\[ \begin{align}\begin{aligned} \label{eq:lorentzeq} m_j \dv[2]{\vec{r}_j}{t} = q_j \qty[ \vec{E}(\vec{r},t) + \dot{\vec{r}_j} \times \vec{B}(\vec{r}, t) ]\\Whenever charge distributions or current densities are signifigant a\end{aligned}\end{align} \]statistical or fluid approach will be required.
Let us start with a static description, \(\vec{E} = 0, \vec{B}(\vec{r}, t) = B \neq 0\). Taking the direction of \(\vec{B}\) to define the \(z\)-axis, so \(\vec{B} = (0,0,B)\), we have the scalar product of \(\hat{z}\) and the Lorentz equation, equation ([eq:lorentzeq]) giving
\(m \ddot{\vec{r}} \cdot \dot{\vec{r}} = 0\), so \(\half m \dot{\vec{r}}^2\) is constant. We can now find, by taking dot products of the unit vectors with equation ([eq:lorentzeq]), that
equation ([eq:lorentzeq]),
Before proceeding, we’ll make the definition of the Larmour Radius,
The Larmour radius, \(r_{\rm L}\) (also gyroradius, or cyclotron radius) is the radius of the circular motion of a charged particle in a uniform magnetic field.
We can now find the full equations of motion by considering
Now let’s elaborate to a situation which involves an electric field. We now have, from equation ([eq:lorentzeq]),
Now we arrange our coordinates such that \(\vec{B}=(0,0, B_z)\), and \(\vec{E} = (E_x, 0, E_z)\), so
The gyrocentre will shift, but to understand how we must solve these equations. To do this we take a similar approach to the last lecture.
Let’s take an alternative approach here, and work entirely with vectors again, so first,
with \(v_{\rm ge}\) the velocity of the guiding centre.
As a result of this drift, the movement of a particle in this situation can be described by
Now, suppose we have a non-uniform field which varies weakly, that is,
for \(L \gg r_{\rm L}\). Again, we set up our coordinates so \(\vec{B} = (0,0, B)\), then we have
In practice magnetic fields are rarely uniform, but are generally space- and time-dependent. Normally this would lead to a numerical treatment being required, but there are cases where it is possible to calculate the variation analytically, by assuming the inhomogeneity to be small.
First, consider the case when a particle moves only parallel to the magnetic field lines. The central force experienced by the particle will be
is moving on. Now, we introduce the quantity
change of the magnetic field along the direction \(\hat{B}\), so
and, since the particle is moving along lines of constant field strength, so that \(\pdv{B}{S}=0,\)
which accompanies the curvature drift, as \(\vec{B}\) must decrease with radius. This is because, in a vacuum we require \(\nabla \times \vec{B} = 0\) (law of conservation of energy) and \(\nabla \cdot \vec{B} = 0\) (Gauss’s Law). Expressing the problem in cylindrical coorinates it is trivial to see that \(\nabla \times \vec{B}\) can only have an \(z\)-component. Now,
\(B = B_{\theta} \sim \frac{1}{R_{\rm c}}\), and, :math:`frac{nabla B}{B} = - frac{vec{R}_{rm
c}}{R_{rm c}^2}`. Using the gradient drift expression from earlier,
\(R_{\rm c}\) and \(\vec{B}\), and so in dipole fields we have drift which is perpendicular to both \(R_{\rm c}\) and \(\vec{B}\), and the drift is charge-dependent, so there is a current, which is known as the ring current.
Consider a non-uniform magnetic field, primarily in the \(z\)-direction, which has a magnitude which varies in the \(z\)-direction. Let the field be axisymmetric, such that \(B_{\theta}=0\), and \(\pdv{\theta}\cdot B=0\). Since the field lines converge and diverge, \(B_r \neq 0\).
in 1,2,3 (7,-) .. controls (3,-) and (2,-) .. (-1,-); (7,) .. controls (3,) and (2,) .. (-1,); (-2,0) – (8,0) node [right] \(z\); (-1,0) – (-1,1) node [midway, left] \(r\);
From \(\nabla \cdot \vec{B} = 0\) we have
If \(\pdv{B_z}{z}\) at \(r=0\) is given, and doesn’t change much with \(r\) we have the relation
drift of the guiding centre about the axis of symmetry, but there is no radial \(\nabla B\) drift, as \(\pdv{B}{\theta}=0\). The Lorentz force is then
\(r \to 0\), \(B_r \to 0\), since \(B_r\) vanishes on the axis. When it doesn’t vanish the azimuthal force leads to a drift in the radial direction. This drift makes the guiding centres follow the magnetic field lines. Consider the \(z\)-component of ([eq:mirroringlorentz]),
The force can then be written
\(\vec{B}\),
particle; \(\vec{B}\) itself is constant. Since \(\frac{mv^2}{2}\) is constant, due to conservation of energy, we have
in time. This invariance allows a plasma to be confined through magnetic mirrors.
\[ \begin{align}\begin{aligned} \frac{1}{2} \frac{m v_{\perp,0}^2}{B_{\rm min}} = \half \frac{m v_{\perp}^{\prime 2}}{B_{\rm max}}\\Then, by conservation of :math:`v_{\perp}^2 + v_{\parallel}^2`,\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}v_0^2 = v_{\perp,0}^2 + v_{\parallel,0}^2 = v_{\perp}^{\prime 2} + v_{\parallel}^{\prime 2}\\and so\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned} \frac{B_{\rm min}}{B_{\rm max}} = \frac{v_{\perp, 0}^{2}}{v_{\perp,0}^{\prime 2}} = \frac{v_{\perp,0}}{v_0^2} = \sin[2](\theta)\\so\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\sin(\theta) := \frac{v_{\perp}}{v_0}\\We can now use :math:`\theta` to describe whether the particle\end{aligned}\end{align} \]escapes from, or is trapped by, the magnetic field.
In a classical system which experiences a periodic motion, the action integral taken over one period of the motion will be constant. That is
If a slow change occurs in the system, such that the motion is (just) non-periodic, the constant of the motion does not change, and is called adiabatically invariant. The slow change is qualified by a change takes longer than one period of the underlying motion, so that the action is well-defined (although the integral is strictly no longer a closed loop integral).
The first adaibatic invariant of a plasma involves the Larmour gyration, so
time, \(\Delta t\) of the magnetic field, \(B\), is long, i.e. \(\Delta t \omega_{\rm c} \gg 1\). Otherwise the magnetic moment is not conserved.
The second adiabatic invariant involves the periodic oscillation of plasma particles between magnetic mirrors. This time
is not exactly periodic, and so the motion is adiabatically invariant. This is also longitudinally invariant \(J\), defined as the half-cycle between the two mirror points, \(a\) and \(b\), with
violated in transit-time magnetic pumping, a method for heating a plasma. This is done by moving \(a\) and \(b\) over time to increase the \(v_{\parallel}\) as the particles approach the mirror points.
Plasma is the “fourth state of matter”. It refers to a state containing enough free charges for its dynamics to be dominated by long-range Coloumb forces, rather than shorter-range binary collisions 1. The presence of charge carriers in the matter cause a plasma to interact strongly with electromagnetic fields.
There are a number of approaches to producing a plasma in the lab:
photoionisation—for this we need photons with sufficient energy to remove electrons from the neutral species, e.g. 13.6 eV for Hydrogen, 24.6 eV for Helium, and 15.6 eV for molecular nitrogen. All of these energies lie within the ultraviolet region of the electromagnetic spectrum. The existence of long-lived, metastable states can help with ionisation processes (He has two, at 19.8 eV (with a half-life of 700s), and 20.61 eV; \({\rm H}_{2}\) at 11.75 eV (\(10^{-3} {\rm s}\)), and H at 0.52 eV (1 month).)
electron impact—for this we accelerate any free electrons (for example, the seed electrons) in an electric field until they reach the ionising threshold. This process makes use of the Townsend process, where \(N\) electrons move along the \(x\)-axis in the presence of a uniform electric field, \(E\); then \(\delta N\) electrons are produced by electron-impact ionisation in a distance \(\dif x\), according to
Thus,
Suppose we have a gas of electrons and protons, i.e. a Hydrogen plasma, at a temperature, \(T\). Consider the situation where a random fluctuation of the electron population exposes some positive particles, thus an unbalanced positive charge. Exposing an unbalanced positve charge will cause a net movement of negative charge (in the form of electrons) to move towards the positive charge. What is the associated scale length for this process; can this be made consistent with thermodynamics?
(-2, 1) – (2, 1); (-2, -1) – (2,-1);
(-1, 1) – (-1, -1); ( 1, 1) – ( 1, -1);
(-1, 1) rectangle (1, -1); (0,0) node [text centered] Zone of depletion; (-1.5, 1.2) node [text centered] neutral; (1.5, 1.2) node [text centered] neutral; (0, -1.2) node [text centered] positive;
The average kinetic energy of the electrons is \(E = \frac{1}{2} k_{B }T\). The fluctuation in the electron denisty leaves behind an unbalanced charge, and therefore an associated electric potential \(\phi\). Poisson’s equation says
boundary conditions: \(\phi(x=\pm d) = 0\). Applying these boundary consitions means that
; ; ;
Now recall that we want to create this region via a small thermal fluctuation; therefore we need \(\frac{1}{2} k_{B}T\) to be the maximum energy of the electrons, and to be the maximum potential energy of the well (otherwise the electrons can’t escape.) Hence \(\frac{ned^2}{2\epsilon_0} e = \frac{1}{2} k_BT\), and so
\(\lambda_{{\rm D}}\). And this is the characteristic screening length for unbalanced charges.
The Debye length in a plasma is the characteristic screening length for an unbalanced charge, which is dictated by the kinetic energy of the plasma.
Plasma |
Density (\(\meter^{-3}\)) |
Electron temperature (\(\kelvin\)) |
Magnetic Field (\(\tesla\)) |
Debye Length (\(\meter\)) |
---|---|---|---|---|
Solar core |
\(10^{32}\) |
\(10^7\) |
\(10^{-11}\) |
|
Tokamak |
\(10^{20}\) |
\(10^7\) |
10 |
\(2.10 \e{-5}\) |
Hot interstellar gas |
\(10^6\) |
\(10^4\) |
\(10^{-10}\) |
\(10\) |
The plasma parameter, \(N\), is the number of particles of a plasma which are contained within the Debye sphere,
For a good plasma we want \(N \gg 1\). For the tokomak, \(N \approx 1\e{6}\), for the interstellar gas, \(N \approx 1\e{9}\). So, the Interstellar medium’s plasma is better than the tokomak’s.
We know that the plasma is electrically neutral over scales around the Debye Length, so there must be a restoring force driving the restoration of charge neutrality. This will produce oscillations about an equilibrium point (think of a swing).
\[ \begin{align}\begin{aligned} \label{eq:density-conservation} \frac{\partial n}{\partial t} + \nabla \cdot (n_{\rm e} \vec u) = 0\\so the change in the electron density over time, plus the flux of\end{aligned}\end{align} \]particles through a volume should be zero.
(0,0) circle (3); (4,1.7) node [black, right] Flux in … – (2,1.7); (2.5,0) – (4,0) node[black, below, right, text width=3.5cm] and Flux out are the only ways to change the internal density.;
Now, a change in electron population or density, relative to the equilibrium, produced by an electric field is
with \(\rho_{\rm f}\) the free-charge density. How do electrons respond to the electric field? We need the fluid momentum equation, which is a restatement of the conservation of momentum,
derivative,
otherwise stationary (and therefore electric-field free) equilibrium.
perturb the full equations to see how our small distribution evolves:
\(\omega_{\rm p}\), the plasma frequency, with \(\nu_{\rm p} = 9 \sqrt{n_0}\). This is an oscillation, but not a wave.
The plasma oscillation has consequences for the propogation of electromagnetic radiation. The restoring force which produces the plasma is a direct consequence of the plasma producing a displacement current. It turns out that we can treat the plasma as a dielectric medium, and that we can see this by considering the plasma’s repsonse to an oscillating imposed electric field,
phase. Charges in motion constitute a current, so for the current density we can write that
\(\epsilon_{\rm r}\) the relative permittivity of the dielectric. The full Maxwell equation reads
\(\epsilon = \epsilon(\omega)\); what’s the connection with refractive index?
notation for waves. Now, \(\nabla \times \vec{E} = - \frac{\partial \vec{B}}{\partial t}\), and we will take \(\vec{B}= \mu_0 \vec{H}\), so,
in the plasma. There is then a dispersion relation,
then
imaginary, and there is no wave propagation. If \(\omega > \omega_{\rm p}\) waves can propagate, but they will be affected. If :math:`omega = omega_{rm
p}`—as we see more closely in the full cold plasma treatment, this
represents wave absorption. The full dispersion relation can then be written
\[ \begin{align}\begin{aligned} \label{eq:momentumelectron} m \vec{\dot{v}} &= q(\dot{\vec{v}} \times \vec{B})\\and assume that :math:`\vec{B}` lies in the\end{aligned}\end{align} \]\(\vec{\hat{z}}\)-direction, then
\[ \begin{align}\begin{aligned}\vec{B} = \vec{\hat{z}} B_0\\so\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\dot{\vec{v}} = \frac{q}{m} \left[ \vec{v} \times ( \hat{\vec{z}} B_0) \right]\\so in components,\end{aligned}\end{align} \]\[\dot{v_x} = \frac{q}{m} \left[ \dot{v_y}B_0 - \dot{v}_z 0\right] = \frac{q B_0}{m} v_y\]\[\dot{v_y} = \frac{q}{m} \left[ \dot{v_z} 0 - \dot{v}_x B_0\right] = - \frac{q B_0}{m} v_x\]\[ \begin{align}\begin{aligned}\dot{v_z} = \frac{q}{m} 0 = {\rm const}\\so, defining the cyclotron frequency\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned} \label{eq:cyclotron} \omega_{\rm c} = \frac{|q|B}{m}\\and for electrons,\end{aligned}\end{align} \]\(\nu_{\rm c} = 28\ \giga \hertz\ \tesla^{-1}\), hence
\[ \begin{align}\begin{aligned}\ddot{v}_{x} = \omega_{\rm c} \dot{v_y} = \omega_{\rm c} [-\omega_{\rm c} v_x]\\that is\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\ddot{v}_x + \omega_{\rm c}^2 v_x = 0\\which is plane perpendicular to the charged particle (e.g. an\end{aligned}\end{align} \]electron), which will thus undergo circular motion. THis motion is, however, uniform along the field. The net effect is that the particle will describe a helix.
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} n_{\rm s} & = n_0_{\rm s}+ n_1(\vec{x},t) \\ \vec{v_{\rm s}} & = \vec{v_{0,{\rm s}}} + \vec{v_1}(\vec{x},t)\end{aligned}\end{split}\\Now, under perturbation, we can linearlise the most important\end{aligned}\end{align} \]equations:
Recall Maxwell’s equations
(\(\vec{v}(E)\)), then, \(\vec{J} = \vec{J}(\vec{v}) = \vec{J}(\vec{E})\), with \(\vec{J} = \vec{\sigma} \vec{E}\), then if we have a conductivity law we can move to a dielectric description.
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} -i\omega v_x &= \frac{q}{m} E_x - \frac{\omega_c}{i \omega} \qty[ \frac{q}{m}E_y - \omega_cv_x] \\ -i \omega v_x - \frac{\omega_c^2}{i \omega} v_x &= \frac{q}{m} E_x - \frac{q \omega_c}{i \omega m}E_y \\ &= \frac{q}{m}\qty[E_x - \frac{\omega_c}{i \omega} E_y] \\ \qty(1 - \frac{\omega_c^2}{\omega^2}) v_x &= - \frac{q}{i \omega m} \qty[E_x - \frac{\omega_c}{i \omega}E_y] \\ \text{since } \qty(1- \frac{\omega_c^2}{c^2}) v_y &= - \frac{q}{i \omega m} \qty[ E_x + \frac{\omega_c}{i \omega} E_y] \\ (1 - \frac{\omega_c^2}{\omega^2} \vec{v} &= M \cdot \vec{E} \\ \vec{v} &= \frac{1}{\qty(1 - \frac{\omega_{c}^2}{\omega^2})} \cdot M \cdot \vec{E}\end{aligned}\end{split}\\So, we know that we can get to this expression for\end{aligned}\end{align} \]\(\vec{J}\)—how does this help? We now bring Maxwell’s equations into the mixture.
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} \nabla \times \vec{E} &= \pdv{\vec{B}}{t} \\ \nabla \times ( \nabla \times \vec{E}) &= - \nabla \times \qty( \pdv{\vec{B}}{t}) \\ &= - \pdv{t} \qty(\nabla \times \vec{B}) \\ &= - \pdv{t} \qty[\mu_0 \vec{J} + \frac{1}{c^2} \pdv{\vec{E}{t}}] \\ &= - \pdv{t} \qty[\mu_0 \vec{\sigma} \cdot \vec{E} + \frac{1}{c^2} \pdv{\vec{E}}{t}]\end{aligned}\end{split}\\We are interested in waves, where solutions are proportional to\end{aligned}\end{align} \]\(e^{i \vec{k}\cdot \vec{r} - i \omega t}\)
\[ \begin{align}\begin{aligned} \begin{aligned} \nabla \times (\nabla \times \vec{E}) &= - \vec{k} \times (\vec{k} \times \vec{E})\end{aligned}\\RHS:\end{aligned}\end{align} \]
\[ \begin{align}\begin{aligned} \begin{aligned} i \omega \qty[ \mu_0 \vec{\sigma} \cdot \vec{E} - \frac{i \omega}{c^2} \vec{E}]\end{aligned}\\So, the full equation is\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\vec{k} \times (\vec{k} \times \vec{E}) + \frac{\omega^2}{c^2} K \cdot \vec{E} &= 0\\with\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}K = I + \frac{i \sigma}{\epsilon_0 \omega}\\being the dielectric tensor.\end{aligned}\end{align} \]
\[ \begin{align}\begin{aligned}\vec{n}= \frac{\vec{k}c}{\omega}\\(a refractive index with “directional complications”),\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\label{eq:genrefind} \vec{n} \times ( \vec{n} \times \vec{E} ) + K \cdot \vec{E} = 0\\To help to understand the significance of equation ([eq:genrefind]),\end{aligned}\end{align} \]let’s choose a geometry—let’s put \(\vec{B}_0 = \vec{\hat{z}} B_0\), and let’s take a wave in the \(\vec{\hat{x}}-\vec{\hat{z}}\)-plane (with one component parallel to \(\vec{B}_0\), and one perpendicular),
\[ \begin{align}\begin{aligned} \begin{aligned} \vec{n} &= \hat{x} n \sin(\theta) + \hat{z} n \cos(\theta)\end{aligned}\\Then expand the vector cross-product\end{aligned}\end{align} \]\(\vec{n} \times (\vec{n} \times \vec{E})\) for this choice of geometry.
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} \begin{pmatrix} S-n^2 \cos^2\theta & - iD & n^2 \cos(\theta) \sin(\theta) \\ iD & S-n^2 & 0 \\ n^2 \cos\theta \sin\theta & 0 & P- n^2 \sin^2 \theta \end{pmatrix} \begin{pmatrix} E_x \\ E_y \\ E_z \end{pmatrix} = 0\end{aligned}\end{split}\\Where we have written\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \label{eq:dielectricten} K= \begin{pmatrix} S & -iD & 0 \\ iD & S & 0 \\ 0 & 0 & P \end{pmatrix}\end{split}\\Where\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} S &= \frac{1}{2}(R+L) \\ D &= \frac{1}{2}(R-L)\end{aligned}\end{split}\\and\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} R &= 1 - \sum_S \frac{\omega^2_{\rm P_s}}{\omega^2} \qty( \frac{\omega}{\omega+\epsilon_0 \omega_{\rm c_s}}) \\ &= 1 - \frac{\omega_{\rm P}^2}{(\omega+\omega_{\rm C_+})(\omega-\omega_{\rm c^-})} \\ L &= 1 - \sum_S \frac{\omega_{\rm P_s}^2}{\omega^2} \qty(\frac{\omega}{\omega-\epsilon_{\rm s} \omega_{\rm c_s}}) \\ &= 1 - \frac{\omega_{\rm p}^2}{(\omega-\omega_{\rm c_+})(\omega+\omega_{\rm c_-})} \\ \epsilon_{\rm s} &= \begin{cases} +1 & \text{ for positive ion } \\ -1 & \text{ for negative electron} \end{cases} \\ \omega_{\rm P}^2 &= \omega_{\rm P_+}^2 + \omega_{\rm P_{-}}^2 \\ P &= 1 - \frac{\omega_{\rm P}}{\omega^2}\end{aligned}\end{split}\\For a non-trivial electric field the determinant of the matrix must\end{aligned}\end{align} \]vanish, giving a relationship between \(\omega\) and \(k\) so th dispertion relation
\[ \begin{align}\begin{aligned} \label{eq:dispertionrelation} An^4 - Bn^2 + C = 0\\with\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} A &= S \sin[2](\theta) + P \cos[2](\theta) \\ B &= RL \sin[2](\theta) + PS (1 + \cos[2](\theta) \\ C &= PRL\end{aligned}\end{split}\\*N.B. There are two spherical cases, :math:`\theta=0`, and\end{aligned}\end{align} \]:math:`theta= frac{pi}{2}`*.
\[ \begin{align}\begin{aligned}\begin{split} \begin{pmatrix} S-n^2 & -iD & 0 \\ iD & S-n^2 & 0 \\ 0 & 0 & P \end{pmatrix} \begin{pmatrix} E_x \\ E_y \\ E_z \end{pmatrix} = 0\end{split}\\i.e. :math:`(S-n^2)^2 - D^2 = 0` if :math:`E_x, E_y \neq 0` or\end{aligned}\end{align} \]:math:` P=0` if \(E_z\neq 0\).
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} n^2 = R & \frac{iE_x}{E_y} = - \frac{S-n^2}{D} = - \frac{S-R}{D} = 1 & \text{right circ. pol.} \\ n^2 = L & \frac{i E_x}{E_y} = - \frac{S-n^2}{D} = - \frac{S-L}{D} = -1 & \text{left circ. pol.}\end{aligned}\end{split}\\*N.B. Recall that\end{aligned}\end{align} \]\(S = \frac{1}{2} (R+L) \therefore S-R = \half{}(-R+L) = -D\) etc.*
\[ \begin{align}\begin{aligned} \label{eq:cutoff} \omega^2 \mp (\omega_{\rm re}-\omega_{\rm c_i}) \omega - (\omega_p^2 + \omega_{c_i}\omega_{c_e})=0\\i.e. at\end{aligned}\end{align} \]\(\omega \sim \mp \half \omega_{\rm c_e}+ \qty(\omega_{\rm p}^2 + \frac{1}{4}\omega_{c_{e}}^2)^{\frac{1}{2}}\). At the low-frequency limit, if \(\omega \ll \omega_{\rm c_i}\) (the lowest natural frequency is \(\omega_{\rm c_i}\), then \(R \sim L \sim 1+ \frac{c^2}{c_A^2}\), where \(C_A^2 = \frac{B_0^2}{\mu_0\rho_0}\) is the Alfven speed. The refractive index is \(n^2 = 1+\frac{c^2}{c_A^2}\), i.e. \(\omega^2 = \frac{k^2 c^2}{1 + \frac{c^2}{c_A^2}} \approx k^2c_A^2\) which descripe non-dispersive Alfven waves (c.f. Magnetohydrodynamics later).
For \(\theta = \frac{\pi}{2}\), waves propagating perpendicular to \(B_0\), the dispertion relation becomes
Characteristics of these modes are best seen in a matrix system.
For \(E_x = E_y = 0\), and \(E_z \neq 0\), we have \(n^2=P\), i.e. we have the same dispertion relation as in the unmagnetic case. Note, \(\vec{k}\cdot \vec{E} = 0\), so we have transverse electormagnetic waves which are independent of \(B_0\), and \(E\) is parallel to \(B_0\), so all motion is aligned with \(B_0\), and therefore
Now suppose we consider \(E_z=0\), \(E_x, E_y \neq 0\), then \(n^2 = \frac{RL}{S}\) in solution,
partly longitudinal, and partly transverse, since both \(E_x, E_y \neq 0\), and thus \(\vec{k} \cdot \vec{E} \neq 0\). The fact that \(E\) is perpendicular to \(B_0\) means that same for the gyration about the magnetic field which is generated, so wave properties depend on \(B_0\),
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} R & = 1 - \frac{\omega_{\rm p}^2}{(\omega+\omega_{\rm c_+})(\omega-\omega_{c_-})} \\ L & = 1 - \frac{\omega_{\rm p}}{(\omega-\omega_{\rm c_+})(\omega+\omega_{\rm c_-})}\end{aligned}\end{split}\\since :math:`\omega \ll \omega_{c_i}`\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} R &\approx 1 - \frac{\omega_p^2}{\omega_{c_i}(-\omega_{c_e})} \\ &= 1 + \frac{\frac{ne^2}{\epsilon_0 m_e} + \frac{n e^2}{\epsilon_0 m_i} } {\frac{eB_0}{m_i} \frac{e B_0}{m_e}} \\ &= 1 + \frac{n(m_i+m_e)}{\epsilon_0 B_0^2} \\ &\approx 1 + \frac{\mu_0 \rho_0 c^2}{B_0^2} \\ &= 1 + \frac{c^2}{c_A^2} \approx \frac{c^2}{c_A^2}\end{aligned}\end{split}\\For :math:`\theta = 0`, :math:`\omega \ll \omega_{c_i}`, which are\end{aligned}\end{align} \]both circular polarisations for transverse Alfven waves. Then,
\[ \begin{align}\begin{aligned}n^2 = \frac{kL}{S} \therefore \omega^2 = k^2 c_A^2\\and\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}S = \half (R+L)\\but :math:`n^2 = P`, so there is a cutoff; no low frequency is\end{aligned}\end{align} \]possible.
\[ \begin{align}\begin{aligned}\begin{split} \begin{pmatrix} S & -iD & 0 \\ iD & S-n^2 & 0 \\ 0 & 0 & P-n^2 \end{pmatrix}\end{split}\\So\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} S E_x - iD E_y & = 0 \\ \frac{i E_x}{E_y} &= - \frac{D}{S}\end{aligned}\end{split}\\So what are the implications for the Alfven wave? Since :math:`R`,\end{aligned}\end{align} \]\(L\) are approximately equal, for \(\omega \ll \omega_c\),
\[ \begin{align}\begin{aligned}\qty|\frac{E_x}{E_y}| \ll 1\\and so :math:`|E_y| \gg |E_x|`. From the equations of motion,\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} \dot{v}_x &= - \frac{e}{m} E_x - \omega_c v_y \\ \dot{v}_y &= - \frac{e}{m} E_y - \omega_c v_x\end{aligned}\end{split}\\we can differentiate the system with respect to :math:`t` to show\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\qty| \frac{v_x}{v_y} | \approx \frac{\omega_c}{\omega} \quad \text{when}\quad \qty| \frac{E_x}{E_y} | \approx 0\\We know that the O-mode cuts off at :math:`\omega=\omega_{\rm p}`.\end{aligned}\end{align} \]For the X-mode, \(n^2= \frac{RL}{S}\), the two cutoffs occur at \(R=0\) or \(L=0\). This is the same as the circuarly polarised waves as before, but we have two cutoff frequencies for the same mode. Resonance occurs at \(S=0\), and again, this occurs at two places, an upper and a lower hybrid frequency:
\[\begin{split}\begin{aligned} \omega_u^2 &= \omega_p^2 + \omega_{c_e}^2 \\ \omega_l^2 &= \omega_p^2 \frac{\omega_{c_i}\omega_{c_e}}{\omega_p^2+\omega_{c_e}^2}\end{aligned}\end{split}\]
Here we consider the behaviour of plasmas at long wavelengths, and low frequencies; in this limit we can retrieve classical thermodynamic relations. Starting with the model equations,
Notably, Ampere’s law contains no mention of displacement current, and there is no allowance for charge separation. An ideal MHD plasma exhibits perfect conductivity, so
at the normal modes, just as with the cold plasma case. Take a perturbation,
Then we assume \(\vec{u}_0 = 0\) (stationary equilibrium), and \(\vec{J}_0 = \frac{\nabla \times \vec{B}}{\mu_0} = 0\).
Then linearise,
then, assuming all perturbed quantities are proportional to \(\exp[ i(\vec{k} \cdot \vec{r} - \omega t)]\), so
\(\nabla \cdot \vec{B} = 0\),
obtain everything in terms of \(u_1\). \(p_1\) can be eliminated using (IV), \(\rho_1\) using (I), and \(B_1\) using (III).
This process yields
Then,
the minimum cyclotron frequency for ions and plasma frequency, and the wavelength is much greater than the Debye length of the Larmour Radius. Consider the component of \(\vec{u}_1\) perpendicular to the direction of motion, \(\hat{z}\). To find this direction take the cross-product of the whole equation ([eq:mhdincoordis]), with \(\hat{z}\). Now take \(\hat{z} \cdot \hat{b} = \cos(\theta)\).
\(\hat{z} \cdot \vec{u}_1 = 0\), and we still have \((\omega^2 - k^2 \cos[2](\theta))(\hat{z} \times \vec{u}_1) = 0\). Clearly a non-trivial solution, governed by the dispertion relation is possible: a wave which satisfies
frequency–for \(\theta=0\) this is just like a cold plasma solution, but at \(\theta=\frac{\pi}{2}\) there is no transverse wave solution, but the cold plasma has the low frequency limit of the X-mode, the compressed Alfvén mode.
The general solution for MHD waves is that we have a dispertion reltion,
Alfven
Fast (+) and Slow (-) Magnetosonic
For the fast MS mode, \(\frac{B^2}{2 \mu_0}\) magnetic pressure enhanes the thermodynamic pressure, \(p\), by varying in phase with it. Fr the slow MS mode the magnetic presure and thermodynamic pressure oppose one another. Magnetic pressure plays a powerful conceptual role in MHD.
Recall the equilibrium; to see how significant the magnetic pressure can be,
\(\vec{B}_0\) and \(p_0\), however, we can generalise this, since
\(\vec{B}_0 \cdot \nabla \vec{B}_0 = 0\), for equilibrium,
Plasma tends to avoid the strongest field regions in equilibrium. This is a possible mechanism for confinement. How could this go wrong? Suppose we have a cylinder of plasma, which is carrying current. It has an azimuthal \(\vec{B}_0\)
(2,0) ellipse (.2 and .5);
(-2,-.5) rectangle (2, .5); (-2,0) ellipse (.2 and .51);
in -1,-.5,…, 1.5 (, -.5) ..controls (+.2, -.3) and (+.2, .3) .. (,.5); (-1.8, -.5) ..controls (-1.8+.2, -.3) and (-1.8+.2, .3) .. (-1.8,.5) node [midway, right] \(\vec{B}_0\);
(0,.9) – (2,.9) node [midway, fill=white] \(\vec{J}\);
If we bend the cylinder, but wish to maintain \(\vec{J}_0\),
The magnetic pressure is stronger on the inside of the bend than the outside. Thus the plasma is push upwards, wosening the distortion. This gives a kink instability.
in 1.5,2,…, 4 (7,-) .. controls (3,-) and (2,-) .. (-1,-) (7,-) .. controls (10,-) and (12,-) .. (15,-); (7,) .. controls (3,) and (2,) .. (-1,) (7,) .. controls (10,) and (12,) .. (15,);
We find that \(\vec{B}_0\) is more intense at the pinch because \(\vec{J}_0\) is larger. Hence the plasma is expelled from the pinched region, drawing the current density up, and making the problem worse. The plasma is described as pinching off.
Consider a binary collision of two plasma particles, where the Coulomb force between the particles is
ion.
(0,0) circle (0.5) node [midway, white] +; (-4,1) circle (0.2) node [white] -;
(-3.7,1) – (0,1) arc (90:40:1) – ++(130:-2);
(0,1) – (4,1) (1,0)–(4,0);
(3,1) – (3,0) node [right, midway] \(b = r \sin(\theta)\);
Consider the change of \(v_{\perp}\), with a massive ion, \(m_{\rm ion} \gg m_{\rm e}\). The change of perpendicular momentum from the equation of motion can be written
\(90^{\circ}\)), the change of \(m_{\rm e} v_{\perp}\) is of the order of \(mv\) itself, so,
The interaction between two particles can be described using the interaction cross-section, \(\sigma\), where
with \(b\) being the area of the disc. We simply assume the interaction is happening for impact parameters
For this interaction the cross-section is
\(n_{\rm i} \approx n_{\rm e}\), and \(q_{\rm i} = q_{\rm e} = e\) is
plasma, and a more rigorous estimate indicates
\(\Lambda \sim n_{\rm e} \lambda^3_{\rm De}\). \(\log(\Lambda)\) is the Coulomb logarithm, and is normally assumed to be constant, with a value in the range \(\log(\Lambda) \in [10,30]\).
Let us estimate the mean free path of an electron in a plasma;
The mean free path of the solar corona. In the plasma composing the solar corona,
For electron-ion collisions in plasma
We define the quantity \(\tau_{\rm ei} := \frac{1}{\nu_{\rm ei}}\) which is the time between collisions, or the mean free time. For the frequency of electron-electron collisions (taking into account the finite mass of the scattering particle, and replacing it with \(m_{\rm e}\)) we have a factor of two, and so,
\(m_{\rm i}\), and so
i}}`, hence
If \(T_{\rm e} \neq T_{\rm i}\), there will be an exchange of temperature caused by the collisions, and the timescales of the interactions are
It is worth noting that \(\nu_{\rm ei}\) is not the rate at which equilibrium is established between electrons and ions, but is instead the rate of momentum transfer from electrons to ions, and not the rate of energy transfer between them. The relaxation time for electorn-ion equiilibrium is given by ion-electron collisions, and \(\frac{m_{\rm e}}{m_{\rm i}} \nu_{\rm ee}\). For a Hydrogen plasma
Consider an unmagnetised quasineutral plasma with \(n_{\rm i} \approx n_{\rm e}\) of electrons and ions both with charge \(q = e\). In response to an applied electric field, \(\vec{E}\), a current will flow in the plasma. The current density will be
is predominantly carried by the electrons, hence,
electron-electron collisions contributing to the resisitivity, and so
and decreases with growing temperature.
The fluid equation of motion including collisions for electrons is
\(n_{\rm e} \approx n_{\rm i} = n\), and \(\nu_{\rm ei}\) is constant. Considering a steady state,
diffusion coefficient. These differ for each species.
The flux, \(\vec{\Gamma}\), is
uncharged we find Fick’s Law,
A plasma should be quasi-neutral, so diffusion of electrons and ions should adjust to some degree to preserve quasineutrality. The fast-moving electrons have higher thermal velocities and tend to leave a plasma first. The positive charge is then left behind, and an electric field is setup to retard the loss of electrons, and accelerate the loss of ions. We let \(\vec{\Gamma}_{\rm e} = \vec{\Gamma}_{\rm i} = \vec{\Gamma}\), so
There are some phenomena which neither MHD nor single-particle descriptions of a plasma can describe. For these situations we need to consider the velocity distribution, \(f(\vec{v})\) of the plasma. In fluid theory the independent variables are functions of \(\vec{r}\) and \(t\) only, which is because the velocity distribution is taken to be Maxwellian everywhere, so can be uniquely speciied by the temperature, \(T\), and the number density, \(n(\vec{r}, t)\),
If \(f\) is correctly normalised it describes the probbaility of finding a particle in the range \(\vec{r} \in (\vec{r}, \vec{r}+\dd{\vec{r}})\) and \(\vec{v} \in (\vec{v}, \vec{v}+\dd{\vec{v}})\). \(f\) is a function in seven variables, and, if it is a Maxwellian distribution, it has the form,
Ignoring collisions, and assuming the plasma to be a closed system, with no sources or sinks of particles, the function will obey the Liouville theorem, so
for the time derivative along a trajectory in \((\vec{r}, \vec{v})\)-phase space, so
electromagnetic the equation takes the form
of Maxwell’s equations, as \(\vec{E}\) and \(\vec{B}\) are the average values of the electric and magnetic fields from particles in the plasma.
If there are collisions in the plasma, \(\dv{f}{t}\neq 0\), and, using the collision integral,
to include sources or sinks of particles by adding terms on the right hand side.
The collision term can sometimes be approximated as
is the collision time. This is the Krook collision term.
As an illustration of the use of the Vlasov equation, we consider the electron plasma oscillations in a uniform plasma with no applied magnetic or electric fields. Consider a first order perturbation,
Poisson’s equation (\(\epsilon_0 \nabla \vec{E} = \rho\)),
waves in the plasma are plane waves in the \(x\)-direction. Then,
\(v_z\) can be carried out, and
\(v_x = \frac{\omega}{k}\). Landau suggested (1946) letting \(\omega \to \omega + i o\) for a small \(o\), which makes the integral
This can be written symbolically as
First, concentrate on the real part, where the integral can be computed by integrating by parts,
velocities), and so we can expand \((v - \frac{\omega}{k})^{-2}\),
And now, using the expansion for the integral,
Then we can write
\(\omega_{\rm pe}^2\) in the second term, so
frac{k_{rm B} T_{rm e}}{m} frac{k^2}{omega_{rm pe}}`). Then, the group velocity is
Finally, the imaginary part. For simplicity ignore the thermal correction, so \(\omega(k) \approx \omega_{\rm pe}\), then
distribution for \(\tilde{f}\),
frac{omega_{rm pe}}{k}`,
described as “probably the most astounding result of plasma physics.”
The damping is not reandomisation by collisions, but a resonant (i.e.phase velocity of waves is the same as the velocity of interacting particles) \(v=\frac{\omega}{k}\) transfer of energy from waves to particles. It can be reversed if \(\pdv*{\tilde{f}_0}{v}>0\).
in 0,30,…,360 (B) (0,0)+(:7); (0,0) +(:7) – +(:10); (0,0) circle (7);
(0,0) circle (5); in 0,30,…,360 (2,0) – (:5) coordinate (A);
(0,0) node [fill=accent-blue, circle] 1; (2,0) node [fill=accent-green, circle] 2;
in 0,30,…,360 (0,0)+(:5) – +(:7);
(0,.6) – (2,.6) node [above, midway] \(\Delta v \ t\);
(0,-.6) – (-5,-.6) node [fill=white, midway] \(r = ct\); (-170:5) – (-170:7) node [below, yshift=-.1cm, midway] \(r^{\prime} = c \Delta t\);
The radial component has the usual form,
while the \(\vec{E}_{\phi}\) component is given by the number of radial field-lines per unit area in the direction of \(\vec{\phi}\). From geometry arguments,
\(\frac{\Delta v}{\Delta t} = \ddot{r}\), and substituting \(\vec{E}_r\),
strength varying with \(\frac{1}{r}\), in contrast to the radial field which varies with \(\frac{1}{r}\). This “kink” is an outward moving pulse of electromagnetic radiation. The power per unit area per second is given by the Poynting vector,
\(\dd{\Omega} = 2 \pi \sin(\phi) \dd{\phi}\),
radiated power from an accelerated charge, and can be applied to our specific case of a particle gyrating in an electric field.
Consider the rest frame of a charge \(q\), called \(S^{\prime}\), and a lab frame \(S\).
\(\theta\) the electron pitch angle) due to the relativistic transformations of \(\vec{E}\) and \(\vec{B}\). Using Larmour’s formula,
is Lorentz invariant (\(P = P^{\prime}\)), so the power in the lab frame must be the same. The total power radiated in the lab frame from a particle with pitch anfgle \(\theta\) is
\(U_{\rm mag} = \frac{B^2}{2 \mu_0}\).
The distribution of the radiation about the moving charge is worth considering;
If the particle is relativistic we must consider the effect on the cyclotron frequency. The relativistic frequency will be
where \(m_0\) is the rest-mass of the particle. This can be decomposed into a series of harmonics, \(\omega_n\),
distinctly different from the delta-function peak we expect from a non-relativistic charge. A distribution of electrons will in fact display a power-law spectrum. As the charge becomes more relativistic the dipole shape of the radiation is deformed, and an effect known as “relativistic beaming” will be observed, with the output of radiation becoming more and more focussed.
\[n^2 = R \qquad n^2 = L \quad L \neq R\]thus different polarisations have different phase speeds—we can exploit this as a remote diagnostic of the plasma conditions.
\[n^2 = \frac{k^2 c^2}{\omega^2}\]Thus it is possible to convert between \(n\) and \(\frac{\omega}{k}\). A superposition of the circuarly polarised waves produces an evolution along the propogation ray of the net \(\vec{E}\) polarisation direction.
(0,-4.4) – (7,-4.4) node [midway, below] \(z\);
(0,0) circle (2); (0,-2) – (0,2) node [midway, right, black] \(\vec{E}\); (2,4) node [text width=3cm, left, text ragged] At \(z=0\) \(R\) and \(L\) are synchronised, producing a vertical polarisation.; (-1,-3) circle (.8) node \(R\); (1,-3) circle (.8) node \(L\); (-1,-2.2) circle (0.1); (1,-2.2) circle (0.1); (-1,-2) arc (90:135:1); (1,-2) arc (270:225:-1);
(0,4) node [text width=3cm, text ragged] As \(z\) increases, \(R\) and \(L\) the faster rotation of the \(L\) mode causes the polarisation to become diagonal.; (0,0) circle (2); (0,-2) – (0,2) node [midway, right, black] \(\vec{E}\);
(-1,-3) circle (.8) node \(R\); (1,-3) circle (.8) node \(L\); (-1,-2.2) circle (0.1); (1,-2.2) circle (0.1); (-1,-2) arc (90:135:1); (1,-2) arc (270:225:-1);
Suppose the polarisation vector \(\vec{E}\) shifts by an angle \(\dd{\theta}\) as a result of the varying phase interference of the superposed \(R\) and \(L\) modes. Then,
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} R & = 1 - \frac{\omega_{\rm p}^2}{\omega(\omega-\omega_{\rm c_e})} \\ L & = 1 - \frac{\omega_{\rm p}^2}{\omega(\omega+\omega_{\rm c_e})}\end{aligned}\end{split}\\with :math:`\omega_{\rm c_+} \to 0`, as :math:`m_+ \to \infty`. Thus\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} \eval{n}_{\rm RCP} &\approx 1 - \frac{\omega_{\rm p}^2}{2 \omega (\omega-\omega_{\rm c_e})} \\ \eval{n}_{\rm LCP} &\approx 1 - \frac{\omega_{\rm p}^2}{2 \omega(\omega+\omega_{\rm c_e})}\end{aligned}\end{split}\\Then\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} (n_R - n_L) &= \frac{\omega_{\rm p}^2}{2 \omega} \qty[ \frac{1}{\omega - \omega_{\rm c_e}} - \frac{1}{\omega(\omega+\omega_{\rm c_e})}]\\ &= \frac{\omega_{\rm p}^2}{2 \omega} \frac{\omega + \omega_{\rm c_e}-(\omega-\omega_{\rm c_e})}{\omega^2 - \omega_{\rm c_e}^2} \\ &= \frac{\omega_{\rm p}^2 \omega_{\rm c_e}}{\omega(\omega^2-\omega_{\rm c_e}^2)} \\ &\approx \frac{\omega_{\rm p}^2 \omega_{\rm c_e}}{\omega^3} \\\end{aligned}\end{split}\\Hence,\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} \dd{\theta} &= \half \qty| (k_R - k_L)| \dd{z} \\ & \approx \half \int_0^2 \frac{\omega_{\rm p}^2 \omega_{\rm c_e} \dd{z}}{c \omega^2} \\ &= \half \int_0^2 \frac{ \frac{ne^2}{m_{\rm e}^2} \frac{eB}{m_{\rm e}}}{ c \omega^2} \dd{z}\\ &= \frac{e^2}{2 \epsilon_0 m_{\rm e}^2 c \omega^2} \int_0^2 nB \dd{z}\end{aligned}\end{split}\\We could allow :math:`n` and :math:`B` to vary slowly, i.e. with the\end{aligned}\end{align} \]gradient-scale length of \(n,B\) much greater than the wavelength of the circular poalrisation modes, and still have a Faraday rotation result which is reasonably slow.
We can generalise this result, with
field direction and the line-of-sight direction.
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} S &= \half \qty( 1 + \frac{\omega_{\rm p}^2}{(\omega + \omega_{\rm ci})(\omega_{\rm ei}-\omega)} + 1 - \frac{\omega_{\rm p}^2}{(\omega - \omega_{\rm ei})(\omega+\omega_{\rm ce})}) \\ &= \half \qty( 2+ \omega_{\rm p}^2 \frac{(\omega-\omega_{\rm ci})(\omega+\omega_{\rm ce}) - (\omega+\omega_{\rm ci})(\omega_{\rm ce}-\omega)}{(\omega^2-\omega_{\rm ci}^2)(\omega_{\rm ce}^2 - \omega^2)}) \\ &= \half \qty( 2+ \omega_{\rm p}^2 \frac{\omega \qty(\omega+\omega_{\rm ce} - (\omega_{\rm ce}-\omega))}{\omega^2(\omega_{\rm ce}^2 - \omega^2)})\\ &= \half \qty( 2+ \omega_{\rm p}^2 \frac{2 \omega^2}{\omega^2 \qty(\omega_{\rm ce}^2 - \omega^2)} )\\ &= \half \qty( 2+ \omega_{\rm p}^2 \frac{2 \omega^2}{\omega^2 \qty(\omega^2_{\rm ce} - \omega^2)}) \\ &= 1 + \frac{\omega_{\rm p}^2}{\qty( \omega_{\rm ce}^2 - \omega^2)}\end{aligned}\end{split}\\Now introducing :math:`\Gamma_{\rm p} := \frac{\omega_{\rm p}^2}{\omega^2_{\rm ce} - \omega^2}`, finally,\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} S &= \half (R+L) = 1+r_{\rm p}\\ D &= \half (R-L) = r_{\rm p} \frac{\omega_{\rm ce}}{\omega}\\ P &= 1 - \frac{\omega_{\rm p}^2}{\omega^2} \approx S\end{aligned}\end{split}\\In the limit :math:`\omega \gg \omega_{\rm ce}`. The general cold\end{aligned}\end{align} \]plasma dispertion relation is equation ([eq:dispertionrelation]), which can be solved for \(n^2\),
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} n^2 &= \frac{2 PS +(RL-PS)(\sin[2](\theta)) \pm \sqrt{(RL-PS)^2 \sin[4](\theta)} }{{2 \qty[ P+(S-P) \sin[2](\theta)]}} \\& \quad+ \frac{{4 P^2D^2 \cos[2](\theta)}}{2 \qty[ P+(S-P) \sin[2](\theta)]} \end{aligned}\end{split}\\In the limit :math:`\omega \gg \omega_{\rm ce}`, :math:`P \approx S`\end{aligned}\end{align} \]and
\[ \begin{align}\begin{aligned}RL - PS \approx RL -S^2 = -D^2\\hence\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned} \begin{aligned} n_{\pm}^2 &= \frac{2 S^2 - D^2 \sin[2](\theta) \pm \sqrt{D^4 \sin[4](\theta) + 4 S^2 D^2 \cos[2](\theta)}}{2S}\end{aligned}\\where the :math:`\pm` corresponds to the right and left polarisations\end{aligned}\end{align} \]respectively. We can further simplify for small angles of wave propogation, so \(\cos(\theta)\approx 1\) and \(v \approx 0\).
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} n_{\pm}^2 &\approx S \pm D \cos(\theta) \\ & \approx 1 + r_{\rm p} \pm r_{\rm p} \frac{\omega_{\rm ce}}{\omega} \cos(\theta) \\ & \approx \frac{\omega_{\rm ce}^2 - \omega^2 + \omega_{\rm p}^2}{\omega_{\rm ce}^2 - \omega^2} \pm \frac{\omega_{\rm p}^2}{\omega_{\rm ce}^2 -\omega^2} \frac{\omega_{\rm ce}}{\omega} \cos(\theta) \\ & \approx 1 \mp \frac{\omega_{\rm p}^2 \omega_{\rm ce}}{\omega^3} \cos(\theta)\end{aligned}\end{split}\\For :math:`\omega^2 \gg \omega_{\rm p}^2`,\end{aligned}\end{align} \]\(\omega^2 \gg \omega_{\rm ce}^2\). Then
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} n_+ & \approx \qty( 1 - \frac{\omega_{\rm p}^2 \omega_{\rm ce}}{\omega^3} \cos(\theta))^{\half} \approx 1 - \frac{\omega_{\rm p}^2 \omega_{\rm ce}}{2 \omega^3 \cos(\theta)} \\ n_- & \approx \qty( 1+ \frac{\omega_{\rm p}^2 \omega_{\rm ce}}{\omega^3} \cos(\theta) )^{\half} \approx 1 + \frac{\omega_{\rm p}^2 \omega_{\rm ce}}{2 \omega^3} \cos(\theta)\end{aligned}\end{split}\\And so,\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\Delta n = n_- - n_+ \approx \frac{\omega_{\rm p}^2 \omega_{\rm ce}}{\omega^3} \cos(\theta)\\Since :math:`n = \frac{kc}{\omega}`, we know\end{aligned}\end{align} \]\(k = \frac{\omega}{c} n\), and so
\[ \begin{align}\begin{aligned} \begin{aligned} K_- - K_+ \approx \frac{\omega_{\rm p}^2 \omega_{\rm ce}}{c \omega^2} \cos(\theta) \end{aligned}\\and so finally, the rotation angle is\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned} \label{eq:13} \dd{\phi} = \half \frac{\omega_{\rm p}^2 \omega_{\rm ce}}{c \omega^2} \cos(\theta) \dd{z}\\At the high wave frequency limit,\end{aligned}\end{align} \]\[ \begin{align}\begin{aligned}\omega = k c = \frac{2 \pi c}{\lambda}\\Hence, as :math:`\omega_{\rm p}\end{aligned}\end{align} \]= sqrt{frac{4 pi e^2 n_{rm e}}{4 pi m epsilon_0}}`, and \(\omega_{\rm c} = \frac{eB}{m}\),
\[ \begin{align}\begin{aligned}\begin{split} \begin{aligned} \dv{\phi}{z} &= \half \frac{e^2 n }{m \epsilon_0} \frac{eB}{m} \frac{1 \lambda^2}{c (2 \pi)^2 c^2} \cos(\theta) \\ &= \frac{e^3 B n_{\rm e} \lambda^2}{8 \pi^2 m^2 \epsilon_0 c^3} \cos(\theta)\end{aligned}\end{split}\\And then, to find the Farday angle after the radiation has travelled\end{aligned}\end{align} \]along a path \(r\),
\[ \begin{align}\begin{aligned} \label{eq:14} \phi_{\rm F} = \frac{e^3 \lambda^2}{8 \pi^2 m^2 \epsilon_0 c^3} \int_0^r B(z) n_{\rm e}(z) \cos(\theta) \dd{z}\\Assuming a uniform number density the Faraday rotation is a measure\end{aligned}\end{align} \]of the magnetic field along the path.
Air generally has around \(10^{9} {\rm m}^{-3}\) ions and electrons which are caused by background sources and friction. These can be harnassed as seeds for further ionisation.