Arithmetic progression

Last updated
Proof without words of the arithmetic progression formulas using a rotated copy of the blocks Arithmetic progression.svg
Proof without words of the arithmetic progression formulas using a rotated copy of the blocks

An arithmetic progression or arithmetic sequence (AP) is a sequence of numbers such that the difference from any succeeding term to its preceding term remains constant throughout the sequence. The constant difference is called common difference of that arithmetic progression. For instance, the sequence 5, 7, 9, 11, 13, 15, . . . is an arithmetic progression with a common difference of 2.

Contents

If the initial term of an arithmetic progression is and the common difference of successive members is , then the -th term of the sequence () is given by:

A finite portion of an arithmetic progression is called a finite arithmetic progression and sometimes just called an arithmetic progression. The sum of a finite arithmetic progression is called an arithmetic series.

History

According to an anecdote of uncertain reliability, [1] young Carl Friedrich Gauss, who was in primary school, reinvented this method to compute the sum of the integers from 1 through 100, by multiplying n/2 pairs of numbers in the sum by the values of each pair n + 1.[ clarification needed ] However, regardless of the truth of this story, Gauss was not the first to discover this formula, and some find it likely that its origin goes back to the Pythagoreans in the 5th century BC. [2] Similar rules were known in antiquity to Archimedes, Hypsicles and Diophantus; [3] in China to Zhang Qiujian; in India to Aryabhata, Brahmagupta and Bhaskara II; [4] and in medieval Europe to Alcuin, [5] Dicuil, [6] Fibonacci, [7] Sacrobosco [8] and to anonymous commentators of Talmud known as Tosafists. [9]

Sum

2+5+8+11+14=40
14+11+8+5+2=40

16+16+16+16+16=80

Computation of the sum 2 + 5 + 8 + 11 + 14. When the sequence is reversed and added to itself term by term, the resulting sequence has a single repeated value in it, equal to the sum of the first and last numbers (2 + 14 = 16). Thus 16 × 5 = 80 is twice the sum.

The sum of the members of a finite arithmetic progression is called an arithmetic series. For example, consider the sum:

This sum can be found quickly by taking the number n of terms being added (here 5), multiplying by the sum of the first and last number in the progression (here 2 + 14 = 16), and dividing by 2:

In the case above, this gives the equation:

This formula works for any real numbers and . For example: this

Derivation

Animated proof for the formula giving the sum of the first integers 1+2+...+n. Animated proof for the formula giving the sum of the first integers 1+2+...+n.gif
Animated proof for the formula giving the sum of the first integers 1+2+...+n.

To derive the above formula, begin by expressing the arithmetic series in two different ways:

Rewriting the terms in reverse order:

Adding the corresponding terms of both sides of the two equations and halving both sides:

This formula can be simplified as:

Furthermore, the mean value of the series can be calculated via: :

The formula is very similar to the mean of a discrete uniform distribution.

Product

The product of the members of a finite arithmetic progression with an initial element a1, common differences d, and n elements in total is determined in a closed expression

where denotes the Gamma function. The formula is not valid when is negative or zero.

This is a generalization from the fact that the product of the progression is given by the factorial and that the product

for positive integers and is given by

Derivation

where denotes the rising factorial.

By the recurrence formula , valid for a complex number ,

,
,

so that

for a positive integer and a positive complex number.

Thus, if ,

,

and, finally,

Examples

Example 1

Taking the example , the product of the terms of the arithmetic progression given by up to the 50th term is

Example 2

The product of the first 10 odd numbers is given by

= 654,729,075

Standard deviation

The standard deviation of any arithmetic progression can be calculated as

where is the number of terms in the progression and is the common difference between terms. The formula is very similar to the standard deviation of a discrete uniform distribution.

Intersections

The intersection of any two doubly infinite arithmetic progressions is either empty or another arithmetic progression, which can be found using the Chinese remainder theorem. If each pair of progressions in a family of doubly infinite arithmetic progressions have a non-empty intersection, then there exists a number common to all of them; that is, infinite arithmetic progressions form a Helly family. [10] However, the intersection of infinitely many infinite arithmetic progressions might be a single number rather than itself being an infinite progression.

See also

Related Research Articles

<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

<span class="mw-page-title-main">Gamma function</span> Extension of the factorial function

In mathematics, the gamma function is one commonly used extension of the factorial function to complex numbers. The gamma function is defined for all complex numbers except the non-positive integers. For every positive integer n,

<span class="mw-page-title-main">Riemann zeta function</span> Analytic function in mathematics

The Riemann zeta function or Euler–Riemann zeta function, denoted by the Greek letter ζ (zeta), is a mathematical function of a complex variable defined as

<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.

<span class="mw-page-title-main">Euler's constant</span> Constant value used in mathematics

Euler's constant is a mathematical constant, usually denoted by the lowercase Greek letter gamma, defined as the limiting difference between the harmonic series and the natural logarithm, denoted here by log:

<span class="mw-page-title-main">Stirling's approximation</span> Approximation for factorials

In mathematics, Stirling's approximation is an asymptotic approximation for factorials. It is a good approximation, leading to accurate results even for small values of . It is named after James Stirling, though a related but less precise result was first stated by Abraham de Moivre.

In mathematics, the falling factorial is defined as the polynomial

<span class="mw-page-title-main">Digamma function</span> Mathematical function

In mathematics, the digamma function is defined as the logarithmic derivative of the gamma function:

A powerful number is a positive integer m such that for every prime number p dividing m, p2 also divides m. Equivalently, a powerful number is the product of a square and a cube, that is, a number m of the form m = a2b3, where a and b are positive integers. Powerful numbers are also known as squareful, square-full, or 2-full. Paul Erdős and George Szekeres studied such numbers and Solomon W. Golomb named such numbers powerful.

In mathematics, a divergent series is an infinite series that is not convergent, meaning that the infinite sequence of the partial sums of the series does not have a finite limit.

<span class="mw-page-title-main">Lambert series</span> Mathematical term

In mathematics, a Lambert series, named for Johann Heinrich Lambert, is a series taking the form

<span class="mw-page-title-main">Barnes G-function</span>

In mathematics, the Barnes G-functionG(z) is a function that is an extension of superfactorials to the complex numbers. It is related to the gamma function, the K-function and the Glaisher–Kinkelin constant, and was named after mathematician Ernest William Barnes. It can be written in terms of the double gamma function.

<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 Engel expansion of a positive real number x is the unique non-decreasing sequence of positive integers such that

The gamma function is an important special function in mathematics. Its particular values can be expressed in closed form for integer and half-integer arguments, but no simple expressions are known for the values at rational points in general. Other fractional arguments can be approximated through efficient infinite products, infinite series, and recurrence relations.

In natural language processing, latent Dirichlet allocation (LDA) is a Bayesian network for modeling automatically extracted topics in textual corpora. The LDA is an example of a Bayesian topic model. In this, observations are collected into documents, and each word's presence is attributable to one of the document's topics. Each document will contain a small number of topics.

<span class="mw-page-title-main">Ramanujan's master theorem</span> Mathematical theorem

In mathematics, Ramanujan's Master Theorem, named after Srinivasa Ramanujan, is a technique that provides an analytic expression for the Mellin transform of an analytic function.

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.

References

  1. Hayes, Brian (2006). "Gauss's Day of Reckoning". American Scientist . 94 (3): 200. doi:10.1511/2006.59.200. Archived from the original on 12 January 2012. Retrieved 16 October 2020.
  2. Høyrup, J. The "Unknown Heritage": trace of a forgotten locus of mathematical sophistication. Arch. Hist. Exact Sci. 62, 613–654 (2008). https://doi.org/10.1007/s00407-008-0025-y
  3. Tropfke, Johannes (1924). Analysis, analytische Geometrie . Walter de Gruyter. pp. 3–15. ISBN   978-3-11-108062-8.
  4. Tropfke, Johannes (1979). Arithmetik und Algebra . Walter de Gruyter. pp. 344–354. ISBN   978-3-11-004893-3.
  5. Problems to Sharpen the Young, John Hadley and David Singmaster, The Mathematical Gazette, 76, #475 (March 1992), pp. 102–126.
  6. Ross, H.E. & Knott, B.I. (2019) Dicuil (9th century) on triangular and square numbers, British Journal for the History of Mathematics, 34:2, 79-94, https://doi.org/10.1080/26375451.2019.1598687
  7. Sigler, Laurence E. (trans.) (2002). Fibonacci's Liber Abaci . Springer-Verlag. pp.  259–260. ISBN   0-387-95419-8.
  8. Katz, Victor J. (edit.) (2016). Sourcebook in the Mathematics of Medieval Europe and North Africa . Princeton University Press. pp. 91, 257. ISBN   9780691156859.
  9. Stern, M. (1990). 74.23 A Mediaeval Derivation of the Sum of an Arithmetic Progression. The Mathematical Gazette, 74(468), 157-159. doi:10.2307/3619368
  10. Duchet, Pierre (1995), "Hypergraphs", in Graham, R. L.; Grötschel, M.; Lovász, L. (eds.), Handbook of combinatorics, Vol. 1, 2, Amsterdam: Elsevier, pp. 381–432, MR   1373663 . See in particular Section 2.5, "Helly Property", pp. 393–394.