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.

September 15, 2024

The Space of Physical Frameworks (Part 4)

Posted by John Baez

In Part 1, I explained my hopes that classical statistical mechanics reduces to thermodynamics in the limit where Boltzmann’s constant kk approaches zero. In Part 2, I explained exactly what I mean by ‘thermodynamics’. I also showed how, in this framework, a quantity called ‘negative free entropy’ arises as the Legendre transform of entropy.

In Part 3, I showed how a Legendre transform can arise as a limit of something like a Laplace transform.

Today I’ll put all the puzzle pieces together. I’ll explain exactly what I mean by ‘classical statistical mechanics’, and how negative free entropy is defined in this framework. Its definition involves a Laplace transform. Finally, using the result from Part 3, I’ll show that as k0k \to 0, negative free entropy in classical statistical mechanics approaches the negative free entropy we’ve already seen in thermodynamics!

Thermodynamics versus statistical mechanics

In a certain important approach to thermodynamics, called classical thermodynamics, we only study relations between the ‘macroscopic observables’ of a system. These are the things you can measure at human-sized distance scales, like the energy, temperature, volume and pressure of a canister of gas. We don’t think about individual atoms and molecules! We say the values of all the macroscopic observables specify the system’s macrostate. So when I formalized thermodynamics using ‘thermostatic systems’ in Part 2, the ‘space of states’ XX was really a space of macrostates. Real-valued functions on XX were macroscopic observables.

I focused on the simple case where the macrostate is completely characterized by a single macroscopic observable called its energy E[0,)E \in [0,\infty). In this case the space of macrostates is X=[0,)X = [0,\infty). If we can understand this case, we can generalize later.

In classical statistical mechanics we go further and consider the set Ω\Omega of microstates of a system. The microstate specifies all the microscopic details of a system! For example, if our system is a canister of helium, a microstate specifies the position and momentum of each atom. Thus, the space of microstates is typically a high-dimensional manifold — where by ‘high’ I mean something like 10 2310^{23}. On the other hand, the space of macrostates is often low-dimensional — where by ‘low’ I mean something between 1 and 10.

To connect thermodynamics to classical statistical mechanics, we need to connect macrostates to microstates. The relation is that each macrostate is a probability distribution of microstates: a probability distribution that maximizes entropy subject to constraints on the expected values of macroscopic observables.

To see in detail how this works, let’s focus on the simple case where our only macroscopic observable is energy.

Classical statistical mechanical systems

Definition. A classical statistical mechanical system is a measure space (Ω,μ)(\Omega,\mu) equipped with a measurable function

H:Ω[0,) H \colon \Omega \to [0,\infty)

We call Ω\Omega the set of microstates, call HH the Hamiltonian, and call H(x)H(x) the energy of the microstate xΩx \in \Omega.

It gets tiring to say ‘classical statistical mechanical system’, so I’ll abbreviate this as classical stat mech system.

When we macroscopically measure the energy of a classical stat mech system to be EE, what’s really going on is that the system is in a probability distribution of microstates for which the expected value of energy is EE. A probability distribution is defined to be a measurable function

p:Ω[0,) p \colon \Omega \to [0,\infty)

with

Ωp(x)dμ(x)=1 \displaystyle{ \int_\Omega p(x) \, d\mu(x) = 1 }

The expected energy in this probability distribution is defined to be

H= ΩH(x)p(x)dμ(x) \displaystyle{ \langle H \rangle = \int_\Omega H(x) \, p(x) \, d \mu(x) }

So what I’m saying is that pp must have

H=E \langle H \rangle = E

But lots of probability distributions have H=E\langle H \rangle = E. Which one is the physically correct one? It’s the one that maximizes the Gibbs entropy:

S=k Ωp(x)lnp(x)dμ(x) \displaystyle{ S = - k \int_\Omega p(x) \, \ln p(x) \, d\mu(x) }

Here kk is a unit of entropy called Boltzmann’s constant. Its value doesn’t affect which probability distribution maximizes the entropy! But it will affect other things to come.

Now, there may not exist a probability distribution pp that maximizes SS subject to the constraint H=E\langle H \rangle = E, but there often is — and when there is, we can compute what it is. If you haven’t seen this computation, you can find it in my book What is Entropy? starting on page 24. The answer is the Boltzmann distribution:

p(x)=e CH(x)/k Ωe CH(x)/kdμ(x) \displaystyle{ p(x) = \frac{e^{-C H(x)/k}}{\int_\Omega e^{-C H(x)/k} \, d \mu(x)} }

Here CC is a number called the inverse temperature. We have to cleverly choose its value to ensure H=E\langle H \rangle = E. That might not even be possible. But if we get that to happen, pp will be the probability distribution we seek.

The normalizing factor in the formula above is called the partition function

Z k(C)= Ωe CH(x)/kdμ(x) Z_k(C) = \int_\Omega e^{-C H(x)/k} \, d\mu(x)

and it turns out to be important in its own right. The integral may not always converge, but when it does not we’ll just say it equals ++\infty, so we get

Z k:[0,)[0,] Z_k \colon [0,\infty) \to [0,\infty]

One reason the partition function is important is that

klnZ k(C)=CHS - k \ln Z_k(C) = C \langle H \rangle - S

where H\langle H \rangle and SS are computed using the Boltzmann distribution for the given value of CC For a proof see pages 67–71 of my book, though beware that I use different notation. The quantity above is called the negative free entropy of our classical stat mech system. In my book I focus on a closely related quantity called the ‘free energy’, which is the negative free entropy divided by CC. Also, I talk about β=1/kT\beta = 1/k T instead of the inverse temperature C=1/TC = 1/T.

Let’s call the negative free entropy Ψ k(C)\Psi_k(C), so

Ψ k(C)=klnZ k(C)=kln Ωe CH(x)/kdμ(x) \displaystyle{ \Psi_k(C) = -k \ln Z_k(C) = - k \ln \int_\Omega e^{-C H(x)/k} \, d\mu(x) }

I’ve already discussed negative free entropy in Part 2, but that was for thermostatic systems, and it was defined using a Legendre transform. This new version of negative free entropy applies to classical stat mech systems, and we’ll see it’s defined using a Laplace transform. But they’re related: we’ll see the limit of the new one as k0k \to 0 is the old one!

The limit as k0k \to 0

To compute the limit of the negative free entropy Ψ k(C)\Psi_k(C) as k0k \to 0 it will help to introduce some additional concepts.

First, given a classical stat mech system with measure space (Ω,μ)(\Omega, \mu) and Hamiltonian H:ΩH \colon \Omega \to \mathbb{R}, let

ν(E)=μ({xΩ|H(x)E} \nu(E) = \mu(\{x \in \Omega \vert \; H(x) \le E \}

be the measure of the set of microstates with energy E\le E. This is an increasing function of EE \in \mathbb{R} which is right-continuous, so it defines a Lebesgue–Stieltjes measure ν \nu on the real line. Yes, I taught real analysis for decades and always wondered when I’d actually use this concept in my own work: today is the day!

The reason I care about this measure ν\nu is that it lets us rewrite the partition function as an integral over the nonnegative real numbers:

Z k(C)= 0 e CE/kdν(E) \displaystyle{ Z_k(C) = \int_0^\infty e^{-C E/k} \, d\nu(E) }

Very often the measure ν\nu is absolutely continuous, which means that

dν(E)=g(E)dE d\nu(E) = g(E) \, d E

for some locally integrable function g:g \colon \mathbb{R} \to \mathbb{R}. I will assume this from now on. We thus have

Z k(C)= 0 e CE/kg(E)dE \displaystyle{ Z_k(C) = \int_0^\infty e^{-C E/k} \, g(E) \, d E }

Physicists call gg the density of states because if we integrate it over some interval [E,E+ΔE][E, E + \Delta E] we get ‘the number of states’ in that energy range. At least that’s what physicists say. What we actually get is the measure of the set

{xX:EH(x)E+ΔE} \{x \in X: \; E \le H(x) \le E + \Delta E \}

Before moving on, a word about dimensional analysis. I’m doing physics, so my quantities have dimensions. In particular, EE and dEd E have units of energy, while the measure dν(E)d\nu(E) is dimensionless, so the density of states g(E)g(E) has units of energy-1.

This matters because right now I want to take the logarithm of g(E)g(E), yet the rules of dimensional analysis include a stern finger-wagging prohibition against taking the logarithm of a quantity unless it’s dimensionless. There are legitimate ways to bend these rules, but I won’t. Instead I’ll follow most physicists and introduce a constant with dimensions of energy, ww, called the energy width. It’s wise to think of this as an arbitrary small unit of energy. Using this we can make all the calculations to come obey the rules of dimensional analysis. If you find that ridiculous, you can mentally set ww equal to 1. In fact I’ll do that later at some point.

With that said, now let’s introduce the so-called microcanonical entropy, often called the Boltzmann entropy:

S micro(E)=kln(wg(E)) S_{\mathrm{micro}}(E) = k \ln (w g(E))

Here we are taking Boltzmann’s old idea of entropy as kk times the logarithm of the number of states and applying it to the density of states. This allows us to define an entropy of our system at a specific fixed energy EE. Physicists call the set of microstates with energy exactly equal to some number EE the microcanonical ensemble, and they say the microcanonical entropy is the entropy of the microcanonical ensemble. This is a bit odd, because the set of microstates with energy exactly EE typically has measure zero. But it’s a useful way of thinking.

In terms of the microcanonical entropy, we have

g(E)=1we S micro(E)/k \displaystyle{ g(E) = \frac{1}{w} e^{S_{\mathrm{micro}}(E)/k} }

Combining this with our earlier formula

Z k(C)= 0 e CE/kg(E)dE \displaystyle{ Z_k(C) = \int_0^\infty e^{-C E/k} g(E) \, d E }

we get this formula for the partition function:

Z k(C)= 0 e (CES micro(E))/kdEw \displaystyle{ Z_k(C) = \int_0^\infty e^{-(C E - S_{\mathrm{micro}}(E))/k} \, \frac{d E}{w} }

Now things are getting interesting!

First, the quantity CES micro(E)C E - S_{\mathrm{micro}}(E) should remind you of the formula we saw in Part 2 for the negative free entropy of a thermostatic system. Remember, that formula was

Ψ(C)=inf E(CES(E)) \Psi(C) = \inf_E (C E - S(E))

Second, we instantly get a beautiful formula for the negative free entropy of a classical stat mech system:

Ψ k(C)=klnZ k(C)=kln 0 e (CES micro(E))/kdEw \displaystyle{ \Psi_k(C) = - k \ln Z_k(C) = - k \ln \int_0^\infty e^{-(C E - S_{\mathrm{micro}}(E))/k} \, \frac{d E}{w} }

Now for the climax of this whole series so far. We can now prove that as k0k \to 0, the negative free entropy of a classical stat mech system approaches the negative free entropy of a thermostatic system!

To state and prove this result, I will switch to treating the microcanonical entropy S microS_{\mathrm{micro}} as fundamental, rather than defining it in terms of the density of states. This means we can now let k0k \to 0 while holding the function S microS_{\mathrm{micro}} fixed. I will also switch to units of energy where w=1w = 1. Thus, starting from S microS_{\mathrm{micro}}, I will define the negative free entropy Ψ k\Psi_k by

Ψ k(C)=kln 0 e (CES micro(E))/kdE \displaystyle{ \Psi_k(C) = - k \ln \int_0^\infty e^{-(C E - S_{\mathrm{micro}}(E))/k} \, d E}

We can now study the limit of Ψ k(C)\Psi_k(C) as k0k \to 0.

Main Result. Suppose S micro:[0,)S_{\mathrm{micro}} \colon [0,\infty) \to \mathbb{R} is a concave function with continuous second derivative. Suppose that for some C>0C \gt 0 the quantity CES micro(E)C E - S_{\mathrm{micro}}(E) has a unique minimum as a function of EE, and S micro<0S''_{\mathrm{micro}} \lt 0 at that minimum. Then

lim k0Ψ k(C)=inf E(CES micro(E)) \displaystyle{ \lim_{k \to 0} \Psi_k(C) \quad = \quad \inf_E \left(C E - S_{\mathrm{micro}}(E)\right) }

The quantity at right deserves to be called the microcanonical negative free entropy. It’s the negative free entropy of the thermostatic system whose entropy function is S microS_{\mathrm{micro}}.

So, when the hypotheses hold,

As k0k \to 0, the free entropy of a classical statistical mechanical system approaches its microcanonical free entropy!

Here I’ve left off the word ‘negative’ twice, which is fine. But this sentence still sounds like a mouthful. Don’t feel bad if you find it confusing. But it could be the result we need to see how classical statistical mechanics approaches classical thermodynamics as k0k \to 0. So I plan to study this result further, and hope to explain it much better!

But today I’ll just prove the main result and quit. I figure it’s good to get the math done before talking more about what it means.

Proof of the main result

Suppose all the hypotheses of the main result hold. Spelling out the definition of the negative free entropy Ψ k(C)\Psi_k(C), what we need to show is

lim k0kln 0 e (CES micro(E))dE=inf E(CES micro(E)) \displaystyle{ \lim_{k \to 0} - k \ln \int_0^\infty e^{-(C E - S_{\mathrm{micro}}(E))} \, d E \quad = \quad \inf_E \left(C E - S_{\mathrm{micro}}(E)\right) }

To do this, we use a theorem from Part 3. My argument for that theorem was not a full mathematical proof — I explained the hole I still need to fill — so I cautiously called it an ‘almost proved theorem’. Here it is:

Almost Proved Theorem. Suppose that f:[0,)f \colon [0,\infty) \to \mathbb{R} is a concave function with continuous second derivative. Suppose that for some s>0s \gt 0 the function sxf(x)s x - f(x) has a unique minimum at x 0x_0, and f(x 0)<0f''(x_0) \lt 0. Then as β+\beta \to +\infty we have

lim β+1βln 0 e β(sxf(x))dx=inf x(sxf(x)) \displaystyle{ \lim_{\beta \to +\infty} -\frac{1}{\beta} \ln \int_0^\infty e^{-\beta (s x - f(x))} \, d x \; = \; \inf_x \left( s x - f(x)\right) }

Now let’s use this to prove our main result! To do this, take

s=C,x=E,f(x)=S micro(E),β=1/k s = C, \quad x = E, \quad f(x) = S_{\mathrm{micro}}(E), \quad \beta = 1/k

Then we get

lim k0kln 0 e (CES micro(E))/kdE=inf E(CES micro(E)) \displaystyle{\lim_{k \to 0} - k \ln \int_0^\infty e^{(C E - S_{\mathrm{micro}}(E))/k} \, d E \quad = \quad \inf_E \left(C E - S_{\mathrm{micro}}(E) \right) }

and this is exactly what we want.       ∎

Posted at September 15, 2024 8:00 PM UTC

TrackBack URL for this Entry:   https://golem.ph.utexas.edu/cgi-bin/MT-3.0/dxy-tb.fcgi/3557

5 Comments & 1 Trackback

Re: The Space of Physical Frameworks (Part 4)

Did you swap from using w to using k for the energy width in the proof at the end?

Posted by: Layra on September 16, 2024 3:59 AM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 4)

Thanks for noticing! I wrote kk a few times when I meant ww, which is very confusing since kk is Boltzmann’s constant and it appears all over the place. I think I’ve fixed them all now.

Posted by: John Baez on September 16, 2024 5:22 AM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 4)

Long time lurking chemist here with a bit of theoretical chemistry under his belt. The series, as well as the recent book on entropy vaguely reminds me of one of Lisi‘s papers, where he draws some parallels between path integrals in qm and statistical thermodynamics.

It is probably known in your circles anyway, but I thought i put it just out there:

Posted by: Matt on September 17, 2024 8:59 PM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 4)

Thanks! I’m friends with Garrett Lisi, and he brought this paper to my attention when I started blogging about the analogy between path integrals in quantum mechanics and state sums in statistical mechanics. I mention his paper here:

Quantropy (part 4), Azimuth, November 11, 2013.

See also his comments on this blog article!

There’s a lot left to figure out, that’s for sure.

Posted by: John Baez on September 30, 2024 9:51 PM | Permalink | Reply to this

Re: The Space of Physical Frameworks (Part 4)

I fixed a subtle problem in my writeup. I had defined the microcanonical entropy S microS_{\mathrm{micro}} in terms of the density of states in a way which depends on Boltzmann’s constant kk. Then my ‘main result’ involved the k0k \to 0 limit of a quantity Ψ k\Psi_k defined in terms of S microS_{\mathrm{micro}}. But when taking this limit, I held S microS_{\mathrm{micro}} fixed, ignoring its dependence on kk.

In fact this is the right thing to do, but first one needs to come out and say one is changing one’s attitude toward S microS_{\mathrm{micro}}!

Now I say this:

Ψ k(C)=kln 0 e (CES micro(E))/kdEw \displaystyle{ \Psi_k(C) = - k \ln \int_0^\infty e^{-(C E - S_{\mathrm{micro}}(E))/k} \, \frac{d E}{w} }

Using this we can prove that as k0k \to 0, the negative free entropy of a classical stat mech system approaches the negative free entropy we’ve already seen in thermodynamics!

To state and prove this result, I will switch to treating the microcanonical energy S microS_{\mathrm{micro}} as fundamental, rather than defining it in terms of the density of states. This means I can now let k0k \to 0 while holding the function S microS_{\mathrm{micro}} fixed. I will also switch to units of energy where w=1w = 1. Thus, starting from S microS_{\mathrm{micro}}, I will define the negative free entropy Ψ k\Psi_k by

Ψ k(C)=kln 0 e (CES micro(E))/kdE \displaystyle{ \Psi_k(C) = - k \ln \int_0^\infty e^{-(C E - S_{\mathrm{micro}}(E))/k} \, d E}

We can now study the limit of Ψ k(C)\Psi_k(C) as k0k \to 0.

Posted by: John Baez on September 30, 2024 9:28 PM | Permalink | Reply to this
Read the post The Space of Physical Frameworks (Part 5)
Weblog: The n-Category Café
Excerpt: Let's think about how classical statistical mechanics reduces to thermodynamics in the limit where Boltzmann's constant \(k\) approaches zero, by looking at an example.
Tracked: October 2, 2024 4:14 AM

Post a New Comment