Fact-checked by Grok 2 weeks ago

Pascal's rule

Pascal's rule, also known as Pascal's identity or Pascal's formula, is a fundamental combinatorial identity that expresses the relationship between consecutive . It states that for non-negative integers n and k with $1 \leq k \leq n, the \binom{n}{k} equals the sum of \binom{n-1}{k-1} and \binom{n-1}{k}, or mathematically,
\binom{n}{k} = \binom{n-1}{k-1} + \binom{n-1}{k}.
This forms the basis for constructing , where each entry is the sum of the two entries directly above it in the previous row, and it directly follows from the definition of as \binom{n}{k} = \frac{n!}{k!(n-k)!}.
The rule's significance lies in its applications to and , particularly in proving the , which expands (x + y)^n = \sum_{k=0}^{n} \binom{n}{k} x^{n-k} y^k. Binomial coefficients \binom{n}{k} count the number of ways to choose k items from n without regard to order, and Pascal's rule provides an efficient recursive method to compute them, avoiding direct calculations for large values. It also underpins symmetries in , such as the equality \binom{n}{k} = \binom{n}{n-k}, and extends to generalizations in higher-order binomial coefficients. Although named after the French mathematician (1623–1662), who systematically explored its properties in his 1653 Treatise on the Arithmetical Triangle, the underlying triangle and recurrence were known much earlier across cultures. Chinese mathematician Jia Xian described an early version in the 11th century, Indian Jain texts from the 2nd century BCE featured a similar "Meru prastara" structure for permutations and combinations, and Persian scholar al-Karaji presented it in the 10th century. Pascal's contributions formalized its use in probability and binomial expansions, influencing later developments like Isaac Newton's generalized .

Background and Statement

Binomial Coefficients

The binomial coefficient \binom{n}{k} is defined as the number of ways to select k items from a set of n distinct items without regard to the order of selection, representing a fundamental concept in known as combinations. This quantity arises naturally in counting problems where the arrangement of chosen elements does not matter. For non-negative integers n \geq k \geq 0, the binomial coefficient is computed using the factorial formula: \binom{n}{k} = \frac{n!}{k!(n-k)!}, where n! denotes the factorial of n, the product of all positive integers up to n. This expression provides an explicit way to calculate the value, leveraging the multiplicative nature of factorials to account for the total permutations divided by the internal arrangements of the selected and remaining items. A key property of binomial coefficients is their , given by \binom{n}{k} = \binom{n}{n-k}, which reflects the between choosing k items to include and choosing n-k items to exclude from the set. For example, \binom{5}{2} = 10, illustrating the 10 possible ways to choose 2 cards from a deck of 5 specific cards, such as selecting the and alongside other pairs. Binomial coefficients serve as the coefficients in the expansion provided by the , which states that for any real numbers x and y, and non-negative n, (x + y)^n = \sum_{k=0}^n \binom{n}{k} x^{n-k} y^k. This theorem expands powers of binomials and highlights the central role of these coefficients in algebraic expansions. For n=3, the theorem yields (x + y)^3 = x^3 + 3x^2 y + 3x y^2 + y^3, where the coefficients 1, 3, 3, and 1 correspond to \binom{3}{0}, \binom{3}{1}, \binom{3}{2}, and \binom{3}{3}, respectively.

Statement of Pascal's Rule

Pascal's rule, also known as Pascal's identity, provides a recursive relation for binomial coefficients. For positive integers n and k with k \leq n, \binom{n}{k} = \binom{n-1}{k-1} + \binom{n-1}{k}. This identity is valid for non-negative integers n \geq 0 and integers k, where the binomial coefficients satisfy the boundary conditions \binom{n}{0} = 1 and \binom{n}{n} = 1 for all n \geq 0, and \binom{n}{k} = 0 if k < 0 or k > n. Combinatorially, the rule reflects that the number of ways to choose k elements from a set of n elements equals the number of such subsets that include a particular fixed element (requiring k-1 more choices from the remaining n-1 elements) plus the number that exclude it (requiring k choices from the remaining n-1 elements). For instance, applying the rule yields \binom{4}{2} = \binom{3}{1} + \binom{3}{2}. With \binom{3}{1} = 3 and \binom{3}{2} = 3, this simplifies to $3 + 3 = 6.

Proofs

Combinatorial Proof

A of Pascal's rule relies on interpreting the s as the number of ways to select s, providing an intuitive counting argument. Consider a set S with n elements, labeled $1 through n. The \binom{n}{k} counts the total number of s of S with exactly k elements. To count these k-s in another way, partition them based on whether they include the element n or not. The s that exclude element n are precisely the k-s of the remaining set \{1, 2, \dots, n-1\}, of which there are \binom{n-1}{k}. The s that include element n consist of \{n\} a (k-1)- of \{1, 2, \dots, n-1\}, of which there are \binom{n-1}{k-1}. Since these two collections of s are disjoint and cover all k-s of S, it follows that \binom{n}{k} = \binom{n-1}{k} + \binom{n-1}{k-1}. This directly establishes Pascal's rule through double counting. For a concrete illustration, take n=4 and k=2, so S = \{1, 2, 3, 4\}. The subsets excluding $4 are the $2-subsets of \{1, 2, 3\}: \{1,2\}, \{1,3\}, \{2,3\} (3 ways, matching \binom{3}{2} = 3). The subsets including $4 are \{1,4\}, \{2,4\}, \{3,4\} (3 ways, matching \binom{3}{1} = 3). The total of 6 subsets equals \binom{4}{2} = 6. This approach highlights the rule's combinatorial essence by linking it to subset selection without resorting to algebraic expansions, offering clear insight into why the identity holds.

Algebraic Proof

The algebraic proof of Pascal's rule proceeds by direct manipulation of the expressions for the s. Recall that the is defined as \dbinom{n}{k} = \frac{n!}{k!(n-k)!} for nonnegative integers n and $0 \leq k \leq n. To verify \dbinom{n}{k} = \dbinom{n-1}{k-1} + \dbinom{n-1}{k}, begin by expressing the right-hand side using the definition: \dbinom{n-1}{k-1} + \dbinom{n-1}{k} = \frac{(n-1)!}{(k-1)!(n-k)!} + \frac{(n-1)!}{k!(n-1-k)!}. The common denominator is k!(n-k)!. Rewrite the first term by multiplying the numerator and denominator by k: \frac{k \cdot (n-1)!}{k!(n-k)!}, and the second term by multiplying the numerator and denominator by (n-k), noting that (n-k)! = (n-k) \cdot (n-1-k)!: \frac{(n-k) \cdot (n-1)!}{k!(n-k)!}. Adding these yields \frac{[k + (n-k)] (n-1)!}{k!(n-k)!} = \frac{n (n-1)!}{k!(n-k)!} = \frac{n!}{k!(n-k)!} = \dbinom{n}{k}. This confirms the identity for $1 \leq k \leq n-1. The rule holds trivially at the boundaries: for k=0, \dbinom{n}{0} = 1 and \dbinom{n-1}{-1} + \dbinom{n-1}{0} = 0 + 1 = 1 by convention; similarly for k=n, \dbinom{n}{n} = 1 = \dbinom{n-1}{n-1} + \dbinom{n-1}{n} = 1 + 0.

Historical Development

Pre-Pascal Discoveries

The earliest known precursors to and its recursive properties emerged in ancient through work in . Around 200 BCE, the mathematician and poet , in his treatise Chandaḥśāstra, implicitly employed a method for counting combinations of long and short syllables in poetic meters, which corresponds to the generation of binomial coefficients via additive , later formalized as the Meru Prastara or "staircase of ." This approach, while not presenting the full triangle, laid foundational combinatorial insights by enumerating patterns of syllable arrangements. In during the 11th century, mathematician Jia Xian advanced these ideas significantly. In his lost work Shi Suo Suan Fa (Method of Interpolation for the Nine Chapters on the Mathematical Art), composed around 1050 , Jia constructed an explicit table of binomial coefficients up to the sixth power, using the additive property to generate entries for extracting roots of polynomials. This table demonstrated the recursive summation where each entry is the sum of the two above it, applied practically to algebraic computations. Persian mathematicians further refined the triangle's construction and applications in the . Al-Karaji, working in around 1000 CE, included a sideways version of the triangle in his algebraic texts Al-Fakhri and Al-Badi, employing it to derive coefficients for expansions such as (a + b)^n up to n=4, recognizing the pattern through successive additions. Building on this, al-Samaw'al in the , in Al-Bahir fi'l-Jabr, attributed the method to al-Karaji and extended its use for higher powers, illustrating the summing property in proofs without formal induction. The Islamic scholarly tradition played a crucial role in preserving, refining, and transmitting these mathematical patterns across . From the onward, works by figures like al-Karaji and al-Samaw'al integrated the triangle into broader algebraic studies, facilitating its dissemination through trade routes and translations to regions including by the late medieval period. Despite these developments, pre-Pascal sources emphasized practical construction and applications of the triangle—such as in prosody, root extraction, and expansions—without explicitly stating or naming a general recursive rule for binomial coefficients.

Pascal's Contribution

formalized the properties of the arithmetical triangle in his treatise Traité du triangle arithmétique, written around 1654 and published posthumously in 1665. In this work, he systematically described the triangle's construction and numerous properties, including its representation of figurate numbers, combinations, and coefficients, with each entry formed by summing the two adjacent numbers above it. Pascal explicitly stated and proved the recursive relation—now known as Pascal's rule—both algebraically through manipulation of expansions and combinatorially by interpreting the entries as counts of ways to choose subsets, thereby linking the triangle to combinatorial problems. Pascal's innovations were particularly evident in his applications of the triangle to probability, motivated by practical issues such as dividing stakes in interrupted games of chance, as posed by the gambler Chevalier de Méré in 1654. He connected the triangle's coefficients to calculating fair divisions based on remaining plays, using the to compute probabilities efficiently, which formed a key part of his correspondence with that year. This exchange established foundational principles of , with Pascal's rigorous approach influencing Fermat's methods and laying groundwork for modern probabilistic reasoning. In the European context of the , Pascal built upon earlier Latin translations of mathematical texts from the , which had preserved knowledge of similar triangular arrays from Islamic scholars like Ibn Munʿim, though these had not been widely integrated into Western mathematics. His key contribution was providing formal proofs, including the first explicit use of in , and extending the triangle's utility to infinite series and figurate numbers, thereby revitalizing and systematizing the structure for contemporary analysis. Despite its ancient origins in non-Western traditions, the triangle and its recursive rule became known as Pascal's in Western mathematical literature due to his comprehensive treatment and popularization.

Generalizations

Multinomial Extension

The multinomial coefficient, denoted \binom{n}{k_1, k_2, \dots, k_m}, is defined as \frac{n!}{k_1! k_2! \dots k_m!} for nonnegative integers k_1, k_2, \dots, k_m satisfying \sum_{i=1}^m k_i = n. Pascal's rule extends naturally to multinomial coefficients via the recursive identity \binom{n}{k_1, \dots, k_m} = \sum_{i=1}^m \binom{n-1}{k_1, \dots, k_i-1, \dots, k_m}, where the sum is over all terms in which exactly one k_i is decremented by 1 (assuming that k_i \geq 1 for the corresponding term; otherwise, that term is zero). This holds for n \geq 1 and m \geq 2, with the binomial case arising as the special instance when m=2. Combinatorially, the multinomial coefficient \binom{n}{k_1, \dots, k_m} counts the number of ways to a set of n distinct items into m labeled groups of sizes k_1, \dots, k_m. The arises by considering the placement of the nth item: it can join any of the m groups, increasing the size of that group by 1, which reduces the problem to partitioning the remaining n-1 items into the adjusted group sizes. For example, consider m=3, n=3, and k_1 = k_2 = k_3 = 1. The multinomial coefficient \binom{3}{1,1,1} = 6 equals the sum \binom{2}{0,1,1} + \binom{2}{1,0,1} + \binom{2}{1,1,0}, where each term is 2 (e.g., \binom{2}{0,1,1} = \frac{2!}{0!1!1!} = 2), reflecting the 3 choices for assigning the third item to one of the three groups, each yielding 2 ways to partition the first two items. A combinatorial proof follows analogously to the binomial case: the total partitions of n items into the specified groups equal the disjoint union over the m possibilities for the group containing the nth item, each recursively partitioning the rest.

Complex Number Extension

The gamma function extends the factorial to complex numbers, satisfying \Gamma(z+1) = z \Gamma(z) for complex z not equal to a non-positive integer, where \Gamma(z) is defined by the integral \Gamma(z) = \int_0^\infty t^{z-1} e^{-t} \, dt for \operatorname{Re}(z) > 0 and by analytic continuation elsewhere. This relation generalizes n! to non-integer and complex arguments, enabling the definition of binomial coefficients beyond non-negative integers. The generalized binomial coefficient for complex z and w is given by \binom{z}{w} = \frac{\Gamma(z+1)}{\Gamma(w+1) \Gamma(z - w + 1)}, provided the gamma functions are defined (i.e., avoiding poles). Using the recurrence \Gamma(z+1) = z \Gamma(z), Pascal's rule extends analytically to this setting: \binom{z}{w} = \binom{z-1}{w-1} + \binom{z-1}{w}, as the algebraic manipulation of the gamma function ratios preserves the identity where defined. For instance, direct evaluation yields \binom{1/2}{1/2} = \frac{\Gamma(3/2)}{\Gamma(3/2) \Gamma(1)} = 1, since \Gamma(3/2) = \frac{1}{2} \Gamma(1/2) = \frac{\sqrt{\pi}}{2} and \Gamma(1) = 1. The identity facilitates iterative computation of such coefficients; starting from base cases like \binom{z}{0} = 1 and \binom{z}{1} = z, one can recursively find values such as \binom{1/2}{2} = \frac{1/2 \cdot (-1/2)}{2!} = -\frac{1}{8} by applying the rule forward for integer steps in w. These extensions are particularly useful in series expansions, such as the generalized (1 + x)^z = \sum_{k=0}^\infty \binom{z}{k} x^k for |x| < [1](/page/1) and non-integer z, which arises in solutions to differential equations and approximations for non-integer powers. However, the generalized coefficients exhibit limitations due to the poles of the gamma function at non-positive integers; for example, if z is a negative integer and w is not an integer, \binom{z}{w} becomes infinite as the numerator diverges while the denominator remains finite.

Applications

In Pascal's Triangle Construction

Pascal's triangle is an infinite triangular array of numbers where the entries in row n (indexed starting from n=0) are the binomial coefficients \binom{n}{0}, \binom{n}{1}, \dots, \binom{n}{n}, arranged from left to right. Each interior entry in row n is formed by adding the two entries directly above it from row n-1, embodying Pascal's rule as the computational basis for generating these coefficients. The construction algorithm initiates with row 0 as a single entry: . For each subsequent row n \geq 1, the row begins and ends with 1, and each interior entry at position k (where $1 \leq k \leq n-1) is the sum of the entry at position k-1 and position k from row n-1. This row-by-row summation process systematically builds the triangle, enabling the computation of without direct factorial calculations. To illustrate, the first five rows (0 through 4) of Pascal's triangle are constructed as follows:
RowEntries
01
11 1
21 2 1
31 3 3 1
41 4 6 4 1
In row 4, for example, the entry \binom{4}{2} = 6 is computed as $3 + 3, summing the adjacent entries from row 3. This recursive method yields several key properties: all entries are positive integers, arising from repeated additions starting from 1, and the leftmost and rightmost entries in every row are invariably 1, as they have only one "parent" entry above (or none for row 0). The efficiency of this construction for the triangle up to row n is O(n^2) time complexity, since each of the \frac{(n+1)(n+2)}{2} entries requires a constant-time addition based on prior rows, making it a practical tool for computing iteratively.

In Probability Distributions

Pascal's rule finds application in probability theory through its role in deriving recursive relations for the probability mass function (PMF) of the , which models the number of successes in a fixed number of independent Bernoulli trials. The PMF is given by P(X = k) = \binom{n}{k} p^k (1-p)^{n-k}, where n is the number of trials, p is the success probability on each trial, k = 0, 1, \dots, n, and X is the number of successes. Substituting Pascal's rule \binom{n}{k} = \binom{n-1}{k-1} + \binom{n-1}{k} into the PMF yields a recursion relating the probabilities for n trials to those for n-1 trials: P(X = k \mid n, p) = p \cdot P(X = k-1 \mid n-1, p) + (1-p) \cdot P(X = k \mid n-1, p), with boundary conditions P(X = k \mid n, p) = 0 if k < 0 or k > n. This relation arises from conditioning on the outcome of the nth trial: a success contributes to k successes with probability p if there were k-1 in the first n-1 trials, while a failure contributes with probability $1-p if there were k in the first n-1 trials. The recursion allows computation of the full PMF starting from the base case of n=0, where P(X=0 \mid 0, p) = 1, and preserves the property that probabilities sum to 1 across values of k. This recursive structure extends to the (CDF) F(n, k; p) = P(X \leq k \mid n, p), which can be computed iteratively from smaller values of n using relations derived from the PMF recursion, such as integrating the probabilities up to k. For instance, the CDF for n trials can be expressed in terms of the CDFs and PMFs for n-1 trials, facilitating efficient numerical evaluation especially for large n. To illustrate, consider n=3 trials with p=0.5. Starting from n=1: P(X=0)=0.5, P(X=1)=0.5. For n=2: P(X=0)=0.5 \cdot 0.5 = 0.25, P(X=1)=0.5 \cdot 0.5 + 0.5 \cdot 0.5 = 0.5, P(X=2)=0.5 \cdot 0.5 = 0.25. For n=3: P(X=0)=0.5 \cdot 0.25 = 0.125, P(X=1)=0.5 \cdot 0.25 + 0.5 \cdot 0.5 = 0.375, P(X=2)=0.5 \cdot 0.5 + 0.5 \cdot 0.25 = 0.375, P(X=3)=0.5 \cdot 0.25 = 0.125. The probabilities sum to $0.125 + 0.375 + 0.375 + 0.125 = 1, as expected from the recursive preservation of the total probability. A similar recursive structure applies to the , which models the number of failures before a fixed number r of successes in trials. The PMF is P(Y = k) = \binom{k + r - 1}{k} p^r (1-p)^k for k = 0, 1, \dots, where the generalized \binom{k + r - 1}{k} satisfies an analogous form of Pascal's rule: \binom{k + r - 1}{k} = \binom{k + r - 2}{k-1} + \binom{k + r - 2}{k}. This leads to a for the PMF relating probabilities for k failures to those for k-1 failures, mirroring the case in structure.

References

  1. [1]
    Pascal's Formula -- from Wolfram MathWorld
    Each subsequent row of Pascal's triangle is obtained by adding the two entries diagonally above. This follows immediately from the binomial coefficient ...
  2. [2]
    1.3 Binomial coefficients
    The entries in Pascal's Triangle are the coefficients of the polynomial produced by raising a binomial to an integer power.
  3. [3]
    Blaise Pascal - Biography - MacTutor - University of St Andrews
    He discovered that the sum of the angles of a triangle are two right angles and, when his father found out, he relented and allowed Blaise a copy of Euclid. At ...
  4. [4]
    Jia Xian (1010 - 1070) - Biography - MacTutor History of Mathematics
    Jia Xian was a Chinese scholar who wrote some early arithmetic books. He had an early version of Pascal's triangle. Biography. Jia Xian is also known as Chia ...
  5. [5]
    Chronology for 900 - 1100 - MacTutor History of Mathematics
    Al-Karaji writes Al-Fakhri in Baghdad which develops algebra. He gives Pascal's triangle. 1000. Ibn al-Haytham (often called Alhazen) writes works on optics ...
  6. [6]
    Binomial Coefficient -- from Wolfram MathWorld
    The binomial coefficient (n; k) is the number of ways of picking k unordered outcomes from n possibilities, also known as a combination or combinatorial number.
  7. [7]
    Binomial coefficient - StatLect
    In combinatorics, the binomial coefficient is used to denote the number of possible ways to choose a subset of objects of a given numerosity from a larger set.
  8. [8]
    Binomial Coefficient - an overview | ScienceDirect Topics
    The binomial coefficient refers to the number of ways of choosing 'r' objects out of 'n' without regard to order, calculated using the formula (nCr) = n!
  9. [9]
    Binomial Theorem - BYJU'S
    Binomial Theorem Statement. The binomial theorem is the method of expanding an expression that has been raised to any finite power. A binomial theorem is a ...
  10. [10]
    [PDF] Ch 6.4: Binomial Coefficients and Identities
    Kyle Berney – Ch 6.4: Binomial Coefficients and Identities. 3-1. Pascal's Identity. Theorem 1: (Pascal's Identity) Let n and k be positive integers, such that ...
  11. [11]
    [PDF] Problem Solving in Math (Math 43900) Fall 2013
    ) = 0 if k>n or if k < 0, and that ( n. 0. ) = (n n. ) = 1 (and so in ... Express an,k as a (single) binomial coefficient. (c) Use the results of the ...
  12. [12]
    [PDF] n! (n-r)! n! r!(n-r)! n r - andrew.cmu.ed
    Pascal's identity. The art of combinatorial proof. How many ways we can create an even size committee out of n people? LHS : There are so many such committees.
  13. [13]
    ECS20 homepage / Supplement
    Examples of Combinatorial Proof. Pascal's identity: C(n+1,k) = C(n,k-1) + C(n,k). Proof: Suppose T is a set of containing n+1 elements.
  14. [14]
    Combinatorial Proofs - Discrete Mathematics - An Open Introduction
    These proofs can be done in many ways. One option would be to give algebraic proofs, using the formula for ...Missing: factorials | Show results with:factorials
  15. [15]
    Earliest Known Uses of Some of the Words of Mathematics (P)
    Pascal's triangle appears in 1886 in Algebra by George Chrystal (1851-1911). While triangle research in Western Europe only gained momentum in the Renaissance, ...
  16. [16]
    [PDF] Computational Aspects of the Aryabhata Algorithm
    The Meru Prastara is a procedure to find combinations that was described by Pingala in 200 B.C.¹¹ Algebra that appears in Aryabhaṭiya can be seen to be an ...<|control11|><|separator|>
  17. [17]
    al-Karaji (953 - 1029) - Biography - MacTutor History of Mathematics
    The table al-Karaji constructed looks like the Pascal triangle on its side. The sum of the squares of the numbers that follow one another in natural order from ...
  18. [18]
    Al-Samawal (1130 - Biography - MacTutor History of Mathematics
    Also in this book is al-Samawal's description of the binomial theorem where the coefficients are given by the Pascal triangle. The method is attributed by al ...
  19. [19]
    [PDF] INDIAN NUMERALS AND CALCULATION IN THE ISLAMIC WORLD
    ... transmitted directly to Europe ... T11is document, which uses "Hindi" numerals to reproduce what i, known as "Pascal's triangle~ shows that Muslim mathematicians ...
  20. [20]
    [PDF] Pascal's Treatise on the Arithmetical Triangle
    Around 1654 Pascal conducted his studies on the Arithmetical Triangle (“Pascal's Triangle”) and its relationship to probabilities. His correspondence with ...
  21. [21]
    [PDF] FERMAT AND PASCAL ON PROBABILITY - University of York
    ” The correspondence which ensued between Fermat and Pascal, was fundamental in the development of modern concepts of probability, and it is unfortunate ...
  22. [22]
    July 1654: Pascal's Letters to Fermat on the "Problem of Points"
    Jul 1, 2009 · Fermat's approach rested on a complete enumeration of the possible outcomes. For example, if the winner of a coin toss game needs to win the ...
  23. [23]
    Rediscovering Arabic Science - Muslim HeritageMuslim Heritage
    Aug 12, 2009 · Four hundred years later, the 17th-century French mathematician Blaise Pascal reinvented Ibn Munim's numerical grid. The famous last theorem by ...
  24. [24]
    From Pascal's Triangle to the Bell-shaped Curve
    Blaise Pascal (1623-1662) did not invent his triangle. It was at least 500 years old when he wrote it down, in 1654 or just after, in his Traité du triangle ...
  25. [25]
    [PDF] Chapter 4 Some Counting Problems; Multinomial Coefficients, The ...
    It can be shown that Snp satisfies the following pecu- liar version of Pascal's recurrence formula: ... This equation is often called the absorption identity.
  26. [26]
    DLMF: §5.2 Definitions ‣ Properties ‣ Chapter 5 Gamma Function
    When R ⁡ z ≤ 0 , Γ ⁡ ( z ) is defined by analytic continuation. It is a meromorphic function with no zeros, and with simple poles of residue.
  27. [27]
    Calculus II - Binomial Series - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will give the Binomial Theorem and illustrate how it can be used to quickly expand terms in the form (a+b)^n when n is an ...
  28. [28]
    [PDF] Arithmetic Triangles and Pascal-Type Recurrence Relations
    Pascal's rule for Binomial coefficients says that n + 1 k. = n k − 1. + n k . We notice that this is exactly the recurrence relation that generates the ...
  29. [29]
    [PDF] an introduction to pascal's triangle - Carroll Collected
    This lesson will focus on counting techniques used in later lessons to understand better and define Pascal's Triangle, The Binomial Theorem, and Binomial ...
  30. [30]
    [PDF] Pascal's Triangle and Mathematical Induction
    The project covers Blaise Pascal's (1623–1662) arrangement of the figurate numbers (binomial coefficients) into one table, known as Pascal's triangle [7].
  31. [31]
  32. [32]
    Understanding the Binomial Recursively
    ### Summary of Binomial Recursion from https://jtr13.github.io/website/Binomial.html
  33. [33]
    Efficient computation of the P.M.F. and the C.D.F. of the generalized ...
    It is possible to compute E(k,n,p) and B(k,n,p) via computer implementations of recursive functions that are directly based on the aforementioned recursive ...
  34. [34]
    Negative Binomial Distribution - VrcAcademy
    Recurrence Relation for the probability of Negative Binomial Distribution ... Proof. The p.m.f. of negative binomial distribution is. P ( X = x ) = ( ...Missing: recursion | Show results with:recursion