1. Introduction
Much of the history of math is trying to broaden our definition of numbers to help solve more equations. Is it possible to set up another number system? A branch of mathematics called abstract algebra makes it possible to understand this question. The biggest motivation in the history of the expansion of numbers is solving equations. When we consider the natural numbers, the equation
can be solved, but for the equation
, we encounter the need to define a new number system, the integers,
that include the solution of this equation. The same applies with the equation
when we consider the integers, so we define the rationals as a new number system where the solution of
lies. The rationals do not include the solutions of
, we define the real numbers by completing the rationals. The fundamental components of abstract algebra are groups, rings, and fields. In field theory, if we can perform basic arithmetic on a set, addition, subtraction, division and multiplication, the set is referred to as a field ( [1] ), the real numbers are an example of a field. Now when one takes into account the field of real numbers,
and the well-known simple equation
with coefficients in
, the solutions of the equation do not lie in
. A natural question that arises is whether there is a larger field where the solutions of this equation lie. The answer proves to be yes, and this larger field is a field of complex numbers. Based on this, we regard the field of complex numbers,
as a field that extends the field of real numbers,
. We’re interested in field extensions because we are assured that given any equation with coefficients in a field, say L, the solutions of this equation lie in a larger field if it does not lie in L. We briefly give an insight into every chapter and highlight some points in each section.
Chapter 1, underlines some important concepts in our paper. We introduce the concept of a ring, which is key in understanding the concept of a field. Finally, the concept of polynomial rings is introduced, which will be encountered more often in our study of field extension, highlighting only key results.
In Chapter 2, we introduce the concept of field extensions, from which we discuss various kinds of field extensions and some results.
In Chapter 3, we sail through the notion of an algebraic extension, here we discuss what it takes to have an algebraically closed field and an algebraic closure of a field. To complement our discussion of algebraic extension, we briefly discuss transcendental extensions.
Chapter 4 introduces the notion of an absolute value on a field and the criterion for a field to be complete. Here we demonstrate that the notion of completion, is one way to obtain field extensions.
In Chapter 5, we conclude our discussion of field extensions and give suggestions of the possible future work.
Basics and Results
We’ll look at some basic ring and field properties in this section. These will come in handy when researching extension fields. Many results, theorems and definitions are taken from [1] and [2] .
Definition 1.1. ( [1] ) Create a set
with the addition and multiplication binary operations. If the following conditions are met,
is referred to as a ring.
1)
is group and for
we have
.
2) Multiplication in
is associative, for all
,
.
3) In
, the two distributive laws apply, for all
,
and
.
Remark 1.2. Suppose that
is a ring and
be elements. If
,
is considered to be a commutative ring. If a multiplicative identity 1 exists such that for any
then
is termed a ring with unity.
is a finite ring if it has a finite number of elements; otherwise, it is infinite. If
is a subset of
and a ring with the same operations as
, it is called a subring of
. We’ll assume the ring
is a ring with unity throughout this discussion.
Example 1.3. The set denoted by,
,
, and
, all contain an element, 1, and they satisfy the axioms stated in Definition 1.1, they are rings with unity and they are infinite.
Definition 1.4. ( [1] ) Let two nonzero elements
and
be elements of ring
. If
is commutative, then
is referred to as a zero divisor if
.
Suppose that
is a commutative ring, If
contains no zero divisors, it is referred to as an integral domain, say ID. If in an ID, every nonzero element can be expressed uniquely as a product of irreducible elements (or prime elements), it is referred to as a Unique Factorization Domain, and we write (UFD).
Definition 1.5. ( [1] ) A mapping between two rings
is referred to as a ring homomorphism if the following conditions are met, for all
, then
1)
2)
Elements of a ring
that map to the additive identity, 0 form a set and it is referred to as the kernel of the ring homomorphism,
. If the mapping in Definition 1.5 is bijective, then it referred to as an isomorphism we write
Definition 1.6. ( [1] ) A subring of a ring
is said to be an ideal, I, of
, if
for ever
and
.
For a ring
and its ideal I, then the following two operations are defined in the quotient group
, suppose
, we have
and
.
is referred to as the quotient ring of
by the ideal I. It is considered to be a principle ideal if I is created by a single element, say
, and we write
. We refer to an integral domain that contains ideals that are principal, as a principal ideal domain. Suppose that there are no ideas between the ideal I and it's ring
, the ideal is maximal ideal of
.
Definition 1.7. ( [1] ) A field F is a nonzero commutative ring such that
is a group under multiplication.
In any field, we are assured of 0 and 1, and
, we therefore have that a field contains at least two elements. A subset of a field, L is has the operation of L and it is referred to as a subfield. A mapping of fields satisfying the axioms stated in Definition 1.5 is called a homomorphism of fields and it is injective because its kernel is a proper ideal.
Example 1.8. Since
,
contains the field
, hence its a subfield. Also
contains
, thus,
is also subfield of
, by transitivity property.
Suppose the multiplicative identity and the additive identity of a field L are denoted by 1L and 0 respectively, then the field L contains elements of the additive group generated by 1L. As we generate the elements of L, we might get an additive identity in the process. Suppose the least number of times we have to add 1L to get 0 is k, then k is the characteristics of the field, L. If such a positive integer does not exist, the characteristic of, L, is said to be zero.
Definition 1.9. ( [1] ) Let L be a field. The least positive integer, k that satisfies
is referred to as the characteristic of L, and we write
.
if such a k does not exist.
Proposition 1.10. ( [1] ) Let F be a field. The characteristic of F is either 0 or a prime number p.
Proof. Assume that the characteristic of F,
, assuming
, then
. Now let
, where
, then we have that
if and only if
or
, the fact that k is the lowest positive integer, we have a contradiction. □
Theorem 1.11. ( [2] ) For a maximal ideal I and the ring
,
is a field.
Proof. See [2] . □
Polynomial Rings
Definition 1.12 [2] Consider a ring
. If
is commutative and has 1 then an object
over
, where
is referred to as a polynomial.
From
, we call
the leading coefficient and
is an n-degree. For
to be monic,
in
and the collection of all polynomials is denoted by
. The addition is done component-wise and when we consider,
from Definition 1.12 and
, multiplication is defined in this manner,
where
,
,
, when we generalize, we have
. From the above argument, we have that
, satisfies the axioms stated in Definition 1.1 and hence its a ring and its referred to as a polynomial ring.
is referred to as an integral domain, provided
defines a ring [1] .
Proposition 1.13. ( [1] ) Suppose that L satisfies the axioms of a field. Then its polynomial ring,
qualifies to be to be called a principle ideal domain and hence a unique factorization domain.
Remark 1.14. For the polynomial
, the solutions of this polynomial given by
and
are referred to as roots of the polynomial. If
we say,
is a constant polynomial.
Definition 1.15 (contemporary) Suppose that L satisfies the axiom of field and from its polynomial ring,
, let
be a nonconstant polynomial. If we cannot write
in the form
where
are of degree less than
, we say,
is irreducible.
We can express
with coefficients in the rationals as
.
is irreducible in the rationals but not in the reals.
Proposition 1.16. ( [1] ) Suppose that the non constant polynomial
is irreducible, then the ideal,
is a maximal ideal.
2. Extension Fields
The concepts discussed in the previous chapter help us to introduce extension fields. As the main component of this section, we familiarize ourselves with field extensions by discussing simple extensions, finite extensions, and splitting fields. Many results, definitions, and theorems in this section are those in [1] [2] [3] and [4] .
2.1. Construction of Extension Fields
Definition 2.1. ( [1] ) Suppose that the field L contains the field K, we refer to the field L as extension field of the field that it contains.
The commonly used notations are
,
and a diagram that depicts a the larger field on top of the base field as shown below. We adopt the first notation for our discussion.
We now give an example of field extensions.
Example 2.2. The field
contains the fields,
and
, so we write
and
. Similarly the field
is contained in the field
, so we have
.
Definition 2.3. ( [1] ) Let F, L and K be fields and
. Then the field L is referred to as a subextension of the K that extends F.
Remark 2.4. From Example 2.2, the field
extends the fields
and
and since we have the inclusion
we have that
is a subextension of the field
that extends
.
We now present a statement and demonstration of the Fundamental Theorem of Field Theory, also known as Kronecker’s Theorem, which establishes the existence of an extension field.
Theorem 2.5 (Fundamental Theorem of Field Theory). [1] Let L represent a field. From the polynomial ring
, let
be nonconstant. Then there exists a field, K that extends the field E. if
, then
Proof. Suppose that L is a field, its polynomial ring
is a UFD. From
, let
be a non-constant, it can be expressed in terms of some irreducible polynomial in
. If one of these irreducible polynomials is,
, then the ideal
is a maximal ideal. And we have that the field
We asserts that L is contained in K and we now define the map
given by
We see that for all
and also
The map is a ring homomorphism. Because we associate L with its image
in K, we’ll use
instead of
for
. We have
, so that s. Since L is contained in K, K is a field that extends L.
For an element
in K, we write
. Now from the polynomial ring,
, let
we have
Therefore
satisfies
.
Example 2.6. For the field
, Consider the polynomial
, and from
, let
, since
is irreducible. The ideal,
is maximal. so we have the field
. By Theorem 2.5, K is the field that extends
and contains the solution of
. We can write
If
satisfies
we van write it as
in K then we have,
Then solving for
in
we have
in K and so
Remark 2.7. The example above gives an illustration of the construction of complex numbers with the help of Theorem 2.5. From the construction, we have an isomorphism between the field
and the field
.
2.2. Simple Extension
In the previous section, we have considered a type of field extension obtained by considering a polynomial and an element that is a root of this polynomial. Can we have a construction where we do not consider the polynomial but the element from the larger field? This question is worth exploring. In this section, we consider this kind of construction, where we only consider an element and adjoin it to the base field. Many results, definitions, and theorems in this section are those in [1] [3] and [5] . We now give a lemma which is a motivation of the definition of simple extensions.
Lemma 2.8 ( [1] ). Suppose that F is a subfield of E and
. Then there exists a unique smallest subfield of E containing both F and
.
Proof. Suppose that
and an element
. Then define the set,
by
as the collection of subfields that contains the field F and an element
. Define the intersection
then for an element,
, there exists elements
and
such that
Now
and
for all i. This implies that
and by the uniqueness of the inverse,
turns out to be a subfield of an extension, E that contains the field F and the element
and it is the smallest such subfield that contains them both. Since
, F is a subfield and the intersection of these subfields is , L is the unique smallest subfield of the extension, E.
The lemma above holds true if
is replaced by
Definition 2.9 ( [1] ). For an extension field,
, consider the elements
. Then E has a subfield with the notation
that contains the elements
and the field F. It is the smallest and it is therefore to referred to as field generated by
.
Definition 2.10 ( [1] ). Suppose that the elements
, in Definition 2.9 are replaced by
and we write
. Then
is referred to as a simple extension with the base field F. The element,
that generates the field is called a primitive element.
Remark 2.11. For the fields
, and K, consider
. We have
as the intermediate field. The elements of
are polynomials in
, for
, we consider an n degree polynomial,
. Suppose that
, we write
where now the coefficients
. Some of the elements of
are not polynomials in
, if
in
. We can express the field
as a field containing the ratios of the polynomials defined above. We have
Example 2.12. Consider an extension field,
and an element i such that
, then we have the inclusion
The field
contains the element i and
. Since i is a root of some polynomial in
, we can write the elements of
as,
thus
. The simple extension generated by
is expressed as
From the field K that extends the field L, we can construct the smallest field
that contains F and the element
in the extension field K. By Theorem 2.5, if
for some irreducible polynomial,
, there exists a field that extends the field, L and contains
, and the field is,
From the field L, we have two extensions field that contains the root
,
and
. We now explore the relationship between the two fields.
Theorem 2.13 ( [1] ). Suppose that E is a field and the nonconstant polynomial,
with coefficients in E is irreducible. Then if
and
, we have
. Define
as the simple extension over E. Then
That is up to isomorphism, the smallest extension of E,
contains a root of the polynomial
.
Proof. Define an evaluation homomorphism
by
. When we restrict
on E, we have the identity map,
and
is a ring homomorphism. Since
,
and so
, therefore
. We now define the induced homomorphism
by
where
is any arbitrary polynomial and
is well defined. Since any homomorphism between two fields is identically zero or injective, we have that
is either a zero map or is injective. Now since
(1)
and
we have that
is injective. We now have
We now show that
is the image of
. From Expression 1,
Now
,
is the smallest field that contains the field E and
, then we have
But
has
has its subfield, so we have that
Therefore
Example 2.14. Consider
. Since
is irreducible, the quotient,
forms a field. Now if
, then
. By Theorem 2.13, we have by taking the positive root of 2 in
Remark 2.15. By Theorem 2.13, we have the isomorphism between the field
and the field
and by Example 2.14 we have the relation
. Transitivity tells us there will be an isomorphism, and we have that
under the isomorphism
Corollary 2.16. Let
be an irreducible polynomial. Isomorphic fields are those generated by adjoining roots of
.
Lemma 2.17 ( [1] ). for the field L, the field
generated by x and y over L is the field
generated by y over the simple extension
.
Proof. For an extension field
, let the elements
, then by Definition 2.9, the field
, generated by the elements x and y contains the field L, and the elements x and y, thus it is smallest among the subfields that contains the two elements and the field L, hence contains the simple extension
. Since the field contains the simple extension
and the element y it contains the field
and we have
(2)
Similarly, the field
generated by x is the smallest among the subfields and contains the simple extension
and y and thus contains L, x, and y. we have
(3)
From Equation (2) and Equation (3), we have the equality of the two fields
□
Example 2.18. Consider the extension field
over
, then the field
contains
and
and it is the smallest among the subfields of the extension field
having
,
and the field
. Also
is an extension over
and it is obtained as follows
2.3. Finite Extension Fields
Since any field say L contains a prime field, the multiplication defined in F makes L into vector space over it’s prime field. Similarly when we consider the extension of the field E over L, E is a vector space over L. Most results, definitions, and theorems in this section are those in [1] [3] and [5] .
First we give the definition of the degree of an extension before we define finite extension fields.
Definition 2.19. Suppose
is an extension field. Then K is vector space over L and the dimension of this vector space is referred to as the degree of the extension, K and we write
.
Definition 2.20. Let K be a field that extends the field L. If the degree defined in Definition 2.19, is finite, then K is finite, otherwise, K is infinite.
Suppose that for the extension
,
, then K has degree one. If the degree of the extension is 2 and 3, the extension is called a cubic and quadratic extension respectively. Suppose that
satisfies a n-degree minimal polynomial,
over a field
, then we write
Theorem 2.21 ( [1] ). Let K be a field that extends the field L and
. If
, then the simple extension,
is a vector space over L has the basis
.
Proof. Suppose that
and
satisfies a polynomial
in the polynomial ring
. Then consider the set
If the set
were linearly dependent, then there exists some polynomial say
with the degree of h less than n. Since
is the minimal degree for a nonzero polynomial in
. If given
, there exist polynomials
and when we employ the division algorithm,
and
is a minimal polynomial of the root
over the base field L. Clearly
, and
is the image of
and we write
It then follows that
is a spanning set and therefore a basis of
. □
Example 2.22. Consider a field extension
then the element
is in the extension
. We have that
implies
, so
. Therefore the
. By Theorem 2.21, the basis of an extension
is
.
Theorem 2.23 ( [1] ). Suppose that the extensions E and K are finite extensions of the fields L and E respectively, we have the multiplicative relation given by
Proof. We define a basis for K over E as the set
and for E as
. It suffices to prove that the basis for K over L is the multiplication of the two basis given by
For the elements
and
, we have a linear combination,
there exist elements
and we have the linear combination
Therefore,
which proves that AB is a spanning set of the
. We assume that
Since each
and that the basis for an extension
is A,
for each i. Now we have that each
and the basis for an extension
is B, it implies that each
which proves we have linearly independent set and it given by AB. □
Remark 2.24. We note that for the extension fields
and
, if
and
where I is the indexing set are basis of the extensions respectively, then for fields
, the set
of length mn is a basis for K over L.
Corollary 2.25 ( [3] ). Suppose that
are finite field extensions, then
Proof. When we employ induction, the proof clearly follows from Theorem 2.23,
□
Example 2.26. Consider an extension
over
. Then we have the inclusion
By Theorem 2.23,
Let
. Then
we have that
and degree
. By Theorem 2.21, we have that the extension
has a basis
and
. We now have,
Similarly, let
, that is
and
. So
. By Theorem 2.21,
is a 3-dimensional vector space
. Now by Theorem 2.23, we have that
The product of the basis
and
gives the basis of an extension
and it is given by
2.4. Splitting Fields
Theorem 2.5, gives us a criterion to construct extension fields. If the field K extends L and
is a nonconstant polynomial from the polynomial ring
, K has a root of
. If the root of
is
, then we have a factor
. For the nonconstant polynomial,
, we can therefore write
. K being a field implies that we can repeat the process, in doing so, we obtain a field E that extends K, and for
we have
. In so doing, we can find a field, En that extends other fields and this fields contains all the roots of
, which can be expressed as a product of linear factor ( [1] ).
Definition 2.27. Let the field E extend the field L and the nonconstant polynomial
. if
can be factored completely into linear factors in the polynomial ring
and over any proper subfields of the base field,
fails to factor completely into linear factors, E is the splitting field for
and is the smallest such extension that contains all the roots of the
.
We now give the following example to demonstrate the above definition.
Example 2.28. Consider field
that extends the field
. Then the polynomial
can be expressed as
. Thus
has been expressed into linear factors, we conclude that the extension field
is a splitting of the polynomial,
. Similarly, for extension field
over
generated by
and
consider the nonconstant polynomial,
, this polynomial can be written as
. We have that
is the splitting field for
.
Theorem 2.29 (Existence of splitting fields) ( [1] ) Given any field, say L and the nonconstant polynomial
with coefficients in L, there exists a field, K that extends L. K is the splitting field for
.
Proof. It suffices to show first that there is a field, K that extends the field, L in which the polynomial,
of degree n factors completely into linear factors. This is accomplished through induction on the of
, if we have that
, then the degree of
is one. Assume
has degree more than n and F being a field implies that the polynomial ring
is a unique factorization domain and so we can express
into linear factors completely as product of irreducible polynomial. If all the irreducible factors are of degree 1 again. Otherwise at least one factor say
has degree more than 1. Let the degree be 2. By Theorem 2.5, we have a field K1 that extends the field L and contains the root of
. This implies that over K1, the two polynomial
and
has a linear factor say
. If
is the remaining factor of
,
is its degree. By induction, we have a field K that extends the field K1 and
factors completely. Since
, this implies that
and K being an extension of L,
has all its roots in K.
Suppose we have a field E that extends all subfields of K and contains L in which
has it’s root, then E is the splitting field
. □
We now provide the following theorem without proof which tells us that splitting fields are unique.
Proposition 2.30 ( [1] ). If there are two splitting fields, they are isomorphic.
Proof. See ( [1] ) □
We now present a statement and demonstration of the theorem relating the degree of a nonconstant polynomial,
, and that of its splitting field, and a demonstration of the theorem follows.
Theorem 2.31. Suppose that a polynomial,
of degree n splits over the field L, then its splitting field has a degree of at most
.
Proof. Suppose that
is an n-degree polynomial in
and
. Then the simple extension,
is of degree atmost n. If
is irreducible over
, then
has degree less than n. Over
,
has at least one linear factor, so any other root of
say
satisfies an
degree polynomial
, thus,
Using Theorem 2.23, if
splits in K, then
□
Example 2.32. From Example (2.28), consider an extension field
over
, then the polynomial
of degree 2 in the polynomial ring
splits in
. By Example (2.22),
. Similarly, the polynomial,
with coefficients in
is of degree 4 and it splits in
. Thus,
has degree 4.
3. Algebraic Extension
Having set up the theory of field extensions, we have considered different kinds of field extensions. By Theorem 2.5, we where able to find an extension field say E over F by considering a polynomial
and an element
such that
. Can we have field, K that extends L containing only those elements that satisfy some polynomial with coefficients in L? In this Chapter, we explore this kind of fields and it’s associated properties. Results, definitions and theorems are mostly from [1] [3] [5] and [6] .
3.1. Properties of Algebraic Extension
We can classify the elements of an extension field into two categories, and with this classification, we formulate the definition of the algebraic extension.
Definition 3.1 ( [1] ). Let K be a field that extends L. If
and there exists a nonzero polynomial
such that
satisfies
. then
is algebraic over L.
We say that
transcendental if it is does not satisfy any polynomial coefficients in the base field.
Example 3.2. Consider the field
that extends the base field
. From the polynomial ring
, let
and since
,
in
is algebraic over the base field. Whereas
and e are transcendental over
since
and
.
Remark 3.3. An element from the extension field is algebraic or transcendental, depending on the base field. From Example 3.2, the elements
, and e are transcendental over
, now for
and
both in
they are algebraic over
since
and e satisfy the polynomial
and
respectively.
Now we’ll look at the subfield of algebraic elements.
Corollary 3.4 ( [1] ). For a field K that extends L, there exists fields contained in K, that contains the set of all elements of K defined in Definition 3.1.
Proof. Suppose that the field K that extends L contains elements defined in Definition 3.1 and these elements are
and
. We now show that
are elements defined in Definition 3.1. To do this, it suffices to show that the field
has a finite degree since it contains all these elements. By Theorem 3.22, we have
Now since the element
is an element defined in Definition 3.1, it is algebraic the simple extension
. Therefore, we have that the field,
that extends the field
and the
that extends L are of finite degree.
□
The subfield of elements from Corollary 3.4 defines what we call a relative algebraic closure of field within a field that extends it and this, we'll present in the preceding section. With the tools at handy we describe the notion of an algebraic extension and find it’s relation with finite extension.
Definition 3.5 ( [1] ). Suppose that field K that extends the field L contains elements defined in Definition 3.1, then K is referred to as an algebraic extension of L.
Note that if the field that extends the base field is a not an extension defined in Definition 3.5, it is said to be a transcendental extension.
Example 3.6. Consider the element i such that
in the field,
that extends
and the polynomial ring
. From
define
We now have that
, hence i defines an element in Definition 3.1. However not all elements of
are algebraic over the base
, e.g.
, hence
is a field defined in Definition 3.5. From Example 3.2, we have that over
, neither
nor
is a field defined in Definition 3.5 .
Definition 3.7 ( [3] ). Let K be the field that extends the field L and let
be an element defined in Definition 3.1. From the polynomial ring
, let
be an irreducible monic polynomial and
. Then the polynomial
is said to be minimal.
Proposition 3.8 ( [3] ). Over the field L and element
, define a finite simple extension generated by
,
. Then the element
is algebraic over L.
Proof. Suppose that the element
satisfies some polynomial with coefficients in L, then over L , let
be a polynomial defined in Definition 3.7, the degree of this polynomial is equivalent to that of
. Hence the field extension is finite, and has degree at most n if the element
satisfies a polynomial of degree k. Conversely, suppose now that over L,
is an element of a k-degree simple extension. Then the extension
has
roots and they are linearly dependent. Suppose the roots are
and for all
in L , we have a linear combination,
For a nonzero polynomial,
of degree at most n with coefficients in L,
, thus
is algebraic over L. □
The next theorem shows that finite implies algebraic.
Theorem 3.9 ( [3] ). Let K be a field that extend the field L. If
algebraic extension, then
is finite.
Proof. Suppose that the field K extends the field L and contains the element
. Over L, the simple extension
is contained in the field K and defines a subspace of the vector space K. Hence the field
has degree at most the degree of the K. By Proposition 3.8, the element
satisfies some polynomial with coefficients in L. □
Theorem 3.10 ( [1] ). The extension field E over the field F is finite if and only if E is generated by a finite number of algebraic elements over F.
Proof. See [1]
□
In our next theorem, we prove the statement that algebraic over algebraic is Algebraic
Theorem 3.11 ( [1] ). Suppose that over the field L, the field K is an algebraic extension and over the field E , the field L is an algebra algebraic extension . Then over E, L is an algebraic extension.
Proof. Let the field K, contain the element
. Then
satisfies some polynomial,
with coefficients in L,
(4)
where
. Now over E, consider the field
generated by
the coefficients
’s of the polynomial
.
Since over E, L is an algebraic extension, the elements
’s satisfies some polynomial with coefficients in E, and so the extension
,
is finite by Theorem 3.10. Now by Equation 4, we have that the element
generates an extension field of degree at most k, since it’s minimal polynomial over this field is a divisor of the polynomial above, we have that
where
is also finite and we have that over E, K is an algebraic extension since the element
satisfies some polynomial with coefficients in E. □
3.2. Algebraic Closure of a Field and Algebraically Closed Fields
Definition 3.12. [1] Let L and
be two fields. Then if over L,
is an algebraic extension field and from the polynomial ring,
, all polynomials splits completely.
is said to the algebraic closure of L.
We can also define an algebraic closure in terms of a set.
Definition 3.13. Suppose that the field K extends the field of the L. We define the algebraic closure of L in the field K as
The element
satisfies some polynomial with coefficients in L, thus all elements contained in
are algebraic over the base field, L.
Definition 3.14 ( [1] ). Suppose that L is field and from the polynomial ring,
, every nonconstant polynomial has a root and this root is contained in K. Then K is referred to as an algebraically closed.
Lemma 3.15. Suppose that the field L equals to its algebraic closure
, then L is algebraically closed.
Proof. When we assert that L is algebraically closed, then from the polynomial ring
, we can choose any polynomial,
. Suppose that
, then
is a factor contained in
. For
, we can now express
as
All the roots are in L. Thus, over L,
splits completely and hence L equals
. Conversely, if we assert that
, it follows immediately that L is algebraically closed. □
Corollary 3.16. Suppose that the field
is the algebraic closure of the field L. Then we have that it is algebraically closed.
Example 3.17. Consider the field,
that extends the field
, then,
algebraic closure of
, and hence its algebraically closed. The
is not an algebraic closure of itself, since when we consider the polynomial
has roots
which are not contained in
, hence its not algebraically closed.
From the uniqueness of splitting fields, we have that algebraic closure of a field is unique up to isomorphism and we know that that if the field
is an algebraic closure of the field L, then
and this field is algebraically closed. Now, the question that may arise is, can we have an algebraically closed field for an given field? Our next proposition helps us understand this question.
Proposition 3.18 ( [1] ). Suppose that L is a field, then exists an algebraically closed field, K that extends the field L.
Proof. See [1] □
Example 3.19. Consider the field
, that extends the field,
and
.
is the algebraic closure of
and
, hence its algebraically closed.
Proposition 3.20. Suppose that the field K contains the field L. If K is algebraically closed, then we call the set
an algebraic closure of L and its defined as
Proof. Suppose that
contains the elements that satisfy some polynomial with coefficients in L. Over L, Definition 3.5 implies that
is algebraic. From the polynomial ring
, every polynomial, say
factors completely into linear factors over K. This also holds true for every polynomial in the polynomial ring
. Now
is algebraic over L since
. We have that
is contained in
. We now have that all linear factors have coefficient in
, which implies that that
factors completely in
and this means that we have an algebraic closure,
of L. □
Algebraic Closure of Finite Fields
In order to describe the algebraic of finite fields, we first presents some results on finite fields. If the elements in a field,
are of a finite number, we say that
is a finite field. Consider the following set of integers
, these set of integers are maximal ideal of the ring of integers and when we get the quotient with ring
, we have the fields
for
. The fields obtained have a finite number of elements
respectively. From the construction, we see that if we consider any prime we can construct these fields and we write
to denote these kind of fields.
Lemma 3.21 ( [1] ). Suppose that the field
contains the field
. If
has q number of elements and over
, k is the degree of
. We have that
has
elements.
Proof. Suppose that
has a finite number of elements, then over
,
is a finite extension field and hence a vector space with a finite dimension. Assume that over
, k is a dimension of
that is
. We have that over
, the set
is basis and the
’s are linearly independent. For any
and
we have a linear combination of the basis, and we write
There are q choices for each
and there are
’s. We have that
has
choices and we conclude that the finite field F must have
elements. □
Theorem 3.22 ( [1] ). Suppose that the field
has a finite number of elements and let the prime p denote its characteristic. Over its prime subfields, let k be the degree of
, then
has
elements.
Proof. Suppose that
has a finite number of elements, then its characteristic is a prime, p. There is an isomorphism between
that contains p number of elements and the prime subfield that has characteristic p. Then we have that
and
. By Lemma 3.21, we have that the field F must have
elements. Hence we conclude that all finite fields must have prime power orders.
□
Remark 3.23. From Theorem 3.22, a finite field cannot have 6 elements because the number 6 is not a power of any prime.
Lemma 3.24 ( [1] ). Suppose that the field
has q elements and for any
. Then
Proof. For the case
, it follows that
. Under multiplication, the nonzero elements of
forms a group and its order is
. We denote this group by
and we have that
. From group theory, we have that
which implies that
and we conclude that the lemma holds. □
Lemma 3.25 ( [1] ). Suppose that the field
has q elements and contains the field
. From the polynomial ring,
, let the polynomial
. Then
splits in
, and for
, we write
Proof. Suppose that
splits in
,
is a splitting field of
. There are at most q roots in
since
is a q-degree polynomial. By the previous lemma, Lemma 3.24, the polynomial,
is satisfied by all the elements of
. Therefore for
, we write
Over a proper subfield of
, that contains
the
does not split, as such a field would have fewer than q elements. □
Theorem 3.26 (Existence and Uniqueness of Finite Fields). [1] Consider an integer
and prime number p there exists a field with
elements and its referred to as a finite field. Let
and from the polynomial ring
, let
. Then there is an isomorphism between the splitting field of
and the field
.
Proof. See [1]
Theorem 3.27 (Subfield Criterion for finite fields). Let
be a finite field with
elements. Then every subfield of
has order
where m is a positive divisor of n then there is exactly one subfield of
with
elements.
Proof. see [1] .
□
Example 3.28. The subfield of the field
are
,
,
,
,
,
and
since 30, 15, 10, 5, 3, 2 and 1 are positive divisors of 30.
Lemma 3.29. Suppose that the field
is finite with k elements. Then not a every polynomial with coefficients in
has a root, that is
is not algebraically closed.
Proof. We asserts that
has k elements, then
is finite. Let the elements be
. From the polynomial ring
define the polynomial
by
Then for all
,
. Therefore, there is no element of
that is a root of
. Since
exists, we conclude that
is not algebraically closed. □
For a finite field of prime order. We consider an integer
and a prime p. Suppose that
, we have the field
. Since
is a field, it has an algebraic closure, say
and so,
contains
by definition.
that contains
. Also since there is a unique field extension
over
and
is contained in
. There is an isomorphism between the
and one field contained in
. If we assume that this subfield is isomorphic to
, we have the inclusion
The union
is an algebraic closure of
. This field is a countable union of arbitrarily large finite fields ( [7] ).
3.3. Transcendental Extensions
The previous section introduced an algebraic extension in which all its elements are algebraic over the base field. If an extension field is not algebraic, then it’s said to be transcendental, and in this section, we introduce this type of field extension. Results, definitions and theorems are similar to those in [8] and [9] .
Definition 3.30 ( [10] ). Let the field
extend the field
. Then we say that
is a transcendental extension if there exists at least one element,
that is not a root some polynomial with coefficients in
.
Example 3.31. Consider the two extensions
over
and
over
, then
and
are transcendental extension of
since from Remark 3.3, the elements
and e are transcendental over
.
Definition 3.32 ( [10] ). Suppose X is a transcendental element over a field
. Then the rational function field over
in one variable Z is the field containing all rational functions and it is defined as
Definition 3.33 ( [10] ). Suppose the field
extends the field
. We define an algebraically dependent set if from the extension
, a subset
satisfies some polynomial,
in
, that is,
A subset S is said to algebraically independent over
if it not algebraically dependent over
. A given set of elements is said to algebraically independent if it does not satisfy some algebraic relation and it is said to be algebraically dependent if it satisfies a polynomial or some algebraic relation.
Example 3.34. Consider the two sets
and
, then over the field
, these two sets are algebraically dependent, since they satisfy the polynomials
and
respectively from the polynomial ring
. But the sets
and
are algebraically independent over
and
respectively.
Remark 3.35. If an element say
is transcendental over any given field, then the set containing
is algebraically independent over the given field. If the set S is algebraically independent over any given field, then it’s subset is also algebraically independent. Thus, we can say that an empty set is algebraically independent over any given field. On the other hand, if S is any set that contains an algebraically dependent set, then S is algebraically dependent.
Corollary 3.36 ( [10] ). Suppose that
is a field and over
, define the set
. If
is algebraically independent, then we have an isomorphism defined by
Definition 3.37 ( [11] ). Suppose that the field
extends the field
and the set
. If
is the maximum of all the algebraically independent subsets of
, we say that
is a transcendental basis of
over
.
If the basis of an extension defined above is empty,
is an algebraic extension.
Example 3.38. Consider the extension field
over
, then set defined by
is a basis of this extension. over
, the function field
, has the set
has its transcendental basis.
Proposition 3.39 ( [11] ). Suppose that the field
extends the field
and over
, define an algebraically independent set,
. Over
, the set
is algebraically independent
an element
, is transcendental over
.
Proof. We assert that over
, the set
is algebraically independent and
satisfies some polynomial with coefficients in
. We therefore have for
The relation shows that
is algebraically dependent which is a contradiction. Conversely, suppose that over
,
satisfies some polynomial,
. Now each coefficient of this polynomial is an element of
, Clearing denominators yields a nonzero polynomial
. Therefore this polynomial yield and algebraic dependence in
, and over
, we have that
is algebraically independent if and only if
is transcendental.
Corollary 3.40 ( [10] ). Suppose that the field
extends the field
and over
, define an algebraically independent set,
. If
is algebraic over
,
is a transcendental basis.
Theorem 3.41. [11] Suppose that
and
are two transcendental basis. Then cardinality of
equals the cardinality of
.
Proof. See [11] □
Definition 3.42. Suppose that the field
extends the field
and
defines a transcendental basis over
. If
we say that
is purely transcendental.
Definition 3.43 ( [10] ). Suppose that the field
extends the field
and
defines a transcendental basis over
. Then the cardinality of
is referred to as the transcendental degree of
over
. We write
to denote the transcendental degree.
Example 3.44. From Example 3.38, the transcendental basis of the extension field
over
is the empty set
which implies that the transcendental degree of
is zero. In general, the transcendental degree of an algebraic extension is zero and we can conclude that all extensions are transcendental extensions.
Theorem 3.45. Suppose that
and
are fields and the inclusion
be a tower of fields. Then
Proof. See [11] □
4. Absolute Values and Completions
In the previous chapters, we have considered different kinds of field construction in which the base field is taken to be any arbitrary field. In this chapter, we consider two kinds of base field,
and
and introduce the notion of an absolute value on these base field and use this concept to construct an extension field. Most results, definition and theorems are taken from [12] [13] [14] [15] .
4.1. Absolute Value over
and Completion of Fields
Definition 4.1. [14] Let E be a field. A mapping
defined by
where
,
is called an absolute value over E, if it has the following properties;
1)
and
if, and only if,
2)
, for all
3)
for all
(Triangle inequality).
Definition 4.2. [14] Suppose the absolute value in Definition 4.1 has an extra property called the strong triangle inequality for any
given by,
then, it is called nonarchimedean. If an absolute value is not nonarchimedean, it is said to be archmedean.
Example 4.3. Consider the usual absolute value defined by
which is called the trivial absolute value. For real numbers and rational numbers, we have the mappings
and
both defined by
Remark 4.4. The absolute value on the complex numbers is defined as
If we assume that both a and b are the same, then the strong triangle inequality is not satisfied in each case and we therefore have that these absolute values we have defined are all archimedean.
We now present the general properties of absolute values on a field.
Lemma 4.5. [13] For a field F. Let
be the absolute value on F. Then we have,
1)
2)
3)
Proof.
1)
2)
3) We assert that
and
, so that
we obtain,
Definition 4.6. Suppose that
is a field and on
, define the two nontrivial absolute values
and
. we say that
and
are equivalent if, and only if, for every
there is a positive real number,
such that,
An immediate result from Definition 4.6 is the following corollary.
Corollary 4.7. For a field
, two nontrivial absolute values
and
on F are equivalent if for all
we have,
We now introduce the notion of a metric space, convergence of a sequence and Cauchy sequence.
Definition 4.8. Suppose
is a set. A mapping
satisfying the following axioms, for every
;
1)
and
2)
3)
is called the metric on
.
Remark 4.9. A metric space denoted by
is a set
equipped with a metric
. The Definition 4.1 of an absolute value on a field looks similar to Definition 4.8 of a metric and thus we can employ the notion of an absolute value together with a field to form a metric space. Let us consider an ordinary absolute value on
, then we have a metric space
and we can define the metric on the rational numbers
as
Definition 4.10. [15] Suppose a field F is equipped with
. A sequence
with its elements taken from F converges to a limit
, if there is an element
such that for any
there exists a natural number M, for all
we have,
If the limit in the above definition exists, it is unique.
We now provide a theorem below without proof.
Lemma 4.11. [15] Suppose F is a field and the sequences
and
in F converge to s and t respectively. Then, the sequence
and
converges to
and
respectively.
Proof. See [15] .
Definition 4.12. Suppose F is a field and
a sequence in F.
is called a Cauchy sequence if for every
, there is a positive integer M such that for any
,
The notion of a Cauchy sequence can be thought of as a sequence in which elements are closer to each other, that is, the difference between the elements is minimal. We now find the relation between a sequence that converges and Cauchy sequence.
Theorem 4.13 ( [15] ). Suppose that the sequence
, then
is Cauchy sequence.
Proof. Suppose the sequence
converges to a limit say s, then for every
there exists a positive integer M such that for all
Now for any
, we have,
□
The converse does not necessarily hold true, it depends on the field.
Definition 4.14. [15] Two sequences
and
are equivalent if for every
, there is a natural number M such that for every natural number
,
From Definition 4.14, we can derive the definition of an equivalence relation on the set of all sequences in the field F. We denote the equivalence classes of the sequence,
by
.
Definition 4.15 ( [15] ). Let L be a field. The field
whose elements are equivalence classes of Cauchy sequences in L is called the completion of L.
Remark 4.16. The field we have just defined is indeed a field since we have all the operations, that is
1)
2)
3)
4)
5) for any nonzero element
of the completion field, we define another element say
in the completion field by
such that
.
Definition 4.17. [15] Let S be any subset of a field L equipped with
. The set S is dense in L if for every element
and every
, there exists an element
such that
Theorem 4.18 ( [15] ). Let L be a field equipped with
and
it’s completion. Then L is dense in
.
Proof. Let
be the equivalence class of Cauchy sequences
in L. Then for every
, there exists an
such that for all
we have the property
. It follows that
, where
is just the equivalence class of
. □
Theorem 4.19. [15] Every Cauchy sequence in the completion field
is equivalent to a Cauchy sequence whose elements lie in F.
Proof. Let
be a Cauchy sequence in
. Since F is dense in
, for each
we pick
so that
Then for any
, we pick M such that for all
,
and
.
It then follows from the triangle inequality that for any
,
thus, the sequence
is Cauchy. □
Theorem 4.20. [15] Suppose L is a field equipped with
. Then there is a complete field
with
that extends L. This completion
is unique up to isomorphism. Moreover on L,
restricts to
. Lastly, L is dense in
.
Proof. (Sketch of the proof) Since some Cauchy sequence in L does not have a limit, the limit should exits in the completion field. So general idea is for each element of the completion field
to be a limit of Cauchy sequence of elements in L. □
Throughout our discussion, we have been talking about the ordinary absolute and the trivial absolute value and we have shown that the ordinary absolute value is archimedean. Can we have a nonarchimedean absolute value on the field of rational numbers? The definition below helps us understand this question.
Definition 4.21. [12] For a prime p, define an absolute value on the rational numbers
as follows, for all
, let
with
, p divides neither
nor
and
then,
is the p-adic absolute value on the rational numbers
.
The p-adic absolute value is indeed an absolute value and it also satisfies the strong triangle inequality condition for absolute values and hence it is nonarchimedean.
Example 4.22. Consider a prime number say
and
. Then we have that
We now have an idea of a p-adic absolute value on
. Does there exist another absolute value on
apart from the p-adic absolute value and the ordinary absolute value? The next theorem tells us more about this question.
Theorem 4.23 (Ostrowski). [10] Let
be an absolute value on the rational numbers. Then
is trivial or it’s either equivalent to the usual absolute value or some p-adic absolute value.
Proof. See [10] □
The Theorem 4.23 tell us that we can only find two absolute values on
, namely the ordinary and the p-adic absolute value.
4.2. Finite Extension of
After establishing the notion of completion and absolute value on
, we can think of completion of a field as an extension of a field.
Example 4.24. The rational numbers with the ordinary absolute value
is not complete. Consider a sequence
. The sequence
is a Cauchy sequence in
but it is not convergent as the limit of this sequence is
which is not in
.
With the tools at hand, we can now complete the rationals with respect to the ordinary absolute value and the p-adic absolute value.
Remark 4.25. Consider the rational numbers
and the ordinary absolute value
on
, we have a metric space
. From Example 4.24, we see that
is not in
so we now get a set of all Cauchy sequence of rational numbers. For
to be included and all the missing limit, we use the equivalence class of a Cauchy sequence with respect to the equivalence relation. We now obtain a new field which we call the field of real numbers.
Corollary 4.26. [14]
is the completion of
.
Remark 4.27. From Corollary 4.26, we have the statements below:
1) The field
of real numbers with respect to the ordinary absolute value is complete,
2) Real numbers
are an extension field of rational numbers
,
3)
is dense in
.
Instead of the ordinary absolute value, we now consider the p-adic absolute value and run through the same process of constructing the completion field.
Corollary 4.28.
is not complete
Corollary 4.29. [14] The completion of
is the field
called the p-adic field.
Remark 4.30. From Corollary 4.29, we have the following statements:
1) the p-adic field with respect to the p-adic absolute value is complete,
2)
is an extension field of
,
3)
is dense in
.
The elements of the above obtained field can be written as,
where
and
for all
.
Example 4.31. Consider the 7-adic field, then we can write two as
in
.
We have constructed the p-adic field,
, this field has a lot of properties and they discussed in [14] . We now give the general definition of all finite field extensions over the rational numbers
.
Definition 4.32. [14] An algebraic number field is a finite extension of
.
If F is an algebraic number field over
, then F has finite degree.
Example 4.33. The field
is a finite extension of itself. Gaussians
and all simple extension of
.
Definition 4.34. We call the complex number,
an algebraic number if it satisfies some monic polynomial with coefficients in
.
4.3. Finite Extension of
In the previous section, we considered
as the base field, and from this consideration, we constructed algebraic number fields. In this section, we take the rational function field over the finite field as our base field.
Definition 4.35. Let
be a prime,
and
. We define the rational function field as
Definition 4.36. [12] A formal Laurent series
is an infinite series of the form
with
,
for all j.
Definition 4.37. [12] Given
and for polynomials
let
such that t divides neither g nor h, then an absolute value
is defined by,
The above absolute is called the t-adic absolute value and it is nonarchimedean like the p-adic absolute value. The t-adic absolute value can also be defined in terms of other parameters than p.
Theorem 4.38. [12] The field
is the completion field of the field
with respect to
.
Proof. Consider the set T of distinct limits of Cauchy sequences in
. We represent each element in T as a unique Cauchy series of the form
with
,
for all i. Thus the completion of
is the field of formal Laurent series denoted by
. □
Remark 4.39. From the two previous sections, we can draw some important conclusions about the fields that we have constructed. Since
is an extension of
which has characteristic zero, the field
is of characteristic zero. Similarly, since
is an extension of the finite field
, it is of characteristic p. The other notable thing is that both fields are constructed with the respect to a nonarchimedean absolute value. The elements are written in form of power series in these fields.
5. Conclusion
In conclusion, we have shown that with the study of field extensions, considering any polynomial with coefficients in the field, we can find the roots of the polynomial. With the notion of algebraically closed fields, we have one field, F, where we can find the roots of any polynomial with coefficients in F. We have also shown that the concept of field extensions can be accounted for by field completion.
Acknowledgements
The authors wish to acknowledge the support of the University of Lusaka and the refereed authors for their helpful work towards this paper. They are also grateful to the anonymous peer-reviewers for their valuable comments and suggestions towards the improvement of the original manuscript.