summaryrefslogtreecommitdiff
diff options
context:
space:
mode:
-rw-r--r--Gemfile1
-rw-r--r--README.md4
-rw-r--r--_config.yml2
-rw-r--r--source/know/concept/canonical-ensemble/index.md2
-rw-r--r--source/know/concept/electromagnetic-wave-equation/index.md347
-rw-r--r--source/know/concept/euler-equations/index.md2
-rw-r--r--source/know/concept/fundamental-relation-of-thermodynamics/index.md326
-rw-r--r--source/know/concept/fundamental-thermodynamic-relation/index.md54
-rw-r--r--source/know/concept/laws-of-thermodynamics/index.md104
-rw-r--r--source/know/concept/martingale/index.md2
-rw-r--r--source/know/concept/nonlinear-schrodinger-equation/index.md675
-rw-r--r--source/know/concept/ritz-method/index.md201
-rw-r--r--source/know/concept/thermodynamic-potential/index.md2
-rw-r--r--source/know/concept/triple-product-rule/index.md97
14 files changed, 1405 insertions, 414 deletions
diff --git a/Gemfile b/Gemfile
index 46a8096..5190e36 100644
--- a/Gemfile
+++ b/Gemfile
@@ -3,6 +3,7 @@ source "https://rubygems.org"
gem "jekyll"
gem "webrick"
gem "json"
+gem "erb"
# https://github.com/rubyjs/mini_racer
gem "mini_racer"
diff --git a/README.md b/README.md
index 75bb8fc..e9c112d 100644
--- a/README.md
+++ b/README.md
@@ -14,9 +14,13 @@ all its extensions needed by this website as listed in the `Gemfile`:
```sh
$ bundle config set --local path .bundle
+$ bundle config build.posix-spawn --with-cflags="-Wno-incompatible-pointer-types"
$ bundle install
```
+First try to run `bundle install` without changing any `CFLAGS`,
+it used to work like that, but not anymore as of mid-2024.
+
Now you should be able to run a local development version with:
```sh
diff --git a/_config.yml b/_config.yml
index 5023f33..250ed0c 100644
--- a/_config.yml
+++ b/_config.yml
@@ -16,7 +16,7 @@ kramdown:
macros:
"\\Real": "\\mathop{\\mathrm{Re}}"
"\\Imag": "\\mathop{\\mathrm{Im}}"
- "\\pv": "\\:\\mathop{\\mathcal{P}}"
+ "\\pv": "\\:\\mathcal{P}\\!"
"\\cross": "\\times"
"\\va": "\\vec{\\mathbf{#1}}"
"\\vb": "\\mathbf{#1}"
diff --git a/source/know/concept/canonical-ensemble/index.md b/source/know/concept/canonical-ensemble/index.md
index 8a96e91..da7d436 100644
--- a/source/know/concept/canonical-ensemble/index.md
+++ b/source/know/concept/canonical-ensemble/index.md
@@ -178,7 +178,7 @@ $$\begin{aligned}
\end{aligned}$$
Rearranging and substituting
-the [fundamental thermodynamic relation](/know/concept/fundamental-thermodynamic-relation/)
+the [fundamental thermodynamic relation](/know/concept/fundamental-relation-of-thermodynamics/)
then gives:
$$\begin{aligned}
diff --git a/source/know/concept/electromagnetic-wave-equation/index.md b/source/know/concept/electromagnetic-wave-equation/index.md
index a27fe6f..559d943 100644
--- a/source/know/concept/electromagnetic-wave-equation/index.md
+++ b/source/know/concept/electromagnetic-wave-equation/index.md
@@ -1,7 +1,7 @@
---
title: "Electromagnetic wave equation"
sort_title: "Electromagnetic wave equation"
-date: 2021-09-09
+date: 2024-09-08 # Originally 2021-09-09, major rewrite
categories:
- Physics
- Electromagnetism
@@ -9,236 +9,281 @@ categories:
layout: "concept"
---
-The electromagnetic wave equation describes
-the propagation of light through various media.
-Since an electromagnetic (light) wave consists of
+Light, i.e. **electromagnetic waves**, consist of
an [electric field](/know/concept/electric-field/)
and a [magnetic field](/know/concept/magnetic-field/),
-we need [Maxwell's equations](/know/concept/maxwells-equations/)
-in order to derive the wave equation.
+one inducing the other and vice versa.
+The existence and classical behavior of such waves
+can be derived using only [Maxwell's equations](/know/concept/maxwells-equations/),
+as we will demonstrate here.
-
-## Uniform medium
-
-We will use all of Maxwell's equations,
-but we start with Ampère's circuital law for the "free" fields $$\vb{H}$$ and $$\vb{D}$$,
-in the absence of a free current $$\vb{J}_\mathrm{free} = 0$$:
-
-$$\begin{aligned}
- \nabla \cross \vb{H}
- = \pdv{\vb{D}}{t}
-\end{aligned}$$
-
-We assume that the medium is isotropic, linear,
-and uniform in all of space, such that:
+We start from Faraday's law of induction,
+where we assume that the system consists of materials
+with well-known (linear) relative magnetic permeabilities $$\mu_r(\vb{r})$$,
+such that $$\vb{B} = \mu_0 \mu_r \vb{H}$$:
$$\begin{aligned}
- \vb{D} = \varepsilon_0 \varepsilon_r \vb{E}
- \qquad \quad
- \vb{H} = \frac{1}{\mu_0 \mu_r} \vb{B}
+ \nabla \cross \vb{E}
+ = - \pdv{\vb{B}}{t}
+ = - \mu_0 \mu_r \pdv{\vb{H}}{t}
\end{aligned}$$
-Which, upon insertion into Ampère's law,
-yields an equation relating $$\vb{B}$$ and $$\vb{E}$$.
-This may seem to contradict Ampère's "total" law,
-but keep in mind that $$\vb{J}_\mathrm{bound} \neq 0$$ here:
+We move $$\mu_r(\vb{r})$$ to the other side,
+take the curl, and insert Ampère's circuital law:
$$\begin{aligned}
- \nabla \cross \vb{B}
- = \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdv{\vb{E}}{t}
+ \nabla \cross \bigg( \frac{1}{\mu_r} \nabla \cross \vb{E} \bigg)
+ &= - \mu_0 \pdv{}{t} \big( \nabla \cross \vb{H} \big)
+ \\
+ &= - \mu_0 \bigg( \pdv{\vb{J}_\mathrm{free}}{t} + \pdvn{2}{\vb{D}}{t} \bigg)
\end{aligned}$$
-Now we take the curl, rearrange,
-and substitute $$\nabla \cross \vb{E}$$ according to Faraday's law:
+For simplicity, we only consider insulating materials,
+since light propagation in conductors is a complex beast.
+We thus assume that there are no free currents $$\vb{J}_\mathrm{free} = 0$$, leaving:
$$\begin{aligned}
- \nabla \cross (\nabla \cross \vb{B})
- = \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdv{}{t}(\nabla \cross \vb{E})
- = - \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{B}}{t}
+ \nabla \cross \bigg( \frac{1}{\mu_r} \nabla \cross \vb{E} \bigg)
+ &= - \mu_0 \pdvn{2}{\vb{D}}{t}
\end{aligned}$$
-Using a vector identity, we rewrite the leftmost expression,
-which can then be reduced thanks to Gauss' law for magnetism $$\nabla \cdot \vb{B} = 0$$:
+Having $$\vb{E}$$ and $$\vb{D}$$ in the same equation is not ideal,
+so we should make a choice:
+do we restrict ourselves to linear media
+(so $$\vb{D} = \varepsilon_0 \varepsilon_r \vb{E}$$),
+or do we allow materials with more complicated responses
+(so $$\vb{D} = \varepsilon_0 \vb{E} + \vb{P}$$, with $$\vb{P}$$ unspecified)?
+The former is usually sufficient:
$$\begin{aligned}
- - \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{B}}{t}
- &= \nabla (\nabla \cdot \vb{B}) - \nabla^2 \vb{B}
- = - \nabla^2 \vb{B}
+ \boxed{
+ \nabla \cross \bigg( \frac{1}{\mu_r} \nabla \cross \vb{E} \bigg)
+ = - \mu_0 \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{E}}{t}
+ }
\end{aligned}$$
-This describes $$\vb{B}$$.
-Next, we repeat the process for $$\vb{E}$$:
-taking the curl of Faraday's law yields:
+This is the general linear form of the **electromagnetic wave equation**,
+where $$\mu_r$$ and $$\varepsilon_r$$
+both depend on $$\vb{r}$$ in order to describe the structure of the system.
+We can obtain a similar equation for $$\vb{H}$$,
+by starting from Ampère's law under the same assumptions:
$$\begin{aligned}
- \nabla \cross (\nabla \cross \vb{E})
- = - \pdv{}{t}(\nabla \cross \vb{B})
- = - \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{E}}{t}
+ \nabla \cross \vb{H}
+ = \pdv{\vb{D}}{t}
+ = \varepsilon_0 \varepsilon_r \pdv{\vb{E}}{t}
\end{aligned}$$
-Which can be rewritten using same vector identity as before,
-and then reduced by assuming that there is no net charge density $$\rho = 0$$
-in Gauss' law, such that $$\nabla \cdot \vb{E} = 0$$:
+Taking the curl and substituting Faraday's law on the right yields:
$$\begin{aligned}
- - \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{E}}{t}
- &= \nabla (\nabla \cdot \vb{E}) - \nabla^2 \vb{E}
- = - \nabla^2 \vb{E}
+ \nabla \cross \bigg( \frac{1}{\varepsilon_r} \nabla \cross \vb{H} \bigg)
+ &= \varepsilon_0 \pdv{}{t} \big( \nabla \cross \vb{E} \big)
+ = - \varepsilon_0 \pdvn{2}{\vb{B}}{t}
\end{aligned}$$
-We thus arrive at the following two (implicitly coupled)
-wave equations for $$\vb{E}$$ and $$\vb{B}$$,
-where we have defined the phase velocity $$v \equiv 1 / \sqrt{\mu_0 \mu_r \varepsilon_0 \varepsilon_r}$$:
+And then we insert $$\vb{B} = \mu_0 \mu_r \vb{H}$$ to get the analogous
+electromagnetic wave equation for $$\vb{H}$$:
$$\begin{aligned}
\boxed{
- \pdvn{2}{\vb{E}}{t} - \frac{1}{v^2} \nabla^2 \vb{E}
- = 0
- }
- \qquad \quad
- \boxed{
- \pdvn{2}{\vb{B}}{t} - \frac{1}{v^2} \nabla^2 \vb{B}
- = 0
+ \nabla \cross \bigg( \frac{1}{\varepsilon_r} \nabla \cross \vb{H} \bigg)
+ = - \mu_0 \varepsilon_0 \mu_r \pdvn{2}{\vb{H}}{t}
}
\end{aligned}$$
-Traditionally, it is said that the solutions are as follows,
-where the wavenumber $$|\vb{k}| = \omega / v$$:
-
-$$\begin{aligned}
- \vb{E}(\vb{r}, t)
- &= \vb{E}_0 \exp(i \vb{k} \cdot \vb{r} - i \omega t)
- \\
- \vb{B}(\vb{r}, t)
- &= \vb{B}_0 \exp(i \vb{k} \cdot \vb{r} - i \omega t)
-\end{aligned}$$
-
-In fact, thanks to linearity, these **plane waves** can be treated as
-terms in a Fourier series, meaning that virtually
-*any* function $$f(\vb{k} \cdot \vb{r} - \omega t)$$ is a valid solution.
+This is equivalent to the problem for $$\vb{E}$$,
+since they are coupled by Maxwell's equations.
+By solving either, subject to Gauss's laws
+$$\nabla \cdot (\varepsilon_r \vb{E}) = 0$$ and $$\nabla \cdot (\mu_r \vb{H}) = 0$$,
+the behavior of light in a given system can be deduced.
+Note that Gauss's laws enforce that the wave's fields are transverse,
+i.e. they must be perpendicular to the propagation direction.
-Keep in mind that in reality $$\vb{E}$$ and $$\vb{B}$$ are real,
-so although it is mathematically convenient to use plane waves,
-in the end you will need to take the real part.
-## Non-uniform medium
+## Homogeneous linear media
-A useful generalization is to allow spatial change
-in the relative permittivity $$\varepsilon_r(\vb{r})$$
-and the relative permeability $$\mu_r(\vb{r})$$.
-We still assume that the medium is linear and isotropic, so:
+In the special case where the medium is completely uniform,
+$$\mu_r$$ and $$\varepsilon_r$$ no longer depend on $$\vb{r}$$,
+so they can be moved to the other side:
$$\begin{aligned}
- \vb{D}
- = \varepsilon_0 \varepsilon_r(\vb{r}) \vb{E}
- \qquad \quad
- \vb{B}
- = \mu_0 \mu_r(\vb{r}) \vb{H}
+ \nabla \cross \big( \nabla \cross \vb{E} \big)
+ &= - \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{E}}{t}
+ \\
+ \nabla \cross \big( \nabla \cross \vb{H} \big)
+ &= - \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{H}}{t}
\end{aligned}$$
-Inserting these expressions into Faraday's and Ampère's laws
-respectively yields:
+This can be rewritten using the vector identity
+$$\nabla \cross (\nabla \cross \vb{V}) = \nabla (\nabla \cdot \vb{V}) - \nabla^2 \vb{V}$$:
$$\begin{aligned}
- \nabla \cross \vb{E}
- = - \mu_0 \mu_r(\vb{r}) \pdv{\vb{H}}{t}
- \qquad \quad
- \nabla \cross \vb{H}
- = \varepsilon_0 \varepsilon_r(\vb{r}) \pdv{\vb{E}}{t}
+ \nabla (\nabla \cdot \vb{E}) - \nabla^2 \vb{E}
+ &= - \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{E}}{t}
+ \\
+ \nabla (\nabla \cdot \vb{H}) - \nabla^2 \vb{H}
+ &= - \mu_0 \mu_r \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{H}}{t}
\end{aligned}$$
-We then divide Ampère's law by $$\varepsilon_r(\vb{r})$$,
-take the curl, and substitute Faraday's law, giving:
+Which can be reduced using Gauss's laws
+$$\nabla \cdot \vb{E} = 0$$ and $$\nabla \cdot \vb{H} = 0$$
+thanks to the fact that $$\varepsilon_r$$ and $$\mu_r$$ are constants in this case.
+We therefore arrive at:
$$\begin{aligned}
- \nabla \cross \Big( \frac{1}{\varepsilon_r} \nabla \cross \vb{H} \Big)
- = \varepsilon_0 \pdv{}{t}(\nabla \cross \vb{E})
- = - \mu_0 \mu_r \varepsilon_0 \pdvn{2}{\vb{H}}{t}
+ \boxed{
+ \nabla^2 \vb{E} - \frac{n^2}{c^2} \pdvn{2}{\vb{E}}{t}
+ = 0
+ }
\end{aligned}$$
-Next, we exploit linearity by decomposing $$\vb{H}$$ and $$\vb{E}$$
-into Fourier series, with terms given by:
-
$$\begin{aligned}
- \vb{H}(\vb{r}, t)
- = \vb{H}(\vb{r}) \exp(- i \omega t)
- \qquad \quad
- \vb{E}(\vb{r}, t)
- = \vb{E}(\vb{r}) \exp(- i \omega t)
+ \boxed{
+ \nabla^2 \vb{H} - \frac{n^2}{c^2} \pdvn{2}{\vb{H}}{t}
+ = 0
+ }
\end{aligned}$$
-By inserting this ansatz into the equation,
-we can remove the explicit time dependence:
+Where $$c = 1 / \sqrt{\mu_0 \varepsilon_0}$$ is the speed of light in a vacuum,
+and $$n = \sqrt{\mu_0 \varepsilon_0}$$ is the refractive index of the medium.
+Note that most authors write the magnetic equation with $$\vb{B}$$ instead of $$\vb{H}$$;
+both are correct thanks to linearity.
+
+In a vacuum, where $$n = 1$$, these equations are sometimes written as
+$$\square \vb{E} = 0$$ and $$\square \vb{H} = 0$$,
+where $$\square$$ is the **d'Alembert operator**, defined as follows:
$$\begin{aligned}
- \nabla \cross \Big( \frac{1}{\varepsilon_r} \nabla \cross \vb{H} \Big) \exp(- i \omega t)
- = \mu_0 \varepsilon_0 \omega^2 \mu_r \vb{H} \exp(- i \omega t)
+ \boxed{
+ \square
+ \equiv \nabla^2 - \frac{1}{c^2} \pdvn{2}{}{t}
+ }
\end{aligned}$$
-Dividing out $$\exp(- i \omega t)$$,
-we arrive at an eigenvalue problem for $$\omega^2$$,
-with $$c = 1 / \sqrt{\mu_0 \varepsilon_0}$$:
+Note that some authors define it with the opposite sign.
+In any case, the d'Alembert operator is important for special relativity.
+
+The solution to the homogeneous electromagnetic wave equation
+are traditionally said to be the so-called **plane waves** given by:
$$\begin{aligned}
- \boxed{
- \nabla \cross \Big( \frac{1}{\varepsilon_r(\vb{r})} \nabla \cross \vb{H}(\vb{r}) \Big)
- = \Big( \frac{\omega}{c} \Big)^2 \mu_r(\vb{r}) \vb{H}(\vb{r})
- }
+ \vb{E}(\vb{r}, t)
+ &= \vb{E}_0 e^{i \vb{k} \cdot \vb{r} - i \omega t}
+ \\
+ \vb{B}(\vb{r}, t)
+ &= \vb{B}_0 e^{i \vb{k} \cdot \vb{r} - i \omega t}
\end{aligned}$$
-Compared to a uniform medium, $$\omega$$ is often not arbitrary here:
-there are discrete eigenvalues $$\omega$$,
-corresponding to discrete **modes** $$\vb{H}(\vb{r})$$.
+Where the wavevector $$\vb{k}$$ is arbitrary,
+and the angular frequency $$\omega = c |\vb{k}| / n$$.
+We also often talk about the wavelength, which is $$\lambda = 2 \pi / |\vb{k}|$$.
+The appearance of $$\vb{k}$$ in the exponent
+tells us that these waves are propagating through space,
+as you would expect.
+
+In fact, because the wave equations are linear,
+any superposition of plane waves,
+i.e. any function of the form $$f(\vb{k} \cdot \vb{r} - \omega t)$$,
+is in fact a valid solution.
+Just remember that $$\vb{E}$$ and $$\vb{H}$$ are real-valued,
+so it may be necessary to take the real part at the end of a calculation.
-Next, we go through the same process to find an equation for $$\vb{E}$$.
-Starting from Faraday's law, we divide by $$\mu_r(\vb{r})$$,
-take the curl, and insert Ampère's law:
+
+
+## Inhomogeneous linear media
+
+But suppose the medium is not uniform, i.e. it contains structures
+described by $$\varepsilon_r(\vb{r})$$ and $$\mu_r(\vb{r})$$.
+If the structures are much larger than the light's wavelength,
+the homogeneous equation is still a very good approximation
+away from any material boundaries;
+anywhere else, however, they will break down.
+Recall the general equations from before we assumed homogeneity:
$$\begin{aligned}
- \nabla \cross \Big( \frac{1}{\mu_r} \nabla \cross \vb{E} \Big)
- = - \mu_0 \pdv{}{t}(\nabla \cross \vb{H})
- = - \mu_0 \varepsilon_0 \varepsilon_r \pdvn{2}{\vb{E}}{t}
+ \nabla \cross \bigg( \frac{1}{\mu_r} \nabla \cross \vb{E} \bigg)
+ &= - \frac{\varepsilon_r}{c^2} \pdvn{2}{\vb{E}}{t}
+ \\
+ \nabla \cross \bigg( \frac{1}{\varepsilon_r} \nabla \cross \vb{H} \bigg)
+ &= - \frac{\mu_r}{c^2} \pdvn{2}{\vb{H}}{t}
\end{aligned}$$
-Then, by replacing $$\vb{E}(\vb{r}, t)$$ with our plane-wave ansatz,
-we remove the time dependence:
+In theory, this is everything we need,
+but in most cases a better approach is possible:
+the trick is that we only rarely need to explicitly calculate
+the $$t$$-dependence of $$\vb{E}$$ or $$\vb{H}$$.
+Instead, we can first solve an easier time-independent version
+of this problem, and then approximate the dynamics
+with [coupled mode theory](/know/concept/coupled-mode-theory/) later.
+
+To eliminate $$t$$, we make an ansatz for $$\vb{E}$$ and $$\vb{H}$$, shown below.
+No generality is lost by doing this;
+this is effectively a kind of [Fourier transform](/know/concept/fourier-transform/):
$$\begin{aligned}
- \nabla \cross \Big( \frac{1}{\mu_r} \nabla \cross \vb{E} \Big) \exp(- i \omega t)
- = - \mu_0 \varepsilon_0 \omega^2 \varepsilon_r \vb{E} \exp(- i \omega t)
+ \vb{E}(\vb{r}, t)
+ &= \vb{E}(\vb{r}) e^{- i \omega t}
+ \\
+ \vb{H}(\vb{r}, t)
+ &= \vb{H}(\vb{r}) e^{- i \omega t}
\end{aligned}$$
-Which, after dividing out $$\exp(- i \omega t)$$,
-yields an analogous eigenvalue problem with $$\vb{E}(r)$$:
+Inserting this ansatz and dividing out $$e^{-i \omega t}$$
+yields the time-independent forms:
$$\begin{aligned}
\boxed{
- \nabla \cross \Big( \frac{1}{\mu_r(\vb{r})} \nabla \cross \vb{E}(\vb{r}) \Big)
- = \Big( \frac{\omega}{c} \Big)^2 \varepsilon_r(\vb{r}) \vb{E}(\vb{r})
+ \nabla \cross \bigg( \frac{1}{\mu_r} \nabla \cross \vb{E} \bigg)
+ = \Big( \frac{\omega}{c} \Big)^2 \varepsilon_r \vb{E}
}
\end{aligned}$$
-Usually, it is a reasonable approximation
-to say $$\mu_r(\vb{r}) = 1$$,
-in which case the equation for $$\vb{H}(\vb{r})$$
-becomes a Hermitian eigenvalue problem,
-and is thus easier to solve than for $$\vb{E}(\vb{r})$$.
+$$\begin{aligned}
+ \boxed{
+ \nabla \cross \bigg( \frac{1}{\varepsilon_r} \nabla \cross \vb{H} \bigg)
+ = \Big( \frac{\omega}{c} \Big)^2 \mu_r \vb{H}
+ }
+\end{aligned}$$
-Keep in mind, however, that in any case,
-the solutions $$\vb{H}(\vb{r})$$ and/or $$\vb{E}(\vb{r})$$
-must satisfy the two Maxwell's equations that were not explicitly used:
+These are eigenvalue problems for $$\omega^2$$,
+which can be solved subject to Gauss's laws and suitable boundary conditions.
+The resulting allowed values of $$\omega$$ may consist of
+continuous ranges and/or discrete resonances,
+analogous to *scattering* and *bound* quantum states, respectively.
+It can be shown that the operators on both sides of each equation
+are Hermitian, meaning these are well-behaved problems
+yielding real eigenvalues and orthogonal eigenfields.
+
+Both equations are still equivalent:
+we only need to solve one. But which one?
+In practice, one is usually easier than the other,
+due to the common approximation that $$\mu_r \approx 1$$ for many dielectric materials,
+in which case the equations reduce to:
$$\begin{aligned}
- \nabla \cdot (\varepsilon_r \vb{E}) = 0
- \qquad \quad
- \nabla \cdot (\mu_r \vb{H}) = 0
+ \nabla \cross \big( \nabla \cross \vb{E} \big)
+ &= \Big( \frac{\omega}{c} \Big)^2 \varepsilon_r \vb{E}
+ \\
+ \nabla \cross \bigg( \frac{1}{\varepsilon_r} \nabla \cross \vb{H} \bigg)
+ &= \Big( \frac{\omega}{c} \Big)^2 \vb{H}
\end{aligned}$$
-This is equivalent to demanding that the resulting waves are *transverse*,
-or in other words,
-the wavevector $$\vb{k}$$ must be perpendicular to
-the amplitudes $$\vb{H}_0$$ and $$\vb{E}_0$$.
+Now the equation for $$\vb{H}$$ is starting to look simpler,
+because it only has an operator on *one* side.
+We could "fix" the equation for $$\vb{E}$$ by dividing it by $$\varepsilon_r$$,
+but the resulting operator would no longer be Hermitian,
+and hence not well-behaved.
+To get an idea of how to handle $$\varepsilon_r$$ in the $$\vb{E}$$-equation,
+notice its similarity to the weight function $$w$$
+in [Sturm-Liouville theory](/know/concept/sturm-liouville-theory/).
+
+Gauss's magnetic law $$\nabla \cdot \vb{H} = 0$$
+is also significantly easier for numerical calculations
+than its electric counterpart $$\nabla \cdot (\varepsilon_r \vb{E}) = 0$$,
+so we usually prefer to solve the equation for $$\vb{H}$$.
+
## References
diff --git a/source/know/concept/euler-equations/index.md b/source/know/concept/euler-equations/index.md
index 2654d2b..415e2f1 100644
--- a/source/know/concept/euler-equations/index.md
+++ b/source/know/concept/euler-equations/index.md
@@ -146,7 +146,7 @@ When the fluid gets compressed in a certain location, thermodynamics
states that the pressure, temperature and/or entropy must increase there.
For simplicity, let us assume an *isothermal* and *isentropic* fluid,
such that only $$p$$ is affected by compression, and the
-[fundamental thermodynamic relation](/know/concept/fundamental-thermodynamic-relation/)
+[fundamental thermodynamic relation](/know/concept/fundamental-relation-of-thermodynamics/)
reduces to $$\dd{E} = - p \dd{V}$$.
Then the pressure is given by a thermodynamic equation of state $$p(\rho, T)$$,
diff --git a/source/know/concept/fundamental-relation-of-thermodynamics/index.md b/source/know/concept/fundamental-relation-of-thermodynamics/index.md
new file mode 100644
index 0000000..a51c231
--- /dev/null
+++ b/source/know/concept/fundamental-relation-of-thermodynamics/index.md
@@ -0,0 +1,326 @@
+---
+title: "Fundamental relation of thermodynamics"
+sort_title: "Fundamental relation of thermodynamics"
+date: 2024-07-21 # Originally 2021-07-07, major rewrite
+categories:
+- Physics
+- Thermodynamics
+layout: "concept"
+---
+
+In most areas of physics,
+we observe and analyze the behaviour
+of physical systems that have been "disturbed" some way,
+i.e. we try to understand what is *happening*.
+In thermodynamics, however,
+we start paying attention once the disturbance has ended,
+and the system has had some time to settle down:
+when nothing seems to be happening anymore.
+
+Then a common observation is that the system "forgets" what happened earlier,
+and settles into a so-called **equilibrium state**
+that appears to be independent of its history.
+No matter in what way you stir your tea, once you finish,
+eventually the liquid stops moving, cools down,
+and just... sits there, doing nothing.
+But how does it "choose" this equilibrium state?
+
+
+
+## Thermodynamic equilibrium
+
+This history-independence suggests that equilibrium
+is determined by only a few parameters of the system.
+Prime candidates are the **mole numbers** $$N_1, N_2, ..., N_n$$
+of each of the $$n$$ different types of particles in the system,
+and its **volume** $$V$$.
+Furthermore, the microscopic dynamics
+are driven by energy differences between components,
+and obey the universal principle of energy conservation,
+so it also sounds reasonable to define a total
+**internal energy** $$U$$.
+
+Thanks to many decades of empirical confirmations,
+we now know that the above arguments can be combined into a postulate:
+the equilibrium state of a closed system with fixed $$U$$, $$V$$ and $$N_i$$
+is completely determined by those parameters.
+The system then "finds" the equilibrium
+by varying its microscopic degrees of freedom
+such that the **entropy** $$S$$ is maximized
+subject to the given values of $$U$$, $$V$$ and $$N_i$$.
+This statement serves as a definition of $$S$$,
+and explains the **second law of thermodynamics**:
+the total entropy never decreases.
+
+We do not care about those microscopic degrees of freedom,
+but we do care about how $$U$$, $$V$$ and $$N_i$$ influence the equilibrium.
+For a given system, we want a formula $$S(U, V, N_1, ..., N_n)$$,
+which contains all thermodynamic information about the system
+and is therefore known as its **fundamental relation**.
+
+The next part of our definition of $$S$$
+is that it must be invertible with respect to $$U$$,
+meaning we can rearrange the fundamental relation
+to $$U(S, V, N_1, ... N_n)$$ without losing any information.
+Specifically, this means that $$S$$ must be continuous,
+differentiable, and monotonically increasing with $$U$$,
+such that $$S(U)$$ can be inverted to $$U(S)$$ and vice versa.
+
+The idea here is that maximizing $$S$$ at fixed $$U$$
+should be equivalent to minimizing $$U$$ for a given $$S$$
+(we prove this later).
+Often it is mathematically more convenient
+to choose one over the other,
+but by definition both approaches are equally valid.
+And because $$S$$ is rather abstract,
+it may be preferable to treat it as a parameter
+for a more intuitive quantity like $$U$$.
+
+Next, we demand that $$S$$ is additive over subsystems,
+so $$S = S_1 + S_2 + ...$$, with $$S_1$$ being the entropy of subsystem 1, etc.
+Consequently, $$S$$ is an **extensive** quantity of the system,
+just like $$U$$ (and $$V$$ and $$N_i$$),
+meaning they satisfy for any constant $$\lambda$$:
+
+$$\begin{aligned}
+ S(\lambda U, \lambda V, \lambda N_1, ..., \lambda N_n)
+ &= \lambda S(U, V, N_1, ..., N_n)
+ \\
+ U(\lambda S, \lambda V, \lambda N_1, ..., \lambda N_n)
+ &= \lambda U(S, V, N_1, ..., N_n)
+\end{aligned}$$
+
+For $$U$$, this makes intuitive sense:
+the total energy in two identical systems
+is double the energy of a single of those systems.
+Actually, reality is a bit hazier than this:
+dynamics are governed by energy *differences* only,
+so an offset $$U_0$$ can be added without a consequence.
+We should choose an offset and a way to split the system into subsystems
+such that the above relation holds for our convenience.
+Fortunately, this choice often makes itself.
+
+$$S$$ does not suffer from this ambiguity,
+since the **third law of thermodynamics** clearly defines
+where $$S = 0$$ should occur: at a temperature of absolute zero.
+In this article we will not explore the reason for this requirement,
+which is also known as the **Nernst postulate**.
+Furthermore, in most situations this law can simply be ignored.
+
+Since $$U$$, $$S$$, $$V$$ and $$N_i$$ are all extensive,
+the partial derivatives of the fundamental relation are **intensive** quantities,
+meaning they do not depend on the size of the system.
+Those derivatives are very important,
+since they are usually the equilibrium properties we want to find.
+
+
+
+## Energy representation
+
+When we have a fundamental relation of the form $$U(S, V, N_1, ..., N_n)$$,
+we say we are treating the system's thermodynamics
+in the **energy representation**.
+
+The following derivatives of $$U$$ are used as the thermodynamic *definitions*
+of the **temperature** $$T$$, the **pressure** $$P$$,
+and the **chemical potential** $$\mu_k$$ of the $$k$$th particle species:
+
+$$\begin{aligned}
+ \boxed{
+ \begin{aligned}
+ T
+ &\equiv \bigg( \pdv{U}{S} \bigg)_{V, N_i}
+ \\
+ P
+ &\equiv - \bigg( \pdv{U}{V} \bigg)_{S, N_i}
+ \\
+ \mu_k
+ &\equiv \bigg( \pdv{U}{N_k} \bigg)_{S, V, N_{i \neq k}}
+ \end{aligned}
+ }
+\end{aligned}$$
+
+The resulting expressions of the form $$T(S, V, N_1, ..., N_n)$$ etc.
+are known as the **equations of state** of the system.
+Unlike the fundamental relation, a single equation of state
+is not a complete thermodynamic description of the system.
+However, if *all* equations of state are known
+(for $$T$$, $$P$$, and all $$\mu_k$$),
+then the fundamental relation can be reconstructed.
+
+As explained above, physical dynamics are driven by energy differences only,
+so we expand an infinitesimal difference $$\dd{U}$$ as:
+
+$$\begin{aligned}
+ \dd{U}
+ = \bigg( \pdv{U}{S} \bigg)_{V, N_i} \!\dd{S}
+ \:\:+\:\: \bigg( \pdv{U}{V} \bigg)_{S, N_i} \!\dd{V}
+ \:\:+\:\: \sum_{k}^{} \bigg( \pdv{U}{N_k} \bigg)_{S, V, N_{i \neq k}} \!\dd{N_k}
+\end{aligned}$$
+
+Those partial derivatives look familiar.
+Substituting $$T$$, $$P$$ and $$\mu_k$$ gives a result
+that is also called the **fundamental relation of thermodynamics**
+(as opposed to the fundamental relation of the system only,
+just to make things confusing):
+
+$$\begin{aligned}
+ \boxed{
+ \dd{U}
+ = T \dd{S} - P \dd{V} + \sum_{k}^{} \mu_k \dd{N_k}
+ }
+\end{aligned}$$
+
+Where the first term represents heating/cooling (also written as $$\dd{Q}$$),
+and the second is physical work done on the system
+by compression/expansion (also written as $$\dd{W}$$).
+The third term is the energy change due to matter transfer and is often neglected.
+Hence this relation can be treated as a form
+of the **first law of thermodynamics** $$\Delta U = \Delta Q + \Delta W$$.
+
+Because $$T$$, $$P$$ and $$\mu_k$$ generally depend on $$S$$, $$V$$ and $$N_k$$,
+integrating the fundamental relation can be tricky.
+Fortunately, the fact that $$U$$ is extensive offers a shortcut.
+Recall that:
+
+$$\begin{aligned}
+ \lambda U(S, V, N_1, ..., N_n)
+ &= U(\lambda S, \lambda V, \lambda N_1, ..., \lambda N_n)
+\end{aligned}$$
+
+For any $$\lambda$$.
+Let us differentiate this equation with respect to $$\lambda$$, yielding:
+
+$$\begin{aligned}
+ U
+ &= \pdv{}{\lambda} U(\lambda S, \lambda V, \lambda N_1, ..., \lambda N_n)
+ \\
+ &= \pdv{U(\lambda S)}{(\lambda S)} \pdv{(\lambda S)}{\lambda}
+ + \pdv{U(\lambda V)}{(\lambda V)} \pdv{(\lambda V)}{\lambda}
+ + \sum_{k} \pdv{U(\lambda N_k)}{(\lambda N_k)} \pdv{(\lambda N_k)}{\lambda}
+ \\
+ &= \pdv{U(S)}{S} S
+ + \pdv{U(V)}{V} V
+ + \sum_{k} \pdv{U(N_k)}{N_k} N_k
+\end{aligned}$$
+
+Where we once again recognize the derivatives.
+The resulting equation is known as the **Euler form**
+of the fundamental relation of thermodynamics:
+
+$$\begin{aligned}
+ \boxed{
+ U
+ = T S - P V + \sum_{k} \mu_k N_k
+ }
+\end{aligned}$$
+
+Plus a constant $$U_0$$ of course,
+although $$U_0 = 0$$ is the most straightforward choice.
+
+
+
+## Entropy representation
+
+If the system's fundamental relation
+instead has the form $$S(U, V, N_1, ..., N_i)$$,
+we are treating it in the **entropy representation**.
+Isolating the above fundamental relation of thermodynamics
+for $$\dd{S}$$ yields its equivalent form in this representation:
+
+$$\begin{aligned}
+ \boxed{
+ \dd{S}
+ = \frac{1}{T} \dd{U} + \frac{P}{T} \dd{V} - \sum_{k}^{} \frac{\mu_k}{T} \dd{N_k}
+ }
+\end{aligned}$$
+
+From which we can then read off the standard partial derivatives of $$S(U, V, N_1, ..., N_n)$$:
+
+$$\begin{aligned}
+ \boxed{
+ \begin{aligned}
+ \frac{1}{T}
+ &= \bigg( \pdv{S}{U} \bigg)_{V, N_i}
+ \\
+ \frac{P}{T}
+ &= \bigg( \pdv{S}{V} \bigg)_{U, N_i}
+ \\
+ \frac{\mu_k}{T}
+ &= - \bigg( \pdv{S}{N_k} \bigg)_{U, V, N_{i \neq k}}
+ \end{aligned}
+ }
+\end{aligned}$$
+
+Note the signs: the parameters $$U$$, $$V$$ and $$N_i$$ are implicitly related
+by our requirement that $$S$$ is stationary at a maximum,
+so the [triple product rule](/know/concept/triple-product-rule/)
+must be used, which brings some perhaps surprising sign changes.
+Reading them off in this way is easier.
+
+And of course, since $$S$$ is defined to be an extensive quantity,
+it also has an Euler form:
+
+$$\begin{aligned}
+ \boxed{
+ S
+ = \frac{1}{T} U + \frac{P}{T} V - \sum_{k} \frac{\mu_k}{T} N_k
+ }
+\end{aligned}$$
+
+Finally, it is worth proving that minimizing $$U$$
+is indeed equivalent to maximizing $$S$$.
+For simplicity, we consider a system
+where only the volume $$V$$ can change
+in order to reach an equilibrium;
+the proof is analogous for all other parameters.
+Clearly, $$S$$ is stationary at its maximum:
+
+$$\begin{aligned}
+ 0
+ &= \bigg( \pdv{S}{V} \bigg)_{U, N_i}
+ = - \frac{ \bigg( \displaystyle\pdv{U}{V} \bigg)_{S, N_i} }{ \bigg( \displaystyle\pdv{U}{S} \bigg)_{V, N_i} }
+ = - \frac{1}{T} \bigg( \pdv{U}{V} \bigg)_{S, N_i}
+\end{aligned}$$
+
+Where we have used the triple product rule.
+This can only hold if $$(\ipdv{U}{S})_{S, N_i} = 0$$,
+meaning $$U$$ is also at an extremum.
+But $$S$$ is not just at any extremum: it is at a *maximum*, so:
+
+$$\begin{aligned}
+ 0
+ > \bigg( \pdvn{2}{S}{V} \bigg)_{U, N_i}
+ &= \bigg( \pdv{}{V} \Big( \frac{P}{T} \Big) \bigg)_{U, N_i}
+ \\
+ &= \bigg( \pdv{}{V} \Big( \frac{P}{T} \Big) \bigg)_{S, N_i}
+ + \bigg( \pdv{}{S} \Big( \frac{P}{T} \Big) \bigg)_{V, N_i} \bigg( \pdv{S}{V} \bigg)_{U, N_i}
+ \\
+ &= \bigg( \pdv{}{V} \Big( \frac{P}{T} \Big) \bigg)_{S, N_i}
+ + \frac{P}{T} \bigg( \pdv{}{S} \Big( \frac{P}{T} \Big) \bigg)_{V, N_i}
+ \\
+ &= \frac{1}{T} \bigg( \pdv{P}{V} \bigg)_{S, N_i}
+ - \frac{P}{T^2} \bigg( \pdv{T}{V} \bigg)_{S, N_i}
+ + \frac{P}{T} \bigg( \pdv{}{S} \Big( \frac{P}{T} \Big) \bigg)_{V, N_i}
+ \\
+ &= - \frac{1}{T} \bigg( \pdvn{2}{U}{V} \bigg)_{S, N_i}
+ + \frac{P}{T} \bigg[ \bigg( \pdv{}{S} \Big( \frac{P}{T} \Big) \bigg)_{V, N_i}
+ - \frac{1}{T} \bigg( \pdv{T}{V} \bigg)_{S, N_i} \bigg]
+\end{aligned}$$
+
+Because $$S$$ is at a maximum, we know that $$P/T = 0$$,
+and $$T$$ is always above absolute zero
+(since we defined $$S$$ to be monotonically increasing with $$U$$),
+which leaves $$(\ipdvn{2}{U}{V})_{S, N_i} > 0$$
+as the only way to satisfy this inequality.
+In other words, $$U$$ is at a minimum, as expected.
+
+
+
+## References
+1. H.B. Callen,
+ *Thermodynamics and an introduction to thermostatistics*, 2nd edition,
+ Wiley.
+2. H. Gould, J. Tobochnik,
+ *Statistical and thermal physics*, 2nd edition,
+ Princeton.
diff --git a/source/know/concept/fundamental-thermodynamic-relation/index.md b/source/know/concept/fundamental-thermodynamic-relation/index.md
deleted file mode 100644
index 0d945fa..0000000
--- a/source/know/concept/fundamental-thermodynamic-relation/index.md
+++ /dev/null
@@ -1,54 +0,0 @@
----
-title: "Fundamental thermodynamic relation"
-sort_title: "Fundamental thermodynamic relation"
-date: 2021-07-07
-categories:
-- Physics
-- Thermodynamics
-layout: "concept"
----
-
-The **fundamental thermodynamic relation** combines the first two
-[laws of thermodynamics](/know/concept/laws-of-thermodynamics/),
-and gives the change of the internal energy $$U$$,
-which is a [thermodynamic potential](/know/concept/thermodynamic-potential/),
-in terms of the change in