Also available at http://math.ucr.edu/home/baez/week190.html December 26, 2002 This Week's Finds in Mathematical Physics (Week 190) John Baez Fall quarter was very busy for me. Next quarter I'll be on sabbatical. I just graded all my final exams, turned in the grade reports, cleaned up my house for the people who will be renting it, and left town. Now I'm in Hong Kong, away from all my usual duties, and I have time to catch up on various things. For example: in November I went to the annual meeting of the Philosophy of Science Association, which was held in Milwaukee. I've never gotten around to talking about this yet. I spoke in a session on Structuralist Approaches to Quantum Gravity, organized by Steven French. "Structuralism" means a lot of things, but as far as I can tell, in the philosophy of physics, it's an attempt to understand how terms gain their meaning as part of a physical theory, and the subtle sense in which they can retain some of their meaning as theories evolve. I ran into this problem in a very practical way when a mathematician once asked me to "define an electron". I was reduced to incoherent sputtering: physics ain't math! The electron in Bohr's model of the hydrogen atom is a very different thing than the electron in Dirac's hydrogen atom, which in turn is very different from the electron in QED... but still, there's something "the same" about them, even apart from the fact that they're all attempts to model the "same thing out there in the real world". How does this work, exactly? It's way too complicated for me, but you can try reading this and see if it helps: 1) Heinz-Juergen Schmidt, Structuralism in physics, The Stanford Encyclopedia of Philosophy (Winter 2002 Edition), ed. Edward N. Zalta, http://plato.stanford.edu/entries/physics-structuralism/ To be honest, what I really like about structuralism is that it makes philosophers think a lot about things like "mappings between theories", which gets them interested in category theory, which in turn gets them interested in other good stuff: background-free physics, n-categories, and so on. It can't be all bad if it does that! I liked this conference because I met quite a few philosophers who are well-versed in the technical aspects of physics and busy thinking about interesting things. Perhaps the most obvious example is John Earman, who gave a big plenary talk on spontaneous symmetry breaking. I'm really fond of a paper he wrote with his student Gordon Belot on the problem of time in quantum gravity: 2) John Earman and Gordon Belot, Pre-Socratic quantum gravity, in Physics Meets Philosophy at the Planck Scale, eds. Chris Callender and Nick Huggett, Cambridge U. Press, Cambridge, 2001. and also this paper on the C*-algebraic approach to quantum field theory on curved spacetime: 3) Aristidis Arageorgis, John Earman, and Laura Ruetsche, Weyling the time away: the non-unitary implementability of quantum field dynamics on curved spacetime, in Studies in the History and Philosophy of Modern Physics, in press. Laura Ruetsche is another student of Earman. At the conference, she gave a nice talk about what it means when a quantum theory formulated in terms of C*-algebras has many inequivalent Hilbert space representations. Here's another paper by the Earman gang: 4) Gordon Belot, John Earman and Laura Ruetsche, The Hawking information loss paradox: the anatomy of a controversy, British Journal for the Philosophy of Science, 50 (1999), 189-230. I haven't read it yet, but I heard John Earman talk about the subject in Vancouver in 1999 and he made a lot of sense. He emphasized that talk of a "paradox" is overblown: there's no reason information *needs* to be conserved in the process of Hawking radiation. Most physicists wish it were, though. Another philosopher I enjoyed speaking to was Alisa Bokulich of Boston University. She mentioned some fascinating things about how one can calculate the spectrum of the helium atom in terms of the dynamics of the *classical* three-body problem. This is precisely where the "old quantum mechanics" of Bohr and Sommerfeld gave up - their ideas only worked for completely integrable systems, where all the orbits are periodic. The three-body problem is not completely integrable: it exhibits chaos, so there are lots of nonperiodic orbits, with periodic orbits densely woven among them. But the old quantum mechanice has recently experienced a kind of renaissance thanks to work on quantum chaos. Apparently now people can compute the energy levels of the *quantum* version of a chaotic system in terms of a sum over the periodic orbits of the corresponding *classical* system! You use something called the "Gutzwiller trace formula", and maybe some other stuff... I don't understand this, but I want to - especially because thanks to this "trace formula" business, there are tantalizing connections to the Riemann hypothesis! People like Michael Berry have hinted that maybe someone could solve this famous open problem if they found a chaotic dynamical system with orbits having periods related to the prime numbers in the right way... or *something* like that; by now I'm just babbling half-forgotten second-hand gossip. Anyway, Bokulich gave me some references that I plan to read. First, a thorough historical review of the subject: 5) G. Tanner, K. Richter and J. Rost, The theory of two-electron atoms: between ground state and complete fragmentation, Reviews of Modern Physics 72 (2000), 497-544. Then, a classic paper extolling the forgotten virtues of the old quantum theory: 6) J. Leopold and I. Percival, The semiclassical two-electron atom and the old quantum theory, Jour. Phys. B13 (1980) 1037-1047. Next, a paper containing the first successful semiclassical quantization of helium: 7) G. Ezra, K. Richter, G. Tanner, and D. Wintgen, Semiclassical cycle expansion for the helium atom, Journal of Physics B 24 (1991), L413-L420. If you don't know anything about the old quantum mechanics, here's a good place to start - it begins with a long explanation and then has translations of original papers: 8) D. ter haar, The Old Quantum Theory, Pergamon Press, London, 1967. And finally, a here's a modern online book on semiclassical methods and quantum chaos, in the process of construction: 9) Predrag Cvitanovic, Roberto Artuso, Per Dahlqvist, Ronnie Mainieri, Gregor Tanner, Gabor Vattay, Niall Whelan and Andreas Wirzba, Chaos: Classical and Quantum, available online at http://www.nbi.dk/ChaosBook/ Cvitanovic is really big on these online books: he's almost done writing one about diagrammatic methods in group representation theory. I should talk about this soon, because it contains some exciting new insights on the exceptional groups. But I'm not really ready yet, so for now I'll just throw you the reference: 10) Predrag Cvitanovic, Group Theory: http://www.nbi.dk/GroupTheory/ Instead, let me talk some more about structure types and their generating functions. I described these concepts in "week185", but I didn't give many examples, which is a real pity. I want to make up for that omission now. First remember the basic idea. A "structure type" is any sort of structure that we can put on a finite set. Given any structure type F, we let F_k be the *set* of ways we can put this structure on a k-element set, and let |F_k| be the *number* of ways we can do it. We define the "generating function" |F| to be the formal power series |F_k| |F|(x) = sum ----- x^k k! Nice operations on generating functions come from nice operations on structure types, so we use the same notation for both. For example: given structure types G and H, we define the structure type G+H by saying an G+H-structure on the set S consists of either a G-structure on S or an H-structure on S. This definition gives: |G+H| = |G| + |H| Or: we define the structure type GH by saying an GH-structure on S consists of a way of chopping S into two disjoint subsets and putting a G-structure on the first subset and an H-structure on the second. If we make this definition, we get: |GH| = |G| |H| Or: we can define a structure GoH by saying an GoH-structure on S consists of a way of partitioning S into disjoint parts, putting a G-structure on the set of parts, and putting an H-structure on each part. Then we get: |GoH| = |G| o |H| where on the right "o" means that we're composing the generating functions |G| and |H|. Here we have to be a bit careful: the composite of formal power series is not always a well-defined formal power series, so the above equation only works when the right-hand side makes sense. It's easy and highly instructive to check all the claims I just made. But let's see what cool stuff we can do with them! First, consider the structure of "being a totally ordered n-element set". There are *no* ways to put this structure on a k-element set if k is different from n, and there are n! ways to put it on an n-element set. So the generating function of this structure type is just x^n If we call this structure type X^n, we get this cute equation: |X^n| = x^n Next, suppose G is the structure "being a totally ordered set". This is the same as being a totally ordered 0-element set or being a totally ordered 1-element set or being a totally ordered 2-element set or... you get the idea. So, we have G = 1 + X + X^2 + ... and thus |G|(x) = 1 + x + x^2 + ... 1 = ----- 1 - x Next, suppose H is the structure "being a totally ordered set with 1 or 2 elements". This has the generating function |H|(x) = x + x^2 Now let's consider the structure type F = GoH To put a structure of this type on a set, we partition the set, order the parts, and give each part the structure of being a totally ordered set with 1 or 2 elements. This sounds a bit weird! But if you think about it, it means: "To put an F-structure on a set, order it and then partition it into parts of size 1 or 2." And we can count the ways of doing this by using this generating function: |F|(x) = |G| o |H| (x) 1 = ----------- 1 - x - x^2 = 1 + (x + x^2) + (x + x^2)^2 + (x + x^2)^3 + (x + x^2)^4 + ... = 1 + x + 2x^2 + 3x^3 + 5x^4 + ... Hey! Fibonacci numbers! It looks like the kth coefficient of this generating function is just the kth Fibonacci number! Now, remember that generating functions have a factorial built into them: |F_k| |F|(x) = sum ----- x^k k! So apparently in this example |F_k| is k! times the kth Fibonacci number. Of course, k! is the number of ways to order a k-element set. So apparently the kth Fibonacci number is just the number of ways to chop a k-element set into parts of size 1 or 2. But how can be *sure* we're getting the Fibonacci numbers as coefficients? Well, after the checking the first couple of coefficients, we just need to make sure that each coefficient in our generating function is the sum of the previous two. And that follows straight from this equation: 1 x x^2 ----------- = ----------- + ----------- + 1 1 - x - x^2 1 - x - x^2 1 - x - x^2 Even better, the above equation comes from an isomorphism between structure types: F = X F + X^2 F + 1 Since X^n is the structure "being a totally ordered n-element set", this isomorphism says: "To put an F-structure on a set S, either remove one element from S and put an F-structure on the rest of S, or remove two elements, order them, and put an F-structure on the rest of S, or check to see if S is the empty set - in which case it has exactly one F-structure, by definition." This recursive definition of F is a categorified version of the recursive definition of the Fibonacci numbers. It gives perhaps the most direct way to see that the number of ways of chopping an n-element set into parts of size 1 or 2 is equal to the nth Fibonacci number. It's pretty simple, and we might have discovered it without structure types - but we can get this sort of thing *systematically* if we use structure types. We also get other spinoffs. For example, the pole of this function 1 ----------- 1 - x - x^2 that's closest to zero occurs at the reciprocal of the golden ratio: 1/G = 0.6180339... So, by a theorem of Hadamard, the nth coefficient of the corresponding series 1 + x + 2x^2 + 3x^3 + 5x^4 + 8x^5 + ... must grow roughly like G^n. In other words, the Fibonacci numbers grow roughly like powers of the golden ratio. Now, this should not be news to any true lover of mathematics! And you can get far more precise information along these lines without much more work. But I'm just trying to make a general point: in combinatorics, we can estimate how fast the number of ways of doing something grows by studying poles of generating functions. For example, suppose you wanted to know approximately how many ways there are to take a million dollars and break it down into 1, 5, and 10 dollar bills. The generating function that solves this problem is 1 1 1 ----- ------- ---------- 1 - x 1 - x^5 1 - x^{10} I'll let you do the rest. Here's another classic example. The number of binary trees with n leaves is called (annoyingly) the (n-1)st Catalan number. There is a structure type T where a T-structure on a set is a way of making it into the leaves of a binary tree. For example, here's a T-structure on the set {a,b,c,d}: b d c a \ \ / / \ \/ / \ \ / \ \/ \ / \/ The number of T-structures on an n-element set is n! times the (n-1)st Catalan number, thanks to the different orderings. To put a T-structure on a set, we either check to see that it has one element, in which case there's a single T-structure, or chop it into two parts and put a T-structure on each part. This means that T = X + T^2 and thus |T| = x + |T|^2 so |T|(x) = (1 - sqrt(1 - 4x))/2 = x + x^2 + 2x^3 + 5x^4 + 14x^5 + 42x^6 + .... so, for example, there are 42 binary trees with 6 leaves. In fact, I did this calculation already in "week144", but I didn't explain it in terms of structure types. You can learn more about Catalan numbers there. If you think this stuff is fun, ponder T(1). This corresponds naturally to the set of all trees. What's the cardinality of this set? Well, the sensible answer is to sum the series: |T|(1) = 1 + 1 + 2 + 5 + 14 + 42 + .... In other words, infinity! But if we were feeling quite relaxed about everything, we might use the other formula for |T|(x) and guess |T|(1) = (1 - sqrt(-3))/2 = exp(-i pi/6) This is pretty odd: it's a complex number! The problem is, we're outside the radius of convergence of the power series. However, this answer is not completely crazy: we can use it to guess things that would be hard to guess otherwise! For example, this number is a sixth root of unity, so if we raise it to the seventh power, we get the same number back again: |T|(1)^7 = |T|(1) Categorifying this fact, Lawvere guessed there was indeed a nice isomorphism T(1)^7 = T(1) In other words: one can take this weird calculation and use it to construct a one-to-one correspondence between trees and 7-tuples of trees! For a good treatment see this paper by Blass: 11) Andreas Blass, Seven trees in one, Jour. Pure Appl. Alg. 103 (1995), 1-21. Also available at http://www.math.lsa.umich.edu/~ablass/cat.html Recently, Leinster and Fiore have proved a very general theorem on how to reason rigorously with complex-valued "cardinalities": 12) Marcelo Fiore and Tom Leinster, Objects of categories as complex numbers, available as math.CT/0212377. This explains the curious result of Lawvere and Blass, and should be a good clue when it comes to a favorite puzzle of mine: how can we categorify the complex numbers? There's much more to say: I should discuss all this using more category theory, say how structure types are related to "operads", and so on... but I'm sitting in a coffee shop and I shouldn't keep hogging this computer, so I'll quit now. Happy Boxing Day! ----------------------------------------------------------------------- Afterword: Gordon McCabe sent me an email with some useful extra references. The second paper here is the talk John Earman gave at the Philosophy of Science Association meeting described above. John, Pleased to see Philosophy of Science making an appearance in the latest 'This Week's Finds'! I noticed that there were no http references to the papers by Earman and Belot that you allude to. Philosophers do seem to have been very slow to catch on to this business of Internet preprints, but there is a growing archive of electronic preprints hosted by the University of Pittsburgh at http://philsci-archive.pitt.edu/. You can find a number of papers here by Earman and Belot, which you might want to add as http references to 'This Week's Finds': Earman, John (2001) Gauge Matters. http://philsci-archive.pitt.edu/documents/disk0/00/00/00/70/index.html Earman, John (2002) Laws, Symmetry, and Symmetry Breaking; Invariance, Conservation Principles, and Objectivity. http://philsci-archive.pitt.edu/documents/disk0/00/00/08/78/index.html Belot, Gordon (2002) Symmetry and Gauge Freedom. http://philsci-archive.pitt.edu/documents/disk0/00/00/05/27/index.html Regards, Gordon McCabe Also, someone noticed something funny about the following: > "To put an F-structure on a set, order it and then partition > it into parts of size 1 or 2." > >And we can count the ways of doing this by using this generating function: > >|F|(x) = |G| o |H| (x) > > 1 > = ----------- > 1 - x - x^2 > > = 1 + (x + x^2) + (x + x^2)^2 + (x + x^2)^3 + (x + x^2)^4 + ... > > = 1 + x + 2x^2 + 3x^3 + 5x^4 + ... > >Hey! Fibonacci numbers! It looks like the kth coefficient of >this generating function is just the kth Fibonacci number! > >Now, remember that generating functions have a factorial built into >them: > > |F_k| >|F|(x) = sum ----- x^k > k! > >So apparently in this example |F_k| is k! times the kth Fibonacci >number. Of course, k! is the number of ways to order a k-element set. >So apparently the kth Fibonacci number is just the number of ways to >chop a k-element set into parts of size 1 or 2. What's funny is how the choice of orderings introduces a factor of k! whose only purpose in life is to cancel the 1/k! in the definition of "generating function". This guy knew that besides the generating functions I was discussing - sometimes called "exponential generating functions" - there are some other generating functions - sometimes called "ordinary generating functions" - whose definition doesn't have that 1/k! in it. If I'd used those, I wouldn't have needed to play this cancellation game! I knew that already, but I didn't want to confuse people by introducing two flavors of generating function. But now that the subject has come up, I might as well say something about it. The way I like to think about it, structure types are really functors F: C -> Set where C is the category of finite sets and bijections. But we also have "structure types on ordered sets" (don't know a good name for them) F: D -> Set where D is the category of linearly ordered finite sets and order-preserving bijections. The exponential generating function applies to structure types, and is defined as above. The ordinary generating function applies to structure types on ordered sets, and is defined by |F|(x) = sum |F_k| x^k It has many of the same nice properties as the exponential generating function, as long as we careful to adapt everything to the category D. You can read all about this in the book by Bergeron et al cited in "week185". I claim that it's best to always insist on this viewpoint: exponential generating functions for structure types, ordinary generating functions for structure types on ordered sets. However, if you want to have fun (i.e. get confused) you can convert structure types into structure types on ordered sets, or vice versa, before you take the generating function! After all, there is a forgetful functor from D to C. This induces a functor from hom(C,Set) to hom(D,Set): given a structure on a set S, we automatically get a structure on any linearly ordered set we obtain by slapping an ordering on S. Furthermore, *this* functor has an adjoint - in fact, both right and left adjoints. In short, there are three ways to hop back and forth between structure types and structure types on ordered sets, which allow you to get very confused about which you are working with at any given moment. To add to the fun (i.e. confusion), there are some formulas relating the exponential generating functions of the former to the ordinary generating functions of the latter. I was implicitly using one of these above. So, if you want to become deconfused, you should figure out these formulas. And if you want to do it in an elegant way: Both "structure types" and "ordered structure types" form 2-rigs - i.e. categories with + and x, satisfying some obvious ring-ish axioms up to isomorphism, but without additive inverses. Let's call these 2-rigs hom(C,Set) and hom(D,Set). If we decategorify a 2-rig we get a rig, so there are rigs I'll call |hom(C,Set)| and |hom(D,Set)|. Elements of the first are just isomorphism classes of structure types; elements of the second are isomorphism classes of ordered structure types; in both cases the + and x operations are hopefully obvious. Now, the exponential generating function is best thought of as a rig homomorphism egf: |hom(C,Set)| -> N{{x}} where N{{x}} is the rig of formal power series where the coefficient of the nth term is a natural number divided by n!, while the ordinary generating function is best thought of as a rig homomorphism ogf: |hom(D,Set)| -> N[[x]] The relations between exponential and ordinary generating functions are really relations between the rigs |hom(C,Set)| and |hom(D,Set)|. And these, in turn, are REALLY relations between the 2-rigs hom(C,Set) and hom(D,Set). I've already said that there is a functor hom(C,Set) -> hom(D,Set) and two going the other way. The question is, which of these functors are 2-rig homomorphisms? I.e., which get along with + and x? These are the ones where there will be *very* nice relations between generating functions - namely, relations that get along with + and x. I leave this as a little puzzle, partially because I am too lazy to work out the answer and explain it nicely. But for category mavens, here's an extra hint. To see if these functors between hom(C,Set) and hom(D,Set) are 2-rig homomorphisms, we need to see whether they preserve + (colimits) and x (the monoidal structure). Preserving colimits is a very general question. Given a functor from D to C we always get three functors going between the categories hom(C,Set) and hom(D,Set), and the question is: which of these preserve colimits? Preserving the monoidal structure is a slightly less general (but still bloody frigging general!) question. The point is that C and D are monoidal categories and hom(C,Set) and hom(D,Set) get their multiplication from that, via a trick called "Day convolution", which is just a categorified version of ordinary convolution of functions. (By now I'm at Macquarie University in Australia, and Brian Day's office is right across the hall, so I had to say this.) So, here the question is: when you have a *monoidal* functor from D to C, as we do here, which of the three functors between hom(C,Set) and hom(D,Set) are monoidal with respect to Day convolution? As usual, I learned most of this category theory stuff from James Dolan, so any errors in the above are his fault, not mine. ----------------------------------------------------------------------- Previous issues of "This Week's Finds" and other expository articles on mathematics and physics, as well as some of my research papers, can be obtained at http://math.ucr.edu/home/baez/ For a table of contents of all the issues of This Week's Finds, try http://math.ucr.edu/home/baez/twf.html A simple jumping-off point to the old issues is available at http://math.ucr.edu/home/baez/twfshort.html If you just want the latest issue, go to http://math.ucr.edu/home/baez/this.week.html