Prime-counting function




In mathematics, the prime-counting function is the function counting the number of prime numbers less than or equal to some real number x.[1][2] It is denoted by π(x) (unrelated to the number π).




The values of π(n) for the first 60 integers




Contents






  • 1 History


    • 1.1 Exact form




  • 2 Table of π(x), x / ln x, and li(x)


  • 3 Algorithms for evaluating π(x)


    • 3.1 The Meissel–Lehmer algorithm




  • 4 Other prime-counting functions


  • 5 Formulas for prime-counting functions


  • 6 Inequalities


  • 7 The Riemann hypothesis


  • 8 See also


  • 9 References


  • 10 External links





History


Of great interest in number theory is the growth rate of the prime-counting function.[3][4] It was conjectured in the end of the 18th century by Gauss and by Legendre to be approximately


xln⁡(x){displaystyle {frac {x}{ln(x)}}}{displaystyle {frac {x}{ln(x)}}}

in the sense that


limx→π(x)x/ln⁡(x)=1.{displaystyle lim _{xrightarrow infty }{frac {pi (x)}{x/ln(x)}}=1.}{displaystyle lim _{xrightarrow infty }{frac {pi (x)}{x/ln(x)}}=1.}

This statement is the prime number theorem. An equivalent statement is


limx→π(x)/li⁡(x)=1{displaystyle lim _{xrightarrow infty }pi (x)/operatorname {li} (x)=1!}lim _{xrightarrow infty }pi (x)/operatorname {li} (x)=1!

where li is the logarithmic integral function. The prime number theorem was first proved in 1896 by Jacques Hadamard and by Charles de la Vallée Poussin independently, using properties of the Riemann zeta function introduced by Riemann in 1859. Proofs of the prime number theorem not using the zeta function or complex analysis were found around 1948 by Atle Selberg and by Paul Erdős (for the most part independently).[5]


More precise estimates of π(x){displaystyle pi (x)!}pi (x)! are now known; for example[citation needed]


π(x)=li⁡(x)+O(xe−ln⁡x/15){displaystyle pi (x)=operatorname {li} (x)+O{bigl (}xe^{-{sqrt {ln x}}/15}{bigr )}!}pi (x)=operatorname {li} (x)+O{bigl (}xe^{-{sqrt {ln x}}/15}{bigr )}!

where the O is big O notation. For most values of x{displaystyle x}x we are interested in (i.e., when x{displaystyle x}x is not unreasonably large) li⁡(x){displaystyle operatorname {li} (x)!}operatorname {li} (x)! is greater than π(x){displaystyle pi (x)!}pi (x)!, but infinitely often the opposite is true. For a discussion of this, see Skewes' number.



Exact form


Of profound importance, if the Riemann hypothesis is true then the prime-counting function equals[citation needed]


π(x)=R⁡(x)−ρR⁡(xρ){displaystyle pi (x)=operatorname {R} (x)-sum _{rho }operatorname {R} (x^{rho })}{displaystyle pi (x)=operatorname {R} (x)-sum _{rho }operatorname {R} (x^{rho })}

where



R⁡(x)=∑n=1∞μ(n)nli⁡(x1/n){displaystyle operatorname {R} (x)=sum _{n=1}^{infty }{frac {mu (n)}{n}}operatorname {li} (x^{1/n})}{displaystyle operatorname {R} (x)=sum _{n=1}^{infty }{frac {mu (n)}{n}}operatorname {li} (x^{1/n})},

μ(n) is the Möbius function, li(x) is the logarithmic integral function, and ρ are the complex zeros of the Riemann zeta function which lie along Re(s) = 1/2.



Table of π(x), x / ln x, and li(x)


The table shows how the three functions π(x), x / ln x and li(x) compare at powers of 10. See also,[3][6][7] and[8]




































































































































































































































x

π(x)

π(x) − x / ln x
li(x) − π(x)

x / π(x)
x / ln x % Error
10
4
−0.3
2.2
2.500
-7.5%
102
25
3.3
5.1
4.000
13.20%
103
168
23
10
5.952
13.69%
104
1,229
143
17
8.137
11.64%
105
9,592
906
38
10.425
9.45%
106
78,498
6,116
130
12.740
7.79%
107
664,579
44,158
339
15.047
6.64%
108
5,761,455
332,774
754
17.357
5.78%
109
50,847,534
2,592,592
1,701
19.667
5.10%
1010
455,052,511
20,758,029
3,104
21.975
4.56%
1011
4,118,054,813
169,923,159
11,588
24.283
4.13%
1012
37,607,912,018
1,416,705,193
38,263
26.590
3.77%
1013
346,065,536,839
11,992,858,452
108,971
28.896
3.47%
1014
3,204,941,750,802
102,838,308,636
314,890
31.202
3.21%
1015
29,844,570,422,669
891,604,962,452
1,052,619
33.507
2.99%
1016
279,238,341,033,925
7,804,289,844,393
3,214,632
35.812
2.79%
1017
2,623,557,157,654,233
68,883,734,693,281
7,956,589
38.116
2.63%
1018
24,739,954,287,740,860
612,483,070,893,536
21,949,555
40.420
2.48%
1019
234,057,667,276,344,607
5,481,624,169,369,960
99,877,775
42.725
2.34%
1020
2,220,819,602,560,918,840
49,347,193,044,659,701
222,744,644
45.028
2.22%
1021
21,127,269,486,018,731,928
446,579,871,578,168,707
597,394,254
47.332
2.11%
1022
201,467,286,689,315,906,290
4,060,704,006,019,620,994
1,932,355,208
49.636
2.02%
1023
1,925,320,391,606,803,968,923
37,083,513,766,578,631,309
7,250,186,216
51.939
1.93%
1024
18,435,599,767,349,200,867,866
339,996,354,713,708,049,069
17,146,907,278
54.243
1.84%
1025
176,846,309,399,143,769,411,680
3,128,516,637,843,038,351,228
55,160,980,939
56.546
1.77%
1026
1,699,246,750,872,437,141,327,603
28,883,358,936,853,188,823,261
155,891,678,121
58.850
1.70%

1027
16,352,460,426,841,680,446,427,399
267,479,615,610,131,274,163,365
508,666,658,006
61.153
1.64%



Graph showing ratio of the prime-counting function π(x) to two of its approximations, x/ln x and Li(x). As x increases (note x axis is logarithmic), both ratios tend towards 1. The ratio for x/ln x converges from above very slowly, while the ratio for Li(x) converges more quickly from below.


In the On-Line Encyclopedia of Integer Sequences, the π(x) column is sequence OEIS: A006880, π(x) − x/ln x is sequence OEIS: A057835, and li(x) − π(x) is sequence OEIS: A057752.


The value for π(1024) was originally computed by J. Buethe, J. Franke, A. Jost, and T. Kleinjung assuming the Riemann hypothesis.[9]
It was later verified unconditionally in a computation by D. J. Platt.[10]
The value for π(1025) is due to J. Buethe, J. Franke, A. Jost, and T. Kleinjung.[11]
The value for π(1026) was computed by D. B. Staple.[12] All other entries in this table were also verified as part of that work.



Algorithms for evaluating π(x)


A simple way to find π(x){displaystyle pi (x)}pi (x), if x{displaystyle x}x is not too large, is to use the sieve of Eratosthenes to produce the primes less than or equal to x{displaystyle x}x and then to count them.


A more elaborate way of finding π(x){displaystyle pi (x)}pi (x) is due to Legendre (using the inclusion–exclusion principle): given x{displaystyle x}x, if p1,p2,…,pn{displaystyle p_{1},p_{2},ldots ,p_{n}}p_{1},p_{2},ldots ,p_{n} are distinct prime numbers, then the number of integers less than or equal to x{displaystyle x}x which are divisible by no pi{displaystyle p_{i}}p_{i} is


x⌋i⌊xpi⌋+∑i<j⌊xpipj⌋−i<j<k⌊xpipjpk⌋+⋯{displaystyle lfloor xrfloor -sum _{i}leftlfloor {frac {x}{p_{i}}}rightrfloor +sum _{i<j}leftlfloor {frac {x}{p_{i}p_{j}}}rightrfloor -sum _{i<j<k}leftlfloor {frac {x}{p_{i}p_{j}p_{k}}}rightrfloor +cdots }{displaystyle lfloor xrfloor -sum _{i}leftlfloor {frac {x}{p_{i}}}rightrfloor +sum _{i<j}leftlfloor {frac {x}{p_{i}p_{j}}}rightrfloor -sum _{i<j<k}leftlfloor {frac {x}{p_{i}p_{j}p_{k}}}rightrfloor +cdots }

(where x⌋{displaystyle lfloor {x}rfloor }{displaystyle lfloor {x}rfloor } denotes the floor function). This number is therefore equal to


π(x)−π(x)+1{displaystyle pi (x)-pi left({sqrt {x}}right)+1}pi (x)-pi left({sqrt {x}}right)+1

when the numbers p1,p2,…,pn{displaystyle p_{1},p_{2},ldots ,p_{n}}p_{1},p_{2},ldots ,p_{n} are the prime numbers less than or equal to the square root of x{displaystyle x}x.



The Meissel–Lehmer algorithm



In a series of articles published between 1870 and 1885, Ernst Meissel described (and used) a practical combinatorial way of evaluating π(x){displaystyle pi (x)}pi (x). Let p1{displaystyle p_{1}}p_{1}p2,…,pn{displaystyle p_{2},ldots ,p_{n}}p_{2},ldots ,p_{n} be the first n{displaystyle n}n primes and denote by Φ(m,n){displaystyle Phi (m,n)}Phi (m,n) the number of natural numbers not greater than m{displaystyle m}m which are divisible by no pi{displaystyle p_{i}}p_{i}. Then


Φ(m,n)=Φ(m,n−1)−Φ(mpn,n−1).{displaystyle Phi (m,n)=Phi (m,n-1)-Phi left({frac {m}{p_{n}}},n-1right).}{displaystyle Phi (m,n)=Phi (m,n-1)-Phi left({frac {m}{p_{n}}},n-1right).}

Given a natural number m{displaystyle m}m, if n=π(m3){displaystyle n=pi left({sqrt[{3}]{m}}right)}n=pi left({sqrt[{3}]{m}}right) and if μ(m)−n{displaystyle mu =pi left({sqrt {m}}right)-n}mu =pi left({sqrt {m}}right)-n, then


π(m)=Φ(m,n)+n(μ+1)+μ2−μ2−1−k=1μπ(mpn+k).{displaystyle pi (m)=Phi (m,n)+n(mu +1)+{frac {mu ^{2}-mu }{2}}-1-sum _{k=1}^{mu }pi left({frac {m}{p_{n+k}}}right).}{displaystyle pi (m)=Phi (m,n)+n(mu +1)+{frac {mu ^{2}-mu }{2}}-1-sum _{k=1}^{mu }pi left({frac {m}{p_{n+k}}}right).}

Using this approach, Meissel computed π(x){displaystyle pi (x)}pi (x), for x{displaystyle x}x equal to 5×105, 106, 107, and 108.


In 1959, Derrick Henry Lehmer extended and simplified Meissel's method. Define, for real m{displaystyle m}m and for natural numbers n{displaystyle n}n and k{displaystyle k}k, Pk(m,n){displaystyle P_{k}(m,n)}P_{k}(m,n) as the number of numbers not greater than m with exactly k prime factors, all greater than pn{displaystyle p_{n}}p_{n}. Furthermore, set P0(m,n)=1{displaystyle P_{0}(m,n)=1}P_{0}(m,n)=1. Then


Φ(m,n)=∑k=0+∞Pk(m,n){displaystyle Phi (m,n)=sum _{k=0}^{+infty }P_{k}(m,n)}{displaystyle Phi (m,n)=sum _{k=0}^{+infty }P_{k}(m,n)}

where the sum actually has only finitely many nonzero terms. Let y{displaystyle y}y denote an integer such that m3≤y≤m{displaystyle {sqrt[{3}]{m}}leq yleq {sqrt {m}}}{sqrt[{3}]{m}}leq yleq {sqrt {m}}, and set n=π(y){displaystyle n=pi (y)}n=pi (y). Then P1(m,n)=π(m)−n{displaystyle P_{1}(m,n)=pi (m)-n}P_{1}(m,n)=pi (m)-n and Pk(m,n)=0{displaystyle P_{k}(m,n)=0}P_{k}(m,n)=0 when k{displaystyle k}k ≥ 3. Therefore,


π(m)=Φ(m,n)+n−1−P2(m,n){displaystyle pi (m)=Phi (m,n)+n-1-P_{2}(m,n)}pi (m)=Phi (m,n)+n-1-P_{2}(m,n)

The computation of P2(m,n){displaystyle P_{2}(m,n)}P_{2}(m,n) can be obtained this way:



P2(m,n)=∑y<p≤m(π(mp)−π(p)+1){displaystyle P_{2}(m,n)=sum _{y<pleq {sqrt {m}}}left(pi left({frac {m}{p}}right)-pi (p)+1right)}{displaystyle P_{2}(m,n)=sum _{y<pleq {sqrt {m}}}left(pi left({frac {m}{p}}right)-pi (p)+1right)},

where the sum is over prime numbers.


On the other hand, the computation of Φ(m,n){displaystyle Phi (m,n)}Phi (m,n) can be done using the following rules:



  1. Φ(m,0)=⌊m⌋{displaystyle Phi (m,0)=lfloor mrfloor }Phi (m,0)=lfloor mrfloor

  2. Φ(m,b)=Φ(m,b−1)−Φ(mpb,b−1){displaystyle Phi (m,b)=Phi (m,b-1)-Phi left({frac {m}{p_{b}}},b-1right)}Phi (m,b)=Phi (m,b-1)-Phi left({frac {m}{p_{b}}},b-1right)


Using his method and an IBM 701, Lehmer was able to compute π(1010){displaystyle pi left(10^{10}right)}pi left(10^{10}right).


Further improvements to this method were made by Lagarias, Miller, Odlyzko, Deléglise and Rivat.[13]



Other prime-counting functions


Other prime-counting functions are also used because they are more convenient to work with. One is Riemann's prime-counting function, usually denoted as Π0(x){displaystyle Pi _{0}(x)}Pi _{0}(x) or J0(x){displaystyle J_{0}(x)}J_{0}(x). This has jumps of 1/n for prime powers pn, with it taking a value halfway between the two sides at discontinuities. That added detail is used because then the function may be defined by an inverse Mellin transform. Formally, we may define Π0(x){displaystyle Pi _{0}(x)}Pi _{0}(x) by


Π0(x)=12(∑pn<x1n +∑pn≤x1n){displaystyle Pi _{0}(x)={frac {1}{2}}{bigg (}sum _{p^{n}<x}{frac {1}{n}} +sum _{p^{n}leq x}{frac {1}{n}}{bigg )}}Pi _{0}(x)={frac {1}{2}}{bigg (}sum _{p^{n}<x}{frac {1}{n}} +sum _{p^{n}leq x}{frac {1}{n}}{bigg )}

where p is a prime.


We may also write


Π0(x)=∑2xΛ(n)ln⁡n−12Λ(x)ln⁡x=∑n=1∞1nπ0(x1/n){displaystyle Pi _{0}(x)=sum _{2}^{x}{frac {Lambda (n)}{ln n}}-{frac {1}{2}}{frac {Lambda (x)}{ln x}}=sum _{n=1}^{infty }{frac {1}{n}}pi _{0}(x^{1/n})}Pi _{0}(x)=sum _{2}^{x}{frac {Lambda (n)}{ln n}}-{frac {1}{2}}{frac {Lambda (x)}{ln x}}=sum _{n=1}^{infty }{frac {1}{n}}pi _{0}(x^{1/n})

where Λ(n) is the von Mangoldt function and


π0(x)=limε(x−ε)+π(x+ε)2.{displaystyle pi _{0}(x)=lim _{varepsilon rightarrow 0}{frac {pi (x-varepsilon )+pi (x+varepsilon )}{2}}.}{displaystyle pi _{0}(x)=lim _{varepsilon rightarrow 0}{frac {pi (x-varepsilon )+pi (x+varepsilon )}{2}}.}

The Möbius inversion formula then gives


π0(x)=∑n=1∞μ(n)nΠ0(x1/n){displaystyle pi _{0}(x)=sum _{n=1}^{infty }{frac {mu (n)}{n}}Pi _{0}(x^{1/n})}{displaystyle pi _{0}(x)=sum _{n=1}^{infty }{frac {mu (n)}{n}}Pi _{0}(x^{1/n})}

Knowing the relationship between log of the Riemann zeta function and the von Mangoldt function Λ{displaystyle Lambda }Lambda , and using the Perron formula we have


ln⁡ζ(s)=s∫0∞Π0(x)x−s−1dx{displaystyle ln zeta (s)=sint _{0}^{infty }Pi _{0}(x)x^{-s-1},dx}ln zeta (s)=sint _{0}^{infty }Pi _{0}(x)x^{-s-1},dx

The Chebyshev function weights primes or prime powers pn by ln(p):



θ(x)=∑p≤xln⁡p{displaystyle theta (x)=sum _{pleq x}ln p}theta (x)=sum _{pleq x}ln p

ψ(x)=∑pn≤xln⁡p=∑n=1∞θ(x1/n)=∑n≤(n).{displaystyle psi (x)=sum _{p^{n}leq x}ln p=sum _{n=1}^{infty }theta (x^{1/n})=sum _{nleq x}Lambda (n).}psi (x)=sum _{p^{n}leq x}ln p=sum _{n=1}^{infty }theta (x^{1/n})=sum _{nleq x}Lambda (n).



Formulas for prime-counting functions


Formulas for prime-counting functions come in two kinds: arithmetic formulas and analytic formulas. Analytic formulas for prime-counting were the first used to prove the prime number theorem. They stem from the work of Riemann and von Mangoldt, and are generally known as explicit formulas.[14]


We have the following expression for ψ:


ψ0(x)=x−ρρln⁡12ln⁡(1−x−2),{displaystyle psi _{0}(x)=x-sum _{rho }{frac {x^{rho }}{rho }}-ln 2pi -{frac {1}{2}}ln(1-x^{-2}),}{displaystyle psi _{0}(x)=x-sum _{rho }{frac {x^{rho }}{rho }}-ln 2pi -{frac {1}{2}}ln(1-x^{-2}),}

where


ψ0(x)=limε(x−ε)+ψ(x+ε)2.{displaystyle psi _{0}(x)=lim _{varepsilon rightarrow 0}{frac {psi (x-varepsilon )+psi (x+varepsilon )}{2}}.}{displaystyle psi _{0}(x)=lim _{varepsilon rightarrow 0}{frac {psi (x-varepsilon )+psi (x+varepsilon )}{2}}.}

Here ρ are the zeros of the Riemann zeta function in the critical strip, where the real part of ρ is between zero and one. The formula is valid for values of x greater than one, which is the region of interest. The sum over the roots is conditionally convergent, and should be taken in order of increasing absolute value of the imaginary part. Note that the same sum over the trivial roots gives the last subtrahend in the formula.


For Π0(x){displaystyle Pi _{0}(x)}Pi _{0}(x) we have a more complicated formula


Π0(x)=li⁡(x)−ρli⁡(xρ)−ln⁡2+∫x∞dtt(t2−1)ln⁡t.{displaystyle Pi _{0}(x)=operatorname {li} (x)-sum _{rho }operatorname {li} (x^{rho })-ln 2+int _{x}^{infty }{frac {dt}{t(t^{2}-1)ln t}}.}{displaystyle Pi _{0}(x)=operatorname {li} (x)-sum _{rho }operatorname {li} (x^{rho })-ln 2+int _{x}^{infty }{frac {dt}{t(t^{2}-1)ln t}}.}



Riemann's explicit formula using the first 200 non-trivial zeros of the zeta function


Again, the formula is valid for x > 1, while ρ are the nontrivial zeros of the zeta function ordered according to their absolute value, and, again, the latter integral, taken with minus sign, is just the same sum, but over the trivial zeros. The first term li(x) is the usual logarithmic integral function; the expression li(xρ) in the second term should be considered as Ei(ρ ln x), where Ei is the analytic continuation of the exponential integral function from negative reals to the complex plane with branch cut along the positive reals.


Thus, Möbius inversion formula gives us[15]


π0(x)=R⁡(x)−ρR⁡(xρ)−1ln⁡x+1πarctan⁡πln⁡x{displaystyle pi _{0}(x)=operatorname {R} (x)-sum _{rho }operatorname {R} (x^{rho })-{frac {1}{ln x}}+{frac {1}{pi }}arctan {frac {pi }{ln x}}}{displaystyle pi _{0}(x)=operatorname {R} (x)-sum _{rho }operatorname {R} (x^{rho })-{frac {1}{ln x}}+{frac {1}{pi }}arctan {frac {pi }{ln x}}}

valid for x > 1, where


R⁡(x)=∑n=1∞μ(n)nli⁡(x1/n)=1+∑k=1∞(ln⁡x)kk!kζ(k+1){displaystyle operatorname {R} (x)=sum _{n=1}^{infty }{frac {mu (n)}{n}}operatorname {li} (x^{1/n})=1+sum _{k=1}^{infty }{frac {(ln x)^{k}}{k!kzeta (k+1)}}}{displaystyle operatorname {R} (x)=sum _{n=1}^{infty }{frac {mu (n)}{n}}operatorname {li} (x^{1/n})=1+sum _{k=1}^{infty }{frac {(ln x)^{k}}{k!kzeta (k+1)}}}

is the so-called Riemann's R-function[16] and μ(n) is the Möbius function. The latter series for it is known as Gram series[17] and converges for all positive x.




Δ-function (red line) on log scale


The sum over non-trivial zeta zeros in the formula for π0(x){displaystyle pi _{0}(x)}pi_0(x) describes the fluctuations of π0(x),{displaystyle pi _{0}(x),}{displaystyle pi _{0}(x),} while the remaining terms give the "smooth" part of prime-counting function,[18] so one can use


R⁡(x)−1ln⁡x+1πarctan⁡πln⁡x{displaystyle operatorname {R} (x)-{frac {1}{ln x}}+{frac {1}{pi }}arctan {frac {pi }{ln x}}}{displaystyle operatorname {R} (x)-{frac {1}{ln x}}+{frac {1}{pi }}arctan {frac {pi }{ln x}}}

as the best estimator of π(x){displaystyle pi (x)}pi (x) for x > 1.


The amplitude of the "noisy" part is heuristically about x/ln⁡x,{displaystyle {sqrt {x}}/ln x,}{displaystyle {sqrt {x}}/ln x,} so the fluctuations of the distribution of primes may be clearly represented with the Δ-function:


Δ(x)=(π0(x)−R⁡(x)+1ln⁡x−arctan⁡πln⁡x)ln⁡xx.{displaystyle Delta (x)=left(pi _{0}(x)-operatorname {R} (x)+{frac {1}{ln x}}-{frac {1}{pi }}arctan {frac {pi }{ln x}}right){frac {ln x}{sqrt {x}}}.}{displaystyle Delta (x)=left(pi _{0}(x)-operatorname {R} (x)+{frac {1}{ln x}}-{frac {1}{pi }}arctan {frac {pi }{ln x}}right){frac {ln x}{sqrt {x}}}.}

An extensive table of the values of Δ(x) is available.[7]



Inequalities


Here are some useful inequalities for π(x).


xln⁡x<π(x)<1.25506xln⁡x{displaystyle {frac {x}{ln x}}<pi (x)<1.25506{frac {x}{ln x}}}{displaystyle {frac {x}{ln x}}<pi (x)<1.25506{frac {x}{ln x}}}

for x ≥ 17.


The left inequality holds for x ≥ 17 and the right inequality holds for x > 1. The constant 1.25506 is 30ln⁡113113{displaystyle {frac {30ln 113}{113}}}{displaystyle {frac {30ln 113}{113}}} to 5 decimal places, as π(x)ln⁡xx{displaystyle {frac {pi (x)ln x}{x}}}{displaystyle {frac {pi (x)ln x}{x}}} has its maximum value at x = 113.[19]


Pierre Dusart proved in 2010:



xln⁡x−1<π(x){displaystyle {frac {x}{ln x-1}}<pi (x)}{frac {x}{ln x-1}}<pi (x) for x≥5393{displaystyle xgeq 5393}xgeq 5393, and


π(x)<xln⁡x−1.1{displaystyle pi (x)<{frac {x}{ln x-1.1}}}pi (x)<{frac {x}{ln x-1.1}} for x≥60184{displaystyle xgeq 60184}xgeq 60184.[20]

Here are some inequalities for the nth prime, pn. The upper bound is due to Rosser (1941),[21] the lower one to Dusart (1999):[22]


n(ln⁡(nln⁡n)−1)<pn<nln⁡(nln⁡n){displaystyle n(ln(nln n)-1)<p_{n}<n{ln(nln n)}!}n(ln(nln n)-1)<p_{n}<n{ln(nln n)}! for n ≥ 6.


The left inequality holds for n ≥ 2 and the right inequality holds for n ≥ 6.


An approximation for the nth prime number is


pn=n(ln⁡(nln⁡n)−1)+n(ln⁡ln⁡n−2)ln⁡n+O(n(ln⁡ln⁡n)2(ln⁡n)2).{displaystyle p_{n}=n(ln(nln n)-1)+{frac {n(ln ln n-2)}{ln n}}+Oleft({frac {n(ln ln n)^{2}}{(ln n)^{2}}}right).}p_{n}=n(ln(nln n)-1)+{frac {n(ln ln n-2)}{ln n}}+Oleft({frac {n(ln ln n)^{2}}{(ln n)^{2}}}right).

In his well-known notebooks, Ramanujan[23] proves that the inequality


π(x)2<exlog⁡(xe){displaystyle pi (x)^{2}<{frac {ex}{log x}}pi {bigg (}{frac {x}{e}}{bigg )}}{displaystyle pi (x)^{2}<{frac {ex}{log x}}pi {bigg (}{frac {x}{e}}{bigg )}}

holds for all sufficiently large values of x{displaystyle x}x.



The Riemann hypothesis


The Riemann hypothesis is equivalent to a much tighter bound on the error in the estimate for π(x){displaystyle pi (x)}pi (x), and hence to a more regular distribution of prime numbers,


π(x)=li⁡(x)+O(xlog⁡x).{displaystyle pi (x)=operatorname {li} (x)+O({sqrt {x}}log {x}).}pi (x)=operatorname {li} (x)+O({sqrt {x}}log {x}).

Specifically,[24]


(x)−li⁡(x)|<18πxlog⁡x,for all x≥2657.{displaystyle |pi (x)-operatorname {li} (x)|<{frac {1}{8pi }}{sqrt {x}},log {x},qquad {text{for all }}xgeq 2657.}|pi (x)-operatorname {li} (x)|<{frac {1}{8pi }}{sqrt {x}},log {x},qquad {text{for all }}xgeq 2657.


See also



  • Foias constant

  • Bertrand's postulate

  • Oppermann's conjecture

  • Ramanujan prime



References





  1. ^ Bach, Eric; Shallit, Jeffrey (1996). Algorithmic Number Theory. MIT Press. volume 1 page 234 section 8.8. ISBN 0-262-02405-5..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. ^ Weisstein, Eric W. "Prime Counting Function". MathWorld.


  3. ^ ab "How many primes are there?". Chris K. Caldwell. Retrieved 2008-12-02.


  4. ^ Dickson, Leonard Eugene (2005). History of the Theory of Numbers, Vol. I: Divisibility and Primality. Dover Publications. ISBN 0-486-44232-2.


  5. ^ Ireland, Kenneth; Rosen, Michael (1998). A Classical Introduction to Modern Number Theory (Second ed.). Springer. ISBN 0-387-97329-X.


  6. ^ "Tables of values of pi(x) and of pi2(x)". Tomás Oliveira e Silva. Retrieved 2008-09-14.


  7. ^ ab "Values of π(x) and Δ(x) for various values of x". Andrey V. Kulsha. Retrieved 2008-09-14.


  8. ^ "A table of values of pi(x)". Xavier Gourdon, Pascal Sebah, Patrick Demichel. Retrieved 2008-09-14.


  9. ^ "Conditional Calculation of pi(1024)". Chris K. Caldwell. Retrieved 2010-08-03.


  10. ^ Platt, David J. (2012). "Computing π(x) Analytically)". arXiv:1203.5712 [math.NT].


  11. ^ "How Many Primes Are There?". J. Buethe. Retrieved 2015-09-01.


  12. ^ "The combinatorial algorithm for computing pi(x)". Dalhousie University. Retrieved 2015-09-01.


  13. ^ "Computing π(x): The Meissel, Lehmer, Lagarias, Miller, Odlyzko method" (PDF). Marc Deléglise and Jöel Rivat, Mathematics of Computation, vol. 65, number 33, January 1996, pages 235–245. Retrieved 2008-09-14.


  14. ^ Titchmarsh, E.C. (1960). The Theory of Functions, 2nd ed. Oxford University Press.


  15. ^ Riesel, Hans; Göhl, Gunnar (1970). "Some calculations related to Riemann's prime number formula". Mathematics of Computation. American Mathematical Society. 24 (112): 969–983. doi:10.2307/2004630. ISSN 0025-5718. JSTOR 2004630. MR 0277489.


  16. ^ Weisstein, Eric W. "Riemann Prime Counting Function". MathWorld.


  17. ^ Weisstein, Eric W. "Gram Series". MathWorld.


  18. ^ "The encoding of the prime distribution by the zeta zeros". Matthew Watkins. Retrieved 2008-09-14.


  19. ^ Rosser, J. Barkley; Schoenfeld, Lowell (1962). "Approximate formulas for some functions of prime numbers". Illinois J. Math. 6: 64–94. ISSN 0019-2082. Zbl 0122.05001.


  20. ^ Dusart, Pierre. "Estimates of Some Functions Over Primes without R.H." (PDF). Retrieved 22 April 2014.


  21. ^ Rosser, Barkley (1941). "Explicit bounds for some functions of prime numbers". American Journal of Mathematics. 63 (1): 211–232. doi:10.2307/2371291. JSTOR 2371291.


  22. ^ Dusart, Pierre (1999). "The kth prime is greater than k(lnk+lnlnk-1) for k>=2". Mathematics of Computation. 68 (225): 411–415. doi:10.1090/S0025-5718-99-01037-6.


  23. ^ Berndt, Bruce C. (2012-12-06). Ramanujan’s Notebooks, Part IV. Springer Science & Business Media. pp. 112–113. ISBN 9781461269328.


  24. ^ Schoenfeld, Lowell (1976). "Sharper bounds for the Chebyshev functions θ(x) and ψ(x). II". Mathematics of Computation. American Mathematical Society. 30 (134): 337–360. doi:10.2307/2005976. ISSN 0025-5718. JSTOR 2005976. MR 0457374.




External links



  • Chris Caldwell, The Nth Prime Page at The Prime Pages.

  • Tomás Oliveira e Silva, Tables of prime-counting functions.




Popular posts from this blog

Lambaréné

Chris Pine

Kashihara Line