Factorial









































































Selected members of the factorial sequence (sequence A000142 in the OEIS); values specified in scientific notation are rounded to the displayed precision

n

n!
01
11
22
36
424
5120
6720
7
7003504000000000000♠5040
8
7004403200000000000♠40320
9
7005362880000000000♠362880
10
7006362880000000000♠3628800
11
7007399168000000000♠39916800
12
7008479001600000000♠479001600
13
7009622702080000000♠6227020800
14
7010871782912000000♠87178291200
15
7012130767436800000♠1307674368000
16
7013209227898880000♠20922789888000
17
7014355687428096000♠355687428096000
18
7015640237370572800♠6402373705728000
19
7017121645100408832♠121645100408832000
20
7018243290200817664♠2432902008176640000
25

7025155112100400000♠1.551121004×1025
50

7064304140932000000♠3.041409320×1064
70

7100119785716700000♠1.197857167×10100
100

7157933262154400000♠9.332621544×10157
450

9000000000000000000♠1.733368733×101000

7003100000000000000♠1000

9000000000000000000♠4.023872601×102567

7003324900000000000♠3249

9000000000000000000♠6.412337688×1010000

7004100000000000000♠10000

9000000000000000000♠2.846259681×1035659

7004252060000000000♠25206

9000000000000000000♠1.205703438×10100000

7005100000000000000♠100000

9000000000000000000♠2.824229408×10456573

7005205023000000000♠205023

9000000000000000000♠2.503898932×101000004

7006100000000000000♠1000000

9000000000000000000♠8.263931688×105565708
7100100000000000000♠10100107101995657055180894♠10101.9981097754820

In mathematics, the factorial of a positive integer n, denoted by n!, is the product of all positive integers less than or equal to n. For example,


5!=5×4×3×2×1=120.displaystyle 5!=5times 4times 3times 2times 1=120,.displaystyle 5!=5times 4times 3times 2times 1=120,.

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


The factorial operation is encountered in many areas of mathematics, notably in combinatorics, algebra, and mathematical analysis. Its most basic occurrence is the fact that there are n! ways to arrange n distinct objects into a sequence. These arrangements are called the permutations of the set of objects.


The definition of the factorial function can also be extended to non-integer arguments, while retaining its most important properties; this involves more advanced mathematics, notably techniques from mathematical analysis.




Contents





  • 1 History


  • 2 Definition

    • 2.1 Factorial of zero


    • 2.2 Factorial of a non-integer



  • 3 Applications


  • 4 Rate of growth and approximations for large n


  • 5 Computation


  • 6 Number theory


  • 7 Series of reciprocals


  • 8 Extension of factorial to non-integer values of argument

    • 8.1 The gamma and pi functions


    • 8.2 Applications of the gamma function


    • 8.3 Factorial at the complex plane


    • 8.4 Approximations of the factorial


    • 8.5 Non-extendability to negative integers



  • 9 Factorial-like products and functions

    • 9.1 Double factorial


    • 9.2 Multifactorials


    • 9.3 Primorial


    • 9.4 Superfactorial

      • 9.4.1 Pickover’s superfactorial



    • 9.5 Hyperfactorial



  • 10 See also


  • 11 References

    • 11.1 Sources


    • 11.2 Further reading



  • 12 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


n!=1⋅2⋅3⋅…⋅(n−2)⋅(n−1)⋅n,displaystyle n!=1cdot 2cdot 3cdot ldots cdot (n-2)cdot (n-1)cdot n,displaystyle n!=1cdot 2cdot 3cdot ldots cdot (n-2)cdot (n-1)cdot n,

initially for integer n ≥ 1. This may be written in the Pi product notation as


n!=∏i=1ni.displaystyle n!=prod _i=1^ni.displaystyle n!=prod _i=1^ni.

From these formulas, one may derive the recurrence relation


n!=n⋅(n−1)!.displaystyle n!=ncdot (n-1)!.displaystyle n!=ncdot (n-1)!.

For example, one has


5!=5⋅4!6!=6⋅5!50!=50⋅49!displaystyle beginaligned5!&=5cdot 4!\6!&=6cdot 5!\50!&=50cdot 49!endaligneddisplaystyle beginaligned5!&=5cdot 4!\6!&=6cdot 5!\50!&=50cdot 49!endaligned

and so on.



Factorial of zero


In order for the recurrence relation to be extended to n = 0, it is necessary to define


0!=1displaystyle 0!=1displaystyle 0!=1

so that


1!=1⋅0!=1.displaystyle 1!=1cdot 0!=1.displaystyle 1!=1cdot 0!=1.

There are several independent reasons to view this definition as harmonious. These include the following:


  • In the case 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


(00)=0!0!0!=1displaystyle binom 00=frac 0!0!0!=1displaystyle binom 00=frac 0!0!0!=1.

More generally, the number of ways to choose all n elements among a set of n is

(nn)=n!n!0!=1displaystyle binom nn=frac n!n!0!=1displaystyle binom nn=frac n!n!0!=1.

  • It allows for the compact expression of many formulae, such as the exponential function, as a power series:
ex=∑n=0∞xnn!.displaystyle e^x=sum _n=0^infty frac x^nn!.displaystyle e^x=sum _n=0^infty frac x^nn!.


Factorial of a non-integer


The factorial function can also be defined for non-integer values using more advanced mathematics (the gamma function n! = Γ(n + 1)), detailed in the section below. This more generalized definition is used by advanced calculators and mathematical software such as Maple, Mathematica, or APL.



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

(n−0)(n−1)(n−2)⋯(n−(k−1))=n!(n−k)!=nk_displaystyle (n-0)(n-1)(n-2)cdots left(n-(k-1)right)=tfrac n!(n-k)!=n^underline kdisplaystyle (n-0)(n-1)(n-2)cdots left(n-(k-1)right)=tfrac n!(n-k)!=n^underline k

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
n(n−1)(n−2)⋯(n−k+1)k(k−1)(k−2)⋯1=nk_k!=n!(n−k)!k!=(nk).displaystyle frac n(n-1)(n-2)cdots (n-k+1)k(k-1)(k-2)cdots 1=frac n^underline kk!=frac n!(n-k)!k!=binom nk.displaystyle frac n(n-1)(n-2)cdots (n-k+1)k(k-1)(k-2)cdots 1=frac n^underline kk!=frac n!(n-k)!k!=binom nk.

This number is known[8] as the binomial coefficient, because it is also the coefficient of xk in (1 + x)n. The term nk_displaystyle n^underline kn^underline k 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]

  • Factorials can be useful to facilitate expression manipulation. For instance the number of k-permutations of n can be written as

nk_=n!(n−k)!;displaystyle n^underline k=frac n!(n-k)!,;displaystyle n^underline k=frac n!(n-k)!,;

while this is inefficient as a means to compute that number, it may serve to prove a symmetry property[7][8] of binomial coefficients:
(nk)=nk_k!=n!(n−k)!k!=nn−k_(n−k)!=(nn−k).displaystyle binom nk=frac n^underline kk!=frac n!(n-k)!k!=frac n^underline n-k(n-k)!=binom nn-k,.displaystyle binom nk=frac n^underline kk!=frac n!(n-k)!k!=frac n^underline n-k(n-k)!=binom nn-k,.

  • The factorial function can be shown, using the power rule, to be
n!=Dnxn=dndxnxndisplaystyle n!=D^n,x^n=frac d^ndx^n,x^ndisplaystyle n!=D^n,x^n=frac d^ndx^n,x^n

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


Rate of growth and approximations for large n




Plot of the natural logarithm of the factorial


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


Most approximations for n! are based on approximating its natural logarithm


ln⁡n!=∑x=1nln⁡x.displaystyle ln n!=sum _x=1^nln x,.displaystyle ln n!=sum _x=1^nln x,.

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 simplest approximations for ln n! by bounding the sum with an integral from above and below as follows:


∫1nln⁡xdx≤∑x=1nln⁡x≤∫0nln⁡(x+1)dxdisplaystyle int _1^nln x,dxleq sum _x=1^nln xleq int _0^nln(x+1),dxdisplaystyle int _1^nln x,dxleq sum _x=1^nln xleq int _0^nln(x+1),dx

which gives us the estimate


nln⁡(ne)+1≤ln⁡n!≤(n+1)ln⁡(n+1e)+1.displaystyle nln left(frac neright)+1leq ln n!leq (n+1)ln left(frac n+1eright)+1,.displaystyle nln left(frac neright)+1leq ln n!leq (n+1)ln left(frac n+1eright)+1,.

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


(ne)ne≤n!≤(n+1e)n+1e.displaystyle left(frac neright)^neleq n!leq left(frac n+1eright)^n+1e,.displaystyle left(frac neright)^neleq n!leq left(frac n+1eright)^n+1e,.

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




Comparison of Stirling's approximation with the factorial


For large n we get a better estimate for the number n! using Stirling's approximation:


n!∼2πn(ne)n.displaystyle n!sim sqrt 2pi nleft(frac neright)^n,.displaystyle n!sim sqrt 2pi nleft(frac neright)^n,.

This in fact comes from an asymptotic series for the logarithm, and n factorial lies between this and the next approximation:


2πn(ne)n<n!<2πn(ne)ne112n.displaystyle sqrt 2pi nleft(frac neright)^n<n!<sqrt 2pi nleft(frac neright)^ne^frac 112n,.displaystyle sqrt 2pi nleft(frac neright)^n<n!<sqrt 2pi nleft(frac neright)^ne^frac 112n,.

Another approximation for ln n! is given by Srinivasa Ramanujan (Ramanujan 1988)


ln⁡n!≈nln⁡n−n+ln⁡(n(1+4n(1+2n)))6+ln⁡π2⟹n!≈2πn(ne)n(1+12n+18n2)16.displaystyle beginalignedln n!&approx nln n-n+frac ln Bigl (nbigl (1+4n(1+2n)bigr )Bigr )6+frac ln pi 2\[6px]Longrightarrow ;n!&approx sqrt 2pi nleft(frac neright)^nleft(1+frac 12n+frac 18n^2right)^frac 16,.endaligneddisplaystyle beginalignedln n!&approx nln n-n+frac ln Bigl (nbigl (1+4n(1+2n)bigr )Bigr )6+frac ln pi 2\[6px]Longrightarrow ;n!&approx sqrt 2pi nleft(frac neright)^nleft(1+frac 12n+frac 18n^2right)^frac 16,.endaligned

Both this and the previous approximation give a relative error on the order of 1/n3, but Ramanujan's is about four times more accurate. However, if we use two correction terms (as in Ramanujan's approximation) the relative error will be of order 1/n5:


n!≈2πn(ne)nexp⁡(112n−1360n3).displaystyle n!approx sqrt 2pi nleft(frac neright)^nexp left(frac 112n-frac 1360n^3right),.displaystyle n!approx sqrt 2pi nleft(frac neright)^nexp left(frac 112n-frac 1360n^3right),.


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 7005249999000000000♠249999, and up to 7007200000000000000♠20000000! 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


(n−1)!≡0(modn).displaystyle (n-1)!equiv 0pmod n.displaystyle (n-1)!equiv 0pmod n.

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


(p−1)!≡−1(modp)displaystyle (p-1)!equiv -1pmod pdisplaystyle (p-1)!equiv -1pmod p

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


∑i=1∞⌊npi⌋displaystyle sum _i=1^infty leftlfloor frac np^irightrfloor displaystyle sum _i=1^infty leftlfloor frac np^irightrfloor

or, equivalently,


n−sp(n)p−1,displaystyle frac n-s_p(n)p-1,displaystyle frac n-s_p(n)p-1,

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 divisible by a prime 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:


∑n=0∞1n!=11+11+12+16+124+1120+⋯=e.displaystyle sum _n=0^infty frac 1n!=frac 11+frac 11+frac 12+frac 16+frac 124+frac 1120+cdots =e,.displaystyle sum _n=0^infty frac 1n!=frac 11+frac 11+frac 12+frac 16+frac 124+frac 1120+cdots =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:


∑n=0∞1(n+2)n!=12+13+18+130+1144+⋯=1.displaystyle sum _n=0^infty frac 1(n+2)n!=frac 12+frac 13+frac 18+frac 130+frac 1144+cdots =1,.displaystyle sum _n=0^infty frac 1(n+2)n!=frac 12+frac 13+frac 18+frac 130+frac 1144+cdots =1,.

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]



Extension of factorial to non-integer values of argument



The gamma and pi functions




The gamma function interpolates the factorial function to non-integer values. The main clue is the recurrence relation generalized to a continuous domain.



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


Γ(z)=∫0∞tz−1e−tdt.displaystyle Gamma (z)=int _0^infty t^z-1e^-t,dt,.displaystyle Gamma (z)=int _0^infty t^z-1e^-t,dt,.

Its relation to the factorial is that for any natural number n


n!=Γ(n+1).displaystyle n!=Gamma (n+1),.displaystyle n!=Gamma (n+1),.

Euler's original formula for the gamma function was


Γ(z)=limn→∞nzn!∏k=0n(z+k).displaystyle Gamma (z)=lim _nto infty frac n^zn!displaystyle prod _k=0^n(z+k).displaystyle Gamma (z)=lim _nto infty frac n^zn!displaystyle prod _k=0^n(z+k).

Another function that also "fills in" the values of the factorial (but with no shift in the argument), originally introduced by Carl Friedrich Gauss, that is sometimes used, is called the pi function, denoted Π(z) for real numbers z ≥ 0. It is defined by


Π(z)=∫0∞tze−tdt.displaystyle Pi (z)=int _0^infty t^ze^-t,dt,.displaystyle Pi (z)=int _0^infty t^ze^-t,dt,.

In terms of the gamma function, the pi function is


Π(z)=Γ(z+1).displaystyle Pi (z)=Gamma (z+1),.Pi (z)=Gamma (z+1),.


The factorial function, generalized to all real numbers except negative integers. For example, 0! = 1! = 1, (−1/2)! = π, 1/2! = π/2.


The pi function properly extends the factorial in that


n!=Π(n) for n∈N.displaystyle n!=Pi (n)quad text for nin mathbf N ,.displaystyle n!=Pi (n)quad text for nin mathbf N ,.

In addition to this, the pi function satisfies the same recurrence as factorials do, but at every complex value z where it is defined


Π(z)=zΠ(z−1).displaystyle Pi (z)=zPi (z-1),.Pi (z)=zPi (z-1),.

In fact, this is no longer a recurrence relation but a functional equation. Expressed in terms of the gamma function this functional equation takes the form


Γ(n+1)=nΓ(n).displaystyle Gamma (n+1)=nGamma (n),.Gamma (n+1)=nGamma (n),.

Since the factorial is extended by the pi function, for every complex value z where it is defined, we can write:


z!=Π(z)displaystyle z!=Pi (z)displaystyle z!=Pi (z)

The values of these functions at half-integer values is therefore determined by a single one of them; one has


Γ(12)=(−12)!=Π(−12)=π,displaystyle Gamma left(frac 12right)=left(-frac 12right)!=Pi left(-frac 12right)=sqrt pi ,,displaystyle Gamma left(frac 12right)=left(-frac 12right)!=Pi left(-frac 12right)=sqrt pi ,,

from which it follows that for nN,


Γ(12+n)=(−12+n)!=Π(−12+n)=π∏k=1n2k−12=(2n)!4nn!π=(2n−1)!22n−1(n−1)!π.displaystyle beginaligned&Gamma left(frac 12+nright)=left(-frac 12+nright)!=Pi left(-frac 12+nright)\[5pt]=&sqrt pi prod _k=1^nfrac 2k-12=frac (2n)!4^nn!sqrt pi =frac (2n-1)!2^2n-1(n-1)!sqrt pi ,.endaligneddisplaystyle beginaligned&Gamma left(frac 12+nright)=left(-frac 12+nright)!=Pi left(-frac 12+nright)\[5pt]=&sqrt pi prod _k=1^nfrac 2k-12=frac (2n)!4^nn!sqrt pi =frac (2n-1)!2^2n-1(n-1)!sqrt pi ,.endaligned

For example,


Γ(92)=72!=Π(72)=12⋅32⋅52⋅72π=8!444!π=7!273!π=10516π≈11.631728…displaystyle Gamma left(frac 92right)=frac 72!=Pi left(frac 72right)=frac 12cdot frac 32cdot frac 52cdot frac 72sqrt pi =frac 8!4^44!sqrt pi =frac 7!2^73!sqrt pi =frac 10516sqrt pi approx 11.631,728ldots displaystyle Gamma left(frac 92right)=frac 72!=Pi left(frac 72right)=frac 12cdot frac 32cdot frac 52cdot frac 72sqrt pi =frac 8!4^44!sqrt pi =frac 7!2^73!sqrt pi =frac 10516sqrt pi approx 11.631,728ldots

It also follows that for nN,


Γ(12−n)=(−12−n)!=Π(−12−n)=π∏k=1n21−2k=(−4)nn!(2n)!π.displaystyle Gamma left(frac 12-nright)=left(-frac 12-nright)!=Pi left(-frac 12-nright)=sqrt pi prod _k=1^nfrac 21-2k=frac left(-4right)^nn!(2n)!sqrt pi ,.displaystyle Gamma left(frac 12-nright)=left(-frac 12-nright)!=Pi left(-frac 12-nright)=sqrt pi prod _k=1^nfrac 21-2k=frac left(-4right)^nn!(2n)!sqrt pi ,.

For example,


Γ(−52)=(−72)!=Π(−72)=2−1⋅2−3⋅2−5π=(−4)33!6!π=−815π≈−0.945308…displaystyle Gamma left(-frac 52right)=left(-frac 72right)!=Pi left(-frac 72right)=frac 2-1cdot frac 2-3cdot frac 2-5sqrt pi =frac left(-4right)^33!6!sqrt pi =-frac 815sqrt pi approx -0.945,308ldots displaystyle Gamma left(-frac 52right)=left(-frac 72right)!=Pi left(-frac 72right)=frac 2-1cdot frac 2-3cdot frac 2-5sqrt pi =frac left(-4right)^33!6!sqrt pi =-frac 815sqrt pi approx -0.945,308ldots

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:


n!=Π(n)=∏k=1∞(k+1k)nkn+k=[(21)n1n+1][(32)n2n+2][(43)n3n+3]⋯displaystyle beginalignedn!=Pi (n)&=prod _k=1^infty left(frac k+1kright)^n!!frac kn+k\&=left[left(frac 21right)^nfrac 1n+1right]left[left(frac 32right)^nfrac 2n+2right]left[left(frac 43right)^nfrac 3n+3right]cdots endaligneddisplaystyle beginalignedn!=Pi (n)&=prod _k=1^infty left(frac k+1kright)^n!!frac kn+k\&=left[left(frac 21right)^nfrac 1n+1right]left[left(frac 32right)^nfrac 2n+2right]left[left(frac 43right)^nfrac 3n+3right]cdots endaligned

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


Vn=πn2Γ(n2+1)Rn.displaystyle V_n=frac pi ^frac n2Gamma left(frac n2+1right)R^n,.displaystyle V_n=frac pi ^frac n2Gamma left(frac n2+1right)R^n,.


Factorial at the complex plane




Amplitude and phase of factorial of complex argument


Representation through the gamma function allows evaluation of factorial of complex argument. Equilines of amplitude and phase of factorial are shown in figure. Let


f=ρeiφ=(x+iy)!=Γ(x+iy+1).displaystyle f=rho e^ivarphi =(x+rm iy)!=Gamma (x+iy+1),.displaystyle f=rho e^ivarphi =(x+rm iy)!=Gamma (x+iy+1),.

Several levels of constant modulus (amplitude) ρ and 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 argument.


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


z!=∑n=0∞gnzn.displaystyle z!=sum _n=0^infty g_nz^n,.displaystyle z!=sum _n=0^infty g_nz^n,.

The first coefficients of this expansion are


















n

gn
approximation
0
1
1
1

γ

3000422784335100000♠−0.5772156649
2

π2/12 + γ2/2

6999989055995500000♠0.9890559955
3

ζ(3)/3π2/12γ3/6

3000092520924000000♠−0.9074790760

where γ is the Euler–Mascheroni constant and ζ(z) 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).[citation needed] Writing z! = eP(z) where P(z) is


P(z)=p(z)+ln⁡2π2−z+(z+12)ln⁡(z),displaystyle P(z)=p(z)+frac ln 2pi 2-z+left(z+frac 12right)ln(z),,displaystyle P(z)=p(z)+frac ln 2pi 2-z+left(z+frac 12right)ln(z),,

Stieltjes gave a continued fraction for p(z):


p(z)=a0z+a1z+a2z+a3z+⋱displaystyle p(z)=cfrac a_0z+cfrac a_1z+cfrac a_2z+cfrac a_3z+ddots p(z)=cfrac a_0z+cfrac a_1z+cfrac a_2z+cfrac a_3z+ddots

The first few coefficients an are[21]



















n

an
0

1/12
1

1/30
2

53/210
3

195/371
4

7004229990000000000♠22999/7004227370000000000♠22737
5

7007299445230000000♠29944523/7007197331420000000♠19733142
6

7011109535241009000♠109535241009/7010482642754620000♠48264275462

There is a misconception that ln z! = P(z) or ln Γ(z + 1) = P(z) for any complex z ≠ 0.[citation needed] 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;[citation needed] 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:


(n−1)!=n!n.displaystyle (n-1)!=frac n!n.(n-1)!=frac n!n.

Note, however, that 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 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,


(2k−1)!!=∏i=1k(2i−1)=(2k)!2kk!=2kPk2k=(2k)k_2k.displaystyle (2k-1)!!=prod _i=1^k(2i-1)=frac (2k)!2^kk!=frac _2kP_k2^k=frac left(2kright)^underline k2^k,.displaystyle (2k-1)!!=prod _i=1^k(2i-1)=frac (2k)!2^kk!=frac _2kP_k2^k=frac left(2kright)^underline k2^k,.

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, 7004103950000000000♠10395, 7005135135000000000♠135135,... (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


n!(k)={nif 0<n≤kn((n−k)!(k))if n>k.displaystyle n!^(k)=begincasesn&textif 0<nleq k\nleft((n-k)!^(k)right)&textif n>kendcases.displaystyle n!^(k)=begincasesn&textif 0<nleq k\nleft((n-k)!^(k)right)&textif n>kendcases.

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


n!(k)=1if −k<n≤0displaystyle n!^(k)=1quad textif -k<nleq 0displaystyle n!^(k)=1quad textif -k<nleq 0

For sufficiently large n ≥ 1, the ordinary single factorial function is expanded through the multifactorial functions as follows:


n!=n!!⋅(n−1)!!,n≥1=n!!!⋅(n−1)!!!⋅(n−2)!!!,n≥2=∏i=0k−1(n−i)!(k), for k∈Z+,n≥k−1.displaystyle beginalignedn!&=n!!cdot (n-1)!!,,&n&geq 1\[5px]&=n!!!cdot (n-1)!!!cdot (n-2)!!!,,&n&geq 2\[5px]&=prod _i=0^k-1(n-i)!^(k),quad text for kin mathbb Z ^+,,&n&geq k-1,.endaligneddisplaystyle beginalignedn!&=n!!cdot (n-1)!!,,&n&geq 1\[5px]&=n!!!cdot (n-1)!!!cdot (n-2)!!!,,&n&geq 2\[5px]&=prod _i=0^k-1(n-i)!^(k),quad text for kin mathbb Z ^+,,&n&geq k-1,.endaligned

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 (sequence A002110 in the OEIS) is similar to the factorial, but with the product taken only over the prime numbers.



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


sf⁡(4)=1!×2!×3!×4!=288.displaystyle operatorname sf (4)=1!times 2!times 3!times 4!=288,.displaystyle operatorname sf (4)=1!times 2!times 3!times 4!=288,.

In general


sf⁡(n)=∏k=1nk!=∏k=1nkn−k+1=1n⋅2n−1⋅3n−2⋯(n−1)2⋅n1.displaystyle beginalignedoperatorname sf (n)=prod _k=1^nk!&=prod _k=1^nk^n-k+1\&=1^ncdot 2^n-1cdot 3^n-2cdots (n-1)^2cdot n^1,.endaligneddisplaystyle beginalignedoperatorname sf (n)=prod _k=1^nk!&=prod _k=1^nk^n-k+1\&=1^ncdot 2^n-1cdot 3^n-2cdots (n-1)^2cdot n^1,.endaligned

Equivalently, the superfactorial is given by the formula


sf⁡(n)=∏0≤i<j≤n(j−i)displaystyle operatorname sf (n)=prod _0leq i<jleq n(j-i)displaystyle operatorname sf (n)=prod _0leq i<jleq n(j-i)

which is the determinant of a Vandermonde matrix.


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


1, 1, 2, 12, 288, 7004345600000000000♠34560, 7007248832000000000♠24883200, 7011125411328000000♠125411328000,... (sequence A000178 in the OEIS)

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


sfk⁡(n)={nif k=0∏r=1nsfk−1⁡(r)if k≥1displaystyle operatorname sf _k(n)=begincasesn&textif k=0\prod _r=1^noperatorname sf _k-1(r)&textif kgeq 1endcasesdisplaystyle operatorname sf _k(n)=begincasesn&textif k=0\prod _r=1^noperatorname sf _k-1(r)&textif kgeq 1endcases

The 2-superfactorials of n are


1, 1, 2, 24, 7003691200000000000♠6912, 7008238878720000000♠238878720, 7015594406696550400♠5944066965504000, 7026745453331864786♠745453331864786829312000000,... (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


n$≡n!n!⋅⋅⋅n!⏟n! copies of n!.displaystyle n$equiv beginmatrixunderbrace n!^n!^cdot ^cdot ^cdot ^n! \n!mbox copies of n!endmatrix.displaystyle n$equiv beginmatrixunderbrace n!^n!^cdot ^cdot ^cdot ^n! \n!mbox copies of n!endmatrix.

This sequence of superfactorials starts


1$=1,2$=22=4,3$=666666.displaystyle beginaligned1$&=1,,\2$&=2^2=4,,\3$&=6^6^6^6^6^6,.endaligneddisplaystyle beginaligned1$&=1,,\2$&=2^2=4,,\3$&=6^6^6^6^6^6,.endaligned

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


This operation may also be expressed as the tetration


n$=n!(n!),displaystyle n$=^n!(n!),,displaystyle n$=^n!(n!),,

or using Knuth's up-arrow notation as


n$=(n!)↑↑(n!).displaystyle n$=(n!)uparrow uparrow (n!),.displaystyle n$=(n!)uparrow uparrow (n!),.


Hyperfactorial


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


H(n)=∏k=1nkk=11⋅22⋅33⋯(n−1)n−1⋅nn.displaystyle beginalignedH(n)&=prod _k=1^nk^k\&=1^1cdot 2^2cdot 3^3cdots (n-1)^n-1cdot n^n.endaligneddisplaystyle beginalignedH(n)&=prod _k=1^nk^k\&=1^1cdot 2^2cdot 3^3cdots (n-1)^n-1cdot n^n.endaligned

For n = 1, 2, 3, 4,... the values of H(n) are 1, 4, 108, 7004276480000000000♠27648,... (sequence A002109 in the OEIS).


The asymptotic growth rate is


H(n)∼An6n2+6n+112e−n24displaystyle H(n)sim An^frac 6n^2+6n+112e^-frac n^24displaystyle H(n)sim An^frac 6n^2+6n+112e^-frac n^24

where A = 1.2824... is the Glaisher–Kinkelin constant.[27]H(14) ≈ 7099184740000000000♠1.8474×1099 is already almost equal to a googol, and H(15) ≈ 7116808960000000000♠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

  • Factorial prime

  • 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 – via ScienceDirect..mw-parser-output cite.citationfont-style:inherit.mw-parser-output .citation qquotes:"""""""'""'".mw-parser-output .citation .cs1-lock-free abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .citation .cs1-lock-subscription abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registrationcolor:#555.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration spanborder-bottom:1px dotted;cursor:help.mw-parser-output .cs1-ws-icon abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center.mw-parser-output code.cs1-codecolor:inherit;background:inherit;border:inherit;padding:inherit.mw-parser-output .cs1-hidden-errordisplay:none;font-size:100%.mw-parser-output .cs1-visible-errorfont-size:100%.mw-parser-output .cs1-maintdisplay:none;color:#33aa33;margin-left:0.3em.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-formatfont-size:95%.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-leftpadding-left:0.2em.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-rightpadding-right:0.2em


  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. ^ ab Conway, John H.; Guy, Richard (1998-03-16). The Book of Numbers. Springer Science & Business Media. ISBN 9780387979939.


  8. ^ ab 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. 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: 376–380.


  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. ^ ab 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, 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


.mw-parser-output .refbeginfont-size:90%;margin-bottom:0.5em.mw-parser-output .refbegin-hanging-indents>ullist-style-type:none;margin-left:0.mw-parser-output .refbegin-hanging-indents>ul>li,.mw-parser-output .refbegin-hanging-indents>dl>ddmargin-left:0;padding-left:3.2em;text-indent:-3.2em;list-style:none.mw-parser-output .refbegin-100font-size:100%


  • 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





  • 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.








Popular posts from this blog

How to check contact read email or not when send email to Individual?

Bahrain

Postfix configuration issue with fips on centos 7; mailgun relay