Gamma function









The gamma function along part of the real axis


In mathematics, the gamma function (represented by Γ, the capital Greek alphabet letter gamma) is one of a number of extensions of the factorial function with its argument shifted down by 1, to real and complex numbers. Derived by Daniel Bernoulli, if n is a positive integer,


Γ(n)=(n−1)!{displaystyle Gamma (n)=(n-1)!}Gamma (n)=(n-1)!

Although other extensions do exist, this particular definition is the most popular and useful. The gamma function is defined for all complex numbers except the non-positive integers. For complex numbers with a positive real part, it is defined via a convergent improper integral:


Γ(z)=∫0∞xz−1e−xdx{displaystyle Gamma (z)=int _{0}^{infty }x^{z-1}e^{-x},dx}{displaystyle Gamma (z)=int _{0}^{infty }x^{z-1}e^{-x},dx}

This integral function is extended by analytic continuation to all complex numbers except the non-positive integers (where the function has simple poles), yielding the meromorphic function we call the gamma function. It has no zeroes, so the reciprocal gamma function 1/Γ(z) is a holomorphic function. In fact the gamma function corresponds to the Mellin transform of the negative exponential function:


Γ(z)={Me−x}(z){displaystyle Gamma (z)={{mathcal {M}}e^{-x}}(z)}{displaystyle Gamma (z)={{mathcal {M}}e^{-x}}(z)}

The gamma function is a component in various probability-distribution functions, and as such it is applicable in the fields of probability and statistics, as well as combinatorics.




Contents






  • 1 Motivation


  • 2 Definition


    • 2.1 Main definition


    • 2.2 Alternative definitions


      • 2.2.1 Euler's definition as an infinite product


      • 2.2.2 Weierstrass's definition


      • 2.2.3 In terms of generalized Laguerre polynomials






  • 3 Properties


    • 3.1 General


    • 3.2 Inequalities


    • 3.3 Stirling's formula


    • 3.4 Residues


    • 3.5 Minima


    • 3.6 Integral representations


    • 3.7 Fourier series expansion


    • 3.8 Raabe's formula


    • 3.9 Pi function


    • 3.10 Relation to other functions


    • 3.11 Particular values




  • 4 The log-gamma function


    • 4.1 Properties


    • 4.2 Integration over log-gamma




  • 5 Approximations


  • 6 Applications


    • 6.1 Integration problems


    • 6.2 Calculating products


    • 6.3 Analytic number theory




  • 7 History


    • 7.1 18th century: Euler and Stirling


    • 7.2 19th century: Gauss, Weierstrass and Legendre


    • 7.3 19th–20th centuries: characterizing the gamma function


    • 7.4 Reference tables and software




  • 8 See also


  • 9 Notes


  • 10 Further reading


  • 11 External links





Motivation




The gamma function interpolates the factorial function to non-integer values.


The gamma function can be seen as a solution to the following interpolation problem:


"Find a smooth curve that connects the points (x, y) given by y = (x − 1)! at the positive integer values for x."

A plot of the first few factorials makes clear that such a curve can be drawn, but it would be preferable to have a formula that precisely describes the curve, in which the number of operations does not depend on the size of x. The simple formula for the factorial, x! = 1 × 2 × … × x, cannot be used directly for fractional values of x since it is only valid when x is a natural number (i.e., a positive integer). There are, relatively speaking, no such simple solutions for factorials; no finite combination of sums, products, powers, exponential functions, or logarithms will suffice to express x!; but it is possible to find a general formula for factorials using tools such as integrals and limits from calculus. A good solution to this is the gamma function.[1]


There are infinitely many continuous extensions of the factorial to non-integers: infinitely many curves can be drawn through any set of isolated points. The gamma function is the most useful solution in practice, being analytic (except at the non-positive integers), and it can be characterized[ambiguous] in several ways. However, it is not the only analytic function which extends the factorial, as adding to it any analytic function which is zero on the positive integers, such as k sin mπx, will give another function with that property.[1]




The gamma function, Γ(z) in blue, plotted along with Γ(z) + sin(πz) in green. Notice the intersection at positive integers, both are valid analytic continuations of the factorials to the non-integers


A more restrictive property than satisfying the above interpolation is to satisfy the recurrence relation defining a translated version of the factorial function,



f(1)=1,{displaystyle f(1)=1,}{displaystyle f(1)=1,}

f(x+1)=xf(x),{displaystyle f(x+1)=xf(x),}{displaystyle f(x+1)=xf(x),}


for x equal to any positive real number. But this would allow for multiplication by any periodic analytic function which evaluates to one on the positive integers, such as ek sin mπx. There's a final way to solve all this ambiguity: Bohr–Mollerup theorem states that when the condition that f be logarithmically convex (or "super-convex"[2]) is added, it uniquely determines f for positive, real inputs. From there, the gamma function can be extended to all real and complex values (except the negative integers and zero) by using the unique analytic continuation of f.[3] Also see Euler's infinite product definition below where the properties f(1) = 1 and f(x+1) = x f(x) together with the asymptotic requirement that limn→+∞ (n−1)! nx / f(n+x) = 1 uniquely define the same function.



Definition



Main definition


The notation Γ(z) is due to Legendre.[1] If the real part of the complex number z is positive (Re(z) > 0), then the integral


Γ(z)=∫0∞xz−1e−xdx{displaystyle Gamma (z)=int _{0}^{infty }x^{z-1}e^{-x},dx}{displaystyle Gamma (z)=int _{0}^{infty }x^{z-1}e^{-x},dx}

converges absolutely, and is known as the Euler integral of the second kind (the Euler integral of the first kind defines the beta function).[1] Using integration by parts, one sees that:


Γ(z+1)=∫0∞xze−xdx=[−xze−x]0∞+∫0∞zxz−1e−xdx=limx→(−xze−x)−(0e−0)+z∫0∞xz−1e−xdx{displaystyle {begin{aligned}Gamma (z+1)&=int _{0}^{infty }x^{z}e^{-x},dx\[4pt]&={Big [}-x^{z}e^{-x}{Big ]}_{0}^{infty }+int _{0}^{infty }zx^{z-1}e^{-x},dx\[4pt]&=lim _{xto infty }(-x^{z}e^{-x})-(0e^{-0})+zint _{0}^{infty }x^{z-1}e^{-x},dxend{aligned}}}{displaystyle {begin{aligned}Gamma (z+1)&=int _{0}^{infty }x^{z}e^{-x},dx\[4pt]&={Big [}-x^{z}e^{-x}{Big ]}_{0}^{infty }+int _{0}^{infty }zx^{z-1}e^{-x},dx\[4pt]&=lim _{xto infty }(-x^{z}e^{-x})-(0e^{-0})+zint _{0}^{infty }x^{z-1}e^{-x},dxend{aligned}}}

Recognizing that xze−x→0{displaystyle -x^{z}e^{-x}to 0}{displaystyle -x^{z}e^{-x}to 0} as x→,{displaystyle xto infty ,}{displaystyle xto infty ,}


Γ(z+1)=z∫0∞xz−1e−xdx=zΓ(z){displaystyle {begin{aligned}Gamma (z+1)&=zint _{0}^{infty }x^{z-1}e^{-x},dx\[6pt]&=zGamma (z)end{aligned}}}{displaystyle {begin{aligned}Gamma (z+1)&=zint _{0}^{infty }x^{z-1}e^{-x},dx\[6pt]&=zGamma (z)end{aligned}}}

We can calculate Γ(1):{displaystyle Gamma (1){text{:}}}{displaystyle Gamma (1){text{:}}}


Γ(1)=∫0∞x1−1e−xdx=[−e−x]0∞=limx→(−e−x)−(−e−0)=0−(−1)=1{displaystyle {begin{aligned}Gamma (1)&=int _{0}^{infty }x^{1-1}e^{-x},dx\[6pt]&={Big [}-e^{-x}{Big ]}_{0}^{infty }\[6pt]&=lim _{xto infty }(-e^{-x})-(-e^{-0})\[6pt]&=0-(-1)\[6pt]&=1end{aligned}}}{displaystyle {begin{aligned}Gamma (1)&=int _{0}^{infty }x^{1-1}e^{-x},dx\[6pt]&={Big [}-e^{-x}{Big ]}_{0}^{infty }\[6pt]&=lim _{xto infty }(-e^{-x})-(-e^{-0})\[6pt]&=0-(-1)\[6pt]&=1end{aligned}}}

Given that Γ(1)=1{displaystyle Gamma (1)=1}Gamma (1)=1 and Γ(n+1)=nΓ(n),{displaystyle Gamma (n+1)=nGamma (n),}{displaystyle Gamma (n+1)=nGamma (n),}


Γ(n)=1⋅2⋅3⋯(n−1)=(n−1)!{displaystyle Gamma (n)=1cdot 2cdot 3cdots (n-1)=(n-1)!}{displaystyle Gamma (n)=1cdot 2cdot 3cdots (n-1)=(n-1)!}

for all positive integers n. This can be seen as an example of proof by induction.


The identity Γ(z)=Γ(z+1)z{displaystyle Gamma (z)={frac {Gamma (z+1)}{z}}}{displaystyle Gamma (z)={frac {Gamma (z+1)}{z}}} can be used (or, yielding the same result, analytic continuation can be used) to uniquely extend the integral formulation for Γ(z) to a meromorphic function defined for all complex numbers z, except integers less than or equal to zero.[1] It is this extended version that is commonly referred to as the gamma function.[1]



Alternative definitions



Euler's definition as an infinite product


When seeking to approximate z! for a complex number z, it turns out that it is effective to first compute n! for some large integer n, then use that to approximate a value for (n+z)!, and then use the recursion relation m! = m (m−1)! backwards n times, to unwind it to an approximation for z!. Furthermore, this approximation is exact in the limit as n goes to infinity.


Specifically, for a fixed integer m, it is the case that


limn→n!(n+1)m(n+m)!=1,{displaystyle lim _{nto infty }{frac {n!;(n+1)^{m}}{(n+m)!}}=1,,}{displaystyle lim _{nto infty }{frac {n!;(n+1)^{m}}{(n+m)!}}=1,,}

and we can ask that the same formula is obeyed when the arbitrary integer m is replaced by an arbitrary complex number z


limn→n!(n+1)z(n+z)!=1.{displaystyle lim _{nto infty }{frac {n!;(n+1)^{z}}{(n+z)!}}=1,.}{displaystyle lim _{nto infty }{frac {n!;(n+1)^{z}}{(n+z)!}}=1,.}

Multiplying both sides by z! gives


z!=limn→n!z!(n+z)!(n+1)z=limn→1⋯n(1+z)⋯(n+z)(n+1)z=limn→1⋯n(1+z)⋯(n+z)[(1+11)(1+12)⋯(1+1n)]z=∏n=1∞[11+zn(1+1n)z].{displaystyle {begin{aligned}z!&=lim _{nto infty }n!{frac {z!}{(n+z)!}}(n+1)^{z}\[8pt]&=lim _{nto infty }{frac {1cdots n}{(1+z)cdots (n+z)}}(n+1)^{z}\[8pt]&=lim _{nto infty }{frac {1cdots n}{(1+z)cdots (n+z)}}left[left(1+{frac {1}{1}}right)left(1+{frac {1}{2}}right)cdots left(1+{frac {1}{n}}right)right]^{z}\[8pt]&=prod _{n=1}^{infty }left[{frac {1}{1+{frac {z}{n}}}}left(1+{frac {1}{n}}right)^{z}right].end{aligned}}}{displaystyle {begin{aligned}z!&=lim _{nto infty }n!{frac {z!}{(n+z)!}}(n+1)^{z}\[8pt]&=lim _{nto infty }{frac {1cdots n}{(1+z)cdots (n+z)}}(n+1)^{z}\[8pt]&=lim _{nto infty }{frac {1cdots n}{(1+z)cdots (n+z)}}left[left(1+{frac {1}{1}}right)left(1+{frac {1}{2}}right)cdots left(1+{frac {1}{n}}right)right]^{z}\[8pt]&=prod _{n=1}^{infty }left[{frac {1}{1+{frac {z}{n}}}}left(1+{frac {1}{n}}right)^{z}right].end{aligned}}}

This infinite product formula converges for all complex numbers z except the negative integers, which fail because trying to use the recursion relation m! = m (m − 1)! backwards through the value m = 0 involves a division by zero.


Similarly for the gamma function, the definition as an infinite product due to Euler is valid for all complex numbers z{displaystyle z}z except the non-positive integers:


Γ(z)=1z∏n=1∞(1+1n)z1+zn.{displaystyle Gamma (z)={frac {1}{z}}prod _{n=1}^{infty }{frac {left(1+{frac {1}{n}}right)^{z}}{1+{frac {z}{n}}}},.}{displaystyle Gamma (z)={frac {1}{z}}prod _{n=1}^{infty }{frac {left(1+{frac {1}{n}}right)^{z}}{1+{frac {z}{n}}}},.}

By this construction, the gamma function is the unique function that simultaneously satisfies Γ(1)=1{displaystyle Gamma (1)=1}Gamma (1)=1, Γ(z+1)=zΓ(z){displaystyle Gamma (z+1)=zGamma (z)}Gamma (z+1)=zGamma (z) for all complex numbers z{displaystyle z}z except the non-positive integers, and limn→Γ(n+z)(n−1)!nz=1{displaystyle lim _{nto infty }{frac {Gamma (n+z)}{(n-1)!;n^{z}}}=1}{displaystyle lim _{nto infty }{frac {Gamma (n+z)}{(n-1)!;n^{z}}}=1} for all complex numbers z{displaystyle z}z.[1]



Weierstrass's definition


The definition for the gamma function due to Weierstrass is also valid for all complex numbers z except the non-positive integers:


Γ(z)=e−γzz∏n=1∞(1+zn)−1ez/n{displaystyle Gamma (z)={frac {e^{-gamma z}}{z}}prod _{n=1}^{infty }left(1+{frac {z}{n}}right)^{-1}e^{z/n}}{displaystyle Gamma (z)={frac {e^{-gamma z}}{z}}prod _{n=1}^{infty }left(1+{frac {z}{n}}right)^{-1}e^{z/n}}

where γ0.577216{displaystyle gamma approx 0.577216}{displaystyle gamma approx 0.577216} is the Euler–Mascheroni constant.[1]



In terms of generalized Laguerre polynomials


A parameterization of the incomplete gamma function in terms of generalized Laguerre polynomials is


Γ(z,x)=xze−x∑n=0∞Ln(z)(x)n+1,{displaystyle Gamma (z,x)=x^{z}e^{-x}sum _{n=0}^{infty }{frac {L_{n}^{(z)}(x)}{n+1}},}{displaystyle Gamma (z,x)=x^{z}e^{-x}sum _{n=0}^{infty }{frac {L_{n}^{(z)}(x)}{n+1}},}

which converges for Re⁡(z)>−1{displaystyle operatorname {Re} (z)>-1}{displaystyle operatorname {Re} (z)>-1} and x>0{displaystyle x>0}x>0.[4]


A somewhat unusual parametrization of the gamma function in terms of Laguerre polynomials is given by



Γ(z)=tz∑n=0∞Ln(z)(t)z+n,{displaystyle Gamma (z)=t^{z}sum _{n=0}^{infty }{frac {L_{n}^{(z)}(t)}{z+n}},,}{displaystyle Gamma (z)=t^{z}sum _{n=0}^{infty }{frac {L_{n}^{(z)}(t)}{z+n}},,} which converges for Re(z) < 1/2.[citation needed]


Properties



General


Other important functional equations for the gamma function are Euler's reflection formula


Γ(1−z)Γ(z)=πsin⁡z),z∉Z{displaystyle Gamma (1-z)Gamma (z)={pi over sin {(pi z)}},qquad znot in mathbb {Z} }{displaystyle Gamma (1-z)Gamma (z)={pi  over sin {(pi z)}},qquad znot in mathbb {Z} }

which implies


Γn)=(−1)n−(−ε(1+ε(n+1−ε),{displaystyle Gamma (varepsilon -n)=(-1)^{n-1};{frac {Gamma (-varepsilon )Gamma (1+varepsilon )}{Gamma (n+1-varepsilon )}},}Gamma (varepsilon -n)=(-1)^{n-1};{frac {Gamma (-varepsilon )Gamma (1+varepsilon )}{Gamma (n+1-varepsilon )}},

and the duplication formula


Γ(z)Γ(z+12)=21−2zπΓ(2z).{displaystyle Gamma (z)Gamma left(z+{tfrac {1}{2}}right)=2^{1-2z};{sqrt {pi }};Gamma (2z).}{displaystyle Gamma (z)Gamma left(z+{tfrac {1}{2}}right)=2^{1-2z};{sqrt {pi }};Gamma (2z).}

The duplication formula is a special case of the multiplication theorem (See [5], Eq. 5.5.6)


k=0m−(z+km)=(2π)m−12m12−mzΓ(mz).{displaystyle prod _{k=0}^{m-1}Gamma left(z+{frac {k}{m}}right)=(2pi )^{frac {m-1}{2}};m^{{frac {1}{2}}-mz};Gamma (mz).}{displaystyle prod _{k=0}^{m-1}Gamma left(z+{frac {k}{m}}right)=(2pi )^{frac {m-1}{2}};m^{{frac {1}{2}}-mz};Gamma (mz).}

A simple but useful property, which can be seen from the limit definition, is:


Γ(z)¯(z¯)⇒Γ(z)Γ(z¯)∈R.{displaystyle {overline {Gamma (z)}}=Gamma ({overline {z}});Rightarrow ;Gamma (z)Gamma ({overline {z}})in mathbb {R} .}{displaystyle {overline {Gamma (z)}}=Gamma ({overline {z}});Rightarrow ;Gamma (z)Gamma ({overline {z}})in mathbb {R} .}

In particular, with z = a + bi, this product is


(a+bi)|2=|Γ(a)|2∏k=0∞11+b2(a+k)2|Γ(bi)|2=πbsinh⁡b)|Γ(12+bi)|2=πcosh⁡b){displaystyle {begin{aligned}|Gamma (a+bi)|^{2}&=|Gamma (a)|^{2}prod _{k=0}^{infty }{frac {1}{1+{frac {b^{2}}{(a+k)^{2}}}}}\[4pt]|Gamma (bi)|^{2}&={frac {pi }{bsinh {(pi b)}}}\[6pt]|Gamma left({tfrac {1}{2}}+biright)|^{2}&={frac {pi }{cosh {(pi b)}}}end{aligned}}}{displaystyle {begin{aligned}|Gamma (a+bi)|^{2}&=|Gamma (a)|^{2}prod _{k=0}^{infty }{frac {1}{1+{frac {b^{2}}{(a+k)^{2}}}}}\[4pt]|Gamma (bi)|^{2}&={frac {pi }{bsinh {(pi b)}}}\[6pt]|Gamma left({tfrac {1}{2}}+biright)|^{2}&={frac {pi }{cosh {(pi b)}}}end{aligned}}}

Perhaps the best-known value of the gamma function at a non-integer argument is


Γ(12)=π,{displaystyle Gamma left({tfrac {1}{2}}right)={sqrt {pi }},}{displaystyle Gamma left({tfrac {1}{2}}right)={sqrt {pi }},}

which can be found by setting z = 1/2 in the reflection or duplication formulas, by using the relation to the beta function given below with x = y = 1/2, or simply by making the substitution u = x in the integral definition of the gamma function, resulting in a Gaussian integral. In general, for non-negative integer values of n we have:


Γ(12+n)=(2n)!4nn!π=(2n−1)!!2nπ=(n−12n)n!πΓ(12−n)=(−4)nn!(2n)!π=(−2)n(2n−1)!!π(−12n)n!{displaystyle {begin{aligned}Gamma left({tfrac {1}{2}}+nright)&={(2n)! over 4^{n}n!}{sqrt {pi }}={frac {(2n-1)!!}{2^{n}}}{sqrt {pi }}={n-{frac {1}{2}} choose n}n!{sqrt {pi }}\[8pt]Gamma left({tfrac {1}{2}}-nright)&={(-4)^{n}n! over (2n)!}{sqrt {pi }}={frac {(-2)^{n}}{(2n-1)!!}}{sqrt {pi }}={frac {sqrt {pi }}{{-{frac {1}{2}} choose n}n!}}end{aligned}}}{displaystyle {begin{aligned}Gamma left({tfrac {1}{2}}+nright)&={(2n)! over 4^{n}n!}{sqrt {pi }}={frac {(2n-1)!!}{2^{n}}}{sqrt {pi }}={n-{frac {1}{2}} choose n}n!{sqrt {pi }}\[8pt]Gamma left({tfrac {1}{2}}-nright)&={(-4)^{n}n! over (2n)!}{sqrt {pi }}={frac {(-2)^{n}}{(2n-1)!!}}{sqrt {pi }}={frac {sqrt {pi }}{{-{frac {1}{2}} choose n}n!}}end{aligned}}}

where n!! denotes the double factorial of n and, when n = 0, n!! = 1. See Particular values of the gamma function for calculated values.


It might be tempting to generalize the result that Γ(1/2) = π by looking for a formula for other individual values Γ(r) where r is rational. However, these numbers are not known to be expressible by themselves in terms of elementary functions. It has been proved that Γ(n + r) is a transcendental number and algebraically independent of π for any integer n and each of the fractions r = 1/6, 1/4, 1/3, 2/3, 3/4, 5/6.[6] In general, when computing values of the gamma function, we must settle for numerical approximations.


Another useful limit for asymptotic approximations is:


limn→Γ(n+α(n)nα=1,αC{displaystyle lim _{nto infty }{frac {Gamma (n+alpha )}{Gamma (n)n^{alpha }}}=1,qquad alpha in mathbb {C} }{displaystyle lim _{nto infty }{frac {Gamma (n+alpha )}{Gamma (n)n^{alpha }}}=1,qquad alpha in mathbb {C} }

The derivatives of the gamma function are described in terms of the polygamma function. For example:


Γ′(z)=Γ(z)ψ0(z).{displaystyle Gamma '(z)=Gamma (z)psi _{0}(z).}Gamma '(z)=Gamma (z)psi _{0}(z).

For a positive integer m the derivative of the gamma function can be calculated as follows (here γ is the Euler–Mascheroni constant):


Γ′(m+1)=m!(−γ+∑k=1m1k).{displaystyle Gamma '(m+1)=m!left(-gamma +sum _{k=1}^{m}{frac {1}{k}}right),.}{displaystyle Gamma '(m+1)=m!left(-gamma +sum _{k=1}^{m}{frac {1}{k}}right),.}

For Re(x) > 0 the nth derivative of the gamma function is:




Derivative of the function Γ(z)


dndxnΓ(x)=∫0∞tx−1e−t(ln⁡t)ndt.{displaystyle {frac {d^{n}}{dx^{n}}}Gamma (x)=int _{0}^{infty }t^{x-1}e^{-t}(ln t)^{n},dt.}{displaystyle {frac {d^{n}}{dx^{n}}}Gamma (x)=int _{0}^{infty }t^{x-1}e^{-t}(ln t)^{n},dt.}

(This can be derived by differentiating the integral form of the gamma function with respect to x, and using the technique of differentiation under the integral sign.)


Using the identity


Γ(n)(1)=(−1)nn!∑πn∏i=1rζ(ai)ki!⋅aiζ(x):={ζ(x)x≠x=1{displaystyle Gamma ^{(n)}(1)=(-1)^{n}n!sum limits _{pi ,vdash ,n},prod _{i=1}^{r}{frac {zeta ^{*}(a_{i})}{k_{i}!cdot a_{i}}}qquad zeta ^{*}(x):={begin{cases}zeta (x)&xneq 1\gamma &x=1end{cases}}}{displaystyle Gamma ^{(n)}(1)=(-1)^{n}n!sum limits _{pi ,vdash ,n},prod _{i=1}^{r}{frac {zeta ^{*}(a_{i})}{k_{i}!cdot a_{i}}}qquad zeta ^{*}(x):={begin{cases}zeta (x)&xneq 1\gamma &x=1end{cases}}}

where ζ(z) is the Riemann zeta function, with partitions[clarification needed]


π=(a1,…,a1⏟k1,…,ar,…,ar⏟kr),{displaystyle pi =(underbrace {a_{1},dots ,a_{1}} _{k_{1}},dots ,underbrace {a_{r},dots ,a_{r}} _{k_{r}}),}{displaystyle pi =(underbrace {a_{1},dots ,a_{1}} _{k_{1}},dots ,underbrace {a_{r},dots ,a_{r}} _{k_{r}}),}

we have in particular


Γ(z)=1z−γ+12(γ2+π26)z−16(γ3+γπ22+2ζ(3))z2+O(z3).{displaystyle Gamma (z)={frac {1}{z}}-gamma +{tfrac {1}{2}}left(gamma ^{2}+{frac {pi ^{2}}{6}}right)z-{tfrac {1}{6}}left(gamma ^{3}+{frac {gamma pi ^{2}}{2}}+2zeta (3)right)z^{2}+O(z^{3}).}{displaystyle Gamma (z)={frac {1}{z}}-gamma +{tfrac {1}{2}}left(gamma ^{2}+{frac {pi ^{2}}{6}}right)z-{tfrac {1}{6}}left(gamma ^{3}+{frac {gamma pi ^{2}}{2}}+2zeta (3)right)z^{2}+O(z^{3}).}


Inequalities


When restricted to the positive real numbers, the gamma function is a strictly logarithmically convex function. This property may be stated in any of the following three equivalent ways:



  • For any two positive real numbers x1 and x2, and for any t ∈ [0, 1],
    Γ(tx1+(1−t)x2)≤Γ(x1)tΓ(x2)1−t.{displaystyle Gamma (tx_{1}+(1-t)x_{2})leq Gamma (x_{1})^{t}Gamma (x_{2})^{1-t}.}{displaystyle Gamma (tx_{1}+(1-t)x_{2})leq Gamma (x_{1})^{t}Gamma (x_{2})^{1-t}.}
    Moreover, the inequality is strict for t ∈ (0, 1).

  • For any two positive real numbers x and y with y > x,
    (y)Γ(x))1y−x>exp⁡′(x)Γ(x)).{displaystyle left({frac {Gamma (y)}{Gamma (x)}}right)^{frac {1}{y-x}}>exp left({frac {Gamma '(x)}{Gamma (x)}}right).}{displaystyle left({frac {Gamma (y)}{Gamma (x)}}right)^{frac {1}{y-x}}>exp left({frac {Gamma '(x)}{Gamma (x)}}right).}


  • For any positive real number x,
    Γ″(x)Γ(x)>Γ′(x).{displaystyle Gamma ''(x)Gamma (x)>Gamma '(x).}{displaystyle Gamma ''(x)Gamma (x)>Gamma '(x).}



The last of these statements is, essentially by definition, the same as the statement that ψ(1)(x)>0{displaystyle psi ^{(1)}(x)>0}{displaystyle psi ^{(1)}(x)>0}, where ψ(1){displaystyle psi ^{(1)}}{displaystyle psi ^{(1)}} is the polygamma function of order 1. To prove the logarithmic convexity of the gamma function, it therefore suffices to observe that ψ(1){displaystyle psi ^{(1)}}{displaystyle psi ^{(1)}} has a series representation which, for positive real x, consists of only positive terms.


Logarithmic convexity and Jensen's inequality together imply, for any positive real numbers x1,…,xn{displaystyle x_{1},ldots ,x_{n}}x_{1},ldots ,x_{n} and a1,…,an{displaystyle a_{1},ldots ,a_{n}}a_1, ldots, a_n,


Γ(a1x1+⋯+anxna1+⋯+an)≤(x1)a1⋯Γ(xn)an)1a1+⋯+an.{displaystyle Gamma left({frac {a_{1}x_{1}+cdots +a_{n}x_{n}}{a_{1}+cdots +a_{n}}}right)leq {bigl (}Gamma (x_{1})^{a_{1}}cdots Gamma (x_{n})^{a_{n}}{bigr )}^{frac {1}{a_{1}+cdots +a_{n}}}.}{displaystyle Gamma left({frac {a_{1}x_{1}+cdots +a_{n}x_{n}}{a_{1}+cdots +a_{n}}}right)leq {bigl (}Gamma (x_{1})^{a_{1}}cdots Gamma (x_{n})^{a_{n}}{bigr )}^{frac {1}{a_{1}+cdots +a_{n}}}.}

There are also bounds on ratios of gamma functions. The best-known is Gautschi's inequality, which says that for any positive real number x and any s ∈ (0, 1),


x1−s<Γ(x+1)Γ(x+s)<(x+1)1−s.{displaystyle x^{1-s}<{frac {Gamma (x+1)}{Gamma (x+s)}}<(x+1)^{1-s}.}{displaystyle x^{1-s}<{frac {Gamma (x+1)}{Gamma (x+s)}}<(x+1)^{1-s}.}


Stirling's formula




Representation of the gamma function in the complex plane. Each point z{displaystyle z}z is colored according to the argument of Γ(z){displaystyle Gamma (z)}Gamma (z). The contour plot of the modulus (z)|{displaystyle |Gamma (z)|}{displaystyle |Gamma (z)|} is also displayed.




The absolute value of the gamma function on the complex plane.


The behavior of Γ(z){displaystyle Gamma (z)}Gamma (z) for an increasing positive variable is simple: it grows quickly — faster than an exponential function. Asymptotically as z→{displaystyle zto infty }{displaystyle zto infty }, the magnitude of the gamma function is given by Stirling's formula


Γ(z+1)∼z(ze)z,{displaystyle Gamma (z+1)sim {sqrt {2pi z}}left({frac {z}{e}}right)^{z},}{displaystyle Gamma (z+1)sim {sqrt {2pi z}}left({frac {z}{e}}right)^{z},}

where the symbol {displaystyle sim }sim means that the ratio of the two sides converges to 1[1] or asymptotically converges.



Residues


The behavior for non-positive z{displaystyle z}z is more intricate. Euler's integral does not converge for z≤0{displaystyle zleq 0}{displaystyle zleq 0}, but the function it defines in the positive complex half-plane has a unique analytic continuation to the negative half-plane. One way to find that analytic continuation is to use Euler's integral for positive arguments and extend the domain to negative numbers by repeated application of the recurrence formula,[1]


Γ(z)=Γ(z+n+1)z(z+1)⋯(z+n),{displaystyle Gamma (z)={frac {Gamma (z+n+1)}{z(z+1)cdots (z+n)}},}{displaystyle Gamma (z)={frac {Gamma (z+n+1)}{z(z+1)cdots (z+n)}},}

choosing n{displaystyle n}n such that z+n{displaystyle z+n}{displaystyle z+n} is positive. The product in the denominator is zero when z{displaystyle z}z equals any of the integers 0,−1,−2,⋯{displaystyle 0,-1,-2,cdots }{displaystyle 0,-1,-2,cdots }. Thus, the gamma function must be undefined at those points to avoid division by zero; it is a meromorphic function with simple poles at the non-positive integers.[1]


We can rewrite this definition as:


(z+n)Γ(z)=Γ(z+n+1)z(z+1)⋯(z+n−1).{displaystyle (z+n)Gamma (z)={frac {Gamma (z+n+1)}{z(z+1)cdots (z+n-1)}}.}{displaystyle (z+n)Gamma (z)={frac {Gamma (z+n+1)}{z(z+1)cdots (z+n-1)}}.}

For a function f{displaystyle f}f of a complex variable z{displaystyle z}z, at a simple pole c{displaystyle c}c, the residue of f{displaystyle f}f is given by:


Res⁡(f,c)=limz→c(z−c)f(z).{displaystyle operatorname {Res} (f,c)=lim _{zto c}(z-c)f(z).}operatorname {Res} (f,c)=lim _{zto c}(z-c)f(z).

When z=−n,{displaystyle z=-n,}{displaystyle z=-n,}


Γ(z+n+1)=Γ(1)=1{displaystyle Gamma (z+n+1)=Gamma (1)=1}{displaystyle Gamma (z+n+1)=Gamma (1)=1}

and


z(z+1)⋯(z+n−1)=(−1)nn!{displaystyle z(z+1)cdots (z+n-1)=(-1)^{n}n!}{displaystyle z(z+1)cdots (z+n-1)=(-1)^{n}n!}

So the residues of the gamma function at those points are:



Res⁡,−n)=(−1)nn!.{displaystyle operatorname {Res} (Gamma ,-n)={frac {(-1)^{n}}{n!}}.}operatorname {Res} (Gamma ,-n)={frac {(-1)^{n}}{n!}}.[7]

The gamma function is non-zero everywhere along the real line, although it comes arbitrarily close to zero as z → −∞. There is in fact no complex number z{displaystyle z}z for which Γ(z)=0{displaystyle Gamma (z)=0}{displaystyle Gamma (z)=0}, and hence the reciprocal gamma function (z){displaystyle {frac {1}{Gamma (z)}}}{displaystyle {frac {1}{Gamma (z)}}} is an entire function, with zeros at z=0,−1,−2,⋯{displaystyle z=0,-1,-2,cdots }{displaystyle z=0,-1,-2,cdots }.[1]



Minima


The gamma function has a local minimum at zmin≈1.46163{displaystyle z_{min }approx 1.46163}{displaystyle z_{min }approx 1.46163} where it attains the value Γ(zmin)≈0.885603{displaystyle Gamma left(z_{min }right)approx 0.885603}{displaystyle Gamma left(z_{min }right)approx 0.885603}. The gamma function must alternate sign between the poles because the product in the forward recurrence contains an odd number of negative factors if the number of poles between z{displaystyle z}z and z+n{displaystyle z+n}{displaystyle z+n} is odd, and an even number if the number of poles is even.[7]



Integral representations


There are many formulas, besides the Euler integral of the second kind, that express the gamma function as an integral. For instance, when the real part of z is positive,[8]


Γ(z)=∫01(log⁡1t)z−1dt.{displaystyle Gamma (z)=int _{0}^{1}left(log {frac {1}{t}}right)^{z-1},dt.}{displaystyle Gamma (z)=int _{0}^{1}left(log {frac {1}{t}}right)^{z-1},dt.}

Binet's first integral formula for the gamma function states that, when the real part of z is positive, then:[9]


log⁡Γ(z)=(z−12)log⁡z−z+12log⁡(2π)+∫0∞(12−1t+1et−1)e−tztdt.{displaystyle log Gamma (z)=left(z-{frac {1}{2}}right)log z-z+{frac {1}{2}}log(2pi )+int _{0}^{infty }left({frac {1}{2}}-{frac {1}{t}}+{frac {1}{e^{t}-1}}right){frac {e^{-tz}}{t}},dt.}{displaystyle log Gamma (z)=left(z-{frac {1}{2}}right)log z-z+{frac {1}{2}}log(2pi )+int _{0}^{infty }left({frac {1}{2}}-{frac {1}{t}}+{frac {1}{e^{t}-1}}right){frac {e^{-tz}}{t}},dt.}

The integral on the right-hand side may be interpreted as a Laplace transform. That is,


log⁡(z)(ez)z2πz)=L(12t−1t2+1t(et−1))(z).{displaystyle log left(Gamma (z)left({frac {e}{z}}right)^{z}{sqrt {2pi z}}right)={mathcal {L}}left({frac {1}{2t}}-{frac {1}{t^{2}}}+{frac {1}{t(e^{t}-1)}}right)(z).}{displaystyle log left(Gamma (z)left({frac {e}{z}}right)^{z}{sqrt {2pi z}}right)={mathcal {L}}left({frac {1}{2t}}-{frac {1}{t^{2}}}+{frac {1}{t(e^{t}-1)}}right)(z).}

Binet's second integral formula states that, again when the real part of z is positive, then:[10]


log⁡Γ(z)=(z−12)log⁡z−z+12log⁡(2π)+2∫0∞arctan⁡(t/z)e2πt−1dt.{displaystyle log Gamma (z)=left(z-{frac {1}{2}}right)log z-z+{frac {1}{2}}log(2pi )+2int _{0}^{infty }{frac {arctan(t/z)}{e^{2pi t}-1}},dt.}{displaystyle log Gamma (z)=left(z-{frac {1}{2}}right)log z-z+{frac {1}{2}}log(2pi )+2int _{0}^{infty }{frac {arctan(t/z)}{e^{2pi t}-1}},dt.}

Let C be a Hankel contour, meaning a path that begins and ends at the point on the Riemann sphere, whose unit tangent vector converges to −1 at the start of the path and to 1 at the end, which has winding number 1 around 0, and which does not cross [0, ∞). Fix a branch of log⁡(−t){displaystyle log(-t)}{displaystyle log(-t)} by taking a branch cut along [0, ∞) and by taking log⁡(−t){displaystyle log(-t)}{displaystyle log(-t)} to be real when t is on the negative real axis. Assume z is not an integer. Then Hankel's formula for the gamma function is:[11]


Γ(z)=−12isin⁡πz∫C(−t)z−1e−tdt,{displaystyle Gamma (z)=-{frac {1}{2isin pi z}}int _{C}(-t)^{z-1}e^{-t},dt,}{displaystyle Gamma (z)=-{frac {1}{2isin pi z}}int _{C}(-t)^{z-1}e^{-t},dt,}

where (−t)z−1{displaystyle (-t)^{z-1}}{displaystyle (-t)^{z-1}} is interpreted as exp⁡((z−1)log⁡(−t)){displaystyle exp((z-1)log(-t))}{displaystyle exp((z-1)log(-t))}. The reflection formula leads to the closely related expression


(z)=i2πC(−t)−ze−tdt,{displaystyle {frac {1}{Gamma (z)}}={frac {i}{2pi }}int _{C}(-t)^{-z}e^{-t},dt,}{displaystyle {frac {1}{Gamma (z)}}={frac {i}{2pi }}int _{C}(-t)^{-z}e^{-t},dt,}

again valid whenever z is not an integer.



Fourier series expansion


The logarithm of the gamma function has the following Fourier series expansion for 0<z<1:{displaystyle 0<z<1:}{displaystyle 0<z<1:}


ln⁡Γ(z)=(12−z)(γ+ln⁡2)+(1−z)ln⁡π12ln⁡sin⁡z)+1πn=1∞ln⁡nnsin⁡(2πnz),{displaystyle ln Gamma (z)=left({frac {1}{2}}-zright)(gamma +ln 2)+(1-z)ln pi -{frac {1}{2}}ln sin(pi z)+{frac {1}{pi }}sum _{n=1}^{infty }{frac {ln n}{n}}sin(2pi nz),}{displaystyle ln Gamma (z)=left({frac {1}{2}}-zright)(gamma +ln 2)+(1-z)ln pi -{frac {1}{2}}ln sin(pi z)+{frac {1}{pi }}sum _{n=1}^{infty }{frac {ln n}{n}}sin(2pi nz),}

which was for a long time attributed to Ernst Kummer, who derived it in 1847.[12][13] However, Iaroslav Blagouchine discovered that Carl Johan Malmsten first derived this series in 1842.[14][15]



Raabe's formula


In 1840 Joseph Ludwig Raabe proved that


aa+1ln⁡Γ(z)dz=12ln⁡+aln⁡a−a,a>0.{displaystyle int _{a}^{a+1}ln Gamma (z),dz={tfrac {1}{2}}ln 2pi +aln a-a,quad a>0.}{displaystyle int _{a}^{a+1}ln Gamma (z),dz={tfrac {1}{2}}ln 2pi +aln a-a,quad a>0.}

In particular, if a = 0 then


01ln⁡Γ(z)dz=12ln⁡.{displaystyle int _{0}^{1}ln Gamma (z),dz={tfrac {1}{2}}ln 2pi .}{displaystyle int _{0}^{1}ln Gamma (z),dz={tfrac {1}{2}}ln 2pi .}


Pi function


An alternative notation which was originally introduced by Gauss and which was sometimes used is the pi function, which in terms of the gamma function is


Π(z)=Γ(z+1)=zΓ(z)=∫0∞e−ttzdt,{displaystyle Pi (z)=Gamma (z+1)=zGamma (z)=int _{0}^{infty }e^{-t}t^{z},dt,}{displaystyle Pi (z)=Gamma (z+1)=zGamma (z)=int _{0}^{infty }e^{-t}t^{z},dt,}

so that Π(n) = n! for every non-negative integer n.


Using the pi function the reflection formula takes on the form


Π(z)Π(−z)=πzsin⁡z)=1sinc⁡(z){displaystyle Pi (z)Pi (-z)={frac {pi z}{sin(pi z)}}={frac {1}{operatorname {sinc} (z)}}}{displaystyle Pi (z)Pi (-z)={frac {pi z}{sin(pi z)}}={frac {1}{operatorname {sinc} (z)}}}

where sinc is the normalized sinc function, while the multiplication theorem takes on the form


Π(zm)Π(z−1m)⋯Π(z−m+1m)=(2π)m−12m−z−12Π(z).{displaystyle Pi left({frac {z}{m}}right),Pi left({frac {z-1}{m}}right)cdots Pi left({frac {z-m+1}{m}}right)=(2pi )^{frac {m-1}{2}}m^{-z-{frac {1}{2}}}Pi (z).}{displaystyle Pi left({frac {z}{m}}right),Pi left({frac {z-1}{m}}right)cdots Pi left({frac {z-m+1}{m}}right)=(2pi )^{frac {m-1}{2}}m^{-z-{frac {1}{2}}}Pi (z).}

We also sometimes find


π(z)=1Π(z),{displaystyle pi (z)={frac {1}{Pi (z)}},}pi (z)={frac {1}{Pi (z)}},

which is an entire function, defined for every complex number, just like the reciprocal gamma function. That π(z){displaystyle pi left(zright)}{displaystyle pi left(zright)} is entire entails it has no poles, so Π(z){displaystyle Pi left(zright)}{displaystyle Pi left(zright)}, like Γ(z){displaystyle Gamma left(zright)}Gamma left(zright), has no zeros.


The volume of an n-ellipsoid with radii r1, …, rn can be expressed as


Vn(r1,…,rn)=πn2Π(n2)∏k=1nrk.{displaystyle V_{n}(r_{1},dotsc ,r_{n})={frac {pi ^{frac {n}{2}}}{Pi left({frac {n}{2}}right)}}prod _{k=1}^{n}r_{k}.}{displaystyle V_{n}(r_{1},dotsc ,r_{n})={frac {pi ^{frac {n}{2}}}{Pi left({frac {n}{2}}right)}}prod _{k=1}^{n}r_{k}.}


Relation to other functions



  • In the first integral above, which defines the gamma function, the limits of integration are fixed. The upper and lower incomplete gamma functions are the functions obtained by allowing the lower or upper (respectively) limit of integration to vary.

  • The gamma function is related to the beta function by the formula


B(x,y)=Γ(x)Γ(y)Γ(x+y).{displaystyle mathrm {B} (x,y)={frac {Gamma (x);Gamma (y)}{Gamma (x+y)}}.}mathrm {B} (x,y)={frac {Gamma (x);Gamma (y)}{Gamma (x+y)}}.


  • The logarithmic derivative of the gamma function is called the digamma function; higher derivatives are the polygamma functions.

  • The analog of the gamma function over a finite field or a finite ring is the Gaussian sums, a type of exponential sum.

  • The reciprocal gamma function is an entire function and has been studied as a specific topic.

  • The gamma function also shows up in an important relation with the Riemann zeta function, ζ(z).



πz2Γ(z2)ζ(z)=π1−z2Γ(1−z2)ζ(1−z).{displaystyle pi ^{-{frac {z}{2}}};Gamma left({frac {z}{2}}right)zeta (z)=pi ^{-{frac {1-z}{2}}};Gamma left({frac {1-z}{2}}right);zeta (1-z).}pi ^{-{frac {z}{2}}};Gamma left({frac {z}{2}}right)zeta (z)=pi ^{-{frac {1-z}{2}}};Gamma left({frac {1-z}{2}}right);zeta (1-z).

It also appears in the following formula:
ζ(z)Γ(z)=∫0∞uzeu−1duu,{displaystyle zeta (z)Gamma (z)=int _{0}^{infty }{frac {u^{z}}{e^{u}-1}},{frac {du}{u}},}{displaystyle zeta (z)Gamma (z)=int _{0}^{infty }{frac {u^{z}}{e^{u}-1}},{frac {du}{u}},}


which is valid only for Re(z) > 1.

The logarithm of the gamma function satisfies the following formula due to Lerch:
log⁡Γ(x)=ζH′(0,x)−ζ′(0),{displaystyle log Gamma (x)=zeta _{H}'(0,x)-zeta '(0),}log Gamma (x)=zeta _{H}'(0,x)-zeta '(0),


where ζH is the Hurwitz zeta function, ζ is the Riemann zeta function and the prime () denotes differentiation in the first variable.


  • The gamma function is related to the stretched exponential function. For instance, the moments of that function are

τn⟩0∞dttn−1e−(tτΓ(nβ).{displaystyle langle tau ^{n}rangle equiv int _{0}^{infty }dt,t^{n-1},e^{-left({frac {t}{tau }}right)^{beta }}={frac {tau ^{n}}{beta }}Gamma left({n over beta }right).}{displaystyle langle tau ^{n}rangle equiv int _{0}^{infty }dt,t^{n-1},e^{-left({frac {t}{tau }}right)^{beta }}={frac {tau ^{n}}{beta }}Gamma left({n over beta }right).}


Particular values



Some particular values of the gamma function are:


Γ(−32)=43π2.363271801207Γ(−12)=−3.544907701811Γ(12)=π1.772453850906Γ(1)=0!=1Γ(32)=12π0.886226925453Γ(2)=1!=1Γ(52)=34π1.329340388179Γ(3)=2!=2Γ(72)=158π3.323350970448Γ(4)=3!=6{displaystyle {begin{array}{rcccl}Gamma left(-{tfrac {3}{2}}right)&=&{tfrac {4}{3}}{sqrt {pi }}&approx &2.363,271,801,207\Gamma left(-{tfrac {1}{2}}right)&=&-2{sqrt {pi }}&approx &-3.544,907,701,811\Gamma left({tfrac {1}{2}}right)&=&{sqrt {pi }}&approx &1.772,453,850,906\Gamma (1)&=&0!&=&1\Gamma left({tfrac {3}{2}}right)&=&{tfrac {1}{2}}{sqrt {pi }}&approx &0.886,226,925,453\Gamma (2)&=&1!&=&1\Gamma left({tfrac {5}{2}}right)&=&{tfrac {3}{4}}{sqrt {pi }}&approx &1.329,340,388,179\Gamma (3)&=&2!&=&2\Gamma left({tfrac {7}{2}}right)&=&{tfrac {15}{8}}{sqrt {pi }}&approx &3.323,350,970,448\Gamma (4)&=&3!&=&6end{array}}}{displaystyle {begin{array}{rcccl}Gamma left(-{tfrac {3}{2}}right)&=&{tfrac {4}{3}}{sqrt {pi }}&approx &2.363,271,801,207\Gamma left(-{tfrac {1}{2}}right)&=&-2{sqrt {pi }}&approx &-3.544,907,701,811\Gamma left({tfrac {1}{2}}right)&=&{sqrt {pi }}&approx &1.772,453,850,906\Gamma (1)&=&0!&=&1\Gamma left({tfrac {3}{2}}right)&=&{tfrac {1}{2}}{sqrt {pi }}&approx &0.886,226,925,453\Gamma (2)&=&1!&=&1\Gamma left({tfrac {5}{2}}right)&=&{tfrac {3}{4}}{sqrt {pi }}&approx &1.329,340,388,179\Gamma (3)&=&2!&=&2\Gamma left({tfrac {7}{2}}right)&=&{tfrac {15}{8}}{sqrt {pi }}&approx &3.323,350,970,448\Gamma (4)&=&3!&=&6end{array}}}

The complex-valued gamma function is undefined for non-positive integers, but in these cases the value can be defined in the Riemann sphere as . The reciprocal gamma function is well defined and analytic at these values (and in the entire complex plane):


(−3)=1Γ(−2)=1Γ(−1)=1Γ(0)=0.{displaystyle {frac {1}{Gamma (-3)}}={frac {1}{Gamma (-2)}}={frac {1}{Gamma (-1)}}={frac {1}{Gamma (0)}}=0.}{displaystyle {frac {1}{Gamma (-3)}}={frac {1}{Gamma (-2)}}={frac {1}{Gamma (-1)}}={frac {1}{Gamma (0)}}=0.}


The log-gamma function




The analytic function log Γ(z)


Because the gamma and factorial functions grow so rapidly for moderately large arguments, many computing environments include a function that returns the natural logarithm of the gamma function (often given the name lgamma or lngamma in programming environments or gammaln in spreadsheets); this grows much more slowly, and for combinatorial calculations allows adding and subtracting logs instead of multiplying and dividing very large values. It is often defined as[16]


ln⁡Γ(z)=−γz−ln⁡z+∑k=1∞[zk−ln⁡(1+zk)]{displaystyle ln Gamma (z)=-gamma z-ln z+sum _{k=1}^{infty }left[{frac {z}{k}}-ln left(1+{frac {z}{k}}right)right]}{displaystyle ln Gamma (z)=-gamma z-ln z+sum _{k=1}^{infty }left[{frac {z}{k}}-ln left(1+{frac {z}{k}}right)right]}

The digamma function, which is the derivative of this function, is also commonly seen.
In the context of technical and physical applications, e.g. with wave propagation, the functional equation


ln⁡Γ(z)=ln⁡Γ(z+1)−ln⁡z{displaystyle ln Gamma (z)=ln Gamma (z+1)-ln z}{displaystyle ln Gamma (z)=ln Gamma (z+1)-ln z}

is often used since it allows one to determine function values in one strip of width 1 in z from the neighbouring strip. In particular, starting with a good approximation for a z with large real part one may go step by step down to the desired z. Following an indication of Carl Friedrich Gauss, Rocktaeschel (1922) proposed for ln(Γ(z)) an approximation for large Re(z):


ln⁡Γ(z)≈(z−12)ln⁡z−z+12ln⁡(2π).{displaystyle ln Gamma (z)approx (z-{tfrac {1}{2}})ln z-z+{tfrac {1}{2}}ln(2pi ).}{displaystyle ln Gamma (z)approx (z-{tfrac {1}{2}})ln z-z+{tfrac {1}{2}}ln(2pi ).}

This can be used to accurately approximate ln(Γ(z)) for z with a smaller Re(z) via (P.E.Böhmer, 1939)


ln⁡Γ(z−m)=ln⁡Γ(z)−k=1mln⁡(z−k).{displaystyle ln Gamma (z-m)=ln Gamma (z)-sum _{k=1}^{m}ln(z-k).}{displaystyle ln Gamma (z-m)=ln Gamma (z)-sum _{k=1}^{m}ln(z-k).}

A more accurate approximation can be obtained by using more terms from the asymptotic expansions of ln(Γ(z)) and Γ(z), which are based on Stirling's approximation.



Γ(z)∼zz−12e−z2π(1+112z+1288z2−13951840z3−5712488320z4){displaystyle Gamma (z)sim z^{z-{frac {1}{2}}}e^{-z}{sqrt {2pi }}left(1+{frac {1}{12z}}+{frac {1}{288z^{2}}}-{frac {139}{51,840z^{3}}}-{frac {571}{2,488,320z^{4}}}right)}{displaystyle Gamma (z)sim z^{z-{frac {1}{2}}}e^{-z}{sqrt {2pi }}left(1+{frac {1}{12z}}+{frac {1}{288z^{2}}}-{frac {139}{51,840z^{3}}}-{frac {571}{2,488,320z^{4}}}right)}

as |z| → ∞ at constant |arg(z)| < π.


In a more "natural" presentation:



ln⁡Γ(z)∼zln⁡(z)−z−12ln⁡(z2π)+112z−1360z3+11260z5{displaystyle ln Gamma (z)sim zln(z)-z-{tfrac {1}{2}}ln left({frac {z}{2pi }}right)+{frac {1}{12z}}-{frac {1}{360z^{3}}}+{frac {1}{1260z^{5}}}}{displaystyle ln Gamma (z)sim zln(z)-z-{tfrac {1}{2}}ln left({frac {z}{2pi }}right)+{frac {1}{12z}}-{frac {1}{360z^{3}}}+{frac {1}{1260z^{5}}}}

as |z| → ∞ at constant |arg(z)| < π.


The coefficients of the terms with k > 1 of zk + 1 in the last expansion are simply


Bkk(k−1){displaystyle {frac {B_{k}}{k(k-1)}}}{frac {B_{k}}{k(k-1)}}

where the Bk are the Bernoulli numbers.



Properties


The Bohr–Mollerup theorem states that among all functions extending the factorial functions to the positive real numbers, only the gamma function is log-convex, that is, its natural logarithm is convex on the positive real axis.


In a certain sense, the ln(Γ) function is the more natural form; it makes some intrinsic attributes of the function clearer. A striking example is the Taylor series of ln(Γ) around 1:


ln⁡Γ(1+z)=−γz+∑k=2∞ζ(k)k(−z)k∀|z|<1{displaystyle ln Gamma (1+z)=-gamma z+sum _{k=2}^{infty }{frac {zeta (k)}{k}},(-z)^{k}qquad forall ;|z|<1}{displaystyle ln Gamma (1+z)=-gamma z+sum _{k=2}^{infty }{frac {zeta (k)}{k}},(-z)^{k}qquad forall ;|z|<1}

with ζ(k) denoting the Riemann zeta function at k.


So, using the following property:


ζ(s)Γ(s)=∫0∞tset−1dtt{displaystyle zeta (s)Gamma (s)=int _{0}^{infty }{frac {t^{s}}{e^{t}-1}},{frac {dt}{t}}}{displaystyle zeta (s)Gamma (s)=int _{0}^{infty }{frac {t^{s}}{e^{t}-1}},{frac {dt}{t}}}

we can find an integral representation for the ln(Γ) function:


ln⁡Γ(1+z)=−γz+∫0∞e−zt−1+ztt(et−1)dt{displaystyle ln Gamma (1+z)=-gamma z+int _{0}^{infty }{frac {e^{-zt}-1+zt}{t(e^{t}-1)}},dt}{displaystyle ln Gamma (1+z)=-gamma z+int _{0}^{infty }{frac {e^{-zt}-1+zt}{t(e^{t}-1)}},dt}

or, setting z = 1 and calculating γ:


ln⁡Γ(1+z)=∫0∞e−zt−ze−t−1+zt(et−1)dt{displaystyle ln Gamma (1+z)=int _{0}^{infty }{frac {e^{-zt}-ze^{-t}-1+z}{t(e^{t}-1)}},dt}{displaystyle ln Gamma (1+z)=int _{0}^{infty }{frac {e^{-zt}-ze^{-t}-1+z}{t(e^{t}-1)}},dt}

There also exist special formulas for the logarithm of the gamma function for rational z.
For instance, if k{displaystyle k}k and n{displaystyle n}n are integers with k<n{displaystyle k<n}k<n and k≠n/2{displaystyle kneq n/2}{displaystyle kneq n/2}, then


ln⁡Γ(kn)=(n−2k)ln⁡2n+12{ln⁡πln⁡sin⁡πkn}+1πr=1n−+ln⁡rr⋅sin⁡rkn−12πsin⁡kn⋅0∞e−nx⋅ln⁡xcosh⁡x−cos⁡kndx{displaystyle {begin{aligned}ln Gamma {biggl (}!{frac {k}{n}}!{biggr )}={}&{frac {(n-2k)ln 2pi }{2n}}+{frac {1}{2}}left{ln pi -ln sin {frac {pi k}{n}}right}+{frac {1}{pi }}!sum _{r=1}^{n-1}{frac {gamma +ln {r}}{r}}cdot sin {frac {2pi rk}{n}}\&{}-{frac {1}{2pi }}sin {frac {2pi k}{n}}cdot !int _{0}^{infty }!!{frac {,e^{-nx}!cdot ln {x},}{,cosh {x}-cos {dfrac {2pi k}{n}},}},dxend{aligned}}}{displaystyle {begin{aligned}ln Gamma {biggl (}!{frac {k}{n}}!{biggr )}={}&{frac {(n-2k)ln 2pi }{2n}}+{frac {1}{2}}left{ln pi -ln sin {frac {pi k}{n}}right}+{frac {1}{pi }}!sum _{r=1}^{n-1}{frac {gamma +ln {r}}{r}}cdot sin {frac {2pi rk}{n}}\&{}-{frac {1}{2pi }}sin {frac {2pi k}{n}}cdot !int _{0}^{infty }!!{frac {,e^{-nx}!cdot ln {x},}{,cosh {x}-cos {dfrac {2pi k}{n}},}},dxend{aligned}}}

see.[17]
This formula is sometimes used for numerical computation, since the integrand decreases very quickly.



Integration over log-gamma


The integral


0zln⁡Γ(x)dx{displaystyle int _{0}^{z}ln Gamma (x),dx}{displaystyle int _{0}^{z}ln Gamma (x),dx}

can be expressed in terms of the Barnes G-function[18][19] (see Barnes G-function for a proof):


0zln⁡Γ(x)dx=z2ln⁡(2π)+z(1−z)2+zln⁡Γ(z)−ln⁡G(z+1){displaystyle int _{0}^{z}ln Gamma (x),dx={frac {z}{2}}ln(2pi )+{frac {z(1-z)}{2}}+zln Gamma (z)-ln G(z+1)}{displaystyle int _{0}^{z}ln Gamma (x),dx={frac {z}{2}}ln(2pi )+{frac {z(1-z)}{2}}+zln Gamma (z)-ln G(z+1)}

where Re(z) > −1.


It can also be written in terms of the Hurwitz zeta function:[20][21]


0zln⁡Γ(x)dx=z2ln⁡(2π)+z(1−z)2−ζ′(−1)+ζ′(−1,z).{displaystyle int _{0}^{z}ln Gamma (x),dx={frac {z}{2}}ln(2pi )+{frac {z(1-z)}{2}}-zeta '(-1)+zeta '(-1,z).}{displaystyle int _{0}^{z}ln Gamma (x),dx={frac {z}{2}}ln(2pi )+{frac {z(1-z)}{2}}-zeta '(-1)+zeta '(-1,z).}


Approximations




Comparison gamma (blue line) with the factorial (blue dots) and Stirling's approximation (red line)


Complex values of the gamma function can be computed numerically with arbitrary precision using Stirling's approximation or the Lanczos approximation.


The gamma function can be computed to fixed precision for Re(z) ∈ [1,2] by applying integration by parts to Euler's integral. For any positive number x the gamma function can be written


Γ(z)=∫0xe−ttzdtt+∫x∞e−ttzdtt=xze−x∑n=0∞xnz(z+1)⋯(z+n)+∫x∞e−ttzdtt.{displaystyle {begin{aligned}Gamma (z)&=int _{0}^{x}e^{-t}t^{z},{frac {dt}{t}}+int _{x}^{infty }e^{-t}t^{z},{frac {dt}{t}}\&=x^{z}e^{-x}sum _{n=0}^{infty }{frac {x^{n}}{z(z+1)cdots (z+n)}}+int _{x}^{infty }e^{-t}t^{z},{frac {dt}{t}}.end{aligned}}}{displaystyle {begin{aligned}Gamma (z)&=int _{0}^{x}e^{-t}t^{z},{frac {dt}{t}}+int _{x}^{infty }e^{-t}t^{z},{frac {dt}{t}}\&=x^{z}e^{-x}sum _{n=0}^{infty }{frac {x^{n}}{z(z+1)cdots (z+n)}}+int _{x}^{infty }e^{-t}t^{z},{frac {dt}{t}}.end{aligned}}}

When Re(z) ∈ [1,2] and x ≥ 1, the absolute value of the last integral is smaller than (x + 1)ex. By choosing a large enough x, this last expression can be made smaller than 2N for any desired value N. Thus, the gamma function can be evaluated to N bits of precision with the above series.


A fast algorithm for calculation of the Euler gamma function for any algebraic argument (including rational) was constructed by E.A. Karatsuba,[22][23][24]


For arguments that are integer multiples of 1/24, the gamma function can also be evaluated quickly using arithmetic–geometric mean iterations (see particular values of the gamma function and Borwein & Zucker (1992)).



Applications


One author describes the gamma function as "Arguably, the most common special function, or the least 'special' of them. The other transcendental functions […] are called 'special' because you could conceivably avoid some of them by staying away from many specialized mathematical topics. On the other hand, the gamma function y = Γ(x) is most difficult to avoid."[25]



Integration problems


The gamma function finds application in such diverse areas as quantum physics, astrophysics and fluid dynamics.[26] The gamma distribution, which is formulated in terms of the gamma function, is used in statistics to model a wide range of processes; for example, the time between occurrences of earthquakes.[27]


The primary reason for the gamma function's usefulness in such contexts is the prevalence of expressions of the type f(t) eg(t) which describe processes that decay exponentially in time or space. Integrals of such expressions can occasionally be solved in terms of the gamma function when no elementary solution exists. For example, if f is a power function and g is a linear function, a simple change of variables gives the evaluation


0∞tbe−atdt=Γ(b+1)ab+1.{displaystyle int _{0}^{infty }t^{b}e^{-at},dt={frac {Gamma (b+1)}{a^{b+1}}}.}int _{0}^{infty }t^{b}e^{-at},dt={frac {Gamma (b+1)}{a^{b+1}}}.

The fact that the integration is performed along the entire positive real line might signify that the gamma function describes the cumulation of a time-dependent process that continues indefinitely, or the value might be the total of a distribution in an infinite space.


It is of course frequently useful to take limits of integration other than 0 and to describe the cumulation of a finite process, in which case the ordinary gamma function is no longer a solution; the solution is then called an incomplete gamma function. (The ordinary gamma function, obtained by integrating across the entire positive real line, is sometimes called the complete gamma function for contrast.)


An important category of exponentially decaying functions is that of Gaussian functions


ae−(x−b)2c2{displaystyle ae^{-{frac {(x-b)^{2}}{c^{2}}}}}ae^{-{frac {(x-b)^{2}}{c^{2}}}}

and integrals thereof, such as the error function. There are many interrelations between these functions and the gamma function; notably, π obtained by evaluating Γ(1/2) is the "same" as that found in the normalizing factor of the error function and the normal distribution.


The integrals we have discussed so far involve transcendental functions, but the gamma function also arises from integrals of purely algebraic functions. In particular, the arc lengths of ellipses and of the lemniscate, which are curves defined by algebraic equations, are given by elliptic integrals that in special cases can be evaluated in terms of the gamma function. The gamma function can also be used to calculate "volume" and "area" of n-dimensional hyperspheres.


Another important special case is that of the beta function


B(x,y)=∫01tx−1(1−t)y−1dt=Γ(x)Γ(y)Γ(x+y).{displaystyle mathrm {B} (x,y)=int _{0}^{1}t^{x-1}(1-t)^{y-1},dt={frac {Gamma (x),Gamma (y)}{Gamma (x+y)}}.}{displaystyle mathrm {B} (x,y)=int _{0}^{1}t^{x-1}(1-t)^{y-1},dt={frac {Gamma (x),Gamma (y)}{Gamma (x+y)}}.}


Calculating products


The gamma function's ability to generalize factorial products immediately leads to applications in many areas of mathematics; in combinatorics, and by extension in areas such as probability theory and the calculation of power series. Many expressions involving products of successive integers can be written as some combination of factorials, the most important example perhaps being that of the binomial coefficient


(nk)=n!k!(n−k)!.{displaystyle {binom {n}{k}}={frac {n!}{k!(n-k)!}}.}{displaystyle {binom {n}{k}}={frac {n!}{k!(n-k)!}}.}

The example of binomial coefficients motivates why the properties of the gamma function when extended to negative numbers are natural. A binomial coefficient gives the number of ways to choose k elements from a set of n elements; if k > n, there are of course no ways. If k > n, (nk)! is the factorial of a negative integer and hence infinite if we use the gamma function definition of factorials—dividing by infinity gives the expected value of 0.


We can replace the factorial by a gamma function to extend any such formula to the complex numbers. Generally, this works for any product wherein each factor is a rational function of the index variable, by factoring the rational function into linear expressions. If P and Q are monic polynomials of degree m and n with respective roots p1, …, pm and q1, …, qn, we have


i=abP(i)Q(i)=(∏j=1mΓ(b−pj+1)Γ(a−pj))(∏k=1nΓ(a−qk)Γ(b−qk+1)).{displaystyle prod _{i=a}^{b}{frac {P(i)}{Q(i)}}=left(prod _{j=1}^{m}{frac {Gamma (b-p_{j}+1)}{Gamma (a-p_{j})}}right)left(prod _{k=1}^{n}{frac {Gamma (a-q_{k})}{Gamma (b-q_{k}+1)}}right).}prod _{i=a}^{b}{frac {P(i)}{Q(i)}}=left(prod _{j=1}^{m}{frac {Gamma (b-p_{j}+1)}{Gamma (a-p_{j})}}right)left(prod _{k=1}^{n}{frac {Gamma (a-q_{k})}{Gamma (b-q_{k}+1)}}right).

If we have a way to calculate the gamma function numerically, it is a breeze to calculate numerical values of such products. The number of gamma functions in the right-hand side depends only on the degree of the polynomials, so it does not matter whether ba equals 5 or 105. By taking the appropriate limits, the equation can also be made to hold even when the left-hand product contains zeros or poles.


By taking limits, certain rational products with infinitely many factors can be evaluated in terms of the gamma function as well. Due to the Weierstrass factorization theorem, analytic functions can be written as infinite products, and these can sometimes be represented as finite products or quotients of the gamma function. We have already seen one striking example: the reflection formula essentially represents the sine function as the product of two gamma functions. Starting from this formula, the exponential function as well as all the trigonometric and hyperbolic functions can be expressed in terms of the gamma function.


More functions yet, including the hypergeometric function and special cases thereof, can be represented by means of complex contour integrals of products and quotients of the gamma function, called Mellin–Barnes integrals.



Analytic number theory


An elegant and deep application of the gamma function is in the study of the Riemann zeta function. A fundamental property of the Riemann zeta function is its functional equation:


Γ(s2)ζ(s)πs2=Γ(1−s2)ζ(1−s)π1−s2.{displaystyle Gamma left({frac {s}{2}}right)zeta (s)pi ^{-{frac {s}{2}}}=Gamma left({frac {1-s}{2}}right)zeta (1-s)pi ^{-{frac {1-s}{2}}}.}Gamma left({frac {s}{2}}right)zeta (s)pi ^{-{frac {s}{2}}}=Gamma left({frac {1-s}{2}}right)zeta (1-s)pi ^{-{frac {1-s}{2}}}.

Among other things, this provides an explicit form for the analytic continuation of the zeta function to a meromorphic function in the complex plane and leads to an immediate proof that the zeta function has infinitely many so-called "trivial" zeros on the real line. Borwein et al. call this formula "one of the most beautiful findings in mathematics".[28] Another champion for that title might be


ζ(s)Γ(s)=∫0∞tset−1dtt.{displaystyle zeta (s);Gamma (s)=int _{0}^{infty }{frac {t^{s}}{e^{t}-1}},{frac {dt}{t}}.}{displaystyle zeta (s);Gamma (s)=int _{0}^{infty }{frac {t^{s}}{e^{t}-1}},{frac {dt}{t}}.}

Both formulas were derived by Bernhard Riemann in his seminal 1859 paper "Über die Anzahl der Primzahlen unter einer gegebenen Größe" ("On the Number of Prime Numbers less than a Given Quantity"), one of the milestones in the development of analytic number theory—the branch of mathematics that studies prime numbers using the tools of mathematical analysis. Factorial numbers, considered as discrete objects, are an important concept in classical number theory because they contain many prime factors, but Riemann found a use for their continuous extension that arguably turned out to be even more important.



History


The gamma function has caught the interest of some of the most prominent mathematicians of all time. Its history, notably documented by Philip J. Davis in an article that won him the 1963 Chauvenet Prize, reflects many of the major developments within mathematics since the 18th century. In the words of Davis, "each generation has found something of interest to say about the gamma function. Perhaps the next generation will also."[1]



18th century: Euler and Stirling





Daniel Bernoulli's letter to Christian Goldbach, October 6, 1729


The problem of extending the factorial to non-integer arguments was apparently first considered by Daniel Bernoulli and Christian Goldbach in the 1720s, and was solved at the end of the same decade by Leonhard Euler. Euler gave two different definitions: the first was not his integral but an infinite product,


n!=∏k=1∞(1+1k)n1+nk,{displaystyle n!=prod _{k=1}^{infty }{frac {left(1+{frac {1}{k}}right)^{n}}{1+{frac {n}{k}}}},,}n!=prod _{k=1}^{infty }{frac {left(1+{frac {1}{k}}right)^{n}}{1+{frac {n}{k}}}},,

of which he informed Goldbach in a letter dated October 13, 1729. He wrote to Goldbach again on January 8, 1730, to announce his discovery of the integral representation


n!=∫01(−ln⁡s)nds,{displaystyle n!=int _{0}^{1}(-ln s)^{n},ds,,}{displaystyle n!=int _{0}^{1}(-ln s)^{n},ds,,}

which is valid for n > 0. By the change of variables t = −ln s, this becomes the familiar Euler integral. Euler published his results in the paper "De progressionibus transcendentibus seu quarum termini generales algebraice dari nequeunt" ("On transcendental progressions, that is, those whose general terms cannot be given algebraically"), submitted to the St. Petersburg Academy on November 28, 1729.[29] Euler further discovered some of the gamma function's important functional properties, including the reflection formula.


James Stirling, a contemporary of Euler, also attempted to find a continuous expression for the factorial and came up with what is now known as Stirling's formula. Although Stirling's formula gives a good estimate of n!, also for non-integers, it does not provide the exact value. Extensions of his formula that correct the error were given by Stirling himself and by Jacques Philippe Marie Binet.



19th century: Gauss, Weierstrass and Legendre



De progressionibus transcendentibus, seu quarum termini generales algebraicae dari nequeunt

The first page of Euler's paper


Carl Friedrich Gauss rewrote Euler's product as


Γ(z)=limm→mzm!z(z+1)(z+2)⋯(z+m){displaystyle Gamma (z)=lim _{mto infty }{frac {m^{z}m!}{z(z+1)(z+2)cdots (z+m)}}}{displaystyle Gamma (z)=lim _{mto infty }{frac {m^{z}m!}{z(z+1)(z+2)cdots (z+m)}}}

and used this formula to discover new properties of the gamma function. Although Euler was a pioneer in the theory of complex variables, he does not appear to have considered the factorial of a complex number, as instead Gauss first did.[30] Gauss also proved the multiplication theorem of the gamma function and investigated the connection between the gamma function and elliptic integrals.


Karl Weierstrass further established the role of the gamma function in complex analysis, starting from yet another product representation,


Γ(z)=e−γzz∏k=1∞(1+zk)−1ezk,{displaystyle Gamma (z)={frac {e^{-gamma z}}{z}}prod _{k=1}^{infty }left(1+{frac {z}{k}}right)^{-1}e^{frac {z}{k}},}{displaystyle Gamma (z)={frac {e^{-gamma z}}{z}}prod _{k=1}^{infty }left(1+{frac {z}{k}}right)^{-1}e^{frac {z}{k}},}

where γ is the Euler–Mascheroni constant. Weierstrass originally wrote his product as one for 1/Γ, in which case it is taken over the function's zeros rather than its poles. Inspired by this result, he proved what is known as the Weierstrass factorization theorem—that any entire function can be written as a product over its zeros in the complex plane; a generalization of the fundamental theorem of algebra.


The name gamma function and the symbol Γ were introduced by Adrien-Marie Legendre around 1811; Legendre also rewrote Euler's integral definition in its modern form. Although the symbol is an upper-case Greek gamma, there is no accepted standard for whether the function name should be written "gamma function" or "Gamma function" (some authors simply write "Γ-function"). The alternative "pi function" notation Π(z) = z! due to Gauss is sometimes encountered in older literature, but Legendre's notation is dominant in modern works.


It is justified to ask why we distinguish between the "ordinary factorial" and the gamma function by using distinct symbols, and particularly why the gamma function should be normalized to Γ(n + 1) = n! instead of simply using "Γ(n) = n!". Consider that the notation for exponents, xn, has been generalized from integers to complex numbers xz without any change. Legendre's motivation for the normalization does not appear to be known, and has been criticized as cumbersome by some (the 20th-century mathematician Cornelius Lanczos, for example, called it "void of any rationality" and would instead use z!).[31] Legendre's normalization does simplify a few formulae, but complicates most others. From a modern point of view, the Legendre normalization of the Gamma function is the integral of the additive character ex against the multiplicative character xz with respect to the Haar measure dxx{displaystyle {tfrac {dx}{x}}}{displaystyle {tfrac {dx}{x}}} on the Lie group R+. Thus this normalization makes it clearer that the gamma function is a continuous analogue of a Gauss sum.



19th–20th centuries: characterizing the gamma function


It is somewhat problematic that a large number of definitions have been given for the gamma function. Although they describe the same function, it is not entirely straightforward to prove the equivalence. Stirling never proved that his extended formula corresponds exactly to Euler's gamma function; a proof was first given by Charles Hermite in 1900.[32] Instead of finding a specialized proof for each formula, it would be desirable to have a general method of identifying the gamma function.


One way to prove would be to find a differential equation that characterizes the gamma function. Most special functions in applied mathematics arise as solutions to differential equations, whose solutions are unique. However, the gamma function does not appear to satisfy any simple differential equation. Otto Hölder proved in 1887 that the gamma function at least does not satisfy any algebraic differential equation by showing that a solution to such an equation could not satisfy the gamma function's recurrence formula, making it a transcendentally transcendental function. This result is known as Hölder's theorem.


A definite and generally applicable characterization of the gamma function was not given until 1922. Harald Bohr and Johannes Mollerup then proved what is known as the Bohr–Mollerup theorem: that the gamma function is the unique solution to the factorial recurrence relation that is positive and logarithmically convex for positive z and whose value at 1 is 1 (a function is logarithmically convex if its logarithm is convex).


The Bohr–Mollerup theorem is useful because it is relatively easy to prove logarithmic convexity for any of the different formulas used to define the gamma function. Taking things further, instead of defining the gamma function by any particular formula, we can choose the conditions of the Bohr–Mollerup theorem as the definition, and then pick any formula we like that satisfies the conditions as a starting point for studying the gamma function. This approach was used by the Bourbaki group.


Borwein & Corless[33] review three centuries of work on the gamma function.



Reference tables and software


Although the gamma function can be calculated virtually as easily as any mathematically simpler function with a modern computer—even with a programmable pocket calculator—this was of course not always the case. Until the mid-20th century, mathematicians relied on hand-made tables; in the case of the gamma function, notably a table computed by Gauss in 1813 and one computed by Legendre in 1825.




A hand-drawn graph of the absolute value of the complex gamma function, from Tables of Higher Functions by Jahnke and Emde.


Tables of complex values of the gamma function, as well as hand-drawn graphs, were given in Tables of Higher Functions by Jahnke and Emde, first published in Germany in 1909. According to Michael Berry, "the publication in J&E of a three-dimensional graph showing the poles of the gamma function in the complex plane acquired an almost iconic status."[34]


There was in fact little practical need for anything but real values of the gamma function until the 1930s, when applications for the complex gamma function were discovered in theoretical physics. As electronic computers became available for the production of tables in the 1950s, several extensive tables for the complex gamma function were published to meet the demand, including a table accurate to 12 decimal places from the U.S. National Bureau of Standards.[1]


Abramowitz and Stegun became the standard reference for this and many other special functions after its publication in 1964.


Double-precision floating-point implementations of the gamma function and its logarithm are now available in most scientific computing software and special functions libraries, for example TK Solver, Matlab, GNU Octave, and the GNU Scientific Library. The gamma function was also added to the C standard library (math.h). Arbitrary-precision implementations are available in most computer algebra systems, such as Mathematica and Maple. PARI/GP, MPFR and MPFUN contain free arbitrary-precision implementations. A little-known feature of the calculator app included with the Android operating system is that it will accept fractional values as input to the factorial function and return the equivalent gamma function value. The same is true for Windows Calculator (in scientific mode).



See also




  • Ascending factorial

  • Elliptic gamma function

  • Gauss's constant

  • Hadamard's gamma function

  • Multiple gamma function

  • Multivariate gamma function


  • p-adic gamma function

  • Pochhammer k-symbol


  • q-gamma function

  • Ramanujan's master theorem

  • Spouge's approximation




Notes





  1. ^ abcdefghijklmn Davis, P. J. (1959). "Leonhard Euler's Integral: A Historical Profile of the Gamma Function". American Mathematical Monthly. 66 (10): 849–869. doi:10.2307/2309786. Retrieved 3 December 2016..mw-parser-output cite.citation{font-style:inherit}.mw-parser-output .citation q{quotes:"""""""'""'"}.mw-parser-output .citation .cs1-lock-free a{background: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 a{background: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 a{background: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-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-ws-icon a{background: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-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-maint{display:none;color:#33aa33;margin-left:0.3em}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}


  2. ^ Kingman, J. F. C. (1961). "A Convexity Property of Positive Matrices". The Quarterly Journal of Mathematics. 12 (1): 283–284. Bibcode:1961QJMat..12..283K. doi:10.1093/qmath/12.1.283.


  3. ^ Weisstein, Eric W. "Bohr–Mollerup Theorem". MathWorld.


  4. ^ Askey, R. A.; Roy, R. (2010), "Series Expansions", in Olver, Frank W. J.; Lozier, Daniel M.; Boisvert, Ronald F.; Clark, Charles W., NIST Handbook of Mathematical Functions, Cambridge University Press, ISBN 978-0521192255, MR 2723248


  5. ^ Askey, R. A.; Roy, R. (2010), "Series Expansions", in Olver, Frank W. J.; Lozier, Daniel M.; Boisvert, Ronald F.; Clark, Charles W., NIST Handbook of Mathematical Functions, Cambridge University Press, ISBN 978-0521192255, MR 2723248


  6. ^ Waldschmidt, M. (2006). "Transcendence of Periods: The State of the Art" (PDF). Pure Appl. Math. Quart. 2 (2): 435–463. doi:10.4310/pamq.2006.v2.n2.a3.
    open access publication – free to read



  7. ^ ab Weisstein, Eric W. "Gamma Function". MathWorld.


  8. ^ Whittaker and Watson, 12.2 example 1.


  9. ^ Whittaker and Watson, 12.31.


  10. ^ Whittaker and Watson, 12.32.


  11. ^ Whittaker and Watson, 12.22.


  12. ^ Bateman, Harry; Erdélyi, Arthur (1955). Higher Transcendental Functions. McGraw-Hill.


  13. ^ Srivastava, H. M.; Choi, J. (2001). Series Associated with the Zeta and Related Functions. The Netherlands: Kluwer Academic.


  14. ^ Blagouchine, Iaroslav V. (2014). "Rediscovery of Malmsten's integrals, their evaluation by contour integration methods and some related results". Ramanujan J. 35 (1): 21–110. doi:10.1007/s11139-013-9528-5.


  15. ^ Blagouchine, Iaroslav V. (2016). "Erratum and Addendum to "Rediscovery of Malmsten's integrals, their evaluation by contour integration methods and some related results"". Ramanujan J. doi:10.1007/s11139-015-9763-z.


  16. ^ "Log Gamma Function". Wolfram MathWorld. Retrieved 3 January 2019.


  17. ^ Blagouchine, Iaroslav V. (2015). "A theorem for the closed-form evaluation of the first generalized Stieltjes constant at rational arguments and some related summations". Journal of Number Theory. 148: 537–592. arXiv:1401.3724. doi:10.1016/j.jnt.2014.08.009.


  18. ^ Alexejewsky, W. P. (1894). "Über eine Classe von Funktionen, die der Gammafunktion analog sind" [On a class of functions analogous to the gamma function]. Leipzig Weidmanncshe Buchhandluns. 46: 268–275.


  19. ^ Barnes, E. W. (1899). "The theory of the G-function". Quart. J. Math. 31: 264–314.


  20. ^ Adamchik, Victor S. (1998). "Polygamma functions of negative order". J. Comput. Appl. Math. 100 (2): 191–199. doi:10.1016/S0377-0427(98)00192-7.


  21. ^ Gosper, R. W. (1997). "n/4m/6log⁡F(z)dz{displaystyle textstyle int _{n/4}^{m/6}log F(z),dz}{displaystyle textstyle int _{n/4}^{m/6}log F(z),dz} in special functions, q-series and related topics". J. Am. Math. Soc. 14.


  22. ^ E.A. Karatsuba, Fast evaluation of transcendental functions. Probl. Inf. Transm. Vol.27, No.4, pp. 339–360 (1991).


  23. ^ E.A. Karatsuba, On a new method for fast evaluation of transcendental functions. Russ. Math. Surv. Vol.46, No.2, pp. 246–247 (1991).


  24. ^ E.A. Karatsuba "Fast Algorithms and the FEE Method".


  25. ^ Michon, G. P. "Trigonometry and Basic Functions". Numericana. Retrieved May 5, 2007.


  26. ^ Chaudry, M. A. & Zubair, S. M. (2001). On A Class of Incomplete Gamma Functions with Applications. p. 37


  27. ^ Rice, J. A. (1995). Mathematical Statistics and Data Analysis (Second Edition). p. 52–53


  28. ^ Borwein, J., Bailey, D. H. & Girgensohn, R. (2003). Experimentation in Mathematics. A. K. Peters. p. 133. ISBN 1-56881-136-5.CS1 maint: Multiple names: authors list (link)


  29. ^ Euler's paper was published in Commentarii academiae scientiarum Petropolitanae 5, 1738, 36–57. See E19 -- De progressionibus transcendentibus seu quarum termini generales algebraice dari nequeunt, from The Euler Archive, which includes a scanned copy of the original article.


  30. ^ Remmert, R. (2006). Classical Topics in Complex Function Theory. Translated by Kay, L. D. Springer. ISBN 0-387-98221-3.


  31. ^ Lanczos, C. (1964). "A precision approximation of the gamma function". J. SIAM Numer. Anal. Ser. B. 1.


  32. ^ Knuth, D. E. (1997). The Art of Computer Programming, Volume 1 (Fundamental Algorithms). Addison-Wesley.


  33. ^ Borwein, Jonathan M.; Corless, Robert M. (2017). "Gamma and Factorial in the Monthly". American Mathematical Monthly. Mathematical Association of America. 125: 400–24. arXiv:1703.05349. Bibcode:2017arXiv170305349B. doi:10.1080/00029890.2018.1420983.


  34. ^ Berry, M. (April 2001). "Why are special functions special?". Physics Today.




Further reading


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



  • Abramowitz, Milton; Stegun, Irene A., eds. (1972). "Chapter 6". Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. New York: Dover.


  • Andrews, G. E.; Askey, R.; Roy, R. (1999). "Chapter 1 (Gamma and Beta functions)". Special Functions. New York: Cambridge University Press. ISBN 978-0-521-78988-2.


  • Artin, Emil (2006). "The Gamma Function". In Rosen, Michael. Exposition by Emil Artin: a selection. History of Mathematics. 30. Providence, RI: American Mathematical Society.


  • Askey, R.; Roy, R. (2010), "Gamma function", in Olver, Frank W. J.; Lozier, Daniel M.; Boisvert, Ronald F.; Clark, Charles W., NIST Handbook of Mathematical Functions, Cambridge University Press, ISBN 978-0521192255, MR 2723248


  • Birkhoff, George D. (1913). "Note on the gamma function". Bull. Amer. Math. Soc. 20 (1): 1–10. doi:10.1090/s0002-9904-1913-02429-7. MR 1559418.


  • Böhmer, P. E. (1939). Differenzengleichungen und bestimmte Integrale [Differential Equations and Definite Integrals]. Leipzig: Köhler Verlag.


  • Davis, Philip J. (1959). "Leonhard Euler's Integral: A Historical Profile of the Gamma Function". American Mathematical Monthly. 66: 849–869. doi:10.2307/2309786.


  • Press, W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P. (2007). "Section 6.1. Gamma Function". Numerical Recipes: The Art of Scientific Computing (3rd ed.). New York: Cambridge University Press. ISBN 978-0-521-88068-8.


  • Rocktäschel, O. R. (1922). Methoden zur Berechnung der Gammafunktion für komplexes Argument [Methods for Calculating the Gamma Function for Complex Arguments]. Dresden: Technical University of Dresden.


  • Temme, Nico M. (1996). Special Functions: An Introduction to the Classical Functions of Mathematical Physics. New York: John Wiley & Sons. ISBN 0-471-11313-1.


  • Whittaker, E. T.; Watson, G. N. (1927). A Course of Modern Analysis. Cambridge University Press. ISBN 978-0-521-58807-2.




External links







  • NIST Digital Library of Mathematical Functions:Gamma function

  • Pascal Sebah and Xavier Gourdon. Introduction to the Gamma Function. In PostScript and HTML formats.

  • C++ reference for std::tgamma

  • Examples of problems involving the gamma function can be found at Exampleproblems.com.


  • Hazewinkel, Michiel, ed. (2001) [1994], "Gamma function", Encyclopedia of Mathematics, Springer Science+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4

  • Wolfram gamma function evaluator (arbitrary precision)


  • "Gamma". Wolfram Functions Site.


  • Volume of n-Spheres and the Gamma Function at MathPages


This article incorporates material from the Citizendium article "Gamma function", which is licensed under the Creative Commons Attribution-ShareAlike 3.0 Unported License but not under the GFDL.









這個網誌中的熱門文章

12.7 cm/40 Type 89 naval gun

Shark

Wiciokrzew