8.2: Momentum #
In this chapter we talk about:
- Electromagnetic momentum
- Maxwell stress tensor
- Conservation of electromagnetic momentum
- Angular momentum in EM fields
8.2.1: Electromagnetic Momentum #
As it turns out, if you disregard the momentum associated with electromagnetic fields, Newton’s laws appear not to work out! Consider a basic system in cartesian coordinates of two moving point charges:
What happens between the two charges? Well, the magnetic field of \( q_1 \) points into the page at \( q_2 \), so the magnetic force on \( q_2 \) is to the right, and the magnetic field of \( q_2 \) is out of the page at \( q_1 \) , so the magnetic force on \( q_1 \) is upward. The net electric force between the two charges is repulsive and opposite, but the magnetic forces aren’t, so the electromagnetic force on \( q_1 \) on \( q_2 \) is equal but not opposite to the force of \( q_2 \) on \( q_1 \), in violation of Newton’s third law! We’ve got a problem, and we’re going to solve it by invoking the momentum of the EM field.
8.2.2 The Maxwell Stress Tensor #
The way to recover conservation of momentum proceeds the same way we recovered the conservation of energy via the Poynting vector. Starting with the basic Coulomb/Lorentz laws, we’ll write down an expression for the electromagnetic force on charges in a volume. We’re going to integrate that over all space, which can have any distribution of charge, and relate that expression for force to an expression which only involves the field. In the interest of brevity, we’ll skip around a bit and leave the full derivations for the real textbook.
Suppose we have a volume \( V \) containing some distribution of charge, current, and electromagnetic fields. The total force on that volume is
\[\vec{F} = \int_V ( \vec E + \vec v \cross \vec B) \rho \dd \tau \qquad (\dd q = \rho \dd \tau) \\ = \int _V (\rho \vec E + \vec J \cross \vec B ) \dd \tau \qquad (\vec J = \rho \vec v)\]Again, the goal is to replace anything that looks like a source in favor of fields, using Maxwell’s equations. It’s handy to define the force per unit volume \( f \):
\[\vec f \equiv \rho \vec E + \vec J \cross \vec B\] \[\rho = \epsilon_0 ( \div \vec E ) \quad \text{(Gauss' Law)} \\ \vec J = \frac{1}{\mu_0} \curl \vec B - \epsilon_0 \pdv{\vec E}{t}\] \[\rightarrow \vec f = \epsilon_0 ( \div \vec E) \vec E + \left( \frac{\curl \vec B}{\mu_0} - \epsilon_0 \pdv{\vec E}{t} \right) \cross \vec B\]Skipping through a few steps, we cut to the chase. Similar to the derivation of the Poynting theorem, also using the other two Maxwell equations we haven’t yet, we get
\[\vec f = \div \overline{\vec T} - \epsilon_0 \mu_0 \pdv{\vec S}{t}\] \[\overline{\vec T} \equiv \begin{pmatrix} T_{xx} & T_{xy} & T_{xz} \\ T_{yx} & T_{yy} & T_{yz} \\ T_{zx} & T_{zy} & T_{zz} \end{pmatrix} \\ T_{ij} \equiv \epsilon_0 \left( E_i E_j - \frac{1}{2} \delta_{ij} E^2 \right) + \frac{1}{\mu_0} \left(B_i B_j - \frac{1}{2} \delta_{ij} B^2 \right)\]where \( \vec S \) is the Poynting vector and \( \overline{\vec T} \) is the so-called “Maxwell stress tensor.” To keep in mind what kind of units we’re talking about here, \( \vec f \) has units force per unit volume, and the divergence will strip one spatial dimension, so the Maxwell stress tensor will have units of stress (force per unit area).
The tensor has diagonal “pressure” terms and off-diagonal “shear” terms. For “pressure” forces, the force and area are in the same direction, and in the “shear” case the force and area are orthogonal.
The divergence term \( \div \overline{\vec T} \) is itself a vector
\[\div \overline{\vec T} = \left( \vu{i} \pdv{}{x} + \vu j \pdv{}{y} + \vu k \pdv{}{z} \right) \cdot \overline{\vec T} \\ = \epsilon_0 \left[ (\div \vec E) E_j + (\vec E \cdot \grad) E_j - \frac{1}{2} \grad _j E^2 \right] \\ + \frac{1}{\mu_0} \left[ ( \div \vec B) B_j + (\vec B \cdot \grad) B_j - \frac{1}{2} \grad _j B^2 \right]\]As we do the volume integral to go from \( \vec f \) to \( \vec F \)
\[\vec F = \int _V \left(\div\overline{\vec T} - \epsilon_0 \mu_0 \pdv{\vec S}{t} \right) \dd \tau \\ = \oint \overline{\vec{T}} \cdot \dd \vec a - \epsilon_0 \mu_0 \pdv{}{t} \int \vec{S} \dd \tau\]8.2.3 Conservation of Electromagnetic Momentum #
\[\vec F = \dv{\vec{p}_{mech}}{t} = - \epsilon_0 \mu_0 \dv{}{t} \int_V \vec S \dd \tau + \oint \overline{\vec T} \cdot \dd \vec a\]The first term on the right is related to the momentum stored in the electromagnetic field. The second term is the rate at which momentum flows across the surface, and we describe the left-hand-side as the rate of change of the momentum of charges within the volume.
We identify another useful term as the first integrand on the right:
\[\vec g \equiv \epsilon_0 \mu_0 \vec S = \epsilon_0 (\vec E \cross B) \quad \text{(momentum density in EM fields)}\]which is the momentum density within the fields. Just as a note, the signs here are swapped from the Poynting theorem - the Maxwell stress tensor is defined such that momentum flowing into the region corresponds with increasing \( \overline{\vec T} \), and vice-versa, opposite the case we had with \( \vec S \).
In a charge-free region,
\[- \dv{}{t} \int_V \vec g \dd \tau + \oint_S \overline{\vec T} \cdot \dd \vec a = 0\]and since the above is true for all regions \( V \), we have our familiar continuity-type equation
\[- \pdv{\vec g}{t} + \div \overline{\vec T} = 0\]Example: Problem 8.7 #
8.2.4 Angular Momentum in EM Fields #
As a reminder, we associate with the electromagnetic fields an energy density
\[u_{em} = \frac{1}{2} \epsilon_0 E^2 + \frac{\mu_0}{2} B^2\]and a momentum density
\[\vec g = \epsilon_0 ( \vec E \cross \vec B)\]For that matter, we define angular momentum in the normal fashion
\[\vec l = \vec r \cross \vec g = \epsilon_0 [ \vec r \cross ( \vec E \cross \vec B) ]\]where the presence of \( \vec r \) means it’s defined about some point or axis.