What's on my mind? March 2019

What's going on in the curious canteloupe of my head? Lots of bizarre things pass through my mind, and I thought it'd be fruitful to document the sorts of topics I'm thinking about. This week, the theme is trajectories in high-dimensional spaces…and a little taste of complex anlysis and Fourier Transforms as well.

Statistical Mechanics and Ergodicity

Here's a question that's been nagging at me for years. When we do statistical mechanics, we take averages over phase space langle A rangle = int A(x,p) dx , dp, but in reality, Nature performs averages over time, overline A = int_0^T A(t) dt. So why the hell do the time averages we observe have anything to do with the abstract ensemble averages that we compute in statistical mechanics?

Furthermore, if we think about an ensemble as a probability distribution over phase space, then the laws of time evolution do a terrible job of sampling from this probability distribution! In particular, in a system of N particles, the phase space is 6N-dimensional (3 position coordinates, 3 momentum coordinates per particle). And thus this probability distribution is defined over a volume of O(e^N), a ridiculously huge number. On the other hand, the number of states sampled by a time trajectory is only of order O(N). So it's all the more puzzling why ensemble averages have anything to do with physical observables at all.

This week, I started rereading Shang-keng Ma's statistical mechanics book. He has a careful and unorthodox viewpoint towards our understanding of statistical mechanics that centers on the key difference between abstract ensemble averages and physical time averages. In particular, he stresses that the concept of equilibrium necessarily involves a timescale of observation over which macroscopic properties don't change.

For instance, he spends a whole chapter discussing impurities in a material,such as solutes dissolved in a (liquid) solution, or weird impurities inside the crystal lattice. Depending on the timescale of diffusion for these impurities, the thermodynamics of such systems are completely different. In one extreme, the mobile impurities thoroughly traverse the entire system during the observation timescale, so when we compute thermodynamic quantities, the relevant volume is V per particle, and we have to include a factor of 1/N! to account for indistinguishability. In the other extreme, the impurities are ‘‘trapped’’ and don't move around much during the observation timescale, so now the volume per impurity atom is V/N, and since they can't switch places, we don't need to introduce the 1/N! factor.

The first time I read through this book, a lot of the chapters didn't make much sense. But now, with a bit more statistical mechanical sophistication under my belt, I think this book is becoming far more rewarding to read and understand. Hopefully it will give me some satisfactory answers about this question of ergodicity.

Chaos theory

Often, people justify the ergodic assumption by appealing to chaos theory. So I checked out a book called ‘‘Chaos and Coarse Graining in Statistical Mechanics’’ from the library, and I'm skimming through to see whether I'm convinced by these arguments. I didn't read the book too carefully, because I don't think it's worth my time to understand all the details. But here were some key takeaways I got:

Here's a way to understand Lyapunov exponents. Say that we know the current state of the system to within an accuracy of delta_0, and we want to extrapolate the trajectory into the future. Because of the chaotic behavior of such a system, a tiny difference between two initial points can lead to an exponentially diverging distance between them. Hence we can only trust our predicted trajectory for a certain amount of time T_p, before our uncertainty in the initial state leads to complete ignorance. To be precise, if our maximum tolerable separation between trajectories is Delta, then we can only trust our predictions to within a time

 T_p sim frac 1 lambda log left(frac Delta {delta_0} right),

where lambda is the Lyapunov exponent, delta_0 is the initial separation between trajectories, Delta is our maximum toleration between trajectories, and T_p is the time range when we can trust our predictions.

A particularly philosophically pithy take-away point is that deterministic, chaotic systems look a lot like stochastic systems when you

For instance, turbulence is a completely deterministic phenomenon – if you know the velocity field at some point in time, you can apply the time-evolution laws to figure out what it is at some later point. Yet you can model turbulence as a random process!

On the other hand, the pseudo-random number generators in our computers are exactly deterministic: they result in the same ‘‘random’’ numbers if you seed them they same way. Yet their outpus have the statistical properties of a random sequence!

Fourier modes and translational invariance

I've wanted to build a better intuition of the Fourier Transform for the longest time, and finally, it's all sort of coming together and making sense. In particular, on my thermodynamics problem set this week, we're solving the classic problem of a ‘‘mattress’’ of coupled harmonic oscillators. We're given a bunch of masses in a square d-dimensional grid, each of which has a spring attached to its nearest neighbors in the form of H = sum_{<ij>} J (phi_i - phi_j)^2, and we want to find its thermodynamic properties.

The trick to solving this problem is to rewrite the energy in terms of normal modes. We've seen such an approach multiple times throughout our physics education – in finding the phonons of a crystal lattice, in solving general small-oscillation problems, in vibrating beer bottles, and more. But the interesting fact is that the modes which diagonalize this Hamiltonian turn out to be Fourier modes.

This Fourier approach seemed intuitive to me, but I don't know why. I'm struggling to remind myself of what examples I'm drawing an analogy to. Let me see…

Anyways, I found it a bit annoying that I couldn't remember where exactly I was drawing my intuition from. I'm a strong believer that ‘‘gut instinct’’ is built up from seeing specific examples, so hopefully I can come up with some solid example to explain why Fourier modes diagonalize translationally invariant systems.

High-Dimensional Spaces

There's lots of counter-intuitive behaviors of high-dimensional spaces that come into play in all areas of science. As humans living in a measly three-dimensional world, we're forced to reason by analogy into spaces of higher dimension, but sometimes, the results of our intuition just don't hold true.

Some concrete examples of high-dimensional spaces include:

Properties of hyperspheres

To think about high-dimensional spaces a bit more, let's consider the geometric object of a hypersphere, the set of all points that are equidistant from the origin in a N-dimensional space. Let's say that points in our space are described by coordinates vec x = (x_1, x_2, x_3, ldots, x_N), and that distance is measured by a typical L2 norm of |vec x|^2 = x_1^2 + x_2^2 + x_3^2 + ldots + x_N^2 = sum_i x_i^2. A hypersphere of radius R is given by the set of all points that are a distance R from the center; i.e., R^2 = sum_i x_i^2.

In Ma's statistical mechanics book, he discussed the non-intuitive properties of hyperspheres. For instance:

Saddle points

For some reason, I also stumbled into a paper from Surya Ganguli's lab about saddle points in high-dimensional spaces. Here the question about training statistical models with a huge buttload of parameters. While training neural networks with optimization algorithms such as stochastic gradient descent (SGD), researchers typically find that the loss function plateaus out. So there's lots of talk into better optimization schemes that don't get ‘‘stuck’’ in the landscape of the loss function.

Often we hear talk about ‘‘getting stuck in local minima’’. That is, if we think about the loss function as a hilly landscape in a high-dimensional parameter space, we imagine that it's got all sorts of hills and saddles and local minima. If we initialize the training trajectory at different places in this parameter space, we observe that the loss function converges to different values, suggesting that we get ‘‘stuck’’ in different minima. Typically, we mumble some justification such as ‘‘some minima are shitty local minima, whereas other are better minima’’.

However, this understanding about minima comes from our intuition of one-dimensional and two-dimensional functions, and it's not obvious whether the intuition in low dimensions can generalize to behavior in higher dimensions. And indeed, as this paper suggested, we've been chanting the wrong mantra for years; statistical models get appear to stuck in saddle points rather than in local minima, because of the nature of high-dimensional parameter space.

There's a simple argument for why saddle points are much more common than minima in high dimensions. Let's imagine that we're exploring the landscape of a high-dimensional function f, and that we've stumbled across a point where the first derivative partial_i f is zero in all directions (i.e. the gradient vanishes). Remember from multivariable calclus that such a point is either a local min, a local max, or a saddle point, depending on what the second derivative is. More precisely, the eigenvalues of the second derivative matrix partial_i partial_j f tell you whether the function decreases or increases as you walk in different directions. If all the eigenvalues are positive, then the function increases in all directions you walk and you're at a local minimum; if they're all negative, then it decreases in all directions and you're at a local max; and if some are positive and some are negative, then it increases or decreases depending on your direction, i.e., you're at a saddle point.

So the question of ‘‘am I at a minimum or at a saddle point’’ corresponds to counting how many of the eigenvalues of the second derivative matrix are positive or negative. Notice that you have to be very lucky to have eigenvalues that are all positive. If you assume that each eigenvalue has equal chances of being positive and negative, then the chance that all the N eigenvalues are positive is given by (1/2)^N, which is an exceedingly small number if N is on the order of millions! And thus, we expect minima to be extremely rare in high dimensions compared to saddle points.

Beyond the statistical unlikelihood of minima, the paper also argued that all minima tend to be around the same, and that the difference between good and mad local minima was dwarfed by the difference between critical points with different amounts of negative eigenvalues. That is, all minima are about the same, all first-order saddles (with one positive eigenvalue) are about the same, and so on.

To summarize, I find that it's rather intriguing to think about high dimensional spaces, mainly because we tend to reason by analogy into a regime where our intuition might actually be wrong.

Winding Numbers, Vortices, Poles, and all that jazz

I've also been dabbling in complex analysis again. For some reason, this flavor of math – back-seat topology, winding around poles, drawing loops and counting directions, conformal maps, projections onto spheres – it all reminds me days of the good old internet, with janky html websites and cute math applets, to simpler times with slower web browsers and non-pretentious little webpages…so I always a warm nostalgic feeling whenever I read about this kind of math.

It's often said that complex analysis is a beautiful subject, and I'm starting to see why it is! Analytic functions are so well behaved and satisfy so many great properties, and they tie together so many cute aspects of math: the topology of winding around poles, the group theory of mappings of a plane, and more…

This week, I finished a cursory scan-through of Visual Complex Analysis, which takes a very geometric and visual approach to the subject. It's a super relaxing storybook to read before going to bed ;)

To be quite frank, I don't identify too strongly with this flavor of math, and I'm not too sure how useful all the topics on projective geometry will be for my future. As a physicist, I think I'll mainly need to learn to do contour integrals, and maybe I'll need to understand analytic continuation as well. It might also be nice to learn some of the topological concepts of winding around vortices and all that jazz too. But it seems a bit excessive to prove everything by drawing pictures (although it's really cute!)

Anyways, I still don't have a very firm grasp of this subject, so I'm going to dive back through the book again, going through the derivations carefully and doing the interesting problems. I'm excited: complex analysis is a cute topic, and it brings up lots of intellectual nostalgia.


Back to my home page

Leave a Comment Below!

Comment Box is loading comments...