You are on page 1of 394

Published Works Of Srinivasa Ramanujan

Srinivasa Ramanujan (born December 22, 1887, Erode, India — died April 26, 1920, Kumbakonam) was an Indian

mathematician who lived during the British Rule in India. Though he had almost no formal training in pure

mathematics, he made substantial contributions to the theory of numbers include pioneering discoveries of the

properties of the partition function.


Srinivasa Ramanujan

Born 22 December 1887

Erode, Madras Presidency, British India (present-

day Tamil Nadu, India)

Died 26 April 1920 (aged 32)

Kumbakonam, Madras Presidency, British

India (present-day Tamil Nadu, India)

 Kumbakonam, Madras Presidency, British


Residence
India (present-day Tamil Nadu, India)

 Madras, Madras Presidency, British

India (present-day Chennai, Tamil

Nadu, India)

 London, England, United Kingdom of Great

Britain and Ireland (present-day United

Kingdom)

Nationality Indian
Education  Government Arts College (no degree)

 Pachaiyappa's College (no degree)

 Trinity College, Cambridge (BSc, 1916)

Known for  Landau–Ramanujan constant

 Mock theta functions

 Ramanujan conjecture

 Ramanujan prime

 Ramanujan–Soldner constant

 Ramanujan theta function

 Ramanujan's sum

 Rogers–Ramanujan identities

 Ramanujan's master theorem

 Ramanujan–Sato series

Awards Fellow of the Royal Society

Scientific career

Fields Mathematics

Institutions Trinity College, Cambridge

Thesis Highly Composite Numbers (1916)

Academic  G. H. Hardy

advisors  J. E. Littlewood

Influences G. S. Carr

Influenced G. H. Hardy

Signature
Some properties of Bernoulli’s numbers

Journal of the Indian Mathematical Society, III, 1911, 219 – 234

1. Let the well-known expansion of x cot x (vide Edwards’ Differential Calculus, §149) be
written in the form
B2 B4 B6
x cot x = 1 − (2x)2 − (2x)4 − (2x)6 − · · · , (1)
2! 4! 6!
from which we infer that B0 may be supposed to be −1. Now

x2 x4 x6
1− + − + ···
cos x 2! 4! 6!
cot x = =
sin x x3 x5 x7
x− + − + ···
3! 5! 7!
2x (2x)3 (2x)5
− + − ···
sin 2x 1! 3! 5!
= =
1 − cos 2x (2x)2 (2x)4 (2x)6
− + − ···
2! 4! 6!
(2x)2 (2x)4 (2x)6
2− + − + ···
1 + cos 2x 2! 4! 6!
= =
sin 2x (2x)3 (2x)5 (2x)7
2x − + − + ···
3! 5! 7!
Multiplying both sides in each of the above three relations by the deno minator of the
right-hand side and equating the coefficients of xn on both sides, we can write the results
thus:
1
Bn−1 Bn−3 Bn−5 (−1) 2 (n−1) n 1
c1 − c3 3 + c5 5 − · · · + n
B0 + n (−1) 2 (n−1) = 0, (2)
2 2 2 2 2
where n is any odd integer;
1 n 1
c2 Bn−2 − c4 Bn−4 + c6 Bn−6 − · · · + (−1) 2 (n−2) B0 + (−1) 2 (n−2) = 0, (3)
2
where n is any even integer;
1 n 1
c1 Bn−1 − c3 Bn−3 + c5 Bn−5 − · · · + (−1) 2 (n−1) B0 + (−1) 2 (n−1) = 0, (4)
2
where n is any odd integer greater than unity.
From any one of (2), (3), (4) we can calculate the B’s. But as n becomes greater and
greater the calculation will get tedious. So we shall try to find simpler methods.
2 Paper 1

2. We know  
d cot x
(x cot x)2 = −x2 1 + .
dx
Using (1) and equating the coefficients of xn on both sides, and simplifying, we have
1
2 (n + 1)Bn = c2 B2 Bn−2 + c4 B4 Bn−4 + c6 B6 Bn−6 + · · · ,
the last term being c n2 −1 B n2 −1 B n2 +1 or 12 c n2 (B n2 )2 according as n2 is odd or even. . . . (5)
A similar result can be obtained by equating the coefficients of xn in the identity
d tan x
= 1 + tan2 x.
dx

3. Again
− 12 x(cot 21 x + coth 21 x) = − 12 x(cot 21 x + i cot 12 ix)

x4 x8
 
= 2 B0 + B4 + B8 + · · · ,
4! 8!
by using (1). The expression may also be written
(cos 12 x sin 21 ix + i sin 21 x cos 21 ix) (1 + i) sin 12 x(1 + i) − (1 − i) sin 21 x(1 − i)
− 12 x = − 21 x
sin 21 x sin 12 ix cos 12 x(1 − i) − cos 21 x(1 + i)

x x5 x9
− 2 + 4 − ···
= −x 1! 2 · 5! 2 · 9!
x2 x6 x10
− 2 + 4 − ···
2! 2 · 6! 2 · 10!
by expanding the numerator and the denominator, and simplifying by De Moivre’s theorem.
Hence
x x5

x 4 x 8  − 2
+ ···
2 B0 + B4 + B8 + · · · = −x 1!2 2 ·65! . (6)
4! 8! x x
− 2 + ···
2! 2 · 6!
Similarly
x2 x6 x10
 
1 1 1

− 2 x cot 2 x − coth 2 x = 2 B2 + B6 + B10 + ···
2! 6! 10!
1
(1 − i) sin 12 x(1 + i) − 21 (1 + i) sin 12 x(1 − i)
= x2
cos 12 x(1 + i) − cos 12 x(1 − i)

x3 x7 x11
− 3 + 5 + ···
= x 2 · 23! 2 6· 7! 2 10· 11! . (7)
x x x
− 2 + 4 − ···
2! 2 · 6! 2 · 10!
Some properties of Bernoulli’s numbers 3

Proceeding as in §1 we have, if n is an even integer greater than 2,


Bn−2 Bn−6 Bn−10 n 1 n 1
c2 − c6 3 + c10 5
− ··· + 1 (−1) 4 n or 1 (−1) 4 (n−2) = 0, (8)
2 2 2 22 (n+2)
22 (n+2)

according as n or n − 2 is a multiple of 4.
Analogous results can be obtained from tan 12 x ± tanh 12 x.
In (2), (3) and (4) there are 21 n terms, while in (5) and (8) there are 41 n or 41 (n − 2) terms.
Thus Bn can be found from only half of the previous B’s.

4. A still simpler method can be deduced from the following identities.


If 1, ω, ω 2 be the three cube roots of unity, then

4 sin x sin xω sin xω 2 = −(sin 2x + sin 2xω + sin 2xω 2 ),

as may easily be verified.


By logarithmic differentiation, we have
cos 2x + ω cos 2xω + ω 2 cos 2xω 2
cot x + ω cot xω + ω 2 cot xω 2 = 2 .
sin 2x + sin 2xω + sin 2xω 2
Writing 21 x for x,

cos x + ω cos xω + ω 2 cos xω 2


− 12 x(cot 12 x + ω cot 12 xω + ω 2 cot 12 xω 2 ) = −x
sin x + sin xω + sin xω 2
and, proceeding as in §3, we get
x2 x8 x14

x6 x12
 − + − ···
3 B0 + B6 + B12 + ··· = −x 2!3 8! 14! . (9)
6! 12! x x9 x15
− + − ···
3! 9! 15!
Again
cos xω 2 − cos xω 2(cos xω − cos xω 2 )
cot 12 xω − cot 12 xω 2 = = .
2 sin 12 x sin 21 xω sin 12 xω 2 sin x + sin xω + sin xω 2

Multiplying both sides by − 12 x(ω 2 −ω) and adding to the corresponding sides of the previous
result, we have
cos x + ω 2 cos xω + ω cos xω 2
− 21 x(cot 21 x + ω 2 cot 21 xω + ω cot 12 xω 2 ) = −x .
sin x + sin xω + sin xω 2
Hence, as before,
x4 x10 x16

x2 x8 x14
 − + − ···
3 B2 + B8 + B14 + ··· = x 4!3 10! 16! . (10)
2! 8! 14! x x9 x15
− + − ···
3! 9! 15!
4 Paper 1

Similarly

cos x + cos xω + cos xω 2 − 3


−x(cot 12 x + cot 21 xω + cot 12 xω 2 ) = x ,
sin x + sin xω + sin xω 2
and therefore
x6 x12 x18

x4 x10 x16
 − + − ···
6 B4 + B10 + B16 + ··· = x 6!3 12! 18! . (11)
4! 10! 16! x x9 x15
− + − ···
3! 9! 15!
Multiplying up and equating coefficients in (9), (10) and (11) as usual, we have,

c3 Bn−3 − c9 Bn−9 + c15 Bn−15 − · · · = 0, (12)


1 1 1
the last term being 16 n(−1) 6 (n−1) , 13 n(−1) 6 (n+1) , 31 n(−1) 6 (n−3) .
Again, dividing both sides in (10) by x and differentiating, we have
 4 10 16

x x x

1 x6 x12

d  − + − · · ·
3 B2 + 7B8 + 13B14 + ··· =  4! 10! 16! 
2! 8! 14! dx  x3 x9 x15 
− + − ···
3! 9! 15!
x 2 x 8 x 14 x4 x10 x16
− + − ··· − + − ···
= 1 − 2!3 8!
9
14!
15 · 4!3 10! 9
16!
15 .
x x x x x x
− + − ··· − + − ···
3! 9! 15! 3! 9! 15!
Hence by (9) and (10),

x2 x8 x14
 
3 B2 + 7B8 + 13B14 + ···
2! 8! 14!
x6 x12 x2 x8 x14
  
2
= x + 9 B0 + B6 + B12 + ··· B2 + B8 + B14 + ··· .
6! 12! 2! 8! 14!

Equating the coefficients of xn we have, if n > 2 and n − 2 is a multiple of 6,


1
3 (n + 2)Bn = c6 Bn−6 B6 + c12 Bn−12 B12 + c18 Bn−18 B18 + · · · . (13)

From (12) the B’s can be calculated very quickly and (13) may prove useful in checking
the calculations. The number of terms is one-third of that in (4); thus B24 is found from
B18 , B12 and B6 .

5. We shall see later on how the B’s can be obtained from their properties only. But to know
these properties, it will be convenient to calculate a few B’s by substituting 3, 5, 7, 9, · · · ,
Some properties of Bernoulli’s numbers 5

for n in succession in (12). Thus

B0 = −1; B2 = 16 ; B4 = 1
30 ; B6 = 1
42 ; B8 − 31 B2 = − 45
1
;

B2
B10 − 52 B4 = − 132
1 4
; B12 − 11B6 = − 455 ; B14 − 143
4 B8 + = 1
120 ;
5
286 1 204 3
B16 − 3 B10 + 4B4 = 306 ; B18 − 221B12 + 5 B6 = 665 ;

3230 1938 B2 1
B20 − 7 B14 + 7 B8 − = − 231 ;
7
3553 7106 11 1
B22 − 4 B16 + 5 B10 − 2 B4 = − 552 ;

and so on. Hence we have finally the following values:

B2 = 16 ; B4 = 1
30 ; B6 = 1
42 ; B8 = 1
30 ; B10 = 5
66 ; B12 = 691
2730 ; B14 = 76 ;

3617 43867 174611 854513 236364091


B16 = 510 ; B18 = 798 ; B20 = 330 ; B22 = 138 ; B24 = 2730 ;

8553103 23749461029 8615841276005


B26 = 6 ; B28 = 870 ; B30 = 14322 ;

7709321041217 2577687858367 26315271553053477373


B32 = 510 ; B34 = 6 ; B36 = 1919190 ;

2929993913841559 261082718496449122051
B38 = 6 ; B40 = 13530 ; · · · , B∞ = ∞.

6. It will be observed∗ that, if n is even but not equal to zero,


(i) Bn is a fraction and the numerator of Bn /n in its lowest terms is
a prime number, (14)
(ii) the denominator of Bn contains each of the factors 2 and 3 once
and only once, (15)
(iii) 2n (2n − 1)Bn /n is an integer and consequently 2(2n − 1)Bn is an
odd integer. (16)
From (16) it can easily be shewn that the denominator of 2(2n − 1)Bn /n in its lowest terms
is the greatest power of 2 which divides n; and consequently, if n is not a multiple of 4,
then 4(2n − 1)Bn /n is an odd integer. . . . (17)
It follows from (14) that the numerator of Bn in its lowest terms is divisible by the greatest
measure of n prime to the denominator, and the quotient is a prime number. . . . (18)
Examples: (a) 2 and 3 are the only prime factors of 12, 24 and 36 and they are found in
the denominators of B12 , B24 and B36 and their numerators are prime numbers.

See §12 below
6 Paper 1

(b) 11 is not found in the denominator of B22 , and hence its numerator is divisible by 11;
similarly, the numerators of B26 , B34 , B38 are divisible by 13, 17, 19, respectively and the
quotients in all cases are prime numbers.
(c) 5 is found in the denominator of B20 and not in that of B30 , and consequently the
numerator of B30 is divisible by 5 while that of B20 is a prime number. Thus we may say
that if a prime number appearing in n is not found in the denominator it will appear in
the numerator, and vice versa.

7. Next, let us consider the denominators.


All the denominators are divisible by 6; those of B4 , B8 , B12 , · · · by 5; those of B6 , B12 ,
B18 , · · · by 7; those of B10 , B20 , B30 , · · · by 11; but those of B8 , B16 , B24 , · · · are not divisible
by 9; and those of B14 , B28 , · · · are not divisible by 15. Hence we may infer that:
the denominator of Bn is the continued product of prime numbers which are the next
numbers (in the natural order) to the factors of n (including unity and the number itself)
. . . (19)
As an example take the denominator of B24 . Write all the factors of 24, viz. 1, 2, 3, 4, 6,
8, 12, 24. The next numbers to these are 2, 3, 4, 5, 7, 9, 13, 25. Strike out the composite
numbers and we have the prime numbers 2, 3, 5, 7, 13. And the denominator of B24 is the
product of 2, 3, 5, 7, 13, i.e., 2730.
It is unnecessary to write the odd factors of n except unity, as the next numbers to these
are even and hence composite.
The following are some further examples:

Even factors of n and unity Denominator of Bn


B2 ··· 1, 2, ··· ··· ··· ··· 2·3=6
B6 ··· 1, 2, 6 · · · ··· ··· ··· 2 · 3 · 7 = 42
B12 ··· 1, 2, 4, 6, 12 ··· ··· ··· 2 · 3 · 5 · 7 · 13 = 2730
B20 ··· 1, 2, 4, 10, 20 ··· ··· ··· 2 · 3 · 5 · 11 = 330
B30 ··· 1, 2, 6, 10, 30 ··· ··· ··· 2 · 3 · 7 · 11 · 31 = 14322
B42 ··· 1, 2, 6, 14, 42 ··· ··· ··· 2 · 3 · 7 · 43 = 1806
B56 ··· 1, 2, 4, 8, 14, 28, 56 ··· ··· 2 · 3 · 5 · 29 = 870
B72 ··· 1, 2, 4, 6, 8, 12, 18, 24, 36, 72 ··· 2 · 3 · 5 · 7 · 13 · 19 · 37 · 73 = 140100870
B90 ··· 1, 2, 6, 10, 18, 30, 90 · · · ··· 2 · 3 · 7 · 11 · 19 · 31 = 272118
B110 ··· 1, 2, 10, 22, 110 · · · · · · ··· 2 · 3 · 11 · 23 = 1518

8. Again taking the fractional part of any B and splitting it into partial fractions, we see
that:
1
the fractional part of Bn = (−1) 2 n {the sum of the reciprocals of the prime factors of the
1
denominator of Bn } − (−1) 2 n . (20)
1 1 1 1 47
Thus the fractional part of B16 = 2 + 3 + 5 + 17 −1= 510 ;
Some properties of Bernoulli’s numbers 7

1 1 1 17
that of B22 = 1 − 2 − 3 − 23 = 138 ;
1 1 1 1 59
that of B28 = 2 + 3 + 5 + 29 −1= 870 ; and so on.

9. It can be inferred from (20) that:


If G be the G.C.M. and L the L.C.M. of the denominators of Bm and Bn , then L/G is
1
the denominator of Bm − (−1) 2 (m−n) Bn , and hence, if the denominators of Bm and Bn are
1
equal, then Bm − (−1) 2 (m−n) Bn is an integer. (21)
Example: B24 − B12 and B32 − B16 are integers, while the denominator of B10 + B20 is 5.
It will be observed that:
1
(1) If n is a multiple of 4, then the numerator of Bn − 30 in its lowest terms is divisible by
Bn 1
20; but if n is not a multiple of 4 then that of n − 12 in its lowest terms is divisible by 5;
(22)
(2) If n is any integer, then

B4n+2 B8n+4 B16n+8


2(24n+2 − 1) , 2(28n+4 − 1) , 2(28n+4 − 1)
2n + 1 2n + 1 2n + 1
are integers of the form 30p + 1. (23)

10. If a B is known to lie between certain limits, then it is possible to find its exact value
from the above properties.
Suppose we know that B22 lies between 6084 and 6244; its exact value can be found as
follows.
17
The fractional part of B22 = 138 by (20), also B22 is divisible by 11 by (18). And by (22)
11 17
B22 − 6 must be divisible by 5. To satisfy these conditions B23 must be either 6137 138 or
17
6192 138 .
But according to (18) the numerator of B22 should be a prime number after it is divided
17
by 11; and consequently B22 must be equal to 6192 138 or 854513
138 , since the numerator of
17
6137 138 is divisible not only by 11 but also by 7 and 17.

11. It is known (Edward’s Differential calculus, Ch. v, Ex.29) that


 
2 · n! 1 1 1
Bn = + + + ··· ,
(2π)n 1n 2n 3n
or    
2 · n! 1 1 1
= B n 1 − 1 − 1 − (24)
(2π)n 2n 3n 5n
where 2, 3, 5, . . . are prime numbers.
Also
Z∞ n−1
Bn x
= 2πx
dx. (25)
2n e −1
0
8 Paper 1

For
Z∞ Z∞
xn−1
dx = xn−1 (e−2πx + e−4πx + · · ·)dx
e2πx − 1
0 0
 
(n − 1)! 1 1 1 Bn
= + + + ··· =
(2π)n 1n 2n 3n 2n
by (24). In a similar manner
Z∞
xn Bn
dx = ,
(eπx − e−πx )2 4π
0

and
Z∞
πBn
xn−2 log(1 − e−2πx )dx = − . (26)
n(n − 1)
0

Take logarithms of both sides in (24) and write for loge n! the well known expansion of
loge Γ(n + 1), as in Carr’s Synopsis, viz.
B2 B4 B6
(n + 21 ) log n − n + 12 log 2π + − 3
+ − ···
1 · 2n 3 · 4n 5 · 6n5
B2p θ
− (−1)p , (27)
(2p − 1)2pn2p−1
where 0 < θ < 1, and where
B2p θ B2p B2p+2
= − + ···
(2p − 1)2pn2p−1 (2p − 1)2pn2p−1 (2p + 1)(2p + 2)n2p+1
Z∞ 2p−2 Z∞ 2p
1 x −2πx 1 x
= − log(1 − e )dx + log(1 − e−2πx )dx − · · ·
π n2p−1 π n2p+1
0 0
Z∞ 
x2p−2 x2p

1
= − − + · · · log(1 − e−2πx )dx
π n2p−1 n2p+1
0
Z∞
1 x2p−2
= − log(1 − e−2πx )dx
π n2p−3 (n2 + x2 )
0
Z∞
x2p−2 log(1 − e−2πnx )
= − dx.
π(1 + x2 )
0

We can find the integral part of Bn , and since the fractional part can be found, as shewn
in §8, the exact value of Bn is known. Unless the calculation is made to depend upon the
Some properties of Bernoulli’s numbers 9

values of log10 e, log e 10, π, . . . , which are known to a great number of decimal places, we
should have to find the logarithms of certain numbers whose values are not found in the
tables to as many places of decimals as we require. Such difficulties are removed by the
method given in §13.

12. Results (14) to (17), (20) and (21) can be obtained as follows. We have

1 1 1 1 1
+ + + + + ···
2x2 (x + 1)2 (x + 2)2 (x + 3)2 (x + 4)2
1 1 1 1 1
− − + + + + ···
x 6(x3 − x) 5(x5 − x) 7(x7 − x) 11(x11 − x)
1 7 55 529
= − + − 21 + · · · (28)
x15 x17 x19 x

where 5, 7, 11, 13, are prime numbers above 3. If we can prove that the left-hand side of
(28) can be expanded in ascending powers of 1/x with integral coefficients, then (20) and
(21) are at once deduced as follows.
From (27) we have

d2 log Γ(n + 1) 1 1 1
= + + + ···
dn2 (n + 1)2 (n + 2)2 (n + 3)2
1 1 B2 B4 B6 B8 B2p θ
= − 2 + 3 − 5 + 7 − 9 + · · · − (−)p 2p+1 , (29)
n 2n n n n n n

where
B2p θ B2p B2p+2
= − 2p+3 + · · ·
n2p+1 n2p+1 n
Z∞ Z∞
x2p x2p+2
= 4π dx − 4π dx + · · ·
n2p+1 (eπx − e−πx )2 n2p+3 (eπx − e−πx )2
0 0
Z∞ 
x2p x2p+2

dx
= π − 2p+3 + · · ·
n 2p+1 n sinh2 πx
0
Z∞ Z∞
x2p dx πx2p
= π = dx.
n 2p−1 (n + x ) sinh2 πx
2 2
(1 + x2 ) sinh2 πnx
0 0

Substituting the result of (29) in (28) we see that


B2 B4 B6 1 1 1 1
3
− 5 + 7 − ··· − 3
+ 5
+ 7
+ 11
+ ···,
x x x 6(x − x) 5(x − x) 7(x − x) 11(x − x)
10 Paper 1

where 5, 7, 11, . . . are prime numbers, can be expanded in ascending powers of 1/x with
integral coefficients.
Therefore B2 − 16 , −B4 − 61 + 15 , B6 − 61 + 17 , −B8 − 61 + 51 , B10 − 61 + 11
1
, · · · , which are the
3 5 7
coefficients of 1/x , 1/x , 1/x , · · · , are integers.
Writing 12 + 13 − 1 for − 16 we get the results of (20) and (21).
Again changing n to 12 n in (29), and subtracting half of the result from (29), we have

1 1 1 1 (22 − 1)B2 (24 − 1)B4


2
− 2
+ 2
− ··· = 2 − +
(n + 1) (n + 2) (n + 3) 2n n3 n5

(26 − 1)B6 B2p θ


− 7
+ · · · + (−1)p (22p − 1) 2p+1 , (30)
n n
where 0 < θ < 1, and also, by (29),
Z∞
2p B2p θ πx2p cosh πnx
(2 − 1) 2p+1 = dx.
n (1 + x2 ) sinh2 πnx
0

Thus we see that, if we can prove that twice the left hand side of (30) can be expanded in
ascending powers of 1/n with integral coefficients, then the second part of (16) is at once
proved.
Again from (27) we have

d log Γ(n + 1) 1 1 1 1
= 1+ 2 + 3 + −γ4 + ··· +
dn n
1 B2 B4 B6 B8 B2p θ
= log n + − 2 + 4 − 6 + 8 − · · · + (−1)p , (31)
2n 2n 4n 6n 8n 2pn2p

where 0 < θ < 1; and also, by (25),


Z∞
B2p θ 2x2p−1
= dx,
2pn2p (1 + x2 )(e2πnx − 1)
0

from which it can easily be shewn that

1 1 1 1 1
− + − + − ···
n+2 n+4 n + 6 n + 8 n + 10
1 B2 B4 B6
= − 2(22 − 1) 2 + 23 (24 − 1) 4 − 25 (26 − 1) 6 + · · ·
2n 2n 4n 6n
B 2p θ
+(−1)p 22p−1 (22p − 1) − ···, (32)
2pn2p
Some properties of Bernoulli’s numbers 11

where 0 < θ < 1; and also, by (31),

Z∞
2p−1 2p B2p θ x2p−1
2 (2 − 1) = dx.
2pn2p 2(1 + x2 ) sinh 21 (πnx)
0

From the above theorem we see that, if we can prove that


 
1 1 1
2 − + − ··· ,
n+2 n+4 n+6

can be expanded in ascending powers of 1/n with integral coefficients, then the first part
of (16) at once follows.

13. The first few digits, and the number of digits in the integral part as well as in the
numerator of Bn , can be found from the approximate formula:

log10 Bn = (n + 12 ) log10 n − 1.2324743503n + 0.700120,

the true value being greater by about 0.0362/n when n is great. (33)
This formula is proved as follows: taking logarithms of both sides in (24),

loge Bn = (n + 21 ) loge n − n(1 + loge 2π) + 21 loge 8π

nearly. Multiplying both sides by log10 e or .4342944819, and reducing, we can get the
result.

14. Changing n to n − 2 in (24) and taking the ratio of the two results, we have

22 − 1 32 − 1 52 − 1
   
n(n − 1)
Bn = Bn−2 1 − n 1− n 1− n ···, (34)
4π 2 2 −1 3 −1 5 −1

where 2, 3, 5, . . . are prime numbers.


Hence we see that BBn−2 n
approaches n(n−1)
4π 2
very rapidly as n becomes greater and greater.
(35)
From the value of π, viz. 3.14159, 26535, 89793, 23846, 26433, 83279, 50288, 41971, 69399,
. . ., the integral part of any B can be found from the previous B; and from the integral
part the exact value can at once be written by help of (20) as follows:
12 Paper 1

Approximate ratio of any lies between∗ Hence the exact value is


B to previous B
1 1 1
B2 ··· ··· 0 and 1 ··· 1− 2 − 3 = 6
3·4 1 1 1 1
B4 = B
4π 2 2
··· 0 and 1 ··· 2 + 3 + 5 −1 = 30
5·6 1 1 1 1
B6 = B
4π 2 4
··· 0 and 1 ··· 1− 2 − 3 − 7 = 42
7·8 1 1 1 1
B8 = 4π 2 B6 ··· 0 and 1 ··· 2 + 3 + 5 −1 = 30
9·10 1 1 1 5
B10 = 4π 2 B8 ··· 0 and 1 ··· 1− 2 − 3 − 11 = 66
11·12 1 1 1 1 1 691
B12 = 4π 2
B10 ··· 0 and 1 ··· 2 + 3 + 5 + 7 + 13 −1 = 2730
13·14 1 1
B14 = 4π 2
B12 ··· 0 and 2 ··· 2− 2 − 3 = 67
15·16 1 1 1 1
B16 = 4π 2
B14 ··· 7 and 8 ··· 6+ 2 + 3 +
17 5 + = 3617
510
17·18
B18 = 4π 2
B16 ··· 54 and 55 ··· 56 − 21 − 13 − 17 − 191 43867
= 798
19·20
B20 = 4π 2
B18 ··· 529 and 530 ··· 528 + 21 + 13 + 15 + 11 1
= 174611
330
··· ··· ···

15. From the preceding theorems we know some of the properties of Bn for all positive
even values of n. As an example let us take B444 = N/D.
The fractional part of B444 is 23975417
90709710 by (20). The numerator of B444 is divisible by 37
and the quotient is a prime number by (18). Again log10 B444 = 630 · 2433, nearly, by (33).
Therefore the integral part of B444 contains 631 digits, the first 4 digits being 1751. Again

log10 N = log10 B444 + log10 D = 630 · 2433 + log10 90709710 = 638 · 2010

nearly. Therefore N contains 639 digits, the first four digits being 1588.
1
Again the numerator of B444 − 30 is divisible by 20; that is to say, if [B444 ] is the integral
23975417 1 698392
part of B444 , [B444 ] + 90709710 − 30 = [B444 ] + 3023657 has a numerator divisible by 20.
Therefore the integral part of B444 ends with 4 and also the figure next to the last is even.
Hence N ends with 57 and also the third figure from the last is even.

16. Instead of starting with cot x as in §§ 2 and 3, we may start with tan x or cosec x and
get other similar results.
Thus

4 n
(i) 3 Bn (2 − 1) = c6 Bn−6 (2n−6 − 1) − c12 Bn−12 (2n−12 − 1) + · · ·
1 1 1
1 (n−2) 1 (n−4) 1 (n−6)
+ 3 n(−1)
6 or 6 n(−1)
6 or 3 n(−1)
6 ···, (36)

These integral limits are got from a rough calculation of any B from the preceding B by the formula
given in the first column.
Some properties of Bernoulli’s numbers 13

     
1 1 1
(ii) c3 1 − Bn−3 − c9 1 − Bn−9 + c15 1 − Bn−15 − · · ·
2n−4 2n−10 2n−16
n 1 (n−1) 1 1 1
= 2
3 · n
{3 2 + (−1) 6 (n−3) or (−1) 6 (n+1) or (−1) 6 (n+5) }. (37)
2

17. The formulæ obtained in §§1, 3, 4 may be called the one interval, two interval and
three interval formula respectively. The p interval formulæ can be got by taking the pth
roots of unity or of i according as p is odd or even.
For example, let us take the fifth roots of unity (1, α, α2 , α3 , α4 ), and find the 5 interval
formulæ.
Let
φ(x) = sin x + sin xα + sin xα2 + sin xα3 + sin xα4
 5
x15 x25

x
= 5 − + − ··· .
5! 15! 25!
Then it can easily be shewn that
16 sin x sin xα sin xα2 sin xα3 sin xα4 = φ(2x) − φ{2x(α + α4 )} − φ{2x(α2 + α3 )}
√ √
= φ(2x) + φ{x(1 + 5)} + φ{x(1 − 5)}.
Taking logarithms and differentiating both sides, we have
x4 x14

x10 x20 x30
 (1 + α5 ) − (1 + α15 ) + · · ·
5 B0 + B10 + B20 + B30 + ··· = −x 4!5 14! , (38)
10! 20! 30! x x15
(1 + α5 ) − (1 + α15 ) + · · ·
5! 15!
where
√ !n √ !n
1+ 5 1− 5
αn = + , so that αn αm = αn+m + (−1)n αm−n .
2 2

Similarly
x6 x16

x2 x12 x22
 (1 + α7 ) − (1 + α17 ) + · · ·
(i) 5 B2 + B12 + B22 + ··· = x 6!5 16! , (39)
2! 12! 22! x x15
(1 + α5 ) − (1 + α15 ) + · · ·
5! 15!
x8 x18

x4 x14 x24
 (α7 − 1) − (α17 − 1) + · · ·
(ii) 5 B4 + B14 + B24 + ··· = x 8!5 18! , (40)
4! 14! 24! x x15
(1 + α5 ) − (1 + α15 ) + · · ·
5! 15!
14 Paper 1

x6 x16 x26
 
(iii) 10 B6 + B16 + B26 + ···
6! 16! 26!

x10 x20 x30


(α10 − 3) − (α20 − 3) + (α30 − 3) − · · ·
= x 10!5 20! 30! , (41)
x x15 x25
(1 + α5 ) − (1 + α15 ) + (1 + α25 ) − · · ·
5! 15! 25!

x8 x18 x28
 
(iv) 5 B8 + B18 + B28 + ···
8! 18! 28!

x12 x22 x32


(α11 − 1) − (α21 − 1) + (α31 − 1) − · · ·
= x 12!5 22! 32! . (42)
x x15 x25
(1 + α5 ) − (1 + α15 ) + (1 + α25 ) − · · ·
5! 15! 25!

Again from

16 cos x cos xα cos xα2 cos xα3 cos xα4

= 1 + Ψ[2x] + Ψ[2x(α + α4 )] + Ψ[2x(α2 + α3 )]


where

Ψ(x) = cos x + cos xα + cos xα2 + cos α3 + cos xα4


x10 x20 x30
 
= 5 1− + − + ··· ,
10! 20! 30!

and similar relations, we can get many other identities.

18. The four interval formulæ can be got from the following identities: If
 n  n  n  n
1 1 1 1
an = 1+ √ + 1− √ and bn = 1+ √ − 1− √ ,
2 2 2 2

so that am+n = am an − am−n /2n ; and bm+n = am bn + bm−n /2n ; then:

x8 x16 x24
 
(i) 4 B0 + B8 + B16 + B24 + ···
8! 16! 24!

x3 x11 x19
a2 − a6 + a10 − · · ·
= −x 3!4 11! 19! , (43)
x x12 x20
a2 − a6 + a10 − · · ·
4! 12! 20!
Some properties of Bernoulli’s numbers 15

x2 x10 x18
 
(ii) 4 B2 + B10 + B18 + ···
2! 10! 18!

x5 x13 x21
a3 − a7 + a11 − · · ·
= −x 5!4 13! 21! , (44)
x x12 x20
a2 − a6 + a10 − · · ·
4! 12! 20!

√ x4 x12 x20
 
(iii) 4 2 B4 + B12 + B20 + ···
4! 12! 20!

x7 x15 x23
b3 − b7 + b11 − · · ·
= x 7!4 15! 23! , (45)
x x12 x20
a2 − a6 + a10 − · · ·
4! 12! 20!

√ x6 x14 x22
 
(iv) 4 2 B6 + B14 + B22 + ···
6! 14! 22!

x9 x17 x25
b4 − b8 + b12 − · · ·
= x 9!4 17! 25! . (46)
x x12 x20
a2 − a6 + a10 − · · ·
4! 12! 20!
On question 330 of Professor Sanjana
Journal of the Indian Mathematical Society, IV, 1912, 59 – 61

1. Prof. Sanjana remarks that it is not easy to evaluate the series


1 1 1 1·3 1 1·3·5 1
n
+ n
+ n
+ + · · · ad inf.,
1 23 2·45 2 · 4 · 6 7n
if n > 3. In attempting to sum the series for a⁀ ll values of n, I have arrived at the following
results:
Let
1 1 1 1·3 1
f (p) = + + + ···
1+p 23+p 2·45+p
Z1  
1 1·3 4
= xp 1 + x2 + x + · · · dx
2 2·4
0
Z1 Z1
xp 1 1
= √ dx = 21 x 2 (p−1) (1 − x)− 2 dx
1−x 2
0 0
   
p+1 1 p+1
Γ 2 Γ( 2 ) π2
1
Γ 2
= 12   =  .
Γ p+2 2 Γ p+2
2 2
1

p+1 π 2 Γ(p + 1)

But Γ 2= p  
2 Γ p+2
2
(vide Williamson, Integral Calculus, p. 164).
π Γ(p + 1)
Therefore f (p) = p+1   .
2 {Γ p+2 }2
2
Therefore
log{f (p)} = log( 21 π) − p log 2
p2 p3
   
1 1
+ 1− S2 − 1 − 2 S3 + · · · , (1)
2 2 3 2
where Sn ≡ 11n + 21n + 31n + · · · ad inf. (vide Carr’s Synopsis, 2295).
Again, by expanding f (p) in ascending powers of p, we have
   
11 1·31 1 1 1·3 1
f (p) = 1+ + + ··· − p 1 + + + ···
23 2·45 2 32 2 · 4 52
 
2 1 1 1·3 1
+p 1 + + + ··· − ···
2 33 2 · 4 53
π
= {φ(0) − pφ(1) + p2 φ(2) − p3 φ(3) + · · ·},
2
On question 330 of Professor Sanjana 17

where
1 1 1 1·3 1 π
+ + + · · · ≡ φ(n).
1n+1 23n+1 2·45n+1 2
Hence (1) may be written
log 21 π + log{φ(0) − p · φ(1) + p2 · φ(2) − · · ·}
p2 p3
   
1 1 1
= log( 2 π) − p log 2 + 1− S2 − 1 − 2 S3 + · · ·
2 2 3 2
p 2 p 3
= log( 21 π) − pσ1 + σ2 − σ3 + · · · ,
2 3
where
1 1 1
σn ≡ 1 − n
+ n − n + ···.
2 3 4
Differentiating with respect to p, and equating the coefficients of pn−1 , we have
nφ(n) ≡ σ1 φ(n − 1) + σ2 φ(n − 2) + σ3 φ(n − 3) + · · · to n terms.
Thus we see that
π 11 1·31 π
φ(0) ≡ 1 + + + ··· = ,
2 23 2·45 2
π 1 1 1·3 1 π
φ(1) ≡ 1 + + + ··· = (log 2),
2 2 32 2 · 4 52 2
π 1 1 1·3 1 π3 π
φ(2) ≡ 1 + + + ··· = + (log 2)2 ,
2 2 33 2 · 4 53 48 4
π 1 1 1·3 1 π3 π3 π
φ(3) ≡ 1 + + + ··· = log 2 + (log 2)3 + σ3
2 2 34 2 · 4 54 48 12 6
π3 π3 π
= log 2 + (log 2)3 + S3 ,
48 12 8
and so on.

2. More generally, consider the series


1 a 1 a(a − 1) 1
n
− n
+ − ···.
b 1! (b + 1) 2! (b + 2)n
Writing
Γ(b)Γ(a + 1)
φ(n − 1)
Γ(a + b + 1)
for this, and taking the identity
1 a 1 a(a − 1) 1
− + − ···
b + p 1! b + 1 + p 2! b+2+p
Z1
Γ(b + p)Γ(a + 1)
= xb+p−1 (1 − x)a dx = ,
Γ(a + b + p + 1)
0
18 Paper 2

we find
nφ(n) = σ1 φ(n − 1) + σ2 φ(n − 2) + σ3 φ(n − 3) + · · · to n terms,
where
1 1 1 1
σn ≡ − + − + ··· .
bn (a + b + 1)n (b + 1)n (a + b + 2)n
Examples: Put a = − 12 , b = 41 . Then we see that

11 1·31 1·3·5 1 {Γ( 41 )}2


(i) 1+ + + + ··· = p ,
25 2·49 2 · 4 · 6 13 4 (2π)
1 1 1·3 1 1·3·5 1 {Γ( 14 )}2 π
(ii) 1+ + + + · · · = ,
2 52 2 · 4 92 2 · 4 · 6 132
p
4 (2π) 4
{Γ( 14 )}2 π 2 1 ′
 
1 1 1·3 1 1·3·5 1
(iii) 1 + + + + · · · = + S 2 ,
2 53 2 · 4 93 2 · 4 · 6 133
p
4 (2π) 32 2
{Γ( 14 )}2 5π 3 π ′
 
1 1 1·3 1 1·3·5 1 1 ′
(iv) 1 + + + + ··· = p + S2 + S3 ,
2 54 2 · 4 94 2 · 4 · 6 134 4 (2π) 384 8 3

1 1 1 1
where Sr′ = − r + r − r + ···.
1r 3 5 7
Note on a set of simultaneous equations∗
Journal of the Indian Mathematical Society, IV, 1912, 94 – 96

1. Consider the equations


x1 + x2 + x3 + · · · +xn = a1 ,
x1 y1 + x2 y2 + x3 y3 + · · · +xn yn = a2 ,
x1 y12 + x2 y22 + x3 y32 + · · · +xn yn2 = a3 ,
x1 y13 + x2 y23 + x3 y33 + · · · +xn yn3 = a4 ,
... ... ... ... ... ... ...
x1 y12n−1 + x2 y22n−1 + x3 y32n−1 + ··· +xn yn2n−1 = a2n ,
where x1 , x2 , x3 , . . . , xn and y1 , y2 , y3 , . . . , yn are 2n unknown quantities.
Now, let us take the expression
x1 x2 x3 xn
φ(θ) ≡ + + + ··· + (1)
1 − θy1 1 − θy2 1 − θy3 1 − θyn
and expand it in ascending powers of θ. Then we see that the expression is equal to
a1 + a2 θ + a3 θ 2 + · · · + a2n θ 2n−1 + · · · . (2)
But (1), when simplified, will have for its numerator an expression of the (n − 1)th degree
in θ, and for its denominator an expression of the nth degree in θ.
Thus we may suppose that
A1 + A2 θ + A3 θ 2 + · · · + An θ n−1
φ(θ) = (3)
1 + B1 θ + B2 θ 2 + B3 θ 3 + · · · + Bn θ n
= a1 + a2 θ + a3 θ 2 + · · · a2n θ 2n−1 + · · · ;
and so (1 + B1 θ + · · ·)(a1 + a2 θ + · · ·) = A1 + A2 θ + · · ·.
Equating the coefficients of like powers of θ, we have
A1 = a1 ,
A2 = a2 + a1 B1 ,
A3 = a3 + a2 B1 + a1 B2 ,
An = an + an−1 B1 + an−2 B2 + · · · + a1 Bn−1 ,
0 = an+1 + an B1 + · · · + a1 Bn ,
0 = an+2 + an+1 B1 + · · · + a2 Bn ,
0 = an+3 + an+2 B1 + · · · + a3 Bn ,
.. ..
. .
0 = a2n + a2n−1 B1 + · · · + an Bn .

For a solution, by determinants, of a similar set of equations, see Burnside and Panton, Theory of
Equations, Vol, II, p.106, Ex.3. [Editor, J.Indian Math. Soc.]
20 Paper 3

From these B1 , B2 , . . . Bn can easily be found, and since A1 , A2 , . . . , An depend upon these
values they can also be found.
Now, splitting (3) into partial fractions in the form
p1 p2 p3 pn
+ + + ··· + ,
1 − q1 θ 1 − q2 θ 1 − q3 θ 1 − qn θ
and comparing with (1), we see that
x1 = p 1 , y 1 = q 1 ;
x2 = p 2 , y 2 = q 2 ;
x3 = p 3 , y 3 = q 3 ;
... ...
2. As an example we may solve the equations:
x + y + z + u + v = 2,
px + qy + rz + su + tv = 3,
p x + q 2 y + r 2 z + s2 u + t2 v = 16,
2

p3 x + q 3 y + r 3 z + s3 u + t3 v = 31,
p4 x + q 4 y + r 4 z + s4 u + t4 v = 103,
p5 x + q 5 y + r 5 z + s5 u + t5 v = 235,
p6 x + q 6 y + r 6 z + s6 u + t6 v = 674,
p7 x + q 7 y + r 7 z + s7 u + t7 v = 1669,
p8 x + q 8 y + r 8 z + s8 u + t8 v = 4526,
p9 x + q 9 y + r 9 z + s9 u + t9 v = 11595,
where x, y, z, u, v, p, q, r, s, t are the unknowns. Proceeding as before, we have
x y z u v
+ + + +
1 − θp 1 − θq 1 − θr 1 − θs 1 − θt
= 2 + 3θ + 16θ 2 + 31θ 3 + 103θ 4 + 235θ 5 + 674θ 6 + 1669θ 7 + 4526θ 8 + 11595θ 9 + · · ·
By the method of indeterminate coefficients, this can be shewn to be equal to
2 + θ + 3θ 2 + 2θ 3 + θ 4
.
1 − θ − 5θ 2 + θ 3 + 3θ 4 − θ 5
Splitting this into partial fractions, we get the values of the unknowns, as follows :
x= − 35 √, p= −1√
18+ 5 3+ 5
y= 10√ , q= 2√
18− 5 3− 5
z= 10 √ , r= √2
8+√ 5 5−1
u= − 2 5 , s= 2 ,
√ √
8−√ 5
v= 2 5
, t= − 5+1
2 .
Irregular numbers

Journal of the Indian Mathematical Society, V, 1913, 105 – 106

1. Let a2 , a3 , a5 , a7 , . . . denote numbers less than unity, where the subscripts 2,3,5,7, . . .
are the series of prime numbers. Then
1 1 1
· · . . . = 1 + a2 + a3 + a2 · a2 + a5
1 − a2 1 − a3 1 − a5
+a2 · a3 + a7 + a2 · a2 · a2 + a3 · a3 + . . . , (1)

the terms being so arranged that the products obtained by multiplying the subscripts are
the series of natural numbers 2, 3, 4, 5, 6, 7, 8, 9, . . . .
The above result is easily got if we remember that the natural numbers are formed by
multiplying primes and their powers.

2. Similarly, we have
1 1 1
· · . . . = 1 − a2 − a3 + a2 · a2 − a5
1 + a2 1 + a3 1 + a5
+a2 · a3 − a7 − a2 · a2 · a2 + a3 · a3 + . . . , (2)

where the sign is negative whenever a term contains an odd number of prime subscripts.

3. Put a2 = 1/2n , a3 = 1/3n , a5 = 1/5n , . . . in (1), and we get


    
1 1 1 1 1
1− n 1− n 1− n 1 − n ... = , (3)
2 3 5 7 Sn

where Sn denotes 1/1n + 1/2n + 1/3n + 1/4n + . . ..


Changing n into 2n in (3) and dividing by the original, we obtain
    
1 1 1 1 Sn
1+ n 1+ n 1+ n 1 + n ... = . (4)
2 3 5 7 S2n

Examples:

   
1 1 1 15
(i) 1+ 2 1+ 2 1+ 2 ... = , (5)
2 3 5 π2

   
1 1 1 105
(ii) 1+ 4 1+ 4 1+ 4 ... = , (6)
2 3 5 π4
22 Paper 4

since
S2 = π 2 /6, S4 = π 4 /90, S8 = π 8 /9450.

4. Subtract (2) from (1) and put a2 = 2−n . . . ; then


1 1 1 1 1 1 1 Sn2 − S2n
+ + + + + + + . . . = ,
2n 3n 5n 7n 8n 11n 12n 2Sn
where the numbers 2, 3, 5, 7, 8, . . . contain an odd number of prime divisors.
Examples:

1 1 1 1 1 π2
(i) + + + + + · · · = , (7)
22 32 52 72 82 20

1 1 1 1 1 π4
(ii) + + + + + · · · = . (8)
24 34 54 74 84 1260

5. Again (2,3,5,7, . . . being the prime numbers)

(1 + a2 )(1 + a3 )(1 + a5 )(1 + a7 ) · · · = 1 + a2 + a3 + a5


+a2 · a3 + a7 + a2 · a5 + a11 + a13 + · · · , (9)

where the product of the subscripts in any term is a natural number containing dissimilar
prime divisors; and

(1 − a2 )(1 − a3 )(1 − a5 )(1 − a7 ) . . . = 1 − a2 − a3 − a5 + a2 · a3 − a7 , (10)

where the signs are negative whenever the number of factors is odd.

6. Replacing as before a2 , a3 , a5 , . . . by the values given in § 3 and using (4), we deduce


that
1 1 1 1 Sn
1+ + n + n + n + ... = , (11)
2n 3 5 6 S2n
where 2, 3, 5, 6, 7, . . . are the numbers containing dissimilar prime divisors.

7. Also taking half the difference between (3) and (4),


1 1 1 1 1 1 1 1 1 1
+ n + + n+ n+ n+ n+ n+ n+ n
2n 3 5n 7 11 13 17 19 23 29
1 1 Sn2 − S2n
+ + + ... = , (12)
30n 31n 2Sn S2n
where 2, 3, 5, . . . are numbers containing an odd number of dissimilar prime divisors.
Irregular numbers 23

Examples:
1 1 1 9
(i) 2
+ 2 + 2 + ... = 2, (13)
2 3 5 2π

1 1 1 15
(ii) + + + ... = 4. (14)
24 34 54 2π

8. Subtracting (11) from Sn , we have


1 1 1 1 Sn (S2n − 1)
+ n + n + n + ... = , (15)
4n 8 9 12 S2n
where 4, 8, 9, . . . are composite numbers having at least two equal prime divisors.
Squaring the circle
Journal of the Indian Mathematical Society, V, 1913, 132
Let P QR be a circle with center O, of which a diameter is P R. Bisect P O at H and let
T be the point of trisection of OR nearer R. Draw T Q perpendicular to P R and place the
chord RS = T Q.
Join P S, and draw OM and T N parallel to RS. Place a chord P K = P M, and draw the
tangent P L = M N . Join RL, RK and KL. Cut off RC = RH. Draw CD parallel to KL,
meeting RL at D.

Q
S
N

P H O T R

L
C

Then the square on RD will be equal to the circle P QR approximately. For


5 2
RS 2 =
d ,
36
where d is the diameter of the circle. Therefore
31
P S 2 = d2 .
36
But P L and P K are equal to M N and P M respectively. Therefore
31 2 31 2
P K2 = d , and P L2 = d .
144 324
Squaring the circle 25

Hence
113 2
RK 2 = P R2 − P K 2 = d ,
144
and
355 2
RL2 = P R2 + P L2 = d .
324
But r
RK RC 3 113
= = ,
RL RD 2 355
and
3
RC = d.
4
Therefore r
d 355 √
RD = = r π, very nearly
2 113
Note: If the area of the circle be 140,000 square miles, then RD is greater than the true
length by about an inch.
Modular equations and approximations to π

Quarterly Journal of Mathematics, XLV, 1914, 350 – 372

1. If we suppose that
√ √ √ 1 √
n n n n/24
(1 + e−π )(1 + e−3π )(1 + e−5π ) · · · = 2 4 e−π Gn (1)

and
√ √ √ 1 √
n n n n/24
(1 − e−π )(1 − e−3π )(1 − e−5π ) · · · = 2 4 e−π gn , (2)

then Gn and gn can always be expressed as roots of algebraical equations when n is any
rational number. For we know that
1 1 1
(1 + q)(1 + q 3 )(1 + q 5 ) . . . = 2 6 q 24 (kk′ )− 12 (3)

and
1 1 1 1
(1 − q)(1 − q 3 )(1 − q 5 ) · · · = 2 6 q 24 k− 12 k′ 6 . (4)

Now the relation between the moduli k and l, which makes


K′ L′
n = ,
K L
where n = r/s, r and s being positive integers, is expressed by the modular equation of the
rsth degree. If we suppose that k = l′ , k′ = l, so that K = L′ , K ′ = L, then


q = e−πL /L = e−π n
,

and the corresponding value of k may be found by the solution of an algebraical equation.
From (1), (2), (3) and (4) it may easily be deduced that
1
g4n = 2 4 gn Gn , (5)

Gn = G1/n , 1/gn = g4/n , (6)

1
(gn Gn )8 (G8n − gn8 ) = . (7)
4
I shall consider only integral values of n. It follows from (7) that we need consider only one
of Gn or gn for any given value of n; and from (5) that we may suppose n not divisible by
4. It is most convenient to consider gn when n is even, and Gn when n is odd.
Modular equations and approximations to π 27

2. Suppose then that n is odd. The values of Gn and g2n are got from the same modular
equation. For example, let us take the modular equation of the 5th degree, viz.
 u 3  v 3  
2 2 1
+ =2 u v − 2 2 , (8)
v u u v
where
1 1
2 4 q 24 u = (1 + q)(1 + q 3 )(1 + q 5 ) · · ·
and
1 5
2 4 q 24 v = (1 + q 5 )(1 + q 15 )(1 + q 25 ) · · ·
By changing q to −q the above equation may also be written as
 v 3  u 3  
2 2 1
− =2 u v + 2 2 , (9)
u v u v
where
1 1
2 4 q 24 u = (1 − q)(1 − q 3 )(1 − q 5 ) · · ·
and
1 5
2 4 q 24 v = (1 − q 5 )(1 − q 15 )(1 − q 25 ) · · ·

If we put q = e−π/ 5 in (8), so that u = G 1 and v = G5 , and hence u = v, we see that
5

4
v − v −4 = 1.
Hence
√ √ !1
4 1+ 5 1+ 5 4
v = , G5 = .
2 2
q
2
−π
Similarly, by putting q = e 5 , so that u = g 2 and v = g10 , and hence u = 1/v, we see
5
that
v 6 − v −6 = 4.
Hence
√ s √
2 1 + 5 1+ 5
v = , g10 = .
2 2
Similarly it can be shewn that
√ ! 13
1+ 3 √ √ 1
G9 = √ , g18 = ( 2 + 3) 3 ,
2
√ ! u √
v v !
u
u 5 + 17 u 17 − 3
G17 = t + t ,
8 8
√ ! u √
v v !
u
u 7 + 17 u 17 − 1
g34 = t + t ,
8 8
28 Paper 6

and so on.

3. In order to obtain approximations for π we take logarithms of (1) and (2). Thus
1

24
π= √
n
log(2 4 Gn ) 
1 , (10)
24
π= √
n
log(2 4 gn ) 

24 −π n
approximately, the error being nearly √
n
e in both cases. These equations may also
be written as
√ 1 √ 1
eπ n/24
= 2 4 Gn , eπ n/24
= 2 4 gn (11)

In those cases in which G12 12


n and gn are simple quadratic surds we may use the forms
1 1
(G12 −12 12 12 −12 12
n + Gn ) , (gn + gn ) ,

instead of Gn and gn , for we have


1 1 π √n 3 − 1 π √n
gn12 = e2 − e 2 ,
8 2
approximately, and so
1 1 √ 13 1 √
gn12 + gn−12 = e 2 π n + e− 2 π n ,
8 2
approximately, so that
2
π = √ log{8(gn12 + gn−12 )}, (12)
n

the error being about 104
√ e−π n , which is of the same order as the error in the formulæ (10).
n
The formula (12) often leads to simpler results. Thus the second of formulæ (10) gives
√ 1
eπ 18/24
= 2 4 g18

or √ √ √
1
e4π 18
= 10 2 + 8 3.
But if we use the formula (12), or
√ 1 1
eπ n/24
= 2 4 (gn12 + gn−12 ) 12 ,

we get a simpler form, viz. √ √


1
e8π 18
= 2 7.

4. The values of g2n and Gn are obtained from the same equation. The approximation by
means of g2n is preferable to that by Gn for the following reasons.
Modular equations and approximations to π 29


(a) It is more accurate. Thus the error when√we use G65 contains a factor e−π 65 , whereas
that when we use g130 contains a factor e−π 130 .
(b) For many values of n, g2n is simpler in form than Gn ; thus

√ !)
v(

u
u 3 + 13
g130 = t (2 + 5) ,
2

while
v v
√ ! √ √ ! u √ !
( !) 1 u v 
4 u u
u
1+ 5 3 + 13 u t 9 + 65 u 1 + 65 
G65 = t +t .
2 2  8 8 

(c) For many values of n, g2n involves quadratic surds only, even when Gn is a root of
an equation of higher order. Thus G23 , G29 , G31 are roots of cubic equations, G47 , G79 are
those of quintic equations, and G71 is that of a septic equation, while g46 , g58 , g62 , g94 , g142
and g158 are all expressible by quadratic surds.

5. Since Gn and gn can be expressed as roots of algebraical equations with rational coeffi-
cients, the same is true of G24 24
n or gn . So let us suppose that

1 = agn−24 − bgn−48 + · · · ,

or
gn24 = a − bgn−24 + · · · .
But we know that
√ √ √
n 24 n n
64e−π gn = 1 − 24e−π + 276e−2π − ···,
√ √
64gn24 = eπ n − 24 + 276e−π n − ···,
√ √
64a − 64bgn−24 + · · · = eπ n − 24 + 276e−π n − ···,
√ √ √
−π n π n −π n
64a − 4096be + ··· = e − 24 + 276e − ···,

that is
√ √
eπ n
= (64a + 24) − (4096b + 276)e−π n
+ ··· (13)

Similarly, if
1 = aG−24
n − bG−48
n + ···,
then
√ √
eπ n
= (64a − 24) − (4096b + 276)e−π n
+ ··· (14)
30 Paper 6


From (13) and (14) we can find whether eπ n is very nearly an integer for given values of
n, and ascertain also the number of 9’s or 0’s in the decimal part. But if Gn and gn be
simple quadratic surds we may work independently as follows. We have, for example,
q √
g22 = (1 + 2).

Hence
√ √
24
64g22 = eπ 22
− 24 + 276e−π 22
− ···,

−24 −π 22
64g22 = 4096e + ···,

so that
24
√ √ √ √
64(g22 −24
+ g22 ) = eπ 22
− 24 + 4372e−π 22
+ · · · = 64{(1 + 2)12 + (1 − 2)12 }.

Hence √
eπ 22
= 2508951.9982 . . . .
Again
√ 1
G37 = (6 + 37) 4 ,

√ √
64G24
37 = e
π 37
+ 24 + 276e−π 37
+ ···,

37
64G−24
37 = 4096e−π − ···,

so that
√ √ √ 6 √
64(G24 −24
37 + G37 ) = e
π 37
+ 24 + 4372e−π 37
− · · · = 64{(6 + 37) + (6 − 37)6 }.

Hence √
eπ 37
= 199148647.999978 . . . .
Similarly, from
√ !
v
u
u 5 + 29
g58 = t ,
2

we obtain

 5 + √29 12 √
 ! !12 
√ √ 5− 29 
24
64(g58 −24
+ g58 ) = eπ 58
− 24 + 4372e−π 58
+ · · · = 64 + .
 2 2 

Hence √
eπ 58
= 24591257751.99999982 . . . .
Modular equations and approximations to π 31

6. I have calculated the values of Gn and gn for a large number of values of n. Many of
these results are equivalent to results given by Weber; for example,
√ √
4 3 + 13 1+ 5
G13 = , G25 = ,
2 2
6
√ √ √
g30 = (2 + 5)(3 + 10), G437 = 6 + 37,
1
q √ √
7 4 + (4 + 7) 5 + 29
2
G49 = , g58 = ,
2 2
√ √
2 (3 + 5)(1 + 2)
g70 = ,
2
√ ! u √ !
v v
u
u 9 + 73 u 1 + 73
G73 = t +t ,
8 8

√ ! √ !1
4
1+ 5 9+ 85
G85 = ,
2 2

√ ! u √ !
v v
u
u 13 + 97 u 5 + 97
G97 = t +t ,
8 8

2
√ √
g190 = (2 + 5)(3 + 10),
1 √ √ √ √ √ √
G2385 = (3 + 11)( 5 + 7)( 7 + 11)(3 + 5),
8
and so on. I have also many results not given by Weber. I give a complete table of new
− 14 − 14
p p
results. In Weber’s notation, Gn = 2 f { (−n)} and gn = 2 f1 { (−n)}.

TABLE I

1 1
q √ q √
g62 + = { (1 + 2) + (9 + 5 2)},
g62 2
√ ! √ !) u √ ! u √ !
v( v v 
u
u 1+ 5 3 + 13 t 1 + 65 + t 9 + 65
u u
G265 = t ,
2 2  8 8 

√ ! u √
v v !
q√ √ √ √ 1
u
2
 u 7 + 33 u 33 − 1 
g66 = ( 2 + 3)(7 2 + 3 11) 6 t +t ,
 8 8 
32 Paper 6

√ ! 16 u √ ! u √ !
v v 
√ √ 1 5 + 23 t 6+3 3 +t 2+3 3
u u
G269 = (3 3 + 23) 4 ,
4  4 4 

√ ! u √ !
v v 
1 √ √ √ 1
u
 u 6 + 11 u 2 + 11
G277 = { ( 7 + 11)(8 + 3 7)} 4 t +t ,
2  4 4 

√ 1
(2 3 + 2) 3 + 1
G381 = √ 1 ,
(2 3 − 2) 3 − 1
√ ! u √
v v !
√ √ √ 1
u
u 3+ 6 u 6−1 
g90 = {(2 + 5)( 5 + 6)} 6 t +t ,
 4 4 

1 1
q √ q √
g94 + = { (7 + 2) + (7 + 5 2)},
g94 2
1 1 √ √
q
g98 + = { 2 + (14 + 4 14)},
g98 2
√ ! u √ !
v v 
q√ √ √ √ 1 u 23 + 3 57
u
u 15 + 3 57 
2
g114 = ( 2 + 3)(3 2 + 19) 6 t +t ,
 8 8 

√ !1
1 3+ 13 4
√ √ 1 1 q √
G117 = (2 3 + 13) 6 {3 4 + (4 + 3)},
2 2
  1 ( 1  1 ) 
1 11 6 1 3 1 3
G121 + = 3+ √ + 3− √



G121 2 3 3 3 3 ,
1 1
√ 1 √ √ 1 √ √ 1 
= [(11 − 3 11) 3 {(3 11 + 3 3 − 4) 3 + (3 11 − 3 3 − 4) 3 } − 2]



G121 3 2

u √
v
√ !
v
√ ! u √
v ! 2
√ √
u
u 3+ 7 1
 u 3+ 2 u 2−1 
g126 = t ( 6 + 7) 6 t +t ,
2  4 4 

u √ √ ! √ ! u √ !
v v v 
√ √
u
u 3 3 + 23 1
 u 5+2 6 u 1+2 6
2
g138 = t (78 2 + 23 23) 6 × t +t ,
2  4 4 

√ ! 61 u √ ! u √ !
v v 
√ √ 1 7 + 47 t 18 + 9 3 + t 14 + 9 3
u u
G2141 = (4 3 + 47) 4 √ ,
2  4 4 
Modular equations and approximations to π 33

√ √ ) u √ √
v( v ! v !
u u
u (2 + 5)(5 + 29) u 17 + 145 u 9 + 145 
G2145 = t t +t ,
2  8 8 
" (s  )#
1 1
− 12 1 1 7 1
= 2 +√ − (28) 6 ,
G147 2 3 4
v
√ ! u √
v !2
uu 5 + 17 u 17 − 3 
G153 = t +t
 8 8 

1
√ ! u √ ! 3
v v
uu 37 + 9 17 u 33 + 9 17 
× t +t ,
 4 4 

√ √ !)
v(
√ √
u
2
u 7 + 11
g154 = t (2 2 + 7)
2

√ ! u √ !
v v 
uu 13 + 2 22 u 9 + 2 22 
× t +t ,
 4 4 

1 1
q √ q √
g158 + = { (9 + 2) + (17 + 13 2)},
g158 2

  1 ( 1   1 )2 
1 13 6 1 3 1 3 
G169 + = 1+ √ + 1− √


G169 4

3 3 3 3



√ ! 31
 

1 √

1 13 − 3 13


= ( 13 − 2) + ,
G169 3 2 


√ !1 √ !1
  

 √ 11 − 13 3 √ 11 − 13 3 



× 3 3− − 3 3+ 

2 2
  

√ ! u √ !
v v 
√ √ √ 1 u 9 + 33
u
q u 1 + 33 
g198 = (1 + 2)(4 2 + 33) 6 t +t ,
 8 8 

√ ! √ √ ! 14 u √ ! u √
v v !
1+ 5 3 5 + 41 t 7 + 41 + t
u u 41 − 1 
G205 = ,
2 2  8 8 
34 Paper 6

√ !1
√ √ 1 59 + 7 71 6
G2213 = (5 3 + 71) 4
4

√ ! u √ !
v v 
uu 21 + 12 3 u 19 + 12 3 
× t +t ,
 2 2 

√ ! u √ !
v v 
u
u 9+4 7 u 11 + 4 7 
G2217 = t +t
 2 2 

√ ! u √ !
v v 
uu 12 + 5 7 u 16 + 5 7 
× t +t ,
 4 4 

√ !
1+ 5 √ 1 q √ 1
G225 = (2 + 3) 3 { (4 + 15) + 15 4 },
4

√ ! u √ !
v v 
u
u 1+2 2 u 5+2 2 
g238 = t +t
 4 4 

√ ! u √ !
v v 
uu 1+3 2 u 5+3 2 
× t +t ,
 4 4 

√ √ √
v( !) v ! v !

u u u
u 7 + 53 
t 89 + 5 265 + t 81 + 5 265
u u 
G2265 = t (2 + 5) ,
2  8 8 

√ √ ) u( √ √ ) 2
v
u(
v 
u 17 + 17 + 17 14 (5 + 17) u 1 + 17 + 17 14 (5 + 17)
G289 = t +t  ,
16 16

( √ √ !) 41
√ 23 43 + 57 7
G2301 = (8 + 3 7)
2

√ ! u √ !
v v 
u
u 46 + 7 43 u 42 + 7 43 
× t +t ,
 4 4 
Modular equations and approximations to π 35

√ !q √ ! u √ !
v v 

u
1+ 5  u 7 + 2 10 u 3 + 2 10
g310 = (1 + 2) t +t ,
2  4 4 

√ !1 
4
3+ 13


G325 = t, where



2 


√ !2 √ !2

1 − 13 1 + 13 ,
t3 + t2 +t +1 
2 2



√ n  √   √  o 

= 5 t3 − t2 1+2 13 + t 1−2 13 − 1


1 √ 1 √ √ 1
q √ q √
G333 = (6 + 37) 4 (7 3 + 2 37) 6 { (7 + 2 3) + (3 + 2 3)},
2

5

G363 = 2 12 t, where 


√ √
q 

2t3 − t2 {(4 + 33) + (11 + 2 33)} ,

q √ 

− t{1 + (11 + 2 33)} − 1 = 0

√ √ !
 √ q √  q √ √ 1
3+ 7 √ 1
 2 + 7 + (7 + 4 7)   (3 + 7) + (6 7) 4 
G2441 = (2 + 3) 3 q √ √ 1 ,
2 2
(3 + 7) − (6 7) 4
 

v
√ √ ! u √ !
!1 u v 
q √ 21 + 445 4 u
u 13 + 89 u 5 + 89
G445 = (2 + 5) t +t ,
2  8 8 

√ ! √ √ !)
v(
√ √ √ 1
u
u 1+ 5 3 3 + 31
G2465 = t (2 + 3) (5 5 + 2 31) 6
2 2

√ ! u √ !
v v 
uu 2 + 31 u 6 + 31 
× t +t
 4 4 

√ ! u √ !
v v 
uu 11 + 2 31 u 13 + 2 31 
× t +t ,
 2 2 
36 Paper 6

√ !
v( )
√ u √
u
1+ 5
G2505 = (2 + 5)t (10 + 101)
2

 5√5 + √101 √
 ! v
u !
u 105 + 505 
× +t ,
 4 8 

√ ! √ ! u √ !
v v v 
√ √ 1 u 9 + 3 6
u u
u 5 + 29 u 5+3 6 
g522 = t (5 29 + 11 6) 6 t +t ,
2  4 4 

√ ! u √ !
v v 
uu 96 + 11 79 u 100 + 11 79 
G2553 = t +t
 4 4 

√ ! u √ !
v v 
uu 141 + 16 79 u 143 + 16 79 
× t +t ,
 2 2 

√ ! √ √ !)
v(
√ √ 1u √
u
3+ 5 3+ 7
g630 = ( 14 + 15) 6 t (1 + 2)
2 2

√ √ u √ √
v ! v !
uu 15 + 7 + 2 u 15 + 7 − 2 
× t +t
 4 4 

√ √ u √ √
v ! v !
uu 15 + 7 + 4 u 15 + 7 − 4 
× t +t ,
 8 8 

√ ! √ !)
v(
√ √
u
3+ 5 1u 9+ 85
G2765 = (16 + 255) 6 t (4 + 15)
2 2

√ ! u √ !
v v 
uu 6 + 51 u 10 + 51 
× t +t
 4 4 

√ ! u √ !
v v 
uu 18 + 3 51 u 22 + 3 51 
× t +t ,
 4 4 
Modular equations and approximations to π 37

√ √ !)
v(
√ √ √ √
u
u 3+ 7 1
G2777 = t (2 + 3)(6 + 37) (246 7 + 107 37) 6
2

√ ! u √ !
v v 
uu 6+3 7 u 10 + 3 7 
× t +t
 4 4 

√ ! u √ !
v v 
uu 15 + 6 7 u 17 + 6 7 
× t +t ,
 2 2 

√ !
 1 q √  32
1+ 5 √ 1 7 + (4 + 7) 
 4
G1225 = (6 + 35) 4
2  2 

v
u 43 + 15√7 + (8 + 3√7) (10√7) 
u q 
u
× t
 8 

v
u 35 + 15√7 + (8 + 3√7) (10√7) 
u q 
u
+t ,

 8 


v( !)
√ √
u
u 11 + 123
G21353 = t (3 + 11)(5 + 3 3)
2

√ !1
6
6817 + 321 451
×
4

√ ! u √ !
v v 
uu 17 + 3 33 u 25 + 3 33 
× t +t
 8 8 

√ ! u √ !
v v 
uu 561 + 99 33 u 569 + 99 33 
× t +t ,
 8 8 

√ !) √ √
v( !1
√ u √
u
7+ 47 73 5 + 9 329 4
G21645 = (2 + 5)t (3 + 7)
2 2
38 Paper 6

√ √
v ! v !
uu 119 + 7 329
u
u 127 + 7 329 
× t +t
 8 8 

√ √
v ! v !
uu 743 + 41 329
u
u 751 + 41 329 
× t +t .
 8 8 

7. Hence we deduce the following approximate formulæ

TABLE II

1
√ √ √ √ 1
√ √ √
e8π = 2 7, eπ 22/12 = 2 + 2, e 4 π 30 = 20 3 + 16 6,
18

1
√ √ 1
√ √
e 4 π 34 = 12(4 + 17), e 2 π 46 = 144(147 + 104 2)
√ √
1

π 42
√ π 58/12 5 + 29
e 4 = 84 + 32 6, e = √ ,
2
1
√ √ √ 1
√ √ √
e 4 π 70 = 60 35 + 96 14, e 4 π 78 = 300 3 + 208 6,
q √
√ 1 + (3 + 2 5) 1
√ √ √
eπ 55/24 = √ , e 4 π 102 = 800 3 + 196 51,
2
1
√ √ √ √ √ √
e 4 π 130 = 12(323 + 40 65), eπ 190/12 = (2 2 + 10)(3 + 10),
( √ √ )
12 (2 + 5)(3 + 13)
π = √ log √ ,
130 2

√ ! u √ !
v v 
u
24  u 10 + 11 2 u 10 + 7 2
π = √ log t +t ,
142  4 4 

12 √ √ √
π = √ log{(2 2 + 10)(3 + 10)},
190
12 1 √ √ √ q √
π = √ log[ (3 + 5)(2 + 2){(5 + 2 10) + (61 + 20 10)}],
310 4
√ !3

4 5 + 29 √ √
π = √ log  √ (5 29 + 11 6)
522 2
Modular equations and approximations to π 39

√ ! 6
 
√ ! u
v v
uu 9+3 6 u 5+3 6  
× t +t .
 4 4 

The last five formulæ are correct to 15, 16, 18, 22 and 31 places of decimals respectively.

8. Thus we have seen how to approximate to π by means of logarithms of surds. I shall


now shew how to obtain approximations in terms of surds only. If

K′ L′
n = ,
K L
we have
ndk dl
2
= ′2 2 .
kk K
′2 ll L
But, by means of the modular equation connecting k and l, we can express dk/dl as an al-
gebraic function of k, a function moreover in which all coefficients which occur are algebraic
numbers. Again,
′ ′
q = e−πK /K , q n = e−πL /L ,

1  1 s
q 12 (1 − q 2 )(1 − q 4 )(1 − q 6 ) · · ·
 
kk′ 6 K
1 = . (15)
q 12 n (1 − q 2n )(1 − q 4n )(1 − q 6n ) · · · ll′ L

Differentiating this equation logarithmically, and using the formula

dq π2q
= ,
dk 2kk′2 K 2
we see that
 2n
2q 4n
 2
2q 4
   
q q
n 1 − 24 + + ··· − 1 − 24 + + ···
1 − q 2n 1 − q 4n 1 − q2 1 − q4
KL
= A(k), (16)
π2
where A(k) denotes an algebraic function of the special class described above. I shall use
the letter A generally to denote a function of this type.
Now, if we put k = l′ and k′ = l in (16), we have
  
1 2
n 1 − 24 2π√n + √ + ···
e − 1 e4π n − 1
    2
1 2 K
− 1 − 24 2π/√n + 4π/√n + ··· = A(k). (17)
e −1 e −1 π
40 Paper 6

The algebraic function A(k) of course assumes a purely numerical form when we substitute
the value of k deduced from the modular equation. But by substituting k = l′ and k′ = l
in (15) we have
1 √ √ √ √
n/12 n n n
n 4 e−π (1 − e−2π )(1 − e−4π )(1 − e−6π )···
√ √ √ √
n) n n n
= e−π/(12 (1 − e−2π/ )(1 − e−4π/ )(1 − e−6π/ )···

Differentiating the above equation logarithmically we have


  
1 2
n 1 − 24 2π√n + √ + ···
e − 1 e4π n − 1
   √
1 2 6 n
+ 1 − 24 2π/√n + 4π/√n + ··· = . (18)
e −1 e −1 π

Now, adding (17) and (18), we have


   2
3 1 2 K
1 − √ − 24 2π n √ + 4π n
√ + ··· = A(k). (19)
π n e −1 e −1 π
But it is known that
2
3q 3 5q 5
  
q 2K
1 − 24 + + + ··· = (1 − 2k2 ),
1 + q 1 + q3 1 + q5 π
so that
   2
1 3 K
1 − 24 √ + √ + ··· = A(k). (20)
eπ n +1 e3π n +1 π

Hence, dividing (19) by (20), we have


 
3 1 2
1 − π√ n
− 24 √
e2π n −1
+ √
e4π n −1
+ · · ·
  = R, (21)
1 − 24 eπ√1n +1 + e3π√3n +1 + · · ·

where R can always be expressed in radicals if n is any rational number. Hence we have
3
√ , π= (22)
(1 − R) n
√ √
nearly, the error being about 8πe−π n (π n − 3).

9. We may get a still closer approximation from the following results. It is known that
r=∞ 4
r 3 q 2r

X 2K
1 + 240 = (1 − k2 k′2 ),
1 − q 2r π
r=1
Modular equations and approximations to π 41

and also that


r=∞ 6
r 5 q 2r

X 2K 1
1 − 504 = (1 − 2K 2 )(1 + k2 k′2 ).
r=1
1 − q 2r π 2
Hence, from (19), we see that
r=∞ r=∞
( )( )
3 X r X r3
1 − √ − 24 √ 1 + 240 √
π n e2πr n − 1 e2πr n − 1
r=1 r=1
r=∞
( )

X r5
= R 1 − 504 √ , (23)
r=1
e2πr n −1

where R′ can always be expressed in radicals for any rational value of n. Hence
3
π= √ , (24)
(1 − R′ ) n
√ √
nearly, the error being about 24π(10π n − 31)e−2π n
It will be seen that the error in (24) is much less than that in (22), if n is at all large.

10. In order to find R and R′ the series in (16) must be calculated in finite terms. I shall
give the final results for a few values of n.

Table III
′ ′
q = e−πK /K , q n = e−πL /L ,

! ∞
!
X q 2mn X q 2m
f (q) = n 1 − 24 − 1 − 24 ,
1
1 − q 2mn 1
1 − q 2m

4KL ′
f (2) = (k + l),
π2
4KL
f (3) = (1 + kl + k′ l′ ),
π2
4KL √ ′ √ 2
f (4) = ( k + l) ,
π2
s 
4KL ′ ′ 1 + kl + k′ l′
f (5) = (3 + kl + k l ) ,
π2 2

12KL
f (7) = (1 + kl + k′ l′ ),
π2
8KL ′ ′
p p p
f (11) = {2(1 + kl + k l ) + (kl) + (k ′ l′ ) − (kk′ ll′ )},
π2
42 Paper 6

4KL 1 1
f (15) = 2
[{1 + (kl) 4 + (k′ l′ ) 4 }4 − {1 + kl + k′ l′ }],
π
4KL √
f (17) = {44(1 + k2 l2 + k′2 l′2 ) + 168(kl + k′ l′ − kk′ ll′ )
π2
1 2
− 102(1 − kl − k′ l′ )(4kk′ ll′ ) 3 − 192(4kk′ ll′ ) 3 },
24KL p p p
f (19) = 2
{(1 + kl + k′ l′ ) + (kl) + (k′ l′ ) − (kk′ ll′ )},
π
4KL 1 p p 1
f (23) = 2
[11(1 + kl + k′ l′ ) − 16(4kk′ ll′ ) 6 {1 + (kl) + (k′ l′ )} − 20(4kk′ ll′ ) 3 ],
π
12KL p p p
f (31) = 2
[3(1 + kl + k′ l′ ) + 4{ (kl) + (k′ l′ ) + (kk′ ll′ )}
π
1 1 1
− 4(kk′ ll′ ) 4 {1 + (kl) 4 + (k′ l′ ) 4 }],
4KL p p p
f (35) = 2
[2{ (kl) + (k′ l′ ) − (kk′ ll′ )}
π
1
+ (4kk′ ll′ )− 6 {1 − (kl) − (k′ l′ )}3 ].
p p

Thus the sum of the series (19) can be found in finite terms, when n = 2, 3, 4, 5, . . . ,
from the equations in Table III. We can use the same table to find the sum of (19) when
n = 9, 25, 49, . . . ; but then we have also to use the equation
 
3 1 2 3
= 1 − 24 2π + 4π + 6π + ··· ,
π e −1 e −1 e −1

which is got by putting k = k′ = 1/ 2 and n = 1 in (18).
Similarly we can find the sum of (19) when n = 21, 33, 57, 93, . . . , by combining the values
of f (3) and f (7), f (3) and f (11), and so on, obtained from Table III.

11. The errors in (22) and (24) being about



n √ √ √
8πe−π (π n − 3), 24π(10π n − 31)e−2π n ,

we cannot expect a high degree of approximation for small values of n. Thus, if we put
n = 7, 9, 16, and 25 in (24), we get
19 √
7 = 3.14180 . . . ,
16
√ !
7 3
1+ = 3.14162 . . . ,
3 5
 
99 7
√ = 3.14159274 . . . ,
80 7 − 3 2
Modular equations and approximations to π 43

√ !
63 17 + 15 5
√ = 3.14159265380 . . . ,
25 7 + 15 5

while

π = 3.14159265358 . . . .

But if we put n = 25 in (22), we get only

r
9 9
+ = 3.14164 . . . .
5 5

12. Another curious approximation to π is

 41
192

2
9 + = 3.14159265262 . . . .
22

This value was obtained empirically, and it has no connection with the preceding theory.
The actual value of π, which I have used for purposes of calculation, is

 
355 .0003
1− = 3.1415926535897943 . . . ,
113 3533

which is greater than π by about 10−15 . This is obtained by simply taking the reciprocal
of 1 − (113π/355).
In this connection it may be interesting to note the following simple geometrical construc-
tions for π. The first merely gives the ordinary value 355/113. The second gives the value
1
(92 + 192 /22) 4 mentioned above.
(1) Let AB (Fig.1) be a diameter of a circle whose centre is O. Bisect AO at M and trisect
OB at T . Draw T P perpendicular to AB and meeting the circumference at P . Draw a
chord BQ equal to P T and join AQ. Draw OS and T R parallel to BQ and meeting AQ
at S and R respectively. Draw a chord AD equal to AS and a tangent AC = RS. Join
BC, BD, and CD; cut off BE = BM, and draw EX, parallel to CD, meeting BC at X.
44 Paper 6

P
Q
R

A M O T B

X
C
E
D

Fig. 1.
Then the square on BX is very nearly equal to the area of the circle, the error being less
than a tenth of an inch when the diameter is 40 miles long.
(2) Let AB (Fig.2) be a diameter of a circle whose centre is O. Bisect the arc ACB at C
and trisect AO at T . Join BC and cut off from it CM and M N equal to AT . Join AM
and AN and cut off from the latter AP equal to AM . Through P draw P Q parallel to M N
and meeting AM at Q. Join OQ and through T draw T R, parallel to OQ and meeting AQ
at R. Draw AS perpendicular to AO and equal to AR, and join OS.
C
Q M
N
R P

A B
T O

Fig. 2.
Then the mean proportional between OS and OB will be very nearly equal to a sixth of
the circumference, the error being less than a twelfth of an inch when the diameter is 8000
miles long.
Modular equations and approximations to π 45

13. I shall conclude this paper√by giving a few series for 1/π.
It is known that, when k ≤ 1/ 2,

2K 2
 3
1·3 3
   
1
=1+ (2kk′ )2 + (2kk′ )4 + · · · (25)
π 2 2·4
Hence we have
1
q 3 (1 − q 2 )4 (1 − q 4 )4 (1 − q 6 )4 · · ·
 2 (  3  3 )
1 ′ 3 1 1 · 3
= kk 1+ (2kk′ )2 + (2kk′ )4 + · · · . (26)
4 2 2·4

Differentiating both sides in (26) logarithmically with respect to k, we can easily shew that
 2
2q 4 3q 6

q
1 − 24 + + + ···
1 − q2 1 − q4 1 − q6
(  3  3 )
1 1 · 3
= (1 − 2k2 ) 1 + 4 (2kk′ )2 + 7 (2kk′ )4 + · · · . (27)
2 2·4

But it follows from (19) that, when q = e−π n , n being a rational number, the left-hand
side of (27) can be expressed in the form

2K 2 B
 
A + ,
π π

where A and B are algebraic numbers expressible by surds. Combining (25) and (27) in
such a way as to eliminate the term (2K/π)2 , we are left with a series for 1/π. Thus, for
example,

7 1 3 13 1 · 3 3 19 1 · 3 · 5 3
     
4
=1+ + 2 + 3 + ···,
π 4 2 4 2·4 4 2·4·6
√ 1
3
(q = e−π , 2kk′ = ), (28)
2

 3
89 1 · 3 3 131 1 · 3 · 5 3
   
16 47 1
=5+ + 2 + 3 + ···,
π 64 2 64 2·4 64 2·4·6
√ 1
7
(q = e−π , 2kk′ = ), (29)
8

√ √ !8
√ 47 5 + 29 1 3
 
32 5−1
= (5 5 − 1) +
π 64 2 2
46 Paper 6

√ !16√
89 5 + 59 1 · 3 3
 
5−1
+ + ···,
642 2·4 2
"

√ !#
1 5 − 1
q = e−π 15 , 2kk′ = ; (30)
8 2
√ √ √
here 5 5 − 1, 47 5 + 29, 89 5 + 59, . . . are in arithmetical progression.

14. The ordinary modular equations express the relations which hold between k and l
when nK ′ /K = L′ /L, or q n = Q, where
′ /K ′
q = e−πK , Q = e−πL /L ,
 2
1·3 2 4
 
1 2
K =1+ k + k + ···.
2 2·4
There are corresponding theories in which q is replaced by one or other of the functions

√ ′
√ ′
2/K1
q1 = e−πK1 , q2 = e−2πK2 /(K2 3)
, q3 = e−2πK3 /K3 ,

where
1 · 3 2 1 · 3 · 5 · 7 4 1 · 3 · 5 · 7 · 9 · 11 6
K1 = 1 + k + k + k + ···,
42 42 · 82 42 · 82 · 122
1·2 2 1·2·4·5 4 1·2·4·5·7·8 6
K2 = 1 + k + k + k + ···
32 32 · 62 32 · 62 · 92
1 · 5 2 1 · 5 · 7 · 11 4 1 · 5 · 7 · 11 · 13 · 17 6
K3 = 1 + k + k + k + ···.
62 62 · 122 62 · 122 · 182
From these theories we can deduce further series for 1/π, such as
   2
27 112 2 1·31·42·5 2
= 2 + 17 + 32 + ···, (31)
4π 233 27 2·43·63·6 27


4 2
   
15 3 112 4 1·31·42·5
= 4 + 37 + 70 + ···, (32)
2π 2 3 3 125 2 · 4 3 · 6 3 · 6 125


4 2
   
5 5 115 4 1 · 3 1 · 7 5 · 11
√ = 1 + 12 + 23 + ···, (33)
2π 3 2 6 6 125 2 · 4 6 · 12 6 · 12 125


115 4 3 1 · 3 1 · 7 5 · 11 4 6
   
85 85
√ = 8 + 141 + 274 + ···, (34)
18π 3 2 6 6 85 2 · 4 6 · 12 6 · 12 85
Modular equations and approximations to π 47

4 3 23 1 1 · 3 43 1 · 3 1 · 3 · 5 · 7
= − 3 + 5 − ···, (35)
π 2 2 2 42 2 2 · 4 42 · 82

4 3 31 1 1 · 3 59 1 · 3 1 · 3 · 5 · 7
√ = − 3 2 42
+ 2 5 − ···, (36)
π 3 4 3 · 4 3 · 4 2 · 4 42 · 82

4 23 283 1 1 · 3 543 1 · 3 1 · 3 · 5 · 7
= − + 5 − ···, (37)
π 18 183 2 42 18 2 · 4 42 · 82

4 41 685 1 1 · 3 1329 1 · 3 1 · 3 · 5 · 7
√ = − 3 2
+ 2 − ···, (38)
π 5 72 5 · 72 2 4 5 · 725 2 · 4 42 · 82

4 1123 22583 1 1 · 3 44043 1 · 3 1 · 3 · 5 · 7


= − + − ···, (39)
π 882 8823 2 42 8825 2 · 4 42 · 82


2 3 9 1 1 · 3 17 1 · 3 1 · 3 · 5 · 7
=1+ + 2 + ···, (40)
π 9 2 42 9 2 · 4 42 · 82

1 1 11 1 1 · 3 21 1 · 3 1 · 3 · 5 · 7
√ = + 3 + 5 + ···, (41)
2π 2 9 9 2 42 9 2 · 4 42 · 82

1 3 43 1 1 · 3 83 1 · 3 1 · 3 · 5 · 7
√ = + 3 2
+ 5 + ···, (42)
3π 3 49 49 2 4 49 2 · 4 42 · 82

2 19 299 1 1 · 3 579 1 · 3 1 · 3 · 5 · 7
√ = + + 5 + ···, (43)
π 11 99 993 2 42 99 2 · 4 42 · 82

1 1103 27493 1 1 · 3 53883 1 · 3 1 · 3 · 5 · 7


√ = + + + ···. (44)
2π 2 992 996 2 42 9910 2 · 4 42 · 82

In all these series the first factors in each term form an arithmetical progression; e.g. 2,
17, 32, 47, . . . , in (31), and 4, 37, 70, 103, . . . , in (32). The first two series belong to the
theory of q2 , the next two to that of q3 , as the rest to that of q1 .
The last series (44) is extremely rapidly convergent. Thus, taking only the first term, we
see that
1103
= .11253953678 . . . ,
992
1
√ = .11253953951 . . . .
2π 2
48 Paper 6

15. In concluding this paper I have to remark that the series


q2 2q 4 3q 6
 
1 − 24 + + + ··· ,
1 − q2 1 − q4 1 − q6

which has been discussed in §§ 8-13, is very closely connected with the perimeter of an
ellipse whose eccentricity is k. For, if a and b be the semi-major and the semi-minor axes,
it is known that
12 · 3 4 12 · 32 · 5 6
 
1 2
p = 2πa 1 − 2 k − 2 2 k − 2 4 2 k − · · · , (45)
2 2 ·4 2 ·4 ·6

where p is the perimeter and k the eccentricity. It can easily be seen from (45) that
 
′2 dK
p = 4ak K +k . (46)
dk

But, taking the equation


1 1
q 12 (1 − q 2 )(1 − q 4 )(1 − q 6 ) · · · = (2kk′ ) 6
p
(K/π),

and differentiating both sides logarithmically with respect to k, and combining the result
with (46) in such a way as to eliminate dK/dk, we can shew that
 2
2q 4
  
4a 2 ′2 1 2 q
p= K (1 + k ) + ( π) 1 − 24 + + ··· . (47)
3K 2 1 − q2 1 − q4

But we have shewn


√ already that the right-hand side of(47) can be expressed in terms of
K if q = e−π n , where n is any rational number. It can also √be shewn that √ K can be
expressed in terms of Γ-functions if q be of the forms e −πn ,e −πn 2 and e−πn 3 , where n is
rational. Thus, for example, we have

k = sin π4 , q = e−π ,



q  1  
Γ( 4 ) Γ( 4 )
3


π
p=a + ,


2 Γ( 4 )
3
Γ( 4 )
5 




π 2
−π

k = tan 8 , q=e ,




  

Γ( 8 )
1
Γ( 8 )
5
q  
π

p=a 4 Γ( 8 )
5 + Γ( 8 )
9 , 


π
√ (48)
k = sin 12 , q = e−π 3 , 

r    

( ) Γ( 65 )
Γ 1 
π 1
p=a 1 + + 2 ,
3

√ √ 
3 Γ( 6 )5
Γ( 3 )
1

3 


b 2 π −2π

a = tan 8 , q = e 



  

( 1
) ( 3
)
q 
Γ Γ

π 1
p = (a + b) + ,
4 4


2 2 Γ( )
3
( 5
)

4
Γ 4
Modular equations and approximations to π 49

and so on.

16. The following approximations for p were obtained empirically:


p
p = π[3(a + b) − {(a + 3b)(3a + b)} + ǫ], (49)

where ǫ is about ak 12 /1048576;


( )
3(a − b)2
p=π (a + b) + p +ǫ , (50)
10(a + b) + (a2 + 14ab + b2 )

where ǫ is about 3ak 20 /68719476736.


Rx tan−1 t
On the integral t dt
0

Journal of the Indian Mathematical Society, VII,1915, 93 – 96

1. Let
Zx
tan−1 t
φ(x) = dt. (1)
t
0

Then it is easy to see that

φ(x) + φ(−x) = 0; (2)

and that
x x3 x5 x7
φ(x) = − + 2 − 2 + ... (3)
12 32 5 7
provided that |x| ≤ 1.
Changing t into 1/t in (1), we obtain
 
1 1
φ(x) − φ = π log x, (4)
x 2

provided that the real part of x is positive.


The results in the following two sections can be very easily proved by differentiating both
sides with respect to x.

2. If 0 < x < 21 π, then


sin 2x sin 6x sin 10x
+ + + . . . = φ(tan x) − x log(tan x). (5)
12 32 52

If, in particular, we put x = 81 π and 1


12 π in (5), we obtain

1 1 1 1 1 √ √ π √
2
+ 2 − 2 − 2 + 2 + · · · = 2φ( 2 − 1) + √ log(1 + 2); (6)
1 3 5 7 9 4 2
and
√ 1 √
2φ(1) = 3φ(2 − 3) + π log(2 + 3). (7)
4
Rx tan−1 t
On the integral t
dt 51
0

If − 21 π < x < 12 π, then


 x 2 sin3 x 2 · 4 sin5 x
2φ tan = sin x + + + ···. (8)
2 3 3 3·5 5
If 0 < x < 12 π, then

sin x sin 2x sin 3x


2
cos x + 2
cos2 x + cos3 x + · · ·
1 2 32
1
= φ(tan x) + π log cos x − x log sin x; (9)
2
and
cos x + sin x 1 cos3 x + sin3 x 1 · 3 cos5 x + sin5 x
+ + + ···
12 2 32 2·4 52
1
= φ(tan x) + π log(2 cos x). (10)
2
If − 12 π < x < 12 π and a be any number such that
 
1
|(1 − a) sin x| ≤ 1, 1 −
cos x ≤ 1,
a

then

1 2 1 3
     
sin x 1 sin 2x sin 3x
1− cos x + 1− cos2 x + 1 − cos3 x + · · ·
12 a 22 a 32 a
sin(x + 12 π) sin 2(x + 12 π)
+ (1 − a) sin x − (1 − a)2 sin2 x + · · ·
12 22
= φ(tan x) − φ(a tan x) + x log a. (11)

3. Let R(x) and I(x) denote the real and the imaginary parts of x respectively. Then, if
−1 < R(x) < 1,

x2 x2 x2
     
log 1 − 2 − 3 log 1 − 2 + 5 log 1 − 2 − · · ·
1 3 5
4 1 1
= [φ(1) − φ{tan π(1 − x)}] + log tan π(1 − x). (12)
π 4 4
2
Putting x = 3 in (12) and using (7), we obtain
5  9

  −3  −7 
4 4 4 4 4 2
1− 2 1− 2 1− 2 1− 2 1− 2 · · · = (2 − 3) 3 en ,
3 9 15 21 27
52 Paper 7

where
4
φ(1) n= (13)

Again, subtracting log(1 − x) from both sides in (12) and making x → 1, we obtain
 −3  5  −7  9
1 1 1 1 π 3n
1− 2 1− 2 1− 2 1− 2 ··· = e . (14)
3 5 7 9 8

If −1 < I(x) < 1, then

x2 x2 x2
     
log 1 + 2 − 3 log 1 + 2 + 5 log 1 + 2 − · · ·
1 3 5
4 1 1
{φ(1) − φ(e− 2 πx )} − 2x tan−1 e− 2 πx .
= (15)
π

From this and (7) we see that, if 21 πx = log(2 + 3), then
5 
x2 x2 x2 x2
  −3  −7
1+ 2 1+ 2 1+ 2 1+ 2 · · · = en , (16)
1 3 5 7

where n is the same as in (13).


It follows at once from (12) and (15) that, if −1 < R(β) < 1, − 1 < I(α) < 1, then
3  5  7
12 + α2 32 − β 2 52 + α2 72 − β 2
 
1
e 2 παβ = ···, (17)
12 − β 2 32 + α2 52 − β 2 72 + α2

provided that cosh 12 πα = sec 12 πβ.

4. Now changing x into 2x(1 + i) in (15), we have


8ix2 8ix2 8ix2
     
log 1 + 2 − 3 log 1 + 2
+ 5 log 1 + 2 − ···
1 3 5
 
4 4 1 −πx(1+i) 1 −3πx(1+i)
= φ(1) − 4x(1 + i) tan−1 e−πx(1+i) − e − e + . . . .
π π 12 32

Equating real and imaginary parts we see that, if x is positive, then

64x4 64x4 64x4


     
log 1 + 4 − 3 log 1 + 4 + 5 log 1 + 4 − ···
1 3 5
 
8 cosh πx + sin πx  cos πx 
= φ (1) − 2x log − 4x tan−1
π cosh πx − sin πx sinh πx
 
8 cos πx −πx cos 3πx −3πx cos 5πx −5πx
− e − e + e − ··· ; (18)
π 12 32 52
Rx tan−1 t
On the integral t
dt 53
0

and
8x2 −1 8x
2
−1 8x
2
tan−1 − 3 tan + 5 tan − ···
12 32 52
 
cosh πx + sin πx  cos πx 
= x log − 2x tan−1
cosh πx − sin πx sinh πx
 
4 sin πx −πx sin 3πx −3πx sin 5πx −5πx
+ e − e + e − ··· . (19)
π 12 32 52
It follows from (18) that, if n is a positive odd integer, then
5 
4n4 4n4 4n4 4n4
  −3  −7
1+ 4 1+ 4 1+ 4 1+ 4 ···
1 3 5 7
!2n cos 1 (n−1)π
− 12 πn 2
8
φ(1) 1 − e
=e π
1 , (20)
1 + e− 2 πn
and, if n is any even integer, then
5 
4n4 4n4 4n4 4n4
  −3  −7
1+ 4 1+ 4 1+ 4 1+ 4 ···
1 3 5 7
 
8 8 1
n − 12 πn 1 −1 − 12 πn
= exp φ(1) − (−1) [φ(e
2 ) + πn tan e ] . (21)
π π 2
Similarly from (19) we see that, if n is any positive odd integer, then
2n2 −1 2n
2
−1 2n
2
tan−1 − 3tan + 5 tan − ···
12 32 52
( ! )
− 12 πn
4 1 πn 1 + e 1 1 1 3 1 5
= (−1) 2 (n−1) log 1 + 2 e− 2 πn + 2 e− 2 πn + 2 e− 2 πn + · · · ; (22)
π 4 1 − e− 2 πn 1 3 5

and, if n is a positive even integer, then


2n2 −1 2n
2
−1 2n
2 1 1
tan−1 2
− 3 tan 2
+ 5 tan 2
− . . . = 2n(−1) 2 n−1 tan−1 e− 2 πn . (23)
1 3 5
In this connection it may be interesting to note that
1 π  n n n 
tan−1 e− 2 πn = − tan−1 − tan−1 + tan−1 − . . . (24)
4 1 3 5
for all real values of n.
π 1 3 5
5. Remembering that = 2 2
− 2 2
+ 2 − . . . we have
4 cosh πx 1 + 4x 3 + 4x 5 + 4x2
∞ n=∞
X 
πX 1 1 3
= − + ···
4 n2 cosh πnx n2 (12 + 4n2 x2 ) n2 (32 + 4n2 x2 )
1 n=1
54 Paper 7

!
π3 1 x2 π
coth 2x coth 3π coth 5π
 
2x 2x
= + − πx − + − ··· . (25)
8 3 2 12 32 52

That is to say, if α and β are real and αβ = π 2 , then

φ(1) + 2φ(e−α ) + 2φ(e−2α ) + 2φ(e−3α ) + · · ·


   
π α β π 1 1
= + − + + · · · . (26)
8 3 2 4β 12 cosh β 22 cosh 2β

If, in particular, we put α = β = π in (26), we obtain

5π 2
 
1 1 1
φ(1) = −2 − + ...
48 12 (eπ − 1) 32 (e3π − 1) 52 (e5π − 1)
 
1 1 1 1
− + + + ...
2 (12 eπ + e−π ) 22 (e2π + e−2π ) 32 (e3π + e−3π )
= .9159655942, (27)

approximately.
On the number of divisors of a number

Journal of the Indian Mathematical Society, VII, 1915, 131 – 133

1. If δ be a divisor of N , then there is a conjugate divisor δ′ such that δδ′ = N . Thus we


see that √ √
the number of divisors from 1 to N is equal to the number of divisors from N to N . (1)
From this it evidently follows that

d(N ) < 2 N , (2)

where d(N ) denotes the number of divisors of N (including √ unity and the number itself).
This is only a trivial result, as all the numbers from 1 to N cannot be divisors of N .
So let us try to find the best possible superior limit for d(N ) by using purely elementary
reasoning.

2. First let us consider the case in which all the prime divisors of N are known. Let

N = pa11 · pa22 · pa33 · · · pann ,


where p1 , p2 , p3 . . . pn are a given set of n primes . Then it is easy to see that

d(N ) = (1 + a1 )(1 + a2 )(1 + a3 ) · · · (1 + an ). (3)

But
1
{(1 + a1 ) log p1 + (1 + a2 ) log p2 + · · · + (1 + an ) log pn }
n
1
> {(1 + a1 )(1 + a2 ) · · · (1 + an ) log p1 log p2 · · · log pn } n , (4)

since the arithmetic mean of unequal positive numbers is always greater than their geometric
mean. Hence
1 1
{log p1 + log p2 + · · · + log pn + log N } > {log p1 log p2 · · · log pn · d(N )} n .
n
In other words
1 n
log(p1 p2 p3 · · · pn N )
n
d(N ) < , (5)
log p1 log p2 log p3 · · · log pn
for all values of N whose prime divisors are p1 , p2 , p3 , . . . , pn .

3. Next let us consider the case in which only the number of prime divisors of N is known.
Let
N = pa11 pa22 pa33 · · · pann ,
56 Paper 8

where n is a given number; and let

N ′ = 2a1 · 3a2 · 5a3 · · · pan ,

where p is the natural nth prime. Then it is evident that

N′ ≤ N; (6)

and

d(N ′ ) = d(N ). (7)

But
1 n
′ n log(2 · 3 · 5 · · · p · N ′ )
d(N ) < , (8)
log 2 log 3 log 5 · · · log p

by virtue of (5). It follows from (6) to (8) that, if p be the natural nth prime, then
1 n
n log(2 · 3 · 5 · · · p · N )
d(N ) < , (9)
log 2 log 3 log 5 · · · log p
for all values of N having n prime divisors.

4. Finally, let us consider the case in which nothing is known about N . Any integer N can
be written in the form
2a2 · 3a3 · 5a5 · · · ,
where aλ ≥ 0. Now let

xh = 2, (10)

where h is any positive number. Then we have

d(N ) 1 + a2 1 + a3 1 + a5
h
= ha2 · ha3 · ha5 (11)
N 2 3 5
But from (10) we see that, if q be any prime greater than x, then
1 + aq 1 + aq 1 + aq
≤ haq = ≤ 1. (12)
q haq x 2aq

It follows from (11) and (12) that, if p be the largest prime not exceeding x, then

d(N ) 1 + a2 1 + a3 1 + a5 1 + ap
≤ · ha3 · ha5 · · · hap
Nh 2ha2 3 5 p
1 + a2 1 + a3 1 + a5 1 + ap
≤ · ha3 · ha5 · · · hap . (13)
2ha2 2 2 2
On the number of divisors of a number 57

2h
But it is easy to shew that the maximum value of (1 + a)2−ha for the variable a is he log 2 .
Hence
ω(x)
2h

d(N )
≤ , (14)
Nh he log 2

where ω(x) denotes the number of primes not exceeding x. But from (10) we have

log 2
h= .
log x

Substituting this in (14), we obtain


( log 2 )ω(x)
log 2 2 log x log x
d(N ) ≤ N . (15)
log x e(log 2)2

But it is easy to verify that, if x ≥ 6.05, then

2h < e(log 2)2 .

From this and (15) it follows that, if x ≥ 6.05, then

d(N ) < 2(log N )/(log x) (log x)ω(x) (16)

for all values of N , ω(x) being the number of primes not exceeding x.

5. The symbol “O” is used in the following sense:

φ(x) = O{(Ψ(x)}

means that there is a positive constant K such that



φ(x)
Ψ(x) ≤ K

for all sufficiently large values of x (see Hardy, Orders of Infinity, pp. 5 et seq.). For
example:

1 √
5x = O(x); x = O(x); x sin x = O(x); x = O(x); log x = O(x);
2

but x2 6= O(x); x log x 6= O(x).


Hence it is obvious that

ω(x) = O(x). (17)


58 Paper 8

Now, let us suppose that


log N
x=
(log log N )2
in (16). Then we have
log x = log log N + O(log log log N );
and so
 
log N log N log N log log log N
= +O . (18)
log x log log N (log log N )2

Again
 
log N log log log N
ω(x) log log x = O(x log log x) = O . (19)
(log log N )2

It follows from (16), (18) and (19), that


 
log 2 log N log N log log log N
log d(N ) < +O (20)
log log N (log log N )2

for all sufficiently large values of N .


On the sum of the square roots of the first n natural numbers

Journal of the Indian Mathematical Society, VII, 1915, 173 – 175

1. Let
√ √ √ √ √ √ 
φ1 (n) = 1 + 2 + 3 + · · · + n − C1 + 23 n n + 12 n
ν=∞
X p p
− 61 { (n + ν) + (n + ν + 1)}−3 ,
ν=0

where C1 is a constant such that φ1 (1) = 0. Then we see that


p h p p i
φ1 (n) − φ1 (n + 1) = − (n + 1) + 23 (n + 1) (n + 1) + 12 (n + 1)
√ √ √ p
−( 32 n n + 12 n) + 16 { n − (n + 1)}3 = 0.

But φ1 (1) = 0. Hence φ1 (n) = 0 for all values of n. That is to say


√ √ √ √ √ √ √ √
n o−3
p
2 1 1
1 + 2 + 3 + 4 + · · · + n = C1 + 3 n n + 2 n + 6 n + (n + 1)
np o−3 np o−3 
p p
+ (n + 1) + (n + 2) + (n + 2) + (n + 3) + · · · . (1)

But it is known that


 
1 1 1 1
C1 = √ + √ + √ + ··· . (2)
4π 1 1 2 2 3 3

Putting n = 1 in (1) and using (2), we obtain


 
1 1 1 1
2π √ + √ √ + √ √ + √ √ + ···
( 1)3 ( 1 + 2)3 ( 2 + 3)3 ( 3 + 4)3
 
1 1 1 1
=3 √ + √ + √ + √ + ··· . (3)
( 1)3 ( 2)3 ( 3)3 ( 4)3

2. Again let
√ √ √ √ √ √ 
φ2 (n) = 1 1 + 2 2 . . . + n n − C2 + 25 n2 n + 12 n n + 81 n
ν=∞
X p p
1
− 40 [ (n + ν) + (n + ν + 1)]−5 ,
ν=0
60 Paper 9

where C2 is a constant such that φ2 (1) = 0. Then we have


p
φ2 (n) − φ2 (n + 1) = −(n + 1) (n + 1)
p p p
+{ 52 (n + 1)2 (n + 1) + 21 (n + 1) (n + 1) + 81 (n + 1)}
√ √ √ √ p
−{ 25 n2 n + 12 n n + 18 n} + 40 1
{ n − (n + 1)}5 = 0.

But φ2 (1) = 0. Hence φ2 (n) = 0. In other words


√ √ √ √ √ √ √ h √ p
1 1 + 2 2 + 3 3 + · · · n n = C2 + 52 n2 n + 21 n n + 18 n + 1
{ n + (n + 1)}−5
40
p p p p i
+{ (n + 1) + (n + 2)}−5 + { (n + 2) + (n + 3)}−5 + · · · . (4)

But it is known that


 
3 1 1 1
C2 = − √ + √ + √ + ··· . (5)
16π 2 12 1 22 2 32 3

It is easy to see from (4) and (5) that


 
2 1 1 1 1
2π √ + √ √ + √ √ + √ √ + ···
( 1)5 ( 1 + 2)5 ( 2 + 3)5 ( 3 + 4)5 
1 1 1 1
= 15 √ + √ + √ + √ + ··· . (6)
( 1)5 ( 2)5 ( 3)5 ( 4)5

3. The corresponding results for higher powers are not so neat as the previous ones. Thus
for example
√ √ √ √ √
12 1 + 22 2 + 32 3 + · · · + n2 n = C3 + n( 27 n3 + 12 n2 + 24
5
n)
1
√ p p p
− 96 [{ n + (n + 1)}−3 + { (n + 1) + (n + 2)}−3 + · · ·]
h √ p p p
1
+ 224 { n + (n + 1)}−7 + { (n + 1) + (n + 2)}−7
p p i
+{ (n + 2) + (n + 3)}−7 + · · · ; (7)

√ √ √ √
13 1 + 23 2 + · · · + n3 n = C4 + n( 92 n4 + 12 n3 + 24
7 2 7
n − 384 )
h √ p p p i
1
− 192 { n + (n + 1)}−5 + { (n + 1) + (n + 2)}−5 + · · ·
1
√ p p p
+ 1152 [{ n + (n + 1)}−9 + { (n + 1) + (n + 2)}−9 + · · ·]; (8)

and so on.
On the sum of the square roots of the first n natural numbers 61

The constants C3 , C4 , . . . can be ascertained from the well-known result that the constant
in the approximate summation of the series 1r−1 + 2r−1 + 3r−1 + · · · + nr−1 is
 
2Γ(r) 1 1 1 1
+ r + r + r + · · · cos 12 πr, (9)
(2π)r 1r 2 3 4
provided that the real part of r is greater than 1.

4. Similarly we can shew, by induction, that


1 1 1 1 √ 1
√ + √ + √ + · · · + √ = C0 + 2 n + √
1 2 3 n 2 n
( √ p p p )
{ n + (n + 1)} −3 { (n + 1) + (n + 2) −3 }
− 12 p + p + ··· , (10)
{n(n + 1)} {(n + 1)(n + 2)}
The value of C0 can be determined as follows: from (10) we have
1 1 1 1 p
√ + √ + √ + ··· + p − 2 (2n) → C0 , (11)
1 2 3 (2n)
as n → ∞. Also
!
1 1 1 1 p √
2 √ + √ + √ + ··· + p − 2 (2n) → C0 2, (12)
2 4 6 (2n)
as n → ∞.
Now subtracting (12) from (11) we see that
1 1 1 1 1 √
√ − √ + √ − √ + ··· − p → C0 (1 − 2), as n → ∞.
1 2 3 4 (2n)
That is to say

 
1 1 1 1
C0 = −(1 + 2) √ − √ + √ − √ + ··· . (13)
1 2 3 4
We can also shew, by induction, that
√ √ √ √ √ √ 1
1 + 2 + 3 + · · · + n = C1 + 32 n n + 12 n + √
24 n
" √ p p p #
{ n + (n + 1)}−5 { (n + 1) + (n + 2)}−5
1
− 24 p + p + · · · . (14)
{n(n + 1)} {(n + 1)(n + 2)}
√ √ √ √
The asymptotic expansion of 1 + 2 + 3 + · · · + n for large values of n can be shewn
to be
√ √
 
2 1 1 1 1 1
C1 + 3 n n + 2 n + √ − + − ··· , (15)
n 24 1920n2 9216n4
by using the Euler-Maclaurin sum formula.
n=∞
"  3#
Y x
On the product 1+
n=0
a + nd

Journal of the Indian Mathematical Society, VII, 1915, 209 – 211

1. Let
( 3 ) (  )
α+β 3
 
α+β
φ(α, β) = 1+ 1+ . (1)
1+α 2+α

It is easy to see that


(  )(  )
α+β 3 α+β 3
 
1+ 1+
n+α n+β
√ )2
   
1 + α+2β 1 + β+2α (
=
n n
3 1 − (α − β) + i(α + β) 3 
2n

3
1 + αn 1 + βn


√ )2
 ( 
(α − β) − i(α + β) 3
× 1 − ; (2)
2n

  
α+2β
n=∞
Y  1+ n 1 + β+2α
n

 {Γ(1 + α)Γ(1 + β)}3
3 = ; (3)
 Γ(1 + α + 2β)Γ(1 + β + 2α)
3  
1 + αn 1 + βn
 
n=1 

and
√ )2 √ )2
  
n=∞
( (
Y
1 − (α − β) + i(α + β) 3   (α − β) − i(α + β) 3 
1−
n=1
2n 2n

cosh π(α + β) 3 − cos π(α − β)
= . (4)
2π 2 (α2 + αβ + β 2 )

It follows from (1) - (4) that

φ(α, β)φ(β, α)
( √ )
{Γ(1 + α)Γ(1 + β)}3 cosh π(α + β) 3 − cos π(α − β)
= . (5)
Γ(1 + α + 2β)Γ(1 + β + 2α) 2π 2 (α2 + αβ + β 2 )
Qn=∞ h x
3 i
On the product n=0 1+ a+nd
63

But it is evident that, if α − β be any integer, then φ(α, β)/φ(β, α) can be expressed in
finite terms. From this and (5) it follows that φ(α, β) can be expressed in finite terms, if
α − β be any integer. That is to say
 ( 3 ) ( 3 )
 x 3  
x

x
1+ 1+ 1+ ···
a a+d a + 2d

can be expressed in finite terms if x − 2a be a multiple of d.

2. Suppose now that α = β in (5). We obtain


( 3 ) (  )(  )
2α 3 2α 3
  

1+ 1+ 1+ ···
1+α 2+α 3+α

{Γ(1 + α)}3 sinh πα 3
= √ . (6)
Γ(1 + 3α) πα 3

Similarly, putting β = α + 1 in (5), we obtain


(  )(  )
2α + 1 3 2α + 1 3
 
1+ 1+ ···
1+α 2+α

{Γ(1 + α)}3 cosh π( 21 + α) 3
= . (7)
Γ(2 + 3α) π

Again, since

α
 α2 4

1+ 1+ + αn4
 ( 2 )
 α 3  
α n n2
1+ 1+3 = ,
n 2n + α α 2

1+ 2n

it is easy to see that


 "( 2 ) ( 2 ) #
α3 α3
    
α α
1+ 3 1+ 3 ··· 1+3 1+3 ...
1 2 2+α 4+α
√ !
Γ( 21 α) cosh πα 3 − cos πα
= √ . (8)
Γ{ 12 (1 + α)} 2α+2 πα π

3. It is known that, if the real part of α is positive, then


Z∞
1 1 tan−1 (x/α)
log Γ(α) = (α − 2 ) log α −α+ 2 log 2π + 2 dx. (9)
e2πx − 1
0
64 Paper 10

From this we can shew that, if the real part of α is positive, then

α3 α3 α3
    
1 πα
2 log 2πα +
√ + log 1+ 3 1+ 3 1 + 3 ...
3 1 2 3
√ ! ∞
cosh πα 3 − cos πα tan−1 (x/α)3
Z
= log +2 dx. (10)
πα e2πx − 1
0

From this and the previous section it follows that


Z∞
tan−1 x3
dx
e2πnx − 1
0

can be expressed in finite terms if n is a positive integer. Thus, for example,


Z∞
tan−1 x3 1 π 1

−π 3
dx = log 2π − √ − log(1 + e ); (11)
e2πx − 1 4 4 3 2
0

Z∞
tan−1 x3 1 π 1 √
4πx
dx = log 12π − √ − log(1 − e−2π 3 ); (12)
e −1 8 4 3 4
0

and so on.

4. It is also easy to see that


12 22 32 42
− + − + ···
13 + n3 23 + n3 33 + n3 43 + n3
 
1 1 1 1 1
= − + − + ...
3 1+n 2+n 3+n 4+n
 
4 2−n 4−n 6−n
+ − + − ... . (13)
3 (2 − n)2 + 3n2 (4 − n)2 + 3n2 (6 − n)2 + 3n2

Since
π 1 3 5
1 = 2 2
− 2 2
+ 2 − ...,
4 cosh 2 πx 1 +x 3 +x 5 + x2
it is clear that the left-hand side of (13) can be expressed in finite terms if n is any odd
integer. For example,

12 22 32 42 1 √
3
− 3
+ 3
− 3
+ . . . = (1 − log 2 + π sech 12 π 3). (14)
1 +1 2 +1 3 +1 4 +1 3
Qn=∞ h x
3 i
On the product n=0 1+ a+nd
65

The corresponding integral in this case is


Z∞ Z∞ ( ν=∞
X (−1)ν
)
x5 dx 2 1 x6 dx
6 6
= +
sinh πx n + x π 2x2 ν=1
ν 2 + x2 n 6 + x6
0 0
 
1 11 1 1
−= + − + ...
3
n n+1 n+2 n+3
 
4 n+2 n+4 n+6
− − + − . . . ; (15)
3 (n + 2)2 + 3n2 (n + 4)2 + 3n2 (n + 6)2 + 3n2

and so the integral on the left-hand side of (15) can be expressed in finite terms if n is any
odd integer. For example,
Z∞ √
x5 dx 1
6
= (log 2 − 1 + π sech 12 π 3). (16)
sinh πx 1 + x 3
0
Some definite integrals

Messenger of Mathematics, XLIV, 1915, 10 – 18

1. Consider the integral


Z∞
cos 2mx dx
,
{1 + x2 /a2 }{1 + x2 /(a + 1)2 }{1 + x2 /(a + 1)2 } · · ·
0

where m and a are positive.


It can be easily proved that
(  2 ) (  2 ) (  2 ) (  2 )
t t t t
1− 1− 1− ··· 1 −
a a+1 a+2 a+n−1

Γ(a + n − t)Γ(a + n + t){Γ(a)}2


= ,
Γ(a − t)Γ(a + t){Γ(a + n)}2

where n is any positive integer. Hence, by splitting


1
{1 + x2 /a2 }{1 + x2 /(a + 1)2 } · · · {1 + x2 /(a + n − 1)2 }

into partial fractions, we see that it is equal to

2Γ(2a){Γ(a + n)}2

a 2a n − 1 a+1

{Γ(a)}2 Γ(n)Γ(2a + n) a2 + x2 1! n + 2a (a + 1)2 + x2

2a(2a + 1) (n − 1)(n − 2) a+2
+ − ··· .
2! (n + 2a)(n + 2a + 1) (a + 2)2 + x2

Multiplying both sides by cos 2mx and integrating from 0 to ∞ with respect to x, we have
Z∞
cos 2mx dx
{1 + x2 /a2 }{1 + x2 /(a + 1)2 } · · · {1 + x2 /(a + n − 1)2 }
0

πΓ(2a){Γ(a + n)}2
 
2a n − 1 −2(a+1)m
= e−2am
− e + ··· .
{Γ(a)}2 Γ(n)Γ(2a + n) 1! n + 2a

The limit of the right-hand side, as n → ∞, is


 
πΓ(2a) −2am 2a −2(a+1)m 2a(2a + 1) −2(a+2)m
e − e + e − ···
{Γ(a)}2 1! 2!
Some definite integrals 67

√ Γ(a + 12 )
= 1
2 π sech2a m.
Γ(a)

Hence
Z∞
cos 2mx dx 1
√ Γ(a + 12 ) 2a
= 2 π Γ(a) sech m. (1)
{1 + x2 /a2 }{1 + x2 /(a + 1)2 } · · ·
0

Since
( 2 ) ( 2 )
{Γ(a)}2
  x 2   
x x
1+ 1+ 1+ ··· = ,
a a+1 a+2 Γ(a + ix)Γ(a − ix)

the formula (1) is equivalent to


Z∞

|Γ(a + ix)|2 cos 2mx dx = 1
2 πΓ(a)Γ(a + 12 )sech2a m. (2)
0

2. In a similar manner we can prove that


Z∞ 
1 + x2 /b2 1 + x2 /(b + 1)2 1 + x2 /(b + 2)2
  
· · · cos mx dx
1 + x2 /a2 1 + x2 /(a + 1)2 1 + x2 /(a + 2)2
0

πΓ(2a){Γ(b)}2

2a b − a − 1 −(a+2)m
= e−am − e
{Γ(a)}2 Γ(b + a)Γ(b − a) 1! b + a

2a(2a + 1) (b − a − 1)(b − a − 2) −(a+2)m
+ e − ··· ,
2! (b + a)(b + a + 1)

where m is positive and 0 < a < b. When 0 < a < b − 12 , the integral and the series remain
convergent for m = 0, and we obtain the formulæ
Z∞ 
1 + x2 /b2 1 + x2 /(b + 1)2 1 + x2 /(b + 2)2
  
· · · dx
1 + x2 /a2 1 + x2 /(a + 1)2 1 + x2 /(a + 2)2
0

1
√ Γ(a + 21 )Γ(b)Γ(b − a − 12 )
= 2 π , (3)
Γ(a)Γ(b − 21 )Γ(b − a)

Z∞
Γ(a + ix) 2 √ 1 1

dx = 1 π Γ(a)Γ(a + 2 )Γ(b − a − 2 ) . (4)
2
Γ(b + ix) Γ(b − 12 )Γ(b)Γ(b − a)
0
68 Paper 11

If a1 , a2 , a3 , . . . , an be n positive numbers in arithmetical progression, then


Z∞
dx
(a21 + x2 )(a22 + x2 )(a23 + x2 ) · · · (a2n + x2 )
0

is a particular case of the above integral, and its value can be written down at once. Thus,
for example, by putting a = 11 61
10 and b = 10 , we obtain

Z∞
dx
(x2 + 112 )(x2 + 212 )(x2 + 312 )(x2 + 412 )(x2 + 512 )
0

= .
12 · 13 · 16 · 17 · 18 · 22 · 23 · 24 · 31 · 32 · 41

3. It follows at once from equation (1), by applying Fourier’s theorem


Z∞ Z∞
cos nydy φ(x) cos xy dx = 12 πφ(n),
0 0

that, when a and n are positive,


Z∞
sech2a x cos 2nx dx
0

1
√ Γ(a) 1
= 2 π 1 {1 + n2 /a2 }{1 + n2 /(a + 1)2 }{1 + n2 /(a + 2)2 } · · ·
Γ(a + 2 )

1
√ |Γ(a + in)|2
= 2 π . (5)
Γ(a)Γ(a + 21 )

Hence the function


Z∞
φ(a) = secha x cos nx dx (0 < a < 2)
0
satisfies the functional equation
π sin πa
φ(a)φ(2 − a) = .
2(1 − a)(cosh πn − cos πa)

4. Let
Zb
f (x)F (nx) dx = Ψ(n),
a
Some definite integrals 69

and

φ(x)F (nx) dx = χ(n).
α
Then, if we suppose the functions f, φ, and F to be such that the order of integration is
indifferent, we have

Zb Zβ Zb
f (x)χ(nx) dx = dy f (x)φ(y)F (nxy) dx
a α a


= φ(y)Ψ(ny)dy. (6)
α

A number of curious relations between definite integrals may be deduced from this result.
We have, for example, the formulæ
Z∞
cos 2nx 1
dx = , (7)
cosh πx 2 cosh n
0

Z∞ √
cos 2nx dx 3
2 = , (8)
1 + 2 cosh 3 πx 2(1 + 2 cosh 2n)
0

Z∞
2
1
√ −n2
e−x cos 2nx dx = 2 πe . (9)
0

By applying the general result (6) to the integrals (7) and (8), we obtain
∞ Z∞
√ Z dx dx
3 = ;
cosh πx(1 + 2 cosh 2nx) cosh nx(1 + 2 cosh 32 πx)
0 0

or, in other words, if αβ = 34 π 2 , then


Z∞
√ dx
α
cosh αx(1 + 2 cosh πx)
0

p Z dx
= β . (10)
cosh βx(1 + 2 cosh πx)
0
70 Paper 11

In the same way, from (8) and (9), we obtain


Z∞ 2 Z∞ 2
√ e−x dx p e−x dx
α = β , (11)
1 + 2 cosh αx 1 + 2 cosh βx
0 0

with the condition αβ = 34 π; and, from (7) and (9),


Z∞ 2 Z∞ 2
√ e−x p e−x
α dx = β dx, (12)
cosh αx cosh βx
0 0

with the conditions αβ = π. ∗


Similarly, by taking the two integrals
Z∞  
sin nx 1 1 1 1
2πx
dx = 2 n
+ 2 − ,
e −1 e −1 n
0

and
Z∞
2 1√ 1 2
xe−x sin nx dx = πne− 4 n ,
4
0

we can prove that, if αβ = π 2 , then


 
Z∞
1  e−αx 
√ 1 + 2α √ dx
4
α e2π x − 1 
0
 
Z∞
1  e−βx 
= √ 1 + 2β √ dx ; (13)
4
β e2π x − 1 
0

and so on.

5. Suppose now that a, b and n are positive, and


Z∞
cos
φ(a, x) nx dx = Ψ(a, n). (14)
sin
0

Then, if the conditions of Fourier’s double integral theorem are satisfied, we have
Z∞
cos
Ψ(b, x) nx dx = 12 πφ(b, n). (15)
sin
0

Formulæ equivalent to (11) and (12) were given by Hardy, Quarterly Journal, XXXV, p.193
Some definite integrals 71

Applying the formula (6) to (14) and (15), we obtain


Z∞ Z∞
1
2π φ(a, x)φ(b, nx) dx = Ψ(b, x)Ψ(a, nx) dx. (16)
0 0

Thus, when a = b, we have the formula


Z∞ Z∞
1
2π φ(x)φ(nx) dx = Ψ(x)Ψ(nx) dx,
0 0

where
Z∞
cos
ψ(t) = φ(x) tx dx;
sin
0
and, in particular, if n = 1, then
Z∞ Z∞
1

2
{φ(x)} dx = {Ψ(x)}2 dx.
0 0

if
1
φ(a, x) = (a > 0),
{1 + x2 /a2 }{1 + x2 /(a + 1)2 } · · ·
then, by (1),
√ Γ(a + 12 )
Ψ(a, x) = 1
2 π sech2a 12 x.
Γ(a)
Hence, by (16),
Z∞ Z∞
Γ(a + 12 )Γ(b + 21 )
φ(a, x)φ(b.x) dx = sech2a+2b 21 x dx;
2Γ(a)Γ(b)
0 0

and so
Z∞
dx
{1 + x2 /a2 }{1 + x2 /(a + 1)2 } · · · {1 + x2 /b2 }{1 + x2 /(b + 1)2 } · · ·
0

1
√ Γ(a + 12 )Γ(b + 21 )Γ(a + b)
= 2 π , (17)
Γ(a)Γ(b)Γ(a + b + 12 )
a and b being positive: or
Z∞
√ Γ(a)Γ(a + 21 )Γ(b)Γ(b + 12 )Γ(a + b)
|Γ(a + ix)Γ(b + ix)|2 dx = 1
2 π . (18)
Γ(a + b + 12 )
0
72 Paper 11

As particular cases of the above result, we have, when b = 1,


Z∞
x dx a
2 2 2 2
= ;
sin hπx {1 + x /a }{1 + x /(a + 1) } · · · 2(1 + 2a)
0

when b = 2,
Z∞
x3 dx a2
= ;
sin hπx {1 + x2 /a2 }{1 + x2 /(a + 1)2 } · · · 2(1 + 2a)(3 + 2a)
0

and so on. Since Π{1+ x2 /(a+ n)2 } can be expressed in finite terms by means of hyperbolic
functions when 2a is an integer, we can deduce a large number of special formulæ from the
preceding results.

6. Another curious formula is the following. If 0 < r < 1, n > 0, and 0 < a < r n−1 , then
Z∞
(1 + arx)(1 + ar 2 x) · · ·
xn−1 dx
(1 + x)(1 + rx)(1 + r 2 x) · · ·
0
m=∞
Y (1 − r m−n )(1 − ar m )
π
= , (19)
sin nπ (1 − r m )(1 − ar m−n )
m=1

unless n is an integer or a is of the form r p , where p is a positive integer.


If a = r p , the formula reduces to
Z∞
xn−1 dx
(1 + x)(1 + rx) · · · (1 + r p x)
0

π (1 − r 1−n )(1 − r 2−n ) · · · (1 − r p−n )


= . (20)
sin nπ (1 − r)(1 − r 2 ) · · · (1 − r p )
If n is an integer, the value of the integral is in any case

log r (1 − r)(1 − r 2 ) · · · (1 − r n−1 )


− .
1 − a (r − a)(r 2 − a) · · · (r n−1 − a)
My own proofs of the above results make use of a general formula, the truth of which de-
pends on conditions which I have not yet investigated completely. A direct proof depending
on Cauchy’s theorem will be found in Mr Hardy’s note which follows this paper. The final
formula used in Mr Hardy’s proof can be proved as follows. Let
m=∞
Y  1 − btxm 
f (t) = = A0 + A1 t + A2 t2 + · · · .
1 − atxm
m=0
Some definite integrals 73

Then it is evident that


(1 − at)f (t) = (1 − bt)f (tx).
That is
(1 − at)(A0 + A1 t + A2 t2 + · · ·) = (1 − bt)(A0 + A1 tx + A2 t2 x2 + · · ·).
Equating the coefficients of tn , we obtain
1 − bxn−1
An = An−1 ;
1 − xn
and A0 is evidently 1. Hence we have
a−b (a − b)(a − bx)
f (t) = 1 + t + t2 + ···. (21)
1−x (1 − x)(1 − x2 )

7. As a particular case of (19), we have, when a = 0,


Z∞
xn−1 dx π 1 − r 1−n 1 − r 2−n
= ···. (22)
(1 + x)(1 + rx)(1 + r 2 x) · · · sin nπ 1 − r 1 − r2
0

When n is an integer, the value of the integral reduces to


1
−r − 2 n(n−1) (1 − r)(1 − r 2 ) · · · (1 − r n−1 ) log r.
1
When we put n = 2 in (19), we have
Z∞
1 1 + ar 2 x2 1 + ar 4 x2
· · · dx
1 + x2 1 + r 2 x2 1 + r 4 x2
0

1 − ar 2 1 − ar 4 1 − r 1 − r3
= 12 π · · · ···. (23)
1 − r2 1 − r4 1 − ar 1 − ar 3
1
If, in particular, n = 2 in (22), or a = 0 in (23), then
Z∞
dx
(1 + x2 )(1 + r 2 x2 )(1 + r 4 x2 ) · · ·
0

1 − r 1 − r3 1 − r5 π
= 12 π ··· = , (24)
1 − r2 1 − r4 1 − r6 2(1 + r + r 3 + r 6 + r 10 + · · ·)
1
the nth term in the denominator being r 2 n(n−1) . Thus, for example, when r = e−5π , we
have
Z∞
dx
2
(1 + x )(1 + e−10π x2 )(1 + e−20x2 ) · · ·
0
74 Paper 11

π
=
2(1 + e−5π + e−15π + e−30π + · · ·)
3 √ √ √ √ 1
= π 4 Γ 43 5 2 12 (1 + 5){ 21 (1 + 5)} 2 e−5π/8 .
8 4

Similarly
Z∞
dx
(1 + x2 )(1 + e−20π x2 )(1 + e−40π x2 ) · · ·
0
3 √ √ √ √
3
5 2 12 (1 + 5)2 { 12 (1 + 5)} 52 e−5π/4 ;
4 4
= π4Γ 4

and
Z∞
dx 10 1110 111110
= 12 π ···
(1 + x2 )(1 2 2
+ .001x )(1 + .00001x ) · · · 11 1111 111111
0
π
= .
2.202 002 000 200 002 000 002 . . .
Some definite integrals connected with Gauss’s sums

Messenger of Mathematics, XLIV, 1915, 75 – 85

1. If n is real and positive, and |I(t)|, where I(t) is the imaginary part of t, is less than
either n or 1, we have
Z∞ Z∞ Z∞
cos πtx −iπnx2 cos πtx cos 2πxy −iπnx2
e dx = 2 e dxdy
cosh πx cosh πy
0 0 0

2
 Z∞

 
t cos πtx iπnx2
= n exp − 14 iπ 1 − e dx. (1)
n cosh πnx
0

When n = 1 the above formula reduces to


Z∞ Z∞
cos πtx cos πtx
sin πx2 dx = tan{ 18 π(1 − t2 )} cos πx2 dx. (2)
cosh πx cosh πx
0 0

if t = 0, and

Z∞
cos πnx2 

φ(n) = dx, 

cosh πx




0
(3)
Z∞
sin πnx2



Ψ(n) = dx, 

cosh πx



0

then q 
2 1
 
φ(n) = n Ψ n + Ψ(n), 

q (3′ )
2 1
 
Ψ(n) = φ − φ(n). 

n n

1

Similarly, if 2 3|I(t)| is less than either 1 or n, we have

Z∞
cos πtx 2
√ eiπnx dx
1 + 2 cosh(2πx/ 3)
0
 Z∞
√ t2
 
1 cos πtx 2
= n exp − 4 iπ 1 − √ e−iπnx dx. (4)
n 1 + 2 cosh(2πnx/ 3)
0
76 Paper 12

If in (4) we suppose n = 1, we obtain


Z∞ Z∞
cos πtx sin πx2 1 2 cos πtx cos πx2
√ dx = tan{ 8 π(1 − t )} √ dx; (5)
1 + 2 cosh(2πx/ 3) 1 + 2 cosh(2πx/ 3)
0 0
and if t = 0, and

Z∞
cos πnx2 

φ(n) = √ dx, 


1 + 2 cosh(2πx/ 3) 


0
(6)
Z∞
sin πnx2



Ψ(n) = √ dx, 

1 + 2 cosh(2πx/ 3)



0
then q 
2 1
 
φ(n) = n Ψ n + Ψ(n), 

q (6′ )
2 1
 
Ψ(n) = φ − φ(n). 

n n
In a similar manner we can prove that
Z∞  Z∞
√ t2
 
sin πtx −iπnx2 1 sin πtx iπnx2
e dx = − n exp 4 iπ 1 + e dx. (7)
tanh πx n tanh πnx
0 0
If we put n = 1 in (7), we obtain
Z∞ Z∞
sin πtx 2 1 2 sin πtx
cos πx dx = tan{ π(1 + t )} sin πx2 dx. (8)
tanh πx 8 tanh πx
0 0
Now
Z∞ Z∞ Z∞
1 sin atx icx2 1 2 sin atx icx2 sin atx icx2
lim e dx = lim e dx + lim e dx
t→0 t tanh bx t→0 t e2bx − 1 t→0 t
0 0 0
Z∞
aeicx ia
= √ dx + . (9)
e2b x − 1 2c
0
Hence, dividing both sides of (7) by t, and making t → 0, we obtain the result corresponding
to (3) and (6), viz.: if

Z∞
cos πnx 

φ(n) = dx,



e2π x −1




0
(10)
Z∞
1 sin πnx



Ψ(n) = + √ dx, 

2πn e2π x −1



0
Some definite integrals connected with Gauss’s sums 77

then q 
1 2 1
 
φ(n) = n n Ψ n − Ψ(n), 

q (10′ )
1 2 1
 
Ψ(n) = φ + φ(n). 

n n n

2. I shall now shew that the integral (1) may be expressed in finite terms for all rational
values of n. Consider the integral
Z∞
cos tx dx
J(t) = .
cosh 21 πx a2 + x2
0

If R(a) and t are positive, we have


Z∞ r=∞
4 X (−1)r (2r + 1) cos tx
J(t) = dx
π r=0
x2 + (2r + 1)2 a2 + x2
0
r=∞
(−1)r
 
X
−(2r+1)t 1 −at
= 2 e − (2r + 1)e
a2 − (2r + 1)2 a
r=0
r=∞
X (−1)r e−(2r+1)t
πe−at
= 1 +2 , (11)
2a cos 2 πa r=0
a2 − (2r + 1)2

and it is easy to see that this last equation remains true when t is complex, provided
R(t) > 0 and |I(t)| ≤ 12 π. Thus the integral J(t) can be expressed in finite terms for all
rational values of a. Thus, for example, we have

Z∞
cos tx dx 

= cosh t log(2 cosh t) − t sinh t,


1 2
cosh 2 πx 1 + x




0
(12)
Z∞
cos 2tx dx


= 2 cosh t − (e2t tan−1 e−t + e−2t tan−1 et ), 


cosh πx 1 + x2



0

and so on. Now let


Z∞
cos 2tx −iπnx2
F (n) = e dx. (13)
cosh πx
0

Then, if R(a) > 0,


Z∞ Z∞
−an cos 2tx dx
e F (n)dn = . (14)
cosh πx a + iπx2
0 0
78 Paper 12

Now let
r=∞
X
f (n) = (−1)r exp{−(2r + 1)t + 14 (2r + 1)2 iπn}
r=0
 r=∞
t2
   
1 1
X
r t 1 2 iπ
+ √ exp −i 4 π − (−1) exp −(2r + 1) − 4 (2r + 1) . (15)
n πn n n
r=0

Then
Z∞ r=∞ p
(−1)r e−(2r+1)t
r
−an
X π  exp{− (2a/π)(1 − i)t}
e f (n)dn = +
a − 41 (2r + 1)2 iπ
q
r=0
2a (1 + i) cosh{(1 + i) ( 1 πa)}
0 2

Z∞
cos 2tx dx
= , (16)
cosh πx a + iπx2
0

in virtue of (11); and therefore

Z∞
e−an {F (n) − f (n)}dn = 0. (17)
0

Now it is known that, if φ(n) is continuous and

Z∞
e−an φ(n)dn = 0,
0

for all positive values of a (or even only for an infinity of such values in arithmetical
progression), then
φ(n) = 0,
for all positive values of n. Hence

F (n) = f (n). (18)

Equating the real and imaginary parts in (13) and (15) we have

Z∞  
cos 2tx πn 9πn 25πn
cos πnx2 dx = e−t cos − e−3t cos + e−5t cos − ···
cosh πx 4 4 4
0

t2 t2
     
1 −t/n π π −3t/n π 9π
+√ e cos − + −e cos − + + ··· , (19)
n 4 πn 4n 4 πn 4n
Some definite integrals connected with Gauss’s sums 79

Z∞  
cos 2tx 2 −t πn −3t 9πn −5t 25πn
sin πnx dx = − e sin − e sin + e sin − ···
cosh πx 4 4 4
0

t2 t2
     
1 −t/n π π −3t/n π 9π
+√ e sin − + −e sin − + + ··· . (20)
n 4 πn 4n 4 πn 4n

We can verify the results (18), (19) and (20) by means of the equation (1). This equation
can be expressed as a functional equation in F (n), nd it is easy to see that f (n) satisfies
the same equation.
The right-hand side of these equations can be expressed in finite terms if n is any rational
number. For let n = a/b, where a and b are any two positive integers and one of them is
odd. Then the results (19) and (20) reduce to

Z∞
πax2
 
cos 2tx
2 cosh bt cos dx
cosh πx b
0

= [cosh{(1 − b)t} cos(πa/4b) − cosh{(3 − b)t} cos(9πa/4b)


+ cosh{(5 − b)t} cos(25πa/4b) − · · · to b terms]
s  
π bt2 πb
    
b 1
+ cosh 1− bt cos − +
a a 4 πa 4a
π bt2 9πb
     
3
− cosh 1− bt cos − + + · · · to a terms , (21)
a 4 πa 4a

Z∞
πax2
 
cos 2tx
2 cosh bt sin dx
cosh πx b
0

= −[cosh{(1 − b)t} sin(πa/4b) − cosh{(3 − b)t} sin(9πa/4b)


+ cosh{(5 − b)t} sin(25πa/4b) − · · · to b terms]
s  
π bt2 πb
    
b 1
+ cosh 1− bt sin − +
a a 4 πa 4a
π bt2 9πb
     
3
− cosh 1− bt sin − + + · · · to a terms . (22)
a 4 πa 4a

Thus, for example, we have, when a = 1 and b = 1,

Z∞ √
cos πx2 1 + 2 sin πt2
cos 2πtx dx = √ , (23)
cosh πx 2 2 cosh πt
0
80 Paper 12

Z∞ √
sin πx2 −1 + 2 cos πt2
cos 2πtx dx = √ . (24)
cosh πx 2 2 cosh πt
0

It is easy to verify that (23) and (24) satisfy the relation (2).
The values of the integrals
Z∞ Z∞
cos πnx2 sin πnx2
dx, dx
cosh πx cosh πx
0 0

can be obtained easily from the preceding results by putting t = 0, and need no special
discussion. By successive differentiations of the results (19) and (20) with respect to t and
n, we can evaluate the integrals

Z∞
sin tx cos

x2m−1 πnx2 dx, 



cosh πx sin 


0
(25)
Z∞
cos tx cos


x2m πnx2 dx, 


cosh πx sin



0

for all rational values of n and all positive integral values of m. Thus, for example, we have

Z∞ 2
cos πx 1 1 
x2

dx = √ − , 

cosh πx 8 2 4π 



0
(26)
Z∞ 2
sin πx 1 1


x2

dx = − √ .  
cosh πx 8 8 2 


0

3. We can get many interesting results by applying the theory of Cauchy’s reciprocal
functions to the preceding results. It is known that, if
Z∞
φ(x) cos knx dx = Ψ(n), (27)
0

then (i) 12 α{ 12 φ(0) + φ(α) + φ(2α) + φ(3α) + · · ·}


= 21 Ψ(0) + Ψ(β) + Ψ(2β) + Ψ(3β) + · · · , (27)
with the
√ condition αβ = 2π/k;
(ii) α 2{φ(α) − φ(3α) − φ(5α) + φ(7α) + φ(9α) − · · ·}
= Ψ(β) − Ψ(3β) − Ψ(5β) + Ψ(7β) + Ψ(9β) − · · · , (27)
with the condition αβ = π/4k;
Some definite integrals connected with Gauss’s sums 81


(iii) α 3{φ(α) − φ(5α) − φ(7α) + φ(11α) + φ(13α) − · · ·}
= Ψ(β) − Ψ(5β) − Ψ(7β) + Ψ(11β) + Ψ(13β) − · · · , (27)
with the condition αβ = π/6k, where 1, 5, 7, 11, 13, . . . are the odd natural numbers
without the multiples of 3.
There are of course corresponding results for the function
Z∞
φ(x) sin knx dx = Ψ(n), (28)
0

such as
α{φ(α) − φ(3α) + φ(5α) − · · ·} = Ψ(β) − Ψ(3β) + Ψ(5β) − · · · ,
with the condition αβ = π/2k.
Thus from (23) and (27) (i) we obtain the following results. If
r=∞ p r=∞
( )
√ 1 X cos r 2 πα2 X sin r 2 πβ 2
F (α, β) = α + − β , (29)
2 cosh rπα cosh rπβ
r=1 r=1

then
1
(2α){ + e−πα + e−4πα + e−9πα + · · ·}2 ,
p
F (α, β) = F (β, α) =
2
provided that αβ = 1.

4. If, instead of starting with the integral (11), we start with the corresponding sine inte-
gral, we can shew that, when R(a) and R(t) are positive and |I(t)| ≤ π,
Z∞ r=∞
X (−1)r e−rt
sin tx dx 1 πe−at
= − + . (30)
sinh πx a2 + x2 2a2 2a sin πa r=1
a 2 − r2
0

Hence the above integral can be expressed in finite terms for all rational values of a. For
example, we have
Z∞
sin tx dx
1 = et tan−1 e−t − e−t tan−1 et . (31)
sinh 2 πx 1 + x2
0

From (30) we can deduce that


Z∞
sin 2tx −iπnx2 1
e dx = − e−2t+iπn + e−4t+4iπn − e−6t+9iπn + · · ·
sinh πx 2
0

t2
  
1 1 9
− √ exp 1
4π + i {e−(t+ 4 iπ)/n + e−(3t+ 4 iπ)/n + · · ·}, (32)
n πn
82 Paper 12

R(t) being positive and |I(t)| ≤ 12 π. The right-hand side can be expressed in finite terms
for all rational values of n. Thus, for example, we have
Z∞
cos πx2 cosh πt − cos πt2
sin 2πtx dx = , (33)
sinh πx 2 sinh πt
0
Z∞
sin πx2 sin πt2
sin 2πtx dx = , (34)
sinh πx 2 sinh πt
0

and so on.
Applying the formula (28) to (33) and (34), we have, when αβ = 41 ,

√ r=∞ p r=∞

X cos{(2r + 1)2 πα2 } X cos{(2r + 1)2 πβ 2 }
(−1)r (−1)r

α + β 

sinh{(2r + 1)πα} sinh{(2r + 1)πβ}


r=0 r=0




√ 1 
= 2 α{ + e−2πα + e−8πα + e−18πα + · · ·}2 ; (35)
2 

√ r=∞ p r=∞

sin{(2r + 1)2 πα2 } sin{(2r + 1)2 πβ 2 }
X X 

(−1)r (−1)r

α = β .


sinh{(2r + 1)πα} sinh{(2r + 1)πβ}


r=0 r=0

By successive differentiation of (32) with respect to t and n we can evaluate the integrals

Z∞
cos tx cos 
x2m−1 πnx2 dx, 


sinh πx sin




0
(36)
Z∞
sin tx

cos 
x2m πnx2 dx



sinh πx sin



0

for all rational values of n and all positive integral values of m. Thus, for example, we have

Z∞ 2 Z∞ 2
cos πx 1 sin πx 1 

x dx = , x dx = , 

sinh πx 8 sinh πx 4π




0 0
(37)
Z∞ 2
  Z∞ 2
cos πx 1 3 sin πx 1


x3 1
x3

dx = − , dx = , 
16 4 π 2

sinh πx sinh πx 16π 


0 0

and so on.
The denominators of the integrands in (25) and (36) are cosh πx and sinh πx. Similar
integrals having the denominators of their integrands equal to
r
Y
cosh πar x sinh πbr x
1
Some definite integrals connected with Gauss’s sums 83

can be evaluated, if ar and br are rational, by splitting up the integrand into partial frac-
tions.

5. The preceding formulæ may be generalised. Thus it may be shewn that, if R(a) and
R(t) are positive, |I(t)| ≤ π, and −1 < R(θ) < 1, then

Z∞
cos tx dx
sin πθ
cosh πx + cos πθ a2 + x2
0
r=∞
( )
π e−at sin πθ X e−(2r+1−θ)t e−(2r+1+θ)t
= + − . (38)
2a cos πa + cos πθ a2 − (2r + 1 − θ)2 a2 − (2r + 1 + θ)2
r=0

From (38) it can be deduced that, if n and R(t) are positive, |I(t)| ≤ π, and −1 < θ < 1,
then

Z∞
cos tx 2
sin πθ e−iπnx dx
cosh πx + cos πθ
0
r=∞
X
= {e−(2r+1−θ)t+(2r+1−θ)2iπn − e−(2r+1+θ)t+(2r+1+θ)2iπn }
r=0
 r=∞
t2
 
1 X 2
1
+ √ exp − 4 i π − (−1)r−1 sin rπθe−(2rt+r iπ)/4n . (39)
n πn
r=1

The right-hand side can be expressed in finite terms if n and θ are rational. In particular,
when θ = 13 , we have

Z∞
cos tx 2
√ e−iπnx dx
1 + 2 cosh(2πx/ 3)
0
1 1
√ 1
√ 1

= {e− 3 (t 3−iπn) − e− 3 (2t 3−4iπn) + e− 3 (4t 3−16iπn) − · · ·}
2
t2
  
1 1
+ √ exp − 4 i π −
2 n πn
√ √ √
3+iπ)/3n 3+4iπ)/3n 3+16iπ)/3n
{e−(t − e−(2t + e−(4t − · · ·}, (40)

where 1,2,4,5, . . . are the natural numbers without the multiples of 3.


84 Paper 12

As an example, when n = 1, we have



Z∞
cos πx2 cos πtx 1 − 2 sin{(π − 3πt2 )/12} 

√ dx = √ ,



1 + 2 cosh(2πx/ 3) 8 cosh(πt/ 3) − 4 


0
√ (41)
Z∞
sin πx2 cos πtx − 3 + 2 cos{(π − 3πt2 )/12}



√ dx = √ . 

1 + 2 cosh(2πx/ 3) 8 cosh(πt/ 3) − 4



0

6. The formula (32) assumes a neat and elegant form when t is changed to t + 12 iπ. We
have then
Z∞
sin tx −iπnx2
e dx (n > 0, t > 0)
tanh πx
0
1
= { + e−t+iπn + e−2t+4iπn + e−3t+9iπn + · · ·}
2
t2
  
1 1 1
− √ exp 4 i π + { + e−(t+iπ)/n + e−(2t+4iπ)/n + · · ·}. (42)
n πn 2

In particular, when n = 1, we have



Z∞
cos πx2 1 
sin 2πtx dx = tanh πt{1 − cos( 41 π + πt2 )},



tanh πx 2




0
(43)
Z∞
sin πx2 1


tanh πt sin( 14 π + πt2 ).

sin 2πtx dx = 

tanh πx 2



0

We shall now consider an important special case of (42). It can easily be seen from (9) that
the left-hand side of (42), when divided by t, tends to
 
Z∞ Z∞
cos πnx  1 sin πnx 
√ dx − i + √ dx (44)
e2π x − 1  2πn e2π x − 1 
0 0

as t → 0. But the limit of the right-hand side of (42) divided by t can be found when n is
rational. Let then n = a/b, where a and b are any two positive integers, and let

Z∞ Z∞
cos πnx 1 sin πnx
φ(n) = √ dx, Ψ(n) = + √ dx.
e2π x −1 2πn e2π x −1
0 0
Some definite integrals connected with Gauss’s sums 85

The relation between φ(n) and Ψ(n) has been stated already in (10′ ). From (42) and (44)
it can easily be deduced that, if a and b are both odd, then

s 
r=b r=a
 2  
r 2 bπ
a  
1
X r πa b b X
(a − 2r) sin 41 π +

φ =4 (b − 2r) cos − ,


b b 4a a a



r=1 r=1
s  (45)
r=b r=a
r 2 πa r 2 πb
a    
b b X
X 

1 1
Ψ = −4 (b − 2r) sin + (a − 2r) cos 4 π + ,



b b 4a a a 
r=1 r=1

It can easily be seen that these satisfy the relation (10′ ). Similarly, when one of a and b is
odd and the other even, it can be shewn that

Pr=b  2  
a σ√
− 21 r=1 r 1 − rb cos r bπa
 
φ b = 4πa a





q P   
b b r=a r 1 r 2 πb
 
+ 2a r 1 − sin π + ,


a r=1 a 4 a 
  (46)
′ Pr=b 2
Ψ ab = 4πaσ√a + 21 r=1 r 1 − rb sin r bπa
  




q P   

b b r=a r 1 r 2 πb
 
− 2a r 1 − cos π + ,


a r=1 a 4 a

where

√ Pa  2
 √ P  2  
σ= b 1 cos 14 π + r aπb = a b1 sin r bπa , 

√ P   √ P  2  (47)
2
σ ′ = b a1 sin 41 π + r aπb = a b1 cos r bπa .

Thus, for example, we have

√ √ 
1 2− 2 1 3− 2
φ(0) = 12 , φ(1) = 8 , φ(2) = 16 , φ(4) = 32 ,

√ √ (48)
13−4 3 1 1 2 8−3 5
 
φ(6) = 144 , φ 2 = 4π , φ 5 = 16 ,

and so on.
By differentiating (42) with respect to n, we can evaluate the integrals

Z∞
xm cos
√ πnx dx (49)
e2π x − 1 sin
0
86 Paper 12

for all rational values of n and positive integral values of m. Thus, for example, we have
Z∞ 
xcos 21 πx 13 − 4π 
dx = ,

√ 
2π x 8π 2 
e −1 


0 




   
x cos 2πx 1 3 5
Z 
1
√ dx = 64 − + , (50)
e2π x − 1 2 π π2 
0




Z∞ 2

  
x cos 2πx 5 5


1 
√ dx = 256 1 − + 2
, 

e2π x −1 π π 

0

and so on.
Summation of a certain series

Messenger of Mathematics, XLIV, 1915, 157 – 160

1. Let
n=∞
X p p
Φ(s, x) = { (x + n) + (x + n + 1)}−s
n=0
n=∞
X p p
= { (x + n + 1) − (x + n)}s .
n=0

The object of this paper is to give a finite expression of Φ(s, 0) in terms of Riemann ζ-
functions, when s is an odd integer greater than 1.
Let ζ(s, x), where x > 0, denote the function expressed by the series

x−s + (x + 1)−s + (x + 2)−s + · · · ,

and its analytical continuations. Then

ζ(s, 1) = ζ(s), ζ(s, 12 ) = (2s − 1)ζ(s), (1)

where ζ(s) is the Riemann ζ-function;

ζ(s, x) − ζ(s, x + 1) = x−s ; (2)

)
1s + 2s + 3s + · · · + ns = ζ(−s) − ζ(−s, n + 1),
, (3)
1s + 3s + 5s + · · · + (2n − 1)s = (1 − 2s )ζ(−s) − 2s ζ(−s, n + 12 )

if n is a positive integer; and


  1−s
1 −s x s s(s + 1)(s + 2) −s−3
lim ζ(s, x) − 2 x + − B2 x−s−1 + B4 x
x→∞ 1−s 2! 4!

s(s + 1)(s + 2)(s + 3)(s + 4) −s−5
−B6 x + · · · to n terms = 0, (4)
6!

if n is a positive integer, −(2n − 1) < s < 1, and B2 = 61 , B4 = 30


1
, B6 = 1
42 , B8 = 1
30 , . . .,
are Bernoulli’s numbers.
Suppose now that √
Ψ(x) = 6ζ(− 12 , x) + (4x − 3) x + Φ(3, x).
Then from (2) we see that
√ √ √
Ψ(x) − Ψ(x + 1) = 6 x + (4x − 3) x − (4x + 1) (x + 1) + { (x + 1) − x}3 = 0;
p p
88 Paper 13

and from (4) that Ψ(x) → 0 as x → ∞. It follows that Ψ(x) = 0. That is to say,

6ζ(− 21 , x) + (4x − 3) x + Φ(3, x) = 0. (5)

Similarly, we can shew that


3 √
40ζ(− , x) + (16x2 − 20x + 5) x + Φ(5, x) = 0. (6)
2

2. Remembering the functional equation satisfied by ζ(s), viz.,


ζ(1 − s) = 2(2π)−s Γ(s)ζ(s) cos 21 πs, (7)

we see from (3) and (5) that


√ √ √ √ 2 3 √ 1 3 1
1 + 2 + 3 + · · · + n = n 2 + 12 n − ζ( ) + Φ(3, n); (8)
3 4π 2 6
and
√ √ √ p
1 + 3 + 5 + · · · + (2n − 1)

1 3 2−1 3 1
= (2n − 1) 2 + 21 (2n − 1) + ζ( ) + √ Φ(3, n − 21 ).
p
(9)
3 4π 2 3 2
Similarly from (6), we have
√ √ √ √
1 1 + 2 2 + 3 3 + ··· + n n
2 5 3 1√ 3 5 1
= n 2 + 12 n 2 + n− 2
ζ( ) + Φ(5, n); (10)
5 8 16π 2 40
and
√ √ √ p
1 1 + 3 3 + 5 5 + · · · + (2n − 1) (2n − 1)
1 5 3 1p
= (2n − 1) 2 + 21 (2n − 1) 2 + (2n − 1)
5 4

3(2 2 − 1 5 1
+ 2
ζ( 2 ) + √ Φ(5, n − 12 ). (11)
16π 10 2
It also follows from (5) and (6) that
p p p p
(a + d) + (a + 2d) + (a + 3d) + · · · + (a + nd)
2 3 1√
= C + (a + nd) 2 + 12 (a + nd) +
p
dΦ(3, n + a/d); (12)
3d 6
and
3 3 3 3
(a + d) 2 + (a + 2d) 2 + (a + 3d) 2 + · · · + (a + nd) 2
Summation of a certain series 89

2 5 3 1 p 1 √
= C′ + (a + nd) 2 + 12 (a + nd) 2 + d (a + nd) + d dΦ(5, n + a/d), (13)
5d 8 40
where C and C ′ are independent of n.
Putting n = 1 in (8) and (10), we obtain

3 3 15 5
Φ(3, 0) = ζ ( ), Φ(5, 0) = 2 ζ ( ). (14)
2π 2 2π 2

3. The preceding results may be generalised as follows. If s be an odd integer greater than
1, then
√ p √ p
Φ(s, x) + 21 { x + (x − 1)}s + 21 { x − (x − 1)}s
s s−2 s(s − 4)(s − 5) s−6
+ 2 ζ(1 − 21 s, x) + 2 ζ(3 − 21 s, x)
1! 3!
s(s − 6)(s − 7)(s − 8)(s − 9) s−10
+ 2 ζ(5 − 12 s, x)
5!
s(s − 8)(s − 9)(s − 10)(s − 11)(s − 12)(s − 13) s−14
+ 2
7!
×ζ(7 − 12 s, x) + · · · to [ 14 (s + 1)] terms = 0, (15)

where [x] denotes, as usual, the integral part of x. This can be proved by induction, using
the formula
√ p √ p
{ x + (x ± 1)}s + { x − (x ± 1)}s
√ s √ s(s − 3) √ s−4
= (2 x)s ± (2 x)s−2 + (2 x)
1! 2!
s(s − 4)(s − 5) √ s−6
± (2 x) + · · · to [1 + 12 s] terms, (16)
3!
which is true for all positive integral values of s.
Similarly, we can shew that if s is a positive even integer, then
s s−2
2 {ζ(1 − 21 s) − ζ(1 − 12 s, x)}
1!
s(s − 4)(s − 5) s−6
+ 2 {ζ(3 − 12 s) − ζ(3 − 21 s, x)}
3!
s(s − 6)(s − 7)(s − 8)(s − 9) s−10
+ 2 {ζ(5 − 12 s) − ζ(5 − 21 s, x)}
5!
+ · · · to [ 14 (s + 2)] terms
√ p √ p
= 21 { x + (x − 1)}s + 12 { x − (x − 1)}s − 1. (17)
90 Paper 13

Now, remembering (7) and putting x = 1 in (15), we obtain


s 1
Φ(s, 0) = − √ π − 2 (1+s) cos 14 πs{1 · 3 · 5 · · · (s − 2)πζ( 21 s)
2
− 3 · 5 · 7 · · · (s − 4) 21 (s − 5) 13 π 3 ζ( 12 s − 2)
+ 5 · 7 · 9 · · · (s − 6) 21 (s − 7) 14 (s − 9) 15 π 5 ζ( 12 s − 4)
− 7 · 9 · 11 · · · (s − 8) 12 (s − 9) 14 (s − 11) 16 (s − 13) 17 π 7 ζ( 12 s − 6)
+ 9 · 11 · 13 · · · (s − 10) 21 (s − 11) 14 (s − 13) 16 (s − 15) 18 (s − 17)
× 19 π 9 ζ( 12 s − 8) − · · · to [ 41 (s + 1)] terms }, (18)

If s is an odd integer greater than 1. Similarly, putting x = 12 in (15), we can express


Φ(s, 12 ) in terms of ζ-functions, if s is an odd integer greater than 1.

4. It is also easy to shew that, if


n=∞ p p
X { (x + n) + (x + n + 1)}−s
Ψ(s, x) = p ,
n=0 {(x + n)(x + n + 1)}

then
√ p √ p
{ x + (x − 1)}s − { x − (x − 1)}s
1
Ψ(s, x) − 2
p
{x(x − 1)}
s − 2 s−2 (s − 4)(s − 5)(s − 6) s−6
= 2 ζ(2 − 21 s, x) + 2 ζ(4 − 12 s, x)
1! 3!
(s − 6)(s − 7)(s − 8)(s − 9)(s − 10) s−10
+ 2 ζ(6 − 21 s, x)
5!
+ · · · to [ 14 (s + 1)] terms, (19)

provided that s is a positive odd integer. For example

√1 ,

Ψ(1, x) = x 




Ψ(3, x) = 4 x − √1x + 2ζ( 12 , x), (20)
√ √


Ψ(5, x) = 16x x − 12 x + √1x + 24ζ(− 21 , x), 

and so on.
New expressions for Riemann’s functions ξ(s) and Ξ(t)

Quarterly Journal of Mathematics, XLVI, 1915, 253 – 260

1. The principal object of this paper is to prove that if the real parts of α and β are
positive, and αβ = π 2 , and t is real, then
 
Z∞  2 2 4

1
 1 1 α x α x x dx 
α− 4 + 4α − + − · · ·
 1 + t2 32 + t2 1! 72 + t2 2! 112 + t2 e2πx − 1 
0
 
Z∞ 
x2 β2 x4

1
 1 1 β x dx 
− β− 4 + 4β − + − ···
 1 + t2 32 + t2 1! 72 + t2 2! 112 + t2 e2πx − 1 
0
− 34        
π −1 + it −1 − it t t β
= Γ Γ Ξ sin log . (1)
4t 4 4 2 8 α

Consider the integral


Z∞ 2
xe−πux
J(u) = dx,
e2πx − 1
0

where the real part of u is positive. Since

Z∞
sin πnx 1 1 1
πx
dx = 2πn + − ,
e −1 e − 1 2 2πn
0

we have
Z∞  
1 1 −πux2 1 1 1
J(u) + − √ = xe 2πx
+ − dx
4πu 4π u e − 1 2 2πx
0
Z∞ Z∞ Z∞ 2
2 sin πxy − 32 xe−πx /u
= xe−πux πy dx dy = u dx; (2)
e −1 e2πx − 1
0 0 0

and so
Z∞  
1 3
−πx2 /u 1 1
J(u) − √ = u− 2 xe 2πx
− dx. (3)
4π u e − 1 2πx
0
92 Paper 14

Suppose now that s = σ + it, where 0 < σ < 1. Then, from (3), we have
Z1 ( )
1
(s−1) 1
u2 J(nu) − p du
4π (nu)
0

Z1 Z∞  
− 23 1
(s−4) 2 /nu 1 1
=n u 2 du xe−πx − dx. (4)
e2πx − 1 2πx
0 0

Changing u into 1/v, we obtain


Z∞ Z∞  
− 23 − 12 s 2 /n 1 1
n v dv xe−πvx − dx
e2πx − 1 2πx
1 0
∞
Z∞  
− 23
Z
− 21 s −πvx2 /n 1 1
=n v dv xe 2πx
− dx
 e − 1 2πx
0 0

Z1

Z∞  
1 2 /n 1 1 
− v − 2 s dv xe−πvx − dx .
e2πx −1 2πx 
0 0
But
Z1 ( )
1
(s−1) 1
u 2 J(nu) − p du
4π (nu)
0

Z1
1 1
= u 2 (s−1) J(nu) du − √
2πs n
0

Z1 Z∞ 2
1 1
(s−1) xe−πnux
=− √ + u 2 du dx
2πs n e2πx − 1
0 0

Z∞ Z1
1 x dx 1 2
=− √ + 2πx
u 2 (s−1) e−πnux du
2πs n e −1
0 0
Z∞ 
πnx2 (πnx2 )2

1 1 x dx
=− √ +2 − + − ··· . (5)
2πs n 1 + s 1!(3 + s) 2!(5 + s) e2πx−1
0
Also
Z∞ Z∞  
− 23 − 21 s −πvx2 /n 1
1
n v dv xe 2πx
− dx
e − 1 2πx
0 0
New expressions for Riemann’s functions ξ(s) and Ξ(t) 93

Z∞   Z∞
− 23 1 1 1 2 /n
=n x 2πx
− dx v − 2 s e−πvx dv
e − 1 2πx
0 0
Z∞  
1 1 1 1 1
= π 2 (s−2) n− 2 (s+1) Γ(1 − s) xs−1 − dx
2 e2πx − 1 2πx
0
− 21 (s+1)
 s s − 1
n
=− √ Γ − Γ ξ(s), (6)
4π π 2 2

where
1 1
ξ(s) = (s − 1)Γ(1 + s)π − 2 s ζ(s).
2
Finally

Z1 Z∞  
− 23 − 21 s 2 /n 1 1
n v dv xe−πvx − dx
e2πx − 1 2πx
0 0

3 Z1 Z∞ Z1 Z∞ 2
n− 2 − 12 s −πvx2 /n − 23 − 12 s xe−πvx /n
=− v dv e dx + n v dv dx
2π e2πx − 1
0 0 0 0

Z1 Z∞ Z1
1 − 21 (1+s) − 32 x dx 1 2 /n
=− v dv + n 2πx
v − 2 s e−πvx dv
4πn e −1
0 0 0
Z∞ 
πx2 /n (πx2 /n)2

1 3 1 x dx
=− + 2n− 2 − + − ··· . (7)
2πn(1 − s) 2 − s 1!(4 − s) 2!(6 − s) e2πx−1
0

All the inversions of the order of integration, effected in the preceding argument, are easily
justified, since every integral remains convergent when the subject of integration is replaced
by its modulus.
It follows from (4) - (7) that, if the real parts of α and β are positive, and αβ = π 2 , then
 
Z∞  2 2 4

1
 1 1 α x α x x dx 
α− 4 − 4α − + − · · · 2πx
1 − s 1 + s 1! 3 + s 2! 5 + s e − 1
0
 
Z∞ 
x2 β2 x4
1 
1 1 β x dx 
+β − 4 − 4β − + − ···
s 2 − s 1! 4 − s 2! 6 − s e2πx − 1 
0
 1−1s   
1 3 α 8 4 s s−1
= π− 4 Γ − Γ ξ(s). (8)
2 β 2 2
94 Paper 14

Changing s to 21 (1 + it) in (8), and writing as usual


1 1 1
ξ( + it) = Ξ( t),
2 2 2
and equating the real and imaginary parts, we obtain the formula (1), and also the formula
 
Z∞  2 2 4

1
 1 3 α 7x α 11x x dx 
α− 4 − 4α − + − · · ·
 1 + t2 32 + t2 1! 72 + t2 2! 112 + t2 e2πx − 1 
0
 
Z∞ 
7x2 β2 11x4

− 41
 1 3 β x dx 
+β − 4β − + − ···
 1 + t2 32 + t2 1! 72 + t2 2! 112 + t2 e2πx − 1 
0
     
1 −3 −1 + it −1 − it 1 t α
= π 4Γ Γ Ξ( t) cos log . (9)
4 4 4 2 8 β

2. We have proved (8) on the assumption that 0 < σ < 1. But it can be shewn that the
formula is true for all values of s other than integral values.
Suppose first that −1 > σ < 0. The formula (3) is equivalent to
Z∞  
− 32 −πx2 /u 1 1 1
J(u) = u xe 2πx
− + dx. (10)
e − 1 2πx 2
0

Using this formula as we used (3) in the previous section, we can shew that (8) is true in
the strip −1 < σ < 0 also. In the right-hand side of (3), the first term in the expansion
of 1/(e2πx − 1), viz. 1/(2πx), is removed, and in that of (10) two terms are removed. By
considering the corresponding formulæ in which more and more terms in the expansion of
1/(e2πx − 1), viz.

1 1 πx π 3 x3 π 5 x5 π 7 x7 π 9 x9
− + − + − + − ···,
2πx 2 6 90 945 9450 93555
are removed,we can shew that the formula (8) is true in the strips −2 < σ < 1, −3 < σ < −2,
and so on. That it is also true in the strips 1 < σ < 2, 2 < σ < 3, . . . is easily deduced from
the functional equation ξ(s) = ξ(1 − s).
The formula also holds on the lines which divide the strips, except at the special points
s = k, where k is an integer. This follows at once from the continuity of ξ(s) and the
uniform convergence of the integrals in question.

3. As a particular case of (9) we have, when α = β = π,


Z∞ 
π 7x2 π 2 11x4

1 3 x dx
− 4π − + − ···
1 + t2 32 + t2 1! 72 + t2 2! 112 + t2 e2πx −1
0
New expressions for Riemann’s functions ξ(s) and Ξ(t) 95

   
1 −1 + it −1 − it 1
= √ Γ Γ Ξ( t). (11)
8 π 4 4 2
But the left-hand side of (11) is equal to
 
Z∞  Z∞  2 2 4 
πx −7z π x −11z x dx 
e − 4π
−z
e−3z
− e + e − · · · 2πx cos tz dz.
 1! 2! e − 1
0 0

Hence,
 
Z∞  Z∞ 2 −4z
xe−3z−πx e 
e−z − 4π dx cos tz dz
 e2πx − 1 
0 0
   
1 −1 + it −1 − it 1
= √ Γ Γ Ξ( t). (12)
8 π 4 4 2
It follows from this and Fourier’s theorem that
Z∞ 2 −4n
xe−πx e
e−n
− 4πe −3n
dx
e2πx − 1
0
Z∞    
1 −1 + it −1 − it 1
= √ Γ Γ Ξ( t) cos nt dt. (13)
4π π 4 4 2
0

But it is easily seen from (2) that, if α and β are positive and αβ = π 2 , then
   
Z∞ −αx2 Z∞ −βx2
1
 xe  1
 xe 
α− 4 1 + 4α dx = β −4
1 + 4β dx . (14)
 e2πx − 1   e2πx − 1 
0 0

From this it follows that the left-hand side of (13) is an even function of n. and so the
formula (13) is true for all real values of n.

4. It can easily be shewn that, if αβ = 4π 2 and R(s), where R(s) is the real part of s, is
greater than -1, then
ζ(1 − s) 1 (s−1) ζ(−s) 1 (s+1)
1 α2 + α2
4 cos 2 πs 8 sin 12 πs
Z∞ Z∞
1
(s+1) xs sin αxy
+α 2 dx dy
(e2πx − 1)(e2πy − 1)
0 0

ζ(1 − s) 1 (s−1) ζ(−s) 1 (s+1)


= 1 β2 + β2
4 cos 2 πs 8 sin 12 πs
96 Paper 14

Z∞ Z∞
1
(s+1) xs sin βxy
+β 2 dx dy. (15)
(e2πx − 1)(e2πy − 1)
0 0

From this we can shew, by arguments similar to those of §§ 1- 2, that if αβ = 4π 2 and


R(s) > −1 then
1 1 Z∞Z∞ 
ζ(1 − s) α 2 (s−1) ζ(−s) α 2 (s+1) 1 αxy (αxy)3
1 + 1 + α 2 (s+1) −
4 cos 2 πs s − 1 − t 8 sin 2 πs s + 1 − t 1!(s + 3 − t) 3!(s + 7 − t)
0 0
1 1
(αxy)5 xs dx dy ζ(1 − s) β 2 (s−1) β 2 (s+1)

ζ(−s)
+ − ··· + +
5!(s + 11 − t) (e2πx − 1)(e2πy − 1) 4 cos 12 πs s − 1 + t 8 sin 12 πs (s + 1 + t)
Z∞Z∞ 
(βxy)3 (βxy)5 xs dx dy

1
(s+1) βxy
+β 2 − + − ···
1!(s + 3 + t) 3!(s + 7 + t) 5!(s + 11 + t) (e2πx − 1)(e2πy − 1)
0 0
  1 t 1 (s−3) 1
α 4 22 Γ{ 4 (s − 1 + t)}Γ{ 14 (s − 1 − t)}
   
1+s+t 1+s−t
= ×ξ ξ . (16)
β π (s + 1)2 − t2 2 2

From this we deduce that, if αβ = 4π 2 and R(s) > −1, then

ζ(1 − s) s−1 1 1 ζ(−s) s+1 1 1


1 2 2
{α 2 (s−1) + β 2 (s−1) } + 1 2 2
{α 2 (s+1) + β 2 (s+1) }
4 cos 2 πs (s − 1) + t 8 sin 2 πs (s + 1) + t
Z∞ Z∞ 
(αxy)3 xs dx dy

1
(s+1) αxy s+3 s+7
+α 2
2 2
− + ···
1! (s + 3) + t 3! (s + 7)2 + t2 (e2πx − 1)(e2πy − 1)
0 0
Z∞ Z∞ 
(βxy)3 xs dx dy

1
(s+1) βxy s+3 s+7
+β 2
2 2
− + ···
1! (s + 3) + t 3! (s + 7)2 + t2 (e2πx − 1)(e2πy − 1)
0 0
1
2 2 Γ{ 14 (s − 1 + it)}Γ{ 41 (s − 1 − it)}
(s−3)
=
π (s + 1)2 + t2
     
t + is t − is 1 α
× Ξ Ξ cos t log ; (17)
2 2 4 β

and
ζ(1 − s) 1 1 1 ζ(−s) 1 1 1
1 2 2
{α 2 (s−1) − β 2 (s−1) } + 1 2 2
{α 2 (s+1) − β 2 (s+1) }
4 cos 2 πs (s − 1) + t 8 sin 2 πs (s + 1) + t
Z∞ Z∞ 
(αxy)3 xs dx dy

1
(s+1) αxy 1 1
+α 2
2 2
− + ···
1! (s + 3) + t 3! (s + 7)2 + t2 (e2πx − 1)(e2πy − 1)
0 0
New expressions for Riemann’s functions ξ(s) and Ξ(t) 97

Z∞ Z∞ 
(βxy)3 xs dx dy

1
(s+1) βxy 1 1
−β 2
2 2
− + ···
1! (s + 3) + t 3! (s + 7)2 + t2 (e2πx − 1)(e2πy − 1)
0 0
1
2 2
(s−3) Γ{ 41 (s − 1 + it)}Γ{ 14 (s − 1 − it)}
=
π (s + 1)2 + t2
     
t + is t − is 1 α
× Ξ Ξ sin t log . (18)
2 2 4 β

5. Proceeding as in § 3 we can shew that, if n is real, and


Z∞
Γ{ 14 (s − 1 + it)}Γ{ 41 (s − 1 − it)}
   
t + is t − is
F (n) = Ξ Ξ cos ntdt,
(s + 1)2 + t2 2 2
0

then, if R(s) > 1,


∞ 
1 1
Z x s dx 
F (n) = (4π)− 2 (s−3) − 2Γ(s)ζ(s) cosh n(1 − s) ; (19)
8  (exen − 1)(exe−n −1) 
0

if −1 < R(s) < 1,


Z∞   
1 1 1 1 1 1
F (n) = (4π)− 2 (s−3) x s
xe
− n − −n dx; (20)
8 e − 1 xe exe − 1 xe
n −n

if −3 < R(s) < −1,


∞
 
1 1
Z 1 1 1
F (n) = (4π)− 2 (s−3) x s
− +
8  exen − 1 xen 2
0
  
1 1 1
× − −n + dx − Γ(1 + s)ζ(1 + s) cosh n(1 + s) ; (21)
exe−n − 1 xe 2

and so on. If, in particular, we put s = 0 in (20), we obtain


Z∞    
−1 + it −1 − it 1 cos nt
Γ Γ {Ξ( t)}2 dt
4 4 2 1 + t2
0
Z∞ 

 
1 1 1 1
=π π − − dx. (22)
exen − 1 xen exe−n − 1 xe−n
0
Highly composite numbers
Proceedings of the London Mathematical Society, 2, XIV, 1915, 347 – 409

CONTENTS
I. §§ 1. Introduction and summary of results.

II. 2–5. Elementary results concerning the order of d(N ).


2. The order of d(N ) when the prime divisors of N are known.
3. The order of d(N ) when the number of prime divisors is known.
4–5. The order of d(N ) when nothing is known about N .

III. 6–31. The structure of highly composite numbers.


6. Definition of a highly composite number.
7. A table of the first 102 highly composite numbers.
8. First standard form of a highly composite number. The index of
the largest prime divisor.
9. Alternative standard form of a highly composite number.
10–24. The indices of the prime divisors in the first standard form.
25–27. The final primes in the second standard form.
28. The ratio of two consecutive highly composite numbers.
29. The form of d(N ) for highly composite values of N .
30. The order of dd(N ) for highly composite values of N .
31. Some apparently paradoxical forms of d(N ) in the table.

IV. 32–38. Superior highly composite numbers.


32. Definition of a superior highly composite number.
33. The exact form of a superior highly composite number.
34. The number of divisors of a superior highly composite number.
35. The maximum value of d(N )N 1/x for a given value of x.
36. Consecutive superior highly composite numbers.
37. The number of superior highly composite numbers less than a given
number. A table of the first 50 superior highly composite numbers.
38. Superior highly composite numbers and the maximum order of d(N ).

V. 39–45. Application to the determination of the order of d(N ).


39. The maximum order of d(N ), assuming the prime number theorem.
40–43. The maximum order of d(N ), assuming the Riemann hypothesis.
44. The order of the number of superior highly composite numbers
less than N .
45. Highly composite numbers between consecutive superior highly
composite numbers. General remarks.
Highly composite numbers 99

VI. §§ 46–51. Special forms of N .


46. The maximum order of d(N ) for a special form of N .
47. The maximum order of d(N ) when d(N ) is a power of 2.
48. The maximum order of d(N ) when the indices of the prime
divisors of N are never less than a given integer.
49. The maximum order of d(N ) when N is a perfect nth power.
50. The number of divisors of the least common multiple of the
first N natural numbers.
51. The number of divisors of N !.
52. On dd(n) and ddd(n).

This paper, as it appeared orginally, is not complete. Since the London Mathematical Society was in
some financial difficulty at that time, Ramanujan had to suppress part of what he had written in order
to save expense. The unpublished part of the manuscript with annotation has been reproduced: Highly
Composite Numbers by Srinivasa Ramanujan, Jean-Louis Nicolas, Guy Robin, The Ramanujan Journal,
Volume 1, Issue 2, 1997, pp. 119 – 153.
100 Paper 15

I.
Introduction and Summary of Results

1. The number d(N ) of divisors of N varies with extreme irregularity as N tends to infinity,
tending itself to infinity or remaining small according to the form of N . In this paper I
prove a large number of results which add a good deal to our knowledge of the behaviour
of d(N ).
It was proved by Dirichlet ∗ that
 
d(1) + d(2) + d(3) + · · · + d(N ) 1 †
= log N + 2γ − 1 + O √
N N

where γ is the Eulerian constant. Voronöi ‡ and Landau § have shewn that the error term
2 2
may be replaced by O(N − 3 +ǫ ), or indeed O(N − 3 log N ). It seems not unlikely that the real
3
value of the error is of the form O(N − 4 +ǫ ), but this is as yet unproved. Mr. Hardy has,
however, shewn recently ¶ that the equation

d(1) + d(2) + d(3) + · · · d(N ) 3


= log N + 2γ − 1 + o(N − 4 )
N
is certainly false. He has also proved that

d(1) + d(2)) + · · · + d(N − 1) + 21 d(N )) − N log N − (2γ − 1)N − 1


4
√ ∞
X d(n) p p
= N √ [H1 {4π (nN )} − Y1 {4π (nN )}],
n
1

where Yn is the ordinary second solution of Bessel’s equation, and


Z∞
2 we−xw dw
H1 (x) = p ;
π (w2 − 1)
1

and that the series on the right-hand side is the sum of the series
1 ∞
N 4 X d(n) p
√ 3 cos{4π (nN ) − 41 π},
π 2 1 n4

and an absolutely and uniformly convergent series.



Werke, Vol.2, p.49.

f = O(φ) means that a constant exists such that |f | < Kφ : f = o(φ) means that f /φ → 0.

Crelle’s Journal, Vol. 126. p. 241.
§
Göttinger Nachrichten, 1912.

Comptes Rendus May 10, 1915.
Highly composite numbers 101

The “average” order of d(N )) is thus known with considerable accuracy. In this paper I
consider, not the average order of d(N )), but its maximum order. This problem has been
much less studied. It is obvious that

d(N ) < 2 N .

It was shewn by Wigert∗ that


log N
(1+ǫ)
d(N ) < 2 log log N (i)
for all positive values of ǫ and all sufficiently large values of N , and that
log N
d(N ) > 2 log log N (1−ǫ) (ii)

for an infinity of values of N . From (i) it follows in particular that

d(N ) < N δ

for all positive values of δ and all sufficiently large values of N .


Wigert proves (i) by purely elementary reasoning, but uses the “Prime Number Theo-
remӠ to prove (ii). This is, however, unnecessary, the inequality (ii) being also capable of
elementary proof. In § 5 I shew, by elementary reasoning, that
log N log N
+O
d(N ) < 2 log log N (log log N)2

for all values of N , and


log N log N
+O
d(N ) > 2 log log N (log log N)2

for an infinity of values of N . I also shew later on that, if we assume all known results
concerning the distribution of primes, then

−a (log log N)
d(N ) < 2Li(log N )+O[log N e ]

for all values of N , and



−a (log log N)
d(N ) > 2Li(log N )+O[log N e ]

for an infinity of values of N , where a is a positive constant.


I then adopt a different point of view, I define a highly composite number as a number
whose number of divisors exceeds that of all its predecessors. Writing such a number in the
form
N = 2a2 · 3a3 · 5a5 · · · pap ,

Arkiv för Matematik, Vol. 3, No.18

The theorem that π(x) ∼ logx x , π(x) being the number of primes not exceeding x.
102 Paper 15

I prove that
a2 ≥ a3 ≥ a5 ≥ · · · ≥ ap ,
and that
ap = 1,
for all highly composite values of N except 4 and 36.
I then go on to prove that the indices near the beginning form a decreasing sequence in the
stricter sense, i.e., that
a 2 > a3 > a5 > · · · > a λ ,
where λ is a certain function of p.
Near the end groups of equal indices may occur, and I prove that there are actually groups
of indices equal to
1, 2, 3, 4, . . . , µ,
where µ again is a certain function of p. I also prove that if λ is fairly small in comparison
with p, then
log p
aλ log λ ∼ ;
log 2
and that the later indices can be assigned with an error of at most unity.
I prove also that two successive highly composite numbers are asymptotically equivalent,
i.e., that the ratio of two consecutive such numbers tends to unity. These are the most
striking results. More precise ones will be found in the body of the paper. These results
give us a fairly accurate idea of the structure of a highly composite number.
I then select from the general aggregate of highly composite numbers a special set which I
call “superior highly composite numbers”. I determine completely the general form of all
such numbers, and I shew how a combination of the idea of a superior highly composite
number with the assumption of the truth of the Riemann hypothesis concerning the roots
of the ζ-function leads to even more precise results concerning the maximum order of d(N ).
These results naturally differ from all which precede in that they depend on the truth of a
hitherto unproved hypothesis.

II.
Elementary Results concerning the Order of d(N ).

2. Let d(N ) denote the number of divisors of N , and let

N = pa11 pa22 pa33 · · · pann , (1)

where p1 , p2 , p3 , . . . , pn are a given set of n primes. Then

d(N ) = (1 + a1 )(1 + a2 )(1 + a3 ) · · · (1 + an ). (2)


Highly composite numbers 103

From (1) we see that

(1/n) log(p1 p2 p3 · · · pn N )
= (1/n){(1 + a1 ) log p1 + (1 + a2 ) log p2 + · · · + (1 + an ) log pn }
> {(1 + a1 )(1 + a2 )(1 + a3 ) · · · (1 + an ) log p1 log p2 · · · log pn }1/n .

Hence we have
{(1/n) log(p1 p2 p3 · · · pn N )}n
d(N ) < , (3)
log p1 log p2 log p3 · · · log pn
for all values of N .
We shall now consider how near to this limit it is possible to make d(N ) by choice of the
indices a1 , a2 , a3 , . . . , an . Let us suppose that
log pn
1 + am = v + ǫm (m = 1, 2, 3, . . . , n), (4)
log pm

where v is a large integer and − 21 < ǫm < 12 . Then, from (4), it is evident that

ǫn = 0. (5)

Hence, by a well-known theorem due to Dirichlet∗ , it is possible to choose values of v as


large as we please and such that

|ǫ1 | < ǫ, |ǫ2 | < ǫ, |ǫ3 | < ǫ, . . . , |ǫn−1 | < ǫ, (6)

where ǫ ≤ v −1/(n−1) . Now let

t = v log pn , δm = ǫm log pm . (7)

Then from (1), (4) and (7), we have


n
X
log(p1 p2 p3 · · · pn N ) = nt + δm . (8)
1

Similarly, from (2), (4) and (7) we see that

(t + δ1 )(t + δ2 ) · · · (t + δn )
d(N ) =
log p1 log p2 log p3 · · · log pn
P P 2 P 3 
δm δm δm
tn exp − + − · · ·
t 2t2 3t3
=
log p1 log p2 log p3 · · · log pn

Werke, Vol. 1, p. 635.
104 Paper 15

P 2
− ( δm )2 n2 δm
P 3
− ( δm )3
 P P 
n δm
n exp − + − ···
2nt2 3n2 t3
 P
δm
= t+
n log p1 log p2 log p3 · · · log pn
{(1/n) log(p1 p2 p3 · · · pn N )}n
=
log p1 log p2 · · · log pn
  X X 2  
1 −2 2 2
1 − 2 (log N ) n δm − n δm + ··· , (9)

in virtue of (8)). From (6), (7) and (9) it follows that it is possible to choose the indices
a1 , a2 , . . . , an so that

{(1/n) log(p1 p2 p3 · · · pn N )}n


d(N ) = {1 − O(log N )−2n/(n−1) }, (10)
log p1 log p2 · · · log pn

where the symbol O has its ordinary meaning.


The following examples shew how close an approximation to d(N ) may be given by the
right-hand side of (3). If
N = 272 · 725 ,
then, according to (3), we have

d(N ) < 1898.00000685 . . . ; (11)

and as a matter of fact d(N ) = 1898. Similarly, taking

N = 2568 · 3358 ,

we have, by (3),

d(N )) < 204271.000000372 . . . ; (12)

while the actual value of d(N ) is 204271. In a similar manner, when

N = 264 · 340 · 527 ,

we have, by (3),

d(N ) < 74620.00412 . . . ; (13)

while actually
d(N ) = 74620.

3. Now let us suppose that, while the number n of different prime factors of N remains
fixed, the primes pν , as well as the indices aν , are allowed to vary. It is evident that d(N ),
Highly composite numbers 105

considered as a function of N , is greatest when the primes pν are the first n primes, say
2, 3, 5, . . . , p, where p is the nth prime. It therefore follows from (3) that

{(1/n) log(2 · 3 · 5 · · · p · N )}n


d(N ) < , (14)
log 2 log 3 log 5 · · · log p

and from (10) that it is possible to choose the indices so that

{(1/n) log(2 · 3 · 5 · · · p · N )}n


d(N ) = {1 − O(log N )−2n/(n−1) }. (15)
log 2 log 3 log 5 · · · log p

4. Before we proceed to consider the most general case, in which nothing is known about
N , we must prove certain preliminary results. Let π(x) denote the number of primes not
exceeding x, and let
ϑ(x) = log 2 + log 3 + log 5 + · · · + log p,
and
̟(x) = log 2 · log 3 · log 5 · · · log p,
where p is the largest prime not greater than x; also let φ(t) be a function of t such that
φ′ (t) is continuous between 2 and x. Then

Zx Z3 Z5 Z7
′ ′ ′
π(t)φ (t)dt = φ (t)dt + 2 φ (t)dt + 3 φ′ (t)dt
2 2 3 5

Z11 Zx
+4 φ′ (t)dt + · · · + π(x) φ′ (t)dt
7 p

= {φ(3) − φ(2)} + 2{φ(5) − φ(3)} + 3{φ(7) − φ(5)}


+ 4{φ(11) − φ(7)} + · · · + π(x){φ(x) − φ(p)}
= π(x))φ(x) − {φ(2) + φ(3) + φ(5) + · · · + φ(p)}. (16)

As an example let us suppose that φ(t) = log t. Then we have


Zx
π(t)
π(x) log x − ϑ(x) = dt. (17)
t
2

Again let us suppose that φ(t) = log log t. Then we see that
Zx
π(t)
π(x) log log x − log ̟(x) = dt. (18)
t log t
2
106 Paper 15

But
Zx Zx Zx Zu
 
π(t) 1 π(t) 1 π(t) 
dt = dt +  dt du.
t log t log x t u(log u)2 t
2 2 2 2
Hence we have
 
ϑ(x)
π(x) log − log ̟(x)
π(x)
Zx Zx Zu
 
 
ϑ(x) 1 π(t) 1 π(t) 
= π(x) log + dt +  dt du. (19)
π(x) log x log x t u(log u)2 t
2 2 2

But
   
ϑ(x) π(x) log x − ϑ(x)
π(x) log = π(x) log 1 −
π(x) log x π(x) log x
Zx
 
 1 π(t) 
= π(x) log 1 − dt
 π(x) log x t 
2
Zx
1 π(t)
< − dt;
log x t
2

and so
  Zx
ϑ(x) 1 π(t)
π(x) log + dt < 0. (20)
π(x) log x log x t
2

Again,
   
ϑ(x) π(x) log x − ϑ(x)
π(x) log = −π(x) log 1 +
π(x) log x ϑ(x)
Zx Zx
 
 1 π(t)  π(x) π(t)
= −π(x) log 1 + dt > − dt;
 ϑ(x) t  ϑ(x) t
2 2

and so
  Zx Zx
ϑ(x) 1 π(t) π(x) log x − ϑ(x) π(t)
π(x) log + dt > − dt
π(x) log x log x t ϑ(x) log x t
2 2
 x 2
1 π(t) 
 Z
= − dt . (21)
ϑ(x) log x  t 
2
Highly composite numbers 107

It follows from (19), (20) and (21) that

Zx Zu
 
 
1 π(t)  ϑ(x)
 dt du > π(x) log − log ̟(x)
u(log u)2 t π(x)
2 2
Zx Zu
 
1 π(t) 
>  dt du
u(log u)2 t
2 2
 x 2
1 π(t) 
 Z
− dt .
ϑ(x) log x  t 
2

Now it is easily proved by elementary methods∗ that


   
x 1 1
π(x) = O , =O ;
log x ϑ(x) x

and so
Zx  
π(t) x
dt = O .
t log x
2

Hence
Zx Zu Zx 
 
  
1 π(t)  1 x
 dt du = O du = O ;
u(log u)2 t (log u)3 (log x)3
2 2 2

and  x 2
x2
   
1 π(t)  1 x
Z
dt = O =O .
ϑ(x) log x  t  ϑ(x) log x (log x)2 (log x)3
2

Hence we see that

{ϑ(x)/π(x)}π(x) 3
= eO[x/(log x) ] . (22)
̟(x)

5. We proceed to consider the case in which nothing is known about N . Let

N ′ = 2a1 · 3a2 · 5a3 · · · pan .

Then it is evident that d(N ) = d(N ′ ), and that

ϑ(p) ≤ log N ′ ≤ log N. (23)



See Landau, Handbuch, pp. 71 et seq.
108 Paper 15

It follows from (3) that

ϑ(p) + log N ′ π(p)


 
′1
d(N ) = d(N ) <
̟(p) π(p)
π(p)
{ϑ(p)/π(p)}π(p)

log N
≤ 1+
ϑ(p) ̟(p)
π(p) 3
log N π(p)+O[p/(log p) ]
  
log N O[p/(log p)3 ]
= 1+ e = 1+ , (24)
ϑ(p) ϑ(p)

in virtue of (22) and (23). But from (17) we know that


 
p
π(p) log p − ϑ(p) = O ;
log p

and so
ϑ(p) = π(p){log p + O(1)} = π(p){log ϑ(p) + O(1)}.
Hence
 
1 1
π(p) = ϑ(p) +O . (25)
log ϑ(p) {log ϑ(p)}2

It follows from (24) and (25) that


 ϑ(p) ϑ(p)
log N log ϑ(p) +O [log ϑ(p)]2

d(N ) ≤ 1 + .
ϑ(p)

Writing t instead of ϑ(p), we have


t
  +O t 2
log N log t (log t)
d(N ) ≤ 1+ ; (26)
t

and from (23) we have

t ≤ log N. (27)

Now, if N is a function of t, the order of the right-hand side of (26), considered as a


function of N , is increased when N is decreased in comparison with t, and decreased when
N is increased in comparison with t. Thus the most unfavourable hypothesis is that N ,
considered as a function of t, is as small as is compatible with the relation (27). We may
therefore write log N for t in (26). Hence
log N log N
+O
d(N ) < 2 log log N (log log N)2 , (28)
Highly composite numbers 109

for all values of N ∗


The inequality (28) has been proved by purely elementary reasoning. We have not assumed,
for example, the prime number theorem, expressed by the relation
x †
π(x) ∼ .
log x

We can also, without assuming this theorem, shew that the right-hand side of (28) is
actually the order of d(N ) for an infinity of values of N . Let us suppose that

N = 2 · 3 · 5 · 7 · · · p.

Then t t
+O
d(N ) = 2π(p) = 2 log t (log t)2 ,
in virtue of (25). Since log N = ϑ(p) = t, we see that
log N log N
+O
d(N ) = 2 log log N (log log N)2 ,

for an infinity of values of N . Hence the maximum order of d(N ) is


log N log N
+O
2 log log N (log log N)2 .

III.
The Structure of Highly Composite Numbers.

6. A number N may be said to be a highly composite number, if d(N ′ ) < d(N ) for all
values of N ′ less than N . It is easy to see from the definition that, if N is highly composite
and d(N ′ ) > d(N ), then there is at least one highly composite number M , such that

N < M ≤ N ′. (29)

If we assume nothing about π(x), we can shew that
log N log N log log log N
+O
d(N ) < 2 log log N (log log N )2 .

If we assume the prime number theorem, and nothing more, we can shew that
log N log N
+[1+O(1)]
d(N ) < 2 log log N (log log N )2 .
x x
If we assume that π(x) = +O ,
log x (log x)2
we can shew that
log N log N log N
+ +O
d(N ) < 2 log log N (log log N )2 (log log N )3 .


φ(x) ∼ Ψ(x) means that φ(x)/Ψ(x) → 1 as x → ∞.
110 Paper 15

if N and N ′ are consecutive highly composite numbers, then d(M ) ≤ d(N ) for all values of
M between N and N ′ . It is obvious that

d(N ) < d(2N ) (30)

for all values of N . It follows from (29) and (30) that, if N is highly composite, then there
is at least one highly composite number M such that N < M ≤ 2N . That is to say, there
is at least one highly composite number N , such that

x < N ≤ 2x, (31)

if x ≥ 1.

7. I do not know of any method for determining consecutive highly composite numbers
except by trial. The following table gives the consecutive highly composite values of N ,
and the corresponding values of d(N ) and dd(N ), up to d(N ) = 10080.
The numbers marked with the asterisk in the table are called superior highly composite
numbers. Their definition and properties will be found in §§ 32, 33.

dd(N ) d(N ) N
2 2=2 *2 = 2
2 3=3 4 = 22
3 4 = 22 *6 = 2·3
4 6=2·3 *12 = 22 · 3
4 8 = 23 24 = 23 · 3
3 9 = 32 36 = 22 · 32
4 10 = 2 · 5 48 = 24 · 3
6 12 = 22 · 3 *60 = 22 · 3 · 5
5 16 = 24 *120 = 23 · 3 · 5
6 18 = 2 · 32 180 = 22 · 32 · 5
6 20 = 22 · 5 240 = 24 · 3 · 5
8 24 = 23 · 3 ∗360 = 23 · 32 · 5
8 30 = 2 · 3 · 5 720 = 24 · 32 · 5
6 32 = 25 840 = 23 · 3 · 5 · 7
9 36 = 22 · 32 1260 = 22 · 32 · 5 · 7
8 40 = 23 · 5 1680 = 24 · 3 · 5 · 7
10 48 = 24 · 3 ∗2520 = 23 · 32 · 5 · 7
12 60 = 22 · 3 · 5 ∗5040 = 24 · 32 · 5 · 7
7 64 = 26 7560 = 23 · 33 · 5 · 7
12 72 = 23 · 32 10080 = 25 · 32 · 5 · 7
10 80 = 24 · 5 15120 = 24 · 33 · 5 · 7
12 84 = 22 · 3 · 7 20160 = 26 · 32 · 5 · 7
12 90 = 2 · 32 · 5 25200 = 24 · 32 · 52 · 7
Highly composite numbers 111

dd(N ) d(N ) N
12 96 = 25 · 3 27720 = 23 · 32 · 5 · 7 · 11
9 100 = 22 · 52 45360 = 24 · 34 · 5 · 7
12 108 = 22 · 33 50400 = 25 · 32 · 52 · 7
16 120 = 23 · 3 · 5 *55440 = 24 · 32 · 5 · 7 · 11
8 128 = 27 83160 = 23 · 33 · 5 · 7 · 11
15 144 = 24 · 32 110880 = 25 · 32 · 5 · 7 · 11
12 160 = 25 · 5 166320 = 24 · 33 · 5 · 7 · 11
16 168 = 23 · 3 · 7 221760 = 26 · 32 · 5 · 7 · 11
18 180 = 22 · 32 · 5 277200 = 24 · 32 · 52 · 7 · 11
14 192 = 26 · 3 332640 = 25 · 33 · 5 · 7 · 11
12 200 = 23 · 52 498960 = 24 · 34 · 5 · 7 · 11
16 216 = 23 · 33 554400 = 25 · 32 · 52 · 7 · 11
12 224 = 25 · 7 665280 = 26 · 33 · 5 · 7 · 11
20 240 = 24 · 3 · 5 ∗720720 = 24 · 32 · 5 · 7 · 11 · 13
9 256 = 28 1081080 = 23 · 33 · 5 · 7 · 11 · 13
18 288 = 25 · 32 ∗1441440 = 25 · 32 · 5 · 7 · 11 · 13
14 320 = 26 · 5 2162160 = 24 · 33 · 5 · 7 · 11 · 13
20 336 = 24 · 3 · 7 2882880 = 26 · 32 · 5 · 7 · 11 · 13
24 360 = 23 · 32 · 5 3603600 = 24 · 32 · 52 · 7 · 11 · 13
16 384 = 27 · 3 *4324320 = 25 · 33 · 5 · 7 · 11 · 13
15 400 = 24 · 52 6486480 = 24 · 34 · 5 · 7 · 11 · 13
20 432 = 24 · 33 7207200 = 25 · 32 · 52 · 7 · 11 · 13
14 448 = 26 · 7 8648640 = 26 · 33 · 5 · 7 · 11 · 13
24 480 = 25 · 3 · 5 10810800 = 24 · 33 · 52 · 7 · 11 · 13
24 504 = 23 · 32 · 7 14414400 = 26 · 32 · 52 · 7 · 11 · 13
10 512 = 29 17297280 = 27 · 33 · 5 · 7 · 11 · 13
21 576 = 26 · 32 ∗21621600 = 25 · 33 · 52 · 7 · 11 · 13
24 600 = 23 · 3 · 52 32432400 = 24 · 34 · 52 · 7 · 11 · 13
16 640 = 27 · 5 36756720 = 24 · 33 · 5 · 7 · 11 · 13 · 17
24 672 = 25 · 3 · 7 43243200 = 26 · 33 · 52 · 7 · 11 · 13
30 720 = 24 · 32 · 5 61261200 = 24 · 32 · 52 · 7 · 11 · 13 · 17
18 768 = 28 · 3 73513440 = 25 · 33 · 5 · 7 · 11 · 13 · 17
18 800 = 25 · 52 110270160 = 24 · 34 · 5 · 7 · 11 · 13 · 17
24 864 = 25 · 33 122522400 = 25 · 32 · 52 · 7 · 11 · 13 · 17
16 896 = 27 · 7 147026880 = 26 · 33 · 5 · 7 · 11 · 13 · 17
28 960 = 26 · 3 · 5 183783600 = 24 · 33 · 52 · 7 · 11 · 13 · 17
30 1008 = 24 · 32 · 7 245044800 = 26 · 32 · 52 · 7 · 11 · 13 · 17
11 1024 = 210 294053760 = 27 · 33 · 5 · 7 · 11 · 13 · 17
24 1152 = 27 · 32 ∗367567200 = 25 · 33 · 52 · 7 · 11 · 13 · 17
30 1200 = 24 · 3 · 52 551350800 = 24 · 34 · 52 · 7 · 11 · 13 · 17
112 Paper 15

dd(N ) d(N ) N
18 1280 = 28 · 5 698377680 = 24 · 33 · 5 · 7 · 11 · 13 · 17 · 19
28 1344 = 26 · 3 · 7 735134400 = 26 · 33 · 52 · 7 · 11 · 13 · 17
36 1440 = 25 · 32 · 5 1102701600 = 25 · 34 · 52 · 7 · 11 · 13 · 17
20 1536 = 29 · 3 1396755360 = 25 · 33 · 5 · 7 · 11 · 13 · 17 · 19
21 1600 = 26 · 52 2095133040 = 24 · 34 · 5 · 7 · 11 · 13 · 17 · 19
40 1680 = 24 · 3 · 5 · 7 2205403200 = 26 · 34 · 52 · 7 · 11 · 13 · 17
28 1728 = 26 · 33 2327925600 = 25 · 32 · 52 · 7 · 11 · 13 · 17 · 19
18 1792 = 28 · 7 2793510720 = 26 · 33 · 5 · 7 · 11 · 13 · 17 · 19
32 1920 = 27 · 3 · 5 3491888400 = 24 · 33 · 52 · 7 · 11 · 13 · 17 · 19
36 2016 = 25 · 32 · 7 4655851200 = 26 · 32 · 52 · 7 · 11 · 13 · 17 · 19
12 2048 = 211 5587021440 = 27 · 33 · 5 · 7 · 11 · 13 · 17 · 19
27 2304 = 28 · 32 ∗6983776800 = 25 · 33 · 52 · 7 · 11 · 13 · 17 · 19
36 2400 = 25 · 3 · 52 10475665200 = 24 · 34 · 52 · 7 · 11 · 13 · 17 · 19
32 2688 = 27 · 3 · 7 ∗13967553600 = 26 · 33 · 52 · 7 · 11 · 13 · 17 · 19
42 2880 = 26 · 32 · 5 20951330400 = 25 · 34 · 52 · 7 · 11 · 13 · 17 · 19
22 3072 = 210 · 3 27935107200 = 27 · 33 · 52 · 7 · 11 · 13 · 17 · 19
48 3360 = 25 · 3 · 5 · 7 41902660800 = 26 · 34 · 52 · 7 · 11 · 13 · 17 · 19
32 3456 = 27 · 33 48886437600 = 25 · 33 · 52 · 72 · 11 · 13 · 17 · 19
20 3584 = 29 · 7 64250746560 = 26 · 33 · 5 · 7 · 11 · 13 · 17 · 19 · 23
45 3600 = 24 · 32 · 52 73329656400 = 24 · 34 · 52 · 72 · 11 · 13 · 17 · 19
36 3840 = 28 · 3 · 5 80313433200 = 24 · 33 · 52 · 7 · 11 · 13 · 17 · 19 · 23
42 4032 = 26 · 32 · 7 97772875200 = 26 · 33 · 52 · 72 · 11 · 13 · 17 · 19
13 4096 = 212 128501483120 = 27 · 33 · 5 · 7 · 11 · 13 · 17 · 19 · 23
48 4320 = 25 · 33 · 5 146659312800 = 25 · 34 · 52 · 72 · 11 · 13 · 17 · 19
30 4608 = 29 · 32 160626866400 = 25 · 33 · 52 · 7 · 11 · 13 · 17 · 19 · 23
42 4800 = 26 · 3 · 52 240940299600 = 24 · 34 · 52 · 7 · 11 · 13 · 17 · 19 · 23
60 5040 = 7 · 5 · 32 · 24 293318625600 = 26 · 34 · 52 · 72 · 11 · 13 · 17 · 19
36 5376 = 28 · 3 · 7 *321253732800 = 26 · 33 · 52 · 11 · 13 · 17 · 19 · 23
48 5760 = 27 · 32 · 5 481880599200 = 25 · 34 · 52 · 7 · 11 · 13 · 17 · 19 · 23
24 6144 = 211 · 3 642507465600 = 27 · 33 · 52 · 7 · 11 · 13 · 17 · 19 · 23
56 6720 = 26 · 3 · 5 · 7 963761198400 = 26 · 34 · 52 · 7 · 11 · 13 · 17 · 19 · 23
36 6912 = 28 · 33 1124388064800 = 25 · 33 · 52 · 72 · 11 · 13 · 17 · 19 · 23
22 7168 = 210 · 7 1606268664000 = 26 · 33 · 53 · 7 · 11 · 13 · 17 · 19 · 23
54 7200 = 25 · 32 · 52 1686582097200 = 24 · 34 · 52 · 72 · 11 · 13 · 17 · 19 · 23
40 7680 = 29 · 3 · 5 1927522396800 = 27 · 34 · 52 · 7 · 11 · 13 · 17 · 19 · 23
48 8064 = 27 · 32 · 7 ∗2248776129600 = 26 · 33 · 52 · 72 · 11 · 13 · 17 · 19 · 23
14 8192 = 213 3212537328000 = 27 · 33 · 53 · 7 · 11 · 13 · 17 · 19 · 23
56 8640 = 26 · 33 · 5 3373164194400 = 25 · 34 · 52 · 72 · 11 · 13 · 17 · 19 · 23
33 9216 = 210 · 32 4497552259200 = 27 · 33 · 52 · 72 · 11 · 13 · 17 · 19 · 23
72 10080 = 25 · 32 · 5 · 7 6746328388800 = 26 · 34 · 52 · 72 · 11 · 13 · 17 · 19 · 23
Highly composite numbers 113

8. Now let us consider what must be the nature of N in order that N should be a highly
composite number. In the first place it must be of the form
ap
2a2 · 3a3 · 5a5 · 7a7 · · · p1 1 ,

where

a2 ≥ a3 ≥ a5 ≥ · · · ≥ ap1 ≥ 1. (32)

This follows at once from the fact that


ap ap
d(̟2a2 ̟3a3 ̟5a5 · · · ̟p1 1 ) = d(2a2 · 3a3 · 5a5 · · · p1 1 ),

for all prime values of ̟2 , ̟3 , ̟5 , . . . , ̟p1 .


It follows from the definition that, if N is highly composite and N ′ < N , then d(N ′ ) must
be less that d(N ). For example, 65 N < N , and so d( 65 N ) < d(N ). Hence
    
1 1 1
1+ 1+ > 1+ ,
a2 a3 1 + a5
provided that N is a multiple of 3.
It is convenient to write

aλ = 0 (λ > p1 ). (33)

Thus if N is not a multiple of 5 then a5 should be considered as 0.


Again, ap1 must be less than or equal to 2 for all values of p1 . For let P1 be the prime next
above p1 . Then it can be shewn that P1 < p21 for all values of p1 .∗
Now, if ap1 is greater than 2, let
N P1
N′ = .
P12
Then N ′ is an integer less than N , and so d(N ′ ) < d(N ). Hence

(1 + ap1 ) > 2(ap1 − 1),

or
3 > ap 1 ,

It can be proved by elementary methods that, if x ≥ 1, there is at least one prime p such that x < p ≤ 2x.
This result is known as Bertrand’s Postulate: for a proof, see Landau, Handbuch, p. 89. It follows at once
that P1 < p21 , if p1 > 2; and the inequality is obviously true when p1 = 2. Some similar results used later in
this and the next section may be proved in the same kind of way. It is for some purposes sufficient to know
that there is always a prime p such that x < p < 3x, and the proof of this is easier than that of Bertrand’s
Postulate. These inequalities are enough, for example, to shew that

log P1 = log p1 + O(1).


114 Paper 15

which contradicts our hypothesis. Hence

ap1 ≤ 2, (34)

for all values of p1 .


′′
Now let p1 , p′1 , p1 , P1 , P1′ be consecutive primes in ascending order. Then, if p1 ≥ 5, ap′′
1
must be less than or equal to 4. For, if this were not so, we could suppose that
N P1
N′ = ′′ .
(p1 )3
But it can easily be shewn that, if p1 ≥ 5, then
′′
(p1 )3 > P1 ;

and so N ′ < N and d(N ′ ) < d(N ). Hence

(1 + ap′′ ) > 2(ap′′ − 2). (35)


1 1

But since ap′′ ≥ 5, it is evident that


1

(1 + ap′′ ) ≤ 2(ap′′ − 2),


1 1

which contradicts (35); therefore, if p1 ≥ 5, then

ap′′ ≤ 4. (36)
1

Now let ′′
N p1 P1
N′ = .
p′1 p1
It is easy to verify that, if 5 ≤ p1 ≤ 19, then
′′
p′1 p1 > p1 P1 ;

and so N ′ < N and d(N ′ ) < d(N ). Hence

(1 + ap1 )(1 + ap′1 )(1 + ap′′ ) > 2ap1 ap′1 (2 + ap′′ ),


1 1

or   ! !
1 1 1
1+ 1+ >2 1+ ,
ap 1 ap′1 1 + ap′′
1

But from (36) we know that 1 + ap′′ ≤ 5. Hence


1

  !
1 1
1+ 1+ > 2 25 . (37)
ap 1 ap′1
Highly composite numbers 115

From this it follows that ap1 = 1. For, if ap1 ≥ 2, then


  !
1 1
1+ 1+ ≤ 2 14 ,
ap1 ap′1

in virtue of (32). This contradicts (37). Hence, if 5 ≤ p1 ≤ 19, then

ap1 = 1. (38)

Next let
′′
N ′ = N P1 P1′ /(p1 p′1 p1 ).
It can easily be shewn that, if p1 ≥ 11, then
′′
P1 P1′ < p1 p′1 p1 ;

and so N ′ < N and d(N ′ ) < d(N ). Hence

(1 + ap1 )(1 + ap′1 )(1 + ap′′ ) > 4ap1 ap′1 ap′′ ,


1 1

or
  ! !
1 1 1
1+ 1+ 1+ > 4. (39)
ap1 ap′1 ap′′
1

From this we infer that ap1 must be 1. For, if ap1 ≥ 2, it follows from (32) that
  ! !
1 1 1
1+ 1+ 1+ ≤ 3 38 ,
ap1 ap′1 ap′′
1

which contradicts (39). Hence we see that, if p1 ≥ 11, then

ap1 = 1. (40)

It follows from (38) and (40) that, if p1 ≥ 5, then

ap1 = 1. (41)

But if p1 = 2 or 3, then from (34) it is clear that

ap1 = 1 or 2. (42)

It follows that ap1 = 1 for all highly composite numbers, except for 22 , and perhaps for
certain numbers of the form 2a · 32 . In the latter case a ≥ 2. It is easy to shew that, if
a ≥ 3, 2a .32 cannot be highly composite. For if we suppose that

N ′ = 2a−1 · 3 · 5,
116 Paper 15

then it is evident that N ′ < N and d(N ′ ) < d(N ), and so

3(1 + a) > 4a,

or
a < 3.
Hence it is clear that a cannot have any other value except 2. Moreover we can see by
actual trial that 22 and 22 · 32 are highly composite. Hence

ap 1 = 1 (43)

for all highly composite values of N save 4 and 36, when

ap1 = 2.

Hereafter when we use this result it is to be understood that 4 and 36 are exceptions.

9. It follows from (32) and (43) that N must be of the form


2 · 3 · 5 · 7 · · · p1
× 2 · 3 · 5 · 7 · · · p2
× 2 · 3 · 5 · · · p3
× ..., (44)

where p1 > p2 ≥ p3 ≥ p4 ≥ · · · and the number of rows is a2 .


Let Pr be the prime next above pr , so that

log Pr = log pr + O(1), (45)

in virtue of Bertrand’s Postulate. Then it is evident that

apr ≥ r, aPr ≤ r − 1; (46)

and so

aPr ≤ apr − 1. (47)

It is to be understood that

aP1 = 0, (48)

in virtue of (33).
It is clear from the form of (44) that r can never exceed a2 , and that

paλ = λ. (49)
Highly composite numbers 117

10. Now let


N [log ν/ log λ] ∗
N′ =
λ ,
ν
where ν ≤ p1 so that N ′ is an integer. Then it is evident that N ′ < N and d(N ′ ) < d(N ),
and so   
log ν
(1 + aν )(a + aλ ) > aν 1 + aλ + ,
log λ
or
 
log ν
1 + a λ > aν . (50)
log λ
Since the right-hand side vanishes when ν > p1 , we see that (50) is true for all values of λ
and ν † .
Again let
−1−[log µ/ log λ]
N ′ = Nµλ ,
where [log µ/ log λ] < aλ , so that N ′ is an integer. Then it is evident that N ′ < N and
d(N ′ ) < d(N ), and so
  
log µ
(1 + aµ )(1 + aλ ) > (2 + aµ ) aλ − . (51)
log λ
Since the right-hand side is less than or equal to 0 when

aλ ≤ [log µ/ log λ],

we see that (51) is true for all values of λ and µ. From (51) it evidently follows that
 
log(λµ)
(1 + aλ ) < (2 + aµ ) . (52)
log λ
From (50) and (52) it is clear that
   
log ν log µ
aν ≤ aλ ≤ aµ + (2 + aµ ) , (53)
log λ log λ
for all values of λ, µ and ν.
Now let us suppose that ν = p1 and µ = P1 , so that aν = 1 and aµ = 0. Then we see that
   
log p1 log P1
≤ aλ ≤ 2 , (54)
log λ log λ
for all values of λ. Thus, for example, we have

p1 = 3, 1 ≤ a2 ≤ 4;

[x] denotes as usual the integral part of x.

That is to say all prime values of λ and ν, since λ in aλ is be definition prime.
118 Paper 15

p1 = 5, 2 ≤ a2 ≤ 4;
p1 = 7, 2 ≤ a2 ≤ 6;
p1 = 11, 3 ≤ a2 ≤ 6;

and so on. It follows from (54) that, if λ ≤ p1 , then

aλ log λ = O(log p1 ), aλ log λ 6= o(log p1 ). (55)

11. Again let


√ √
′ [ {(1+aλ +aµ ) log µ/ log λ}] −1−[ {(1+aλ +aµ ) log λ/ log µ}]
N = Nλ µ ,

and let us assume for the moment that


q
aµ > {(1 + aλ + aµ ) log λ/ log µ},

in order that N ′ may be an integer. Then N ′ < N and d(N ′ ) < d(N ), and so
p
(1 + aλ )(1 + aµ ) > {1 + aλ + [ {(1 + aλ + aµ ) log µ/ log λ}]}
p
×{aµ − [ {(1 + aλ + aµ ) log λ/ log µ}]}
p
> {aλ + {(1 + aλ + aµ ) log µ/ log λ}}
p
×{aµ − {(1 + aλ + aµ ) log λ/ log µ}}. (56)

It is evident that the right-hand side of (56) becomes negative when


q
aµ < {(1 + aλ + aµ ) log λ/ log µ},

while the left-hand side remains positive, and so the result is still true. Hence
q
aµ log µ − aλ log λ < 2 {(1 + aλ + aµ ) log λ log µ}, (57)

for all values of λ and µ. Interchanging λ and µ in (57), we obtain


q
aλ log λ − aµ log µ < 2 {(1 + aλ + aµ ) log λ log µ}. (58)

From (57) and (58) it evidently follows that


q
|aλ log λ − aµ log µ| < 2 {(1 + aλ + aµ ) log λ log µ}, (59)

for all values of λ and µ. It follows from this and (55) that, if λ and µ are neither greater
than p1 , then
p
aλ log λ − aµ log µ = O {log p1 log(λµ)}, (60)
Highly composite numbers 119

and so that, if log λ = o(log p1 ), then

a2 log 2 ∼ a3 log 3 ∼ a5 log 5 ∼ · · · ∼ aλ log λ. (61)

12. It can easily be shewn by elementary algebra that, if x, y, m and n are not negative,
and if p
|x − y| < 2 (mx + ny + mn),
then
p p p )
| (x + n) − (y + m)| < (m + n);
p p p (62)
| (x + n) − (m + n)| < (y + m).

From (62) and (59) it follows that


p q p
| {(1 + aλ ) log λ} − {(1 + aµ ) log µ}| < {log(λµ)}, (63)

and
p p q
| {(1 + aλ ) log λ} − {log(λµ)}| < {(1 + aµ ) log µ}, (64)

for all values of λ and µ. If, in particular, we put µ = 2 in (63), we obtain


p p p
{(1 + a2 ) log 2} − {log(2λ)} < {(1 + aλ ) log λ}
p p
< {(1 + a2 ) log 2} + {log(2λ)}, (65)

for all values of λ. Again, from (63), we have

(1 + aλ ) log λ < ( {(1 + aν ) log ν} + {log(λν)})2 ,


p p

or
p
aλ log λ < (1 + aν ) log ν + log ν + 2 {(1 + aν ) log ν log(λν)}. (66)

Now let us suppose that λ ≤ µ. Then, from (66), it follows that


p
aλ log λ + log µ < (1 + aν ) log ν + log(µν) + 2 {(1 + aν ) log ν log(λν)}
p
≤ (1 + aν ) log ν + log(µν) + 2 {(1 + aν ) log ν log(µν)}
= { {(1 + aν ) log ν} + log(µν)}2 ,
p p
(67)

with the condition that λ ≤ µ. Similarly we can shew that

aλ log λ + log µ > { {(1 + aν ) log ν} − log(µν)}2 ,


p p
(67′ )
120 Paper 15

with the condition that λ ≤ µ.

13. Now let


N [log λ/{π(µ) log 2}] [log λ/{π(µ) log 3}]
N′ = 2 3 · · · µ[log λ/{π(µ) log µ}] ,
λ
where π(µ) log µ < log λ ≤ log p1 . Then it is evident that N ′ is an integer less than N , and
so d(N ′ ) < d(N ). Hence
 
1
1+ (1 + a2 )(1 + a3 )(1 + a5 ) · · · (1 + aµ )

    
log λ log λ log λ
> a2 + a3 + · · · aµ + ;
π(µ) log 2 π(µ) log 3 µ(µ) log µ
that is
    
log λ log λ log λ
a2 log 2 + a3 log 3 + · · · aµ log µ +
π(µ) π(µ) π(µ)
 
1
< 1+ (a2 log 2 + log 2)(a3 log 3 + log 3) · · · (aµ log µ + log µ)

 
1
≤ 1+ (a2 log 2 + log µ)(a3 log 3 + log µ) · · · (aµ log µ + log µ).

In other words
 
1
1+

( log λ )( log λ ) ( log λ )
π(µ) − log µ π(µ) − log µ π(µ) − log µ
> 1+ 1+ ··· 1 +
a2 log 2 + log µ a3 log 3 + log µ aµ log µ + log µ
( log λ )π(µ)
− log µ
π(µ)
> 1+ p p , (68)
{ {(1 + aν ) log ν} + log(µν)}2

w here ν is any prime, in virtue of (67). From (68) it follows that


v 
u
log λ
π(µ) − log µ
u 

p p u
{(1 + aν ) log ν} + log(µν) > t 
u 1/π(µ) , (69)
 1+ 1

−1


provided that π(µ) log µ < log λ ≤ log p1 .

14. Again let


N ′ = N λ2−1−[log λ/{π(µ) log 2}] 3−1−[log λ/{π(µ) log 3}] · · · µ−1−[log λ/{π(µ) log µ}] ,
Highly composite numbers 121

where µ ≤ p1 and λ > µ. Let us assume for the moment that


log λ
aκ log κ > ,
π(µ)

for all values of κ less than or equal to µ, so that N ′ is an integer. Then, by arguments
similar to those of the previous section, we can shew that
( log λ )π(µ)
1 + aλ + log µ
π(µ)
> 1− p p . (70)
2 + aλ { {(1 + aν ) log ν} − log(µν)}2

From this it follows that


v 
u
log λ
π(µ) + log µ
u 

p p u
| {(1 + aν ) log ν} − log(µν)| < t
u  1/π(µ) , (71)
 1 − 1+aλ
 
2+aλ

provided that µ ≤ p1 and µ < λ. The condition that

aκ log κ > {log λ/π(µ)}

is unnecessary because we know from (67′ ) that


s 
p p p log λ
| {(1 + aν ) log ν} − log(µν)| < (aκ log κ + log µ) ≤ + log µ , (72)
π(µ)

when
aκ log κ ≤ {log λ/π(µ)},
and the last term in (72) is evidently less than the right-hand side of (71).

15. We shall consider in this and the following sections some important deductions from
the preceding formulæ. Putting ν = 2 in (69) and (71), we obtain
v 
u
log λ
π(µ) − log µ
u 
p u  p
{(1 + a2 ) log 2} > t 
u 1/π(µ)
− log(2µ), (73)
 1+ 1

− 1


provided that π(µ) log µ < log λ ≤ log p1 , and


v 
u
log λ
π(µ) + log µ
u 
 p
p u
{(1 + a2 ) log 2} < t
u  1/π(µ)  + log(2µ), (74)
 1 − 1+aλ
 
2+aλ
122 Paper 15

provided that µ ≤ p1 , and µ < λ. Now supposing that λ = p1 in (73), and λ = P1 in (74),
we obtain
v(
u log p1 )
p u π(µ) − log µ p
{(1 + a2 ) log 2} > t − log(2µ), (75)
21/π(µ) − 1

provided that π(µ) log µ < log p1 , and


v(
u log P1 )
p u π(µ) + log µ p
{(1 + a2 ) log 2} < t −1/π(µ)
+ log(2µ), (76)
1−2

provided that µ ≤ p1 . In (75) and (76) µ can be so chosen as to obtain the best possible
inequality for a2 . If p1 is too small, we may abandon this result in favour of
   
log p1 log P1
≤ a2 ≤ 2 , (77)
log 2 log 2

which is obtained from (54) by putting λ = 2.


After having obtained in this way what information we can about a2 , we may use (73) and
(74) to obtain information about aλ . Here also we have to choose µ so as to obtain the
best possible inequality for aλ . But if λ is too small we may, instead of this, use
p p p
{(1 + a2 ) log 2} − log(2λ) < {(1 + aλ )logλ}
p p
< {(1 + a2 ) log 2} + log(2λ), (78)

which is obtained by putting µ = 2 in (63).

16. Now let us consider the order of a2 . From (73) it is evident that, if π(µ) log µ < log λ ≤
log p1 , then
log λ
p π(µ) − log µ
(1 + a2 ) log 2 + log(2µ) + 2 {(1 + a2 ) log 2 log(2µ)} >  1/π(µ) (79)
1 + a1λ −1

But we know that for positive values of x,


 
1 1 1 1
x
= + O(1), x =O .
e −1 x e −1 x

Hence
 
log λ 1 log λ  π(µ) 
1/π(µ) =   + O(1)
π(µ)  1 π(µ)  log 1 + 1
1+ −1

aλ aλ
Highly composite numbers 123

 
log λ log λ
=   +O ;
log 1 + a1λ π(µ)

and  
log µ  π(µ) log µ 
1/π(µ) =O   = O(µaλ ).

1  log 1 + 1 
1+ aλ −1 aλ

Again from (55) we know that a2 = O(log p1 ). Hence (79) may be written as
p
a2 log 2 + O (log p1 log µ) + O(log µ)
 
log λ log λ
≥   +O + O(µaλ ). (80)
log 1 + a1λ π(µ)

But
log µ = O(µaλ ),
 
µ µ log p1
µaλ = · aλ log λ = O ,
log λ log λ
 
log λ log λ log µ
=O .
π(µ) µ
Again
log λ log µ µ log p1 p
+ > 2 (log p1 log µ);
µ log λ
and so  
p log λ log µ µ log p1
(log p1 log µ) = O + .
µ log λ
Hence (80) may be replaced by
 
log λ log λ log µ µ log p1
a2 log 2 ≥   +O + , (81)
log 1 + a1λ µ log λ

provided that π(µ) log µ < log λ ≤ log p1 . Similarly, from (74), we can shew that
 
log λ log λ log µ µ log p1
a2 log 2 ≤   +O + , (82)
log 1 + 1+a1 µ log λ
λ

provided that µ ≤ p1 and µ < λ. Now supposing that λ = p1 in (81), and λ = P1 in (82),
and also that p p
µ = O (log p1 log log p1 ),µ 6= o (log p1 log log p1 ), ∗

f 6= o(φ) is to be understood as meaning that |f | > Kφ, where K is a constant, and f 6= O(φ) as
meaning that |f |/φ → ∞. They are not the mere nagations of f = o(φ) and f = O(φ).
124 Paper 15

we obtain

log p1
p
a2 log 2 ≥ log 2 +O (log p1 log log p1 ), 
(83)
log p1
p
a2 log 2 ≤ log 2 +O (log p1 log log p1 ). 

From (83) it evidently follows that

log p1 p
a2 log 2 = + O (log p1 log log p1 ). (84)
log 2

And it follows from this and (60) that if λ ≤ p1 then

log p1 p p
aλ log λ = + O{ log p1 log λ) + (log p1 log log p1 )}. (85)
log 2

Hence, if log λ = o(log p1 ), we have

log p1
a2 log 2 ∼ a3 log 3 ∼ a5 log 5 ∼ · · · ∼ aλ log λ ∼ . (86)
log 2

17. The relations (86) give us information about the order of aλ when λ is sufficiently small
compared to p1 , in fact, when λ is of the form pǫ1 , where ǫ → 0. Such values of λ constitute
but a small part of its total range of variation, and it is clear that further formulæ must
be proved before we can gain an adequate idea of the general behaviour of aλ . From (81),
(82) and (84) it follows that
 
log λ log p1 log λ log µ µ log p1 p

 ≤ +O + + (log p1 log log p1 ) ,  
1 log 2 µ log λ

log 1 +


aλ 
  (87)
log λ log p1 log λ log µ µ log p1 p 
 ≥ +O + + (log p1 log log p1 ) , 

log 2 µ log λ

1
log 1 + 1+aλ

provided that π(µ) log µ log λ ≤ log p1 . From this we can easily shew that if

π(µ) log µ log λ ≤ log p1

then
( p ) 
log µ µ log p 1 (log p 1 log log p 1 ) 
aλ ≤ (2log λ/ log p1 − 1)−1 + O + + ,


µ (log λ)2 log λ



( p ) (88)
log µ µ log p 1 (log p 1 log log p 1 ) 
aλ ≥ (2log λ/ log p1 − 1)−1 − 1 + O

+ + .


µ (log λ)2 log λ


Highly composite numbers 125

Now let us suppose that s 


log p1
log λ 6= o .
log log p1
Then we can choose µ so that
( s )
log log p1
µ = O log λ ,
log p1
( s )
log log p1
µ 6= o log λ .
log p1
Now it is clear that log µ = O(log log p1 ), and so
  (p )
log µ log log p1 (log p1 log log p1 )
=O =O ;
µ µ log λ

and (p )
µ log p1 (log p1 log log p1 )
=O .
(log λ)2 log λ
From this and (88) it follows that, if
s 
log p1
log λ 6= o ,
log log p1

then
(p ) 
(log p1 log log p1 )

aλ ≤ (2log λ/ log p1 − 1)−1 +O , 

log λ



(p ) (89)
(log p1 log log p1 ) 
aλ ≥ (2log λ/ log p1 − 1)−1 + O

. 


log λ 

Now we shall divide that primes from 2 to p1 into five ranges thus

V IV III II RANGE I

2 p1
(  1 )
1 lp 2 1
(lp1 )κ e{κ(lp) }e κ
3 e{κ(lpl2 p) 2 }
l2 p
126 Paper 15

We shall use the inequalities (89) to specify the behavior of aλ in ranges I and II, and the
formula (85) in ranges IV and V. Range III we shall deal with differently, by a different
choice of µ in the inequalities (88). We can easily see that each result in the following
sections gives the most information in its particular range.

18. Range I: p
log λ 6= O (log p1 log log p1 ).∗
Let
Λ = [(2log λ/ log p1 − 1)−1 ],
and let
(2log λ/ log p1 − 1)−1 + ǫλ ,
where − 21 < ǫλ < 21 , be an integer, so that

(2log λ/ log p1 − 1)−1 = Λ + 1 − ǫλ (90)

when ǫλ > 0, and

(2log λ/ log p1 − 1)−1 = Λ − ǫλ (91)

when ǫλ < 0. By our supposition we have


p
(log p1 log log p1 )
= o(1). (92)
log λ
First let us consider the case in which
(p )
(log p1 log log p1 )
ǫλ 6= O ,
log λ

so that
p
(log p1 log log p1 )
= o(ǫλ ). (93)
log λ
It follows from (89), (90), and (93) that, if ǫλ > 0, then
)
aλ ≤ Λ + 1 − ǫλ + o(ǫλ ),
(94)
aλ ≥ Λ − ǫλ + o(ǫλ ).

We can with a little trouble replace all equations of the type f = O(φ) which occur by inequalities
of the type |f | < Kφ, with definite numerical constants. This would enable us to extend all the different
ranges a little. For example, an equation true for
p
log λ 6= O (log p1 )
p
would bepreplaced by an inequality true for log λ >pK (log p1 ), where K is definite constant, and similarly
log λ = o (log p1 ) would be replaced by log λ < k (log p1 ).
Highly composite numbers 127

Since 0 < ǫλ < 12 , and aλ and Λ are integers, it follows from (94) that

aλ ≤ Λ, aλ > Λ − 1. (95)

Hence

aλ = Λ. (96)

Similarly from (89) (91) and (93) we see that, if ǫλ < 0, then
)
aλ ≤ Λ − ǫλ + o(ǫλ ),
(97)
aλ ≥ Λ − 1 − ǫλ + o(ǫλ ).

Since − 12 < ǫλ < 0, it follows from (97) that the inequalities (95), and therefore the equation
(96), still hold. Hence (96) holds whenever
(p )
(log p1 log log p1 )
ǫλ 6= O . (98)
log λ

In particular it holds whenever

ǫλ 6= 0(1), (99)

Now let us consider the case in which


(p )
(log p1 log log p1 )
ǫλ = 0 , (100)
log λ

so that ǫλ = o(1), in virtue of (92). It follows from this and (89) and (90) that, if ǫλ > 0,
then
)
aλ ≤ Λ + 1 + o(1),
(101)
aλ ≥ Λ + o(1).

Hence
aλ ≤ Λ + 1, aλ ≥ Λ;
and so

aλ = Λ or Λ + 1. (102)

Similarly from (89), (91), and (100), we see that, if ǫλ < 0, then
)
aλ ≤ Λ + o(1),
(103)
aλ ≥ Λ − 1 + o(1).
128 Paper 15

Hence
aλ ≤ Λ, aλ ≥ Λ − 1;
and so

aλ = Λ or Λ − 1. (104)
1
For example, let us suppose that it is required to find aλ when λ ∼ p18 . We have

(2log λ/ log p1 − 1)−1 = (21/8 − 1)−1 + o(1) = 11.048 · · · + o(1).

It is evident that Λ = 11 and ǫλ 6= o(1). Hence aλ = 11.

19. The results in the previous section may be rewritten with slight modifications, in order
that the transition of aλ from one value to another may be more clearly expressed. Let
log(1+1/x)
log 2
λ = p1 , (105)

and let x + ǫx , where − 21 < ǫx < 12 , be an integer. Then the range of x which we are now
considering is
s 
log p1
x=o , (106)
log log p1

and the results of the previous section may be stated as follows. If


( s )
log log p1
ǫx 6= O x , (107)
log p1

then

aλ = [x]. (108)

As a particular case of this we have


aλ = [x],
when ǫx 6= o(1). But if
( s )
log log p1
ǫx = O x , (109)
log p1

then when ǫx > 0

aλ = [x] or [x + 1]; (110)


Highly composite numbers 129

and when ǫx < 0


aλ = [x] or [x − 1]. (110′ )

20. Range II:


p 
log λ = O (log p1 log log p1 ), 

s 

log p1
log λ 6= o . 

log log p1 

From (89) it follows that


(p )
(log p1 log log p1 )
aλ = (2log λ/ log p1 − 1)−1 + O . (111)
log λ

But
log p1
(2log λ/ log p1 − 1)−1 = + O(1).
log 2 log λ
Hence
log p1 p
aλ log λ = + O (log p1 log log p1 ). (112)
log 2

As an example we may suppose that



λ ∼ e (log p1 ) .

Then from (112) it follows that


p
(log p1 ) p
aλ = + O (log log p1 ).
log 2

21. Range III:


s  
log p1 
log λ = O ,


log log p1

1 
log λ 6= o(log p1 ) 3 .

Let us suppose that µ = O(1) in (88). Then we see that


( p )
log p1 log µ µ log p1 (log p1 log log p1 )
aλ = + O(1) + O + + , (113)
log 2 log λ µ (log λ)2 log λ
130 Paper 15

or
 
log p1 log µ log λ µ log p1 p
aλ log λ = +O + + (log p1 log log p1 ) . (114)
log 2 µ log λ

Now  
log µ log λ log p1
= O(log λ) = o ,
µ log λ
 
µ log p1 log p1
=O ,
log λ log λ
 
p log p1
(log p1 log log p1 ) = O .
log λ
Hence
 
log p1 log p1
aλ log λ = +O . (115)
log 2 log λ

For example, when


3
λ ∼ e(log p1 ) 8 ,
we have 5
(log p1 ) 8 1
aλ = + O(log p1 ) 4 .
log 2

22. Range IV:


1 )
log λ = O(log p1 ) 3 ,
log λ 6= o(log log p1 ).

In this case it follows from (85) that

log p1 p
aλ log λ = + O (log p1 log λ). (116)
log 2

As an example in this range, when we suppose that


1
λ ∼ e(log p1 ) 4 ,

we obtain from (116)


3
(log p1 ) 4 3
aλ = + O(log p1 ) 8 .
log 2

23. Range V: log λ = O(log log p1 ).


Highly composite numbers 131

From (85) it follows that


log p1 p
aλ log λ = + O (log p1 log log p1 ). (117)
log 2
For example, we may suppose that

λ ∼ e (log log p1 ) .

Then
log p1 p
aλ = p + O (log p1 ).
log 2 (log log p1 )

24. Let λ′ be the prime next below λ, so that λ′ ≤ λ − 1. Then it follows from (63) that
p p p
{(1 + aλ′ ) log λ′ } − {(1 + aλ ) log λ} > − log(λλ′ ). (118)

Hence
p p p
{(1 + aλ′ log(λ − 1)} − {(1 + aλ ) log λ} > − {2 log λ}. (119)

But  2
1 1
log(λ − 1) < log λ − < log λ 1 − ;
λ 2λ log λ
and so (119) may be replaced by

(1 + aλ′ ) √
p
p p
(1 + aλ′ ) − (1 + aλ ) > − 2. (120)
2λ log λ
But from (54) we know that
 
log p1 log p1 log p1
1 + aλ′ ≥1+ ′
> ′
> .
log λ log λ log λ

From this and (120) it follows that


p
p p (log p1 )
(1 + aλ′ ) − (1 + aλ ) > 3 − 2. (121)
2λ(log λ) 2

Now let us suppose that λ2 (log λ)3 < 18 log p1 . Then, from (121), we have
p p
(1 + aλ′ ) − (1 + aλ ) > 0,

or

aλ′ > aλ . (122)


132 Paper 15

1
From (122) it follows that, if λ2 (log λ)3 < 8 log p1 , then

a2 > a3 > a5 > a7 > · · · > aλ . (123)

In other words, in a large highly composite number

2a2 · 3a3 · 5a5 · 7a7 · · · p1 ,

the indices comparatively near the beginning form a decreasing sequence in the strict sense
which forbids equality. Later on groups of equal indices will in general occur.
To sum up, we have obtained fairly accurate information about aλ for all possible values of
λ. The range I is by far the most extensive, and throughout this range aλ is known with
an error never exceeding 1. The formulæ (86) hold throughout a range which includes all
the remaining ranges II - V, and a considerable part of I as well, while we have obtained
more precise formulæ for each individual range II-V.

25. Now let us consider the nature of pr . It is evident that r cannot exceed a2 ; i.e., r
cannot exceed
log p1 p
+ O (log p1 log log p1 ). (124)
(log 2)2

From (55) it evidently follows that


)
apr log pr = O(log p1 ),
(125)
apr log pr 6= o(log p1 );

)
(1 + aPr ) log pr = O(log p1 ),
(126)
(1 + aPr ) log pr 6= o(log p1 ).

But from (46) we know that


)
apr log pr ≥ r log pr ,
(127)
(1 + aPr ) log pr ≤ r log pr .

From (125) - (127) it follows that


)
r log pr = O(log p1 ),
(128)
r log pr 6= o(log p1 );

and
)
apr = O(r),
(129)
apr 6= o(r).
Highly composite numbers 133

26. Supposing that λ = pr in (81) and λ = Pr in (82), and remembering (128), we see
that, if rµ = o(log p1 ), then
    
1 log pr log µ rµ
log 1 + ≥ 1+O + , (130)
ap r a2 log 2 rµ log p1
and
    
1 log Pr log µ rµ
log 1 + ≤ 1+O + . (131)
a + aP r a2 log 2 rµ log p1
But, from (47), we have
   
1 1
log 1 + ≤ log 1 + .
ap r 1 + aP r
Also we know that
     
1 r
log Pr = log pr + O(1) = log pr 1 + O = log pr 1 + O .
log pr log p1
Hence (131) may be replaced by
    
1 log pr log µ rµ
log 1 + ≤ 1+O + . (132)
ap r a2 log 2 rµ log p1

From (130) and (132) it is evident that


    
1 log pr log µ rµ
log 1 + = 1+O + . (133)
ap r a1 log 2 rµ log p1
In a similar manner
    
1 log pr log µ rµ
log 1 + = 1+O + . (134)
1 + aP r a2 log 2 rµ log p1
Now supposing that

rµ = o(log p1 ),
(135)
rµ 6= O(log µ),

and dividing (134) by (133), we have


 
log 1 + 1+a1 P 
log µ rµ

r
  =1+O + ,
log 1 + a1p rµ log p1
r

or   
1 1 log µ rµ
1+ =1+ +O + /apr ,
1 + aP r apr rµ log p1
134 Paper 15

that is   
1 1 log µ rµ
= 1+O + .
1 + aP r apr rµ log p1
Hence
r2µ
 
log µ
apr = aPr + 1 + O + , (136)
µ log p1

in virtue of (129). But aPr ≤ r − 1, and so

r2µ
 
log µ
ap r ≤ r + O + . (137)
µ log p1

But we know that apr ≥ r. Hence it is clear that

r2µ
 
log µ
ap r = r + O + . (138)
µ log p1

From this and (136) it follows that

r2µ
 
log µ
aP r = r − 1 + O + , (139)
µ log p1

provided that the conditions (135)


p are satisfied.
Now let us suppose that r = o (log p1 ). Then we can choose µ such that r 2 µ = o(log p1 )
and µ 6= O(1). Consequently we have

log µ r2µ
= o(1), = o(1);
µ log p1

and so it follows from (138) and (139) that

apr = 1 + aPr = r, (140)


p p
provided that r = o (log p1 ). From this it is clear that, if r = o (log p1 ), then

p1 > p2 > p3 > p4 > · · · > pr . (141)

In other words, in a large highly composite number

2a2 · 3a3 · 5a5 · · · p1 ,

the indices comparatively near the end form a sequence of the type

. . . 5 . . . 4 . . . 3 . . . 2 . . . 1.

Near the beginning gaps in the indices will in general occur.


Highly composite numbers 135

p
Again, let us suppose that r = o(log p1 ), r 6= o (log p1 ), and µ = O(1) in (138) and (139).
Then we see that
 2  
r
ap r = r + O , 

log p1


 2  (142)
r 
aP r = r + O ; 


log p1
p
provided that r = o(log p1 ) and r 6= o (log p1 ). But when r 6= o(log p1 ), we shall use the
general result, viz.,

apr = O(r), apr 6= o(r),
(143)
aPr = O(r), aPr 6= o(r),

which is true for all values of r except 1.

27. It follows from (87) and (128) that


 
log pr log p1 log p1 log µ p 
 ≤ +O + rµ + (log p1 log log p1 ) , 

log 1 + a1p log 2 rµ




r
  (144)
log Pr log p1 log p1 log µ p 
 ≥ +O + rµ + (log p1 log log p1 ) ,


log 2 rµ

1
log 1 +


1+aPr

with the condition that rµ = o(log p1 ). From this it can easily be shewn, by arguments
similar to those used in the beginning of the previous section, that
 
log pr log p1 log p1 log µ p
= +O + rµ + (log p1 log log p1 ) , (145)
log(1 + 1/r) log 2 rµ

provided that rµ = o(log p1 ).


Now let us suppose that r = o(log p1 ); then we can choose µ such that
 
log p1
µ=o , µ 6= O(1).
r

Consequently rµ = o(log p1 ) and log µ = o(µ), and so

log p1 log µ
= o(log p1 ).

From these relations and (145) it follows that, if r = o(log p1 ), then

log pr log p1
∼ ; (146)
log(1 + 1/r) log 2
136 Paper 15

that is to say that, if r = o(log p1 ), then

log p1 log p2 log p3 log pr


∼ 1 ∼ 1 ∼ ··· ∼ . (147)
log 2 log(1 + 2 ) log(1 + 3 ) log(1 + 1/r)
p
Again let us suppose that r = O (log p1 log log p1 ) in (145). Then it is possible to choose
µ such that
p 
rµ = Op (log p1 log log p1 ),
(148)
rµ 6= o (log p1 log log p1 ).

It is evident that log µ = O(log log p1 ), and so


 
log p1 log µ log p1 log log p1 p
=O = O (log p1 log log p1 ),
rµ rµ

in virtue of (148). Hence

log pr log p1 p
= + O (log p1 log log p1 ), (149)
log(1 + 1/r) log 2

provided that p
r=O (log p1 log log p1 ).
p
Now let us suppose that r = o(log p1 ), r 6= o (log p1 log log p1 ) and µ = O(1), in (145).
Then it is evident that

log p1 = O(r 2 ),
p
(log p1 log log p1 = O(r),

and  
log p1 log µ log p1 )
=O = O(r).
rµ r
Hence we see that
log pr log p1
= + O(r), (150)
log(1 + 1/r) log 2

if p
r = o(log p1 ), r 6= o (log p1 log log p1 ).
But, if r 6= o(log p1 ), we see from (128) that

log pr

= O(log p1 ), 
log(1 + 1/r)


(151)
log pr 
6= o(log p1 ). 

log(1 + 1/r)
Highly composite numbers 137

p
From (150) and (151) it follows that, if r 6= o (log p1 log log p1 ), then
log pr log p1
= + O(r); (152)
log(1 + 1/r) log 2
and from (149) and (152) that, if r = o(log p1 ), then
log pr log p1
∼ ,
log(1 + 1/r) log 2
in agreement with (147). This result will, in general, fail for the largest possible values of
r, which are of order log p1 .
It must be remembered that all the results involving p1 may be written in terms of N , since
p1 = O(log N ) and p1 6= o(log N ), and consequently

log p1 = log log N + O(1). (153)

28. We shall now prove that successive highly composite numbers are asymptotically equiv-
alent. Let m and n be any two positive integers which are prime to each other, such that

log mn = o(log p1 ) = o(log log N ); (154)

and let
m
= 2δ2 · 3δ3 · 5δ5 · · · ℘δ℘ . (155)
n
Then it is evident that

mn = 2|δ2 | · 3|δ3 | · 5|δ5 | · · · ℘|δ℘ | . (156)

Hence

δλ log λ = O(log mn) = o(log p1 ) = o(aλ log λ); (157)

so that δλ = o(aλ ).
Now
m      
δ2 δ3 δ℘
d N = d(N ) 1 + 1+ ··· 1 + . (158)
n 1 + a2 1 + a3 1 + a℘
But, from (60), we know that
p
aλ log λ = a2 log 2 + O (log p1 log λ).

Hence
(  3 )
δλ δλ log λ log λ 2
1+ = 1+ +O |δλ |
1 + aλ a2 log 2 log p1
138 Paper 15

( s )
δλ log λ log λ log ℘
= 1+ + O |δλ |
a2 log 2 log p1 log p1
( s  )
δλ log λ 2
 
δλ log λ |δλ | log λ log ℘
= exp +O +O
a2 log 2 log p1 log p1 log p1
( s  )
δλ log λ |δλ | log λ log mn
= exp +O . (159)
a2 log 2 log p1 log p1

It follows from (155), (156), (158) and (159) that



m  δ2 log 2 + δ3 log 3 + · · · + δ℘ log ℘
d N = d(N ) exp
n a2 log 2
s )
|δ2 | log 2 + |δ3 | log 3 + · · · + |δ℘ | log ℘ log mn
+O
log p1 log p1
 3
log(m/n) log mn 2
+O
= d(N )e a2 log 2 log p1

 r 
1 log mn
a2 log 2
logm
n +O log mn log p1
= d(N )e . (160)

Putting m = n + 1, we see that, if

log n = o(log p1 ) = o(log log N ),

then
   n  o
1
q
1 1 log n
log(1+ n )+O log n log
d N 1+ = d(N )e a2 log 2 p 1
n
 q 
1+O n log n ( logloglognN )
 
1 a2 log 2
= d(N ) 1 + (161)
n
Now it is possible to choose n such that
3 p
n(log n) 2 6= o (log log N ),

and ( s )
log n
1 + O n log n > 0;
log log N
that is to say
  
1
d N 1+ > d(N ). (162)
n
Highly composite numbers 139

From this and (29) it follows that, if N is a highly composite number, then the next highly
composite number is of the form
3
( )
N (log log log N ) 2
N +O p (163)
(log log N )

Hence the ratio of two consecutive highly composite numbers tends to unity.
It follows from (163) that the number of highly composite numbers not exceeding x is not
of the form ( p )
log x (log log x)
o 3 .
(log log log x) 2

29. Now let us consider the nature of d(N ) for highly composite values of N . From (44)
we see that

d(N ) = 2π(p1 )−π(p2 ) · 3π(p2 )−π(p3 ) · 4π(p3 )−π(p4 ) · · · (1 + a2 ). (164)

From this it follows that

d(N ) = 2a2 · 3a3 · 5a5 · · · ̟ a̟ , (165)

where ̟ is the largest prime not exceeding 1 + a2 ; and

αλ = π(pλ−1 ) + O(pλ ). (166)

It also follows that, if ℘1 , ℘2 , ℘3 , . . . , ℘λ are a given set of primes, then a number µ̄ can be
found such that the equation
β
d(N ) = ℘β1 1 · ℘β2 2 · ℘β3 3 · · · ℘µµ · · · ℘βλλ

is impossible if N is a highly composite number and βµ > µ̄. We may state this roughly
by saying that as N (a highly composite number) tends to infinity, then, not merely in N
itself, but also in d(N ), the number of prime factors, as well as the indices, must tend to
infinity. In particular such an equation as

d(N ) = k · 2m , (167)

where k is fixed, becomes impossible when m exceeds a certain limit depending on k.


It is easily seen from (153), (164), and (165) that

̟ = O(a2 ) = O(log p1 ) = O(log log N ) = O{log log d(N )}, ∗ (168)
̟ 6= o(a2 ) = o(log p1 ) = o(log log N ) = o{log log d(N )}.
140 Paper 15

It follows from (147) that if λ = o(log p1 ) then


log α2 log α3 log α5 log αλ
∼ ∼ ∼ ··· . (169)
log(1 − 12 ) log(1 − 31 ) log(1 − 51 ) log(1 − λ1 )
p
Similarly, from (149), it follows that if λ = O (log p1 log log p1 ) then

log(1 + αλ ) log p1 p
=− + O (log p1 log log p1 ). (170)
log(1 − 1/λ) log 2
p
Again, from (152), we see that if λ 6= o (log p1 log log p1 ) then

log(1 + αλ ) log p1
=− + O(λ). (171)
log(1 − 1/λ) log 2
In the left-hand side we cannot write αλ instead of 1 + αλ , as αλ may be zero for a few
values of λ.
From (165) and (170) we can shew that

log d(N ) = α2 log 2 + O(α3 ), log d(N ) 6= α2 log 2 + o(α3 );

and so
log 3
2

log p1 +O (log p1 log log p1 )
log d(N ) = α2 log 2 + e log 2 (172)

But from (163) we see that

log log d(N ) = log p1 + O(log log p1 ).

From this and (172) it follows that


log 3
rn o
2 log log log d(N)
log 2
+O log log d(N)
a2 log 2 = log d(N ) − {log d(N )} . (173)

30. Now we shall consider the order of dd(N ) for highly composite values of N . It follows
from (165) that

log dd(N ) = log(1 + α2 ) + log(1 + α3 ) + · · · + log(1 + α̟ ). (174)

Now let λ, λ′ , λ′′ , . . . be consecutive primes in ascending order, and let


p
λ = O (log p1 log log p1 ),

More precisely ̟ ∼ a2 . But this involves the assumption that two consecutive primes are asymptotically
equivalent. This follows at once from the prime number theorem. It appears probable that such a result
cannot really be as deep as the prime number theorem, but nobody has succeeded up to now in proving it
by elementary reasoning.
Highly composite numbers 141

p
λ 6= o (log p1 log log p1 ).

Then, from (174), we have

log dd(N ) = log(1 + α2 ) + log(1 + α3 ) + · · · + log(1 + αλ )


+ log(1 + αλ′ ) + log(1 + aλ′′ ) + · · · + log(1 + α̟ ). (175)

But, from (170), we have

log(1 + α2 ) + log(1 + α3 ) + · · · + log(1 + αλ )


  
log p1 1 1 1 1
= − log (1 − 2 )(1 − 3 )(1 − 5 ) · · · 1 −
log 2 λ
  
p 1
+ O (log p1 log log p1 ) log (1 − 21 )(1 − 13 ) · · · 1 − . (176)
λ

It can be shewn, without assuming the prime number theorem∗ , that


    
1 1
− log (1 − 12 )(1 − 13 )(1 − 15 ) · · · 1 − = log log p + γ + O , (177)
p log p

where γ is the Eulerian constant. Hence


  
1
log (1 − 21 (1 − 13 )(1 − 51 ) · · · 1 − = O(log log p).
p

From this and (176) it follows that

log(1 + α2 ) + log(1 + α3 ) + · · · + log(1 + αλ )


  
log p1 1 1 1
= − log (1 − 2 )(1 − 3 ) · · · 1 −
log 2 λ
p
+ O{ (log p1 log log p1 ) log log λ}
  
log p1 1 1 1
= − log (1 − 2 )(1 − 3 ) · · · 1 −
log 2 λ
p
+ O{ (log p1 log log p1 ) log log log p1 }. (178)

Again, from (152), we see that

log(1 + αλ′ ) + log(1 + αλ′′ ) + · · · + log(1 + α̟ )


    
log p1 1 1 1
=− log 1− ′ 1 − ′′ · · · 1 −
log 2 λ λ ̟

See Landau, Handbuch, p.139.
142 Paper 15

      
1 ′′ 1 1
+ O λ′ log 1 − ′ + λ log 1 − ′′ + · · · + ̟ log 1 −
λ λ ̟
    
log p1 1 1 1
=− log 1− ′ 1 − ′ ··· 1 − + O{π(̟) − π(λ)}
log 2 λ λ ̟
      
log p1 1 1 1 log p1
=− 1− ′ 1 − ′′ · · · 1 − +O . (179)
log 2 λ λ ̟ log log p1

From (175), (178) and (179) it follows that


  
log p1 1 1 1
log dd(N ) = − log (1 − 2 )(1 − 3 ) · · · 1 −
log 2 λ
p
+ O{ (log p1 log log p1 ) log log log p1 }
      
log p1 1 1 1 log p1
− log 1− ′ 1 − ′′ · · · 1 − +O
log 2 λ λ ̟ log log p1
    
log p1 1 1 1 log p1
= − log (1 − 2 )(1 − 3 ) · · · 1 − +O
log 2 ̟ log log p1
    
log p1 1 log p1
= log log ̟ + γ + O +O
log 2 log ̟ log log p1
    
log p1 1 log p1
= log log log p1 + γ + O +O
log 2 log log p1 log log p1
  
log log N 1
= log log log log N + γ + O , (180)
log 2 log log log N

in virtue of (177), (168), and (163). Hence, if N is a highly composite number, then
n  o
1
log log log log N +γ+O log log1 log N
dd(N ) = (log N ) log 2
. (181)

31. It may be interesting to note that, as far as the table is constructed,

2, 22 , 23 , . . . , 213 , 3, 3 · 2, 3 · 22 , . . . , 3 · 211 , 5 · 2, 5 · 22 , . . . , 5 · 28 ,
7 · 25 , 7 · 26 , . . . , 7 · 210 , 9, 9 · 2, 9 · 22 , . . . , 9 · 210 ,

and so on, occur as values of d(N ). But we know from § 29 that k · 2m cannot be the value
of d(N ) for sufficiently large values of m; and so numbers of the form k · 2m which occur as
the value of d(N ) in the table must disappear sooner or later when the table is extended.
Thus numbers of the form 5 · 2m have begun to disappear in the table itself. The powers
of 2 disappear at any rate from 218 onwards. The least number having 218 divisors is

27 · 33 · 53 · 7 · 11 · 13 · · · 41 · 43,
Highly composite numbers 143

while the smaller number, viz.,

28 · 34 · 53 · 72 · 11 · 13 · · · 41

has a larger number of divisors. viz. 135 · 211 . The numbers of the form 7 · 2m disappear
at least from 7 · 213 onwards. The least number having 7 · 213 divisors is

26 · 33 · 53 · 7 · 11 · 13 · · · 31 · 37,

while the smaller number, viz.

29 · 34 · 52 · 7 · 11 · 13 · · · 31

has a larger number of divisors, viz. 225 · 28 .

IV
Superior Highly Composite Numbers

32. A number N may be said to be a superior highly composite number if there is a positive
number ǫ such that
d(N ) d(N ′ )
≥ , (182)
Nǫ (N ′ )ǫ

for all values of N ′ less that N , and

d(N ) d(N ′ )
> (183)
Nǫ (N ′ )ǫ

for all values of N ′ greater that N .


All superior highly composite numbers are also highly composite. For, if N ′ < N , it follows
from (182) that  ǫ
′ N
d(N ) ≥ d(N ) > d(N ′ );
N′
and so N is highly composite.

33. Now let us consider what must be the nature of N in order that it should be a superior
highly composite number. In the first place it must be of the form

2a2 · 3a3 · 5a5 · · · pa1 , (184)

or of the form

2 · 3 · 5 · 7 · · · p1
× 2 · 3 · 5 · 7 · · · p2
144 Paper 15

× 2 · 3 · 5 · · · p3
× ...... :

i.e. it must satisfy the conditions for a highly composite number. Now let

N ′ = N/λ,

where λ ≤ p1 . Then from (182) it follows that

1 + aλ aλ
≥ ǫ(a −1) ,
λǫaλ λ λ
or
 
ǫ 1
λ ≤ 1+ . (185)

Again let
N ′ − Nλ .
Then, from (183), we see that
1 + aλ 2 + aλ
ǫa
> ǫ(a +1) ,
λ λ λ λ
or
 
ǫ 1
λ > 1+ . (186)
1 + aλ

Now supposing that λ = p1 in (185) and λ = P1 in (186), we obtain

log 2 log 2
<ǫ≤ . (187)
log P1 log p1

Now let us suppose that ǫ = 1/x. Then, from (187), we have

p1 ≤ 2x < P1 . (188)

That is, p1 is the largest prime not exceeding 2x . It follows from (185) that

aλ ≤ (λ1/x − 1)−1 . (189)

Similarly, from (186),

aλ > (λ1/x − 1)−1 − 1. (190)

From (189) and (190) it is clear that

aλ = [λ1/x − 1)−1 ]. (191)


Highly composite numbers 145

Hence N is of the form


1/x −1)−1 ] 1/x −1)−1 ] 1/x −1)−1 ]
2[(2 · 3[(3 · 5[(5 . . . p1 , (192)

where p1 is the largest prime not exceeding 2x .

34. Now let us suppose that λ = pr in (189). Then


apr ≤ (p1/x
r − 1)−1 .

But we know that r ≤ apr . Hence

r ≤ (p1/x
r − 1)−1 ,

or
 x
1
pr ≤ 1+ . (193)
r

Similarly by supposing that λ = Pr in (190), we see that

aPr > (Pr1/x − 1)−1 − 1.

But we know that r − 1 ≥ apr . Hence

r > (Pr1/x − 1)−1

or
 x
1
Pr > 1+ . (194)
r

From (193) and (194) it is clear that pr is the largest prime not exceeding (1 + 1/r)x . Hence
N is of the form

2 · 3 · 5 · 7 · · · p1
× 2 · 3 · 5 · 7 · · · p2
× 2 · 3 · 5 · · · p3
× ........., (195)

where x
 p1 is the largest prime not greater than 2 , p2 is the largest prime not greater than
3 x
2 , and so on. In other words N is of the form
x )+ϑ( 3 )x +ϑ( 4 )x +···.
eϑ(2 2 3 , (196)

and d(N ) is of the form


x) 3 x 4 x
2π(2 · ( 23 )π( 2 ) · ( 43 )π( 3 ) · · · . (197)
146 Paper 15

Thus to every value of x not less than 1 correspondence one, and only one, value of N .

35. Since
d(N ) d(N ′ )
≥ ,
N 1/x (N ′ )1/x
for all values of N ′ , it follows from (196) and (197) that
x) 3 x 4 x
1/x 2π(2 ( 32 )π( 2 ) ( 43 )π( 3 )
d(N ) ≤ N ···, (198)
e(1/x)ϑ(2x ) e(1/x)ϑ( 32 )x e(1/x)ϑ( 43 )x

for all values of N and x; and d(N ) is equal to the right-hand side when
x )+ϑ( 3 )x +ϑ( 4 )x +···
N = eϑ(2 2 3 . (199)

Thus, for example, putting x = 2, 3, 4 in (198), we obtain


p 
d(N ) ≤ (3N ), 

1
d(N ) ≤ 8(3N/35) , 3 (200)
1 
d(N ) ≤ 96(3N/50050) 4 , 
1
for all values of N ; and d(N ) = (3N ) when N = 22 · 3; d(N ) = 8(3N/35) 3 when N =
p
1
23 · 32 · 5 · 7; d(N ) = 96(3N/50050) 4 when

N = 25 · 33 · 52 · 7 · 11 · 13.

36. M and N are consecutive superior highly composite numbers if there are no superior
highly composite numbers between M and N .
From (195) and (196) it is easily seen that, if M and N are any two superior highly
composite numbers, and if M > N , then M is a multiple of N ; and also that, if M and
N are two consecutive superior highly composite numbers, and if M > N , then M/N is
a prime number. From this it follows that consecutive superior highly composite numbers
are of the form

π1 , π1 π2 , π1 π2 π3 , π1 π2 π3 π4 , . . . , (201)

where π1 , π2 , π3 . . . are primes. In order to determine π1 , π2 , . . . we proceed as follows. Let


x′1 be the smallest value of x such that [2x ] is prime x′2 the smallest value of x such that
x
[ 32 ] is prime, and so on; and let x1 , x2 , . . . be the numbers x′1 , x′2 . . . arranged in order of
magnitude. Then πn is the prime corresponding to xn , and

N = π1 π2 π3 · · · πn , (202)

if xn ≤ x < xn+1 .
Highly composite numbers 147

37. From the preceding results we see that the number of superior highly composite num-
bers not exceeding

x )+ϑ( 3 )x +ϑ( 4 )x +···


eϑ(2 2 3 (203)

is

π(2x ) + π( 23 )x + π( 34 )x + · · · .

In other words if xn ≤ x < xn+1 then

n = π(2x ) + π( 32 )x + π( 34 )x + · · · . (204)

It follows from (192) and (202) that, of the primes π1 , π2 , π3 , . . . , πn , the number of primes
which are equal to a given prime ̟ is equal to

[(̟ 1/x − 1)−1 ]. (205)

Further, the greatest of the primes π1 , π2 , π3 . . . , πn is the largest prime not greater than
2x , and is asymptotically equivalent to the natural nth prime, in virtue of (204).
The following table gives the values of πn and xn for the first 50 values of n, that is till xn
reaches very nearly 7.

log 2
π1 = 2 x1 = log 2 =1
log 3
π2 = 3 x2 = log 2 = 1.5849 · · ·
log 2
π3 = 2 x3 = log( 32 )
= 1.7095 · · ·
log 5
π4 = 5 x4 = log 2 = 2.3219 . . .
log 2
π5 = 2 x5 = log( 43 )
= 2.4094 . . .
148 Paper 15

log 3
π6 = 3 x6 = log( 23 )
= 2.7095 . . .
log 7
π7 = 7 x7 = log 2 = 2.8073 . . .
log 2
π8 = 2 x8 = log( 45 )
= 3.1062 . . .
log 11
π9 = 11 x9 = log 2 = 3.4594 . . .
log 13
π10 = 13 x10 = log 2 = 3.7004 . . .
log 2
π11 = 2 x11 = log( 56 )
= 3.8017 . . .
log 3
π12 = 3 x12 = log( 34 )
= 3.8188 . . .
log 5
π13 = 5 x13 = log( 23 )
= 3.9693 · · ·
log 17
π14 = 17 x14 = log 2 = 4.0874 . . .
log 19
π15 = 19 x15 = log 2 = 4.2479 . . .
log 2
π16 = 2 x16 = log( 67 )
= 4.4965 . . .
log 23
π17 = 23 x17 = log 2 = 4.5235 . . .
log 7
π18 = 7 x18 = log( 23 )
= 4.7992 . . .
log 29
π19 = 29 x19 = log 2 = 4.8579 . . .
log 3
π20 = 3 x20 = log( 45 )
= 4.9233 . . .
log 31
π21 = 31 x21 = log 2 = 4.9541 . . .
log 2
π22 = 2 x22 = log( 78 )
= 5.1908 . . .
log 37
π23 = 37 x23 = log 2 = 5.2094 . . .
log 41
π24 = 41 x24 = log 2 = 5.3575 . . .
log 43
π25 = 43 x25 = log 2 = 5.4262 . . .
log 47
π26 = 47 x26 = log 2 = 5.5545 . . .
Highly composite numbers 149

log 5
π27 = 5 x27 = log( 43 )
= 5.5945 . . .
log 53
π28 = 53 x28 = log 2 = 5.7279 . . .
log 59
π29 = 59 x29 = log 2 = 5.8826 . . .
log 2
π30 = 2 x30 = log( 98 )
= 5.8849 . . .
log 11
π31 = 11 x31 = log( 32 )
= 5.9139 . . .
log 61
π32 = 61 x32 = log 2 = 5.9307 . . .
log 3
π33 = 3 x33 = log( 65 )
= 6.0256 . . .
log 67
π34 = 67 x34 = log 2 = 6.0660 . . .
log 71
π35 = 71 x35 = log 2 = 6.1497 . . .
log 73
π36 = 73 x36 = log 2 = 6.1898 . . .
log 79
π37 = 79 x37 = log 2 = 6.3037 . . .
log 13
π38 = 13 x38 = log( 32 )
= 6.3259 . . .
log 83
π39 = 83 x39 = log 2 = 6.3750 . . .
log 89
π40 = 89 x40 = log 2 = 6.4757 . . .
log 2
π41 = 2 x41 = log( 10 )
= 6.5790 . . .
9
log 97
π42 = 97 x42 = log 2 = 6.5999 . . .
log 101
π43 = 101 x43 = log 2 = 6.6582 . . .
log 103
π44 = 103 x44 = log 2 = 6.6724 . . .
log 107
π45 = 107 x45 = log 2 = 6.7414 . . .
log 7
π46 = 7 x46 = log( 43 )
= 6.7641 . . .
log 109
π47 = 109 x47 = log 2 = 6.7681 . . .
log 113
π48 = 113 x48 = log 2 = 6.8201 . . .
log 17
π49 = 17 x49 = log( 32 )
= 6.9875 . . .
log 127
π50 = 127 x50 = log 2 = 6.9886 . . .
150 Paper 15

38. It follows from (17) and (198) that log d(N ) ≤ F (x), where
( 32 )x ( 43 )x
(Z )
2x
1 1 π(t) π(t) π(t)
Z Z
F (x) = log N + dt + dt + dt + · · · , (206)
x x 2 t 2 t 2 t

for all values of N and x. In order to obtain the best possible upper limit for log d(N ), we
must choose x so as to make the right-hand side a minimum.
The function F (x) is obviously continuous unless (1 + 1r)x = p, where r is a positive
integer and p a prime. It is easily seen to be continuous even then, and so continuous
without exception. Also
(Z x
2 Z ( 3 )x )
1 1 π(t) 2 π(t)
F ′ (x) = − 2 log N − 2 dt + dt + · · ·
x x 2 t 2 t
1
+ {π(2x ) log 2 + π( 32 )x log 32 + · · ·}
x
1
= {ϑ(2x ) + ϑ( 32 )x + ϑ( 43 )x + · · · − log N }, (207)
x2
unless (1 + 1/r)x = p, in virtue of (17).
Thus we see that F (x) is continuous, and F ′ (x) exists and is continuous except at certain
isolated points. The sign of F ′ (x), where it exists, is that of

ϑ(2x ) + ϑ( 32 )x + ϑ( 34 )x + · · · − log N,

and ϑ(2x ) + ϑ( 32 )x + ϑ( 43 )x + · · · ,
is a monotonic function. Thus F ′ (x) is first negative and then positive, changing sign once
only, and so F (x) has a unique minimum. Thus F (x) is a minimum when x is a function
of N defined by the inequalities
 
v 3 v 4 v < log N (y < x)
ϑ(2 ) + ϑ( 2 ) + ϑ( 3 ) + · · · . (208)
> log N (y > x)

Now let D(N ) be a function of N such that


x 3 x 4 x
D(N ) = 2π(2 ) ( 32 )π( 2 ) ( 43 )π( 3 ) · · · , (209)

where x is the function of N defined by the inequalities (208). Then, from (198), we see
that

d(N ) ≤ D(N ), (210)

for all values of N ; and d(N ) = D(N ) for all superior highly composite values of N . Hence
D(N ) is the maximum order of d(N ). In other words, d(N ) will attain its maximum order
when N is a superior highly composite number.
Highly composite numbers 151

V.
Application to the Order of d(N ).

39. The most precise result known concerning the distribution of the prime numbers is
that
√ )
π(x) = Li(x) + O(xe−a log x ),
√ (211)
ϑ(x) = x + O(xe−a log x ),

where x
dt
Z
Li(x) =
log t
and a is a positive constant.
In order to find the maximum order of d(N ) we have merely to determine the order of
D(N ) from the equations (208) and (209). Now, from (208), we have

log N = ϑ(2x ) + O( 23 )x = ϑ(2x ) + o(22x/3 );

and so
2
ϑ(2x ) = log N + o(log N ) 3 ; (212)

and similarly from (209) we have

log D(N ) 2
π(2x ) = + o(log N ) 3 . (213)
log 2
It follows from (211) - (213) that the maximum order of d(N ) is

(log log N)
Li(log N )+O[log N e−a ]
2 . (214)

It does not seen to be possible to obtain an upper limit for d(N ) notably more precise
than (214) without assuming results concerning the distribution of primes which depend
on hitherto unproved properties of the Riemann ζ-function.

40. We shall now assume that the “Riemann hypothesis” concerning the ζ-function is true,
1
i.e., that all the complex roots of ζ(s) have their real part equal to 2. Then it is known
that
√ X xρ 1
ϑ(x) = x − x − + O(x 3 ), (215)
ρ

where ρ is a complex root of ζ(s), and that


√ X 1
π(x) = Li(x) − 12 Li( x) − Li(xρ ) + O(x 3 )
152 Paper 15

√ √  √
1 X xρ X xρ 
x 2 x 1 x
= Li(x) − − − − +O , (216)
log x (log x)2 log x ρ (log x)2 ρ2 (log x)3
P xρ
since ρk
is absolutely convergent when k > 1. Also it is known that
X xρ p
= O{ x(log x)2 }; (217)
ρ
and so
p
ϑ(x) − x = O{ x(log x)2 }. (218)

From (215) and (216) it is clear that


 √ 
ϑ(x) − x x
π(x) = Li(x) + − R(x) + O , (219)
log x log x)3
where
√ P xρ
2 x+ ρ2
R(x) = (220)
(log x)2
But it follows from Taylor’s theorem and (218) that
ϑ(x) − x
Liϑ(x) − Li(x) = + O(log x)2 , (221)
log x
and from (219) and (221) it follows that
 √ 
x
π(x) = Liϑ(x) − R(x) + O . (222)
(log x)3

41. It follows from the functional equation satisfied by ζ(s), viz.,


(2π)−s Γ(s)ζ(s) cos 12 πs = 21 ζ(1 − s), (223)

that √   √ 


− 14 s 1+ s 1+ s
(1 − s)π Γ ζ
4 2
is an integral function of s whose apparent order is less than 1, and hence is equal to
Yn o
Γ( 14 )ζ( 12 ) s
1 − (2ρ−1)2 ,

[where ρ runs through the complex roots of ζ(s) whose imaginary parts are positive]. From
this we can easily deduce that
  Y 
− 1+s 1+s s
s(1 + s)π 2 Γ ζ(1 + s) = 1+ , (224)
2 ρ
Highly composite numbers 153

[where ρ now runs through all the roots]. Subtracting 1 from both sides, dividing the result
by s, and then making s → 0, we obtain
X1
= 1 + 12 (γ − log 4π), (225)
ρ
where γ is the Eulerian constant. Hence we see that
X ρ X ρ
x √ X √ X 1
 
x 1 1
≤ ρ2 = x
= x +
ρ2 ρ(1 − ρ) ρ 1−ρ
√ X1 √
=2 x = x(2 + γ − log 4π). (226)
ρ
It follows from (220) and (226) that
√ √
(log 4π − γ) x ≤ R(x)(log x)2 ≤ (4 + γ − log 4π) x. (227)

It can easily be verified that



log 4π − γ = 1.954,
(228)
4 + γ − log 4π = 2.046
approximately.

42. Now √
2 x + S(x)
R(x) = ,
(log x)2
where X xρ
S(x) = ;
ρ2
so that, considering R(x) as a function of a continuous variable, we have

′ 1 4 x + 2S(x) S ′ (x)
R (x) = √ − +
x(log x)2 x(log x)3 (log x)2
S ′ (x)
 
1
= 2
+O √ ,
(log x) x(log x)2
for all values of x for which S(x) possesses a differential coefficient.
Now the derived series of S(x), viz.,
1 X xρ
S̄(x) = ,
x ρ
is uniformly convergent throughout any interval of positive values of x which does not
include any value of x of the form x = pm ; and S(x) is continuous for all values of x. It
follows that Z x2
S(x1 ) − S(x2 ) = S̄(x)dx,
x1
154 Paper 15

for all positive values of x1 and x2 , and that S(x) possesses a derivative

S ′ (x) = S̄(x),

whenever x is not of the form pm . Also

(log x)2
 
S̄(x) = O √ .
x

Hence
x+h    
1 h
Z
R(x + h) = R(x) + O √ dt = R(x) + O √ . (229)
x t x

43. Now

log N = ϑ(2x ) + ϑ( 32 )x + O( 34 )x
1
= ϑ(2x ) + ( 23 )x + O{x2 ( 32 ) 2 x } + O( 43 )x
= ϑ(2x ) + ( 32 )x + O(25x/12 ).

Similarly log D(N ) = log 2 · π(2x ) + log( 32 )Li( 23 )x + O(25x/12 ). Writing X for 2x , we have
3 5

log N = ϑ(X) + X log( 2 )/ log 2 + O{(log N ) 12 }; 
3 5
(230)
log D(N ) = log 2.π(X) + log( 32 )Li{X log( 2 )/ log 2 } + O(X 12 . 

It follows that
3
log N = X + O[X log( 2 / log 2 ];
and so
3
X = log N + O[(log N )log( 2 )/ log 2 ]. (231)

Again, from (230) and (231), it follows that


3 5
log N = ϑ(X) + (log N )log( 2 )/ log 2 + O{(log N ) 12 }, (232)

and
3 5
log D(N ) = log 2.π(X) + log( 32 )Li(log N )log( 2 )/ log 2 + O{(log N ) 12 }
( " √ #)
X
= log 2 Liϑ(X) − R(X) + O
(log X)3
3 5
+ log( 23 )Li{(log N )log( 2 )/ log 2 } + O{(log N ) 12 }, (233)
Highly composite numbers 155

in virtue of (222). From (231) and (233) it evidently follows that


3
log D(N ) = log 2 · Liϑ(X) − log 2 · R(X) + log( 23 )Li{(log N )log( 2 )/ log 2 }
( p )
(log N )
+O
(log log N )3
3 5
= log 2 · Li{log N − (log N )log( 2 )/ log 2 + O(log N ) 12 }
3
− log 2 · R{log N + O(log N )log( 2 / log 2 }
( p )
3 (log N )
+ log( 32 )Li{(log N )log( 2 )/ log 2 } + O , (234)
(log log N )3

in virtue of (231) and (232). But


3 5
Li{log N − (log N )|log( 2 )/ log 2 } + O(log N ) 12 }
3 5 3
( ) ( )
(log N )log( 2 )/ log 2 (log N ) 12 (log N ){2 log( 2 )/ log 2}−1
= Li(log N ) − +O +O
log log N log log N (log log N )2
3
(log N )log( 2 )/ log 2 5
= Li(log N ) − + O(log N ) 12 ;
log log N
and
3 3 1
R{log N + O(log N )log( 2 )/ log 2 } = R(log N ) + O{(log N ){log( 2 )/ log 2}− 2 }
1
= R(log N ) + O(log N ) 10 ,

in virtue of (229). Hence (234) may be replaced by


3
log D(N ) = log 2 · Li(log N ) + log( 32 )Li{(log n)log( 2 ) log 2 }
3
( p )
(log N )log( 2 )/ log 2 (log N )
− log 2 − log 2.R(log N ) + O . (235)
log log N (log log N )3

That is to say the maximum order of d(N ) is

2Li(log N )+φ(N ) , (236)

where
3
log( 23 ) 3 (log N )log( 2 )/ log 2
φ(N ) = Li{(log N )log( 2 )/ log 2 } − − R(log N )
log 2 log log N
( p )
(log N )
+O .
(log log N )3
156 Paper 15

This order is actually attained for an infinity of values of N .

44. We can now find the order of the number of superior highly composite numbers not
exceeding a given number N . Let N ′ be the smallest superior highly composite number
greater than N , and let
x )+ϑ( 3 )x +ϑ( 4 )x +···
N ′ = eϑ(2 2 3 .
Then, from § 37, we know that

2N ≤ N ′ ≤ 2x N, (237)

so that N ′ = 0(N log N ); and also that the number of superior highly composite numbers
not exceeding N ′ is
n = π(2x ) + π( 23 )x + π( 43 )x + · · · .
By arguments similar to those of the previous section we can shew that
3
log( 23 )/ log 2 (log N )log( 2 )/ log 2
n = Li(log N ) + Li(log N ) − − R(log N )
log log N
( p )
(log N )
+O . (238)
(log log N )3

It is easy to see from § 37 that, if the largest superior highly composite number not exceeding
N is
2a2 · 3a3 · 5a5 · · · pap ,
then the number of superior highly composite numbers not exceeding N is the sum of all
the indices, viz,
a2 + a3 + a5 + · · · + ap

45. Proceeding as in § 28, we can shew that, if N is a superior highly composite number
and m and n are any two positive integers such that [n is a divisor of N , and]

log mn = o(log log N ),

then
m  2
 log(m/n) log mn
+o
d N = d(N )2 log log N log log N
. (239)
n
From this we can easily shew that the next highly composite number is of the form

N (log log log N )2


 
N +O . (240)
log log N
Highly composite numbers 157

Again, let us S and S be any two consecutive superior highly composite numbers, and let
x )+ϑ( 3 )+ϑ( 4 )x +···
S = eϑ(2 2 3

Then it follows from § 35 that


 1/x
N
d(N ) < d(S), (241)
S

for all values of N except S and s′ . Now, if S be the nth superior highly composite number,
so that
xn ≤ x < xn+1 ,
where xn is the same as in § 36, we see that
 1/xn
N
d(N ) < d(S), (241′ )
S

for all values of N except S and s′ . If N is S or S ′ , then the inequality becomes an equality.
It follows from § 36 that d(S) ≤ 2d(S ′ ). Hence, if N be highly composite and S ′ < N < S,
so that dS ′ < d(N ) < d(S), then
1
2 d(S) < d(N ) < d(S ′ ), d(S ′ ) < d(N ) < 2d(S ′ ).

From this it is easy to see that the order (236) is actually attained by d(N ) whenever N is
a highly composite number. But it may also be attained when N is not a highly composite
number. For example, if

N = (2 · 3 · 5 · · · p1 ) × (2 · 3 · 5 · · · p2 ),

where p1 is the largest prime not greater than 2x , and p2 the largest prime not greater than
( 32 )x , it is easily seen that d(N ) attains the order (236): and N is not highly composite.

VI.
Special Forms of N .

46. In §§-38 we have indirectly solved the following problem: to find the relations which
must hold between x1 , x2 , x3 , . . . in order that

2π(x1 ) · ( 23 )x(x2 ) · ( 43 )π(x3 ) · · ·

may be a maximum, when it is given that

ϑ(x1 ) + ϑ(x2 ) + ϑ(x3 ) + · · ·


158 Paper 15

is a fixed number. The relations which we obtained are


log 2 log( 32 ) log( 43 )
= = = ···.
log x1 log x2 log x3
This suggests the following more general problem. If N is an integer of the form

ec1 ϑ(x1 ) + c2 ϑ(x2 ) + c3 ϑ(x3 ) + · · · , (242)

where c1 , c2 , c3 , . . . are any given positive integers, it is required to find the nature of N ,
that is to say the relations which hold between x1 , x2 , x3 , . . . , when d(N ) is of maximum
order. From (242) we see that
 π(x2 )  π(x3 )
1 + c1 + c2 1 + c1 + c2 + c3
d(N ) = (1 + c1 )π(x1 ) (242′ )
1 + c1 1 + c1 + c2

If we define the “superior” numbers of the class (242) by the inequalities

d(N ) d(N ′ )
≥ ,
Nǫ (N ′ )ǫ

for all values of N ′ less than N , and

d(N ) d(N ′ )
> ,
Nǫ (N ′ )ǫ

for all values of N ′ greater than N, N and N ′ in the two inequalities being of the form
(242), and proceed as in § 33, we can shew that
 x/c2 
 π 1+c1 +c2
1+c1 +c2 1+c1
x/c
(1 + c1 )π{(1+c1 ) 1 } 1+c1
d(N ) ≤ N 1/x x/c
 x/c2  ..., (243)
e(c1 /x)ϑ{(1+c1 ) 1 } (c2 /x)ϑ
1+c1 +c2
1+c1
e
for all values of x, and for all values of N of the form (242). From we can shew, by arguments
similar to those of § 38, that N must be of the form
 x/c2   x/c3 
1+c1 +c2 1+c1 +c2 +c3
c1 ϑ{(1+c1 )x/c1 }+c2 ϑ 1+c1
+c3 ϑ 1+c1 +c2
+···
e , (244)

and d(N ) of the form


  x    x 
 x

 π 1+c1 +c2 c2  π 1+c1 +c2 +c3 c3
π (1+c1 ) c1 1 + c1 + c2 1+c1 1 + c1 + c2 + c3 1+c1 +c2
(1 + c1 )
1 + c1 1 + c1 + c2
(244′ )
From (244) and (244’) we can find the maximum order of d(N ), as in § 43.
Highly composite numbers 159

47. We shall now consider the order of d(N ) for some special forms of N . The simplest
case is that in which N is of the form

2 · 3 · 5 · 7 · · · p;

so that
log N = ϑ(p),
and
d(N ) = 2π(p) .
It is easy to shew that
 √ 
(log N)
Li(log N )−R(log N )+O
(log log N)3
d(N ) = 2 . (245)

In this case d(N ) is exactly a power of 2, and this naturally suggests the question: what
is the maximum order of d(N ) when d(N ) is exactly a power of 2?
It is evident that, if d(N ) is a power of 2, the indices of the prime divisors of N cannot be
any other numbers except 1,3,7,15,31, . . . ; an so in order that d(N ) should be of maximum
order, N must be of the form

eϑ(x1 )+2ϑ(x2 )+4ϑ(x3 )+8ϑ(x4 )+···,

and d(N ) of the form


2π(x1 )+π(x2 )+π(x3 )+··· .
It follows from § 46 that, in order that d(N ) should be of maximum order N , must be of
the form
√ 1 1
eϑ(x)+2ϑ( x)+4ϑ(x 4 )+8ϑ(x 8 )+···,
(246)

and d(N ) of the form


√ 1 1
2π(x)+π( x)+π(x 4 )+π(x 8 )+···.
(247)

Hence the maximum order of d(N ) can easily be shewn to be


√  √ 
4 (log N) (log N)
Li(log N )+ −R(log N )+O
(log log N)2 (log log N)3
2 . (248)

It is easily seen from (246) that the least number having 2n divisors is

2 · 3 · 4 · 5 · 7 · 9 · 11 · 13 · 16 · 17 · 19 · 23 · 25 · 29 · · · to n factors, (249)

where 2, 3, 4, 5, 7, . . . are the natural primes, their squares, fourth powers and so on,
arranged according to order of magnitude.
160 Paper 15

48. We have seen that the last indices of the prime divisors of N must be 1, if d(N ) is of
maximum order. Now we shall consider the maximum order of d(N ) when the indices of
the prime divisors of N are never less than an integer n. In the first place, in order that
d(N ) should be of maximum order, N must be of the form

enϑ(x1 )+ϑ(x2 )+ϑ(x3 )+···,

and d(N ) of the form


 π(x2 )  π(x3 )
π(x1 ) 2+n 3+n
(1 + n) ···.
1+n 2+n

It follows from §46 that N must be of the form


x x
enϑ{(1+n)
x/n }+ϑ
{( 2+n
1+n ) }
3+n
+ϑ{( 2+n ) }+··· , (250)

and d(N ) of the form


 π{( 1+n )x }  π{( 3+n )x }
π{(1+n)x/n } 2+n 1+n 3+n 2+n
(1 + n) ··· (251)
1+n 2+n

Then, by arguments similar to those of § 43, we can shew that the maximum order of d(N )
is

(n + 1)Li{(1/n) log N }+φ(N ) , (252)

where
 
   n log(n+2) −n 
log(n + 2) 1 log(n−1)
φ(N ) = − 1 Li log N
log(n + 1)  n 

1
n log(n+2) −n  ( p )
log N log(n−1)

n 1 (log N )
− 1
 −R log N +) .
n log n log N
n (log log N )3

if n ≥ 3, it is easy to verify that

log(n + 2)
n − n < 12 ;
log(n + 1)

and so (252) reduces to


 √ 
(log N)
Li{(1/n) log N }−R{(1/n) log N }+O
(log log N)3
(n + 1) , (253)

provided that n ≥ 3.
Highly composite numbers 161

49. Let us next consider the maximum order of d(N ) when N is a perfect nth power. In
order that d(N ) should be of maximum order, N must be of the form

enϑ(x2 )+nϑ(x2 )+nϑ(x3 )+···

and d(N ) of the form


 π(x2 )  π(x3 )
1 + 2n 1 + 3n
(1 + n)π(x1 ) ···.
1+n 1 + 2n
It follows from § 46 that N must be of the form
x x
enϑ{(1+n)
x }+nϑ
{( 1+2n
1+n ) }
+nϑ{( 1+3n
1+2n ) }
+···
, (254)

and d(N ) of the form


 π{( 1+2n )x }  π{( 1+3n )x }
pi{(1+n)x } 1 + 2n 1+n 1 + 3n 1+2n
(1 + n) ··· (255)
1+n 1 + 2n

Hence we can shew that the maximum order of d(N ) is


 √ 
(log N)
Li{(1/n) log N }−R{(1/n) log N }+O
(log log N)3
(n + 1) , (256)

provided that n > 1.

50. Let l(N ) denote the least common multiple of the first N natural numbers. Then it
can easily be shewn that

l(N ) = 2[log N log 2] · 3[log N/ log 3] · 5[log N/ log 5] · · · p, (257)

where p is the largest prime not greater than N . From this we can shew that
√ 1 1
l(N ) = eϑ(N )+ϑ( N )+ϑ(N 3 )+ϑ(N 4 )+···
; (258)

and so
1
d{l(N )} = 2π(N ) ( 32 )π(N ) ( 34 )π(N 3 ) · · · . (259)

From (258) and (259) we can shew that, if N is of the form l(M ), then

d(N ) = 2Li(log N )+φ(N ) , (260)

where
( p )
log( 98 ) (log N ) 4 log( 23 )
p
(log N )
φ(N ) = + − R(log N ) + O .
log 2 log log N (log log N )2 (log log N )3
162 Paper 15

It follows from (258) that


√ 2
l(N ) = eN +O{ N (log) } ; (261)

and from (259) that



d{l(N )} = 2Li(n)+O( N log N )
. (262)

51. Finally, we shall consider the number of divisors of N ! It is easily seen that

N ! = 2a2 · 3a3 · 5a5 · · · pap , (263)

where p is the largest prime not greater than N , and


     
N N N
aλ = + 2 + 3 + ···.
λ λ λ

It is evident that the primes greater than 12 N and not exceeding N appear once in N !, the
primes greater than 13 N and not exceeding 12 N appear twice, and so on up to those greater
√ √ √
than N/[ N ] and not exceeding N/( N ] − 1), appearing [ N ] − 1 times∗ . The indices of
the smaller primes cannot be specified so simply. Hence it is clear that
1 1
 
ϑ(N )+ϑ( 2 N )+ϑ( 3 N )+···+ϑ √N
N! = e [ N]−1 × 2a2 · 3a3 · 5a5 · · · ̟ a̟ , (264)

where ̟ is the largest prime not greater than N , and

     
N N N
aλ − 1 + [ N ] = + 2 + 3 + ···.
λ λ λ

From (264) we see that


1 1 √
d(N !) = 2π(N ) ( 23 )π( 2 N ) ( 43 )π( 3 N ) · · · to [ N ] − 1 factors
×eO{log(1+a2 )+log(1+a3 )+···+log(1+a √
̟ )}
1 1
= 2π(N ) ( 32 )π( 2 N ) ( 43 )π( 3 N ) · · · to [ N ] − 1 factors
×eO{̟ log(1+a2 )} √
1 1
= 2Li(N ) ( 32 )Li( 2 N ) ( 43 )Li( 3 N ) · · · to [ N ] factors

×eO( N log N ) . (265)

Since  
N N
Li(N ) = +O ,
log N (log N )2

Strictly speaking, this is true only when N ≥ 4.
Highly composite numbers 163

we see that
n o
N N
+O
d(N !) = C log N (log N)2 , (266)

where
1 1 1
C = (1 + 1)1 (1 + 12 ) 2 (1 + 31 ) 3 (1 + 14 ) 4 . . . .
From this we can easily deduce that, if N is of the form M !, then
n o
log N
+ 2 log N log log log N
+O log N
d(N ) = C (log log N)2 (log log N)3 (log log N)3 , (267)

where C is the same constant as in (266).

52. It is interesting in this connection to shew how, by considering numbers of certain


special forms, we can obtain lower limits for the maximum orders of the iterated functions
dd(n) and ddd(n). By supposing that

N = 22−1 · 33−1 · · · pp−1 ,

we can shew that



(2 log n)
dd(n) > 4 log log n (268)

for an infinity of values of n. By supposing that


a2 −1 a3 −1 ap
N = 22 · 33 · · · pp −1
,

where  
log p
aλ = −1
log λ
we can shew that

ddd(n) > (log n)log log log log n (269)

for an infinity of values of n.


On certain infinite series

Messenger of Mathematics, XLV, 1916, 11 – 15

1. This paper is merely a continuation of the paper on “Some definite integrals” published
in this Journal∗ . It deals with some series which resemble those definite integrals not merely
in form but in many other respects. In each case there is a functional relation. In the case
of the integrals there are special values of a parameter for which the integrals may be
evaluated in finite terms. In the case of the series the corresponding results involve elliptic
functions.

2. It can be shewn, by the theory of residues, that if α and β are real and αβ = 14 π 2 , then

α 3α 5α
− + − ···
(α + t) cosh α (9α + t) cosh 3α (25α + t) cosh 5α
β 3β 5β
+ − + − ···
(β − t) cosh β (9β − t) cosh 3β (25β − t) cosh 5β
π
= p p . (1)
4 cos (αt) cosh (βt)

Now let

αeinα 3αe9inα 5αe25inα


 
F (n) = − + − ···
cosh α cosh 3α cosh 5α
 −inβ 
βe 3βe−9inβ 5βe−25inβ
− − + − ··· (2)
cosh β cosh 3β cosh 5β

Then we see that, if t is positive,


Z∞
πi
e−2tn F (n)dn = p p (3)
4 cosh{(1 − i) (αt)} cosh{(1 + i) (βt)}
0

in virtue of (1). Again, let


r
1 π XX 1 √
(−1) 2 (µ+ν) {µ(1 + i) α − ν(1 − i) β}
p
f (n) = −
2n 2n
2 α+iν 2 β)/4n
×e−(πµν−iµ (µ = 1, 3, 5, . . . ; ν = 1, 3, 5, . . .). (4)

[No. 11 of this volume (pp. 66 – 74); see also No.12 (pp. 75 – 86).]
On certain infinite series 165

Then it is easy to shew that


Z∞
πi
e−2tn f (n)dn = p p . (5)
4 cosh{(1 − i) (αt)} cosh{(1 + i) (βt)}
0

Hence, by a theorem due to Lerch∗ , we obtain


F (n) = f (n) (6)
for all positive values of n, provided that αβ = 14 π 2 . In particular, when α = β = 21 π, we
have
sin 12 πn 3 sin 29 πn 5 sin 25 2 πn
1 − 3 + − ···
cosh 2 π cosh 2 π cosh 25 π
1 XX 1
=− √ (−1) 2 (µ+ν) e−πµν/4n
4n n
π(µ2 − ν 2 ) π(µ2 − ν 2 )
 
(µ + ν) cos + (µ − ν) sin
8n 8n
(µ = 1, 3, 5, . . . ; ν = 1, 3, 5, . . .) (7)
for all positive values of n. As particular cases of (7), we have
sin( 12 π/a) 3 sin( 29 π/a) 5 sin( 25
2 π/a)
1 − 3 + − ···
cosh 2 π cosh 2 π cosh 52 π
!
1 √ 1 3 5
= a a − + − ···
4 cosh 4 πa cosh 4 πa cosh 45 πa
1 3

1 √ 1 9 25 49
= a a(e− 16 πa − e− 16 πa − e− 16 πa + e− 16 πa + · · ·)4 , (8)
2
if a is a positive even integer; and
sin( 12 π/a) 3 sin( 92 π/a)
− + ···
cosh 21 π cosh 32 π
!
1 √ cosh 41 πa 3 cosh 34 πa 5 cosh 45 πa 7 cosh 74 πa
= a a + − − + ··· , (9)
2 cosh 12 πa cosh 32 πa cosh 25 πa cosh 72 πa

if a is a positive odd integer; and so on.

3. It is also easy to shew that if αβ = π 2 , then


 
α 2α 3α
− + − ···
(α + t) sinh α (4α + t) sinh 2α (9α + t) sinh 3α

See Mr. Hardy’s note at the end of my previous paper [Messenger of Mathematics, XLIV, pp. 18 – 21].
166 Paper 16

 
β 2β 3β
− − + − ···
(β − t) sinh β (4β − t) sinh 2β (9β − t) sinh 3β
1 π
= − p p . (10)
2t 2 sin (αt) sinh (βt)

From this we can deduce, as in the previous section, that if αβ = π 2 , then

αeinα 2αe4inα 3αe9inα βe−inβ 2βe−4inβ 3βe−9inβ


− + − ··· + − + − ···
sinh α sinh 2α sinh 3α sinh β sinh 2β sinh 3β
r
1 1  π  XX √ 2 2
{µ(1 − i) α + ν(1 + i) β}e−(2πµν−iµ α+iν β)/4n
p
= − ×
2 n 2n
(µ = 1, 3, 5, . . . ; ν = 1, 3, 5, . . .) (11)

for all positive values of n. If, in particular, we put α = β = π, we obtain


1 cos πn 2 cos 4πn 3 cos 9πn
− + − + ···
4π sinh π sinh 2π sinh 3π
π(µ2 − ν 2 ) π(µ2 − ν 2 )
 
1 XX
−πµν/2n
= p e (µ + ν) cos +(µ − ν) sin
2n (2n) 4n 4n
(µ = 1, 3, 5, . . . ; ν = 1, 3, 5, . . .) (12)

for all positive values of n. Thus, for example, we have


1 cos(2π/a) 2 cos(8π/a) 3 cos(18π/a)
− + − + ···
4π sinh π sinh 2π sinh 3π
!
1 √ 1 3 5
= a a + + + ...
8 sinh 4 πa sinh 4 πa sinh 54 πa
1 3

1 √ 1 9 25 49
= a a(e− 16 πa + e− 16 πa + e− 16 πa + e− 16 πa + · · ·)4 , (13)
4
if a is a positive even integer; and
1 cosh(2π/a) 2 cosh(8π/a)
− + − ···
4π sinh π sinh 2π
!
1 √ sinh 41 πa 3 sinh 43 πa 5 sinh 45 πa 7 sinh 74 πa
= a a − − + + ··· (14)
4 cosh 12 πa cosh 32 πa cosh 25 πa cosh 72 πa

if a is a positive odd integer.

4. In a similar manner we can shew that, if αβ = π 2 , then


αeinα 2αe4inα 3αe9inα βe−inβ 2βe−4inβ 3βe−9inβ
+ + + · · · + + + + ···
e2α − 1 e4α − 1 e6α − 1 e2β − 1 e4β − 1 e6β − 1
On certain infinite series 167

Z∞ 2 Z∞ inβx2
xe−inαx xe 1
=α 2πx
dx + β

2πx
dx −
e −1 e −1 4
0 0
r  ∞ ∞
1 π XX √ 2 2
{µ(1 − i) α + ν(1 + i) β}e−(2πµν−iµ α+iν β)/n
p
+ (15)
n 2n
µ=1 ν=1

for all positive values of n. Putting α = β = π in (15) we see that, if n > 0, then
1 cos πn 2 cos 4πn 3 cos 9πn
+ 2π + 4π + 6π + ···
8π e − 1 e −1 e −1
Z∞ ∞ X∞
x cos πnx2 1 X
= dx + e−2πµν/n
e2πx − 1
p
2n (2n) µ=1 ν=1
0

π(µ2 − ν 2 ) π(µ2 − ν 2 )
    
× (µ + ν) cos + (µ − ν) sin . (16)
n n

As particular cases of (16) we have

1 cos(π/a) 2 cos(4π/a) 3 cos(9π/a)


+ 2π + + + ···
8π e −1 e4π − 1 e6π − 1
Z∞
x cos(πx2 /a)
 
1 2 3
q
1
= dx + a ( 2 a) 2πa + + + ··· , (17)
e2πx − 1 e − 1 e4πa − 1 e6πa − 1
0

If a is a positive even integer;


1 cos(π/a) 2 cos(4π/a) 3 cos(9π/a)
+ 2π + + + ···
8π e −1 e4π − 1 e6π − 1
Z∞
x cos(πx2 /a)
 
1 2 3
q
1
= dx + a ( 2 a) 2πa − + − ··· , (18)
e2πx − 1 e + 1 e4πa + 1 e6πa + 1
0

if a is a positive odd integer; and


Z∞
1 cos(2π/a) 2 cos(8π/a) 3 cos(18π/a) x cos(2πx2 /a) √
+ 2π + + + ··· = dx + 14 a aS
8π e −1 e4π − 1 e6π − 1 e2πx −1
0

(−1)n−1 n

X X n
where S = nπa n
or S = nπa
(19)
e + (−1) e + (−1)n−1
n=1 n=1

according as a ≡ 1 or a ≡ 3 (mod 4).



I shewed in my former paper [No.12 of the present volume] that this integral can be calculated in finite
terms whenever nα is a rational multiple of π.
168 Paper 16

It may be interesting to note that different functions dealt with in this paper have the same
asymptotic expansion for small values of n. For example, the two different functions
1 cos n 2 cos 4n 3 cos 9n
+ 2π + 4π + 6π + ···
8π e − 1 e −1 e −1
and
Z∞
x cos nx2
dx
e2πx − 1
0

have the same asymptotic expansion, viz.

1 n2 n4 n6
− + − + ···∗ (20)
24 1008 6336 17280


This series (in spite of the appearance of the first few terms) diverges for all values of n.
Some formulæ in the analytic theory of numbers
Messenger of Mathematics, XLV, 1916, 81 – 84
I have found the following formulæ incidentally in the course of other investigations. None
of them seem to be of particular importance, nor does their proof involve the use of any
new ideas, but some of them are so curious that they seem to be worth printing. I denote
by d(x) the number of divisors of x, if x is an integer, and zero otherwise, and by ζ(s) the
Riemann Zeta-function.
ζ 4 (s)
(A) = 1−s d2 (1) + 2−s d2 (2) + 3−s d2 (3) + · · · , (1)
ζ(2s)

η 4 (s)
= 1−s d2 (1) − 3−s d2 (3) + 5−s d2 (5) − · · · , (2)
(1 − 2−2s )ζ(2s)
where
η(s) = 1−s − 3−s + 5−s − 7−s + · · · .

(B) d2 (1) + d2 (2) + d2 (3) + . . . + d2 (n)


3
= An(log n)3 + Bn(log n)2 + Cn log n + Dn + O(n 5 +ǫ ), ∗ (3)

where
1 12γ − 3 36 ′
A= 2
,B= − 4 ζ (2),
π π2 π
γ is Euler’s constant, C, D more complicated constants, and ǫ any positive number.

n n n


(C) d3 + d3 + d3 + ···
1 3 2
n n n n o2
= d +d +d + ··· ,† (4)
1 2 3


X
n−s dr (n) = {ζ(s)}2 φ(s),
r
(5)
1

where φ(s) is absolutely convergent for R(s) > 21 , and in particular


∞ Y  
X 1 s 1 p
= p log = {ζ(s)}φ(s). (6)
ns d(n) p
1 − p−s
1

∗ 1
If we assume the Riemann hypothesis, the error term here is of the form O(n 2 +ǫ ).

Mr Hardy has pointed out to me that this formula has been given already by Liouville, Journal de
Mathématiques, Ser.2, Vol. II (1857), p.393.
170 Paper 17

1 1 1 1
(D) + + + ··· +
d(1) d(2) d(3) d(n)
( )
A1 A2 Ar 1
=n 1 + 3 + ··· + 1 +O 1 , (7)
(log n) 2 (log n) 2 (log n)r− 2 (log n)r+ 2

where   
1 Y p 2 p
A1 = √ (p − p) log
π p p−1

and A2 , A3 , . . . Ar are more complicated constants.


More generally

ds (1) + ds (2) + ds (3) + . . . + ds (n)


1
= n{A1 (log n)2 + A2 (log n)2 + . . . + A2s } + O(n 2 +ǫ ), ∗
s −1 s −2
(8)

if 2s is an integer, and

ds (1) + ds (2) + ds (3) + . . . + ds (n)


  
2s −1 2s −2 Ar+2s 1
= n A1 (log n) + A2 (log n) + ... + +O , (9)
(log n)r (log n)r+1

if 2s is not an integer, the A’s being constants.

(E) d(1)d(2)d(3) . . . d(n) = 2n(log log n+C)+φ(n) , (10)

where
∞    
X 1 1
C=γ+ log2 1 + − (2−ν + 3−ν + 5−ν + . . .).
ν ν
2

Here 2, 3, 5, . . . are the primes and

φ(n) γ−1 1! 2!
= + (γ + γ1 − 1) + (γ + γ1 + γ2 − 1) + . . .
n log n (log n)2 (log n)3  
(r − 1)! 1
+ (γ + γ1 + γ2 + . . . + γr−1 − 1) + O ,
(log n)r (log n)r+1

where
1
ζ(1 + s) = + γ − γ1 s + γ2 s2 − γ3 s3 + . . .
s
or  
r 1 r 1 r 1 r+1
r!γr = lim (log 1) + (log 2) + . . . + (log ν) − (log ν) .
ν→∞ 2 ν r+1

Assuming the Riemann hypothesis.
Some formulæ in the analytic theory of numbers 171


X u v  X u v 
(F) d(uv) = µ(n)d d = µ(δ)d d , (11)
n n δ δ
1

where δ is a common factor of u and v, and



1 X µ(n)
= .
ζ(s) ns
1

(G) If Dv (n) = d(v) + d(2v) + . . . + d(nv),


we have
X v n
Dv (n) = µ(δ)d D1 , (12)
δ δ
where δ is a divisor of v, and

Dv (n) = α(v)n(log n + 2γ − 1) + β(v)n + ∆v (n), (13)

where
∞ ∞
X α(ν) ζ 2 (s) X β(ν) ζ 2 (s)ζ ′ (1 + s)
= , = − ,
νs ζ(1 + s) νs ζ 2 (1 + s)
1 1
and
1
∆v (n) = O(n 3 log n) ∗

(H) d(v + c) + d(2v + c) + d(3v + c) + . . . + d(nv + c)


= αc (v)n(log n + 2γ − 1) + βc (v)n∆v,c (n), (14)

where

X αc (ν) ζ(s)σ−s (|c|)
= ,
νs ζ(1 + s)
1

ζ ′ (s) ζ ′ (1 + s) σ−s ′ (|c|)
 
X βc (ν) ζ(s)σ−s (|c|)
s
= + + ,
1
ν ζ(1 + s) ζ(s) ζ(1 + s) σ−s (|c|)
σs (n) being the sum of the sth powers of the divisors of n and σs′ (n) the derivative of σs (n)
with respect to s, and
1
∆v,c (n) = O(n 3 log n).†
(I) The formulæ(1) and (2) are special cases of
ζ(s)ζ(s − a)ζ(s − b)ζ(s − a − b)
ζ(2s − a − b)
∗ 1
It seems not unlikely that ∆v (n) is of the form O(n 4 +ǫ ). Mr Hardy has recently shewn that ∆1 (n) is
1
not of the form o{(n log n) 4 log log n}. The same is true in this case also.

It is very likely that the order of ∆v,c (n) is the same as that of ∆1 (n).
172 Paper 17

= 1−s σa (1)σb (1) + 2−s σa (2)σb (2) + 3−s σa (3)σb (3) + · · · ; (15)

η(s)η(s − a)η(s − b)η(s − a − b)


(1 − 2−2s+a+b )ζ(2s − a − b)
= 1−s σa (1)σb (1) − 3−s σa (3)σb (3) + 5−s σa (5)σb (5) − · · · (16)

It is possible to find an approximate formula for the general sum

σa (1)σb (1) + σa (2)σb (2) + . . . + σa (n)σb (n). (17)

The general formula is complicated, The most interesting cases are a = 0, b = 0, when the
formula is (3); a = 0, b = 1, when it is

π 4 n2
(log n + 2c) + nE(n), (18)
72ζ(3)

where
1 ζ ′ (2) ζ ′ (3)
c=γ− + − ,
4 ζ(2) ζ(3)
and the order of E(n) is the same as that of ∆1 (n); and a = 1, b = 1, when it is

5 3
n ζ(3) + E(n), (19)
6
where
E(n) = O{n2 (log n)2 }, E(n) 6= o(n2 log n).
(J) If s > 0, then

σs (1)σs (2)σs (3)σs (4) . . . σs (n) = θcn (n!)s , (20)

where
1 > θ > (1 − 2−s )(1 − 3−s )(1 − 5−s ) . . . (1 − ̟ −s ),
̟ is the greatest prime not exceeding n, and
2 3
Y  p2s − 1 1/p  p3s − 1 1/p  p4s − 1 1/p
( )
c= ··· .
p
p2s − ps p3s − ps p4s − ps

P∞
(K) If ( 12 + q + q 4 + q 9 + q 16 + · · ·)2 = 1
4 + 1 r(n)q n ,
so that

X
ζ(s)η(s) = r(n)n−s ,
1
Some formulæ in the analytic theory of numbers 173

then
ζ 2 (s)η 2 (s)
= 1−s r 2 (1) + 2−s r 2 (2) + 3−s r 2 (3) + · · · (21)
(1 + 2−s )ζ(2s)

n 3
r 2 (1) + r 2 (2) + r 2 (3) + · · · + r 2 (n) = (log n + C) + O(n 5 +ǫ ), (22)
4
where
1 3 12
C = 4γ − 1 + log 2 − log π + 4 log Γ( ) − 2 ζ ′ (2).
3 4 π
These formulæ are analogous to (1) and (3).
On certain arithmetical functions
Transactions of the Cambridge Philosophical Society, XXII, No.9, 1916, 159 – 184

1. Let σs (n) denote the sum of the sth powers of the divisors of n (including 1 and n), and
let
σs (0) = 12 ζ(−s),
where ζ(s) is the Riemann Zeta-function. Further let
X
(n) = σr (0)σs (n) + σr (1)σs (n − 1) + · · · + σr (n)σs (0). (1)
r,s

In this paper I prove that


X Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)
(n) = σr+s+1 (n)
r,s
Γ(r + s + 2) ζ(r + s + 2)

ζ(1 − r) + ζ(1 − s) 2
+ nσr+s−1 (n) + O{n 3 (r+s+1) }, (2)
r+s
whenever r and s are positive odd integers. I also prove that there is no error term on the
right-hand side of (2) in the following nine cases: r = 1, s = 1; r = 1, s = 3; r = 1, s =
5; r = 1, sP= 7; r = 1, s = 11, r = 3, s = 3; r = 3, s = 5; r = 3, , s = 9; r = 5, s = 7. That
is to say r,s (n) has a finite expression in terms of σr+s+1 (n) and σr+s−1 (n) in these nine
cases; but for other values of r and s it involves other arithmetical functions as well.
It appears probable, from the empirical results I obtain in §§ 18-23, that the error term on
the right-hand side of (2) is of the form
1
O{n 2 (r+s+1+ǫ) }, (3)

where ǫ is any positive number, and not of the form


1
o{n 2 (r+s+1) }. (4)

But all I can prove rigorously is (i) that the error is of the form
2
O{n 3 (r+s+1) }

in all cases, (ii) that it is of the form


2 3
O{n 3 (r+s+ 4 ) } (5)

if r + s is of the form 6m, (iii) that it is of the form


2 1
O{n 3 (r+s+ 2 ) } (6)
On certain arithmetical functions 175

if r + s is of the form 6m + 4, and (iv) that it is not of the form


1
o{n 2 (r+s) }. (7)

It follows from (2) that, if r and s are positive odd integers, then
X Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)
(n) ∼ σr+s+1 (n). (8)
r,s
Γ(r + s + 2) ζ(r + s + 2)

It seems very likely that (8) is true for all positive values of r and s, but this I am at present
unable to prove.
P
2. If r,s (n)/σr+s+1 (n) tends to a limit, then the limit must be

Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)


.
Γ(r + s + 2) ζ(r + s + 2)

For then
P P P P
r,s (n) r,s (1) + r,s (2) + ··· + r,s (n)
lim = lim
n→∞ σr+s+1 (n) n→∞ σr+s+1 (1) + σr+s+1 (2) + · · · + σr+s+1 (n)
2
P P P
r,s (0) + r,s (1)x + r,s (2)x + · · ·
= lim
x→1 σr+s+1 (0) + σr+s+1 (1)x + σr+s+1 (2)x2 + · · ·
Sr Ss
= lim ,
x→1 Sr+s+1

where

1r x 2r x2 3r x3
Sr = 12 ζ(−r) + + + + ··· (9)
1 − x 1 − x2 1 − x3
Now it is known that, if r > 0, then

Γ(r + 1)ζ(r + 1)
Sr ∼ , (10)
(1 − x)r+1

as x → 1 ∗ . Hence we obtain the result stated.

3. It is easy to see that

σr (1) + σr (2) + σr (3) + · · · + σr (n)


= u1 + u2 + u3 + u4 + · · · + un ,

Knopp, Dissertation (Berlin, 1907), p.34.
176 Paper 18

where h n ir
ut = 1r + 2r + 3r + · · · + .
t
From this it is easy to deduce that
nr+1 ∗
σr (1) + σr (2) + · · · + σr (n) ∼ ζ(r + 1) (11)
r+1
and
Γ(r + 1)Γ(s + 1)
σr (1)(n − 1)s + σr (2)(n − 2)s + · · · + σr (n − 1)1s ∼ ζ(r + 1)nr+s+1 ,
Γ(r + s + 2)
provided r > 0, s ≥ 0. Now
σs (n) > ns ,
and
σs (n) < ns (1−s + 2−s + 3−s + · · ·) = ns ζ(s).
From these inequalities and (1) it follows that
P
r,s (n) Γ(r + 1)Γ(s + 1)
lim r+s+1 ≥ ζ(r + 1), (12)
n Γ(r + s + 2)
if r > 0 and s ≥ 0; and
P
r,s (n)
Γ(r + 1)Γ(s + 1)
lim r+s+1
≤ ζ(r + 1)ζ(s), (13)
n Γ(r + s + 2)

if r > 0 and s > 1. Thus n−r−s−1 r,s (n) oscillates between limits included in the interval
P

Γ(r + 1)Γ(s + 1) Γ(r + 1)Γ(s + 1)


ζ(r + 1), ζ(r + 1)ζ(s).
Γ(r + s + 2) Γ(r + s + 2)

On the other hand n−r−s−1 σr+s+1 (n) oscillates between 1 and ζ(r + s + 1), assuming values
as near as we please to either of
Pthese limits. The formula (8) shews that the actual limits
of indetermination of n−r−s−1 r,s (n) are

Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)


,
Γ(r + s + 2) ζ(r + s + 2)
Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)ζ(r + s + 1)
. (14)
Γ(r + s + 2) ζ(r + s + 2)
Naturally
ζ(r + 1)ζ(s + 1) ζ(r + 1)ζ(s + 1)ζ(r + s + 1)
ζ(r + 1) < < < ζ(r + 1)ζ(s) .†
ζ(r + s + 2) ζ(r + s + 2)

(10) follows from this as an immediate corollary.
On certain arithmetical functions 177

What is remarkable about the formula (8) is that it shews the asymptotic equality of two
functions neither of which itself increases in a regular manner.

4. It is easy to see that, if n is a positive integer, then

cot 21 θ sin nθ = 1 + 2 cos θ + 2 cos 2θ + · · · + 2 cos(n − 1)θ + cos nθ.

Suppose now that


2
x sin θ x2 sin 2θ x3 sin 3θ

1 1
4 cot 2θ + + + + ···
1−x 1 − x2 1 − x3
= ( 14 cot 12 θ)2 + C0 + C1 cos θ + C2 cos 2θ + C3 cos 3θ + · · · ,

where Cn is independent of θ. Then we have

x2 x3
 
1 x
C0 = 2 + + + ···
1 − x 1 − x2 1 − x3
2 2 3 2
( 2   )
x x x
+ 21 + + + ···
1−x 1 − x2 1 − x3

x2 x3
 
1 x
= 2 + + + ···
(1 − x)2 (1 − x2 )2 (1 − x3 )2
2x2 3x3
 
x
= 21 + + + · · · . (15)
1 − x 1 − x2 1 − x3

Again

1 xn xn+1 xn+2 xn+3


Cn = 21 + + + + ···
− xn 1 − xn+1 1 − xn+2 1 − xn+3
x xn+1 x2 xn+2 x3 xn+3
+ · + · + · + ···
1 − x 1 − xn+1 1 − x2 1 − xn+2 1 − x3 1 − xn+3
xn−1 x2 xn−2 xn−1
 
1 x x
− 2 · + · + ··· + · .
1 − x 1 − xn−1 1 − x2 1 − xn−2 1 − xn−1 1 − x

Hence
xn+1 x2 xn+2
   
Cn x
(1 − xn ) = 1
2 + − + − + ···
xn 1 − x 1 − xn+1 1 − x2 1 − xn+2
xn−1 x2 xn−2
   
1 x
− 2 1+ + + 1+ +
1 − x 1 − xn−1 1 − x2 1 − xn−2

For example when r = 1 and s = 9 this inequality becomes 1.64493 . . . < 1.64616 . . . < 1.64697 . . . <
1.64823 . . ..
178 Paper 18

xn−1
 
x
+··· + 1+ +
1 − xn−1 1 − x
1 n
= − .
1 − xn 2
That is to say
xn nxn
Cn = − . (16)
(1 − xn )2 2(1 − xn )
It follows that
2
x sin θ x2 sin 2θ x3 sin 3θ

1
4 cot 12 θ + + + + ···
1−x 1 − x2 1 − x3

1 x cos θ x2 cos 2θ
2 x3 cos 3θ
= 4 cot 21 θ
+ + + + ···
(1 − x)2 (1 − x2 )2 (1 − x3 )2
2x2 3x3
 
1 x
+2 (1 − cos θ) + (1 − cos 2θ) + (1 − cos 3θ) + · · · . (17)
1−x 1 − x2 1 − x3
Similarly, using the equation
cot2 12 θ(1 − cos nθ) =
(2n − 1) + 4(n − 1) cos θ + 4(n − 2) cos 2θ + · · · + 4 cos(n − 1)θ + cos nθ,

we can shew that


2x2

x
1
8 cot2 21 θ + 1
+
(1 − cos θ) +
12 (1 − cos 2θ)+
1−x 1 − x2
2
3x3 1 2
(1 − cos 3θ) + · · · = 18 cot2 12 θ + 12

1−x 3
 3
23 x2 33 x3

1 1 x
+ 12 (5 + cos θ) + (5 + cos 2θ) + (5 + cos 3θ) + · · · . (18)
1−x 1 − x2 1 − x3

For example, putting θ = 32 π and θ = 21 π in (17), we obtain


2
x2 x4 x5

1 x
+6 − + − + ···
1 − x 1 − x2 1 − x4 1 − x5
2x2 4x4 5x5
 
1 1 x
= 36 +3 + + + + ··· , (19)
1 − x 1 − x2 1 − x4 1 − x5
where 1, 2, 4, 5, . . . are the natural numbers without the multiples of 3; and
2
x3 x5 x7

1 x
4 + 1 − x − 1 − x3 + 1 − x5 − 1 − x7 + · · ·
On certain arithmetical functions 179

2x2 3x3 5x5


 
1 1 x
= 16 + 2 + + + + ··· , (20)
1 − x 1 − x2 1 − x3 1 − x5
where 1, 2, 3, 5, . . . are the natural numbers without the multiples of 4.

5. It follows from (18) that


2
θ2 θ4 θ6

1
+ S3 − S5 + S7 − · · ·
2θ 2 2! 4! 6!
 2
θ4 θ6

1 1 1 θ
= 4 + 2 S3 − 12 S5 − S7 + S9 − · · · , (21)
4θ 2! 4! 6!
where Sr is the same as in (9). Equating the coefficients of θ n in both sides in (21), we
obtain
   
(n − 2)(n + 5) n n
Sn+3 = S3 Sn−1 + S5 Sn−8 +
12(n + 1)(n + 2) 2 4
   
n n
S7 Sn−5 + · · · + Sn−1 S3 , (22)
6 n−2
where  
n n!
= ,
r r!(n − r)!
if n is an even integer greater than 2.
Let us now suppose that
m=∞
X n=∞
X
Φr,s (x) = mr ns xmn , (23)
m=1 n=1

so that
Φr,s (x) = Φs,r (x),
and

1s x 2s x2 3s x3
Φ0,s (x) = + + + · · · = Ss − 12 ζ(−s),


1 − x 1 − x2 1 − x3


. (24)
1s x 2s x2 3s x3 
Φ1,s (x) = + + + ···


(1 − x)2 (1 − x2 )2 (1 − x3 )2

Further let
2x2 3x3
  
x
P = −24S1 = 1 − 24 + + + · · · ∗,


1 − x 1 − x2 1 − x3




 3 
23 x2 33 x3
 
1 x

Q = 240S3 = 1 + 240 + + + · · · , . (25)
1 − x 1 − x2 1 − x3 

 5 
25 x2 35 x3
 
1 x


R = −540S5 = 1 − 504 + + + ···


1 − x 1 − x2 1 − x3

180 Paper 18

The putting n = 4, 6, 8, . . . in (22) we obtain the results contained in the following table.

TABLE 1

1. 1 − 24Φ0,1 (x) = P.
2. 1 + 240Φ0,3 (x) = Q.
3. 1 − 504Φ0,5 (x) = R.
4. 1 + 480Φ0,7 (x) = Q2 .
5. 1 − 264Φ0,9 (x) = QR.
6. 691 + 65520Φ0,11 (x) = 441Q3 + 250R2 .
7. 1 − 24Φ0,13 (x) = Q2 R.
8. 3617 + 16320Φ0,15 (x) = 1617Q4 + 2000QR2 .
9. 43867 − 28728Φ0,17 (x) = 38367Q3 R + 5500R3 .
10. 174611 + 13200Φ0,19 (x) = 53361Q5 + 121250Q2 R2 .
11. 77683 − 552Φ0,21 (x) = 57183Q4 R + 20500QR3 .
12. 236364091 + 131040Φ0,23 (x) = 49679091Q6 + 176400000Q3 R2
+10285000R4 .
13. 657931 − 24Φ0,25 (x) = 392931Q5 R + 265000Q2 R3 .
14. 3392780147 + 6960Φ0,27 (x) = 489693897Q7 + 2507636250Q4 R2
+395450000QR4 .
15. 1723168255201 − 171864Φ0,29 (x) = 815806500201Q6 R
+881340705000Q3 R3 + 26021050000R5 .
16. 7709321041217 + 32640Φ0,31 (x) = 764412173217Q8
+5323905468000Q5 R2 + 1621003400000Q2 R4 .
In general
X
1
2 ζ(−s) + Φ0,s (x) = Km,n Qm Rn , (26)

where Km,n is a constant and m and n are positive integers (including zero) satisfying the
equation
4m + 6n = s + 1.
This is easily proved by induction, using (22).

If x = q 2 , then in the notation of elliptic functions
 2  
12ηω 2K 3E 2
P = = + k − 2 ,
π2 π K
4
12g2 ω 4

2K
Q = = (1 − k2 + k4 ),
π4 π
6
216g3 ω 6

2K
R = = (1 + k2 )(1 − 2k2 )(1 − 12 k2 ).
π6 π
On certain arithmetical functions 181

6. Again from (17) we have


2
θ3 θ5

1 θ
+ S1 − S3 + S5 − · · ·
2θ 1! 3! 5!
1 θ2 θ4 θ6
= 2
+ S1 − Φ1,2 (x) + Φ1,4 (x) − Φ1,6 (x) + · · ·
4θ 2! 4! 6!
 2 4 θ6

1 θ θ
+2 S3 − S5 + S7 − · · · . (27)
2! 4! 6!
Equating the coefficients of θ n in both sides in (27) we obtain
   
n+3 n n
Sn+1 − Φ1,n (x) = S1 Sn−1 + S3 Sn−3 +
2(n + 1) 1 3
   
n n
S5 Sn−5 + · · · + Sn−1 S1 , (28)
5 n−1
if n is a positive even integer. From this we deduce the results contained in Table II.

TABLE II
1. 288Φ1,2 (x) = Q − P 2 .
2. 720Φ1,4 (x) = P Q − R.
3. 1008Φ1,6 (x) = Q2 − P R.
4. 720Φ1,8 (x) = Q(P Q − R).
5. 1584Φ1,10 (x) = 3Q3 + 2R2 − 5P QR.
6. 65520Φ1,12 (x) = P (441Q3 + 250R2 ) − 691Q2 R.
7. 144Φ1,14 (x) = Q(3Q3 + 4R2 − 7P QR).
In general
X
Φ1,s (x) = Kl,m,n P l Qm Rn , (29)

where l ≤ 2 and 2l + 4m + 6n = s + 2. This is easily proved by induction, using (28).

7. We have
dP P2 − Q

x = − 24Φ1,2 (x) = , 

dx 12




dQ PQ − R

x = 240Φ1,4 (x) = , (30)
dx 3 


dR P R − Q2


x = − 504Φ1,6 (x) =


dx 2
Suppose now that r < s and that r + s is even. Then
d r
 
Φr,s (x) = x Φ0,s−r (x), (31)
dx
182 Paper 18

and Φ0,s−r (x) is a polynomial in Q and R. Also


dP dQ dR
x ,x ,x
dx dx dx
are polynomials in P, Q and R. Hence Φr,s (x) is a polynomial in P, Q and R. Thus we
deduce the results contained in Table III.

TABLE III
1. 1728Φ2,3 (x) = 3P Q − 2R − P 3 .
2. 1728Φ2,5 (x) = P 2 Q − 2P R + Q2 .
3. 1728Φ2,7 (x) = 2P Q2 − P 2 R − QR.
4. 8640Φ2,9 (x) = 9P 2 Q2 − 18P QR + 5Q3 + 4R2 .
5. 1728Φ2,11 (x) = 6P Q3 − 5P 2 QR + 4P R2 − 5Q2 R.
6. 6912Φ3,4 (x) = 6P 2 Q − 8P R + 3Q2 − P 4 .
7. 3456Φ3,6 (x) = P 3 Q − 3P 2 R + 3P Q2 − QR.
8. 5184Φ3,8 (x) = 6P 2 Q2 − 2P 3 R − 6P QR + Q3 + R2 .
9. 20736Φ4,5 (x) = 15P Q2 − 20P 2 R + 10P 3 Q − 4QR − P 5 .
10. 41472Φ4,7 (x) = 7(P 4 Q − 4P 3 R + 6P 2 Q2 − 4P QR) + 3Q3 + 4R2 .
In general
X
Φr,s (x) = Kl,m,n P l Qm Rn , (32)

where l − 1 does not exceed the smaller of r and s and

2l + 4m + 6n = r + s + 1.

The results contained in these three tables are of course really results in the theory of
elliptic functions. For example Q and R are substantially the invariants g2 and g3 , and the
formulæ of Table I are equivalent to the formulæ which express the coefficients in the series
1 g2 u2 g3 u4 g22 u6 3g2 g3 u8
℘(u) = + + + + + ···
u2 20 28 1200 6160
in terms of g2 and g3 . The elementary proof of these formulæ given in the preceding sections
seems to be of some interest in itself.

8. In what follows we shall require to know the form of Φ1,s (x) more precisely than is
shewn by the formula (29).
We have
X
1
2 ζ(−s) + Φ0,s (x) = Km,n Qm Rn , (33)

where s is an odd integer greater than 1 and 4m + 6n = s + 1. Also


d m n m n 
x (Qm Rn ) = + P Qm R n − Qm−1 Rn+1 + Qm+2 Rn−1 . (34)
dx 3 2 3 3
On certain arithmetical functions 183

Differentiating (33) and using (34) we obtain


X
Φ1,s+1 (x) = 1
12 (s + 1)P { 12 ζ(−s) + Φ0,s (x)} + Km,n Qm Rn , (35)

where s is an odd integer greater than 1 and 4m + 6n = s + 3. But when s = 1 we have

Q − P2
Φ1,2 (x) = . (36)
288

9. Suppose now that


Fr,s (x) = { 21 ζ(−r) + Φ0,r (x)}{ 12 ζ(−s) + Φ0,s (x)}
ζ(1 − r) + ζ(1 − s) Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)
− Φ1,r+s (x) −
r+s Γ(r + s + 2) ζ(r + s + 2)
×{ 12 ζ(−r − s − 1) + Φ0,r+s+1 (x)}. (37)

Then it follows from (33), (35) and (36) that, if r and s are positive odd integers,
X
Fr,s (x) = Km,n Qm Rn , (38)

where
4m + 6n = r + s + 2.
But it is easy to see, from the functional equation satisfied by ζ(s), viz.

(2π)−s Γ(s)ζ(s) cos 12 πs = 21 ζ(1 − s), (39)

that

Fr,s (0) = 0. (40)

Hence Q3 − R2 is a factor of the right-hand side in (38), that is to say


X
Fr,s (x) = (Q3 − R2 ) Km,n Qm Rn , (41)

where
4m + 6n = r + s − 10.

10. It is easy to deduce from (30) that


d
x log(Q3 − R2 ) = P. (42)
dx
But it is obvious that
d
P =x log[x{(1 − x)(1 − x2 )(1 − x3 ) · · ·}24 ]; (43)
dx
184 Paper 18

and the coefficient of x in Q3 − R2 = 1728. Hence

Q3 − R2 = 1728x{(1 − x)(1 − x2 )(1 − x3 ) · · ·}24 . (44)

But it is known that

{(1 − x)(1 − x2 )(1 − x3 )(1 − x4 ) · · ·}3


= 1 − 3x + 5x3 − 7x6 + 9x10 − · · · (45)

Hence

Q3 − R2 = 1728x(1 − 3x + 5x3 − 7x6 + · · ·)8 . (46)

The coefficient of xν−1 in 1−3x+5x3 −· · · is numerically less than (8ν), and the coefficient
p

of xν in Q3 − R2 is therefore numerically less than that of xν in

1728x{ (8ν)(1 + x + x3 + x6 + · · ·)}8 .


p

But
13 x 23 x2 33 x3
x(1 + x + x3 + x6 + · · ·)8 = 2
+ 4
+ + ···, (47)
1−x 1−x 1 − x6
and the coefficient of xν in the right-hand side is positive and less than
 
1 1 1
ν3 + + + · · · .
13 33 53

Hence the coefficient of xν in Q3 − R2 is of the form

ν 4 O(ν 3 ) = O(ν 7 ).

That is to say
X
Q3 − R 2 = O(ν 7 )xν . (48)

Differentiating (48) and using (42) we obtain


X
P (Q3 − R2 ) = O(ν 8 )xν . (49)

Differentiating this again with respect to x we have


X
A(P 2 − Q)(Q3 − R2 ) + BQ(Q3 − R2 ) = O(ν 9 )xν ,

where A and B are constants. But


12 x 22 x2
 
2
P − Q = −288Φ1,2 (x) = −288 + + ··· ,
(1 − x)2 (1 − x2 )2
On certain arithmetical functions 185

and the coefficient of xν in the right-hand side is a constant multiple of νσ1 (ν). Hence
X X
(P 2 − Q)(Q3 − R2 ) = Oνσ1 (ν)xν O(ν 7 )xν
X
= O(ν 8 ){σ1 (1) + σ1 (2) + · · ·
X
+σ1 (ν)}xν = O(ν 10 )xν ,

and so
X
Q(Q3 − R2 ) = O(ν 10 )xν . (50)

Differentiating this again with respect to x and using arguments similar to those used above,
we deduce
X
R(Q3 − R2 ) = O(ν 12 )xν . (51)

Suppose now that m and n are any two positive integers including zero, and that m + n is
not zero. Then

Qm Rn (Q3 − R2 ) = Q(Q3 − R2 )Qm−1 Rn


X X X
= O(ν 10 )xν { O(ν 3 )xν }m−1 { O(ν 5 )xν }n
X X X
= O(ν 10 )xν O(ν 4m−5 )xν O(ν 6n−1 )xν

X
= O(ν 4m+6n+6 )xν ,

If m is not zero, Similarly we can shew that

Qm Rn (Q3 − R2 ) = R(Q3 − R2 )Qm Rn−1


X
= O(ν 4m+6n+6 )xν ,

if n is not zero. Therefore in any case


X
(Q3 − R2 )Qm Rn = O(ν 4m+6n+6 )xν . (52)

11. Now let r and s be any two positive odd integers including zero. Then, when r + s is
equal to 2,4,6,8 or 12, there are no values of m and n satisfying the relation

4m + 6n = r + s − 10

in (41); consequently in these cases

Fr,s (x) = 0. (53)


186 Paper 18

When r + s = 10, m and n must both be zero, and this result does not apply; but it follows
from (41) and (48) that

X
Fr,s (x) = O(ν 7 )xν . (54)

And when r + s ≥ 14 it follows from (52) that

X
Fr,s (x) = O(ν r+s−4 )xν . (55)

Equating the coefficients of xν in both sides in (53), (54) and (55) we obtain

X Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)


(n) = σr+s+1 (n)
r,s
Γ(r + s + 2) ζ(r + s + 2)

ζ(1 − r) + ζ(1 − s)
+ nσr+s−1 (n) + Er,s (n), (56)
r+s

where

Er,s (n) = 0, r + s = 2, 4, 6, 8, 12;


Er,s (n) = O(n7 ), r + s = 10;
Er,s (n) = O(nr+s−4 ), r + s ≥ 14.

Since σr+s+1 (n) is of order nr+s+1 , it follows that in all cases

X Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)


(n) ∼ σr+s+1 (n). (57)
r,s
Γ(r + s + 2) ζ(r + s + 2)

P
The following table gives the values of r,s (n) when r + s = 2, 4, 6, 8, 12.

TABLE IV
P 5σ3 (n)−6nσ1 (n)
1. 1,1 (n) = 12 .
P 7σ5 (n)−10nσ3 (n)
2. 1,3 (n) = 80 .
On certain arithmetical functions 187

P σ7 (n)
3. 3,3 (n) = 120 .

P 10σ7 (n)−21nσ5 (n)


4. 1,5 (n) = 252 .
P 11σ9 (n)
5. 3,5 (n) = 5040 .

P 11σ9 (n)−30nσ7 (n)


6. 1,7 (n) = 480 .
P σ13 (n)
7. 5,7 (n) = 10080 .

P σ13 (n)
8. 3,9 (n) = 2640 .

P 691σ13 (n)−2730nσ11 (n)


9. 1,11 (n) = 65520 .

12. In this connection it may be interesting to note that

σ1 (1)σ3 (n) + σ1 (3)σ3 (n − 1) + σ1 (5)σ3 (n − 2) + · · ·


1
+σ1 (2n + 1)σ3 (0) = 240 σ5 (2n + 1). (58)

This formula may be deduced from the identity

15 x 35 x2 55 x3
+ + + ···
1 − x 1 − x3 1 − x5
3x2 5x3
 
x
=Q + + + ··· , (59)
1 − x 1 − x3 1 − x5

which can be proved by means of the theory of elliptic functions or by elementary methods.

13. More precise results concerning the order of Er,s (n) can be deduced from the theory
of elliptic functions. Let
x = q2.
Then we have
Q = φ8 (q){1 − (kk′ )2 }




R = φ12 (q)(k′2 − k2 ){1 + 12 (kk′ )2 } , (60)

= φ12 (q){1 + 12 (kk′ )2 } {1 − (2kk′ )2 }
p 

where φ(q) = 1 + 2q + 2q 4 + 2q 9 + · · ·
But, if
1
f (q) = q 24 (1 − q)(1 − q 2 )(1 − q 3 ) · · · ,
188 Paper 18

then we know that


1 1 1

2 6 f (q)= k 12 k′ 3 φ(q) 



1 1 
6 ′
2 f (−q) = (kk ) φ(q) 
12

1 1
(61)
2 3 f (q 2 ) = (kk′ ) 6 φ(q) 



2 1 1


2 3 f (q 4 ) ′
= k 3 k 12 φ(q)

It follows from (41), (60) and (61) that, if r + s is of the form 4m + 2, but not equal to 2
or to 6, then
1
(r+s−6)
f 4(r+s−4) (−q) 4 X f 24n (q 2 )
Fr,s (q 2 ) = 2(r+s−10) 2 Kn , (62)
f (q ) 1
f 24n (−q)

and if r + s is of the form 4m, but not equal to 4, 8 or 12, then


1
(r+s−8)
f 4(r+s−6) (−q) 4 X
f 24n (q 2 )
Fr,s (q 2 ) = 2(r+s−10) 2 {f 8 (q) − 16f 8 (q 4 )} Kn , (63)
f (q ) f 24n (−q)
1

when Kn depends on r and s only. Hence it is easy to see that in all cases Fr,s (q 2 ) can be
expressed as
b  c  d
f 5 (−q) f 5 (q 2 ) f 5 (q) 3
X 
3 a
Ka,b,c,d,e,h,k {f (−q)} f (q)
f 2 (q 2 ) f 2 (−q) f 2 (q 2 )
e
f 5 (q 4 ) 3 4

× 2 2
f (q ) f h (−q)f k (q 2 ), (64)
f (q )

where a, b, c, d, e, h, k are zero or positive integers such that

a + b + c + 2(d + e) = [ 32 (r + s + 2)],
h + k = 2(r + s + 2) − 3[ 32 (r + s + 2)],

and [x] denotes as usual the greatest integer in x. But

12 52 72 112

f (q) = q 24 − q 24 − q 24 + q 24 + ··· 




12 32 52 72

f 3 (q) = q − 3q + 5q − 7q + ···

8 8 8 8 

12 52 72 112
, (65)
f 5 (q)
f 2 (q 2 ) = q 24 − 5q 24 + 7q 24 − 11q + ··· 
24 




f 5 (q 2 ) 12 22 42 52


f 2 (−q) = q 3 − 2q 3 + 4q 3 − 5q 3 + ··· 
On certain arithmetical functions 189

where 1, 2, 4, 5, · · · are the natural numbers without the multiples of 3, and 1, 5, 7, 11, · · ·
are the natural odd numbers without the multiples of 3.
Hence it is easy to see that
1
n− 2 (a+b+c)−d−e Er,s (n)
is not of higher order than the coefficient of q 2n in
1 1 1 1 1 2 1 1 1
φa (q 8 )φb (q 24 )φc (q 3 ){φ(q 24 )φ(q 8 )}d {φ(q 3 )φ(q 2 )}e φh (q 24 )φk (q 12 ),

or the coefficient of q 48n in

φa+d (q 3 )φb+d+h (q)φc (q 8 )φe (q 16 )φe (q 12 )φk (q 2 ).

But the coefficient of q ν in φ2 (q 2 ) cannot exceed that of q ν in φ2 (q), since

φ2 (q) + φ2 (−q) = 2φ2 (q 2 ); (66)

and it is evident that the coefficient of q ν in φ(q 4λ ) cannot exceed that of q ν in φ(q λ ). Hence
it follows that
1 2
n− 2 [ 3 (r+s+2)] Er,s (n)
is not of higher order than the coefficient of q 48n in

φA (q)φB (q 3 )φC (q 2 ),

where A, B, C are zero or positive integers such that

A + B + C = 2(r + s + 2) − 2[ 23 (r + s + 2)],

and C is 0 or 1.
Now, if r + s ≥ 14, we have
A + B + C ≥ 12,
and so
A + B ≥ 11.
Therefore one at least of A and B is greater than 5. But

X
6
φ (q) = O(ν 2 )q ν .∗ (67)
0

Hence it is easily deduced that


X 1
φA (q)φB (q 3 )φC (q 2 ) = O{ν 2 (A+B+C)−1 }q ν . (68)

See §§24–25.
190 Paper 18

It follows that
1 2
Er,s (n) = O{nr+s− 2 [ 3 (r+s−1)] }, (69)

If r + s ≥ 14. We have already shewn in § 11 that, if r + s = 10, then

Er,s (n) = O(n7 ). (70)

This agrees with (69). Thus we see that in all cases


2
Er,s (n) = O{n 3 (r+s+1) }; (71)

and that, if r + s is of the form 6m, then


2 3
Er,s (n) = O{n 3 (r+s+ 4 ) }, (72)

and if of the form 6m + 4, then


2 1
Er,s (n) = O{n 3 (r+s+ 2 ) }. (73)

1
14. I shall now prove that the order of Er,s (n) is not less than that of n 2 (r+s) . In order to
prove this result I shall follow the method used by Messrs Hardy and Littlewood in their
paper “Some problems of Diophantine approximation” (II) ∗ .
Let
q = eπiτ , q ′ = eπiT ,
where
c + dτ
T = ,
a + bτ
and
ad − bc = 1.
Also let
v
V = .
a + bτ
Then we have
√ √
ω veπibvV ϑ1 (v, τ ) = V ϑ1 (V, T ), (74)

where ω is an eighth root of unity and


1
2n 2n
ϑ1 (v, τ ) = 2 sin πv · q 4 Π∞
1 (1 − q )(1 − 2q cos 2πv + q 4n ). (75)

Acta Mathematica, Vol. XXXVII, pp. 193 – 238.
On certain arithmetical functions 191

From (75) we have


∞ 2n
X q (1 + 2 cos 2nπv)
log ϑ1 (v, τ ) = log(2, sin πv) + 41 log q − . (76)
1
n(1 − q 2n )

It follows from (74) and (76) that


∞ 2n
X q (1 + 2 cos 2nπv)
log sin πv + 12 log v + 14 log q + log ω −
n(1 − q 2n )
1
∞ ′ 2n
1
X q (1 + 2 cos 2nπV )
= log sin πV + 2 log V + 14 log q ′ − πibvV − . (77)
1
n(1 − q ′ 2n )

Equating the coefficients of v 8+1 on the two sides of (77), we obtain

1s q 2 2s q 4 3s q 6
 
s+1 1
(a + bτ ) 2 ζ(−s) + 1 − q 2 + 1 − q 4 + 1 − q 6 + · · ·

1s q ′2 2s q ′4 3s q ′6
= 21 ζ(−s) + + + + ···, (78)
1 − q ′2 1 − q ′4 1 − q ′6

provided that s is an odd integer greater than 1. If, in particular,we put s = 3 and s = 5
in (78) we obtain
 3 2
23 q 4 33 q 6
 
4 1 q
(a + bτ ) 1 + 240 + + + ···
1 − q2 1 − q4 1 − q6
 3 ′2
23 q ′4 33 q ′6
 
1 q
= 1 + 240 + + + ··· , (79)
1 − q ′2 1 − q ′4 1 − q ′6

and
 5 2
25 q 4 35 q 6
 
6 1 q
(a + bτ ) 1 − 504 + + + ···
1 − q2 1 − q4 1 − q6
 5 ′2
25 q ′4 35 q ′6
 
1 q
= 1 − 504 + + + · · · . (80)
1 − q ′2 1 − q ′4 1 − q ′6

It follows from (38), (79) and (80) that

(a + bτ )r+s+2 Fr,s (q 2 ) = Fr,s (q ′2 ). (81)

It can easily be seen from (56) and (37) that



X
Fr,s (x) = Er,s (n)xn . (82)
1
192 Paper 18

Hence

X ∞
X
(a + bτ )r+s+2 Er,s (n)q 2n = Er,s (n)q ′2n . (83)
1 1

It is important to observe that


ζ(−r) + ζ(−s) ζ(1 − r) + ζ(1 − s)
Er,s (1) = −
2 r+s
Γ(r + 1)Γ(s + 1) ζ(r + 1)ζ(s + 1)
− 6= 0, (84)
Γ(r + s + 2) ζ(r + s + 2)
if r + s is not equal to 2,4,6,8 or 12. This is easily proved by the help of the equation (39).

15. Now let


τ = u + iy, t = e−πy (u > 0, y > 0, 0 < t < 1),
so that
q = eπiu−πy = teπiu ;
and let us suppose that pn /qn is a convergent to
1 1 1
u= ,
a1 + a2 + a3 + · · ·
so that
ηn = pn−1 qn − pn qn−1 = ±1.
Further, let us suppose that
a = pn , b = −qn ,
c = ηn pn−1 , d = −ηn qn−1 ,
so that
ad − bc = ηn2 = 1.
Furthermore, let

y = 1/(qn qn+1 ),
where

qn+1 = a′n+1 qn + qn−1 ,
and a′n+1 is the complete quotient corresponding to an+1 .
Then we have

| ± 1 − i| 2
|a + bτ | = |pn − qn u − iqn y| = ′ = ′ , (85)
qn+1 qn+1

and
|q ′ | = e−πλ ,
On certain arithmetical functions 193

where
   
c + dτ d 1
λ = I(T ) = I +I −
a + bτ b b(a + bτ )

qn+1
y
= ′ = , (86)
(1/qn+1 )2 + qn2 y 2 2qn

and I(T ) is the imaginary part of T . It follows from (83), (85) and (86) that
∞ ′ r+s+2 X∞
qn+1
X 
| Er,s (n)q 2n | = √ | Er,s (n)q ′2n |
1
2 1
′ r+s+2
qn+1

≥ √ {|Er,s (1)|e−2πλ − |Er,s (2)e−4πλ − |Er,s (3)|e−6πλ − · · ·}. (87)
2
We can choose a number λ0 , depending only on r and s, such that

|Er,s (1)|e−2πλ > 2{|Er,s (2)|e−4πλ + |Er,s (3)|e−6πλ + · · ·}

for λ ≥ λ0 . Let us suppose λ0 > 10. Let us also suppose that the continued fraction for u
satisfies the condition

4λ0 qn > qn+1 > 2λ0 qn (88)

for an infinity of values of n. Then


∞  ′ r+s+2
X
2n qn+1
E (n)q ≥ 1
|E (1)| √ e−4πλ0 > K(qn+1

)r+s+2 , (89)

r,s 2 r,s

1
2

where K depends on r and s only. Also



qn qn+1 = 1/y,
s 
′ 1 π K
qn+1 >√ = >p .
y log(1/t) (1 − t)
It follows that, if u is an irrational number such that the condition (88) is satisfied for an
infinity of values of n, then

X∞ 1
Er,s (n)q 2n > K(1 − t)− 2 (r+s+2) (90)



1

for an infinity of values of t tending to unity. But if we had


1
Er,s (n) = o{n 2 (r+s) }
194 Paper 18

then we should have



X 1
| Er,s (n)q 2n | = o{(1 − t)− 2 (r+s+2) },
1
P
which contradicts (90). It follows that the error term in r,s (n) is not of the form
1
o{n 2 (r+s) }. (91)

The arithmetical function τ (n).

16. We have seen that


Er,s (n) = 0,
P
if r + s is equal to 2,4,6,8, or 12. In these P
cases r,s (n) has a finite expression in terms of
σr+s+1 (n) and σr+s−1 (n). In other cases r,s (n) involves other arithmetical functions as
well. The simplest of these is the function τ (n) defined by

X
τ (n)xn = x{(1 − x)(1 − x2 )(1 − x3 ) · · ·}24 . (92)
1

These cases arise when r + s has one of the values 10, 14, 16, 18, 20 or 24.
Suppose that r + s has one of these values. Then
1728 ∞ n
P
1 Er,s (n)x
(Q3 − R2 )Er,s (1)
is, by (41) and (82), equal to the corresponding one of the functions

1, Q, R, Q2 , QR, Q2 R.

In other words

X ∞
X
n
Er,s (n)x = Er,s (1) τ (n)xn
1 1

( )
2 X xn
1+ nr+s−11 . (93)
ζ(11 − r − s) 1 − xn
1

We thus deduce the formulæ

Er,s (n) = Er,s (1)τ (n), (94)

if r + s = 10; and

σr+s−11 (0)Er,s (n) = Er,s (1){σr+s−11 (0)τ (n)


On certain arithmetical functions 195

+σr+s−11 (1)τ (n − 1) + · · · + σr+s−11 (n − 1)τ (1)}, (95)

if r + s is equal to 14, 16, 18, 20 or 24. It follows from (94) and (95) that, if r + s = r ′ + s′ ,
then

Er,s (n)Er′ ,s′ (1) = Er,s (1)Er′ ,s′ (n), (96)

and in general

Er,s (m)Er′ ,s′ (n) = Er,s (n)Er′ ,s′ (m), (97)

when r + s has one of the values in question. The different cases in which r + s has the
same value are therefore not fundamentally distinct.

17. The values of τ (n) may be calculated as follows: differentiating (92) logarithmically
with respect to x, we obtain


X ∞
X
n
nτ (n)x = P τ (n)xn . (98)
1 1

Equating the coefficients of xn in both sides in (98),we have

24
τ (n) = {σ1 (1)τ (n − 1) + σ1 (2)τ (n − 2) + · · · + σ1 (n − 1)τ (1)}. (99)
1−n

If, instead of starting with (92), we start with


X
τ (n)xn = x(1 − 3x + 5x3 − 7x6 + · · ·)8 ,
1

we can shew that

(n − 1)τ (n) − 3(n − 10)τ (n − 1) + 5(n − 28)τ (n − 3) − 7


(n − 55)τ (n − 6) + · · · to [ 12 {1 + (8n − 7)}] terms = 0,
p
(100)

where the rth term of the sequence 0,1,3,6, . . . is 12 r(r − 1), and the rth term of the
sequence 1,10,28,55, . . . is 1 + 29 r(r − 1). We thus obtain the values of τ (n) in the following
table.

TABLE V
196 Paper 18

n τ (n) n τ (n)
1 +1 16 +987136
2 −24 17 −6905934
3 +252 18 +2727432
4 −1472 19 +10661420
5 +4830 20 −7109760
6 −6048 21 −4219488
7 −16744 22 −12830688
8 +84480 23 +18643272
9 −113643 24 +21288960
10 −115920 25 −25499225
11 +534612 26 +13865712
12 −370944 27 −73279080
13 −577738 28 +24647168
14 +401856 29 +128406630
15 +1217160 30 −29211840

18. Let us consider more particularly the case in which r + s = 10. The order of Er,s (n)
is then the same as that of τ (n). The determination of this order is a problem interesting
in itself. We have proved that Er,s (n), and therefore τ (n), is of the form O(n7 ) and not of
11
the form o(n5 ). There is reason for supposing that τ (n) is of the form O(n 2 +ǫ ) and not of
11
the form o(n 2 ). For it appears that

X τ (n) Y 1
= . (101)
nt p
1− τ (p)p−t + p11−2t
1

This assertion is equivalent to the assertion that, if

n = pa11 pa22 pa33 · · · par r ,

where p1 , p2 , . . . , pr are the prime divisors of n, then


11 sin(1 + a1 )θp1 sin(1 + a2 )θp2 sin(1 + ar )θpr
n− 2 τ (n) = ··· , (102)
sin θp1 sin θp2 sin θpr
where
11
cos θp = 12 p− 2 τ (p).
It would follow that, if n and n′ are prime to each other, we must have

τ (nn′ ) = τ (n)τ (n′ ). (103)

Let us suppose that (102) is true, and also that (as appears to be highly probable)

{2τ (p)}2 ≤ p11 , (104)


On certain arithmetical functions 197

so that θp is real. Then it follows from (102) that


11
n− 2 |τ (n)| ≤ (1 + a1 )(1 + a2 ) · · · (1 + ar ),

that is to say
11
|τ (n)| ≤ n 2 d(n), (105)

where d(n) denotes the number of divisors of n.


Now let us suppose that n = pa , so that
11 sin(1 + a)θp
n− 2 τ (n) = .
sin θp
Then we can choose a as large as we please and such that

sin(1 + a)θp
≥ 1.
sin θp

Hence
11
|τ (n)| ≥ n 2 (106)

for an infinity of values of n.

19. It should be observed that precisely similar questions arise with regard to the arith-
metical function Ψ(n) defined by

X
Ψ(n)xn = f a1 (xc1 )f a2 (xc2 ) · · · f ar (xcr ), (107)
0

where
1
f (x) = x 24 (1 − x)(1 − x2 )(1 − x3 ) · · · ,
the a’s and c’s are integers, the latter being positive,
1
24 (a1 c1 − a2 c2 + · · · + ar cr )

is equal to 0 or 1, and  
a1 a2 ar
l + + ··· + ,
c1 c2 cr
where l is the least common multiple of c1 , c2 , . . . , cr , is equal to 0 or to a divisor of 24.
The arithmetical functions χ(n), P (n), χ4 (n), Ω(n) and Θ(n), studied by Dr. Glaisher in
the Quarterly Journal, Vols. XXXVI-XXXVIII, are of this type. Thus

X
χ(n)xn = f 6 (x4 ),
1
198 Paper 18


X
P (n)xn = f 4 (x2 )f 4 (x4 ),
1

X
χ4 (n)xn = f 4 (x)f 2 (x2 )f 4 (x4 ),
1

X
Ω(n)xn = f 12 (x2 ),
1

X
Θ(n)xn = f 8 (x)f 8 (x2 ).
1

20. The results (101) and (104) may be written as



X Er,s (n) Y 1
= Er,s (1) , (108)
nt p
1 − 2cp p−t + pr+s+1−2t
1

where
c2p ≤ pr+s+1 ,
and
2cp Er,s (1) = Er,s (p).
It seems probable that the result (108) is true not only for r + s = 10 but also when r + s
is equal to 14, 16, 18, 20 or 24, and that

Er,s (n) 1
(r+s+1)
Er,s (1) ≤ n 2 d(n) (109)

for all values of n, and



Er,s (n) 1
(r+s+1)
Er,s (1) ≥ n 2 (110)

for an infinity of values of n. If this be so, then


1 1
Er,s (n) = O{n 2 (r+s+1+ǫ) }, Er,s (n) 6= o{n 2 (r+s+1) }. (111)

And it seems very likely that these equations hold generally, whenever r and s are positive
odd integers.

21. It is of some interest to see what confirmation of these conjectures can be found from
a study of the coefficients in the expansion of

X
24/α 48/α 72/α a
x{(1 − x )(1 − x )(1 − x ) · · ·} = Ψα (n)xn ,
1
On certain arithmetical functions 199

where α is a divisor of 24. When α = 1 and α = 3 we know the actual value of Ψα (n). For
we have

2 2 2 2 2 2
X
Ψ1 (n)xn = x1 − x5 − x7 + x11 + x13 − x17 − · · · , (112)
1

where 1, 5, 7, 11, . . . the natural odd numbers without the multiples of 3; and

2 2 2 2
X
Ψ3 (n)xn = x1 − 3x3 + 5x5 − 7x7 + · · · (113)
1

The corresponding Dirichlet’s series are



X Ψ1 (n) 1
= , (114)
ns (1 + 5−2s )(1 + 7−2s )(1 − 11−2s )(1 − 13−2s ) · · ·
1

where 5, 7, 11, 13, . . . are the primes greater than 3, those of the form 12n ± 5 having the
plus sign and those of the form 12n ± 1 the minus sign; and

X Ψ3 (n) 1
= (115)
ns (1 + 31−2s )(1 − 51−2s )(1 + 71−2s )(1 + 111−2s ) · · ·
1

where 3, 5, 7, 11, . . . are the odd primes, those of the form 4n − 1 having the plus sign and
those of the form 4n + 1 the minus sign.
It is easy to see that

|Ψ1 (n)| ≤ 1, |Ψ3 (n)| ≤ n (116)

for all values of n, and



|Ψ1 (n)| = 1, |Ψ3 (n)| = n (117)

for an infinity of values of n.


The next simplest case is that in which α = 2. In this case it appears that

X Ψ2 (n)
= Π1 Π2 , (118)
1
ns

where
1
Π1 = ,
(1 + 5−2s )(1 − 7−2s )(1 − 11−2s )(1 + 17−2s ) · · ·
5, 7, 11, . . . being the primes of the forms 12n − 1 and 12n ± 5, those of the form 12n + 5
having the plus sign and the rest the minus sign; and
1
Π2 = ,
(1 + 13−s )2 (1 − 37−s )2 (1 − 61−s )2 (1 + 73−s )2 · · ·
200 Paper 18

13, 37, 61, . . . being the primes of the form 12n + 1, those of the form m2 + (6n − 3)2
having the plus sign and those of the form m2 + (6n)2 the minus sign.
This is equivalent to the assertion that if

n = (5a5 · 7a7 · 11a11 · 17a17 · · ·)2 13a13 · 37a37 · 61a61 · 73a73 · · · ,

where ap is zero or a positive integer, then

Ψ2 (n) = (−1)a5 +a13 +a17 +a29 +a41 +··· (1 + a13 )(1 + a37 )(1 + a61 ) · · · , (119)

where 5, 13, 17, 29, . . . are the primes of the form 4n + 1, excluding those of the form
m2 + (6n)2 ; and that otherwise

Ψ2 (n) = 0. (120)

It follows that

|Ψ2 (n)| ≤ d(n) (121)

for all values of n, and

|Ψ2 (n)| ≥ 1 (122)

for an infinity of values of n. These results are easily proved to be actually true.

22. I have investigated also the cases in which α has one of the values 4, 6, 8 or 12. Thus
for example, when α = 6, I find

X Ψ6 (n)
= Π1 Π2 , ∗ (123)
1
ns

where
1
Π1 = ,
(1 − 32−2s )(1 − 72−2s )(1 − 112−2s ) · · ·
3, 7, 11, . . . being the primes of the form 4n − 1; and
1
Π2 = ,
(1 − 2c5 · 5−s + 52−2s )(1 − 2c13 · 13−s + 132−2s ) · · ·

5, 13, 17, . . . being the primes of the form 4n + 1, and cp = u2 − (2v)2 , where u and v are the
unique pair of positive integers for which p = u2 + (2v)2 . This is equivalent to the assertion
that if
n = (3a3 · 7a7 · 11a11 · · ·)2 · 5a5 · 13a13 · 17a17 · · · ,

Ψ6 (n) is Dr. Glaisher’s λ(n).
On certain arithmetical functions 201

then
Ψ6 (n) sin(1 + a5 )θ5 sin(1 + a13 )θ13 sin(1 + a17 )θ17
= · · ···, (124)
n sin θ5 sin θ13 sin θ17
where
u
tan 12 θp =
(0 < θp < π),
2v
and that otherwise Ψ6 (n) = 0. From these results it would follow that

|Ψ6 (n)| ≤ nd(n) (125)

for all values of n, and

|Ψ6 (n)| ≥ n (126)

for an infinity of values of n. What can actually be proved to be true is that

|Ψ6 (n) < 2nd(n)

for all values of n, and


|Ψ6 (n)| ≥ n
for an infinity of values of n.

23. In the case in which α = 4 I find that, if


n = (5a5 · 11a11 · 17a17 · · ·)2 · 7a7 · 13a13 · 19a19 · · · ,

where 5, 11, 17, . . . are the primes of the form 6m − 1 and 7, 13, 19, . . . are those of the
form 6m + 1, then
Ψ4 (n) sin(1 + a7 )θ7 sin(1 + a13 )θ13
√ = (−1)a5 +a11 +a17 +··· · ···, (127)
n sin θ7 sin θ13
where √
u 3
tan θp = (0 < θp < π),
1 ± 3v
and u and v are the unique pair of positive integers for which p = 3u2 + (1 ± 3v)2 ; and that
Ψ4 (n) = 0 for other values.
In the case in which α = 8 I find that, if

n = (2a2 · 5a5 · 11a11 · · ·)2 · 7a7 · 13a13 · 19a19 · · · ,

where 2, 5, 11, . . . are the primes of the form 3m − 1 and 7, 13, 19, . . . are those of the
form 6m + 1, then
Ψ8 (n) sin 3(1 + a7 )θ7 sin 3(1 + a13 )θ13
√ = (−1)a2 +a5 +a11 +··· · ···, (128)
n n sin 3θ7 sin 3θ13
202 Paper 18

where θp is the same as in (127); and that Ψ8 (n) = 0 for other values.
The case in which α = 12 will be considered in § 28.
In short, such evidence as I have been able to find, while not conclusive, points to the truth
of the results conjectured in § 18.

24. Analysis similar to that of the preceding sections may be applied to some interesting
arithmetical functions of a different kind. Let

X
φs (q) = 1 + 2 rs (n)q n , (129)
1

where
φ(q) = 1 + 2q + 2q 4 + 2q 9 + · · · ,
so that rs (n) is the number of representations of n as the sum of s squares. Further let

q3 q5
 
X
n q
δ2 (n)q = 2 − + − ···
1 − q 1 − q3 1 − q5
1

q2 q3
 
q
= 2 + + + ··· ; (130)
1 + q2 1 + q4 1 + q6


1s−1 q 2s−1 q 2 3s−1 q 3
X  
s n
(2 − 1)Bs δ2s (n)q = s + + + ··· , (131)
1+q 1 − q2 1 + q3
1

when s is a multiple of 4;

1s−1 q 2s−1 q 2 3s−1 q 3
X  
s n
(2 − 1)Bs δ2s (n)q = s + + + ··· , (132)
1
1−q 1 + q2 1 − q3

when s + 2 is a multiple of 4;

1s−1 q 2s−1 q 2 3s−1 q 3
X  
n s
Es δ2s (n)q = 2 + + + ···
1 + q2 1 + q4 1 + q6
1
 s−1 s−1 3
5s−1 q 5

1 q3 q
+2 + − ··· , (133)
1 − q 1 − q3 1 − q5
when s − 1 is a multiple of 4;

1s−1 q 2s−1 q 2 3s−1 q 3
X  
n s
Es δ2s (n)q = 2 + + + ···
1
1 + q2 1 + q4 1 + q6
 s−1
1 q 3s−1 q 3 5s−1 q 5

−2 − + − ··· , (134)
1−q 1 − q3 1 − q5
On certain arithmetical functions 203

when s + 1 is a multiple of 4. In these formulæ

B2 = 61 , B4 = 1
30 , B6 = 1
42 , B8 = 1
30 , B10 = 5
66 , . . .

are Bernoulli’s numbers, and

E1 = 1, E3 = 1, E5 = 5, E7 = 61, E9 = 1385, . . .

are Euler’s numbers. Then δ2s (n) is in all cases an arithmetical function depending on the
real divisors of n; thus, for example, when s + 2 is a multiple of 4, we have

(2s − 1)Bs δ2s (n) = s{σs−1 (n) − 2s σs−1 ( 14 n)}, (135)

where σs (x) should be considered as equal to zero if x is not an integer.


Now let

r2s (n) = δ2s (n) + e2s (n). (136)

Then I can prove (see § 26) that

e2s (n) = 0 (137)

if s = 1, 2, 3, 4 and that
1 2
e2s (n) = O(ns−1− 2 [ 3 s]+ǫ) (138)

for all positive integral values of s. But it is easy to see that, if s ≥ 3, then

Hns−1 < δ2s (n) < Kns−1 , (139)

where H and K are positive constants. It follows that

r2s (n) ∼ δ2s (n) (140)

for all positive integral values of s.


It appears probable, from the empirical results I obtain at the end of this paper, that
1
e2s (n) = O{n 2 (s−1)+ǫ } (141)

for all positive integral values of s; and that


1
e2s (n) 6= o{n 2 (s−1) } (142)

if s ≥ 5. But all that I can actually prove is that


1 2
e2s (n) = O(ns−1− 2 [ 3 s] ) (143)
204 Paper 18

if s ≥ 9 and that
1
e2s (n) 6= o(n 2 s−1 ) (144)

if s ≥ 5.

25. Let

X ∞
X
f2s (q) = e2s (n)q n = {r2s (n) − δ2s (n)}q n . (145)
1 1

Then it can be shewn by the theory of elliptic functions that


X
f2s (q) = φ2s (q) Kn (kk′ )2n , (146)
1
1≤n≤ 4 (s−1)

that is to say that

f 4s (−q) X f 24n (q 2 )
f2s (q) = Kn , (147)
f 2s (q 2 ) 1
f 24n (−q)
1≤n≤ 4 (s−1)

where φ(q) and f (q) are the same as in § 13. We thus obtain the results contained in the
following table.

TABLE VI

1. f2 (q) = 0, f4 (q) = 0, f6 (q) = 0, f8 (q) = 0.


f 14 (q 2 )
2. 5f10 (q) = 16 f 4 (−q) , f12 (q) = 8f 12 (q 2 ).
3. 61f14 (q) = 728f 4 (−q)f 10 (q 2 ), 17f16 (q) = 256f 8 (−q)f 8 (q 2 ).
30 2
4. 1385f18 (q) = 24416f 12 (−q)f 6 (q 2 ) − 256 ff12 (−q)
(q )
.
28 2
5. 31f20 (q) = 616f 16 (−q)f 4 (q 2 ) − 128 ff 8 (−q)
(q )
.
26 2
6. 50521f22 (q) = 1103272f 20 (−q)f 2 (q 2 ) − 821888 ff 4 (−q)
(q )
.
7. 691f24 (q) = 16576f 24 (−q) − 32768f 24 (q 2 ).
It follows from the last formula of Table VI that
691
64 e24 (n) = (−1)n−1 259τ (n) − 512τ ( 12 n), (148)

where τ (n) is the same as in § 16, and τ (x) should be considered as equal to zero if x is
not an integer.
On certain arithmetical functions 205

Results equivalent to 1,2,3,4 of Table VI were given by Dr. Glaisher in the Quarterly
Journal, Vol. XXXVIII. The arithmetical functions called by him

χ4 (n), Ω(n), W (n), Θ(n), U (n)

are the coefficients of q n in

f 14 (q 2 ) 12 2 4
, f (q ), f (−q)f 10 (q 2 ), f 8 (q)f 8 (q 2 ), f 12 (−q)f 6 (q 2 ).
f 4 (−q)

He gave reduction formulæ for these functions and observed how the functions which I call
e10 (n), e12 (n) and e16 (n) can be defined by means of the complex divisors of n. It is very
likely that τ (n) is also capable of such a definition.

26. Now let us consider the order of e2s (n). It is easy to see from (147) that f2s (q) can be
expressed in the form
b  c
f 5 (−q) f 5 (q 2 )
X 
3 a
Ka,b,c,h,k {f (−q)} f h (−q)f k (q 2 ), (149)
f 2 (q 2 ) f 2 (−q)

where a, b, c, h, k are zero or positive integers, such that

a + b + c = [ 23 s], h + k = 2s − 3[ 32 s].

Proceeding as in § 13 we can easily shew that


1 2
n− 2 [ 3 s] e2s (n)

cannot be of higher order than the coefficient of q 24n in

φA (q)φB (q 3 )φC (q 2 ), (150)

where C is 0 or 1 and
A + B + C = 2s − 2[ 32 s].
Now, if s ≥ 5, A + B + C ≥ 4; and so A + B ≥ 3. Hence one at least of A and B is greater
than 1. But we know that X
φ2 (q) = O(ν ǫ )q ν .

It follows that the coefficient of q 24n in (150) is of order not exceeding


1
n 2 (A+B+C)−1+ǫ .

Thus
1 2
e2s (n) = O(ns−1− 2 [ 3 s]+ǫ ) (151)
206 Paper 18

for all positive integral values of s.

27. When s ≥ 9 we can obtain a slightly more precise result.


If s ≥ 16 we have A + B + C ≥ 12; and so A + B ≥ 11. Hence one at least of A and B is
greater than 5. But X
φ6 (q) = O(ν 2 )q ν .

It follows that the coefficient of q 24n in (150) is of order not exceeding


1
n 2 (A+B+C)−1 ,

or that
1 2
e2s (n) = O(ns−1− 2 [ 3 s] ), (152)

if s ≥ 16. We can easily shew that (152) is true when 9 ≤ s ≤ 16 considering all the cases
separately, using the identities.

f 12 (−q)f 6 (q 2 ) = {f 3 (−q)}4 , {f 3 (q 2 )}2 ,


 5 2 6
f 30 (q 2 ) f (q )
= ,
f 12 (−q) f 2 (−q)
 5 4  5 2 2
f (−q) f (q )
f 16 (−q)f 4 (q 2 ) = f 2 (q 2 ),
f 2 (q 2 ) f 2 (−q)
 5 2 4
f 28 (q 2 ) f (q )
8
= {f 3 (q 2 )}2 f 2 (q 2 ), · · · ,
f (−q) f 2 (−q)

and proceeding as in the previous two sections.


The argument of §§ 14-15 may also be applied to the function e2s (n). We find that
1
e2s (n) 6= o(n 2 s−1 ). (153)

I leave the proof to the reader.

28. There is reason to suppose that


1

e2s (n) = O{n 2 (s−1+ǫ) } 
1
, (154)
e2s (n) 6= o{n 2 (s−1) } 

if s ≥ 5. I find, for example, that



X e10 (n) e10 (1)
= Π1 Π2 , (155)
ns 1 + 22−s
1
On certain arithmetical functions 207

where
1
Π1 = ,
(1 − 34−2s )(1 − 74−2s )(1 − 114−2s ) · · ·
3, 7, 11, . . . being the primes of the form 4n − 1, and
1
Π2 = ,
(1 − 2c5 · 5−s + 54−2s )(1 − 2c13 · 13−s + 134−2s ) · · ·
5, 13, 17, . . . being the primes of the form 4n + 1, and

cp = u2 − (4v)2 ,

where u and v are the unique pair of positive integers satisfying the equation

u2 + (4v)2 = p2 .

The equation (155) is equivalent to the assertion that, if

n = (3a3 · 7a7 · 11a11 · · ·)2 · 2a2 · 5a5 · 13a13 · · · ,

where ap is zero or a positive integer, then

e10 (n) sin 4(1 + a5 )θ5 sin 4(1 + a13 )θ13


2
= (−1)a2 · ···, (156)
n e10 (1) sin 4θ5 sin 4θ13
where
u
tan θp =(0 < θp < 12 π),
v
u and v being integers satisfying the equation u2 + v 2 = p; and e10 (n) = 0 otherwise. If
this is true then we should have

e10 (n) 2
e10 (1) ≤ n d(n) (157)

for all values of n, and



e10 (n) 2
e10 (1) ≥ n (158)

for an infinity of values of n. In this case we can prove that, if n is the square of a prime
of the form 4m − 1, then
e10 (n) 2
e10 (1) = n .

Similarly I find that


∞ Y 
X e12 (n) 1
= e12 (1) , (159)
1
ns p
1 + 2cp · p−s + p5−2s
208 Paper 18

p being an odd prime and c2p ≤ p5 . From this it would follow that

e12 (n) 5
e12 (1) ≤ n 2 d(n) (160)

for all values of n, and



e12 (n) 5
e12 (1) ≥ n 2 (161)

for an infinity of values of n.


Finally I find that
∞  
X e16 (n) e16 (1) Y 1
= , (162)
1
ns 1 + 23−s p 1 + 2cp · p−s + p7−2s

p being an odd prime and c2p ≤ p7 . From this it would follow that

e16 (n) 7
e16 (1) ≤ n 2 d(n) (163)

for all values of n, and



e16 (n) 7
e16 (1) ≥ n 2 (164)

for an infinity of values of n.


In the case in which 2s = 24 we have
691
64 e24 (n) = (−1)n−1 259τ (n) − 512τ ( 21 n).

I have already stated the reasons for supposing that


11
|τ (n)| ≤ n 2 d(n)

for all values of n, and


11
|τ (n)| ≥ n 2
for an infinity of values of n.
A series for Euler’s constant γ

Messenger of Mathematics, XLVI, 1917, 73 – 80

1. In a paper recently published in this Journal (Vol.XLIV, pp. 1 – 10), Dr. Glaisher
proves a number of formulæ of the type
 
S3 S5 S7
γ =1−2 + + + ··· ,
3·4 5·6 7·8

where
Sn = 1−n + 2−n + 3−n + 4−n + · · · ,
and conjectures the existence of a general formula

γ = λr − (r + 1)(r + 2) · · · (2r)
 
S3 S5
× + + ··· ,
3(r + 3)(r + 4) · · · (2r + 2) 5(r + 5)(r + 6) · · · (2r + 4)
where λr is a rational number. I propose now to prove the general formula of which Dr.
Glaisher’s are particular cases: this formula is itself a particular case of still more general
formulæ.

2. Let r and t be any two positive numbers. Then


Z1 Z1
r−1 t−1
x (1 − x) log Γ(1 − x) dx = xt−1 (1 − x)r−1 log Γ(x) dx
0 0

Z1 Z1
t−1 r−1
= x (1 − x) log Γ(1 + x) dx − xt−1 (1 − x)r−1 log x dx (1)
0 0

But
Z1
xr−1 (1 − x)t−1 log Γ(1 − x) dx
0

Z1
x2 x3
 
r−1 t−1
= x (1 − x) γx + S2 + S3 + · · · dx
2 3
0

Γ(1 + r)Γ(t) Γ(2 + r)Γ(t) S2 Γ(3 + r)Γ(t) S3


= γ+ + + ··· (2)
Γ(1 + r + t) Γ(2 + r + t) 2 Γ(3 + r + t) 3
210 Paper 19

Similarly
Z1
xt−1 (1 − x)r−1 log Γ(1 + x) dx
0
Γ(1 + t)Γ(r) Γ(2 + t)Γ(r) S2 Γ(3 + t)Γ(r) S3
=− γ+ − + ··· (2′ )
Γ(1 + r + t) Γ(2 + r + t) 2 Γ(3 + r + t) 3
And also
Z1 Z1  
t−1 r−1 d t−1 r−1 d Γ(t)Γ(r)
x (1 − x) log x dx = x (1 − x) dx =
dt dt Γ(r + t)
0 0
 
Γ(r)Γ(t) Γ′ (t) Γ′ (r + t)
= −
Γ(r + t) Γ(t) Γ(r + t)
Z1
Γ(r)Γ(t) 1 − xr
= − xt−1 dx (3)
Γ(r + t) 1−x
0

It follows from (1)–(3) that, if r and t are positive, then


r r(r + 1) r(r + 1)(r + 2)
γ+ S2 + S3 + · · ·
1(r + t) 2(r + t)(r + t + 1) 3(r + t)(r + t + 1)(r + t + 2)
t t(t + 1) t(t + 1)(t + 2)
+ γ− S2 + S3 − · · ·
1(r + t) 2(r + t)(r + t + 1) 3(r + t)(r + t + 1)(r + t + 2)
Z1
xt−1 (1 − xr )
= dx (4)
1−x
0

Now, interchanging r and t in (4), and taking the sum and the difference of the two results,
we see that, if r and t are positive, then
r+t r(r + 1)(r + 2) + t(t + 1)(t + 2)
γ+ S3 + · · ·
1(r + t) 3(r + t)(r + t + 1)(r + t + 2)
Z1
1 xr−1 + xt−1 − 2xr+t−1
= dx; (5)
2 1−x
0

and
r(r + 1) − t(t + 1) r(r + 1)(r + 2)(r + 3) − t(t + 1)(t + 2)(t + 3)
S2 + S4 + · · ·
2(r + t)(r + t + 1) 4(r + t)(r + t + 1)(r + t + 2)(r + t + 3)
Z1
1 xt−1 − xr−1
= dx. (6)
2 1−x
0
A series for Euler’s constant γ 211

The right-hand sides of (5) and (6) can be expressed in finite terms if r and t are rational.
If, in particular, r and t are integers, then

Z1
xr−1 + xt−1 − 2xr+t−1 1 1 1 1
dx = + + + ··· +
1−x r r+1 r+2 r+t−1
0
1 1 1 1
+ + + + ··· + ;
t t+1 t+2 r+t−1
and
Z1
xt−1 − xr−1
 
1 1 1 1
dx = 1+ + + + ··· +
1−x 2 3 4 r−1
0
 
1 1 1 1
− 1 + + + + ··· + .
2 3 4 t−1

3. Let us now suppose that t = r in (5). Then it is clear that

(r + 1)(r + 2) (r + 1)(r + 2)(r + 3)(r + 4)


γ+ S3 + S5 + · · ·
3(2r + 1)(2r + 2) 5(2r + 1)(2r + 2)(2r + 3)(2r + 4)
Z1 Z1
xr−1 (1 − xr ) 1 + x2r−1
= dx = dx, (7)
1−x 1+x
0 0

if r > 0. If we suppose, in (7), that r is an integer, we obtain the formula conjectured by


Dr Glaisher, the value of λr being

Z1
1 + x2r−1 1 1 1 1
dx = 1 − + − + · · · + .
1+x 2 3 4 2r − 1
0

Again, dividing both sides in (6) by r − t and making t → r, we see that, if r > 0, then
 
r+1 1 1
+ S2
2(2r + 1) r r+1
 
(r + 1)(r + 2)(r + 3) 1 1 1 1
+ + + + S4 + · · ·
4(2r + 1)(2r + 2)(2r + 3) r r+1 r+2 r+3
Z1
xr−1 log x 1 1 1 1
=− dx = 2 + 2
+ 2
+ + (8)
1−x r (r + 1) (r + 2) (r + 3)2
0
212 Paper 19

Thus for example we have


π2 S2 S4 S6
= (1 + 12 ) + (1 + 1
2 + 1
3 + 14 ) (1 + 1
2 + 1
3 + 1
4 + 1
5 + 61 ) + ···.
12 2···3 4·5 6·7

4. If we start with the integral


Z1  x
xr−1 (1 − x)t−1 log Γ 1 − dx,
2
0

and proceed as in § 2, we can shew that, if r and t are positive, then


r r(r + 1) r(r + 1)(r + 2)
S′ + S′ + S′ + · · ·
1(r + t) 1 2(r + t)(r + t + 1) 2 3(r + t)(t + t + 1)(r + t + 2) 3
t t(t + 1) t(t + 1)(t + 2)
− S1′ + S2′ − S′ + · · ·
1(r + t) 2(r + t)(r + t + 1) 3(r + t)(r + t + 1)(r + t + 2) 3
Z1
xt−1 (1 − xr ) π
= dx − log , (9)
1−x 2
0

where
Sn′ = 1−n − 2−n + 3−n − 4−n + · · ·
From (9) we can easily deduce that, if r and t are positive, then
r(r + 1) + t(t + 1) ′
S
2(r + t)(r + t + 1) 2
r(r + 1)(r + 2)(r + 3) + t(t + 1)(t + 2)(t + 3) ′
+ S4 + · · ·
4(r + t)(r + t + 1)(r + t + 2)(r + t + 3)
Z1
1 xr−1 + xt−1 − 2xr+t−1 π
= dx − log ; (10)
2 1−x 2
0

and
Z1
r−t ′ r(r + 1)(r + 2) − t(t + 1)(t + 2) ′ 1 xt−1 − xr−1
S1 + S3 + · · · = dx. (11)
1(r + t) 3(r + t)(r + t + 1)(r + t + 2) 2 1−x
0

As particular cases of (10) and (11), we have


Z1
π r+1 (r + 1)(r + 2)(r + 3) 1 + x2r−1
log + S2′ + S′ + · · · = dx; (12)
2 2(2r + 1) 4(2r + 1)(2r + 2)(2r + 3) 4 1+x
0
A series for Euler’s constant γ 213

and
 
1 ′ (r + 1)(r + 2) 1 1 1
S1 + + + S3′
r 3(2r + 1)(2r + 2) r r+1 r+2
(r + 1)(r + 2)(r + 3)(r + 4)
+
5(2r + 1)(2r + 2)(2r + 3)(2r + 4)
 
1 1 1 1 1
× + + + + S5′ + · · ·
r r+1 r+2 r+3 r+4
1 1 1
= + + + ···, (13)
r 2 (r + 1)2 (r + 2)2

provided that r > 0. Thus for example we have


 
π S2′ S4′ S6′
1 = log + 2 + + + ··· ;
2 2.3 4.5 6.7
π2 S′ S′ 1 1 S′ 1 1 1 1
= 1 + 3 (1 + + ) + 5 (1 + + + + ) + · · · .
12 1.2 3.4 2 3 5.6 2 3 4 5

5. The preceding results may be generalised as follows. Let ζ(s, x) denote the function
represented by the series

x−s + (x + 1)−s + (x + 2)−s + (x + 3)−s + · · · (x > 0)

and its analytical continuations, so that ζ(s, 1) = ζ(s) and ζ(s, 12 ) = (2s − 1)ζ(s), ζ(s) being
the Riemann ζ-function. Then

Z1 Z1
r−1 t−1
x (1 − x) ζ(s, 1 − x) dx = xt−1 (1 − x)r−1 ζ(s, x) dx
0 0

Z1 Z1
t−1 r−1
= x (1 − x) ζ(s, 1 + x) dx + xt−s−1 (1 − x)r−1 dx, (14)
0 0

provided that r and t are positive. But we know that, if |x| < 1, then

s s(s + 1)
ζ(s, 1 − x) = ζ(s) + ζ(s + 1)x + ζ(s + 2)x2 + · · · ; (15)
1! 2!
and that
Z1
Γ(t − s)Γ(r)
xt−s−1 (1 − x)r−1 dx = , (16)
Γ(r − s + t)
0
214 Paper 19

provided that t > s. It follows from (14)–(16) that, if r and t are positive and t > s, then
 
s r s(s + 1) r(r + 1)
ζ(s) + ζ(s + 1) + ζ(s + 2) + · · ·
1! r + t 2! (r + t)(r + t + 1)
 
s t s(s + 1) t(t + 1)
− ζ(s) − ζ(s + 1) + ζ(s + 2) − · · ·
1! r + t 2! (r + t)(r + t + 1)
Γ(r + t)Γ(t − s)
= . (17)
Γ(t)Γ(r − s + t)

As particular cases of (17), we have

s r+t s(s + 1)(s + 2) r(r + 1)(r + 2) + t(t + 1)(t + 2)


ζ(s + 1) + ζ(s + 3) + · · ·
1! r + t 3! (r + t)(r + t + 1)(r + t + 2)
 
1 Γ(r + t) Γ(t − s) Γ(r − s)
= + , (18)
2 Γ(r − s + t) Γ(t) Γ(r)

and

s(s + 1) r(r + 1) − t(t + 1)


ζ(s + 2)
2! (r + t)(r + t + 1)
s(s + 1)(s + 2)(s + 3) r(r + 1)(r + 2)(r + 3) − t(t + 1)(t + 2)(t + 3)
+ ζ(s + 4) + · · ·
4! (r + t)(r + t + 1)(r + t + 2)(r + t + 3)
 
1 Γ(r + t) Γ(t − s) Γ(r − s)
= − , (19)
2 Γ(r − s + t) Γ(t) Γ(r)

provided that r and t are positive and greater than s. From (18) and (19) we deduce that,
if r is positive and greater than s, then

s r s(s + 1)(s + 2) r(r + 1)(r + 2)


ζ(s + 1) + ζ(s + 3) + · · ·
1! 2r 3! 2r(2r + 1)(2r + 2)
1 Γ(2r)Γ(r − s)
= , (20)
2 Γ(r)Γ(2r − s)

and
 
s(s + 1) r(r + 1) 1 1
+ ζ(s + 2)
2! 2r(2r + 1) r r+1
s(s + 1)(s + 2)(s + 3) r(r + 1)(r + 2)(r + 3)
+
4! 2r(2r + 1)(2r + 2)(2r + 3)
 
1 1 1 1
× + + + ζ(s + 4) + · · ·
r r+1 r+2 r+3
A series for Euler’s constant γ 215

Z1
1 Γ(2r)Γ(r − s) xr−s−1 (1 − xs )
= dx. (21)
2 Γ(r)Γ(2r − s) 1−x
0

6. If we start with the integral


Z1  x
xr−1 (1 − x)t−1 ζ s, 1 − dx,
2
0

and proceed as in § 5, we can shew that, if r and t are positive and t > s, then
s r s(s + 1) r(r + 1)
ζ1 (s) + ζ1 (s + 1) ζ1 (s + 2) + · · ·
1! r + t 2! (r + t)(r + t + 1)
s t s(s + 1) t(t + 1)
+ζ1 (s) − ζ1 (s + 1) + ζ1 (s + 2) − · · ·
1! r + t 2! (r + t)(r + t + 1)
Γ(r + t)Γ(t − s)
= , (22)
Γ(t)Γ(r − s + t)

where ζ1 (s) is the function represented by the series

1−s − 2−s + 3−s − 4−s + · · ·

and its analytical continuations. From (22) we deduce that, if r and t are positive and
greater than s, then
s(s + 1) r(r + 1) + t(t + 1)
(1 + 1)ζ1 (s) + ζ1 (s + 2) + · · ·
2! (r + t)(r + t + 1)
 
1 Γ(r + t) Γ(t − s) Γ(r − s)
= + ; (23)
2 Γ(r − s + t) Γ(t) Γ(r)
and
s r−t
ζ1 (s + 1)
1! r + t
s(s + 1)(s + 2) r(r + 1)(r + 2) − t(t + 1)(t + 2)
+ ζ1 (s + 3) + · · ·
3! (r + t)(r + t + 1)(r + t + 2)
 
1 Γ(r + t) Γ(t − s) Γ(r − s)
= − . (24)
2 Γ(r − s + t) Γ(t) Γ(r)

As particular cases of (23) and (24), we have

s(s + 1) r(r + 1) 1 Γ(2r)Γ(r − s)


ζ1 (s) + ζ1 (s + 2) + · · · = , (25)
2! 2r(2r + 1) 2 Γ(r)Γ(2r − s)
216 Paper 19

and
 
s r 1 s(s + 1)(s + 2) r(r + 1)(r + 2) 1 1 1
ζ1 (s + 1) + + + ζ1 (s + 3) + · · ·
1! 2r r 3! 2r(2r + 1)(2r + 2) r r+1 r+2
Z1
1 Γ(2r)Γ(r − s) sr−s−1 (1 − xs )
= dx, (26)
2 Γ(r)Γ(2r − s) 1−x
0

provided that r is positive and greater than s.


On the expression of a number in the form
ax2 + by 2 + cz 2 + du2

Proceedings of the Cambridge Philosophical Society, XIX, 1917, 11 – 21

1. It is well known that all positive integers can be expressed as the sum of four squares.
This naturally suggests the question: For what positive integral values of a, b, c, d, can all
positive integers be expressed in the form

ax2 + by 2 + cz 2 + du2 ? (1.1)

I prove in this paper that there are only 55 sets of values of a, b, c, d for which this is true.
The more general problem of finding all sets of values of a, b, c, d for which all integers with
a finite number of exceptions can be expressed in the form (1.1), is much more difficult and
interesting. I have considered only very special cases of this problem, with two variables
instead of four; namely, the cases in which (1.1) has one of the special forms

a(x2 + y 2 + z 2 ) + bu2 , (1.2)

and
a(x2 + y 2 ) + b(z 2 + u2 ). (1.3)
These two cases are comparatively easy to discuss. In this paper I give the discussion of
(1.2) only, reserving that of (1.3) for another paper.

2. Let us begin with the first problem. We can suppose, without loss of generality, that

a ≤ b ≤ c ≤ d. (2.1)

if a > 1, then 1 cannot be expressed in the form (1.1); and so

a=1 (2.2)

If b > 2, then 2 is an exception; and so

1 ≤ b ≤ 2. (2.3)

We have therefore only to consider the two cases in which (1.1) has one or other of the
forms
x2 + y 2 + cz 2 + du2 , x2 + 2y 2 + cz 2 + du2 .
In the first case, If c > 3, then 3 is an exception; and so

1 ≤ c ≤ 3. (2.31)
218 Paper 20

In the second case, if c > 5, then 5 is an exception; and so

2 ≤ c ≤ 5. (2.32)

We can now distinguish 7 possible cases.

(2.41) x2 + y 2 + z 2 + du2 .

If d > 7, 7 is an exception; and so


1 ≤ d ≤ 7. (2.411)
(2.42) x2 + y 2 + 2z 2 + du2 .
If d > 14, 14 is an exception; and so

2 ≤ d ≤ 14. (2.421)

(2.43) x2 + y 2 + 3z 2 + du2 .
If d > 6, 6 is an exception; and so
3 ≤ d ≤ 6. (2.431)
(2.44) x2 + 2y 2 + 2z 2 + du2 .
If d > 7, 7 is an exception; and so
2 ≤ d ≤ 7. (2.441)
(2.45) x2 + 2y 2 + 3z 2 + du2 .
If d > 10, 10 is an exception; and so

3 ≤ d ≤ 10. (2.451)

(2.46) x2 + 2y 2 + 4z 2 + du2 .
If d > 14, 14 is an exception; and so

4 ≤ d ≤ 14. (2.461)

(2.47) x2 + 2y 2 + 5z 2 + du2 .
If d > 10, 10 is an exception; and so

5 ≤ d ≤ 10. (2.471)

We have thus eliminated all possible sets of values of a, b, c, d, except the following 55∗ :

L. E. Dickson (Bulletin of the Ameraican Math. Soc. [vol XXXIII (1927), pp. 63 – 70]) has pointed
out that Ramanujan has overlooked the fact that (1, 2, 5, 5) does not represent 15. Consequently, there are
only 54 forms.
On the expression of a number in the form ax2 + by 2 + cz 2 + du2 219

1, 1, 1, 1 1, 2, 3, 5 1, 2, 4, 8
1, 1, 1, 2 1, 2, 4, 5 1, 2, 5, 8
1, 1, 2, 2 1, 2, 5, 5 1, 1, 2, 9
1, 2, 2, 2 1, 1, 1, 6 1, 2, 3, 9
1, 1, 1, 3 1, 1, 2, 6 1, 2, 4, 9
1, 1, 2, 3 1, 2, 2, 6 1, 2, 5, 9
1, 2, 2, 3 1, 1, 3, 6 1, 1, 2, 10
1, 1, 3, 3 1, 2, 3, 6 1, 2, 3, 10
1, 2, 3, 3 1, 2, 4, 6 1, 2, 4, 10
1, 1, 1, 4 1, 2, 5, 6 1, 2, 5, 10
1, 1, 2, 4 1, 1, 1, 7 1, 1, 2, 11
1, 2, 2, 4 1, 1, 2, 7 1, 2, 4, 11
1, 1, 3, 4 1, 2, 2, 7 1, 1, 2, 12
1, 2, 3, 4 1, 2, 3, 7 1, 2, 4, 12
1, 2, 4, 4 1, 2, 4, 7 1, 1, 2, 13
1, 1, 1, 5 1, 2, 5, 7 1, 2, 4, 13
1, 1, 2, 5 1, 1, 2, 8 1, 1, 2, 14
1, 2, 2, 5 1, 2, 3, 8 1, 2, 4, 14
1, 1, 3, 5

Of these 55 forms, the 12 forms

1, 1, 1, 2 1, 1, 2, 4 1, 2, 4, 8
1, 1, 2, 2 1, 2, 2, 4 1, 1, 3, 3
1, 2, 2, 2 1, 2, 4, 4 1, 2, 3, 6
1, 1, 1, 4 1, 1, 2, 8 1, 2, 5, 10

have been already considered by Liouville and Pepin∗ .

3. I shall now prove that all integers can be expressed in each of the 55 forms. In order to
prove this we shall consider the seven cases (2.41)–(2.47) of the previous section separately.
We shall require the following results concerning ternary quadratic arithmetical forms.

There are a large number of short notes by Liouville in Vols. V–VIII of the second series of his Journal.
See also Pepin, ibid., Ser.4, Vol. VI, pp.1 – 67. The object of the work of Liouville and Pepin is rather
different from mine, viz., to determine, in a number of special cases, explicit formulæ for the number of
representations, in terms of other arithmetical functions.
220 Paper 20

The necessary and sufficient condition that a number cannot be expressed in the form

x2 + y 2 + z 2 (3.1)

is that it should be of the form

4λ (8µ + 7), (λ = 0, 1, 2, · · · , µ = 0, 1, 2, · · ·). (3.11)

Similarly the necessary and sufficient conditions that a number cannot be expressed in the
forms
x2 + y 2 + 2z 2 , (3.2)
x2 + y 2 + 3z 2 , (3.3)
x2 + 2y 2 + 2z 2 , (3.4)
x2 + 2y 2 + 3z 2 , (3.5)
x2 + 2y 2 + 4z 2 , (3.6)
x2 + 2y 2 + 5z 2 , (3.7)
are that it should be of the forms
4λ (16µ + 14), (3.21)
9λ (9µ + 6), (3.31)
4λ (8µ + 7), (3.41)
4λ (16µ + 10), (3.51)
4λ (16µ + 14), (3.61)
25λ (25µ + 10) or 25λ (25µ + 15).∗ (3.71)


Results (3.11)–(3.71) may tempt us to suppose that there are similar simple results for the form ax2 +
by + cz 2 , whatever are the values of a, b, c. It appears, however, that in most cases there are no such simple
2

results. For instance, the numbers which are not of the form x2 + 2y 2 + 10z 2 are those belonging to one or
other of the four classes

25λ (8µ + 7), 25λ (25µ + 5), 25λ (25µ + 15), 25λ (25µ + 20).

Here some of the numbers of the first class belong also to one of the next three classes.
Again, the even numbers which are not of the form x2 + y 2 + 10z 2 are the numbers

4λ (16µ + 6),

while the odd numbers that are not of the form, viz.

3, 7, 21, 31, 33, 43, 67, 79, 87, 133, 217, 219, 223, 253, 307, 391, · · ·

do not seem to obey any simple law.


On the expression of a number in the form ax2 + by 2 + cz 2 + du2 221

The result concerning x2 + y 2 + z 2 is due to Cauchy: for a proof see. Landau, Handbuch
der Lehre von der Verteilung der Primzahlen, p.550. The other results can be proved in an
analogous manner. The form x2 + y 2 + 2z 2 has been considered by Lebesgue, and the form
x2 + y 2 + 3z 2 by Dirichlet. For references see Bachmann, Zahlentheorie, Vol,IV, p.149.

4. We proceed to consider the seven cases (2.41)–(2.47). In the first case we have to shew
that any number N can be expressed in the form

N = x2 + y 2 + z 2 + du2 , (4.1)

d being any integer between 1 and 7 inclusive.


If N is not of the form 4λ (8µ + 7), we can satisfy (4.1) with u = 0. We may therefore
suppose that N = 4λ (8µ + 7).
First, suppose that d has one of the values 1,2,4,5,6. Take u = 2λ . Then the number

N − du2 = 4λ (8µ + 7 − d)

is plainly not of the form 4λ (8µ + 7), and is therefore expressible in the form x2 + y 2 + z 2 .
Next, let d = 3. If µ = 0, take u = 2λ . Then

N − du2 = 4λ+1 .

If µ ≥ 1, take u = 2λ+1 . Then

N − du2 = 4λ (8µ − 5).


I have succeeded in finding a law in the following six simple cases:

x2 + y 2 + 4z 2 ,

x2 + y 2 + 5z 2 ,
x2 + y 2 + 6z 2 ,
x2 + y 2 + 8z 2 ,
x2 + 2y 2 + 6z 2 ,
x2 + 2y 2 + 8z 2 .
The numbers which are not of these forms are the numbers

4λ (8µ + 7) or 8µ + 3,
4λ (8µ + 3),
9λ (9µ + 3),
4λ (16µ + 14), 16µ + 6, or 4µ + 3,
4λ (8µ + 5),
4λ (8µ + 7) or 8µ + 5.
222 Paper 20

In neither of these cases is N − du2 of the form 4λ (8µ + 7), and therefore in either case it
can be expressed in the form x2 + y 2 + z 2 .
Finally, let d = 7. If µ is equal to 0,1, or 2, take u = 2λ . Then N − du2 is equal to 0, 2 · 4λ+1 ,
or 4λ+2 . If µ ≥ 3, take u = 2λ+1 . Then

N − du2 = 4λ (8µ − 21).

Therefore in either case N − du2 can be expressed in the form x2 + y 2 + z 2 .


Thus in all cases N is expressible in the form (4.1). Similarly we can dispose of the
remaining cases, with the help of the results stated in § 3. Thus in discussing (2.42) we use
the theorem that every number not of the form (3.21) can be expressed in the form (3.2).
The proofs differ only in detail, and it is not worth while to state them at length.

5. We have seen that all integers without any exception can be expressed in the form

m(x2 + y 2 + z 2 ) + nu2 , (5.1)

when
m = 1, 1 ≤ n ≤ 7,
and
m = 2, n = 1.
We shall now consider the values of m and n for which all integers with a finite number of
exceptions can be expressed in the form (5.1).
In the first place m must be 1 or 2. For, if m > 2, we can choose an integer ν so that

nu2 6≡ ν(mod m)

for all values of u. Then


(mµ + ν) − nu2
,
m
where µ is any positive integer, is not an integer; and so mµ + ν can certainly not be
expressed in the form (5.1).
We have therefore only to consider the two cases in which m is 1 or 2. First let us consider
the form
x2 + y 2 + z 2 + nu2 . (5.2)
I shall shew that, when n has any of the values

1, 4, 9, 17, 25, 36, 68, 100, (5.21)

or is of any of the forms

4k + 2, 4k + 3, 8k + 5, 16k + 12, 32k + 20, (5.22)


On the expression of a number in the form ax2 + by 2 + cz 2 + du2 223

then all integers save a finite number, and in fact all integers from 4n onwards at any rate,
can be expressed in the form (5.2); but that for the remaining values of n there is an infinity
of integers which cannot be expressed in the form required.
In proving the first result we need obviously only consider numbers of the form 4λ (8µ + 7)
greater than n, since otherwise we may take u = 0. The numbers of this form less than n
are plainly among the exceptions.

6. I shall consider the various cases which may arise in order of simplicity.

(6.1) n ≡ 0(mod 8).

There are an infinity of exceptions. For suppose that

N = 8µ + 7.

Then the number


N − nu2 ≡ 7(mod 8)
cannot be expressed in the form x2 + y 2 + z 2 .

(6.2) n ≡ 2(mod 4).

There is only a finite number of exceptions. In proving this we may suppose that N =
4λ (8µ + 7). Take u = 1. Then the number

N − nu2 = 4λ (8µ + 7) − n

is congruent to 1, 2, 5 or 6 to modulus 8, and so can be expressed in the form x2 + y 2 + z 2 .


Hence the only numbers which cannot be expressed in the form (5.2) in this case are the
numbers of the form 4λ (8µ + 7) not exceeding n.

(6.3) n ≡ 5(mod 8).

There is only a finite number of exceptions. We may suppose again that N = 4λ (8µ + 7).
First, let λ 6= 1. Take u = 1. Then

N − nu2 = 4λ (8µ + 7) − n ≡ 2 or 3(mod 8).

If λ = 1 we cannot take u = 1, since

N − n ≡ 7(mod 8);

so we take u = 2. Then

N − nu2 = 4λ (8µ + 7) − 4n ≡ 8(mod 32).

In either of these cases N − nu2 is of the form x2 + y 2 + z 2 .


224 Paper 20

Hence the only numbers which cannot be expressed in the form (5.2) are those of the form
4λ (8µ + 7) not exceeding n, and those of the form 4(8µ + 7) lying between n and 4n.

(6.4) n ≡ 3(mod 4).

There is only a finite number of exceptions. Take

N = 4λ (8µ + 7).

if λ ≥ 1, take u = 1. Then
N − nu2 ≡ 1 or 5(mod 8).
if λ = 0, take u = 2. Then
N − nu2 ≡ 3(mod 8).
In either case the proof is completed as before.
In order to determine precisely which are the exceptional numbers, we must consider more
particularly the numbers between n and 4n for which λ = 0. For these u must be 1, and

N − nu2 ≡ 0(mod 4).

But the numbers which are multiples of 4 and which cannot be expressed in the form
x2 + y 2 + z 2 are the numbers

4κ (8ν + 7), (κ = 1, 2, 3, . . . , ν = 0, 1, 2, 3, . . .).

The exceptions required are therefore those of the numbers

n + 4κ (8ν + 7) (6.41)

which lie between n and 4n and are of the form

8µ + 7 (6.42).

Now in order that (6.41) may be of the form (6.42), κ must be 1 if n is of the form 8k + 3,
and κ may have any of the values 2, 3, 4, . . . if n is of the form 8k + 7. Thus the only
numbers which cannot be expressed in the form (5.2), in this case, are those of the form
4λ (8µ + 7) less than n and those of the form

n + 4κ (8ν + 7), (ν = 0, 1, 2, 3, · · ·),

lying between n and 4n, where κ = 1 if n is of the form 8k + 3, and κ > 1 if n is of the
form 8k + 7.
(6.5) n ≡ 1(mod 8).
In this case we have to prove that
(i) if n ≥ 33, there is an infinity of integers which cannot be expressed in the form (5.2);
On the expression of a number in the form ax2 + by 2 + cz 2 + du2 225

(ii) if n is 1, 9, 17, or 25, there is only a finite number of exceptions.


In order to prove (i) suppose that N = 7.4λ . Then obviously u cannot be zero. But if u is
not zero u2 is always of the form 4κ (8ν + 1). Hence

N − nu2 = 7 · 4λ − n · 4κ (8ν + 1).

Since n ≥ 33, λ must be greater than or equal to κ + 2, to ensure that the right-hand side
shall not be negative. Hence
N − nu2 = 4κ (8k + 7),
where
1
k = 14 · 4λ−κ−2 − nν − (n + 7)
8
is an integer; and so N − nu is not of the form x + y 2 + z 2 .
2 2

In order to prove (ii) we may suppose, as usual, that

N = 4λ (8µ + 7).

If λ = 0, take u = 1. Then

N − nu2 = 8µ + 7 − n ≡ 6 (mod 8).

If λ ≥ 1, take u = 2λ−1 . Then

N − nu2 = 4λ−1 (8k + 3),

where
1
k = 4(µ + 1) − (n + 7).
8
In either case the proof may be completed as before. Thus the only numbers which cannot
be expressed in the form (5.2), in this case, are those of the form 8µ + 7 not exceeding n.
In other words, there is no exception when n = 1; 7 is the only exception when n = 9; 7
and 15 are the only exceptions when n = 17; 7, 15 and 23 are the only exceptions when
n = 25.
(6.6) n ≡ 4(mod 32).
By arguments similar to those used in (6.5), we can shew that
(i) if n ≥ 132, there is an infinity of integers which cannot be expressed in the form (5.2);
(ii) if n is equal to 4, 36, 68, or 100, there is only a finite number of exceptions, namely the
numbers of the form 4λ (8µ + 7) not exceeding n.

(6.7) n ≡ 20(mod 32).

By arguments similar to those used in (6.3), we can shew that the only numbers which
cannot be expressed in the form (5.2) are those of the form 4λ (8µ + 7) not exceeding n, and
those of the form 42 (8µ + 7) lying between n and 4n.

(6.8) n ≡ 12(mod 16).


226 Paper 20

By arguments similar to those used in (6.4), we can shew that the only numbers which
cannot be expressed in the form (5.2) are those of the form 4λ (8µ + 7) less than n, and
those of the form
n + 4κ (8ν + 7), (ν = 0, 1, 2, 3, . . .),

lying between n and 4n where κ = 2 if n is of the form 4(8k + 3) and κ > 2 if n is of the
form 4(8k + 7).
We have thus completed the discussion of the form (5.2), and determined the exceptional
values of N precisely whenever they are finite in number.

7. We shall proceed to consider the form

2(x2 + y 2 + z 2 ) + nu2 . (7.1)

In the first place n must be odd; otherwise the odd numbers cannot be expressed in this
form. Suppose then that n is odd. I shall shew that all integers save a finite number can
be expressed in the form (7.1); and that the numbers which cannot be so expressed are
(i) the odd numbers less than n,
(ii) the numbers of the form 4λ (16µ + 14) less that 4n,
(iii) the numbers of the form n + 4λ (16µ + 14) greater than n less than 9n,
(iv) the numbers of the form

cn + 4κ (16ν + 14), (ν = 0, 1, 2, 3, . . .),

greater than 9n and less than 25n, where c = 1 if n ≡ 1 (mod 4), c = 9 if n ≡ 3 (mod 4),
κ = 2 if n2 ≡ 1 (mod 16), and κ > 2 if n2 ≡ 9 (mod 16).
First, let us suppose N even. Then, since n is odd and N is even, it is clear that u must
be even. Suppose then that
u = 2v, N = 2M.

We have to shew that M can be expressed in the form

x2 + y 2 + z 2 + 2nv 2 . (7.2)

Since 2n ≡ 2 (mod 4), it follows from (6.2) that all integers except those which are less
than 2n and of the form 4λ (8µ + 7) can be expressed in the form (7.2). Hence the only even
integers which cannot be expressed in the form (7.1) are those of the form 4λ (16µ + 14) less
than 4n.
This completes the discussion of the case in which N is even. If N is odd the discussion
is more difficult. In the first place, all odd numbers less than n are plainly among the
exceptions. Secondly, since n and N are both odd, u must also be odd. We can therefore
suppose that
N = n + 2M, u2 = 1 + 8∆,
On the expression of a number in the form ax2 + by 2 + cz 2 + du2 227

where ∆ is an integer of the form 12 k(k + 1), so that ∆ may assume the values 0, 1, 3, 6,
. . . And we have to consider whether n + 2M can be expressed in the form

2(x2 + y 2 + z 2 ) + n(1 + 8∆),

or M in the form
x2 + y 2 + z 2 + 4n∆. (7.3)
If M is not of the form 4λ (8µ + 7), we can take ∆ = 0. If it is of this form, and less than
4n, it is plainly an exception. These numbers give rise to the exceptions specified in (iii) of
section 7. We may therefore suppose that M is of the form 4λ (8µ + 7) and greater than 4n.

8. In order to complete the discussion, we must consider the three cases in which n ≡ 1
(mod 8), n ≡ 5 (mod 8), and n ≡ 3 (mod 4) separately.

(8.1) n ≡ 1 mod 8).

If λ is equal to 0,1, or 2, take ∆ = 1. Then

M − 4n∆ = 4λ (8µ + 7) − 4n

is of one of the forms


8ν + 3, 4(8ν + 3), 4(8ν + 6).
If λ ≥ 3 we cannot take ∆ = 1, since M − 4n∆ assumes the form 4(8ν + 7); so we take
∆ = 3. Then
M − 4n = ∆4λ (8µ + 7) − 12n
is of the form 4(8ν + 5). In either of these cases M − 4n∆ is of the form x2 + y 2 + z 2 . Hence
the only values of M , other than those already specified which cannot be expressed in the
form (7.3), are those of the form

4κ (8ν + 7), (ν = 0, 1, 2, . . . , κ > 2),

lying between 4n and 12n. In other words, the only numbers greater than 9n which cannot
be expressed in the form (7.1), in this case, are the numbers of the form

n + 4κ (8ν + 7), (ν = 0, 1, 2, . . . , κ > 2),

lying between 9n and 25n.


(8.2) n ≡ 5 mod 8).
if λ 6= 2, take ∆ = 1. Then

M − 4n∆ = 4λ (8µ + 7) − 4n

is of one of the forms


8ν + 3, 4(8ν + 2), 4(8ν + 3).
228 Paper 20

If λ = 2, we cannot take ∆ = 1, since M − 4n∆ assumes the form 4(8ν + 7); so we take
∆ = 3. Then
M − 4n∆ = 4λ (8µ + 7) − 12n
is of the form 4(8ν + 5). In either of these cases M − 4n∆ is of the form x2 + y 2 + z 2 . Hence
the only values of M , other than those already specified, which cannot be expressed in the
form (7.3), are those of the form 16(8µ + 7) lying between 4n and 12n. In other words, the
only numbers greater than 9n which cannot be expressed in the form (7.1), in this case, are
the numbers of the form n + 42 (16µ + 14) lying between 9n and 25n.
(8.3) n ≡ 3 (mod 4).
If λ 6= 1, take ∆ = 1. Then
M − 4n∆ = 4λ (8µ + 7) − 4n
is of one of the forms
8ν + 3, 4(4ν + 1).
If λ = 1, take ∆ = 3. Then
M − 4n∆ = 4(8µ + 7) − 12n
is of the form 4(4ν + 2). In either of these cases M − 4n∆ is of the form x2 + y 2 + z 2 .
This completes the proof that there is only a finite number of exceptions. In order to
determine what they are in this case, we have to consider the values of M , between 4n and
12n, for which ∆ = 1 and
M − 4n∆ = 4(8µ + 7 − n) ≡ 0 mod 16).
But the numbers which are multiples of 16 and which cannot be expressed in the form
x2 + y 2 + z 2 are the numbers
4κ (8ν + 7), (κ = 2, 3, 4, . . . , ν = 0, 1, 2, . . .).
The exceptional values of M required are therefore those of the numbers
4n + 4κ (8ν + 7) (8.31)
which lie between 4n and 12n and are of the form
4(8µ + 7). (8.32)
But in order that (8.31) may be of the form (8.32), κ must be 2 if n is of the form 8k + 3,
and κ may have any of the values 3, 4, 5, . . . if n is of the form 8k + 7. It follows that the
only numbers greater than 9n which cannot be expressed in the form (7.1), in this case, are
the numbers of the form
9n + 4κ (16ν + 14), (ν = 0, 1, 2, . . .),
lying between 9n and 25n, where κ = 2 if n is of the form 8k + 3, and κ > 2 if n is of the
form 8k + 7.
This completes the proof of the results stated in section 7.
On certain trigonometrical sums and their applications in the
theory of numbers

Transactions of the Cambridge Philosophical Society, XXII, No.13, 1918, 259 – 276

1. The trigonometrical sums with which this paper is concerned are of the type
X 2πλn
cs (n) = cos ,
s
λ

where λ is prime to s and not greater than s. It is plain that


X
cs (n) = αn ,

where α is a primitive root of the equation

xs − 1 = 0.

These sums are obviously of very great interest, and a few of their properties have been
discussed already∗ . But, so far as I know, they have never been considered from the point
of view which I adopt in this paper; and I believe that all the results which it contains are
new.
My principal object is to obtain expressions for a variety of well-known arithmetical func-
tions of n in the form of a series X
as cs (n).
s

A typical formula is

π2n
 
c1 (n) c2 (n) c3 (n)
σ(n) = + 2 + 2 + ··· ,
6 12 2 3

where σ(n) is the sum of the divisors of n. I give two distinct methods for the proof of this
and a large variety of similar formulæ. The majority of my formulæ are “elementary” in
the technical sense of the word — they can (that is to say) be proved by a combination
of processes involving only finite algebra and simple general theorems concerning infinite
series. These are however some which are of a “deeper” character, and can only be proved
by means of theorems which seem to depend essentially on the theory of analytic functions.
A typical formula of this class is

c1 (n) + 12 c2 (n) + 13 c3 (n) + · · · = 0,



See, e.g., Dirichlet-Dedekind, Vorlesungen über Zahlentheorie, ed. 4, Supplement VII, pp. 360 – 370.
230 Paper 21

a formula which depends upon, and is indeed substantially equivalent to, the “Prime Num-
ber Theorem” of Hadamard and de la Vallée Poussin.
Many of my formulæ are intimately connected with those of my previous paper “On certain
arithmetical functions”, published in 1916 in these Transactions∗ . They are also connected
(in a manner pointed out in § 15) with a joint paper by Mr Hardy and myself, “Asymptotic
Formulæ in Combinatory Analysis”, in course of publication in the Proceedings of the
London Mathematical Society† .

2. Let F (u, v) be any function of u and v, and let


X
(2.1) D(n) = F (δ, δ′ ),
δ

where δ is a divisor of n and δδ′ = n. For instance


D(1) = F (1, 1); D(2) = F (1, 2) + F (2, 1);
D(3) = F (1, 3) + F (3, 1); D(4) = F (1, 4) + F (2, 2) + F (4, 1);
D(5) = F (1, 5) + F (5, 1); D(6) = F (1, 6) + F (2, 3) + F (3, 2) + F (6, 1); . . . . . .
It is clear that D(n) may also be expressed in the form
X
(2.2) D(n) = F (δ′ , δ).
δ

Suppose now that


s−1
X 2πνn
(2.3) ηs (n) = cos ,
0
s

so that ηs (n) = s if s is a divisor of n and ηs (n) = 0 otherwise. Then


t
X 1  n
(2.4) D(n) = ην (n)F ν, ,‡
1
ν ν

where t is any number not less than n. Now let


X 2πλn
(2.5) cs (n) = cos ,
s
λ

where λ is prime to s and does not exceed s; e.g.

c1 (n) = 1; c2 (n) = cos nπ; c3 (n) = 2 cos 32 nπ;



[No.18 of this volume].

[No. 36 of this volume].
Xt X
[t]

is to be understood as meaning , where [t] denotes as usual the greatest integer in t.
1 1
On certain trigonometrical sums and their applications in the theory of numbers 231

c4 (n) = 2 cos 21 nπ; c5 (n) = 2 cos 25 nπ + 2 cos 54 nπ;


c6 (n) = 2 cos 31 nπ; c7 (n) = 2 cos 27 nπ + 2 cos 74 nπ + 2 cos 76 nπ;
c8 (n) = 2 cos 41 nπ + 2 cos 34 nπ; c9 (n) = 2 cos 92 nπ + 2 cos 94 nπ + 2 cos 89 nπ;
c10 (n) = 2 cos 51 nπ + 2 cos 35 nπ; . . . .

It follows from (2.3) and (2.5) that


X
(2.6) ηs (n) = cδ (n),
δ

where δ is a divisor of s; and hence∗ that


X
(2.7) cs (n) = µ(δ′ )ηδ (n),
δ

where δ is a divisor of s, δδ′ = s, and


X µ(ν) 1
(2.8) = ,
νs ζ(s)

ζ(s) being the Riemann Zeta-function. In particular

c1 (n) = η1 (n); c2 (n) = η2 (n) − η1 (n); c3 (n) = η3 (n) − η1 (n);


c4 (n) = η4 (n) − η2 (n); c5 (n) = η5 (n) − η1 (n); . . . . . .

But from (2.3) we know that ηδ (n) = 0 if δ is not a divisor of n; and so we can suppose
that, in (2.7), δ is a common divisor of n and s. It follows that
X
|cs (n)| ≤ δ,

where δ is a divisor of n; so that

(2.9) cν (n) = O(1)

if n is fixed and ν → ∞. Since

ηs (n) = ηs (n + s); cs (n) = cs (n + s),

the values of cs (n) for n = 1, 2, 3, . . . can be shewn conveniently by writing

c1 (n) = 1; c2 (n) = −1, 1; c3 (n) = −1, −1, 2;


c4 (n) = 0, −2, 0, 2; c5 (n) = −1, −1, −1, −1, 4;

See Landau, Handbuch der Lehre von der Verteilung der Primzahlen, p. 577.
232 Paper 21

c6 (n) = 1, −1, −2, −1, 1, 2 : c7 (n) = −1, −1, −1, −1, −1, −1, 6;
c8 (n) = 0, 0, 0, −4, 0, 0, 0, 4; c9 (n) = 0, 0, −3, 0, 0, −3, 0, 0, 6;
c10 (n) = 1, −1, 1, −1, −4, −1, 1, −1, 1, 4; . . . . . .

the meaning of the third formula, for example, being that c3 (1) = −1, c3 (2) = −1, c3 (3) = 2,
and that these values are then repeated periodically.
It is plain that we have also

(2.91) cν (n) = O(1).

when ν is fixed and n → ∞.

3. Substituting (2.6) in (2.4) and collecting the coefficients of c1 (n), c2 (n), c3 (n) . . . , we
find that

(3.1)
1 1
t 2t 3t
X 1  n X 1  n X 1  n
D(n) = c1 (n) F ν, + c2 (n) F 2ν, + c3 (n) F 3ν, + ···,
ν ν 2ν 2ν 3ν 3ν
1 1 1

where t is any number not less than n. If we use (2.2) instead of (2.1) we obtain another
expression, viz.

(3.2)
1 1
t 2t 3t
X 1 n  X 1 n  X 1 n 
D(n) = c1 (n) F , ν + c2 (n) F , 2ν , +c3 (n) F , 3ν + · · · ,
1
ν ν 1
2ν 2ν 1
3ν 3ν

where t is any number not less that n.


Suppose now that

F1 (u, v) = F (u, v) log u, F2 (u, v) = F (u, v) log v.

Then we have
X X
D(n) log n = F (δ, δ′ ) log n = F (δ, δ′ ) log(δδ′ )
δ δ
X X

= F1 (δ, δ ) + F2 (δ, δ′ ),
δ δ

where δ is a divisor of n and δδ′ = n.


On certain trigonometrical sums and their applications in the theory of numbers 233

Now for δ F1 (δ, δ′ ) we shall write the expression corresponding to (3.1) and for δ F2 (δ, δ′ )
P P

the expression corresponding to (3.2). Then we have


1
r 2r
X log ν  n X log 2ν  n
(3.3) D(n) log n = c1 (n) F ν, + c2 (n) F 2ν,
ν ν 2ν 2ν
1 1
1
3 r t
X log 3ν  n X log ν  n 
+ c3 (n) F 3ν, + · · · + c1 (n) F ,ν
1
3ν 3ν 1
ν ν
1 1
2t 3t
X log 2ν  n  X log 3ν  n 
+ c2 (n) F , 2ν + c3 (n) F , 3ν + · · · ,
2ν 2ν 3ν 3ν
1 1

where r and t are any two numbers not less than n. If, in particular, F (u, v) = F (v, u),
then (3.3) reduces to
t
1
X log ν  n
(3.4) 2 D(n) log n = c1 (n) F ν,
1
ν ν
1 1
2t 3t
X log 2ν  n X log 3ν  n
+ c2 (n) F 2ν, + c3 (n) F 3ν, + ···,
2ν 2ν 3ν 3ν
1 1

where t is any number not less than n.

4. We may also write D(n) in the form


u
X r
X
(4.1) D(n) = F (δ, δ′ ) + F (δ′ , δ),
δ=1 δ=1

where δ is a divisor of n, δδ′ = n, and u, v are may two positive numbers such that uv = n,
it being understood that, if u and v are both integral, a term F (u, v) is to be subtracted
from the right-hand side. Hence (with the same conventions)
u v
X 1  n X 1 n 
D(n) = ην (n)F ν, + ην (n)F ,ν ,
ν ν ν ν
1 1

Applying to this formula transformations similar to those of §3, we obtain


1
u 2u
X 1  n X 1  n
(4.2) D(n) = c1 (n) F ν, + c2 (n) F 2ν, + ···
1
ν ν 1
2ν 2ν
1
v 2v
X 1 n  X 1 n 
+ c1 (n) F , ν + c2 (n) F , 2ν + · · · ,
ν ν 2ν 2ν
1 1
234 Paper 21

where u and v are positive numbers such that uv = n. If u and v are integers then a term
F (u, v) should be subtracted from the right-hand side.
If we suppose that 0 < u ≤ 1 then (4.2) reduces to (3.2), and if 0 < v ≤ 1 it reduces to
(3.1). Another particular case of interest is that in which u = v. Then

n
X 1 n  n  n o
(4.3) D(n) = c1 (n) F ν, +F ,ν
ν ν ν
1
1√
2Xn
1 n  n n o
+ c2 (n) F 2ν, +F , 2ν + ···.
1
2ν 2ν 2ν
√ √
If n is a perfect square then F ( n, n) should be subtracted from the right hand side.

5. We shall now consider some special forms of these general equations. Suppose that
F (u, v) = v s , so that D(n) is the sum σs (n) of the sth powers of the divisors of n. Then
from (3.1) and (3.2) we have
1 1
t 2 t 3 t
σs (n) X 1 X 1 X 1
(5.1) = c1 (n) + c2 (n) + c3 (n) + ···,
ns 1
ν s+1 1
(2ν)s+1 1
(3ν)s+1

1 1
t 2 t 3 t
X X X
(5.2) σs (n) = c1 (n) ν s−1 + c2 (n) (2ν)s−1 + c3 (n) (3ν)s−1 + · · · ,
1 1 1

where t is any number not less than n: from (3.3)


1
r 2 r
X X
(5.3) σs (n) log n = c1 (n) ν s−1 log ν + c2 (n) (2ν)s−1 log 2ν + · · ·
1 1
1
 
 t 2 t 
s
 X log ν X log 2ν 
+ n c1 (n) + c2 (n) + · · · ,

 1
ν s+1 1
(2ν)s+1 

where r and t are two numbers not less than n: and from (4.2)
1 1
u 2 u 3 u
X X X
s−1
(5.4) σs (n) = c1 (n) ν + c2 (n) (2ν)s−1 + c3 (n) (3ν)s−1 + · · ·
1 1 1
1 1
 
 v 2v 3v 
 X 1 X 1 X 1 
+ ns c1 (n) + c2 (n) + c3 (n) + ··· ,
 ν s+1 (2ν)s+1 (3ν)s+1 
 1 1 1 
On certain trigonometrical sums and their applications in the theory of numbers 235

where uv = n. If u and v are integers then us should be subtracted from the right-hand
side.
Let d(n) = σ0 (n) denote the number of divisors of n and σ(n) = σ1 (n) the sum of the
divisors of n. Then from (5.1) – (5.4) we obtain
1 1
t 2 3 t t
X 1 X 1 X 1
(5.5) d(n) = c1 (n) + c2 (n) + c3 (n) + ···,
1
ν 1
2ν 1

(5.6) σ(n) = c1 (n)[t] + c2 (n)[ 12 t] + c3 (n)[ 31 t] + · · · ,

1 1
t 2 t 3 t
1
X log ν X log 2ν X log 3ν
(5.7) 2 d(n) log n = c1 (n) + c2 (n) + c3 (n) + ···,
1
ν 1
2ν 1

1 1

( u v
)  2 u 2 v 
X 1 X 1 X 1 X 1
(5.8) d(n) = c1 (n) + + c2 (n) +
ν ν  2ν 2ν 
1 1  1  1
1 1

 3u 3v 
X 1 X 1
+ c3 (n) + + ···
 1 3ν

1
3ν 

where t ≥ n and uv = n. If u and v are integers then 1 should be subtracted from the

right-hand side of (5.8). Putting u = v = n in (5.8) we obtain
√ 1√ 1√
n 2X 3X n n
1
X 1 1 1
(5.9) 2 d(n) = c1 (n) + c2 (n) + c3 (n) + ···,
1
ν 1
2ν 1

unless n is a perfect square, when 12 should be subtracted from the right-hand side. It may
be interesting to note that, if we replace the left-hand side in (5.9) by
1 1 
2 + 2 d(n) ,

then the formula is true without exception.

6. So far our work has been based on elementary formal transformations, and no questions
of convergence have arisen. We shall now consider the equation (5.1) more carefully. Let
us suppose that s > 0. Then
t/k ∞    
X 1 X 1 1 1 1
s+1
= +O = ζ(s + 1) + O .
(kν) (kν)s+1 kts ks+1 kts
1 1
236 Paper 21

The number of terms in the right-hand side of (5.1) is [t]. Also we know that cν (n) = O(1)
as n → ∞. Hence
t t ∞
( )  
σs (n) X cν (n) 1 X1 X cν (n) log t
= ζ(s + 1) +O s = ζ(s + 1) +O .
ns ν s+1 t ν ν s+1 ts
ν=1 ν=1 1

Making t → ∞, we obtain
 
s c1 (n) c2 (n) c3 (n)
(6.1) σs (n) = n ζ(s + 1) + s+1 + s+1 + · · · ,
1s+1 2 3

if s > 0. Similarly, if we make t → ∞ in (5.3), we obtain


1
r 2 r
X X
s−1
σs (n) log n = c1 (n) ν log ν + c2 (n) (2ν)s−1 log 2ν + · · ·
1 1
∞ ∞
( )
X log ν X log 2ν
+ns c1 (n) + c2 (n) + ··· .
1
ν s+1 1
(2ν)s+1

But

X log kν log k 1
s+1
= s+1 ζ(s + 1) − s+1 ζ ′ (s + 1).
(kν) k k
1

It follows from this and (6.1) that


t
ζ ′ (s + 1)
  X
(6.2) σs (n) + log n = c1 (n) ν s−1 log ν
ζ(s + 1) 1
1
2t
X
+ c2 (n) (2ν)s−1 log 2ν + · · ·
1
 
s c1 (n) log 1 c2 (n) log 2 c3 (n) log 3
+ n ζ(s + 1) + + + ··· ,
1s+1 2s+1 3s+1

where s > 0 and t ≥ n. Putting s = 1 in (6.1) and (6.2) we obtain

π2
 
c1 (n) c2 (n) c3 (n)
(6.3) σ(n) = n + 2 + 2 + ··· ,
6 12 2 3

ζ ′ (2) π2
   
c1 (n) c2 (n)
(6.4) σ(n) + log n = n 2
log 1 + 2 log 2 + · · ·
ζ(2) 6 1 2
+ c1 (n)[t] log 1 + c2 (n)[ 12 t] log 2 + · · ·
+ c1 (n) log[t]! + c2 (n) log[ 21 t]! + · · · ,
On certain trigonometrical sums and their applications in the theory of numbers 237

where t ≥ n.

7. Since

(7.1) σs (n) = ns σ−s (n),

we may write (6.1) in the form

σ−s (n) c1 (n) c2 (n) c3 (n)


(7.2) = s+1 + s+1 + s+1 + · · · ,
ζ(s + 1) 1 2 3

where s > 0. This result has been proved by purely elementary methods. But in order
to know whether the right-hand side of (7.2) is convergent or not for values of s less than
or equal to zero we require the help of theorems which have only been established by
transcendental methods.
Now the right-hand side of (7.2) is an ordinary Dirichlet’s series for
1
σ−s (n) × .
ζ(s + 1)

The first factor is a finite Dirichlet’s series and so an absolutely convergent Dirichlet’s series.
It follows that the right-hand side of (7.2) is convergent whenever the Dirichlet’s series for
1/ζ(s + 1), viz.
X µ(n)
(7.3) ,
n1+s
is convergent. But it is known∗ that the series (7.3) is convergent when s = 0 and that its
sum is 0.
It follows from this that

(7.4) c1 (n) + 12 c2 (n) + 13 c3 (n) + · · · = 0.

Nothing is known about the convergence of (7.3) when − 12 < s < 0. But with the assump-
tion of the truth of the hitherto unproved Riemann hypothesis it has been proved† that
(7.3) is convergent when s > − 12 . With this assumption we see that (7.2) is true when
s > − 21 . In other words, if − 12 < s < 12 , then
 
c1 (n) c2 (n) c3 (n)
(7.5) σs (n) = ζ(1 − s) + 1−s + 1−s + · · ·
11−s 2 3
 
s c1 (n) c2 (n) c3 (n)
= n ζ(1 + s) + 1+s + 1+s + · · · .
11+s 2 3


Landau, Handbuch, p. 591.

Littlewood, Comptes Rendus, 29 Jan. 1912.
238 Paper 21

8. It is known∗ that all the series obtained from (7.3) by term-by-term differentiation with
respect to s are convergent when s = 0, and it is obvious that the derivatives of σ−s (n)
with respect to s are all finite Dirichlet’s series and so absolutely convergent. It follows
that all the derivatives of the right-hand side of (7.2) are convergent when s = 0; and so
we can equate the coefficients of like powers of s from the two sides of (7.2). Now
1
(8.1) = s − γs2 + · · · ,
ζ(s + 1)
where γ is Euler’s constant. And
X X X
σ−s (n) = δ−s = 1−s log δ + · · · ,
δ δ δ

where δ is a divisor of n. But


X X X
1
log δ = log δ′ = 2 log(δδ′ ) = 12 d(n) log n,
δ δ δ

where δδ′ = n. Hence

(8.2) σ−s (n) = d(n) − 21 sd(n) log n + · · · .

Now equating the coefficients of s and s2 from the two sides of (7.2) and using (8.1) and
(8.2), we obtain

(8.3) c1 (n) log 1 + 12 c2 (n) log 2 + 31 c3 (n) log 3 + · · · = −d(n),

(8.4) c1 (n)(log 1)2 + 12 c2 (n)(log 2)2 + 13 c3 (n)(log 3)2 + · · · = −d(n)(2γ + log n).

9. I shall now find an expression of the same kind for φ(n), the number of numbers prime
to and not exceeding n. Let p1 , p2 , p3 , . . . be the prime divisors of n, and let

(9.1) φs (n) = ns (1 − p−s −s −s


1 )(1 − p2 )(1 − p3 ) · · · ,

so that φ1 (n) = φ(n). Suppose that

F (u.v) = µ(u)v s .

Then it is easy to see that


D(n) = φs (n).
Hence, from (3.1), we have
1
t 2 t
φs (n) X µ(ν) X µ(2ν)
(9.2) s
= c1 (n) s+1
+ c2 (n) + ···,
n 1
ν 1
(2ν)s+1

Landau, Handbuch, p.594
On certain trigonometrical sums and their applications in the theory of numbers 239

where t is any number not less than n. If s > 0 we can make t → ∞, as in §6. Then we
have
∞ ∞
φs (n) X µ(ν) X µ(2ν)
(9.3) s
= c1 (n) s+1
+ c2 (n) + ···.
n ν (2ν)s+1
1 1

But it can easily be shewn that



X µ(nν) µ(n)
(9.4) = ,
1
νs ζ(s)(1 − p−s
1 )(1 − p−s −s
2 )(1 − p3 ) · · ·

where p1 , p2 , p3 , . . . are the prime divisors of n. In other words



X µ(nν) µ(n)ns
(9.5) = .
νs φs (n)ζ(s)
1

It follows from (9.3) and (9.5) that

φs (n)ζ(s + 1) µ(1)c1 (n) µ(2)c2 (n) µ(3)c3 (n)


(9.6) s
= + + + ···.
n φs+1 (1) φs+1 (2) φs+1 (3)

In particular

π2 c2 (n) c3 (n) c5 (n)


(9.7) φ(n) = c1 (n) − 2 − 2 − 2
6n 2 −1 3 −1 5 −1
c6 (n) c7 (n) c10 (n)
+ 2 − 2 + 2 − ···.
(2 − 1)(3 − 1) 7 − 1 (2 − 1)(52 − 1)
2

10. I shall now consider an application of the main formulæ to the problem of the number
of representations of a number as the sum of 2, 4, 6, 8, . . . squares. We require the following
preliminary results.
(1) Let
X 1s−1 x 2s−1 x2 3s−1 x3
(10.1) D(n)xn = X1 = + + + ···.
1+x 1 − x2 1 + x3
We shall choose
F (u, v) = v s−1 , u ≡ 1( mod 2),
F (u, v) = −v s−1 , u ≡ 2( mod 4),
F (u, v) = (2s − 1)v s−1 , u ≡ 0( mod 4).
240 Paper 21

Then from (3.1) we can shew, by arguments similar to those used in §6, that

D(n) = ns−1 (1−s + 3−s + 5−s + · · ·){1−s c1 (n) + 2−s c4 (n) + 3−s c3 (n)
(10.11)
+4−s c8 (n)5−s c5 (n) + 6−s c12 (n) + 7−s c7 (n) + 8−s c16 (n) + · · ·}.

if s > 1.
(2) Let
X 1s−1 x 2s−1 x2 3s−1 x3
(10.2) D(n)xn = X2 = + + + ···.
1−x 1 + x2 1 − x3
We shall choose
F (u, v) = v s−1 , u ≡ 1( mod 2),
F (u, v) = v s−1 , u ≡ 2( mod 4),
F (u, v) = (1 − 2s )v s−1 , u ≡ 0( mod 4).
Then we obtain as before
D(n) = ns−1 (1−s + 3−s + 5−s + · · ·){1−s c1 (n) − 2−s c4 (n) + 3−s c3 (n)
(10.21)
−4−s c8 (n) + 5−s c5 (n) − 6−s c12 (n) + 7−s c7 (n) − 8−s c16 (n) + · · ·}.

(3) Let
X 1s−1 x 2s−1 x2 3s−1 x3
(10.3) D(n)xn = X3 = + + + ···.
1 + x2 1 + x4 1 + x6
We shall choose
F (u, v) = 0, u ≡ 0( mod 2),
F (u, v) = v s−1 , u ≡ 1( mod 4),
F (u, v) = −v s−1 , u ≡ 3( mod 4).
Then we obtain as before

(10.31) D(n) = ns−1 (1−s − 3−s + 5−s − · · ·){1−s c1 (n) − 3−s c3 (n) + 5−s c5 (n) − · · ·}.

(4) We shall also require a similar formula for the function D(n) defined by
X 1s−1 x 3s−1 x3 5s−1 x5
(10.4) D(n)xn = X4 = − + − ···.
1−x 1 − x3 1 − x5
The formula required is not a direct consequence of the preceding analysis, but if, instead
of starting with the function
X 2πnλ
cr (n) = cos ,
r
λ
On certain trigonometrical sums and their applications in the theory of numbers 241

we start with the function


1 2πnλ
(−1) 2 (λ−1) sin
X
sr (n) = ,
r
λ

where λ is prime to r and does not exceed r, and proceed as in §§2–3, we can shew that

(10.41) D(n) = 21 ns−1 (1−s − 3−s + 5−s − · · ·){1−s s4 (n) + 2−s s8 (n) + 3−s s12 (n) + · · ·}.

It should be observed that there is a correspondence between cr (n) and the ordinary ζ-
function of the one hand and sr (n) and the function

η(s) = 1−s − 3−s + 5−s − · · ·

on the other. It is possible to define an infinity of systems of trigonometrical sums such as


cr (n), sr (n), each corresponding to one of the general class of “L-functions∗ ” of which ζ(s)
and η(s) are the simplest members.
We have shewn that (10.31) and (10.41) are true when s > 1. But if we assume that the
Dirichlet’s series for 1/η(s) is convergent when s = 1, a result which is precisely of the
same depth as the prime number theorem and has only been established by transcendental
methods, then we can shew by arguments similar to those of §7 that (10.31) and (10.41)
are true when s = 1.

11. I have shewn elsewhere† that if s is a positive integer and


X
1+ rs (n)xn = (1 + 2x + 2x4 + 2x9 + · · ·)s ,

then
r2s (n) = δ2s (n) + e2s (n),
where e2s (n) = 0 when s = 1, 2, 3 or 4 and is of lower order ‡ than δ2s (n) in all cases; that
if s is a multiple of 4 then
X πs
(11.1) (1−s + 3−s + 5−s + · · ·) δ2s (n)xn = X1 ;
(s − 1)!
if s is of the form 4k + 2 then
X πs
(11.2) (1−s + 3−s + 5−s + · · ·) δ2s (n)xn = X2 ;
(s − 1)!
if s is of the form 4k + 1 then
X πs
(11.3) (1−s − 3−s + 5−s − · · ·) δ2s (n)xn = (X3 + 21−s X4 );
(s − 1)!

See Landau, Handbuch, pp. 414 et seq.

Transactions of the Cambridge Philosophical Society Vol.XXII, 1916, pp. 159 – 184. [No.18 of this
volume; see in particular §§ 24 – 28, pp. 202 – 208].

For a more precise result concerning the order of e2s (n) see § 15.
242 Paper 21

except when s = 1; and if s is of the form 4k + 3 then


X πs
(11.4) (1−s − 3−s + 5−s − · · ·) δ2s (n)xn = (X3 − 21−s X4 ),
(s − 1)!

X1 , X2 , X3 , X4 being the same as in §10.


In the case in which s = 1 it is well known that

x3 x5
 
P n x
δ2 (n)x = 4 − + − ···
1 − x 1 − x3 1 − x5
(11.5)
x2 x3
 
x
= 4 + + + · · ·
1 + x2 1 + x4 1 + x6

It follows from §10 that, if s is a multiple of 4 then

π s ns−1 −s
δ2s (n) = {1 c1 (n) + 2−s c4 (n) + 3−s c3 (n) + 4−s c8 (n)
(11.11) (s − 1)!
+5−s c5 (n) + 6−s c12 (n) + 7−s c7 (n) + 8−s c16 (n) + · · ·};

if s is of the form 4k + 2 then

π s ns−1 −s
δ2s (n) = {1 c1 (n) − 2−s c4 (n) + 3−s c3 (n) − 4−s c8 (n)
(11.21) (s − 1)!
+5−s c5 (n) − 6−s c12 (n) + 7−s c7 (n) − 8−s c16 (n) + · · ·};

if s is of the form 4k + 1 then

π s ns−1 −s
δ2s (n) = {1 c1 (n) + 2−s s4 (n) − 3−s c3 (n) + 4−s s8 (n)
(11.31) (s − 1)!
+5−s c5 (n) + 6−s s12 (n) − 7−s c7 (n) + 8−s s16 (n) + · · ·};

except when s = 1; and if s is of the form 4k + 3 then

π s ns−1 −s
δ2s (n) = {1 c1 (n) − 2−s s4 (n) − 3−s c3 (n) − 4−s s8 (n)
(11.41) (s − 1)!
+5−s c5 (n) − 6−s s12 (n) − 7−s c7 (n) − 8−s s16 (n) + · · ·};

From (11.5) and the remarks at the end of the previous section, it follows that

r2 (n) = δ2 (n) = π{c1 (n) − 13 c3 (n) + 15 c5 (n) − · · ·}


(11.51)
= π{ 21 s4 (n) + 41 s8 (n) + 61 s12 (n) + · · ·},

but this of course not such an elementary result as the preceding ones.
On certain trigonometrical sums and their applications in the theory of numbers 243

We can combine all the formulæ (11.11) – (11.41) in one by writing

π s ns−1 −s
δ2s (n) = {1 c1 (n) + 2−s c4 (n) + 3−s c3 (n) + 4−s c8 (n)
(11.6) (s − 1)!
+5−s c5 (n) + 6−s c12 (n) + 7−s c7 (n) + 8−s c16 (n) + · · ·},

where s is an integer greater than 1 and

cr (n) = cr (n) cos 12 πs(r − 1) − sr (n) sin 21 πs(r − 1).

12. We can obtain analogous results concerning the number of representations of a number
as the sum of 2, 4, 6, 8, . . . triangular numbers. Equation (147) of my former paper∗ is
equivalent to

X
(12.1) (1 − 2x + 2x4 − 2x9 + · · ·)2s = 1 + δ2s (n)(−x)n
1

f 4s (x) X f 24n (x2 )


+ Kn (−x)n ,
f 2s (x2 ) f 24n (x)
1
1≤n≤ 4 (s−1)

where Kn is a constant and

f (x) = (1 − x)(1 − x2 )(1 − x3 ) · · · .

Suppose now that


x = e−πα , x′ = e−2π/α .
Then we know that
√ ′1 ′ ′
(12.2) α(1 − 2x + 2x4 − 2x9 + · · ·) = 2x 8 (1 + x′ + x 3 + x 6 + · · ·),

q
1 ′ 1 ′ √ 1 ′ 1
(12.3) ( 21 α)x 24 f (x) = x 12 f (x 2 ), αx 12 f (x2 ) = x 24 f (x′ ).
P∞
Finally 1 + 1 δ2s (n)(−x)n can be expressed in powers of x′ by using the formula

12s−1 22s−1 32s−1


 
1
αs 2 ζ(1 − 2s) + + + + ···
e2α − 1 e4α − 1 e6α − 1
(12.4)
12s−1 22s−1 32s−1
 
1
= (−β)s 2 ζ(1 − 2s) + 2β + + + ··· ,
e − 1 e4β − 1 e6β − 1

Loc. cit., p.181 [p. 204 of this volume].
244 Paper 21

where αβ = π 2 and s is an integer greater than 1; and


12s 22s 32s
 
1
s+ 2
(2α) + + + ···
eα + e−α e2α + e−2α e3α + e−3α
(12.5)
12s 32s 52s
 
s 1
p
= (−β) (2β) 2 η(−2s) + β − + − ··· ,
e − 1 e3β − 1 e5β − 1
where αβ = π 2 , s is any positive integer, and η(s) is the function represented by the series
1−s − 3−s + 5−s − · · · and its analytical continuations.
It follows from all these formulæ that, if s is a positive integer and
X X X
(12.6) (1 + x + x3 + x6 + · · ·)2s = ′
r2s (n)xn = ′
δ2s (n)xn + e′2s (n)xn ,

then
X f 4s (x2 ) X f 24n (x)
e′2s (n)xn = Kn (−x)−n ,
f 2s (x) f 24n (x2 )
1
1≤n≤ 4 (s−1)

where Kn is a constant, and f (x) is the same as in (12.1);

′ (n)xn =
( 21 π)s − 1 s
(1−s + 3−s + 5−s + · · ·)
P
δ2s x 4
(s − 1)!
(12.61)  s−1
2s−1 x2 3s−1 x3

1 x
+ + + ···
1 − x2 1 − x4 1 − x6
if s is a multiple of 4;

′ (n)xn =
2( 14 π)s − 1 s
(1−s + 3−s + 5−s + · · ·)
P
δ2s x 4
(s − 1)!
(12.62)  1 3 5

s−1 s−1 s−1
 1 x + 3 x + 5 x + · · ·
2 2 2

1−x 1 − x3 1 − x5

if s is of the form 4k + 2;
X 2( 18 π)s − 1 s
(12.63) (1−s − 3−s + 5−s − · · ·) ′
δ2s (n)xn = x 4
(s − 1)!
1 3 5 1 3 5
!
1s−1 x 4 3s−1 x 4 5−s x 4 1s−1 x 4 3s−1 x 4 5s−1 x 4
× 1 + 3 + 5 + ··· + 1 − 3 + 5 − ···
1 + x2 1 + x2 1 + x2 1 − x2 1 − x2 1 − x2
if s is of the form 4k + 1 (except when s = 1); and
X 2( 18 π)s − 1 s
(12.64) (1−s − 3−s + 5−s − · · ·) ′
δ2s (n)xn = x 4
(s − 1)!
1 3 5 1 3 5
!
1s−1 x 4 3s−1 x 4 5−s x 4 1s−1 x 4 3s−1 x 4 5−s x 4
× 1 + 3 + 5 + ··· − 1 + 3 − 5 + ···
1 + x2 1 + x2 1 + x2 1 − x2 1 − x2 1 − x2
On certain trigonometrical sums and their applications in the theory of numbers 245

if s is of the form 4k + 3. In the case in which s = 1 we have


1 3 5
!
− 41 x4 x4 x4
δ2′ (n)xn
P
= x 1 + 3 + 5 + ···
1 + x2 1 + x2 1 + x2
(12.65) 1 3 5
!
− 14 x4 x4 x4
= x 1 − 3 + 5 − ··· .
1 + x2 1 + x2 1 + x2

It is easy to see that the principal results proved about e2s (n) in my former paper are also
true of e′2s (n), and in particular that

e′2s (n) = 0

when s = 1, 2, 3 or 4, and
′ ′
r2s (n) ∼ δ2s (n)
for all values of s.

13. It follows from (12.62) that, if s is of the form 4k + 2, then

(1−s + 3−s + 5−s + · · ·)δ2s



(n)

is the coefficient of xn in
1 3
!
2( 14 π)s − 1 s 1s−1 x 2 2s−1 x 3s−1 x 2
(13.1) x 4 + + + ··· .
(s − 1)! 1−x 1 − x2 1 − x3

Similarly from (12.63) and (12.64) it follows that, if s is an odd integer greater than 1, then
′ (n) is the coefficient of xn in
(1−s − 3−s + 5−s − · · ·)δ2s
1 1 3
!
4( 18 π)s − 1 s 1s−1 x 4 2s−1 x 2 3s−1 x 4
(13.2) x 4 1 + + 3 + ··· .
(s − 1)! 1 + x2 1+x 1 + x2

Now by applying our main formulæ to (12.61) and (13.1) and (13.2) we obtain:

( 12 π)s
′ (n)
δ2s = (n + 41 s)s−1
(13.3) (s − 1)!
{1−s c1 (n + 14 s) + 3−s c3 (n + 14 s) + 5−s c5 (n + 14 s) + · · ·}

if s is a multiple of 4;

( 21 π)s
′ (n) =
δ2s (n + 14 s)s−1
(13.4) (s − 1)!
{1−s c1 (2n + 12 s) + 3−s c3 (2n + 21 s) + 5−s c5 (2n + 12 s) + · · ·}
246 Paper 21

if s is twice an odd number; and


( 12 π)s

δ2s (n) = (n + 41 s)s−1
(13.5) (s − 1)!
{1−s c1 (4n + s) − 3−s c3 (4n + s) + 5−s c5 (4n + s) − · · ·}
if s is an odd number greater than 1.
Since the coefficient of xn in (1 + x + x3 + · · ·)2 is that of x4n+1 in
1 4
2
2 + x + x + ··· ,

it follows from (11.51) that


π
(13.6) {c1 (4n + 1) − 31 c3 (4n + 1) + 15 c5 (4n + 1) − · · ·}.
r2′ (n) = δ2′ (n) =
4
This result however depends on the fact that the Dirichlet’s series for 1/η(s) is convergent
when s = 1.
′ (n) may be arrived at by another method.
14. The preceding formulæ for σs (n), δ2s (n), δ2s
We understand by
sin nπ
(14.1)
k sin(nπ/k)
the limit of
sin xπ
k sin(xπ/k)
when x → n. It is easy to see that, if n and k are positive integers, and k odd, then (14.1)
is equal to 1 if k is a divisor of n and to 0 otherwise.
When k is even we have (with similar conventions)
sin nπ
(14.2) = 1 or 0
k tan(nπ/k)
according as k is a divisor of n or not. It follows that
(   !
sin nπ sin nπ
σs−1 (n) = ns−1 1−s + 2−s
sin nπ tan 12 nπ
(14.3) ! ! )
sin nπ sin nπ
+3−s + 4−s + ··· .
sin 31 nπ tan 14 nπ
′ (n) we find that
Similarly from the definitions of δ2s (n) and δ2s
π s ns−1
  
−s −s −s −s −s sin nπ
(14.4) {1 + (−3) + 5 + (−7) + · · ·}δ2s (n) = 1
(s − 1)! sin nπ
! ! ! )
sin nπ sin nπ sin nπ
+2−s +3−s + 4−s + ···
sin( 21 nπ + 12 sπ) sin( 13 nπ + sπ) sin( 14 nπ + 32 sπ)
On certain trigonometrical sums and their applications in the theory of numbers 247

if s is an integer greater than 1;


(  ! ! )
sin nπ 1 sin nπ 1 sin nπ
r2 (n) = δ2 (n) = 4 − 3 + 5 − ···
sin nπ sin 13 nπ sin 51 nπ
(14.5) ( ! ! ! )
1 sin nπ 1 sin nπ 1 sin nπ
=4 2 − 4 + 6 − ··· ;
cos 12 nπ cos 14 nπ cos 16 nπ

( 21 π)s
(1−s + 3−s + 5−s + · · ·)δ2s
′ (n) = (n + 14 s)s−1
(s − 1)!
( ! !
sin(n + 41 s)π sin(n + 41 s)π
(14.6) 1−s + 3−s
sin(n + 41 s)π sin 13 (n + 41 s)π
! )
−s sin(n + 14 s)π
+5 + ···
sin 15 (n + 41 s)π
if s is a multiple of 4;
( 12 π)s
(1−s + 3−s + 5−s + · · ·)δ2s
′ (n) = (n + 14 s)s−1
(s − 1)!
( ! !
−s sin(2n + 12 s)π sin(2n + 12 s)π
(14.7) 1 + 3−s
sin(2n + 12 s)π sin 31 (2n + 12 s)π
! )
−s sin(2n + 21 s)π
+5 + ···
sin 15 (2n + 21 s)π
if s is twice an odd number;
(14.8)
( 12 π)s
(1−s − 3−s + 5−s − · · ·)δ2s
′ (n) = (n + 14 s)s−1
(s − 1)!
(   !
sin(4n + s)π sin(4n + s)π
1−s − 3−s
sin(4n + s)π sin 13 (4n + s)π
! )
−s sin(4n + s)π
+5 − ···
sin 15 (4n + s)π
if s is an odd number greater than 1; and
  !
′ ′ sin(4n + 1)π 1 sin(4n + 1)π
r2 (n) = δ2 (n) = − 3
sin(4n + 1)π sin 31 (4n + 1)π
(14.9) ! )
sin(4n + 1)π
+ 15 − ··· .
sin 51 (4n + 1)π
248 Paper 21

In all these equations the series on the right hand are finite Dirichlet’s series and therefore
absolutely convergent.
But the series (14.3) is (as easily shewn by actual multiplication) the product of the two
series
1−s c1 (n) + 2−s c2 (n) + · · ·
and
ns−1 (1−s + 2−s + 3−s + · · ·).
We thus obtain an alternative proof of the formulæ (7.5). Similarly taking the previous
expression of δ2s (n). viz. the right-hand side of (11.6), and multiplying it by the series

1−s + (−3)−s + 5−s + (−7)−s + · · ·

we can shew that the product is actually the right-hand side of (14.4). The formulæ for
′ (n) can be disposed off similarly.
δ2s

15. The formulæ which I have found are closely connected with a method used for another
purpose by Mr. Hardy and myself∗ . The function
X
(15.1) (1 + 2x + 2x4 + 2x9 + · · ·)2s = r2s (n)xn

has every point of the unit circle as a singular point. If x approaches a “rational point”
exp(−2pπi/q) on the circle, the function behaves roughly like

π s (ωp,q )s
(15.2) ,
{−(2pπi/q) − log x}s

where ωp,q = 1, 0, or −1 according as q is of the form 4k + 1, 4k + 2 or 4k + 3, while if q is


of the form 4k then ωp,q = −2i or 2i according as p is of the form 4k + 1 or 4k + 3.
Following the argument of our paper referred to, we can construct simple functions of x
which are regular except at one point of the circle of convergence, and there behave in a
manner very similar to that of the function (15.1); for example at the point exp(−2pπi/q)
such a function is

π s (ωp,q )s X s−1 2npπi/q n
(15.3) n e x .
(s − 1)!
1

The method which we used, with particular reference to the function

1 X
(15.4) = p(n)xn ,
(1 − x)(1 − x2 )(1 − x3 ) · · ·

“Asymptotic formulæ in Combinatory Analysis”, Proc. London Math. Soc., Ser.2, Vol. XVII, 1918, pp.
75 – 115 [No.36 of this volume].
On certain trigonometrical sums and their applications in the theory of numbers 249

was to approximate to the coefficients by means of a sum of a large number of the coefficients
of these auxiliary functions. This method leads, in the present problem, to formulæ of the
type
1
r2s (n) = δ2s (n) + O(n 2 s ),
the first term on the right-hand side presenting itself precisely in the form of the series
(11.11) etc.
It is a very interesting problem to determine in such cases whether the approximate formula
gives an exact representation of such an arithmetical function. The results proved here shew
that, in the case of r2s (n), this is in general not so. The formula represents not r2s (n) but
(except when s = 1) its dominant term δ2s (n), which is equal to r2s (n) only when s = 1, 2, 3
or 4. When s = 1 the formula gives 2δ2 (n)∗ .

16. We shall now consider the sum


(16.1) σs (1) + σs (2) + · · · + σs (n).

Suppose that
X  sin{(2n + 1)πλ/r} 
1
Tr (n) = 2 −1 ,
sin(πλ/r)
λ
(16.2)
X sin{(2n + 1)πλ/r}
1
Ur (n) = 2 ,
sin(πλ/r)
λ

where λ is prime to r and does not exceed r, so that

Tr (n) = cr (1) + cr (2) + · · · + cr (n)

and
Ur (n) = Tr (n) + 21 φ(r),
where φ(n) is the same as in §9. Since cr (n) = O(1) as r → ∞, it follows that

(16.21) Tr (n) = O(1), Ur (n) = O(r),

as r → ∞. It follows from (7.5) that, if s > 0, then

σ−s (1) + σ−s (2) + · · · + σ−s (n) = ζ(s + 1)


(16.3)
 
T2 (n) T3 (n) T4 (n)
n + s+1 + s+1 + s+1 + · · · .
2 3 4

The method is also applicable to the problem of the representation of a number by the sum of an odd
number of squares, and gives an exact result when the number of squares is 3, 5, or 7. See G.H. Hardy, “On
the representation of a number as the sum of any number of squares, and in particular of five or seven,”
Proc. London Math. Soc. (Records of proceedings at meetings, March 1918). A fuller account of this paper
will appear shortly in the Proceedings of the National Academy of Sciences (Washington, D.C.) [loc. cit.,
Vol.IV, 1918, 189-193].
250 Paper 21

Since

X φ(n) ζ(s)
=
1
ν s+1 ζ(s + 1)
if s > 1, (16.3) can be written as
(16.31)
σ−s (1)σ−s (2) + · · · + σ−s (n) = ζ(s + 1)
 
U2 (n) U3 (n) U4 (n)
1
n + 2 + s+1 + s+1 + s+1 + · · · − 12 ζ(s),
2 3 4
if s > 1. Similarly from (8.3), (8.4) and (11.51) we obtain

d(1) + d(2) + · · · + d(n)


(16.4)
= − 21 T2 (n) log 2 − 13 T3 (n) log 3 − 14 T4 (n) log 4 − · · · ,

d(1) log 1 + d(2) log 2 + · · · + d(n) log n


(16.5)
= 21 T2 (n){2γ log 2 − (log 2)2 } + 31 T3 (n){2γ log 3 − (log 3)2 } + · · · ,

(16.6) r2 (1) + r2 (2) + · · · + r2 (n) = π{n − 13 T3 (n) + 15 T5 (n) − 71 T7 (n) + · · ·}.

Suppose now that


X 2πλ 4πλ 2nπλ

s
Tr,s (n) = 1 cos + 2s cos + · · · + ns cos ,
r r r
λ

where λ is prime to r and does not exceed r, so that

Tr,s (n) = 1s cr (1) + 2s cr (2) + · · · + ns cr (n).

Then it follows from (7.5) that

σs (1) + σs (2) + · · · + σs (n)


(16.7)
 
s s s T2,s (n) T3,s (n) T4,s (n)
= ζ(s + 1) (1 + 2 + · · · + n ) + s+1 + s+1 + s+1 + · · ·
2 3 4
if s > 0. Putting s = 1 in (16.3) and (16.7), we find that

(n − 1)σ−1 (1) + (n − 2)σ−1 (2) + · · · + (n − n)σ−1 (n)


(16.8) π 2 n(n − 1) ν2 (n) ν3 (n) ν4 (n)
 
= + + + + ··· ,
6 2 22 32 42
where
X  sin2 (πnλ/r) 
1
νr (n) = 2 −n ,
λ
sin2 (πλ/r)
On certain trigonometrical sums and their applications in the theory of numbers 251

λ being prime to r and not exceeding r.


It has been proved by Wigert ∗ , by less elementary methods, that the left-hand side of
(16.8) is equal to
√ X∞
π2 2 1 1 n σ−1 (ν) p
(16.9) n − 2 n(γ − 1 + log 2nπ) − 24 + √ J1 {4π (νn)},
12 2π 1 ν

where J1 is the ordinary Bessel’s function.

17. We shall now find a relation between the functions (16.1) and (16.3) which enables us
to determine the behaviour of the former for large values of n. It is easily shewn that this
function is equal to
√ √ √
n  h n is  X n h i √ X n
s n
X
s s s
(17.1) 1 + 2 + 3 + ··· + + ν − [ n] ν s.
ν=1
ν ν=1
ν ν=1

Now
(k + 12 )s+1
1s + 2s + · · · + ks = ζ(−s) + + O(ks−1 )
s+1
for all values of s, it being understood that

(k + 12 )s+1
ζ(−s) +
s+1

denotes γ + log(k + 21 ) when s = −1. Let


hni n √ √
= − 21 + ǫν , [ n] = t = n − 1
2 + ǫ.
ν ν
Then we have
ns−1
h n is  
s s 1  n s+1  n s
1 + 2 + ··· + = ζ(−s) + + ǫν +O
ν s+1 ν ν ν s−1

and hni
νs= nν s−1 − 12 ν s + ǫν ν s .
ν
It follows from these equations and (17.1) that
t 
X 1  n s+1
σs (1) + σs (2) + · · · + σs (n) = ζ(−s) + + nν s−1
(17.2) s+1 ν
ν=1
s − (√n + ǫ)ν s + O ns−1
 o
n s

+ǫν ν + ǫ ν ν ν s−1 .

Acta Mathematica, Vol.XXXVII, 1914, pp. 113 – 140(p.140).
252 Paper 21

Changing s to −s in (17.2) we have


(17.21) ns {σ−s (1) + σ−s (2) + · · · + σ−s (n)}
t 
X nν s−1  n s+1  n s
= ns ζ(s) + + + ǫν ν s + ǫν
1−s ν ν
ν=1
 s+1 
√  n s ν
− ( n + ǫ) +O .
ν n
It follows that
(17.3) ns {σ−s (1) + σ−s (2) + · · · + σ−s (n)} − {σs (1) + σs (2) + · · · + σs (n)}
t 
X s  n s+1 s √
= ns ζ(s) − ζ(−s) + + nν s−1 + ( n + ǫ)ν s
ν=1
1+s ν 1−s
 s−1 s+1


 n s n ν
−( n + ǫ) +O s−1
+ .
ν ν n

Suppose now that s > 0. Then, since ν varies from 1 to t, it is obvious that
ν s+1 ns−1
< s−1
n ν
and so
ν s+1 ns+1
   
O =O .
n ν s−1
. Also
t
X √
{ns ζ(s) − ζ(−s)} = ( n − 1
2 + ǫ){ns ζ(s) − ζ(−s)};
ν=1
t
X s  n s+1 sns+1 ns+1 √ 1
= ζ(1 + s) − ( n + ǫ)−s + O(n 2 s );
1+s ν 1+s s+1
ν=1
t
s ns n √ 1
( n + ǫ)s + O(n 2 s );
X
nν s−1 = ζ(1 − s) +
ν=1
1−s 1−s 1−s
t √
√ √ ( n + ǫ)2+s 1
+ O(n 2 s );
X
s
( n + ǫ)ν = ( n + ǫ)ζ(−s) +
ν=1
1+s
t
√  n s √ ns √ 1
( n + ǫ)2−s + O(n 2 s );
X
( n + ǫ) = ns ( n + ǫ)ζ(s) +
ν 1−s
ν=1

and
t
ns−1
X  
O = O(m),
ν s−1
ν=1
On certain trigonometrical sums and their applications in the theory of numbers 253

where
1
(17.4) m = n 2 s (s < 2), m = n log n(s = 2), m = ns−1 (s > 2).

It follows that the right-hand side of (17.3) is equal to


√ √
sn1+s sn 1 s ( n + ǫ)2+s − ns+1 ( n + ǫ)−s
ζ(1 + s) + ζ(1 − s) − 2 n ζ(s) +
1+s 1−s 1+s
√ √
n( n + ǫ) − n ( n + ǫ)2−s
s s
+ + O(m).
1−s
But
√ √
( n + ǫ)2+s − ns+1 ( n + ǫ)−s 1 1
= 2ǫn 2 (1+s) + O(n 2 s );
1+s
√ √
n( n + ǫ) − ns ( n + ǫ)2−s
s 1 1
= −2ǫn 2 (1+s) + O(n 2 s ).
1−s
It follows that
σs (1) + σs (2) + · · · + σs (n) = ns {σ−s (1) + σ−s (2) + · · · + σ−s (n)}
(17.5) sn1+s sn
− ζ(1 + s) + 12 ns ζ(s) − ζ(1 − s) + O(m)
1+s 1−s
if s > 0, m being the same as in (17.4). If s = 1, (17.5) reduces to

(n − 1)σ−1 (1) + (n − 2)σ−1 (2) + · · · + (n − n)σ−1 (n)


(17.6) 2 √
= π12 n2 − 12 n(γ − 1 + log 2nπ) + O( n)∗ .

From (16.2) and (17.5) it follows that

n1+s
(17.7) σs (1) + σs (2) + · · · + σs (n) = ζ(1 + s) + 21 ns ζ(s)
1+s
 
sn s T2 (n) T3 (n) T4 (n)
+ ζ(1 − s) + n ζ(1 + s) + s+1 + s+1 + · · · + O(m),
s−1 2s+1 3 4
for all positive values of s. If s > 1, the right-hand side can be written as
ns
ζ(1 − s) + ns ζ(1 + s)
s−1
(17.8)  
n 1 U2 (n) U3 (n) U4 (n)
+ + s+1 + s+1 + s+1 + · · · + O(m).
1+s 2 2 3 4

This result has been proved by Landau. See his report on Wigert’s memoir in the Göttingische gelehrte

Anzeigen, 1915, pp. 377 – 414 (p.402). Landau has also, by a more transcendental method, replaced O( n)
2
by O(n 5 ) (loc.cit., p.414).
254 Paper 21

Putting s = 1 in (17.7) we obtain

π2 2 1
σ1 (1) + σ1 (2) + · · · + σ1 (n) = n + 2 n(γ − 1 + log 2nπ)
12
(17.9)
π 2 n T2 (n) T3 (n) T4 (n) √
 
+ + + + · · · + O( n).
6 22 32 42

Additional note to §7 (May 1, 1918).


From (7.2) it follows that
∞ ∞
1 X X
{1−s σ1−r (1) + 2−s σ1−r (2) + · · ·} = 1−s m−r cm (1) + 2−s m−r cm (2) + · · · ,
ζ(r)
1 1

or
∞ ∞
ζ(s)ζ(r + s − 1) X X cm (n)
= r ns
,
ζ(r) 1 1
m
from which we deduce
X cm (1) cm (2) cm (3)
ζ(s) µ(δ)δ′1−s = + + + ···,
1s 2s 3s
δ

δ being a divisor of m and δ′ its conjugate. The series on the right-hand side is convergent
for s > 0 (except when m = 1, when it reduces to the ordinary series for ζ(s)).
When s = 1, m > 1 we have to replace the left-hand side by its limit as s → 1. We find
that

(18) cm (1) + 21 cm (2) + 13 cm (3) + · · · = −Λ(m),

Λ(m) being the well-known arithmetical function which is equal to log p if m is a power of
a prime p and to zero otherwise.
Some definite integrals

Proceedings of the London Mathematical Society, 2, XVII, 1918,


Records for 17 Jan. 1918

Typical formulæ are:


Z∞
enix dx (2 cos 12 n)α+β−2 1 n(β−α)i
= e2 (or 0), (1)
Γ(α + x)Γ(β − x) Γ(α + β − 1)
−∞

Z∞
Γ(α + x) nix 2πi(2 sin 21 N )β−α−1 −nαi+ 1 (π−N )(β−α−1)i
e dx = ± e 2 (or 0), (2)
Γ(β + x) Γ(β − α)
−∞

Z∞
Γ(α + x)Γ(β − x)enix dx
−∞

2πiΓ(α + β) 1 n(β−α)i h
kπ(α+β)i −kπ(α+β)i
i
= e 2 ǫ n (β)e − ǫ n (−α)e . (3)
(2 sin 12 N )α+β

Here n is real, n = 2kπ + N (0 ≤ N < 2π) in (2), and n = (2k − 1)π + N (0 ≤ N < 2π)
in (3). In (1) the zero value is to be taken if |n| ≥ π, the non-zero value otherwise. In (2)
α must be complex: the zero value is to be taken if n and I(α) have the same sign, the
positive sign if n ≥ 0 and I(α) < 0, and the negative sign if n ≤ 0 and I(α) > 0. In (3) α
and β must both be complex; and ǫn (ζ) is 0,1, or −1 according as (i) π − n and J(ζ) have
the same sign, (ii) n ≤ π and I(ζ) < 0, (iii) n ≥ π and I(ζ) > 0.
The convergence conditions are, in general,
(1) R(α + β) > 1, (2) R(α − β) < 0, (3)R(α + β) < 1.
But there are certain special cases in which a more stringent condition is required.
A formula of a different character, deduced from (1), is

Z∞ ! 1 (α+β)
Jα+x (λ) Jβ−x (µ) nix 2 cos 21 n 2

e dx =
λα+x µβ−x Ω
−∞
q
1
e 2 n(β−α)i Jα+β { (2Ω cos 21 n)} (or 0).

Here
1 1
Ω = λ2 e 2 ni + µ2 e− 2 ni ;
256 Paper 22

the zero value is to be taken if |n|+ ≥ π, the non-zero value otherwise; and the condition
of convergence is, in general, that

R(α + β) > −1.

The formulæ include a large number of interesting special cases, such as


Z∞
dx 2α+β−2
= ,
Γ(α + x)Γ(β − x) Γ(α + β − 1)
−∞
Z∞
sin πxdx 22k−1 π
= (−1)k ,
x(x2 − 12 )(x2 − 22 ) · · · (x2 − k2 ) (2k)!
0
Z∞
Jα+x (λ)Jβ−x (λ)dx = Jα+β (2λ).
−∞

The formula
Z∞
dx
Γ(α + x)Γ(β − x)Γ(γ + x)Γ(δ − x)
−∞

Γ(α + β + γ + δ − 3)
= ,
Γ(α + β − 1)Γ(β + γ − 1)Γ(γ + δ − 1)Γ(δ + α − 1)

may also be mentioned : it holds, in general, if

R(α + β + γ + δ) > 3.

A fuller account of these formulæ will be published in the Quarterly Journal of Mathemat-
ics∗


[See No.27 of this volume]
Some definite integrals
Journal of the Indian Mathematical Society, XI, 1919, 81 – 87
I have shewn elsewhere∗ that the definite integrals
Z∞
cos πtx −πwx2
φw (t) = e dx,
cosh πx
0

Z∞
sin πtx −πwx2
ψw (t) = e dx
sinh πx
0
can be evaluated in finite terms if w is any rational multiple of i.
In this paper I shall shew, by a much simpler method, that these integrals can be evaluated
not only for these values but also for many other values of t and w.
Now we have
Z ∞Z ∞
cos 2πxz 2
φw (t) = 2 cos πtxe−πwx dxdz
0 0 cosh πz
1 2 ′
e− 4 πt w ∞
cosh πtxw′ −πx2 w′
Z
= √ e dx
w 0 cosh πx
where w′ stands for 1/w.
It follows that
1 1 2 ′
φw (t) = √ e− 4 πt w φw′ (itw′ ). (1)
w
Again
1 1 2 ′
φw (t + w) = √ e− 4 π(t+w) w
w
Z ∞
cosh(πtx/w) cosh πx + sinh πtx/w sinh πx −πx2 /w
× e dx
0 cosh πx
1 1 2
= √ e− 4 π(t+w) /w
w
Z ∞Z ∞
√ 1 πt2 /w
 
1 sin 2πxz πtx −πx2 /w
× 2 we 4 +2 sinh e dxdz
0 0 sinh πz w

1 1 2
= √ e− 4 π(t+w) /w
w

Messenger of Mathematics, Vol,44, 1915, pp. 75 – 85 [No.12 of this volume].
258 Paper 23


√ √ ∞ 
sin πtx −πwx2
Z
1 2 /w 1 2 /w
× 1
2 we 4 πt + we 4 πt e dx .
0 sinh πx

In other words
1 2 /w 1 2/w
e 4 πt { 12 + ψw (t)} = e 4 π(t+w) φw (t + w). (2)

It is obvious that
)
φw (t) = φw (−t)
. (3)
ψw (t) = −ψw (−t)

From (1), (2) and (3) we easily find that


  
i 1 2 it
1
2 + ψw (t + i) = √ e− 4 πt /w 1
2 − ψw′ +i . (4)
w w
It is easy to see that
1 i 1
φw (i) = √ ; ψw (i) = √ ; φw (w) = 21 e− 4 πw ;
2 w 2 w
 
− 14 πw 1 i 1
1
2 − ψw (w) = e φw (0); φw (w ± i) = √ + e− 4 πw ;
2 w 2
i 1
ψw (w ± i) = 1
2 ± √ e− 4 πw ; φw ( 12 w) + ψw ( 12 w) = 21 .
2 w
Again we see that
1 1 2
φw (t + i) + φw (t − i) = √ e− 4 πt /w ; (5)
w
and
i 1 2
ψw (t + i) − ψw (t − i) = √ e− 4 πt /w . (6)
w

From (1) and (5) we deduce that


1 2 /w 1 2 /w 1 2 /w
e 4 π(t+w) φw (t + w) + e 4 π(t−w) φw (t − w) = e 4 πt . (7)

Similarly from (4) and (6) we obtain


1 2 /w 1 2
e 4 π(t+w) − ψw (t + w) = e 4 π(t−w) /w { 21 + ψw (t − w)}.
1
2 (8)

It is easy to deduce from (5) that if n is a positive integer, then

φw (t) + (−1)n+1 φw (t ± 2ni)


Some definite integrals 259

n 1 2 /w 1 2 /w 1 2 /w
o
= √1 e− 4 π(t±i) − e− 4 π(t±3i) + e− 4 π(t±5i) − · · · to n terms . (9)
w

Similarly from (6) we have

ψw (t) − ψw (t ± 2ni)
n 1 2 1 2 1 2
o
= ∓ √iw e− 4 π(t+i) /w + e− 4 π(t+3i) /w + e− 4 π(t+5i) /w + · · · to n terms .

(10)

Again from (7) we have


1 2 /w 1 2 /w
e 4 πt φw (t) + (−1)n+1 e 4 π(t+2nw) φw (t + 2nw)
1 1 1
π(t+w)2 /w π(t+3w)2 /w π(t+5w)2 /w
=e 4 −e 4 +e 4 − · · · to n terms; (11)

and from (8)


1 2 /w 1 2 /w
e 4 πt { 12 + ψw (t)} + (−1)n+1 e 4 π(t+2nw) { 12 + ψw (t + 2nw)}
1 2 /w 1 2 /w 1 2 /w
= e 4 π(t+2w) − e 4 π(t+4w) + e 4 π(t+6w) − · · · to n terms. (12)

Now, combining (9) and (11), we deduce that, if m and n are positive integers and s =
t + 2mw ± 2ni, then
1
φw (s) + (−1)(m+1)(n+1) e− 2 πm(s+t) φw (t)
1 2
n 1 2 1 2 1 2
o
= e− 4 πs /w e 4 π(s−w) /w − e 4 π(s−3w) /w + e 4 π(s−5w) /w − · · · to m terms

(−1)(m+1)(n+1) − 1 πm(s+t)
+ √ e 2
w
n 1 2 1 2 1 2
o
× e− 4 π(t±i) /w − e− 4 π(t±3i) /w + e− 4 π(t±5i) /w − · · · to n terms . (13)

Similarly, combining (10) and (12), we obtain


1
1
2 − ψw (s) + (−1)mn+m+1 e− 2 πm(s+t) { 21 − ψw (t)}
1 2
n 1 2 1 2 1 2
o
= e− 4 πs /w e 4 π(s−2w) /w − e 4 π(s−4w) /w + e 4 π(s−6w) /w − · · · to m terms
i 1
±(−1)mn+m+1 √ e− 2 πm(s+t)
w
n 1 2 1 2 1 2
o
× e− 4 π(t±i) /w + e− 4 π(t±3i) /w + e− 4 π(t±5i) /w + · · · to n terms , (14)

where s and t have the same relation as in (13).


260 Paper 23

Suppose now that s = t in (13) and (14). Then we see that, if w = in/m, then

φw (t){1 + (−1)(m+1)(n+1) e−πmt }


1 2
n 1 2 1 2 1 2
o
= e− 4 πt /w e 4 π(t−w) /w − e 4 π(t−3w) /w + e 4 π(t−5w) /w − · · · to m terms
(m+1)(n+1)
n 1 2 1 2
o
+ (−1) √w e−πmt e− 4 π(t−i) /w − e− 4 π(t−3i) /w + · · · to n terms ; (15)

{ 21 − ψw (t)}{1 + (−1)mn+m+1 e−πmt }


1 2
n 1 2 1 2
o
= e− 4 πt /w e 4 π(t−2w) /w − e 4 π(t−4w) /w + · · · to m terms
n 1 2 1 2
o
+(−1)mn+m √iw e−πmt e− 4 π(t−i) /w − e− 4 π(t−3i) /w + · · · to n terms . (16)

where w should be taken as r
1
πi n
e 4 .
m
In (15) and (16) there is no loss of generality in supposing that one of the two numbers m
and n is odd.
Now equating the real and imaginary parts in (15), we deduce that, if m and n are positive
integers of which one is odd, then
Z ∞
πmx2
 
cos 2tx
2 cosh nt cos dx
0 cosh πx n
= [cosh{(1 − n)t} cos(πm/4n) − cosh{(3 − n)t} cos(9πm/4n) + · · · to n terms]
nt2
r       
n 1 π πn
+ cosh 1− nt cos − +
m m 4 πm 4m
nt2
     
3 π 9πn
− cosh 1− nt cos − + + · · · to m terms ; (17)
m 4 πm 4m
and

πmx2
 
cos 2tx
Z
2 cosh nt sin dx
0 cosh πx n
= −[cosh{(1 − n)t} sin(πm/4n) − cosh{(3 − n)t} sin(9πm/4n)
+ cosh{(5 − n)t} sin(25π/4n) − · · · to n terms]
nt2
r       
n 1 π πn
+ cosh 1− nt sin − +
m m 4 πm 4m
nt2
     
3 π 9πn
− cosh 1− nt sin − + + · · · to n terms . (18)
m 4 πm πm
Some definite integrals 261

Equating the real and imaginary parts in (16), we can find similar expressions for the
integrals
Z ∞ Z ∞
πmx2 πmx2
   
sin tx sin tx
sin dx, cos dx.
0 sinh πx n 0 sinh πx n
From these formulæ we can evaluate a number of definite integrals, such as
Z ∞ √
cos 2πtx 2 1 + 2 sin πt2
cos πx dx = √ ,
0 cosh πx 2 2 cosh πt
Z ∞ √
cos 2πtx 2 −1 + 2 cos πt2
sin πx dx = √ ,
0 cosh πx 2 2 cosh πt
Z ∞
sin 2πtx cosh πt − cos πt2
cos πx2 dx = ,
0 sinh πx 2 sinh πt
Z ∞
sin 2πtx sin πt2
sin πx2 dx = ,
0 sinh πx 2 sinh πt

and so on.
Again supposing that s = −t in (13), we deduce that if t = mw ± ni, where m and n are
positive integers of which one at least is odd, then
1 2
n 1 2 1 2
o
φw (t) = 12 e− 4 πt /w e 4 π(t−w) /w − e 4 π(t−3w) /w + · · · to m terms
1 n − 1 π(t∓i)2 /w 1 2
o
+ √ e 4 − e− 4 π(t∓3i) /w + · · · to n terms . (19)
2 w

This formula is not true when both m and n are even.


If t = mw ± ni, where m and n are both even, then
1 1 1
φw (t) + (−1)(1+ 2 m)(1+ 2 n) e− 4 πmt φw (0)
1 2
n 1 2 1 2
o
= e− 4 πt /w e 4 π(t−w) /w − e 4 π(t−3w) /w + · · · to 1
2m terms
1 1
(−1)1+ 2 m)(1+ 2 n) − 1 πmt n 1 π/w 9 25
o
+ √ e 4 e4 − e 4 π/w + e 4 π/w − · · · to 1
2n terms . (20)
w

This is easily obtained by putting t = 0 and then changing s to t in (13). Similarly from
(14) we deduce that if t = mw ± ni, where m and n are both even, or both odd, or m is
even and n is odd, then
1 2
n 1 2 1 2
o
ψw (t) = − 12 e− 4 πt /w e 4 π(t−2w) /w − e 4 π(t−4w) /w + · · · to m terms
i n − 1 π(t∓i)2 /w 1 2
o
± √ e 4 + e− 4 π(t∓3i) /w + · · · to n terms . (21)
2 w
262 Paper 23

If t = mw ± ni, where m is odd and n is even, then


1 1
1
2 − ψw (t) + {(−1)1+ 4 (m−1)(n+2) e− 4 π{(m−1)t+mw} φw (0)
1 2
n 1 2 1 2
o
= e− 4 πt /w e 4 π(t−2w) /w − e 4 π(t−4w) /w · · · + to 12 (m − 1) terms
1 i 1
±(−1)1+ 4 (m−1)(n+2) √ e− 4 π(m−1)(t+w)
w
n 1 2 1 2
o
× e− 4 π(w±i) /w − e− 4 π(w±3i) /w + · · · to 1
2n terms . (22)

This is obtained by putting t = w in (14). A number of definite integrals such as the


following can be evaluated with the help of the above formulæ:
Z ∞ ( )
cos πtx −π(t+i)x2 1 + i − 1 πt i
e dx = √ e 4 1− p ,
0 cosh πx 2 2 (t + i)
Z ∞ 1
sin πtx −π(t+i)x2 1 1+i e− 4 πt
e dx = − √ ·p ,
0 sinh πx 2 2 2 (t + i)

and so on.
A proof of Bertrand’s postulate

Journal of the Indian Mathematical Society, XI, 1919, 181 – 182

1. Landau in his Handbuch, pp. 89 – 92, gives a proof of a theorem the truth of which was
conjectured by Bertrand: namely that there is at least one prime p such that x < p ≤ 2x,
if x ≥ 1. Landau’s proof is substantially the same as that given by Tschebyschef. The
following is a much simpler one.
Let ν(x) denote the sum of the logarithms of all the primes not exceeding x and let
1 1
Ψ(x) = ν(x) + ν(x 2 ) + ν(x 3 ) + · · · , (1)

log[x]! = Ψ(x) + Ψ( 12 x) + Ψ( 31 x) + · · · , (2)

where [x] denotes as usual the greatest integer in x.


From (1) we have
√ 1 1
Ψ(x) − 2Ψ( x) = ν(x) − ν(x 2 ) + ν(x 3 ) − · · · , (3)

and from (2)

log[x]! − 2 log[ 12 x]! = Ψ(x) − Ψ( 12 x) + Ψ( 31 x) − · · · . (4)

Now remembering that ν(x) and Ψ(x) are steadily increasing functions, we find from (3)
and (4) that

Ψ(x) − 2Ψ( x) ≤ ν(x) ≤ Ψ(x); (5)

and

Ψ(x) − Ψ( 12 x) ≤ log[x]! − 2 log[ 21 x]! ≤ Ψ(x) − Ψ( 12 x) + Ψ( 31 x). (6)

But it is easy to see that

log Γ(x) − 2 log Γ( 12 x + 21 ) ≤ log[x]! − 2 log[ 12 x]!


≤ log Γ(x + 1) − 2 log Γ( 12 x + 12 ). (7)

Now using Stirling’s approximation we deduce from (7) that

log[x]! − 2 log[ 21 x]! < 43 x, if x > 0; (8)

and

log[x]! − 2 log[ 12 x]! > 32 x, if x > 300. (9)


264 Paper 24

It follows from (6), (8) and (9) that

Ψ(x) − Ψ( 12 x) < 43 x, if x > 0; (10)

and

Ψ(x) − Ψ( 12 x) + Ψ( 31 x) > 32 x, if x > 300. (11)

Now changing x to 12 x, 41 x, 18 x, . . . in (10) and adding up all the results, we obtain

Ψ(x) < 23 x, if x > 0. (12)

Again we have

Ψ(x) − Ψ( 21 x) + Ψ( 31 x) ≤ ν(x) + 2Ψ( x) − ν( 21 x) + Ψ( 31 x)

< ν(x) − ν( 12 x) + 21 x + 3 x, (13)

in virtue of (5) and (12).


It follows from (11) and (13) that

ν(x) − ν( 12 x) > 61 x − 3 x, if x > 300. (14)

But it is obvious that 61 x − 3 x ≥ 0, if x ≥ 324. Hence

ν(2x) − ν(x) > 0, if x ≥ 162. (15)

In other words there is at least one prime between x and 2x if x ≥ 162. Thus Bertrand’s
Postulate is proved for all values of x not less than 162; and, by actual verification, we find
that it is true for smaller values.

2. Let π(x) denote the number of primes not exceeding x. Then, since π(x) − π( 12 x) is the
number of primes between x and 12 x, and ν(x) − ν( 12 x) is the sum of logarithms of primes
between x and 12 x, it is obvious that

ν(x) − ν( 12 x) ≤ {π(x) − π( 21 x)} log x, (16)

for all values of x. It follows from (14) and (16) that


1 1 √
π(x) − π( 12 x) > ( 6 x − 3 x), if x > 300. (17)
log x
From this we easily deduce that

π(x) − π( 21 x) ≥ 1, 2, 3, 4, 5, . . . , if x ≥ 2, 11, 17, 29, 41, . . . , (18)

respectively.
Some properties of p(n), the number of partitions of n∗

Proceedings of the Cambridge Philosophical Society, XIX, 1919, 207 – 210

1. A recent paper by Mr Hardy and myself † contains a table, calculated by Major MacMa-
hon, of the values of p(n), the number of unrestricted partitions of n, for all values of n from
1 to 200. On studying the numbers in this table I observed a number of curious congruence
properties, apparently satisfied by p(n). Thus
(1) p(4), p(9), p(14), p(19), · · · ≡ 0 (mod 5),
(2) p(5), p(12), p(19), p(26), · · · ≡ 0 (mod 7),
(3) p(6), p(17), p(28), p(39), · · · ≡ 0 (mod 11),
(4) p(24), p(49), p(74), p(99), · · · ≡ 0 (mod 25),
(5) p(19), p(54), p(89), p(124), · · · ≡ 0 (mod 35),
(6) p(47), p(96), p(145), p(194), · · · ≡ 0 (mod 49),
(7) p(39), p(94), p(149), · · · ≡ 0 (mod 55),
(8) p(61), p(138), · · · ≡ 0 (mod 77),
(9) p(116), · · · ≡ 0 (mod 121),
(10) p(99), ··· ≡ 0 (mod 125).
From these data I conjectured the truth of the following theorem: if δ = 5a 7b 11c and
24λ ≡ 1 (mod δ) then

p(λ), p(λ + δ), p(λ + 2δ), · · · ≡ 0 (mod δ).

This theorem is supported by all the available evidence; but I have not yet been able to
find a general proof.
I have, however, found quite simple proofs of the theorems expressed by (1) and (2), viz.

(1) p(5m + 4) ≡ 0 (mod 5)

and

(2) p(7m + 5) ≡ 0 (mod 7).

From these

(5) p(35m + 19) ≡ 0 (mod 35)

follows at once as a corollary. These proofs I give in § 2 and § 3. I can also prove

(4) p(25n + 24) ≡ 0 (mod 25)



[See also Ramanujan’s posthumous paper “Congruence properties of partitions” in the Math. Zeitschrift,
[No.30 of this volume].

G. H. Hardy and S.Ramanujan, “Asymptotic formulæ in Combinatory Analysis,” Proc. London Math.
Soc., Ser. 2, Vol. XVII, 1918, pp. 75 – 115 (Table IV, pp. 114 – 115) [No.36 of this volume].
266 Paper 25

and

(6), p(49n + 47) ≡ 0 (mod 49),

but only in a more recondite way, which I sketch in § 4.

2. Proof of (1). We have

(11) x{(1 − x)(1 − x2 )(1 − x3 ) · · ·}4


= x(1 − 3x + 5x3 − 7x6 + · · ·)(1 − x − x2 + x5 + · · ·)
X 1 1
= (−1)µ+ν (2µ + 1)x1+ 2 µ(µ+1)+ 2 ν(3ν+1) ,

the summation extending from µ = 0 to µ = ∞ and from ν = −∞ to ν = ∞.


Now if
1 + 12 µ(µ + 1) + 12 ν(3ν + 1) ≡ 0 (mod 5),
then
8 + 4µ(µ + 1) + 4ν(3ν + 1) ≡ 0 (mod 5),
and therefore

(12) (2µ + 1)2 + 2(ν + 1)2 ≡ 0 (mod 5).

But (2µ + 1)2 is congruent to 0,1 or 4, and 2(ν + 1)2 to 0, 2, or 3. Hence it follows from
(12) that 2µ + 1 and ν + 1 are both multiplies of 5. That is to say, the coefficient of x5n in
(11) is a multiple of 5.
Again, all the coefficients in (1 − x)−5 are multiples of 5, except those of 1, x5 , x10 , . . . ,
which are congruent to 1: that is to say

1 1
5
≡ (mod 5),
(1 − x) 1 − x5
or
1 − x5
≡ 1 (mod 5).
(1 − x)5
Thus all the coefficients in

(1 − x5 )(1 − x10 )(1 − x15 ) · · ·


{(1 − x)(1 − x2 )(1 − x3 ) · · ·}5

(except the first) are multiples of 5. Hence the coefficient of x5n in

x(1 − x5 )(1 − x10 ) · · · 2


5 10
4 (1 − x )(1 − x ) · · ·
= x{(1 − x)(1 − x ) · · ·}
(1 − x)(1 − x2 )(1 − x3 ) · · · {(1 − x)(1 − x2 ) · · ·}5
Some properties of p(n), the number of partitions of n 267

is a multiple of 5. And hence, finally, the coefficient of x5n in


x
(1 − x)(1 − x2 )(1 − x3 ) · · ·
is a multiple of 5; which proves (1).

3. Proof of (2). The proof of (2) is very similar. We have


(13) x2 {(1 − x)(1 − x2 )(1 − x3 ) · · ·}6
= x2 (1 − 3x + 5x3 − 7x6 + · · ·)2
X 1 1
= (−1)µ+ν (2µ + 1)(2ν + 1)x2+ 2 µ(µ+1)+ 2 ν(ν+1) ,

the summation now extending from 0 to ∞ for both µ and ν. If


2 + 21 µ(µ + 1) + 21 ν(ν + 1) ≡ 0 (mod 7),
then 16 + 4µ(µ + 1) + 4ν(ν + 1) ≡ (mod 7),
(2µ + 1)2 + (2ν + 1)2 ≡ 0 (mod 7),
and 2µ + 1 and 2ν + 1 are both divisible by 7. Thus the coefficient of x7n in (13) is divisible
by 49.
Again, all the coefficients in
(1 − x7 )(1 − x14 )(1 − x21 ) · · ·
{(1 − x)(1 − x2 )(1 − x3 ) · · ·}7
(except the first) are multiples of 7. Hence (arguing as in § 2) we see that the coefficient of
x7n in
x2
(1 − x)(1 − x2 )(1 − x3 ) · · ·
is a multiple of 7; which proves (2). As I have already pointed out, (5) is a corollary.

4. The proofs of (4) and (6) are more intricate, and in order to give them I have to consider
a much more difficult problem. viz. that of expressing
p(λ) + p(λ + δ)x + p(λ + 2δ)x2 + · · ·
in terms of Theta-functions, in such a manner as to exhibit explicitly the common factors
of the coefficients, if such common factors exist. I shall content myself with sketching the
method of proof, reserving any detailed discussion of it for another paper.
It can be shewn that
(1 − x5 )(1 − x10 )(1 − x15 ) · · · 1
(14) 1 2 3 = 1 2
(1 − x )(1 − x )(1 − x ) · · ·
5 5 5 ξ −1 − x 5 − ξx 5
1 2 3 4
ξ −4 − 3xξ + x 5 (ξ −3 + 2xξ 2 ) + x 5 (2ξ −2 − xξ 3 ) + x 5 (3ξ −1 + xξ 4 ) + 5x 5
= ,
ξ −5 − 11x − x2 ξ 5
268 Paper 25

where
(1 − x)(1 − x4 )(1 − x6 )(1 − x9 ) · · ·
ξ= ,
(1 − x2 )(1 − x3 )(1 − x7 )(1 − x8 ) · · ·
the indices of the powers of x, in both numerator and denominator of ξ, forming two
arithmetical progressions with common difference 5. It follows that
5
(15) (1 − x5 )(1 − x10 )(1 − x15 ) · · · {p(4) + p(9)x + p(14)x2 + · · ·} = −5 .
ξ − 11x − x2 ξ 5
1 1 1 1 1
Again, if in (14) we substitute ωx 5 , ω 2 x 5 , ω 3 x 5 , and ω 4 x 5 , where ω 5 = 1, for x 5 , and
multiply the resulting five equations, we obtain
6
(1 − x5 )(1 − x10 )(1 − x15 ) · · ·

1
(16) 2 3
= −5 .
(1 − x)(1 − x )(1 − x ) · · · ξ − 11x − x2 ξ 5
From (15) and (16) we deduce
{(1 − x5 )(1 − x10 )(1 − x15 ) · · ·}5
(17) p(4) + p(9)x + p(14)x2 + · · · = 5 ;
{(1 − x)(1 − x2 )(1 − x3 ) · · ·}6
from which it appears directly that p(5m + 4) is divisible by 5.
The corresponding formula involving 7 is
{(1 − x7 )(1 − x14 )(1 − x21 ) · · ·}3
(18) p(5) + p(12)x + p(19)x2 + · · · = 7
{(1 − x)(1 − x2 )(1 − x3 ) · · ·}4
{(1 − x7 )(1 − x14 )(1 − x21 ) · · ·}7
+49x ,
{(1 − x)(1 − x2 )(1 − x3 ) · · ·}8
which shews that p(7m + 5) is divisible by 7.
From (16) it follows that
p(4)x + p(9)x2 + p(14)x3 + · · ·
5{(1 − x5 )(1 − x10 )(1 − x15 ) · · ·}4
x {(1 − x5 )(1 − x10 )(1 − x15 ) · · ·
= .
(1 − x)(1 − x2 )(1 − x3 ) · · · {(1 − x)(1 − x2 )(1 − x3 ) · · ·}5
As the coefficient of x5n on the right-hand side is a multiple of 5, it follows that p(25m + 24)
is divisible by 25.
Similarly
p(5)x + p(12)x2 + p(19)x3 + · · ·
7{(1 − x7 )(1 − x14 )(1 − x21 ) · · ·}2
(1 − x7 )(1 − x14 ) · · · 7 14
2 {(1 − x )(1 − x ) · · ·}
5
= x(1 − 3x + 5x3 − 7x6 + · · ·) + 7x ;
{(1 − x)(1 − x2 ) · · ·}7 {(1 − x)(1 − x2 ) · · ·}8
from which it follows that p(49m + 47) is divisible by 49.
Another proof of (1) and (2) has been found by Mr. H. B. C. Darling, to whom my conjecture had been
communicated by Major MacMahon. This proof will also be published in these Proceedings. I have since
found proofs of (3), (7) and (8).
Proof of certain identities in combinatory analysis

Proceedings of the Cambridge Philosophical Society, XIX, 1919,214 – 216

Let

(1 − xq)(1 − xq 2 ) · · · (1 − xq ν−1 )

X 1
G(x) = 1 + (−1)ν x2ν q 2 ν(5ν−1) (1 − xq 2ν )
(1 − q)(1 − q 2 )(1 − q 3 ) · · · (1 − q ν )
1
1 1 − xq
= 1 − x2 q 2 (1 − xq 2 ) + x4 q 9 (1 − xq 4 ) − ···. (1)
1−q (1 − q)(1 − q 2 )
If we write
1 − xq 2ν = 1 − q ν + q ν (1 − xq ν ),
every term in (1) is split up into two parts. Associating the second part of each term with
the first part of the succeeding term, we obtain

1 − xq (1 − xq)(1 − xq 2 )
G(x) = (1 − x2 q 2 ) − x2 q 3 (1 − x2 q 6 ) + x4 q 11 (1 − x2 q 10 ) − · · · . (2)
1−q (1 − q)(1 − q 2 )

Now consider
G(x)
H(x) = − G(xq). (3)
1 − xq
Substituting for the first term from (2) and for the second term from (1), we obtain

x2 q 3 x4 q 11 (1 − xq 2 )
H(x) = xq − {(1 − q) + xq 4 (1 − xq 2 )} + {(1 − q 2 ) + xq 7 (1 − xq 3 )}
1−q (1 − q)(1 − q 2 )
x6 q 24 (1 − xq 2 )(1 − xq 3 )
− {(1 − q 3 ) + xq 10 (1 − xq 4 )} + · · · .
(1 − q)(1 − q 2 )(1 − q 3 )
Associating, as before, the second part of each term with the first part of the succeeding
term, we obtain

1 − xq 3

1
H(x) = xq(1 − xq 2 ) 1 − x2 q 6 (1 − xq 4 ) + x4 q 17 (1 − xq 6 )
1−q (1 − q)(1 − q 2 )
(1 − xq 3 )(1 − xq 4 )

6 33 8
−x q (1 − xq ) ···
(1 − q)(1 − q 2 )(1 − q 3 )+
= xq(1 − xq 2 )G(xq 2 ). (4)

If now we write
G(x)
K(x) = ,
(1 − xq)G(xq)
270 Paper 26

we obtain, from (3) and (4),


xq
K(x) = 1 + ,
K(xq)
and so
xq xq 2 xq 3
K(x) = 1 + . (5)
1+ 1+ 1 + · · ·
In particular we have
1 q q2 1 (1 − q)G(q)
= = ; (6)
1+ 1+ 1 + · · · K(1) G(1)
or
1 q q2 1 − q − q 4 + q 7 + q 13 − · · ·
= . (7)
1+ 1+ 1 + · · · 1 − q 2 − q 3 + q 9 + q 11 − · · ·
This equation may also be written in the form
1 q q2 (1 − q)(1 − q 4 )(1 − q 6 )(1 − q 9 )(1 − q 11 ) · · ·
= . (8)
1+ 1+ 1 + · · · (1 − q 2 )(1 − q 3 )(1 − q 7 )(1 − q 8 )(1 − q 12 ) · · ·
If we write
G(x)
F (x) = ,
(1 − xq)(1 − xq 2 )(1 − xq 3 ) · · ·
then (4) becomes
F (x) = F (xq) + xqF (xq 2 ),
from which it readily follows that
xq x2 q 4 x3 q 9
F (x) = 1 + + + + ···. (9)
1 − q (1 − q)(1 − q ) (1 − q)(1 − q 2 )(1 − q 3 )
2

In particular we have
q q4 G(1)
1+ + + ··· =
1 − q (1 − q)(1 − q 2 ) (1 − q)(1 − q 2 )(1 − q 3 ) + · · ·
1 − q 2 − q 3 + q 9 + q 11 − · · ·
=
(1 − q)(1 − q 2 )(1 − q 3 ) · · ·
1
= ,(10)
(1 − q)(1 − q 4 )(1 − q 6 )(1 − q 9 )(1 − q 11 ) · · ·
and
q2 q6 (1 − q)G(q)
1+ + 2
+ ··· = ···
1 − q (1 − q)(1 − q ) (1 − q)(1 − q 2 )(1 − q 3 )
1 − q − q 4 + q 7 + q 13 − · · ·
=
(1 − q)(1 − q 2 )(1 − q 3 ) · · ·
1
= . (11)
(1 − q 2 )(1 − q 3 )(1 − q 7 )(1 − q 8 )(1 − q 12 ) · · ·
A class of definite integrals

Quarterly Journal of Mathematics, XLVIII, 1920, 294 – 310

1. It is well known that


1
π
π Γ(1 + m)
Z
2
(1.1) (cos x)m einx dx =
− 21 π 2m Γ{1 + 21 (m + n)}Γ{1 + 12 (m − n)}

if R(m) > −1. It follows from this and Fourier’s Theorem that, if n is any real number
except ±π and R(α + β) > 1, or if n = ±π and R(α + β) > 2, then

einx (2 cos 21 n)α+β−2 1 in(β−α)
Z
(1.2) dx = e2 or 0,
−∞ Γ(α + x)Γ(β − x) Γ(α + β − 1)

according as |n| < π or |n| ≥ π. In particular we have



dx 2α+β−2
Z
(1.21) =
−∞ Γ(α + x)Γ(β − x) Γ(α + β − 1)

if R(α + β) > 1; and



dx 22α−3
Z
(1.22) =
0 Γ(α + x)Γ(α − x) Γ(2α − 1)

if R(α) > 12 . If α is an integer n + 1, (1.22) reduces to



sin πx π 22n (n!)2
Z
(1.23) dx = .
0 x{1 − (x2 /12 )}{1 − (x2 /22 )} · · · {1 − (x2 /n2 )} 2 (2n)!

Again, if m is a positive integer, we have

1 + 2 cos 2πx + 2 cos 4πx + · · · to 21 (m + 1) terms


(
sin mπx or
= ,
sin πx 2 cos πx + 2 cos 3πx + · · · to 12 m terms

according as m is odd or even. It follows from this and (1.2) that, if R(α) > 1,

22α−3
Z ∞
sin mπx dx
= or 0,
0 sin πx Γ(α + x)Γ(α − x) Γ(2α − 1)

according as m is odd or even. Hence, if m and n are positive integers, we have

π 22n (n!)2
Z ∞
sin mπx
(1.24) dx = or 0,
0 x{1 − (x2 /12 )}{1 − (x2 /22 )} · · · {1 − (x2 /n2 )} 2 (2n)!
272 Paper 27

according as m is odd or even. From this we easily deduce that, if l, m, and n are positive
integers,
(1.25)
 2
(sin mπx)2l+1
Z ∞
2(n−l)−1 (2l)! n!
2 2 2 2 2 2
dx = 2 π or 0,
0 x{1 − (x /1 )}{1 − (x /2 )} · · · {1 − (x /n )} (2n)! l!
according as m is odd or even. It follows that
(sin mπx)2n+1
Z ∞
π
(1.26) 2 /12 )}{1 − (x2 /22 )} · · · {1 − (x2 /n2 )}
dx = or 0,
0 x{1 − (x 2
according as m is odd or even. Similarly we can shew that
(sin mπx)2l
Z ∞
(1.27) dx = 0
0 {1 − (x2 /12 )}{1 − (x2 /22 )} · · · {1 − (x2 /n2 )}
for all positive integral values of l, m, and n.
In this connection it is interesting to note the following results:
(i) If R(α + β) > 1 and R(γ + δ) > 1, then
Z 1
2 (2 cos πx)γ+δ−2 eiπ(γ−δ)x
(1.3) Γ(α + β − 1) dx
− 12 Γ(α + x)Γ(β − x)
Z 1
2 (2 cos πx)α+β−2 eiπ(α−β)x
= Γ(γ + δ − 1) dx.
−1 Γ(γ + x)Γ(δ − x)
2

This is easily proved by writing


Z 1
2
(2 cos πz)α+β−2 eiπz(α−β+2x) dz
− 12

instead of
Γ(α + β − 1)
Γ(α + x)Γ(β − x)
in the left-hand side of (1.3).
(ii) If m and n are integers of which one is odd and the other even, and m ≥ 0 and
R(α + β) > 2, then
Z ξ+1
(cos πx)m einπx (cos πx)m einπx
Z ∞
(1.4) (α + β − 2) dx = dx
ξ Γ(α + x)Γ(β − x) ξ Γ(α − 1 + x)Γ(β − x)
where ξ is any real number. This is proved as follows: Suppose that the left-hand side,
minus the right-hand side, is f (ξ). Then
(α + β − 2)(cos πξ)m inπξ
f ′ (ξ) = − e
Γ(α + ξ)Γ(β − ξ)
(−1)m+n (cos πξ)m einπξ (cos πξ)m einπξ
− + = 0.
Γ(α + ξ)Γ(β − ξ − 1) Γ(α − 1 + ξ)Γ(β − ξ)
A class of definite integrals 273

Hence f (ξ) is a constant which is easily seen to be zero.

2. Before proceeding further I shall give a few general rules for generalising the results in
the previous and the following sections. If
Z η
(x + ǫ1 )(x + ǫ2 ) · · · (x + ǫr )
fr (ζ) = F (x) dx,
ξ Γ(ζ ± x)

where r is zero or a positive integer, and the ǫ’s, ξ, η, ζ and F (x) are all arbitrary, then it
is easy to see that

(2.1) fr+1 (ζ + 1) = ±{fr (ζ) − (ζ ∓ ǫr+1 )fr (ζ + 1)},

provided the necessary convergence conditions are satisfied.


Similarly if Z η
fr (ζ) = (x + ǫ1 )(x + ǫ2 ) · · · (x + ǫr )Γ(ζ ± x)F (x) dx,
ξ

then

(2.2) fr+1 (ζ) = ±{fr (ζ + 1) − (ζ ∓ ǫr+1 )fr (ζ)}.

Thus we see that, if f0 (ζ) is known, fr (ζ) can be easily determined.


Suppose now that P (x) is a polynomial of the rth degree and N any integer greater than
or equal to r. Let D, E and ∆ denote the usual operators so that

E = 1 + ∆ = eD .

Then, if
η
F (x)
Z
f (ζ) = dx,
ξ Γ(ζ ± x)

η N
P (x)F (x) f (ζ − ν)
Z X
(±∆)ν P − 12 ν ± 1 − ζ + 12 ν ,
 
(2.3) dx =
ξ Γ(ζ ± x) 0
ν!

as easily seen by replacing P (x) by


1 1
(1 ± ∆E − 2 ± 2 )ζ±x−1 P {±(1 − ζ)} .

Similarly using the equation


1 1
P (x) = (1 ∓ ∆E − 2 ∓ 2 )−(ζ±x) P (±ζ),

we find that, if Z η
f (ζ) = Γ(ζ ± x)F (x) dx,
ξ
274 Paper 27

then
η N
f (ζ + ν)
Z X
(±∆)ν P − 12 ν ∓ ζ + 12 ν .
 
(2.4) Γ(ζ ± x)P (x)F (x) dx =
ξ 0
ν!

As an illustration let us apply (2.3) to (1.2). We find that if n is any real number except
±π, and R(α + β) > 1 + r, or if n = ±π, and R(α + β) > 2 + r, and N is any positive
integer greater than or equal to r, where r is the degree of the polynomial P (x), then

 0 or
P (x)einx
Z ∞ 

(2.5) dx = N
kν (2 cos 12 n)α+β−ν−2 1 in(β−α) ,
−∞ Γ(α + x)Γ(β − x)
X
 e2
ν! Γ(α + β − ν − 1)


0

1
according as |n| ≥ π or |n| < π, kν being either e 2 inν ∆ν P (1 − α) or
1
e− 2 inν (−∆)ν P (β − ν − 1).

It is immaterial which value of kν we take.


If
Γ(ζ1 + x)
P (x) = ,
Γ(ζ2 + x)
where ζ1 − ζ2 is a positive integer, then it is well known that

(ζ1 − ζ2 )! Γ(ζ1 + x)
(2.6) ∆ν P (x) = .
(ζ1 − ζ2 − ν)! Γ(ζ2 + x + ν)

This affords a very good example for the previous formulæ.


It is easy to see that (1.2) can be restated as

einx
Z
(2.7) dx = 0
−∞ Γ(α + x)Γ(β − x)
or


einx ein(ν−α)
Z X
(2.71) dx =
−∞ Γ(α + x)Γ(β − x) Γ(ν)Γ(α + β − ν)
1

X ein(β−ν)
= ,
1
Γ(ν)Γ(α + β − ν)

according as |n| ≥ π or |n| < π (n being of course real). But



P (x)einx ∞
e(D+in)x P (0)
Z Z
dx = dx.
−∞ Γ(α + x)Γ(β − x) −∞ Γ(α + x)Γ(β − x)
A class of definite integrals 275

Hence, if the conditions stated for (2.5) are satisfied,



P (x)einx
Z
(2.8) dx = 0
−∞ Γ(α + x)Γ(β − x)
or


P (x)einx ein(ν−α) P (ν − α)
Z X
(2.81) dx =
−∞ Γ(α + x)Γ(β − x) Γ(ν)Γ(α + β − ν)
1

X ein(β−ν) P (β − ν)
= ,
Γ(ν)Γ(α + β − ν)
1

according as |n| ≥ π or |n| < π.

3. We shall now consider an important extension of (1.2). Let [x] denote the greatest
integer not exceeding x, so that (e.g.) [−5 12 ] = −6. Let us agree further that
ν
X
= 0,
µ

if ν < µ. Then if n and s are real and φ(z) is a function that can be expanded in the form

X
Cν z ν
−∞

when |z| = 1, we have


Z ∞
φ(eisx )einx X {2 cos 1 (n + νs)}α+β−2 1
(3.1) dx = Cν 2
e 2 i(β−α)(n+νs) ,
−∞ Γ(α + x)Γ(β − x) Γ(α + β − 1)

where the summation is bounded by


   
π n π n
− + ≤ν≤ − ,
|s| s |s| s

provided that either


(i) π + n and π − n are not multiples of s, and R(α + β) > 1,
or (ii) π + n and π − n are multiples of s, and R(α + β) > 2,
or (iii) C(π−n)/s = 0 and C−(π+n)/s = 0, whenever one or the other or both of the suffixes
of C happen to be integral, and R(α + β) > 1,
or (iv) π + n and π − n are multiples of s, α + β is an integer greater than 1, and
C(π−n)/s = e2iπα C−(π+n)/s .
276 Paper 27

The formula (3.1) is easily obtained by substituting the series for φ and integrating term-
by-term, using (1.2).
It should be remembered that (3.1) is not true if n or s ceases to be real, though the
integral may be convergent. In such cases, generally, the integral cannot be evaluated in
finite terms.
The following integrals can be evaluated at once, with the help of (3.1):
Z ∞
dx
(3.11) imx inx
,
−∞ (pe + qe )Γ(α + x)Γ(β − x)
where m and n are real and |p| =
6 |q|;
 m
cos
∞ sx einx
sin
Z
(3.12) dx,
−∞ Γ(α + x)Γ(β − x)
where n and s are real and m is a positive integer;
(1 + ǫeisx )m einx
Z ∞
(3.13) dx,
−∞ Γ(α + x)Γ(β − x)

where m is real and |ǫ| < 1;



ep cos sx+iq sin sx+inx
Z
(3.14) dx.
−∞ Γ(α + x)Γ(β − x)
For instance the value of (3.14) is

X q + p2ν
1
np o 2 cos 1 (n + νs) α+β−2 1
Jν (q 2 − p2 ) 2
e 2 i(β−α)(n+νs) ,
q−p Γ(α + β − 1)
1
where Jν (x) is the ordinary Bessel function of the νth order, and {(q + p)/(q − p)} 2 ν should
be interpreted so that the first term in the expansion of
1
np o
{(q + p)/(q − p)} 2 ν Jν (q 2 − p2 )

is 1 ν
+ q)2 (p
.
ν!
Putting s = 2π in (3.1) we obtain the following corollary. If φ is the same function as in
(3.1), and n is any real number, then

φ(e2iπx )
Z ∞
(3.2) einx dx
−∞ Γ(α + x)Γ(β − x)
 α+β−2
2 cos 12 n − π[(π + n)/2π]

1
n o
i(β−α) 2 n−π[(π+n)/2π]
= Cλ e ,
Γ(α + β − 1)
A class of definite integrals 277

where  
π+n
λ=− ,

provided that either
(i) n is not an odd multiple of π, and R(α + β) > 1,
or (ii) n is an odd multiple of π , and R(α + β) > 2,
or (iii) n is an odd multiple of π, C(π−n)/2π = 0 and C−(π+n)/2π = 0, and R(α + β) > 1,
or (iv) n is an odd multiple of π, α + β is an integer greater than 1, and
C(π−n)/2π = e2iπα C−(π+n)/2π .
Thus we see that the value of each integrals (3.12) – (3.14), when s = 2π, reduces to a
single term.
The next section will be denoted to the application of (3.2) in evaluating some special
integrals.

4. Suppose that α is not real and

φ(z) = 1 + e−2iπα z + e−4iπα z 2 + · · · (I(α) < 0),

and φ(z) = −e2iπα z −1 − e4iπα z −2 + · · · (I(α) > 0),


so that φ(z) is convergent when |z| = 1. Then it is easy to see that

φ(e2iπx )
Z ∞ Z ∞
Γ(α + x) inx
e dx = 2iπe−iπα eix(π−n) dx.
−∞ Γ(β + x) −∞ Γ(1 − α + x)Γ(β − x)

It follows from (3.2) that if α is not real, n real, and

(i) n is neither 0 nor any multiple of 2π, and R(α − β) < 0,


or (ii) n has the same sign as I(α) and R(α − β) < 0,
or (iii) n is 0, or a multiple of 2π, having the sign opposite to that of I(α), and
R(α − β) < −1,
or (iv) n is not 0 and α − β is a negative integer, then
Z ∞
Γ(α + x) inx
(4.1) e dx = 0 or
−∞ Γ(β + x)

2iπ
 
π−n h n iβ−α−1
= ± 2 cos +π
Γ(β − α) 2 2π
  h n i
π−n
× exp −inα + i(β − α − 1) +π ,
2 2π

the zero value being taken when n and I(α) have the same sign, the plus sign when n ≥ 0
and I(α) < 0, and the minus sign when n ≤ 0 and I(α) > 0.
278 Paper 27

As a particular cases of (4.1) we have



Γ(α + x)
Z
(4.11) dx = 0,
−∞ Γ(β + x)

if α is not real and R(α − β) < −1. If α is not real, n real, and (i) r is a positive integer,
or (ii) r = 0 and n 6= 0, then

einx dx
Z
(4.12) =0
−∞ (x + α)(x + α + 1) · · · (x + α + r)
2iπ r 1
or ± 2 sin 21 n e 2 ir(π−n)−inα ,
r!
the different values being selected as in (4.1).
Similarly we can shew that if α and β are not real and n is real, and

(i) n is not an odd multiple of π and R(α + β) < 1,


or (ii) n is an odd multiple of π which has either the sign of I(β) or the sign opposite
to that of I(α), and R(α + β) < 0,
or (iii) n is an odd multiple of π which has neither the sign of I(β) nor the sign opposite
to that of I(α), and R(α + β) < 1, then

2iπΓ(α + β) 1 in(β−α)
Z
(4.2) Γ(α + x)Γ(β − x)einx dx = e2
−∞ |2 cos 21 n|α+β
      
(π + n) (π + n)
× ηn (β) exp iπ(α + β) − ηn (−α) exp −iπ(α + β) ,
2π 2π

where ηn (ζ) is equal to 0 when π − n and I(ζ) have the same sign, to 1 when n ≤ π and
I(ζ) < 0, and to −1 when n ≥ π and I(ζ) > 0.
It should be remembered that for real values of n
  
2 cos 1 n = 2 cos 1 n − π (π + n)

2 2 .

It follows, in particular, that if α and β are not real, and R(α + β) < 1, then
Z ∞
(4.21) Γ(α + x)Γ(β − x) dx = 0 or ± 21−α−β iπΓ(α + β),
−∞

the zero value being chosen when I(α) and I(β) have different signs, the plus sign when
I(α) and I(β) are both negative, and the minus sign when I(α) and I(β) are both positive.
The following results can either be deduced from (1.5) or be proved independently in the
same way as (1.4).
A class of definite integrals 279

If n is zero or any multiple of 2π, ξ real, α any number except the real numbers less than
or equal to −ξ, and β is any number such that R(β − α) > 1, then
∞ ξ+1
Γ(α + x) inx Γ(α + x) inx
Z Z
(4.3) (β − α − 1) e dx = e dx.
ξ Γ(β + x) ξ Γ(β − 1 + x)

If n is any odd multiple of π, α and ξ are the same as in (4.3), and β is not real and
R(α + β) < 0, then
Z ∞ Z ξ+1
inx
(4.4) (α + β) Γ(α + x)Γ(β − x)e dx = Γ(α + x)Γ(β + 1 − x)einx dx.
ξ ξ

5. We now proceed to consider an application of (1.2) to some other functions. Suppose


that Us (x), Vs (x) and Ws (x) are many-valued functions of x defined by

u0 x s u1 xs+ǫ u2 xs+2ǫ
Us (x) = + + + ···,
Γ(1 + s) Γ(1 + s + ǫ) Γ(1 + s + 2ǫ)

v0 x s v1 xs+ǫ v2 xs+2ǫ
Vs (x) = + + + ···,
Γ(1 + s) Γ(1 + s + ǫ) Γ(1 + s + 2ǫ)
w0 xs w1 xs+ǫ w2 xs+2ǫ
Ws (x) = + + + ···,
Γ(1 + s) Γ(1 + s + ǫ) Γ(1 + s + 2ǫ)
where R(ǫ) ≥ 0, the u’s, v’s and w’s are any numbers connected by the relation to be found
by equating the coefficients of the various powers of k in the equation

w0 + w1 k + w2 k2 + · · · = (u0 + u1 k + u2 k2 + · · ·)(v0 + v1 k + v2 k2 + · · ·)∗ ,

and the series Us (x), Vs (x) and Ws (x) are convergent at least for the values of s and x
that appear in the equation (5.2).
The functions U, V and W are many valued. If |x/y| = 1 and | arg(x/y)| < π, then one
value of arg(x + y) is given by the equation

(5.1) arg x + arg y = 2 arg(x + y).

If we choose arg x and arg y arbitrarily, and agree that

xs+µǫ = exp{(s + µǫ)(log |x| + i arg x)},

and that of y s+µǫ and (x + y)s+µǫ are to be interpreted similarly, that value of arg(x + y)
being chosen which is given by (5.1), then a definite branch of W is associated with any
arbitrary pair of branches of U and V .

These series need not, of course, be convergent for any value of k.
280 Paper 27

If α, β, x, y are any numbers such that |x/y| = 1, and R(α + β) > 0 when | arg(x/y)| = π
and R(α + β) > −1 otherwise, then
Z ∞
(5.2) Uα+ξ (x)Vβ−ξ (y)(x/y)ξ dξ = 0 or Wα+β (x + y),
−∞

according as | arg(x/y)| ≥ π or | arg(x/y)| < π, whatever be the branches of U (x) and


V (y), provided that the corresponding branch of W (x + y) is fixed in accordance with the
conventions explained above. This is proved as follows. Suppose that
1 1
x = te 2 in , y = te− 2 in ,

where t is arbitrary and n is any real number. Then the integral becomes
Z ∞
1 1
Uα+ξ (te 2 in )Vβ−ξ (te− 2 in )einξ dξ.
−∞

If we expand the integral in powers of t, and integrate term by term with the help of (1.2)
and then make use of the relations between the u’s, v’s and w’s, the result will be

Wα+β 2t cos 12 n = Wα+β (x + y),




or zero, according to the conditions stated with regard to (5.2).


In particular, if R(α + β) > −1, we have
Z ∞
(5.21) Uα+ξ (x)Vβ−ξ (x)dξ = Wα+β (2x).
−∞

Suppose now that Gs (p, x) is a many-valued function of x defined by

xs p xs+1 p(p + 1) xs+2


Gs (p, x) = − + − ···.
Γ(s + 1) 1! Γ(s + 2) 2! Γ(s + 3)

Then it follows from (5.2) that if α, β, x, y are any numbers such that |x/y| = 1, and
R(α + β) > 0 when | arg(x/y)| = π and R(α + β) > −1 otherwise, we have
Z ∞
(5.3) Gα+ξ (p, x)Gβ−ξ (q, y)(x/y)ξ dξ = 0 or Gα+β (p + q, x + y),
−∞

according as | arg(x/y)| ≥ π or < π, whatever be the branches of G(p, x) and G(q, y),
provided that the corresponding branch of G(p + q, x + y) is chosen according to our former
convention. If, in particular, R(α + β) > −1, then
Z ∞
(5.31) Gα+ξ (p, x)Gβ−ξ (q, x)dξ = Gα+β (p + q, 2x).
−∞
A class of definite integrals 281

It may be interesting to note that the right-hand sides of (5.3) and (5.31) are of the form
Gα+β (p + q, z), which reduces to
z α+β
Γ(α + β + 1)
when p = −q, becoming independent of p or q.
The ordinary Bessel’s functions are particular cases of the function Gs (p, x). Hence we
have the following particular results. If n is real, and R(α + β) > 0 when n = ±π and
R(α + β) > −1 otherwise, then


Jα+ξ (x) Jβ−ξ (y) inξ
Z
(5.4) e dξ = 0
−∞ xα+ξ y β−ξ
! 1 (α+β) "r #
2 cos 12 n 2
1
in(β−α) 1

− 1
in 1
in

or 1 1 e 2 Jα+β 2 cos 2n x2 e 2 + y 2 e 2 ,
x2 e− 2 in + y 2 e 2 in

according as |n| ≥ π or < π. If n is real, and R(α+β) > 0 when n = ±π and R(α+β) > −1
otherwise, then
Z ∞
1
Jα+ξ (x)Jβ−ξ (x)einξ dξ = 0 or e 2 in(β−α) Jα+β 2x cos 21 n ,

(5.41)
−∞

according as |n| ≥ π or < π. If R(α + β) > −1, then


np o
Z ∞
Jα+ξ (x) Jβ−ξ (y) Jα+β (2x2 + 2y 2 )
(5.42) dξ =
xα+ξ y β−ξ 1 2
1
1 2 2 (α+β)
2x + 2y
−∞

and
Z ∞
(5.43) Jα+ξ (x)Jβ−ξ (x)dξ = Jα+β (2x).
−∞

6. We shall now consider some special cases of the integral



einx
Z
(6.1) dx,
−∞ Γ(α + x)Γ(β − x)Γ(γ + lx)Γ(δ − lx)

l and n being real numbers.


Replacing 1/{Γ(γ + lx)Γ(δ − lx)} by
1
π
1
Z
2
(2 cos z)γ+δ−2 e−iz(γ−δ+2lx) dz,
πΓ(γ + δ − 1) − 21 π
282 Paper 27

it follows from (1.2) that (6.1) is equal to


Z v
1  α+β−2
2 cos 12 n − lz (2 cos z)γ+δ−2

(6.11)
πΓ(α + β − 1)Γ(γ + δ − 1) u
× exp i(β − α) 21 n − lz + i(δ − γ)z dz,
 

where u and v are the lower and upper extremities of the common part of the intervals

− 21 π < z < 21 π, − 12 π < 21 n − lz < 12 π.

If the intervals do not overlap, the value of (6.1) is zero. It is easy to see that if

(6.12) |n| ≥ π(1 + |l|),

the intervals do not overlap; and that, if they do overlap and l > 0, then

π n − π π n − π
u = + − − ,
4 4l 4 4l

and
π n + π π n + π
v= +
− − .
4 4l 4 4l
It should also be observed that, though (6.11) may not be convergent for all the values of
α, β, γ and δ for which (6.1) is convergent, yet we may evaluate (6.11) when it is conver-
gent, and so obtain a formula for (6.1) which may be extended, by the theory of analytic
continuation, to all values of the parameters for which the integral converges.
From (6.12) we see that, if l and n are any real numbers such that |n| ≥ π(1 + |l|), then

einx
Z ∞
(6.2) dx = 0,
−∞ Γ(α + x)Γ(β − x)Γ(γ + lx)Γ(δ − lx)

provided that (i) R(α + β + γ + δ) > 2 when |n| > π(1 + |l|),
and (ii) R(α + β + γ + δ) > 3 when |n| = π(1 + |l|).

7. Suppose now that l = 1 and n = 0; then (6.11) reduces to


1
π
1
Z
2
(2 cos z)α+β+γ+δ−4 eiz(α−β−γ+δ) dz,
πΓ(α + β − 1)Γ(γ + δ − 1) − 12 π

which is easily evaluated by the help of (1.1). Hence


Z ∞
dx
(7.1)
−∞ Γ(α + x)Γ(β − x)Γ(γ + x)Γ(δ − x)
Γ(α + β + γ + δ − 3)
= ,
Γ(α + β − 1)Γ(β + γ − 1)Γ(γ + δ − 1)Γ(δ + α − 1)
A class of definite integrals 283

provided that (i) R(α + β + γ + δ) > 3, or (ii) 2(α − γ) and 2(β − δ) are odd integers and
R(α + β + γ + δ) > 2.
It should be noted that the formula fails when α + β + γ + δ = 3 and 2(α − γ) and 2(β − δ)
are odd integers. The value of the integral in this case is some times 1/2π and some times
−1/2π. The value to be selected may be fixed as follows. It is easy to see that, in this case,
one and only one of the numbers α+ β − 1, β + γ − 1, γ + δ − 1 and δ + α− 1 will be an integer
less than or equal to zero. If α + β − 1 or β + γ − 1 happens to be such a number, then
the value of the integral is ±1/2π, according as 2(β − δ) ≡ ∓1 (mod 4). But if γ + δ − 1
or δ + α − 1 happens to be such a number, the value of the integral is ±1/2π, according as
2(β − δ) ≡ ±1 (mod 4).
As particular cases of (7.1), we have

1 Γ(2α + 2β − 3)
Z
(7.11) 2
dx = ,
−∞ {Γ(α + x)Γ(β − x)} {Γ(α + β − 1)}4

provided that R(α + β) > 23 ,



dx Γ(2α + 2β − 3)
Z
(7.12) = ,
0 Γ(α + x)Γ(α − x)Γ(β + x)Γ(β − x) 2Γ(2α − 1)Γ(2β − 1){Γ(α + β − 1)}2

provided that (i) R(α + β) > 32 , or (ii) 2(α − β) is an odd integer and R(α + β) > 1. If
2(α + β) = 3 and 2(α − β) is an odd integer, then the value of the integral (7.12), when
α ≥ 1, is ±1/2π, according as 2(α − β) ≡ ±1 (mod 4), and when α < 1 it is ±1/2π,
according as 2(α − β) ≡ ∓1 (mod 4).
Putting α = β in (7.11) or in (7.12), we obtain

dx Γ(4α − 3)
Z
(7.13) 2
= ,
0 {Γ(α + x)Γ(α − x)} 2{Γ(2α − 1)}4

if R(α) > 43 . Suppose again that l = 1, n = π, and α + δ = β + γ. Then (6.11) reduces to


1 1
e 2 iπ(β−α) π
Z
2
(2 sin z)α+β−2 (2 cos z)γ+δ−2 dz.
πΓ(α + β − 1)Γ(γ + δ − 1) 0

Hence we see that


(7.2)
1
e± 2 iπ(β−α)
Z ∞
e±iπx dx
= 1 1 ,
−∞ Γ(α + x)Γ(β − x)Γ(γ + x)Γ(δ − x) 2Γ 2 (α + β) Γ 2 (γ + δ) Γ(α + δ − 1)

if α + δ = β + γ and R(α + β + γ + δ) > 2. In particular


1

e±iπx e± 2 iπ(β−α)
Z
(7.21) dx = 2 ,
{Γ(α + x)Γ(β − x)}2 2Γ(α + β − 1) Γ 12 (α + β)
 
−∞
284 Paper 27

if R(α + β) > 1, and


Z ∞
cos πx 1
(7.22) dx = ,
0 {Γ(α + x)Γ(α − x)}2 4Γ(2α − 1) {Γ(α)}2

if R(α) > 21 .

8. It follows from (6.2), (7.1), and (7.2) that, if



X ∞
X
φ(z) = c2ν z 2ν , ψ(z) = c2ν+1 z 2ν+1 ,
−∞ −∞

the series being convergent when |z| = 1, then

φ(eiπx )
Z ∞
(8.1) dx
−∞ Γ(α + x)Γ(β − x)Γ(γ + x)Γ(δ − x)
c0 Γ(α + β + γ + δ − 3)
= ,
Γ(α + β − 1)Γ(β + γ − 1)Γ(γ + δ − 1)Γ(δ + α − 1)

provided that (i)R(α + β + γ + δ) > 3 or (ii)R(α + β + γ + δ) > 2 and

c2 eiπ(β+δ) + c−2 e−iπ(β+δ) = 2c0 cos π(β − δ),

c2 e−iπ(α+γ) + c−2 eiπ(α+γ) = 2c0 cos π(α − γ);


and that

ψ(eiπx )
Z
(8.2) dx
−∞ Γ(α + x)Γ(β − x)Γ(γ + x)Γ(δ − x)
1 1
c1 e 2 iπ(β−α) + c−1 e− 2 iπ(β−α)
= ,
2Γ 2 (α + β) Γ 12 (γ + δ) Γ(α + δ − 1)
1 

provided that α + δ = β + γ and R(α + β + γ + δ) > 2.


If α + δ − β − γ is an integer other than zero, it is possible to evaluate the integrals (7.2)
and (8.2) in finite terms, but not as a single term.
The following integrals can be evaluated as a single term, with the help of (8.1) and (8.2),
whenever they are convergent:

Γ(δ + x)einx
Z ∞
(8.3) dx,
−∞ Γ(α + x)Γ(β − x)Γ(γ + x)

where (i) n is an odd multiple of π, or (ii) n is an even multiple of π and α + 1 = β + γ + δ;


Z ∞
Γ(γ + x)Γ(δ ± x) inx
(8.4) e dx,
−∞ Γ(α + x)Γ(β ± x)
A class of definite integrals 285

where (i) n is an even multiple of π, or (ii) n is an odd multiple of π and α + δ = β + γ;


Z ∞
Γ(β + x)Γ(γ − x)Γ(δ + x) inx
(8.5) e dx,
−∞ Γ(α + x)

where (i) n is an odd multiple of π, or (ii) n is an even multiple of π and α + β + γ = 1 + δ;


Z ∞
(8.6) Γ(α + x)Γ(β − x)Γ(γ + x)Γ(δ − x)einx dx,
−∞

where (i) n is an even multiple of π, or (ii) n is an odd multiple of π and α + δ = β + γ.


Thus for instance, if δ is not real, α + 1 = β + γ + δ, and R(α + β + γ − δ) > 1, then
Z ∞
Γ(δ + x)
(8.31) dx
−∞ Γ(α + x)Γ(β − x)Γ(γ + x)
1
πe± 2 iπ(δ−γ)
= ,
Γ(α − δ)Γ 2 (α + β) Γ 12 (γ − δ + 1)
1 

according as I(δ) is positive or negative; and if γ and δ are not real and R(α+β −γ −δ) > 1,
then
Z ∞
Γ(γ + x)Γ(δ + x)
(8.41) dx = 0
−∞ Γ(α + x)Γ(β + x)
2iπ 2 Γ(α + β − γ − δ − 1)
or ± ,
sin π(γ − δ) Γ(α − γ)Γ(α − δ)Γ(β − γ)Γ(β − δ)

the zero value being taken when I(γ) and I(δ) have the same sign, the plus sign when
I(γ) > 0 and I(δ) < 0, and the minus sign when I(γ) < 0 and I(δ) > 0.

9. The following results are easily obtained with the help of (6.11). If 2(α − β) = γ − δ
and R(α + β + γ + δ) > 3, then
Z ∞
e±iπx
(9.1) dx
−∞ Γ(α + x)Γ(β − x)Γ(γ + 2x)Γ(δ − 2x)
1
2α+β+γ+δ−5 e± 2 iπ(β−α) Γ 12 (α + β + γ + δ − 3)

= √ .
π Γ 12 (α + β) Γ(γ + δ − 1)Γ(2α + δ − 2)


If R(α + β) > 23 , then


Z ∞
cos πx
(9.11) dx
0 Γ(α + x)Γ(α − x)Γ(β + 2x)Γ(β − 2x)
22α+2β−6 Γ α + β − 32

=√ .
π Γ(α)Γ(2β − 1)Γ(2α + β − 2)
If α + β + γ + δ = 4, then
286 Paper 27


cos π(x + β + γ)
Z
(9.12) dx
−∞ Γ(α + x)Γ(β − x)Γ(γ + 2x)Γ(δ − 2x)
1
= .
2Γ(γ + δ − 1)Γ(2α + δ − 2)Γ(2β + γ − 2)
If 2(α − β) = γ − δ + k, where k is ±1 or ±2, then
Z ∞
sin π(2x + α − β)
(9.2) dx
−∞ Γ(α + x)Γ(β − x)Γ(γ + 2x)Γ(δ − 2x)

22α−γ−3
= ±√ ,
π Γ β + γ − α + 21 Γ(2α + δ − 2)


provided that R(α + β + γ + δ) > 2.


If 3(α − β) = γ − δ + k, where k is ±1 or ±2, then
Z ∞
sin π(2x + α − β)
(9.3) dx
−∞ Γ(α + x)Γ(β − x)Γ(γ + 3x)Γ(δ − 3x)
33α+δ−4 Γ(2α − β + δ − 2)
=± ,
4π Γ(γ + δ − 1)Γ(3α + δ − 3)
provided that (i) R(α + β + γ + δ) > 3, or (ii)β + γ − 2α is integral and R(α + β + γ + δ) > 2.
In (9.2) and (9.3) the plus sign or minus sign on the right-hand side is to be taken according
as k is positive or negative. If k is an integer other than ±1 or ±2, the integrals in (9.2)
and (9.3) can still be evaluated in finite terms, but in a less simple form.

10. In this connection, it may be interesting to note that, if n is an even multiple of π,


and α + β + γ + δ = 4, then


einx
Z
(10.1) (α + β − 2)(β + γ − 2) dx
ξ Γ(α + x)Γ(β − x)Γ(γ + x)Γ(δ − x)
ξ+1
einx
Z
= dx
ξ Γ(α − 1 + x)Γ(β − x)Γ(γ − 1 + x)Γ(δ − x)

for all real values of ξ. The proof of this is the same as that of (1.4).
Finally I may mention the formula
Z ∞
α+β+γ+δ
(10.2) Jα+ξ (x)Jβ−ξ (x)Jγ+ξ (x)Jδ−ξ (x)dξ = 12 x
−∞
ν=∞ ν−1
X − 14 x2 {Γ(α + β + γ + δ + 2ν − 1)}2
× ,
Γ(ν)Γ(α + β + γ + δ + ν)Γ(α + β + ν)Γ(β + γ + ν)Γ(γ + δ + ν)Γ(δ + α + ν)
ν=1

which holds if (i) R(α + β + γ + δ) > −1, or (ii) 2(α − γ) and 2(β − δ) are odd integers
and R(α + β + γ + δ) > −2.
Congruence properties of partitions
Proceedings of the London Mathematical Society, 2, XVIII, 1920,
Records for 13 March 1919
In a paper published recently in the Proceedings of the Cambridge Philosophical Society∗ , I
proved a number of arithmetical properties of p(n), the number of unrestricted partitions
of n, and in particular that
p(5n + 4) ≡ 0 (mod 5),
and
p(7n + 5) ≡ 0 (mod 7).
Alternative proofs of these two theorems were found afterwards by Mr. H. B. C. Darling† .
I have since found another method which enables me to prove all these properties and a
variety of others, of which the most striking is

p(11n + 6) ≡ 0 (mod 11).

There are also corresponding properties in which the moduli are powers of 5, 7, or 11; thus

p(25n + 24) ≡ 0 (mod 25),

p(49n + 19), p(49n + 33), p(49n + 40), p(49n + 47) ≡ 0 (mod 49),
p(121n + 116) ≡ 0 (mod 121).
It appears that there are no equally simple properties for any moduli involving primes other
than these three.
The function τ (n) defined by the equation
X

τ (n)xn = x{(1 − x)(1 − x2 )(1 − x3 ) · · ·}24 ,
1

also possesses very remarkable arithmetical properties. Thus

τ (5n) ≡ 0 (mod 5),

τ (7n), τ (7n + 3), τ (7n + 5), τ (7n + 6) ≡ 0 (mod 7),


while
τ (23n + ν) ≡ 0 (mod 23),
if ν is any one of the numbers

5, 7, 10, 11, 14, 15, 17, 19, 20, 21, 22.


Vol. XIX, 1919, pp. 207 – 210 [No. 25 of this volume; see also No. 30] .

Ibid., pp. 217, 218.
Algebraic relations between certain infinite products

Proceedings of the London Mathematical Society, 2, XVIII, 1920,


Records for 13 March 1919

It was proved by Prof. L. J. Rogers∗ that

1 x4 x9
G(x) = 1 + + + + ···
1 − x (1 − x)(1 − x2 ) (1 − x)(1 − x2 )(1 − x3 )
1 1
= ··· × ,
(1 − x)(1 − x6 )(1 − x11 ) (1 − x4 )(1 − x9 )(1 − x14 ) · · ·

and
x2 x6 x12
H(x) = 1 + + + + ···
1 − x (1 − x)(1 − x2 ) (1 − x)(1 − x2 )(1 − x3 )
1 1
= 2 7 12
··· × .
(1 − x )(1 − x )(1 − x ) (1 − x )(1 − x8 )(1 − x13 ) · · ·
3

Simpler proofs were afterwards found Prof. Rogers and myself.†


I have now found an algebraic relation between G(x) and H(x), viz.:

H(x){G(x)}11 − x2 G(x){H(x)}11 = 1 + 11x{G(x)H(x)}6 .

Another noteworthy formula is

H(x)G(x11 ) − x2 G(x)H(x11 ) = 1.

Each of these formulæ is the simplest of a large class.


Proc. London Math. Soc., Ser. 1, Vol. XXV, 1894, pp. 318 – 343.

Proc. Camb. Phil. Soc., Vol. XIX, 1919, pp. 211 – 216. A short account of the history of the theorems
is given by Mr. Hardy in a note attached to this paper. [For Ramanujan’s proofs see No. 26 of this volume.]
Congruence properties of partitions
Mathematische Zeitschrift, IX, 1921, 147 – 153
[Extracted from the manuscripts of the author by G. H. Hardy]∗

1. Let
2x2 3x3
 
x
(1.11) P = 1 − 24 + + + ··· ,
1 − x 1 − x2 1 − x3

23 x2 33 x3
 
x
(1.12) Q = 1 + 240 + + + ··· ,
1 − x 1 − x2 1 − x3

25 x2 35 x3
 
x
(1.13) R = 1 − 504 + + + ··· ,
1 − x 1 − x2 1 − x3

(1.2) f (x) = (1 − x)(1 − x2 )(1 − x3 ) · · · .

Then it is well known that



X 1 1
(1.3) f (x) = 1 − x − x2 + x5 + x7 − · · · = 1 + (−1)n (x 2 n(3n−1) + x 2 n(3n+1) ),
n=1

Srinivasa Ramanujan, Fellow of Trinity College, Cambridge, and of the Royal Society of London, died
in India on 26 April, 1920, aged 32. The manuscript from which this note is derived is a sequel to a short
memoir “Some properties of p(n), the number of partitions of n,” Proceedings of the Cambridge Philosophical
Society, Vol.XIX (1919), 207-210 [No.25 of this volume]. In this memoir Ramanujan proves that
p(5n + 4) ≡ 0 (mod 5)
and
p(7n + 5) ≡ 0 (mod 7),
and states without proof a number of further congruences to moduli of the form 5a 7b 11c of which the most
striking is
p(11n + 6) ≡ 0 (mod 11).
Here now proofs are given of the first two congruences, and the first published proof of the third.
The manuscript contains a large number of further results. It is very incomplete, and will require very
careful editing before it can be published in full. I have taken from it the three simplest and most striking
results, as a short but characteristic example of the work of a man who was beyond question one of the
most remarkable mathematicians of his time.
I have adhered to Ramanujan’s notation, and followed his manuscript as closely as I can. A few insertions
of my own are marked by brackets. The most substantial of these is in § 5, where Ramanujan’s manuscript
omits the proof of (5.4). Whether I have reconstructed his argument correctly I cannot say.
The references given in the footnotes to “Ramanujan” are to his memoir “On certain arithmetical func-
tions,” Transactions of the Cambridge Philosophical Society, Vol. XXII, No.9 (1916), 159-184 [No.18 of this
volume].
290 Paper 30

(1.4) Q3 − R2 = 1728x(f (x))24 .

Further, let
∞ X
X ∞ ∞
X
(1.51) Φr,s (x) = mr ns xmn = nr σs−r (n)xn ,
m=1 n=1 n=1

where σk (n) is the sum of the kth powers of the divisors of n; so that

x 2s x2 3s x3
(1.52) Φ0,s (x) = + + + ···,
1 − x 1 − x2 1 − x3
and in particular

(1.53) P = 1 − 24Φ0,1 (x), Q = 1 + 240Φ0,3 (x), R = 1 − 504Φ0,5 (x).

Then [it may be deduced from the theory of the elliptic modular functions, and has been
shewn by the author in a direct and elementary manner ∗ , that, when r + s is odd, and
r < s, Φr,s (x) is expressible as a polynomial in P, Q, and R, in the form
X
Φr,s(x) = kl,m,n P l Qm Rn ,

where
l − 1 ≤ Min (r, s), 2l + 4m + 6n = r + s + 1.
In particular † ]

27 x2
 
2 x
(1.61) Q = 1 + 480Φ0,7 (x) = 1 + 480 + + ··· ,
1 − x 1 − x2

29 x2
 
x
(1.62) QR = 1 − 264Φ0,9 (x) = 1 − 264 + + ··· ,
1 − x 1 − x2

(1.63) 441Q3 + 250R2 = 691 + 65520Φ0,11 (x)


211 x2
 
x
= 691 + 65520 + + ··· ,
1 − x 1 − x2

Ramanujan, p. 165 [pp. 181 – 183].

Ramanujan, pp. 163 – 165 [pp. 180 – 181] (Tables I to III). Ramanujan carried the calculation of
formulæ of this kind to considerable lengths, the formula of Table I being

7709321041217 + 32640Φ0,31 (x) = 764412173217Q8


+5323905468000Q5 R2 + 1621003400000Q2 R4 .

It is worth while to quote one such formula; for it is impossible to understand Ramanujan without realising
his love of numbers for their own sake.
Congruence properties of partitions 291

(1.71) Q − P 2 = 288Φ1,2 (x),

(1.72) P Q − R = 720Φ1,4 (x),

(1.73) Q2 − P R = 1008Φ1,6 (x),

(1.74) Q(P Q − R) = 720Φ1,8 (x),

(1.81) 3P Q − 2R − P 3 = 1728Φ2,3 (x),

(1.82) P 2 Q − 2P R + Q2 = 1728Φ2,5 (x),

(1.83) 2P Q2 − P 2 R − QR = 1728Φ2,7 (x),

(1.91) 6P 2 Q − 8P R + 3Q2 − P 4 = 6912Φ3,4 (x),

(1.92) P 3 Q − 3P 2 R + 3P Q2 − QR = 3456Φ3,6 (x),

(1.93) 15P Q2 − 20P 2 R + 10P 3 Q − 4QR − P 5 = 20736Φ4,5 (x).

Modulus 5

2. We denote generally by J an integral power-series in x whose coefficients are integers.


It is obvious from (1.12) that
Q = 1 + 5J.
Also n5 − n ≡ 0 (mod 5), and so, from (1.11) and (1.13),

R = P + 5J.

Hence
Q3 − R2 = Q(1 + 5J)2 − (P + 5J)2 = Q − P 2 + 5J.
292 Paper 30

Using (1.4), (1.71), and (1.51), we obtain



X
(2.1) 1728x(f (x))24 = 288 nσ1 (n)xn + 5J.
n=1

Also
(1 − x)25 = 1 − x25 + 5J,
and so
(f (x))25 = f (x25 ) + 5J,

f (x25 )
(2.2) (f (x))24 = + 5J.
f (x)
But
1
= 1 + p(1)x + p(2)x2 + · · · ,
f (x)
and therefore, by (2.1) and (2.2),

(2.3) 1728xf (x25 )(1 + p(1)x + p(2)x2 + · · ·)


f (x25 )
= 1728x = 1728x(f (x))24 + 5J
f (x)

X
= 288 nσ1 (n)xn + 5J.
n=1

Multiplying by 2, rejecting multiples of 5, and replacing f (x25 ) by its expansion given by


(1.3), we obtain

(x − x26 − x51 + x126 + · · ·)(1 + p(1)x + p(2)x2 + · · ·)


X∞
= nσ1 (n)xn + 5J.
n=1

Hence
p(n − 1) − p(n − 26) − p(n − 51) + p(n − 126) + p(n − 176)
(2.4)
−p(n − 301) − · · · ≡ nσ1 (n) (mod 5),

the numbers 1,26, 51, . . . being the numbers of the forms


25 25
n(3n − 1) + 1, n(3n + 1) + 1,
2 2
or, what is the same thing, of the forms
1 1
(5n − 1)(15n − 2), (5n + 1)(15n + 2).
2 2
Congruence properties of partitions 293

In particular it follows from (2.3) that

(2.5) p(5m − 1) ≡ 0 (mod 5).

Modulus 7

3. It is obvious from (1.13) that


R = 1 + 7J.
Also n7 − n ≡ 0 (mod 7), and so, from (1.11) and (1.61),

Q2 = P + 7J.

Hence

(Q3 − R2 )2 = (P Q − 1 + 7J)2 = P 2 Q2 − 2P Q + 1 + 7J
= P 2 − 2P Q + R + 7J.

But, from (1.72) and (1.81),



X
P 3 − 2P Q + R = 144 (5nσ3 (n) − 12n2 σ1 (n))xn
n=1

X
= (n2 σ1 (n) − nσ3 (n))xn + 7J.
n=1

And therefore

X
3 2 2
(3.1) (Q − R ) = (n2 σ1 (n) − nσ3 (n))xn + 7J.
n=1

Again (by the same argument which led to (2.2)) we have

f (x49 )
(3.2) (f (x))48 = + 7J.
f (x)

Combining (3.1) and (3.2), we obtain

f (x49 )
x2 = x2 (f (x))48 + 7J = 17282 x2 (f (x))48 + 7J
f (x)
(3.3)
(Q3 − R2 )2 + 7J
= P
∞ 2 n
= n=1 (n σ1 (n) − nσ3 (n))x + 7J.
294 Paper 30

From (3.3) it follows (just as (2.4) and (2.5) followed from (2.3)) that

p(n − 2) − p(n − 51) − p(n − 100) + p(n − 247) + p(n − 345)


(3.4)
− p(n − 590) − · · · ≡ n2 σ1 (n) − nσ3 (n) (mod 7),

the numbers, 2, 51, 100, . . . being those of the forms

1 1
(7n − 1)(21n − 4), (7n + 1)(21n + 4);
2 2
and that

(3.5) p(7m − 2) ≡ 0 (mod 7).

Modulus 11.

4. It is obvious from (1.62) that

(4.1) QR = 1 + 11J.

Also n11 − n ≡ 0 (mod 11), and so, from (1.11) and (1.63),

(4.2) Q3 − 3R2 = 441Q3 + 250R2 + 11J


211 x2
 
x
= 691 + 65520 + + · · · + 11J
1 − x 1 − x2
2x2
 
x
= −2 + 48 + + · · · + 11J
1 − x 1 − x2
= −2P + 11J.

It is easily deduced that

(Q3 − R2 )5 = (Q3 − 3R2 )5 − Q(Q3 − 3R2 )3


(4.3) −R(Q3 − 3R2 )2 + 6QR + 11J
= P 5 − 3P 3 Q − 4P 2 R + 6QR + 11J.

[For

(Q3 − 3R2 )5 − Q(Q3 − 3R2 )3 − R(Q3 − 3R2 )2 + 6QR


= (Q3 − 3R2 )5 − Q3 R2 (Q3 − 3R2 )3 − Q3 R4 (Q3 − 3R2 )2 + 6Q6 R6 + 11J
= Q15 − 16Q12 R2 + 98Q9 R4 − 285Q6 R6 + 423Q3 R8 − 243R10 + 11J
= (Q3 − R2 )5 + 11J

by (4.1), and (4.3) then follows from (4.2).]


Congruence properties of partitions 295

Again, [if we multiply (1.74), (1.83), (1.92), and (1.93) by −1, 3, −4, and −1, and add, we
obtain, on rejecting multiples of 11,]

P 5 − 3P 3 Q − 4P 2 R + 6QR = −5Φ1,8 + 3Φ2,7 + 3Φ3,6 − Φ4,5 + 11J;

and from this and (4.3) follows


P∞
(Q3 − R2 )5 = − n=1 (5nσ7 (n) − 3n2 σ3 (n) − 3n3 σ5 (n)
(4.4)
+n4 σ1 (n))xn + 11J.

But (by the same argument which led to (2.2) and (3.2)) we have

f (x121 )
(4.5) (f (x))120 = + 11J.
f (x)

From (4.4) and (4.5)

f (x121 )
x5 = x5 (f (x))120 + 11J = 17285 x5 (f (x))120 + 11J
f (x)
= (Q3 − R2 )5 + 11J

X
= − (5nσ7 (n) − 3n2 σ5 (n) − 3n3 σ3 (n) + n4 σ1 (n))xn + 11J.
n=1

It now follows as before that


p(n − 5) − p(n − 126) − p(n − 247) + p(n − 610) + p(n − 852)
(4.6) − p(n − 1457) − · · · ≡ −n4 σ1 (n) + 3n3 σ3 (n) + 3n2 σ5 (n)
− 5nσ7 (n) (mod 11),

5, 126, 247, . . . being the numbers of the forms


1 1
(11n − 2)(33n − 5), (11n + 2)(33n + 5);
2 2
and in particular that

(4.7) p(11m − 5) ≡ 0 (mod 11).

5. If we are only concerned to prove (4.7), it is not necessary to assume quite so much.
d
Let us write ϑ for the operation x dx . Then∗ we have

1
(5.11) ϑP = (P 2 − Q),
12

Ramanujan, p.165 [pp. 181].
296 Paper 30

1
(5.12) ϑQ = (P Q − R),
3
1
(5.13) (P R − Q2 ).
ϑR =
2
From these equations we deduce [by straightforward calculation
864ϑ4 P = P 5 − 10P 3 Q − 15P Q2 + 20P 2 R + 4QR,
72ϑ3 Q = 5P 3 Q + 15P Q2 − 15P 2 R − 5QR,
2
24ϑ R = −14P Q2 + 7P 2 R + 7QR.
The left-hand side of each of these equations is of the form
dJ
x .
dx
Multiplying by 1,8, and 2, adding and rejecting multiples of 11, we find
dJ
(5.2) P 5 − 3P 3 Q + 2P 2 R = x + 11J.
dx
We have also, by (5.11),
dP
6P 2 R − 6QR = 72xR .
dx
But, differentiating (4.2), and using (4.1), we obtain
 
dP 2 dQ dR
72xR = 36xR −3Q + 6R + 11J
dx dx dx
dQ dR
= −108xQ + 216xR2 + 11J
dx dx
dJ
= x + 11J.
dx
Hence
dJ
(5.3) 6P 2 R − 6QR = x + 11J.
dx
From (5.2) and (5.3) we deduce
dJ
P 5 − 3P 3 Q − 4P 2 R + 6QR = x + 11J,
dx
and from (4.3)]
dJ
(5.4) (Q3 − R2 )5 = x + 11J.
dx
Finally, from (4.5) and (5.4),
f (x121 )
x5 = x5 (f (x))120 + 11J = (Q3 − R2 )5 + 11J
f (x)
dJ
= x + 11J.
dx
As the coefficient of x11m on the right-hand side is a multiple of 11, (4.7) follows immediately.
Papers written in collaboration with
G. H. Hardy
Papers 31 to 37
298 Paper 30
Une formule asymptotique pour le nombre des partitions de n

Comptes Rendus, 2 Jan. 1917

1. Les divers Problèmes de la théorie de la partition des nombres ont été étudiés surtout
par les mathématiciens anglais, Cayley, Sylvester et Macmahon∗ , qui les ont abordés d’un
point de vue purement algébrique. Ces auteurs n’y ont fait aucune application des méthodes
de la théorie des fonctions, de sorte qu’on ne trouve pas, dans la théorie en question, de
formules asymptotiques, telles qu’on en rencontre, par exemple, dans la théorie des nombres
premiers. Il nous semble donc que les résultats que nous allons faire connaı̂tre peuvent
présenter quelque nouveauté.

2. Nous nous sommes occupés surtout de la fonction p(n), nombre des partitions de n. On
a

1 X
f (x) = 2 3
= p(n)xn (|x| < 1).
(1 − x)(1 − x )(1 − x ) · · ·
0

Nous avons pensé d’abord à faire usage de quelque théorème de caractère Taubérien: on
désigne ainsi les théorèmes réciproques du théorème classique d’Abel et de ses générali-
sations. A cette catégorie appartient l’énoncé suivant:
an xn une série de puissances à coefficients POSITIFS, telle qu’on ait
P
Soit g(x) =

A
log g(x) ∼ ,
1−x
quand x tend vers un par valeurs positives. Alors on a
p
log sn = log (a0 + a1 + · · · an ) ∼ 2 (An),

quand n tend vers l’infini† .


En posant g(x) = (1 − x)f (x), on a
π2
A= ;
6
et nous en tirons
q
p(n) = e
π ( 23 n)(1+ǫ) , (1)

où ǫ tend vers zéro avec 1/n.



Voir le grand traité Combinatory Analysis de M. P. A. Macmahon (Cambridge, 1915-16).

Nous avons donné des généralisations étendues de ce théorẽme dans un mémoire qui doit paraı̂tre dans
un autre recueil. [The paper referred to is No. 34 of this volume; see, in particular, pp. 314 – 322.]
300 Paper 31

3. Pour pousser l’approximation plus loin, il faut recourir au théorème de Cauchy. Des
formules
1 f (x)
Z
p(n) = dx,
2πi xn+1
avec un chemin d’intégration convenable intérieur au cercle re rayon un, et
1
s
π2 4π 2
     
x 24 1
f (x) = p log exp f exp − (2)
(2π) x 6log (1/x) log (1/x)

(fournie par la théorie de la transformation linéaire des fonctions elliptiques), nous avons
tiré, en premier lieu, la formule vraiment asymptotique

1
q
p(n) ∼ P (n) = √ e ( 3 ) .
2
π n
(3)
4n 3
On a
p(10) = 42, p(20) = 627, p(50) = 204226, p(80) = 15796476;
P (10) = 48, P (20) = 692, P (50) = 217590, P (80) = 16606781.
Les valeurs correspondantes de P (n) : p(n) sont

1.145; 1.104; 1.065; 1.051 :

la valeur approximative est toujours en excès.

4. Mais nous avòns abouti plus tard à des résultats beaucoup plus satisfaisants. Nous
considérons la fonction
 q 
∞ 2 1
1 X d  cosh[π { 3 (n − 24 )}] − 1 
F (x) = √ q xn . (4)
π 2 1 dn  (n − ) 1 
24

En faisant usage des formules sommatoires que démontre M. E. Lindelöf dans son beau
livre Le calcul des résidus, on trouve aisément que F (x) (on parle, il va sans dire, de la
branche principale) a pour seul point singulier le point x = 1, et que la fonction
1
s
π2
   
x 24 1
F (x) − p log exp −1
(2π) x 6log (1/x)

est régulière pour x = 1. On est conduit naturellement à appliquer le théorème de Cauchy


à la fonction f (x) − F (x), et l’on trouve
q
π 1
{ 2 (n− 24 )} √ √
1 d e 3
p(n) = √ q + O(ek n
) = Q(n) + O(ek n
), (5)
2π 2 dn (n − 1
24 )
Une formule asymptotique pour le nombre des partitions de n 301


où k désigne un nombre quelconque supérieur à π/ 6. L’apporoximation, pour des valeurs
assez grandes de n, est très bonne: on trouve, en effet,

p(61) = 1121505, p(62) = 1300156, p(63) = 1505499;


Q(61) = 1121539, Q(62) = 1300121, Q(63) = 1505536.

La valeur approximative est, pour les valeurs suffsamment grandes de n, alternativement


en excès et en défaut.

5. On peut pousser ces calculs beaucoup plus loin. On forme des foncttions, analogues à
F (x), qui présentent, pour les valeurs
2 2 2
x = −1, e 3 πi , e− 3 πi , i, −i, e 5 πi , . . . ,

des singularités d’un type très analogue à celles que présente f (x). On soustrait alors
de f (x) une somme d’un nombre fini convenable de ces fonctions. On trouve ainsi, par
exemple,
q q
π 1
{ 23 (n− 24 )} 1
π 1
{ 2 (n− 24 )}
1 d e (−1)nd e 32
p(n) = √ q + q
2π 2 dn (n − 1 2π dn 1
24 ) (n − 24 )

q
1
  1
π { 23 (n− 24 )} √
3 2nπ π d e3
+ √ cos − q + O(ek n ), (6)
π 2 3 18 dn (n − 1 ) 24
q
où k désigne un nombre quelconque plus grand que 14 π 2
3.
Proof that almost all numbers n are composed of about
log log n prime factors

Proceedings of the London Mathematical Society, 2, XVI,1917,


Records for 14 Dec. 1916

A number n is described in popular language as a round number if it is composed of a


considerable number of comparatively small factors: thus 1200 = 24 · 3 · 52 would generally
be said to be a round number. It is a matter of common experience that round numbers
are exceedingly rare. The fact may be verified by anybody who will make a practice
of factorising numbers, such as the numbers of taxi-cab or railway carriages, which are
presented to his attention in moments of leisure. The object of this paper∗ is to provide
the mathematical explanation of this phenomenon.
Let πν (x) denote the number of numbers which do not exceed x and are formed of exactly
ν prime factors. There is an ambiguity in this definition, for we may count multiple factors
multiply or not. But the results are substantially the same on either interpretation.
It has been proved by Landau† that

x (log log x)ν−1


πν (x) ∼ , (1)
log x (ν − 1)!

as x → ∞, for every fixed value of ν. It is moreover obvious that

[x] = π1 (x) + π2 (x) + π3 (x) + · · · , (2)

and

(log log x)2


 
x
x= 1 + log log x + + ··· . (3)
log x 2!

Landau’s result shews that there is certain correspondence between the terms of the series
(2) and (3). The correspondence is far from exact. The first series is finite, for it is obvious
that πν (x) = 0 if ν > (log x/log 2; and the second is infinite. But it is reasonable to
anticipate a correspondence accurate enough to throw considerable light on the distribution
of the numbers less than x in respect of number of their prime factors.
The greatest term of the series (3) occurs when ν is about log log x. And if we consider
the block of terms for which
p p
log log x − φ(x) (log log x) < ν < log log x + φ(x) (log log x), (4)

The paper has been published in the Quarterly Journal of Mathematics, Vol. XLVIII, pp. 76 – 92[No.
35 of this volume].

See Handbuch der Lehre von Verteilung der Primzahalen, pp. 203 – 213.
Proof that almost all numbers n are composed of about log log n prime factors 303

where φ(x) is any function of x which tends to infinity with x, we find without difficulty
that it is these terms which contribute almost all the sum of the series: the ratio of their
sum to that of the remaining terms tends to infinity with x.
In this paper we shew that the same conclusion holds for the series (2). Let us consider all
numbers n which do not exceed x, and denote by xP the number of them which possess a
property P (n, x): this property may be a function of both n and x, or of one variable only.
If then xP /x → 1 when x → ∞, we say that almost all numbers less than x possess the
property P. And if P is a function of n only, we say simply that almost all numbers possess
the property P. This being so, we prove the following theorems.

1. Almost all numbers n less than x are formed of more than


p
log log x − φ(x) (log log x)

and less than p


log log x + φ(x) (log log x)
prime factors.

2. Almost all numbers n are formed of more than


p
log log n − φ(n) (log log n)

and less than p


log log n + φ(n) (log log n)
prime factors.

In these theorems φ is any function of x (or n) which tends to infinity with its argument: and
either theorem is true in whichever manner the factors of n are counted. The only serious
difficulty in the proof lies in replacing Landau’s asymptotic relations (1) by inequalities
valid for all values of ν and x.
Since log log n tends to infinity with extreme slowness, the theorems are fully sufficient to
explain the observations which suggested them.
Asymptotic formulæ in combinatory analysis

Proceedings of the London Mathematical Society, 2, XVI, 1917,


Records for 1 March 1917

A preliminary account of some of the contents of this paper∗ appeared in the Comptes
Rendus of January 2nd, 1917. The paper contains a full discussion and proof of the results
there stated. The asymptotic formula for p(n), the number of unrestricted partitions of n,
of which only the first three terms were given, is completed; and it is shewn that, by taking

a number of terms of order n, the exact value of p(n) can be obtained for all sufficiently
large values of n. Some account is also given of actual or possible applications of the method
used to other problems in Combinatory Analysis or the Analytic Theory of Numbers.


[No. 36 of this volume.]
Asymptotic formulæ for the distribution of integers of various
types∗

Proceedings of the London Mathematical Society, 2, XVI, 1917, 112 – 132

1. Statement of the problem.

1.1 We denote by q a number of the form

(1.11) 2a2 3a3 5a5 · · · pap ,

where 2, 3, 5, . . . , p are primes and

(1.111) a2 ≥ a3 ≥ a5 ≥ . . . ≥ ap ;

and by Q(x) the number of such numbers which do not exceed x: and our problem is that
of determining the order of Q(x). We prove that

" s #
2π log x
(1.12) Q(x) = exp {l + o(l)} √ ,
3 log log x

that is to say that to every positive ǫ corresponds an x0 = x0 (ǫ), such that


  s    s 
2π log x 2π log x
(1.121) √ −ǫ < log Q(x) < √ + ǫ ,
3 log log x 3 log log x

for x > x0 . The function Q(x) is thus of higher order than any power of log x, but of lower
order than any power of x.
The interest of the problem is threefold. In the first place the result itself, and the method
by which it is obtained, are curious and interesting in themselves. Secondly, the method of
proof is one which, as we shew at the end of the paper, may be applied to a whole class of
problems in the analytic theory of numbers: it enables us, for example, to find asymptotic
formulæ for the number of partitions of n into positive integers, or into different positive
integers, or into primes. Finally , the class of numbers q includes as a sub-class the “highly
composite” numbers recently studied by Mr.Ramanujan in an elaborate memoir in these
Proceedings† The problem of determining, with any precision, the number H(x) of highly

This paper was originally communicated under the title “A problem in the Analytic Theory of Numbers.”

Ramanujan, “Highly Composite Numbers,” Proc. London Math. Soc., Ser.2, Vol.XIV 1915, pp. 347 –
409 [No. 15 of this volume].
306 Paper 34

composite numbers not exceeding x appears to be one of extreme difficulty. Mr Ramanujan


has proved, by elementary methods, that the order of H(x) is at any rate greater than that
of log x∗ : but it is still uncertain whether or not the order of H(x) is greater than that of
any power of log x. In order to apply transcendental methods to this problem, it would be
necessary to study the properties of the function
X 1
H(s) = ,
hs
where h is a highly composite number, and we have not been able to make any progress
in this direction. It is therefore very desirable to study the distribution of wider classes of
numbers which include the highly composite numbers and possess some at any rate of their
characteristic properties. The simplest and most natural such class is that of the numbers
q; and here progress is comparatively easy, since the function
X 1
(1.13) Q(s) =
qs
possesses a product expression analogous to Euler’s product expression for ζ(s), viz.
∞  
Y 1
(1.14) Q(s) = ,
1
1 − ln−s

where ln = 2 · 3 · 5 · · · pn is the product of the first n primes.


We have not been able to apply to this problem the methods, depending on the theory of
functions of a complex variable, by which the Prime Number Theorem was proved. The
function Q(s) has the line σ = 0† as a line of essential singularities, and we are not able to
obtain sufficiently accurate information concerning the nature of these singularities. But
it is easy enough to determine the behaviour of Q(s) as a function of the real variable s;
and it proves sufficient for our purpose to determine an asymptotic formula for Q(s) when
s → 0, and then to apply a “Tauberian” theorem similar to those proved by Messrs Hardy
and Littlewood in a series of papers published in these Proceedings and elsewhere‡ .
This “Tauberian” theorem is in itself of considerable interest as being (so far as we are
aware) the first such theorem which deals with functions or sequences tending to infinity
more rapidly than any power of the variable.

2. Elementary results.
2.1 Let us consider, before proceeding further, what information concerning the order of
Q(x) can be obtained by purely elementary methods.
p
∗ log x (log log x)
As great as that of 3 : see p.385 of his memoir [p. 139 of this volume].
(log log log x) 2

We write as usual s = σ + it.

See, in particular, Hardy and Littlewood, “Tauberian theorems concerning power series and Dirichlet’s
series whose coefficients are positive,” Proc. London Math. Soc., Ser.2, Vol XIII, 1914, pp. 174 – 191; and
“Some theorems concerning Dirichlet’s series,” Messenger of Mathematics, Vol.XLIII, 1914, pp. 134 – 147.
Asymptotic formulæ for the distribution of integers of various types 307

Let

(2.11) ln = 2 · 3 · 5 · · · pn = eϑ(pn ) ,

where ϑ(x) is Tschebyshef’s function


X
ϑ(x) = log p .
p≤x

The class of numbers q is plainly identical with the class of numbers of the form

(2.12) l1b1 l2b2 · · · lnbn ,

where b1 ≥ 0, b2 ≥ 0, . . . , bn ≥ 0.
Now every b can be expressed in one and only one way in the form

(2.13) bi = ci,m 2m + ci,m−1 2m−1 + · · · + ci,0 ,

where every c is equal to zero or to unity. We have therefore


 P
m 
n ci,j 2j m Y
n m
c 2j j
Y Y Y
(2.14) q= lj=0
i
= li i,j = rj2 ,
i=1 j=0 i=1 j=0

say, where
c c c
(2.141) rj = l11,j l22,j · · · lnn,j .

Let r denote , generally, a number of the form

(2.15) r = l1c1 l2c2 · · · lncn ,

where every c is zero or unity : and R(x) the number of such numbers which do not exceed
x. If q ≤ x, we have
r0 ≤ x, r12 ≤ x, r24 ≤ x, . . .
The number of possible values of r0 , in formula (2.14), cannot therefore exceed R(x); the
1
number of possible values of r1 cannot exceed R(x 2 ); and so on. The total number of values
of q can therefore not exceed
1 1 −̟
(2.16) S(x) = R(x)R(x 2 )R(x 4 ) · · · R(x2 ),

where ̟ is the largest number such that


−̟ ̟
(2.161) x2 ≥ 2, x ≥ 22 .
308 Paper 34

Thus

(2.17) Q(x) ≤ S(x).

2.2 We denote by f and g the largest numbers such that

(2.211) lf ≤ x ,

(2.212) l1 l2 · · · lg ≤ x .
It is known∗ (and may be proved by elementary methods ) that constants A and B exist,
such that

(2.221) ϑ(x) ≥ Ax (x ≥ 2) ,

and

(2.222) pn ≥ Bn log n (n ≥ 1) .

We have therefore eAp f ≤ x ,

f log f = O(log x) ,

(2.23) log f = O(log log x) ;


g
P g
P
and ϑ(pν ) ≤ log x , pν = O(log x) ,
1 1

g
X
ν log ν = O(log x) , g 2 log g = O(log x) ,
1
s 
log x
(2.24) g=O .
log log x
But it is easy to obtain an upper bound for R(x) in terms of f and g. The number of
numbers l1 , l2 , . . . , not exceeding x, is not greater than f ; the number of products not
exceeding x, of pairs of such numbers, is a fortiori not greater than 12 f (f − 1); and so on.
Thus
f (f − 1) f (f − 1)(f − 2)
R(x) ≤ f + + + ··· ,
2! 3!
where the summation need be extended to g terms only, since

l1 l2 · · · lg lg+1 > x .

See Landau, Handbuch, pp. 79, 83, 214.
Asymptotic formulæ for the distribution of integers of various types 309

A fortiori, we have
f2 fg
R(x) ≤ 1 + f + + ··· + < (1 + f )g = eg log(1+f ) .
2! g!
Thus

(2.25) R(x) = eO(g log f ) = eO( log x log log x)
,
by (2.23) and (2.24). Finally, since
√ √ 1
log x log log x< log x log log x ,
2
it follows from (2.16) and (2.17) that
   
1 1 1 p
(2.26) Q(x) = exp O 1 + + + ··· + ̟ ( log x log log x)
2 4 2

= eO{ ( log x log log x)} .

2.3 A lower bound for Q(x) may be found as follows. If g is defined as in 2.2, we have

l1 l2 · · · lg ≤ x < l1 l2 · · · lg lg+1 .
It follows from the analysis of 2.2 that
lg+1 = eϑ(pg+1 ) = eO(g log g) ,
( g )
2
X
and l1 l2 · · · lg = exp ϑ(pν ) = eO(g log g) .
1
2
Thus x < eO(g log g) ;
which is only possible if g is greater than a constant positive multiple of
s 
log x
.
log log x
Now the numbers l1 , l2 , . . . , lg can be combined in 2g different ways, and each such combi-
nation gives a number q not greater than x. Thus
( s )
log x
(2.31) Q(x) ≥ 2g > exp K ,
log log x

where K is a positive constant. From (2.26) and (2.31) it follows that there are positive
constants K and L such that
s 
log x p
(2.32) K < log Q(x) < L ( log x log log x) .
log log x
310 Paper 34

The inequalities (2.32) give a fairly accurate idea as to the order of magnitude of Q(x).
But they are much less precise than the inequalities (1.121). To obtain these requires the
use of less elementary methods.

3. The behaviour of Q(s) when s → 0 by positive values.


3.1. From the fact, already used in 2.1, that the class of numbers q is identical with the
class of numbers of the form (2.12), it follows at once that

X 1 ∞  
Y 1
(1.14) Q(s) = = .
qs 1 − ln−s 1

Both series and product are absolutely convergent for σ > 0, and
1 1
(3.11) log Q(s) = φ(s) + φ(2s) + φ(3s) + · · · ,
2 3
where

X
(3.111) φ(s) = ln−s .
1

We have also
1 − 2−s −s 1 − 3
−s
−s −s 1 − 5
−s
(3.12) φ(s) = + 2 + 2 3 + ···
2s − 1 3s − 1 5s − 1
   
1 −s 1 1 −s −s 1 1
= s +2 − +2 3 − + ···
2 −1 3s − 1 2s − 1 5s − 1 3s − 1
∞ pn+1  
1 X −sϑ(pn ) d 1
Z
= s e s
dx
2 −1 1 pn dx x −1


1 xs−1
Z
= s −s e−sϑ(x) dx .
2 −1 2 (xs − 1)2

3.2. Lemma. If x > 1, s > 0, then


1 1 xs 1
(3.21) 2
− < s 2
< .
(s log x) 12 (x − 1) (s log x)2

Write xs = eu : then we have to prove that


1 1 eu 1
(3.22) 2
− < u 2
< 2
u 12 (e − 1) u
Asymptotic formulæ for the distribution of integers of various types 311

for all positive values of u ; or (writing w for 21 u) that


1 1 1 1
(3.23) 2
− < 2 < 2
w 3 sinh w w

for all positive values of w. But it is easy to prove that the function
1 1
f (w) = −
w 2 sinh2 w
is a steadily decreasing function of w, and that its limit when w → 0 is 31 ; and this establishes
the truth of the lemma.

3.3. We have therefore


1 1
(3.31) φ(s) = − φ1 (s) = − φ1 (s) + O(1) ,
2s −1 s log 2

where
∞ ∞
s2 e−sϑ(x) dx

1 1 dx 1
Z Z
(3.311) − e−sϑ(x) < φ1 (s) < .
s 2 (log x)2 12 x s 2 (log x)2 x

From the second of these inequalities, and (2.221), it follows that


1 ∞ e−Asx dx
Z ∞ −Asx
e−2As e
Z
φ1 (s) < 2
= −A dx
s 2 (log x) x s log 2 2 log x
Z ∞ −Asx
1 e
= −A dx + O(1) ;
s log 2 2 log x

and so that

e−Asx
Z
(3.32) φ(s) > A dx + O(1) .
2 log x

On the other hand there is a positive constant B, such that


ϑ(x) < Bx (x ≥ 2)∗ .
Thus
∞ ∞
s2 e−Bsx dx

1 1 −Bsx dx 1
Z Z
φ1 (s) > − e = + O(1)
s 2 (log x)2 12 x s 2 (log x)2 x
Z ∞ −Bsx
1 e
= −B dx + O(1) ;
s log 2 2 log x

Landau, Handbuch, loc.cit
312 Paper 34

and so

e−Bsx
Z
(3.33) φ(s) < B dx + O(1) .
2 log x

3.4. Lemma. If H is any positive number, then


Z ∞ −Hsx
e 1
J(s) = H dx ∼ ,
2 log x s log(1/s)
when s → 0.
Given any positive number ǫ, we can choose ξ and X, so that
Z ξ Z ∞
−Hx
He dx < ǫ , He−Hx dx < ǫ .
0 X

Now
Z ∞ Z √s Z ξ Z X Z ∞
He−Hu log(1/s)
 
1
s log J(s) = du = + √ + +
s 2s log u + log(1/s) 2s s ξ X
= j1 (s) + j2 (s) + j3 (s) + j4 (s) ,

say. And we have


Z √s
log(1/s) √
0 < j1 (s) < He−Hu du = O{ s log(1/s)} = o(1) ,
log 2 2s
Z ξ
0 < j2 (s) < 2 He−Hu du < 2ǫ ,
0
Z X
j3 (s) = He−Hu du + o(1) ,
ξ
Z ∞
0 < j4 (s) < He−Hu du < ǫ ;
X

and so
  Z ∞
1 − s log 1 J(s) = −Hu

He du − j1 (s) − j2 (s) − j3 (s) − j4 (s)
s
0
< 5ǫ + o(1) < 6ǫ ,

for all sufficiently small values of s.

3.5. From (3.32), (3.33), and lemma just proved, it follows that
X 1
(3.51) φ(s) = ln−s ∼ .
s log(1/s)
Asymptotic formulæ for the distribution of integers of various types 313

From this formula we can deduce an asymptotic formula for log Q(s). We choose N so
that
X 1
(3.52) < ǫ,
n2
N <n

and we write
X1 X X X X
(3.53) log Q(s) = φ(ns) = + + +
n √ √
1≤n≤N N <n<1/ s 1/ s≤n≤1/s 1/s<n

= Φ1 (s) + Φ2 (s) + Φ3 (s) + Φ4 (s) ,

say.
In the first place
N
1 + o(1) X 1
(3.541) Φ1 (s) = .
s log(1/s) 1 n2

In the second place  


1
φ(ns) = O ,
ns log(1/ns)
and
1
log(1/ns) > log(1/s),
2

if N < n < 1/ s. It follows that a constant K exists such that

K X 1 Kǫ
(3.542) Φ2 (s) < < .
s log(1/s) n2 s log(1/s)
N <n

Thirdly, s ≤ ns ≤ 1 in Φ3 (s), and a constant L exists such that

L
φ(ns) < √ .
s log(1/s)

Thus
1/s
L X1 2L
(3.543) Φ3 (s) < √ <√ ,
s log(1/s) n s
1

for all sufficiently small values of s.


Finally, in Φ4 (s) we have ns > 1, and a constant M exists such that

φ(ns) < M 2−ns .


314 Paper 34

Thus
X 2−ns X sM
(3.544) Φ4 (s) < M < sM 2−ns < = O(1) .
n 1 − 2−s
1/s<n 1/s<n

From (3.53), (3.541)–(3.544), and (3.52) it follows that


  2  
1 π ′
(3.55) log Q(s) = {1 + o(1)} +ρ +ρ
s log(1/s) 6
 
1
+O √ + O(1) ,
s log(1/s)

where
|ρ| < ǫ , |ρ′ | < Kǫ.
Thus
π2
(3.56) log Q(s) ∼ ,
6s log(1/s)
or
π2
 
(3.57) Q(s) = exp {1 + o(1)} .
6s log(1/s)

4. A Tauberian theorem.
4.1. The passage from (3.57) to (1.12) depends upon a theorem of the “Tauberian” type.
Theorem A. Suppose that

(1) λ1 ≥ 0 , λn > λn−1 , λn → ∞ ;

(2) λn /λn−1 → 1 ;

(3) an ≥ 0 ;

(4) A > 0 , a > 0 ;

an e−λn s is convergent for s > 0 ;


P
(5)
h −β i
(6) f (s) = an e−λn s = exp {1 + o(1)}As−α log 1s
P 
,

when s → 0. Then

An = a1 + a2 + · · · + an = exp[{1 + o(1)}Bλα/(1+α)
n (log λn )−β/(1+α) ] ,
Asymptotic formulæ for the distribution of integers of various types 315

where
B = A1/(1+α) α−α/(1+α) (1 + α)1+[β/(1+α)] ,
when n → ∞.
We are given that
 −β  −β
−α 1 −α 1
(4.11) (1 − δ)As log < log f (s) < (1 + δ)As log ,
s s

for every positive δ and all sufficiently small values of s; and we have to shew that

(4.12) (1 − ǫ)Bλnα/(1+α) (log λn )−β/(1+α) < log An


< (1 + ǫ)Bλα/(1+α)
n (log λn )−β/(1+α) ,

for every positive ǫ and all sufficiently large values of n.


In the argument which follows we shall be dealing with three variables, δ, s and n (or m),
the two latter variables being connected by an equation or by inequalities, and with an
auxiliary parameter ζ. We shall use the letter η, without a suffix, to denote generally a
function of δ, s and n (or m)∗ , which is not the same in different formulæ , but in all cases
tends to zero when δ and s tend to zero and n (or m) to infinity; so that, given any positive
ǫ, we have
0 < |η| < ǫ ,
for 0 < δ < δ0 , 0 < s < s0 , n > n0 .
We shall use the symbol ηζ to denote a function of ζ only which tends to zero with ζ, so
that
0 < |ηζ | < ǫ ,
if ζ is small enough. It is to be understood that the choice of a ζ to satisfy certain conditions
is in all cases prior to that of δ, s and n (or m). Finally, we use the letters H, K, . . . to
denote positive numbers independent of these variables and of ζ.
The second of the inequalities (4.12) is very easily proved. For

(4.131) An e−λn s < a1 e−λ1 s + a2 e−λ2 s + · · · + an e−λn s


(  −β )
1
< f (s) < exp (1 + δ)As−α log ,
s

(4.132) An < exp χ ,


where
 −β
−α 1
(4.1321) χ = (1 + δ)As log + λn s.
s

η may, of course, in some cases be a function of some of these variables only.
316 Paper 34

We can choose a value of s, corresponding to every large value of n, such that

1 −β 1 −β
   
(4.14) (1 − δ)Aαs−1−α log < λn < (1 + δ)Aαs−1−α log .
s s

From these inequalities we deduce, by an elementary process of approximation,

1 β/(1+α)
 
(4.151) (1 − η)(Aα)−1/(1+α) λn1/(1+α) log
s
1 β/(1+α)
 
1 −1/(1+α) 1/(1+α)
< < (1 + η)(Aα) λn log ,
s s

1−η 1 1+η
(4.152) log λn < log < log λn ,
1+α s 1+α

αB −1/(1+α)
(4.153) (1 − η) λ (log λn )−β/(1+α)
1+α n
αB −1/(1+α)
< s < (1 + η) λ (log λn )−β/(1+α) ,
1+α n

(4.154) χ < (1 + η)Bλα/(1+α)


n (log λn )−β/(1+α) .
We have therefore

(4.16) log An < (1 + ǫ)Bλα/(1+α)


n (log λn )−β/(1+α) ,

for every positive ǫ and all sufficiently large values of n∗ .

4.2. We have
X X
(4.21) f (s) = an e−λn s = An (e−λn s − ǫ−λn+1 s )

X Z λn+1 Z ∞
−sx
=s An e dx = s A(x)e−sx dx ,
1 λn 0

where A(x) is the discontinuous function defined by

A(x) = An (λn ≤ x < λn+1 )† ,



We use the second inequality (4.12) in the proof of the first. It would be sufficient for our purpose to
begin by proving a result cruder than (4.16), with any constant K on the right-hand side instead of (1 + ǫ)B.
But it is equally easy to obtain the more precise inequality. Compare the argument in the second of the
two papers by Hardy and Littlewood quoted on p.114 [p. 306 of this volume] (pp. 143 et seq.)

Compare Hardy and Riesz, “The General Theory of Dirichlet’s Series,” Cambridge Tracts in Mathe-
matics, No.18, 1915, p.24.
Asymptotic formulæ for the distribution of integers of various types 317

so that, by (4.16),

(4.22) log A(x) < (1 + ǫ)Bxα/(1+α) (log x)−β/(1+α)

for every positive ǫ and all sufficiently large values of x. We have therefore
(  )
1 −β
 Z ∞
−α
(4.23) exp (1 − δ)As log <s A(x)e−sx dx
s 0
(  −β )
1
< exp (1 + δ)As−α log
s

for every positive δ and all sufficiently small values of s.


We define λx , a steadily increasing and continuous function of the continuous variable x,
by the equation
λx = λn + (x − n)(λn+1 − λn ) (n ≤ x ≤ n + 1) .
We can then choose m so that
1 1 + α 1/(1+α)
(4.24) = λ (log λm )β/(1+α) .
s αB m
We shall now shew that the limits of the integral in (4.23) may be replaced by (1 − ζ)λm
and (1 + ζ)λm , where ζ is an arbitrary positive number less than unity.
We write
Z ∞
(4.25) J(s) = s A(x)e−sx dx
0
(Z )
λm /H
Z (1−ζ)λm
Z Z (1+ζ)λm Z Hλm ∞
=s + + + +
0 λm /H (1−ζ)λm (1+ζ)λm Hλm

= J1 + J2 + J3 + J4 + J5 ,

where H is a constant, in any case greater than 1, and large enough to satisfy certain
further conditions which will appear in a moment; and we proceed to shew that J1 , J2 , J4
and J5 are negligible in comparison with the exponentials which occur in (4.23), and so in
comparison with J3 .

4.3 The integrals J1 and J5 are easily disposed of. In the first place we have
λm /H  
λm
Z
−sx
(4.31) J1 = s A(x)e dx < A
0 H
( α/(1+α)  )
λm −β/(1+α)
 
λm
< exp (1 + δ)B log ,
H H
318 Paper 34

by (4.22)∗ . It will be found, by a straightforward calculation, that this expression is less


than
(  −β )
1
(4.32) exp (1 + η)A(1 + α)H −α/(1+α) s−α log ,
s

and is therefore certainly negligible if H is sufficiently large.


Thus J1 is negligible. To prove that J5 is negligible we prove first that

sx > 4Bxα/(1+α) (log x)−β/(1+α) ,

if x > Hλm and H is large enough† . It follows that


Z ∞ Z∞
−sx
J5 = s A(x)e dx < s exp{(1 + δ)Bxα/(1+α) (log x)−β/(1+α) − sx}dx
Hλm
Hλm

1
Z
1
< s e− 2 sx dx = ,
0 2
and is therefore negligible.

4.4 The integrals J2 and J4 may be discussed in practically the same way, and we may
confine ourselves to the latter. We have
Z Hλm Z Hλm
(4.41) J4 (s) = s A(x)e−sx dx < s eψ dx ,
(1+ζ)λm (1+ζ)λm

where

(4.411) ψ = (1 + δ)Bxα/(1+α) (log x)−β/(1+α) − sx .

The maximum of the function ψ occurs for x = x0 , where


1 1 + α 1/(1+α)
(4.42) = (1 + η) x (log x0 )β/(1+α) .
s αB 0
From this equation, and (4.24), it plainly results that

(4.43) (1 − η)λm < x0 < (1 + η)λm ,

and that x0 falls (when δ and s are small enough) between (1 − ζ)λm and (1 + ζ)λm .
Let us write x = x0 + ξ in J4 . Then

1 d2 α/(1+α)
ψ(x) = ψ(x0 ) + (1 + δ)Bξ 2 2 {x1 (log x1 )−β/(1+α) } ,
2 dx1

With δ in the place of ǫ.

We suppress the details of the calculation, which is quite straightforward.
Asymptotic formulæ for the distribution of integers of various types 319

where x0 < x1 < x and a fortiori

(1 − ζ)λm < x1 < Hλm .

It follows that
d2 α/(1+α)
(4.44) 2 {x1 (log x1 )−β/(1+α) } < −Kλm
α/(1+α)−2
(log λm )−β/(1+α) .
dx1
On the other hand, an easy calculation shews that

1 −β 1 −β
   
−α −α
(4.45) (1 − η)As log < ψ(x0 ) < (1 + η)As log .
s s
Thus
(  −β )
−α 1
(4.46) J4 < exp (1 + η)As log
s
Z ∞
× exp{−Lξ 2 λm
α/(1+α)−2
(log λm )−β/(1+α) }dξ
(ζ−η)λm
(  −β )
−α 1
< exp (1 + η)As log − M ζ 2 λα/(1+α)
m (log λm )−β/(1+α)
s
(  −β )
1
< exp (1 + η − N ζ 2 )As−α log .
s

Since ζ is independent of δ and s, this inequality shews that J4 is negligible; and a similar
argument may be applied to J2 .

4.5 We may therefore replace the inequalities (4.23) by


(  −β )
1
(4.51) exp (1 − δ)As−α log
s
( −β )
(1+ζ)λm 
1
Z
−sx −α
<s A(x)e dx < exp (1 + δ)As log .
(1−ζ)λm s

Since A(x) is a steadily increasing function of x, it follows that


(  −β ) Z (1+ζ)λm
−α 1
(4.521) exp (1 − δ)As log < sA{(1 + ζ)λm } e−sx dx ,
s (1−ζ)λm

( −β )
 (1+ζ)λm
1
Z
−α
(4.522) exp (1 + δ)As log > sA{(1 − ζ)λm } e−sx dx ;
s (1−ζ)λm
320 Paper 34

or
(  −β )
ζsλm −ζsλm −α 1
(4.531) (e −e )A{(1 − ζ)λm } < exp (1 + δ)As log + λm s ,
s
(  −β )
1
(4.532) (eζsλm − e−ζsλm )A{(1 + ζ)λm } > exp (1 − δ)As−α log + λm s .
s
If we substitute for s, in terms of λm , in the right-hand sides of (4.531) and (4.532), we
obtain expressions of the form

exp{(1 + η)Bλα/(1+α)
m (log λm )−β/(1+α) } .

On the other hand


eζsλm − e−ζsλm
is of the form
exp{ηζ λα/(1+α)
m (log λm )−β/(1+α) }.
We have thus

(4.541) A{(1 − ζ)λm } < exp{(1 + ηζ + η)Bλα/(1+α)


m (log λm )−β/(1+α) } ,

(4.542) A{(1 + ζ)λm } > exp{(1 − ηζ − η)Bλα/(1+α)


m (log λm )−β/(1+α) } .
Now let ν be any number such that

(4.55) (1 − ζ)λm ≤ λν ≤ (1 + ζ)λm .

Since λn /λn−1 → 1, it is clear that all numbers n from a certain point onwards will fall
among the numbers ν. It follows from (4.541) and (4.542) that

(4.56) exp{(1 − ηζ − η)(1 − ηζ )Bλα/(1+α)


ν (log λν )−β/(1+α) } < A(λν )
< exp{(1 + ηζ + η)(1 + ηζ )Bλα/(1+α)
ν (log λν )−β/(1+α) } ;

and therefore that, given ǫ we can choose first ζ and then n0 so that

(4.57) exp{(1 − ǫ)Bλα/(1+α)


n (log λn )−β/(1+α) } < A(λn )
< exp{(1 + ǫ)Bλα/(1+α)
n (log λn )−β/(1+α) } ,

for n ≥ n0 . This completes the proof of the theorem.

4.6 There is of course a corresponding “Abelian” theorem, which we content ourselves with
enunciating. This theorem is naturally not limited by the restriction that the coefficients
an are positive.
Asymptotic formulæ for the distribution of integers of various types 321

Theorem B. Suppose that

(1) λ1 ≥ 0 , λn > λn−1 , λn → ∞ ;

(2) λn /λn−1 → 1 ;

(3) A > 0 , 0 < α < 1 ;

(4) An = α1 + α2 + · · · + αn = exp {1 + o(1)}Aλαn (log λn )−β ,


 

an e−λn s is convergent for s > 0, and


P
when n → ∞. Then the series
 −β/(1−α)
X
−λn s −α/(1−α) 1
f (s) = an e = exp[{1 + o(1)}Bs log ],
s

where
B = A1/(1−α) αα/(1−α) (1 − α)1+[β/(1−α)] ,
when s → 0.
The proof of this theorem, which is naturally easier than that of the correlative Taube-
rian theorem, should present no difficulty to anyone who has followed the analysis which
precedes.

4.7 The simplest and most interesting cases of Theorems A and B are those in which

λn = n , β = 0.

It is then convenient to write x for e−s . We thus obtain

Theorem C. If A > 0, 0 < α < 1, and

log An = log(a1 + a2 + · · · + an ) ∼ Anα ,

an xn is convergent for |x| < 1, and


P
then the series
X
log f (x) = log( an xn ) ∼ B(1 − x)−α/(1−α) ,

where
B = (1 − α)αα/(1−α) A1/(1−α) ,
when x → 1 by real values.
If the coefficients are positive the converse inference is also correct. That is to say, if

A > 0,α > 0,

and
log f (x) ∼ A(1 − x)−α ,
322 Paper 34

then
log An ∼ Bnα/(1+α) ,
where
B = (1 + α)α−α/(1+α) A1/(1+α) .

5. Application to our problem, and to other problems in the Theory of Numbers.

5.1 We proved in 3 that

π2
(3.56) log Q(s) ∼ .
6s log (1/s)

In Theorem A take
π2
λn = log n , A= , α = 1, β = 1.
6
Then all the conditions of the theorem are satisfied. And An is Q(n), the number of numbers
q not exceeding n. We have therefore
s 
log n
(5.11) log Q(n) ∼ B ,
log log n

where
s
π2

3 2π
(5.12) B=2 2 =√ .
6 3

5.2 The method which we have followed in solving this problem is one capable of many
other interesting applications.
Suppose, for example, that Rr (n) is the number of ways in which n can be represented as
the sum of any number of r th powers of positive integers∗ .
We shall prove that
    r/(r+1)
1 1 1
(5.21) log Rr (n) ∼ (r + 1) Γ +1 ζ +1 n1/(r+1) .
r r r

Thus 28 = 33 + 13 = 3 · 23 + 4 · 13 = 2 · 23 + 12 · 13 = 23 + 20 · 13 = 28 · 13 : and R3 (28) = 5.
The order of the powers is supposed to be indifferent, so that (e.g.) 33 + 13 and 13 + 33 are not reckoned as
separate representations.
Asymptotic formulæ for the distribution of integers of various types 323

In particular, if P (n) = R1 (n) is the number of partitions of n, then


r
2n
(5.22) log P (n) ∼ π .
3

We need only sketch the proof, which is in principle similar to the main proof of this paper.
We have
∞  
X 1
Rr (n)e−ns = Π∞ 1 ,
1 − e−sν r
1

and so
∞  
X
−ns ∞ 1
(5.23) f (s) = {Rr (n) − Rr (n − 1)}e = Π2 .∗
1 − e−sν r
1

It is obvious that Rr (n) increases with n and that all the coefficients in f (s) are positive.
Again,
∞   ∞
X 1 X r 1 r
(5.24) log f (s) = log r = (e−sν + e−2sν + · · ·)
2
1 − e−sν 2
2

X 1
= φ(ks) ,
k
k=1

where

r
X
(5.241) φ(s) = e−sν .
2

But
 
1
(5.25) φ(s) ∼ Γ + 1 s−1/r ,
r

when s → 0; and we can deduce, by an argument similar to that of 3.5, that


   
1 1
(5.26) log f (s) ∼ Γ +1 ζ + 1 s−1/r .
r r

We now obtain (5.21) by an application of Theorem A, taking


   
1 1 1
λn = n , α = , β = 0 , A = Γ +1 ζ +1 .
r r r

Rr (0) is to be interpreted as zero.
324 Paper 34

In a similar manner we can shew that, if S(n) is the number of partitions of n into different
positive integers, so that
X
S(n)e−ns = (1 + e−s )(1 + e−2s )(1 + e−3s ) + · · ·
1
= ,
(1 − e−s )(1 − e−3s )(1 − e−5s ) · · ·
then
r
n
(5.27) log S(n) ∼ π ;
3

that if Tr (n) is the number of representations of n as the sum of r th powers of primes,


then
    r/(r+1)
1 1
(5.28) log Tr (n) ∼ (r + 1) Γ +2 ζ +1 n1/(r+1) (log n)−r/(r+1) ;
r r

and, in particular, that if T (n) = T1 (n) is the number of partitions of n into primes, then
s 
2π n
(5.281) log T (n) ∼ √ .
3 log n

Finally, we can shew that if r and s are positive integers, a > 0, and 0 ≤ b ≤ 1, and
X {(1 + ax)(1 + ax2 )(1 + ax3 ) · · ·}r
(5.291) φ(n)xn = ,
{(1 + bx)(1 + bx2 )(1 + bx3 ) · · ·}s
then
p
(5.292) log φ(n) ∼ 2 (cn) ,

where
a b
log(1 + t) log(1 − t)
Z Z
(5.2921) c=r dt − s dt .
0 t 0 t
In particular, if a = 1, b = 1 , and r = s, we have
X
(5.293) φ(n)xn = (1 − 2x + 2x4 − 2x9 + · · ·)−r ,

p
(5.294) log φ(n) ∼ π (rn) .

[Added March 28th , 1917. – Since this paper was written M.G.Valiron (“Sur la croissance
du module maximum des séries entières,” Bulletin de la Société mathématique de France,
Asymptotic formulæ for the distribution of integers of various types 325

Vol.XLIV, 1916, pp. 45 – 64) has published a number of very interesting theorems con-
cerning power-series which are more or less directly related to ours. M.Valiron considers
power-series only, and his point of view is different from ours, in some respects more re-
stricted and in others more general.
He proves in particular that the necessary and sufficient conditions that
A
log M (r) ∼
,
(1 − r)α

an xn for |x| = r, are that


P
where M (r) is the maximum modulus of f (x) =
 n α/(1+α)
log |an | < (1 + ǫ)(1 + α)A1/(1+α)
α
for n > n0 (ǫ), and
 n α/(1+α)
log |an | > (1 − ǫp )(1 + α)A1/(1+α)
α
for n = np (p = 1, 2, 3, . . .), where np+1 /np → 1 and ǫp → 0 as p → ∞.
M.Valiron refers to previous, but less general or less precise, results given by Borel (Leçons
sur les séries à termes posı́tifs, 1902, Ch. V) and by Wiman (“Über dem Zusammen-
hang zwischen dem Maximal-betrage einer analytischen Funktion und dem grössten Gliede
der zugehörigen Taylor’schen Reihe,” Acta Mathematica, Vol.XXXVII, 1914, pp. 305 –
326). We may add a reference to Le Roy, “Valeurs asymptotiques de certaines séries
procédant suivant les puissances entières et positives d’une variable réelle,” Bulletin des
sciences mathématiques, Ser.2, Vol. XXIV, 1900, pp.245 – 268.
We have more recently obtained results concerning P (n), the number of partitions of n, far
more precise than (5.22). A preliminary account of these researches has appeared, under
the title “Une formule asymptotique pour le nombre des partitions de n,” in the Comptes
Rendus of January 2nd , 1917∗ ; and a fuller account has been presented to the Society. See
Records of Proceedings at Meetings, March 1st , 1917† .]


[No. 31 of this volume.]

[No. 33 of this volume; see also No. 36.]
The normal number of prime factors of a number n

Quarterly Journal of Mathematics, XLVIII, 1917, 76 – 92

I.
Statement of the problem.

1.1. The problem with which we are concerned in this paper may be stated roughly as
follows: What is the normal degree of compositeness of a number n? We shall prove a
number of theorems the general result of which is to shew that n is, as a rule, composed of
about log log n factors.
These statements are vague, and we must define our problem more precisely before we
proceed further.
1.2. There are two ways in which it is natural to measure the “degree of compositeness”
of n, viz.

(1) by its number of divisors,


(2) by its number of prime factors.

In this paper we adopt the second point of view. A distinction arises according as multiple
factors are or are not counted multiply. We shall denote the number of different prime
factors by f (n), and the total number of prime factors by F (n), so that (e.g.)

f (23 32 5) = 3, F (23 32 5) = 6.

With regard to these functions (or any other arithmetical functions of n) four questions
naturally suggest themselves.
(1) In the first place we may ask what is the minimum order of the function considered. We
wish to determine an elementary function of n, with as low a rate of increase as possible,
such that (e.g.)
f (n) ≤ φ(n)
for an infinity of values of n. This question is, for the functions now under consideration,
trivial; for it is plain that
f (n) = F (n) = 1
when n is a prime.
(2) Secondly, we may ask what is the maximum order of the function. This question also
is trivial for F (n) : we have
log n
F (n) =
log 2
whenever n is of the form 2k ; and there is no φ(n) of slower increase which satisfies the
conditions. The answer is less obvious in the case of f (n) : but, supposing n to be the
The normal number of prime factors of a number n 327

product of the first k primes, we can shew (by purely elementary reasoning) that, if ǫ is
any positive number, we have
log n
f (n) < (1 + ǫ)
log log n
for all sufficiently large values of n, and
log n
f (n) > (1 − ǫ)
log log n

for an infinity of values; so that the maximum order of f (n) is

log n
.
log log n

It is worth mentioning, for the sake of comparison, that the minimum order of d(n), the
number of divisors of n, is 2, while the maximum order lies between

2(1−ǫ) log n/ log log n , 2(1+ǫ) log n/ log log n

for every positive value of ǫ.∗


(3) Thirdly, we may ask what is the average order of the function. It is well known – to
return for a moment to the theory of d(n) – that

(1.21) d(1) + d(2) + · · · + d(n) ∼ n log n.

This result, indeed, is almost trivial, and far more is known; it is known in fact that
1
(1.22) d(1) + d(2) + · · · + d(n) = n log n + (2γ − 1)n + O(n 3 log n),

this result being one of the deepest in the analytic theory of numbers. It would be natural,
then, to say that the average order of d(n) is log n.
A similar problem presents itself for f (n) and F (n); and it is easy to shew that, in this
sense, the average order of f (n) and F (n) is log log n. In fact it may be shewn, by purely
elementary methods, that
 
n
(1.23) f (1) + f (2) + · · · + f (n) = n log log n + An + O ,
log n
 
n
(1.24) F (1) + F (2) + · · · + F (n) = n log log n + Bn + O ,
log n
where A and B are certain constants.

Wigret, “Sur l’ordre de grandeur du nombre des diviseurs d’un entier,” Arkiv for matematik, Vol. III
(1907), No. 18, pp. 1 – 9. See Ramanujan, “Highly Composite Numbers,” Proc. London math. soc., Ser.
2, Vol. XIV (1915), pp. 347 – 409 [No. 15 of this volume] for more precise results.
328 Paper 35

This problem, however, we shall dismiss for the present, as results still more precise than
(1.23) and (1.24) can be found by transcendental methods.
(4) Fourthly, we may ask what is the normal order of the function. This phrase requires a
little more explanation.
Suppose that N (x) is the number of numbers, not exceeding x, which possess a certain
property P . This property may be a function of n only, or of x only, of both n and x:
we shall be concerned only with cases in which it is a function of one variable alone. And
suppose further that

(1.25) N (x) ∼ x

when x → ∞. Then we shall say, if P is a function of n only, that almost all numbers
possess the property, and, if P is a function of x only, that almost all numbers less than x
possess the property. Thus, to take a trivial example, almost all numbers are composite.
If then g(n) is an arithmetical function of n and φ(n) an elementary increasing function,
and if, for every positive ǫ, we have

(1.26) (1 − ǫ)φ(n) < g(n) < (1 + ǫ)φ(n)

for almost all values of n, we shall say that the normal order of g(n) is φ(n).
It is in no way necessary that a function should possess either a determinate average order
or a determinate normal order, or that one should be determinate when the other is, or
that, if both are determinate, they should be the same. But we shall find that each of the
functions f (n) and F (n) has both the average order and the normal order log log n.
The definitions just given may be modified so as to apply only to numbers of a special class.
If Q(x) is the number of numbers of the class not exceeding x, and N (x) the number of
these numbers which possess the property P , and

(1.27) N (x) ∼ Q(x),

then we shall say that almost all of the numbers (or almost all of the numbers less than
x) possess the property.

II.
“Quadratfrei” numbers.

2.1. It is most convenient to begin by confining our attention to quadratfrei numbers,


numbers, that is to say, which contain no prime factor raised to a power higher than the
first. It is well known that the number Q(x) of such numbers, not exceeding x, satisfies the
relation
The normal number of prime factors of a number n 329

6 ∗
(2.11) Q(x) ∼ x
π2
We shall denote by πν (x) the number of numbers which are products of just ν different
prime factors and do not exceed x, so that π1 (x) = π(x) is the number of primes not
exceeding x, and

X
πν (x) = Q(x).
1

It has been shewn by Landau† that

x (log log x)ν−1


(2.12) πν (x) ∼
log x (ν − 1)!

for every fixed value of ν. In what follows we shall require, not an asymptotic equality,
but an inequality satisfied for all values of ν and x.

2.2. Lemma A. There are absolute constants C and K such that


Kx (log log x + C)ν
(2.21) πν+1 (x) <
log x ν!

for
ν = 0, 1, 2, . . . , x ≥ 2.
The inequality (2.21) is certainly true when ν = 0, whatever the value of C. It is known
(and may be proved by elementary methods‡ ) that
X 1
(2.221) < log log x + B
p
p≤x

and
X log p
(2.222) < H log x,
p
p≤x

where B and H are constants. We shall prove by induction that (2.21) is true if

(2.223) C > B + H.


Landau, Handbuch, p. 581.

Landau, Handbuch, pp. 203 et seq.

In applying (2.21) we assume that x/p ≥ 2. If x/p < 2, πν (x/p) is zero.
330 Paper 35

Consider the numbers which do not exceed x and are comprised in the table
2 · p1 · p2 · · · pν ,
3 · p1 · p2 · · · pν ,
5 · p1 · p2 · · · pν ,
······,
P · p1 · p2 · · · pν ,
where p1 , p2 , . . . , pν are, in each row, different primes arranged in ascending order of mag-
nitude, and where pν ≥ 2 in the first row, pν ≥ 3 in second, pν ≥ 5 in the third, and so on.

It is plain that P ≤ x. The total number of numbers in the table is plainly not greater
than  
X x
πν .
2
p
p ≤x

If now ω1 , ω2 , . . . are primes and


ω1 < ω2 · · · < ων+1, ω1 ω2 · · · ων+1≤x,

the number ω1 ω2 · · · ων+1 will occur at least ν times in the table, once in the row in which
the first figure is ω1 once in that in which it is ω2 , . . ., once in that in which it is ων . We
have therefore
 
X x
(2.23) νπν+1 (x) ≤ πν .
2
p
p ≤x

Assuming, then, that (2.21) is true when ν is replaced by ν − 1, we obtain∗


   ν−1
Kx X 1 x
(2.24) πν+1 (x) < log log +C
ν! 2 p log(x/p) p
p ≤x

Kx(log log x + C)ν−1 X 1


< .
ν! 2
p log(x/p)
p ≤x

But
 
1 1 log p log p 1 2 log p
(2.25) = + 1+ + ··· ≤ + ,
log x − log p log x (log x)2 log x log x (log x)2
1
since log p ≤ 2 log x; and so
X 1 1 X 1 2 X log p
(2.26) ≤ +
p log(x/p) log x 2 p (log x)2 2 p
p2 ≤x p ≤x p ≤x

log log x + B H log log x + C


< + < .
log x log x log x

It may be well to observe explicitly that log log 2 + C is positive.
The normal number of prime factors of a number n 331

Substituting in (2.24), we obtain (2.21).

2.3. We can now prove one of our main theorems.


Theorem A. Suppose that φ is a function of x which tends steadily to infinity with x.
Then
√ √
(2.31) log log x − φ (log log x) < f (n) < log log x + φ (log log x)

for almost all quadratfrei numbers n less than x.


It is plainly enough to prove that
X
(2.321) S1 = πν+1 (x) = o(x),

ν<llx−φ (llx)

X
(2.322) S2 = πν+1 (x) = o(x).∗

ν>llx+φ (llx)

It will be sufficient to consider one of the two sums S1 , S2 , say the latter; the discussion
of S1 proceeds on the same lines and is a little simpler. We have

x X (llx + C)ν
(2.33) S2 < K .
log x √ ν!
ν>llx+φ (llx)

Write
log log x + C = ξ.
Then the condition that p
ν > log log x + φ log log x,
where φ is some function of x which tends steadily to infinity with x, is plainly equivalent
to the condition that

ν > ξ + Ψ ξ,
where Ψ is some function of ξ which tends steadily to infinity with ξ; and so what we have
to prove is that
X ξν
(2.34) S= = o(eξ ).
√ ν!
ν>ξ+Ψ ξ

We choose a positive number δ so small that

δ δ2 δ3
(2.35) + + + · · · < 14 ,
2·3 3·4 4·5

We shall sometimes write lx, llx, . . . , instead of log x, log log x, . . . , in order to shorten our formulæ.
332 Paper 35

and write
X ξν X ξν
(2.36) S= + = S ′ + S ′′ ,
√ ν! ν!
ξ+Ψ ξ<ν≤(1+δ)ξ ν>(1+δ)ξ

say. In the first place we have


ξν ξ2
 
ξ
(2.371) S < 1
′′
1+ + + ···
ν1 ! ν1 + 1 (ν1 + 1)(ν1 + 2)
ξ1ν 1 + δ ξ ν1
 
1 1
< 1+ + + · · · = ,
ν1 ! 1 + δ (1 + δ)2 δ ν1 !

where ν1 is the smallest integer greater than (1 + δ)ξ. It follows that


K
(2.372) S ′′ < √ eν1 (log ξ−log ν1 +1)
δ ν1
K
< √ e∆ξ,
δ ξ
where the K’s are absolute constants and

∆ = (1 + δ) log(1 + δ) − δ.

And since

(1 + δ) log(1 + δ) − δ > (1 + δ)(δ − 12 δ2 ) − δ = 12 δ2 (1 − δ) = η,

say, we obtain
K
(2.373) S ′′ < √ e(1−η)ξ ,
δ ξ
where η is positive: so that

(2.374) S ′′ = o(eξ ).

In S ′ , we write ν = ξ + µ, so that Ψ ξ < µ ≤ δξ. Then

ξν ξ ξ+µ
(2.381) =
ν! (ξ + µ)!
K
< √ exp{(ξ + µ) log ξ − (ξ + µ) log(ξ + µ) + ξ + µ}
ξ
µ2
  
K µ
= √ exp (ξ + µ) 1 − + 2 − · · ·
ξ ξ 2ξ
K 2
< √ eξ−(µ /4ξ) .∗
ξ
The normal number of prime factors of a number n 333

Thus
 
δξ
eξ X 2 /4ξ
(2.382) S′ = O  √ e−µ  ∗
ξ √
Ψ ξ
( )
eξ ∞
Z
−t2 /4ξ
=O √ √
e dt
ξ Ψ ξ
 Z ∞ 
−u2
= O eξ e du = o(eξ ),
Ψ

since Ψ → ∞. From (2.36), (2.374), and (2.382) follows the truth of (2.34), and so that
of (2.322). As (2.321) may be proved in the same manner, the proof of Theorem A is
completed.

2.4. Theorem A′ . If φ is a function of n which tends steadily to infinity with n, then


almost all quadratfrei numbers have between
p p
log log n − φ log log n and log log n + φ log log n

prime factors.
This theorem is a simple corollary of Theorem A. Consider the numbers not exceeding x.

We may plainly neglect numbers less than x; so that
1
2 log x ≤ log n ≤ log x,

(2.41) log log x − log 2 ≤ log log n ≤ log log x.

Given φ, we can determine a function Ψ(x) such that Ψ tends steadily to infinity with x
and

(2.42) Ψ(x) < 21 φ( x).

Then

Ψ(x) < 12 φ(n) ( x ≤ n ≤ x).

Since the exponent is equal to
µ2 µ3
 
µ
ξ−µ − + − · · · ,
1·2·ξ 2 · 3 · ξ2 3 · 4 · ξ3
and
µ2 µ3 µ
2
+ + ··· <
2·3·ξ 3 · 4 · ξ3 4ξ
in virtue of (2.35)

In this sum the values of µ are not, in general, integral: ξ + µ is integral.
334 Paper 35

In the first place p p


φ log log n > Ψ log log x
if
1
log log n > 4 log log x,
which is certainly true, in virtue of (2.41), for sufficiently large values of x. Thus
p p
(2.43) log log n − φ (log log n) < log log x − Ψ (log log x).

In the second place the corresponding inequality


p p
(2.44) log log n + φ (log log n) > log log x + Ψ (log log x)

is certainly true if p p
φ log log n > Ψ log log x + log 2;
and this also is true, in virtue of (2.41) and (2.42), for sufficiently large values of x. From
(2.43), (2.44), and Theorem A, Theorem A′ follows at once.
As a corollary we have
Theorem A′′ . The normal order of the number of prime factors of a quadratfrei number
is log log n.

III.
The normal order of f (n).

3.1. So far we have confined our attention to numbers which have no repeated factors.
When we remove this restriction, the functions f (n) and F (n) have to be distinguished
from one another.
We shall denote by ̟ν (x) the number of numbers, not exceeding x, for which

f (n) = ν.

It is obvious that
̟ν (x) ≥ πν (x).
We require an inequality for ̟ν (x) similar to that for πν (x) given by Lemma A.
Lemma B. There are absolute constants D and L such that
Lx (log log x + D)ν
(3.11) ̟ν+1 (x) <
log x ν!

for
ν = 0, 1, 2, . . . , x ≥ 2.
The normal number of prime factors of a number n 335

It is plain that
√ √
 
x
̟1 (x) = π(x) + π( x) + π( 3 x) + · · · = O .
log x
The inequality is therefore true for ν = 0, whatever the value of D. We shall prove by
induction that it is true in general if

(3.12) D > B + H + J,

where B and H have the same values as in 2.2, and



X
(3.13) J= (s + 1)(2−s + 3−s + 5−s + · · ·).
2

Consider the numbers which do not exceed x and are comprised in the table

2a · pa11 · pa22 · · · paνν ,


3a · pa11 · pa22 · · · paνν ,
...............,
P · pa11 · pa22 · · · paνν ,
a

where p1 , p2 , . . . , pν satisfy the same conditions as in the table of 2.2. It is plain that
√ √
P ≤ x if a = 1, P ≤ 3 x if a = 2, and so on, so that the total number of numbers in the
table does not exceed  
X x
̟ν .
pa
p
a+1 ≤x

If now ω1 , ω2 , . . . are primes and


a
ω1 < ω2 < · · · < ων+1 , ω1a1 ω2a2 · · · ων+1
ν+1
≤ x,
a
the number ω1a1 ω2a2 · · · ων+1
ν+1
will occur at least ν times in one of the tables which correspond
to different values of a. We have therefore
  X  
X x x
(3.14) ν̟ν+1 (x) ≤ ̟ν + ̟ν + ···.
2
p 3
p2
p ≤x p ≤x

Now let us suppose that (3.11) is true when ν − 1 is substituted for ν. Then it is plain∗
that
 
Lx(log log x + D) ν−1  X 1 X 1 
(3.15) ̟ν+1 (x) < + + · · · .
ν!  2 p log(x/p) 3
p2 log(x/p2 ) 
p ≤x p ≤x


See the footnote to p. 81 [footnote * on 330].
336 Paper 35

Now
X 1 log log x + B + H
(3.16) < ,
p log(x/p) log x
p2 ≤x

as we have already seen in 2.2. Also, if ps+1 ≤ x, we have

x x log x
≥ x1/(s+1) , log ≥ ;
ps p s s+1

and so
X 1 s + 1 −s
(3.17) ≤ (2 + 3−s + 5−s + · · ·),
ps log(x/ps ) log x
ps+1 ≤x

if s ≥ 2. Hence
X X 1 log log x + B + H + J log log x + D
(3.18) < < .
s
ps log(x/ps ) log x log x
ps+1 ≤x

From (3.15) and (3.18) the truth of (3.11) follows immediately.

3.2. We can now argue with ̟ν (x) as we argued with πν (x) in 2.3 and the paragraphs
which follow; and we may, without further preface, state the following theorems.
Theorem B. If φ is a function x which tends steadily to infinity to x, then
p p
log log x − φ (log log x) < f (n) < log log x + φ (log log x)

for almost all numbers n less than x.


Theorem B′ . If φ is a function of n which tends steadily to infinity with n, then almost
all numbers have between
p p
log log n − φ (log log n) and log log n + φ (log log n)

different prime factors.


Theorem B′′ . The normal order of the number of different prime factors of a number is
log log n.

IV.
The normal order of F (n).

4.1. We have now to consider the corresponding theorems for F (n), the number of prime
factors of n when multiple factors are counted multiply. These theorems are slightly more
The normal number of prime factors of a number n 337

difficult. The additional difficulty occurs, however, only in the first stage of the argument,
which requires the proof of some inequality analogous to those given by Lemmas A and B.
We denote by
Πν (x)
the number of numbers, not exceeding x, for which

F (n) = ν.

It would be natural to expect an inequality of the same form as those of Lemmas A and B,
though naturally with different values of the constants. It is easy to see, however, that no
such inequality can possibly be true.
For, if ν is greater than a constant multiple of log x, the function

x (log log x + C)ν


log x ν!
is — as may be seen at once by a simple approximation based upon Stirling’s theorem —
exceedingly small; and Πν+1 (x), being an integer, cannot be small unless it is zero. Thus
such an inequality as is suggested would shew that F (n) cannot be of order as high as log x
for any n less than x; and this is false, as we can see by taking
log n
x = 2k + 1, n = 2k , F (n) = .
log 2
the inequality required must therefore be of a less simple character.

4.2. Lemma C. Suppose that K and C have the same meaning as in Lemma A, and that
9
(4.21) 10 ≤ λ < 1.

Then
Y Kx
(4.22) (x) <
ν+1
log x

(log log x + C)ν (log log x + C)ν−1 ν−2


 
2 (log log x + C)
× +λ +λ + ··· ,
ν! (ν − 1)! (ν − 2)!

the series being continued to the term λν .


9
It is plainly sufficient to prove this inequality when λ = 10 . In what follows we shall suppose
that λ has this particular value.
We require a preliminary inequality analogous to (2.23) and (3.14). Consider the numbers
which do not exceed x and are comprised in the table

2a · p1 · p2 · · · pν+1−a ,
338 Paper 35

3a · p1 · p2 · · · pν+1−a ,
............,
P a · p1 · p2 · · · pν+1−a ,

where now p1 , p2 , . . . , pν+1−α are primes, arranged in ascending order of magnitude, but
not necessarily different, and where pν+1−α ≥ 2 in the first row, pν+1−α ≥ 3 in the second,
and so forth. Arguing as in 2.2. and 3.1, we now find
  X   X  
X x x x
(4.23) νΠν+1 (x) < Πν + Πν−1 2
+ Πν−2 + ···.
2
p 2
p 2
p3
p ≤x p ≤x p ≤x

This inequality differs formally from (3.14) in that the suffixes of the functions on the
right-hand side sink by one from term to term.
We can now prove (4.22) by a process of induction similar in principle to that of 2.2 and
3.1. Let us suppose that the inequality is true when ν is replaced by ν − 1, and let us write

log log x + C = ξ,

as in 2.3. Substituting in (4.23), and observing that


x
log log + C ≤ ξ,
ps

we obtain
 ν−1
ξ ν−2

Kx X 1 ξ
(4.24) Πν+1 (x) < +λ + ···
ν p log(x/p) (ν − 1)! (ν − 2)!
p2 ≤x
 ν−2
ξ ν−3

Kx X 1 ξ
+ + λ + · · ·
ν p2 log(x/p2 ) (ν − 2)! (ν − 3)!
p3 ≤x
 ν−3
ξ ν−4

Kx X 1 ξ
+ +λ + ··· + ···
ν 4
p3 log(x/p3 ) (ν − 3)! (ν − 4)!
p ≤x

Now, as we have seen already,


X 1 log log x + C
(4.251) < ;
p log(x/p) log x
p2 ≤x

and
X 1 s + 1 −s
(4.252) < (2 + 3−s + 5−s + · · ·)
ps log(x/ps ) log x
ps+1 ≤x
The normal number of prime factors of a number n 339

if s ≥ 2. It is moreover easy to prove that

(4.2531) (s + 1)(2−s + 3−s + 5−s + · · ·) < 2λs = 2( 10


9 s
)

if s = 2, and

(4.2532) (s + 1)(2−s + 3−s + 5−s + · · ·) < λs = ( 10


9 s
)

if s > 2.∗ From (4.24) – (4.2532) it follows that

ξν ξ ν−1 ν−2
 
Kx 2 ξ
(4.25) Πν+1 (x) < +λ +λ + ···
ν log x (ν − 1)! (ν − 2)! (ν − 3)!
 ν−2
ξ ν−3

2Kx 2 ξ
+ λ +λ + ···
ν log x (ν − 2)! (ν − 3)!
 ν−3
ξ ν−4

Kx 3 ξ
+ λ +λ + ···
ν log x (ν − 3)! (ν − 4)!
+ · · · †.

When we collect together the various terms on the right-hand side which involve the same
powers of ξ, it will be found that the coefficient of ξ ν−p is exactly
Kx λp
,
log x (ν − p)!
except when p = 1, when it is
 
1 Kx λ
1− .
ν log x (ν − 1)!

We thus obtain (4.22).

4.3. We may now argue substantially as in 2.3. We have to shew, for example, that
X
(4.31) S2 = Πν+1 (x) = o(x);

ν>llx+φ (llx)

and this is equivalent to proving that


X
(4.32) S= Πν+1 (x) = o(x),

ν>ξ+Ψ ξ

where Ψ is any function of x which tends to infinity with x.



The inequalities may be verified directly for s = 2 and s = 3, and then proved to be true generally by
induction

The factor 2 occurs in the second line only.
340 Paper 35

We choose δ so that
1
(4.33) − 1 < δ < 91 , ∗
9 log(1 + 91 )
and write
X X
(4.34) S= Πν+1 (x) + Πν+1 (x) = S ′ + S ′′ ,

ξ+Ψ ξ<ν≤(1+δ)ξ ν>(1+δ)ξ

say. In S ′ , we have
Kx ξ ν K1 x ξ ν
 
9 ν 9 2 ν(ν − 1)
(4.351) Πν+1 (x) < 1+ 10 ξ + ( 10 ) + ··· < ,
log x ν! ξ2 log x ν!
where
K
(4.352) K1 = 9 ;
1− 10 (1 + δ)
and by means of this inequality we can shew, just as in 2.3, that
(4.353) S ′ = o(x).
In discussing S ′′ we must use (4.22) in a different manner. We have
2
 
Kx ν ν−1 ξ ν−2 ξ Kx ν ξ/λ
(4.361) Πν+1 (x) < λ +λ +λ + ··· < λ e ;
log x 1! 2! log x
and so
 
Kx ξ/λ
′′
X
ν K x ξ/λ (1+δ)ξ x
(4.362) S < e λ < e λ =O ,
log x 1 − λ log x (log x)η
ν>(1+δ)ξ

where
10 1
(4.363) η = 1 − (1/λ) + (1 + δ) log(1/λ) = (1 + δ) log 9 − 9 > 0,
by (4.33). Hence
(4.364) S ′′ = o(x).
From (4.34), (4.353), and (4.364), it follows that S = o(x).

4.4. We have therefore the following theorems.


Theorem C. The result of Theorem B remains true when F (n) is substituted for f (n).
Theorem C′ , The result of Theorem B′ remains true when the word “different” is omitted.
Theorem C′′ . The normal order of the total number of prime factors of a number is
log log n.

1 1

1)
9 log(1+ 9
−1 < 1
1− 18
− 1 < 19 .
The normal number of prime factors of a number n 341

V.
The normal order of d(n).

5.1. It is natural to ask whether similar theorems cannot be proved with regard to some
of the other standard arithmetical functions of n, such as d(n).
If
n = pa11 pa22 · · · paνν ,
we have
d(n) = (1 + a1 )(1 + a2 ) · · · (1 + aν ).
Since
2 ≤ 1 + a ≤ 2a
if a ≥ 1, we obtain at once
2ν ≤ d(n) ≤ 2a1 +a2 +···+aν ,
or

(5.11) 2f (n) ≤ d(n) ≤ 2F (n) .

From (5.11), and Theorems B ′ and C ′ , we obtain at once


Theorem D′ . The inequalities
√ √
(5.12) 2log log n−φ (log log n)
< d(n)2log log n+φ (log log n)
,

where φ is any function of n which tends to infinity with n, are satisfied for almost all
numbers n.
The inequalities (5.12) are of a much less precise type than (1.26): we cannot say that the
normal order of d(n) is 2log log n . We can however say that (to put it roughly) the normal
order of d(n) is about
2log log n = (log n)log 2 = (log n)0.69... .
It should be observed that this order is far removed from log n, the average order f d(n).
The explanation of this apparent paradox is simple. The majority of numbers have about
(log n)log 2 divisors. But those which have an abnormal number may have a very much
larger number indeed: the excess of the maximum order over the normal order is so great
that, when we compute the average order, it is the numbers with an abnormal number
of divisors which dominate the calculation. The maximum order of the number of prime
factors is not large enough to give rise to a similar phenomenon.
Asymptotic formulæ in combinatory analysis∗

Proceedings of the London Mathematical Society, 2, XVII, 1918, 75 — 115

1. Introduction and summary of results


1.1 The present paper is the outcome of an attempt to apply to the principal problems of
the theory of partitions the methods, depending upon the theory analytic functions, which
have proved so fruitful in the theory of the distribution of primes and allied branches of
the analytic theory of numbers.
The most interesting functions of the theory of partitions appear as the coefficients in the
power-series which represents certain elliptic modular functions. Thus p(n), the number of
unrestricted partitions of n, is the coefficient of xn in the expansion of the function

X 1
(1.11) f (x) = 1 + p(n)xn = . †
(1 − x)(1 − x2 )(1 − x3 ) · · ·
1

If we write

(1.12) x = q 2 = e2πiτ ,

where the imaginary part of τ is positive, we see that f (x) is substantially the reciprocal
of the modular function called by Tannery and Molk‡ h(τ ); that, in fact,
∞ 1
1 1 Y x 24
(1.13) h(τ ) = q 12 q0 = q 12 (1 − q 2n ) = .
f (x)
1

The theory of partitions has, from the time of Euler onwards, been developed from an almost
exclusively algebraical point of view. It consists of an assemblage of formal identities – many
of them, it need hardly be said, of an exceedingly ingenious and beautiful character. Of
asymptotic formluæ, one may fairly say, there are none§ . So true is this, in fact, that we

A short abstract of the contents of part of this paper appeared under the title “Une formule asymptotique
pour le nombre des partitions de n, ” in the Comptes Rendus, January 2nd, 1917 [No. 31 of this volume].

P. A. MacMahon, Combinatory Analysis, Vol. II, 1916, p. 1.

J. Tannery and J. Molk, Fonctions elliptiques, Vol. II, 1896, pp, 31 et seq. We shall follow the notation
of this work whenever we have to quote formulæ from the theory of elliptic functions.
§
We should mention one exception to this statement, to which our attention was called by Major MacMa-
hon. The number of partitions of n into parts none of which exceed r is the coefficient pr (n) in the series

X 1
1+ pr (n)xn = .
1
(1 − x)(1 − x2 ) · · · (1 − xr )

This function has been studied in much detail, for various special values of r, by Cayley, Sylvester and
Glaisher: we may refer in particular to J. J. Sylvester, “On a discovery in the theory of partitions,” Quarterly
Journal, Vol. I, 1857, pp. 81 – 85, and “On the partition of numbers,” ibid., pp. 141 – 152 (Sylvester’s
Asymptotic formulæ in combinatory analysis 343

have been unable to discover in the literature of the subject any allusion whatever to the
question of the order of magnitude of p(n).
1.2 The function p(n) may, of course, be expressed in the form of an integral

1 f (x)
Z
(1.21) p(n) = dx,
2πi Γ xn+1

by means of Cauchy’s theorem, the path Γ enclosing the origin and lying entirely inside the
unit circle. The idea which dominates this paper is that of obtaining asymptotic formulæ
for p(n) by a detailed study of the integral (1.21). This idea is an extremely obvious one;
it is the idea which has dominated nine-tenths of modern research in analytic theory of
numbers: and it may seem very strange that it should never have been applied to this
particular problem before. Of this there are no doubt two explanations. The first is that
the theory of partitions has received its most important developments, since its foundation
by Euler, at the hands of a series of mathematicians whose interests have lain primarily in
algebra. The second and more fundamental reason is to be found in the extreme complexity
of the behavior of the generating function f (x) near a point of the unit circle.
It is instructive to contrast this problem with the corresponding problems which arise
for the arithmetical functions π(n), ϑ(n), Ψ(n), µ(n), d(n), . . . which have their genesis in
Riemann’s Zeta-function and the functions allied to it. In the latter problems we are
dealing with functions defined by Dirichlet’s series. The study of such functions presents
difficulties far more fundamental than any which confront us in the theory of the modular
functions. These difficulties, however, relate to the distribution of the zeros of the functions
and their general behavior at infinity: no difficulties whatever are occasioned by the crude
singularities of the functions in the finite part of the plane. The single finite singularity
of ζ(s), for example, the pole at s = 1, is a singularity of the simplest possible character.
Works, Vol. II, pp. 86 – 89 and 90 – 99); J. W. L. Glaisher, “On the number of partitions of a number into a
given number of parts”, Quarterly Journal, Vol. XL, 1909, pp. 57 – 143; “Formulæ for partitions into given
elements, derived from Sylvester’s Theorem”, ibid., pp. 275 – 348; “Formulæ for the number of partitions
of a number into the elements 1, 2, 3, . . . , n upto n = 9”, ibid., Vol. XLI, 1910, pp. 94 – 112; and further
references will be found in MacMahon, loc. cit., pp. 59 – 71, and E. Netto, Lehrbuch der Combinatorik,
1901, pp. 146 – 158. Thus, for example, the coefficient of xn in
1
(1 − x)(1 − x2 )(1 − x3 )
is
2nπ
p3 (n) = 1
12
(n + 3)2 − 7
72
+ 18 (−1)n + 2
9
cos ;
3
as is easily found by separating the function into partial fractions. This function may also be expressed in
the forms
1
12
(n + 3)2 + ( 12 cos 21 πn)2 − ( 32 sin 31 πn)2 ,
(n + 3)2 },
1  1
1 + 12 n(n + 6) , { 12
where [n] and {n} denote the greatest integer contained in n and the integer nearest to n. These formulæ
do, of course, furnish incidentally asymptotic formulæ for the functions in question. But they are, from this
point of view, of a very trivial character: the interest which they possess is algebraical.
344 Paper 36

It is this pole which gives rise to the dominant terms in the asymptotic formulæ for the
arithmetical functions associated with ζ(s). To prove such a formula rigorously is often
exceedingly difficult; to determine precisely the order of the error which it involves is in
many cases a problem which still defies the utmost resources of analysis. But to write
down the dominant terms involves, as a rule, no difficulty more formidable than that of
deforming a path of integration over a pole of the subject of integration and calculating the
corresponding residue.
In the theory of partitions, on the other hand, we are dealing with functions which do not
exist at all outside the unit circle. Every point of the circle is an essential singularity of the
function, and no part of the contour of integration can be deformed in such a manner as to
make its contribution obviously negligible. Every element of the contour requires special
study; and there is no obvious method of writing down a “dominant term”.
The difficulties of the problem appear then, at first sight, to be very serious. We possess,
however, in the formulæ of the theory of linear transformation of the elliptic functions, an
extremely powerful analytical weapon by means of which we can study the behavior of f (x)
near any assigned point of the unit circle∗ . It is to an appropriate use of these formulæ that
the accuracy of our final results, an accuracy which will, we think, be found to be quite
startling, is due.
1.3 It is very important, in dealing with such a problem as this, to distinguish clearly the
various stages to which we can progress by arguments of progressively “deeper” and less
elementary character. The earlier results are naturally (so far as the particular problem
is concerned) superseded by the later. But the more elementary methods are likely to be
applicable to other problems in which the more subtle analysis is impracticable.
We have attacked this particular problem by a considerable number of different methods,
and cannot profess to have reached any very precise conclusions as to the possibilities of
each. A detailed comparison of the results to which they lead would moreover expand this
paper to a quite unreasonable length. But we have thought it worth while to include a
short account of two of them. The first is quite elementary; it depends only on Euler’s
identity

1 x x4
(1.31) = 1 + + + ···
(1 − x)(1 − x2 )(1 − x3 ) · · · (1 − x)2 (1 − x)2 (1 − x2 )2

– an identity capable of wide generalisation – and on elementary algebraical reasoning. By


these means we shew, in section 2, that
√ √
(1.32) eA n
< p(n) < eB n
,

where A and B are positive constants, for all sufficiently large values of n.

See G. H. Hardy and J. E. Littlewood, “Some problems of Diophantine approximation (II: The trigono-
metrical series associated with the elliptic Theta-functions),” Acta Mathematica, Vol. XXXVII, 1914, pp.
193 – 238, for applications of the formulæ to different but not unrelated problems.
Asymptotic formulæ in combinatory analysis 345

It follows that
√ √
(1.33) A n < log p(n) < B n;

and the next question which arises is the question whether a constant C exists such that

(1.34) log p(n) ∼ C n.

We prove that this is so in section 3. Our proof is still, in a sense, “elementary.” It does not
appeal to the theory of analytic functions, depending only on a general arithmetic theorem
concerning infinite series; but this theorem is of the difficult and delicate type which Messrs
Hardy and Littlewood have called “Tauberian.” The actual theorem required was proved
by us in a paper recently printed in these Proceedings∗ . It shews that


(1.35) C=√ ;
6

in other words that


( s  )
2n
(1.36) p(n) = exp π (1 + ǫ) ,
3

where ǫ is small when n is large. This method is one of very wide application. It may
be used, for example, to prove that, if p(s) (n) denotes the number of partitions of n into
perfect s-th powers, then
    s/(s+1)
(s) 1 1 1
log p (n) ∼ (s + 1) Γ 1+ ζ 1+ n1/(s+1) .
s s s

It is certainly possible to obtain, by means of arguments of this general character, informa-


tion about p(n) more precise than that furnished by the formula (1.36). And it is equally
possible to prove (1.36) by reasoning of a more elementary, though more special, character:
we have a proof, for example, based on the identity
n
X
np(n) = σ(ν)p(n − ν),
ν=1

where σ(ν) is the sum of divisors of ν, and a process of induction. But we are at present
unable to obtain, by any method which does not depend upon Cauchy’s theorem, a result
as precise as that which we state in the next paragraph, a result, that is to say, which is
“vraiment asymptotique.”

G. H. Hardy and S. Ramanujan, “Asymptotic formulæ for the distribution of integers of various types,”
Proc. London Math. Soc., Ser. 2, Vol. XVI, 1917, pp. 112 – 132 [No. 34 of this volume].
346 Paper 36

1.4 Our next step was to replace (1.36) by the much more precise formula
( s )
1 2n
(1.41) p(n) ∼ √ exp π .
4n 3 3

The proof of this formula appears to necessitate the use of much more powerful machinery,
Cauchy’s integral (1.21) and the functional relation
s
x1/24 π2
  
1
(1.42) f (x) = p log exp f (x′ ),
(2π) x 6log (1/x)

where
4π 2
 

(1.43) x = exp − .
log (1/x)
This formula is a merely a statement in different notation of the relation between h(τ ) and
h(T ) where
c + dτ
T = , a = d = 0, b = 1, c = −1;
a + bτ
viz. s 
i
h(τ ) = h(T ).∗
τ

It is interesting to observe the correspondence between (1.41) and the results of numerical
computation. Numerical data furnished to us by Major MacMahon gave the following
results: we denote the right-had side of (1.41) by ̟(n).

n p(n) ̟(n) ̟/p


10 42 48.104 1.145
20 627 692.385 1.104
50 204226 217590.499 1.065
80 15796476 16606781.567 1.051

It will be observed that the progress of ̟/p towards its limit unity is not very rapid, and
that ̟ − p is always positive and appears to tend rapidly to infinity.
1.5 In order to obtain more satisfactory results it is necessary to construct some auxiliary
function F (x) which is regular at all points of the unit circle save x = 1, and has there a
singularity of a type as near as possible to that of the singularity of f (x). We may then
hope to obtain a much more precise approximation by applying Cauchy’s theorem to f − F
instead of to f . For although every point of the circle is a singular point of f , x = 1 is, to

Tannery and Molk, loc. cit., p. 265 (Table XLV, 5).
Asymptotic formulæ in combinatory analysis 347

put it roughly, much the heaviest singularity. When x −→ 1 by real values, f (x) tends to
infinity like an exponential
π2
 
exp ;
6(1 − x)
when
x = re2pπi/q ,
p and q being co-prime integers, and r −→ 1, |f (x)| tends to infinity like an exponential

π2
 
exp ;
6q 2 (1 − r)

while, if
x = re2θπi ,
where θ is irrational, |f (x)| can become infinite at most like an exponential of the type
  
1
exp o .∗
1−r

The function required is



1 X
(1.51) F (x) = √ Ψ(n)xn ,
π 2 1
where  
d cosh Cλn − 1
(1.52) Ψ(n) = ,
dnq λn q

(1.53) C = 2π/ 6 = π 23 , λn = (n − 24 1
).
This function may be transformed into an integral by means of a general formula given by
Lindelöf† ; and it is then easy to prove that the “principal branch” of F (x) is regular all
over the plane except at x = 1‡ ; and that

F (x) − χ(x),

where

The statements concerning the “rational” points are corollaries of the formulæ of the transformation
theory, and proofs of them are contained in the body of the paper. The proposition concerning “irrational”
points may be proved by arguments similar to those used by Hardy and Littlewood in their memoir already
quoted. It is not needed for our present purpose. As pa matter of fact it is generally true that f (x) −→ 0
when θ is irrational, and very nearly as rapidly as 4 (1 − r). It is in reality owing to this that our final
method is so successful.

E. Lindelöf, Le calcul de résidus et ses applications à la théorie des fonctions, (Gauthier-Villars, Col-
lection Borel, 1905), p. 111.

We speak, of course, of the principal branch of the function, viz., that represented by the series (1.51)
when x is small. The other branches are singular at the origin.
348 Paper 36

s
x1/24 π2
    
1
(1.54) χ(x) = p log exp −1
(2π) x 6log (1/x)
is regular for x = 1. If we compare (1.42) and (1.54), and observe that f (x′ ) tends to unity
with extreme rapidity when x tends to 1 along any regular path which does not touch the
circle of convergence, we can see at once the very close similarity between the behaviour of
f and F inside the unit circle and in the neighbourhood of x = 1.
It should be observed that the term −1 in (1.52) and (1.54) is-so far as our present assertions
are concerned-otiose: all that we have said remains true if it is omitted; the resemblance
between the singularities of f and F becomes indeed even closer. The term is inserted
merely in order to facilitate some of our preliminary analysis, and will prove to be without
influence on the final result.
Applying Cauchy’s theorem to f − F , we obtain
1 d eCλn √
 
(1.55) p(n) = √ + O(eD n ),
2π 2 dn λn
where D is any number greater than
q
1 1 2
2 C = 2 π ( 3 ).

1.6 The formula (1.55) is an asymptotic formula of a type far more precise than that of
(1.41). The error term is, however, of an exponential type, and may be expected ultimately
to increase with very great rapidity. It was therefore with considerable surprise that we
found what exceedingly good results the formula gives for fairly large values of n. For
n = 61, 62, 63 it gives

1121538.972, 1300121.359, 1505535.606,

while the correct values are

1121505, 1300156, 1505499.

The errors
33.972, −34.641, 36.606
are relatively very small, and alternate in sign.
The next step is naturally to direct our attention to the singular point of f (x) next in
importance after that at x = 1, viz., that at x = −1; and to subtract from f (x) a second
auxiliary function, related to this point as F (x) is to x = 1. No new difficulty of principle
is involved, and we find that  1 
 Cλn  n Cλn √
1 d e (−1) d  e 2  + O(eD n ),
(1.61) p(n) = √ +
2π 2 dn λn 2π dn λn
where D is now any number greater than 31 C. It now becomes obvious why our earlier
approximation gave errors alternately of excess and of defect.
Asymptotic formulæ in combinatory analysis 349

It is obvious that this process may be repeated indefinitely. The singularities next in
2 4
importance are those at x = e 3 πi and x = e 3 πi ; the next those at x = i and x = −i; and
so on. The next two terms in the approximate formula are found to be
√ 1
!
3 2 1
 d e 3 Cλn
√ cos 3 nπ − 18 π dn
π 2 λn

and
√ 1
!
2 1 1
 d e 4 Cλn
cos 2 nπ − 8 π dn .
π λn

As we proceed further, the complexity of the calculations increases. The auxiliary function
associated with the point x = e2pπi/q involves a certain 24q-th root of unity, connected with
the linear transformation which must be used in order to elucidate the behaviour of f (x)
near the point; and the explicit expression of this root in terms of p and q, though known,
is somewhat complex. But it is plain that, by taking a sufficient number of terms, we can
find a formula in which the error is

O(eCλn /ν ),

where ν is a fixed but arbitrarily large integer.


1.7 A final question remains. We have still the resource of making ν a function of n, that
is to say of making the number of terms in our approximate formula itself a function of n.
In this way we may reasonably hope, at any rate, to find a formula in which the error is
of order less than that of any exponential of the type ean ; of the order of a power of n, for
example, or even bounded.
When, however, we proceeded to test this hypothesis by means of the numerical data most
kindly provided for us by Major MacMahon, we found a correspondence between the real
and the approximate values of such astonishing accuracy as to lead us to hope for even
more. Taking n = 100, we found that the first six terms of our formula gave

190568944.783
+348.872
−2.598
+.685
+.318
−.064

190569291.996,

while p(100) = 190569292; so that the error after six terms is only .004. We then proceeded
to calculate p(200) and found
350 Paper 36

3, 972, 998, 993, 185.896


+36, 282.978
−87.555
+5.147
+1.424
+0.071
+0.000 ∗
+0.043

3, 972, 999, 029, 388.004,


and Major MacMahon’s subsequent calculations shewed that p(200) is, in fact,

3, 972, 999, 029, 388.

These results suggest very forcibly that it is possible to obtain a formula for p(n), which
not only exhibits its order of magnitude and structure, but may be used to calculate its
exact value for any value of n. That this is in fact so is shewn by the following theorem.
Statement of the main theorem.
THEOREM. Suppose that
√ !
q d eCλn /q
(1.71) φq (n) = √ ,
2π 2 dn λn
where C and λn are defined by the equations (1.53), for all positive integral values of q; that
p is a positive integer less than and prime to q; that ωp,q is a 24q-th root of unity, defined
when p is odd by theformula
 h n   o i
(1.721) ωp,q = −q p exp − 1
4 (2 − pq − p) + 1
12 q − 1
q (2p − p ′ + p2 p′ ) πi ,

and when q is odd by the   formulah n   o i


(1.722) ωp,q = −p q exp − 1
4 (q − 1) + 1
12 q − 1
q (2p − p ′ + p2 p′ ) πi ,

where ab is the symbol of Legendre and Jacobi† , and p′ is any positive integer such that


1 + pp′ is divisible by q; that


Aq (n) = (p) ωp,q e−2npπi/q ;
P
(1.73)

and that α is any positive constant, and ν the integral part of α n.
Then
1
p(n) = ν1 Aq φq + O(n− 4 ),
P
(1.74)
so that p(n) is, for all sufficiently large values of n, the integer nearest to

(1.75) 1 Aq φq .

This term vanishes identically.

See Tannery and Molk, loc. cit., pp. 104 – 106, for a complete set of rules for the calculation of the
value of ab , which is, of course, always 1 or −1. When both p and q are odd it is indifferent which formula


is adopted.
Asymptotic formulæ in combinatory analysis 351

It should be observed that all the numbers Aq are real. A table of Aq from q = 1 to q = 18
is given at the end of the paper (Table II).
The proof of this main theorem is given in section 5; section 4 being devoted to a number
of preliminary lemmas. The proof is naturally somewhat intricate; and we trust that we
have arranged it in such a form as to be readily intelligible. In section 6 we draw attention
to one or two questions which our theorem, in spite of its apparent completeness, still
leaves open. In section 7 we indicate some other problems in combinatory analysis and
the analytic theory of numbers to which our method may be applied; and we conclude
by giving some functional and numerical tables: for the latter we are indebted to Major
MacMahon and Mr. H. B. C. Darling. To Major MacMahon in particular we owe many
thanks for the amount of trouble he has taken over very tedious calculations. It is certain
that, without the encouragement given by the results of these calculations, we should never
have attempted to prove theoretical results at all comparable in precision with those which
we have enunciated.

√ √
2. Elementary proof that eA n < p(n) < eB n for sufficiently large values of n.

2.1 In this section we give the elementary proof of the inequalities (1.32). We prove, in
fact, rather more, viz., that positive constants H and K exist such that
H 2√n K √
(2.11) e < p(n) < e2 2n
n n
for n ≥ 1∗ . We shall use in our proof only Euler’s formula (1.31) and a debased form of
Stirling’s theorem, easily demonstrable by quite elementary methods: the proposition that
1
n!en /nn+ 2

lies between two positive constants for all positive integral values of n.
2.2 The proof of the first of the two inequalities is slightly the simpler. It is obvious that
if X 1
pr (n)xn =
(1 − x)(1 − x2 ) · · · (1 − xr )
so that pr (n) is the number of partitions of n into parts not exceeding r, then
(2.21) pr (n) = pr−1 (n) + pr−1 (n − r) + pr−1 (n − 2r) + · · ·
We shall use this equation to prove, by induction, that
rnr−1
(2.22) pr (n) ≥ .
(r!)2

Somewhat inferior inequalities, of the type
√ √
2A[ n]
< p(n) < nB[ n]
,

may be proved by entirely elementary reasoning; by reasoning, that is to say, which depends only on the
arithmetical definition of p(n) and on elementary finite algebra, and does not presuppose the notion of a
limit or the definition of the logarithmic or exponential functions.
352 Paper 36

It is obvious that (2.22) is true for r = 1. Assuming it to be true for r = s, and using
(2.21), we obtain
s
ps+1 (n) ≥ {ns−1 + (n − s − 1)s−1 + (n − 2s − 2)s−1 + · · ·}
(s!)2
 s
n − (n − s − 1)s (n − s − 1)s − (n − 2s − 2)s

s
≥ + + ···
(s!)2 s(s + 1) s(s + 1)
ns (s + 1)ns
= = .
(s + 1)(s!)2 {(s + 1)!}2
This proves (2.22). Now p(n) is obviously not less than pr (n), whatever the value of r.

Take r = [ n]: then
√ √

[ n] n[ n] H 2√n
p(n) ≥ p[ n] (n) ≥ √ > e ,
n {[ n]!}2 n
by a simple application of the degenerate form of Stirling’s theorem mentioned above.
2.3 The proof of the second inequality depends upon Euler’s identity. If we write
X 1
qr (n)xn = ,
(1 − x) (1 − x2 )2 · · · (1 − xr )2
2

we have
(2.31) qr (n) = qr−1 (n) + 2qr−1 (n − r) + 3qr−1 (n − 2r) + · · ·,
and
(2.32) p(n) = q1 (n − 1) + q2 (n − 4) + q3 (n − 9) + · · ·
We shall first prove by induction that
(n + r 2 )2r−1
(2.33) qr (n) ≤ .
(2r − 1)!(r!)2
This is obviously true for r = 1. Assuming it to be true for r = s, and using (2.31), we
obtain
1
qs+1 (n) ≤ {(n + s2 )2s−1 + 2(n + s2 − s − 1)2s−1
(2s − 1)!(s!)2
+ 3(n + s2 − 2s − 2)2s−1 + · · ·}.

Now
m(m − 1)am−2 b2 ≤ (a + b)m − 2am + (a − b)m ,
if m is a positive integer, and a, b, and a − b are positive, while if a − b ≤ 0, and m is odd,
the term (a − b)m may be omitted. In this inequality write

m = 2s + 1, a = n + s2 − ks − k (k = 0, 1, 2, . . .), b = s + 1,

and sum with respect to k. We find that

(2s + 1)2s(s + 1)2 {(n + s2 )2s−1 + 2(n + s2 − s − 1)2s−1 + · · ·} ≤ (n + s2 + s + 1)2s+1 ;


Asymptotic formulæ in combinatory analysis 353

and so
(n + s2 + s + 1)2s+1 {n + (s + 1)2 }2s+1
qs+1 (n) ≤ ≤ .
(2s + 1)2s(s + 1)2 (2s − 1)!(s!)2 (2s + 1)!{(s + 1)!}2
Hence (2.33) is true generally.
It follows from (2.32) that

X n2r−1
p(n) = q1 (n − 1) + q2 (n − 4) + · · · ≤ .
(2r − 1)!(r!)2
1

But, using the degenerate form of Stirling’s theorem once more, we find without difficulty
that
1 26r K
< ,
(2r − 1)!(r!)2 4r!
where K is a constant. Hence
∞ ∞ 1
X (8n)2r−1 X (8n) 2 r−1 K 2√2n
p(n) < 8K < 8K < e .
4r! r! n
1 1

This is the second of the inequalities (2.11).

3. Application of a Tauberian theorem to the determination of the constant


C.

3.1 The value of the constant


log p(n)
C = lim √ ,
n
is most naturally determined by the use of the following theorem.
n
P
If g(x) = an x is a power-series with positive coefficients, and

A
log g(x) ∼
1−x
when x → 1, then √
log sn = log (a0 + a1 + · · · + an ) ∼ 2 An
when n → ∞.
This theorem is a special case∗ of Theorem C in our paper already referred to.
Now suppose that
X 1
g(x) = (1 − x)f (x) = {p(n) − p(n − 1)}xn = .
(1 − x2 )(1 − x3 )(1 − x4 ) · · ·

Loc. cit., p. 129 (with α = 1) [p. 321 of this volume].
354 Paper 36

Then
an = p(n) − p(n − 1)
is plainly positive. And
∞ ∞
X 1 x2ν ∞
X 1 1 X 1 π2
(3.11) log g(x) = log µ
= ν
∼ 2
= ,
2
1−x 1
ν 1 − x 1 − x 1
ν 6(1 − x)
when x → 1∗ . Hence √
(3.12) log p(n) = a0 + a1 + · · · + an ∼ C n,
√ q
where C = 2π/ 6 = π 23 , as in (1.53).

3.2 There is no doubt that it is possible, by “Tauberian” arguments, to prove a good


deal more about p(n) than is asserted by (3.12). The functional equation satisfied by f (x)
shews, for example, that

(1 − x)3/2 π2
 
g(x) ∼ √ exp ,
2π 6(1 − x)

a relation far more precise than (3.11). From this relation, and the fact that the coefficients
in g(x) are positive, it is certainly possible to deduce more than (3.12). But it hardly
seems likely that arguments of this character will lead us to a proof of (1.41). It would
be exceedingly interesting to know exactly how far they will carry us, since the method
is comparatively elementary, and has a much wider range of application than the more
powerful methods employed later in this paper. We must, however, reserve the discussion
of this question for some future occasion.

4. Lemmas preliminary to the proof of the main theorem.

4.1 We proceed now to the proof of our main theorem. The proof is somewhat intricate,
and depends on a number of subsidiary theorems which we shall state as lemmas.


This is a special case of a much more general theorems: see K. Knopp, “Grenzwerte von Reihen bei
der Annäherung an die Konvergenzgrenze,” Inaugural Dissertation, Berlin, 1907, pp. 25 et seq.; K. Knopp,
“Über Lambertsche Reihen,” Journal für Math., Vol. CXLII, 1913, pp. 283 – 315; G. H. Hardy, “Theorems
connected with Abel’s Theorem on the continuity of power series,” Proc. London Math. Soc., Ser. 2, Vol.
IV, 1906, pp. 247 – 265 (pp. 252, 253); G. H. Hardy, “Some theorems concerning infinite series,” Math.
Ann., Vol. LXIV, 1907, pp. 77 – 94; G. H. Hardy, “Note on Lambert’s series,” Proc. London Math. Soc.,
Ser. 2, Vol. XIII, 1913, pp. 192 – 198.
A direct proof is very easy: for

νxν−1 (1 − x) < 1 − xν < ν(1 − x),

1 X x2ν 1 X xν+1
2
< log g(x) < .
1−x ν 1−x ν2
Asymptotic formulæ in combinatory analysis 355

Lemmas concerning Farey’s series.


4.21 The Farey’s series of order m is the aggregate of irreducible rational fractions

p/q (0 ≤ p ≤ q ≤ m),

arranged in ascending order of magnitude. Thus


0 1 1 1 1 2 1 2 3 1 4 3 2 5 3 4 5 6 1
, , , , , , , , , , , , , , , , , ,
1 7 6 5 4 7 3 5 7 2 7 5 3 7 4 5 6 7 1
is the Farey’s series of order 7.
Lemma 4.21. If p/q, p′ /q ′ are two successive terms of a Farey’s series, then
(4.211) p′ q − pq ′ = 1.
This is, of course, a well-known theorem, first observed by Farey and first proved by
Cauchy∗ . Th following exceedingly simple proof is due to Hurwitz† .
The result is plainly true when m = 1. Let us suppose it true for m = k; and let p/q, p′ /q ′
be two consecutive terms in the series of order k.
Suppose now that p′′ /q ′′ is a term of the series of order k + 1 which falls between p/q, p′ /q ′ .
Let
p′′ q − pq ′′ = λ > 0, p′ q ′′ − p′′ q ′ = µ > 0.
Solving these equations for p′′ , q ′′ and observing that p′ q − pq ′ = 1, we obtain

p′′ = µp + λp′ , q ′′ = µq + λq ′ .

Consider now the aggregate of fractions

(µp + λp′ )/(µq + λq ′ ),

where λ and µ are positive integers without common factor. All of these fractions lie between
p/q and p′ /q ′ ; and all are in their lowest terms, since a factor common to numerator and
denominator would divide

λ = q(µp + λp′ ) − p(µq + λq ′ ),

and µ = p′ (µq + λq ′ ) − q ′ (µp + λp′ ).


Each of them first makes its appearance in the Farey’s series of order µq + λq ′ , and the first
of them to make its appearance must be that for which λ = 1, µ = 1. Hence

p′′ = p + p′ , q ′′ = q + q ′ ,

J. Farey, “On a curious property of vulgar fraction,” Phil. Mag., Ser. 1, Vol. XLVII, 1816, pp. 385,
386; A. L. Cauchy, “Démonstration d’un théorème curieux sur les nombres,” Exercices de mathématiques,
Vol. I, 1826, pp. 114 – 116. Cauchy’s proof was first published in the Bulletin de la Société Philomatique
in 1816.

A. Hurwitz, “Ueber die angenäherte Darstellung der Zahlen durch rationale Brüche,” Math. Ann., Vol.
XLIV, 1894, pp 417-436.
356 Paper 36

p′′ q − pq ′′ = p′ q ′′ − p′′ q ′ = 1.
The lemma is consequently proved by induction.
Lemma 4.22. Suppose that p/q is a term of the Farey’s series of order m, and p′′ /q ′′ , p′ /q ′
the adjacent terms on the left and right: and let jp,q denote the interval
p 1 p 1
− , + .∗
q q(q + q ′′ ) q q(q + q ′ )
Then (i) the intervals jp,q exactly fill up the continuum (0, 1), and (ii) the length of each
of the parts into which jp,q is divided by p/q † is greater than 1/2mq and less than 1/mq.
(i) Since
1 1 1 p′ q − pq ′ p′ p
+ = = = − ,
q(q + q ′ ) q ′ (q ′ + q) qq ′ qq ′ q′ q
the intervals just fill up the continuum.
(ii) Since neither q nor q ′ exceeds m, and one at least must be less than m, we have
1 1
> .
q(q + q ′ ) 2mq
Also q + q ′ > m, since otherwise (p + p′ )/(q + q ′ ) would be a term in the series between p/q
and p′ /q ′ . Hence
1 1
< .
q(q + q ′ ) mq

Standard dissection of a circle.


4.22 The following mode of dissection of a circle, based upon Lemma 4.22, is of fundamental
importance for our analysis.
Suppose that the circle is defined by

x = Re2πiθ (0 ≤ θ ≤ 1).

Construct the Farey’s series of order m, and the corresponding intervals jp,q . When these
intervals are considered as intervals of variation of θ, and the two extreme intervals, which
correspond to abutting arcs on the circle, are regarded as constituting a single interval ξ1,1 ,
the circle is divided into a number of arcs

ξp,q ,

where q ranges from 1 to m and p through the numbers not exceeding and prime to q ‡ . We
call this dissection of the circle the dissection Ξm .


When p/q is 0/1 or 1/1, only the part of this interval inside (0, 1) is to be taken; thus j0,1 is 0, 1/(m + 1)
and j1,1 is 1 − 1/(m + 1), 1.

See the preceding footnote

p = 0 occurring with q = 1 only.
Asymptotic formulæ in combinatory analysis 357

Lemmas from the theory of the linear transformation of the


elliptic modular functions.

4.3 Lemma 4.31. Suppose that q is a positive integer; that p is a positive integer not
exceeding and prime to q; that p′ is a positive integer such that 1 + pp′ is divisible by q; that
ωp,q is defined by the formulæ (1.721) or (1.722) ; that

2π 2p′ πi
   
2πz 2pπi ′
x = exp − + , x = exp − + ,
q q qz q

where the real part of z is positive; and that


1
f (x) = .
(1 − x)(1 − x2 )(1 − x3 ) · · ·

Then

 
π πz
f (x) = ωp,q z exp − f (x′ ).
12qz 12q

This lemma is merely a restatement in a different notation of well-known formulæ in the


transformation theory.
Suppose, for example, that p is odd. If we take

1 + pp′
a = p, b = −q, c = , d = −p′ ,
q

so that ad − bc = 1; and write

x = q 2 = e2πiτ , x′ = Q2 = e2πiT ,

so that
p iz p′ i
τ= + , T = + ;
q q q qz
then we can easily verify that
c + dτ
T = .
a + bτ
Also, in the notation of Tanner and Molk, we have
1 1
q 12 Q 12
f (x) = , f (x′ ) = ;
h(τ ) h(T )

and the formula for the linear transformation of h(τ ) is


 
b
[a(b − c) + bd(a2 − 1)]}πi
p
exp { 14 (a − 1) − 12
1

h(T ) = (a + bτ ) h(τ ),
a
358 Paper 36

where (a + bτ ) has its real part positive∗ . A little elementary algebra will shew the
p

equivalence of this result and ours.


The other formula for ωp,q may be verified similarly, but in this case we must take

1 + pp′
a = −p, b = q, c = , d = p′ .
q
We have included in the Appendix (Table I) a short table of some values of ωp,q , or rather
of (log ωp,q )/πi.
Lemma 4.32. The function sf (x) satisfies the equation
π2
  1  
q 1
(4.321) f (x) = ωp,q log 24
xp,q exp f (x′p,q ),
2π xp,q 6q 2 log(1/xp,q )
where
4π 2 2p′ πi
 
−2pπi/q ′
(4.322) xp,q = xe , xp,q = exp − 2 + .
q log(1/xp,q ) q
This is an immediate corollary from Lemma 4.31, since
 
q 1 1
z= log , e−πz/12q = xp,q 24
,
2π xp,q

π2 2π 2p′ πi
 
π ′
= 2 , x = exp − + = x′p,q .
12qz 6q log(1/xp,q ) qz q
If we observe that
f (x′p,q ) = 1 + p(1)x′p,q + · · · ,
we see that, if x tends to e2pπi/q along a radius vector, or indeed any regular path which
does not touch the circle of convergence, the difference
s
π2
  1  
q 1 24
f (x) − ωp,q log xp,q exp
2π xp,q 6q 2 log(1/xp,q )

tends to zero with great rapidity. It is on this fact that our analysis is based.

Lemmas concerning the auxiliary function Fa (x).

4.41. The auxiliary function Fa (x) is defined by the equation



X
Fa (x) = Ψa (n)xn ,
1

where
d cosh aλn − 1
Ψa (n) = ,
dn λn

Tannery and Molk, loc. cit., pp. 113, 267.
Asymptotic formulæ in combinatory analysis 359

q
1
λn = (n − 24 ), a > 0.

Lemma 4.41. Suppose that a cut is made along the segment (1, ∞) in the plane f x.
Then Fa (x) is regular at all points inside the region thus defined.
This lemma is an immediate corollary of a general theorem proved by Lindelöf on pp. 109
et seq. of his Calcul des residus∗ .
The function q
1
d cosh a (z − 24 ) − 1
Ψa (z) = q
dz (z − 1 ) 24

satisfies the conditions imposed upon it by Lindelöf, if the number which he calls α is
1
greater than 24 ; and
Z a+i∞
xz
(4.411) Fa (x) 2πiz
φ(z) dz,
a−i∞ 1 − e
if
x = reiθ , 0 < θ < 2π, xz = exp{z(log r + iθ)}.

4.42. Lemma 4.42. Suppose that D is the region defined by the inequalities

−π < −θ0 < θ < θ0 < π, r0 < r, 0 < r0 < 1,

and that log(1/x) has its principal value, so that log(1/x) is one-valued, and its square root
two-valued, in D. Further, let

a2
   
p 1
χa (x) = {π log(1/x)}x 24 exp −1 ,
4 log(1/x)

that value of the square root being chosen which is positive when 0 < x < 1. Then

Fa (x) − χa (x)

is regular inside D † .
We observe first that, whenθ has fixed value between 0 and 2π, the integral on the right-
1 1
hand side of (4.411) is uniformly convergent for 24 ≤ α ≤ α0 . Hence we may take α = 24
in (4.411). We thus obtain
∞ ∞
xit x−it
Z Z
1 1
1 1
Fa (x) = ix 24
1 Ψa ( 24 + it) dt + ix 24
1 Ψa ( 24 − it) dt,
πi−2πt
0 1−e 12 0 1 − e 12 πi+2πt

Lindelöf gives references to Mellin and Le Roy, who had previously established the theorem in less
general forms.

Both Fa (x) and χa (x) are two-valued in D. The value of Fa (x) contemplated is naturally that repre-
sented by the power series.
360 Paper 36

√ √ 1 1
where the it and −it which occur in Ψa ( 24 + it) and Ψa ( 24 − it) are to be interpreted
(1/4)πi
√ −(1/4)πi

as e t and e t respectively. We write this in the form
Z ∞
1 xit 1
(4.421) Fa (x) = Xa (x) + ix 24
1 Ψa ( 24 + it) dt
0 e− 12 πi+2πt − 1
Z ∞
1 x−it 1
+ix 24 1 Ψa ( 24 − it) dt
πi+2πt
0 1 − e 12
= Xa (x) + Θ1 (x) + Θ2 (x),

say, where
1
Z ∞
Xa (x) = ix 24 xit Ψa ( 24
1
+ it) dt.
0
Now, since
|xit | = e−θt , |x−it | = eθt ,
the functions Θ are regular throughout the angle of Lemma 4.42. And
1
∞  √ 
x 24 d cosh µ t − 1
Z
−λt
Xa (x) = √ e √ dt,
i 0 dt t
where
1 √
λ = i log , µ = a i.
x
The form of this integral may be calculated by supposing λ and µ positive, when we obtain
Z ∞   Z ∞
−λw 2 d cosh µw − 1 2
e dw = 2λ e−λw (cosh µw − 1) dw
0 dw w 0
2
(λπ)(eµ /4λ − 1).
p
=

Hence
a2
   
p 1
(4.422) Xa (x) = {π log(1/x)}x 24 exp − 1 = χa (x),
4 log(1/x)
and the proof of the lemma is completed.
Lemmas 4.41 and 4.42 shew that x = 1 is the sole finite singularity of the principal branch
of Fa (x).
4.43 Lemma 4.43 Suppose that P, θ1 , and A are positive constants, θ1 being less than π.
Then
|Fa (x)| < K = K(P, θ1 , A),
for
0 ≤ r ≤ P, θ1 ≤ θ ≤ 2π − θ1 , 0 < a ≤ A.
Asymptotic formulæ in combinatory analysis 361

We use K generally to denote a positive number independent of x and of a. We may employ


the formula (4.411). It is plain that
xz

−θ1 |η|

1 − e2πiz < Ke ,

 q 
 cosh a (z − 1 ) − 1  √
d 24
< KeK |η| ,

|Ψa (z)| = q
dz  (z − 1 ) 
24

where η is the imaginary part of z. Hence


Z ∞ √
|Fa (x)| < K eK |η|−θ1 |η| dη < K.
−∞

4.44 Lemma 4.44. Let c be a circle whose centre is x = 1, and whose radius δ is less than
unity. Then
|Fa (x) − χa (x)| < Ka2 ,
is x lies in c and 0 < a ≤ A, K = K(δ, A) being as before independent of x and of a.
If we refer back to (4.421) and (4.422), we see that it is sufficient to prove that

|Θ1 (x)| < Ka2 , |Θ2 (x)| < Ka2 ;

and we may plainly confine ourselves to the first of these inequalities. We have
1 Z
( )
x 24 ∞
p
xit d cosh a (it) − 1
Θ1 (x) = √ 1 p dt.
i 0 e− 12 πi+2πt − 1 dt (t)

Rejecting the extraneous factor, which is plainly without importance, and integrating by
parts, we obtain
Z ∞ p
cosh a (it) − 1
Θ(x) = Φ(t) p dt,
0 (t)
where 1
ixit log x 2πxit e− 12 πi+2πt
Φ(t) = − 1 + 1 .
e− 12 πi+2πt − 1 (e− 12 πi+2πt − 1)2
1
Now |θ| < 12 π and |xit | < Ke 2 πt . It follows that

|Φ(t)| < Ke−πt ;

and

e−πt
Z
√ | sinh2 21 a (it)| dt
p
|Θ(x)| < K
0 t
362 Paper 36


e−πt
Z q q
< K √ {cosh a ( 2 t) − cos a ( 12 t)} dt
1
0 t
Z ∞  
−πw 2 aw aw
< K e cosh √ − cos √ dw
0 2 2
2 /8π 2 /8π
= K(ea − e−a ) < Ka2 .

5. Proof of the main theorem.


5.1 We write √
q
(5.11) Fp,q (x) = ωp,q √ FC/q (xp,q ),
q π 2
2 −2pπi/q
where C = π 3, xp,q = xe ; and
XX
(5.12) Φ(x) = f (x) − Fp,q (x),
q p
where the summation applies to all values of p not exceeding q and prime to q, and to all
values of q such that √
(5.13) 1 ≤ q ≤ ν = [α n],
α being positive and independent of n. If then X
(5.14) Fp,q (x) = cp,q,n xn ,
we have
1 Φ(x)
XX Z
(5.15) p(n) − cp,q,n = dx,
q p
2πi Γ xn+1
where Γ is a circle whose centre is the origin and whose radius R is less than unity. We
take
β
(5.16) R=1− ,
n
where β also is positive and independent of n.
Our object is to shew that the integral on the right-hand side of (5.15) is of the form
1
O(n− 4 ); the constant implied in the O will of course be a function of α and β. It is to
be understood throughout that O’s are used in this sense; O(1), for instance, stands for
a function of x, n, p, q, α, and β (or some only of these variables) which is less in absolute
value than a number K = K(α, β) independent of x, n, p, and q.
We divide up the circle Γ by means of dissection Ξν of 4.22, into arcs ξp,q each associated
with a point Re2pπi/q ; and we denote by ηp,q the arc of Γ complementary to ξp,q . This being
so, we have
Φ(x) f (x) − Fp,q (x) Fp,q (x)
Z XZ XZ
(5.17) n+1
dx = n+1
dx − n+1
dx
Γ x ξp,q x ηp,q x
X X
= Jp,q − jp,q ,
1
say. We shall prove that each of these sums is of the form O(n− 4 ); and we shall begin
with the second sum, which only involves the auxiliary functions F .
Asymptotic formulæ in combinatory analysis 363

1
jp,q = O(n− 4 ).
P
Proof that
5.21. We have, by Cauchy’s theorem,
Fp,q (x) Fp,q (x)
Z Z
(5.211) jp,q = n+1
dx = n+1
dx,
ηp,q x ζp,q x
where ζp,q consists of the contour LM N M ′ L′ shewn in the figure. Here L and L′ are the
ends of ξp,q , LM and M ′ L′ are radii vectors, and M N M ′ is part of a circle Γ1 whose radius
R1 is greater than 1. P is the point e2pπi/q ; and we suppose that R1 is small enough to
ensure that all points of LM and M ′ L′ are at a distance from P less than 12 . The other
circle c shewn in the figure has P as its centre and radius 21 . We denote LM by ̟p,q , M ′ L′

by ̟p,q and M N M ′ by γp,q : and we write Z Z Z Z
1 2 3
(5.212) jp,q = = + + = jp,q + jp,q + jp,q .
ζp,q γp,q ̟p,q ′
̟p,q

M
̟ c

P M′
L ξ
Γ ̟′
L′
Γ1

N η O R l R1

1 .
P
The contribution of jp,q
5.22. Suppose first that x lies on γp,q and outside c. Then, in virtue of (5.11) and Lemma
4.43, we have

(5.221) Fp,q (x) = O( q).
If on the other hand x lies on γp,q , but inside c, we have, by (5.11) and Lemma 4.44,
3
(5.222) Fp,q (x) − χp,q (x) = O(q − 2 ),
364 Paper 36

where √
q
(5.2221) χp,q (x) = ωp,q √ χC/q (xp,q ).
π 2
But, if we refer to the definition of χa (x) in Lemma 4.42, and observe that
a2 a2 log(1/r)


exp = exp <1
4 log(1/x) 4[{log(1/r)}2 + θ 2 ]
if x = reiθ and r > 1, we see that

(5.223) χp,q (x) = O( q)
on the part of γp,q in question. Hence (5.221) holds for all γp,q . It follows that

1
jp,q = O(R1−n q),
X X 3 5
(5.224) jp,q = O(R1−n q 2 ) = O(n 4 R1−n ) ∗
q
This sum tends to zero more rapidly than any power of n, and is therefore completely
trivial.
P 2 P 3
The contributions of jp,q and jp,q .
5.231. We must now consider the sums which arise from the integrals along ̟p,q and
′ ; and it is evident that we need consider in detail only the first of these two lines. We
̟p,q
write
Fp,q (x) − χp,q (x) χp,q (x)
Z Z
2 ′ ′′
(5.2311) jp,q = n+1
dx + n+1
dx = jp,q + jp,q ,
̟p,q x ̟p,q x
say.
In the first place we have, from (5.222),
 Z R1 
′ − 23 dr 3
jp,q = O q n+1
= O(q − 2 n−1 ),
R r
since
β −n

−n
(5.2312) R = 1− = O(1).
n
Thus X X 1 3

(5.2313) jp,q = O{n−1 q − 2 } = O(n− 4 ).

q<O( n)
5.232. In the second place we have
√ Z
′′ q χC/q (xp,q )
jp,q = ωp,q √ dx.
π 2 ̟p,q xn+1

Here, and in many passages in our subsequent argument, it is to be remembered that the number √ of
values of p, corresponding to a given q is less than q, and that the number of values of q is of order n.
Thus we have generally  
X s
X s+1 1
O(q ) = O  q  = O(n 2 s+1 ).

q<O( n)
Asymptotic formulæ in combinatory analysis 365

It is plain that, if we substitute y for xe−2pπi/q , then write x again for y, and finally
substitute for χC/q its explicit expression as an elementary function, given in Lemma 4.42,
we obtain s 
√ 1 −n− 23 √
Z
′′
(5.2321) jp,q = O( q) {E(x) − 1} log x 24 dx = O( q)J,
x
say, where
π2
 
(5.23211) E(x) = exp ,
6q 2 log(1/x)
and the path of integration is now a line related to x = 1 as ̟p,q is to x = e2pπi/q : the line
defined by x = reiθ , where R ≤ r ≤ R1 , and θ is fixed and (by Lemma 4.22) lies between
1/2qν and 1/qν.
Integrating J by parts, we find
" s  #r=R1
1 1 −n+ 1
(5.2322) (n − 24 )J = − {E(x) − 1} log x 24
x
r=R
 1
1 − 2 −n− 23
Z 
− 21 {E(x) − 1} log x 24 dx
x
 3
π2 1 − 2 −n− 23
Z 
+ 2 E(x) log x 24 dx = J1 + J2 + J3 ,
6q x
say.
5.233. In estimating J1 , J2 , and J3 , we must bear the following facts in mind.
(1) Since |x| ≥ R, it follows from (5.2312) that |x|−n = O(1) throughout the range of
integration.

(2) Since 1 − R = β/n and 1/2qν < θ < 1/qν, where ν = [α n], we have
   
1 1
log =O √ ,
x q n

when r = R, and
1 √
= O(q n),
log(1/x)
throughout the range of integration.
(3) Since
π 2 log(1/r)
|E(x)| = exp ,
6q 2 [{log(1/r)}2 + θ 2 ]
E(x) is less than 1 in absolute value when r > 1. And, on the part of the path for which
r < 1, it is of the form  
1
exp O = exp O(1) = O(1).
q 2 nθ 2
366 Paper 36

It is accordingly of the form O(1) throughout the range of integration.


5.234. Thus we have, first
1 1 1 1
(5.2341) J1 = O(1)O(1)O(R1−n ) + O(1)O(q − 2 n− 4 )O(1) = O(q − 2 n− 4 ),
secondly
R1
dr
Z
1 1 1 3
(5.2342) J2 = O(1)O(q 2 n 4 ) 23 = O(q 2 n− 4 ),
R r n+ 24
and thirdly
R1
dr
Z
3 3 1 1
(5.2343) J3 = O(q −2
)O(1)O(q n ) 2 4
23 = O(q − 2 n− 4 ).
R r n+ 24
From (5.2341), (5.2342), (5.2343), and (5.2322), we obtain
1 5 1 7
J = O(q − 2 n− 4 ) + O(q 2 n− 4 );
5 7
′′
and, from (5.2321), jp,q = O(n− 4 ) + O(qn− 4 ).
Summing, we obtain
X 5 X 7 X
′′
(5.2344) jp,q = O(n− 4 q) + O(n− 4 q2)
√ √
q<O( n) q<O( n)
1 1 1
= O(n− 4 ) + O(n− 4 ) = O(n− 4 ).

5.235. From (5.2311), (5.2313), and (5.2344), we obtain


X 1
2
(5.2351) jp,q = O(n− 4 );
and in exactly the same way we can proveX 1
3
(5.2352) jp,q = O(n− 4 ).
And from (5.212), (5.224), (5.2351), and (5.2352), we obtain, finally,
X 1
(5.2353) jp,q = O(n− 4 ).
1
Jp,q = O(n− 4 ).
P
Proof that
5.31. We turn now to the discussion of
f (x) − Fp,q (x)
Z
(5.311) Jp,q = dx
ξp,q xn+1
f (x) − Xp,q (x) Fp,q (x) − χp,q (x)
Z Z
= n+1
dx − dx
ξp,q x ξp,q xn+1
ρp,q (x)
Z
+ n+1
dx
ξp,q x
1 2 3
= Jp,q + Jp,q + Jp,q ,
Asymptotic formulæ in combinatory analysis 367

say, where s  1
q 1 24
ρp,q (x) = ωp,q log xp,q ,
2π xp,q

Xp,q (x) = χp,q (x) + ρp,q (x) = ρp,q (x)E(xp,q ),


E(x) being defined as in (5.23211).
2 and 3 .
P P
Discussion of Jp,q Jp,q
5.32. The discussion of these two sums is, after the analysis which precedes, a simple

matter. The arc ξp,q is less than a constant multiple of 1/q n; and x−n = O(1) on ξp,q .
Also
3
|Fp,q (x) − χp,q (x)| = O(q − 2 ),
by (5.222); and s 
1 1 1
(5.321) log = O(q − 2 n− 4 ),
xp,q
since |xp,q | = R = 1 − (β/n), |amxp,q | < 1/qν.
Hence
5 1
2
Jp,q = O(q − 2 n− 2 ),
1 3 1
= O(n− 2 q − 2 ) = O(n− 2 );
X X
2
(5.322) Jp,q

q<O( n)
and
3
3
Jp,q = O(q −1 n− 4 ),
X 3 X 1
3
(5.323) Jp,q = O(n− 4 1) = O(n− 4 ).

q<O( n)
1 .
P
Discussion of Jp,q
5.33. From (4.321) and (5.2221), we haves
  1
q 1
(5.331) f (x) − Xp,q (x) = ωp,q log 24
xp,q E(xp,q )Ω(x′p,q ),
2π xp,q
where
∞   ∞
Y 1 X
Ω(z) = f (z) − 1 = −1= p(ν)z ν ,
1
1 − zν 1

if |z| < 1 and


4π 2 2πip′
 
x′p,q = exp − + .
q 2 log(1/xp,q ) q
Now
4π 2 log(1/R)
 
|x′p,q | = exp − 2 ,
q {[log(1/R)]2 + θ 2 }
368 Paper 36

where θ is the amplitude of xp,q . Also


    
2 2 2
2 1 1 1
q {[log(1/R)] + θ } = O q 2
+ 2 =O ,
n q n n

while log(1/R) is greater than a constant multiple of 1/n. There is therefore a positive
number δ, less than unity and independent of n and of q, such that

|x′p,q | < δ;

and we may write Ω(x′p,q ) = O(|x′p,q |).


We have therefore
1 23
E(xp,q )Ω(x′p,q ) = O(|x′p,q |− 24 )O(|x′p,q |) = O(|x′p,q | 24 ) = O(1);

and so, by (5.321),


s !
√ 1 O(1) = O(n− 41 ).

f (x) − χp,q (x) = O( q)O log
xp,q

And hence, as the length of ξp,q is of the form O(1/q n), we obtain
3
1
Jp,q = O(q −1 n− 4 ),
X 3 X 1
1
(5.332) Jp,q = O(n− 4 1) = O(n− 4 ).

q<O( n)
5.34. From (5.311), (5.322), (5.323), and
X(5.332), we obtain
1
(5.341) Jp,q = O(n− 4 ).
Completion of the proof.
5.4. From (5.15), (5.17), (5.2353), and X
(5.341),
X we obtain 1
(5.41) p(n) − cp,q,n = O(n− 4 ).
q p
But √
X q d cosh(Cλn /q) − 1
cp,q,n = √ Aq ,
p π 2 dn λn

where X
Aq = ωp,q e−2npπi/q .
p

All that remains, in order to complete the proof of the theorem, is to shew that

cosh(Cλn /q) − 1

may be replaced by 12 eCλn /q ;


Asymptotic formulæ in combinatory analysis 369

and in order to prove this it is only necessary to shew that


X 3 d 12 eCλn /q − cosh(Cλn /q) + 1 1
q2 = O(n− 4 ).
√ dn λn
q<O( n)

On differentiating we find that the sum is of the form


 
    
X 3 1 1 1 X 1
 1
q 2 O +O 3 =O q2 = O(n− 4 ).
√ qn n2 n √ 
q<O( n) q<O( n)

Thus the theorem is proved.

6. Additional remarks on the theorem.


6.1. The theorem which we have proved gives information about p(n) which is in some
ways extraordinarily exact. We are for this reason the more anxious to point out explicitly
two respects in which the results of our analysis are incomplete.
6.21. We have proved that
X 1
p(n) = Aq φq + O(n− 4 ),

where the summation extends over the values of q specified in the theorem, for every fixed
value of α; that is to say that, when α is given, a number K = K(α) can be found such
that X 1
|p(n) − Aq φq | < Kn− 4
for every value of n. It follows that X
(6.211) p(n) = { Aq φq },
where {x} denotes the integer nearest to x, for n ≥ n0 , where n0 = n0 (α) is a certain
function of α.
The question remains whether we can, by an appropriate choice of α, secure the truth of
(6.211) for all values of n, and not merely for all sufficiently large values. Our opinion is that
this is possible, and that it could be proved to be possible without any fundamental change
in our analysis. Such a proof would however involve a very careful revision of our argument.
It would be necessary to replace all formulæ involving O’s by inequalities, containing only
numbers expressed explicitly as functions of the various parameters employed. This process
would certainly add very considerably to the length and the complexity of our argument. It
is, as it stands, sufficient to prove what is, from our point of view, of the greatest interest;
and we have not thought it worth while to elaborate it further.
6.22. The second point of incompleteness of our results is of much greater interest and
importance. We have not proved either that the series

X
Aq φq
1
370 Paper 36

is convergent, or that, if it is convergent, it represents p(n). Nor does it seem likely that
our method is one intrinsically capable of proving these results, if they are true – a point
on which we are not prepared to express any definite opinion.
It should be observed in this connection that we have not even discovered anything definite
concerning the order of magnitude of Aq for large values of q. We can prove nothing better
than the absolutely trivial equation Aq = O(q). On the other hand we can assert that Aq
is, for an infinity of values of q, effectively of an order as great as q, or indeed even that it
does not tend to zero (though of course this is most unlikely).
6.3. Our formula directs us, if we wish to obtain the exact value of p(n) for a large value of

n, to take a number of terms of order n. The numerical data suggest that a considerably
smaller number of terms will be equally effective; and it is easy to see that this conjecture
is correct.
Let us write  √ 
β n
q 
2
β = 4π 3 = 4C, µ= ,
log n
and let us suppose that α < 2. Then
 
ν ν   √ ν √
X X 3 1 1 X √ C n/q 
Aq φq = O(q 2 )O O(eC n/q
) = O qe
qn n
µ+1 µ+1 µ+1
 Z ν
√ C √n/x

1
= O xe dx ,
n µ
√ √
since qeC n/q decreases steadily throughout the range of summation∗ .

Writing n/y for x, we obtain
Z √n/µ ! (  √ − 5 √ )
− 41 − 25 Cy − 41 n 2 1 5 1
O n y e dy = O n eC n/µ = O{n− 4 (log n)− 2 e 4 log n }
1/α µ
5
= O(log n)− 2 = o(1).

It follows that it is enough, when n is sufficiently large, to take


 √ 
β n
log n

terms of the series. It is probably also necessary to take a number of terms of order

n/(log n); but it is not possible to prove this rigorously without a more exact knowledge
of the properties of Aq than we possess.
6.4. We add a word on certain simple approximate formulæ for log p(n) found empirically
by Major MacMahon and by ourselves. Major MacMahon found that if
∗ √
The minimum occurs when q is about equal to 2C n.
Asymptotic formulæ in combinatory analysis 371

p
(6.41) log10 p(n) = (n + 4) − an ,
then an is approximately equal to 2 within the limits of his table of values of p(n) (Table
IV). This suggested to us that we should endeavour to find more accurate formulæ of the
same type. The most striking that we have found p is
10
(6.42) log10 p(n) = 9 { (n + 10) − an };
the mode of variation of an is shewn in Table III.
In this connection it is interesting to observe that the function

13− n
p(n)

(which ultimately tends to infinity with exponential rapidity) is equal to .973 for n =
30000000000.

7. Further applications of the method.


7.1. We shall conclude with a few remarks concerning actual or possible applications of our
method to other problems in Combinatory Analysis or the Analytic Theory of Numbers.
The class of problems in which the method gives the most striking results may be defined
as follows. Suppose that q(n) is the coefficient of xn in the expansion of F (x), where F (x)
is a function of the form ′ ′
{f (±xa )}α {f (±xa )}α · · · ∗
(7.11) ;
{f (±xb )}β {f (±xb′ )}β ′ · · ·
f (x) being the function considered in this paper, The a’s, b’s, α’s, and β’s being positive
integers, and the number of factors in numerator and denominator being finite; and suppose
that |F (x)| tends exponentially to infinity when x tends in an appropriate manner to some
or all the points e2pπi/q . Then our method may be applied in its full power to the asymptotic
study of q(n), and yields results very similar to those which we have found concerning p(n).
Thus, if

f (x) 1
F (x) = 2
= (1 + x)(1 + x2 )(1 + x3 ) · · · = ,
f (x ) (1 − x)(1 − x3 )(1 − x5 ) · · ·

so that q(n) is the number of partitions of n into odd parts, or into unequal parts† , we find
that
1 d
q
q(n) = √ 1
J0 [iπ { 13 (n + 24 )}]
2 dn
√ d
q
+ 2 cos( 3 nπ − 9 π) J0 [ 3 iπ { 31 (n + 24
2 1 1 1
)}] + · · · .
dn

Since
{f (x2 )}3
f (−x) = ,
f (x)f (x4 )
the arguments with a negative sign may be eliminated if this is desired.

Cf. MacMahon, loc. cit, p. 11. We give at the end of the paper a table (Table V) of the values of q(n)
up to n = 100. This table was calculated by Mr. Darling.
372 Paper 36


The error after [α n] terms is of the form O(1). We are not in a position to assert that
the exact value of q(n) can always be obtained from the formula (though this is probable);
but the error is certainly bounded.
If
f (x2 ) f (x)f (x4 )
F (x) = = = (1 + x)(1 + x3 ) + x5 ) · · · ,
f (−x) {f (x2 )}2
so that q(n) is a number of partitions of n into parts which are both odd and unequal, then

d
q
q(n) = J0 [iπ { 16 (n − 24
1
)}]
dn
d
q
+2 cos( 23 nπ − 92 π) J0 [ 13 iπ { 61 (n − 24
1
)}] + · · · .
dn
The error is again bounded (and probably tends to zero).
If
{f (x)}2 1
F (x) = = ,
f (x2 ) 1 − 2x + 2x4 − 2x9 + · · ·
q(n) has no very simple arithmetical interpretation; but the series is none the less, as the
direct reciprocal of simple ϑ-function, of particular interest. In this case we find
√ √ 1 √
1 d eπ n 3 2 1 d e3π n
q(n) = √ + cos( 3 nπ − 6 π) √ + ···.
4π dn n 2π dn n
1
The error here is (as in the partition problem) of order O(n− 4 ), and the exact value can
always be found from the formula.
7.2. The method also be applied to product of form (7.11) which have (to put the matter
roughly) no exponential infinities. In such cases the approximation is of much less exact
character. On the other hand the problems of this character are of even greater arithmetical
interest.
The standard problem of this category is that of the representation of a number as a sum
of s squares, s being any positive integer odd or even∗ . We must reserve the application of
our method to this problem for another occasion; but we can indicate the character of our
main result as follows.
If rs (n) is the number of representations of n as the sum of s squares we have
X {f (x2 )}s {f (x)}2s {f (x4 )}2s
F (x) = rs (n)xn = (1 + 2x + 2x4 + · · ·)s = = .
{f (−x)}2s {f (x2 )}5s

We find that
1
π 2 s 1 s−1 X cq 1
s
(7.21) rs (n) = 1 n
2
1 + O(n ),
4
Γ( 2 s) q2 s


As is well known, the arithmetical difficulties of the problem are much greater when s is odd.
Asymptotic formulæ in combinatory analysis 373

where cq is a function of q and of n of the same general type as the function Aq of this
paper. The series X cq
(7.22) 1
q 2s
is absolutely convergent for sufficiently large values of s, and the summation in (7.21) may

be regarded indifferently as extended over all values of q or only over a range 1 ≤ q ≤ α n.
It should be observed that the series (7.22) is quite different in form from any of the infinite
series which are already known to occur in connection with this problem.
7.3. There is also a wide range of problems to which our methods are partly applicable.
Suppose, for example, that
X 1
F (x) = p2 (n)xn = ,
(1 − x)(1 − x4 )(1 − x9 ) · · ·

so that p2 (n) is the number of partitions of n into squares. Then F (x) is not an elliptic
modular function; it possesses no general transformation theory: and the full force of our
method can not be applied. We can still, however, apply some of our preliminary methods.
Thus the “Tauberian” argument shews that
4 1 2 1
log p2 (n) ∼ 2− 3 3π 3 {ζ( 32 )} 3 n 3 .

And although there is no general transformation theory, there is a formula which enables
us to specify the nature of the singularity at x = 1. This formula is
r   
1 π 2π 1
= 2 exp √ ζ(− 2 )
f (e−πz ) z z
Y∞ √ √
× {1 − 2e−2π (n/z) cos 2π (n/z) + e−4π (n/z) }.
p

By the use of this formula, in conjunction with Cauchy’s theorem, it is certainly possible
to obtain much more precise information about p2 (n) and in particular the formula
1 2 1
7 2 −(4/3) 3π (1/3) {ζ( 3 )} 3 n 3
p2 (n) ∼ 3− 2 (4πn)− 6 {ζ( 32 )} 3 e2 2 .

The corresponding formula for ps (n), the number of partitions of n into perfect s-th powers,
is s 
1 s 1 3 1/(s+1)
s
p (n) ∼ (2π) − 2 (s+1)
kn s+1 − 2 e(s+1)kn ,
s+1
where      s
1 1 1 s+1
k= Γ 1+ ζ 1+ .
s s s
374 Paper 36

The series (7.21) may be written in the form


1
s
π 2 s 1 s−1 X ωp,q −npπi/q
n 2
1 e ,
Γ( 21 s) p,q q 2
s

1 1 1 1
where ωp,q is always one of the five numbers 0, e 4 πi , e− 4 πi , −e 4 πi , −e− 4 πi . When s is even
it begins
1
π 2 s 1 s−1 − 1 s 1 1
1 n
2 {1 2 + 2 cos( 12 nπ − 41 sπ)2− 2 s + 2 cos( 32 nπ − 12 sπ)3− 2 s + · · ·}.
Γ( 2 s)

It has been proved by Ramanujan that the series gives an exact representation of rs (n)
when s = 4, 6, 8; and by Hardy that this also true when s = 3, 5, 7. See Ramanujan, “On
certain trigonometrical sums and their applications in the Theory of Numbers”; Hardy,
“On the expression of a number as the sum of any number of squares, and in particular of
five or seven∗ ”.


[Ramanujan’s paper referred to is No. 21 of this volume. That of Hardy was published, in the first
instance, in Proc. National Acad. of Sciences, (Washington), Vol. IV, 1918, pp. 189 – 193, and later (in
fuller form and with a slightly different title) in Trans. American Math. Soc., Vol. XXI, 1920, pp. 255 –
284.]
Asymptotic formulæ in combinatory analysis 375

Table I: ωp,q .

p q log ωp,q /πi p q log ωp,q /πi p q log ωp,q /πi


1 1 0 3 11 3/22 8 15 7/18
1 2 0 4 11 3/22 11 15 -19/90
1 3 1/18 5 11 −5/22 13 15 -7/18
2 3 −1/18 6 11 5/22 14 15 -1/90
1 4 1/8 7 11 −3/22 1 16 -29/32
3 4 −1/8 8 11 −3/22 3 16 -27/32
1 5 1/5 9 11 −5/22 5 16 -5/32
2 5 0 10 11 −15/22 7 16 -3/32
3 5 0 1 12 55/72 9 16 3/32
4 5 −1/5 5 12 −1/72 11 16 5/32
1 6 5/18 7 12 1/72 13 16 27/32
5 6 −5/18 11 12 −55/72 15 16 29/32
1 7 5/14 1 13 11/13 1 17 -14/17
2 7 1/14 2 13 4/13 2 17 8/17
3 7 −1/14 3 13 1/13 3 17 5/17
4 7 1/14 4 13 −1/13 4 17 0
5 7 −1/14 5 13 0 5 17 1/17
6 7 −5/14 6 13 −4/13 6 17 5/17
1 8 7/16 7 13 4/13 7 17 1/17
3 8 1/16 8 13 0 8 17 -8/17
5 8 −1/16 9 13 1/13 9 17 8/17
7 8 −7/16 10 13 −1/13 10 17 -1/17
1 9 14/27 11 13 −4/13 11 17 -5/17
2 9 4/27 12 13 −11/13 12 17 -1/17
4 9 −4/27 1 14 13/14 13 17 0
5 9 4/27 3 14 3/14 14 17 -5/17
7 9 −4/27 5 14 3/14 15 17 -8/17
8 9 −14/27 9 14 −3/14 16 17 14/17
1 10 3/5 11 14 −3/14 1 18 -20/27
3 10 0 13 14 −13/14 5 18 2/27
7 10 0 1 15 1/90 7 18 -2/27
9 10 −3/5 2 15 7/18 11 18 2/27
1 11 15/22 4 15 19/90 13 18 -2/27
2 11 5/22 7 15 −7/18 17 18 20/27
376 Paper 36

Table II: Aq .

A1 = 1.
A2 = cos nπ.
A3 = 2 cos( 23 nπ − 1
18 π).
A4 = 2 cos( 12 nπ − 18 π).
A5 = 2 cos( 25 nπ − 15 π) + 2 cos 54 nπ.
A6 = 2 cos( 13 nπ − 18 5
π).
A7 = 2 cos( 27 nπ − 14 5
π) + 2 cos( 74 nπ − 14 1
π) + 2 cos( 76 nπ + 14 1
π).
A8 = 2 cos( 14 nπ − 16 7
π) + 2 cos( 43 nπ − 16 1
π).
A9 = 2 cos( 29 nπ − 14 4 4 8
27 π) + 2 cos( 9 nπ − 27 π) + 2 cos( 9 nπ + 27 π).
4

A10 = 2 cos( 15 nπ − 35 π) + 2 cos 53 nπ.


2
A11 = 2 cos( 11 nπ − 15 4 5 6
22 π) + 2 cos( 11 nπ − 22 π) + 2 cos( 11 nπ − 22 π)
3

8 3 10 5
+2 cos( 11 nπ − 22 π) + 2 cos( 11 nπ + 22 π).
A12 = 2 cos( 16 nπ − 55 5 1
72 π) + 2 cos( 6 nπ + 72 π).
2
A13 = 2 cos( 13 nπ − 11 4 4 6
13 π) + 2 cos( 13 nπ − 13 π) + 2 cos( 13 nπ − 13 π)
1

8 1 10 12 4
+2 cos( 13 nπ + 13 π) + 2 cos 13 nπ + 2 cos( 13 nπ + 15 π).
A14 = 2 cos( 71 nπ − 13 3 3 5
14 π) + 2 cos( 7 nπ − 14 π) + 2 cos( 7 nπ − 14 π).
3

2 1 4 7 8 19
A15 = 2 cos( 15 nπ − 90 π) + 2 cos( 15 nπ − 18 π) + 2 cos( 15 nπ − 90 π) + 2 cos( 14 7
15 nπ + 18 π).
A16 = 2 cos( 18 nπ + 29 3 27 5 5 7
32 π) + 2 cos( 8 nπ + 32 π) + 2 cos( 8 nπ + 32 π) + 2 cos( 8 nπ + 32 π).
3

2
A17 = 2 cos( 17 nπ + 14 4 8 6 5
17 π) + 2 cos( 17 nπ − 17 π) + 2 cos( 17 nπ − 17 π) + 2 cos 17 nπ
8

10 1 12 5 14 1 16 8
+2 cos( 17 nπ − 17 π) + 2 cos( 17 nπ − 17 π) + 2 cos( 17 nπ − 17 π) + 2 cos( 17 nπ + 17 π).
A18 = 2 cos( 91 nπ + 20 5 2 7
27 π) + 2 cos( 9 nπ − 27 π) + 2 cos( 9 nπ + 27 π).
2

It may be observed that


A5 = 0 (n ≡ 1, 2 (mod 5)), A7 = 0 (n ≡ 1, 3, 4 (mod 7)),

A10 = 0 (n ≡ 1, 2 (mod 5)), A11 = 0 (n ≡ 1, 2, 3, 5, 7 (mod 11)),

A13 = 0 (n ≡ 2, 3, 5, 7, 9, 10 (mod 13)), A14 = 0 (n ≡ 1, 3, 4 (mod 7)),

A16 = 0 (n ≡ 0 (mod 2)), A17 = 0 (n ≡ 1, 3, 4, 6, 7, 9, 13, 14 (mod 17));


while A1 , A2 , A3 , A4 , A6 , A8 , A9 , A12 , A15 and A18 never vanish.
Asymptotic formulæ in combinatory analysis 377

10
p
Table III: log10 p(n) = 9 { (n + 10) − an }.

n an n an
1 3.317 10000 4.148
3 3.176 30000 4.364
10 3.011 100000 4.448
30 2.951 300000 4.267
100 3.036 1000000 3.554
300 3.237 3000000 2.072
1000 3.537 10000000 -1.188
3000 3.838 30000000 -6.796
∞ −∞

Table IV∗ : p(n).

1 1 21 792 41 44583 61 1121505


2 2 22 1002 42 53174 62 1300156
3 3 23 1255 43 63261 63 1505499
4 5 24 1575 44 75175 64 1741630
5 7 25 1958 45 89134 65 2012558
6 11 26 2436 46 105558 66 2323520
7 15 27 3010 47 124754 67 2679689
8 22 28 3718 48 147273 68 3087735
9 30 29 4565 49 173525 69 3554345
10 42 30 5604 50 204226 70 4087968
11 56 31 6842 51 239943 71 4697205
12 77 32 8349 52 281589 72 5392783
13 101 33 10143 53 329931 73 6185689
14 135 34 12310 54 386155 74 7089500
15 176 35 14883 55 451276 75 8118264
16 231 36 17977 56 526823 76 9289091
17 297 37 21637 57 614154 77 10619863
18 385 38 26015 58 715220 78 12132164
19 490 39 31185 59 831820 79 13848650
20 627 40 37338 60 966467 80 15796476

. . . contd.

The numbers in this table were calculated by Major MacMahon, by means of the recurrence formulæ
obtained by equating the coefficients in the identity

X
(1 − x − x2 + x5 + x7 − x12 − x15 + · · ·) p(n)xn = 1.
0

We have verified the table by direct calculation up to n = 158. Our calculation of p(200) from the asymptotic
formula then seemed to render further verification unnecessary.
378 Paper 36

Table IV (Contd.)

81 18004327 111 679903203 141 16670689208 171 301384802048


82 20506255 112 761002156 142 18440293320 172 330495499613
83 23338469 113 851376628 143 20390982757 173 362326859895
84 26543660 114 952050665 144 22540654445 174 397125074750
85 30167357 115 1064144451 145 24908858009 175 435157697830
86 34262962 116 1188908248 146 27517052599 176 476715857290
87 38887673 117 1327710076 147 30388671978 177 522115831195
88 44108109 118 1482074143 148 33549419497 178 571701605655
89 49995925 119 1653668665 149 37027355200 179 625846753120
90 56634173 120 1844349560 150 40853235313 180 684957390936
91 64112359 121 2056148051 151 45060624582 181 749474411781
92 72533807 122 2291320912 152 49686288421 182 819876908323
93 82010177 123 2552338241 153 54770336324 183 896684817527
94 92669720 124 2841940500 154 60356673280 184 980462880430
95 104651419 125 3163127352 155 66493182097 185 1071823774337
96 118114304 126 3519222692 156 73232243759 186 1171432692373
97 133230930 127 3913864295 157 80630964769 187 1280011042268
98 150198136 128 4351078600 158 88751778802 188 1398341745571
99 169229875 129 4835271870 159 97662728555 189 1527273599625
100 190569292 130 5371315400 160 107438159466 190 1667727404093
101 214481126 131 5964539504 161 118159068427 191 1820701100652
102 241265379 132 6620830889 162 129913904637 192 1987276856363
103 271248950 133 7346629512 163 142798995930 193 2168627105469
104 304801365 134 8149040695 164 156919475295 194 2366022741845
105 342325709 135 9035836076 165 172389800255 195 2580840212973
106 384276336 136 10015581680 166 189334822579 196 2814570987591
107 431149389 137 11097645016 167 207890420102 197 3068829878530
108 483502844 138 12292341831 168 228204732751 198 3345365983698
109 541946240 139 13610949895 169 250438925115 199 3646072432125
110 607163746 140 15065878135 170 274768617130 200 3972999029388
Asymptotic formulæ in combinatory analysis 379

Table V ∗ : q(n).

n cn n cn n cn n cn
1 1 26 165 51 4097 76 53250
2 1 27 192 52 4582 77 58499
3 2 28 222 53 5120 78 64234
4 2 29 256 54 5718 79 70488
5 3 30 296 55 6378 80 77312
6 4 31 340 56 7108 81 84756
7 5 32 390 57 7917 82 92864
8 6 33 448 58 8808 83 101698
9 8 34 512 59 9792 84 111322
10 10 35 585 60 10880 85 121792
11 12 36 668 61 12076 86 133184
12 15 37 760 62 13394 87 145578
13 18 38 864 63 14848 88 159046
14 22 39 982 64 16444 89 173682
15 27 40 1113 65 18200 90 189586
16 32 41 1260 66 20132 91 206848
17 38 42 1426 67 22250 92 225585
18 46 43 1610 68 24576 93 245920
19 54 44 1816 69 27130 94 267968
20 64 45 2048 70 29927 95 291874
21 76 46 2304 71 32992 96 317788
22 89 47 2590 72 36352 97 345856
23 104 48 2910 73 40026 98 376256
24 122 49 3264 74 44046 99 409174
25 142 50 3658 75 48446 100 444793


We are indebted to Mr. Darling for this table.
On the coefficients in the expansions of certain modular
functions

Proceedings of the Royal Society, A, XCV, 1919, 144 – 155

1. A very large proportion of the most interesting arithmetical functions – of the functions,
for example, which occur in the theory of partitions, the theory of the divisors of numbers,
or the theory of the representation of numbers by sums of squares – occur as the coefficients
in the expansions of elliptic modular functions in powers of the variable q = eπiτ . All of
these functions have a restricted region of existence, the unit circle |q| = 1 being a “natural
boundary” or line of essential singularities. The most important of them, such as the
functions∗

(1.1) (ω1 /π)12 ∆ = q 2 {(1 − q 2 )(1 − q 4 ) · · ·}24 ,

(1.2) ϑ3 (0) = 1 + 2q + 2q 4 + 2q 9 + · · · ,

13 q 2 23 q 4
 ω 4  
1
(1.3) 12 g2 = 1 + 240 + + ··· ,
π 1 − q2 1 − q4

15 q 2 25 q 4
 ω 6  
1
(1.4) 216 g3 = 1 − 504 + + ··· ,
π 1 − q2 1 − q4

are regular inside the unit circle; and many, such as the functions (1.1) and (1.2), have the
additional property of having no zeros inside the circle, so that their reciprocals are also
regular.
In a series of recent papers† we have applied a new method to the study of these arithmetical
functions. Our aim has been to express them as series which exhibit explicitly their order
of magnitude, and the genesis of their irregular variations as n increases. We find, for
example, for p(n), the number of unrestricted partitions of n, and for rs (n), the number of
representations of n as the sum of an even number s of squares, the series

We follow, in general, the notation of Tannery and Molk’s Éléments de la théorie des fonctions ellip-
tiques. Tannery and Molk, however, write 16G in place of the more usual ∆.

(1) G. H. Hardy and S. Ramanujan, “Une formule asymptotique pour le nombre des partitions de n,”
Comptes Rendus, January 2, 1917 [No. 31 of this volume]; (2) G. H. Hardy and S. Ramanujan, “Asymptotic
Formulæ in Combinatory Analysis,” Proc. London Math.Soc., Ser. 2, Vol. XVII, 1918, pp. 75 – 115 [No.
36 of this volume]; (3) S. Ramanujan, “On Certain Trigonometrical Sums and their Applications in the
Theory of Numbers,” Trans. Camb. Phil. Soc., Vol.XXII, 1918, pp. 259 – 276 [No. 21 of this volume]; (4)
G. H. Hardy, “On the Expression of a Number as the Sum of any Number of Squares, and in particular of
Five or Seven,” Proc. National Acad. of Sciences, Vol.IV, 1918, pp. 189 – 193: [and G. H. Hardy, “On the
expression of a number as the sum of any number of squares, and in particular of five,” Trans. American
Math. Soc., Vol.XXI, 1920, pp. 255 – 284].
On the coefficients in the expansions of certain modular functions 381

1
!
eCλn (−1)n d e 2 Cλn /2
 
1 d
(1.5) √ +
2π 2 dn λn 2π dn λn
s  !
eCλn /3
 
3 2 1 d
+ π cos nπ − π + ···,
2 3 18 dn λn
s  s 
1 2
where λn = n− and C = π , and
24 3

π s/2 s −1 −s 1 1 −s 2 1 −s
(1.6) n 2 {1 2 + 2 cos( nπ − sπ)2 2 + 2 cos( nπ − sπ)3 2 + · · ·};
Γ(s/2) 2 4 3 2
and our methods enable us to write down similar formulæ for a very large variety of other
arithmetical functions.
The study of series such as (1.5) and (1.6) raises a number of interesting problems, some of
which appear to be exceedingly difficult. The first purpose for which they are intended is
that of obtaining approximations to the functions with which they are associated. Some-
times they give also an exact representation of the functions, and sometimes they do not.
Thus the sum of the series (1.6) is equal to rs (n)if s is 4, 6, or 8, but not in any other
case. The series (1.5) enables us, by stopping after an appropriate number of terms, to
find approximations to p(n) of quite startling accuracy; thus six terms of the series give
p(200) = 3972999029388, a number of 13 figures, with an error of 0.004. But we have never
been able to prove that the sum of the series is p(n) exactly, nor even that it is convergent.
There is one class of series, of the same general character as (1.5) or (1.6), which lends
itself to comparatively simple treatment. These series arise when the generating modular
function f (q) of φ(τ ) satisfies an equation
 
c + dτ
φ(τ ) = (a + bτ )n φ ,
a + bτ
where n is a positive integer, and behaves, inside the unit circle, like a rational function;
that is to say, possesses no singularities but poles. The simplest examples of such functions
are the reciprocals of the functions (1.3) and (1.4). The coefficients in their expansions are
integral, but possess otherwise no particular arithmetical interest. The results, however,
are very remarkable from the point of view of approximation; and it is in any case, well
worthwhile, in view of the many arithmetical applications of this type of series, to study in
detail any example in which the results can be obtained by comparatively simple analysis.
We begin by proving a general theorem (Theorem 1) concerning the expression of a modular
function with poles as a series of partial fractions. This series is (as appears in Theorem
2) a “Poincaré’s series”: what our theorem asserts is, in effect, that the sum of a certain
Poincaré’s series is the only function which satisfies certain conditions. It would, no doubt,
be possible to obtain this result as a corollary from propositions in the general theory
382 Paper 37

of automorphic functions; but we thought it best to give an independent proof, which is


interesting in itself and demands no knowledge of this theory.

2. Theorem 1. Suppose that


(2.1) f (q) = f (eπiτ ) = φ(τ )
is regular for q = 0, has no singularities save poles within the unit circle, and satisfies the
functional equation  
c + dτ
(2.2) φ(τ ) = (a + bτ )n φ = (a + bτ )n φ(T ),
a + bτ
n being a positive integer and, a, b, c, d any integers such that ad − bc = 1. Then
(2.3) f (q) = ΣR,
where R is a residue of f (x)/(q − x) at a pole of f (x), if |q| < 1; while if |q| > 1 the sum
of the series on the right hand side of (2.3) is zero.
The proof requires certain geometrical preliminaries.

3. The half-plane I(τ ) > 0, which corresponds to the inside of the unit circle in the plane
of q, is divided up, by the substitutions of the modular group, into a series of triangles
whose sides are arcs of circles and whose angles are π/3, π/3, and 0∗ . One of these, which
is called the fundamental polygon (P )† , has its vertices at the points ρ, ρ2 , and i∞, where
ρ = eπi/3 , and its sides are parts of the unit circle |τ | = 1 and the lines R(τ ) = ± 21 .
Suppose that Fm is the “Farey’s series” of order m, that is to say the aggregate of the
rational fractions between 0 and 1, whose denominators are not greater than m, arranged
in order of magnitude‡ , and that h′ /k′ and h/k, where 0 < h′ /k′ < h/k < 1, are two
adjacent terms in the series. We shall consider what regions in the τ -plane correspond to
P in the T -plane, when
h′ − k′ τ h − kτ
(3.1) T =− , (3.2) T = ′ .
h − kτ h − k′ τ
′ ′
Both of these substitutions belong to the modular group, since hk − h k = 1. The points
i∞, 1/2, −1/2, in the T -plane correspond to h/k, (h + 2h′ )/(k + 2k′ ), (h − 2h′ )/(k − 2k′ ) in
the τ -plane. Thus the lines R(T ) = 21 , R(T ) = − 21 correspond to semicircles described on
the segments
h h + 2h′ h h − 2h′
   
, , ,
k k + 2k′ k k − 2k′
respectively as diameters. Similarly the upper half of the unit circle corresponds to a
semicircle on the segment

h + h′ h − h′
 
, .
k + k′ k − k′

It is for many purposes necessary to divide each triangle into two, whose angles are π/2, π/3, and 0;
but this further subdivision is not required for our present purpose. For the detailed theory of the modular
group, see Klien-Fricke, Vorlesungen über die Theorie der Elliptischen Modulfunktionen, 1890-1892.

See Fig. 1.

The first and last terms are 0/1 and 1/1. A brief account of the properties of Farey’s series is given in
§4.2 of our paper (2)[pp. 355 – 356 of this volume].
On the coefficients in the expansions of certain modular functions 383

p2 p

−1 − 21 0 1
2
1
Fig 1

The polygon P corresponds to the region bounded by these three semicircles. In particular,
the right hand edge of P corresponds to a circular arc stretching from h/k (where it cuts
the real axis at right angles) to the point √
h′ k′ + hk + 21 (hk′ + h′ k) + 12 i 3
(3.3)
k2 + kk′ + k′2
corresponding to τ = ρ.
Similarly we find that the substitution (3.2) correlates to P a triangle bounded by semicir-
cles on the segments
 ′ ′   ′ ′   ′
h − h h′ + h

h h − 2h h h + 2h
, , , , , .
k′ k′ − 2k k′ k′ + 2k k′ − k k′ + k

In particular, the left hand edge of P corresponds to a circular arc from h′ /k′ to the point
(3.3). These two arcs, meeting at the point (3.3), form a continuous path ω, connecting h/k
and h′ /k′ , every point of which corresponds, in virtue of one or other of the substitutions
(3.1) and (3.2), to a point on one of the rectilinear boundaries of P ∗ .
Performing a similar construction for every pair of adjacent fractions of Fm , we obtain a
continuous path from τ = 0 to τ = 1. This path, and its reflexion in the imaginary axis,

Fig. 2 illustrates the case in which h/k = 35 , h′ /k′ = 21 . These fractions are adjacent in F5 and F6 , but
not in F7 .
384 Paper 37

give a continuous path from τ = −1 to τ = 1, which we shall denote by Ωm . To Ωm


corresponds a path in the q-plane, which we call Hm ; Hm is a closed path, formed entirely
by arcs of circles which cut the unit circle at right angles.

1 5 4 7 3 5 2
2 9 7 12 5 8 3
Fig 2
The region shaded horizontally corresponds to P for the substitution (3.1), that
shaded vertically for the substitution (3.2). The thickest lines shew the path ω;
the line of medium thickness shews the semicircle which corresponds (for either
substitution) to the unit semicircle in the plane of T . The large incomplete
semicircle passes through τ = 1.

Since
h′ h′ + 2h h + 2h′ h
< , < ,
k′ k′ + 2k k + 2k′ k
the path ω from h′ /k′ to h/k is always passing from left to right, and its length is less than
twice that of the semicircle on (h′ /k′ , h/k), i.e., than π/kk ′ . The total length of Ωm is less
than 2π; and, since
dq πiτ
= πie ≤ π,

the length of Hm is less than 2π 2 . Finally, we observe that the maximum distance of Ωm
from the real axis is less than half the maximum distance between two adjacent terms of
Fm , and so less than 1/2m∗ . Hence Ωm tends uniformly to the real axis, and Hm to the
unit circle, when m −→ ∞.

4. The function φ(τ ) can have but a finite number of poles in P ; we suppose, for simplicity,
that none of them lie on the boundary. There is then a constant K such that |f (q)| < K
on the boundary of P .

See Lemma 4.22 of our paper (2) [p. 356 of this volume].
On the coefficients in the expansions of certain modular functions 385

We now consider the integral


1 f (x)
Z
(4.1) dx,
2πi x−q
where |q| < 1 and the contour of integration is Hm ∗ . By Cauchy’s Theorem, the integral is
equal to
f (q) − ΣR,
where R is a residue of f (x)/(q − x) at a pole of f (x) inside Hm † . To prove our theorem,
then, we have merely to shew that the integral (4.1) tends to zero when m −→ ∞.
Let ω1′ and ω1 be the left- and right-hand parts of ω, and ζ1′ , ζ1 and ζ the corresponding
arcs of Hm . The length of ω1 is, as we have seen, less than 21 π/kk′ , and that of ζ1 than
1 2 ′
2 π /kk . Further, we have, on ζ1 ,
n
h h′
 
n K
|f (x)| = |φ(τ )| = |h − kτ | |φ(T )| < K k − ′ = ′n .
k k k
Thus the contribution of ζ1 to the integral is numerically less than C/(kk′n+1 ), where C is
independent of m; and the whole integral (4.1)  is numerically
 less than
1 1 1
(4.2) 2CΣ ′ + ,
kk kn k′n
where the summation extends to all pairs of adjacent terms of Fm .
When ν is fixed and m > ν, the number of terms of Fm whose denominators are less than
ν is a function of ν only, say N (ν). If h/k is one of these, and h′ /k′ is adjacent to it,
k + k′ > m‡ , and so k′ > m − ν. Thus the terms of (4.2) in which either k or k′ is less than
ν contribute less than 8CN (ν)/(m − ν). The remaining terms contribute less than
4C X 1 4C
n ′
= n.
ν kk ν
Hence the sum (4.2) is less than
8CN (ν) 4C
+ n,
m−ν ν
and it is plain that, by choice of first ν and then m, this may be made as small as we
please. Thus (4.1) tends to zero and the theorem is proved. It should be observed that ΣR
must, for the present at any rate, be interpreted as meaning the limit of the sum of terms
corresponding to poles inside Hm ; we have not established the absolute convergence of the
series.
We supposed that no pole of φ(τ ) lies on the boundary of P . This restriction, however,
is in no way essential; if it is not satisfied, we have only to select our “fundamental poly-
gon” somewhat differently. The theorem is consequently true independently of any such
restriction.

Strictly speaking, f (x) is not defined at the points where Hm meets the unit circle, and we should
integrate round a path just inside Hm and proceed to the limit. The point is trivial, as f (x), in virtue of
the functional equation, tends to zero when we approach a cusp of Hm from inside.

We suppose m large enough to ensure that x = q lies inside Hm .

See our paper (2), loc. cit., [p. 356]
386 Paper 37

So far we have supposed |q| < 1. It is plain that, if |q| > 1, the same reasoning proves that
(4.3) ΣR = 0.

5. Suppose in particular that φ(τ ) has one pole only, and that a simple pole at τ = α, with
residue A. The complete system of poles is then given by
c + dα
(5.1) τ =a= (ad − bc = 1),
a + bα
If a and b are fixed, and (c, d) is one pair of solutions of ad − bc = 1, the complete system
of solutions is (c + ma, d + mb), where m is an integer. To each pair (a, b) correspond an
infinity of poles in the plane of τ ; but these poles correspond to two different poles only in
the plane of q, viz,
(5.2) q = ±q = ±eπia ,
the positive and negative signs corresponding to even and odd values of m respectively. It
is to be observed, moreover, that different pairs (a, b) may give rise to the same pole q.
The residue of φ(τ ) for τ = a is , in virtue of the functional equation (2.2),
A
;
(a + bα)n+2

and the residue of f (q) for q = q is


 
A dq πiAq
= .
(a + bα)n+2 dτ τ =a (a + bα)n+2

Thus the sum of the terms of our series which correspond to the poles (5.2) is

q2
 
πiA q q 2πiA
n+2
− = n+2
.
(a + bα) q−q q+q (a + bα) q − q2
2

We thus obtain:
Theorem 2. If φ(τ ) has one pole only in P , viz., a simple pole at τ = α with residue A,
and |q| < 1, then
X 1 q2
(5.3) f (q) = 2πiA ,
(a + bα)n+2 q 2 − q2
where  
c + dα
q = exp πi;
a + bα
c, d being any pair of solutions of ad − bc = 1, and the summation extending over all pairs
a, b, which give rise to distinct values of q. If |q| > 1, the sum of the series on the right-hand
side of (5.3) is zero.
If φ(τ ) has several poles in P , f (q), of course, will be the sum of a number of series such
as (5.3). Incidentally, we may observe that it now appears that the series in question are
absolutely convergent.

6. As an example, we select the function


On the coefficients in the expansions of certain modular functions 387


π6 1 X
(6.1) f (q) = 6 = P∞ r 5 q 2r
= p n xn ,
216ω1 g3 1 − 504 1 0
1−q 2r
say, where x = q2.
It is evident that pn is always an integer; the values of the first 13
coefficients are
p0 = 1,
p1 = 504,
p2 = 270648,
p3 = 144912096,
p4 = 77599626552,
p5 = 41553943041744,
p6 = 22251789971649504,
p7 = 11915647845248387520,
p8 = 6380729991419236488504,
p9 = 3416827666558895485479576,
p10 = 1829682703808504464920468048,
p11 = 979779820147442370107345764512,
p12 = 524663917940510191509934144603104;
so that p12 is a number of 33 figures.
By means of the formulæ∗

8 1
g3 = (e1 − e3 )2 (1 + k2 )(1 − k2 )(1 − 2k2 ),
27 2
 2
π 2K
e1 − e3 = {ϑ3 (0)}4 , = {ϑ3 (0)}2 ,
2ω1 π
we find that
 6
1 2K 1
= (1 + k2 )(1 − k2 )(1 − 2k2 ).
f (q) π 2

The value of n is 6. The poles of f (q) correspond to the value of τ which make K = k2
equal to −1, 2 or 12 . It is easily verified† that these values are given by the general formula

c + di
τ= (ad − bc = 1),
a + bi

so that    
c + di ac + bd π
(6.2) q = exp πi = exp πi − 2 .
a + bi a2 + b2 a + b2

All the formulæ which we quote are given in Tannery and Molk’s Tables; see in particular Tables XXXVI
(3), LXXI (3), XCVI, CX (3).

A full account of the problem of finding τ when κ if given will be found in Tannery and Molk, loc. cit.,
Vol. III, ch. 7 (“On donne k2 ou g2 , g3 ; trouver τ ou ω1 , ω3 ”).
388 Paper 37

The value of α is i∗ . In order to determine A we observe that


 5 2
25 q 4 16 q 2 26 q 4
  
d 1 q 1008
−504 + + ··· = − + + ··· .
dq 1 − q 2 1 − q 4 q (1 − q 2 )2 (1 − q 4 )2

The series in curly brackets is the function called by Ramanujan† Φ1,6 and‡

1008Φ1,6 = Q2 − P R,

where
12η1 ω1  ω 4
1
 ω 6
1
P = , Q = 12g2 , R = 216g3 .
π2 π π
Here R = 0, so that

17 q 2 27 q 4
 
1008Φ1,6 = Q2 = 1 + 480Φ0,7 § = 1 + 480 + + · · · .
1 − q2 1 − q4

Hence we find that


A = i/πC, 2πiA = −2/C,
where
17 27
 
(6.3) C = 1 + 480 + + ··· .
e2π − 1 e4π − 1
Another expression for C is
 8
K0
(6.4) C = 144 ,
π
where
π/2
dθ {Γ(1/4)}2
Z
(6.41) K0 = q = √ .
(1 − 21 sin2 θ)
0 4 π
We have still to consider more closely the values of a and b, over which the summation is
effected. Let us fix k, and suppose that (a, b) is a pair of positive solutions of the equation
a2 + b2 = k. This pair gives rise to a system of eight solutions, viz.,

(±a, ±b), (±b, ±a).

But it is obvious that, if we change the signs of both a and b, we do not affect the aggregate
of values of a. Thus we need only consider the four pairs

(a, b), (a, −b), (b, a), (b, −a).



It will be observed that in this case α is on the boundary of P ; see the concluding remarks of §4. As it
happens, τ = i lies on that edge of P (the circular edge) which was not used in the construction of Hm , so
that our analysis is applicable as it stands.

S. Ramanujan, “On Certain Arithmetical Functions,” Trans. Camb. Phil. Soc., Vol. XXII, pp. 159 –
184 (p. 163) [No. 18 of this volume, p. 179].

Ramanujan, loc. cit., p. 164 [p. 181].
§
Ramanujan, loc. cit., p. 163 [p. 180].
On the coefficients in the expansions of certain modular functions 389

If a or b is zero, or if a = b, these four pairs reduce to two.


It is easily verified that, if (a, b) leads to the pair of poles
 
ac + bd π
q = ±q = ± exp πi − 2 ,
a2 + b2 a + b2

then (a, −b) and (b, a) each lead to q = ±q̄, where q̄ is the conjugate of q. Thus, in general
(a, b) and the solutions derived from it lead to four distinct poles, viz., ±q and ±q̄. These
four reduce to two in two cases, when q is real, so that q = q̄, and when q is purely
imaginary, so that q = −q̄. It is easy to see that the first case can occur only when k = 1,
and the second when k = 2∗ .
If k = 1 we take a = 1, b = 0, c = 0, d = 1; and q = q̄ = e−π . If k = 2 we take
a = 1, b = 1, c = 0, d = 1; and q = −q̄ = ie−π/2 . The corresponding terms in our series are
1 1
, .
1 − q 2 e2π 24 (1 + qeπ )

If k > 2, and is a sum of two coprime squares a2 and b2 , it gives rise to terms
1 1 1 1
8 2
+ 8
.
(a + bi) 1 − (q/q) (a − bi) 1 − (q/q̄)2
There is, of course, a similar pair of terms corresponding to every other distinct represen-
tation of k as a sum of coprime squares. Thus finally we obtain the following result:
Theorem 3. If

π6 1 X
f (q) = 6 = = pn q 2n ,
216ω1 g3 1 − 504 ∞
P r 5 q 2r
0
1 1−q 2r

and |q| < 1, then


1 1 1
(6.5) Cf (q) = +
2 1 − q 2 e2π 24 (1
+ q 2 eπ )
 
X 1 1 1 1
+ + ;
(a + bi)8 1 − (q/q)2 (a − bi)8 1 − (q/q̄)2

When a and b are given, we can always choose c and d so that |ac + bd| ≤ 12 (a2 + b2 ). If q is real, we
have ad − bc = 1 and ac + bd = 0 simultaneously; whence

(a2 + b2 )(c2 + d2 ) = 1.

If q is purely imaginary, we have

ad − bc = 1, 2|ac + bd| = a2 + b2 ,

whence
(c2 + d2 )2 = (|ac + bd| − c2 − d2 )2 + 1.
This is possible only if c2 + d2 = 1 and |ac + bd| = 1, whence a2 + b2 = 2.
390 Paper 37

where
17 27 9π 4
 
C = 1 + 480 2π + 4π + ··· = ,
e −1 e −1 16{Γ(3/4)}16
   
c + di ac + bd π
q = exp πi = exp πi − ,
a + bi a 2 + b2 a2 + b2
and q̄ is the conjugate of q. The summation applies to every pair of coprime positive
numbers a and b, such that k = a2 + b2 ≥ 5, such pairs, however, only being counted
as distinct if they correspond to independent representations of k as a sum of squares. If
|q| > 1, then the sum of the series on the right-hand side of (6.5) is zero.

7. It follows that
(−1)n nπ X
 
1 1 1 X cλ (n)
(7.1) Cpn = e2nπ + e + q−2n + q̄−2n = e2nπ/λ ,
2 24 (a + bi)8 (a − bi)8 λ4
(λ)
say. Here λ is the sum of two coprime squares, so that
λ = 2a2 5a5 13a13 17a17 · · · ,
where a2 is 0 or 1 and 5, 13, 17, . . . are the primes of the form 4k + 1; and the first few
values of cλ (n) are
c1 (n) = 1, c2 (n) = (−1)n , c5 (n) = 2 cos 45 nπ + 8 arctan 2 ,


c10 (n) = 2 cos 35 nπ − 8 arctan 2 , c13 (n) = 2 cos 13 10


 
nπ + 8 arctan 5 .
The approximations to the coefficients given by the formula (7.1) are exceedingly remark-
able. Dividing by 21 C, and taking n = 0, 1, 2, 3, 6, and 12, we find the following results:

(0) 0.944 (1) 505.361 (2) 270616.406


+0.059 −1.365 +31.585
−0.003 +0.004 +0.009
p0 = 1.000 p1 = 504.000 p2 = 270648.000
(3) 144912827.002 (6) 22251789962592450.237
−730.900 +9057051.688
−0.101 +2.081
−0.001 −0.006
p3 = 144912096.000 p6 = 22251789971649504.000
(12) 524663917940510190119197271938395.329
+1390736872662028.140
+2680.418
+0.130
−0.014
−0.003
p12 = 524663917940510191509934144603104.000
On the coefficients in the expansions of certain modular functions 391

An alternative expression for C is

C = 962 e−8π/3 {(1 − e−4π )(1 − e−8π ) · · ·}16 ,

by means of which C may be calculated with great accuracy∗ . To five places we have
2/C = 0.94373, which is very nearly equal to 352/373 = 0.94370.
It is easy to see directly that pn lies between the coefficients of xn in the expansions of
1 1 − 7.5x
, ,
(1 − 535x)(1 + 31x) (1 − 535.5x)(1 + 24x)

and so that
(535)n+1 − (−31)n+1 352(535.5)n + 21(−24)n
≤ pn ≤ .
566 373
The function
X cλ (x)
Ω(x) = e2xπ/λ
λ4
(λ)

has very remarkable properties. It is an integral function of x, whose maximum modulus is


less than a constant multiple of e2π|x| . It is equal to pn , an integer, when x = n, a positive
integer; and to zero when x = −n. But we must reserve the discussion of these peculiarities
for some other occasion.


Gauss, Werke, Vol. III, pp. 418 – 419, gives the values of various powers of e−π to a large number of
figures.

You might also like