Project Name: P-Adic Integers
Project Name: P-Adic Integers
Santacruz(East), Mumbai-400098.
PROJECT NAME :
p-ADIC INTEGERS
SUBMITTED BY :
1. Anjali Singh
2. Kainat Shaikh
3. Mariya Sadaf Shaikh
4. Nasra Momin
5. Sadia Khatoon
6. Safeera Shaikh
7. Shubham Jadhav
8. Suresh Mishra
1
STATEMENT TO BE INCORPORATED BY THE
CANDIDATES IN THE THESIS AS REQUIRED
UNDER REGULATION FOR THE M.SC DEGREE
As required by the University Regulation No. R.1972/ R.2190 we wish
to state that the work embodied in this thesis titled “p-ADIC INTEGERS
” is expository in nature and does not contain any new results. The entire
work is carried out under the guidance of Dr. Anuradha S. Garge at the
Department of Mathematics, University of Mumbai. This work has not been
submitted for any other degree of this or any other University. Whenever
references have been made to previous works of others, it has been clearly
indicated as such and included in the Bibliography.
Signature of Students
Anjali Singh
Kainat Shaikh
Mariya Sadaf Shaikh
Nasra Momin
Sadia Khatoon
Safeera Shaikh
Shubham Jadhav
Suresh Mishra
2
Certificate
External Examiner
Date : / / 2022
3
Acknowledgement
DATE :
PLACE : MUMBAI
Department of Mathematics,
University of Mumbai,
Kalina, Mumbai-400098.
4
Contents
1 Introduction 6
7 Geometry in Qp 18
8 Completions 22
9 p-adic Integers 33
11 Hensel’s Lemma 44
5
1 Introduction
Over the last century, p-adic number and p-adic analysis have come to play a
central role in modern number theory. This importance comes from the fact
that they afford a natural and powerful language for talking about congru-
ences between integers and allow the use of methods borrowed from calculus
and analysis for studying such problems. More recently, p-adic number is
shown up in other areas of mathematics and even in physics. The study
of p-adic number is attractive because it blends together so many part of
mathematics.
The p-adic numbers were first introduced by the German mathematician
Kurt Hensel in 1897 (though they are fore-shadowed in the work of his prede-
cessor E. Kummer). It seems that Hensel’s main motivation was the analogy
between the ring of integers Z, together with its field of fractions Q. The
p-adic numbers were then generalized to valuations by Hungarian mathe-
matician József Kürschák in 1913. More formally, for a given prime p, the
field Qp of p-adic number is completion of the rational numbers with respect
to p-adic metric. The field Qp is also given a topology derived from a metric,
which is itself derived from the p-adic order. This metric space is complete
in the sense that every Cauchy sequence converges to a point in Qp . This is
what allows development of calculus on Qp and it is the interaction of this
analytic and algebraic structure that gives the p-adic numbers system their
power and utility.
The p in “p-adic” is a variable and may be replaced with a prime (yield-
ing, for instance, “the 2-adic numbers”) or another expression representing a
prime number. The “adic” of “p-adic” comes from the ending in words such
as dyadic or triadic. The p-adic numbers are most simply a field extension of
Q, the rational numbers which can be formulated in two ways using either an-
alytic or algebraic methods. The p-adic integers are the p-adic numbers with
non-negative valuation. Every integer is a p-adic integer (including zero).
The p-adic numbers are useful because they provide another toolset for
solving problems, one which is sometimes easier to work with than the real
numbers. Suppose we have a physical or any other system and we make
measurments. To describe results of the measurments, we can always use
rationals. According to the Ostrowski’s theorem there are only two kinds
of completions of the rationals. They give real or p-adic Qp numbers fields,
where p is any prime number with corresponding p-adic norm | x |p which is
6
non-archimedean. Also, p-adic numbers have application in number theory,
where it has a significant role in diophantine geometry and diophantine ap-
proximation, some application required the development of functional anal-
ysis, algebra and more. One example is Hensel’s lemma for finding roots of
the polynomial.
To build a mathematical model of the system we will use real or p-adic
numbers or both, depending on the properties of system. We shall start with
introducing the absolute value on the field Q and then valuation on the field
Q. We then move on to p-adic absolute value of Q and some properties. Next
we state the Ostrowiski’s Theorem. Then we define distance on the field Q:
the p-adic distance etc. Then we will look at how the p-adic geometry differs
from Euclidean geometry. Next, we discuss the completion using the concept
of topology to get a sense of p-adic spaces. We then move to p-adic integers.
Finally, we will look at Hensel’s lemma.
1. | x |≥ 0;
2. | x | = 0 if and only if x = 0;
3. | xy | = | x || y |;
4. | x + y | ≤ | x | + | y |;
5. | x + y | ≤ max{| x |, | y |}.
7
Note that | x | is usual absolute value.
| x |= 1, if x ̸= 0 and | 0 | = 0 is trivial absolute value.
Now for defining p-absolute value, let us explore some definitions and lemma.
Examples :
(i) v3 (402) = 1, multiplicity of 3 as divisor of 402 is 1, since 402 = 31 × 134.
(ii) v3 ( 123
48
) = v3 (123) − v3 (48) = 1 − 1 = 0, multiplicity of 3 as divisor of 123
and 48 is 1, since 123 = 31 × 41, 48 = 31 × 16.
(iii) vp (1) = 0, multiplicity of any prime number as divisor of 1 is 0, since
1 = p0 × 1.
Note that vp (0) = +∞, since any arbitrarily high power of p divides zero.
We observe that for any x ∈ Q, the value of vp (x) does not depend on it’s
representation as a quotient of two integers. In other words, if ab = dc , then
vp (a) − vp (b) = vp (c) − vp (d).
We will prove this as follows:
Let x = ab = dc .
We have x = ab implies that vp (x) = vp (a) − vp (b). Similarly vp (x) =
vp (c) − vp (d). Hence vp (a) − vp (b) = vp (c) − vp (d).
6 12
Example : As 4
= 8
then v2 (6) − v2 (4) = −1 = v2 (12) − v2 (8).
8
Now we will discuss some basic properties of p-adic valuation.
9
Then min{vp (x), vp (y)} = n = m.
Consider,
a c a c
x + y = pn + pm = pn ( + ).
b d b d
If n ̸= m then,
CASE I: If n < m,
Consider,
a c a c a c
x + y = pn + pm = pn + pm+n−n = pn ( + pm−n ).
b d b d b d
ad + pm−n bc
= vp (pn ) + vp ( ) = vp (pn ) + vp (ad + pm−n bc) − vp (bd).
bd
Therefore vp (x + y) = vp (pn ) = n.
10
Thus we get, vp (x + y) = min{vp (x), vp (y)}.
pn−m ad + bc
= vp (pm ) + vp ( ) = vp (pm ) + vp (pn−m ad + bc) − vp (bd).
bd
Therefore vp (x + y) = vp (pm ) = m.
Hence from both the cases 1 and 2 we get vp (x + y) = min{vp (x), vp (y)}.
11
4 p-adic Absolute Value Of Q
Definition : For any x ∈ Q, we define p-adic absolute value of x by
| x |p = p−vp (x) ,
where p ∈ Z is prime.
Examples :
(a) | x |p ≥ 0;
(c) | xy |p = | x |p | y |p ;
12
(c) | xy |p = p−vp (xy)
= p−(vp (x)+vp (y))
= p−vp (x) p−vp (y)
=| x |p | y |p .
(d) If | x |p = | y |p ,
then max{| x |p , | y |p } = | x |p = | y |p .
As | x |p = | y |p .
By p-adic absolute value, p−vp (x) = p−vp (y) .
This gives vp (x) = vp (y).
Since by property 2 of valuation, vp (x + y) ≥ vp (x).
So −vp (x + y) ≤ −vp (x).
This gives p−vp (x+y) ≤ p−vp (x) .
We get that | x + y |p ≤ | x |p .
Hence | x + y |p ≤ max{| x |p , | y |p }.
If | x |p ̸= | y |p ,
CASE I: | x |p > | y |p .
Then max{| x |p , | y |p } = | x |p .
As | x |p > | y |p .
We have p−vp (x) > p−vp (y) , which gives −vp (x) > −vp (y).
We can write vp (x) < vp (y).
Then vp (x) = min{vp (x), vp (y)} = vp (x + y).
We get −vp (x + y) = −vp (x).
This implies p−vp (x+y) = p−vp (x) .
Hence | x + y |p = | x |p = max{| x |p , | y |p }.
13
i.e., vp (x) > vp (y).
Then vp (y) = min{vp (x), vp (y)} = vp (x + y),
i.e., −vp (x + y) = −vp (y).
This implies p−vp (x+y) = p−vp (y) .
Hence | x + y |p = | y |p = max{| x |p , | y |p }.
Now we state the theorem that classifies all non-trivial absolute values
defined on the field of the rational numbers Q.
14
5 Distance On The Field Q
Definition : Let Q be a field of rationals numbers and | | an absolute value
on Q. We define the distance d(x, y) between two elements x, y ∈ Q by
d(x, y) = | x − y | .
Now we see the general properties for a metric d(x, y) as follows, for any
x, y, z ∈ Q:
(a) d(x, y) ≥ 0;
Here the last inequality is called the triangle inequality, since it expresses the
usual fact that the sum of the lengths of two sides of triangle is bigger than
the length of the other sides.
A set on which the metric is defined is called the ‘metric space’.
15
for some fixed prime p.
16
(d) Consider d(x, z) = | x − z |p
= | x + y − y − z |p
= | (x − y) + (y − z) |p
≤ max{| x − y |p , | y − z |p }
≤ | x − y |p + | y − z |p
= d(x, y) + d(y, z).
Proposition : If d(x, z) ≤ max{d(x, y), d(y, z)} then the metric defined
by d is non-archimedean.
Proof : Let x, y, z ∈ Q.
Consider d(x, z) = | x − z |p
= | x + y − y − z |p
= | (x − y) + (y − z) |p
≤ max{| x − y |p , | y − z |p }
= max{d(x, y), d(y, z)}.
Thus,
d(x, z) ≤ max{d(x, y), d(y, z)}.
17
7 Geometry in Qp
Once we have a way to measure distances, we can do the geometry. So to do
the geometry in Qp we start with three following remarks:
Remarks :
(a) A point is an element of Qp .
(c) A triangle abc has sides of lengths given by d(a, b), d(b, c) and d(c, a).
Now we will see a geometry different from the euclidean geometry in
which all triangles are isosceles.
| a − c |p = max{| (a − b) |p , | (b − c) |p },
18
Example : Let Q5 with | |5 .
Let x = 200, y = 185, z = 85 ∈ Q5 be vertices of a triangle.
By 5-adic metric we have,
d(x, y) =| x − y |5 = | 200 − 185 |5 = | 15 |5 = 1/5
and
d(y, z) =| y − z |5 = | 185 − 85 |5 = | 100 |5 = 1/25.
It implies that d(x, y) ̸= d(y, z).
Now consider,
d(x, z) =| x − z |5 = | x − z + y − y |5 = | (x − y) + (y − z) |5
= | (200 − 185) + (185 − 85) |5 = | 15 + 100 |5 .
Since d(x, y) ̸= d(y, z), therefore by using property (d) of p-adic absolute
value of Qp ,
i.e., if | x |p ̸=| y |p then | x + y |p = max{| x |p , | y |p }, we get,
d(x, z) =| 15 + 100 |5 = max{| 15 |5 , | 100 |5 } = max{1/5, 1/25} = 1/5.
Hence we get, d(x, y) = d(x, z).
Therefore △xyz is isosceles triangle.
Theorem : If a triangle is not equilateral, the unequal side has the largest
valuation and hence becomes the shortest side of the triangle.
Proof : Let △abc be a triangle. Let a, b and c be distinct points in Qp .
Then d(a, b), d(b, c) and d(a, c) are the lengths of the sides of the triangle
determined by those points.
Since △abc is not equilateral, without loss of generality, let d(a, b) = d(b, c) ̸=
d(a, c).
Then | a − b |p = | b − c |p ̸= | a − c |p ,
i.e., p−vp (a−b) = p−vp (b−c) ̸= p−vp (a−c) .
We get vp (a − b) = vp (b − c) ̸= vp (a − c).
Then vp (a − c) = vp (a − c + b − b) = vp ((a − b) + (b − c))
≥ min{vp (a − b), vp (b − c)}, by property (3) of valuation on the field Q.
Since vp (a − c) ̸= vp (a − b) = vp (b − c) and the triangle is not equilateral, we
have vp (a − c) > vp (a − b).
Hence the unequal side has largest valuation.
Now, to show that the unequal side has the shortest length.
From above, we have vp (a − c) > vp (a − b) or vp (a − c) > vp (b − c).
19
Let vp (a − c) > vp (a − b).
Then pvp (a−c) > pvp (a−b) .
This gives 1/pvp (a−c) < 1/pvp (a−b) ,
i.e., p−vp (a−c) > p−vp (a−b) ,
i.e., | a − c |p < | a − b |p .
Thus d(a, c) < d(a, b). Hence the unequal side has the shortest length.
Theorem : Given three points a, b and c ∈ Qp , d(a, c) < d(a, b) + d(b, c).
In other words, no three points in Qp are collinear.
Proof : Let △abc be a triangle. Let a, b and c be distinct points in Qp .
Then d(a, b), d(b, c) and d(a, c) are the lengths of the sides of the triangle
determined by points a, b, c.
By triangle inequality of p-adic distances on the field Q, we have
i.e., p−vp (a−c) > p−vp (a−b) i.e., vp (a − c) < vp (a − b). (2)
Similarly,
d(a, c) > d(b, c) i.e., vp (a − c) < vp (b − c). (3)
20
It means that either vp (a − c) ≥ vp (a − b) or vp (a − c) ≥ vp (b − c).
So its a contradiction to equations (2) and (3).
Hence our assumption is wrong.
Therefore d(a, c) ̸= d(a, b) + d(b, c).
Thus by equation (1),
d(a, b) < d(a, c) + d(b, c).
Thus three points a, b and c are not collinear.
21
8 Completions
We are now ready to construct the p-adic field Qp . So first we need to recall
some important concepts from basic topology.
(b) The field K is called Complete with respect to the absolute value | | if
every Cauchy sequence of elements of K are convergent in K (i.e., it
has limit that is also in K).
The first useful thing to note is that the Cauchy sequences can be char-
acterized much more simply when the absolute value is non-archimedean.
lim | xn+1 − xn |p = 0.
n→∞
We already know that Q is not complete with respect to the metric given
by the usual absolute value. Is Q complete with respect to the new distances
or not ?
The answer to this question is no. We will show this in the following lemma:
• a is not a square in Q;
22
• p does not divide a;
Note that a exists as there are exactly half the residues modulo p and half
non-residues modulo p.
Now we can construct a Cauchy sequence (with respect to ||p ) in the following
way:
We can see that such a sequence (xn ) does exist as the initial element x0
exists.
Now we check that the sequence obtained above is a Cauchy sequence. It is
clear from the construction that xn+1 ≡ xn mod pn+1 .
This implies that xn+1 = xn + λpn+1 . Hence
23
Hence by previous lemma we see that the sequence (xn ) is indeed a sequence.
We also know that x2n ≡ a mod pn+1 . So, x2n = a + µpn+1 and hence
Now for completion we emphasize p and Q. For this we need to add more
points to Q as completion of Q depends on the metric used and is based on
the notion of Cauchy sequences.
Proposition : By defining
Now we only need to check that the sum and product of two Cauchy se-
quences, as defined in above proposition, are also Cauchy sequence.
24
n > N.
Simillarly, let (yn ) be a Cauchy sequence in Q with respect to | |p .
By definition, for some ϵ > 0, ∃M ∈ N such that | yn − ym |p < 2ϵ , for all
n > M.
Choose n > max{N, M }.
Consider for ϵ > 0 and n > max{N, M },
| (xn + yn ) − (xm + ym ) |p ≤ | xn − xm |p + | yn − ym |p
ϵ ϵ
< + = ϵ.
2 2
This implies | (xn + yn ) − (xm + ym ) |p < ϵ, for all n > max{N, M }.
Thus (xn + yn ) is a Cauchy sequence in Q with respect to | |p .
Further, for multiplication we have,
| xn yn − xm ym |p = | xn yn − yn xm + yn xm − xm ym |p
= | yn (xn − xm ) + xm (yn − ym ) |p
≤ | yn |p | xn − xm |p + | xm |p | yn − ym |p
ϵ ϵ
< | yn |p + | xm |p
2 2
1ϵ 1ϵ 1 ϵ ϵ 1
< + = ( + ) = ϵ < ϵ,
k2 k2 k 2 2 k
as a Cauchy sequence (xn ) and (yn ) is bounded by some 1/k, where k > 0.
This implies | xn yn − xm ym |p < ϵ, for all n > max{N, M }.
Thus (xn yn ) is also a Cauchy sequence in Q with respect to | |p .
Hence the sum and product of two Cauchy sequences is also a Cauchy se-
quence. Therefore closure property is also satisfied under addition and mul-
tiplication. Thus the set C forms a commutative ring with unity.
Let (xn ) ∈ C, where (xn ) ̸= 0 = (0, 0, 0, ....). If any term in (xn ) equals
25
zero (for example, (1, 0, 1, 0, ...)), then (xn ).(yn ) ̸= 1 = (1, 1, 1, ...) for any
(yn ) ∈ C, also 0 is not invertible, and so (xn )−1 does not exist, for any
(xn ) ∈ C.
Hence we conclude that the ring C does not form a field, because not all
non-zero elements are invertible. Moreover, it contains zero divisors as the
product of the non-zero two sequences in C (for example (0, 1, 1, 1, ...) and
(1, 0, 0, 0, ...)) is clearly 0 = (0, 0, 0, ...).
We say two Cauchy sequences are equivalent when they share the same limit
and we define the set N ⊂ C of sequences that tends to zero.
Definition : We define N ⊂ C to be the ideal
N = {(xn ) : xn → 0} = {(xn ) : lim | xn |p = 0}
n→∞
26
Lemma : N is a maximal ideal of C.
Proof : Let (xn ) ∈ C be a Cauchy sequence that does not tend to zero (i.e.,
does not belong to N).
Let I be the ideal generated by (xn ) and N, i.e., I = ⟨x, N⟩.
We get N ⊂ I ⊂ C, we have to prove that N is maximal ideal, i.e., I = N or
I = C.
So we want to show that I must be all of C if I ̸= N. We will do that by
showing that the unit element (1) (i.e., the constant sequence corresponding
to 1) is in I.
This is enough, because any ideal that contains the unit element must be
the whole ring.
Now, since (xn ) does not tend to zero and is a Cauchy sequence, it must
“eventually” be away from zero, that is, there must exist a number c > 0
and an integer N such that | xn | ≥ c > 0, whenever n ≥ N . Now in particu-
lar this means that xn ̸= 0 for n ≥ N , so that we may define a new sequence
(yn ) setting yn = 0 if n < N and yn = x1n if n ≥ N .
The first thing to check is that (yn ) is a Cauchy sequence. That is clear
because if n ≥ N we have
1 1 | xn+1 − xn |p | xn+1 − xn |p
| yn+1 − yn |p =
− = ≤ → 0,
xn+1 xn p | xn xn+1 |p c2
This means that the product sequence (xn )(yn ) consists of a finite number
of 0’s followed by an infinite string of 1’s,
i.e., (xn )(yn ) = (0, 0, · · · , 0, 1, 1, · · · ). In particular, if we subtract it from the
constant sequence (1), we get a sequence that tends to zero,
i.e., (1, 1, · · · ) − (0, 0, · · · , 0, 1, 1, · · · ) = (0, 0, · · · ).
27
In other words
(1) − (xn )(yn ) ∈ N.
It means that
(1) ∈ (xn )(yn ) + N.
This says that (1) can be written as a multiple of (xn ) plus an element of N.
Hence (1) belongs to I, as we had claimed. Thus N is maximal ideal of C.
Now by taking the quotient of the ring C by the ideal N, we make things
even nicer, i.e., taking a quotient of the ring by a maximal ideal gives us a
field. Momentarily, we define this field:
Let x and y be the two equivalence classes of Cauchy sequences with repre-
sentatives (xn ) and (yn ) respectively, i.e., x, y ∈ Qp .
28
= | x′n (yn′ − yn ) + yn (x′n − xn ) |p
≤ max{| x′n (yn′ − yn ) |p , | yn (x′n − xn ) |p }
= max{| x′n |p | yn′ − yn |p , | yn |p | x′n − xn |p }.
The limit of terms in right hand side of above expression is
| x |p limn→∞ | yn′ − yn |p = 0 and | y |p limn→∞ | x′n − xn |p = 0, (since x
and y are equivalence classes).
This implies that limn→∞ | x′n yn′ − xn yn |p = 0.
Therefore (x′n yn′ ) ∼ (xn yn ).
This implies that x.y ∈ Qp , it means that multiplication is defined on Qp .
29
≤ max{| x′n − xn |p , | yn′ − yn |p }.
Then limit of right hand side of above expressions is,
limn→∞ (| x′n − xn |p ) = 0 and limn→∞ (| yn′ − yn |p ) = 0 (since x and y are
in Qp ).
This implies that limn→∞ | (x′n + yn′ ) − (xn + yn ) |p = 0.
Hence (x′n + yn′ ) ∼ (xn + yn ).
This implies that x + y ∈ Qp . This means that addition is defined on Qp .
=| −(x′n − xn ) |p = | x′n − xn |p .
As x is in Qp , limn→∞ | x′n − xn |p = 0.
Therefore by above expression we have limn→∞ | (−x′n ) − (−xn ) |p = 0.
Hence (−x′n ) ∼ (−xn ).
It means that −x ∈ Qp , i.e., additive inverse is also defined on Qp .
Now let us check for the distributive law; x(y + z) is the equivalence class of,
30
so (x + y) + z is also an equivalence class.
Hence x + (y + z) = (x + y) + z, so the associative law holds.
(xn + yn ) = (yn + xn )
Notice that the two different constant sequences never differ by an element
of N (their difference is just another constant sequence). Hence we still have
an inclusion
Q ,→ Qp
by sending x ∈ Q to the equivalence class of the constant sequence (x).
To check we have indeed obtained the completion, we must now check the
remaining two requirements: that Q is dense in Qp , and that Qp is complete.
31
some fix radius r > 0. Hence we are done.
32
9 p-adic Integers
We shall devote this section to the p-adic integers which we denote by Zp .
Definition : The ring of p-adic integers is the valuation ring
Zp = {x ∈ Qp : | x |p ≤ 1}.
Examples :
(i) For p = 2, | 3 |2 = 2−v2 (3) = 2−0 = 1. Thus 3 ∈ Z2 .
(ii) For p = 2, | 1
|
4 2
= 2−v2 (1/4) = 2−v2 (1)−(−v2 (4)) = 2−(−2) = 4 > 1.
Thus 4 ∈/ Z2 .
Remarks :
Now we discuss more about the ring of the p-adic integers Zp . For this
we first define local ring and then recall the following proposition from the
abstract algebra.
33
Then every x ∈ M ′ \ M is an unit.
This implies M ′ = R,
i.e., M ′ can not be a maximal ideal of R.
Thus we get M ⊂ M ′ .
CASE II: M ′ ⊊ M
This implies every element outside M ′ is a unit, hence 1 ∈ M , which is a
contradiction as M is proper ideal of R.
Thus M ′ ⊂ M .
Hence by both cases I and II we get M = M ′ ,
i.e., M is the unique maximal ideal of R.
Thus R is a local ring.
Lemma : The valuation ring Zp is local ring with unique maximal ideal
pZp = {x ∈ Zp : | x |p < 1}. We shall call pZp the valuation ideal of Zp .
Proof : We will show that pZp is an ideal.
Let x, y ∈ pZp , this implies | x |p < 1, | y |p < 1, for x, y ∈ Qp . So we get
max{| x |p , | y |p } < 1.
Therefore we get | x − y |p ≤ max{| x |p , | −y |p } = max{| x |p , | y |p } < 1,
as | −1 |p = 1.
Thus x − y ∈ pZp .
Now for r ∈ Zp , | r |p ≤ 1 and x ∈ pZp , | x |p < 1.
We consider | rx |p = | r |p | x |p ≤ 1 | x |p < 1.
This implies rx ∈ pZp .
Hence pZp is an ideal of Zp .
Further, only we have to prove that every x ∈ Zp \ pZp is an unit of Zp .
If x ∈
/ pZp but x ∈ Zp =⇒ | x |p = 1.
Thus | 1/x |p = 1.
This implies x−1 ∈ Zp , i.e., x is invertible in Zp .
Hence we get for x ∈ Zp , there exists x−1 in Zp such that x.x−1 = 1.
Since x is arbitrary, therefore we get every element in Zp outside pZp is a
unit of Zp .
Thus by using the proposition of this section, we get that Zp is local ring and
pZp is its maximal ideal.
34
Proposition : The ring Zp of p-adic integers is a local ring whose maxi-
mal ideal is the principal ideal pZp = {x ∈ Zp : | x |p < 1}. Furthermore,
i) Q ∩ Zp = Z(p) := { ab ∈ Q : p ̸ | b}.
ii) The inclusion Z ,→ Zp has dense image. In particular, given x ∈ Zp and
n ≥ 1, there exist α ∈ x ∈ Z, 0 ≤ α ≤ pn − 1, such that | x − α |≤ p−n . The
integer α with these properties is unique.
iii) For any x ∈ Zp , there exists a Cauchy sequence αn converging to x, of
the following type:
• αn ∈ Z satisfies 0 ≤ αn ≤ pn − 1;
35
This implies x ∈ Z(p) . Hence Q ∩ Zp ⊂ Z(p) .
Now to prove that Z(p) ⊂ Q ∩ Zp .
Let x ∈ Z(p) , this implies x = a/b ∈ Q; p ̸ | b, by definition of Zp .
a |a|p |a|p
Now | |
b p
= |b|p
= p−0
, as p ̸ | b
|a|p
= 1
=| a |p . As a ∈ Z, vp (a) ≥ 0. Hence | a |p ≤ 1.
a
Hence | |
b p
≤ 1, this implies | x |p ≤ 1.
Hence x ∈ Zp .
Therefore Z(p) ⊂ Q ∩ Zp . Hence we can conclude that Q ∩ Zp = Z(p) .
The point is to show that we can in fact choose an integer close enough to
x. Notice that for p ̸ | b as above, we will have
a a
| | ≤ max{| x |, | x − |} ≤ 1,
b b
which says that a/b ∈ Z(p) , i.e., p ̸ | b. Now recall that from the elementary
theory of congruences, if p ̸ | b there exits a unique integer b′ ∈ Z such that
bb′ ≡ 1(modpn ) and hence
a
| − ab′ | ≤ p−n ,
b
and of course ab′ ∈ Z. Finally, we need to check that we can find an inte-
ger between zero and pn − 1, but this is clear from the connection between
congruences modulo powers of p-adic absolute value: choosing α to be the
unique integer such that
0 ≤ α ≤ pn − 1
36
and
α ≡ ab′ (modpn ),
gives | x − α | ≤ p−n .
iii) It follows directly from (ii); just use (i), we define αn = α for a se-
quence of integers n = 1, 2, 3, etc.
Further note, αn−1 ≡ ab′ (mod pn−1 ) and αn ≡ ab′ (mod pn ).
This implies αn ≡ ab′ + λpn ≡ ab′ (mod pn−1 ) ≡ αn−1 (mod pn−1 ).
This proposition says several important things. Notice, that the sequence
(iii) is the “coherent sequences”. Now we begin from item (iii) in the above
proposition: for given x ∈ Zp , we can find a special kind of a Cauchy se-
quences converging to x and this sequences has the property of “coherent
sequences”. So we go to obtain a canonical way to represent the elements of
Qp as “power series in p”.
Let x ∈ Zp and (an ) be the Cauchy sequence described in the above item
(iii) of the proposition. We write them in the base p. Let a0 = b0 for some
0 ≤ b0 ≤ p − 1. Since a1 ≡ a0 ( mod p), a1 = b0 + b1 p for some 0 ≤ b1 ≤ p − 1.
Going up the sequences, we get
a0 = b 0 0 ≤ b0 ≤ p − 1
a1 = b 0 + b 1 p 0 ≤ b1 ≤ p − 1
2
a2 = b 0 + b 1 p + b 2 p 0 ≤ b2 ≤ p − 1
2 3
a3 = b 0 + b 1 p + b 2 p + b 3 p 0 ≤ b3 ≤ p − 1
and so on. Putting all of this together, we get an infinitely long expansion
b0 + b1 p + b2 p 2 + · · · + bn p n + · · · .
b0 + b1 p + b2 p 2 + · · · + bn p n + · · ·
37
(an ), which we already know converge to x, by the proof of part (ii) we have
done.
From the above lemma, the ring of the p-adic integers can also be stated
as: Definition: Let p < ∞ be a prime. P We define the p-adic integers, de-
noted Zp , to be the collection of all sums: ∞ n
n=0 bn p , with bn ∈ {0, ..., p − 1}.
In other words, Zp consists of all numbers with no negative powers of p in
its expansion.
x = b0 + b1 p + b2 p 2 + · · · + bn p n + · · ·
38
10 Solving Congruences Modulo pn
The “p-adic numbers” we have just constructed are closely related to the
problem of solving congruences modulo powers of p. We will look at some
examples of this.
Here we consider the system of congruence x2 ≡ 2 mod 7n and we try
to show that 7-adic expansion number
x = 3 + 1 × 7 + 2 × 49 + 6 × 343 + · · ·
satisfies x2 = 2 in Q7 .
For this, we first compute the solution for x2 ≡ 2 mod 7, which is 3 and 4.
Therefore, we get x ≡ 3 mod 7 and x ≡ 4 mod 7 ≡ −3 mod 7.
Now we compute the solution for x2 ≡ 2 mod 7j , for this we lifting our
solution from modulo 7 to modulo 72 to modulo 73 , until we get to the 7j
that is our target.
Let us see this,
for j = 2, i.e.,
x2 ≡ 2 mod 72
x2 ≡ 2 mod 49.
Now to find solution, note that their reductions module 7 must be solution
for j = 1.
Hence we set x = 3 + 7k or x = 4 + 7k, for some k ∈ Z.
Now solve for k,
consider
x = 3 + 7k
(3 + 7k)2 ≡ 2 mod 72
9 + 42k ≡ 2 mod 49
7 + 42k ≡ 0 mod 49
7(1 + 6k) ≡ 0 mod 49
1 + 6k ≡ 0 mod 7
6k ≡ −1 mod 7
6k ≡ 6 mod 7
39
36k ≡ 36 mod 7
k ≡ 1 mod 7.
We get k = 1.
Therefore x = 3 + 7k gives x = 3 + 7 = 10.
Thus,
x ≡ 10 mod 49.
For j = 3, i.e.,
x2 ≡ 2 mod 73
x2 ≡ 2 mod 343.
Now to find solution, note that their reductions module 72 must be solution
for j = 2.
Hence we set x = 10 + 49k.
Solve for k,
(10 + 49k)2 ≡ 2 mod 343
100 + 980k ≡ 2 mod 343
98 + 980k ≡ 0 mod 343
49(2 + 20k) ≡ 0 mod 343
2 + 20k ≡ 0 mod 7
20k ≡ −2 mod 7
6k ≡ 5 mod 7
36k ≡ 30 mod 7
k ≡ 2 mod 7.
We get k = 2.
Therefore x = 10 + 49k gives x = 10 + 49 × 2 = 10 + 98 = 108.
Thus,
x ≡ 108 mod 343.
40
Similarly for j = 4,
we have x = 108 + 343k and we get that
x1 = x0 + 7 j y 0 ,
(1)
where y0 is unknown.
It ensures that x0 mod 7j+1 is a lift of x0 mod 7j .
And
x21 ≡ 2 mod 7j+1
(2)
is the equation we are trying to solve.
41
For this plugging equation (1) into equation (2) gives,
We get,
x20 ≡ 2 mod 7j+1 .
Here we can see that the variable y0 has completely disappeared from the
equation so we cannot solve for it.
Hence by induction hypothesis, equation x2 ≡ 2 mod 7n has solution for all
n.
We now note the interesting fact that the 7-adic field Q7 is strictly bigger
than the field Q. Clearly Q is contained in Q7 , as we know that Q is 7-adic
expansion of Q.
Now only remain to√ show that Q7 is strictly bigger than Q.
For x2 = 2, x = ± 2 ∈ / Q,
2
i.e., x = 2 has no solution in Q.
Note that it has solution in Q7 , as we have seen above.
Hence the field Q7 is strictly bigger than Q.
As we saw above that the x2 = 2 has solution in Q7 but now we will see
that x2 = 2 has no solution in Q5 . For proving this we use contradiction.
We consider the system of congruence
x2 ≡ 2 mod 5n .
Suppose it has solution in Q5 , i.e., that solution will have 5-adic expansion,
i.e., it is of the form a0 + a1 5 + a2 52 + · · · .
42
For finding a0 , the possible values of a0 is 0, 1, 2, 3 and 4.
For a0 = 0, 02 = 0 ̸≡ 2 mod 5 as 5 ̸ | 2;
a0 = 1, 12 = 1 ≡
̸ 2 mod 5 as 5 ̸ | 1;
a0 = 2, 22 = 4 ≢ 2 mod 5 as 5 ̸ | 2;
2
a0 = 3, 3 =9≡ ̸ 2 mod 5 as 5 ̸ | 7;
2
and a0 = 4, 4 = 16 ≡ ̸ 2 mod 5 as 5 ̸ | 14.
43
11 Hensel’s Lemma
In this section we discuss Hensel’s lemma, concerning the solvability of p-adic
polynomials. Let’s first restrict our attention to finding roots in Zp . With
this theorem we test whether a polynomial has root in Zp or not.
44
To show that one can find α2 , we have to show that one can find b1 so that
Now, we know that F (α1 ) ≡ 0 (mod p), so that F (α1 ) = px for some x.
Then the equation becomes px + F ′ (α1 )b1 p ≡ 0 (mod p2 ),
after we divide by p we obtain x + F ′ (α1 )b1 ≡ 0 (mod p).
Note that F ′ (α1 ) is not divisible by p (as F ′ (α1 ) ̸≡ 0 (mod pZp )), and hence
its inverse exists in Zp (invertible), so that we can take
such that 0 ≤ b1 ≤ p − 1.
For this choice of b1 , we set
α2 = α1 + b1 p,
which will have the stated properties. Continuing inductively we can find
α1 , α2 , · · · , αn−1 that satisfies conditions (1) and (2). We want to find αn
that satisfies conditions (1) and (2).
So, αn = αn−1 + bn−1 pn−1 with bn ∈ Zp .
We expand
F (αn ) = F (αn−1 + bn−1 pn−1 ))
≡ F (αn−1 ) + F ′ (αn−1 )bn−1 pn−1 (mod pn ).
45
this choice of bn−1 , we get αn = αn−1 + bn−1 pn−1 and F (αn ) ≡ 0 (mod pn ).
i.e., F (αn ) ≡ 0 (mod pn ) and αn ≡ αn+1 (mod pn ).
Examples :
46
and f ′ (1) = 2(1) = 2 ̸≡ 0 (mod 3).
So there exists a unique 3-adic integer x such that x2 + 7 = 0 and
x ≡ 1 (mod 3).
Therefore x = 1 + b1 3 + b2 32 + b3 33 + · · · , where bi ∈ {0, 1, 2}.
Here we can also find bi ’s as we have found in above example.
i) g1 (X) is monic (i.e., the coefficient of the highest degree term is one),
i) g(X) is monic,
iii) f (X) ≡ gn (X)hn (X) ( mod pn ) (we take the congruences coefficient-by-
coefficient).
If we can find such sequences, we are already done, since going to the limit
gives the desired polynomials g(X) and h(X). (In other words, the coef-
ficients of g(X) will be the limits of the corresponding coefficients of the
gn (X).)
47
We already have g1 (X) and h1 (X). Let us describe how to get g2 (X) and
h2 (X). Since the g’s are to be congruent, we must have
for some polynomial r1 (X) ∈ Zp [X] (since g2 (X) ≡ g1 (X) (mod p), i.e., p
divides g2 (X) − g1 (X)).
Similarly, we must have
which expand to
By multiplying, we get
f (X) ≡ g1 (X)h1 (X) + pr1 (X)h1 (X) + ps1 (X)g1 (X) + p2 r1 (X)s1 (X) (mod p2 )
f (X) ≡ g1 (X)h1 (X) + pr1 (X)h1 (X) + ps1 (X)g1 (X) (mod p2 ).
Given that,
f (X) ≡ g1 (X)h1 (X) (mod p),
so that we have
f (X) − g1 (X)h1 (X) = pk1 (X),
for some k1 (X) ∈ Zp [x]. Rearranging, we get
48
Dividing both sides by p, we get
Now divide r̃1 (X) by g1 (X), and let r1 (X) be the remainder
We know that deg r1 (X) < deg g1 (X). But now, if we set
49
≡ r̃1 (X)h1 (X) + s̃1 (X)g1 (X)
≡ k1 (X) (mod p)
so that our congruence conditions are satisfied, and the fact that the degree
of r1 (X) is smaller than the degree of g1 (X) is enough to guarantee that
g1 (X) + pr1 (X) is monic, and we are done.
This shows that g2 and h2 exist. Since they are congruent to g1 and h1
modulo p, they are also relatively prime modulo p, so that there will be no
difficulty in going on to the next step.
Now we repeat the argument changing the indices and exponents to find
g3 and h3 . Inductively, it can be proved that this produces the sequence
whose convergence proves the theorem.
50
References
[1] Fernando Q. Gouvêa, p-adic Numbers: An Introduction, 2nd edition,
New York: Springer-Verlag, 1997.
[2] William Stein, Elementary Number Theory: Primes, Congruences,
and Secrets. Available at https://www.wstein.org.
[3] Brian Courthoute, Pablo Guzman, Antoine Ronk, The p-adic integers.
Available at https://www.google.com.
[4] Logan Quick, p-adic Absolute Values. Available at https://www.
math.uchicago.edu.
[5] Andrew Baker, An Introduction to p-adic Numbers and p-adic Analy-
sis, (2011) Available at https://scholar.google.co.in.
[6] Alexa Pomerantz, An Introduction to the p-adic Numbers. Available
at http://www.math.uchicago.edu.
[7] Theodor Christian Herwig, The p-adic completion of Q and Hensel’s
Lemma. Available at http://www.math.uchicago.edu.
[8] Neal Koblitz, p-adic Numbers, p-adic Analysis, and Zeta-Functions,
2nd ed., New York: Springer-Verlag, 1984. Available at https://www.
math.mcgill.ca.
[9] Catherine Crompton, “Some Geometry of the p-adic Rationals,” Rose-
Hulman Undergraduate Mathematics Journal: Vol. 8 : Iss. 1 , Article
2, (2007). Available at https://scholar.rose-hulman.edu/rhumj/
vol8/iss1/2.
[10] Charles I. Harrington, “An Introduction to the p-adic Numbers,”
Honours Theses. 992, (2011). Available at https://digitalworks.
union.edu/theses/992.
[11] Article in p-adic Numbers Ultrametric Analysis and Applica-
tions, April 2009. Available at https://www.researchgate.net/
publication/24374483.
[12] Eric W. Weisstein, “p-adic Number,” From MathWorld–A Wol-
fram Web Resource. Available at https://mathworld.wolfram.com/
p-adicNumber.html.
51