Skip to the Main Content

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

August 31, 2024

The Space of Physical Frameworks (Part 1)

Posted by John Baez

Besides learning about individual physical theories, students learn different frameworks in which physical theories are formulated. I’m talking about things like this:

  • classical statics
  • classical mechanics
  • quantum mechanics
  • thermodynamics
  • classical statistical mechanics
  • quantum statistical mechanics

A physical framework often depends on some physical constants that we can imagine varying, and in some limit one framework may reduce to another. This suggests that we should study a ‘moduli space’ or ‘moduli stack’ of physical frameworks. To do this formally, in full generality, we’d need to define what counts as a ‘framework’, and what means for two frameworks to be equivalent. I’m not ready to try that yet. So instead, I want to study an example: a 1-parameter family of physical frameworks that includes classical statistical mechanics — and, I hope, also thermodynamics!

Physicists often say things like this:

“Special relativity reduces to Newtonian mechanics as the speed of light, cc, approaches \infty.”

“Quantum mechanics reduces to classical mechanics as Planck’s constant \hbar approaches 00.”

“General relativity reduces to special relativity as Newton’s constant GG approaches 00.”

Sometimes they try to elaborate this further with a picture called Bronstein’s cube or the CGh cube:

This is presumably hinting at some 3-dimensional space where 1/c,1/c, \hbar and GG can take arbitrary nonnegative values. This would be an example of what I mean by a ‘moduli space of physical frameworks’.

But right now I want to talk about talk about a fourth dimension that’s not in this cube. I want to talk about whether classical statistical mechanics reduces to thermodynamics as k0k \to 0, where kk is Boltzmann’s constant.

Since thermodynamics and statistical mechanics are often taught in the same course, you may be wondering how I distinguish them. Here are my two key principles: anything that involves probability theory or Boltzmann’s constant I will not call thermodynamics: I will call it statistical mechanics. For example, in thermodynamics we have quantities like energy EE, entropy SS, temperature TT, obeying rules like

dE=TdS d E = T d S

But in classical statistical mechanics EE becomes a random variable and we instead have

dE=TdS d \langle E \rangle = T d S

In classical statistical mechanics we can also compute the variance of EE, and this is typically proportional to Boltzmann’s constant. As k0k \to 0, this variance goes to zero and we’re back to thermodynamics! Also, in classical statistical mechanics entropy turns out to be given by

S=k Xp(x)ln(p(x))dμ(x) S = - k \int_X p(x) \ln(p(x)) \, d\mu(x)

where pp is some probability distribution on some measure space of states (X,μ)(X,\mu).

I want to flesh out how classical statistical mechanics reduces to thermodynamics as k0k \to 0, and my hope is that this is quite analogous to how quantum mechanics reduces to classical mechanics as 0\hbar \to 0:

taking the 0\hbar \to 0 limit of taking the k0k \to 0 limit of
Quantum Mechanics Classical Statistical Mechanics
gives gives
Classical Mechanics Thermodynamics

Here’s the idea. Quantum fluctuations are a form of randomness inherent to quantum mechanics, described by complex amplitudes. Thermal fluctuations are a form of randomness inherent to classical statistical mechanics, described by real probabilities. Planck’s constant \hbar sets the scale of quantum fluctuations, and as 0\hbar \to 0 these go away and quantum mechanics reduces to classical mechanics. Boltzmann’s constant kk sets the scale of thermal fluctuations, and as k0k \to 0 these go away and classical statistical mechanics reduces to thermodynamics.

If this idea works, the whole story for how quantum mechanics reduces to classical mechanics as 0\hbar \to 0 may be a Wick rotated version of how classical statistical mechanics reduces to thermodynamics as k0k \to 0. In other words, the two stories may be formally the same if we replace kk everywhere with ii \hbar.

However, there are many obstacles to getting the idea to to work — or at least apparent obstacles, much as walls can feel like ‘obstacles’ when you’re trying to walk through a wide open door at night. Even before we meet the technical problems with Wick rotation, there’s the preliminary problem of getting thermodynamics to actually arise as the k0k \to 0 limit of classical statistical mechanics!

So despite the grand words above, it’s that preliminary problem that I’m focused on now. It’s actually really interesting.

Today I’ll just give a bit of background.

The math: deformation of rigs

Deformation quantization, is a detailed analysis of how quantum mechanics reduces to classical mechanics as 0\hbar \to 0, and how you can try to reverse this process. If you’ve thought about this a lot, may have bumped into ‘idempotent analysis’: a version of analysis where you use minimization as a replacement for addition of real numbers, and addition as a replacement for multiplication. This works because addition distributes over minimization:

x+(yminz)=(x+y)min(x+z) x + (y \, \min \, z) = (x + y) \, \min \, (x + z)

and it’s called ‘idempotent’ because minimization obeys

xminx=x x \, \min \, x = x

When we use minimization as addition we don’t get additive inverses, so numbers form a rig, meaning a ‘ring without negatives’. You also need to include ++\infty to serve as an additive identity for minimization.

Idempotent analysis overlaps with ‘tropical mathematics’, where people use this number system, called the ‘tropical rig’, to simplify problems in algebraic geometry. People who do idempotent analysis are motivated more by applications to physics:

The basic idea is to study a 1-parameter family of rigs R βR_\beta which for finite β>0\beta \gt 0 are all isomorphic to (,](-\infty,\infty] with its usual addition and multiplication, but in the limit β+\beta \to +\infty approach a rig isomorphic to [0,)[0,\infty) with ‘min’ as addition and the usual ++ as multiplication.

Let me describe this in more detail, so you can see exactly how it works. In classical statistical mechanics, the probability of a system being in a state of energy EE decreases exponentially with energy, so it’s proportional to

e βE e^{-\beta E}

where β>0\beta \gt 0 is some constant we’ll discuss later. Let’s write

f β(E)=e βE f_\beta(E) = e^{-\beta E}

but let’s extend f βf_\beta to a bijection

f β:(,][0,)f_\beta: (-\infty, \infty] \to [0,\infty)

sending \infty to 00. This says that states of infinite energy have probability zero.

Now let’s conjugate ordinary addition and multiplication, which make [0,)[0,\infty) into a rig, by this bijection f βf_\beta:

x βy=f β 1(f β(x)+f β(y)) x \oplus_\beta y = f_\beta^{-1} (f_\beta(x) + f_\beta(y))   x βy=f β 1(f β(x)f β(y)) x \odot_\beta y = f_\beta^{-1} (f_\beta(x) \cdot f_\beta(y))

These conjugated operations make (,](-\infty,\infty] into a rig. Explicitly, we have

x βy=1βln(e βx+e βy) x \oplus_\beta y = -\frac{1}{\beta} \ln(e^{-\beta x} + e^{-\beta y})   x βy=x+y x \odot_\beta y = x + y

So the multiplication is always the usual ++ on (,](-\infty,\infty] — yes, I know this is confusing — while the addition is some more complicated operation that depends on β\beta. We get different rig structures on (,](-\infty,\infty] for different values of β>0\beta \gt 0, but these rigs are all isomorphic because we got them all from the same rig structure on [0,)[0,\infty).

However, now we can take the limit as β+\beta \to +\infty and get operations we call \oplus_\infty and \odot_\infty. If we work these out we get

x y=xminy x \oplus_\infty y = x \min y   x y=x+y x \odot_\infty y = x + y

These give a rig structure on (,](-\infty,\infty] that’s not isomorphic to any of those for finite β\beta. This is the tropical rig.

(Other people define the tropical rig differently, using different conventions, but usually theirs are isomorphic to this one.)

The physics: classical statistical mechanics

What does all of this mean for classical statistical mechanics? The idea is that [0,)[0,\infty) with its usual ++ and ×\times is the rig of unnormalized probabilities. I’ll assume you know why we add probabilities for mutually exclusive events and multiply probabilities for independent events. But probabilities lie in [0,1][0,1], which is not closed under addition. To get a rig, we work instead with ‘unnormalized’ probabilities, which lie in [0,)[0,\infty). We add and multiply these just like probabilities. When we have a list of unnormalized probabilities p 1,,p np_1, \dots, p_n, we can convert them to probabilities by dividing each one by their sum. We do this normalization only after all the addition and multiplication is done and we want to make predictions.

In classical statistical mechanics, a physical system has many states, each with its own energy E(,]E \in (-\infty,\infty]. The unnormalized probability that the system is in a state of energy EE is

f β(E)=e βE f_\beta(E) = e^{-\beta E}

The case of infinite energy is not ordinarily considered, but it’s allowed by the the math here, and this gives an unnormalized probability of zero.

We can reason with these unnormalized probabilities using addition and multiplication — or, equivalently, we can work directly with the energies EE using the operations β\oplus_\beta and β\odot_\beta on (,] (-\infty,\infty].

In short, we’ve enhanced the usual machinery of probability theory by working with unnormalized probabilities, and then transferred it over to the world of energies.

The physical meaning of β+\beta \to +\infty

All very nice. But what’s the physical meaning of β\beta and the β+\beta \to +\infty limit? This is where things get tricky.

First, what’s β\beta? In physics we usually take

β=1kT \beta = \frac{1}{k T}

where TT is temperature and kk is Boltzmann’s constant. Boltzmann’s constant has units of energy/temperature — it’s about 1.3810 231.38 \cdot 10^{-23} joules per kelvin — so in physics we use it to convert between energy and temperature. kTk T has units of energy, so β\beta has units of 1/energy, and βE\beta E is dimensionless. That’s important: we’re only allowed to exponentiate dimensionless quantities, and we want e βEe^{-\beta E} to make sense.

One can imagine doing physics using some interpretation of the deformation parameter β\beta other than 1/kT1/k T. But let’s take β\beta to be 1/kT1/k T. Then we can still understand the β+\beta \to +\infty limit in more than one way! We can

  1. hold kk constant and let T0T \to 0
  2. hold TT constant and let k0k \to 0.

We could also try other things, like simply letting kk and TT do whatever they want as long as their product approaches zero. But let’s just consider these two options.

Hold kk constant and let T0T \to 0

This option seems to make plenty of sense. It’s called Boltzmann’s ‘constant’, after all. So maybe we should hold it constant and let TT approach zero. In this case we’re taking the low temperature limit of classical statistical mechanics.

It’s sad that we get the tropical rig as a low-temperature limit: it should have been called the arctic rig! But the physics works out well. At temperature TT, systems in classical statistical mechanics minimize their free energy ETSE - T S where EE is energy and SS is entropy. As T0T \to 0, free energy reduces to simply the energy, EE. Thus, in the low temperature limit, such systems always try to minimize their energy! In this limit we’re doing classical statics: the classical mechanics of systems at rest.

These ideas let us develop this analogy:

taking the 0\hbar \to 0 limit of taking the T0T \to 0 limit of
Quantum Mechanics Classical Statistical Mechanics
gives gives
Classical Mechanics Classical Statics

Blake Pollard and I explored this analogy extensively, and discovered some exciting things:

  • Quantropy:

    • Part 1: the analogy between quantum mechanics and statistical mechanics, and the quantum analogue of entropy: quantropy.

    • Part 2: computing the quantropy of a quantum system starting from its partition function.

    • Part 3: the quantropy of a free particle.

    • Part 4: a paper on quantropy, written with Blake Pollard.

But while this analogy is mathematically rigorous and leads to new insights, it’s not completely satisfying. First, \hbar and TT just feel different. Planck’s constant takes the same value everywhere while temperature is something we can control. It would be great if next to the thermostat on your wall there was a little box where you could adjust Planck’s constant, but it just doesn’t work that way!

Second, there’s a detailed analogy between classical mechanics and thermodynamics, which I explored here:

  • Classical mechanics versus thermodynamics:

    • Part 1: Hamilton’s equations versus the Maxwell relations.

    • Part 2: the role of symplectic geometry.

    • Part 3: a detailed analogy between classical mechanics and thermodynamics.

    • Part 4: what is the analogue of quantization for thermodynamics?

Most of this was about how classical mechanics and thermodynamics share common mathematical structures, like symplectic and contact geometry. These structures arise naturally from variational principles: principle of least action in classical mechanics, and the principle of maximum entropy in thermodynamics. By the end of this series I had convinced myself that thermodynamics should appear as the k0k \to 0 limit of some physical framework, just as classical mechanics appears as the 0\hbar \to 0 limit of quantum mechanics. So, let’s look at option 2.

Hold TT constant and let k0k \to 0

Of course, my remark about holding kk constant because it’s called Boltzmann’s constant was just a joke. Planck’s constant is a constant too, yet it’s very fruitful to imagine treating it as a variable and letting it approach zero.

But what would it really mean to let k0k \to 0? Well, first let’s remember what it would mean to let 0\hbar \to 0.

I’ll pick a specific real-world example: a laser beam. Suppose you’re a physicist from 1900 who doesn’t know quantum mechanics, who is allowed to do experiments on a beam of laser light. If you do crude measurements, this beam will look like a simple solution to the classical Maxwell equations: an electromagnetic field oscillating sinusoidally. So you will think that classical physics is correct. Only when you do more careful measurements will you notice that this wave’s amplitude and phase are a bit ‘fuzzy’: you get different answers when you repeatedly measure these, no matter how good the beam is and how good your experimental apparatus is. These are ‘quantum fluctuations’, related to the fact that light is made of photons.

The product of the standard deviations of the amplitude and phase is bounded below by something proportional to \hbar. So, we say that \hbar sets the scale of the quantum fluctuations. If we could let 0\hbar \to 0, these fluctuations would become ever smaller, and in the limit the purely classical description of the situation would be exact.

Something similar holds with Boltzmann’s constant. Suppose you’re a physicist from 1900 who doesn’t know about atoms, who has somehow been given the ability to measure the pressure of a gas in a sealed container with arbitrary accuracy. If you do crude measurements, the pressure will seem to be a function of the temperature, the container’s volume, and the amount of gas in the container. This will obey the laws of thermodynamics. Only when you do extremely precise experiments will you notice that the pressure is fluctuating. These are ‘thermal fluctuations’, caused by the gas being made of molecules.

The variance of the pressure is proportional to kk. So, we say that kk sets the scale of the thermal fluctuations. If we could let k0k \to 0, these fluctuations would become ever smaller, and in the limit the thermodynamics description of the situation would be exact.

The analogy here is a bit rough at points, but I think there’s something to it. And if you examine the history, you’ll see some striking parallels. Einstein discovered that light is made of photons: he won the Nobel prize for his 1905 paper on this, and it led to a lot of work on quantum mechanics. But Einstein also wrote a paper in 1905 showing how to prove that liquid water is made of atoms! The idea was to measure the random Brownian motion of a grain of pollen in water — that is, thermal fluctuations. In 1908, Jean Perrin carried out this experiment, and he later won the Nobel “for his work on the discontinuous structure of matter”. So the photon theory of light and the atomic theory of matter both owe a lot to Einstein’s work.

Planck had earlier introduced what we now call Planck’s constant in his famous 1900 paper on the quantum statistical mechanics of light, without really grasping the idea of photons. Remarkably, this is also the paper that first introduced Boltzmann’s constant kk. Boltzmann had the idea that entropy was proportional to the logarithm of the number of occupied states, but he never estimated the constant of proportionality or gave it a name: Planck did both! So Boltzmann’s constant and Planck’s constant were born hand in hand.

There’s more to say about how to correctly take the 0\hbar \to 0 limit of a laser beam: we have to simultaneously increase the expected number of photons, so that the rough macroscopic appearance of the laser beam remains the same, rather than becoming dimmer. Similarly, we have to be careful when taking the k0k \to 0 limit of a container of gas: we need to also increase the number of molecules, so that the average pressure remains the same instead of decreasing.

Getting these details right has exercised me quite a bit lately. This is what I want to talk about.

Posted at August 31, 2024 9:30 PM UTC

TrackBack URL for this Entry:   https://golem.ph.utexas.edu/cgi-bin/MT-3.0/dxy-tb.fcgi/3553

8 Comments & 1 Trackback

Re: The Space of Physical Frameworks (Part 1)

That kh 1~20.836×10 9kh^{-1} ~ 20.836 \times 10^9 Hertz per Kelvin is {\bf not} an incomprehensibly-sized ratio seems significant to me.

Inverting the first Newtonian power sum

(1)p 1= k=1 x i=e 1Z[e i|i1] p_1 = \sum_{k=1}^\infty x_i = e_1 \in Z[e_i|i \geq 1]

in the (graded Hopf) algebra of elementary symmetric functions (roughly, assuming that the sum is generically nonzero) defines a (filtered but not graded) ring Z[ *]Z [\wp_*] of `renormalized’ symmetric functions

(2) k=p 1 kp k \wp_k = p_1^{-k} p_k

that seems sometimes to be useful in variations of classical statistical mechanics related to Boltzmannish-looking equations such as

(3) k=0 e kt k=exp( k=1 (1) k+1kp kt k) \sum _{k=0}^{\infty} e_{k}\,t^{k} = \exp \left(\sum _{k=1}^{\infty }{\frac{(-1)^{k+1}}{k}}p_{k}\,t^{k}\right)

\dots

Posted by: jack morava on September 1, 2024 4:08 PM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 1)

I remembered that my distinction between ‘thermodynamics’ and ‘classical statistical mechanics’ was confusing to someone I spoke to, so I added this:

Since thermodynamics and statistical mechanics are often taught in the same course, you may be wondering how I distinguish them. Here are my two key principles: anything that involves probability theory or Boltzmann’s constant I will not call thermodynamics: I will call it statistical mechanics. For example, in thermodynamics we have quantities like energy EE, entropy SS, temperature TT, obeying rules like

dE=TdS d E = T d S

But in classical statistical mechanics EE becomes a random variable and we instead have

dE=TdS d \langle E \rangle = T d S

In classical statistical mechanics we can also compute the variance of EE, and this is typically proportional to Boltzmann’s constant. As k0k \to 0, this variance goes to zero and we’re back to thermodynamics! Also, in classical statistical mechanics entropy turns out to be given by

S=k Xp(x)ln(p(x))dμ(x) S = - k \int_X p(x) \ln(p(x)) \, d\mu(x)

where pp is some probability distribution on some measure space of states (X,μ)(X,\mu).

Posted by: John Baez on September 1, 2024 7:15 PM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 1)

I’ve been annoyed about the tropical vs. arctic thing for a long time. Incidentally, there are old papers (mostly by Japanese authors) that use the terms “tropicalization”/”detropicalization” in the physically reasonable direction you mention. But I think the usage has crystalized along the defective direction at this point.

Black-body radiation involves the ratio k/hbar that Jack Morava mentions. So you’ve got a whole slope worth of choices how to take them both to 0 simultaneously. What do those look like?

Posted by: Allen Knutson on September 1, 2024 7:36 PM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 1)

I wish someone would thaw out the crystallized bad usage of ‘tropical’, change it, and then just chill. But it seems hopeless.

Right now I seem to be studying the ray in (,k)(\hbar,k) quadrant where =0\hbar = 0 and k0k \ge 0 varies. If I can get that under control maybe I can bring them both in. We should get quantum statistical mechanics where the amount of ‘quantumness’ and the amount of ‘statisticalness’ are independently adjustable.

But so far the ray I’m studying has been a lot more problematic than I expected! The reason is that when varying kk you need to adjust other quantities as well, in the correct manner, or the k0k \to 0 limit is very degenerate.

Posted by: John Baez on September 1, 2024 8:00 PM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 1)

There’s a cc vs \hbar quadrant too, come to think of it, well-established in geometric quantization theory - where cc is not the speed of light but a Chern class or topological charge, with large \hbar corresponding to the semiclassical case and low cc corresponding to the quantum case.

People interested in zeta-functions in number theory as counting functions [Connes, Consani,…] are concerned with similar quirky behavior as the primes get small, IIUC ? The issue seems to have something to do with transitions as small groups of things cluster into larger groups of things to form `matter’.

Posted by: jack morava on September 2, 2024 8:04 PM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 1)

There is a typo in the penultimate paragraph: the intriguing ‘nesured’ should be ‘need’ in “we nesured to also increase the number of molecules”.

Posted by: L Spice on September 2, 2024 3:54 AM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 1)

Hmm, maybe I had measures on my brain? Thanks!

Posted by: John Baez on September 2, 2024 9:15 PM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 1)

Very fascinating post, John. What nice little fact that Plank introduced both his and Boltzmann’s constants!

Anyway, the 1-parameter family of rigs you talk about features predominantly in a recent preprint of mine–indeed, if you have references about that family, I’d be interested to read what’s known about them.

The aforementioned work is very preliminary, but it seems crucial that some operation like addition are ‘non-classical’ versions of the ‘classical’ connectives max/min. More excitingly, although I didn’t show it there, maximum entropy principle and Occam’s razors are indeed linked in that setting. By expressing abduction in terms of Kan lifts (an idea I stole from David Corfield) in a putative equipment of quantitive relations, you can show these have the form of Boltzmann distributions. Such an equipment is, however, not a well-defined one at the moment, but I’m confident it can be adequately fixed (I’ll be working on this next year).

So physics, logic… is a new Rosetta stone coming to light?

Posted by: Matteo Capucci on September 3, 2024 12:37 PM | Permalink | Reply to this
Read the post The Space of Physical Frameworks (Part 2)
Weblog: The n-Category Café
Excerpt: What's a thermostatic system, and what are some basic things we can do with them?
Tracked: September 9, 2024 4:04 PM

Post a New Comment