Home Drew Rogers

Geodesic Equation

As a graduate student I gave a talk deriving the geodesic equation using variational calculus. The audience was primarily pure mathematicians, so I also gave some elementary intuition for the classical Euler-Lagrange equation. Here are some of my notes.

Conventions: Unless otherwise noted, the prime notation will indicate a derivative with respect to $x$. That is, $y'(x) = \frac{\mathrm{d}y}{\mathrm{d}x}$. Meanwhile, dot notation will indicate a derivative with respect to time. That is, $\dot{y} = \frac{\mathrm{d}y}{\mathrm{d}t}$

Physical Intuition & Classical Euler-Lagrange: For a particle traveling along some path on a plane, we of course know that the shortest path between two points is a straight line. We calculate this first by observing that the length of any short section of the path is $\dd s = \sqrt{\dd x^2 + \dd y^2} = \sqrt{1 + y'(x)^2} \dd x$. Then if we let our starting point be $1 = (x_1,y_1)$ and our end point be $2=(x_2,y_2)$ we can integrate to find the length of our path as follows:

\[ \begin{align*} \int_1^2 \dd s &= \int_1^2 \sqrt{1+y'(x)^2} \dd x \\ &= \int_1^2 f(x,y,y') \dd x \end{align*} \]

For another example, we of course know (or are about to learn about) Fermat's principle, which states that light takes a time-minimizing path. Lots of great physics comes straight from this, and we will take some of it for granted to skip to the part that is important for us.

Let $n\geq 1$ be the refractive index of some material, $c$ be the speed of light (in a vacuum), and $\dd s$ again be a short snippet of some path. Then $v = c/n \quad \& \quad \dd s = v \dd t \implies \dd t = \frac1v \dd s \quad \& \quad \dd t = \frac nc \dd s $. Additionally, we again let our starting point be $1 = (x_1,y_1)$ and our end point be $2=(x_2,y_2)$. So to find the travel time of our photon (the particle model is easier to imagine, but the wave model works essentially the same way) we have:

\[ \begin{align*} \int_1^2 \dd t &= \frac{1}{c} \int_1^2 n \dd s \\ &= \frac 1c \int_1^2 n \sqrt{1 + y'(x)^2} \dd x \\ &= \int_1^2 f(x,y,y') \dd x \end{align*} \]

Where we can of course pull $c$ out of our integral, since it is a constant, but we could have that $n$ is a function of $x, y $ or both.

It is important to note that we may not always be looking for a minimum value of a function. Oftentimes finding maxima and saddle points is the goal of an optimization problems i.e. maximizing range, or finding Lagrange points, which are points in the Sun-Earth system where the two gravitational fields balance out and are a sort of unstable or semi-stable equilibrium for (relatively) small objects.

In both of the two cases above, we have an integral $\int f(x,y,y')\dd x$ for which we'd like to find the shortest $y(x)$ (in time, length, or whatever else you care to minimize over). To do so, assuming that $y(x)$ is the "correct curve" let $\eta$ be some variation. That is, any curve between 1 and 2 can be represented as $W(x) =y(x) + \alpha\eta(x)$ for some choice of $\eta$. We of course must have that $\eta(1) = \eta(2) = 0$, and we've added a new parameter $\alpha$ so that we can easily obtain the "correct" path by setting $\alpha = 0$. So, $S(\alpha) = \int_1^2 f(x,W,W') \dd x = \int_1^2 f(x, y + \alpha \eta , y' + \alpha \eta') \dd x$. We vary our path in search of some extremum, so $\frac{\mathrm{d}S}{\mathrm{d}\alpha}\Big|_{\alpha = 0} = 0$:

\[ \begin{align*} \frac{\mathrm{d}S}{\mathrm{d}\alpha} \Bigg|_{\alpha = 0} &= \int_1^2 \frac{\partial f}{\partial \alpha} \Bigg|_{\alpha = 0} \mathrm{d}x \\ &= \int_1^2 \bigg( \eta \frac{\partial f}{\partial y} + \eta' \frac{\partial f}{\partial y'} \bigg) \mathrm{d}x = 0 \end{align*} \]

Since:

\[ \begin{align*} \frac{\partial f}{\partial \alpha} \Bigg|_{\alpha = 0} &= \frac{\partial f}{\partial W} \frac{\partial W}{\partial \alpha}\Bigg|_{\alpha = 0} + \frac{\partial f}{\partial W'} \frac{\partial W'}{\partial \alpha}\Bigg|_{\alpha = 0} \\ &= \eta \frac{\partial f}{\partial y} + \eta' \frac{\partial f}{\partial y'} \end{align*} \]

Integrating the second term by parts we get:

\[ \begin{align*} \int_1^2 \eta'\frac{\partial f}{\partial y'} \mathrm{d}x = \bigg(\eta \frac{\partial f}{\partial y'}\bigg)\bigg|_1^2 - \int_1^2 \eta \frac{\mathrm{d}}{\mathrm{d}x} \frac{\partial f}{\partial y'} \mathrm{d}x \end{align*} \]

Now, since $\eta(1) = \eta(2) = 0$ we have:

\[ \begin{align*} \int_1^2 \eta' \frac{\partial f}{\partial y'} \mathrm{d}x &= - \int_1^2 \eta \frac{\mathrm{d}}{\mathrm{d}x} \frac{\partial f}{\partial y'} \mathrm{d}x \end{align*} \]

And so,

\[ \begin{align*} \int_1^2 \bigg( \eta \frac{\partial f}{\partial y} + \eta' \frac{\partial f}{\partial y'}\bigg) &= \int_1^2 \eta \bigg( \frac{\partial f}{\partial y} - \frac{\mathrm{d}}{\mathrm{d}x} \frac{\partial f}{\partial y'}\bigg) \mathrm{d}x = 0 \end{align*} \]

And since this must be true, regardless of our choice of $\eta$, we have

\[ \begin{align*} \frac{\partial f}{\partial y} - \frac{\mathrm{d}}{\mathrm{d}x} \frac{\partial f}{\partial y'} =0 \end{align*} \]

Or, in some general coordinate $q$,

\[ \begin{align*} \frac{\partial L}{\partial q} &= \frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial \dot{q}} \end{align*} \]

We have introduced the Lagrangian $L = T - U$ (kinetic energy minus potential energy). Choosing to have $L$ to be a difference of energies seems arbitrary and as though it would be less useful than the total energy (which we call the Hamiltonian) and there is no simple explanation for this choice. The reasoning that is most straightforward is that this choice reduces to Newton's second law as follows:

\[ \begin{align*} \frac{\partial L}{\partial q} &= -\frac{\partial U}{\partial q} = F_q \\ \frac{\partial L}{\partial \dot{q}} &= \frac{\partial T}{\partial \dot{q}} = m\dot{q} \end{align*} \]

So,

\[ \begin{align*} \frac{\partial L}{\partial q} &= \frac{\mathrm{d}}{\mathrm{d}t} \frac{\partial L}{\partial \dot{q}} \\ \implies F_q &= m\ddot{q} \\ \iff\Vec{F} &= m\Vec{a} \end{align*} \]

Differential Geometry: Consider the Lagrangian:

\[ \begin{align*} L = \frac{1}{2} g_{jk} \dot{x}^j\dot{x}^k \end{align*} \]

Which we can consider to represent a particle with zero potential moving freely on some manifold. Then the Euler-Lagrange equations are:

\[ \begin{align*} \frac{\partial L}{\partial x^i} &= \frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial \dot{x}^i} \qquad \qquad\qquad i = 1,2,\dots,n \end{align*} \]

Solving for each of the terms we have:

\[ \begin{align*} \frac{\partial L}{\partial \dot{x}^i} &= \frac{1}{2}g_{jk}\delta^j_i \dot{x}^k + \frac{1}{2}g_{jk}\dot{x}^j \delta^k_i \\ &= \frac 1 2 g_{ik}\dot{x}^k + \frac 1 2 g_{ij} \dot{x}^j\\ \implies \frac{\mathrm{d}}{\mathrm{d}t} \frac{\partial L}{\partial \dot{x}^i} &= \frac 1 2 \left( g_{ik} \ddot{x}^k + g_{ij}\ddot{x}^j + g_{ik,l}\dot{x}^l\dot{x}^k + g_{ij,l}\dot{x}^l\dot{x}^j \right) \qquad\qquad \qquad l = 1,2,\dots, n\\ \frac{\partial L}{\partial x^i} &= \frac{1}{2} g_{jk,i}\dot{x}^j \dot{x}^k \\ \therefore \frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial \dot{x}^i} - \frac{\partial L}{\partial x^i} &= \frac 1 2 \left( g_{ik} \ddot{x}^k + g_{ij}\ddot{x}^j + g_{ik,l}\dot{x}^l\dot{x}^k + g_{ij,l}\dot{x}^l\dot{x}^j -g_{jk,i}\dot{x}^j \dot{x}^k \right) = 0 \end{align*} \]

Now, re-index as follows. First, let $g_{ik}\ddot{x}^k + g_{ij}\ddot{x}^j = 2g_{lm} \ddot{x}^m$, which is legal since in the first term we sum over $k$ and in the second term we sum over $j$. Then let $g_{ik,l}\dot{x}^l\dot{x}^k= g_{ik,j}\dot{x}^j\dot{x}^k$ and $g_{ij,l}\dot{x}^l\dot{x}^j = g_{ij,k}\dot{x}^j\dot{x}^k$. Again, both are completely legal, since we are changing the name of a summed over index so that we can remove all the $\dot{x}$ terms. Now, with these new indices, we have:

\[ \begin{align*} \frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial \dot{x}^i}-\frac{\partial L}{\partial x^i} &= g_{lm} \ddot{x}^m + \frac 1 2 \left(g_{lk,j} + g_{jl,k} - g_{jk,l}\right) \dot{x}^j \dot{x}^k = 0 \end{align*} \]

We can contract to move all of the metrics into a single term:

\[ \begin{align*} g_{lm} \ddot{x}^m + \frac 1 2 \left(g_{lk,j} + g_{jl,k} - g_{jk,l}\right) \dot{x}^j \dot{x}^k &= g^{il}g_{lm} \ddot{x}^m + \frac 1 2 g^{il}\left(g_{lk,j} + g_{jl,k} - g_{jk,l}\right) \dot{x}^j \dot{x}^k \\ &= \ddot{x}^i + \frac 1 2 g^{il}\left(g_{lk,j} + g_{jl,k} - g_{jk,l}\right) \dot{x}^j \dot{x}^k = 0 \end{align*} \]

From here, it makes perfect mathematical sense to collect the sum/product of metrics in front of the pair of $\dot{x}$s into a single symbol to further simplify this equation. This leads directly to the definition of the Christoffel Symbol $\Gamma^i_{jk} = \frac 1 2 g^{il}\left(g_{lk,j} + g_{jl,k} - g_{jk,l}\right)$ and simplifies the above equation into:

\[ \begin{align*} \ddot{x}^i + \Gamma^i_{jk}\dot{x}^j\dot{x}^k = 0 \end{align*} \]

Both of which match our past results for the geodesic equation and Christoffel symbols.