>p</i>-adic number Number system extending the rational numbers

In number theory, given a prime number p, the p-adic numbers form an extension of the rational numbers which is distinct from the real numbers, though with some similar properties; p-adic numbers can be written in a form similar to decimals, but with digits based on a prime number p rather than ten, and extending to the left rather than to the right. Formally, given a prime number p, a p-adic number can be defined as a series

<span class="mw-page-title-main">Imaginary unit</span> Principal square root of −1

The imaginary unit or unit imaginary number is a solution to the quadratic equation x2 + 1 = 0. Although there is no real number with this property, i can be used to extend the real numbers to what are called complex numbers, using addition and multiplication. A simple example of the use of i in a complex number is 2 + 3i.

<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 mathematics, the harmonic series is the infinite series formed by summing all positive unit fractions:

<span class="mw-page-title-main">Multiplicative inverse</span> Number which when multiplied by x equals 1

In mathematics, a multiplicative inverse or reciprocal for a number x, denoted by 1/x or x−1, is a number which when multiplied by x yields the multiplicative identity, 1. The multiplicative inverse of a fraction a/b is b/a. For the multiplicative inverse of a real number, divide 1 by the number. For example, the reciprocal of 5 is one fifth (1/5 or 0.2), and the reciprocal of 0.25 is 1 divided by 0.25, or 4. The reciprocal function, the function f(x) that maps x to 1/x, is one of the simplest examples of a function which is its own inverse (an involution).

<span class="mw-page-title-main">Egyptian fraction</span> Finite sum of distinct unit fractions

An Egyptian fraction is a finite sum of distinct unit fractions, such as

One half is the irreducible fraction resulting from dividing one (1) by two (2), or the fraction resulting from dividing any number by its double.

<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.

<span class="mw-page-title-main">Fraction</span> Ratio of two numbers

A fraction represents a part of a whole or, more generally, any number of equal parts. When spoken in everyday English, a fraction describes how many parts of a certain size there are, for example, one-half, eight-fifths, three-quarters. A common, vulgar, or simple fraction consists of an integer numerator, displayed above a line, and a non-zero integer denominator, displayed below that line. If these integers are positive, then the numerator represents a number of equal parts, and the denominator indicates how many of those parts make up a unit or a whole. For example, in the fraction 3/4, the numerator 3 indicates that the fraction represents 3 equal parts, and the denominator 4 indicates that 4 parts make up a whole. The picture to the right illustrates 3/4 of a cake.

<span class="mw-page-title-main">Lemniscate constant</span> Ratio of the perimeter of Bernoullis lemniscate to its diameter

In mathematics, the lemniscate constantϖ is a transcendental mathematical constant that is the ratio of the perimeter of Bernoulli's lemniscate to its diameter, analogous to the definition of π for the circle. Equivalently, the perimeter of the lemniscate is 2ϖ. The lemniscate constant is closely related to the lemniscate elliptic functions and approximately equal to 2.62205755. The symbol ϖ is a cursive variant of π; see Pi § Variant pi.

The Erdős–Straus conjecture is an unproven statement in number theory. The conjecture is that, for every integer that is 2 or more, there exist positive integers , , and for which

<span class="mw-page-title-main">Sylvester's sequence</span> Doubly exponential integer sequence

In number theory, Sylvester's sequence is an integer sequence in which each term is the product of the previous terms, plus one. The first few terms of the sequence are

In mathematics, particularly in the area of arithmetic, a modular multiplicative inverse of an integer a is an integer x such that the product ax is congruent to 1 with respect to the modulus m. In the standard notation of modular arithmetic this congruence is written as

<span class="mw-page-title-main">Rational number</span> Quotient of two integers

In mathematics, a rational number is a number that can be expressed as the quotient or fraction of two integers, a numerator p and a non-zero denominator q. For example, is a rational number, as is every integer. The set of all rational numbers, also referred to as "the rationals", the field of rationals or the field of rational numbers is usually denoted by boldface Q, or blackboard bold

In the bin covering problem, items of different sizes must be packed into a finite number of bins or containers, each of which must contain at least a certain given total size, in a way that maximizes the number of bins used.

The configuration linear program (configuration-LP) is a linear programming technique used for solving combinatorial optimization problems. It was introduced in the context of the cutting stock problem. Later, it has been applied to the bin packing and job scheduling problems. In the configuration-LP, there is a variable for each possible configuration - each possible multiset of items that can fit in a single bin. Usually, the number of configurations is exponential in the problem size, but in some cases it is possible to attain approximate solutions using only a polynomial number of configurations.

References

  1. Cavey, Laurie O.; Kinzel, Margaret T. (February 2014), "From whole numbers to invert and multiply", Teaching Children Mathematics, 20 (6): 374–383, doi:10.5951/teacchilmath.20.6.0374, JSTOR   10.5951/teacchilmath.20.6.0374
  2. Solomon, Pearl Gold (2007), The Math We Need to Know and Do in Grades 6 9: Concepts, Skills, Standards, and Assessments, Corwin Press, p. 157, ISBN   978-1-4129-1726-1
  3. 1 2 Betz, William (1957), Algebra for Today, First Year, Ginn, p. 370
  4. Humenberger, Hans (Fall 2014), "Egyptian fractions – representations as sums of unit fractions", Mathematics and Computer Education , 48 (3): 268–283, ProQuest   1622317875
  5. 1 2 Cormen, Thomas H.; Leiserson, Charles E.; Rivest, Ronald L.; Stein, Clifford (2001) [1990], "31.4 Solving modular linear equations", Introduction to Algorithms (2nd ed.), MIT Press and McGraw-Hill, pp. 869–872, ISBN   0-262-03293-7
  6. 1 2 Goodrich, Michael T.; Tamassia, Roberto (2015), "Section 24.2.2: Modular multiplicative inverses", Algorithm Design and Applications, Wiley, pp. 697–698, ISBN   978-1-118-33591-8
  7. 1 2 Brent, Richard P.; Zimmermann, Paul (2010), "2.5 Modular division and inversion", Modern Computer Arithmetic (PDF), Cambridge Monographs on Applied and Computational Mathematics, vol. 18, Cambridge University Press, pp. 65–68, arXiv: 1004.4710 , doi:10.1017/cbo9780511921698.001, ISBN   978-1-139-49228-7, S2CID   441260
  8. Guy, Richard K. (2004), "D11. Egyptian Fractions", Unsolved problems in number theory (3rd ed.), Springer-Verlag, pp. 252–262, ISBN   978-0-387-20860-2
  9. Croot, Ernest S. III (2003), "On a coloring conjecture about unit fractions", Annals of Mathematics , 157 (2): 545–556, arXiv: math.NT/0311421 , doi:10.4007/annals.2003.157.545, MR   1973054, S2CID   13514070
  10. Elsholtz, Christian; Tao, Terence (2013), "Counting the number of solutions to the Erdős–Straus equation on unit fractions" (PDF), Journal of the Australian Mathematical Society, 94 (1): 50–105, arXiv: 1107.1010 , doi:10.1017/S1446788712000468, MR   3101397, S2CID   17233943
  11. Ore, Øystein (1948), "On the averages of the divisors of a number", The American Mathematical Monthly , 55 (10): 615–619, doi:10.2307/2305616, JSTOR   2305616
  12. Magnus, Wilhelm (1974), Noneuclidean Tesselations and their Groups, Pure and Applied Mathematics, vol. 61, Academic Press, p. 65, ISBN   978-0-08-087377-0, MR   0352287
  13. Boas, R. P. Jr.; Wrench, J. W. Jr. (1971), "Partial sums of the harmonic series", The American Mathematical Monthly , 78 (8): 864–870, doi:10.1080/00029890.1971.11992881, JSTOR   2316476, MR   0289994
  14. Freniche, Francisco J. (2010), "On Riemann's rearrangement theorem for the alternating harmonic series" (PDF), The American Mathematical Monthly , 117 (5): 442–448, doi:10.4169/000298910X485969, JSTOR   10.4169/000298910x485969, MR   2663251, S2CID   20575373
  15. Roy, Ranjan (1990), "The discovery of the series formula for π by Leibniz, Gregory and Nilakantha" (PDF), Mathematics Magazine , 63 (5): 291–306, doi:10.1080/0025570X.1990.11977541
  16. Ayoub, Raymond (1974), "Euler and the zeta function", The American Mathematical Monthly , 81 (10): 1067–86, doi:10.2307/2319041, JSTOR   2319041
  17. van der Poorten, Alfred (1979), "A proof that Euler missed ... Apéry's proof of the irrationality of " (PDF), The Mathematical Intelligencer , 1 (4): 195–203, doi:10.1007/BF03028234, S2CID   121589323, archived from the original (PDF) on 2011-07-06
  18. Euler, Leonhard (September 1983), "From Elements of Algebra", Old Intelligencer, The Mathematical Intelligencer , 5 (3): 75–76, doi:10.1007/bf03026580, S2CID   122191726
  19. Choi, Man Duen (1983), "Tricks or treats with the Hilbert matrix", The American Mathematical Monthly , 90 (5): 301–312, doi:10.2307/2975779, JSTOR   2975779, MR   0701570
  20. Richardson, Thomas M. (2001), "The Filbert matrix" (PDF), Fibonacci Quarterly , 39 (3): 268–275, arXiv: math.RA/9905079 , Bibcode:1999math......5079R
  21. 1 2 Ford, L. R. (1938), "Fractions", The American Mathematical Monthly , 45 (9): 586–601, doi:10.1080/00029890.1938.11990863, JSTOR   2302799, MR   1524411
  22. Polkinghorne, Ada R. (May 1935), "Young-children and fractions", Childhood Education, 11 (8): 354–358, doi:10.1080/00094056.1935.10725374
  23. Empson, Susan Baker; Jacobs, Victoria R.; Jessup, Naomi A.; Hewitt, Amy; Pynes, D'Anna; Krause, Gladys (April 2020), "Unit fractions as superheroes for instruction", The Mathematics Teacher, 113 (4): 278–286, doi:10.5951/mtlt.2018.0024, JSTOR   10.5951/mtlt.2018.0024, S2CID   216283105
  24. Wilson, P. Holt; Edgington, Cynthia P.; Nguyen, Kenny H.; Pescosolido, Ryan C.; Confrey, Jere (November 2011), "Fractions: how to fair share", Mathematics Teaching in the Middle School, 17 (4): 230–236, doi:10.5951/mathteacmiddscho.17.4.0230, JSTOR   10.5951/mathteacmiddscho.17.4.0230
  25. Welsh, Alan H. (1996), Aspects of Statistical Inference, Wiley Series in Probability and Statistics, vol. 246, John Wiley and Sons, p. 66, ISBN   978-0-471-11591-5
  26. Saichev, Alexander; Malevergne, Yannick; Sornette, Didier (2009), Theory of Zipf's Law and Beyond, Lecture Notes in Economics and Mathematical Systems, vol. 632, Springer-Verlag, ISBN   978-3-642-02945-5
  27. 1 2 Bar-Noy, Amotz; Ladner, Richard E.; Tamir, Tami (2007), "Windows scheduling as a restricted version of bin packing", ACM Transactions on Algorithms, 3 (3): A28:1–A28:22, doi:10.1145/1273340.1273344, MR   2344019, S2CID   2461059
  28. 1 2 van Stee, Rob (June 2012), "SIGACT news online algorithms column 20: The power of harmony" (PDF), ACM SIGACT News, 43 (2): 127–136, doi:10.1145/2261417.2261440, S2CID   14805804
  29. Yang, Fujia; Hamilton, Joseph H. (2009), Modern Atomic and Nuclear Physics, World Scientific, pp. 81–86, ISBN   978-981-283-678-6
  30. Kilmister, Clive William (1994), Eddington's Search for a Fundamental Theory: A Key to the Universe, Cambridge University Press, ISBN   978-0-521-37165-0