You're reading the public-facing archive of the Category Theory Zulip server.
To join the server you need an invite. Anybody can get an invite by contacting Matteo Capucci at name dot surname at gmail dot com.
For all things related to this archive refer to the same person.
I'm trying to understand the "big Witt ring" of a commutative ring; there are various perspectives on it.
1) One slick abstract perspective is that the forgetful functor from [[lambda-rings]] to commutative rings has both the expected left adjoint but also a right adjoint
and is the big Witt ring. This perspective is helpful in proportion to how well you understand lambda-rings. There are several perspectives on lambda-rings, each of which should give a different outlook on the big Witt ring. My favorite is that when you decategorify a 2-rig by taking its Grothendieck group, you get not only a ring but a lambda-ring. But there is another important perspective based on number theory, where a lambda-ring is a commutative ring equipped with commuting [[Frobenius lifts]], one for each prime .
2) The underlying set of is
where is the underlying ring of the free lambda-ring on one generator. I will explain why later.
is a biring, i.e. a ring object in , and the multiplication in comes from the comultiplication in .
3) Since one can show the underlying commutative ring of the free lambda-ring on one generator is
we have an isomorphism of sets
where
is mapped to
Then the challenge is to describe the addition and multiplication on in these terms. People often do this with explicit formulas, which I tend to find cryptic. The addition in corresponds to the multiplication in , which is why we use this description. The multiplication in is more tricky. For any we get an element of called , defined by
and the multiplication on turns out to be determined once we know
This formula turns out to be very useful but I don't have a good understanding of how it comes from 2).
4) There's a ring isomorphism
where is sent to , and
sends to something called its ith ghost component. Here's an explicit formula for the ghost components that one often sees. Start by using the isomorphism
from 3) to write as a power series in . Then write
This is quite cryptic at first sight, but there has got to be some conceptual interpretation, probably involving [[Adams operations]].
5) Back to conceptual interpretations, Cartier noticed that the big Witt ring can be seen as the ring of all formal curves starting at in the multiplicative group of ... or something like that. This is explained here, where that ring of formal curves starting at is called .
This is probably just another way of thinking about 3), but it connects the big Witt ring to formal group laws, and in particular the "multiplicative formal group law", and I believe this should ultimately clarify the following trio of facts: a) the -theory of any space is a -ring, b) -theory is a complex oriented cohomology theory, 3) such cohomology theories are classified by formal group laws, 4) -theory corresponds to the multiplicative formal group law.
(Cartier generalized the big Witt ring to an arbitrary formal group law.)
It should be possible to clearly see how all these facts follow from the definition 1), but I'm not there yet!
Here's how 1) implies 2). We use the fact that the forgetful functor
has not only a left adjoint
but also a right adjoint
Let's use this to compute the underlying set of for some commutative ring .
The underlying set of any commutative ring is since is the free commutative ring on one generator.
Thus the underlying set of (the underlying commutative ring of) is
but since has a left adjoint this is
and since has a left adjoint this is
By general nonsense is the free lambda-ring on one generator. The underlying commutative ring of this, , is denoted . So, from this we get
The underlying set of is isomorphic to the set of ring homomorphisms from to .
It happens that is the ring of [[symmetric functions]], but we didn't need this in the above argument. It also happens that the ring of symmetric functions is the polynomial ring on generators called [[elementary symmetric functions]]. This is called the fundamental theorem of symmetric functions. We used this to see why 2) implies 3).
Where did this come from original and why is it named after Witt? I've never heard of it, so I'm curious how it connects to other things and why you're interested in understanding it now ;)
There was an exchange between you and James Borger back here which touched on Big Witt vectors. Jack Morava and David Ben-Zvi chip in and get to Adams operations. Maybe a conversation worth mining.
Morgan Rogers (he/him) said:
Where did this come from original and why is it named after Witt? I've never heard of it, so I'm curious how it connects to other things and why you're interested in understanding it now ;)
Let me dig into the history....
Okay, it turns out Witt introduced a kind of "Witt vector" associated to an algebra over a field of characteristic back in 1936, in his paper Zyklische Körper und Algebren der Characteristik vom Grad . Struktur diskret bewerteter perfekter Körper mit vollkommenem Restklassenkörper der Charakteristik . No, that's not the whole paper - it's just the title.
Given a commutative ring and a prime we can make up a "-typical Witt ring" containing these Witt vectors. But then someone noticed you can combine all these -typical Witt rings in a single "big Witt ring", and that's what I'm talking about.
People use these for various mysterious number-theoretic tasks, which I would like to understand someday.
But here's why I'm actually interested!
The big Witt ring is now best understood as the cofree lambda-ring on a commutative ring. Lambda-rings are important in representation theory, where they let us study things like exterior powers and symmetric powers of representations. But also, surprisingly, they're important in number theory, where they let us study things like "Frobenius lifts" - lifting the Frobenius endomorphism of a field to commutative algebras over that field.
For a long time I've been interested, just as a hobby, in understanding why people think the Riemann Hypothesis is connected to the "field with one element". So I was very interested when lambda-rings and the big Witt ring beame important in James Borger's approach to the field with one element:
His idea is that an algebra over the field with one element is a lambda-ring.
Digging into this, I've come to understand fairly well how lambda-rings arise from decategorifying concepts from representation theory. I've always liked representation theory, and @Joe Moeller and @Todd Trimble and I have showed that categories of representations tend to be 2-rigs, and the Grothendieck group of any 2-rig is a lambda-ring.
But I'm less comfortable with how lambda-rings are connected to number theory and Frobenius lifts. Mainly, it seems like a miracle that lambda-rings show up in two different contexts: representation theory in characteristic zero, and number theory in characteristic !
They can't really be different; they must be deeply connected, so I want to understand this.
One way for me to start understanding this is to take the big Witt ring of a ring , which doesn't seem to involve number theory or primes - it's just the cofree lambda-ring on - and see how it's built from -typical lambda rings, one for each prime.
For years I've felt I don't have the intelligence to fully understand the big Witt ring - so I wished there were some sort of "half-Witt ring" to practice on. But it's gradually starting to make sense.
That sounds like fun! Someone suggested to me at some point that I should try and apply some of the stuff I've done with monoids to Lambda-rings, so I'll be keeping an eye on this topic.
What could you do with lambda-rings, speculatively speaking?
David Corfield said:
There was an exchange between you and James Borger back here which touched on Big Witt vectors. Jack Morava and David Ben-Zvi chip in and get to Adams operations. Maybe a conversation worth mining.
Yes! One difficulty I've had is connecting the 'big picture' ideas to the nitty-gritty comutations with lambda-rings and the big Witt ring. So, I've been digging into the nitty-gritty and now maybe I'll be better able to connect it to the big picture.
There's a nice connection between Adams operations and Frobenius operators which is starting to make sense to me.
I haven't thought about this in a bit but I remember that the explanation in the nlab was very useful to understand why lambda rings are interesting
John Baez said:
What could you do with lambda-rings, speculatively speaking?
It's been just long enough that I can't remember; the general idea was to treat them as internal somethings in a topos of suitably chosen monoid actions, but that much is not exactly profound.
Josselin Poiret said:
the explanation in the nlab was very useful to understand why lambda rings are interesting
Yes, thanks! It seems to have improved since I last read it, or maybe I just know more now. This is about what Borger calls the heterodox interpretation of a lambda-ring as a ring with a family of commuting Frobenius lifts, one for each prime.
Equivalently it can be seen as a ring with with commuting [[p-derivations]] - a generalization of the concept of derivation. This allows us to develop a number-theoretic generalization of the concept of Taylor series, building on the known analogy between the ring of formal power series and the ring of p-adic integers, which gives a geometrical interpretation of localizing at a prime. If we go down this fascinating road, it's good to treat a lambda-ring as a [[Joyal delta-ring]].
But I understand a lot more about what Borger calls the orthodox interpretation of lambda-rings, which is a ring equipped with -operations. The idea here is that if we have a category with well-behaved "exterior power" operations, like a category of vector bundles or group representations, these will endow it's Grothendieck ring with -operations making it into a lambda-ring.
I find this to be a great intro to the 'heterodox' interpretation of lambda-rings and the big Witt ring:
Ignore the scary title and read the nice review material!
Let me get back to you later.
Todd and I were talking about how multiplication in the big Witt ring of a commutative ring arises from comultiplication in the biring of symmetric functions , via the formula
Recall that as a commutative ring,
which we could conceivably write as
to emphasizing is the free -ring on one generator . In fact Todd uses such a notation for the free -ring on two generators and , which is .
Since comulltiplication
is a ring homomorphism, it will be determined once we know what it does to each of the generators . I think Todd showed me how it works for and I asked how it worked for , or something like that. Here was his reply - we've decided to talk about this here.
You asked what does the comultiplication
applied to look like. Short answer:
where we think of as the free lambda-ring on two generators . This can also be written more nicely as .
Longer answer: use the splitting principle, which guarantees that the 2-rig map , sending the generator to a sum of independent bosonic subline objects, is an extension when restricted to polynomial functors of degree . Since is degree , this means in effect that we can pretend the generator of is a sum of two bosonic sublines. Then the 2-rig comultiplication , taking to per our first paper, induces the map of lambda-rings that takes to . Since this lambda-ring map preserves the operation, we calculate
and use the exponential law for plus vanishing of to write this out long-hand. Sparing some gory details, this gives the answer in the short identity.
I wonder how much of this needs to be explained further?
To my taste, the symbols and refer primarily to certain functors, and secondarily to symmetric functions; ultimately these two points of view are united via the all-powerful splitting principle. More precisely, the ring is the Grothendieck group or ring based on isomorphism classes of Schur functors, which are essentially those endofunctors on the category of vector spaces (let's say over ) that you can build up using (together with its symmetric monoidal structure), coproducts, and splitting of idempotents. For example, is a Schur functor. If denotes the symmetry that swaps tensor factors, then the endomorphisms
are idempotent maps, and we can split these idempotents to obtain the symmetric square and exterior square . Likewise, the symmetric power and exterior power functors, and , are Schur functors. Tensor products and coproducts of Schur functors are again Schur functors, and Schur functors are closed under composition.
The symbols and can be read as the isomorphism classes and , regarded as elements of the ring (whose addition is induced from taking coproducts, and whose multiplication is induced from taking tensor products). It turns out -- and this is by no means trivial -- that as a ring, is isomorphic to the polynomial algebra . It is also the polynomial algebra : both sets and serve as polynomial bases. There are other famous bases as well, which I won't mention right now, but you can read about them in famous texts such as Representation Theory by Fulton and Harris, and Symmetric Functions and Hall Polynomials by MacDonald.
But there is so much more to ! People go gaga over the richness of its structure and its interplay with the rest of mathematics. I'll try to indicate some main features of this structure by first pointing to similar features of a far simpler structure, namely the polynomial algebra , which represents the forgetful functor
in the sense of a natural isomorphism . Now, we can remind the forgetful functor of its (tautological) ring object structure, by pointing at natural transformations (whose components are the addition functions ) and (multiplication). At the level of the representing object , these transformations are induced by ring homomorphisms
Here we should pause to note that is the coproduct of commutative rings , which means, by the universal property of coproducts, that . Thus is the representing object of .
In any category with coproducts, we can define a notion of co-ring object, dual to the notion of a ring object in a category with products. So what we can say is that according to the above, is a co-ring object in the category of rings. We call it a biring. I'll leave it as a semi-advanced exercise in applying the Yoneda lemma to figure out explicit formulas for the co-ring structure on that we are talking about here.
Stepping back a little: any time a representable functor has ring object structure -- or in other words lifts up through the forgetful functor to give an endofunctor , the representing object becomes a biring. It can be shown that such lifts of representable functors are necessarily limit-preserving, or better yet they are right adjoints. Thus right adjoint endofunctors on are equivalent to biring structures. The case of the biring corresponds to the identity endofunctor on (which of course is a right adjoint).
But wait, there's more! In our example, carries another binary operation , namely polynomial composition (replacing the in by ). If you'll allow me to abuse language and write the identity functor as
for the biring sitting in the contravariant slot, then the polynomial composition can be read as contravariantly inducing a transformation in the other direction,
(these manipulations might seem a tad puzzling, and indeed it takes some fancy footwork to put it just right, but I'm going to skip over that -- you can read Schur Functors and Categorified Plethysm for details). I'll just say that the polynomial composition on corresponds to the identity transformation that is the comultiplication for the tautological comonad structure on .
More generally, if is a biring, and if there is a comonad structure on the lifted endofunctor , then the comultiplication transfers over to an operation (a function) that behaves similarly to polynomial composition. This operation is called plethysm, and a biring equipped with a plethysm operation is called a plethory. So in summary, giving a plethory is equivalent to giving a right adjoint comonad on ; the plethory is the representing object for that comonad.
So now I can tell you that the main structural features of is that it carries a plethory structure. A rather complicated plethory structure that is still very far away from being fully understood.
Where does that plethory structure come from? Well, John mentioned at the top of this thread that the forgetful functor from the category of lambda-rings to the category of commutative rings has both a left adjoint (expected for abstract nonsense reasons) and a right adjoint (a much more specialized circumstance). So and also are right adjoints, hence the comonad is also a right adjoint. The plethory is the representing object for that right adjoint comonad.
This likely still sounds very mysterious, because for example, what are lambda-rings? Instead of saying directly what they are, I'll say in a different way where the comonad comes from. In our paper Schur Functors and Categorified Plethysm, we explain that it is the result of decategorifying a 2-comonad that is much easier (conceptually) to understand. Namely, what we do is categorify the story I was telling, about how with its plethory structure represents the identity comonad on commutative rings. In this particular categorification, commutative rings are replaced by 2-rigs, which are symmetric monoidal -enriched categories which have coproducts and idempotent splittings. The forgetful functor is replaced by the forgetful (2-)functor . This forgetful functor is representable, by a category (rather, 2-rig) of abstract Schur functors, so this category is the replacement of at the categorified level, and it is a 2-birig that represents the identity functor on the 2-category of 2-rigs. It carries a 2-plethory structure, corresponding to the tautological 2-comonad structure on this identity functor.
The air up here on this categorified mountaintop is clean and clear; the view is simple and beautiful. But then we descend, from down to its set of isomorphism classes, or really I mean its Grothendieck ring . That descent process is called 'decategorification'. And it's a little tricky. It took us months of study to make sure of our footing and the right path down to the valley. After all, we start with the "easiest" 2-comonad in the world, the identity functor on 2-rigs, and somehow this decategorifies down to an extremely non-trivial comonad on commutative rings, namely this right adjoint comonad I mentioned. Then we can (and do) define lambda-rings to be the coalgebras of this comonad.
Okay, I've just given a thumbnail sketch of our first paper.
To return to the topic, though: John quoted my reply to a question he asked about a specific calculation, about the biring . As I said, , even just as a biring let alone a plethory, is a pretty complicated beast, and I believe not completely grasped in terms of giving explicit formulas for the comultiplication (co-addition is rather easier to deal with). But some small calculations you can do by hand, and I was describing to him how to go about calculating what the comultiplication does to the element , by exploiting the subject of our second paper, the splitting principle [set in the context of 2-rigs].
All this is related to the second and third posts at the top of this thread, where John is again in search of ways to wrap one's head around the comultiplication, and he was quoting some stuff he saw in a paper by Niranjan Ramachandran which gives some hints. I can come around to discussing that as well, but in order for others to be able to follow along, I thought it would help to give some background, hence this string of ear-bending comments.
Thanks, Todd! I was distracted last night and didn't quite know the best way to kick off the conversation.
I just noticed a small notational point that might puzzle and perturb novices:
Todd Trimble said:
To my taste, the symbols and refer primarily to certain functors, and secondarily to symmetric functions; ultimately these two points of view are united via the all-powerful splitting principle. [...] If denotes the symmetry that swaps tensor factors, then the endomorphisms
are idempotent maps, and we can split these idempotents to obtain the symmetric square and exterior square . Likewise, the symmetric power and exterior power functors, and are Schur functors.
This point is that the symmetric functions are not powers of the "switch" map - instead there is a map sending Schur functors to symmetric functions, and the are the symmetric functions corresponding to the "ith symmetrized tensor power" functors .
In the study of the symmetric group, people like lots of things named , or , or , or , and sometimes we get carried away and the notations conflict!
Yeah, I edited to change notation in that comment from (which is one tradition) to , which we seem to favor.
Also, I did not explain how Schur functors give rise to symmetric functions. I wouldn't think it was completely common knowledge, how that goes. But I mentioned that the connection arises through this splitting principle that keeps coming up.
I also have another slightly less picayune comment. If we went ahead and computed , this formula:
would become the 2nd term in a sequence of similar expressions: polynomials in the where the total degree of the nth polyomial is 2n, it seems, if we count as having degree .
Are these polynomials we should recognize? Are they famous?
Mm, I imagine they're quite famous, and also pretty intensively studied, but my impression is that we ("we" being the community of mathematicians, including the experts) don't know them all yet. I wish I knew what people called them by name. Atiyah and Tall give the lackluster notation to these polynomials (page 258), and I think that notation might be pretty common. But your posts at the top of the thread are all about these polynomials!
So for example, the Encyclopedia of Mathematics uses the same notation .
I've been scanning Hazewinkel for notation, and I guess it's true that the formulas can be made much more explicit if you use a different polynomial basis, like the power sums basis. He gives a number of such formulas around page 46. So maybe I'll eat my words a little, but anyway I don't know if the explicit list of polynomials purely in terms of the are completely known. Maybe someone else can say.
(I suppose I should know Hazewinkel's article a lot better than I do, because there's evidently a lot of great stuff in it. I'm a little put off by his crapola typesetting job, but I guess that'll be more on me than on him.)
And now that I'm looking at Hazewinkel further, I see that he brings up quasisymmetric functions in section 11, which are supposed to be really important in this biz. Joachim Kock gave an interesting talk about quasisymmetric functions at the CT 2024 conference. I should be looking more into these things.
Do you have any hint as to why quasisymmetric functions are important? I don't know any conceptual explanation of them, so I sometimes have cynically wondered if it's a case of "what are you going to do when you're an expert on symmetric functions and you run out of good ideas? Invent quasisymmetric functions and generalize all the theorems to those!" It's probably not true.
(Lusztig once said "Some people like to take theorems about groups and generalize them to quantum groups. I like to find theorems about quantum groups that aren't like theorems about groups." He came up with some amazing results....)
Eh, I don't yet. But Hazewinkel says at the beginning of section 11, "When looking at various universality properties of the Witt vectors and Symm (which is the topic of the next section) one rapidly stumbles over a (maximally) non commutative version, NSymm, and a (maximally) non cocommutative version, QSymm. This section is devoted to a brief discussion of these two objects. Somehow a good many things become easier to see and to formulate in these contexts (including certain explicit calculations). As I have said before, e.g. in [200], p. 56; [199], Ch. H1, p. 1, once one has found the right non commutative version, things frequently become more transparent, easier to understand, and much more elegant.
There's a certain amount of hardcore algebraic combinatorics to all this. Another buzzphrase that seems relevant and important to me here is this so-called Cauchy identity; see around pages 457-458 of Fulton and Harris. (Perhaps I'm jotting this down as a reminder just to myself to come back to it -- it would be very boring to the casual onlooker.)
Thanks - I'll try to find out if anyone has a name for these polynomials . That could unlock a lot of wisdom - or at least piles of cryptic and unmotivated identities that we might find conceptual explanations for. :upside_down:
For other people listening in, let me try to give a conceptual explanation of these polynomials based on algebraic topology. Todd already knows this, at least implicitly, but I feel like saying it.
The operations of direct sum and tensor product can be applied to matrices, and in particular to unitary matrices, so if is the group of n unitary matrices then we get Lie group homomorphisms
We can work with all simultaneously if we use the obvious inclusions
to define the colimit of topological (or even 'smooth') groups
Then we get binary operations which are group homomorphisms
in addition to the group operation, which is another binary operation. (By the way, I believe books on K-theory use an Eckmann-Hiltonesque argument to show that is homotopy equivalent to the group operation, and even better.)
These maps induce maps on the classifying space for stable complex vector bundles, :
and thus we get maps on K-theory going backward, which we can call coaddition and comultiplication:
Since of any space also has a ring structure, these wind up making into a 'biring'. But this biring is just our friend the free -ring on on generator, which @Todd Trimble has been explaining. This is called . As commutative rings we have
Here , thought of as a symmetric function, is the ith elementary symmetric polynomial. But
and thought of as an element of the integral cohomology of , is called the ith Chern class. It's a cohomology class of degree 2i.
So now let's think about what comultiplication
does to , in these terms!
I'll use the fact I hinted at: for a paracompact Hausdorff space , the set of homotopy classes is isomorphic to the set of stable complex vector bundles over : that is, equivalence classes of vector bundles over , where we count two as equivalent if they become isomorphic after summing with the same complex vector bundle.
Using this and the definition of K-theory we get
After all is defined to be the set of stable complex vector bundles over , made into a commutative ring using and .
Thus, is a somewhat self-referential entity: it's the commutative ring of stable vector bundles on . It's a ring because of the operations acting on the covariant argument in , and a coring because of these operations acting on the contravariant argument in
I'm sort of meandering, but from all this we get yet another interpretation of the elements
Namely, they must come from stable vector bundles on !
I should describe these stable vector bundles, but I won't now. Instead, I just want to say what the
does to each element .
It works like this: we can take any stable vector bundle on and pull it back along the tensor product map
and get a stable vector bundle on . This induces a map
and this is just our friend the comultiplication .
So what does mean? We start with the stable vector bundle on , and pull it back along . I believe every stable vector bundle on is an integral linear combination of tensor products of stable vector bundles , where is the 'external' tensor product of stable vector bundles: if you've got one on some space and one on some space , you can tensor them and get one on .
If so, we should be able to take , pull it back along , and write it in terms of the . And I believe Todd's calculation is an example of this. He wrote
but in my current notation I believe this means
The first term here is not manifestly a tensor product of stable vector bundles , but it actually is: it's
where is the trivial line bundle (the identity in the ring ).
So, in short, the polynomial people call answers this question
Given a stable vector bundle on , and pulling it back along , how can we express it in terms of the stable vector bundles ?
We could also frame this in terms of Chern classes.
All this 'fluffy' stuff doesn't help us compute the polynomials . And indeed, Todd already showed one way to do that. It simply says why we should care about these polynomials.
@Todd Trimble had written to me some more about the big Witt ring, in which he starts analyzing some formulas from here:
In particular, this paper discusses a formula for multiplication in the big Witt ring which I mentioned earlier:
John Baez said:
3) Since one can show the underlying commutative ring of the free lambda-ring on one generator is
we have an isomorphism of sets
where
is mapped to
Then the challenge is to describe the addition and multiplication on in these terms. People often do this with explicit formulas, which I tend to find cryptic. The addition in corresponds to the multiplication in , which is why we use this description. The multiplication in is more tricky. For any we get an element of called , defined by
and the multiplication on turns out to be determined once we know
This formula turns out to be very useful but I don't have a good understanding of how it comes from 2).
The first question is: why should we care about these elements ? And the second is: what does the above formula for a product of them mean? And the third is: does it really determine the product on all of the big Witt ring ?
Todd was making progress on all these questions.
Yes, sorry, I was going to say something about that! But let me collect my thoughts.
The paper of Ramachandran that John linked to mentioned that there are several reasonable choices for the (big) Witt ring multiplication. This has a lot to do with how there are various reasonable choices for a nice polynomial basis of .
So going back to the message at the top of the thread: one conceptual way to define is that it is the hom-set . Thanks to the rich plethory structure on that I was sketching earlier, the hom-set acquires a commutative ring structure and even a lambda-ring structure, and indeed furnishes the right adjoint to the forgetful functor from lambda-rings to commutative rings, as John mentioned earlier.
Now it seems that a lot of sources introduce as consisting of formal power series with constant coefficient , i.e., elements in . So there is an isomorphism , and John showed how this might go: using the polynomial basis , we can define this isomorphism as sending to .
Another possibility is to use the polynomial basis , and define the isomorphism so as to send to .
Each of these has its uses. But before tackling what any of this has to do with those formulas John mentioned, it might not be a bad exercise to look for a moment at how addition works in the Witt ring . (No, it is not ordinary addition of power series!) It turns out that the same formula will work whether you use the basis or the basis, and it's based on the co-addition on that was lightly alluded to. Maybe I'll pause a moment.
This sounds like a good plan! Please pause all night long if you want... I'm about to have dinner, and it's 3 hours later for you.
In fact, I can take this opportunity to go a smidge deeper into our first paper. First, how does co-addition work, from first principles? I plan to be very methodical about this, which might make it look heavy in places -- I'll try to ameliorate that by surrounding some of the main conclusions by extra white space, so that readers can skip ahead to get the main points.
The plan is to see how co-addition works in the toy example of , and then categorify that. In discussion above I said that the explicit formula for co-addition can be derived as a "semi-advanced exercise" in using the Yoneda lemma, so I'll start with that. I'll use the notation to denote hom-sets (usually hom-sets that acquire extra structure). The Yoneda lemma is about representable functors. Here we have where is the forgetful functor; evaluated at a ring , the isomorphism takes to . Similarly we have , instantiated by
where the first product projection is induced by the first coproduct coprojection , and is induced by the second coproduct coprojection . Taking , chase the identity element through the sequence
a la the proof of the Yoneda lemma. We get
In other words, the co-addition is the unique ring map taking to .
The same type of calculation shows that the comultiplication is the unique map taking to .
(One could simply guess these formulas and check that they work, but I think it's nice to know how the Yoneda lemma removes any guesswork.)
If one uses , in effect identifying with and with , then the co-addition becomes simply the ring map taking to , which makes everything look simple and obvious. The comultiplication takes to .
Moving in the opposite direction, suppose given a (commutative, cocommutative) bi-ring . The addition on is retrieved from the co-addition as a composite
where the isomorphism obtains by the universal property of as a coproduct. To be explicit, this isomorphism takes a pair of homomorphisms to the composite
where the codiagonal is precisely the multiplication .
Hence the composite takes a pair of morphisms to the composite
and this defines in the ring .
Replacing the co-addition by comultiplication , the same construction as in produces the product in .
Now categorify all this. Replace , the free commutative ring on one generator, with the free 2-rig on one generator, which is the (additive) Cauchy completion of the -linearization of the free symmetric monoidal category on one generator. (To get anywhere interesting in categorifying commutative rings, you should add in some limits/colimits, and Cauchy completeness is a good place to start.) We write this as .
I think I mentioned before that this is the representing 2-rig for the forgetful functor , and on those grounds alone, through abstract nonsense, one can copy over (or categorify) the development above for , to derive a 2-birig structure on , with a categorified co-addition given by the unique (up to isomorphism) 2-rig map
(where the codomain is the free 2-rig on two generators ) that sends the generator of to the formal coproduct in .
It is interesting to watch what does to objects like and in . I'll start with , the symmetric power. For any 2-rig , there is a 2-rig of graded -objects (whose symmetric monoidal tensor is given by Day convolution, coming from ), and one way we can view the symmetric power for an object of , which I sometimes like to write as , is that it is the homogeneous component of a symmetric algebra construction, which I will write as
Here I'm thinking of the object as sitting in grade in , so that sits in grade . This is the free commutative monoid on , and the category of commutative monoid objects has as its coproduct. This free functor gives a (partially defined) left adjoint [I say "partial" because if for example is in degree , then so would be , but maybe the 2-rig we started with doesn't have infinite coproducts like -- there's a better chance of success if the coproduct is spread across grades]. Since left adjoints preserve coproducts, we deduce an isomorphism
Whereupon, focusing on one grade at a time, we further deduce
,
an identity which holds as a natural transformation between Schur functors of type . This holds in particular when .
Decategorifying this "identity" by taking isomorphism classes, i.e., by taking Grothendieck rings, this implies that the induced co-addition satisfies
where the right side is visibly a convolution product.
Consider now the induced addition on ( a commutative ring), sending a pair of homomorphisms to the homomorphism
This composite takes to . In other words, Witt ring addition is defined by
and if we set up an isomorphism by , then the induced Witt ring addition on is given by multiplying power series.
It turns out the same is true if we use instead the isomorphism given by . At the beginning of that story, we have the identity
as a natural transformation between Schur functors on any 2-rig . To see this, we replace the 2-rig of graded objects, with its "vanilla" symmetry for objects in grade and in grade , with the more sophisticated symmetry that introduces a sign factor, . Otherwise, the entire story (symmetric algebra construction as free commutative monoid, etc.) remains the same, mutatis mutandis, hence we get this "exponential identity". Therefore, Witt addition on formal power series is again multiplication, even if we opt for this other identification with .
Now that I've fully explained how addition on the big Witt ring works, I can turn my attention to how multiplication works. Multiplication of is defined by the composite
where this time we have to figure out how comultiplication on works. It is, of course, gotten by decategorifying (taking isomorphism classes) the unique (up to isomorphism) 2-rig map
that takes the generator to the formal tensor product .
Working this out in detail will involve the "splitting principle" for 2-rigs, which John will be discussing at the upcoming Octoberfest meeting, but perhaps I'll pause to take a break from this longish series of comments.
Todd Trimble said:
Therefore, Witt addition on again corresponds to multiplication of formal power series, even if we use this second identification .
That's great! Maybe someday we should write a little exposition of the big Witt ring. (Better than
a big exposition of the little Witt ring, explaining stuff like this. :upside_down:)
It might be fun to see what Witt addition looks like if we use the basis for given by power sum symmetric functions.
Yes, all these seem like good suggestions!
I'm going to push on with this series of posts; the next topic will be this splitting principle that I keep banging on about.
For us, the splitting principle is the idea that to establish isomorphisms between Schur functors, it is permissible to pretend that they can be decomposed as coproducts of "line objects". It is analogous to the splitting principle in K-theory, where equations are verified by acting as if vector bundles split as coproducts of line bundles.
This sounds a little flaky perhaps, so I'll put it more precisely in a moment, but first I want to say that the situation reminds me of how those Italian mathematicians from the late Renaissance who developed the cubic formula -- Cardano, Tartaglia, etc. -- made the bold move to act as if were a thing. Even when all the roots of the cubic polynomial were real, they were arrived at by making use of imaginary elements. (True, they were fairly uncomfortable with the situation, and it took a few centuries before mathematicians felt generally at home with splitting extensions of fields.) The analogy is apt: the coefficients of a polynomial are symmetric functions of their roots, and the roots are analogous to the line objects we are about to discuss. Don't believe the analogy? That's okay. Humor me anyway by considering a linear transformation on a vector space. By passing to an extension of the scalar field if need be, i.e., adjoining roots, we can split into a coproduct of eigenspaces (which generically are lines), by splitting the characteristic polynomial into linear factors.
(At the risk of too much self-indulgence: this also reminds me of splitting light into a spectrum, where the energy levels of photons are given by eigenvalues of a suitable operator.)
Okay, now let me state the splitting principle more precisely. For the purposes of this thread, define a line object in a 2-rig to be a (nonzero) object satisfying . (In our paper, we say instead "bosonic subline object".) Another way of saying it is that the canonical quotient is an isomorphism, or equivalently that the symmetry that transposes factors equals the identity. Or equivalently still, that the tautological action of on is trivial for all . Finally, this last condition is equivalent to saying that for the symmetric powers , we have for all .
Just as with its generator is initial among 2-rigs equipped with an object, so (the linear Cauchy completion of the linearized discrete symmetric monoidal category ) is initial among 2-rigs equipped with a line object. Here is equivalent to the category of -graded vector spaces of finite total dimension, which we in our paper denote as , and the line object in this universal case is a 1-dimensional vector space concentrated in grade . Likewise, the walking 2-rig on line objects , denoted as is equivalent to the category of -graded vector spaces of finite total dimension.
Assume the ground field is of characteristic zero. Suppose given Schur objects , also known as polynomial functors, say of degree or less; the latter means they are valued in finite-dimensional vector spaces, and that their restrictions to the symmetric groups are zero for . Thus is given by (let's say right) linear representations of for . The polynomial or Schur functor itself takes a finite-dimensional vector space to . In fact this formula for the Schur functor makes sense for any 2-rig, taking to be an arbitrary object in the 2-rig. In particular, it can be applied to in .
The splitting principle we use concerns properties of the 2-rig map that takes to . In the form that we use it here, it states that the restriction of this 2-rig map to the subcategory of Schur objects or polynomial functors of degree at most is essentially injective, i.e., if
for of degree at most , then . (Actually we prove more: that this restricted functor is faithful and conservative as well. But the essential injectivity property just stated is the one that is really key.)
We can play with this a bit. For example, we can calculate
(the Schur functor is in the case where , the sign representation of ), by exploiting the exponential identity mentioned earlier:
For a line object , for all , as is easily shown by induction (since is a retract of in any 2-rig). Thus the only summands that survive on the right in the last display line are the ones where all the indices are or (and add up to ). Thus
Letting denote the isomorphism class , the Grothendieck ring of is isomorphic to the polynomial ring , and the isomorphism class of the coproduct on the right is
which is precisely the elementary symmetric polynomial as defined by the identity
Similarly, the exponential identity for the symmetric powers yields
and using the fact noted earlier that for line objects we have , this may be rewritten as
The isomorphism class of the expression on the right is
which is precisely , the complete homogeneous polynomial in variables, as defined by the identity
Letting be the Grothendieck group of the category of Schur objects of degree at most , the splitting principle implies that the induced map
taking to , is an injection. Of course is manifestly invariant under permutations of the elements , so that elements in the image of this map are symmetric polynomials in the (of degree no more than ). Recall that every symmetric polynomial in variables is uniquely a polynomial in the elementary symmetric polynomials . Provided that the total degree of is no more than (where the degree of is of course ), we have
Since the map carries to , we deduce that every class is a polynomial of total degree no more than , which is already a nontrivial theorem. It says that every polynomial functor of degree less than or equal to is isomorphic to a suitable coproduct of tensor powers of exterior power functors.
We are beginning to come full circle. Passing to suitable limits, explained in our paper, we can summarize one of our consequences of the splitting principle by the intuitive formula
so that we are splitting an "infinite polynomial" into linear factors, a la splitting extensions in the sense of Galois theory, and the coefficients are thereby manifestly identified with symmetric functions in the , which play a role similar to roots in the splitting extension.
Similarly, we may write
This goes at least some distance towards answers to some of John's questions at the top of this thread, although we still have some ways to go. I'm going to take a break for the moment.
I will take this opportunity to digress a bit and ponder Todd's suggestion that this formula arising from the splitting principle
amounts to splitting "infinite polynomial" into linear factors. This analogy seems extremely strong. We can take any monic polynomial of degree , factor it as
and then write the coefficients of this polynomial as
where is a degree polynomial in . If I'm doing it right $e_i$$ is the ith elementary symmetric polynomial in variables. E.g.
and so on.
But the , thought of as symmetric functions, are very similar! They're the elementary symmetric functions, which are essentially elementary symmetric polynomials in infinitely many variables.
What exactly is the relation between what I just said and what you explained about
?
There are some differences in convention, e.g. I was talking about monic polynomials while you're looking at a 'comonic' power series (where the coefficient of the constant term is 1), and correspondingly I've got factors of while you've got factors of . Those can presumably be fiddled so things match up better. But what is this business about factoring a formal power series into linear factors?
In my case I think there's really a Galois extension lurking around: if we treat the as formal variables, the field generated by the roots of the polynomial is an extension of the field generated by the coefficients, and the Galois group is . But in your case we seem to be doing a field extension that's not Galois, whose 'Galois group' (group of automorphisms over the base field) is something like .
I think what you wrote, , and what I wrote at 49 past the hour, , are basically the same thing. "My " is times your , is the way I would match things up.
As I've mentioned to you and Joe in private conversation, I very much have in mind that either way, is a splitting field extension of its subfield , with Galois group . When the textbooks talk about unsolvability of the quintic and so on, that's really the formal framework of what they're discussing (maybe replace here by any field). So we start with indeterminates that have no special meaning attached to them, and pass to the splitting field of (alternate those terms if you like) and basically wind up with which is abstractly isomorphic to the original field, but which (now thinking geometrically) fibers over it differently, with fibers given by -torsors.
[I'll say again that I perceive a kind of unity between the various uses of the word "splitting" (splitting a polynomial, splitting into eigenspaces, splitting into line bundles, etc., even splitting field in the sense of representation theory, although I would need to recall the story of why I thought that's similar).]
But anyway, this fibering reminds me of configuration spaces, how -tuples of distinct points in , say, fiber over -element subsets of . Jim D. sometimes talks about this sort of thing, too.
In my write-up here, I'm thinking of as a ring filtered by the , but one can argue that really what we should be thinking of is that the quotient ring is . This would correspond to a lambda-ring generated by Young diagrams with rows or fewer, whereas the filtration component is not a ring, and corresponds to Young diagrams with boxes or fewer.
We didn't quite get to a full explanation of the picture in our paper, which we certainly hold (without proving this) to be the Grothendieck ring of the 2-rig of algebraic representations of the multiplicative monoid where is an -dimensional vector space. Same as the monoid of matrices. We denote this 2-rig of algebraic representations as
and the "splitting principle pretense" of acting as if matrices can be diagonalized, split into 1-dimensional eigenspaces, is formalized by considering the pullback functor
where is the multiplicative submonoid of diagonal matrices, and showing that this pullback functor satisfies our trio of conditions: faithful, conservative, and essentially injective. (For the readers out there, this is one of our main results.) We also observe that is equivalent to , the walking 2-rig on line objects.
Anyway, what I am leading up to here is that maybe in some ways it's better to think of not as a colimit or union of its filtered pieces , as we do in the paper, but as a limit of 2-rig quotients . Not a limit in a "naive" 2-rig sense, but in a graded 2-rig sense. This is analogous to how we typically treat the cohomology algebra : not as an inverse limit of in the category of rings (now interpreting the as Chern classes!), but as an inverse limit in the category of graded rings, leading to the polynomial algebra in infinitely many variables.
I'm thinking that this inverse limit perspective on , placing it in the same neighborhood as how we treat as an inverse limit in a graded sense, might lead to a more harmonious picture of what is going on at the level of Galois groups. For example, it might clarify whether we are thinking of this as the full permutation group, or just the union of the (I'm thinking the former is more appropriate).
All this is really exciting. It's good to sort out what are limits and what are quotients. As far as the 'splitting fields' go, if we want homomorphisms between them, the maps clearly must go this way:
We have similar homomorphisms for the fields these are extending
and it seems all this induces inclusions of Galois groups
Right, if you're using fields, you have to go this direction (maybe you want to use parentheses instead of square brackets for fields). It might be in fact that fields are awkward. Maybe there's an okay sense of speaking of as a "Galois extension" of , though, with Galois group ?
The reason I bring this up is that we go the "bad" direction in our paper. Here "bad" means the wrong direction if we consider fields of fractions. (Of course, the fields of fractions construction is not functorial. I guess it is functorial however on the category of integral domains and injective maps between them.)
Back in week261 of This Week's Finds, I wrote a lot about this stuff in the special case n = 3. I explained how this is related to thinking of as a Coxeter group, and I hint at generalizations of the theory of symmetric polynomials to other Dynkin diagrams:
Imagine we're trying to solve a cubic equation. We can always divide by the coefficient of the cubic term, so it's enough to consider equations like this:
z^3 + Az^2 + Bz + C = 0
If we could solve this and find the roots a, b, and c, we could
write it as:
(z - a)(z - b)(z - c) = 0
But this means
A = -(a + b + c)
B = ab + bc + ca
C = -abc
Note that A, B, and C don't change when we permute a, b, and c. So, they're called "symmetric polynomials" in the variables a, b, and c.
You see this directly, but there's also a better explanation: the coefficients of a polynomial depend on its roots, but they don't change when we permute the roots.
I can't resist mentioning a cool fact, which is deeply related to the trefoil: every symmetric polynomial of a, b, and c can be written as a polynomial in A, B, and C - and in a unique way!
In fact, this sort of thing works not just for cubics, but for polynomials of any degree. Take a general polynomial of degree n and write the coefficients as functions of the roots. Then these functions are symmetric polynomials, and every symmetric polynomial in n variables can be written as a polynomial of these - and in a unique way.
But, back to our cubic. Note that -A/3 is the average of the three roots. So, if we slide z over like this:
x = z + A/3
we get a new cubic equation for which the average of the three roots is zero. This new cubic equation will be of this form:
x^3 + Bx + C = 0
for some new numbers B and C. In other words, the "A" in this new cubic is zero, since we translated the roots to make their average zero.
So, to solve cubic equations, it's enough to solve cubics like x^3 + Bx + C = 0. This is a great simplification. When you first see it, it's really exciting. But then you realize you have no idea what to do next! This must be why it's called a "depressed cubic".
In fact, Scipione del Ferro figured out how to solve the "depressed cubic" shortly after 1500. So, you might think he could solve any cubic. But, NEGATIVE NUMBERS HADN'T BEEN INVENTED YET. This prevented him from reducing any cubic to a depressed one!
It's sort of hilarious that Ferro was solving cubic equations before negative numbers were worked out. It should serve as a lesson: we mathematicians often work on fancy stuff before understanding the basics. Often that's why math seems hard! But often it's impossible to discover the basics except by working on fancy stuff and getting stuck.
Here's one trick for solving the depressed cubic x^3 + Bx + C = 0. Write
x = y - B/(3y)
Plugging this in the cubic, you'll get a quadratic equation in y^3, which you can solve. From this you can figure out y, and then x.
Alas, I have no idea what this trick means. Does anyone know? Ferro and Tartaglia used a more long-winded method that seems just as sneaky. Later Lagrange solved the cubic yet another way. I like his way because it contains strong hints of Galois theory.
You can see all these methods here:
6) Cubic function, Wikipedia,
http://en.wikipedia.org/wiki/Cubic_equation
So, I won't say more about solving the cubic now. Instead, I want to explain the "discriminant". This is a trick for telling when two roots of our cubic are equal. It turns out to be related to the trefoil knot.
For a quadratic equation ax^2 + bx + c = 0, the two roots are equal precisely when b^2 - 4ac = 0. That's why b^2 - 4ac is called the "discriminant" of the quadratic. The same idea works for other equations; let's see how it goes for the cubic.
Suppose we were smart enough to find the roots of our cubic
x^3 + Bx + C = 0
and write it as
(x - a)(x - b)(x - c) = 0
Then two roots are equal precisely when
(a - b)(b - c)(c - a) = 0
The left side isn't a symmetric polynomial in a, b, and c; it changes sign whenever we switch two of these variables. But if we square it, we get a symmetric polynomial that does the same job:
D = (a - b)^2 (b - c)^2 (c - a)^2
This is the discriminant of the cubic! By what I said about symmetric polynomials, it has to be a polynomial in B and C (since A = 0). If you sweat a while, you'll see
D = -4B^3 - 27C^2
So, here's the grand picture: we've got a 2-dimensional space of cubics with coordinates B and C. Sitting inside this 2d space is a curve consisting of "degenerate" cubics - cubics with two roots the same. This curve is called the "discriminant locus", since it's where the discriminant vanishes:
4B^3 + 27C^2 = 0
If we only consider the case where B and C are real, the discriminant
locus looks like this:
|C
o |
o |
o |
-----------o-------------
o | B
o |
o |
|
It's smooth except at the origin, where it has a sharp point called a "cusp".
Now here's where the trefoil knot comes in. The equation for the discriminant locus:
4B^3 + 27C^2 = 0
should remind you of the equation for the trefoil:
u^2 = v^3
Indeed, after a linear change of variables they're the same! But, for the trefoil we need u and v to be complex numbers. We took them to be unit complex numbers, in fact.
So, the story is this: we've got a 2-dimensional complex space of complex cubics. Sitting inside it is a complex curve, the discriminant locus. In our new variables, it's this:
u^2 = v^3
If we intersect this discriminant locus with the torus
|u| = |v| = 1
we get a trefoil knot. But that's not all!
Normal folks think of knots as living in ordinary 3d space, but topologists often think of them as living in a 3-sphere: a sphere in 4d space. That's good for us. We can take this 4d space to be our 2d complex space of complex cubics! We can pick out spheres in this space by equations like this:
|u|^2 + |v|^3 = c (c > 0)
These are not round 3-spheres, thanks to that annoying third power. But, they're topologically 3-spheres. If we take any one of them and intersect it with our discriminant locus, we get a trefoil knot!
This is clear when c = 2, since then we have
|u|^2 + |v|^3 = 2
and
u^2 = v^3
which together imply
|u| = |v| = 1
But if you think about it, we also get a trefoil knot for any other c > 0. This trefoil shrinks as c -> 0, and at c = 0 it reduces to a single point, which is also the cusp here:
|u
| o
| o
| o
-----------o-------------
| o v
| o
| o
|
We don't see trefoil knots in this picture because it's just a real 2d slice of the complex 2d picture. But, they're lurking in the background!
Now let me say how the group of permutations of three things gets into the game. We've already seen the three things: they're the roots a, b, and c of our depressed cubic! So, they're three points on the complex plane that add to zero. Being a physicist at heart, I sometimes imagine them as three equal-mass planets, whose center of mass is at the origin.
The space of possible positions of these planets is a 2d complex vector space, since we can use any two of their positions as coordinates and define the third using the relation
a + b + c = 0
So, there are three coordinate systems we can use: the (a,b) system, the (b,c) system and the (c,a) system. We can draw all three coordinate systems at once like this:
b
\ /
\ /
\ /
\ /
--------o--------a
/ \
/ \
/ \
/ \
c
The group of permutations of 3 things acts on this picture by permuting the three axes. Beware: I've only drawn a 2-dimensional real vector space here, just a slice of the full 2d complex space.
Now suppose we take this 2d complex space and mod out by the permutation symmetries. What do we get? It turns out we get another 2d complex vector space! In this new space, the three coordinate axes shown above become just one thing... but this thing is a curve, like this:
o
o
o
o
o
o
o
Look familiar? Sure! It's just the discriminant locus we've seen before.
Why does it work this way? The explanation is sitting before us. We've got two 2d complex vector spaces: the space of possible ordered triples of roots of a depressed cubic, and the space of possible coefficients. There's a map from the first space to the second, since the coefficients are functions of the roots:
B = ab + bc + ca
C = -abc
These functions are symmetric polynomials: they don't change when we permute a, b, and c. And, it follows from what I said earlier that we can get any symmetric polynomial as a function of these - under the assumption that a+b+c = 0, that is.
So, the map where we mod out by permutation symmetries of the roots is exactly the map from roots to coefficients.
The lines in this picture are places where two roots are equal:
c=a
\ /
\ /
\ /
\ /
--------o-------- b=c
/ \
/ \
/ \
/ \
a=b
So, when we apply the map from roots to coefficients, these lines get mapped to the discriminant locus:
|
o |
o |
o |
-----------o-------------
o |
o |
o |
|
You should now feel happy and quit reading... unless you know a bit of topology. If you do know a little topology, here's a nice spinoff of what we've done. Though I didn't say it using so much jargon, we've already seen that space of nondegenerate depressed cubics is C^2 minus a cone on the trefoil knot. So, the fundamental group of this space is the same as the fundamental group of S^3 minus a trefoil knot. This is a famous group: it has three generators x,y,z, and three relations saying that:
x conjugated by y is z
y conjugated by z is x
z conjugated by x is y
On the other hand, we've seen this space is the space of triples of distinct points in the plane, centered at the origin, mod permutations. The condition "centered at the origin" doesn't affect the fundamental group. So, this fundamental group is another famous group: the "braid group on 3 strands". This has two generators:
\ / |
/ | X
/ \ |
and
| \ /
| / Y
| / \
and one relation, called the "Yang-Baxter equation" or "third Reidemeister move":
\ / | | \ /
/ | | /
/ \ | | / \
| \ / \ / |
| / = / | XYX = YXY
| / \ / \ |
\ / | | \ /
/ | | /
/ \ | | / \
So: the 3-strand braid group is isomorphic to the fundamental group of the complement of the trefoil! You may enjoy checking this algebraically, using generators and relations, and then figuring out how this algebraic proof relates to the geometrical proof.
I find all this stuff very pretty...
... but what's really magnificent is that most of it generalizes to any Dynkin diagram, or even any Coxeter diagram! (See "week62" for those.)
Yes, we've secretly been studying the Coxeter diagram A_2, whose "Coxeter group" is the group of permutations of 3 things, and whose "Weyl chambers" look like this:
\ /
\ /
\ /
\ /
--------o--------
/ \
/ \
/ \
/ \
Let me just sketch how we can generalize this to A_{n-1}. Here the Coxeter group is the group of permutations of n things, which I'll call n!.
Let X be the space of n-tuples of complex numbers summing to 0. X is a complex vector space of dimension n-1. We can think of any point in X as the ordered n-tuple of roots of some depressed polynomial of degree n. Here "depressed" means that the leading coefficient is 1 and the sum of the roots is zero. This condition makes polynomials sad.
The permutation group n! acts on X in an obvious way. The quotient X/n! is isomorphic (as a variety) to another complex vector space of dimension n-1: namely, the space of depressed polynomials of degree n. The quotient map
X -> X/n!
is just the map from roots to coefficients!
Sitting inside X is the set D consisting of n-tuples of roots where two or more roots are equal. D is the union of a bunch of hyperplanes, as we saw in our example:
\ /
\ /
\ /
\ /
--------o--------
/ \
/ \
/ \
/ \
Sitting inside X/n! is the "discriminant locus" D/n!, consisting of degenerate depressed polynomials of degree n - that is, those with two or more roots equal. This is a variety that's smooth except for some sort of "cusp" at the origin:
o
o
o
o
o
o
o
The fundamental group of the complement of the discriminant locus is the braid group on n strands. The reason is that this group describes homotopy classes of ways that n points in the plane can move around and come back to where they were (but possibly permuted). These points are the roots of our polynomial.
On the other hand, the discriminant locus is topologically the cone on some higher-dimensional knot sitting inside the unit sphere in C^{n-1}. So, the fundamental group of the complement of this knot is the braid group on n strands.
This relation between higher-dimensional knots and singularities was investigated by Milnor, not just for the A_n series of Coxeter diagrams but more generally:
7) John W. Milnor, Singular Points of Complex Hypersurfaces, Princeton U. Press, 1969.
The other Coxeter diagrams give generalizations of braid groups called Artin-Brieskorn groups. Algebraically you get them by taking the usual presentations of the Coxeter groups and dropping the relations saying the generators (reflections) square to 1.
Thanks for recalling all this! I've never read this carefully, but now it seems I shall.
This gives a much more elaborated picture (for ) of what I was waving my hands at earlier, when I spoke of configuration spaces of -tuples of distinct points mapping down onto the space of -element subsets (or is the latter called the configuration space?). The fundamental group of the space below is the full braid group, and the fundamental group of the space above is the pure braid group, which is the kernel of the quotient .
Now I really want to find an analogous story for the series of Coxeter groups and their Artin-Brieskorn braid groups. (The Dynkin diagrams have the same Coxeter groups as the , so I'm lumping them together, especially since also has a completely different meaning in what you just wrote!)
The Coxeter group is the symmetry group of the -cube (I hope I'm getting the numbers to match correctly here), or if you prefer, the -dimensional 'orthoplex', the -dimensional generalization of an octahedron. I like to think of it as the full rotation-reflection symmetry group of the coordinate axes in .
This is the symmetry group of pairs of things, where the pairs can be permuted, and the two roots within each pair can be switched, but the roots within each pair are "joined at the hip". It's the wreath product of and .
To get this group as the Galois group of something, maybe I should be looking for a polynomial with roots that come in pairs, something like
Let's see, in the case of the Coxeter group , down below, you localize away from (geometrically, take the complement of) the discriminant locus; up above, you remove some hyperplanes where two coordinates are equated.
So I guess in your proposal, the space above would consist of pairs which are all distinct from each other and their negatives, and also nonzero, and down below, the space of possible coefficients of this polynomial, which again can be given by localizing away from the locus of a suitable variant of a discriminant (so as to forbid any from being a root; I guess it's the usual discriminant times ).
Hmm. I'm confused about a lot of things, but people have written about "symplectic symmetric functions", and I'm hoping these should give a way of working with the cohomology ring just as the usual symmetric functions can be identified with elements of .
Very interesting idea! (Did I say something wrong or confusing in my last message?) Oh well, we can certainly discuss this.
There's a paper that might be relevant:
I haven't read it, but the MathSciNet review says:
The authors extend the use of Young diagram methods from the case of GL(n) to the case of the other classical groups SO(2n+1), Sp(2n) and SO(2n). They establish a number of propositions attesting to the fact that the character rings of rational representations of all the classical groups are polynomial rings over certain irreducible characters associated with totally symmetric and totally antisymmetric irreducible representations. Various determinantal expansions and generating functions for the characters of arbitrary irreducible representations of each of the classical groups [H. Weyl, The classical groups, their invariants and representations, second edition, Princeton Univ. Press, Princeton, N.J., 1946; D. E. Littlewood, The theory of group characters and matrix representation of groups, second edition, Clarendon Press, Oxford, 1950] are discussed and the connection between characters of GL(n) and both Young diagrams and Young tableaux is noted. The authors then introduce the universal character ring, Λ [I. G. Macdonald, Symmetric functions and Hall polynomials, Clarendon Press, New York, 1979; MR0553598], and define bases {χGL(λ)}λ, {χO(λ)}λ and {χSp(λ)}λ from which the characters of irreducible representations of each of the classical groups may be obtained by means of appropriate specialization homomorphisms. The relationship between the bases is established and the branching rules for the restriction from GL(n) to SO(n) and Sp(n) are determined as well as the rules for decomposing the tensor product of representations of both SO(n) and Sp(n).
The results are not new; those not already given in the works of Littlewood [op. cit.] and F. D. Murnaghan[The theory of group representations, Johns Hopkins Press, Baltimore, Md. 1938; Jbuch 64, 964; The unitary and rotation groups.
I'd read in Weyl's book about generalizations of Young diagram techniques for orthogonal and symplectic groups, but for some reason I'd never thought about that stuff in conjunction with symmetric functions. Macdonald's anemic index doesn't contain the words "symplectic" or "orthogonal", so I'm having trouble finding about these various bases. It sounds like he's somehow embedding the rings of orthogonal and symplectic characteristic classes into the ring of characteristic classes for , i.e. the usual ring of symmetric functions. I would like to start by treating them separately, as their own independent rings.
Todd Trimble said:
(Did I say something wrong or confusing in my last message?)
I just don't have a clear idea of what's going on. I guess you're probably right in what you said.
My mind is just a bit blown by the bigger picture we seem to be stumbling on. It gets even bigger if we remember Bott periodicity and thus that looped 4 times gets you , which looped 4 times gets you back to .
Let me now pick up where I left off in the general exposition. I promised to get into some nitty-gritties involving the comultiplication , considering as a biring. Now that we have the splitting principle in hand, we're in a position to do that.
By the way, for anyone reading these posts but who are relatively new to the general topic of lambda rings, I might pose an exercise at this point: what can you say about line objects in ; alternatively, what can you say about elements in such that ? I'll reveal the answer later.
Here's another exercise we can do right now: in any 2-rig, prove that the tensor product of two line objects is also a line object. (Let's drop the nonzero requirement for that one. Actually, in our paper, we don't impose that condition. I only put it in there, parenthetically, as a kind of nod to what is done elsewhere in the literature, but morally it's better to leave it out. That's why in our paper we use the term "subline" by the way, because for example satisfies , but it isn't at all 1-dimensional. Nonetheless, here in this series, I'll keep saying "line" because it's short.)
For this exercise, one should decide which formulation of line object is the most convenient. Our choices are
A line object is one that satisfies .
A line object is one that satisfies .
A line object is one for which the self-symmetry equals the identity.
A line object is one for which acts trivially on , for all .
A line object is one for which for all .
If you chose the third as the most convenient, then I think you chose well, because for one thing you don't need all the infrastructure of a 2-rig to make sense of it; you only use the symmetric monoidal structure. (Recall that in our paper we call these objects "bosonic subline objects", as distinguished from "fermionic subline objects". The terminology is meant to recall the mathematics of supersymmetry, which actually plays an important role, especially in categorified calculations that have anything to do with negatives -- see for example section 7 of our first paper, where we need to transition from a rig-plethory to the ring-plethory .)
Having made this choice, the exercise becomes fairly straightforward. For example, it's easy to give a proof via string diagrams (go ahead and try it!).
That exercise is important for what we do. In the spirit of the splitting principle, given objects in a 2-rig that are finite coproducts of line objects, and , their tensor product is also a coproduct of line bundles,
according to that exercise.
In a moment we'll actually apply the splitting principle to understand comultiplication , but first I want to extend it to say that not only is our canonical (to the category of -graded spaces) essentially injective when restricted to polynomial functors of degree no more than , so are 2-tensor products of such 2-rig maps, i.e.,
is essentially injective when restricted to polynomial functors of degree no more than in and in .
On to comultiplication. Abstractly, we understand . Its effect on isomorphism classes of a Schur object or polynomial functor (polynomial species) is to take it to , where lives in . But now we want to understand how to calculate explicitly this map in terms of a chosen polynomial basis for , say
and for that it suffices to know how to calculate as an element in
I'm not going to do it for all . I just want to explain how I would go about it for small if I were locked in a room with pen and paper and no internet or books. Experts in the area would know how to do it efficiently (as far as the state of the art allows), but that's not the point here; the modest point is simply to understand it.
In fact I'm just going to recall how it goes for (which we will need later anyway), and wave my hands a little at the case for higher . In fact, I can just quote John quoting me:
John Baez said:
You asked what does the comultiplication
applied to look like. Short answer:
where we think of as the free lambda-ring on two generators . This can also be written more nicely as .
Longer answer: use the splitting principle, which guarantees that the 2-rig map , sending the generator to a sum of independent bosonic subline objects, is an extension when restricted to polynomial functors of degree . Since is degree , this means in effect that we can pretend the generator of is a sum of two bosonic sublines. Then the 2-rig comultiplication , taking to per our first paper, induces the map of lambda-rings that takes to . Since this lambda-ring map preserves the operation, we calculate
and use the exponential law for plus vanishing of to write this out long-hand. Sparing some gory details, this gives the answer in the short identity.
I'll just add a few notes to this explanation:
Since is the isomorphism class of a polynomial functor of degree 2, the extended splitting principle guarantees that there is an isomorphism
if there is an isomorphism between these objects as interpreted in under the 2-rig map
that takes to the sum of canonical line objects in the first copy of , and to the sum of canonical line objects in the second copy of .
The comultiplication is a map of -rings because it comes from a 2-rig map (that's the whole point of our first paper, that the Grothendieck ring of a 2-rig is a lambda-ring, and a 2-rig map induces a morphism of lambda-rings).
For generic line objects , we calculate using the exponential law, but we saw earlier that we wind up with the second elementary symmetric polynomial in the :
where I omit symbols, and abbreviate to , to save space. Now interpret as , as , as , and as . The second elementary symmetric polynomial becomes
After some manipulation, one sees that this matches
which is isomorphic to according to the interpretations of as elementary symmetric polynomials and the as complete homogeneous polynomials applied to line objects. (I can't put my finger on it, but something here smells a little like umbral calculus...)
There is an algorithm for writing out a symmetric polynomial as a polynomial in elementary symmetric polynomials; it proceeds by an induction along let's say reverse lexicographic order on the list of exponents in monomial expressions appearing in the given symmetric polynomial. The first such term has , and you subtract off whatever is the correct monomial in the that has the same feature, I think it's
, which [sanity check] has anyway the correct degree:
.
You're left with a symmetric polynomial that starts off with a leading term lower in the well-order.
Anyhow, to sum it all up, what you do to compute the is expand and then write the coefficients as a polynomial in the elementary symmetric polynomials .
Finally: the answer to the exercise: there are no nonzero line objects in !
One might hastily think that the generator itself is surely a line object: as a functor , it vanishes at for every ; at , its value is the 1-dimensional (trivial) representation of . The trouble is that the tensor product on is not the pointwise one (that would be the Hadamard product, as some people say); it's the Day convolution one (aka the Cauchy product). Tensors powers using the Day convolution "spread out", and their retracts are also spread out.
Or, one might think that you can get line objects in by pulling back the canonical line object (in ) along the symmetric monoidal quotient . The trouble is that the resulting pullback functor is only lax monoidal; we require strong monoidality for our 2-rig maps. It's pushforward (i.e., left Kan) along this symmetric monoidal quotient functor that is strong monoidal.
John Baez said:
My mind is just a bit blown by the bigger picture we seem to be stumbling on.
Yes, I get a little sense of the enormity as well.
It gets even bigger if we remember Bott periodicity and thus that looped 4 times gets you , which looped 4 times gets you back to .
Yes, well, I'm not ready to think about that! I suppose if I were trying to return my mind back to real Bott periodicity, I would at first take a Clifford algebra approach, which I probably never properly learned in the first place.
The amount of latent geometry we began to stumble on in the second paper is itself pretty daunting. In the long series of comments I just posted, it seems we're running into the Segre embedding again, induced by
This appeared around the point that we were describing 2-rigs of algebraic representations, but you thought it would be better to remove the words "Segre embedding", so as to not scare off readers. :-)
I said:
Finally: the answer to the exercise: there are no nonzero line objects in !
I lied. The species that takes to and otherwise to , i.e., the monoidal unit, is a nonzero line object. (The monoidal unit in any 2-rig is a line object.) But that's the only one, up to isomorphism.
Todd Trimble said:
John Baez said:
It gets even bigger if we remember Bott periodicity and thus that looped 4 times gets you , which looped 4 times gets you back to .
Yes, well, I'm not ready to think about that! I suppose if I were trying to return my mind back to real Bott periodicity, I would at first take a Clifford algebra approach, which I probably never properly learned in the first place.
I got to know Bott periodicity pretty well blogging and giving talks about the tenfold way, which unifies real and complex Bott periodicity. For some reason I never noticed until now that our friend arises from , which is one of the ten infinite loop spaces in the tenfold way, and that others should give 'mutant' versions of symmetric functions. But I will resist derailing this thread with that.
On with the big Witt ring!
It's now time to look at the big Witt ring , for a commutative ring . This on the other hand has lots of line objects, I mean line elements! But we will need to familiarize ourselves with how its lambda-ring structure works.
There are two abstract ways to define lambda-rings: either as coalgebras of the right adjoint comonad , or as algebras of its left adjoint monad, typically denoted . A standard abuse of language is that may be thought of either as the action constraint of an actegory structure where birings act on rings, or as a monoidal product on the category of birings, but abstractly it is easy to understand: if is a biring, then is a right adjoint, and we define the functor to be its left adjoint. If are two birings, then the composition of right adjoints is again a right adjoint endofunctor on , and therefore is of the form for some new biring . We define to be this biring, so that , naturally in rings . This defines a monoidal (just monoidal, not symmetric monoidal) product on the category of birings, and by abuse of language there is an isomorphism
for birings and a (commutative) ring . A plethory may be defined in at least three ways:
As a biring for which the endofunctor is equipped with a comonad structure;
As a biring for which the left adjoint is equipped with a monad structure;
As a biring equipped with a monoid structure in the monoidal category of birings.
The ur-example, perhaps the original example of a plethory, is , but there is a plethora of plethories. I'll just mention one class of examples quickly. If is a cocommutative -coalgebra, i.e., a cocommutative comonoid in the category of abelian groups under tensor product, then for any ring , the abelian group of additive homomorphisms between underlying abelian groups carries a commutative ring structure, where multiplication of homomorphisms is given by the expected formula
The functor is a right adjoint endofunctor. Since the symmetric -algebra construction is left adjoint to , we have that the right adjoint endofunctor is represented by , which is thereby a biring. But if moreover carries a cocommutative -bialgebra structure, then the multiplication induces a comonad structure on , and in this way becomes a plethory. This type of plethory is called a linear plethory.
Anyway, for technical reasons we chose in our first paper to emphasize the first point of view, defining a lambda-ring as a coalgebra of , or of the big Witt ring comonad . Of course itself would be the cofree lambda-ring cogenerated by . Its coalgebra structure is given by the comonad comultiplication , which is a map of type . If denotes the ordinary hom-set of functions between rings , then sits inside , and sits inside
and it turns out that the comonad comultiplication is a restriction of a map induced by an operation . This map is closely related to "plethystic multiplication", except that one has to be careful to get the order right. Given a pair of isomorphism classes of polynomial functors, the map takes this pair to their composition (aka substitution) as polynomial functors; considered as species, the formula is
where the tensorial exponent refers to the Day convolution on . (See our first paper, top of page 35.) The class is denoted .
We can now say that the comonad comultiplication on the big Witt ring is given by the map
If we want to compute the value for an element of the lambda-ring , this formula tells us what to do: it's the homomorphism defined by
Now, very much in keeping with our ruminations on the connection between the splitting principle and splittings of polynomials into linear factors , we can at least guess what line elements should look like in . First, I ought to define a "line element" in a general lambda-ring . The definition I'll adopt (and I believe something like this appears in the literature; I need to check up on that) is that it's an element such that for all .
The statement is that line elements in are homomorphisms that take to some element , and to for . In other words, under the identification taking to , the line elements in are of the form . In the first place, it is necessary that a line element be of this form, because if for all , then evaluation at yields
for all . For sufficiency, we must show that if for all , then for all , or that for all and . Since is a ring homomorphism, it is enough to see that belongs to the ideal . For now I am going to leave this as an exercise in using the splitting lemma. The basic point is that
where the arguments of are indexed over all cases where , written as a polynomial in the symmetric functions , contain no terms of type , because for example no monomials terms of type can occur in the expansion in terms of the .
Thus, we have identified line elements in with elements of type , , under the identification using the polynomial basis. What do these elements look like when we use the identification using the basis?
I claim that if and for all , then for all . One way to see this is by induction on . It is true for , since . For , use the following beautiful identity that holds in :
(John and Joe and I have discussed this many times; it has a relatively short conceptual proof at the 2-rig level, using ideas of superalgebra. I may include a proof a little later in these notes.) Applying the ring homomorphism to this identity, and the assumption that vanishes for , we see that
and the induction goes through.
Thus, when written in the basis, a line element in necessarily has the form
Finally, if are line elements [again in the basis], I claim that their Witt product is . (I claim) this can be deduced again by exploiting the splitting principle. As a consequence, in the basis, the Witt product of and must be .
Incidentally, there is a famous ring involution that takes to and to . (We explain this in terms of 2-rig theory at the end of our second paper.) This induces a ring involution on that takes to . This allows us to deduce that
even if we use the basis throughout!
(It's possible I'm slipping up somewhere in this last part, but it's late where I am, and I'm going to bed.)
Yes, I retract the claim that even if we use the -basis. If sends to , then certainly the composite
is an involution, and it sends to . But isn't a ring involution itself, because is not a biring involution; it's only a ring involution. (There should be a quick counterexample to show it's not a biring involution; maybe I'll cook one up later.)
So anyway, we do have the explicit formula
under the -basis, i.e., transferring the god-given big Witt multiplication on over to using the identification given by . But I actually like this formula less than the corresponding explicit formula using the -basis, which is
This is because I want to think in terms of good old-fashioned product formulas
and liken these to splitting principles, like , which comes from the 2-rig map that sends the generator to the sum of line objects .
Ramachandran remarks that the explicit formulas (whether in terms of or ) plus functoriality of are enough to pin down the big Witt multiplication on . It would be nice to understand this point better. I'm thinking there must be a splitting principle type of explanation. It might be something like this. A general element in , say where , may be written as where is the unique ring map sending to . In the category of commutative rings, consider the pushout of the span
where on the right we have power series of bounded degree (admitting formal infinite sums like ). This thing on the right is the Grothendieck ring , and the map on the right is obtained by applying to the canonical 2-rig map . The idea is that, for example, the image of under the pushout coprojection should match
where is the other coprojection. Adapting the formal manipulations of Ramachandran to this framework, the idea is that
is the coefficient of of a formal product , and this formal infinite product is supposed to be (if I'm interpreting Ramachandran correctly) an infinite Witt sum of line elements, since Witt addition is given by multiplying power series, as we saw earlier. So, then, the lambda-ring map takes to an infinite Witt sum
('' stands for 'Witt') and similarly would take some other element to some other formal infinite Witt sum
and the Witt product of these two elements (supposing for now they wind up in the same ) is going to be defined by a third infinite Witt sum
which can be expanded and rewritten back in . As you can see, it looks like a bunch of shenanigans, and there is the mild challenge of making honest sense of these formal manipulations, but that seems to be the idea.
Now you've gotten to the stuff I really want to understand! It will take me a while to absorb this and reply.
The pushout of rings I mentioned at the end probably ought to be replaced by the pushout of the span
where are the two elements we want to Witt-multiply.
It's taking me a while to find enough time to think about this stuff, but please don't mistake that for lack of interest. I will get around to it.
No worries at all! I'm mulling some stuff over anyway, about ways of putting those manipulations of Ramachandran on solid ground.
So the more I think about this, the more I suspect what Ramachandran is doing is a kind of "joke" in the sense of Littlewood. (For anyone who doesn't know what is meant by this, see this MO post.)
My own favorite example of a Littlewood joke is the proof of a statement from spectral theory: if are operators on a vector space and is invertible, then so is . Proof: Write
The punchline of the joke is that is in fact the inverse of , as one can easily verify. The telling of the joke is the stuff in the middle, which strictly speaking isn't really legitimate, but it doesn't really matter: it's just a vehicle for how to remember the punchline, which is all that is needed in a rigorous proof. (I'm reminded here of Abel's comment on Gauss's mathematical writing style: "He is like the fox, who effaces his tracks in the sand with his tail"; in other words, he never gives you the inner motivation, or explanations of how he arrived at his insights. Presumably Gauss would have written down without further commentary.)
If I'm right, those infinite Witt sums in Ramachandran's proof are part of a Littlewood joke about how Witt products work, in a way that can be made rigorous but at the cost of using somewhat more roundabout expressions. I'll try to flesh out these thoughts soon.
(I don't think foxes actually do that.)
(Sounds like expressing yourself in a very literary but logically and scientifically discutable way can be seen as cool when you’re a serious math guy.
It reminds me of this extract from the Wikipedia page « Geometric group theory »:
« In the introduction to his book Topics in Geometric Group Theory, Pierre de la Harpe wrote: "One of my personal beliefs is that fascination with symmetries and groups is one way of coping with frustrations of life's limitations: we like to recognize symmetries which allow us to recognize more than what we can see. In this sense the study of geometric group theory is a part of culture, and reminds me of several things that Georges de Rhampracticed on many occasions, such as teaching mathematics, reciting Mallarmé. »
This phenomenon can also be seen in the titles of math papers on the Arxiv. The words that are more rare are more used than the most common ones even when it makes less sense according do dictionaries. For instance the word « via » is more used than « by » or « through » in titles of the form « characterization of … via/by/through … ».)
"Via" and "by" and "through" aren't syntactically interchangeable, to be sure. You shouldn't say "Characterization of a class A by theorem B", whereas "via" works well there. Instead you say "by application of theorem B", or something. This is usually the situation with rarer words; there are few direct substitutions in English or, I'm sure, in most languages. "Utilize" for "use" is a famous exception that proves the rule!
A little point:
Todd Trimble said:
First, I ought to define a "line element" in a general lambda-ring . The definition I'll adopt (and I believe something like this appears in the literature; I need to check up on that) is that it's an element such that for all .
I think I see how to show any object in a 2-rig with automatically has for . So I would hope that in any lambda-ring implies for . However, my ability to do computations in lambda-rings is not up to the job of proving this!
Kevin Carlson said:
"Via" and "by" and "through" aren't syntactically interchangeable, to be sure. You shouldn't say "Characterization of a class A by theorem B", whereas "via" works well there. Instead you say "by application of theorem B", or something. This is usually the situation with rarer words; there are few direct substitutions in English or, I'm sure, in most languages. "Utilize" for "use" is a famous exception that proves the rule!
Ok for « by ». But isn’t « through » better than « via »? If I remember what I found by looking on the internet, it was said that « through » can be interpreted as « by mean of » whereas « via » is also interpreted as « by mean of » today but in the past was used more strictly to mention a city by which you pass on a travel. I guess it is used also in the sense of « by mean of » today because it is a Latin word and so sounds more poetic and sophisticated.
I also discovered that « cf. » is for « confere » which means « compare » in latin. It was used in the past in a more restricted way, in particular in legal texts to mean « compare with the other reference which says something different. » But today it is used to mean « see », almost all the time to point to a reference where more details are given, so to defend a point which is almost the contrary of what it means in Latin.
At least for cf., it shows that people tend to use uncommon words just to sounds cool, distorting their original meaning.
(Sorry for this interruption which has nothing to do with the very cool math you were talking about.)
Responding to John here: frankly, I get confused about some things here myself, especially when it comes to dealing with negatives. A question that's been needling me is whether every lambda-ring comes from a 2-rig, i.e., whether we can cook up a 2-rig whose Grothendieck ring is the given lambda-ring. That would be a big help!
A related observation (which can be made this time on the paper on the splitting principle!) is that people tend to say that two categories are equivalent when in fact they are even isomorphic, because « equivalent » sounds more cool.
I've noticed this thing about "cf." (some people write instead c.f., not knowing where the abbreviation comes from), and sometimes I have pangs of conscience about using it in this slightly looser way, but then again, language is constantly changing, so I waffle back and try not to worry about it much, feeling that most people reading would get the drift.
I wouldn't attribute "equivalence" to wanting to sound cooler, necessarily (and it sounds unkind to think so). For some personalities, it could be an act of hedging one's bets or playing it safe. It's a tricky business.
There is a spot in the splitting principles paper where we deal with a 1-limit in a 2-category (where we introduce ), which requires thinking up to isomorphism with very specific models in mind, but then other ways of referring to the object up to equivalence could be appropriate if one wishes to evoke other ways of thinking about it. But I'm actually not sure what you're talking about. Do you want to tell us more specifically what you had in mind?
Sure. I was thinking about Lemma 5.2. I think the equivalence is actually an isomorphism here.
Todd Trimble said:
I've noticed this thing about "cf." (some people write instead c.f., not knowing where the abbreviation comes from), and sometimes I have pangs of conscience about using it in this slightly looser way, but then again, language is constantly changing, so I waffle back and try not to worry about it much, feeling that most people reading would get the drift.
Ahah, almost the same for me. Now that I know precisely the original meaning, I feel bad about using it to mean "see" but I'm so used to that it's difficult to stop doing it. But I also think that we don't have to be so conservative about words and so it doesn't matter much.
Jean-Baptiste Vienney said:
Sure. I was thinking about Lemma 5.2. I think the equivalence is actually an isomorphism here.
Eh, I think I'll let John or Joe respond. (I have to leave the house!)
Well, now I think that it is probably not an isomorphism :sweat_smile:. It is just that some category theory people defined the category of affine schemes as in talks. So I might be completely wrong.
Hmm, but you actually define the category of affine schemes precisely like this ahah (as the opposite of ). So I'm not wrong about your lemma 5.2 finally.
It must be that you want to give a simple definition but then you feel bad about it in the proof and statement of Lemma 5.2 because you know that this is not the usual definition and the usual definition is only equivalent and not isomorphic to your definition.
I think it is something like this, but maybe upside down. It's morally correct to treat them as equivalent, but it's also good manners to give a definition of things. So we take advantage of the equivalence to shortcut the definition. Then if we asked ourselves if these two things are at odds, we moved on from it quickly. I don't actually see the benefit of the isomorphism. I think of isomorphisms of categories as equivalences with a bijection on objects. Bijections are good for counting, but I don't want to count anything here.
Equivalence of categories confuse me because I'm always thinking to the equivalence between the category of matrices over and the category of finite dimensional vector spaces over which feel quite different to me: the first one is a world where you think in terms of coordinates and need to do arbitrary choices to do so (if you start from abstract vector spaces), in the second one you think without coordinates. You can compute easily in the first one with a computer, but it would be more difficult to work with the second one on a computer. I tend to feel like category theory is too coarse here (like a coarser topology) and puts under the rug some subtleties (I mean, if you consider that two equivalent categories are more or less the same categories). But it's maybe just a psychological problem of me. So when I read "equivalent" but it is in fact "isomorphic" (or when it is not even clear what it should be because of how things are written), it creates these questionings in my mind and distracts me from the real point of the work under consideration.
Jean-Baptiste Vienney said:
Well, now I think that it is probably not an isomorphism :sweat_smile:. It is just that some category theory people defined the category of affine schemes as in talks. So I might be completely wrong.
This is an example of why I say "equivalence": to show you have an isomorphism of categories you have to check everything very carefully.
But also there's no advantage to doing so! If we are doing category theory in the usual style, isomorphisms between objects in a 2-category like are considered no more useful than equivalences. They are merely distracting.
(I am avoiding the word [[evil]], because it's not politically correct. :upside_down:)
I think that saying "isomorphism" instead would also be distracting for many readers, because it would stop them in their tracks and make them think, "really, isomorphic?", when so much else is already going on. I agree with Joe that there's no real benefit to saying "isomorphism" here, and I expect most readers will go along with "equivalence" without a murmur (it's certainly not wrong to say equivalence). The distinction is not worth bothering about.
Jean-Baptiste Vienney said:
It must be that you want to give a simple definition but then you feel bad about it in the proof and statement of Lemma 5.2 because you know that this is not the usual definition and the usual definition is only equivalent and not isomorphic to your definition.
I assure you that I didn't "feel bad" about anything! Well, not here anyway. I feel bad that we didn't finish the proofs of more theorems! :-)
I don't want to think about whether Lemma 5.2 could be stating an isomorphism - it's like thinking about how many threads of cotton are in my sock, when I just want to put on my sock.
Let me just tell lurkers what this lemma says.
This lemma says that the category of monoids internal to the category of affine schemes is equivalent to the category of commutative bialgebras. It's a triviality, but the two viewpoints have a different feel to them. In one case I picture a geometrical object, an affine scheme, which is equipped with a multiplication making it a monoid. In the other case I imagine a vector space which is equipped with a commutative multiplication and a comultiplication that get along with each other. So the first picture is 'geometrical' while the second is 'algebraic'.
For example in the first picture I might imagine the Lie group , which is shaped like some sort of 3-dimensional hyperboloid in 4-space, but I'd view it as an algebraic variety equipped with a group structure. In the second picture I'd imagine the commutative algebra
and think about how to equip it with a comultiplication arising from the group structure of . I find this second picture a lot less intuitive, but it has the advantage that in the end we just have a vector space with some operations and co-operations - so our linear algebra skills become useful.
In a talk I'm giving to some algebraic topologists tomorrow I'm going to talk about the general process of externalization (of which the above amounts to a particular case). These different perspectives on objects can be wildly different and I'm confident there are many cases where the different perspectives have not been adequately exploited :star_struck:
I suspect that many people who will open that paper have never gotten seriously familiar with categories of comodules of coalgebras, or at least not to anything like the extent they are familiar with modules over algebras. I got an inkling of their importance for algebraic representations of algebraic monoids through conversations some years back with Jim Dolan, but even so it was really only while we were developing ideas for this paper that I really began getting my hands dirty with them. (That's just an expression; no shade on comodule theory, which is very clean and beautiful, and sometimes surprising!)
Getting back to our topic --
The jokey aspect is that Ramachandran manipulates certain infinite Witt sums (= infinite products of formal power series with constant coefficient ), but the language of commutative rings doesn't accommodate infinite products in general. So there's a faint odor of bullshit to what he's doing, even though his arguments are succinct and suggestive, in the style of a good Littlewood joke.
Here is my attempt to put what Ramachandran does on grounds that seem more rigorous to me. Taking it from the top: let be two elements in the big Witt ring , and let denote their Witt product that we are trying to describe. Using the fact that is the coproduct of with itself in the category of commutative rings, there is an induced ring map that I'll denote as . As mentioned earlier, this is the composite
Next, the splitting principle gives a 2-rig extension , or better yet an extension of graded 2-rigs. Here is defined to be the 2-rig consisting of functors , where is the commutative monoid of natural number sequences whose sum is finite, and we regard this commutative monoid as a discrete symmetric monoidal category. This is graded by the sum ; the set of sequences with that sum is denoted . The 2-rig extension takes to the functor that is constantly the ground field on elements of the component , and for other . The elements of are , , , etc., and the line object takes the element of this sequence to , and all other elements of to . Hence is . Decategorifying this 2-rig map gives a commutative ring map, even a lambda-ring map , sending to the elementary symmetric function as we have discussed, where is the isomorphism class . So far, this is all on solid ground.
Applying to this ring map , we get the assignment
as long as we say to ourselves that the right side is suggestive shorthand for the well-founded expression , which makes sense in our context (yes, the coefficients are "infinite sums", but they make sense as elements in , just as ordinary formal power series in , which ostensibly are infinite sums , make perfect sense as sequences of elements of ).
[By the way, if we interpret what Ramachandran is doing with his infinite Witt sums, we are led to write down infinite products of type
as an element of . Yes, one can make sense of this, but... I don't know about you, but taking infinite products of linear terms feels more familiar and comfortable to me than taking infinite products of geometric series. I suppose that's silly, since Euler didn't blink an eye writing down
but anyway I'll stick to how I'm setting this up here.]
I said above that I want to consider the pushout , where we push out the injection along the map .
And one could hope, at least for the limited purpose of trying putting a gloss on what Ramachandran is doing, that this is also an injection. (There might be some really principled way of seeing that, but I don't know what it would be.) It would follow that the induced map , which by definition is , is also an injection.
In the pushout square, there's also a map coming "down",
So the plan is to Witt-multiply the elements and , where the first product sits in the image of
with being the first coproduct coprojection in the category of commutative rings, and the second product similarly sits in the image of . Their Witt-product is the simple-looking
which is again a shorthand for something more complicated-looking, but what is going on can be derived at the 2-rig level. (I am tempted to give this 2-rig level explanation now, but I'll resist.) Now push this element down to . The result is a corresponding Witt-product in . This gives the image in of the desired Witt-product in ; this desired element is uniquely determined, by injectivity of .
This joke is by now more like a shaggy dog story, and I think that's about as far as I'll take it for now. I'm hoping it will make sense to John at least, how this account fits with the verbiage set down in Ramachandran's paper. Somehow I imagine the cognoscenti reading his paper smiling and nodding knowingly at this passage, and with others having cartoon question marks popping out of their heads, because what he writes really is cryptic unless you already know (or until you figure out) the story.
Having unraveled what I think he was getting at, I think it makes matters harder than necessary, although all the ideas are there. But again, all we have to do is figure out how comultiplication works, then take the composite
You figure out how comultiplication works using the splitting principle, along the lines sketched way back here in a simple case, which was amplified further here.
Thanks for getting to the bottom of multiplication in the big Witt ring, @Todd Trimble! It looks a lot simpler and less problematics in the basis, which almost makes me wonder: why bother with the basis?
I guess Ramachandran provides one answer to this question. As you pointed out, in the basis a line element in looks like
Ramachandran wants to relate the big Witt ring to zeta functions over finite fields ; one of the simplest of these is the zeta function of the affine line, which is
I suspect the resemblance is no coincidence.
I just now noticed the double appearance of word "line" - line element versus affine line. That could be a coincidence: it's hard for me to connect these two kinds of line.
Oh! Interesting observation. I'll/we'll have to ponder whether there's some reason that zeta functions would jibe better with the .
It's all rather mysterious. But the zeta function of the affine line should not be too mysterious.
In general, the coefficient of in the zeta function is defined to be the number of -points of the scheme .
The number of -points of the affine line is simply . So we get
or in other words
The really big result Ramachandran proves in this area is that
Curious. The refer to the class of , and for a line object we have . But this is in contrast to the cartesian product , whose size is .
I'll have to peer at that paper some more.
That's a very pretty formula for the zeta function!
We can test Ramachandran's formula
in an example:
This is right, since the affine plane has points over .
Yes indeed.
Ah, so the fleeting remark I made about the Euler product formula might not have been too far off the mark.
Right! Actually I'm confused: rereading Ramachandran I think my claim that is just the generating function for the number of -points of is wrong, but weirdly all the specific computations I did seem to work.
The top formula of his equation (11) is the right formula relating to numbers of points.
Yes, that formula is familiar. You and Jim have those nice papers on the nLab, whose titles you will be able to recall more quickly than I can. One of them interprets the Hasse-Weil zeta function. The other is about zeta functions of Z-sets generally.
The stuff on Euler characteristics is extremely interesting. (And the very coarse "cutting apart" of a scheme into and reminds me very much of Schanuel's papers on negative sets and Euler characteristic. I mean like this one. I don't know if there are others particularly, except for "What is the length of a potato?" where the Euler characteristic plays a starring role. Here it is, courtesy of Tom Leinster's website -- thanks @Tom Leinster !)
You were also telling me about motives some months back, which by my memory also involve this coarse cutting apart of schemes, reminiscent of how a projective space is a "sum"
where the left side is the quotient of punctured -space by the action of the multiplicative group , and the right side is a decomposition into Schubert cells.
Hopefully you can remind me at some point of the things you were telling me about motives, if what I said rings any bells (or even if doesn't).
I'll mention one pretty cool result from Schanuel's negative sets paper. It starts off with a reason that the open interval can be thought of as a negative set, indeed a proxy for . Take an open interval , say and divide it into three parts: Hence "". If we could cancel , then "".
Next, consider the category of bounded polyhedra. A polyhedron is by definition a subset of some Euclidean space contained in the smallest Boolean subalgebra of that contains loci of the form where is an affine function. A bounded polyhedron is what you think it is. A morphism between polyhedron is the graph of a function between them that is itself a polyhedron. The category of polyhedra or the subcategory of bounded polyhedra has some good properties, such as extensitivity. The equivalence comes from an isomorphism in this category.
Let be the Burnside rig of
Theorem (Schanuel): The canonical rig map
is an isomorphism. This is well worth pondering.
Everything you say is ringing bells! :bell:
Ramachandran actually talks about zeta functions from a somewhat "motivic" point of view, but this is based on the low-budget approach to motives based on the "Grothendieck ring of varieties", as explained quite tersely on page 6. The idea is that you take the rig category of varieties, decategorify and group complete it to get a commutative ring, and then impose the extra relations
whenever is a subvariety of . (I am simplifying a bit here.)
The high-end, difficult approach to motives seeks instead to define a rig category of motives, rather than merely this commutative ring, which is intended to be some sort of decategorification of that dreamt-of category.
Mm. I think in fact we touched recently on the fact that motives should form a 2-rig, no?
Or something close to it!
Umm, maybe!
Ramachandran uses to denote the Grothendieck ring of varieties over a field , and intriguingly writes:
The genesis of dates back to 1964 (it was considered by Grothendieck [8, p.174] in his letter (dated August 16, 1964) to J.-P. Serre; it is the first written mention of the word ”motives”). The ring is a shadow (decategorification) of the category of motives; some aspects of the yoga of motives are not seen at the level of .
Wow, he even says "decategorification".
I mean, we maybe have to change -linearity to something else, but we should have a symmetric monoidal category with coproducts and idempotent splittings. Just a passing thought for the moment.
I want to get back to the Schanuel paper I was just describing. It must be understood of course that morphisms in the category , let's say isomorphisms, need not be continuous at all -- the graph of a function can be broken up in pieces, as we saw in the case .
Two very interesting definitions: (1) the Euler characteristic of a commutative rig is the universal quotient where to a commutative rig that enjoys additive cancellation; (2) The dimension of a commutative rig is the quotient .
The Euler characteristic of is the expected quotient to (easy to check I think).
I'll copy what Schanuel says about the dimension of :
Equally simple, if less familiar, is : it is
with and . The exponential notation is in keeping with the idea that multiplying poIyhedra adds dimensions, while adding gives the maximum of the two dimensions.
Todd Trimble said:
I mean, we maybe have to change -linearity to something else, but we should have a symmetric monoidal category with coproducts and idempotent splittings.
Yes, I believe motives form a 2-rig. I can't believe I didn't notice that. I often find myself thinking about two things and noticing only later that they're connected.
I was a bit confused about the -linearity but yes, the category of pure motives defined using 'numerical equivalence' is a -linear abelian category; people don't emphasize the symmetric monoidal structure so much but it should exist.
(There are potentially many categories of pure motives defined using different 'adequate equivalence relations' on cycles, but some of the Standard Conjectures say some of these equivalence relations are the same as numerical equivalence... let us not sink into this mire now!)
Indeed, something that confused me for a while (and apparently still) is that the field we're talking about here is typically different than the field our varieties (or schemes) are defined over! You know how cohomology has 'coefficients'. Motives are like a universal cohomology theory for varieties defined over some field with coefficients in some field . Right now we're talking about the field of coefficients.
Instead of typing this out, I can just refer you to page 382 for the demonstration of the theorem (that I just ascribed to Schanuel).
The same considerations apply to other structures, the so-called o-minimal structures. An archetypal example is where semialgebraic sets replace the semilinear sets that constitute the category . These examples are the propositionally (or Boolean-)definable sets of a model of a logical theory where the language for semialgebraic sets, say, would be given by . In the types of theories I have in mind, there is a quantifier elimination theorem (e.g., Tarski-Seidenberg theorem) that says the image of a definable set under a linear projection is itself definable. And there is also an o-minimality, which says that the only definable subsets of the real line are finite unions of points and intervals. Model theorists, those clever devils, know how to tease out an incredible amount of geometric structure from these two conditions.
Anyway, I think the rough upshot is that Schanuel's theorem extends to such cases as semialgebraic sets.
For anyone interested, we're talking about
and
The latter, unfortunately, is paywalled: I hope the Kazakhs have liberated it.
Yes, I believe motives form a 2-rig. I can't believe I didn't notice that. I often find myself thinking about two things and noticing only later that they're connected.
I was a bit confused about the -linearity but yes, the category of pure motives defined using 'numerical equivalence' is a -linear abelian category; people don't emphasize the symmetric monoidal structure so much but it should exist.
Well, this is incredible. I sort of took a seat-of-the-pants guess there. I'm eager to learn more!
Where do you find this?
(The Kazakhs are doing just fine, btw. I have a tab open.)
By the way, I read about 25 minutes ago that the zeta functions under study are valued in . It should be mentioned that is itself a lambda-ring, where the act on by . These functions generate, as a ring, precisely the integer-valued polynomial functions .
Where do you find this?
I've read a lot about motives here and there. I believe you'd like this, which defines pure motives and states their known properties:
He also has more advanced papers but this covers everything I just said.
By the way, when he says the category of motives is not Tannakian, he (like many other people) is sort of crying wolf: they're not Tannakian with a certain bad symmetric monoidal structure, but they are with the good 'super' symmetric monoidal structure, where you stick in minus signs in the expected places.
(So I take it back about the symmetric monoidal structure being less discussed - I seem to have forgotten lots of stuff I knew.)
I'm making a note for later about the occurrence of Chow rings on page 5 of Milne's paper (we were talking about these in connection with Grothendieck-Riemann-Roch).
While I do want to learn algebraic geometry, I don't feel I have much of a chance doing anything new when it comes to Chow rings. To work with those, you need a good understanding of algebraic varieties. For example, to define the intersection map in equation (1) on Milne's page 5, you need Chow's moving lemma, which says that given two subvarieties you can move one a bit so that they're in general position. This is like an algebraic geometry version of the fact that given an -dimensional submanifold and an -dimensional submanifold of an -manifold, you can isotope one of them so that they intersect in finitely many points. I suffered through learning the techniques for doing such things in differential topology, but I have no desire to go through it all again in the more rigid context of algebraic geometry! And this is just the basic stuff: it gets a lot worse. This is why the Standard Conjectures remain conjectures.
Where I think I might contribute is in figuring out how to distill some concepts from algebraic geometry, formulate them using category theory, and prove (easy) things about them.
So, when it comes to Ramachandran's paper, I'm not seriously interested in proving anything about motives or the Grothendieck ring of varieties. But his proof that the zeta function of a variety obeys
is just a calculation which doesn't really use anything about varieties! It probably works for the Hasse-Weil zeta function of any functor from finite commutative rings to finite sets. (We think of such functor as telling us the set of -points of some gadget for each finite commutative ring , but we don't have to say what this gadget is! The functor says it all.)
So, the kind of question I'm really interested in now is what does multiplication in the big Witt ring really mean - and why does it make Ramachandran's identity hold?
And that's what you've been explaining!
Slogan from Ramachandran (p. 15):
Motivic measures are invariants of algebraic varieties that behave like Euler characteristics.
To be honest, I'm having trouble identifying where Ramachandran's proof of Theorem 2.1 begins and ends. I'll continue reading and scanning (between this and the Milne paper you linked to).
It probably starts at the bottom of page 10 where it says Proof (of Thm. 2.1). But it relies on the previous lemmas.
Oh, I glided right over that text. Thanks.
Oh I see, these ghost components are all about the Adams operations.
Good god, it's Adams' ghost! :ghost:
Well, this just looks ridiculously easy, I have to say. :-)
Seriously, that's frigging amazing.
It's another case of how two things I'm struggling to understand turn out to be related.
I mean, I think this is more or less how it goes. Give me a minute to think.
So I'm looking at my notes that I shared with you, split2rigs. (I want time to think whether I want those publicly shared yet.) Notation: . So then the Adams operations can be defined by a generating function
where is this thing, times the logarithmic derivative of . We've talked about these things before. The Adams operations are of course ring homomorphisms.
Well, "of course". I mean that of course you've seen me talk about this before. We recently went through a proof in that split2rigs, where I invoke the beautiful identity
for all .
What the hell, split2rigs.pdf.
Anyway, you see times the logarithmic derivative in the proof of Lemma 2.3 in Ramachandran.
Yes, I think we talked about that. But are the ghost components simply the components of an element of in the basis? Is that what you're about to tell me?
If so, the ring homomorphism property of the should do something good for these ghost components.
Todd Trimble said:
Well, "of course". I mean that of course you've seen me talk about this before. We recently went through a proof in that split2rigs, where I invoke the beautiful identity
for all .
I'll just mention that the "beautiful identity" can be written as , symbolically
and that's where this whole connection with being the basis analogue of the line elements comes into play.
John Baez said:
Yes, I think we talked about that. But are the ghost components simply the components of an element of in the basis? Is that what you're about to tell me?
Yes, I think you could put it that way.
John Baez said:
If so, the ring homomorphism property of the should do something good for these ghost components.
Yes, that's the message! The ring homomorphism property of the gives the first line of the proof of Theorem 2.1.
So I'm speculating out loud here, really just playing around, but looking at Ramachandran's Remark 2.4, I get a sense that a zeta function on motives might be derivable by taking advantage of the 2-rig structure on the category of motives (I'm following Milne and considering rational equivalence classes of algebraic cycles -- he uses this symbol to cover either the rational equivalence case or the numerical equivalence case). I'll let denote this 2-category of motives. Then its Grothendieck ring is a lambda-ring. That means we get a canonical -coalgebra structure (which is a -coalgebra map = lambda-ring map)
and now I'm half-wondering whether composition with the lambda-ring map where I'll describe in a moment, could morally be a zeta function
Here I'm guessing that could be a dimension function, maybe. Or better yet -- Euler characteristic?! It might itself come from a suitable 2-rig map . (Eh, maybe not.)
(As I say, this is playing around at the moment.)
Ha ha ha, see just about the first line of the paper --
Steve Lichtenbaum’s philosophy [38, 37, 39] that special values of arithmetic zeta functions and motivic L-functions are given by suitable Euler characteristics.
And also of course "Euler characteristic" appears in Remark 2.4. (But I need to think about this more slowly and carefully.)
But my god, equation (15) is beautiful!
There is a pretty pregnant comment on page 15:
• Theorem 2.1 says that gives rise to a motivic measure .
and now I remember the slogan I jotted down:
Todd Trimble said:
Slogan from Ramachandran (p. 15):
Motivic measures are invariants of algebraic varieties that behave like Euler characteristics.
(Sorry, this is a real jumble of half-baked thoughts...)
On a different front: Ramachandran talks about the Grothendieck ring of schemes of finite type over a field . He says this is a pre-lambda-ring. So that's a kind of poor cousin of an actually lambda ring. A pre-lambda-ring is given by a map
(so I'm defining the lambda operations in terms of the given structure map) satisfying an exponential law
and I think that's about it. I'm roughly thinking that if is merely a pre-lambda ring, and not a lambda-ring, that may be because lacks good categorical properties (its not being a 2-rig, for instance). Working with a 2-rig of motives could address this. (?)
Indeed, all this stuff is great! What I'd really like to do is "devein" it - as people say of shrimp - and remove the stuff related to varieties, leaving pure 2-rig theory. Of course we may need a 2-rig with extra structure and properties to get various things to work.
Todd Trimble said:
On a different front: Ramachandran talks about the Grothendieck ring of schemes of finite type over a field . He says this is a pre-lambda-ring.
I seem to recall somewhere he says it's an open question whether it's a lambda-ring. Let me see what he says about that...
Yeah, at the bottom of page 18 he says there are 4 pre-lambda-ring structures on , and says for one it's not known if it's a lambda-ring. I guess this one is the only pre-lambda-ring structure he actually discusses.
I'm getting a little lost there...
I'd want to study this thing.
I edited my comment. I'm annoyed that he has a and a running around and I don't see how he defines .
Oh, I see, it's defined by that factorization diagram near the bottom of page 18.
Anyway, I don't understand this portion of the paper yet.
Yeah, I'm still not getting it.