Pascal's triangle

Last updated
A diagram showing the first eight rows of Pascal's triangle.

In mathematics, Pascal's triangle is a triangular array of the binomial coefficients which play a crucial role 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, [1] India, [2] China, Germany, and Italy. [3]

Contents

The kth entry in the nth row of Pascal's triangle is the binomial coefficient "n choose k", written . The rows are enumerated from at the top (the 0th row), and the entries in each row are numbered from on the left to on the right, and are usually staggered left relative to the numbers in the previous row. The triangle may be constructed as follows: the top entry is equal to 1, and each lower entry is the sum of the two entries above it to the left and right, treating blank entries as 0. For example, the initial number of row 1 (or any other row) is 1 (the sum of 0 and 1), whereas the first two numbers 1 and 3 in row 3 are added to produce the number 4 below them in row 4.

Formula

In Pascal's triangle, each number is the sum of the two numbers directly above it. PascalTriangleAnimated2.gif
In Pascal's triangle, each number is the sum of the two numbers directly above it.

In the th row of Pascal's triangle, the th entry is denoted , pronounced "n choose k". For example, the topmost element is . With this notation, the downward addition construction of the previous paragraph may be written as:

,

for any non-negative integer and any integer . [4] This recurrence for the binomial coefficients is known as Pascal's rule.

History

Yang Hui's triangle, as depicted by the Chinese using rod numerals, appears in Jade Mirror of the Four Unknowns, a mathematical work by Zhu Shijie, dated 1303. Yanghui triangle.gif
Yang Hui's triangle, as depicted by the Chinese using rod numerals, appears in Jade Mirror of the Four Unknowns, a mathematical work by Zhu Shijie, dated 1303.
Pascal's version of the triangle TrianguloPascal.jpg
Pascal's version of the triangle

The pattern of numbers that forms Pascal's triangle was known well before Pascal's time. The Persian mathematician Al-Karaji (953–1029) wrote a now-lost book which contained the first formulation of the binomial coefficients and the first description of Pascal's triangle. [5] [6] [7] It was later repeated by Omar Khayyám (1048–1131), another Persian mathematician; thus the triangle is also referred to as Khayyam's triangle (مثلث خیام) in Iran. [8] Several theorems related to the triangle were known, including the binomial theorem. Khayyam used a method of finding nth roots based on the binomial expansion, and therefore on the binomial coefficients. [1]

Pascal's triangle was known in China during the early 11th century through the work of the Chinese mathematician Jia Xian (1010–1070). During the 13th century, Yang Hui (1238–1298) defined the triangle, and it is known as Yang Hui's triangle (杨辉三角; 楊輝三角) in China. [9]

In Europe, Pascal's triangle appeared for the first time in the Arithmetic of Jordanus de Nemore (13th century). [10] The binomial coefficients were calculated by Gersonides during the early 14th century, using the multiplicative formula for them. [11] Petrus Apianus (1495–1552) published the full triangle on the frontispiece of his book on business calculations in 1527. [12] Michael Stifel published a portion of the triangle (from the second to the middle column in each row) in 1544, describing it as a table of figurate numbers. [11] In Italy, Pascal's triangle is referred to as Tartaglia's triangle, named for the Italian algebraist Niccolò Fontana Tartaglia (1500–1577), who published six rows of the triangle in 1556. [11] Gerolamo Cardano also published the triangle as well as the additive and multiplicative rules for constructing it in 1570. [11]

Pascal's Traité du triangle arithmétique (Treatise on Arithmetical Triangle) was published posthumously in 1665. [13] In this, Pascal collected several results then known about the triangle, and employed them to solve problems in probability theory. The triangle was later named for Pascal by Pierre Raymond de Montmort (1708) who called it table de M. Pascal pour les combinaisons (French: Mr. Pascal's table for combinations) and Abraham de Moivre (1730) who called it Triangulum Arithmeticum PASCALIANUM (Latin: Pascal's Arithmetic Triangle), which became the basis of the modern Western name. [14]

Binomial expansions

Visualisation of binomial expansion up to the 4th power Binomial theorem visualisation.svg
Visualisation of binomial expansion up to the 4th power

Pascal's triangle determines the coefficients which arise in binomial expansions. For example, in the expansion:

,

the coefficients are the entries in the second row of Pascal's triangle: , , .

In general, the binomial theorem states that when a binomial like is raised to a positive integer power , the expression expands into:

,

where the coefficients are precisely the numbers in row of Pascal's triangle:

.

The entire left diagonal of Pascal's triangle corresponds to the coefficient of in these binomial expansions, while the next left diagonal corresponds to the coefficient of , and so on.

To see how the binomial theorem relates to the simple construction of Pascal's triangle, consider the problem of calculating the coefficients of the expansion of in terms of the corresponding coefficients of , where we sett for simplicity. Suppose then that

.

Now

Six rows Pascal's triangle as binomial coefficients

The two summations can be reindexed with and combined:

Thus the extreme left and right coefficients remain as 1, and for any given , the coefficient of the term in the polynomial is equal to , the sum of the and coefficients in the previous power . This is indeed the downward-addition rule for constructing Pascal's triangle.

It is not difficult to turn this argument into a proof (by mathematical induction) of the binomial theorem.

Since , the coefficients are identical in the expansion of the general case.

An interesting consequence of the binomial theorem is obtained by setting both variables , so that:

In other words, the sum of the entries in the th row of Pascal's triangle is the th power of 2. This is equivalent to the statement that the number of subsets of an -element set is , as can be seen by observing that each of the elements may be independently included or excluded from a given subset.

Combinations

A second useful application of Pascal's triangle is in the calculation of combinations. The number of combinations of items taken at a time, i.e. the number of subsets of elements from among elements, can be found by the equation

.

This is equal to entry in row of Pascal's triangle. Rather than performing the multiplicative calculation, one can simply look up the appropriate entry in the triangle (constructed by additions). For example, suppose 3 workers need to be hired from among 7 candidates; then the number of possible hiring choices is 7 choose 3, the entry 3 in row 7 of the above table, which is . [15]

Relation to binomial distribution and convolutions

When divided by , the th row of Pascal's triangle becomes the binomial distribution in the symmetric case where . By the central limit theorem, this distribution approaches the normal distribution as increases. This can also be seen by applying Stirling's formula to the factorials involved in the formula for combinations.

This is related to the operation of discrete convolution in two ways. First, polynomial multiplication corresponds exactly to discrete convolution, so that repeatedly convolving the sequence with itself corresponds to taking powers of , and hence to generating the rows of the triangle. Second, repeatedly convolving the distribution function for a random variable with itself corresponds to calculating the distribution function for a sum of n independent copies of that variable; this is exactly the situation to which the central limit theorem applies, and hence results in the normal distribution in the limit. (The operation of repeatedly taking a convolution of something with itself is called the convolution power.)

Patterns and properties

Pascal's triangle has many properties and contains many patterns of numbers.

Each frame represents a row in Pascal's triangle. Each column of pixels is a number in binary with the least significant bit at the bottom. Light pixels represent 1 and dark pixels 0. Pascal's Triangle animated binary rows.gif
Each frame represents a row in Pascal's triangle. Each column of pixels is a number in binary with the least significant bit at the bottom. Light pixels represent 1 and dark pixels 0.
The numbers of compositions of n +1 into k +1 ordered partitions form Pascal's triangle. Pascal triangle compositions.svg
The numbers of compositions of n+1 into k+1 ordered partitions form Pascal's triangle.

Rows

Diagonals

Derivation of simplex numbers from a left-justified Pascal's triangle Pascal triangle simplex numbers.svg
Derivation of simplex numbers from a left-justified Pascal's triangle

The diagonals of Pascal's triangle contain the figurate numbers of simplices:

The symmetry of the triangle implies that the nth d-dimensional number is equal to the dthn-dimensional number.

An alternative formula that does not involve recursion is

where n(d) is the rising factorial.

The geometric meaning of a function Pd is: Pd(1) = 1 for all d. Construct a d-dimensional triangle (a 3-dimensional triangle is a tetrahedron) by placing additional dots below an initial dot, corresponding to Pd(1) = 1. Place these dots in a manner analogous to the placement of numbers in Pascal's triangle. To find Pd(x), have a total of x dots composing the target shape. Pd(x) then equals the total number of dots in the shape. A 0-dimensional triangle is a point and a 1-dimensional triangle is simply a line, and therefore P0(x) = 1 and P1(x) = x, which is the sequence of natural numbers. The number of dots in each layer corresponds to Pd  1(x).

Calculating a row or diagonal by itself

There are simple algorithms to compute all the elements in a row or diagonal without computing other elements or factorials.

To compute row with the elements , begin with . For each subsequent element, the value is determined by multiplying the previous value by a fraction with slowly changing numerator and denominator:

For example, to calculate row 5, the fractions are , , ,  and , and hence the elements are ,  ,  , etc. (The remaining elements are most easily obtained by symmetry.)

To compute the diagonal containing the elements begin again with and obtain subsequent elements by multiplication by certain fractions:

For example, to calculate the diagonal beginning at , the fractions are , and the elements are , etc. By symmetry, these elements are equal to , etc.

Fibonacci sequence in Pascal's triangle Fibonacci in Pascal triangle.png
Fibonacci sequence in Pascal's triangle

Overall patterns and properties

A level-4 approximation to a Sierpinski triangle obtained by shading the first 32 rows of a Pascal triangle white if the binomial coefficient is even and black if it is odd. Sierpinski Pascal triangle.svg
A level-4 approximation to a Sierpinski triangle obtained by shading the first 32 rows of a Pascal triangle white if the binomial coefficient is even and black if it is odd.
As the proportion of black numbers tends to zero with increasing n, a corollary is that the proportion of odd binomial coefficients tends to zero as n tends to infinity. [21]
Chess rll45.svg Chess x1d45.svg Chess x1l45.svg Chess x1d45.svg
Chess x1d45.svg Chess x2l45.svg Chess x3d45.svg Chess x4l45.svg
Chess x1l45.svg Chess x3d45.svg Chess x6l45.svg 10
Chess x1d45.svg Chess x4l45.svg 1020

Pascal's triangle overlaid on a grid gives the number of distinct paths to each square, assuming only rightward and downward movements are considered.

Pascal's Triangle 4 paths.svg
1
11
121
1331
14641
15101051
1615201561
172135352171

Construction as matrix exponential

Binomial matrix as matrix exponential. All the dots represent 0.

Due to its simple construction by factorials, a very basic representation of Pascal's triangle in terms of the matrix exponential can be given: Pascal's triangle is the exponential of the matrix which has the sequence 1, 2, 3, 4, ... on its subdiagonal and zero everywhere else.

Construction of Clifford algebra using simplices

Labelling the elements of each n-simplex matches the basis elements of Clifford algebra used as forms in Geometric Algebra rather than matrices. Recognising the geometric operations, such as rotations, allows the algebra operations to be discovered. Just as each row, n, starting at 0, of Pascal's triangle corresponds to an (n-1)-simplex, as described below, it also defines the number of named basis forms in n dimensional Geometric algebra. The binomial theorem can be used to prove the geometric relationship provided by Pascal's triangle. [22] This same proof could be applied to simplices except that the first column of all 1's must be ignored whereas in the algebra these correspond to the real numbers, , with basis 1.

Relation to geometry of polytopes

Pascal's triangle can be used as a lookup table for the number of elements (such as edges and corners) within a polytope (such as a triangle, a tetrahedron, a square, or a cube).

Number of elements of simplices

Let's begin by considering the 3rd line of Pascal's triangle, with values 1, 3, 3, 1. A 2-dimensional triangle has one 2-dimensional element (itself), three 1-dimensional elements (lines, or edges), and three 0-dimensional elements (vertices, or corners). The meaning of the final number (1) is more difficult to explain (but see below). Continuing with our example, a tetrahedron has one 3-dimensional element (itself), four 2-dimensional elements (faces), six 1-dimensional elements (edges), and four 0-dimensional elements (vertices). Adding the final 1 again, these values correspond to the 4th row of the triangle (1, 4, 6, 4, 1). Line 1 corresponds to a point, and Line 2 corresponds to a line segment (dyad). This pattern continues to arbitrarily high-dimensioned hyper-tetrahedrons (known as simplices).

To understand why this pattern exists, one must first understand that the process of building an n-simplex from an (n − 1)-simplex consists of simply adding a new vertex to the latter, positioned such that this new vertex lies outside of the space of the original simplex, and connecting it to all original vertices. As an example, consider the case of building a tetrahedron from a triangle, the latter of whose elements are enumerated by row 3 of Pascal's triangle: 1 face, 3 edges, and 3 vertices. To build a tetrahedron from a triangle, position a new vertex above the plane of the triangle and connect this vertex to all three vertices of the original triangle.

The number of a given dimensional element in the tetrahedron is now the sum of two numbers: first the number of that element found in the original triangle, plus the number of new elements, each of which is built upon elements of one fewer dimension from the original triangle. Thus, in the tetrahedron, the number of cells (polyhedral elements) is 0 + 1 = 1; the number of faces is 1 + 3 = 4; the number of edges is 3 + 3 = 6; the number of new vertices is 3 + 1 = 4. This process of summing the number of elements of a given dimension to those of one fewer dimension to arrive at the number of the former found in the next higher simplex is equivalent to the process of summing two adjacent numbers in a row of Pascal's triangle to yield the number below. Thus, the meaning of the final number (1) in a row of Pascal's triangle becomes understood as representing the new vertex that is to be added to the simplex represented by that row to yield the next higher simplex represented by the next row. This new vertex is joined to every element in the original simplex to yield a new element of one higher dimension in the new simplex, and this is the origin of the pattern found to be identical to that seen in Pascal's triangle.

Number of elements of hypercubes

A similar pattern is observed relating to squares, as opposed to triangles. To find the pattern, one must construct an analog to Pascal's triangle, whose entries are the coefficients of (x + 2)row number, instead of (x + 1)row number. There are a couple ways to do this. The simpler is to begin with row 0 = 1 and row 1 = 1, 2. Proceed to construct the analog triangles according to the following rule:

That is, choose a pair of numbers according to the rules of Pascal's triangle, but double the one on the left before adding. This results in:

The other way of producing this triangle is to start with Pascal's triangle and multiply each entry by 2k, where k is the position in the row of the given number. For example, the 2nd value in row 4 of Pascal's triangle is 6 (the slope of 1s corresponds to the zeroth entry in each row). To get the value that resides in the corresponding position in the analog triangle, multiply 6 by 2position number = 6 × 22 = 6 × 4 = 24. Now that the analog triangle has been constructed, the number of elements of any dimension that compose an arbitrarily dimensioned cube (called a hypercube) can be read from the table in a way analogous to Pascal's triangle. For example, the number of 2-dimensional elements in a 2-dimensional cube (a square) is one, the number of 1-dimensional elements (sides, or lines) is 4, and the number of 0-dimensional elements (points, or vertices) is 4. This matches the 2nd row of the table (1, 4, 4). A cube has 1 cube, 6 faces, 12 edges, and 8 vertices, which corresponds to the next line of the analog triangle (1, 6, 12, 8). This pattern continues indefinitely.

To understand why this pattern exists, first recognize that the construction of an n-cube from an (n − 1)-cube is done by simply duplicating the original figure and displacing it some distance (for a regular n-cube, the edge length) orthogonal to the space of the original figure, then connecting each vertex of the new figure to its corresponding vertex of the original. This initial duplication process is the reason why, to enumerate the dimensional elements of an n-cube, one must double the first of a pair of numbers in a row of this analog of Pascal's triangle before summing to yield the number below. The initial doubling thus yields the number of "original" elements to be found in the next higher n-cube and, as before, new elements are built upon those of one fewer dimension (edges upon vertices, faces upon edges, etc.). Again, the last number of a row represents the number of new vertices to be added to generate the next higher n-cube.

In this triangle, the sum of the elements of row m is equal to 3m. Again, to use the elements of row 4 as an example: 1 + 8 + 24 + 32 + 16 = 81, which is equal to .

Counting vertices in a cube by distance

Each row of Pascal's triangle gives the number of vertices at each distance from a fixed vertex in an n-dimensional cube. For example, in three dimensions, the third row (1 3 3 1) corresponds to the usual three-dimensional cube: fixing a vertex V, there is one vertex at distance 0 from V (that is, V itself), three vertices at distance 1, three vertices at distance 2 and one vertex at distance 3 (the vertex opposite V). The second row corresponds to a square, while larger-numbered rows correspond to hypercubes in each dimension.

Fourier transform of sin(x)n+1/x

As stated previously, the coefficients of (x + 1)n are the nth row of the triangle. Now the coefficients of (x  1)n are the same, except that the sign alternates from +1 to −1 and back again. After suitable normalization, the same pattern of numbers occurs in the Fourier transform of sin(x)n+1/x. More precisely: if n is even, take the real part of the transform, and if n is odd, take the imaginary part. Then the result is a step function, whose values (suitably normalized) are given by the nth row of the triangle with alternating signs. [23] For example, the values of the step function that results from:

compose the 4th row of the triangle, with alternating signs. This is a generalization of the following basic result (often used in electrical engineering):

is the boxcar function. [24] The corresponding row of the triangle is row 0, which consists of just the number 1.

If n is congruent to 2 or to 3 mod 4, then the signs start with −1. In fact, the sequence of the (normalized) first terms corresponds to the powers of i, which cycle around the intersection of the axes with the unit circle in the complex plane:

Extensions

Pascal's triangle may be extended upwards, above the 1 at the apex, preserving the additive property, but there is more than one way to do so. [25]

To higher dimensions

Pascal's triangle has higher dimensional generalizations. The three-dimensional version is known as Pascal's pyramid or Pascal's tetrahedron, while the general versions are known as Pascal's simplices .

To complex numbers

When the factorial function is defined as , Pascal's triangle can be extended beyond the integers to , since is meromorphic to the entire complex plane. [26]

To arbitrary bases

Isaac Newton once observed that the first five rows of Pascal's Triangle, considered as strings, are the corresponding powers of eleven. He claimed without proof that subsequent rows also generate powers of eleven. [27] In 1964, Dr. Robert L. Morton presented the more generalized argument that each row can be read as a radix numeral, where is the hypothetical terminal row, or limit, of the triangle, and the rows are its partial products. [28] He proved the entries of row , when interpreted directly as a place-value numeral, correspond to the binomial expansion of . More rigorous proofs have since been developed. [29] [30] To better understand the principle behind this interpretation, here are some things to recall about binomials:

By setting the row's radix (the variable ) equal to one and ten, row becomes the product and , respectively. To illustrate, consider , which yields the row product . The numeric representation of is formed by concatenating the entries of row . The twelfth row denotes the product:

with compound digits (delimited by ":") in radix twelve. The digits from through are compound because these row entries compute to values greater than or equal to twelve. To normalize [36] the numeral, simply carry the first compound entry's prefix, that is, remove the prefix of the coefficient from its leftmost digit up to, but excluding, its rightmost digit, and use radix-twelve arithmetic to sum the removed prefix with the entry on its immediate left, then repeat this process, proceeding leftward, until the leftmost entry is reached. In this particular example, the normalized string ends with for all . The leftmost digit is for , which is obtained by carrying the of at entry . It follows that the length of the normalized value of is equal to the row length, . The integral part of contains exactly one digit because (the number of places to the left the decimal has moved) is one less than the row length. Below is the normalized value of . Compound digits remain in the value because they are radix residues represented in radix ten:

See also

Related Research Articles

<span class="mw-page-title-main">Binomial distribution</span> Probability distribution

In probability theory and statistics, the binomial distribution with parameters n and p is the discrete probability distribution of the number of successes in a sequence of n independent experiments, each asking a yes–no question, and each with its own Boolean-valued outcome: success or failure. A single success/failure experiment is also called a Bernoulli trial or Bernoulli experiment, and a sequence of outcomes is called a Bernoulli process; for a single trial, i.e., n = 1, the binomial distribution is a Bernoulli distribution. The binomial distribution is the basis for the popular binomial test of statistical significance.

<span class="mw-page-title-main">Binomial coefficient</span> Number of subsets of a given size

In mathematics, the binomial coefficients are the positive integers that occur as coefficients in the binomial theorem. Commonly, a binomial coefficient is indexed by a pair of integers nk ≥ 0 and is written It is the coefficient of the xk term in the polynomial expansion of the binomial power (1 + x)n; this coefficient can be computed by the multiplicative formula

In elementary algebra, the binomial theorem (or binomial expansion) describes the algebraic expansion of powers of a binomial. According to the theorem, it is possible to expand the polynomial (x + y)n into a sum involving terms of the form axbyc, where the exponents b and c are nonnegative integers with b + c = n, and the coefficient a of each term is a specific positive integer depending on n and b. For example, for n = 4,

In mathematics, a combination is a selection of items from a set that has distinct members, such that the order of selection does not matter. For example, given three fruits, say an apple, an orange and a pear, there are three combinations of two that can be drawn from this set: an apple and a pear; an apple and an orange; or a pear and an orange. More formally, a k-combination of a set S is a subset of k distinct elements of S. So, two combinations are identical if and only if each combination has the same members. If the set has n elements, the number of k-combinations, denoted by or , is equal to the binomial coefficient

In mathematics, a recurrence relation is an equation according to which the th term of a sequence of numbers is equal to some combination of the previous terms. Often, only previous terms of the sequence appear in the equation, for a parameter that is independent of ; this number is called the order of the relation. If the values of the first numbers in the sequence have been given, the rest of the sequence can be calculated by repeatedly applying the equation.

<span class="mw-page-title-main">Catalan number</span> Recursive integer sequence

In combinatorial mathematics, the Catalan numbers are a sequence of natural numbers that occur in various counting problems, often involving recursively defined objects. They are named after the French-Belgian mathematician Eugène Charles Catalan.

In mathematics, a multiset is a modification of the concept of a set that, unlike a set, allows for multiple instances for each of its elements. The number of instances given for each element is called the multiplicity of that element in the multiset. As a consequence, an infinite number of multisets exist which contain only elements a and b, but vary in the multiplicities of their elements:

In algebra, a binomial is a polynomial that is the sum of two terms, each of which is a monomial. It is the simplest kind of a sparse polynomial after the monomials.

In mathematics, the multinomial theorem describes how to expand a power of a sum in terms of powers of the terms in that sum. It is the generalization of the binomial theorem from binomials to multinomials.

<span class="mw-page-title-main">Pascal's pyramid</span>

In mathematics, Pascal's pyramid is a three-dimensional arrangement of the trinomial numbers, which are the coefficients of the trinomial expansion and the trinomial distribution. Pascal's pyramid is the three-dimensional analog of the two-dimensional Pascal's triangle, which contains the binomial numbers and relates to the binomial expansion and the binomial distribution. The binomial and trinomial numbers, coefficients, expansions, and distributions are subsets of the multinomial constructs with the same names.

<span class="mw-page-title-main">Pentatope number</span> Number in the 5th cell of any row of Pascals triangle

In number theory, a pentatope number is a number in the fifth cell of any row of Pascal's triangle starting with the 5-term row 1 4 6 4 1, either from left to right or from right to left. It is named because it represents the number of 3-dimensional unit spheres which can be packed into a pentatope of increasing side lengths.

In mathematics, the Gaussian binomial coefficients are q-analogs of the binomial coefficients. The Gaussian binomial coefficient, written as or , is a polynomial in q with integer coefficients, whose value when q is set to a prime power counts the number of subspaces of dimension k in a vector space of dimension n over , a finite field with q elements; i.e. it is the number of points in the finite Grassmannian .

<span class="mw-page-title-main">Lazy caterer's sequence</span> Counts pieces of a disk cut by lines

The lazy caterer's sequence, more formally known as the central polygonal numbers, describes the maximum number of pieces of a disk that can be made with a given number of straight cuts. For example, three cuts across a pancake will produce six pieces if the cuts all meet at a common point inside the circle, but up to seven if they do not. This problem can be formalized mathematically as one of counting the cells in an arrangement of lines; for generalizations to higher dimensions, see arrangement of hyperplanes.

<span class="mw-page-title-main">Central binomial coefficient</span> Sequence of numbers ((2n) choose (n))

In mathematics the nth central binomial coefficient is the particular binomial coefficient

In mathematics, Pascal's rule is a combinatorial identity about binomial coefficients. It states that for positive natural numbers n and k,

<span class="mw-page-title-main">Cake number</span> Greatest number of regions into which a cube can be partitioned by n planes

In mathematics, the cake number, denoted by Cn, is the maximum of the number of regions into which a 3-dimensional cube can be partitioned by exactly n planes. The cake number is so-called because one may imagine each partition of the cube by a plane as a slice made by a knife through a cube-shaped cake. It is the 3D analogue of the lazy caterer's sequence.

<span class="mw-page-title-main">Ordered Bell number</span> Number of weak orderings

In number theory and enumerative combinatorics, the ordered Bell numbers or Fubini numbers count the weak orderings on a set of elements. Weak orderings arrange their elements into a sequence allowing ties, such as might arise as the outcome of a horse race.

In the context of combinatorial mathematics, stars and bars is a graphical aid for deriving certain combinatorial theorems. It was popularized by William Feller in his classic book on probability. It can be used to solve many simple counting problems, such as how many ways there are to put n indistinguishable balls into k distinguishable bins.

The trinomial triangle is a variation of Pascal's triangle. The difference between the two is that an entry in the trinomial triangle is the sum of the three entries above it:

<span class="mw-page-title-main">Bernoulli's triangle</span> Array of partial sums of the binomial coefficients

Bernoulli's triangle is an array of partial sums of the binomial coefficients. For any non-negative integer n and for any integer k included between 0 and n, the component in row n and column k is given by:

References

  1. 1 2 Coolidge, J. L. (1949), "The story of the binomial theorem", The American Mathematical Monthly , 56 (3): 147–157, doi:10.2307/2305028, JSTOR   2305028, MR   0028222 .
  2. Maurice Winternitz, History of Indian Literature, Vol. III
  3. Peter Fox (1998). Cambridge University Library: the great collections. Cambridge University Press. p. 13. ISBN   978-0-521-62647-7.
  4. The binomial coefficient is conventionally set to zero if k is either less than zero or greater than n.
  5. Selin, Helaine (2008-03-12). Encyclopaedia of the History of Science, Technology, and Medicine in Non-Western Cultures. Springer Science & Business Media. p. 132. Bibcode:2008ehst.book.....S. ISBN   9781402045592.
  6. The Development of Arabic Mathematics Between Arithmetic and Algebra - R. Rashed "Page 63"
  7. Sidoli, Nathan; Brummelen, Glen Van (2013-10-30). From Alexandria, Through Baghdad: Surveys and Studies in the Ancient Greek and Medieval Islamic Mathematical Sciences in Honor of J.L. Berggren. Springer Science & Business Media. p. 54. ISBN   9783642367366.
  8. Kennedy, E. (1966). Omar Khayyam. The Mathematics Teacher 1958. National Council of Teachers of Mathematics. pp. 140–142. JSTOR   i27957284.
  9. Weisstein, Eric W. (2003). CRC concise encyclopedia of mathematics, p. 2169. ISBN   978-1-58488-347-0.
  10. Hughes, Barnabas (1 August 1989). "The arithmetical triangle of Jordanus de Nemore". Historia Mathematica. 16 (3): 213–223. doi: 10.1016/0315-0860(89)90018-9 .
  11. 1 2 3 4 Edwards, A. W. F. (2013), "The arithmetical triangle", in Wilson, Robin; Watkins, John J. (eds.), Combinatorics: Ancient and Modern, Oxford University Press, pp. 166–180.
  12. Smith, Karl J. (2010), Nature of Mathematics, Cengage Learning, p. 10, ISBN   9780538737586 .
  13. Traité du triangle arithmétique, avec quelques autres petits traitez sur la mesme matière at gallica
  14. Fowler, David (January 1996). "The Binomial Coefficient Function". The American Mathematical Monthly . 103 (1): 1–17. doi:10.2307/2975209. JSTOR   2975209. See in particular p. 11.
  15. "Pascal's Triangle in Probability". 5010.mathed.usu.edu. Retrieved 2023-06-01.
  16. Brothers, H. J. (2012), "Finding e in Pascal's triangle", Mathematics Magazine , 85: 51, doi:10.4169/math.mag.85.1.51, S2CID   218541210 .
  17. Brothers, H. J. (2012), "Pascal's triangle: The hidden stor-e", The Mathematical Gazette , 96: 145–148, doi:10.1017/S0025557200004204, S2CID   233356674 .
  18. Foster, T. (2014), "Nilakantha's Footprints in Pascal's Triangle", Mathematics Teacher , 108: 247, doi:10.5951/mathteacher.108.4.0246
  19. Fine, N. J. (1947), "Binomial coefficients modulo a prime", American Mathematical Monthly , 54 (10): 589–592, doi:10.2307/2304500, JSTOR   2304500, MR   0023257 . See in particular Theorem 2, which gives a generalization of this fact for all prime moduli.
  20. Hinz, Andreas M. (1992), "Pascal's triangle and the Tower of Hanoi", The American Mathematical Monthly, 99 (6): 538–544, doi:10.2307/2324061, JSTOR   2324061, MR   1166003 . Hinz attributes this observation to an 1891 book by Édouard Lucas, Théorie des nombres (p. 420).
  21. Ian Stewart, "How to Cut a Cake", Oxford University Press, page 180
  22. Wilmot, G.P. (2023), The Algebra Of Geometry
  23. For a similar example, see e.g. Hore, P. J. (1983), "Solvent suppression in Fourier transform nuclear magnetic resonance", Journal of Magnetic Resonance, 55 (2): 283–300, Bibcode:1983JMagR..55..283H, doi:10.1016/0022-2364(83)90240-8 .
  24. Karl, John H. (2012), An Introduction to Digital Signal Processing, Elsevier, p. 110, ISBN   9780323139595 .
  25. Hilton, P.; et al. (1989). "Extending the binomial coefficients to preserve symmetry and pattern". Symmetry 2. In International Series in Modern Applied Mathematics and Computer Science. Pergamon. pp. 89–102. doi:10.1016/B978-0-08-037237-2.50013-1. ISBN   9780080372372..
  26. Hilton, P.; et al. (1989). "Extending the binomial coefficients to preserve symmetry and pattern". Symmetry 2. In International Series in Modern Applied Mathematics and Computer Science. Pergamon. pp. 100–102. doi:10.1016/B978-0-08-037237-2.50013-1. ISBN   9780080372372..
  27. Newton, Isaac (1736), "A Treatise of the Method of Fluxions and Infinite Series", The Mathematical Works of Isaac Newton: 1:31–33, But these in the alternate areas, which are given, I observed were the same with the figures of which the several ascending powers of the number 11 consist, viz. , , , , , etc. that is, first 1; the second 1, 1; the third 1, 2, 1; the fourth 1, 3, 3, 1; the fifth 1, 4, 6, 4, 1, and so on.
  28. Morton, Robert L. (1964), "Pascal's Triangle and powers of 11", The Mathematics Teacher, 57 (6): 392–394, doi:10.5951/MT.57.6.0392, JSTOR   27957091 .
  29. Arnold, Robert; et al. (2004), "Newton's Unfinished Business: Uncovering the Hidden Powers of Eleven in Pascal's Triangle", Proceedings of Undergraduate Mathematics Day.
  30. Islam, Robiul; et al. (2020), Finding any row of Pascal's triangle extending the concept of power of 11 .
  31. Winteridge, David J. (1984), "Pascal's Triangle and Powers of 11", Mathematics in School, 13 (1): 12–13, JSTOR   30213884 .
  32. Kallós, Gábor (2006), "A generalization of Pascal's triangle using powers of base numbers" (PDF), Annales Mathématiques, 13 (1): 1–15, doi:10.5802/ambp.211 .
  33. Hilton, P.; et al. (1989). "Extending the binomial coefficients to preserve symmetry and pattern". Symmetry 2. In International Series in Modern Applied Mathematics and Computer Science. Pergamon. pp. 89–91. doi:10.1016/B978-0-08-037237-2.50013-1. ISBN   9780080372372..
  34. Mueller, Francis J. (1965), "More on Pascal's Triangle and powers of 11", The Mathematics Teacher, 58 (5): 425–428, doi:10.5951/MT.58.5.0425, JSTOR   27957164 .
  35. Low, Leone (1966), "Even more on Pascal's Triangle and Powers of 11", The Mathematics Teacher, 59 (5): 461–463, doi:10.5951/MT.59.5.0461, JSTOR   27957385 .
  36. Fjelstad, P. (1991), "Extending Pascal's Triangle", Computers & Mathematics with Applications, 21 (9): 3, doi: 10.1016/0898-1221(91)90119-O .