<<

Tom Leinster, University of Edinburgh Version of 13 May 2021 Note to the reader2 1 Overview of Galois theory4 1.1 The view of C from Q ...... 4 1.2 Every has a ...... 9 1.3 . . . which determines whether it can be solved...... 11 2 Rings and fields 14 2.1 Rings ...... 14 2.2 Fields ...... 19 3 24 3.1 The of polynomials...... 24 3.2 Factorizing polynomials ...... 28 3.3 Irreducible polynomials...... 33 4 extensions 38 4.1 Definition and examples ...... 38 4.2 Algebraic and transcendental elements...... 43 4.3 Simple extensions...... 47 5 Degree 51 5.1 Degrees of extensions and polynomials...... 51 5.2 The tower law...... 56 5.3 Algebraic extensions ...... 58 5.4 Ruler and compass constructions...... 60 6 Splitting fields 67 6.1 Extending ...... 68 6.2 Existence and uniqueness of splitting fields ...... 70 6.3 The ...... 76 7 Preparation for the fundamental theorem 82 7.1 Normality...... 83 7.2 Separability...... 91 7.3 Fixed fields ...... 96 8 The fundamental theorem of Galois theory 100 8.1 Introducing the Galois correspondence ...... 100 8.2 The theorem...... 104 8.3 A specific example ...... 109 9 Solvability by radicals 115 9.1 Radicals...... 116 9.2 Solvable polynomials have solvable groups ...... 119 9.3 An unsolvable polynomial ...... 126 10 Finite fields 130 10.1 ?th roots in ? ...... 131 10.2 Classification of finite fields ...... 133 10.3 Multiplicative structure...... 135 10.4 Galois groups for finite fields...... 136

1 Note to the reader

These are the 2020–21 course notes for Galois Theory (MATH10080).

Structure Each chapter corresponds to one week of the semester. You should read Chapter 1 in Week 1, Chapter 2 in Week 2, and so on. I’m writing the notes as we go along, so the chapters will appear one by one: keep your eye on Learn.

Exercises looking like this are sprinkled through the notes. The idea is that you try them immediately, before you continue reading. Most of them are meant to be quick and easy, much easier than as- signment or workshop questions. If you can do them, you can take it as a sign that you’re following. For those that defeat you, talk with your group or ask on Piazza; or if you’re really stuck, ask me. I promise you that if you make a habit of trying every exercise, you’ll enjoy the course more and understand it better than if you don’t bother.

Digressions like this are optional and not examinable, but might interest you. They’re usually on points that I find interesting, and often describe connections between Galois theory and other parts of mathematics. Here you’ll see titles of relevant References to theorem , page numbers, etc., are clickable links. videos What to prioritize You know by now that the most important things in almost any course are the definitions and the results called Theorem. But I also want to emphasize the proofs. This course presents a wonderful body of theory, and the idea is that you learn it all, including the proofs that are its beating heart. A closed-book exam would test that by asking you to reproduce some proofs. Your exam will be open book, so it can’t ask you to reproduce proofs, but it will test something arguably harder: that you understand them. So the proofs will need your attention and energy.

2 Prerequisites You are required to have taken these two courses:

• Honours : We’ll need some linear algebra, corresponding to Chap- ter 1 of that course. For example, you should be able to convince yourself that an endomorphism of a finite-dimensional is injective if and only if it is surjective. We’ll also need everything from Honours Algebra about rings and polyno- mials (Chapter 3 there), including ideals, quotient rings (factor rings), the universal property of quotient rings, and the first theorem for rings.

: From that course, we’ll need fundamentals such as normal , quotient groups, the universal property of quotient groups, and the first isomorphism theorem for groups. You should know lots about the symmetric groups (=, alternating groups =, cyclic groups = and dihedral groups =, and I hope you can list all of the groups of order < 8 without having to think too hard. Chapter 8 of Group Theory, on solvable groups, will be crucial for us! If you skipped it, you’ll need to go back and fix that. For example, you’ll need to understand the statement that (4 is solvable but 5 is not. We won’t need anything on free groups, the Sylow theorems, or the Jordan– Hölder theorem.

If you’re a visiting student and didn’t take those courses, please get in touch so we can decide whether your background is suitable.

Mistakes I’ll be grateful to hear of any mistakes ([email protected]), even if it’s something very small and even if you’re not sure.

3 Chapter 1

Overview of Galois theory

This chapter stands apart from all the others, Modern treatments of Galois theory take advantage of several well-developed branches of algebra: the theories of groups, rings, fields, and vector spaces. This is as it should be! However, assembling all the algebraic apparatus will take us some time, during which it’s easy to lose sight of what it’s all for. Introduction to Galois theory came from two basic insights: Week 1 • every polynomial has a symmetry group; • this group determines whether the polynomial can be solved by radicals (in a sense I’ll define). In this chapter, I’ll explain these two ideas in as short and low-tech a way as I can manage. In Chapter2 we’ll start again, beginning the modern approach that will take up the rest of the course. But I hope that all through that long build-up, you’ll keep in mind the fundamental ideas that you learn in this chapter.

1.1 The view of C from Q

Imagine you lived several centuries ago, before the discovery of complex numbers. Your whole mathematical world is the real numbers, and there is no square root of −1. This situation frustrates you, and you decide to do something about it. So, you invent a new symbol 8 (for ‘imaginary’) and decree that 82 = −1. You still want to be able to do all the usual operations (+, ×, etc.), and you want to keep all the rules that govern them (associativity, commutativity, etc.). So you’re also forced to introduce new numbers such as 2 + 3 × 8, and you end up with what today we call the complex numbers. So far, so good. But then you notice something strange. When you invented the complex numbers, you only intended to introduce one square root of −1. But

4 accidentally, you introduced a second one at the same time: −8. (You wait centuries for a square root of −1, then two arrive at once.) Maybe that’s not so strange in itself; after all, positive reals have two square roots too. But then you realize something genuinely weird:

There’s nothing you can do to distinguish 8 from −8.

Try as you might, you can’t find any reasonable statement that’s true for 8 but not −8. For example, you notice that 8 is a solution of 17 I3 − I2 − I − = , 3 16 3 I but then you realize that −8 is too. Of course, there are unreasonable statements that are true for 8 but not −8, such as ‘I = 8’. We should restrict to statements that only refer to the known world of real numbers. More precisely, let’s consider statements of the form ? (I) ? (I) 1 = 3 , (1.1) ?2(I) ?4(I) where ?1,?2,?3,?4 are polynomials with real coefficients. Any such equation can be rearranged to give ?(I) = 0, where again ? is a polynomial with real coefficients, so we might as well just consider statements of that form. The point is that if ?(8) = 0 then ?(−8) = 0. Let’s make this formal.

Definition 1.1.1 Two complex numbers I and I0 are indistinguishable when seen from R, or conjugate over R, if for all polynomials ? with coefficients in R,

?(I) = 0 ⇐⇒ ?(I0) = 0.

Warning 1.1.2 The standard term is ‘conjugate over R’. ‘Indistin- guishable’ is a term I invented for this chapter only, to express the idea of not being able to tell the two numbers apart.

For example, 8 and −8 are indistinguishable when seen from R. This follows from a more general result:

Lemma 1.1.3 Let I,I0 ∈ C. Then I and I0 are indistinguishable when seen from R if and only if I0 = I or I0 = I.

5 Proof ‘Only if’: suppose that I and I0 are indistinguishable when seen from R. Write I = G + 8H with G,H ∈ R. Then (I − G)2 = −H2. Since G and H are real, indistinguishability implies that (I0 − G)2 = −H2, so I0 − G = ±8H, so I0 = G ± 8H. ‘If’: obviously I is indistinguishable from itself, so it’s enough to prove that I is indinguishable from I. I’ll give two proofs. Each one teaches us a lesson that will be valuable later. First proof: recall that complex conjugation satisfies

F1 + F2 = F1 + F2, F1 · F2 = F1 · F2 for all F1,F2 ∈ C, and 0 = 0 for all 0 ∈ R. It follows by induction that for any polynomial ? over R, ?(F) = ?(F) for all F ∈ C. So

?(I) = 0 ⇐⇒ ?(I) = 0 ⇐⇒ ?(I) = 0.

Second proof: write I = G + 8H with G,H ∈ R. Let ? be a polynomial over R such that ?(I) = 0. We will prove that ?(I) = 0. This is trivial if H = 0, so suppose that H ≠ 0. Consider the polynomial <(C) = (C − G)2 + H2. Then <(I) = 0. You know from Honours Algebra that

?(C) = <(C)@(C) + A(C) (1.2) for some polynomials @ and A with deg(A) < deg(<) = 2 (so A is either a constant or of degree 1). Putting C = I in (1.2) gives A(I) = 0. But it’s easy to see that this is impossible unless A is the zero polynomial (using the assumption that H ≠ 0). So ?(C) = <(C)@(C). But <(I) = 0, so ?(I) = 0, as required. We have just shown that for all polynomials ? over R, if ?(I) = 0 then ?(I) = 0. Exchanging the roles of I and I proves the converse. Hence I and I are indistinguishable when seen from R. 

Exercise 1.1.4 Both proofs of ‘if’ contain little gaps: ‘It follows by induction’ in the first proof, and ‘it’s easy to see’ in the second. Fill them.

Digression 1.1.5 With complex analysis in mind, we could imagine a stricter definition of indistinguishability in which polynomials are replaced by arbi- trary convergent power series (still with coefficients in R). This would allow functions such as exp, cos and sin, and equations such as exp(8c) = −1.

6 But this apparently different definition of indistinguishability is, in fact, equivalent. A complex is still indistinguishable from exactly itself and its complex conjugate. (For example, exp(−8c) = −1 too.) Do you see why?

Lemma 1.1.3 tells us that indistinguishability over R is rather simple. But the same idea becomes much more interesting if we replace R by Q. And in this course, we will mainly focus on polynomials over Q. Define indistinguishability seen from Q, or officially conjugacy over Q, by replacing R by Q in Definition 1.1.1. From now on, I will usually just say ‘indistinguishable’, dropping the ‘seen from Q’. √ √ Example 1.1.6 I claim that 2 and − 2 are indistinguishable. And I’ll give you two different proofs, closely analogous to the two proofs of the ‘if’ part of Lemma 1.1.3. First proof: write √ √ Q( 2) = {0 + 1 2 : 0, 1 ∈ Q}. √ √ √ For F ∈ Q( 2), there are unique 0, 1 ∈ Q such that F = 0 + 1 2, because 2 is irrational. So it is logically valid to define Example 1.1.6 √ √ F 0 1 . e = − 2 ∈ Q( 2) It is straightforward to check that F F F F , F F F F Ÿ1 + 2 = f1 + f2 Ÿ1 · 2 = f1 · f2 √ F ,F 0 0 0 for all 1 2 ∈ Q( 2), and that e = for all ∈ Q. Just as in the proof√ of F F F Lemma 1.1.3,√ it follows that and e are indistinguishable√ for every ∈ Q( 2). In particular, 2 is indistinguishable from − 2. ? ? C √Second proof: let = ( ) be a polynomial with coefficients in Q such that ?( 2) = 0. You know from Honours Algebra that

?(C) = (C2 − 2)@(C) + A(C) √ @ C A C A < C for√ some polynomials√ ( ) and ( ) over Q with deg 2. Putting = 2 gives A A C 0C 1 0, 1 A ( 2) = 0. But 2 is irrational and ( ) is of the form + with √ ∈ Q, so ? C C2 @ C ? must be the zero polynomial. Hence ( ) = ( − 2) ( ), giving (−√ 2) = 0. ? ? We√ have just shown that for all polynomials √over Q, if√ ( 2) = 0 then ? (− 2) = 0. The same√ argument with the roles of 2 and − 2 reversed proves the converse. Hence ± 2 are indistinguishable.

7 l l2

1

l3 l4

Figure 1.1: The 5th roots of unity.

Exercise 1.1.7 Let I ∈ Q. Show that I is distinguishable from I0 for any I0 ≠ I.

One thing that makes indistinguishability more subtle over Q than over R is that over Q, more than two numbers can be indistinguishable:

Example 1.1.8 The 5th roots of unity are

1, l, l2, l3, l4, where l = 42c8/5 (Figure 1.1). Now 1 is distinguishable from the rest, since it is a root of the polynomial C − 1 and the others are not. (See also Exercise 1.1.7.) But it turns out that l, l2, l3, l4 are all indistinguishable from each other. Complex conjugates are indistinguishable when seen from R, so they’re cer- tainly indistinguishable when seen from Q. Since l4 = 1/l = l, it follows that l and l4 are indistinguishable when seen from Q. By the same argument, l2 and l3 are indistinguishable. What’s not so obvious is that l and l2 are indistin- guishable. I know two proofs, which are like the two proofs of Lemma 1.1.3 and Example 1.1.6. But we’re not equipped to do either yet.

Example 1.1.9 More generally, let ? be any prime and put l = 42c8/?. Then l, l2, . . . , l?−1 are indistinguishable.

So far, we have asked when one complex number can be distinguished from another, using only polynomials over Q. But what about more than one? : I ,...,I I0 ,...,I0 : Definition 1.1.10 Let ≥ 0 and let ( 1 : ) and ( 1 : ) be -tuples of I ,...,I I0 ,...,I0 complex numbers. We say that ( 1 : ) and ( 1 : ) are indistinguishable if for all polynomials ?(C1,...,C: ) over Q in : variables, ? I ,...,I ? I0 ,...,I0 . ( 1 : ) = 0 ⇐⇒ ( 1 : ) = 0

8 When : = 1, this is just the previous definition of indistinguishability.

I ,...,I I0 ,...,I0 Exercise 1.1.11 Suppose that ( 1 : ) and ( 1 : ) are in- I I0 distinguishable. Show that 8 and 8 are indistinguishable, for each 8 ∈ {1,...,:}.

Example 1.1.12 For any I1,...,I: ∈ C, the :-tuples (I1,...,I: ) and (I1,..., I: ) are indistinguishable. For let ?(C1,...,C: ) be a polynomial over Q. Then

?(I1,...,I: ) = ?(I1,..., I: ) since the coefficients of ? are real, by a similar argument to the one in the first proof of Lemma 1.1.3. Hence

?(I1,...,I: ) = 0 ⇐⇒ ?(I1,..., I: ) = 0, which is what we had to prove.

Example 1.1.13 Let l = 42c8/5, as in Example 1.1.8. Then

(l, l2, l3, l4) and (l4, l3, l2, l) are indistinguishable, by Example 1.1.12. It can also be shown that

(l, l2, l3, l4) and (l2, l4, l, l3) are indistinguishable, although the proof is beyond us for now. But

(l, l2, l3, l4) and (l2, l, l3, l4)

? C ,C ,C ,C C C2 are distinguishable, since if we put ( 1 2 3 4) = 2 − 1 then

?(l, l2, l3, l4) = 0 ≠ ?(l2, l, l3, l4).

I I0 So the converse of Exercise 1.1.11 is false: just because 8 and 8 are indistinguish- 8 I ,...,I I0 ,...,I0 able for all , it doesn’t follow that ( 1 : ) and ( 1 : ) are indistinguish- able.

1.2 Every polynomial has a symmetry group. . .

We are now ready to describe the first main idea of Galois theory: every polynomial has a symmetry group.

9 Definition 1.2.1 Let 5 be a polynomial with coefficients in Q. Write U1,...,U: for its distinct roots in C. The Galois group of 5 is

Gal( 5 ) = {f ∈ (: : (U1,...,U: ) and (Uf(1),...,Uf(:)) are indistinguishable}. ‘Distinct roots’ means that we ignore any repetition of roots: e.g. if 5 (C) = 5 9 C (C − 1) then : = 2 and {U1,U2} = {0, 1}.

Exercise 1.2.2 Show that Gal( 5 ) is a of (: . (This one is maybe a bit tricky notationally. Stick to : = 3 if you like.)

Exercise 1.2.2 Digression 1.2.3 I brushed something under the carpet. The definition of Gal( 5 ) depends on the order in which the roots are listed. Different orderings gives different subgroups of (: . However, these subgroups are all conjugate to each other, and therefore isomorphic as abstract groups. So Gal( 5 ) is well-defined as an abstract group, independently of the choice of ordering.

Example 1.2.4 Let 5 be a polynomial over Q all of whose complex roots U1,...,U: are rational. If f ∈ Gal( 5 ) then Uf(8) and U8 are indistinguishable for each 8, by Exercise 1.1.11. But since they are rational, that forces Uf(8) = U8 (by Exercise 1.1.7), and since U1,...,U: are distinct, f(8) = 8. Hence f = id. So the Galois group of 5 is trivial. Example 1.2.5 Let 5 be a quadratic over Q. If 5 has rational roots then as we have just seen, Gal( 5 ) is trivial. If 5 has two non-real roots then they are complex conjugate, so Gal( 5 ) = (2 by Example 1.1.12. The remaining case is where 5 has two distinct roots that are real but not rational, and it can be shown that in that case too, Gal( 5 ) = (2.

Warning 1.2.6 On terminology: note that just now I said ‘non-real’. Sometimes people casually say ‘complex’ to mean ‘not real’. But try not to do this yourself. It makes as little sense as saying ‘real’ to mean ‘irrational’, or ‘rational’ to mean ‘not an ’.

Example 1.2.7 Let 5 (C) = C4 + C3 + C2 + C + 1. Then (C − 1) 5 (C) = C5 − 1, so 5 has roots l, l2, l3, l4 where l = 42c8/5. We saw in Example 1.1.13 that 1 2 3 4 1 2 3 4 1 2 3 4 , ∈ Gal( 5 ), ∉ Gal( 5 ). 4 3 2 1 2 4 1 3 2 1 3 4 In fact, it can be shown that 1 2 3 4 Gal( 5 ) =  . 2 4 1 3  4

10 Example 1.2.8 Let 5 (C) = C3 + 1C2 + 2C + 3 be a cubic over Q with no rational roots. Then √ ( 2 3 3 2 2 3 if −273 + 18123 − 42 − 41 3 + 1 2 ∈ Q, Gal( 5 )  (3 otherwise. This appears as Proposition 22.4 in Stewart, but is way beyond us for now: calcu- lating Galois groups is hard. Galois groups, informally 1.3 . . . which determines whether it can be solved

Here we meet the second main idea of Galois theory: the Galois group of a polynomial determines whether it can be solved. More exactly, it determines whether the polynomial can be ‘solved by radicals’. To explain what this means, let’s begin with the . The roots of a quadratic 0C2 + 1C + 2 are √ −1 ± 12 − 02 4 . 20 After much struggling, it was discovered that there is a similar formula for cubics 0C3 + 1C2 + 2C + 3: the roots are given by

q3 √ q3 √ −27023+9012−213+30 3(270232−180123+4023+4133−1222) + −27023+9012−213−30 3(270232−180123+4023+4133−1222) √ . 3 3 20 (No, you don’t need to memorize that!) This is a complicated formula, and there’s also something strange about it. Any nonzero complex number√ has three cube roots, √ 3 and there are two 3 signs in the formula (ignoring the 2 in the denominator), so it looks as if the formula gives nine roots for the cubic. But a cubic can only have three roots. What’s going on? It turns out that some of the nine aren’t roots of the cubic at all. You have to choose your cube roots carefully. Section 1.4 of Stewart’s book has much more on this point, as well as an explanation of how the cubic formula was obtained. (We won’t be going into this ourselves.) As Stewart also explains, there is a similar but even more complicated formula for quartics (polynomials of degree 4).

Digression 1.3.1 Stewart doesn’t actually write out the explicit formula for the cubic, let alone the much worse one for the quartic. He just describes by which they can be solved. But if you unwind the for the cubic, you get the formula above. I have done this exercise once and do not recommend it.

11 Once mathematicians discovered how to solve quartics, they naturally looked for a formula for quintics (polynomials of degree 5). But it was eventually proved by Abel and Ruffini, in the early 19th century, that there is no formula like the quadratic, cubic or quartic formula for polynomials of degree ≥ 5. A bit more precisely, there is no formula for the roots in terms of the coefficients that uses only the usual arithmetic operations (+, −, ×, ÷) and :th roots (for :). Spectacular as this result was, Galois went further, and so will we. Informally, let us say that a complex number is radical if it can be obtained from the rationals using only the usual arithmetic operations and :th roots. For example, √ √ 1 p3 7 2 2 + 2 − 7 r q 4 5 2 6 + 3 is radical, whichever square root, cube root, etc., we choose. A polynomial over Q is solvable (or soluble) by radicals if all of its complex roots are radical.

Example 1.3.2 Every quadratic√ over Q is solvable by radicals. This follows from the quadratic formula: (−1 ± 12 − 402)/20 is visibly a radical number.

Example 1.3.3 Similarly, the cubic formula shows that every cubic over Q is solvable by radicals. The same goes for quartics.

Example 1.3.4 Some quintics are solvable by radicals. For instance,

(C − 1)(C − 2)(C − 3)(C − 4)(C − 5) is solvable by radicals, since all its roots are rational and, therefore, radical. A bit C 5 less trivially, ( − 123) √+ 456 is solvable by radicals, since its roots are the five complex numbers 123 + 5 −456, which are all radical.

What determines whether a polynomial is solvable by radicals? Galois’s amazing achievement was to answer this question completely:

Theorem 1.3.5 (Galois) Let 5 be a polynomial over Q. Then

5 is solvable by radicals ⇐⇒ Gal( 5 ) is a .

Example 1.3.6 Definition 1.2.1 implies that if 5 has degree = then Gal( 5 ) is isomorphic to a subgroup of (=. You saw in Group Theory that (4 is solvable, and that every subgroup of a solvable group is solvable. Hence the Galois group of any polynomial of degree ≤ 4 is solvable. It follows from Theorem 1.3.5 that every polynomial of degree ≤ 4 is solvable by radicals.

12 5 Example 1.3.7 Put 5 (C) = C −6C +3. Later we’ll show that Gal( 5 ) = (5. You saw in Group Theory that (5 is not solvable (or at least, you saw that 5 isn’t solvable, which implies that (5 isn’t either, as (5 contains 5 as a subgroup). Hence 5 is not solvable by radicals.

If there was a quintic formula then all quintics would be solvable by radicals, for the same reason as in Examples 1.3.2 and 1.3.3. But since this is not the case, there is no quintic formula. Galois’s result is much sharper than Abel and Ruffini’s. They proved that there is no formula providing a solution by radicals of every quintic, whereas Galois found a way of determining which quintics (and higher) can be solved by radicals and which cannot.

Digression 1.3.8 From the point of view of modern numerical computation, this is all a bit odd. Computationally speaking, there is probably not much 5 difference√ between solving C + 3 = 0 to 100 decimal places (that is, finding 5 −3) and solving C5 − 6C + 3 = 0 to 100 decimal places (that is, solving a polynomial that isn’t solvable by radicals). Numerical computation and have different ideas about what is easy and what is hard!

∗ ∗ ∗

This completes our overview of Galois theory. What’s next? Mathematics increasingly emphasizes abstraction over calculation. Individual mathematicians’ tastes vary, but the historical trend is clear. In the case of Galois theory, this means dealing with abstract algebraic structures, principally fields, instead of manipulating explicit algebraic expressions such as polynomials. The cubic formula already gave you a taste of how hairy that can get. Developing Galois theory using abstract algebraic structures helps us to see its connections to other parts of mathematics, and also has some fringe benefits. For example, we’ll solve some notorious geometry problems that perplexed the ancient Greeks and remained unsolved for millennia. For that and many other things, we’ll need some ring theory and some field theory—and that’s what’s next.

13 Chapter 2

Rings and fields

We now start again. This chapter is a mixture of revision and material that is likely to be new to you. The revision is from Honours Algebra and Introduction to (if you took it, which I won’t assume). Introduction to Week 2 2.1 Rings

We’ll begin with some stuff you know—but with a twist. In this course, the word ring means with 1 (multiplicative identity). Noncommutative rings and rings without 1 are important in some parts of mathematics, but since we’ll be focusing on commutative rings with 1, it will be easier to just call them ‘rings’. Given rings ' and (, a from ' to ( is a i : ' → ( satisfying the equations i(A + A0) = i(A) + i(A0), i(0) = 0, i(−A) = −i(A), i(AA0) = i(A)i(A0), i(1) = 1 (note this!) for all A,A0 ∈ '. For example, complex conjugation is a homomorphism C → C. It is a very useful lemma that if i(A + A0) = i(A) + i(A0), i(AA0) = i(A)i(A0), i(1) = 1 for all A,A0 ∈ ' then i is a homomorphism. In other words, to show that i is a homomorphism, you only need to check it preserves +, · and 1; preservation of 0 and negatives then comes for free. But you do need to check it preserves 1. That doesn’t follow from the other conditions. A of a ring ' is a subset ( ⊆ ' that contains 0 and 1 and is closed under addition, multiplication and negatives. Whenever ( is a subring of ', the inclusion ]: ( → ' (defined by ](B) = B) is a homomorphism.

14 Warning 2.1.1 In Honours Algebra, rings had 1s but homomorphisms were not required to preserve 1. Similarly, of ' had to have a 1, but it was not required to be the same as the 1 of '. For example, take the ring C of complex numbers, the noncommutative ring " of 2 × 2 matrices over C, and the function i : C → " defined by I 0 i(I) = . 0 0 In the terminology of Honours Algebra, i is a homomorphism and its image im i is a subring of ". But in our terminology, i is not a homomorphism (as i(1) ≠ ) and im i is not a subring of " (as  ∉ im i).

Exercise 2.1.2 Let ' be a ring and let S be any (perhaps infinite) ' Ñ ( of subrings of . Prove that their intersection (∈S is also a subring of '. (In contrast, in the Honours Algebra setup, even the intersection of two subrings need not be a subring.)

Example 2.1.3 For any ring ', there is exactly one homomorphism Z → '. Here is a sketch of the proof. To show there is at least one homomorphism j : Z → ', we will construct one. Define j on nonnegative integers = inductively by j(0) = 0 and j(=+1) = j(=)+1. (Thus, j(=) = 1' + · · · + 1'.) Define j on negative integers = by j(=) = −j(−=). A series of tedious checks shows that j is indeed a ring homomorphism. To show there is only one homomorphism Z → ', let i be any homomorphism Z → '; we have to prove that i = j. Certainly i(0) = 0 = j(0). Next prove by induction on = that i(=) = j(=) for nonnegative integers =. I leave the details to you, but the crucial point is that because homomorphisms preserve 1, we must have i(= + 1) = i(=) + i(1) = i(=) + 1 for all = ≥ 0. Once we have shown that i and j agree on the nonnegative integers, it follows that for negative =,

i(=) = −i(−=) = −j(−=) = j(=).

Hence i(=) = j(=) for all = ∈ Z; that is, i = j. j(=) = · = The meaning of Usually we write as 1', or simply as if it is clear from the context ‘= · 1’, and that = is to be interpreted as an element of '. Exercise 2.2.7 15 Every ring homomorphism i : ' → ( has an image im i, which is a subring of (, and a kernel ker i, which is an ideal of '.

Warning 2.1.4 Subrings in ring theory are analogous to subgroups in group theory, and ideals in ring theory are analogous to normal subgroups in group theory. But whereas normal subgroups are a special kind of subgroup, ideals are not a special kind of subring! Subrings must contain 1, but most ideals don’t.

Exercise 2.1.5 Prove that the only subring of a ring ' that is also an ideal is ' itself.

Given a ring ' and an ideal  ', we obtain the quotient ring or factor ring '/ and the canonical homomorphismP c : ' → '/, which is surjective and has kernel . Quotient rings As explained in Honours Algebra, the quotient ring together with the canonical homomorphism has a ‘universal property’: given any ring ( and any homomor- phism i : ' → ( satisfying ker i ⊇ , there is exactly one homomorphism i¯ : '/ → ( such that this diagram commutes:

c ' / '/

i¯ i (.

(For a diagram to commute means that whenever there are two different paths from one object to another, the composites along the two paths are equal. Here, it means that i = i¯ ◦ c.) The first isomorphism theorem says that if i is surjective and has kernel equal to  then i¯ is an isomorphism. So c : ' → '/ is essentially the only surjective homomorphism out of ' with kernel .

Digression 2.1.6 Loosely, the ideals of a ring ' correspond one-to-one with the surjective homomorphisms out of '. This means four things:

• given an ideal  ', we get a surjective homomorphism out of ' P (namely, c : ' → '/); • given a surjective homomorphism i out of ', we get an ideal of ' (namely, ker i); • if we start with an ideal  of ', take its associated surjective homomor- phism c : ' → '/, then take its associated ideal, we end up where we started (that is, ker(c ) = );

16 • if we start with a surjective homomorphism i : ' → (, take its asso- ciated ideal ker i, then take its associated surjective homomorphism cker i : ' → '/ker i, we end up where we started (at least ‘up to iso- morphism’, in that we have the isomorphism i¯ : '/ker i → ( making the triangle commute). This is the first isomorphism theorem.

Analogous stories can be told for groups and for modules.

0 An integral domain is a ring ' such that 0' ≠ 1' and for A,A ∈ ',

AA0 = 0 ⇒ A = 0 or A0 = 0.

Exercise 2.1.7 The trivial ring or zero ring is the one-element set with its only possible ring structure. Show that the only ring in which 0 = 1 is the trivial ring.

Equivalently, an integral domain is a nontrivial ring in which cancellation is valid: AB = A0B implies A = A0 or B = 0.

Digression 2.1.8 Why is the condition 0 ≠ 1 in the definition of integral domain? My answer begins with a useful general point: the sum of no things should always be interpreted as 0. (The amount you pay in a shop is the sum of the prices of the individual things. If you buy no things, you pay £0.) Similarly, the product of no things should be interpreted as 1. Now consider the following condition on a ring ': for all = ≥ 0 and A1,...,A= ∈ ',

A1A2 ··· A= = 0 ⇒ there exists 8 ∈ {1, . . . , =} such that A8 = 0. (2.1)

In the case = = 0, this says: if 1 = 0 then there exists 8 ∈ œ such that A8 = 0. But any statement beginning ‘there exists 8 ∈ œ’ is false! So in the case = = 0, condition (2.1) states that 1 ≠ 0. And for = = 2, it’s the main condition in the definition of integral domain. So ‘1 ≠ 0’ is the 0-fold analogue of the main condition. On the other hand, if (2.1) holds for = = 0 and = = 2 then a simple induction shows that it holds for all = ≥ 0. Conclusion: an integral domain can equivalently be defined as a ring in which (2.1) holds for all = ≥ 0.

Let ) be a subset of a ring '. The ideal h)i generated by ) is the intersection of all the ideals of ' containing ). You can show that any intersection of ideals is an ideal (much as you did for subrings in Exercise 2.1.2), so h)i is an ideal. It is

17 the smallest ideal of ' containing ). That is, h)i is an ideal containing ), and if  is another ideal containing ) then h)i ⊆ . When ) is a finite set {A1,...,A=}, we write h)i as hA1,...,A=i, and it satisfies

hA1,...,A=i = {01A1 + · · · + 0=A= : 01, . . . , 0= ∈ '}. (2.2) In particular, when = = 1, hAi = {0A : 0 ∈ '}. Ideals of the form hAi are called principal ideals.A principal ideal domain is an integral domain in which every ideal is principal. Example 2.1.9 Z is a principal ideal domain. Indeed, if  Z then either  = {0}, in which case  = h0i, or  contains some positive integer,P in which case we can define = to be the least positive integer in  and use the division algorithm to show that  = h=i.

Exercise 2.1.10 Fill in the details of Example 2.1.9.

Let A and B be elements of a ring '. We say that A divides B, and write A | B, if there exists 0 ∈ ' such that B = 0A. This condition is equivalent to B ∈ hAi, and to hBi ⊆ hAi. An element D ∈ ' is a unit if it has a multiplicative inverse, or equivalently if hDi = '. The units form a group '× under multiplication. For instance, Z× = {1, −1}.

Exercise 2.1.11 Let A and B be elements of an integral domain. Show that A | B | A ⇐⇒ hAi = hBi ⇐⇒ B = DA for some unit D.

Elements A and B of a ring are coprime if for 0 ∈ ', 0 | A and 0 | B ⇒ 0 is a unit. Proposition 2.1.12 Let ' be a principal ideal domain and A,B ∈ '. Then A and B are coprime ⇐⇒ 0A + 1B = 1 for some 0, 1 ∈ '. Proof ⇒: suppose that A and B are coprime. Since ' is a principal ideal domain, hA,Bi = hDi for some D ∈ '. Since A ∈ hA,Bi = hDi, we must have D | A, and similarly D | B. But A and B are coprime, so D is a unit. Hence 1 ∈ hDi = hA,Bi. But by equation (2.2), hA,Bi = {0A + 1B : 0, 1 ∈ '}, and the result follows. ⇐: suppose that 0A + 1B = 1 for some 0, 1 ∈ '. If D ∈ ' with D | A and D | B then D | (0A + 1B) = 1, so D is a unit. Hence A and B are coprime. 

18 2.2 Fields

A field is a ring in which 0 ≠ 1 and every nonzero element is a unit. Equivalently, it is a ring such that × = \{0}. Every field is an integral domain.

Exercise 2.2.1 Write down all the examples of fields that you know.

A field has exactly two ideals: {0} and . For if {0} ≠  then D ∈  for some D ≠ 0; but then D is a unit, so hDi = , so  = . P

Lemma 2.2.2 Every homomorphism between fields is injective.

Proof Let i : → ! be a homomorphism between fields. Then ker i , so ker i is either {0} or . If ker i = then i(1) = 0; but i(1) = 1 by definitionP of homomorphism, so 0 = 1 in !, contradicting the assumption that ! is a field. Hence ker i = {0}, that is, i is injective. 

Warning 2.2.3 With the Honours Algebra definition of homomor- phism, Lemma 2.2.2 would be false, since the map with constant value 0 would be a homomorphism.

Let ' be any ring. By Example 2.1.3, there is a unique homomorphism j : Z → '. Its kernel is an ideal of the principal ideal domain Z. Hence ker j = h=i for a unique integer = ≥ 0. This = is called the characteristic of ', and written as char '. Explicitly, ( the least = > 0 such that = · 1 = 0 , if such an = exists; char ' = ' ' (2.3) 0, otherwise.

Another way to say it: for < ∈ Z, we have < · 1' = 0 if and only if < is a multiple of char '. The concept of characteristic is mostly used in the case of fields.

Examples 2.2.4 i. Q, R and C all have characteristic 0.

ii. For a prime number ?, we write F? for the field Z/h?i of integers modulo ?. Then char F? = ?.

Lemma 2.2.5 The characteristic of an integral domain is 0 or a prime number.

19 Proof Let ' be an integral domain and write = = char '. Suppose that = > 0; we must prove that = is prime. Since 1 ≠ 0 in an integral domain, = ≠ 1. (Remember that 1 is not a prime! So that step was necessary.) Now let :, < > 0 with :< = =. Writing j for the unique homomorphism Z → ', we have

j(:) j(<) = j(:<) = j(=) = 0,

and ' is an integral domain, so j(:) = 0 or j(<) = 0. WLOG, j(:) = 0. But ker j = h=i, so = | :, so : = =. Hence = is prime. 

Examples 2.2.4 show that there exist fields of every possible characteristic. But there is no way of mapping between fields of different characteristics: Lemma 2.2.6 Let i : → ! be a homomorphism of fields. Then char = char !.

Proof Write j and j! for the unique homomorphisms from Z to and !, respectively. Since j! is the unique homomorphism Z → !, the triangle Z j j!

 ! i /

commutes. (Concretely, this says that i(= · 1 ) = = · 1! for all = ∈ Z.) Hence ker(i ◦ j ) = ker j!. But i is injective by Lemma 2.2.2, so ker(i ◦ j ) = ker j . Hence ker j = ker j!, or equivalently, char = char !.  For example, the inclusion Q → R is a homomorphism of fields, and both have characteristic 0.

Exercise 2.2.7 This proof of Lemma 2.2.6 is quite abstract. Find a more concrete proof, taking equation (2.3) as your definition of characteristic. (You will still need the fact that i is injective.) The meaning of ‘= · 1’, and A subfield of a field is a subring that is a field. The prime subfield of Exercise 2.2.7 is the intersection of all the subfields of . It is straightforward to show that any intersection of subfields is a subfield (just as you showed in Exercise 2.1.2 that any intersection of subrings is a subring). Hence the prime subfield is a subfield. It is the smallest subfield of , in the sense that any other subfield of contains it. Concretely, the prime subfield of is   < · 1 : <, = ∈ Z with = · 1 ≠ 0 . = · 1

20 To see this, first note that this set is a subfield of . It is the smallest subfield of : for if ! is a subfield of then 1 ∈ ! by definition of subfield, so < · 1 ∈ ! for all integers <, so (< · 1 )/(= · 1 ) ∈ ! for all integers < and = such that = · 1 ≠ 0. Examples 2.2.8 i. The field Q has no proper subfields, so the prime subfield of Q is Q itself.

ii. Let ? be a prime. The field F? has no proper subfields, so the prime subfield of F? is F? itself.

Exercise 2.2.9 What is the prime subfield of R? Of C?

The prime subfields appearing in Examples 2.2.8 were Q and F?. In fact, these are the only prime subfields of anything:

Lemma 2.2.10 Let be a field.

i. If char = 0 then the prime subfield of is Q.

ii. If char = ? > 0 then the prime subfield of is F?.

In the statement of this lemma, as so often in mathematics, the word ‘is’ means ‘is isomorphic to’. I hope you’re comfortable with that by now.

Proof For (i), suppose that char = 0. By definition of characteristic, = · 1 ≠ 0 for all integers = ≥ 0. One can check that there is a well-defined homomorphism i : Q → defined by 0. By Lemma 2.2.5, ? is prime. The unique homomorphism j : Z → has kernel h?i, by definition. By the first isomorphism theorem, im j  Z/h?i = F?. But im j is a subring of , and in fact a subfield since is it isomorphic to F?. It is the prime subfield, since F? has no proper subfields. 

Lemma 2.2.11 Every finite field has positive characteristic.

Proof By Lemma 2.2.10, a field of characteristic 0 contains a copy of Q and is therefore infinite. 

21 Warning 2.2.12 There are also infinite fields of positive characteris- tic! We haven’t met one yet, but we will soon.

So far, we have rather few examples of fields. The following construction will allow us to manufacture many, many more. An element A of a ring ' is irreducible if A is not 0 or a unit, and if for 0, 1 ∈ ', Building blocks A = 01 ⇒ 0 or 1 is a unit.

For example, the irreducibles in Z are ±2, ±3, ±5,.... An element of a ring is reducible if it is not 0, a unit, or irreducible. So 0 and units count as neither reducible nor irreducible.

Exercise 2.2.13 What are the irreducible elements of a field?

Proposition 2.2.14 Let ' be a principal ideal domain and 0 ≠ A ∈ '. Then

A is irreducible ⇐⇒ '/hAi is a field.

Proof Write c for the canonical homomorphism ' → '/hAi. ⇒: suppose that A is irreducible. To show that 1'/hAi ≠ 0'/hAi, note that since A is not a unit, 1' ∉ hAi = ker c, so

1'/hAi = c(1') ≠ 0'/hAi. Next we have to show that every nonzero element of '/hAi is a unit, or equivalently that c(B) is a unit whenever B ∈ ' with B ∉ hAi. We have A - B, and A is irreducible, so A and B are coprime. Hence by Proposition 2.1.12 (and the assumption that ' is a principal ideal domain), we can choose 0, 1 ∈ ' such that

0A + 1B = 1'. Applying c to each side gives

c(0)c(A) + c(1)c(B) = 1'/hAi. But c(A) = 0, so c(1)c(B) = 1, so c(B) is a unit. ⇐: suppose that '/hAi is a field. Then 1'/hAi ≠ 0'/hAi, that is, c(1') ∉ ker c = hAi, that is, A - 1'. Hence A is not a unit. Next we have to show that if 0, 1 ∈ ' with A = 01 then 0 or 1 is a unit. We have 0 = c(A) = c(0)c(1)

22 and '/hAi is an integral domain, so WLOG c(0) = 0. Then 0 ∈ ker c = hAi, so 0 = A10 for some 10 ∈ '. This gives

A = 01 = A101.

But A ≠ 0 by hypothesis, and ' is an integral domain, so 101 = 1. Hence 1 is a unit. 

Example 2.2.15 When = is an integer, Z/h=i is a field if and only if = is irreducible (that is, ± a prime number).

Proposition 2.2.14 enables us to construct fields from irreducible ele- ments. . . but irreducible elements of a principal ideal domain. Right now that’s not much help, because we don’t have many examples of principal ideal domains. But we will soon.

23 Chapter 3

Polynomials

This chapter revisits and develops some themes you met in Honours Algebra. Although it’s long, it contains material you’ve seen before. Before you begin, it may help you to reread Section 3.3 (Polynomials) of the Honours Algebra notes. Introduction to Week 3 3.1 The ring of polynomials

You already know the definition of polynomial, but I want to make a point by phrasing it in an unfamiliar way. Definition 3.1.1 Let ' be a ring. A polynomial over ' is an infinite sequence (00, 01, 02,...) of elements of ' such that {8 : 08 ≠ 0} is finite. The set of polynomials over ' forms a ring as follows:

(00, 01,...) + (10, 11,...) = (00 + 10, 01 + 11,...), (3.1) (00, 01,...)·(10, 11,...) = (20, 21,...) (3.2) Õ where 2: = 081 9 , (3.3) 8, 9 : 8+ 9=: the zero of the ring is (0, 0,...), and the multiplicative identity is (1, 0, 0,...). 2 Of course, we almost always write (00, 01, 02,...) as 00 + 01C + 02C + · · · , or the same with some other symbol in place of C. In that notation, formulas (3.1) and (3.2) look like the usual formulas for addition and multiplication of polyno- mials. Nevertheless:

Warning 3.1.2 A polynomial is not a function! A polynomial gives rise to a function, as we’ll recall in a moment. But a polynomial itself is a purely formal object.

24 The set of polynomials over ' is written as '[C] (or '[D], '[G], etc.). Since ( = '[C] is itself a ring, we can consider the ring ([D] = ('[C])[D], usually written as '[C,D], and similarly '[C,D,E] = ('[C,D])[E], etc. Why study We use 5 , 6, ℎ, . . . and 5 (C), 6(C), ℎ(C),... interchangeably to denote elements polynomials? of '[C]. A polynomial 5 = (00, 01,...) over ' gives rise to a function ' → ' 2 A 7→ 00 + 01A + 02A + · · · .

(The sum on the right-hand side makes sense because only finitely many 08s are nonzero.) This function is usually called 5 too. But calling it that is slightly dangerous, because:

Warning 3.1.3 Different polynomials can give rise to the same func- 2 tion. For example, consider C,C ∈ F2 [C]. They are different polyno- mials: going back to Definition 3.1.1, they’re alternative notation for the sequences

(0, 1, 0, 0,...) and (0, 0, 1, 0,...),

which are plainly not the same. On the other hand, they induce the 2 same function F2 → F2, because 0 = 0 for all (both) 0 ∈ F2.

Exercise 3.1.4 Show that whenever ' is a finite nontrivial ring, it is possible to find distinct polynomials over ' that induce the same func- tion ' → '. (Hint: are there finitely or infinitely many polynomials over '? Functions ' → '?)

The ring of polynomials has a universal property: a homomorphism from '[C] to some other ring  is determined by its effect on constant polynomials and on C itself, in the following sense. The universal property of '[C] Lemma 3.1.5 (Universal property of the ) Let ' and  be rings. For every homomorphism i : ' →  and every 1 ∈ , there is exactly one homomorphism \ : '[C] →  such that

\(0) = i(0) for all 0 ∈ ', (3.4) \(C) = 1. (3.5)

On the left-hand side of (3.4), the ‘0’ means the polynomial 0 + 0C + 0C2 + · · · .

25 Proof To show there is at most one such \, take any homomorphism \ : '[C] →  Í 0 C8 ' satisfying (3.4) and (3.5). Then for every polynomial 8 8 over ,

Õ 8 Õ 8 \ 08C = \(08)\(C) since \ is a homomorphism 8 8 Õ 8 = i(08)1 by (3.4) and (3.5). 8 So \ is uniquely determined. To show there is at least one such \, define a function \ : '[C] →  by

Õ 8 Õ 8 \ 08C = i(08)1 8 8 Í 0 C8 ' C \ ( 8 8 ∈ [ ]). Then clearly satisfies conditions (3.4) and (3.5). It remains to check that \ is a homomorphism. I will do the worst part of this, which is to check that \ preserves multiplication, and leave the rest to you. 5 C Í 0 C8 6 C Í 1 C 9 5 C 6 C So, take polynomials ( ) = 8 8 and ( ) = 9 9 . Then ( ) ( ) = Í 2 C: 2 : : , where : is as defined in equation (3.3). We have

Õ :  \( 5 6) = \ 2: C : Õ : = i(2: )1 by definition of \ : Õ  Õ  : = i 081 9 1 by definition of 2: : 8, 9 : 8+ 9=: Õ Õ : = i(08)i(1 9 )1 since i is a homomorphism : 8, 9 : 8+ 9=: Õ 8+ 9 = i(08)i(1 9 )1 8, 9 Õ 8 Õ 9  = i(08)1 i(1 9 )1 8 9 = \( 5 )\(6) by definition of \. 

Here are three uses for the universal property of the ring of polynomials. First: Definition 3.1.6 Let i : ' → ( be a ring homomorphism. The induced homo- morphism i∗ : '[C] → ([C] is the unique homomorphism '[C] → ([C] such that i∗(0) = i(0) for all 0 ∈ ' and i∗(C) = C.

26 The universal property guarantees that there is one and only one homomor- phism i∗ with these properties. Concretely,

Õ 8 Õ 8 i∗ 08C = i(08)C 8 8 Í 0 C8 ' C for all 8 8 ∈ [ ]. Second, let ' be a ring and A ∈ '. By the universal property, there is a unique homomorphism evA : '[C] → ' such that evA (0) = 0 for all 0 ∈ ' and evA (C) = A. Concretely, Õ 8 Õ 8 evA 08C = 08A 8 8 Í 0 C8 ' C A for all 8 8 ∈ [ ]. This map evA is called evaluation at . Í 8 (The notation 08C for what is officially (00, 01,...) makes it look obvious that we can evaluate a polynomial at an element and that this gives a homomorphism: of course ( 5 · 6)(A) = 5 (A)6(A), for instance! But that’s only because of the notation: there was actually something to prove here.) Third, let ' be a ring and 2 ∈ '. For any 5 (C) ∈ '[C], we can ‘substitute C = D + 2’ to get a polynomial in D. What exactly does this mean? Formally, there is a unique homomorphism \ : '[C] → '[D] such that \(0) = 0 for all 0 ∈ ' and \(C) = D + 2. Concretely,

Õ 8 Õ 8 \ 08C = 08 (D + 2) . 8 8 This particular substitution is invertible. Informally, the inverse is ‘substitute D = C − 2’. Formally, there is a unique homomorphism \0 : '[D] → '[C] such that \0(0) = 0 for all 0 ∈ ' and \0(D) = C − 2. These maps \ and \0 carrying out the substitutions are inverse to each other, as you can deduce this from either the universal property or the concrete descriptions. So, the substitution maps

\ / '[C] o '[D] (3.6) \0 define an isomorphism between '[C] and '[D]. For example, since isomorphism preserve irreducibility (and everything else that matters!), 5 (C) is irreducible if and only if 5 (C − 2) is irreducible.

Exercise 3.1.7 What happens to everything in the previous paragraph if we substitute C = D2 + 2 instead?

The rest of this section is about degree.

27 Í 8 Definition 3.1.8 The degree deg( 5 ) of a nonzero polynomial 5 (C) = 08C is the largest = ≥ 0 such that 0= ≠ 0. By convention, deg(0) = −∞, where −∞ is a formal symbol which we give the properties −∞ < =, (−∞) + = = −∞, (−∞) + (−∞) = −∞ for all integers =.

Digression 3.1.9 Defining deg(0) like this is helpful because it allows us to make statements about all polynomials without having to make annoying exceptions for the zero polynomial (e.g. Lemma 3.1.10(i)). But putting deg(0) = −∞ also makes intuitive sense. At least for polynomials over R, the degree of a nonzero polynomial tells us how fast it grows: when C is large, 5 (C) behaves roughly like Cdeg( 5 ) . What about the zero polynomial? Well, whether or not C is large, 0(C) = 0, and C−∞ can sensibly be interpreted A as limA→−∞ C = 0. So it makes sense to put deg(0) = −∞.

Lemma 3.1.10 Let ' be an integral domain. Then: i. deg( 5 6) = deg( 5 ) + deg(6) for all 5 , 6 ∈ '[C]; ii. '[C] is an integral domain; iii. when ' is a field, the units in '[C] are the polynomials of degree 0 (that is, the nonzero constants); iv. when ' is a field, 5 (C) ∈ '[C] is irreducible if and only if deg( 5 ) > 0 and 5 cannot be expressed as a product of two polynomials of degree > 0. Proof At least parts (i)–(iii) were in Honours Algebra (Section 3.3). Part (iv) follows from the general definition of irreducible element of a ring. 

3.2 Factorizing polynomials

Every nonzero integer can be expressed as a product of primes in an essentially unique way. But the analogous statement is not true in all rings, or even all integral domains. Some rings have elements that can’t be expressed as a product of irreducibles at all. In other rings, into irreducibles exist but are not unique. (By ‘not unique’ I mean more than just changing the order of the factors or multiplying them by units.) The big theorem of this section is that, happily, every polynomial over a field can be factorized, essentially uniquely, into irreducibles. We begin with a result on division of polynomials from Section 3.3 of Honours Algebra.

28 Proposition 3.2.1 Let be a field and 5 , 6 ∈ [C]. Then there is exactly one pair of polynomials @,A ∈ [C] such that 5 = @6 + A and deg(A) < deg(6). 

We use this to prove an extremely useful fact: Proposition 3.2.2 Let be a field. Then [C] is a principal ideal domain.

Proof First, [C] is an integral domain, by Lemma 3.1.10(ii). Now let  [C]. If  = {0} then  = h0i. Otherwise, put 3 = min{deg( 5 ) : 0 ≠ 5 ∈ } andP choose 6 ∈  such that deg(6) = 3. I claim that  = h6i. To prove this, let 5 ∈ ; we must show that 6 | 5 . By Proposition 3.2.1, 5 = @6 + A for some @,A ∈ [C] with deg(A) < 3. Now A = 5 − @6 ∈  since 5 , 6 ∈ , so the minimality of 3 implies that A = 0. Hence 5 = @6, as required. 

If you struggled with Exercise 2.1.10, that proof should give you a clue.

Warning 3.2.3 Lemma 3.1.10(ii) implies that [C1,...,C=] is an integral domain for all = ≥ 2, but it is not a principal ideal domain. For example, the ideal

hC1,C2i = { 5 (C1,C2) ∈ Q[C1,C2] : 5 has constant term 0}

is not principal. Also, Proposition 3.2.2 really needed the hypothesis that is a field; it’s not enough for it to be a principal ideal domain. For example, Z is a principal ideal domain, but in Z[C], the ideal

h2,Ci = { 5 (C) ∈ Z[C] : the constant term of 5 is even}

is not principal.

Exercise 3.2.4 Prove that the ideals in Warning 3.2.3 are indeed not principal.

Exercise 3.2.4: a At the end of Chapter2, I promised I’d give you a way of manufacturing lots non-principal ideal of new fields. Here it is! Corollary 3.2.5 Let be a field and let 0 ≠ 5 ∈ [C]. Then

5 is irreducible ⇐⇒ [C]/h 5 i is a field.

Proof This follows from Propositions 2.2.14 and 3.2.2. 

29 To manufacture new fields using Corollary 3.2.5, we’ll need a way of knowing which polynomials are irreducible. That’s the topic of Section 3.3, but for now let’s stick to our mission: proving that every polynomial factorizes into irreducibles in an essentially unique way. To achieve this mission, we’ll need two more lemmas.

Lemma 3.2.6 Let be a field and let 5 (C) ∈ [C] be a nonconstant polynomial. Then 5 (C) is divisible by some irreducible in [C].

The word nonconstant just means ‘of degree > 0’.

Proof Let 6 be a nonconstant polynomial of smallest possible degree such that 6 | 5 . (For this to make sense, there must be at least one nonconstant polynomial dividing 5 , and there is: 5 .) I claim that 6 is irreducible. Proof: if 6 = 6162 then each 68 divides 5 , so by the minimality of deg(6), each 68 has degree 0 or deg(6). They cannot both have degree deg(6), since deg(61) + deg(62) = deg(6) > 0. So at least one has degree 0, i.e., is a unit. 

Lemma 3.2.7 Let be a field and 5 , 6, ℎ ∈ [C]. Suppose that 5 is irreducible and 5 | 6ℎ. Then 5 | 6 or 5 | ℎ.

This behaviour is familiar in the integers: if a prime ? divides some product 01, then ? | 0 or ? | 1.

Proof Suppose that 5 - 6. Since 5 is irreducible, 5 and 6 are coprime. Since [C] is a principal ideal domain, Proposition 2.1.12 implies that there are ?,@ ∈ [C] such that ? 5 + @6 = 1. Multiplying both sides by ℎ gives

? 5 ℎ + @6ℎ = ℎ.

But 5 | ? 5 ℎ and 5 | 6ℎ, so 5 | ℎ. 

Theorem 3.2.8 Let be a field and 0 ≠ 5 ∈ [C]. Then

5 = 0 51 52 ··· 5=

for some = ≥ 0, 0 ∈ and monic irreducibles 51, . . . , 5= ∈ [C]. Moreover, = and 0 are uniquely determined by 5 , and 51, . . . , 5= are uniquely determined up to reordering.

30 In the case = = 0, the product 51 ··· 5= should be interpreted as 1 (as in Digression 2.1.8). Monic means that the leading coefficient is 1. Proof First we prove that such a exists, by induction on deg( 5 ). If deg( 5 ) = 0 then 5 is a constant 0 and we take = = 0. Now suppose that deg( 5 ) > 0 and assume the result for polynomials of smaller degree. By Lemma 3.2.6, there is an irreducible 6 dividing 5 , and we can assume that 6 is monic by dividing by a constant if necessary. Then 5 /6 is a nonzero polynomial of smaller degree than 5 , so by inductive hypothesis,

5 /6 = 0ℎ1 ··· ℎ< for some 0 ∈ and monic irreducibles ℎ1, . . . , ℎ<. Rearranging gives

5 = 0ℎ1 ··· ℎ<6, completing the induction. Now we prove uniqueness, again by induction on deg( 5 ). If deg( 5 ) = 0 then 5 is a constant 0 and the only possible factorization is the one with = = 0. Now suppose that deg( 5 ) > 0, and take two factorizations

0 51 ··· 5= = 5 = 161 ··· 6< (3.7) where 0, 1 ∈ and 58, 6 9 are monic irreducible. Since deg( 5 ) > 0, we have =, < ≥ 1. Now 5= | 161 ··· 6<, so by Lemma 3.2.7, 5= | 6 9 for some 9. By rearranging, we can assume that 9 = <. But 6< is also irreducible, so 5= = 26< for some nonzero 2 ∈ , and both 5= and 6< are monic, so 2 = 1. Hence 5= = 6<. Cancelling in (3.7) (which we can do as [C] is an integral domain) gives

0 51 ··· 5=−1 = 161 ··· 6<−1.

By inductive hypothesis, = = <, 0 = 1, and the lists of irreducibles 51, . . . , 5=−1 and 61, . . . , 6<−1 are the same up to reordering. This completes the induction.  One way to find an irreducible factor of a polynomial 5 (C) ∈ [C] is to find a root (an element 0 ∈ such that 5 (0) = 0): Lemma 3.2.9 Let be a field, 5 (C) ∈ [C] and 0 ∈ . Then 5 (0) = 0 ⇐⇒ (C − 0) | 5 (C). Proof ⇒: suppose that 5 (0) = 0. By Proposition 3.2.1, 5 (C) = (C − 0)@(C) + A(C) (3.8) for some @,A ∈ [C] with deg(A) < 1. Then A is a constant, so putting C = 0 in (3.8) gives A = 0. ⇐: if 5 (C) = (C − 0)@(C) for some polynomial @ then 5 (0) = 0. 

31 Definition 3.2.10 Let be a field, let 0 ≠ 5 ∈ [C], and let 0 ∈ be a root of 5 . The multiplicity of 0 is the unique integer < ≥ 1 such that (C − 0)< | 5 (C) but (C − 0)<+1 - 5 (C).

Exercise 3.2.11 This definition assumes that there is a unique < with these properties (that is, there is one and only one such <). Prove it.

Example 3.2.12 Over any field , the polynomial C has a root at 0 with multiplicity 1, and the polynomial C2 has a root at 0 with multiplicity 2. This is true even when 2 = F2, in which case C and C induce the same function → (Warning 3.1.2).

Proposition 3.2.13 Let be a field and let 0 ≠ 5 ∈ [C]. Write 01, . . . , 0: for the distinct roots of 5 in , and <1, . . . , <: for their multiplicities. Then

<1 <: 5 (C) = (C − 01) ···(C − 0: ) 6(C) for some 6(C) ∈ [C] that has no roots. Proof By induction on :. If : = 0, this is immediate (again interpreting an empty < product as 1). Now suppose that : ≥ 1. By definition, (C − 0: ) : | 5 (C), so we can put 5 (C) 5 (C) = ∈ [C]. e < (C − 0: ) :

Any root of e5 is a root of 5 , so it is one of 01, . . . , 0: . But e5 (0: ) ≠ 0: for < +1 if e5 (0: ) = 0 then (C − 0: ) | e5 (C) by Lemma 3.2.9, so (C − 0: ) : | 5 (C), a contradiction. Hence any root of e5 is one of 01, . . . , 0:−1. These are indeed roots of e5 , with multiplicities <1, . . . , <:−1. So by inductive hypothesis,

<1 <:−1 e5 (C) = (C − 01) ···(C − 0:−1) 6(C) for some 6(C) ∈ [C] with no roots, completing the induction.  Corollary 3.2.14 Let be a field and 5 ∈ [C]. Suppose that 5 has distinct roots 01, . . . , 0: ∈ with multiplicities <1, . . . , <: . Then <1 + · · · + <: ≤ deg( 5 ).  In other words, a polynomial of degree = has no more than = roots, even when you count the roots with multiplicities (e.g. count a double root twice). A field is algebraically closed if every nonconstant polynomial has at least one root. For example, C is algebraically closed (the fundamental theorem of algebra). Proposition 3.2.13 implies: Corollary 3.2.15 Let be an algebraically closed field and let 0 ≠ 5 ∈ [C]. Write 01, . . . , 0: for the distinct roots of 5 in , and <1, . . . , <: for their multi- plicities. Then <1 <: 5 (C) = 2(C − 01) ···(C − 0: ) , where 2 is the leading coefficient of 5 . 

32 3.3 Irreducible polynomials

Determining whether an integer is prime is generally hard, and determining whether a polynomial is irreducible is hard too. This section presents a few techniques for doing so. Let’s begin with the simplest cases. Recall Lemma 3.1.10(iv): a polynomial is irreducible if and only if it is nonconstant (has degree > 0) and cannot be expressed as a product of two nonconstant polynomials. Lemma 3.3.1 Let be a field and 5 ∈ [C]. i. If 5 is constant then 5 is not irreducible.

ii. If deg( 5 ) = 1 then 5 is irreducible. iii. If deg( 5 ) ≥ 2 and 5 has a root then 5 is reducible. iv. If deg( 5 ) ∈ {2, 3} and 5 has no root then 5 is irreducible.

Proof Parts (i) and (ii) follow from what we just recalled, and (iii) follows from Lemma 3.2.9. For (iv), suppose for a contradiction that 5 = 6ℎ with deg(6), deg(ℎ) ≥ 1. We have deg(6) + deg(ℎ) ∈ {2, 3}, so without loss of gener- ality, deg(6) = 1. Also without loss of generality, 6 is monic, say 6(C) = C + 0; but then 5 (−0) = 0, a contradiction. 

Warning 3.3.2 The converse of (iii) is false! To show a polynomial is irreducible, it’s not enough to show it has no root (unless it has degree 2 or 3). For instance, (C2 + 1)2 ∈ Q[C] has no root but is reducible.

?−1 Examples 3.3.3 i. Let ? be a prime. Then 5 (C) = 1 + C + · · · + C ∈ F? [C] is reducible, since 5 (1) = 0. ii. Let 5 (C) = C3 − 10 ∈ Q[C]. Then deg( 5 ) = 3 and 5 has no root in Q, so 5 is irreducible by part (iv) of the lemma. iii. Over C or any other algebraically closed field, the irreducibles are exactly the polynomials of degree 1.

Exercise 3.3.4 If I gave you a quadratic over Q, how would you decide whether it is reducible or irreducible?

From now on we focus on = Q. Any polynomial over Q can be multiplied by a nonzero rational constant to get a polynomial over Z, and that’s often a helpful move, so we’ll look at Z[C] too.

33 Definition 3.3.5 A polynomial over Z is primitive if its coefficients have no common divisor except for ±1. For example, 15 + 6C + 10C2 is primitive but 15 + 6C + 30C2 is not. Lemma 3.3.6 Let 5 (C) ∈ Q[C]. Then there exist a primitive polynomial (C) ∈ Z[C] and U ∈ Q such that 5 = U. 5 C Í 0 1 C8 0 1 Proof Write ( ) = 8 ( 8/ 8) , where 8 ∈ Z and 0 ≠ 8 ∈ Z. Take any common Í 8 multiple 1 of the 18s; then writing 28 = 081/18 ∈ Z, we have 5 (C) = (1/1) 28C . Now let 2 be the greatest common divisor of the 28s, put 38 = 28/2 ∈ Z, and put Í 8 (C) = 38C . Then (C) is primitive and 5 (C) = (2/1)(C).  If the coefficients of a polynomial 5 (C) ∈ Q[C] happen to all be integers, the word ‘irreducible’ could mean two things: irreducibility in the ring Q[C] or in the ring Z[C]. We say that 5 is irreducible over Q or Z to distinguish between the two. Suppose we have a polynomial over Z that’s irreducible over Z. In principle it could still be reducible over Q: although there’s no nontrivial way of factorizing it over Z, perhaps it can be factorized when you give yourself the freedom of non-integer coefficients. But the next result tells us that you can’t. Lemma 3.3.7 (Gauss) i. The product of two primitive polynomials over Z is primitive. ii. If a polynomial over Z is irreducible over Z, it is irreducible over Q. Proof For (i), let 5 and 6 be primitive polynomials over Z. Let ? be a prime number. (We’re going to show that ? doesn’t divide all the coefficients of 5 6.) Write c : Z → Z/?Z = F? for the canonical homomorphism, which induces a homomorphism c∗ : Z[C] → F? [C] as in Definition 3.1.6. Since 5 is primitive, ? does not divide all the coefficients of 5 . Equivalently, c∗( 5 ) ≠ 0. Similarly, c∗(6) ≠ 0. But F? [C] is an integral domain, so

c∗( 5 6) = c∗( 5 )c∗(6) ≠ 0, so ? does not divide all the coefficients of 5 6. This holds for all primes ?, so 5 6 is primitive. For (ii), let 5 ∈ Z[C] be a polynomial irreducible over Z. Let 6, ℎ ∈ Q[C] with 5 = 6ℎ. By Lemma 3.3.6, 6 = U and ℎ = V for some U,V ∈ Q and primitive ,  ∈ Z[C]. Then UV =

34 Gauss’s lemma quickly leads to a test for irreducibility. It involves taking a polynomial over Z and reducing it mod ?, for some prime ?. This means applying the map c∗ : Z[C] → F? [C] from the last proof. As we saw after Definition 3.1.6, Í 8 Í 8 if 5 (C) = 08C then c∗( 5 )(C) = c(08)C , where c(08) is the congruence class of 08 mod ?. I’ll write 0 = c(0) and 5 = c∗( 5 ). That is, 5 is ‘ 5 mod ?’. = Proposition 3.3.8 (Mod p method) Let 5 (C) = 00 + 01C + · · · + 0=C ∈ Z[C]. If there is some prime ? such that ? - 0= and 5 ∈ F? [C] is irreducible, then 5 is irreducible over Q.

I’ll give some examples first, then the proof. Examples 3.3.9 i. Let’s use the mod ? method to show that 5 (C) = 9+14C−8C3 3 is irreducible over Q. Take ? = 7: then 5 (C) = 2 − C ∈ F7 [C], so it’s enough 3 to show that 2−C is irreducible over F7. Since this has degree 3, it’s enough 3 to show that C = 2 has no solution in F7 (by Lemma 3.3.1(iv)). And you can easily check this by computing 03, (±1)3, (±2)3 and (±3)3 mod 7.

ii. The condition in Proposition 3.3.8 that ? - 0= can’t be dropped. For instance, consider 5 (C) = 6C2 + C and ? = 2.

Warning 3.3.10 Take 5 (C) as in Example 3.3.9(i), but this time take 3 ? = 3. Then 5 (C) = −C + C ∈ F3 [C], which is reducible. But that doesn’t mean 5 is reducible! The mod ? method only ever lets you show that a polynomial is irreducible over Q, not reducible.

Proof of Proposition 3.3.8 Take a prime ? satisfying the conditions in the Propo- sition. By Gauss’s lemma, it is enough to prove that 5 is irreducible over Z. Since 5 is irreducible, deg( 5 ) > 0, so deg( 5 ) > 0. Now let 5 = 6ℎ in Z[C]. We have 5 = 6ℎ and 5 is irreducible, so without loss of generality, 6 is constant. The leading coefficient of 5 is the product of the leading coefficients of 6 and ℎ, and is not divisible by ?, so the leading coefficient of 6 is not divisible by ?. Hence deg(6) = deg(6) = 0. 

We finish with an irreducibility test that turns out to be surprisingly powerful.

= Proposition 3.3.11 (Eisenstein’s criterion) Let 5 (C) = 00 + · · · + 0=C ∈ Z[C], with = ≥ 1. Suppose there exists a prime ? such that:

• ? - 0=;

• ? | 08 for all 8 ∈ {0, . . . , = − 1};

2 • ? - 00.

35 Then 5 is irreducible over Q.

To prove this, we will use the concept of the codegree codeg( 5 ) of a polynomial 5 C Í 0 C8 8 0 5 ( ) = 8 8 , which is defined to be the least such that 8 ≠ 0 (if ≠ 0), or as the formal symbol ∞ if 5 = 0. For polynomials 5 and 6 over an integral domain,

codeg( 5 6) = codeg( 5 ) + codeg(6).

Clearly codeg( 5 ) ≤ deg( 5 ) unless 5 = 0. Proof By Gauss’s lemma, it is enough to show 5 is irreducible over Z. Let 6, ℎ ∈ Z[C] with 5 = 6ℎ. Continue to write 5 (C) ∈ F? [C] for 5 reduced mod ?; then 5 = 6ℎ. Since 2 ? - 00 = 5 (0) = 6(0)ℎ(0), we may assume without loss of generality that ? - 6(0). Hence codeg(6) = 0. Also, codeg( 5 ) = =, since ? divides each of 00, . . . , 0=−1 but not 0=. So

= = codeg( 5 ) = codeg(6) + codeg(ℎ) = codeg(ℎ) ≤ deg(ℎ) ≤ deg(ℎ), (3.9)

giving = ≤ deg(ℎ). But 5 = 6ℎ with deg( 5 ) = =, so deg(ℎ) = = and deg(6) = 0.

Exercise 3.3.12 The last step in (3.9) was ‘deg(ℎ) ≤ deg(ℎ)’. Why is that true? And when does equality hold?

Example 3.3.13 Let 2 5 1 6(C) = C5 − C4 + C3 + ∈ Q[C]. 9 3 3 Then 6 is irreducible over Q if and only if

96(C) = 2C5 − 15C4 + 9C3 + 3

is irreducible over Q, which it is by Eisenstein’s criterion with ? = 3. Testing for irreducibility Exercise 3.3.14 Use Eisenstein’s criterion to show that for every = ≥ 1, there is an irreducible polynomial over Q of degree =.

I’ll give you one more example, and it’s not just any old polynomial: it’s an important one that we’ll need when we come to think about solvability by radicals. It needs a lemma. ? < 8 < ? ? ? Lemma 3.3.15 Let be a prime and 0 . Then | 8 .

36 For example, the 7th row of Pascal’s triangle is 1, 7, 21, 35, 35, 21, 7, 1, and 7 divides all of these numbers apart from the first and last. 8 ? 8 ? ? ? ? 8 ? 8 ? Proof We have !( − )! 8 = !, and divides ! but not ! or ( − )! (since < 8 < ? ? ? is prime and 0 ), so must divide 8 .  Example 3.3.16 Let ? be a prime. The ?th cyclotomic polynomial is C ? − 1 Φ (C) = 1 + C + · · · + C ?−1 = . (3.10) ? C − 1 I claim that Φ? is irreducible. We can’t apply Eisenstein to Φ? as it stands, because whichever prime we choose (whether it’s ? or another one) doesn’t divide any of the coefficients. However, we saw on p. 27 that Φ? (C) is irreducible if and only if Φ? (C − 2) is irreducible, for any 2 ∈ Q. We’ll take 2 = −1. We have (C + 1) ? − 1 Φ (C + 1) = ? (C + 1) − 1 ? 1 Õ C = C8 C 8 8=1 ?  ?  = ? + C + · · · + C ?−2 + C ?−1. 2 ? − 1

So Φ? (C + 1) is irreducible by Eisenstein’s criterion and Lemma 3.3.15, hence Φ? (C) is irreducible too.

Digression 3.3.17 I defined the ?th cyclotomic polynomial Φ? only when ? is prime. The definition of Φ= for general = ≥ 1 is not the obvious generalization of (3.10). It’s this: Ö Φ= (C) = (C − Z), Z

where the product runs over all primitive =th roots of unity Z. (In this context, ‘primitive’ means that = is the smallest number satisfying Z = = 1; it’s a different usage from ‘primitive polynomial’.)

Many surprising things are true. It’s not obvious that the coefficients of Φ= are real, but they are. Even given that they’re real, it’s not obvious that they’re rational, but they are. Even given that they’re rational, it’s not obvious that they’re integers, but they are. The degree of Φ= is i(=), the number of integers between 1 and = that are coprime with = (Euler’s function). It’s also true that the polynomial Φ= is irreducible for all =, not just primes. Some of these things are quite hard to prove, and results from Galois theory help. We probably won’t get into all of this, but you can read more here.

37 Chapter 4

Field extensions

Roughly speaking, an ‘extension’ of a field is a field " that contains as a subfield. It’s not much of an exaggeration to say that field extensions are the central objects of Galois theory, in much the same way that vector spaces are the central objects of linear algebra. Introduction to It will be a while before it becomes truly clear why field extensions are so Week 4 important, but here are a couple of indications:

• For any polynomial 5 over Q, we can take the smallest subfield " of C that contains all the complex roots of 5 , and that’s an extension of Q.

• For any irreducible polynomial 5 over a field , the quotient ring " = [C]/h 5 i is a field. The constant polynomials form a subfield of " isomor- phic to , so " is an extension of .

It’s important to distinguish between these two types of example. The first extends Q by all the roots of 5 , whereas the second extends by just one root of 5 —as we’ll see. But let’s begin at the beginning.

4.1 Definition and examples

Definition 4.1.1 A field extension consists of a field , a field ", and a homo- morphism ]: → ".

Since homomorphisms between fields are injective (Lemma 2.2.2), is iso- morphic to the subfield im(]) of ". It is usually safe to identify ](0) with 0 for each 0 ∈ —in other words, pretend that is actually a subfield of " and ](0) = 0. We then say that " is an extension of and write " : .

38 It’s worth taking a minute to make sure you understand the relationship between subsets and injections. This is a fundamental point about sets, not fields. Given a set  and a subset  ⊆ , there’s an inclusion function ]:  → , defined by ](1) = 1 for all 1 ∈ . (Remember that any function has a specified domain and codomain, so this isn’t the same thing as the identity on .) This inclusion function is an injective. On the other hand, given any injective function between sets, say i : - → , the image im i is a subset of , and there’s a i0 : - → im i given by i0(G) = i(G) (G ∈ -). Hence - is ‘isomorphic to’ (in bijection with) the subset im i of . This back-and-forth process means that subsets and injections are more or less the same thing.

Digression 4.1.2 If a field extension is basically the same thing as a field " together with a subfield , you might wonder why we bother with the more general Definition 4.1.1, involving an arbitrary homomorphism ]. It turns out that in the long run, it makes things easier. There are two factors at play. The first is purely set-theoretic. The concept of subset isn’t actually as simple as it appears, at least when you look at what mathematicians do rather than what we claim we do. For example, (1) it’s common to define the set C as R2, (2) everyone treats R as a subset of C, but (3) almost no one would say that R is a subset of R2 (if you wrote ‘the point c of R2’ on an exam, you’d be marked wrong). In truth, the common conventions are inconsistent. A good way to make everything respectable is to do everything in terms of injections rather than subsets. It would take up too much space to go into this here, but Definition 4.1.1 is one example of this approach in action. The second factor is algebraic. According to Definition 4.1.1, a field exten- sion is simply a homomorphism between fields, so it includes examples such as the conjugation map : C → C. You might feel this example obeys the letter but not the spirit of Definition 4.1.1. But again, it turns out to be useful to include such examples. When we come to count between fields in a few weeks, you’ll see why.

Examples 4.1.3 i. The inclusion ]: Q → C is a field extension, usually just written as C : Q. Similarly, there are field extensions C : R and R : Q.

ii. Let √ √ Q( 2) = {0 + 1 2 : 0, 1 ∈ Q}. √ Then Q( 2) is a subring of C (easily), and in fact it’s a subfield: for if 0, 1 , ( ) ≠ (0 0) then √ 1 0 − 1 2 √ = 0 + 1 2 02 − 212

39 √ (noting that the denominators√ are not 0 because 2 √is irrational). So we have an extension√ C : Q( 2). Also, because Q ⊆ Q( 2), we have another extension Q( 2) : Q. iii. By direct calculation or later theory (which will make it much easier), √ √ √ Q( 2, 8) = {0 + 1 2 + 28 + 3 28 : 0, 1, 2, 3 ∈ Q} √ is also a subfield of C, so we have an extension Q( 2, 8) : Q.

iv. Let be a field. A rational expression over is a ratio of two polynomials 5 (C) , 6(C)

where 5 (C), 6(C) ∈ [C] with 6 ≠ 0. Two such expressions, 51/61 and 52/62, are regarded as equal if 5162 = 5261 in [C]. So formally, a rational expression is an equivalence class of pairs ( 5 , 6) under the equivalence relation in the last sentence. The set of rational expressions over is denoted by (C). Rational expressions are added, subtracted and multiplied in the ways you’d expect, making (C) into a field. There is a homomorphism ]: → (C) given by ](0) = 0/1 (0 ∈ ). In other words, (C) contains a copy of as the constant rational expressions. So, we have a field extension (C) : .

v. In particular, when = F? for some prime ?, we have the field extension F? (C) : F?. Note that F? (C) is an infinite field of characteristic ?! Fields of positive characteristic don’t have to be finite. So I’ve now fulfilled the promise I made in Warning 2.2.12.

Warning 4.1.4 People sometimes say ‘’ to mean ‘rational expression’. But just as for polynomials (Warnings 3.1.2 and 3.1.3), I want to emphasize that rational expressions are not func- tions. For instance, 1/(C − 1) is a totally respectable element of (C). You don’t have to—and shouldn’t—worry about what happens when C = 1, because C is just a formal symbol (a mark on a piece of paper) rather than a variable, and 1/(C − 1) is just a formal expression, not a function. If this puzzles you, I suggest going back to those warnings about polynomials, which make the same point in a simpler setting.

40

Exercise√ 4.1.5 Find two examples of fields such that Q( ( Q( 2, 8). (The symbol ( means proper subset.)

For any kind of , there is a notion of the ‘substructure generated by’ a given subset. For example, when - is a subset of a group , you know what the subgroup generated by - is: it’s the intersection of all the subgroups containing -. Similarly, when - is a subset of a vector space +, the generated by - (or ‘spanned by -’, as one usually says) is the intersection of all the linear subspaces containing -. We now make a similar definition for fields.

Definition 4.1.6 Let be a field and - a subset of . The subfield of generated by - is the intersection of all subfields of containing -.

You can check that any intersection of subfields of is a subfield of (even if it’s an uncountably infinite intersection). In fact, you already did most of the work for this in Exercise 2.1.2. So, the subfield  of generated by - really is a subfield of . It contains - itself. By definition of intersection,  is the smallest subfield of containing -, in the sense that any subfield of containing - contains .

Exercise 4.1.7 Check the truth of all the statements in the previous paragraph.

Examples 4.1.8 i. The subfield of generated by œ is the prime subfield of .

ii. Let ! be the subfield of C generated by {8}. I claim that

! = {0 + 18 : 0, 1 ∈ Q}.

To prove this, we have to show that ! is the smallest subfield of C containing 8. First, it is a subfield of C (by an argument similar to Example 4.1.3(ii)) and it contains 0 + 18 = 8. Now let !0 be any subfield of C containing 8. Then !0 contains the prime subfield of C (by definition of prime subfield), which is Q. So whenever 0, 1 ∈ Q, we have 0, 1, 8 ∈ !0 and so 0 + 18 ∈ !0. Hence ! ⊆ !0, as required. √ iii. A very similar argument shows√ that the subfield of C generated by 2 is what we have been calling Q( 2).

41 Exercise 4.1.9 What is the subfield of C generated by {7/8}? By {2 + 38}? By R ∪ {8}?

We will be very interested in chains of fields

⊆ ! ⊆ " in which and " are regarded as fixed and ! as variable. You can think of as the floor, " as the ceiling, and ! as varying in between.

Definition 4.1.10 Let " : be a field extension and . ⊆ ". We write (.) for the subfield of " generated by ∪ ., and call it with . adjoined.

So, (.) is the smallest subfield of " containing both and .. When . is a finite set {U1,...,U=}, we write ({U1,...,U=}) as (U1,...,U=). √ " . . Examples 4.1.11 i. Take : to be C√: Q and = { 2}. Then ( ) is the smallest subfield of C containing Q∪ { 2}. But every subfield of C contains . Q: that’s what it means for Q to be√ the prime subfield of C. So, ( ) is the smallest subfield of C containing√ 2. By Example 4.1.8√(iii), that’s exactly√ what we’ve been calling Q( 2) all along. We refer to Q( 2) as ‘Q with 2 adjoined’.

ii. Similarly, Q with 8 adjoined is

Q(8) = {0 + 18 : 0, 1 ∈ Q} √ , 8 (Example√ 4.1.8(ii)), and Q with { 2 } adjoined is the subfield denoted by Q( 2, 8) in Example 4.1.3(iii).

iii. Let " be a field and - ⊆ ". Write for the prime subfield of ". Then (-) is the smallest subfield of " containing and -. But every subfield of " contains , by definition of prime subfield. So (-) is the smallest subfield of " containing -; that is, it’s the subfield of " generated by -. √ We already saw this argument in (i), in the case " = C and - = { 2}.

iv. Let be any field and let " be the field (C) of rational expressions over , which is an extension of . You might worry that there’s some ambiguity in the notation: (C) could either mean the field of rational expressions over (as defined in Example 4.1.3(iv)) or the subfield of (C) obtained by adjoining the element C of (C) to (as in Definition 4.1.10).

42 In fact, they’re the same. In other words, the smallest subfield of (C) containing and C is (C) itself. Or equivalently, the only subfield of (C) containing and C is (C) itself. To see this, let ! be any such subfield. For Í 8 any polynomial 5 (C) = 08C over , we have 5 (C) ∈ !, since 08,C ∈ ! and ! is closed under multiplication and addition. Hence for any polynomials 5 (C), 6(C) over with 6(C) ≠ 0, we have 5 (C), 6(C) ∈ !, so 5 (C)/6(C) ∈ ! as ! is closed under division by nonzero elements. So ! = (C).

Warning 4.1.12 It is not true in general that

(U) = {0 + 1U : 0, 1 ∈ } (false!) (4.1) √ 8 Examples√ like Q( 2) and Q( ) do satisfy this, but that’s only because 2 and 8 satisfy quadratic equations. Certainly the right-hand side is a subset of (U), but in general it’s much smaller, and isn’t a subfield. You’ve just seen an example: the field (C) of rational expressions is much bigger than the set {0 + 1C : 0, 1 ∈ } of polynomials of degree ≤ 1. And that set of polynomials isn’t closed under multiplication. Another example: let U be the real cube root of 2. You can show that U2 cannot be expressed as 0 + 1U for any 0, 1 ∈ Q (a fact we’ll come back to in Example 4.2.9(ii)). But U ∈ Q(U), so U2 ∈ Q(U), so (4.1) fails in this case. In fact,

Q(U) = {0 + 1U + 2U2 : 0, 1, 2 ∈ Q}.

We’ll see why next week.

Exercise 4.1.13 Let " : be a field extension. Show that (. ∪/) = ( (.))(/) whenever .,/ ⊆ ". (For example, (U,V) = ( (U))(V) whenever U,V ∈ ".)

4.2 Algebraic and transcendental elements

A complex number U is said to be ‘algebraic’ if

= 00 + 01U + · · · + 0=U = 0 for some rational numbers 08, not all zero. (You may have seen this definition with ‘integer’ instead of ‘’; it makes no difference, as you can always clear the denominators.) This concept generalizes to arbitrary field extensions:

43 Definition 4.2.1 Let " : be a field extension and U ∈ ". Then U is algebraic over if there exists 5 ∈ [C] such that 5 (U) = 0 but 5 ≠ 0, and transcendental otherwise.

Exercise 4.2.2 Show that every element of is algebraic over .

Examples 4.2.3 i. Let = ≥ 1. Then 42c8/= ∈ C is algebraic over Q, since 5 (C) = C= − 1 is a nonzero polynomial such that 5 (42c8/=) = 0.

ii. The numbers c and 4 are both transcendental over Q. Both statements are hard to prove (and we won’t). By Exercise 4.2.2, any complex number transcendental over Q is irrational. Proving the irrationality of c and 4 is already a challenge; proving they’re transcendental is even harder.

iii. Although c is transcendental over Q, it is algebraic over R, since it’s an element of R. (Again, we’re using Exercise 4.2.2.) Moral: you shouldn’t say an element of a field is just ‘algebraic’ or ‘transcendental’; you should say it’s ‘algebraic/transcendental over ’, specifying your . Or at least, you should do this when there’s any danger of confusion.

iv. Take the field (C) of rational expressions over a field . Then C ∈ (C) is transcendental over , since 5 (C) = 0 ⇐⇒ 5 = 0.

The set of complex numbers algebraic over Q is written as Q. It’s a fact that Q is a subfield of C, but this is extremely hard to prove by elementary means. Next week I’ll show you that with a surprisingly small amount of abstract algebra, you can transform this from a very hard problem into an easy one. So that you appreciate the miracle later, I give you this unusual exercise now.

Exercise 4.2.4 Attempt to prove any part of the statement that Q is a subfield of C. For example, try to show that Q is closed under addition, or multiplication, or reciprocals. I have no idea how to do any of these using only our current tools, but it’s definitely worth a few minutes of doomed effort to get a sense of the difficulties.

Let " : be a field extension and U ∈ ". An annihilating polynomial of U is a polynomial 5 ∈ [C] such that 5 (U) = 0. So, U is algebraic if and only if it has some nonzero annihilating polynomial. It is natural to ask not only whether U is annihilated by some nonzero polyno- mial, but which polynomials annihilate it. The situation is pleasantly simple:

44 Lemma 4.2.5 Let " : be a field extension and U ∈ ". Then there is a polynomial <(C) ∈ [C] such that

h

If U is transcendental over then < = 0. If U is algebraic over then there is a unique < satisfying (4.2).

Proof By the universal property of polynomial rings (Lemma 3.1.5), there is a unique homomorphism \ : [C] → " such that \(0) = 0 for all 0 ∈ and \(C) = U. (Here we’re taking the ‘i’ of Lemma 3.1.5 to be the inclusion → ".) Then

Õ 8 Õ 8 \ 08C = 08U

Í 8 for all 08C ∈ [C], so

ker \ = {annihilating polynomials of U over }.

Now ker \ is an ideal of the principal ideal domain [C] (using Proposition 3.2.2), so ker \ = h

Exercise 4.2.7 What is the minimal polynomial of an element 0 of ?

This is an important definition, so we give some equivalent ways of stating it.

Lemma 4.2.8 Let " : be a field extension, let U ∈ " be algebraic over , and let < ∈ [C] be a monic polynomial. The following are equivalent:

i. < is the minimal polynomial of U over ;

45 ii. <(U) = 0, and < | 5 for all annihilating polynomials 5 of U over ; iii. <(U) = 0, and deg(<) ≤ deg( 5 ) for all nonzero annihilating polynomials 5 of U over ; iv. <(U) = 0 and < is irreducible over . Part (iii) says the minimal polynomial is a monic annihilating polynomial of least degree. Proof (i)⇒(ii) follows from the definition of minimal polynomial. (ii)⇒(iii) because if < | 5 ≠ 0 then deg(<) ≤ deg( 5 ). (iii)⇒(iv): assume (iii). First, < is not constant: for if < is constant then < = 1 (since < is monic); but <(U) = 0, so 1 = 0 in , a contradiction. Next, suppose that < = 5 6 for some 5 , 6 ∈ [C]. Then 0 = <(U) = 5 (U)6(U), so without loss of generality, 5 (U) = 0. By (iii), deg( 5 ) ≥ deg(<), so deg( 5 ) = deg(<) and deg(6) = 0. This proves (iv). (iv)⇒(i): assume (iv), and write