Skip to the Main Content

Note:These pages make extensive use of the latest XHTML and CSS Standards. They ought to look great in any standards-compliant modern browser. Unfortunately, they will probably look horrible in older browsers, like Netscape 4.x and IE 4.x. Moreover, many posts use MathML, which is, currently only supported in Mozilla. My best suggestion (and you will thank me when surfing an ever-increasing number of sites on the web which have been crafted to use the new standards) is to upgrade to the latest version of your browser. If that's not possible, consider moving to the Standards-compliant and open-source Mozilla browser.

August 25, 2023

Grothendieck–Galois–Brauer Theory (Part 2)

Posted by John Baez

Last time we took a tiny taste of Grothendieck’s approach to Galois theory. Now let’s dig in a bit deeper!

Grothendieck’s approach to Galois theory makes heavy use of an analogy with covering spaces in topology, so let’s start there.

Covering spaces

Here is a covering space of the circle:

There are 4 points sitting over any point of the circle. So, if you stand anywhere on the circle, 4 guys could be standing directly above you, one at each of these points.

Suppose now that you play a game: as you walk around the circle, these guys have to follow you, staying directly overhead. If you walk all the way around the circle once, these guys switch places in a certain way. With this particular covering space they get cyclically permuted:

12 1 \mapsto 2 23 2 \mapsto 3 34 3 \mapsto 4 41 4 \mapsto 1

But you could play this sort of game with any covering space of the circle, and you’d get some permutation of some bunch of guys. Even better: for any permutation of any set of guys, we can cook up a covering space of the circle, such that when you play this game and walk once around the circle, the guys will carry out the desired permutation!

To formalize this, you can make up a category of covering spaces of the circle, and a category of ‘sets equipped with permutation’, and prove these categories are equivalent.

But there’s no need to limit yourself to the circle. Suppose you have a connected manifold XX, and suppose you choose a point xXx \in X. Then you can go back and forth between covering spaces

p:EXp \colon E \to X

and actions of the fundamental group π 1(X,x)\pi_1(X,x) on sets. The idea is that sitting over the point xx there is some set of points in EE. That’s your set:

S={eE|p(e)=x} S = \{e \in E \vert \; p(e) = x \}

If you start at xx and walk around a loop in XX you can get little guys standing at these points to follow you overhead in EE, and they’ll come back permuted. That’s how the fundamental group π 1(X,*)\pi_1(X,\ast) acts on the set SS. Conversely — and this takes a bit more work — given any set SS equipped with an action of π 1(X,x)\pi_1(X,x), we can build a covering space p:EXp \colon E \to X.

We can formalize this result as an equivalence of categories. We could also generalize it. For example, we could drop the requirement that XX is connected, and avoid choosing a favorite point xx, by replacing the fundamental group with the fundamental groupoid. There’s also no need for XX to be a manifold — though it does need to be nicer than an arbitrary topological space. To see all these things, try:

But I want to do the opposite of generalizing: I want to limit myself to compact manifolds. Why? To bring in some finiteness conditions that we’ll see analogues of later, when we get to Galois theory.

If we require that our manifold XX and its covering space EE are compact, this forces our set

S={eE|p(e)=x} S = \{e \in E \vert \; p(e) = x \}

to be finite. Conversely if XX is compact and SS is finite, the covering space EE will be compact. So I believe we get this:

Theorem 9. Suppose XX is a connected compact manifold and xXx \in X. Then the the category of compact covering spaces of XX is equivalent to the category of finite sets equipped with an action of π 1(X,x)\pi_1(X,x).

Furthermore, a covering space EE is connected iff the corresponding action of π 1(X,x)\pi_1(X,x) is transitive. This a way of saying a guy up in EE can walk from any point over xx to any other point over xx by following me as I walk around some loop in XX. So, we get

Theorem 10. Suppose XX is a connected compact manifold and xXx \in X. Then the category of connected compact covering spaces of XX is equivalent to the category of finite sets equipped with a transitive action of π 1(X,x)\pi_1(X,x).

By the way, I’m carrying on the numbering of theorems from last time. And I’m mentioning these two results mainly as a warmup for two analogous results in Galois theory. Those are my destination today.

Galois theory

In algebraic geometry we learn that any commutative ring acts like a ring of functions on a kind of space called an affine scheme. You can click the link for the definition, but the upshot is that we can define the category of affine schemes, AffSch\mathrm{AffSch}, to be the opposite of the category of commutative rings, CommRing\mathrm{CommRing}. So the objects in these two categories are really the same, but we think about them differently.

We call the affine scheme associated to a commutative ring its spectrum, ultimately because the spectrum of light produced by a hot gas is an example. But never mind that now. All you need to know is this: any commutative ring RR gives an affine scheme called Spec(R)\mathrm{Spec}(R), and any ring homomorphism

f:RSf \colon R \to S

corresponds to a map of affine schemes

Spec(f):Spec(S)Spec(R) \mathrm{Spec}(f) \colon \mathrm{Spec}(S) \to \mathsf{Spec}(R)

Now, with a lot more work we can define what it means for a map of affine schemes to be like a covering space, and we can copy Theorem 9 in this context. And we get the fundamental theorem of Grothendieck’s Galois theory!

To keep things as lowbrow as possible I want to talk about commutative rings than affine schemes. When a map of affine schemes

Spec(f):Spec(S)Spec(R) \mathrm{Spec}(f) \colon \mathrm{Spec}(S) \to \mathsf{Spec}(R)

is like a covering space, we say the ring homomorphism

f:RS f \colon R \to S

is ‘etale’. So the question is: when is a ring homomorphism etale?

The answer is simplest when RR is a field. In this case any homomorphism f:RSf \colon R \to S is one-to-one, and SS becomes a commutative algebra over the field RR. Then ff is etale iff SS is a separable commutative algebra over RR.

Why is this true? That’s a very important question, which I want tackle someday. But first: you may be wondering what ‘separable’ means. I said a lot about separable algebras last time, but that was a long time ago, so let me remind you. Skip the next section if you remember.

Separable algebras

Let’s change notation now and let AA be any algebra over a field kk. AA is separable if can be equipped with the structure of a special Frobenius algebra. Unpacking the jargon, this means we can find a map

Δ:AAA \Delta \colon A \to A \otimes A

which has

m(Δ(a))=a(1) \qquad \qquad m(\Delta(a)) = a \qquad \qquad \qquad \qquad (1)

and

aΔ(b)=Δ(ab)=Δ(a)b(2) \qquad a \Delta(b) = \Delta(a b) = \Delta(a) b \qquad \qquad (2)

In the first equation

m: AA A ab ab \begin{array}{rccl} m \colon & A \otimes A &\to& A \\ & a \otimes b &\to& a b \end{array}

is the multiplication in AA, and in the second I’m making AAA \otimes A into an (A,A)(A,A)-bimodule in the obvious way:

a(bc)=abc,(bc)a=bca a(b \otimes c) = a b \otimes c, \qquad (b \otimes c) a = b \otimes c a

If we write the multiplication mm like this:

\quad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad

and write the ‘comultiplication’ Δ\Delta like this:

\quad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad

then equation (1), called the special law, says

\quad \qquad \qquad \qquad \qquad \qquad \qquad \qquad

while equation (2), called the Frobenius laws, says

\quad \qquad \qquad \qquad \qquad

Separable field extensions

Last time I explained that when kk is an algebraically closed field, the only commutative separable algebras over kk are finite products

k××k k \times \cdots \times k

Geometrically speaking, this says that the coverings of Spec(k)\mathrm{Spec}(k) are quite boring. It’s like how a covering space of a point is just a bunch of points. But now let’s see what happens for fields that may not be algebraically closed.

Theorem 11. Any commutative separable algebra over a field kk is a finite product of finite separable extensions of kk.

More buzzwords! And worse, the word ‘separable’ in the phrase ‘separable extension’ doesn’t mean the same thing as it does in the phrase ‘separable algebra’. They are related, and indeed the theorem says how. But I should explain.

An extension of a field kk is a field KK that contains kk as a subfield. When this happens KK will be a vector space over kk, indeed an algebra. We say the extension is finite if KK is finite-dimensional as a vector space over kk.

Suppose KK is any extension of kk.

An element xKx \in K is algebraic over kk if it obeys some polynomial equation with coefficients in kk. For example, if KK is a finite extension of kk then every xKx \in K is algebraic. Otherwise all the elements 1,x,x 2,1, x, x^2, \dots would be linearly independent over kk, and KK would not be finite-dimensional!

If xKx \in K is algebraic over kk, there is a unique polynomial PP with coefficients in kk such that

  • P(x)=0P(x) = 0,
  • PP has the lowest possible degree for a nonzero polynomial that vanishes on xx,
  • the leading coefficient of PP is 11.

This is called the minimal polynomial of xx.

If xKx \in K is algebraic over kk, we say it’s separable if the derivative of its minimal polynomial is nonzero. And finally, we say the extension KK of kk is separable if all its elements are separable.

I don’t really like throwing piles of definitions at you like this. I want to explain separable field extensions more conceptually. However, today my main goal is to state the fundamental theorem of Grothendieck’s Galois theory — just to put it on the table. So, for now, I’ll only say a few reassuring words to the beginners out there.

First, saying an algebraic element of an extension of kk is separable is a bit like saying it’s ‘generic’: no weird coincidences are making the derivative of its minimal polynomial zero.

Second, it actually takes some work to find nonseparable finite extensions of a field! For a field of characteristic zero, or a finite field, all finite extensions are separable.

So, you can think of separable finite extensions of fields as the ‘easiest to understand, least annoying’ extensions. And Theorem 11 links them to separable commutative algebras in a very nice way. But this is not getting to the bottom of separability — not by a long shot.

Galois theory in a nutshell

Before diving into Grothendieck’s Galois theory we need to know a bit about ordinary Galois theory. I never took a course on Galois theory because it seemed dry and uninteresting at the time. As a youth I was interested in quarks and black holes, not why you can’t trisect a bloody angle using archaic tools! Now, in my old age, I am condemned not only to study Galois theory but to blog about it. I hope you’re not in the same boat. I hope you know all this stuff and understand it deeply.

Given a finite separable extension KK of a field kk, we can form the group of automorphisms of KK that fix every element of kk. This group is always finite. And when we’re lucky, only elements of kk are fixed by everything in this group. In this case we say KK is a Galois extension of kk, and a bunch of nice things happen.

For starters, in this case we call the group of automorphisms of KK that fix every element of kk the Galois group Gal(K|k)\mathrm{Gal}(K|k). But that’s not the main nice thing!

The first nice thing is that for any subgroup GGal(K|k)G \subseteq \mathrm{Gal}(K|k), there’s a field FF consisting of all elements of KK fixed by everything in GG. This field has

kFK k \subseteq F \subseteq K

so we call it an intermediate field: intermediate between kk and KK, that is.

Conversely, for any intermediate field kFKk \subseteq F \subseteq K we get a group GGal(K|k)G \subseteq \mathrm{Gal}(K|k) consisting of precisely those elements that fix everything in FF.

In fact, this sets up a one-to-correspondence between subgroups

GGal(K|k)G \subseteq \mathrm{Gal}(K|k)

and intermediate fields with

kFKk \subseteq F \subseteq K

Bigger subgroups GG correspond to smaller intermediate fields FF. Furthermore, normal subgroups GGal(K|k)G \subseteq \mathrm{Gal}(K|k) correspond to fields kFKk \subseteq F \subseteq K that are Galois extensions of kk. Even bettter, if GG is normal then the corresponding intermediate field FF has

Gal(F|k)Gal(K|k)/G \mathrm{Gal}(F|k) \cong \mathrm{Gal}(K|k) / G

The analogy between covering spaces and Galois theory

Everything I just said about Galois theory has an analogue for covering spaces. Thanks to stuff I already said, if KK is a finite separable extension of kk then the inclusion

i:kK i \colon k \to K

gives an etale map

Spec(i):Spec(K)Spec(k) \mathrm{Spec}(i) \colon \mathrm{Spec}(K) \to \mathrm{Spec}(k)

and we should think of this as a covering space. It will always be a connected covering space with finitely many sheets. If KK is a Galois extension, the Galois group Gal(K|k)\mathrm{Gal}(K|k) is the group of deck transformations of this cover. Then subgroups GGal(K|k)G \subseteq \mathrm{Gal}(K|k) correspond to intermediate fields

kFK k \subseteq F \subseteq K

which correspond to what you might call ‘intermediate coverings’

Spec(K)Spec(F)Spec(k) \mathrm{Spec}(K) \to \mathrm{Spec}(F) \to \mathrm{Spec}(k)

And so on!

But there are some things missing from this wonderful analogy. First, we don’t yet have an analogue of the universal covering space of Spec(k)\mathrm{Spec}(k), because this may not have finitely many sheets. Second, we don’t have an analogue of the fundamental group of Spec(k)\mathrm{Spec}(k), since this may be an infinite group. And third, I have not described anything analogous to disconnected covering spaces.

All these deficiencies prevent us from stating a Galois theory analogue of the big result we had for covering spaces:

Theorem 9. Suppose XX is a compact connected manifold and xXx \in X. Then there is an equivalence between the category of compact covering spaces of XX and the category of finite sets equipped with an action of π 1(X,x)\pi_1(X,x).

Grothendieck fixed this problem! And not just for fields — he handled all commutative rings and even more. But today I’m just doing fields.

The absolute Galois group

Let’s start by getting our hands on the Galois theory analogues of universal covering spaces and the fundamental group.

Any field kk has a separable closure k sepk^{\mathrm{sep}}, which is a separable extension containing all finite separable extensions of kk. It’s usually not a finite extension of kk, but it’s a union of finite extensions. In fact it’s a union of Galois extensions!

The map

Spec(k sep)Spec(k) \mathrm{Spec}(k^{\mathrm{sep}}) \to \mathrm{Spec}(k)

acts like the universal covering space of Spec(k)\mathrm{Spec}(k). Spec(k sep) \mathrm{Spec}(k^{\mathrm{sep}}) is too big to be a finite-sheeted covering space, but it maps down to any finite-sheeting connected covering space of Spec(k)\mathrm{Spec}(k), since any Galois extension KK of kk is isomorphic to a subfield of k sepk^{\mathrm{sep}}, giving us inclusions

kKk sep k \subseteq K \subset k^{\mathrm{sep}}

and thus maps

Spec(k sep)Spec(K)Spec(k) \mathrm{Spec}(k^{\mathrm{sep}}) \to \mathrm{Spec}(K) \to \mathrm{Spec}(k)

Next, what about the fundamental group? In topology this is isomorphic to the group of deck transformations of the universal covering space. So the fundamental group of Spec(k)\mathrm{Spec}(k) should be like Gal(k sep|k)\mathrm{Gal}(k^{\mathrm{sep}}|k).

Beware: k sepk^{\mathrm{sep}} is not a Galois extension of kk, since it’s not a finite extension. But we proceed with caution, and say the absolute Galois group of kk is the group of automorphisms of k sepk^{\mathrm{sep}} that fix kk. It’s called Gal(k)\mathrm{Gal}(k).

Since this is supposed to be the ‘mother of all Galois groups’, we would hope that any intermediate field

kKk sep k \subseteq K \subseteq k^{\mathrm{sep}}

corresponds to some subgroup of Gal(k)\mathrm{Gal}(k), and vice versa. Or something like that. But we are mainly interested in the case where KK is a finite separable extension of kk, and it’s certainly not true that these correspond bijectively to subgroups of GG. The group Gal(k)\mathrm{Gal}(k) is too infinite, too wild and woolly, for that to be true.

So, what we do is put a topology on Gal(k)\mathrm{Gal}(k) and show that the open subgroups of Gal(k)\mathrm{Gal}(k) correspond to finite separable extensions of kk.

That sounds a bit weird at first… we were trying to do algebra analogous to topology, but now topology has snuck into our algebra. But don’t worry: this topology is secretly just algebra. How do we put this topology on Gal(k)\mathrm{Gal}(k)? To make Gal(k)\mathrm{Gal}(k) into a topological group we just need to give a base of open neighborhoods of the identity. And we’ll take these neighborhoods to be certain normal subgroups of Gal(k)\mathrm{Gal}(k). Namely, for each intermediate field

kKk sep k \subseteq K \subseteq k^{\mathrm{sep}}

that’s a Galois extension, we’ll take the subgroup of Gal(k)\mathrm{Gal}(k) fixing every element of KK!

With this choice, the thing we want happens!

Theorem 12. For any field kk, the subset of k sepk^{\mathrm{sep}} fixed by some open subgroup GGal(k)G \subseteq \mathrm{Gal}(k) is a finite separable extension of kk, and this gives a one-to-one correspondence between

  • intermediate fields kKk sepk \subseteq K \subseteq k^{\mathrm{sep}} that are finite separable extensions of kk

and

  • open subgroups of Gal(k)\mathrm{Gal}(k).

Galois theory versus covering spaces

Theorem 12 is a nice analogue of the theorem we saw about connected covering spaces. Remember:

Theorem 10. Suppose XX is a connected compact manifold and xXx \in X. Then the category of connected compact covering spaces of XX is equivalent to the category of finite sets acted on transitively by π 1(X,x)\pi_1(X,x).

How are they analogous? They look different! Theorem 12 gives one-to-one correspondence, while Theorem 10 gives an equivalence of categories. Theorem 12 talks about subgroups of some group, while Theorem 10 talks about transitive actions of some group. So there is a difference in style between these results, as well as a difference in subject matter.

Part of the point is that subgroups give transitive actions. If HH is a subgroup of a group GG then GG acts transitively on G/HG/H by left multiplication. And any transitive action of GG is isomorphic to one of this form. But it’s only isomorphic, not equal. So when we work with transitive group actions, we need to work with a category of them, not a set.

We could bridge this gap of styles and make Theorem 10 and Theorem 12 look more similar. But while we’re doing it, we might as well get rid of the connectedness assumption in Theorem 12. Doing the analogous thing in Theorem 10 gives us the fundamental theorem of Grothendieck’s Galois theory.

The fundamental theorem of Grothendieck’s Galois theory

Here it is:

Theorem 13 (The Fundamental Theorem of Grothendieck’s Galois Theory). The category of commutative separable algebras over a field kk is equivalent to the opposite of the category of finite sets equipped with a continuous action of the absolute Galois group Gal(k)\mathrm{Gal}(k).

I hope you admire how it looks like the result I started with:

Theorem 9. Suppose XX is a connected compact manifold and xXx \in X. Then the the category of compact covering spaces of XX is equivalent to the category of finite sets equipped with an action of π 1(X,x)\pi_1(X,x).

The opposite shows up in Theorem 13 because of the duality between commutative algebra and topology. The continuity shows up in Theorem 13 because we are putting a topology on the absolute Galois group to keep it under control.

I also hope you understand a bit about what Theorem 13 is claiming. Note that while a finite set XX has the discrete topology, it’s still not trivial to say an action

Gal(k)×XX \mathrm{Gal}(k) \times X \to X

is continuous, since it has to be continuous in the first variable too, and Gal(k)\mathrm{Gal}(k) has a nontrivial topology. Ultimately we’re saying that for some open neighborhood UU of the identity, gUg \in U acts trivially on XX. And by how we’ve defined the topology on Gal(k)\mathrm{Gal}(k), we can assume UU is the subgroup of Gal(k)\mathrm{Gal}(k) that fixes every element of some Galois extension of kk.

This winds up meaning we can build XX from some Galois extension of kk. And ultimately we do it using this:

Theorem 11. Any commutative separable algebra over a field kk is a finite product of finite separable extensions of kk.

since we can find a Galois extension of kk that contains all these finite separable extensions.

I also hope you see what Theorem 13 says when kk is algebraically closed. Then its Galois group is trivial, and we get a result from last time:

Theorem 4\prime. Let kk be an algebraically closed field. Then the category of commutative separable algebras over kk is equivalent to the opposite of the category of finite sets.

Of course, you may still not deeply understand the Fundamental theorem of Grothendieck’s Galois theory. Maybe I don’t either. But I hope you see, at least roughly, that it strengthens the analogy between Galois theory and the theory of covering spaces… and in the case of algebraically closed fields it describes the category of finite sets using linear algebra! That’s a big deal.

There’s a lot left to explain, obviously. But what I really want to do next time is to say more about separable algebras. There are some huge questions I’ve left unanswered, like:

  • Why are separable commutative algebras over a field kk precisely the ones that give covering spaces of Spec(k)Spec(k)?

  • Why are separable commutative algebras over kk the same as finite products of finite separable extensions of kk?

I do want to tackle these. But until then, try this:

Puzzle. Let kk be a field. Guess a recipe for getting a finite set Φ(A)\Phi(A) on which Gal(k)\mathrm{Gal}(k) acts continously from any commutative separable algebra AA over kk. This recipe should be contravariant: a map of algebras f:AAf \colon A \to A' should give a map Φ(f):Φ(A)Φ(A).\Phi(f) \colon \Phi(A') \to \Phi(A). Can you guess a way to go back?

If you can do this, you’re well on your way to proving the fundamental theorem.

References

I think the most useful reference for today is this:

We give an explicit proof of the fundamental theorem of Grothendieck Galois theory: the category of separable algebras over a field kk is anti-equivalent to the category of continuous actions on finite sets of the profinite fundamental group of kk.

That group is the absolute Galois group of kk. For the topology on this group, and basic facts about how that topology works, I also recommend this:

  • Juergen Neukirch, Algebraic Number Theory: Chapter IV.1, Infinite Galois Theory, Springer, Berlin, 2013.

Both these references have the advantage of being quite down-to-earth.

Posted at August 25, 2023 10:29 AM UTC

TrackBack URL for this Entry:   https://golem.ph.utexas.edu/cgi-bin/MT-3.0/dxy-tb.fcgi/3477

0 Comments & 1 Trackback

Read the post Grothendieck--Galois--Brauer Theory (Part 4)
Weblog: The n-Category Café
Excerpt: The word 'separable' means two things, but today we'll use some geometry to show every finite separable extension of a field is a separable algebra over that field. And don't worry, I'll explain what all this stuff means!
Tracked: September 3, 2023 12:32 PM

Post a New Comment