Padic Current
Padic Current
Unreal Analysis
October 8, 2024
In memory of Kurt Hensel, Reinhold Strassmann, and the
countless others persecuted or murdered by the National Socialists.
Preface
A few months ago, on his last visit to In the long history of mathematics,
New Jersey, I was telling [Paul] Erdős ‘number’ has always meant real
something about p-adic analysis. number, and it is only relatively
Erdős was not interested. “You know,” recently that we have become aware
he said about the p-adic numbers, of the world of p-adic numbers. The
“they don’t really exist.” situation is akin to someone who has
only experienced daylight gazing in
Melvyn Nathanson astonishment at the night sky.
Kazuya Kato
Nobushige Kurokawa
Takeshi Saito
p-adic numbers were birthed into the world by Kurt Hensel in 1897, at a lecture
for the annual meeting of the German Mathematical Society. Initially, these
new objects were viewed with some skepticism. Helmut Hasse recollects that in
1920, no less a figure than Richard Courant advised him against studying with
Hensel, dismissing Hensel’s 1913 book on p-adic numbers as an unproductive
detour (“unfruchtbarer Seitenweg”). Yet today p-adic numbers have thoroughly
permeated number theory. Local-global principles, which relate solubility in
number fields to solubility in p-adic fields, are central to modern arithmetic
geometry. Tate’s thesis and the adelic viewpoint on class field theory are bread
and butter to algebraic number theorists. And those on the analytic side of
the fence can point to ‘local densities’ as p-adically motivated objects that
arise in nearly every study of arithmetic counting problems. We live in an age
when activists and arithmeticians universally agree on the need to think (and
act) both globally and locally.
These problem sets, written for a topics course at the 2024 Ross Indiana
summer mathematics camp, do not describe any of these significant modern
applications of p-adic numbers. Rather, they offer a hands-on — better, minds-
on! — approach to the foundational theory of Qp , working towards elementary
but attractive number-theoretic applications. Whenever there was a choice
to be made, I have elected to treat special cases rather than develop general
theory. This has allowed prerequisites to be kept to a minimum. Readers
I
are expected to have seen number theory and rigorous calculus, as well as
both abstract and matrix algebra, but most of what is needed belongs to the
standard undergraduate curriculum.
Notwithstanding the chosen title, these problem sets aim to offer a general in-
troduction to p-adic numbers, privileging neither analysis nor algebra. Analysts
will be disappointed by the absence of material on continuity, differentiability,
and p-adic interpolation. Algebraists will be disheartened that the discussion
never ventures beyond Qp to any of its finite or infinite extensions. Despite
the many omissions, I am optimistic that as individuals engage with these
problem sets, they will catch a glimpse of an alluring territory begging to
be charted. The suggested readings listed below have been selected as partic-
ularly approachable, and we commend their consideration to those with an
adventurous spirit.
II
I should emphasize that this is not meant to be your only book on p-adic
numbers. The author, for instance, learned about p-adic numbers from entirely
different sources! After mastering the material on each problem set, it is
recommended that you look at how the corresponding content is treated in
the suggested references. One only truly understands a topic after examining
it from all angles, and different texts bring out different perspectives.
Exercises are typically introduced as assertions, without the use of “show that”
or “prove.” (The Extra Explorations, embedded in the solutions, are exceptions.)
Problem solvers should recognize they are on the hook to validate every claim.
We take no position on the hotly contested issue of whether 0 is a natural
number. Integers greater than or equal to 0 are — naturally enough — referred
to as nonnegative integers. As is customary in English, positive means strictly
larger than zero. We write Z+ , Q+ , and R+ for the sets of positive integers,
rational numbers, and real numbers, respectively.
Our terminology around quotient rings is slightly unconventional. To start
with, the term ring always refers to a commutative ring with unity. If I is an
ideal of the ring R, and a ∈ R, the class of a with respect to congruence modulo
I is typically denoted “a mod I.” We break this rule when (and only when)
R = Z and I = mZ, instead writing “a mod m.” Additionally, we use “Z/m”
in place of “Z/mZ”; this is a mild concession to the notation Zm appearing
on the first-year Ross Program problem sets.
Acknowledgements
Assembling these problem sets involved much scouring and borrowing. De-
serving of particular mention: Exercise 4.53 was sourced from an article by
Catherine Crompton in the Rose Hulman Undergrad. Mathematics Journal.
Exercises 7.90 and 9.117 were inspired by discussions in Murty’s text [6].
Exercise 11.132 was taken from Koblitz’s book [5]. The proof of Ramanujan’s
conjecture on integer solutions to x2 + 7 = 2m follows closely the exposition in
Cassels’s monograph [1]. The proof “from-the-Book” of Skolem’s theorem on
integer linear recurrences, outlined on Set #11, is due to Tao (loosely based
on an argument of Georges Hansel):
https://terrytao.wordpress.com/2007/05/25/open-question-e
ffective-skolem-mahler-lech-theorem/
In this connection it was also helpful to consult a post of Seewoo Lee:
https://seewoo5.github.io/jekyll/update/2023/02/21/p-adi
c-numbers-application.html
III
Exercises 9.112 and 10.129 are based on a paper of Mahler:
On some irrational decimal fractions. J. Number Theory 13 (1981),
268–269.
The proofs on Set #13 of the Adams and Kummer theorems on Bernoulli
numbers are adapted from a delightful article of Wells Johnson:
p-adic proofs of congruences for the Bernoulli numbers. J. Number
Theory 7 (1975), 251–265.
Various tech tools were employed to produce the book in front of you. The
cover, which features 19th-century illustrations from the Hirayama Fireworks
company∗ , was designed in Canva. Typesetting was done with LATEX, using
Springer’s SVMono document class. Several calculations were farmed out to
the fantastically capable PARI/GP. ChatGPT assisted by offering English
translations, hunting down typos, and suggesting alternative phrasings.
This manuscript would not exist without the backing and encouragement
of the Ross Mathematics Foundation. Site directors Timothy All and Jim
Fowler have my profound appreciation for their inspiring efforts, year after
year, to ensure that the Ross Program provides a supportive, welcoming, and
stimulating environment for all camp participants (students, counselors, and
even us lecturers!). Special thanks to Phoebe Watkins for her service as course
assistant, Paco Adajar for suggesting the term “p-set” as an alternative to the
more mundane “problem set,” and Jacob Bucciarelli for sharing his expertise
in orbital mechanics.
Work on this project was facilitated by a grant from the US National Science
Foundation, award DMS-2001581.
IV
7. W. H. Schikhof, Ultrametric calculus: An introduction to p-adic analysis, Cam-
bridge Studies in Advanced Mathematics, vol. 4, Cambridge University Press,
Cambridge, 2006.
8. J.-P. Serre, A course in arithmetic, Graduate Texts in Mathematics, vol. 7,
Springer-Verlag, New York-Heidelberg, 1973.
9. J. Steuding, Die p-adischen Zahlen. Online survey article. URL: https://www.
uni-marburg.de/de/fb12/fachbereich/profil/geschichte-des-fachberei
chs/biographisches/hensel_p_adischen_zahlen.pdf
V
VI
Part I
Problems
1
2
Problem Set #1
William R. Inge
1.2. Let K be a field. Define |x| by letting |x| = 0 if x = 0 and |x| = 1 for all
x ̸= 0. Show that | · | is an absolute value on K (the trivial absolute value).
1.3. Let p be a prime. For each x ∈ Q, define |x|p = p−vp (x) , where p−∞ is
taken to be 0. Then | · |p is an absolute value on Q (the p-adic absolute value).
3
The absolute value | · |p will play a starring role throughout the course. You
are strongly advised to compute several examples to develop a feel for this
notion of absolute value. Here are a few to get you started:
5 · 210 2
= 2−10 , |3−5 |2 = 1, = 3, |100!|7 = 7−16 .
210 + 1 2 21 3
1.4. For each prime p, the absolute value | · |p on Q defined in Exercise 1.3 is
non-Archimedean.
1.6. Let K be a field equipped with an absolute value | · | for which |2| ≤ 1.
Then |2e1 + 2e2 + · · · + 2en | ≤ n for all nonnegative integers e1 , . . . , en . It
follows that | nk | ≤ n whenever n is a positive integer and 0 ≤ k ≤ n.
1.7 (Product Formula). Let |·|∞ denote the standard Archimedean absolute
value on Q. For every x ∈ Q× ,
Y
|x|∞ |x|p = 1.
p prime
1.9. |n!|∞ grows faster than C n for any fixed C. Hence, the product formula
implies that there are infinitely many primes.
4
‘p’s and Harmony
1.10. Let Hn = 1 + 12 + · · · + n1 , the nth harmonic number. From calculus,
|Hn |∞ tends to infinity. The same holds for |Hn |2 .
1
P
1.11 (Wolstenholme). When p is odd, Hp−1 = p 0<i<p/2 i(p−i) . Hence,
|Hp−1 |p ≤ p−2 for p > 3.
5
6
p-Set #2
Distance Learning
Kurt Hensel
Let (K, | · |) be a valued field. For x0 ∈ K and real r > 0, the open and closed
discs of radius r centered at x0 are defined, respectively, by
D<r (x0 ) = {x ∈ K : |x − x0 | < r} and D≤r (x0 ) = {x ∈ K : |x − x0 | ≤ r}.
2.14. Suppose | · | is non-Archimedean. Then any two open discs are either
disjoint or one contains the other.
7
Valuation Theory
2.16. How many absolute values can you find on F2027 ? (Here and below, Fp
denotes the field with p elements, or equivalently the residue ring Z/p.)
T
2.21. Determine p prime Z(p) .
8
Variations on a Theme of Euclid
2.24 (Schur). Let F (T ) ∈ Z[T ] be nonconstant. There are infinitely many
primes that divide F (n) for some n ∈ Z. In other words: F has a root in Z/p
for infinitely many primes p.
2.26. Let n ∈ Z+ . If p | (2n + 1)2 − 2, then p ≡ ±1 (mod 8). Not every prime
dividing (2n + 1)2 − 2 can be congruent to 1 (mod 8). Hence, (2n + 1)2 − 2 is
always divisible by some prime congruent to −1 (mod 8). By varying n, we
can find infinitely many primes p ≡ −1 (mod 8).
2.27. Every coprime residue class mod 8 contains infinitely many primes.
In this last statement, “mod 8” can be replaced with “mod m,” for any m ∈ Z+ . This is a
celebrated (and difficult!) theorem of Dirichlet that you will meet in courses on analytic number
theory.
9
10
p-Set #3
−1 = 4 + 4 · 5 + 4 · 52 + 4 · 53 + . . . .
For every real number ϵ > 0, there is an N ∈ Z+ with the property that
(The inequality “|xn − x| < ϵ” could have been written as “xn ∈ D<ϵ (x).”)
We say {xn } converges if it converges to some x in K. Those who already
grok convergence in the context of calculus will notice that xn → x in (K, | · |)
precisely when |xn − x| → 0 in the familiar sense.
11
(i) If xn = x for all n, then xn → x.
(ii) If xn → x and yn → y, then xn + yn → x + y.
(iii) If xn → x and yn → y, then xn yn → xy.
3.30. xn = 1 + 3 + 32 + · · · + 3n converges to − 12 in
P∞ (Q, | · |3 ). So defining the
value of a series as the limit of its partial sums, k=0 3k = − 12 . Does {xn }
converge in (Q, | · |p ) for any other values of p?
P∞ P∞
3.31. Evaluate n=0 n · n! and n=0 n2 · 2n in (Q, | · |2 ).
Valuation Theory
3.34. Let (K, | · |) be a valued field. Then | · | is non-Archimedean ⇐⇒ |2| ≤ 1.
(Here 2 means 1 + 1.) Is this equivalence true with 3 in place of 2?
12
m 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
um 0 1 1 −1 −3 −1 5 7 −3 −17 −11 23 45 −1 −91 −89 93
3.39. Z(p) is a principal ideal domain (PID) with pZ(p) its only maximal ideal.
A Question of Ramanujan
When is a power of 2 equal to 7 more than a square? This happens for
23 , 24 , 25 , 27 , and 215 (in this last case, 215 = 32 768 = 7 + 1812 ). In a 1913
issue of the Journal of the Indian Mathematical Society, Ramanujan listed all
of these examples and posed the problem of finding others.
√ √ √
Let R = Z[ 1+ 2 −7 ] (viewed as a subring of C). Since ( 1+ 2 −7 )2 = 1+ −7
2 − 2,
√
√
1+ −7 1
R=Z+Z = (a + b −7) : a, b ∈ Z and a ≡ b (mod 2) .
2 2
On our final problem set (Set #13) you will determine all solutions to um−2 = ±1 by p-adic
methods.
13
14
p-Set #4
Berning Questions
91 409 924 241 424 243 424 241 924 242 500.
Jacob Bernoulli
T2 T3 T4 T2 T3
T = T+ + + + ... B0 + B1 T + B2 + B3 + ... .
2! 3! 4! 2! 3!
T 2
Since e T−1 = 1 + 2!
T
+ T3! + . . . is a power series with rational coefficients and
nonzero constant term, its reciprocal eTT−1 is also a power series with rational
coefficients. Therefore, every Bk ∈ Q.
For positive integers n and k, define Sk (n) = 1k + 2k + · · · + (n − 1)k .
∗
We assume acquaintance with basic facts about formal power series, as found for
instance in §1.1 of Stanley’s Enumerative Combinatorics.
15
k 0 1 2 3 4 5 6 7 8 9 10 11 12
Bk 1 − 12 1
6
0 1
− 30 0 1
42
0 1
− 30 0 5
66
0 691
− 2730
P∞ k
4.41. 1 + eT + e2T + · · · + e(n−1)T = k=0 Sk (n) Tk! (as formal power series).
k
X k nj+1
4.42 (Faulhaber’s Formula). Sk (n) = Bk−j .
j=0
j j+1
2T P∞ k
4.43. If coth T := ee2T −1
+1
, then T coth T = T + e2T2T−1 = T + k=0 Bk (2Tk!) (as
formal series). T coth(T ) is invariant under the substitution T 7→ −T . Hence,
Bk = 0 for every odd k > 1.
22k
4.44. coth T = T1 + k≥1 B2k (2k)! d
T 2k−1 and dT coth T = 1 − coth2 T .
P
Other properties of Bernoulli numbers lie a bit further below the surface.
For instance, based on the above table one might conjecture that Bernoulli
numbers always have squarefree denominators. We will prove this — and much
more — in due time!
4.47. If xn → x in (K, | · |), then |xn | → |x| in the real numbers. In fact, if | · |
is non-Archimedean and xn → x, where x ̸= 0, then |xn | = |x| for all large n.
16
Valuation Theory
4.48. Every nontrivial non-Archimedean absolute value on Q has the form
| · |cp for some prime p and some c > 0.
e.g., B = |2|·|3| m n
|3|−1 works. Since 3 ≤ 2 and n can be taken arbitrarily large, it
must be that c ≤ d. Reversing the roles of 2 and 3 shows d ≤ c, and so c = d.
17
18
p-Set #5
And Introducing. . .
Let p be a prime number. Q∞A p-adic integer is an infinite tuple (a1 mod p, a2 mod
p2 , a3 mod p3 , . . . ) ∈ k=1 Z/pk that satisfies the compatibility condition
Zp Z/pk+1
Z/pk
Take a moment to convince yourself that this last example satisfies our compatibility
condition!
Q∞
5.56. Zp is a subring of k=1 Z/pk .
19
Figure for Exercise 5.64.
5.61. Zp is a principal ideal domain with pZp its only nonzero prime ideal.
5.64. What does the picture on this page — created by TEX StackExchange
user Qrrbrbirlbel∗ — have to do with Z3 ?
5.65 (Steuding). Repeat the last exercise for Salvador Dali’s painting La
Cara de la Guerra.
20
Well, Color Me Impressed!
5.67 (Thomas, Monsky). Consider the following 3-coloring of the rational
plane Q2 : (x, y) is red if |x|2 < 1, |y|2 < 1, blue if |x|2 ≥ 1, |x|2 ≥ |y|2 , and
green if |y|2 ≥ 1, |y|2 > |x|2 . Show that any trio of differently colored points
forms a triangle ∆ with |Area(∆)|2 > 1.
This observation plays a key role in the proof of Monsky’s Theorem: It is impossible to dissect a
square into an odd number of triangles all of which have the same area.
21
22
p-Set #6
Enter Cauchy
Let K be a field equipped with an absolute value | · |. If {xn } converges to x
in (K, | · |), then for any real number ϵ > 0, all terms far enough out in the
sequence {xn } are within 12 ϵ of x. By the triangle inequality, all such terms
are within ϵ of each other. That is:
n
6.69. The sequence {25 } is Cauchy in (Q, | · |5 ).
If R is equipped with the usual absolute value, then every Cauchy sequence
in R converges to an element of R. This need not be the case for a general
valued field. For example, the sequence of rational numbers
23
x1 = 1, x2 = 1.4, x3 = 1.41, x4 = 1.414,
...,
√
obtained by successively truncating the decimal expansion of √2, is Cauchy in
(Q, | · |∞ ) but does not converge to any element of Q, since 2 ∈ / Q. Coping
with this unsettling scenario was one of the motivations behind the invention
(discovery?) of the real numbers in the first place!
Disconcerting examples of this same kind can also be found when K = Q
and | · | = | · |p . Take any sequenceP{ck }k≥0 from {0, 1, . . . , p − 1} that is
n
not eventually periodic. Then xn = k=0 ck pk defines a Cauchy sequence in
(Q, | · |p ). To check the Cauchy condition (C), we may assume that m < n
(why?). Then
p−n Zp = Zp [1/p].
S
6.72. Qp = n≥0
6.74. When x ∈ Q, we have defined vp (x) twice: once on Set #1 and again
just now, since x is also an element of Qp .
Check that when x ∈ Q our two definitions of vp (x) agree (so our sin is venial,
rather than mortal). Furthermore, if we set |x|p = p−vp (x) for x ∈ Qp , then
| · |p defines a non-Archimedean absolute value on Qp . (We continue to call vp
the p-adic valuation and | · |p the p-adic absolute value.)
24
6.75. Zp = {x ∈ Qp : |x|p ≤ 1}. (That is, Zp = D≤1 (0) in Qp .)
Also, Z×
p = {x ∈ Qp : |x|p = 1}.
Curious Congruences
1
6.77 (Stern). Recall that log 1−t = t + 12 t2 + 13 t3 + . . . whenever t is a complex
2 3
number with |t| < 1 (usual absolute value). Exponentiating, eT eT /2 eT /3 · · · =
1 p
1−T , as formal power series. Expanding and comparing coefficients of T shows
1
that |1 − p! − p1 |p ≤ 1. Hence, (p − 1)! ≡ −1 (mod p) (Wilson’s theorem).
ap−1 −1
6.78. For each prime p and each a ∈ Z coprime to p, put qp (a) = p . By
Fermat’s little theorem, qp (a) ∈ Z.
Prove: If a and b are both coprime to p, then qp (ab) ≡ qp (a) + qp (b) (mod p).
25
26
p-Set #7
You Complete Me
Kazuya Kato
7.83. Q is dense in Qp .
Suppose K and L are fields equipped with absolute values and that the absolute
value | · | on L extends the absolute value on K. If (L, | · |) is complete and K
∗
Disclaimer: “Everything in the world” means every occurrence of a sequence being
Cauchy.
27
is a dense subset of L, we call L the completion of K with respect to | · |. For
example, R is the completion of Q with respect to | · |∞ . By Exercises 7.80
and 7.83, Qp is the completion of Q with respect to | · |p .
Why do we say the completion and not a completion? Completions are unique,
up to isometric (absolute-value preserving) isomorphism.
7.84. If (L, | · |) and (L′ , | · |′ ) are two completions of the same valued field
(K, | · |0 ), then there is an isomorphism ϕ : L → L′ that fixes K and satisfies
|ϕ(x)|′ = |x| for all x ∈ L.
The parenthesized terms in the numerator are the (least nonnegative integer)
reductions of x modulo pk+1 Zp and modulo pk Zp . (Make sure you see why!)
So we’ve determined the digits ck in any possible expansion of x. Now that we
know to try these digits, we are home free, as you are asked to check in the
next exercise.
x = c0 + c1 p + c2 p2 + c3 p3 + . . . .
28
Thus, the (finite) base p expansions of the xk coalesce to an infinite base p expansion of x. The
representation produced in this way is not merely an infinite base p expansion of x, but — as the
lead-in to the problem establishes — its unique base p expansion.
Curious Congruences
p−1 p−1
X a −1 (p − 1)! + 1
7.90. For every odd prime p: ≡ (mod p).
a=1
p p
3 + 1 · 7 + 2 · 72 + 6 · 73 + . . . .
29
After that wall of text, you might be wondering what exactly you are being asked to do. Your
job: Check that the process can be continued indefinitely, uniquely determining all the xk , that
setting x = (x1 , x2 , x3 , . . . ) really does define a Z7 -solution to x2 = 2, and finally, that the 7-adic
expansion of x starts the way we claimed.
30
p-Set #8
8.93. Let P
K be a field complete with respect to a non-Archimedean absolute
∞
value. If k=1 akP is a series in K that converges to s ∈ K, then every
∞
rearrangement of k=1 ak also converges to s.
P∞ P∞
8.94. Suppose that n=0 an and n=0 bn are convergent series in the field
K, which is assumed
Pncomplete with respect to a non-ArchimedeanP∞ absolute
ak bn−k , for
value. DefinePcn = k=0P n = 0, 1, 2, . . . . Then n=0 c n converges
∞ ∞ P∞
and in fact n=0 cn = ( n=0 an )( n=0 bn ).
P∞ P∞
Important consequence: If F (T ) = k=0 ak T k , G(T ) = k=0 bk T k ∈ K[[T ]] both converge at
the point z ∈ K, so does F (T )G(T ), and F (T )G(T )|T =z = F (z)G(z).
31
8.96 (a double switch). Continue with the notation and assumptions of
Exercise 8.95. For every nonnegative integer N ,
XX N X
X XX N
XX
ai,j − ai,j ≤ ϵN +1 , ai,j − ai,j ≤ ϵN +1 ,
i j i=0 j j i j i=0
N X
X N X
X N N
XX N X
X N
ai,j − ai,j ≤ ϵN +1 , ai,j − ai,j ≤ ϵN +1 .
i=0 j i=0 j=0 j i=0 j=0 i=0
P P P P
Therefore, i j ai,j = j i ai,j .
1C denotes the indicator function of the condition C. Thus, 1p−1|k is 1 if p − 1 divides k and 0
otherwise.
1p−1|k j
k p
8.98. For p prime, k ∈ Z+ : Bk +
P
p + 0<j<k j Bk−j j+1 ∈ Z(p) .
1p−1|k
8.99. For p an odd prime, k ∈ Z+ : Bk + p ∈ Z(p) .
1
8.100. For k ∈ 2Z+ : Bk + 2 ∈ Z(2) .
X 1
8.101 (Clausen, von Staudt). For k ∈ 2Z+ : Bk + ∈ Z.
p
p prime
p−1|k
It follows from Exercise 8.101 that the denominator of Bk is the (squarefree!) product of the
primes p for which p − 1 | k.
8.103. Find ω(2) (exactly) when p = 3. Then take p = 7 and determine the
digits c0 , c1 , c2 in the canonical expansion ω(2) = c0 + c1 · 7 + c2 · 72 + . . . .
32
p-Set #9
9.106. Let p be odd, and let n ∈ Z be a nonsquare mod p. Then 1, n are coset
representatives for Z× × 2
p /(Zp ) , while 1, n, p, np are coset representatives for
× × 2
Qp /(Qp ) .
9.107. If a ∈ Z×
2 , then a is a square in Q2 ⇐⇒ a ≡ 1 (mod 8Z2 ).
9.109. For a ∈ Q×
p:
√
a ∈ Z×
p ⇐⇒ (some value of)
n
a ∈ Qp for infinitely many n ∈ Z+ .
33
9.110 (a Liouville approximation theorem in Zp ). Suppose α ∈ Zp is
a root of a polynomial F (T ) ∈ Z[T ] of degree d having no integer roots. For
every nonzero n ∈ Z,
Strassmann Series
an T n be a formal power series with Qp -coefficients.
P
Let F (T ) = n≥0
k
P
9.115 (recentering Strassmann series). Let F (T ) = k≥0 ak T be a
Strassmann power series, and let x0 ∈ Zp . For every x ∈ Zp ,
∗
Here we depart from convention; the usual terms are restricted power series or strictly
convergent power series.
34
X
j
X k k−j
F (x + x0 ) = bj x , where bj := ak x .
j 0
j≥0 k≥j
j
P
The recentered series j≥0 bj T is also Strassmann.
Curious Congruences
9.117 (Glaisher). For all primes p: p2 | (p−1)!+1 ⇐⇒ Bp−1 + p1 −1 ∈ pZ(p) .
Primes p for which p2 | (p − 1)! + 1 are known as Wilson primes. The only known examples, and
the only examples smaller than 2 · 1013 , are 5, 13, and 563.
9.118 (Johnson). Recall that when u ∈ Z× p , we are writing ω(u) for the
unique (p − 1)th root of unity congruent to u modulo pZp (see Exercise 8.102).
Show that if u ∈ Z has order 3 modulo p, and v ∈ Z satisfies v ≡ ω(u)
(mod pk Zp ), where k ∈ Z+ , then
u ω(u)
2 2 + 4 · 7 + 6 · 72 + 3 · 73 + 2 · 75 + . . .
3 3 + 4 · 7 + 6 · 72 + 3 · 73 + 2 · 75 + . . .
4 4 + 2 · 7 + 3 · 73 + 6 · 74 + 4 · 75 + . . .
5 5 + 2 · 7 + 3 · 73 + 6 · 74 + 4 · 75 + . . .
Sixth roots of unity in Q7 , omitting ω(±1) = ±1. Notice that ω(3) = 1 + ω(2)
and ω(5) = 1 + ω(4), as guaranteed by Problem 8.104.
35
36
p-Set #10
Edward B. Burger
Thomas Struppeck
To fashion a proof, suppose x is a real number with |x| < 1. Then B 12 (x)
converges absolutely (e.g., by the ratio test). So if we multiply B 12 (x) by
B 12 (x), we can reshuffle the terms as we please. One such regrouping gives us
C(x). On the other hand, our “real world” identity says that B 12 (x)2 = 1 + x
whenever |x| < 1. It follows that C(x) − (1 + x) = 0 when |x| < 1. But a power
series that vanishes on an open interval around 0 has all its coefficients equal
to 0. This forces C(T ) = 1 + T , as formal series.
37
Armed with this formal identity, we can head back to Qp . Suppose we have in
hand an x ∈ Qp for which B 21 (x) converges. If we multiply B 12 (x) by B 12 (x),
we can rearrange the result into C(x) — this time justifying ourselves not on
the basis of absolute convergence but by an appeal to Problem 8.94. Since
C(x) = 1 + x, this shows that B 21 (x) represents a square root of 1 + x in Qp
whenever B 21 (x) converges.
10.119. If p is odd, then B 21 (x) converges when |x|p ≤ 1/p. If p = 2, then B 12 (x)
converges when |x|2 ≤ 1/23 . Are these conditions necessary for convergence?
1
9 9 k 5
in R but to − 54 in Q3 .
P
10.120. B 12 ( 16 )= k≥0
2
k
( 16 ) converges to 4
X F (j) (a)
F (a + T ) = T j.
j!
j≥0
1 (j)
Furthermore: If F (T ) ∈ Zp [T ], so is j! F (T ), for all nonnegative integers j.
38
Don’t Ever Change
10.127. The only ring homomorphism from Qp to Qp is the identity.
There is a field embedding of Qp into C (assuming the Axiom of Choice). In fact, there are
many such embeddings, but none are canonical, and none carry convergent sequences in Qp to
convergent sequences in C. Thus (channeling Hermann Weyl∗ ), choosing such an embedding must
always be regarded as something of a brute act.
∗
Weyl famously wrote “The introduction of numbers as coordinates . . . is an act of
violence.”
39
40
p-Set #11
Passing the familiar identity log xy = log x + log y (valid for x, y ∈ R+ ) from
R to Qp [[X − 1, Y − 1]], and then on to Qp , one can show that
41
Filling in the details here is a bit finicky. If you enjoy this kind of work (you
know who you are. . . ), give it a try!
X (ap)2k+1 X (ap)2k+1
p−1 = | − ap|p = (−1)k − ap = (−1)k ≤ p−2 .
(2k + 1)! (2k + 1)!
k≥0 k≥1
p p
Contradiction!
Strassmann Series
11.133. Let F (T ) = 1 + T + (pT )2 + (pT )4 + (pT )8 + (pT )16 + . . . . There is
exactly one x ∈ Zp with F (x) = 0.
It will be convenient for the next exercise, and certain others afterward, to
name the coefficients of the falling factorial T (T − 1) · · · (T − (N − 1)). For
nonnegative integers N and K, we let s(N, K) be the integer defined by the
formal equality
∞
X
T (T − 1) · · · (T − (N − 1)) = s(N, K)T K .
K=0
∗
“Ordinary analysis has amassed a great stock of identities between power series. Many
of these are valid in p-adic analysis too. But here an identity between power series
yields congruences between partial sums.” — Max Zorn
42
For example, when N = 5, we have T (T − 1)(T − 2)(T − 3)(T − 4) = 24T −
50T 2 + 35T 3 − 10T 4 + T 5 , so that
where
X ak
Ca,j := s(k, j) .
k!
k≥j
Moreover, |Ca,j |p → 0 as j → ∞.
Your job: Fill in the missing details. In particular, justify the swapping of the sums on k and j.
43
0 1 0 ... 0 x0 1
0 0 1 ... 0 x1
0
.. . .. .. .. x2
A=. .. and v= , e = 0.
. . . .. ..
0 0 0 ... 1 . .
ad ad−1 ad−2 . . . a1 xd−1 0
11.139. For each fixed r ∈ {0, 1, . . . , k − 1}, either xn = 0 for all nonnegative
integers n ≡ r (mod k), or xn = 0 for only finitely many n ≡ r (mod k).
Hence: The set of n with xn = 0 is the union of a finite set and a finite
collection of residue classes. (Skolem’s Theorem)
11.141 (Mahler). Does Skolem’s Theorem hold for recurrence sequences over
Q? (“Over Q” means that a1 , . . . , ad and x0 , . . . , xd−1 belong to Q.) Over an
arbitrary finite extension of Q (number field)?
44
p-Set #12
Strassmann Series
12.142. Let p be an odd prime and let a ∈ pZp . The identity
Binom(1 + a; n) = (1 + a)n
proved in Exercise 11.134 for nonnegative integers n holds for all n ∈ Z, since
m
(1 + a)n = lim (1 + a)n+p = lim Binom(1 + a; n + pm ) = Binom(1 + a; n).
m→∞ m→∞
45
For each α ∈ K, the norm Nα of α is defined by Nα = αα′ α′′ . By a tedious
but straightforward calculation,
Hence, ε ∈ Z[θ]× ∩ R+ = U.
Summarizing: U = {ε ∈ Z[θ] : Nε = 1}.
12.145. For each real number R > 0, there are finitely many α ∈ Z[θ] with
|α|, |α′ | ≤ R. (Here and in Problem 12.146, | · | is the usual real/complex
absolute value.)
46
x+y =0
(i) G = {0},
(ii) G is infinite cyclic,
(iii) some (Euclidean) disc centered at 0 has infinite intersection with G.
D µ D µ
2
2 1+θ+θ 7 4 + 2θ + θ2
3 4 + 3θ + 2θ2 9 4 + 2θ + θ2
4 5 + 3θ + 2θ2 10 181 + 84θ + 39θ2
5 41 + 24θ + 14θ2 11 89 + 40θ + 18θ2
6 109 + 60θ + 33θ2 12 9073 + 3963θ + 1731θ2
47
Theorem. Let D be a cubefree integer with D > 1. Then x3 − Dy 3 = 1 has
finitely many integer solutions x, y.
12.149. Let L = Q(θ, ω) (= Q(θ, θ′ , θ′′ )). Fix an odd prime p for which L
embeds into Qp . Then |θ|p = |ω|p = |µ|p = |µ′ |p = |µ′′ |p = 1. That is, all of
θ, ω, µ, µ′ , µ′′ belong to Z×
p.
Here we abuse notation slightly, using the same symbols for elements of L and corresponding
elements of Qp under our embedding.
48
To deal with this, set ν := µp−1 , ν ′ := µ′p−1 , ν ′′ := µ′′p−1 . These belong to
1 + pZp by Fermat’s little theorem. Writing n = (p − 1)m + r,
For each fixed r, there is some m ∈ Z where the right-hand side is nonzero
(check this!). Hence, the RHS vanishes for only finitely many m ∈ Z (Exercise
12.143). The Theorem follows.
Much more is known. For instance, Delaunay and Nagell showed (independently) that x3 −Dy 3 = 1
has at most one integer solution ̸= (1, 0).
49
50
p-Set #13
Bk 1p−1|k (p−1)
13.153. Let βk = k − pk . For each k ∈ Z+ :
p−1
X k − 1 pj X k − 1 pj−1 X
βk + Bk−j + uk−j θ(u)j = 0.
j−1 j(j + 1) j−1 j u=1
0<j≤k 0<j≤k
Pp−1
13.155. Assume p ≥ 5. Then βk + u=1 uk−1 ϑ(u) ∈ pZp for each k ∈ 2Z+ .
51
B4 1 B10 1 B4 B10 1
For example, 4 = − 120 and 10 = 132 are congruent mod 7Z7 . In fact, 4 − 10 = 7 · − 440 .
Strassmann Series
aK = max |ak |p .
k≥0
52
√
√ quadratic field Q( −7) can be viewed as a subfield of Q11 , identifying
The
−7 with the square root of −7 in Z11 that is congruent to 2 modulo 11Z11 .
By hand, or with the aid of software such as PARI/GP, one computes that
√
−7 = 2 + 8 · 11 + 8 · 112 + 7 · 113 + 10 · 114 + 1 · 115 + . . . ,
n n
−β
We study the equation αα−β = ±1 by the method of Exercise 12.150. Put
10 10
A = α and B = β , so that A = 1 + a, B = 1 + b for
53
13.162. F3,− (T ) has a T -coefficient that vanishes mod 112 . Its T 2 -coefficient
is a multiple of 112 but not 113 . Its Strassmann degree is 2.
n n
−β
13.163. There are at most 3 · 1 + 2 = 5 integers n with αα−β = ±1. We know
five such integers — n = 1, 2, 3, 5, and 13 — so those must be all of them.
Remark. The choice to work in Qp for p = 11 was fortuitous. The next prime√after
11 for which −7 is a quadratic residue is p = 23. If we had elected to embed Q( −7)
into Q23 , we would have to bound the number of zeros of 2 · 22 = 44 Strassmann
22m+12
−β 22m+12
series. One of those would correspond to the equation α α−β
= −1. If you
approximate the coefficients sufficiently to apply Strassmann’s theorem, you’ll find
22m+12
−β 22m+12
that this series has at most one zero m ∈ Zp ; hence, α α−β
= −1 has at
most one integer solution m. But in fact (as we know after Exercise 13.163), there
are zero integers m satisfying the equation. So working in Q23 , we fail to rule out an
extraneous zero. What’s going on in this instance is that there really is an m ∈ Z23
where the associated power series vanishes — but this m does not belong to Z!
54
Part II
55
56
Solutions to Set #1
1.1
(a) First off, |1|2 = |1 · 1| = |1| and |1| > 0 (from (i)). Thus, |1| = 1. Next,
| − 1|2 = |(−1)(−1)| = |1| = 1. As | − 1| > 0, we conclude that | − 1| = 1.
(b) The proof is the same for the “standard” absolute value: By the triangle
inequality, |x| = |(x − y) + y| ≤ |x − y| + |y|. Now rearrange.
(c) Since |y −1 | · |y| = |y −1 · y| = |1| = 1, we have |y −1 | = |y|−1 . So |xy −1 | =
|x||y −1 | = |x||y|−1 , as claimed.
1.2 Property (i) in the absolute value definition is clear. Property (ii) is also
easy: When x + y = 0, the inequality is obvious. Otherwise, either x or y is
nonzero, so that |x| or |y| is 1. Hence, 1 = |x + y| ≤ |x| + |y|. To prove (iii),
take cases: If x and y are nonzero, both sides are 1, otherwise both sides are 0.
In this last step we use that fields are integral domains.
1.3 Property (i) is again clear. Of the remaining two properties, (iii) is
quicker to dispense with: If x or y is zero, both sides of (iii) vanish. Otherwise,
write x = pvp (x) ab and y = pvp (y) dc , where p does not divide any of a, b, c, d.
Then xy = pvp (x)+vp (y) ac
bd , and p does not divide either of ac or bd. Hence,
vp (xy) = vp (x) + vp (y) and |xy|p = p−vp (x) p−vp (y) = |x|p |y|p .
To prove (ii) we have to work a bit harder. If x, y, or x + y is zero, (ii)
is trivial. Otherwise, write x = pvp (x) ab and y = pvp (y) dc as above. The
symmetry of (ii) in x and y allows us to assume vp (x) ≤ vp (y). Then
x + y = pvp (x) ab + pvp (y)−vp (x) dc . Since pvp (y)−vp (x) ∈ Z, we can express
pvp (y)−vp (x) dc as a fraction with denominator d. Hence, ab + pvp (y)−vp (x) dc = bd
N
w ′
for some nonzero N ∈ Z. Write N = p N , where w is a nonnegative in-
′
teger and N ′ ∈ Z is not divisible by p. Then x + y = pvp (x)+w N bd , where
neither N ′ nor bd is divisible by p. Hence, vp (x + y) = vp (x) + w ≥ vp (x), and
|x + y|p = p−vp (x) p−w = |x|p p−w ≤ |x|p ≤ |x|p + |y|p .
57
1.5 We may assume without loss of generality that |x| > |y|. A direct
application of the strong triangle inequality gives |x + y| ≤ max{|x|, |y|} = |x|.
The strong triangle inequality also implies that |x| = |(x + y) + (−y)| ≤
max{|x+y|, |−y|} = max{|x+y|, |y|}. (We use in this last step that |−y| = |y|,
which follows from | − 1| = 1.) Since |x| > |y|, the maximum here cannot be
|y|. So it must be |x + y|, yielding |x| ≤ |x + y|. Hence, |x + y| = |x|.
1.6 Since |2| ≤ 1, we see that |2e1 +· · ·+2en | ≤ |2|e1 +· · ·+|2|en ≤ 1+· · ·+1 = n.
To conclude, notice that (a) every positive integer less than 2n is a sum
of
P at most n powers of 2 (e.g., use the binary representation), while (b)
n n
n
n
0≤k≤n k = 2 , so that k < 2 for every k.
1.7 Write x = ± ab where a and b are relatively prime positive integers. Since
a and b share no prime factors,
Y Y Y −vp (a) Y vp (b)
|x|p = |a|p |b|−1
p = p p = |a|−1
∞ |b|∞ ,
p prime p prime p|a p|b
Let P be a finite set of primes. Choose rational numbers xp for each prime p ∈ P,
alongside a rational number x∞ . For each ϵ > 0, there is a rational number x
satisfying
p−n/(p−1) .
58
1
(a) Show that vp (m) = p−1 (sp (m − 1) − sp (m) + 1) for every positive integer m.
1
Deduce that vp (n!) = p−1 (n−sp (n)) for all nonnegative integers n (an alternative
form of Legendre’s formula).
(b) Prove that n!/(−p)vp (n!) ≡ n0 !n1 !n2 !n3 ! · · · (mod p) for all nonnegative n ∈ Z.
(Anton [1], Stickelberger [5, pp. 342–343], Hensel [2])
1.9 We interpret “grows faster” to mean that n!/C n tends to infinity. To prove
that, we may (in fact, should!)
P assume that C > 0. A (canonical) application
of the ratio test shows that n≥0 C n /n! converges. (It converges to eC but we
do not need that here.) So its terms must tend to 0. Since C n /n! is positive
for each n, we deduce that n!/C n → ∞.
On
Q the−1 hand, if there areQonly finitely many primes, then n! = |n!|∞ =
other Q
n/(p−1)
|n!| p ≤ pp = ( p p1/(p−1) )n . That is, n! ≤ C n with C :=
Qp 1/(p−1)
pp . Contradiction!
1.11 The stated expression for Hp−1 is immediate upon pairing the terms
1
i and
1
. To show |Hp−1 |p ≤ p−2 , we must prove that |S|p ≤ p−1 , where
Pp−i 1
S := 0<i<p/2 i(p−i) .
∗
The first equality is not a tautology! What is being claimed is that the integer on
the left has its mod p reduction equal to the element of Fp on the right. The −1st
powers indicate inverses in the field Fp .
59
Here we use that i−2 = (p − i)−2 and that as i runs over all the nonzero
elements of Fp , so does j = i−1 . To finish, observe that since p > 3, the group
F× ×
p has an element other than ±1. So there is an r ∈ Fp with r =
2
̸ 1. Since
×
multiplication by r permutes Fp ,
X X X
r2 j2 = (rj)2 = j2,
j∈F×
p j∈F×
p j∈F×
p
References
1. H. Anton, Die Elferprobe und die Proben für die Modul Neun, Dreizehn und
Hunderteins. Für Volks- und Mittelschulen. Archiv Math. Physik 49 (1869),
241–308.
2. K. Hensel, Über die arithmetischen Eigenschaften der Faktoriellen. Archiv Math.
Physik (third series) 2 (1902), 293–294.
3. J. Kürschák, A harmonikus sorról. Mat. Fiz. Lapok 27 (1918), 299–300.
4. T. Skolem, On certain exponential equations. Norske Vid. Selsk. Forh. 18 (1945),
71–74.
5. L. Stickelberger, Ueber eine Verallgemeinerung der Kreistheilung. Math. Ann.
37 (1890), 321–367.
60
Solutions to Set #2
2.12 We have to show that for any x, y, z ∈ K, at least two of |x − y|, |y − z|,
and |z−x| coincide. This is a simple consequence of “survival of the greatest”: If
|x−y| =
̸ |y −z|, then |z −x| = |x−z| = |(x−y)+(y −z)| = max{|x−y|, |y −z|}.
This argument shows that the largest side length always appears at least twice.
2.15 Let R = {pv : v ∈ Z}. If r ∈/ R, then D<r (x0 ) = D≤r (x0 ), for any center
x0 . If r ∈ R, then D<r (x0 ) = D≤r/p (x0 ) while D≤r (x0 ) = D<pr (x0 ).
n n
n
X n k n−k X n
|x + y| = x y ≤ |x|k |y|n−k
k k
k=0 k=0
n
X n n
≤ max{|x|, |y|}n ≤ max{|x|, |y|}n · (n + 1) max .
k 0≤k≤n k
k=0
61
(Moving from the first to the second line, we used that |x|k |y|n−k ≤
max{|x|, |y|}k · max{|x|, |y|}n−k = max{|x|, |y|}n .) Take nth roots.
Since C∞ > 1, the inequality is preserved upon raising C∞ to both sides. This
gives |x + y|∞ ≤ max{|x|∞ , |y|∞ }, proving the strong triangle inequality.
2.20 To prove that D≤1 (0) is a subring it is enough to argue that 1 ∈ D≤1 (0)
and that D≤1 (0) is closed under multiplication and subtraction. The first
62
requirement is clear, since |1| = 1. Closure under multiplication follows from the
multiplicative property of |·|, as the interval [0, 1] is closed under multiplication.
Closure under subtraction is a consequence of the strong triangle inequality: If
|x|, |y| ≤ 1, then |x − y| ≤ max{|x|, | − y|} = max{|x|, |y|} ≤ 1.
By definition of the p-adic absolute value, x ∈ Z(p) if and only if x = pv a/b for
some nonnegative integer v and some a, b ∈ Z not divisible by p. This happens
precisely when the denominator of x in lowest terms is not a multiple of p.
(−1)k−1
p
≡p (mod p2 Z(p) ).
k k
k−1
2.23 Summing the congruence kp ≡ p (−1)k
(mod p2 Z(p) ) over integers
p k−1
/k (mod p2 Z(p) ). Dividing by p,
P
0 < k < p yields 2 − 2 ≡ p 0<k<p (−1)
2p − 2 1 1 1 1
≡ 1 − + − + ··· − (mod pZ(p) ).
p 2 3 4 p−1
The left and right-hand sides of the displayed congruence have difference
smaller than 1 in terms of p-adic absolute value. So by the strong triangle
inequality, one side has absolute value < 1 if and only if the other does. The
p
solution is concluded by observing that | 2 p−2 |p < 1 ⇐⇒ p2 | 2p − 2.
2.24 Since F has finitely many complex roots, we can fix n0 ∈ Z with
F (n0 ) ̸= 0. Replacing F (T ) with F (T + n0 ), we may assume that F (T ) has
nonzero constant term a0 (say). Then F (a0 T ) = a0 G(T ) for some nonconstant
G(T ) ∈ Z[T ] with G(0) = 1.
63
Let P be the set of primes dividing G(n) for some integer n. Every prime
dividing a value of G also divides a value of F , so it suffices to prove P is
infinite.
We mimic Euclid. Suppose p1 , . . . , pk is any finite list of primes in P. We
choose an integer m with |G(mp1 · · · pk )| > 1. (Such an m surely exists, as the
inequality excludes no more than 3 deg G values of m.) Then G(mp1 · · · pk ) is
divisible by some prime p, but
Extra Exploration 6 (Bauer [1]; see also Nagell [4, §49, pp. 168–169]).
Let F (T ) be a nonconstant polynomial with integer coefficients. Suppose that F has
a real root of odd multiplicity. Show that for each integer m ≥ 3, there are infinitely
many primes p ̸≡ 1 (mod m) for which F has a root mod p.
64
that is not on the list p1 , . . . , pk . As our starting list was arbitrary, there are
infinitely many primes p ≡ −1 (mod 8).
2.27 Problem 2.25 handles the residue class 1 mod 8 while Problem 2.26
handles −1 mod 8. To take care of 5 mod 8, we argue as in Problem 2.26 with
(2n + 1)2 + 4 replacing (2n + 1)2 − 2. Every odd prime p with −4 a square mod p
is congruent to 1 or 5 (mod 8). Since (2n + 1)2 + 4 ≡ 1 + 4 ≡ 5 (mod 8), there
must be some prime congruent to 5 (mod 8) dividing (2n + 1)2 + 4. Following
the solution to Problem 2.26, we obtain infinitely many primes p ≡ 5 (mod 8)
by varying n.
To deal with 3 mod 8, use (2n + 1)2 + 2. Every odd p with −2 a square mod p
is congruent to 1 or 3 (mod 8). As (2n + 1)2 + 2 ≡ 1 + 2 ≡ 3 (mod 8), there
is always some prime congruent to 3 (mod 8) dividing (2n + 1)2 + 2. Varying
n, we obtain infinitely many primes p ≡ 3 (mod 8).
References
1. M. Bauer, Über die arithmetische Reihe. J. Reine Angew. Math. 131 (1906),
265–267.
2. M. R. Murty, Primes in certain arithmetic progressions. J. Madras Univ., Section
B, 51 (1988), 161–169.
3. M. R. Murty and N. Thain, Prime numbers in certain arithmetic progressions.
Funct. Approx. Comment. Math. 35 (2006), 249–259.
4. T. Nagell, Introduction to number theory, Chelsea Publishing Co., New York,
1964.
5. I. Schur, Über die Existenz unendlich vieler Primzahlen in einigen speziellen
arithmetischen Progressionen. Sitzungber. Berliner Math. Ges. 11 (1912), 40–50.
65
66
Solutions to Set #3
3.28 The proof is the same as in calculus: Suppose for a contradiction that
xn → x and xn → x′ , where x′ ̸= x. Then ϵ := 12 |x′ − x| > 0. Since xn → x,
we can choose N ∈ Z+ with |xn − x| < ϵ for all n ≥ N . Similarly, we can
choose N ′ ∈ Z+ with |xn − x′ | < ϵ for all n ≥ N ′ . Taking n ≥ max{N, N ′ },
we find that
Contradiction!
3.29
(a) Let ϵ > 0. Choose N1 = 1. If n ≥ N1 , then |xn − x| = 0 < ϵ.
(b) Let ϵ > 0. Choose N1 , N2 ∈ Z+ so that |xn − x| < 12 ϵ whenever n ≥ N1
and |yn − y| < 12 ϵ whenever n ≥ N2 . For n ≥ max{N1 , N2 },
1 1
|(xn +yn )−(x+y)| = |(xn −x)+(yn −y)| ≤ |xn −x|+|yn −y| < ϵ+ ϵ = ϵ.
2 2
So xn + yn → x + y.
(c) Here we must work a bit harder. When checking the definition of conver-
gence, we can assume that 0 < ϵ < 1. (Larger values of ϵ only make life
1
easier.) Given such an ϵ, choose N1 , N2 ∈ Z+ with |xn − x| < 3(|y|+1) ϵ for
1
all n ≥ N1 and |yn − y| < 3(|x|+1) ϵ for all n ≥ N2 . Write xn = x + dn and
yn = y + en , so that xn yn = xy + xen + ydn + dn en . For n ≥ max{N1 , N2 },
|x| ϵ |y| ϵ
|xen | ≤ ϵ< and |ydn | ≤ ϵ< .
3(|x| + 1) 3 3(|y| + 1) 3
ϵ 1
For these same values of n, we have |dn |, |en | < 3 < 3. So (estimating
crudely) |dn en | ≤ |dn | < 3ϵ . Therefore,
ϵ ϵ ϵ
|xn yn − xy| = |xen + ydn + dn en | ≤ |xen | + |ydn | + |dn en | < + + = ϵ.
3 3 3
67
3.30 Notice that 2xn = 2 + 2 · 3 + 2 · 32 + · · · + 2 · 3n , which is precisely
n+1
the ternary expansion of 3n+1 − 1. So xn = 3 2 −1 = − 12 + 12 3n+1 , and
|xn − (− 12 )|3 = | 12 3n+1 |3 = 3−(n+1) , which tends to 0. Therefore, xn → − 12 in
(Q, | · |3 ).
P∞
The series k=0 3k diverges in Qp for each prime p ̸= 3. To prove this, we
appeal to a result possessing the air of the familiar.
P∞
Lemma (kth term test for a valued field). If k=1 ak converges in (K, |·|),
then ak → 0.
To apply this in our situation, observe that if p ̸= 3, then |3k |p = 1 for every
k, and 1 does not tend to 0 !∗
P∞
Proof. Suppose k=1 ak = P x (where x ∈ K). This means that the sequence
n
{sn } with nth term sn = k=1 ak converges to x. Let ϵ > 0 and choose
N ∈ Z with the property that |sn − x| < 12 ϵ for all n ≥ N . Then for every
+
positive integer n ≥ N + 1,
1
|an | = |sn − sn−1 | = |(sn − x) + (x − sn−1 )| ≤ |sn − x| + |x − sn−1 | < 2 · ϵ = ϵ.
2
We have verified the definition of “an → 0.” ❚
PN
3.31 Some experimentation suggests that n=0 n · n! = (N + 1)! − 1, which
PN
is easily confirmed by induction. Hence, | n=0 n · n! − (−1)|2 = |(N + 1)!|2 .
Since the power of 2 in (N + 1)! tends to infinity, |(N + 1)!|2 → 0, and
PN P∞
n=0 n · n! → −1. That is, n=0 n · n! = −1.
PN PN
Next, we look at n=0 n2 · 2n . Let F (T ) = n=0 T n . Differentiating and
PN
multiplying by T gives T F ′ (T ) = n
n=0 nT . Another round of the same
process yields
N
X
n2 T n = T (T F ′ (T ))′ = T 2 F ′′ (T ) + T F ′ (T ).
n=0
1−T N +1 1 1
Substituting in F (T ) = 1−T = 1−T − T N +1 1−T and simplifying,
N
X T (T + 1) GN (T )
n2 T n = + T N +1
n=0
(1 − T )3 (1 − T )3
68
N
X
n2 2n − (−6) = 2−N −1 · |GN (2)|2 ≤ 2−N −1 .
n=0 2
P∞
We send N to infinity and conclude that n=0 n2 2n = −6 in (Q, | · |2 ).
3.32 With x0 := 117 ∈ Z(3) , each step of the displayed algorithm has the form
xn = dn + 3xn+1 , where dn ∈ {0, 1, 2} and xn+1 ∈ Z(3) .
Repeated substitution reveals that for each nonnegative integer n,
x0 = d0 + 3x1
= d0 + 3(d1 + 3x2 )
..
.
= d0 + 3(d1 + 3(d2 + · · · + 3(dn + 3xn+1 )))
= d0 + 3d1 + 32 d2 + · · · + 3n dn + 3n+1 xn+1 .
Therefore,
n
X
x0 − 3k dk = 3−n−1 |xn+1 |3 ≤ 3−n−1 .
k=0 3
P∞
Sending n to infinity, x0 = k=0 3k dk .
As shown in the problem statement, x1 = x7 , implying that the “digits” di
repeat in blocks of six starting from i = 1.
3.33 We modify the algorithm of Problem 3.32. This time x0 = 27 ∈ Q(5) , and
each dn ∈ {0, 1, 2, 3, 4} is chosen so that xn = dn + 5xn+1 for an xn+1 ∈ Q(5) .
Grinding this out,
2 −1
=1+5·
7 7 −6 −4
=2+5·
−1 −3 7 7
=2+5·
7 7 −4 −5
=3+5·
−3 −2 7 7
=1+5·
7 7 −5 −1
=0+5· .
−2 −6 7 7
=4+5·
7 7
Replicating the logic of the solution to Problem 3.32, we conclude that in
(Q, | · |5 ),
2
= 1 + 2 · 5 + 1 · 52 + 4 · 53 + 2 · 54 + 3 · 55 + 0 · 56 + 2 · 57 + . . . ,
7
where the “digits” follow the eventually periodic pattern 1, 2, 1, 4, 2, 3, 0.
69
Extra Exploration 7 (cf. Burger and Struppeck [1]).
P∞ Show that there is
a sequence of rational numbers {an } with the property that n=1 an converges to 0
with respect to | · |∞ and converges to 1/p with respect to | · |p for every prime p.
70
Hence, max{|m|, |n|} = 1, so that either |m| = 1 or |n| = 1.
3.37 The units in O are precisely the x ∈ K × satisfying both |x| ≤ 1 and
|x−1 | ≤ 1. As |x−1 | = |x|−1 , the last two inequalities are satisfied simultane-
ously precisely when |x| = 1.
Let M be the collection of nonunits in O, so that M = D<1 (0). Clearly 0 ∈ M .
If x, y ∈ M , then |x + y| ≤ max{|x|, |y|} < 1, and so x + y ∈ M . Moreover, if
x ∈ M and r ∈ O, then |rx| = |r||x| ≤ |x| < 1, so that rx ∈ M . Hence, M is
an ideal of O. Since 1 ∈/ M , the ideal M is proper.
Let I be any proper ideal of O. If x ∈ I, then x cannot be a unit in O:
Otherwise I ⊇ xO = O. Thus, x ∈ M . Since this holds for all x ∈ I, we
conclude that I ⊆ M .
Thus, M is a proper ideal of O containing all proper ideals of O. So M cannot
itself be properly contained in a proper ideal of O; that is, M is maximal.
3.39 Let I be any nonzero ideal of Z(p) and choose a nonzero x ∈ I with
vp (x) minimal among nonzero elements of I. Set v = vp (x).
Claim: I = pv Z(p) .
71
3.40 Since x2 + 7 = 2m , the integer x must be odd. Hence, 8 | x2 + 7 = 2m
and m ≥ 3.
√
Since x is an odd number, x± 2 −7 ∈ R for both choices of sign. Furthermore,
√ √
recalling that α = 1+ 2 −7 and β = 1− 2 −7 ,
√ √
x + −7 x − −7
· = 2m−2 = (αβ)m−2
2 2
= αm−2 · β m−2 . (*)
The converse also holds: Every m with um−2 = ±1 gives rise√to an x with x2 +
7 = 2m . If um−2 = ±1, then αm−2 − β m−2 = ±(α − β) = ± −7. On the other
72
√
hand, if we write αm−2 = x+y2 −7 for integers x and y, then (applying complex
√ √
conjugation) β m−2 = x−y2 −7 , so that αm−2 − β m−2 = y −7. Comparing
x2 +7y 2 x2 +7
expressions, y = ±1. Therefore, 2m−2 = αm−2 β m−2 = 4 = 4 , and
x2 + 7 = 2m .
Remarks.
(i) We didn’t need to know R was a UFD to execute our solution. After proving
that α and β are prime, we could have appealed to the following result, valid in
every integral Q domain: If a product U V factors as π1 · · · πk , with all πi prime∗ ,
then U = η i∈S πi and V = η i∈S ′ πi for some units η, η ′ and some partition
′Q
References
∗
we really do mean prime, not merely irreducible!
73
74
Solutions to Set #4
(jT )k
4.41 We have 1 + eT + e2T + · · · + e(n−1)T =
P P
0≤j<n k≥0 k! =
P
k Tk Tk
P P
k≥0 0≤j<n j k! = k≥0 Sk (n) k! .
Tk
Comparing coefficients of k! gives
X (r + s)! r+1 X Bk−r k
Sk (n) = Bs n = nr+1 .
(r + 1)!s! r+1 r
r+s=k 0≤r≤k
r,s≥0
Remark. If you stare carefully, you will notice this argument takes for granted the
identities ejT = (eT )j (for j = 0, 1, 2, 3 . . . ). Here what is important is not that these
identities hold for real numbers T (which is familiar from calculus), but that they
hold as identities of formal power series in the indeterminate T .
75
k
= j k . Therefore, (eT )j =
P
By the multinomial theorem, k1 +···+kj =k k1 ,k2 ,...,kj
(jT )k
= ejT .
P
k≥0 k!
We could also have proved the required identities byP leveraging our prior knowledge
of the “real world.” Expand (formally) ejT − (eT )j = k≥0 ck T k for some coefficients
ck . The exact same transformations you use to put the left side into the form of the
right will show that
X
(ex )j − ejx = ck xk for all real numbers x.
k≥0
Here all manipulations with real numbers can be justified by citing absolute con-
vergence of the relevant series. (Operations on formal power series are defined to
mirror operations that can be performed on numerical series in the presence of
sufficiently good convergence.) Since (ex )j − ejx = 0 for all real numbers x, the series
P k
k≥0 ck x converges everywhere to 0. This forces each ck = 0, which in turn shows
that ejT = (eT )j formally. While gratuitous in this instance, the principle that an
identity of real numbers can (often) be transmogrified into an identity of formal
power series is frequently useful.
4.43 All of the claimed equalities are straightforward to verify, including the
invariance of T coth T under the substitution T 7→ −T (provided we accept
that e2T e−2T = 1, which can be proved by the method of the preceding
Remark). Writing down the power series for −T coth(−T ) and T coth(T ), we
conclude that
X (−2T )k X (2T )k
−T + Bk =T+ Bk .
k! k!
k≥0 k≥0
k
4.44 Starting from T coth T = T + k≥0 Bk (2Tk!) , divide by T and substitute
P
As ck and B2k share the same sign, we will be done if we show that
(−1)k+1 ck > 0 for each k ∈ Z+ . Assuming this claim fails, let K be the minimal
counterexample. For later use, note that K > 1 (since c1 = 2B2 = 13 > 0).
Differentiating the last displayed equation,
76
d 1 X
coth T = − 2 + (2k − 1)ck T 2k−2 .
dT T
k≥1
2K−2
The right-hand Laurent series has T appearing with coefficient (2K −
d
1)cK . On the other hand, the coefficient of T 2K−2 in dT coth T is the same
2
, since dT coth T = 1 − coth2 T and K > 1.
as its coefficient in − coth TP d
(−1)K X 1 X
(−1)K+1 cK = ci cj = (−1)i+1 ci · (−1)j+1 cj > 0.
2K + 1 2K + 1
i+j=K i+j=K
i,j≥1 i,j≥1
We use in the last step that (−1)i+1 ci > 0 and (−1)j+1 cj > 0, since i and j
are positive integers smaller than K.
Remark. A more satisfying explanation for why the even-indexed Bernoulli numbers
alternate in sign is found in Euler’s formula
(2m)!
B2k = (−1)k+1 2ζ(2k) · ,
(2π)2m
where ζ(s) := n≥1 n1s is the Euler–Riemann zeta function. This remarkable relation
P
pins down B2k rather precisely as a real number. Indeed, when conjoined to Stirling’s
estimate on factorials, it implies that
|B2k |
lim √ k 2k
= 1.
k→∞ 4 πk( πe )
Unfortunately, Euler’s formula does not contain any obvious information about the
number-theoretic properties of B2k .
We will not prove Euler’s result here. Interested readers are referred to the exquisitely
written textbook of Ireland and Rosen for a characteristically elegant treatment [2,
pp. 231–232].
1 2
2 22
3 22 · 5/3
4 22 · 8/3
5 28 · 1/15
6 25 · 13/15
5
7 2 · 151/105
8 213 · 1/105
9 29 · 83/315
10 210 · 73/315
77
Already this limited data suggests that v2 ( 1≤k≤n 2k /k) tends to infinity
P
P∞ k
with n (equivalently, that P k=1 2 /k = 0 in (Q, | · |2 )). Being a bit bolder,
we might conjecture that v2 ( 1≤k≤n 2k /k) is bounded below by a function
ever-so-slightly smaller than n. For the resolution of both conjectures, see the
solution to Problem 11.130.
Lemma. For each x ∈ Z(p) , there is a d ∈ {0, 1, 2, . . . , p−1} with x−d ∈ pZ(p) .
With the lemma in hand, we can express r in the desired form by the algorithm
of Problems 3.32 and 3.33. Specifically, let x0 = r, and for n = 0, 1, 2, 3, . . . ,
select dn ∈ {0, 1, 2, . . . , p − 1} so that xn = dn + pxn+1 for some xn+1 ∈ Z(p) .
Then r = k≥0 dk pk .
P
78
(b) each |xn |∞ ≤ M , where M := max{2, |r|∞ }.
Claim (a) is obvious when n = 0, since x0 = r = a/b. Suppose that xn ∈ 1b Z.
Then bpxn+1 = bxn − bdn ∈ Z. As xn+1 ∈ Z(p) , we also have |bpxn+1 |p =
|bp|p |xn+1 |p ≤ p−1 . So in fact bpxn+1 ∈ pZ, and xn+1 ∈ bp 1
(pZ) = 1b Z. This
gives (a). Since x0 = r, (b) is trivial when n = 0. If |xn |∞ ≤ M , then
1 1 1 1
|xn+1 |∞ = |xn − dn |∞ < (|xn |∞ + p) ≤ |xn |∞ + 1 ≤ M + 1 ≤ M.
p p 2 2
So we have (b) as well. From (a) and (b) it is easy to conclude: [−M, M ] has
finite intersection with 1b Z, so the xn cannot all be distinct.
Hence,
−|x − xn | ≤ |x| − |xn | ≤ |x − xn |.
Since |xn − x| → 0, the squeeze theorem implies that |x| − |xn | → 0. Therefore,
|xn | converges to |x|.
Suppose now that | · | is non-Archimedean and that xn → x, where x = ̸ 0. The
limit definition guarantees that |xn − x| < |x| for all sufficiently large n. For
these n, “survival of the greatest” yields |xn | = |(xn − x) + x| = |x|.
4.48 By Problem 3.34 we can choose a prime p with |p| < 1. Then whenever
n is an integer not divisible by p, the integers p and n are relatively prime and
Problem 3.35 tells us that |n|p = 1.
Write |p| = p−c with c > 0. Given x ∈ Q× , we can express x = pvp (x) a/b where
a and b are integers not divisible by p. Then
|x| = |p|vp (x) |a||b|−1 = |p|vp (x) = p−cvp (x) = (p−vp (x) )c = |x|cp .
4.49 From Exercise 4.45, |r| > 1 for each integer r ≥ 2. So there are indeed
real numbers c, d > 0 with |2| = 2c and |3| = 3d .
Pm
If we write 2n = i=0 ϵi 3i , with each ϵi ∈ {0, 1, 2} and ϵm > 0, then
m
X |3|m+1 − 1
2cn = |2|n = |2n | ≤ |ϵi ||3|i ≤ max{|1|, |2|}
i=0
|3| − 1
max{|1|, |2|}|3| |2| · |3| dm
≤ · |3|m = 3 ,
|3| − 1 |3| − 1
79
|2|·|3|
proving the claimed inequality with B := |3|−1 . As ϵm > 0, we have 2n ≥ 3m ,
and thus 3 ≤ 2 . Therefore, 2 ≤ B · 2 , and 2(c−d)n ≤ B. Since n may
dm dn cn dn
4.50 The arguments of Problem 4.49 apply equally well with 3 replaced by
an arbitrary integer r ≥ 3. Writing |2| = 2c and |r| = rd , we find that for each
positive integer n,
4.51 We begin with a simple but useful observation: If (K, | · |) is any valued
field, and x ∈ K, then
xn → x in (Q, | · |) ⇐⇒ |xn − x| → 0
⇐⇒ |xn − x|∞ → 0 ⇐⇒ xn → x in (Q, | · |∞ ).
80
If | · | is non-Archimedean, then | · | = | · |cp for some prime p and some c > 0.
In this case | · | is equivalent to | · |p (same reasoning as displayed above).
It remains to show that none of | · |∞ , | · |2 , | · |3 , | · |5 . . . are equivalent. From
Exercise 4.51, | · |∞ is not equivalent to any of the others, since | · |∞ is
Archimedean while | · |p is non-Archimedean. If p and q are distinct primes,
then pn → 0 in (Q, | · |p ) but pn ̸→ 0 in (Q, | · |q ). So | · |p and | · |q are
inequivalent.
81
allows us to conclude that |Hn |p ≥ p whenever pk+1 ≤ n < pk+2 . Repeating
the reasoning, |Hn |p ≥ p2 whenever pk+2 ≤ n < pk+3 . In general, |Hn |p ≥ pj
for all n ≥ pk+j (for j = 0, 1, 2, 3, . . . ). Therefore, |Hn |p → ∞.
One can check with a simple computer program that k = 2 satisfies our
hypothesis both when p = 3 and when p = 5. So |Hn |3 → ∞ and |Hn |5 → ∞.
Several open questions persist regarding the numerators and denominators of the
harmonic numbers. Here is one that appears deceptively simple: Are there infinitely
many n for which the denominator of Hn (in lowest terms) is the least common
multiple of 1, 2, 3, . . . , n?
References
1. D. W. Boyd, A p-adic study of the partial sums of the harmonic series. Experi-
ment. Math. 3 (1994), 287–302.
2. K. Ireland and M. Rosen, A classical introduction to modern number theory,
second ed., Graduate Texts in Mathematics, vol. 84, Springer-Verlag, New York,
1990.
82
Solutions to Set #5
5.58 For each n ∈ Z+ , the element n·1 of Zp has a nonzero mod pk component
whenever pk > n. In particular, n · 1 ̸= 0. Thus, Zp has characteristic 0, so
that we can (and henceforth will!) view Z as a subring of Zp , identifying each
integer n with n · 1 = (n mod p, n mod p2 , n mod p3 , . . . ).
The rest of the problem asks (essentially) for a description of Q ∩ Zp . It may
not be clear that this task even makes sense: To take this intersection, both Q
and Zp have to be viewed as sitting inside a common superset. What is that
set?
Fortunately, this question has an easy answer: Both Q and Zp are contained in
the fraction field of Zp . Taking this interpretation, if a, b ∈ Z with b nonzero,
83
a
b ∈ Q ∩ Zp precisely when there is a y ∈ Zp with by = a. When does this
happen?
We can assume, without loss of generality, that gcd(a, b) = 1. If p | b, there
is no hope of finding a y with by = a: The mod p component of by is 0 for
every y ∈ Zp , while the mod p component of a is nonzero (remember that
gcd(a, b) = 1, so that p cannot divide a). Suppose instead that p ∤ b. In this
case, we can choose integers Bk with bBk ≡ 1 (mod pk ), for k = 1, 2, 3, . . . .
Observing that bBk+1 ≡ 1 (mod pk+1 ), we find that
84
pv = (a2 mod p, a3 mod p2 , . . . ) = (a1 mod p, a2 mod p2 , . . . ) = u.
(We use here that ak+1 ≡ ak (mod pk ) for each k.) Thus, p | u.
(Here we use that a2v+i ≡ a2v+i−1 ≡ · · · ≡ av+i (mod pv+i ).) So we have
produced a representation x = pv u with u ∈ Z×
p.
′
To prove uniqueness, suppose x = pv u = pv u′ with v, v ′ nonnegative integers
v ′ −v ′
and u, u′ ∈ Z× ′ v
(p) . Assume v ≤ v . Canceling p from both sides, u = p u.
×
Since u ∈ Zp , to avoid a contradiction with Exercise 5.59 we must have
v ′ − v = 0, i.e., v = v ′ . Then pv u = pv u′ , leading to u = u′ .
5.61 For each nonzero x ∈ Zp , define vp (x) as the integer v appearing in the
representation of x described in Problem 5.60. Then proceed as in the solution
to Problem 3.39.
5.62 We preface the solution with some philosophical comments. With the
definition we have given of Zp , there is an obvious informal way to reduce x
modulo powers of p: Write x = (a1 mod p, a2 mod p2 , . . . ) and reduce mod pn
by extracting the nth component, i.e., singling out an mod pn . This defines
the reduction of x mod pn as an element of Z/pn . Since Zp is a ring, there is
also a canonical way to reduce x modulo pn : Take the image of x in Zp /pn Zp .
It would be reassuring if these two ways of reducing mod pn were somehow the
same. As we will see shortly, this turns out to be the case! The isomorphism
between Z/pn and Zp /pn Zp induced by the inclusion Z ,→ Zp matches up
an mod pn and x mod pn Zp . Equivalently: x mod pn Zp = an mod pn Zp .
85
There is no justice in Heaven or Earth,
but there is certainly justice in
MATHEMATICS!
Paul Erdős
Next we show that ϕ, and hence also ϕ̃, is surjective. Let x = (a1 mod
p, a2 mod p2 , . . . ) ∈ Zp . The mod pj -component of an − x is an − aj mod pj ,
which vanishes for all j ≤ n (by the compatibility condition in the definition
of Zp ). So by our solution to Problem 5.60, either an − x = 0 or an − x = pk u
for an integer k ≥ n and a unit u ∈ Z× n
p . In either case, an − x ∈ p Zp , so that
ϕ(an ) = an mod pn Zp = x mod pn Zp . This proves ϕ is surjective, and so ϕ̃ is
an isomorphism.
The isomorphism ϕ̃ carries an mod pn to ϕ(an ) = x mod pn Zp , as we promised
in our initial remarks.
“Straightforward” means that the tedium of writing out the proof outweighs
the enlightenment gained from doing so. (Please walk through the argument
in your head and decide if you agree!)
To complete the problem, it suffices to show that if p is prime and v ∈ Z+ ,
then Zpv ∼= Zp . Here we can map (b1 mod pv , b2 mod p2v , b3 mod p3v , . . . ) to
(b1 mod p, . . . , b1 mod pv , b2 mod pv+1 , . . . , b2 mod p2v , . . . ), each bi repeated
v times. (Again, that this work is “straightforward.” Verify!)
5.64 The outer, red disc represents the entirety of Z3 . The salmon-colored
discs partition Z3 into three parts, based on the mod 3 component. Once the
mod 3 component is fixed, there are 3 possibilities for the mod 32 component,
depicted by the green discs. Finally, having fixed the mod 32 component, there
are three possibilities for the mod 33 component, illustrated by the purple
86
discs. In theory this partitioning process could continue indefinitely, but at
some point our eyes would ask for a break!
Remark. For further discussions around visualizing Zp , see [3], [9], [10, Chapter 2],
and [11, Chapter 1, §2].
Extra Exploration 10. Show that Zp has the same cardinality as R. A useful
tool for this kind of proof is the Cantor–Schröder–Bernstein theorem: If there are
injections from A to B and from B to A, then there is a bijection between A and B.∗
5.67 Let (xr , yr ), (xb , yb ), and (xg , yg ) be the coordinates of the red, blue,
and green vertices of ∆. Then, by a formula well-known to those who know it
well,
xb yb 1
2 · Area(∆) = ± xg yg 1 = ±(xb yg + xr yb + xg yr − xg yb − xb yr − xr yg ).
xr yr 1
We claim that xb yg has 2-adic absolute value strictly larger than the other five
terms. By definition of the coloring, |xb yg |2 = |xb |2 |yg |2 ≥ 1 · 1 = 1. Next, we
observe that
|xr |2 |yb |2 ≤ |xr |2 |xb |2 < |xb |2 ≤ |xb |2 |yg |2 .
If yr = 0, then |xg yr |2 = 0 < |xb yg |2 ; otherwise,
87
|y|2
|x|2
Next,
|xb |2 |yr |2 < |xb |2 ≤ |xb |2 |yg |2 .
Finally,
|xr |2 |yg |2 < |yg |2 ≤ |xb |2 |yg |2 .
Phew!
Falling back on “survival of the greatest,” we conclude that |2 · Area(∆)|2 =
|xb yg |2 ≥ 1, so that |Area(∆)|2 ≥ |2|−1
2 > 1.
Remark. Here is a quick sketch of Monsky’s proof. We can assume the square we
are trying to dissect is S = [0, 1] × [0, 1] ⊆ R2 .
Suppose S has been dissected into finitely many triangles. So that we can bring
the coloring of Q2 into the picture, let’s suppose that the vertices of all triangles
appearing in the dissection have rational coordinates.
Using a version of Sperner’s lemma from combinatorial geometry, along with basic
properties of our coloring, Monsky argues that the dissection must contain an odd
number of rainbow triangles. In particular, there is always at least one rainbow
triangle ∆ involved. As we showed above, |Area(∆)|2 > 1. So if we assume there are
88
n triangles involved in the dissection, each with the same area, then |1/n|2 > 1; that
is, n is even.
This is all well and good, but we are interested in arbitrary (real) dissections of S,
not merely “rational” ones. Well, here is the astounding part: The same argument
applies in the general case! At first glance this claim doesn’t seem to make sense:
Our initial coloring was defined in terms of | · |2 , and | · |2 has domain Q, not R. It
turns out, however, that it is possible to extend | · |2 to an absolute value on R. (This
is far from obvious!) With this extension in hand, everything else in the argument
goes through with zero change.
A complete proof of Monsky’s theorem can be found in [1, Chapter 22]. The discussion
in [1] sidesteps extending | · |2 to an absolute value on R; the reader interested in
seeing such an extension constructed can consult §14 of [12].
Extra Exploration 11 (cf. Hales and Straus [4]). Let K be an infinite field
that can be endowed with a nontrivial, non-Archimedean absolute value. Show that
it is possible to 3-color the affine plane K 2 , using all three colors in an interesting
way, so that every line receives points of at most two colors. In an interesting way
means that each color is assigned to three non-collinear points. (If we remove “in an
interesting way,” this can be done for every infinite field. Can you see why?)
Remark. It seems that neither the definition in [8], nor Hensel’s earlier descriptions
in [6, 7] (similar but less refined), were viewed by Hensel’s contemporaries as entirely
rigorous. In his obituary [5] for Hensel, Hasse describes the p-adic numbers as “a
genuine creation driven by intuition and imagination, which, initially, like every
revolutionary idea, was thrown down bluntly and in raw form and which, similar to
Leibniz’s differential calculus, initially lacked a solid logical foundation.” Apprehen-
sions about the logical standing of p-adic numbers were dispelled only after Kürschák
and Ostrowski’s papers on valuation theory, which began to appear around the same
time as [8].
References
1. M. Aigner and G. M. Ziegler, Proofs from The Book, sixth edition, Springer,
Berlin, 2018.
89
2. R. H. Cox, A proof of the Schroeder-Bernstein theorem. Amer. Math. Monthly
75 (1968), 508.
3. A. A. Cuoco, Visualizing the p-adic integers. Amer. Math. Monthly 98 (1991),
355–36.
4. A. W. Hales and E. G. Straus, Projective colorings. Pacific J. Math. 99 (1982),
31–43.
5. H. Hasse, Kurt Hensel zum Gedächtnis. J. Reine Angew. Math. 187 (1949),
1–13.
6. K. Hensel, Neue Grundlagen der Arithmetik. J. Reine Angew. Math. 127 (1904),
51–84.
7. , Theorie der algebraischen Zahlen, B. G. Teubner, Leipzig and Berlin,
1908.
8. , Zahlentheorie, G. J. Göschen, Berlin and Leipzig, 1913.
9. J. E. Holly, Pictures of ultrametric spaces, the p-adic numbers, and valued fields.
Amer. Math. Monthly 108 (2001), 721–728.
10. S. Katok, p-adic analysis compared with real, Student Mathematical Library,
vol. 37, American Mathematical Society, Providence, RI; Mathematics Advanced
Study Semesters, University Park, PA, 2007.
11. A. M. Robert, A course in p-adic analysis, Grad. Texts in Math., vol. 198,
Springer-Verlag, New York, 2000.
12. W. H. Schikhof, Ultrametric calculus: An introduction to p-adic analysis, Cam-
bridge Studies in Advanced Mathematics, vol. 4, Cambridge University Press,
Cambridge, 2006.
90
Solutions to Set #6
The cases where m > n ≥ N are interchangeable with these, since |xn − xm | =
|xm − xn |. Finally, the inequality |xn − xm | < ϵ is trivial for n = m.
n n+1 n
6.69 Let xn = 25 . By Euler’s theorem, xn+1 − xn = xn · (25 −5 − 1) =
n+1
xn · (2φ(5 ) − 1) ≡ xn · 0 ≡ 0 (mod 5n+1 ). Thus, |xn+1 − xn |5 ≤ 5−n−1 , which
tends to 0, and {xn } is a Cauchy sequence by Exercise 6.68.
6.71 The argument is the same as in calculus. Suppose {xn } is Cauchy and
that the subsequence {xnk } converges to x.
Given ϵ > 0, choose N1 ∈ Z+ with the property that |xk − xℓ | < 12 ϵ whenever
k, ℓ ≥ N1 . Choose N2 ∈ Z+ so that |xnk − x| < ϵ for all k ≥ N2 . If k ≥
max{N1 , N2 }, then
1 1
|xk − x| = |(xk − xnk ) + (xnk − x)| ≤ |xk − xnk | + |xnk − x| < ϵ + ϵ = ϵ.
2 2
91
(We use here that nk ≥ k ≥ N1 .) Hence, xk → x.
6.72 Since p−1 ∈ Qp , both n≥0 p−n Zp and Zp [1/p] are contained in Qp .
S
pv u
By Exercise 5.60, each x ∈ Q×
p can be written as pv ′ u ′
for some nonnegative
v−v ′ ′
integers v, v ′ and some u, u′ ∈ Z× . Then x = p uu′−1 ∈ pv−v Zp . Since
′ S p
pv−v Zp is a subset of both n≥0 p−n Zp and Zp [1/p], it follows that x is
contained in both sets.
92
and Zp = {x ∈ Qp : |x|p ≤ 1}.
Next, observe that for each nonzero x ∈ Qp ,
x ∈ Z×
p ⇐⇒ vp (x) = 0 ⇐⇒ p
−vp (x)
= 1 ⇐⇒ |x|p = 1.
Hence, Z×
p = {x ∈ Qp : |x|p = 1}.
6.76 We address only the part after “thus”; the earlier claims in the problem are
consequences of the Pigeonhole Principle. By construction, ak+1 ≡ ak (mod pk )
for every k, so that x := (a1 mod p, a2 mod p2 , . . . ) ∈ Zp . Furthermore, for
each k, there are infinitely many n ∈ Z+ where the mod pk component of xn
is ak mod pk .
Choose n1 ∈ Z+ so that the mod p component of xn1 is a1 mod p. Then choose
n2 > n1 where the mod p2 component of xn2 is a2 mod p2 , then n3 > n2 where
the mod p3 component of xn3 is a3 mod p3 , etc.
We argue that xnk → x in (Qp , | · |p ). Let k ∈ Z+ . By construction, the mod
pk component of xnk − x vanishes. By the compatibility condition baked into
the definition of Zp , the mod p, mod p2 , . . . , mod pk−1 components also
vanish. Therefore (see the solution to Problem 5.60), either xnk − x = 0 or
xnk −x = pv u for an integer v ≥ k and a u ∈ Z×p . In either case, vp (xnk −x) ≥ k,
−k −k
and |xnk − x|p ≤ p . Since p → 0, we conclude that xnk → x, as required.
Remark. Grouchy experts may complain that Problem 6.76 establishes the sequen-
tial compactness of Zp , not compactness with today’s accepted meaning in general
topology (cf. [1]). But those same experts will know that for metric spaces — such
as Zp — the two concepts coincide. So what exactly are they complaining about?
Such grumblers may find the following Extra Exploration more to their tastes.
j
1
eT /j
Q
6.77 Let’s accept for the time being the identity j≥1 = 1−T in Q[[T ]].
1
The series for 1−T has T p -coefficient (in fact, every coefficient) equal to 1. To
see what the coefficient of T p looks like on the left side of our identity, we
expand
93
∞
Tj 1 T 2j 1 T 3j
Y
1+ + 2
+ + ...
j=1
j 2! j 3! j 3
X X 1 1
= Tk .
e1 !e2 !e3 ! · · · 1 2 3e3 · · ·
e 1 e 2
k≥0 e1 +2e2 +3e3 +···=k
all ei ≥ 0
1 1 (p − 1)! + 1
+ = = p|(p − 1)! + 1|p .
p! p p p! p
Remark. It is a little tricky to write down a rigorous proof of the formal identity
Q T j /j 1
j≥1 e = 1−T . We sketch one argument, which, however, involves some ‘cheating’;
specifically we draw on the theory of complex variables, which is not a prerequisite
for the rest of the text.
j
For each J ∈ Z+ , let FJ (T ) = 1≤j≤J eT /j (as a formal power series). Reasoning as
Q
j
in the Remark to Problem 4.42, FJ (z) = 1≤j≤J ez /j for all complex z. As J → ∞,
Q
1
the functions FJ (z) converge uniformly to 1−z on every compact subset of the open
unit disc |z| < 1. So by Cauchy’s integral formula, if we write FJ (T ) = k≥0 ak,J T k ,
P
then for each fixed k,
Z J→∞
Z
1 1 1
ak,J = FJ (z)z −k−1 dz −−−→ z −k−1 dz = 1.
2πi |z|=9/10 2πi |z|=9/10 1 − z
6.78 Write ap−1 = 1 + pqp (a), bp−1 = 1 + pqp (b). Multiplying, (ab)p−1 =
1 + p(qp (a)qp (b) + pqp (a)qp (b)), so that
(ab)p−1 − 1
qp (ab) = = qp (a) + qp (b) + pqp (a)qp (b) ≡ qp (a) + qp (b) (mod p).
p
94
References
95
96
Solutions to Set #7
7.80 Let {xn } be a Cauchy sequence in Qp . We know from Exercise 6.70 that
{|xn |} is bounded. Hence, if k is sufficiently large, then each |pk xn | ≤ 1, i.e.,
{pk xn } is a sequence in Zp . Since {xn } is Cauchy, so is {pk xn }. By Problem
7.79, pk xn → x for some x ∈ Zp . Then xn → p−k x.
n
7.81 Let xn = 25 . By Problem 6.69, {xn } is a Cauchy sequence in Z5 .
Invoking Exercise 7.79, xn → x for some x ∈ Zp .
By the product rule for limits, lim x5n = (lim xn )5 = x5 . On the other hand,
x5n = xn+1 , so that lim x5n = lim xn+1 = x. Hence, x = x5 .
n 1
As shown in the solution to Problem 6.69, xn = 25 ≡ 25 ≡ 2 (mod 5) for all
n. Therefore, |xn − 2|5 ≤ 15 for all n, and (see Problem 4.47)
1
|x − 2|5 = | lim (xn − 2)|5 = lim |xn − 2|5 ≤ .
5
̸ 0, and x = x5 tells us 1 = x4 . That
Thus, x ≡ 2 (mod 5Z5 ). In particular, x =
is, x is a 4th root of 1, as claimed.
The only 4th roots of 1 belonging to Q are 1 and −1. Neither is congruent to
2 modulo 5Z5 . Therefore, x ∈
/ Q.
97
7.84 We define a candidate isomorphism ϕ : L → L′ as follows. Since K is
dense in L, for each x ∈ L there is a sequence {xn } in K such that xn → x in
L. Since {xn } converges in L, {xn } is Cauchy in L. Then {xn } is also Cauchy
in L′ , as each xn ∈ K and | · | and | · |′ extend the same absolute value on K.
′
Since (L′ , | · |′ ) is complete, {xn } converges in L′ . We define ϕ(x) = lim(L ) xn ,
where the superscript indicates that the limit is taken in L′ .
Honor demands we check that ϕ(x) depends only on x and not on the particular
{xn }. Suppose {xn } and {x̃n } are two sequences in K both converging to x in
L. Then xn − x̃n → 0 in L. As each xn − x̃n ∈ K, and | · | and | · |′ extend the
same absolute value on K, it must be that xn − x̃n → 0 in L′ . Consequently,
′ ′
lim(L ) xn = lim(L ) x̃n . Vindication!
If x ∈ K, we can compute ϕ(x) by taking each xn = x. This shows that
′
ϕ(x) = lim(L ) x = x. So ϕ fixes K.
Let x, y ∈ L and choose sequences {xn }, {yn } in K such that xn → x in L
and yn → y in L. Then xn + yn → x + y in L, and
′ ′ ′
ϕ(x + y) = lim(L ) (xn + yn ) = lim(L ) xn + lim(L ) yn = ϕ(x) + ϕ(y).
and
|x| ≥ |xn | − ϵ = |xn |′ − ϵ ≥ |ϕ(x)|′ − 2ϵ.
Hence |x| and |ϕ(x)|′ are within 2ϵ of each other. Since this holds for each
ϵ > 0, it must be that |x| = |ϕ(x)|′ .
Extra Exploration 14. Show that the map ϕ described above is the unique
isometric isomorphism from L to L′ .
98
Remark. Every valued field (K, | · |) admits a completion (which we have just seen
is then unique up to isometric isomorphism). The usual way to show a completion
exists is to mimic one of the standard constructions of R from Q, due to Cantor and
Méray: Take the ring of Cauchy sequences in K and quotient by the maximal ideal
of sequences tending to 0. For details, see for instance [3, §1.3]. Those with more
exotic tastes might enjoy a recent, very different argument by Kionke [4].
Pn
7.85 Let sn = k=0 ck pk . Then |sn+1 − sn |p = |cn+1 pn+1 |p ≤ p−n−1 . As
p−n−1 → 0, the sequence {sn } is Cauchy (see Exercise 6.68). By Problem
P∞ 7.79,
sn has a limit in Zp . This is precisely what it means to say that k=0 ck pk
converges to an element of Zp .
c0 +c1 p+· · ·+ck−1 pk−1 = (x1 −x0 )+(x2 −x1 )+· · ·+(xk −x1 ) = xk −x0 = xk
99
integer k such that the following holds: The digits d0 , d1 , . . . , dm−1 appear, in
that order, infinitely often in the base p expansion of kx.
cj pj , then cj = d0 , cj+1 = d1 , . . . , cj+m−1 =
P
The last bit means that if we expand kx = j≥0
dm−1 for infinitely many nonnegative integers j. For the proof, consider multiples of k0 x,
where k0 x has long runs of zeros as in (b).
(d) Assume x is an irrational element of Zp . Let m ∈ Z+ . Show that there is a
positive integer k′ such that every sequence of m base p digits appears infinitely
often in the base p expansion of k′ x.
Hint. Can you do this with a not-necessarily-positive (but nonzero) k′ ? If so, what happens
when you replace k′ with −k′ ?
7.87 Start with any x ∈ Qp and select k ∈ Z with pk x ∈ Zp . Take the infinite
base p expansion of pk x constructed in Problem 7.86 and scale by p−k to
obtain an expansion of x satisfying (i)–(iii).
If there were two expansions of x possessing properties (i)—(iii), scaling both
by the same sufficiently large power of p would produce an element of Zp with
two different base p expansions.
7.89 We record a simple observation for later use: For each ℓ ∈ Z+ , the
series 1 + pℓ + p2ℓ + . . . converges to 1/(1 − pℓ ) in (Qp , | · |). We omit the
straightforward proof (compare with Exercise 3.30).
Suppose x has an eventually periodic canonical expansion with (not necessarily
minimal) period ℓ. Scaling x by aPsuitable power of p (which does not affect
rationality), we can assume x = k≥0 ck pk where ck = ck+ℓ for all k ≥ k0 .
k k k
P P P
Then x = 0≤k<k0 ck p + k≥k0 ck p . Clearly, 0≤k<k0 ck p ∈ Q. Less
trivially,
X X
ck pk = lim ck pk
m→∞
k≥k0 k0 ≤k<k0 +mℓ
X
= lim ck pk (1 + pℓ + · · · + p(m−1)ℓ )
m→∞
k0 ≤k<k0 +ℓ
X
= ck pk (1 + pℓ + p2ℓ + . . . )
k0 ≤k<k0 +ℓ
X pk
= ck ∈ Q.
1 − pℓ
k0 ≤k<k0 +ℓ
Therefore, x ∈ Q.
100
Turning to the converse: We have already shown that each x ∈ Z(p) has
an eventually periodic base p expansion. To obtain an eventually periodic
canonical expansion for an arbitrary x ∈ Q, scale x by pk to place pk x ∈ Z(p) ,
then rescale by p−k .
P∞
Open problem: Is the p-adic number k=0 k! rational for some prime p?
P∞
Extra Exploration 19 (Dragovich [2]). Disprove: k=0 k! converges to the
same rational number in Qp for all primes p.
Pp−1 (p−1)!p−1 −1
7.90 By Problem 6.78, a=1 qp (a) ≡ qp ((p − 1)!) ≡ p (mod p). So
(p−1)!p−1 −(p−1)!−2
the claimed congruence is equivalent to p ≡0
(mod p), or
(p − 1)!p−1 − (p − 1)! ≡ 2 (mod p2 ). Using Wilson’s theorem, write (p − 1)! =
−1 + pr with r ∈ Z. Then, working modulo p2 ,
p−1
X p−1
(p − 1)!p−1 = (−1)p−1−j (pr)j ≡ 1 + (p − 1)(−1)pr ≡ 1 + pr,
j=0
j
2 − a2k
2ak q ≡ (mod 7).
7k
Both 2 and ak are coprime to 7. (Since a2k ≡ 2 (mod 7), we cannot have ak ≡ 0
(mod 7).) Thus, the last displayed congruence uniquely determines the residue
class of q modulo 7. Picking any q from this class and defining ak+1 = ak + 7k q
yields our desired lift. Note that since q is uniquely determined mod 7, our lift
is uniquely determined as a residue class mod 7k+1 .
Assume now that a1 mod 7, a2 mod 72 , a3 mod 73 , . . . have been determined
by the above procedure. Let x = (a1 mod 7, a2 mod 72 , a3 mod 73 , . . . ). Then
101
x ∈ Z7 as each ak+1 is a lift of ak . By construction, x2 = (2 mod 7, 2 mod
72 , 2 mod 73 , . . . ) = 2.
We still need to verify that the canonical expansion of x is as stated. The
initial steps of the algorithm, starting with 3 mod 7, are described in the
problem statement. We lifted 3 mod 7 to 3 + 1 · 7 mod 72 , and we lifted that
to 3 + 1 · 7 + 2 · 72 mod 73 . If we run the algorithm one more round, we arrive
at 3 + 1 · 7 + 2 · 72 + 6 · 73 mod 74 (check this!). We can read off from here that
the 7-adic expansion of x starts as 3 + 1 · 7 + 2 · 72 + 6 · 73 + . . . .
Remark. Gauss left us in 1855, four decades before Hensel’s first publication on
the p-adic numbers. It is therefore rather remarkable that one finds in his Nachlass
(manuscripts left behind at death) the claim that
√
5 (mod 11∞ ) = 9.0.4.10.4.4.
References
∗
https://math.stackexchange.com/a/4877205
102
Page from Gauss’s Nachlass; originally scanned by the Göttinger Digital-
isierungszentrum.
103
104
Solutions to Set #8
8.92
(a) In any complete valued field, a sequence {sn } converges ⇐⇒ {sn } is
Cauchy. The forward implication was noted on Set #6 (for every valued
field, not necessarily complete). The backward implication is the definition
of completeness.
P∞
Let k=1 ak be a series in K, a field assumed to P be complete withPrespect
n ∞
to a non-Archimedean absolute value. Put sn = k=1 ak . Then k=1 ak
converges ⇐⇒ {sn } converges ⇐⇒ {sn } is Cauchy ⇐⇒ sn+1 − sn → 0
⇐⇒ an+1 → 0 ⇐⇒ an → 0.
P∞
(b) If we assume k=1 ak converges to s, then |sn | → |s| (see Problem
4.47). By the strong inequality, each |sn | ≤ max{|a1 |, |a2 |, . . . , |an |} ≤
maxk=1,2,3,... |ak |. (That the last max exists is ensured by |ak | tending to
0.) Hence, |s| = lim |sn | ≤ maxk=1,2,3,... |ak |.
P∞ P∞
8.93 Let k=1 bk be any rearrangement of Pk=1 ak ; say bk = P aσ(k) for all
n n
k = 1, 2, 3, . . . , where σ ∈ Sym(Z+ ). Put sn = k=1 ak and tn = k=1 bk . By
assumption, sn → s, while our task is to prove that tn → s.
Given ϵ > 0, select N0 ∈ Z+ with |an | < ϵ whenever n > N0 . Then choose
N ∈ Z+ as the maximum of σ −1 (1), . . . , σ −1 (N0 ). Let us argue that |tn −s| < ϵ
whenever n ≥ N .
Let n ≥ N . By our choice of N , all of a1 , a2 , . . . , aN0 appear among
b1 , b2 , . . . , bN , so that
105
X X X
tn − sN0 = bk − ak = aσ(k) ,
1≤k≤n 1≤k≤N0 1≤k≤n, σ(k)>N0
and
|tn − sN0 | ≤ max{|aσ(k) | : 1 ≤ k ≤ n, σ(k) > N0 } < ϵ.
Moreover,
X
|s − sN0 | = ak ≤ max |ak | < ϵ.
k>N0
k>N0
Hence,
as desired.
Pn Pn Pn
8.94 Put sn = k=0 ak , tn = k=0 bk , and un = k=0 ck . Let s = lim sn
and t = lim tn be the infinite sums of the ak and bk , respectively. We must
show that un → st.
Observe that X X
sn tn = ak bℓ + ak bℓ
0≤k,ℓ≤n 0≤k,ℓ≤n
k+ℓ≤n k+ℓ>n
8.96 We start by noting that all of the double series appearing here converge or
involve finitely many terms. P
This follows from Problem 8.95. ToPtake
P one exam-
N P
ple (the others are similar), i=0 j ai,j can be rewritten as i j ai,j 1i≤N .
106
Since |ai,j 1i≤N | ≤ |ai,j |, the sufficient condition for convergence furnished by
Problem 8.95 is satisfied with the same sequence {ϵN }.
We can now establish our four inequalities. To attack the first of these, notice
that
XX N X
X ∞
X X X
ai,j − ai,j = ai,j ≤ max ai,j .
i≥N +1
i j i=0 j i=N +1 j j
Each term ai,j appearing P in the final sum on j has i ≥ N + 1, and thus
|ai,j | ≤ ϵN +1P
. Hence, | j ai,j | ≤ maxj |ai,j | ≤ ϵN +1 for each i ≥ N + 1, and
maxi≥N +1 | j ai,j | ≤ ϵN +1 . So we have the first inequality.
Analogous reasoning shows that
XX N
XX ∞
X X ∞
X
ai,j − ai,j = ai,j ≤ max ai,j ≤ ϵN +1 ,
j
j i j i=0 j i=N +1 i=N +1
N X
X N X
X N N
X ∞
X ∞
X
ai,j − ai,j = ai,j ≤ max ai,j ≤ ϵN +1 ,
0≤i≤N
i=0 j i=0 j=0 i=0 j=N +1 j=N +1
and
N
XX N X
X N ∞
X N
X N
X
ai,j − ai,j = ai,j ≤ max ai,j ≤ ϵN +1 .
j≥N +1
j i=0 j=0 i=0 j=N +1 i=0 i=0
Now look at the first column of inequalities in the problem statement. These
inequalities, in conjunction with the strong triangle inequality, imply that
XX N X
X N
ai,j − ai,j ≤ ϵN +1 .
i j i=0 j=0
XX N X
X N
ai,j − ai,j ≤ ϵN +1 .
j i j=0 i=0
PN PN PN PN
Since i=0 j=0 ai,j = j=0 i=0 ai,j (it is always OK to swap finite sums!),
a final application of the strong triangle inequality yields
XX XX
ai,j − ai,j ≤ ϵN +1 .
i j j i
107
P P P P
Send N to infinity to conclude that i j ai,j = j i ai,j .
pk pj
8.98 By Faulhaber’s formula, Skp(p) = k+1 + Bk + 0<j<k kj Bk−j j+1
P
.
Rearranging,
1p−1|k X k pj Sk (p) + 1p−1|k pk
Bk + + Bk−j = − .
p j j+1 p k+1
0<j<k
8.99 Suppose the claim is false and let k be the smallest positive integer for
which Bk + p−1 1p−1|k ∈ / Z(p) . Let S be the sum on j appearing in Problem
8.98, so that Bk + p−1 1p−1|k + S ∈ Z(p) . For use momentarily, we rewrite
X k pj−1
S= pBk−j .
j j+1
0<j<k
By the minimality of k, we have Bk−j + p−1 1p−1|k−j ∈ Z(p) for all j in the
j−1
range 0 < j < k. Hence, pBk−j ∈ Z(p) for all these j. Furthermore, pj+1 ∈ Z(p)
in this same range of j: To see why, note that pj ≥ 3j > j + 1, so that
j−1
vp (j + 1) < j. Since vp (j + 1) is an integer, vp (j + 1) ≤ j − 1, and vp ( pj+1 ) ≥ 0.
It follows that each term in our rewritten expression for S belongs to Z(p) , so
that S itself belongs to Z(p) . But if S ∈ Z(p) and Bk + p−1 1p−1|k + S ∈ Z(p) ,
then Bk + p−1 1p−1|k ∈ Z(p) . This contradicts the choice of k.
8.100 This is similar to Problem 8.99. Assuming the claim fails, let k be the
smallest even positive integer for which Bk + 12 ∈
/ Z(2) . From Exercise 8.98,
1
Bk + 2 + S ∈ Z(2) , where
108
X k 2j−1
S= 2Bk−j .
j j+1
0<j<k
Let us argue that each term in this expression for S is in Z(2) (and thus,
S ∈ Z2 ). Let 0 < j < k. If j is odd, then Bk−j = 0 unless j = k − 1. The
k−2
k
2B1 2 k = −2k−2 , which is indeed in Z(2) . (We
j = k − 1 term of S is k−1
recalled here that B1 = − 12 .) If j is even, the minimality of k ensures that
j−1
2Bk−j ∈ Z(2) . As 2j+1 is also in Z(2) (clear, as j + 1 is odd), the jth term is
2-integral in the even case as well.
Since S ∈ Z(2) and Bk + 12 + S ∈ Z(2) , we are forced to have Bk + 1
2 ∈ Z(2) .
This contradicts the choice of k.
8.101 Let k be a positive even integer. From the last two problems we have
that for every prime p,
1p−1|k
Bk + ∈ Z(p) .
p
Put X 1
B̂k := Bk + .
p
p: p−1|k
More recent work on the distribution of the pD can be found in the article [3] of
Pomerance and Wagstaff; among other things, they re-prove (in stronger form) a
109
theorem of Sunseri that 6 is the most popular denominator among even-indexed
Bernoulli numbers.
for each n = 1, 2, 3, . . . . (We may appeal to Euler’s theorem here since the
unit groups of Zp mod powers of pZp are “the same” as the unit groups of Z
mod powers of p, by Problem 5.62.) Hence, |xn+1 − xn |p ≤ p−n−1 for each n,
and {xn } is a Cauchy sequence in Zp .
Let x = lim xn . Since lim xn+1 = lim xpn = (lim xn )p = xp , and lim xn+1 =
lim xn = x, it follows that xp = x.
n n−1
By Fermat’s little theorem, each xn = up ≡ up ≡ · · · ≡ up ≡ u (mod pZp ),
Therefore, x ≡ u (mod pZp ) (cf. the solution to Problem 7.81), and in partic-
ular, x ̸= 0. Thus, we have shown that ω(u) := x is a solution to ω(u)p−1 = 1
with ω(u) ≡ u mod pZp .
To prove ω(u) is the unique (p−1)th root of unity congruent to u mod pZp , it is
enough to argue that no residue class mod pZp contains two different (p − 1)th
roots of unity. This is easy: The (p − 1)th roots of unity ω(1), . . . , ω(p − 1)
belong to different residue classes mod pZp . And these are all of the (p − 1)th
roots of unity, since the degree p − 1 polynomial xp−1 − 1 cannot have more
than p − 1 roots in Qp .
110
Lemma. Let F be a field of characteristic not equal to 2. If ζ is an element
of order 3 in F × , then 1 + ζ is an element of order 6.
Now assume that u ∈ Z has order 3 mod p. There are no elements of order 3
in F×
2 . Hence, p is odd, and we can invoke our lemma to deduce that 1 + u
has order 6 in F×
p . In particular, p ≡ 1 (mod 6), as the order of each element
in F×
p necessarily divides p − 1; this congruence will be needed shortly.
Here ω(u3 ) is the unique (p − 1)th root of unity congruent to u3 mod pZp . But
1 is a (p − 1)th root of unity, and 1 ≡ u3 (mod pZp ). Thus, 1 = ω(u3 ) = ω(u)3 .
If ω(u) = 1, then u ≡ ω(u) ≡ 1 (mod pZp ), contradicting that u has order 3
modulo p. So ω(u) is an element of order 3 in Q× p and, applying the lemma a
second time, 1 + ω(u) is an element of order 6.
In particular, 1 + ω(u) is a (p − 1)th root of unity (since 6 | p − 1). Furthermore,
1 + ω(u) ≡ 1 + u (mod pZp ). Therefore, 1 + ω(u) = ω(1 + u).
111
References
1. M. Baker, Primitive roots, discrete logarithms, and p-adic numbers. URL: https:
//mattbaker.blog/2013/11/07/primitive-roots-discrete-logarithms-and
-the-interplay-between-p-adic-and-complex-numbers/
2. P. Erdős and S. S. Wagstaff, Jr., The fractional parts of the Bernoulli numbers.
Illinois J. Math. 24 (1980), 104—112.
3. C. Pomerance and S. S. Wagstaff, Jr., The denominators of the Bernoulli numbers.
Acta Arith. 209 (2023), 1–15.
112
Solutions to Set #9
9.105 Let a ∈ Z× 2 2
p . If a = x for some x ∈ Qp , then |x|p = |a|p = 1. Therefore,
|x|p = 1; in particular, x ∈ Zp . Reducing the equation x2 = a modulo the ideal
pZp shows that a mod pZp is a square in Zp /pZp . Now recall from Exercise
5.62 that the inclusion of Z into Zp induces an isomorphism between Z/p and
Zp /pZp and that this isomorphism identifies a1 mod p with a mod pZp . Since
a mod pZp is a square in Zp /pZp , it follows that a1 mod p is a square in Z/p.
Conversely, suppose a ∈ Z× p and that a1 mod p is a square in Z/p. Then
a mod pZp is a square in Zp /pZp and we can choose r1 ∈ Zp with r12 ≡ a
(mod pZp ). We construct a Zp -solution to x2 = a following the iterative
procedure of Problem 7.91.
Let k ∈ Z+ . Suppose we have a mod pk Zp -solution rk mod pk Zp to x2 = a;
we lift this to a mod pk+1 Zp solution. Expanding, (rk + pk q)2 ≡ rk2 + 2pk rk q
(mod pk+1 Zp ). The right-hand side is congruent to a mod pk+1 Zp precisely
when
2pk rk q ≡ a − rk2 (mod pk+1 Zp ).
By construction, a − rk2 ∈ pk Zp , so we can rewrite the last congruence as
a − rk2
2rk q ≡ (mod pZp ).
pk
113
Zp . By construction, |rk2 − a|p ≤ p−k . Thus, rk2 → a. Since rk2 also tends to x2 ,
we conclude that x2 = a.
a1
9.106 By Problem 9.105, the map ϕ : Z×
p → {±1} that sends a to p is a
homomorphism with kernel (Z× 2
p ) . This homomorphism is onto, as 1 is sent
to 1 and any n ∈ Z with np = −1 is sent to −1. Therefore, the two cosets
of (Z× 2 ×
p ) in Zp are ϕ
−1
(1) = (Z× 2
p ) and ϕ
−1
(−1) = n(Z× 2
p ) , establishing the
first claim.
Every x ∈ Q× v
p has a unique representation in the form p u where v ∈ Z and
u ∈ Zp . Identifying p u with (v, u) sets up an isomorphism Q×
× v ∼ ×
p = Z × Zp ,
× × 2 ∼ × × 2
which after quotienting by squares becomes Qp /(Qp ) = Z/2 × Zp /(Zp ) .
In this last isomorphism, 1, p, n, np are sent to the four distinct elements of
Z/2×Z× × 2 × × 2
p /(Zp ) . Hence, 1, p, n, and np are coset representatives for Qp /(Qp ) .
2k q ≡ 2b − (rk2 + rk ).
114
9.108 By Problem 9.107, the homomorphism from Z× ×
2 to (Z2 /8Z2 ) send-
× 2
ing a to a mod 8Z2 has kernel (Z2 ) . As this homomorphism is easily seen
× 2 ∼
to be surjective, Z×2 /(Z2 ) = (Z2 /8Z2 )
×
= {1 mod 8Z2 , 3 mod 8Z2 , 5 mod
8Z2 , 7 mod 8Z2 }. Hence, 1, 3, 5, and 7 are coset representatives for (Z2 /8Z2 )× .
Noting that each element of Q× 2 has a unique representation in the form 2 u,
v
×
with v ∈ Z and u ∈ Z2 , we conclude that
1, 3, 5, 7, 2 · 1, 2 · 3, 2 · 5, 2 · 7
9.109 Let a ∈ Q× v × n
p and write a = p u with u ∈ Zp . If a = x for some
x ∈ Qp , then n | nvp (x) = vp (a) = v. For this to hold for infinitely many n
requires
√ v = 0, so that a = pv u = u ∈ Z× p . This proves the “⇐=” implication:
If n a ∈ Qp for infinitely many n, then a ∈ Z× p.
Now suppose that a ∈ Z× p . It suffices to show that if ℓ is any prime not dividing
p(p − 1), then a is an ℓth power in Qp .
Since ℓ ∤ p − 1 = #(Zp /pZp )× , the ℓth power map is an automorphism
of (Zp /pZp )× . So we can find an r1 ∈ Z× ℓ
p with r1 ≡ a mod pZp . We take
r1 mod pZp as our starting point in the method of successive approximation.
Assume we know a mod pk Zp -solution to xℓ = a, say rk mod pk Zp . We seek a
mod pk+1 Zp -solution rk+1 mod pk+1 Zp . Expanding, (rk +pk q)ℓ ≡ rkℓ +ℓrkℓ−1 pk q
(mod pk+1 Zp ). For the right-hand expression to agree with a mod pk+1 , we
require that ℓrkℓ−1 pk q ≡ a − rkℓ (mod pk+1 Zp ), or equivalently
a − rkℓ
ℓrkℓ−1 q ≡ (mod pZp ).
pk
115
F (X) − F (Y ) ≡ F (Y ) − F (Y ) ≡ 0 (mod (X − Y )R[X, Y ]).
Therefore, |F (n)|−1 −d
|ak |∞ )−1 . (The definition of c makes
P
∞ ≥ c|n|∞ for c := ( k
sense since the ak are not all zero.)
Pd
Extra Exploration 23. Let A(T ) = ak T k ∈ Zp [T ]. It is easy to prove (and
k=0
we essentially did this in our solution to Problem 9.110) that |A(x+h)−A(x)|p ≤ |h|p
whenever x, h ∈ Zp . Show that for x ∈ Zp and h ∈ pZp , this bound can be refined to
116
9.111 That α := k≥1 pk! ∈ Zp follows from Problem 8.92. Suppose for a
P
contradiction that α is a root of the nonconstant polynomial F (T ) ∈ Q[T ].
We can assume F (T ) is irreducible over Q and, by clearing denominators, that
F (T ) ∈ Z[T ].
If F has an integer root n0 , irreducibility over Q forces F to be linear with n0
as its only root. In that case, α = n0 ∈ Z. But this is absurd: α’s canonical
expansion is not eventually periodic, so that α is not even rational, let alone a
rational integer.
We can therefore apply the result of Exercise 9.110. Let Pcn be the constant
k!
associated with F in that problem. For each n, let rn = k=1 p ∈ Z. Then
|rn − α|p = | k>n pk! |p ≤ p−(n+1)! . Also, |rn |∞ ≤ 2pn! ; here we use that the
P
largest term in the sum defining rn is pn! and that each term in that sum is at
least twice the preceding one. Therefore,
Rearranging,
pn!(d−(n+1)) ≥ c · 2−d .
But the right-hand side is a positive quantity independent of n, while the
left-hand side tends to 0 as n tends to infinity. Contradiction!
9.112 Given x ∈ Z10 , define the Z2 and Z5 -reductions of x as the first and
second components of the image of x under the isomorphism Z10 ∼ = Z2 ×
Z5 described in the solution to Problem 5.63. Concretely, if x = (a1 mod
10, a2 mod 102 , . . . ), its Z2 -reduction is (a1 mod 2, a2 mod 22 , . . . ), and its Z5 -
reduction is (a1 mod 5, a1 mod 52 , . . . ).
Convergence in Z10 can then be defined in terms of convergence in Q2 and Q5 .
If {xn } is a sequence of elements of Z10 , and x ∈ Z10 , we say xn → x in Z10 if
the Z2 and Z5 -reductions of the xn converge to the Z2 and Z5 -reductions of x
(in Q2 and Q5 ).
With this definition of convergence in place, your work on Set #7 can be
adapted toPshow that every element of Z10 has a unique, convergent 10-adic
expansion k≥0 dk · 10k with each dk ∈ {0, 1, 2, . . . , 9}. (Check this!)
n n
Clearly, 24·5 → 0 in Z2 . By Exercise 7.81, 24·5 → 1 in Z5 . So the Z10 limit
n
x of 24·5 corresponds, underP our isomorphism, to (0, 1) ∈ Z2 × Z5 . Write the
10-adic expansion of x as k≥0 dk · 10k and assume for a contradiction that
the sequence {dk } is eventually periodic, say dk = dk+ℓ for all k ≥ k0 . Working
in Z2 , we find that the Z2 -reduction of x is
117
X X
10k + dk · 10k (1 + 10ℓ + 102ℓ + . . . )
0≤k<k0 k0 ≤k<k0 +ℓ
X X dk · 10k
= 10k + .
1 − 10ℓ
0≤k<k0 k0 ≤k<k0 +ℓ
Let r be the rational number defined by the right-hand side. The exact
samePcalculation shows that the Z5 -reduction of x is also r. It follows that
x = k≥0 dk ·10k is the element of Z10 corresponding under our isomorphism to
(r, r) ∈ Z2 ×Z5 . But we saw already that x corresponds to (0, 1). Contradiction!
Extra Exploration 24. Say that the infinite sequence of decimal digits d0 , d1 ,
d2 , d3 , . . . is self-squaring if, for every k ∈ Z+ ,
For example, the two sequences 0, 0, 0, 0, . . . (all zeros) and 1, 0, 0, 0, . . . (all zeros
past the first term) are trivially self-squaring.
(a) Prove that there is a unique self-squaring sequence starting with d0 = 6.
(b) Show that the sequence in (a) begins 6, 7, 3, 9, 0, 1, 7, 8, 7, 1. (As a spot check,
762 = 5776, while 1093762 = 11963109376.)
(c) How many self-squaring sequences are there?
73990485815519399724778909194186633 = . . . 990485815519399724778909194186633.
For more on this theme, see the papers of Jiménez Urroz & Yebra [2] and Germain [1].
118
Extra Exploration 27. Check that the Strassmann series form a subring of
Qp [[T ]]. This ring is commonly denoted Qp ⟨T ⟩ and referred to as the Tate algebra in
one variable over Qp . Show that for each z ∈ Zp , the evaluation map evalz : Qp ⟨T ⟩ →
Qp sending F (T ) to F (z) is a ring homomorphism.
X m
X
an xn ≤ max |an xn |p ≤ |x|m+1
p < δ|x|m m
p = |am x |p = an xn .
n>m
n>m p n=0 p
Therefore,
m
X X m
X X
|F (x)|p = an xn + a n xn ≥ an xn − an xn > 0,
n=0 n>m n=0 p n>m p
p
so that F (x) ̸= 0.
As a consequence, if F (T ) is Strassmann and not the zero series in Qp [[T ]],
then F (pm ) is nonzero for all sufficiently large m (specifically, for any m with
p−m < δ). Hence, a Strassmann series that vanishes everywhere on Zp must
be the zero series.
9.115 Substituting and applying the binomial theorem, we find that whenever
x, x0 ∈ Zp ,
X XX k j k−j
F (x + x0 ) = ak (x + x0 )k = 1k≥j ak x x0 .
j
k≥0 k≥0 j≥0
119
j j
P P
Since j≥0 bj x converges for all x ∈ Zp (to F (x + x0 )), j≥0 bj T is a
Strassmann series.
By Faulhaber’s formula,
p−1 p−1
pj
X Sp−1 (p) p − 1 1 X p−1
qp (a) = − = Bp−1 + − 1 + Bp−1−j .
a=1
p p p j=1
j j+1
120
For each j = 1, 2, 3, . . . , p−1, the Bernoulli number Bp−1−j ∈ Z(p) ; this is clear
for j = p − 1 (recall that B0 = 1) and for j < p − 1 it follows from Exercise
8.99. In this same range of j, we have pj ≥ 3j > j + 1. Thus, vp (j + 1) < j,
pj
and vp ( j+1 ) = j − vp (j + 1) > 0. Consequently, each term in our sum on j
belongs to pZ(p) , implying that the sum itself belongs to pZ(p) . But that sum
Pp−1
is precisely a=1 qp (a) − (Bp−1 + p1 − 1).
Using once more that 1 + ω(u) = ω(u + 1), we deduce that (1 + v)p ≡ 1 + v p
(mod p2k+1 Zp ). Since both (1 + v)p and 1 + v p lie in Z, this last congruence
in fact holds modulo (p2k+1 Zp ) ∩ Z = p2k+1 Z.
Let p = 7. In the course of solving Problem 8.103, we computed that ω(2) ≡ 324
(mod 73 Z7 ). Taking v = 324 and k = 3 in the result of the last paragraph
“explains” the given example.
References
121
122
Solutions to Set #10
where the divisibility relations are being asserted in the ring Z. Naturally,
these same relations also hold in Zℓ , so that working in Zℓ modulo ℓvℓ (k!) Zℓ ,
1 1 1
− 1 ··· − (k − 1) ≡ A(A − 1) · · · (A − (k − 1)) ≡ 0.
2 2 2
1
It follows that vℓ ( 12 ( 12 − 1) · · · ( 12 − (k − 1))) ≥ vℓ (k!), so that vℓ ( k2 ) ≥ 0. As
1
this holds for all odd primes ℓ, the rational number k2 has denominator a
power of 2.
To determine which power of 2, we look 2-adically. Expanding out,
1
2 1 · (1 − 2) · (1 − 2 · 2) · · · (1 − 2(k − 1))
= .
k 2k k!
The numerator on the right is odd, and so the power of 2 in the denominator
1
of k2 is 2k+v2 (k!) .
1
Collecting what we know so far: k2 is a rational number with denominator
2k+v2 (k!) .
We are now well-positioned to decide when B 12 (x) converges — equivalently,
1
when | k2 xk |p → 0 (as k → ∞). For use momentarily, we recall that v2 (k!) ≤ k,
so that k + v2 (k!) ≤ 2k.
123
1 1
Suppose p is odd. Then | k2 |p ≤ 1, and | k2 xk |p ≤ |x|kp . If |x|p ≤ 1/p, then
|x|kp → 0 as k → ∞. Thus, B 12 (x) converges for these x.
1
Suppose instead that p = 2. If x ∈ Q2 , then | k2 xk |2 = 2k+v2 (k!) |x|k2 ≤ 22k |x|k2 .
This last expression tends to 0 if |x|2 ≤ 1/23 .
These conditions on x turn out to be necessary for convergence.
10.120 When x is real and |x| < 1, the series B 12 (x) converges to the positive
9
square root of 1 + x. So over the real numbers, B 12 ( 16 ) = 54 .
1
Suppose now that p is an odd prime and that x ∈ pZp . Recalling that k2 ∈ Zp
1
for each k = 0, 1, 2, . . . , we find that |B 12 (x) − 1|p ≤ maxk≥1 | k2 xk |p ≤ 1/p.
Thus, B 21 (x) ∈ 1 + pZp .
9
When p = 3 and x = 16 , the (unique) square root of 1 + x belonging to 1 + 3Z3
is − 54 .
124
10.121 These results are not particular to Zp and Qp . Let D be any domain of
characteristic 0 with fraction field K. (For example, we could take D = Zp and
K = Qp .) Then Taylor’s formula holds for polynomials over K: If F (T ) ∈ K[T ]
and a ∈ K, then
X F (j) (a)
F (a + T ) = T j.
j!
j≥0
For the proof, fix a ∈ K. Consider K[T ] as a K-vector space and observe that
both sides of the claimed equation represent K-linear functions of F (T ) ∈ K[T ].
So the identity will be established if it is shown for F (T ) = 1, T, T 2 , . . . (a
K-basis for K[T ]). When F (T ) = T n , the left-hand side is (a + T )n while the
right is
X n(n − 1)(n − 2) . . . (n − (j − 1)) n−j j X n
1n≥j a T = an−j T j .
j! j
j≥0 0≤j≤n
10.122 Taylor’s formula gives F (x̃) = j≥0 j!1 F (j) (x)(−F (x)/F ′ (x))j . The
P
terms of the sum corresponding to j = 0 and j = 1 cancel each other out
(being F (x) and −F (x), respectively), so that
X F (j) (x) j j
F (j) (x)
F (x) F (x)
|F (x̃)|p = − ′ ≤ max − ′ .
j! F (x) j≥2 j! F (x)
j≥2 p
p
125
Furthermore, xn+1 ≡ xn (mod pZp ), implying F ′ (xn+1 ) ≡ F ′ (xn ) ̸≡ 0
(mod pZp ), so that
|F ′ (xn+1 )|p = 1.
Thus, the hypotheses we originally assumed for xn hold for xn+1 , allowing us
to reboot the procedure with xn+1 replacing xn .
Starting from the given x1 , we produce in this way a sequence {xn } of elements
of Zp satisfying
n−1
|F (xn )|p ≤ |F (xn−1 )|2p ≤ · · · ≤ |F (x1 )|2p (*)
for each n = 1, 2, 3, . . . . As |F (x1 )|p < 1, (*) guarantees that F (xn ) converges
(rapidly!) to 0.
Observe that |xn+1 − xn |p = |F (xn )/F ′ (xn )|p = |F (xn )|p , for every n. There-
fore, {xn } is Cauchy, and xn → x for some x ∈ Zp . Hence, F (x) = F (lim xn ) =
lim F (xn ) = 0. (We use here that polynomials preserve limits, which follows
from Exercise 3.29.)
argue that x ≡ x1 (mod pZp ). This is immediate from the
It remains only toP
identity x − x1 = k≥1 (xk+1 − xk ) expressing x − x1 as a sum of terms from
pZp .
10.124 Express the Q[T ]-ideal F (T )Q[T ] + F ′ (T )Q[T ] as D(T )Q[T ], where
D(T ) ∈ Q[T ]. Since F (T ), F ′ (T ) ∈ D(T )Q[T ], each complex root of D(T ) is a
common zero of F (T ) and F ′ (T ) — in other words, a multiple root of F (T ).
We are given that F (T ) has distinct complex roots. Therefore, D(T ) has
no complex roots, meaning that D(T ) is a nonzero constant. As a result,
F (T )Q[T ] + F ′ (T )Q[T ] = D(T )Q[T ] = Q[T ], and there are X(T ), Y (T ) ∈
Q[T ] with F (T )X(T ) + F ′ (T )Y (T ) = 1. Choose R ∈ Z+ so that X̂(T ) :=
RX(T ), Ŷ (T ) := RY (T ) ∈ Z[T ]. Then
F (T )X̂(T ) + F ′ (T )Ŷ (T ) = R.
Taking this last equation mod p, we find that the mod p reductions of F (T )
and F ′ (T ) generate the unit ideal in (Z/p)[T ] whenever p ∤ R. Hence, F (T )
and F ′ (T ) are coprime in (Z/p)[T ] for each prime p not dividing R.
10.125 According to Problem 2.24, there are infinitely many primes p for
which F has a root in Fp . By Problem 10.124, there are only finitely many
126
primes p for which F and F ′ have a common root in Fp . So for infinitely many
p, we can find an x1 ∈ Z with F (x1 ) ≡ 0 (mod p) and F ′ (x1 ) ̸≡ 0 (mod p).
For each of these primes, F has a root in Zp by Problem 10.123.
ϕ(xn ) − ϕ(x) = ϕ(xn − x) = ϕ(pvp (xn −x) )ϕ(un ) = pvp (xn −x) u′n
127
|ϕ(xn ) − ϕ(x)|p = p−vp (xn −x) = |xn − x|p .
Of course, the equation |ϕ(xn ) − ϕ(x)|p = |xn − x|p also holds when xn − x = 0.
We are assuming xn → x. Therefore, |ϕ(xn ) − ϕ(x)|p = |xn − x|p → 0. Hence,
ϕ(xn ) → ϕ(x), as desired.
10.128 Suppose first that p and q are odd primes with q ̸= p. Select n ∈ Z with
n
q = −1, and choose a ∈ Z with a ≡ 1 (mod p) and a ≡ n (mod q). Then a
is a square in Qp but not a square in Qq (see Problem 9.105). Since a ∈ Z, any
homomorphism from Qp to Qq must send a to a. However, homomorphisms
always send squares to squares, so no homomorphism from Qp to Qq can exist.
If p = 2 and q is odd, the same argument works if we select a ≡ 1 (mod 8)
and a ≡ n (mod q) (see Problem 9.107). If p is odd and q = 2, choose a ≡ 1
(mod p) and a ≡ 3 (mod 8).
The same argument works to demonstrate the impossibility of a homomorphism
from Qp to R. If p is odd, select a with a ≡ 1 (mod p) and a < 0. If p = 2,
pick a ≡ 1 (mod 8) with a < 0.
References
128
Solutions to Set #11
11.130 We assume |x − 1|p < 1 and argue that the terms of the series defining
logp x tend to
Q 0. For each positive integer k, the product formula yields
|k|−1
p = |k|∞ prime q̸=p |k|q ≤ |k|∞ = k. Therefore,
(x − 1)k
(−1)k−1 = |x − 1|kp · |k|−1 k
p ≤ k · |x − 1|p ,
k p
11.131 We start by noting that if x ∈ 1 + pZp , then x−1 ∈ 1 + pZp . This is not
difficult to prove directly, futzing about with absolute values, but a more elegant
approach is to observe that (1 + rp)−1 = 1 − rp + r2 p2 − · · · ∈ 1 + pZp whenever
r ∈ Zp . Utilizing the addition rule (†), we deduce that for all x ∈ 1 + pZp ,
∞
X (−1)k−1
1
logp x + logp = logp 1 = (1 − 1)k = 0.
x k
k=1
1
(That logp x + logp x = 0 should not be surprising, but it ought to be comfort-
ing!)
If j ∈ {1, 2, . . . , p − 1}, then
∞ −1 !
1 pk
X p p j
= − logp 1 − = logp 1− = logp .
jk k j j j−p
k=1
Therefore,
p−1 p−1 X
∞ ∞
! p−1
k
X X 1 p =
X 1 pk X j
k k
= logp
j=1
j k j=1
j k j=1
j−p
k=1 k=1
Y j
= logp = logp ((−1)p−1 ).
0<j<p
j − p
129
Here the interchange of the sums on j and k is routine to justify, for example
by setting ak,j = 10<j<p j −k pk k −1 and applying the result of Problem 8.96.
If p is odd, then (−1)p−1 = 1, and logp 1 = 0 gives the desired result. To
finish off, we must show log2 (−1) = 0. But this is easy: log2 (−1) + log2 (−1) =
log2 ((−1)2 ) = log2 (1) = 0.
Remarks.
(i) You may have conjectured in Problem 4.45 thatPv2 ( n k
P
k=1 2 /k) is bounded below
n k
by n − (something small). Once we know that k=1 2 /k = 0, this becomes easy
to show. Indeed,
n ∞
X 2k X 2k 2k k n+1
= − ≤ max ≤ max = n+1 .
k k k≥n+1 k 2
k≥n+1 2k 2
k=1 2 k=n+1 2
(To estimate |2k /k|2 we used the displayed inequality in the solution to Problem
11.130, for x = −1.) Converting this from a statement about absolute values to
log(n+1)
one about valuations, v2 ( n k
P
k=1 2 /k) ≥ (n + 1) − log 2
. A slight modification
of this argument will show that equality holds when n + 1 is a power of 2.
(ii) In the next Extra Exploration, we outline one proof of the addition law (†).
Arguments for (†) based on different principles can be found in the books of
Gouvêa [2, §5.7] and Robert [10, Chapter 5, §4].
Our proof of (†) is motivated by the following formula from (real) calculus: For
each real number x > 0,
√
log x = lim k( k x − 1).
k→∞
d t
(Check this yourself, possibly starting from dt x t=0 = log x.) This identity is
surprisingly powerful; Landau shows in [5] that all of the familiar theory of the
natural logarithm can be developed taking the right-hand side as the definition
of log x. We will establish and apply a p-adic analogue of this relation.
The rest of this problem is devoted to showing that Lp and logp coincide; once this
is known, (†) follows from (c).
1 pm
(d) For x ∈ 1 + pZp and m ∈ Z+ : (x − 1)k .
P
Lp,m (x) = k≥1 pm k
130
1 p m (−1)k−1 pm
(e) For m, k ∈ Z+ :
Q
pm k
= k 1≤j<k 1− j
. Therefore, Lp,m (x) =
logp (x) + ep,m (x), where
X (−1)k−1 Y p m
ep,m (x) = 1− − 1 (x − 1)k .
k j
k≥2 1≤j<k
(f) For each x ∈ 1 + pZp , we have limm→∞ ep,m (x) = 0. Thus, Lp (x) = logp x.
− pm /j) is congruent to 1 modulo a large
Q
Hint. If k is small in terms of m, then 1≤j<k (1
power of pZp . On the other hand, when k is large, vp ((x − 1)k /k) is also large.
11.132 Quite a lot of what is claimed here is true — just none of the important
bits!
T 2k+1
Let p be an odd prime and define sinp (T ) = k≥0 (−1)k (2k+1)!
P
∈ Qp [[T ]]. It is
straightforward to check that sinp x converges when |x|p ≤ 1/p; consequently,
sinp (pa) is a well-defined element of Qp for all a ∈ Z. It is also true that if
a is an integer not divisible by p, then sinp (ap) ̸= 0. Otherwise the display
following “Therefore” indeed lays out a contradiction.
Unfortunately, none of this has much to do with the real number π! Let’s
assume π = ab , P
with a, b ∈ Z, b ̸= 0. Choose an odd prime p not dividing a,
Then the series k≥0 (−1)k (ap)k /(2k + 1)! converges to sin(pbπ) = 0 in R. So
far, so good. But the fact that a series converges to 0 in R and converges to
something in Qp — something that could (misleadingly?) be labeled sinp (pbπ)
— doesn’t imply it converges to 0 in Qp . (Cf. Exercise 10.120.)
Remark. An error of a similar nature crept into the work of Hensel himself. In 1905,
Hensel claimed to prove that [Qp (e) : Q] = p for each odd prime p, from which he
derived the corollary that e is transcendental over Q [3]. While the transcendence of
e had already been shown by Hermite in 1873, Hensel’s reasoning was much shorter
and simpler; his proof has an unmistakable air of elegance. Unfortunately, it is also
fundamentally flawed.∗ See [12] and [9, §5.6] for discussion (and compare with [4,
Exercise 9, p. 84]).
∗
“For every complex problem there is an answer that is clear, simple, and wrong.” —
H. L. Mencken
131
j j j
Since pk | ((xk + pk h)2 − x2k ) and p | p2 , the sum on j belongs to pk+1 Zp .
Hence, F (xk + pk h) ≡ F (xk ) + pk h (mod pk+1 ), and F (xk + pk h) ≡ 0
(mod pk+1 ) provided we choose h ≡ −F (xk )/pk (mod pZp).
If x1 , x2 , x3 , . . . are constructed as above, then the {xn } form a Cauchy se-
quence in Zp , so that xn → x for some x ∈ Zp . For each n,
Since the right-hand sum is divisible by p, the only way the p-adic absolute
values of both sides can agree is if x′ − x = 0, so that x′ = x.
11.133 (yes, déjà vu all over again!) We will factor F (T ) so as to make obvious
that F (x) = 0 for a unique x ∈ Zp . The factors are constructed by a variant
of the method of successive approximation.
The only property of F (T ) we will use is that F (T ) = 1 + T + pG(T ), where
G(T ) ∈ Zp [[T ]] is a Strassmann series. Since G(T ) is Strassmann, for any
k ∈ Z+ the power series F (T ) ∈ Zp [[T ]] is congruent modulo pk Zp [[T ]] to a
polynomial with Zp coefficients. This will be crucial.
We begin by locating r1 ∈ Zp and q1 (T ) ∈ Zp [T ] with
F (T ) − (1 + T )
(1 + T )q1 (T ) + r1 ≡ (mod pZp [[T ]]). (*)
p
132
(1 + T )q̃1 (T ) + r̃1 in (Zp /pZp )[T ]. Then (*) holds if we choose q1 (T ) ∈ Zp [T ]
and r1 ∈ Zp to lift q̃1 (T ) and r̃1 .
Continuing, suppose we have already found r1 , r2 , . . . , rk ∈ Zp and q1 (T ),
q2 (T ), . . . , qk (T ) ∈ Zp [T ] with
k
X k
X
1+T + pj rj 1+ pj qj (T ) ≡ F (T ) (mod pk+1 Zp [[T ]]).
j=1 j=1
F (T ) = U (T )V (T ),
∞
X ∞
X
where U (T ) := 1 + T + pj rj , V (T ) := 1 + pj qj (T ).
j=1 j=1
P∞
There’s one problem with this proposal. While j=1 pj rj is a well-defined
element ofPZp (the sum being obviously convergent), it is not clear what is
meant by j=1 pj qj (T ), an infinite sum of elements of Zp [T ].∗ Fortunately
∞
∗
This should be a power series over Zp . But its precise interpretation requires some
care: Convergence in power series rings is usually defined by the requirement that
the coefficient on each fixed power of T stabilizes, i.e., is eventually constant. But
that is not the intended meaning here.
133
this difficulty is not so serious: For each fixed nonnegative integer r, the sum
r
of the TP∞coefficients of pj qj (TP
) converges, to γr ∈ Zp (say), and we simply
j r
define j=1 p qj (T ) to mean r≥0 γr T . Having made this definition, the
factorization F (T ) = U (T )V (T ) follows easily (check that the two sides are
congruent modulo pk Zp [[T ]], for every k, and convince yourself that this implies
equality in Zp [[T ]]).
So far we have shown F (T ) = U (T )V (T ), as formal power series.
P∞ As each qj (T )
is a polynomial, a moment’s thought shows that V (T ) = 1 + j=1 pj qj (T ) is
a Strassmann series. Invoking Problem 8.94, F (x) = U (x)V (x) for all x ∈ Zp .
Since V (x) ∈ 1 + pZp for each x ∈ Zp , the Zp -zeros of F are precisely the
P∞ of U . But it is obvious U (x) = 0 for a unique x ∈ Zp , namely
same as those
x = −1 − j=1 pj rj .
Remark. This took quite a bit longer than our first approach! However, if you
understand this argument, you are well on your way to proving (one form of) the
Weierstrass preparation theorem for Qp . See the remark after the solution to Problem
13.159.
11.134 We explain how to transform the double sum on k and j into the
k
claimed sum on j. Put uk,j = 1k≥j s(k, j)nj ak! . Since p is odd and |a|p ≤ 1/p,
134
ak
1k≥j s(k, j)nj ≤ 1k≥j p−k |k!|−1
p ≤ 1k≥j p
−k k/(p−1)
p ≤ 1k≥j p−k/2 .
k! p
k
So if we set ϵN := p−N/2 , then |1k≥j s(k, j) ak! |p ≤ ϵN whenever k ≥ N or
j ≥ N . By Problem 8.96,
∞ k ∞ X ∞ ∞ X ∞
X X ak X X
s(k, j)nj = uk,j = uk,j
j=0
k!
k=0 k=0 j=0 j=0 k=0
∞ k ∞
X X a X
= nj s(k, j) = Ca,j nj .
j=0
k! j=0
k≥j
P∞ P∞
When n = 1, this argument shows that j=0 Ca,j = 1 + a. Since j=0 Ca,j
converges, |Ca,j |p → 0 as j → ∞.
11.136 Computing the determinant by expanding along the first column gives
det(A) = ±ad . Therefore, A is invertible over any ring in which a−1
d exists,
such as Fp when p ∤ ad .
11.137 We have
11.138 Let s(N, K) be the Stirling numbers of the first kind. For each non-
negative integer m, Problem 11.137 gives
135
∞
X m(m − 1)(m − 2) · · · (m − (j − 1))
xkm+r = pj ⟨Ar B j v, e⟩
j=0
j!
∞ j
!
X pj X
= s(j, ℓ)mℓ ⟨Ar B j v, e⟩.
j=0
j!
ℓ=0
j
If we put uj,ℓ = 1ℓ≤j pj! s(j, ℓ)mℓ ⟨Ar B j v, e⟩, then |uj,ℓ |p ≤ 1ℓ≤j |j!|−1
p ·p
−j
≤
j/(p−1) −j −j/2 −N/2
1ℓ≤j p p ≤ 1ℓ≤j p . Hence, if ϵN := p , then |uj,ℓ |p ≤ ϵN when-
ever j ≥ N or ℓ ≥ N . Because ϵN → 0 as N → ∞, Exercise 8.96 allows us to
write
∞ X
X ∞ ∞ X
X ∞ ∞
X X pj
xkm+r = uj,ℓ = uj,ℓ = mℓ s(j, ℓ)⟨Ar B j v, e⟩.
j=0 ℓ=0
j!
ℓ=0 j=0 ℓ=0 j≥ℓ
So if we define
∞
X X pj
Fk,r (T ) = s(j, ℓ)⟨Ar B j v, e⟩ T ℓ ,
j!
ℓ=0 j≥ℓ
then Fk,r (m) converges to xkm+r for every nonnegative integer m. In particular,
Fk,r (1) converges, so that the coefficients of T ℓ tend to 0 as ℓ → ∞. Thus,
Fk,r (T ) is a Strassmann series.
Remarks.
(i) Imagine you are presented with an integer linear recurrence sequence, specified
by a list of coefficients a1 , . . . , ad ∈ Z (where ad ̸= 0) and a collection of initial
values x0 , . . . , xd−1 ∈ Z. The Skolem problem asks whether xn = 0 for some
nonnegative integer n. Frustratingly, while we have algorithms to decide this for
any sequence with d ≤ 4 (“algorithm” here meaning a procedure that is proven
to always terminate), no general algorithm is known for d ≥ 5. See [8] for a
discussion.
(ii) The Monthly article [7] of Myerson and Van der Poorten provides a highly
entertaining account of the Skolem–Mahler–Lech theorem, addressing several
aspects not discussed here.
136
Extra Exploration 39 (Shapiro [11], Myerson and Van der Poorten
[7]). Let {xn }n≥0 be an integer linear recurrence sequence. Call m ∈ Z a repeat
offender (relative to {xn }) if xn = m for infinitely many nonnegative integers n. Show
that each integer linear recurrence sequence is associated with only finitely many
repeat offenders.
11.140 These are easy to find once you go looking. For instance, let x0 = 0,
x1 = 1, and set xn = −xn−2 for all n ≥ 2. Then xn = 0 precisely when n is
even.
Here is a more compelling example, yoinked from [7]. Consider the integer
recurrence sequence {xn } satisfying
11.141 Yes, the same result holds for recurrence sequences defined over an
arbitrary number field K.
The proof is almost the same as that given for integer recurrences. An obvious
complication is that when K = ̸ Q, a general element of K does not carry
a preassigned meaning as an element of Qp . This obstacle is by no means
insurmountable; it is resolved by embedding K into Qp , which we know is
possible for infinitely many primes p (Exercise 10.126). Let’s suppose we have
not just an embedding but an embedding+ (TM), meaning an embedding of
K into Qp , p odd, where (identifying elements of K with their images in Qp )
a1 , . . . , ad , x0 , . . . , xd−1 ∈ Zp , and ad ∈ Z×
p . Then it is not hard to convince
yourself that our proof of Skolem’s theorem goes through nearly verbatim.
(Check this!)
We know there are always embeddings, but is there always an embedding+ ?
Yes! Since K embeds into Qp for infinitely many p, it will suffice to prove the
following lemma.
Proof. Let m(T ) be the minimal polynomial of α over Q, scaled to have integer
coefficients. Write m(T ) = cd T d + cd−1 T d−1 + · · · + c1 T + c0 . If c0 = 0, then
T divides m(T ), and the irreducibility of m(T ) over Q forces m(T ) to be a
constant multiple of T . But then α = 0, contrary to hypothesis. So c0 ̸= 0.
137
Assume now that K has been embedded into Qp , where p is a prime not
dividing cd c0 . We will show that |α|p = 1. Since only finitely many primes
divide cd c0 , the lemma follows.
Suppose instead that |α|p > 1. Since m(α) = 0, we have cd αd = −(cd−1 αd−1 +
· · · + c0 ). However (keeping in mind that p ∤ cd ),
If |α|p < 1, we argue similarly, exploiting that 1/α is a root of the reciprocal
polynomial T d m(T −1 ) = c0 T d + c1 T d−1 + · · · + cd−1 T + cd . In this situation,
c0 (1/α)d = −(c1 (1/α)d−1 + · · · + cd−1 (1/α) + cd ), and (keeping in mind that
p ∤ c0 )
d d d−1
1 1 1
≥ max |c1 (1/α)d−1 |p , . . . , |cd−1 (1/α)|p , |cd |p
c0 = >
α p α p α p
d−1 j
X 1
≥ cd−j .
j=0
α p
Remark. Lech has shown that Skolem’s theorem holds for recurrence sequences
defined over arbitrary fields of characteristic 0. Today one usually refers to this
general statement as the Skolem–Mahler–Lech theorem. In this setting the embedding
required to make the proof work is guaranteed by a lemma of Cassels (see [1, Chapter
5]): Let K be a finitely generated extension of Q and let S be a finite set of nonzero
elements of K. For infinitely many primes p, there is an embedding of K into Qp
with the property that |x|p = 1 for all x ∈ S. To prove the Skolem–Mahler–Lech
Theorem, apply this with K = Q(a1 , . . . , ad , x0 , . . . , xd−1 ) (details left to you!).
References
138
6. H.-W. Leopoldt, Zur Approximation des p-adischen Logarithmus. Abh. Math.
Sem. Univ. Hamburg 25 (1961), 77–81.
7. G. Myerson and A. van der Poorten, Some problems concerning recurrence
sequences. Amer. Math. Monthly 102 (1995), 698–705.
8. J. Ouaknine and J. Worrell, Decision problems for linear recurrence sequences.
In: Reachability Problems, Lecture Notes in Comput. Sci., vol. 7550, Springer,
Heidelberg, 2012, pp. 21–28.
9. B. Petri, Perioden, Elementarteiler, Transzendenz. Kurt Hensels Weg zu den
p-adischen Zahlen, Dr. Hut, München, 2011. URL: http://tuprints.ulb.tu-d
armstadt.de/2785/1/DissPetri.pdf
10. A. M. Robert, A course in p-adic analysis, Grad. Texts in Math., vol. 198,
Springer-Verlag, New York, 2000.
11. H. N. Shapiro, On a theorem concerning exponential polynomials. Comm. Pure
Appl. Math. 12 (1959), 487–500.
12. P. Ullrich, Der Henselsche Beweisversuch für die Transzendenz von e, in: Mathe-
matik im Wandel, bd. 1, Franzbecker, Hildesheim, 1998, pp. 320–330.
139
140
Solutions to Set #12
p
Proof. Write x = 1+pk r, where r ∈ Zp . Then xp = 1+pk+1 r+ pjk rj ≡
P
j≥2 j
1 (mod pk+1 Zp ), as desired. ❚
Now we return to the problem at hand. Since a ∈ pZp , applying the lemma m
m m
times will show (1 + a)p ∈ 1 + pm+1 Zp . Hence, |(1 + a)p − 1|p ≤ p−m−1 , so
pm
that (1 + a) → 1. Therefore, for any n ∈ Z,
m m
(1 + a)n = (1 + a)n lim (1 + a)p = lim (1 + a)n+p .
m→∞ m→∞
Extra Exploration 40. Let p be an odd prime, and let a ∈ pZp . You have just
demonstrated one way to extend (1 + a)x from a function of x defined on Z to one
defined on Zp : Set (1 + a)x := Binom(1 + a; x). But there is another approach you
could take to extend the domain of (1 + a)x to Zp . Namely, you might define (1 + a)x
as the limit of (1 + a)xn , where xn is any sequence of integers converging to x in Qp .
This should remind you of how exponentiation of real numbers is defined when the
exponent is irrational.
141
Check that this idea works (i.e., leads to a well-defined value of (1+a)x , independent of
the particular sequence xn ) but doesn’t give anything new: (1+a)x = Binom(1+a; x),
again. What happens when p = 2?
Pm
12.143 Put F (T ) = i=1 ai ·Binom(βi ; T )−A. The results of Problems
Pm 11.134
and 12.142 imply that F (T ) is a Strassmann series with F (n) = i=1 ai βin −A
for every n ∈ Z.
Every n ∈ Z satisfying a1 β1n + · · · + am βmn
= A is a zero of F in Zp . So if
this equation has infinitely many integer solutions, then F has infinitely many
zeros in Zp . In that case, F (T ) = 0 in Qp [[T ]] (by Exercise 9.116). Then
A = F (n) + A = a1 β1n + · · · + am βm
n
for all n ∈ Z.
Pm n
Remark. It is often easy to rule out that i=1 ai βi = A for all n ∈ Z. Let’s
assume, as is natural to do, that no ai = 0.
̸ 0. If m n
P
Similar reasoning can be applied Pwhen A= i=1 ai βi = A for all nonnegative
A m ai
integers n, we find that 1−T = i=1 1−βi T . This leads to the conclusion that every
element in the list 1, β1 , . . . , βm is repeated.
142
12.145 Suppose |α|, |α′ | ≤ R. Since α′′ is the complex conjugate of α′ ,
we also have |α′′ | ≤ R. By Exercise 12.144, |x| ≤ 13 (|α| + |α′ | + |α′′ |) ≤ R,
1
|y| ≤ 3|θ| (|α|+|ω 2 α′ |+|ωα′′ |) ≤ |θ|
R 1
, and |z| ≤ 3|θ| ′ 2 ′′ R
2 (|α|+|ωα |+|ω α |) ≤ |θ|2 .
So there are finitely many possibilities for the integers x, y, and z, and therefore
finitely many possibilities for α = x + yθ + zθ2 .
x0
12.147 We show that if (i) and (iii) both fail, then (ii) must hold. Since (i)
fails, we may choose an x0 ∈ G with x0 ̸= 0. Let D denote the closed disc
about 0 passing through x0 . The set (G \ {0}) ∩ D is nonempty (containing,
for example, x0 ). Since (iii) fails, this set is also finite. Let g be any element
of (G \ {0}) ∩ D minimizing the Euclidean distance to 0 (among elements of
(G \ {0}) ∩ D).
143
As g is a nonzero vector belonging to the one-dimensional subspace V :=
{(x, y) : x + y = 0} of R2 , we have that V = Rg. We leverage this observation
to show that g generates G. (Hence, (ii) holds!) It is enough to show that an
arbitrarily chosen x ∈ G belongs to ⟨g⟩. Since G ⊆ V , we may write x = tg,
where t ∈ R. Then {t}g = x − ⌊t⌋g ∈ G. (Here {t} = t − ⌊t⌋ denotes the
fractional part of t.) Since 0 ≤ {t} < 1, the vector {t}g is an element of G
closer to the origin than g. This contradicts the choice of g unless x − ⌊t⌋g = 0.
But then x = ⌊t⌋g ∈ ⟨g⟩.
12.148 In view of Problem 12.147, it is enough to show that each closed disc
about 0 intersects L(U) in a finite set.
Let D be the closed disc of radius R centered at 0. Every element of D ∩ L(U)
can be written as L(ε) where 1 ≤ ε ≤ eR and |ε′ | ≤ eR/2 . As a consequence of
Problem 12.145, there are only finitely many possibilities for ε and hence only
finitely many possibilities for L(ε).
144
bounds over C that we derived above. Those bounds imply that E(m) = 0
for only finitely many positive integers m, but they say nothing useful about
negative integers m.) Finally, letting r range from 0 to p − 1, we see that
µn + ωµ′n + ω 2 µ′′n = 0 for only finitely many n ∈ Z.
Remark. We can shorten the argument using the Remark following the solution
to Problem 12.143. If E(m) = 0 for all m ∈ Z, then each term in the list ν, ν ′ , ν ′′ is
repeated. But, working again in C,
12.151 Suppose n3 +1 has all prime factors from P. Then n3 +1 = p∈P pvp for
Q
some nonnegative integers vp . Writing each vp = 3ep + rp , where rp ∈ {0, 1, 2},
Y Y
n3 + 1 = (−D)y 3 for D= p rp , y = − pep .
p∈P p∈P
Extra Exploration 41. Keep the same assumptions and notation from the proof
of the theorem. In this exercise, you will show that for each nonzero integer k, the
equation x3 − Dy 3 = k has finitely many integer solutions x, y. Equivalently, there
are finitely many pairs of integers x, y with N(x − yθ) = k.
(a) Show that the ring Z[θ]/kZ[θ] is finite, and that in fact #Z[θ]/kZ[θ] = |k|3∞ .
(b) Prove that if κ ∈ Z[θ] has Nκ = k, then κZ[θ] ⊇ kZ[θ]. Deduce from (a) that
there are finitely many possibilities for κ, up to associates.
(c) Suppose κ ∈ Z[θ] has Nκ = k. Prove that there are finitely many u ∈ U (same
meaning as above — positive units in Z[θ]) for which uκ has the form x − yθ for
some x, y ∈ Z. Conclude!
145
Remark. As stated on Set #12, Delaunay and Nagell showed that x3 − Dy 3 = 1
has at most one integer solution with y = ̸ 0. It was discovered by Skolem [2] that
their theorem can be established by p-adic methods. Skolem’s proof uses arithmetic
in extensions of Q3 ; a more elementary p-adic proof can be found in lecture notes of
Ljunggren [1].
References
146
Solutions to Set #13
Pp−1
13.152 When p − 1 | k, every term in the sum is 1, and u=1 ω(u)k = p − 1 =
1p−1|k (p − 1).
Now suppose that p − 1 ∤ k. Recall that every finite subgroup of the multi-
plicative group of a field is cyclic. Thus, we may choose a generator ζ for the
group µp−1 = {ω(1), . . . , ω(p − 1)} of (p − 1)th roots of unity in Qp . Since
ζ (p−1)k = 1 and ζ k ̸= 1,
p−1
X p−1
X p−1
X
ω(u)k = (ζ j )k = (ζ k )j
u=1 j=1 j=1
1 − ζ (p−1)k
= ζ k (1 + ζ k + · · · + ζ (p−2)k ) = ζ k = 0,
1 − ζk
which agrees with 1p−1|k (p − 1) in this case.
This becomes the relation claimed in the problem statement after the substi-
tution kj = kj k−1
j−1 .
147
13.154 Suppose the claim is false and let k be the smallest positive integer
with βk ∈
/ Zp . We will derive a contradiction from the relation
p−1
X k − 1 pj X k − 1 pj−1 X
βk + Bk−j + uk−j ϑ(u)j = 0 (*)
j−1 j(j + 1) j−1 j u=1
0<j≤k 0<j≤k
148
P
Since k is even, the only odd integer j which contributes to 1 is j = k − 1.
j k−1 k−1
p p 1p
For this j, we have k−1
j−1 Bk−j j(j+1) = (k − 1)B1 k(k−1) = − 2 k . We already
k−1
observed
P that p k ∈ pZp , while clearly − 12 ∈ Zp (p is odd). Hence, the term
of 1 corresponding to j = k − 1 belongs to pZp . Suppose now that j is even
and 0 < j ≤ k. If 0 < j < k, the p-integrality of βk−j established in Problem
13.154 implies that
pj−1
vp = j − 1 − vp (j(j + 1)) ≥ j − 1 − vp ((j + 1)!)
j(j + 1)
j+1 j+1
>j−1− ≥j−1− > 0,
p−1 4
keeping in mind for the last step that j ≥ 2.
13.156 We are assuming that p − 1 does not divide the even integer k, so
that p ≥ 5. Since k and k ′ are congruent mod p − 1, the integer k ′ is also not
divisible by p − 1. By Problem 13.155,
p−1
Bk X
= βk ≡ − uk−1 ϑ(u) (mod pZp ),
k u=1
p−1
Bk ′ X ′
′
= β k ′ ≡ − uk −1 ϑ(u) (mod pZp ).
k u=1
Remark. Call an odd prime p regular if p does not divide the numerator of any of
B2 , B4 , . . . , Bp−3 . In the middle of the 19th century, Kummer showed that Fermat’s
last theorem for the exponent p holds for all regular primes p. That is, if p is a
regular prime, then xp + y p = z p has no solutions in integers x, y, z with xyz ̸= 0
(see the books of Ribenboim [8] and Washington [12] for accounts of this work).
All primes smaller than 37 are regular, while 37 is not: B32 = −37 · 683·305065927
2·3·5·17
.
Via Kummer’s congruence (the result of Problem 13.156), it is possible to show that
there are infinitely many irregular primes (odd primes that are not regular). We
sketch an argument for this due to Carlitz [2], which nicely illustrates how the facts
we have built up about Bernoulli numbers can be put to use.
149
Looking at the Remark following the solution to Problem 4.44, we see that |B2m |∞
tends to infinity faster than any power of m. In particular, as we will use momentarily,
| B2m
2m
|∞ > 1 for all large values of m.
From the perspective of progress towards Fermat’s Last Theorem, it would certainly
be more encouraging to know that there are infinitely many regular primes. Unfor-
tunately, this question remains wide open! Of course, the urgency of the problem
has diminished somewhat in the wake of the full proof of Fermat’s Last Theorem by
Wiles and Taylor–Wiles.∗
In this last expression, the term pBk has been dropped from Faulhaber’s
formula, which is harmless as k is odd and larger than 1.
Continuing, we argue that every term of the sum on j belongs to p3 Zp . Suppose
j ≥ 3. Since pBk−j ∈ Zp (as follows from Problem 8.99 or 13.154),
pj+1 pj
k k
vp Bk−j = vp pBk−j ≥ j − vp (j + 1)
j j+1 j j+1
j+1 j+1
≥ j − vp ((j + 1)!) > j − ≥j− ≥ 2.
p−1 4
∗
The story of which is gloriously recounted in Joshua Rosenblum and Joanne Sydney
Lessner’s 2000 musical Fermat’s Last Tango.
150
This shows that each term with j ≥ 3 belongs to p3 Zp . The remaining term,
3
corresponding to j = 2, is k2 Bk−2 p3 . As k − 2 ≡ p − 4 ̸≡ 0 (mod p − 1), we
3
have that Bk−2 ∈ Zp . Hence, k2 Bk−2 p3 ∈ p3 Zp .
2
Collecting our results so far, Hp−1 ≡ k p2 Bk−1 (mod p3 Zp ). Kummer, as
incarnated in Problem 13.156, now steps in to tell us that
Bk−1 Bp−3
≡ (mod pZp ).
k−1 p−3
Therefore,
Bp−3 2
Bk−1 ≡ (k − 1) ≡ Bp−3 (mod pZp ),
p−3 3
and
p2 p2 p2
Hp−1 ≡ k Bk−1 ≡ k Bp−3 ≡ − Bp−3 (mod p3 Zp ).
2 3 3
From this last congruence for Hp−1 , we have Hp−1 ∈ p3 Zp ⇐⇒ Bp−3 ∈ pZp .
This equivalence is the concluding assertion of the problem statement.
Remark. This problem brings to a close our study of the Bernoulli numbers. For
everything you ever wanted to know about this subject but were afraid to ask, see
Chapter 15 in the book of Ireland and Rosen [6], Chapter 9 in Cohen’s volume [4], and
the recent book [1] by T. Arakawa, T. Ibukiyama, and M. Kaneko. A bibliography
with ≈ 3000 books and papers related to Bernoulli numbers has been compiled by
Karl Dilcher, Ladislav Skula, and Ilja Sh. Slavutskii [5].
ak (xk − rk ) =
P
13.158 Since F (r) = 0, we have F (x) = F (x) − F (r) = k≥0
P Pk−1 j k−1−j
k≥0 ak (x − r) j=0 x r .
Moreover,
P P ϵN → 0P(asPF is a Strassmann series). So by Exercises 8.95 and 8.96,
k j k,j and
u j k uk,j both converge, and to the same value. Hence,
XX XX
F (x) = (x − r) uk,j = (x − r) uk,j
k j j k
X X
= (x − r) xj ak rk−1−j .
j≥0 k>j
151
Now we suppose that the Strassmann degree K of F (T ) is at least 1 and prove
that the Strassmann degree of G(T ) is K − 1. If j < K − 1, then
When j = K − 1,
X
|bK−1 |p = aK + ak rk−K .
k>K p
Hence, the maximum value of |bj |p is |aK |p , and this maximum is attained for
the last time at j = K − 1. Therefore, G(T ) has Strassmann degree K − 1.
Extra Exploration 43. Recall our notation Qp ⟨T ⟩ for the ring of Strassmann
series. Problem 13.158 establishes a version of the Root-Factor theorem for Strass-
mann series viewed as functions on Zp . Show that the Root-Factor theorem is also
valid at the level of formal power series. More precisely, show that if r ∈ Zp is a root
of F (T ) ∈ Qp ⟨T ⟩, then F (T ) = (T − r)G(T ), where G(T ) ∈ Qp ⟨T ⟩ is the power
series constructed in the preceding solution.
X X
|F (x)|p = a0 + ak xk ≥ |a0 |p − ak xk > 0.
k≥1 p k≥1 p
152
Assuming the general claim fails, choose a counterexample F (T ) whose Strass-
mann degree K is as small as possible. Then K ≥ 1. By assumption, F has
more than K distinct zeros in Zp ; pick one of these and call it r. By Exercise
13.158, we can find a Strassmann series G(T ) of Strassmann degree K − 1
with F (x) = (x − r)G(x) for all x ∈ Zp . As K − 1 < K, we know that G has
at most K − 1 distinct zeros in Zp . But each zero of F , other than (possibly)
r, is a zero of G. Thus, there are at most 1 + (K − 1) = K zeros of F in Zp ,
contradicting the choice of F (T ).
13.160 Referring back to Set #11, for each A ∈ pZp the constant term CA,0
of Binom(1 + A, T ) is
X Ak A0
CA,0 = s(k, 0) = s(0, 0) = 1.
k! 0!
k≥0
153
r 0 1 2 3 4 5 6 7 8 9
Fr,+ (0) mod 11Z11 9 0 0 7 3 7 8 1 3 8
r 0 1 2 3 4 5 6 7 8 9
Fr,− (0) mod 11Z11 2 4 4 0 7 0 1 5 7 1
ak bk
X
αr Ca,j − β r Cb,j = s(k, j) αr − βr .
k! k!
k≥j
k
For each k ≥ j, we have v11 (ak /k!), v11 (bk /k!) ≥ k − v11 (k!) > k − 10 = 0.9k ≥
0.9j. Therefore,
k k
ra rb
r r
|α Ca,j − β Cb,j |11 ≤ max s(k, j) α −β ≤ 11−0.9j .
k≥j k! k! 11
Plugging our approximations of α and a into (*), we find that the T -coefficient
154
So the T -coefficients are divisible by 11 but not 112 for (r, ±) = (1, +), (2, +),
and (5, −).
Let’s take stock of what we’ve shown in this exercise and the last. If (r, ±)
is any of (1, +), (2, +), (5, −), then the constant term of Fr,± (T ) is 0. The T -
coefficient has 11-adic absolute value 11−1 . And the T j coefficient has 11-adic
absolute value at most 11−0.9·2 < 11−1 , for each j ≥ 2. It follows that Fr,± (T )
has Strassmann degree 1.
13.162 We saw already in the solution to Problem 13.161 that F3,− (T ) has
T -coefficient divisible by 112 .
The T 2 -coefficient of F3,− (T ) is given by
k k
3a 3b
X
3 3
α Ca,2 − β Cb,2 = s(k, 2) α −β .
k! k!
k≥2
155
Extra Exploration 45 (Skolem, Chowla, and Lewis [10]). Prove that
n n
−β
each integer appears at most three times in the sequence { αα−β }n≥0 (for the same
α, β as above).
Extra Exploration 46 (Cohen and Ljunggren [3]). Find all integer solu-
tions to x2 + 11 = 3m .
Extra Exploration 47. What are all of the positive integer solutions to 2x2 +1 =
3m ? Equivalently: Which squares have ternary expansions consisting entirely of the
digit 1?
References
156
Index
157
B Fermat’s last theorem 149
Bernoulli numbers H
Adams’ theorem 51, 148
harmonic number 5, 17, 52, 59, 81, 150
alternate in sign 16, 76
Hensel’s lemma 38, 125, 126
associated characterization of Wilson
primes 35, 120 I
associated congruence for Hp−1 52,
150 irregular prime 149
Clausen–von Staudt theorem 32, 109
definition 15 L
distribution of denominators 109
La Cara de la Guerra 20
Faulhaber’s formula 16, 75, 108, 120,
Legendre’s formula for vp (n!) 58
147
limits (sequential) 11, 16, 67, 79
in terms of ζ-values 77
Liouville’s approximation theorem in Zp
Kummer’s congruence 51, 149
34, 115
vanish for odd indices > 1 16, 76
logarithm, on Qp see p-adic logarithm
C M
canonical expansions of elements of Qp method of successive approximation
29, 100 29, 33, 101, 113, 131
Cantor–Schröder–Bernstein theorem Monsky’s theorem 21, 88
87
Cauchy sequence 23, 91 N
Clausen–von Staudt theorem 32, 109
Newton’s method 38, 125
completeness of a valued field 27
non-Archimedean absolute value 4
completion of a valued field 28
is unique 28, 98 O
convergence of infinite product 105
convergence of infinite series 12, 31, 68, Ostrowski’s theorem 17, 80
105
convergence of sequences see limits P
(sequential)
p-adic
cubic Pellian equation 49, 144
absolute value 3, 24
valuation 3, 24
D
p-adic exponential 134
discs, open and closed 7, 61 p-adic integer see Zp , ring of p-adic
integers
E p-adic logarithm
definition and convergence on 1 + pZp
equivalent absolute values 17, 80 41, 129
exponential function, on Qp see p-adic proof of additivity 130
exponential sets up isomorphism between 1 + pZp
and pZp 134
F p-adic number see Qp , field of p-adic
numbers
Faulhaber’s formula 16, 75, 108, 120, p-integral rational number see Z(p) ,
147 ring of p-integral rationals
158
primes in arithmetic progressions 9, 65 Strassmann degree of 52
product formula for absolute values on Strassmann’s theorem 52, 152
Q 4, 58 strictly convergent power series see
Strassmann series
R strong triangle inequality 4
extension to infinite series in Qp 31,
Ramanujan–Nagell equation 13, 52, 72,
105
155
survival of the greatest 4
regular prime 149
respectable mathematician 11
T
non-example, see bottom of 32
restricted power series see Strassmann
Taylor’s formula 38, 125
series
Teichmüller representatives 32, 35, 51,
S 110, 121, 147
trivial absolute value 3
sadness in our hearts 27
simultaneous approximation 58 V
Skolem–Mahler–Lech theorem 44, 136
Strassmann degree 52 valued field 3
Strassmann series
definition 34 W
form a subring of Qp [[T ]] 119
if nonzero has finitely many zeros in Weierstrass preparation theorem for Qp
Zp 35, 120 134, 153
number of zeros bounded by Wieferich prime 8
Strassmann degree 52, 152 Wilson prime 35
representing (1 + a)n 43, 45, 134, 141 Wilson’s theorem 25, 94, 101
representing terms of a linear Wolstenholme prime 52
recurrence 44, 135 Wolstenholme’s theorem 5, 59
159