Algebraic Number Theory

2.1 Rings of integers

Definition 2.1.1
#

Let \(K\) be a field extension of \(\mathbb {Q}\) (not necessarily finite). An element \(\alpha \) is called an algebraic integer if it is a root of a monic polynomial with coefficients in \(\mathbb {Z}\).

Lemma 2.1.2

If \(A\) is a matrix with integer coefficients, then its eigenvalues are algebraic integers.

Proof

If the matrix is integral, its characteristic polynomial is monic with integer coefficients, and thus the eigenvalues are algebraic integers.

Example 2.1.3
#
  1. \(\sqrt{2}\) is an algebraic integer as it satisfies \(x^2-2\).

  2. Any integer \(n \in \mathbb {Z}\) is an algebraic integer as they satisfy \(x-n\).

Non-example 2.1.4
#

\(\pi \),\(\epsilon \) are not algebraic integers or even algebraic numbers.

Exercise 2.1.5
#

Let \(K=\mathbb {Q}(\sqrt{5})\), is \(\frac{1+\sqrt{5}}{2}\) an algebraic integer?

Notation 2.1.6
#

If \(K\) is a field extension of \(\mathbb {Q}\) we denote the set of algebraic integers in \(K\) by \(\mathcal{O}_K\). We will show later that this is in fact a ring.

Remark 2.1.7
#

Recall the we defined \(\overline{\mathbb {Q}}\) to be the subfield of all algebraic numbers in \(\mathbb {C}\). One can similarly define \(\overline{\mathbb {Z}}\) to be the ring (we will see later why this is a ring) of all algebraic integers. From this, one then gets \(\mathcal{O}_K=K \cap \overline{\mathbb {Z}}\) for \(K\) a number field.

In general it can be hard to prove something isn’t an algebraic integer, since you’d have to prove it satisfies no monic polynomial with inter coefficients. But we have the following result that helps in some cases:

Proposition 2.1.8

Let \(\alpha \in K\) be an algebraic number with minimal polynomial \(m_\alpha \). Then \(\alpha \) is an algebraic integer if and only if \(m_{\alpha }\) has integer coefficients (note that it will be monic by the definition of minimal polynomial).

Proof

If \(\alpha \) is an algebraic integer, then its an algebraic number, so I claim that \(m_\alpha \) is monic with integer coefficient. We know \(\alpha \) satisfies some some monic polynomial \(f(x)\) with integer coefficients and therefore \(m_\alpha \) divides \(f(x)\). Now use the monic Gauss lemma 1.2.11 to get that \(m_\alpha \) is in \(\mathbb {Z}[x]\).

Conversely, if \(m_\alpha \) is has integer coefficients, then \(\alpha \) satisfies a monic polynomial with integer coefficients, thus is an algebriac integer.

Corollary 2.1.9

Let \(K\) be a number field and \(\alpha \in K\). Then \(\alpha \) is an algebraic integer if and only if \(C_\alpha \) has integer coefficients.

Proof

This follows from Corollary 1.6.8 and Proposition 2.1.8.

Corollary 2.1.10

Let \(K\) be a number fields and \(\alpha \in K\). Then there exists a \(n \in \mathbb {Z}\backslash \{ 0\} \) such that \(n \alpha \) is an algebraic integer.

Proof

Let \(A_\alpha \) be the standard representation of \(\alpha \). Then this is a matrix with rational entries. So if we clear denominators by multiplying with a suitable integer \(n\) we get that \(n A_\alpha =A_{n\alpha }\) is an matrix with integer entries and therefore \(n\alpha \) is an algebraic integer by Lemma 2.1.2.

Theorem 2.1.11 Integers in quadratic fields

Let \(d\) be a square free integer and let \(K=\mathbb {Q}(\sqrt{d})\) then

\[ \mathcal{O}_K= \begin{cases} \mathbb {Z}[\sqrt{d}] & \text{ if } d\equiv 2 \text{ or } 3 \pmod4 \\ \mathbb {Z}[\frac{1+\sqrt{d}}{2}] & \text{ if } d\equiv 1 \pmod4 \end{cases} \]

Proof

Let \(\alpha =a+b\sqrt{d}\) with \(a,b \in \mathbb {Q}\). If \(b=0\) then we are just asking which rational numbers are algebraic integers, which we have seen are only the integers. So assume \(b \neq 0\). Then the minimal polynomial of \(\alpha \) over \(\mathbb {Q}\) is

\[ x^2-2ax+ (a^2-db^2). \]

Therefore, \(\alpha \) is an algebraic integer if and only if \(2a\) and \((a^2-db^2)\) are integers.

Now, since \(2a\) is an integer, we have \(4a^2\) is an integer, and thus \(d(2b)^2\) must be an integer. If \(2b\) was not an integer, then a \(p^2\) (for some prime \(p\)) would appear in the denominator of \((2b)^2\). This would force \(d\) to be divisible by \(p^2\) contradicting the square-free assumption. Thus \(2b \in \mathbb {Z}\).

So let \(u=2a,v=2b\). Then we have \(u^2-dv^2 \equiv 0 \pmod4\). Now, if \(v\) is even, then so is \(u\) in which case \(a,b \in \mathbb {Z}\). So assume \(v\) is odd, we need to show this can only happen if \(d \equiv 1 \mod 4\). Now, \(v^2 \equiv 1 \pmod4\), and \(u^2\) is either \(0,1 \pmod4\). But note it can’t be zero since \(d\) cant be \(0 \pmod4\) as its square-free, therefore \(u^2 \equiv 1 \mod 4\) and hence \(d \equiv 1 \mod 4\).

Exercise 2.1.12
#

Check that if \(d\) is square-free and congruent to \(1\) modulo \(4\) then \(\mathbb {Z}[\frac{1+\sqrt{d}}{2}]=\{ \frac{a+b\sqrt{d}}{2} \mid a,b \in \mathbb {Z}\text{ and } a \equiv b \pmod2 \} \)

Definition 2.1.13
#

Let \(\alpha \in \mathbb {C}\). We let \(\mathbb {Z}[\alpha ]=\{ f(\alpha ) \mid f \in \mathbb {Z}[x]\} \), in other words its the smallest ring containing both \(\mathbb {Z}\) and \(\alpha \).

If we think of \(\mathbb {Z}[\alpha ]\) simply as an additive group, then it is generated by \(\{ 1,\alpha ,\alpha ^2,\dots \} \).

Definition 2.1.14
#

Let \(\alpha \in \mathbb {C}\). We say \(\mathbb {Z}[\alpha ]\) is finitely generated as an abelian group if there exits a finite set \(\{ b_1,\dots ,b_n\} \) of elements \(b_i \in \mathbb {Z}[\alpha ]\) such that every element of \(\mathbb {Z}[\alpha ]\) may be written in the form

\[ x_1b_1+\cdots +x_nb_n \]

with \(x_i \in \mathbb {Z}\).

Theorem 2.1.15

Let \(\alpha \in \mathbb {C}\). The following are equivalent:

  1. \(\alpha \) is an algebraic integer.

  2. As an additive group, \(\mathbb {Z}[\alpha ]\) is finitely generated.

  3. \(\alpha \) is an element of some subring of \(\mathbb {C}\) having a finitely generated additive group.

  4. \(\alpha A \subset A\) with \(A \subset \mathbb {C}\) some finitely generated additive subgroup. Here \(\alpha A\) is the set of elements of the form \(\alpha x\) for \(x \in A\).

Proof

\((1) \implies (2)\) : Since \(\alpha \) is an algebraic integer, it is the root of the monic polynomial \(m_\alpha \) of degree \(n\), say, which has integer coefficients. In this case we claim that \(\{ 1,\alpha ,\cdots ,\alpha ^{n-1}\} \) is generates \(\mathbb {Z}[\alpha ]\). For this we note that if \(f(\alpha ) \in \mathbb {Z}[\alpha ]\), then by doing polynomial long division we have \(f(x)=q(x)m_\alpha (x)+r(x)\) with \(\deg (r){\lt}n\). Now if evaluate at \(\alpha \) we get \(f(\alpha )=r(\alpha )\). But since \(\deg (r){\lt}n\) we see that \(r(\alpha )\) is in the \(\mathbb {Z}\) span of \(\{ 1,\alpha ,\cdots ,\alpha ^{n-1}\} \). Thus giving the claim.

\((2)\implies (3) \implies (4)\) : This is trivial.

It remains to prove \((4) \implies (1)\). Let \(\{ a_1,\dots ,a_r\} \) be a basis of \(A\). Then as each \(\alpha a_i \in A\) we can again write it in terms of this basis, so we get a system of equations

\begin{align*} \alpha a_1 & =x_{1,1} a_1+\cdots +x_{1,r}a_r \\ \alpha a_2 & =x_{2,1} a_1+\cdots +x_{2,r}a_r \\ & \vdots \\ \alpha a_r & =x_{r,1} a_1+\cdots +x_{r,r}a_r \\ \end{align*}

with the \(x_{i,j} \in \mathbb {Z}\). If we write this in matrix form by setting \(X={x_{i,j}}\), \(a=(a_1,\dots ,a_r)\) then we have

\[ X a^{T}=\alpha a^{T} \]

(here \(()^T\) denotes transpose). Therefore, \(\alpha \) is an eigenvalue of \(X\), which is a matrix with integer entries. Thus by Lemma 2.1.2 we must have that \(\alpha \) is an algebraic integer.

Using this we can now prove that \(\mathcal{O}_K\) is actually a ring.

Corollary 2.1.16

Let \(\alpha ,\beta \) be algebraic integers, then so are \(\alpha +\beta \) and \(\alpha \beta \).

Proof

Since \(\alpha ,\beta \) are algebraic integers, then by Theorem 2.1.15 \((2)\) we have \(\mathbb {Z}[\alpha ]\) and \(\mathbb {Z}[\beta ]\) are finitely generated (as additive groups). Moreover, the ring \(\mathbb {Z}[\alpha ,\beta ]\) (take this to be defined as the smallest ring containing \(\mathbb {Z},\alpha ,\beta \)) is also finitely generated, since if \(\{ a_1,\dots ,a_n\} \) generate \(\mathbb {Z}[\alpha ]\) and \(\{ b_1,\dots ,b_m\} \) generates \(\mathbb {Z}[\beta ]\) then \(\{ a_ib_j\} _{i,j}\) generates \(\mathbb {Z}[\alpha ,\beta ]\). Now, \(\mathbb {Z}[\alpha ,\beta ]\) contains both \(\alpha +\beta \) and \(\alpha \beta \), then 2.1.15 \((3)\) tells us that they must also be algebraic integers.

Corollary 2.1.17

If \(K\) is a number field, then \(\mathcal{O}_K\) is a ring.

Proof

This follows at once from the above, since if \(\alpha ,\beta \) are algebraic integers in \(K\), then \(\alpha +\beta \in K, \alpha \beta \in K\) and by the above, they are both algebraic integers.

Exercise 2.1.18
#

Let \(K/F\) be an extension of number fields and assume that \(\alpha \in K\) is a root of a monic polynomial with coefficients in \(\mathcal{O}_F\). Prove that \(\alpha \) is an algebraic integer.

Proposition 2.1.19

If \(K\) is a number field and \(\alpha \in \mathcal{O}_K\) then \(\operatorname {Tr}_{K/\mathbb {Q}}(\alpha )\) and \(N_{K/\mathbb {Q}}(\alpha )\) are both in \(\mathbb {Z}\).

Proof

By Corollary 2.1.9 we know \(C_\alpha \) has integer coefficients and therefore by Proposition 1.7.3 we get the result.

Warning 2.1.20
#

Note that if \(\alpha \) is an algebraic integer then \(A_\alpha \) need not have integer entries. It possible that in different bases the matrix \(A_\alpha \) does not have integer entries. What is true is that in any basis the norm and trace will always be integers as the corollary shows.

Warning 2.1.21
#

The converse is not true. For example \(\frac{\sqrt{1+\sqrt{17}}}{2}\) has integer norm and trace, but it is not an algebraic integer.

Proposition 2.1.22

Let \(\alpha \) be an algebraic integer and \(K=\mathbb {Q}(\alpha )\), then \(\mathbb {Z}[\alpha ] \subset \mathcal{O}_K\).

Proof

Since \(\alpha \) is an algebraic integer and by definition it is contained in \(K\), we have \(\alpha \in \mathcal{O}_K\), from which the result follows by Definition 2.1.13.

Warning 2.1.23
#

If \(K=\mathbb {Q}(\alpha )\) is a number field, then is it not always the case that \(\mathcal{O}_K=\mathbb {Z}[\alpha ]\) as we can see from Theorem 2.1.11. You might ask if its possible to always find some algebraic integer \(\alpha '\) such that \(\mathcal{O}_K=\mathbb {Z}[\alpha ']\) but this is also not true. There are rings of integers that are not generated by a single element. For example, if \(\alpha \) is a root of \(x^3+x^2-2x-8\), then one can show that if \(K=\mathbb {Q}(\alpha )\) then \(\mathcal{O}_K\) is never of the form \(\mathbb {Z}[\alpha ']\) for any algebraic integer \(\alpha '\).

So you may ask, why do we care about \(\mathcal{O}_K\) instead of \(\mathbb {Z}[\alpha ]\)? we’ll it turns out \(\mathcal{O}_K\) is a better invariant of \(K\) as we will see later.

Proposition 2.1.24

Let \(K\) be a number field and let \(R\) be a subring of \(\mathcal{O}_K\) which generates \(K\) as a field (i.e. \(Frac(R)=K\)). If \(R\) has unique factorization, then \(R=\mathcal{O}_K\).

Proof

We know that \(R \subset \mathcal{O}_K\), we will show the opposite inclusion. Let \(\alpha \in \mathcal{O}_K\). Then since \(K\) is the field of fractions of \(R\) we can find \(\delta ,\gamma \in R\) such that \(\alpha =\frac{\gamma }{\delta }\), with \(\delta ,\gamma \) sharing no common factors other than a unit.

Now, since \(\alpha \) is an algebraic integer, we can find some monic polynomial such that

\[ \alpha ^n+a_{n-1}\alpha ^{n-1}+\cdots +a_0=0. \]

If we now multiply through by \(\delta ^n\) we get

\[ \gamma ^n+a_{n-1}\gamma ^{n-1}\delta +\cdots +a_0\delta ^n=0. \]

Since \(R\) is assumed to have unique factorization, we see that if \(\varpi \) is an irreducible factor of \(\delta \), then \(\varpi \) is an irreducible factor of \(\gamma ^n\) and thus is a factor of \(\gamma \). But we assumed that \(\delta ,\gamma \) shared no common factors other than units. So \(\delta \) has no irreducible factors and is therefore a unit. So \(\mathcal{O}_K \subset R\), giving the result.

So, lets try to find a basis for \(\mathcal{O}_K\).

Definition 2.1.25
#

Let \(K\) be a number field and let \(\{ b_1,\dots ,b_n\} \) be a basis for \(K/\mathbb {Q}\). We call this an integral basis if \(\mathcal{O}_K=\mathbb {Z}[b_1,\dots ,b_n]=\{ x_1b_1+\cdots x_nb_n \mid x_i \in \mathbb {Z}\} \) as an additive group. In practice, the set \(\{ b_i\} \) will be some thing like \(\{ 1,\alpha ,\alpha ^2,\dots ,\alpha ^{n-1}\} \) in which case, as rings, we have \(\mathbb {Z}[\alpha ]=\mathbb {Z}[1,\alpha ,\alpha ^2,\dots ,\alpha ^{n-1}]\)

Ok, so we know what we are looking for, so how are we going to make such a basis. For this we will study the discriminant.