Вы находитесь на странице: 1из 14

Factorial

In mathematics, the factorial of a positive integer n, denoted by n!, is Selected members of the
the product of all positive integers less than or equal to n: factorial sequence (sequence A000142 in
the OEIS); values specified in scientific
notation are rounded to the displayed
precision

For example, n n!
0 1
1 1
The value of 0! is 1, according to the convention for an empty product.[1] 2 2

The factorial operation is encountered in many areas of mathematics, 3 6


notably in combinatorics, algebra, and mathematical analysis. Its most 4 24
basic use counts the possible distinct sequences – the permutations –
5 120
of n distinct objects: there are n!.
6 720
The factorial function can also be extended to non-integer 7 5 040
argumentswhile retaining its most important properties by
defining x! = Γ(x + 1), where Γ is the gamma function; this is 8 40 320
undefined when x is a negative integer. 9 362 880
10 3 628 800
Contents 11 39 916 800

History 12 479 001 600

Definition 13 6 227 020 800


Factorial of zero 14 87 178 291 200
Applications 15 1 307 674 368 000
Rate of growth and approximations for large n 16 20 922 789 888 000
Computation 17 355 687 428 096 000
Number theory 18 6 402 373 705 728 000
Series of reciprocals 19 121 645 100 408 832 000
Factorial of non-integer values 20 2 432 902 008 176 640 000
The gamma and pi functions
25 1.551 121 004 × 1025
Applications of the gamma function
Factorial in the complex plane 50 3.041 409 320 × 1064

Approximations of the factorial 70 1.197 857 167 × 10100


Non-extendability to negative integers 100 9.332 621 544 × 10157
Factorial-like products and functions 450 1.733 368 733 × 101 000
Double factorial
1 000 4.023 872 601 × 102 567
Multifactorials
Primorial 3 249 6.412 337 688 × 1010 000
Superfactorial 10 000 2.846 259 681 × 1035 659
Pickover’s superfactorial
25 206 1.205 703 438 × 10100 000
Hyperfactorial
100 000 2.824 229 408 × 10456 573
See also
205 023 2.503 898 932 × 101 000 004
References
Sources 1 000 000 8.263 931 688 × 105 565 708
Further reading 10100 1010
101.998 109 775 4820

External links
History
Factorials were used to count permutations at least as early as the 12th century, by Indian scholars.[2] In
1677, Fabian Stedman described factorials as applied to change ringing, a musical art involving the ringing of
many tuned bells.[3] After describing a recursive approach, Stedman gives a statement of a factorial (using the
language of the original):

Now the nature of these methods is such, that the changes on one number comprehends [includes]
the changes on all lesser numbers ... insomuch that a compleat Peal of changes on one number
seemeth to be formed by uniting of the compleat Peals on all lesser numbers into one entire
body.[4]

The notation n! was introduced by the French mathematician Christian Kramp in 1808.[5]

Definition
The factorial function is defined by the product

for integer n ≥ 1. This may be written in pi product notation as

This leads to the recurrence relation

For example,

and so on.

Factorial of zero

The factorial of 0 is 1, or in symbols, 0! = 1.

There are several motivations for this definition:

For n = 0, the definition of n! as a product involves the product of no numbers at all, and so is an example
of the broader convention that the product of no factors is equal to the multiplicative identity (see Empty
product).
There is exactly one permutation of zero objects (with nothing to permute, the only rearrangement is to do
nothing).
It makes many identities in combinatorics valid for all applicable sizes. The number of ways to choose 0
elements from the empty set is given by the binomial coefficient

More generally, the number of ways to choose all n elements among a set of n is
It allows for the compact expression of many formulae, such as the exponential function, as a power
series:

It extends the recurrence relation to 0.

Applications
Although the factorial function has its roots in combinatorics, formulas involving factorials occur in many
areas of mathematics.

There are n! different ways of arranging n distinct objects into a sequence, the permutations of those
objects.[6][7]
Often factorials appear in the denominator of a formula to account for the fact that ordering is to be
ignored. A classical example is counting k-combinations (subsets of k elements) from a set
with n elements. One can obtain such a combination by choosing a k-permutation: successively selecting
and removing one element of the set, k times, for a total of

possibilities. This, however, produces the k-combinations in a particular order that one wishes to ignore;
since each k-combination is obtained in k! different ways, the correct number of k-combinations is

This number is known[8] as the binomial coefficient, because it is also the coefficient of xk in (1 + x)n.
The term is often called a falling factorial (pronounced "n to the falling k").

Factorials occur in algebra for various reasons, such as via the already mentioned coefficients of
the binomial formula, or through averaging over permutations for symmetrization of certain operations.
Factorials also turn up in calculus; for example, they occur in the denominators of the terms of Taylor's
formula,[9] where they are used as compensation terms due to the nth derivative of xn being equivalent
to n!.
Factorials are also used extensively in probability theory[10] and number theory (see below).
Factorials can be useful to facilitate expression manipulation. For instance the number of k-permutations
of n can be written as

while this is inefficient as a means to compute that number, it may serve to prove a symmetry
property[7][8] of binomial coefficients:

The factorial function can be shown, using the power rule, to be


where Dn xn is Euler's notation for the nth derivative of xn.[11]

Rate of growth and approximations for large n


As n grows, the factorial n! increases faster than
all polynomialsand exponential functions (but slower than
and double exponential functions) in n.

Most approximations for n! are based on approximating


its natural logarithm

The graph of the function f(n) = ln n! is shown in the figure on


the right. It looks approximately linear for all reasonable
values of n, but this intuition is false. We get one of the
Plot of the natural logarithm of the factorial
simplest approximations for ln n! by bounding the sum with
an integralfrom above and below as follows:

which gives us the estimate

Hence ln n! ∼ n ln n (see Big O notation). This result plays a key role in the analysis of the computational
complexity of sorting algorithms (see comparison sort). From the bounds on ln n! deduced above we get that

It is sometimes practical to use weaker but simpler estimates. Using the above formula it is easily shown that
n n
for all n we have ( )n < n!, and for all n ≥ 6 we have n! < ( )n.
3 2

For large n we get a better estimate for the


number n! using Stirling's approximation:

This in fact comes from an asymptotic series for the logarithm,


and n factorial lies between this and the next approximation:

Another approximation for ln n! is given by Srinivasa Comparison of Stirling's approximation with the
Ramanujan (Ramanujan 1988) factorial
1
Both this and Stirling's approximation give a relative error on the order of 3 , but Ramanujan's is about four
n
times more accurate. However, if we use two correction terms in a Stirling-type approximation, as with
1
Ramanujan's approximation, the relative error will be of order n5 :

Computation
If efficiency is not a concern, computing factorials is trivial from an algorithmic point of view: successively
multiplying a variable initialized to 1 by the integers up to n (if any) will compute n!, provided the result fits in
the variable. In functional languages, the recursive definition is often implemented directly to illustrate
recursive functions.

The main practical difficulty in computing factorials is the size of the result. To assure that the exact result will
fit for all legal values of even the smallest commonly used integral type (8-bit signed integers) would require
more than 700 bits, so no reasonable specification of a factorial function using fixed-size types can avoid
questions of overflow. The values 12! and 20! are the largest factorials that can be stored in, respectively, the
32-bit and 64-bit integers commonly used in personal computers, however many languages support variable
length integer types capable of calculating very large values.[12] Floating-point representation of an
approximated result allows going a bit further, but this also remains quite limited by possible overflow.
Most calculators use scientific notation with 2-digit decimal exponents, and the largest factorial that fits is
then 69!, because 69! < 10100 < 70!. Other implementations (such as computer software such as spreadsheet
programs) can often handle larger values.

Most software applications will compute small factorials by direct multiplication or table lookup. Larger
factorial values can be approximated using Stirling's formula. Wolfram Alpha can calculate exact results for
the ceiling function and floor function applied to the binary, natural and common logarithm of n! for values
of n up to 249 999, and up to 20 000 000! for the integers.

If the exact values of large factorials are needed, they can be computed using arbitrary-precision arithmetic.
Instead of doing the sequential multiplications ((1 × 2) × 3) × 4..., a program can partition the sequence into
two parts, whose products are roughly the same size, and multiply them using a divide-and-conquer method.
This is often more efficient.[13]

The asymptotically best efficiency is obtained by computing n! from its prime factorization. As documented
by Peter Borwein, prime factorization allows n! to be computed in time O(n(log n log log n)2), provided that
a fast multiplication algorithm is used (for example, the Schönhage–Strassen algorithm).[14] Peter Luschny
presents source code and benchmarks for several efficient factorial algorithms, with or without the use of
a prime sieve.[15]

Number theory
Factorials have many applications in number theory. In particular, n! is necessarily divisible by all prime
numbers up to and including n. As a consequence, n > 5 is a composite number if and only if

A stronger result is Wilson's theorem, which states that


if and only if p is prime.[16][17]

Legendre's formula gives the multiplicity of the prime p occurring in the prime factorization of n! as

or, equivalently,

where sp(n) denotes the sum of the standard base-p digits of n.

Adding 1 to a factorial n! yields a number that is only divisible by primes that are larger than n. This fact can
be used to prove Euclid's theorem that the number of primes is infinite.[18] Primes of the form n! ± 1 are
called factorial primes.

Series of reciprocals
The reciprocals of factorials produce a convergent series whose sum is the exponential base e:

Although the sum of this series is an irrational number, it is possible to multiply the factorials by positive
integers to produce a convergent series with a rational sum:

The convergence of this series to 1 can be seen from the fact that its partial sums are less than one by an
inverse factorial. Therefore, the factorials do not form an irrationality sequence.[19]

Factorial of non-integer values

The gamma and pi functions

Besides nonnegative integers, the factorial can also be defined for non-integer values, but this requires more
advanced tools from mathematical analysis.

One function that fills in the values of the factorial (but with a shift of 1 in the argument), that is often used, is
called the gamma function, denoted Γ(z). It is defined for all complex numbers z except for the non-positive
integers, and given when the real part of z is positive by

Its relation to the factorial is that n! = Γ(n + 1) for every nonnegative integer n.

Euler's original formula for the gamma function was


Carl Friedrich Gauss used the notation Π(z) to denote the
same function, but with argument shifted by 1, so that it
agrees with the factorial for nonnegative integers. This pi
function is defined by

The pi function and gamma function are related by the The gamma function interpolates the factorial
formula Π(z) = Γ(z + 1). Likewise, Π(n) = n! for any function to non-integer values. The main clue is
nonnegative integer n. the recurrence relation generalized to a continuous
domain.
In addition to this, the pi function satisfies the same
recurrence as factorials do, but at every complex
value zwhere it is defined

This is no longer a recurrence relation but a functional


equation. In terms of the gamma function, it is

The values of these functions at half-integer values is


therefore determined by a single one of them:

The factorial function, generalized to all real numbers


except negative integers. For
example, 0! = 1! = 1, (− 1 )! = √ π , 1 ! = √ π .
from which it follows that for n ∈ N, 2 2 2

For example,

It also follows that for n ∈ N,

For example,
The pi function is certainly not the only way to extend factorials to a function defined at almost all complex
values, and not even the only one that is analytic wherever it is defined. Nonetheless it is usually considered
the most natural way to extend the values of the factorials to a complex function. For instance, the Bohr–
Mollerup theorem states that the gamma function is the only function that takes the value 1 at 1, satisfies the
functional equation Γ(n + 1) = nΓ(n), is meromorphic on the complex numbers, and is log-convex on the
positive real axis. A similar statement holds for the pi function as well, using the Π(n) = nΠ(n − 1) functional
equation.

However, there exist complex functions that are probably simpler in the sense of analytic function theory and
which interpolate the factorial values. For example, Hadamard's 'gamma' function (Hadamard 1894) which,
unlike the gamma function, is an entire function.[20]

Euler also developed a convergent product approximation for the non-integer factorials, which can be seen to
be equivalent to the formula for the gamma function above:

However, this formula does not provide a practical means of computing the pi function or the gamma
function, as its rate of convergence is slow.

Applications of the gamma function

The volume of an n-dimensional hypersphere of radius R is

Factorial in the complex plane

Representation through the gamma function


allows evaluation of factorial of complex
argument. Equilines of amplitude and phase of
factorial are shown in figure. Let

Several levels of constant modulus


(amplitude) ρand constant phase φ are shown.
The grid covers the
range −3 ≤ x ≤ 3, −2 ≤ y ≤ 2, with unit steps.
The scratched line shows the level φ = ±π.

Thin lines show intermediate levels of constant


modulus and constant phase. At the poles at
every negative integer, phase and amplitude are
not defined. Equilines are dense in vicinity of
singularities along negative integer values of the Amplitude and phase of factorial of complex argument
argument.

For |z| < 1, the Taylor expansions can be used:


The first coefficients of this expansion are

n gn Approximation

0 1 1
1 −γ −0.577 215 6649

2 π2 γ2 0.989 055 9955


12
+ 2
2 3
3 − ζ(3)
3
π
− 12 − γ6 −0.907 479 0760

where γ is the Euler–Mascheroni constant and ζ is the Riemann zeta function. Computer algebra systems such
as SageMath can generate many terms of this expansion.

Approximations of the factorial

For the large values of the argument, the factorial can be approximated through the integral of the digamma
function, using the continued fraction representation. This approach is due to T. J.
Stieltjes (1894). Writing z! = eP(z) where P(z) is

Stieltjes gave a continued fraction for p(z):

The first few coefficients an are[21]

n an
1
0 12
1
1 30
53
2 210
195
3 371
22 999
4 22 737
29 944 523
5 19 733 142
109 535 241 009
6 48 264 275 462

There is a misconception that ln z! = P(z) or ln Γ(z + 1) = P(z) for any complex z ≠ 0. Indeed, the relation
through the logarithm is valid only for a specific range of values of z in the vicinity of the real axis,
where −π < Im(Γ(z + 1)) < π. The larger the real part of the argument, the smaller the imaginary part should
be. However, the inverse relation, z! = eP(z), is valid for the whole complex plane apart from z = 0. The
convergence is poor in the vicinity of the negative part of the real axis; it is difficult to have good convergence
of any approximation in the vicinity of the singularities. When |Im z| > 2 or Re z > 2, the six coefficients
above are sufficient for the evaluation of the factorial with complex double precision. For higher precision
more coefficients can be computed by a rational QD scheme (Rutishauser's QD algorithm).[22]
Non-extendability to negative integers

The relation n! = n × (n − 1)! allows one to compute the factorial for an integer given the factorial for a
smaller integer. The relation can be inverted so that one can compute the factorial for an integer given the
factorial for a larger integer:

However, this recursion does not permit us to compute the factorial of a negative integer; use of the formula to
compute (−1)! would require a division of a nonzero value by zero, and thus blocks us from computing a
factorial value for every negative integer. Similarly, the gamma function is not defined for zero or negative
integers, though it is defined for all other complex numbers.

Factorial-like products and functions


There are several other integer sequences similar to the factorial that are used in mathematics:

Double factorial

The product of all the odd integers up to some odd positive integer n is called the double factorial of n, and
denoted by n!!.[23] That is,

For example, 9!! = 1 × 3 × 5 × 7 × 9 = 945.

The sequence of double factorials for n = 1, 3, 5, 7,... starts as

1, 3, 15, 105, 945, 10 395, 135 135,... (sequence A001147 in the OEIS)

Double factorial notation may be used to simplify the expression of certain trigonometric integrals,[24] to
provide an expression for the values of the gamma function at half-integer arguments and the volume
of hyperspheres,[25] and to solve many counting problems in combinatorics including counting binary
trees with labeled leaves and perfect matchings in complete graphs.[23][26]

Multifactorials

A common related notation is to use multiple exclamation points to denote a multifactorial, the product of
integers in steps of two (n!!), three (n!!!), or more (see generalizations of the double factorial). The double
factorial is the most commonly used variant, but one can similarly define the triple factorial (n!!!) and so on.
One can define the k-tuple factorial, denoted by n!(k), recursively for positive integers as

1!
In addition, similarly to 0! = 1 = 1, one can define:

For sufficiently large n ≥ 1, the ordinary single factorial function is expanded through the multifactorial
functions as follows:
In the same way that n! is not defined for negative integers, and n!! is not defined for negative even
integers, n!(k) is not defined for negative integers divisible by k.

Primorial

The primorial of a natural number n (sequence A002110 in the OEIS), denoted n#, is similar to the factorial,
but with the product taken only over the prime numbers less than or equal to n. That is,

where p ranges over the prime numbers less than or equal to n. For example, the primorial of 11 is

Superfactorial

Neil Sloane and Simon Plouffe defined a superfactorial in The Encyclopedia of Integer Sequences
(Academic Press, 1995) to be the product of the first n factorials. So the superfactorial of 4 is

In general

Equivalently, the superfactorial is given by the formula

which is the determinant of a Vandermonde matrix.

The sequence of superfactorials starts (from n = 0) as

1, 1, 2, 12, 288, 34 560, 24 883 200, 125 411 328 000,... (sequence A000178 in the OEIS)

By this definition, we can define the k-superfactorial of n (denoted sfk(n)) as:

The 2-superfactorials of n are

1, 1, 2, 24, 6 912, 238 878 720, 5 944 066 965 504 000, 745 453 331 864 786 829 312 000 000,...
(sequence A055462 in the OEIS)

The 0-superfactorial of n is n.
Pickover’s superfactorial

In his 1995 book Keys to Infinity, Clifford Pickover defined a different function n$ that he called the
superfactorial. It is defined by

This sequence of superfactorials starts

(Here, as is usual for compound exponentiation, the grouping is understood to be from right to
c c
left: ab = a(b ).)

This operation may also be expressed as the tetration

or using Knuth's up-arrow notation as

Hyperfactorial

Occasionally the hyperfactorial of n is considered. It is written as H(n) and defined by

For n = 1, 2, 3, 4,... the values of H(n) are 1, 4, 108, 27 648,... (sequence A002109 in the OEIS).

The asymptotic growth rate is

where A = 1.2824... is the Glaisher–Kinkelin constant.[27] H(14) ≈ 1.8474 × 1099 is already almost equal to
a googol, and H(15) ≈ 8.0896 × 10116 is almost of the same magnitude as the Shannon number, the theoretical
number of possible chess games. Compared to the Pickover definition of the superfactorial, the hyperfactorial
grows relatively slowly.

The hyperfactorial function can be generalized to complex numbers in a similar way as the factorial function.
The resulting function is called the K-function.

See also
Alternating factorial
Bhargava factorial
Digamma function
Exponential factorial
Factorial number system
Factorion
List of factorial and binomial topics
Pochhammer symbol, which gives the falling or rising factorial
Subfactorial
Trailing zeros of factorial
Triangular number, the additive analogue of factorial

References
1. Graham, Knuth & Patashnik 1988, p. 111.
2. Biggs, Norman L. (May 1979). "The roots of combinatorics". Historia Mathematica. 6 (2): 109–
136. doi:10.1016/0315-0860(79)90074-0. ISSN 0315-0860.
3. Stedman 1677, pp. 6–9.
4. Stedman 1677, p. 8.
5. Higgins 2008, p. 12
6. Cheng, Eugenia (2017-03-09). Beyond Infinity: An expedition to the outer limits of the mathematical
universe. Profile Books. ISBN 9781782830818.
7. Conway, John H.; Guy, Richard (1998-03-16). The Book of Numbers. Springer Science & Business
Media. ISBN 9780387979939.
8. Knuth, Donald E. (1997-07-04). The Art of Computer Programming: Volume 1: Fundamental Algorithms.
Addison-Wesley Professional. ISBN 9780321635747.
9. "18.01 Single Variable Calculus, Lecture 37: Taylor Series". MIT OpenCourseWare. Fall
2006. Archived from the original on 2018-04-26. Retrieved 2017-05-03.
10. Kardar, Mehran (2007-06-25). "Chapter 2: Probability". Statistical Physics of Particles. Cambridge
University Press. pp. 35–56. ISBN 9780521873420.
11. "18.01 Single Variable Calculus, Lecture 4: Chain rule, higher derivatives". MIT OpenCourseWare. Fall
2006. Archived from the original on 2018-04-26. Retrieved 2017-05-03.
12. "wesselbosman/nFactorial". GitHub. 2017-12-25. Archived from the original on 26 April 2018.
Retrieved 26 April 2018.
13. "Factorial Algorithm". GNU MP Software Manual. Archived from the original on 2013-03-14.
Retrieved 2013-01-22.
14. Borwein, Peter (1985). "On the Complexity of Calculating Factorials". Journal of Algorithms. 6 (3): 376–
380. doi:10.1016/0196-6774(85)90006-9.
15. Luschny, Peter. "Fast-Factorial-Functions: The Homepage of Factorial Algorithms". Archived from the
original on 2005-03-05.
16. O'Connor, John J.; Robertson, Edmund F., "Abu Ali al-Hasan ibn al-Haytham", MacTutor History of
Mathematics archive, University of St Andrews.
17. Weisstein, Eric W. "Wilson's Theorem". MathWorld. Retrieved 2017-05-17.
18. Bostock, Chandler & Rourke 2014, pp. 168.
19. Guy 2004, p. 346.
20. Luschny, Peter. "Hadamard versus Euler – Who found the better Gamma function?". Archived from the
original on 2009-08-18.
21. "5.10". Digital Library of Mathematical Functions. Archived from the original on 2010-05-29.
Retrieved 2010-10-17.
22. Luschny, Peter. "On Stieltjes' Continued Fraction for the Gamma Function". Archived from the original on
2011-05-14.
23. Callan, David (2009), A combinatorial survey of identities for the double
factorial, arXiv:0906.1317, Bibcode:2009arXiv0906.1317C.
24. Meserve, B. E. (1948), "Classroom Notes: Double Factorials", The American Mathematical
Monthly, 55 (7): 425–426, doi:10.2307/2306136, JSTOR 2306136, MR 1527019
25. Mezey, Paul G. (2009), "Some dimension problems in molecular databases", Journal of Mathematical
Chemistry, 45 (1): 1–6, doi:10.1007/s10910-008-9365-8.
26. Dale, M. R. T.; Moon, J. W. (1993), "The permuted analogues of three Catalan sets", Journal of Statistical
Planning and Inference, 34 (1): 75–87, doi:10.1016/0378-3758(93)90035-5, MR 1209991.
27. Weisstein, Eric W. "Glaisher–Kinkelin Constant". MathWorld.
Sources
Bostock, Linda; Chandler, Suzanne; Rourke, C. (2014-11-01), Further Pure Mathematics, Nelson
Thornes, ISBN 9780859501033
Graham, Ronald L.; Knuth, Donald E.; Patashnik, Oren (1988), Concrete Mathematics, Reading, MA:
Addison-Wesley, ISBN 0-201-14236-8
Guy, Richard K. (2004), "E24 Irrationality sequences", Unsolved problems in number theory (3rd
ed.), Springer-Verlag, ISBN 0-387-20860-7, Zbl 1058.11001
Higgins, Peter (2008), Number Story: From Counting to Cryptography, New York: Copernicus, ISBN 978-
1-84800-000-1
Stedman, Fabian (1677), Campanalogia, London The publisher is given as "W.S." who may have been
William Smith, possibly acting as agent for the Society of College Youths, to which society the "Dedicatory"
is addressed.

Further reading
Hadamard, M. J. (1894), "Sur L'Expression Du Produit 1·2·3· · · · ·(n−1) Par Une Fonction
Entière" (PDF), Œuvres de Jacques Hadamard (in French), Paris (1968): Centre National de la
Recherche Scientifiques
Ramanujan, Srinivasa (1988), The Lost Notebook and Other Unpublished Papers, Springer Berlin,
p. 339, ISBN 3-540-18726-X

External links
Factorial (mathematics) at the Encyclopædia Britannica
Hazewinkel, Michiel, ed. (2001) [1994], "Factorial", Encyclopedia of Mathematics, Springer
Science+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4
Weisstein, Eric W. "Factorial". MathWorld.
Factorial at PlanetMath.org.

Вам также может понравиться