sniffnoy: (Sonic)
Apparently this was recently proven by Thomas Browning, and the paper also proves that the set is closed! Wow!

I haven't yet had time to actually read this, but I've already gone and updated the page on my website... (Though as of when I'm writing this, the update hasn't gone through yet. :P )
sniffnoy: (SMPTE)
So! I've made a whole bunch of updates to my equational probabilities page. As a result I've eliinated the "???" category, because I managed to rule out nearly everything in that category; the exception, exponential rings, I decided to reclassify as "plausible??".

You can see the page for the details of all the things I've now ruled out that I previously hadn't. I do like how I managed to rule out both sloops and squags by switching betwenen them. :) But the most interesting case is the case of MV-algebras.

Now, there are two other cases on there -- near-rings with identity, and Kleene algebras in the lattice theory sense -- that I classified as "false for well-ordering", because I could rule out well-ordering, but couldn't rule out gaps; but in those cases I assume that the failure of well-ordering means gaps are unlikely. Whereas in the case of MV-algebras, I can prove that well-ordering fails, but gaps hold!

Or at least, they do in the finite case. And that's what I really want to discuss -- is there a compact case to speak of? By which I mean, can we meaningfully define probability in this case? Now of course, any probability measure we define won't be translation-invariant, but that doesn't mean there's no natural way to define one.

For instance: It turns out that MV-algebras are equivalent to lattice-ordered abelian group together with a designated nonnegative element (that also has to meet certain conditions if we want to keep things canonical). However, does that equivalence carry over when working over the category of topological spaces, rather than sets? I found a paper saying it carries over to topoi, but topological spaces don't form a topos, so that's no good.

Well, in truth, I don't particularly care about the answer to that question per se; I just care if we can canonically identify MV-algebras with lattice-ordered abelian groups with a designated nonegative element, I don't particularly care if this actually forms an equivalence of categories or not. So y'know I want the two constructions to be inverse to each other, but if they're not actually functorial, that's fine by me?

Anyway, the idea is that if this works, then perhaps a locally compact Hausdorff MV-algebra, and in particular a compact one, could be shown to necessarily come from a locally compact Hausdorff group, and necessarily has positive Haar measure (obviously it has to be finite measure since it's compact), and so we can get a probability measure on the MV-algebra that way.

...unfortunately, uh, there's a lot that would need to be proven to make that work. Topological spaces, and in particular quotients of topological spaces, can be nasty. (One step of the construction for going from MV-algebra to lattice-ordered abelian group requires taking a Grothendieck group, and that means a quotient.) Um, at least if either is Hausdorff than the other is, that I can show. And I think I can show that going from group to MV-algebra is left-inverse to going from MV-algebra to group (that is, it still is once topology is involved, since after all it's already known to be in the usual non-topological case!). But is it right-inverse? And what about the local compactness issue...?

Yeah, this has been bugging me. I might just not think too much more about this, since this is getting a bit far afield from the original question -- even if the infinite case is meaningful, I don't see how one would prove gaps there, or even have any idea if gaps would hold -- but maybe I should ask about this on MathOverflow...?

-Harry
sniffnoy: (SMPTE)
OK, I've turned a bunch of my recent thinking about equational probabilities into a webpage!
sniffnoy: (SMPTE)
So I was talking to Julian yesterday and as a result I decided to go back and look at groups some more. I figured I'd focus on words of length 4 or less, as well as ak for k composite (since primes have already been looked at so much) and not divisible by any primes larger than 3 (since GAP is unlikely to give me the real maxima for such numbers).

For words of length 4 or less... well, it turns out there just aren't that many of them that aren't in some way trivial or reducible to a shorter one. Other than a4, which I'll cover below, there's aba-1b-1 (i.e. the commutator, already well-studied), and abab-1, and... that's it. You may say, wait, what about a²b², but nope, that's equivalent to the latter of these.

So, uh, as for abab-1, it also has a gap of 5/8, achieved by D8, and the proof is just an easy elaboration on the usual 5/8 proof. And then as mentioned a²b² is equivalent, although I didn't realize this at first and initially proved that it has a gap of 5/8 by a separate means.

OK, so what about ak for k small, composite, and 3-smooth? Here of course I didn't prove anything, but I decided to run the numbers with GAP and see what I'd get.

Let's start with the least surprising one; for k=6, I got an empirical gap of 8/9, achieved by a group that could be described as ((C9×C3)⋊C3)⋊C2. (Note: I didn't bother checking exactly how C2 acts on the rest.) Note that ((C9×C3)⋊C3) is the group that gets you the 7/9 gap for k=3. (Or at least, I assume it's that same group! I'll admit I didn't actually check that it's the same semidirect product.) So, 8/9 as the fraction is a bit surprising, but that group makes some sense... it's 3ish and it's 2ish, you know?

OK, but what about k=4? Here I *also* got an empirical gap of 8/9! Yes, that is a 9 in the denominator! The group that achieves this is a group that could be described as (C3×C3)⋊Q8 -- and make sure you use the right one of those, there are multiple groups matching that description and not all of them have this property! So, we got a factor of 3 seemingly out of nowhere, and not even any powers of 2 in the denominator...

This same group will also yield 8/9 for k=8, and for k any larger power of 2, as there are no elements of order 8.

So what about when I tried k=9? This got me an empirical gap of 17/19, which I really have to wonder if it's right, due to the involvement of a prime larger than 3, which as mentioned above I'm skeptical of. (But that just means I'm skeptical that this is the exact right number... not that it'll be something weird! If it were something non-weird, I would have expected to see that appear as better. :P ) Here, that fraction is achieved by C19⋊C9. Huh!

I guess it is worth noting that in both these cases, of k=8 and k=9, we have a semidirect product where the expected p-part is the quotient, and the normal subgroup is something else entirely. I would not have much confidence in such a pattern holding up, but it's worth paying some attention to I think...

So uh yeah. I guess this is part of why even for groups this is a hard problem that is unlikely to be solved anytime soon!

I'll see if I can quickly prove that 8/9 gap for k=4, but my guess is that if nobody else has so far (and I don't think anyone has), then I won't be able to either, as, uh, I'm not much of a group theorist...

-Harry
sniffnoy: (Chu-Chu Zig)
OK, just a quick followup to these entries: Loops with the inverse property are not enough to get gaps or well-ordering. You can iterate the "generalized dihedral loop" construction, and if the input is a loop with the inverse property, then so is the output. Meanwhile the fraction of involutions approaches 1.

So: It might work for groups. It might work for Moufang loops. (It definitely works for abelian groups.) But it doesn't work for monoids (you can't get rid of inverses) and it doesn't work for loops with the inverse property (you do need some sort of associativity). (Notionally, one could try it for Bol loops or Bruck loops? I am probably not going to do that.)

Meanwhile, over on the ring-like-things side of things, it currently seems entirely plausible to me that all this works even for rngs without associativity, i.e., having absolutely no constraints on the multiplication other than distributivity. Being built on top of an abelian group does a lot, I guess?

-Harry
sniffnoy: (Chu-Chu Zig)
Following up on this post, Drake asked me, what about probabilities in rings? (Or commutative rings?)

I initially dismissed this as not worth thinking about, but... it seems there could be something there? I dunno, it's hard to say. I certainly haven't found any counterexamples for commutative rings or rings more generally (i.e., any equations for which no gap exists, or more generally for which an infinite ascending chain exists), but I feel like mostly the only cases where I can show a gap *does* exist are cases where a gap exists for, like, stupid reasons. :P

The one nonstupid case I know of, in the not-necessarily-commutative setting, is once again commuting probabilities. It seems those once again have a 5/8 gap, for an exactly analogous reason -- I didn't come up with this myself, I found a paper stating the 5/8 gap and then noticed the proof was just the same as in the group case. Of course, if you want an example to show that 5/8 is achieved, that is different!

(EDIT January 1: Actually, I should note that one of the cases I considered "stupid" above is actually somewhat interesting -- considering x²=x in commutative rings.)

I had been kind of hoping there was a classification of finite commutative rings that I could go by, for the commutative setting, but it seems there isn't. The fact that finite rings are necessarily Artin rings of course tells you a lot immediately, but, finite commutative local rings can vary a fair bit, it seems. Oh well. (Of course, if you want to generalize to compact rings, those don't have to be Artin, so...)

I should note that if you try to generalize to semirings (or commutative semirings), then you don't get gaps; you can show that x+y=xy has no gap there, for instance.

So, hm. Probably not really going to go anywhere with this, but, I guess there may be something here after all?

(I'm kind of wondering now whether I should in fact attempt to generalize beyond Moufang loops, to inverse-property loops or something, but, uh, I don't really have a good idea of how I'd get any sense of that... IDK, I'll probably make some little attempt at this later to see what I can find but I don't expect to find much.)

-Harry
sniffnoy: (Sonic)
So continuing from last entry, I thought I'd see what are the biggest actual non-1 probabilities I can get for small Moufang loops (for involutions, commutativity, and associativity); so if there are indeed gaps, we can get an idea what they might be.

Searching for small finite Moufang loops quickly turned up this paper; according to it, most of the non-associative Moufang loops of small order are what I'm going to call "generalized dihedral Moufang loops" (the paper, oddly, doesn't seem to give them a name).

Basically, if you have a group G, you can define a Moufang loop L of order 2|G| by adjoining an element x and defining (for g,h∈G)

gh = gh
g(hx) = (hg)x
(gx)h = (gh-1)x
(gx)(hx) = h-1g.

You'll notice that if G is abelian, then L is just the generalized dihedral group G⋊C2. If G is nonabelian, however, then L is a nonassociative Moufang loop.

Because these are most of the ones of small order, I decided to focus on these, in the hopes that these would give me the best non-1 probabilities. Also, in the group case, we know that the best non-1 probabilities for a²=1 and for ab=ba both come from the dihedral group D8, so this seemed like a decent direction to try.

The results are interesting. First off, involutions and commutativity. These follow exactly the same formulas as they do in the group case... except since the input no longer has to be abelian, we can now get larger numbers out!

Specifically,
PL(a²=1) = 1/2 + 1/2 * PG(a²=1).

Picking G=D8, so PG(a²=1) = 3/4 (i.e. the maximum), we get PL(a²=1) = 7/8, which is what's better than what's possible for a group!

Again, this is the same formula as in the case where L is a group and G is abelian; but in that case our maximum non-1 input would be 1/2.

As for commutativity, we get
PL(ab=ba) = 1/4 * PG(ab=ba) + 3/4 * PG(a²=1).

So again picking G=D8, yielding probabilities of 5/8 and 3/4 respectively (i.e. the maxima once again), we get a commuting probability of 23/32, again better than the 5/8 that's possible for a group!

Once again, this is the same formula as where L is a group and G is abelian, just, now we can pick larger inputs.

Edit: Worth noting also, that when we look at the "commutalizers", we see that some of them are of order 12, and so not subloops. So that's why it's possible to do better than the 3/4 and 5/8 bounds, it would seem.

Finally, what about associativity? What do we get for that? Well, what we end up getting is
PL((ab)c=a(bc)) = 1/8 + 3/4 * PG(ab=ba) + 1/8 * PG(abc=cba).

(Note there's no parentheses on the equation abc=cba, as it's taking place in a group.)

Now, that last probability is a new one, but it's not a difficult one. It's pretty easy to adapt the usual 5/8 commutativity argument to this case and see that this probability also can never be more than 5/8 if the group is not abelian; and that 5/8 is once again achieved by D8. (Indeed, for a generalized dihedral group, this probability will always be equal to the commutation probability.)

So, plugging this in, we get an association probability of 9/16. So, it is possible to have a nonassociative Moufang loop where the probability that 3 random elements associate is at least 9/16.

Of course, this is just what's possible with these generalized dihedral Moufang loops! It's possible the real numbers could be higher... or that there could not be a gap at all. But hey, these could be the correct numbers, right? :) (Hell, I'd say there's a decent chance they are!) And they certainly constrain what's possible...

Edit again: Actually, with regard to associativity, this is not the best possible! If we look at the octonion Moufang loop of 16 elements, then if I've done my calculations correctly, this has an association probability of 43/64, which is bigger than 9/16. Of course, its commutation probability is much worse...

-Harry
sniffnoy: (Chu-Chu Zig)
So, I've written before -- both on my website and here -- about equational probabilities in groups. Beren Gunsolus recently posed to me the question, what if we generalized beyond groups?

Well, equational probabilities in monoids don't really seem to be interesting as best I can tell, or at least not interesting in this way. One can pretty easily find equations that have no gap; e.g., a²=1, or a²=a³ if one insists on equations that don't use 1. (I'll skip listing examples; they're not too hard to construct.)

But what if instead of getting rid of inverses, we weaken associativity? Say, loops? Or, to be a bit more restrictive, Moufang loops?

That equational probabilities in finite Moufang loops might exhibit gaps or well-ordering seems... kind of plausible. So -- I don't know much about Moufang loops, but like... apparently they satisfy Lagrange's Theorem, the order of the subloop divides the order of the loop?? Of course this is true for some complicated reason, not the easy reason it's true for groups, but it's apparently true. So that's a point in favor.

Another reason to look at something more restrictive than just loops is that -- so, in groups, we can look not only at the finite case, but also the compact case. Well, it turns out that locally compact inverse-property loops have Haar measure as well! So if we're looking at Moufang loops, or inverse-property loops more generally, we can again ask about the compact case, not just the finite case. I'm not going to go all the way out to inverse-property loops though; I'm going to keep things to Moufang loops, I mean those are hard enough!

So let's look at the simplest nontrivial cases from groups, x²=1 and xy=yx. (Another really interesting case to look at in Moufang loops would likely be associativity, x(yz)=(xy)z! But obviously that one is trivial in groups.) Do the elementary gap proofs (3/4 for x²=1 and 5/8 for xy=xy) carry over to the Moufang loop case? (Possibly with a weaker gap? In particular, it wouldn't surprise me if that 5/8 for commutativity in the group case becomes 2/3 or 3/4 in the Moufang loop case.)

As best as I can tell, the answer is no. The ordinary proofs for groups are all about looking at centers and centralizers. Now, in a Moufang loop, the commutant -- it's called the "commutant" rather than the "center", because the "center" refers to the intersection of the commutant with the nucleus, the nucleus being the subloop of all the elements that associate with everything -- is indeed a subloop. (Although it's apparently not always a normal subloop.) But as best I can tell, you don't get centralizers (commutalizers?) as subloops. So, as best I can tell, those proofs don't work.

(And even if these proofs do work, they might not extend to the compact case like the group proofs do... finite Moufang loops satisfy the Lagrange property, sure, but is the "index" meaningful? If you had a compact Moufang loops, would a subloop necessarily have measure 1/n or 0?)

...of course, maybe they are subloops, and it just doesn't have an easy proof! That does seem to be how Moufang loops seem to go, things are true but the proofs are hard. And, well, even if they're not subloops, a gap theorem could still hold for another reason!

So, it's an interesting question. I don't really intend to work on this much more, and this is even more speculative than the case of probabilities in groups, but it's certainly an interesting question...

-Harry
sniffnoy: (SMPTE)
So the past few weeks I've gotten sucked back into thinking about probabilities of equations being satisfied in groups after deciding to get in touch with Sean Eberhard. (You may notice I've added a bit to that page.) I know, I know, I should be thinking about integer complexity, but, y'know, sometimes something else calls.

By the way, a question I forgot to ask about on the original page is whether these sets are closed. That question's been added now.

Here are some things I've learned -- some of which has been added to my website page, some of which has not:

1. I asked what happens if we allow for compact groups instead of just finite groups. Well, for the case of [a,b], i.e. commuting probabilities, what happens is nothing! (Aside from the inclusion of 0.) Eberhard shows in this blog post that given any compact group, with commuting probability greater than 0, there's a finite group with the same commuting probability.

2. For ap, p an odd prime, the set is definitely not closed... at least not for finite groups. The point (p-1)/p is a limit point, but not in the set. However, it is in the set if we allow more general compact groups.

This raises the question of whether, for a given word, the set for compact groups might be equal to the closure of the set for finite groups. (Remember, in the case of [a,b], it's conjectured that the set for finite groups might be already closed, aside from 0.)

If the set for [a,b] is closed, it would have to be for a pretty crazy reason. Like, for [a,b], the limit point 1/2 is approached by 1/2 + 1/22n+1; the latter is realized by a group G which consists of the product of n copies of D8, but with their centers identified. But 1/2 is instead realized by S3. Yeah...

3. If you do want to consider compact groups, connected compact groups will get you nowhere. In a connected compact group, regardless of the word, you'll only ever get 0 or 1.

Unfortunately, that doesn't mean you can just pass to G modulo the connected component of the identity, and thereby only consider profinite groups; taking this quotient may change the probability. (E.g., consider if G is S1⋊C2, the generalized dihedral group over the circle, and your word is a²; the correct probability is 1/2, but if you mod out you'll get a probability of 1.)

Still, it does simplify things somewhat, as it does mean you only need consider each way of assigning a connected component to each variable, and seeing whether that gets you 0 or 1; you can then average that over the connected components.

4. A thing I didn't make clear last time -- I mentioned that the 5/8-gap argument for [a,b] is simple enough that it obviously applies to compact groups; I didn't mention that the same is obviously true of the 3/4-gap argument for a². (Unfortunately, the only 7/9-gap arguments I can find for a³ seem specific to finite groups.) But way more is true!

In their paper Groups with automorphisms inverting most elements, Liebeck and MacHale give a classification of finite groups with over half the elements being involutions; in particular, the probabilities you get this way are 1/2 + 1/(2n). So, that's our top ω for a², or ω+1 if you include 1/2 itself.

But what the authors don't note is that -- as far as I can tell -- their proofs work just as well (with only a little modification) for compact groups as well! Their classification goes through with only minimal modification, and you get the same probabilities! (Basically: In Type I*, A can now be infinite, although [A:CA(x)] must still be finite, and in the other types, Z can now be infinite. If we're concerned with counting involutions, then for the other types that means E can now be infinite.)

The most significant thing to note about the necessary modifications, FWIW, is that for Theorem 3.5, which is pretty key, you don't need to assume that H is finite index; rather, you can conclude it! If H is infinite index, you get a contradiction. Similarly with 3.6; you don't need to assume that the qi are finite, you can conclude it. Note that you'll need Zorn's Lemma to prove the existence of a maximal abelian subgroup.

Unfortunately, it's probably going to be really hard for anyone to get that into the literature...

5. If we're looking at top ω's, then for [a,b], this is known as well, plus a little more. In his paper What is the probability that two elements of a finite group commute?, David Rusin finds that, for finite groups, the top probabilities are 1/2 + 1/22n+1 (as mentioned above), then 1/2, then 7/16, then 11/27, then 2/5, then 25/64, then 3/8, then 5/14, then 11/32. So, that's the top ω+8.

Except... he doesn't actually include 5/14 (realized by D14); I added that in. This doesn't seem to be a fundamental mistake, just a minor omission in case 2; he failed to realize that setting p=7, n=2 yields 5/14 > 11/32; he claimed that p=7 always yields something less than 11/32, which is false.

Dunno if anyone's corrected that in the literature; doubt it though.

Rusin's methods seem specific to finite groups, but by Eberhard's work above we know that doesn't matter!

6. Eberhard's proof that the probabilities for [a,b] are reverse-well-ordered really doesn't seem to imply that the order type is ωω, even though it seems like it should. :-/ (The main problem seems to be that the mapping p↦q need not be injective.)

There's more I could say, but I think this is a good place to stop.

-Harry
sniffnoy: (Kirby)
Following up to this entry, I want to report: Yes, we can indeed construct a group of order pp+1 where the fraction of elements of order p is 1-1/p+1/p². Thanks to Hunter and to Julian for figuring out an important part of the construction.

Although, first I should point out David Speyer's construction from the comments, which does achieve the desired fraction but uses a larger order to do so; it has order p(p choose 2) + 2. The group consists of all (p+1)×(p+1) upper-triangular matrices over Fp with 1's on the diagonal such that the first superdiagonal is an arithmetic progression.

But, OK, how can you do it with only pp+1 elements?

EDIT: Actually Julian came up with an even simpler description; skip to the end if you just want to see that.

Well, I was inspired by the p=3 case. In the p=3 case, there's a unique group of order 81 that achieves the fraction 7/9, and it has the form (C9×C3)⋊C3. And of course for p=2 our group D8 has the form C4⋊C2. So, I decided to try to generalize this. I wondered, can we do it as a group of the form (C×Cpp-2)⋊Cp?

Let's say A = C×Cpp-2 and G is our group. We need to specify some automorphism of A of order p. But that's not enough; in order to get it to work, we need that all the elements of G\A have order p, which is not something that happens automatically.

So, what do we actually need here? Let's consider A additively; then our automorphism is given by a matrix M. We can consider this as a matrix over Z, if we want to be lazy, or over Z/(p²), if we want to be slightly less lazy, but really it's a sort of mixed matrix; the first row consists of elements of Z/(p²), while the other rows can be taken to be over Z/(p). Moreover, in the first row, all the off-diagonal elements must be divisible by p. (And also the diagonal element in the first row had better not be divisible by p, but that's not exactly a tough restriction and also it's redundant with the additional requirement I'm about to tack on.)

Now, we said it has to be of order p, right? So we need Mp=1. Except as mentioned above, that's not strong enough; in order to get that all elements of G\A have order p, what we want is the stronger condition 1 + M + M2 + ... + Mp-1 = 0. Except, what do I mean by these matrix equations? Equal mod what? Well, they need to be equal mod p² in the first row, and equal mod p in the rest.

So the problem remains to construct such a matrix! Whether considered as a mixed matrix like I described above or as a matrix over Z or whatever. This is the problem as I posed it to various people I know the other day. (Figuring I'd take it to MathOverflow if I didn't get an answer, but I did.)

Hunter came up with a way of doing it first, and his way is definitely more direct conceptually, but I'm going to go with Julian's construction because it's really nice and also easier to carry out; no need to go finding eigenvectors or extending things to free bases. Note that this construction (both Hunter's and Julian's, actually) only work if p>2, but of course we already know how to handle the p=2 case (you take the 1x1 matrix (-1) and get the group D8).

Julian's construction is as follows: Let's construct the transpose MT instead. Take the ring Zp], where ζp is a primitive p'th root of unity. As an abelian group this has a free basis 1, ζp-1, (ζp-1)2, ..., (ζp-1)p-2. Let's reverse that, so (ζp-1)p-2 is the first basis element and 1 is the last. Now consider the additive endomorphism that is multiplication by (1-p)ζp. Then the matrix of this, with respect to the basis I just gave, satisfies the transposed version of all the requirements. (It's a matrix over Z of course but you reduce it down.) So you transpose it and you get M.

I'll skip his proof, I don't think it should be too hard to rederive. :) (If you just use the fact that vpp-1) = 1/(p-1).)

So there you go! Take M as above, take the semidirect product, and you've got your group. (Julian also asked an interesting question: Since he found a way to construct M by means of Zp], can the group G be naturally expressed in terms of this ring somehow?)

EDIT: Yes it can, at least somewhat. We can write A as Zp]/((ζp-1)p), and then our automorphism is just multiplication by ζp, and then we construct the semidirect product with that. No need to construct any fancy matrix, hooray! And unlike the other two constructions, this one still works when p=2. :)

Anyway, that question at least is answered: You can achieve 1-1/p+1/p², and you can do it with only pp+1 elements. Unfortunately that was the easy half, of a watered-down version, of a special case, of my original question! But hey, like I said, I'm not really planning on tackling any of the hard versions, so I'll take it. :)

-Harry
sniffnoy: (Chu-Chu Zig)
I posted a series of tweets about this the other day, but, y'know what, let's write this up a little more properly on DW.

Whenever I see a theorem or conjecture about a gap, between the smallest element of a set and the second-smallest, I ask if this is the start of a well-ordering. Obviously, this applies in reverse too; if there's a gap between the largest and second-largest, I ask if it's a reverse well-ordering.

(And remember, when asking about your set's order type, don't forget to ask if it's closed, or to consider the closure if it's not!)

What are some examples of this? Well, of course, there's all my theorems and/or conjectures on integer complexity and addition chains (and all sorts of variants), but I'll skip going into that for now, because I've discussed that at length here already.

There's the group theory problem I mentioned in my last entry. That's certainly a good one; the numerical data looks promising there.

Here are some others I can think of, off the top of my head:

The 1/3-2/3 conjecture: Given a finite poset P, define δ(P) to be the maximum over all pairs of distinct elements (x,y) of the minimum of the fraction of linearizations in which x<y and the fraction of linearizations in which y<x. (So, δ(P)=0 iff P is already a total order.) Then the 1/3-2/3 conjecture says that any P with δ(P)>0 in fact has δ(P)≥1/3. I.e., it's a gap conjecture. Could the values of δ be well-ordered?

(Maybe they can't! I don't know a lot about this sort of thing, maybe there's already a counterexample.)

(EDIT October 2021: Yup, they're not.)

What about various number-theoretic sets of real algebraic integers? Like, there's Lehmer's conjecture, which conjectures that there's a gap between 1 and the next-smallest Mahler measure. However, Mahler measures can't be well-ordered, because the include the Pisot numbers. That said, we can still ask about order type! In particular, the order type of the Pisot numbers is known, and it's a nifty one! Though it begins with an initial ω, it isn't a well-order after that. Still, like I said, it's a nifty order.

In fact, I'd conjecture that if you considered integer complexity or addition chains with subtraction allowed, and considered the defects arising from those, and then either took the closure or added whole numbers (because those should probably give the same set), that the result would be order-isomorphic to the Pisot numbers. But I'm not trying to prove that right now; I tried briefly and found that, total orders get much harder once they're not well-orders! Someday, I hope to get back to this, but certainly not now.

Anyway, they won't be well-ordered, but one could still ask about order types of (or closures of) Salem numbers or Mahler measures or Perron numbers.

(EDIT October 2021: No, not really; it turns out that each Pisot number is approached on boy sides by Salem numbers, and so Salem numbers are not well-ordered either; and Mahler measures and Perron numbers include the Salem and Pisot numbers and so are definitely not well-ordered either. Indeed, I already noted this for Mahler measures above...? This entry seems a bit inconsistent...)

Of course maybe you get nothing like that. Like in the theorem mentioned in this MathOverflow answer, where you start out with a gap, then another gap, but then you just get all real numbers past that point. What can I say, nothing's guaranteed.

Still, I think this is a question people should ask more often. So: What are other examples of sets of numbers with gap theorems, that we could ask if they're well-ordered? I'm sure I must have missed a ton.

-Harry
sniffnoy: (Chu-Chu Zig)
So I've recently been wondering again about this question (yes, this was originally a post here, but I decided it deserved a page on my website). Just mostly idly wondering about it, not seriously studying it; I'm no group theorist, and I don't really have time to think about such a problem what with everything else that needs doing (both in terms of old math that needs getting back to or writing up, and non-math stuff).

Anyway. As mentioned in the linked page, the simplest place to start would be words of the from an, so we're counting all elements of order dividing n. Or really the simplest place to start would be words of the form ap, for p prime, so you're just counting elements of order p (together with the identity). (Or maybe prime power order if you want to be a little more general.)

Also as mentioned, for p=2, it's known that there really is a gap between ¾ and 1, with ¾ being achievable as D8; I'm using the convention here that the dihedral group with 2n elements is D2n, not Dn.

What I'm wondering is, what's the best known fraction (less than 1) for other p (or maybe other n)? Even if we can't prove a gap, it would still be nice to see how high we can find, to maybe get some idea of what the best might be (assuming that there really is a best like I'm hypothesizing).

Like, sticking to primes first -- for p=2, the best is ¾, that's a theorem. What about for p=3? I can again get ¾ by looking at A4; I don't know offhand how to do better than that. And for p=5, uh, no clue. Well, I mean, obviously one can get ½ for any n (take C2n), but I'm considering that baseline -- can we do better?

Since for p=2 you get a good fraction by looking at D8, i.e., C4⋊C2, you might think more generally to try looking at C⋊Cp; however, this doesn't work so well when p≠2 -- you don't get lots of elements of order p like you do when p=2. Like so many other differences between 2 and the other primes, this one once again basically comes down to the fact that 2 is the only prime p for which p does not divide (p choose 2). (I really think this is an underappreciated point, it really does seem to underlie so many of the differences, but I've never seen anyone else point this out!)

If we allow more general n then of course for any even n we can get ¾ by looking at D4n. But mostly I don't want to look at more general n because this quickly gets kind of silly. I guess powers of primes might be kind of noteworthy? In which case that's how to get ¾ for arbitrary powers of 2? But like let's say you look at powers of 3 -- well then A4 still does the job, you don't need to change the example at all! This sort of thing is part of why I say more general n is kind of silly. I guess more generally this gets you anything divisible by 3 but not 2, so we've now covered all n divisible by either 2 or 3. And of course in the trivial case of n=1 the best you can do is the baseline ½. But this is kind of why I said more general n is kind of silly.

Anyway, I dunno, that's what I've got. Anyone know anything about this?

ADDENDUM: OK, doing some quick computations with GAP here's what I get for some small primes based on groups of size less than 256 (haven't attempted to determine what groups witness these; also man I really do not know GAP, using it is a pain):

p=2: 3/4 (as we know)
p=3: 7/9
p=5: 9/11
p=7: 7/8
p=11: 21/23
p=13: 1/2
p=17: 1/2
p=19: 1/2

(EDIT: See the comments for some updates here)

Surprising, if you ask me! My offhand suspicion is that those ½'s at the end are spurious, an artifact of me not using large enough groups, but who knows? And it's not at all clear what pattern might be going on in the first few...

FURTHER ADDENDUM: Wow, after plotting the results from above for p=2 on a graph and doing the same for p=3... yeah, I'd totally bet on well-ordering, for this simple case at least. But I don't think I can go further and actually attempt to discern the order type from the graph, as I'd like to.

(Like I said I have no intent of trying to prove this. Maybe I will run some more numerical experiments to gather more empirical evidence, but that's it.)

YET MORE: You know what, just see the comments for more regarding the numbers.

-Harry
sniffnoy: (Sonic)
By "irresponsible speculation", I mean "speculation without having done my homework first", i.e., without having even tried to look at the existing literature beyond this paper or actually really tried anything at all.

So, Juan Arias de Reyna recently pointed out to me the following paper: Commuting Probabilities of Finite Groups, by Sean Eberhard.

EDIT DEC 1: The paper has been updated; the following refers to the original version of the paper. See the comments for discussion of the updated version.

So: If you have a finite group, you can consider the probability that a randomly chosen ordered pair of elements from it commutes. You could then consider the set of all probabilities obtained this way; this is some set of rational numbers between 0 and 1.

In this paper, Eberhard shows that this set is in fact reverse well-ordered! Its order type is of the form ωα where 2≤α≤ω². (Though Juan Arias de Reyna points out that it's not too hard to find examples that show that in fact α≥ω, so the order type is at least ωω.) He also shows that all the limit points of this set are rational. I think it should be pretty obvious why I'm interested in this! (Even though I have no plans to actually study this, my hands are full as it is.)

Now for the irresponsible speculation:

1. Instead of just doing finite groups, one could generalize to compact groups; one can consider probabilities there as well. How would this affect the set? It would be really nice if this just gave you the closure of the probability set for finite groups! Though Arias de Reyna tells me it's conjectured that the finite group commuting probability set is closed in (0,1], so that would be saying that using general compact groups only gives you 0 in addition. (I mean, it certainly does give you 0 in addition!)

2. I'm reminded of some work of John Wiltshire-Gordon and Gene Kopp. They considered probabilities of randomly chosen elements from a compact group satisfying some general word; the case of commuting probabilities is the case where the word w is aba-1b-1. I wonder if the same phenomenon might be seen for other words.

Probably it would be best to first look at words in one variable. Obviously using w=a generates an ω if we stick to finite groups and an ω+1 if we allow general compact groups -- not ωω, but still well-ordered. As for a² or a³... well, I don't know, and I don't really intend to work on it as I said above, but it seems vaguely plausible and it's an interesting question!

So yeah that's basically all I have to say on the matter.

-Harry

June 2025

S M T W T F S
1234567
891011121314
15161718192021
2223 2425262728
2930     

Syndicate

RSS Atom
Page generated Jul. 8th, 2025 11:53 pm
Powered by Dreamwidth Studios