diff options
author | Prefetch | 2021-09-22 15:33:15 +0200 |
---|---|---|
committer | Prefetch | 2021-09-22 15:33:15 +0200 |
commit | bb79f9b1beee85f2290f3bed9b62eacaca445602 (patch) | |
tree | f0bde207897aa6514d91371baa92b9e00c727cb1 | |
parent | 42d409fa774efb8206ae5c701d5cbcc4ae1d9cad (diff) |
Expand knowledge base
-rw-r--r-- | content/know/category/plasma-physics.md | 9 | ||||
-rw-r--r-- | content/know/concept/guiding-center-theory/index.pdc | 522 | ||||
-rw-r--r-- | content/know/concept/lorentz-force/index.pdc | 182 | ||||
-rw-r--r-- | content/know/concept/rabi-oscillation/index.pdc | 218 | ||||
-rw-r--r-- | content/know/concept/time-dependent-perturbation-theory/index.pdc | 5 |
5 files changed, 818 insertions, 118 deletions
diff --git a/content/know/category/plasma-physics.md b/content/know/category/plasma-physics.md new file mode 100644 index 0000000..0d57a87 --- /dev/null +++ b/content/know/category/plasma-physics.md @@ -0,0 +1,9 @@ +--- +title: "Plasma physics" +firstLetter: "P" +date: 2021-09-18T14:33:19+02:00 +draft: false +layout: "category" +--- + +This page will fill itself. diff --git a/content/know/concept/guiding-center-theory/index.pdc b/content/know/concept/guiding-center-theory/index.pdc new file mode 100644 index 0000000..12abac0 --- /dev/null +++ b/content/know/concept/guiding-center-theory/index.pdc @@ -0,0 +1,522 @@ +--- +title: "Guiding center theory" +firstLetter: "G" +publishDate: 2021-09-21 +categories: +- Physics +- Electromagnetism +- Plasma physics + +date: 2021-09-18T13:47:41+02:00 +draft: false +markup: pandoc +--- + +# Guiding center theory + +When discussing the [Lorentz force](/know/concept/lorentz-force/), +we introduced the concept of *gyration*: +a particle in a uniform [magnetic field](/know/concept/magnetic-field/) $\vb{B}$ +*gyrates* in a circular orbit around a **guiding center**. +Here, we will generalize this result +to more complicated situations, +for example involving [electric fields](/know/concept/electric-field/). + +The particle's equation of motion +combines the Lorentz force $\vb{F}$ +with Newton's second law: + +$$\begin{aligned} + \vb{F} + = m \dv{\vb{u}}{t} + = q \big( \vb{E} + \vb{u} \cross \vb{B} \big) +\end{aligned}$$ + +We now allow the fields vary slowly in time and space. +We thus add deviations $\delta\vb{E}$ and $\delta\vb{B}$: + +$$\begin{aligned} + \vb{E} + \to \vb{E} + \delta\vb{E}(\vb{x}, t) + \qquad \quad + \vb{B} + \to \vb{B} + \delta\vb{B}(\vb{x}, t) +\end{aligned}$$ + +Meanwhile, the velocity $\vb{u}$ can be split into +the guiding center's motion $\vb{u}_{gc}$ +and the *known* Larmor gyration $\vb{u}_L$ around the guiding center, +such that $\vb{u} = \vb{u}_{gc} + \vb{u}_L$. +Inserting: + +$$\begin{aligned} + m \dv{t} \big( \vb{u}_{gc} + \vb{u}_L \big) + = q \big( \vb{E} + \delta\vb{E} + (\vb{u}_{gc} + \vb{u}_L) \cross (\vb{B} + \delta\vb{B}) \big) +\end{aligned}$$ + +We already know that $m \: \dv*{\vb{u}_L}{t} = q \vb{u}_L \cross \vb{B}$, +which we subtract from the total to get: + +$$\begin{aligned} + m \dv{\vb{u}_{gc}}{t} + = q \big( \vb{E} + \delta\vb{E} + \vb{u}_{gc} \cross (\vb{B} + \delta\vb{B}) + \vb{u}_L \cross \delta\vb{B} \big) +\end{aligned}$$ + +This will be our starting point. +Before proceeding, we also define +the average of $\expval{f}$ of a function $f$ over a single gyroperiod, +where $\omega_c$ is the cyclotron frequency: + +$$\begin{aligned} + \expval{f} + \equiv \int_0^{2 \pi / \omega_c} f(t) \dd{t} +\end{aligned}$$ + +Assuming that gyration is much faster than the guiding center's motion, +we can use this average to approximately remove the finer dynamics, +and focus only on the guiding center. + + +## Uniform electric and magnetic field + +Consider the case where $\vb{E}$ and $\vb{B}$ are both uniform, +such that $\delta\vb{B} = 0$ and $\delta\vb{E} = 0$: + +$$\begin{aligned} + m \dv{\vb{u}_{gc}}{t} + = q \big( \vb{E} + \vb{u}_{gc} \cross \vb{B} \big) +\end{aligned}$$ + +Dotting this with the unit vector $\vu{b} \equiv \vb{B} / |\vb{B}|$ +makes all components perpendicular to $\vb{B}$ vanish, +including the cross product, +leaving only the (scalar) parallel components +$u_{gc\parallel}$ and $E_\parallel$: + +$$\begin{aligned} + m \dv{u_{gc\parallel}}{t} + = \frac{q}{m} E_{\parallel} +\end{aligned}$$ + +This simply describes a constant acceleration, +and is easy to integrate. +Next, the equation for $\vb{u}_{gc\perp}$ is found by +subtracting $u_{gc\parallel}$'s equation from the original: + +$$\begin{aligned} + m \dv{\vb{u}_{gc\perp}}{t} + = q (\vb{E} + \vb{u}_{gc} \cross \vb{B}) - q E_\parallel \vu{b} + = q (\vb{E}_\perp + \vb{u}_{gc\perp} \cross \vb{B}) +\end{aligned}$$ + +Keep in mind that $\vb{u}_{gc\perp}$ explicitly excludes gyration. +If we try to split $\vb{u}_{gc\perp}$ into a constant and a time-dependent part, +and choose the most convenient constant, +we notice that the only way to exclude gyration +is to demand that $\vb{u}_{gc\perp}$ does not depend on time. +Therefore: + +$$\begin{aligned} + 0 + = \vb{E}_\perp + \vb{u}_{gc\perp} \cross \vb{B} +\end{aligned}$$ + +To find $\vb{u}_{gc\perp}$, we take the cross product with $\vb{B}$, +and use the fact that $\vb{B} \cross \vb{E}_\perp = \vb{B} \cross \vb{E}$: + +$$\begin{aligned} + 0 + = \vb{B} \cross (\vb{E}_\perp + \vb{u}_{gc\perp} \cross \vb{B}) + = \vb{B} \cross \vb{E} + \vb{u}_{gc\perp} B^2 +\end{aligned}$$ + +Rearranging this shows that $\vb{u}_{gc\perp}$ is constant. +The guiding center drifts sideways at this speed, +hence it is called a **drift velocity** $\vb{v}_E$. +Curiously, $\vb{v}_E$ is independent of $q$: + +$$\begin{aligned} + \boxed{ + \vb{v}_E + = \frac{\vb{E} \cross \vb{B}}{B^2} + } +\end{aligned}$$ + +Drift is not specific to an electric field: +$\vb{E}$ can be replaced by a general force $\vb{F}/q$ without issues. +In that case, the resulting drift velocity $\vb{v}_F$ does depend on $q$: + +$$\begin{aligned} + \boxed{ + \vb{v}_F + = \frac{\vb{F} \cross \vb{B}}{q B^2} + } +\end{aligned}$$ + + +## Non-uniform magnetic field + +Next, consider a more general case, where $\vb{B}$ is non-uniform, +but $\vb{E}$ is still uniform: + +$$\begin{aligned} + m \dv{\vb{u}_{gc}}{t} + = q \big( \vb{E} + \vb{u}_{gc} \cross (\vb{B} + \delta\vb{B}) + \vb{u}_L \cross \delta\vb{B} \big) +\end{aligned}$$ + +Assuming the gyroradius $r_L$ is small compared to the variation of $\vb{B}$, +we set $\delta\vb{B}$ to the first-order term +of a Taylor expansion of $\vb{B}$ around $\vb{x}_{gc}$, +that is, $\delta\vb{B} = (\vb{x}_L \cdot \nabla) \vb{B}$. +We thus have: + +$$\begin{aligned} + m \dv{\vb{u}_{gc}}{t} + = q \big( \vb{E} + \vb{u}_{gc} \cross \vb{B} + + \vb{u}_{gc} \cross (\vb{x}_L \cdot \nabla) \vb{B} + + + \vb{u}_L \cross (\vb{x}_L \cdot \nabla) \vb{B} \big) +\end{aligned}$$ + +We approximate this by taking the average over a single gyration, +as defined earlier: + +$$\begin{aligned} + m \dv{\vb{u}_{gc}}{t} + = q \big( \vb{E} + \vb{u}_{gc} \cross \vb{B} + + \vb{u}_{gc} \cross \expval{ (\vb{x}_L \cdot \nabla) \vb{B} } + + \expval{ \vb{u}_L \cross (\vb{x}_L \cdot \nabla) \vb{B} } \big) +\end{aligned}$$ + +Where we have used that $\expval{\vb{u}_{gc}} = \vb{u}_{gc}$. +The two averaged expressions turn out to be: + +$$\begin{aligned} + \expval{ (\vb{x}_L \cdot \nabla) \vb{B} } + = 0 + \qquad \quad + \expval{ \vb{u}_L \cross (\vb{x}_L \cdot \nabla) \vb{B} } + \approx - \frac{u_L^2}{2 \omega_c} \nabla B +\end{aligned}$$ + +<div class="accordion"> +<input type="checkbox" id="proof-nonuniform-B-averages"/> +<label for="proof-nonuniform-B-averages">Proof</label> +<div class="hidden"> +<label for="proof-nonuniform-B-averages">Proof.</label> +We know what $\vb{x}_L$ is, +so we can write out $(\vb{x}_L \cdot \nabla) \vb{B}$ +for $\vb{B} = (B_x, B_y, B_z)$: + +$$\begin{aligned} + (\vb{x}_L \cdot \nabla) \vb{B} + = \frac{u_L}{\omega_c} + \begin{pmatrix} + \sin\!(\omega_c t) \pdv{B_x}{x} + \cos\!(\omega_c t) \pdv{B_x}{y} \\ + \sin\!(\omega_c t) \pdv{B_y}{x} + \cos\!(\omega_c t) \pdv{B_y}{y} \\ + \sin\!(\omega_c t) \pdv{B_z}{x} + \cos\!(\omega_c t) \pdv{B_z}{y} + \end{pmatrix} +\end{aligned}$$ + +Integrating $\sin$ and $\cos$ over their period yields zero, +so the average vanishes: + +$$\begin{aligned} + \expval{ (\vb{x}_L \cdot \nabla) \vb{B} } + = 0 +\end{aligned}$$ + +Moving on, we write out $\vb{u}_L \cross (\vb{x}_L \cdot \nabla) \vb{B}$, +suppressing the arguments of $\sin$ and $\cos$: + +$$\begin{aligned} + \vb{u}_L \cross (\vb{x}_L \cdot \nabla) \vb{B} + &= \frac{u_L^2}{\omega_c} + \begin{pmatrix} + \cos \\ + - \sin \\ + 0 + \end{pmatrix} + \cross + \begin{pmatrix} + \pdv{B_x}{x} \sin + \pdv{B_x}{y} \cos \\ + \pdv{B_y}{x} \sin + \pdv{B_y}{y} \cos \\ + \pdv{B_z}{x} \sin + \pdv{B_z}{y} \cos + \end{pmatrix} + \\ + &= \frac{u_L^2}{\omega_c} + \begin{pmatrix} + - \pdv{B_z}{x} \sin^2 - \pdv{B_z}{y} \sin \cos \\ + - \pdv{B_z}{x} \sin \cos - \pdv{B_z}{y} \cos^2 \\ + \pdv{B_y}{x} \sin \cos + \pdv{B_y}{y} \cos^2 + + \pdv{B_x}{x} \sin^2 + \pdv{B_x}{y} \sin \cos + \end{pmatrix} +\end{aligned}$$ + +Integrating products of $\sin$ and $\cos$ over their period gives us the following: + +$$\begin{aligned} + \expval{\cos^2} = \expval{\sin^2} = \frac{1}{2} + \qquad \quad + \expval{\sin \cos} = 0 +\end{aligned}$$ + +Inserting this tells us that the average +of $\vb{u}_L \cross (\vb{x}_L \cdot \nabla) \vb{B}$ is given by: + +$$\begin{aligned} + \expval{ \vb{u}_L \cross (\vb{x}_L \cdot \nabla) \vb{B} } + &= \frac{u_L^2}{2 \omega_c} + \begin{pmatrix} + - \pdv{B_z}{x} \\ + - \pdv{B_z}{y} \\ + \pdv{B_y}{y} + \pdv{B_x}{x} + \end{pmatrix} +\end{aligned}$$ + +We use [Maxwell's equation](/know/concept/maxwells-equations/) $\nabla \cdot \vb{B} = 0$ +to rewrite the $z$-component, +and follow the convention that $\vb{B}$ +points mostly in the $z$-direction, +such that $B \equiv |\vb{B}| \approx B_z$: + +$$\begin{aligned} + \expval{ \vb{u}_L \cross (\vb{x}_L \cdot \nabla) \vb{B} } + &= - \frac{u_L^2}{2 \omega_c} + \begin{pmatrix} + \pdv{B_z}{x} \\ + \pdv{B_z}{y} \\ + \pdv{B_z}{z} + \end{pmatrix} + \approx - \frac{u_L^2}{2 \omega_c} + \begin{pmatrix} + \pdv{B}{x} \\ + \pdv{B}{y} \\ + \pdv{B}{z} + \end{pmatrix} + = - \frac{u_L^2}{2 \omega_c} \nabla B +\end{aligned}$$ +</div> +</div> + +With this, the guiding center's equation of motion +is reduced to the following: + +$$\begin{aligned} + m \dv{\vb{u}_{gc}}{t} + = q \bigg( \vb{E} + \vb{u}_{gc} \cross \vb{B} - \frac{u_L^2}{2 \omega_c} \nabla B \bigg) +\end{aligned}$$ + +Let us now split $\vb{u}_{gc}$ into +components $\vb{u}_{gc\perp}$ and $u_{gc\parallel} \vu{b}$, +which are respectively perpendicular and parallel +to the magnetic unit vector $\vu{b}$, +such that $\vb{u}_{gc} = \vb{u}_{gc\perp} \!+\! u_{gc\parallel} \vu{b}$. +Consequently: + +$$\begin{aligned} + \dv{\vb{u}_{gc}}{t} + = \dv{\vb{u}_{gc\perp}}{t} + \dv{u_{gc\parallel}}{t} \vu{b} + u_{gc\parallel} \dv{\vu{b}}{t} +\end{aligned}$$ + +Inserting this into the guiding center's equation of motion, +we now have: + +$$\begin{aligned} + \dv{\vb{u}_{gc}}{t} + = m \bigg( \dv{\vb{u}_{gc\perp}}{t} + \dv{u_{gc\parallel}}{t} \vu{b} + u_{gc\parallel} \dv{\vu{b}}{t} \bigg) + = q \bigg( \vb{E} + \vb{u}_{gc} \cross \vb{B} - \frac{u_L^2}{2 \omega_c} \nabla B \bigg) +\end{aligned}$$ + +The derivative of $\vu{b}$ can be rewritten as follows, +where $R_c$ is the radius of the field's [curvature](/know/concept/curvature/), +and $\vb{R}_c$ is the corresponding vector from the center of curvature: + +$$\begin{aligned} + \dv{\vu{b}}{t} + \approx - u_{gc\parallel} \frac{\vb{R}_c}{R_c^2} +\end{aligned}$$ + +<div class="accordion"> +<input type="checkbox" id="proof-nonuniform-B-curvature"/> +<label for="proof-nonuniform-B-curvature">Proof</label> +<div class="hidden"> +<label for="proof-nonuniform-B-curvature">Proof.</label> +Assuming that $\vu{b}$ does not explicitly depend on time, +i.e. $\pdv*{\vu{b}}{t} = 0$, +we can rewrite the derivative using the chain rule: + +$$\begin{aligned} + \dv{\vu{b}}{t} + = \pdv{\vu{b}}{s} \dv{s}{t} + = u_{gc\parallel} \dv{\vu{b}}{s} +\end{aligned}$$ + +Where $\dd{s}$ is the arc length of the magnetic field line, +which is equal to the radius $R_c$ times the infinitesimal subtended angle $\dd{\theta}$: + +$$\begin{aligned} + \dd{s} + = R_c \dd{\theta} +\end{aligned}$$ + +Meanwhile, across this arc, $\vu{b}$ rotates by $\dd{\theta}$, +such that the tip travels a distance $|\dd{\vu{b}}|$: + +$$\begin{aligned} + |\dd{\vu{b}}\!| + = |\vu{b}| \dd{\theta} + = \dd{\theta} +\end{aligned}$$ + +Furthermore, the direction $\dd{\vu{b}}$ is always opposite to $\vu{R}_c$, +which is defined as the unit vector from the center of curvature to the base of $\vu{b}$: + +$$\begin{aligned} + \dd{\vu{b}} + = - \vu{R}_c \dd{\theta} +\end{aligned}$$ + +Combining these expressions for $\dd{s}$ and $\dd{\vu{b}}$, +we find the following derivative: + +$$\begin{aligned} + \dv{\vu{b}}{s} + = - \frac{\vu{R}_c \dd{\theta}}{R_c \dd{\theta}} + = - \frac{\vu{R}_c}{R_c} + = - \frac{\vb{R}_c}{R_c^2} +\end{aligned}$$ +</div> +</div> + +With this, we arrive at the following equation of motion +for the guiding center: + +$$\begin{aligned} + m \bigg( \dv{\vb{u}_{gc\perp}}{t} + \dv{u_{gc\parallel}}{t} \vu{b} - u_{gc\parallel}^2 \frac{\vb{R}_c}{R_c} \bigg) + = q \bigg( \vb{E} + \vb{u}_{gc} \cross \vb{B} - \frac{u_L^2}{2 \omega_c} \nabla B \bigg) +\end{aligned}$$ + +Since both $\vb{R}_c$ and any cross product with $\vb{B}$ +will always be perpendicular to $\vb{B}$, +we can split this equation into perpendicular and parallel components like so: + +$$\begin{aligned} + m \dv{\vb{u}_{gc\perp}}{t} + &= q \vb{E}_{\perp} - \frac{q u_L^2}{2 \omega_c} \nabla_{\!\perp} B + m u_{gc\parallel}^2 \frac{\vb{R}_c}{R_c} + q \vb{u}_{gc} \cross \vb{B} + \\ + m \dv{u_{gc\parallel}}{t} + &= q E_{\parallel} - \frac{q u_L^2}{2 \omega_c} \nabla_{\!\parallel} B +\end{aligned}$$ + +The parallel part simply describes an acceleration. +The perpendicular part is more interesting: +we rewrite it as follows, defining an effective force $\vb{F}_{\!\perp}$: + +$$\begin{aligned} + m \dv{\vb{u}_{gc\perp}}{t} + = \vb{F}_{\!\perp} + q \vb{u}_{gc} \cross \vb{B} + \qquad \quad + \vb{F}_{\!\perp} + \equiv q \vb{E}_\perp + m u_{gc\parallel}^2 \frac{\vb{R}_c}{R_c} - \frac{q u_L^2}{2 \omega_c} \nabla_{\!\perp} B +\end{aligned}$$ + +To solve this, we make a crude approximation now, and improve it later. +We thus assume that $\vb{u}_{gc\perp}$ is constant in time, +such that the equation reduces to: + +$$\begin{aligned} + 0 + \approx \vb{F}_{\!\perp} + q \vb{u}_{gc} \cross \vb{B} + = \vb{F}_{\!\perp} + q \vb{u}_{gc\perp} \cross \vb{B} +\end{aligned}$$ + +This is analogous to the previous case of a uniform electric field, +with $q \vb{E}$ replaced by $\vb{F}_{\!\perp}$, +so it is also solved by crossing with $\vb{B}$ in front, +yielding a drift: + +$$\begin{aligned} + \vb{u}_{gc\perp} + \approx \vb{v}_F + \equiv \frac{\vb{F}_{\!\perp} \cross \vb{B}}{q B^2} +\end{aligned}$$ + +From the definition of $\vb{F}_{\!\perp}$, +this total $\vb{v}_F$ can be split into three drifts: +the previously seen electric field drift $\vb{v}_E$, +the **curvature drift** $\vb{v}_c$, +and the **grad-$\vb{B}$ drift** $\vb{v}_{\nabla B}$: + +$$\begin{aligned} + \boxed{ + \vb{v}_c + = \frac{m u_{gc\parallel}^2}{q} \frac{\vb{R}_c \cross \vb{B}}{R_c^2 B^2} + } + \qquad \quad + \boxed{ + \vb{v}_{\nabla B} + = \frac{u_L^2}{2 \omega_c} \frac{\vb{B} \cross \nabla B}{B^2} + } +\end{aligned}$$ + +Such that $\vb{v}_F = \vb{v}_E + \vb{v}_c + \vb{v}_{\nabla B}$. +We are still missing a correction, +since we neglected the time dependence of $\vb{u}_{gc\perp}$ earlier. +This correction is called $\vb{v}_p$, +where $\vb{u}_{gc\perp} \approx \vb{v}_F + \vb{v}_p$. +We revisit the perpendicular equation, which now reads: + +$$\begin{aligned} + m \dv{t} \big( \vb{v}_F + \vb{v}_p \big) + = \vb{F}_{\!\perp} + q \big( \vb{v}_F + \vb{v}_p \big) \cross \vb{B} +\end{aligned}$$ + +We assume that $\vb{v}_F$ varies much faster than $\vb{v}_p$, +such that $\dv*{\vb{v}_p}{t}$ is negligible. +In addition, from the derivation of $\vb{v}_F$, +we know that $\vb{F}_{\!\perp} + q \vb{v}_F \cross \vb{B} = 0$, +leaving only: + +$$\begin{aligned} + m \dv{\vb{v}_F}{t} + = q \vb{v}_p \cross \vb{B} +\end{aligned}$$ + +To isolate this for $\vb{v}_p$, +we take the cross product with $\vb{B}$ in front, +like earlier. +We thus arrive at the following correction, +known as the **polarization drift** $\vb{v}_p$: + +$$\begin{aligned} + \boxed{ + \vb{v}_p + = - \frac{m}{q B^2} \dv{\vb{v}_F}{t} \cross \vb{B} + } +\end{aligned}$$ + +In many cases $\vb{v}_E$ dominates $\vb{v}_F$, +so in some literature $\vb{v}_p$ is approximated as follows: + +$$\begin{aligned} + \vb{v}_p + \approx - \frac{m}{q B^2} \dv{\vb{v}_E}{t} \cross \vb{B} + = - \frac{m}{q B^2} \Big( \dv{t} (\vb{E}_\perp \cross \vb{B}) \Big) \cross \vb{B} + = - \frac{m}{q B^2} \dv{\vb{E}_\perp}{t} +\end{aligned}$$ + +The polarization drift stands out from the others: +it has the opposite sign, +it is proportional to $m$, +and it is often only temporary. +Therefore, it is also called the **inertia drift**. + + + +## References +1. F.F. Chen, + *Introduction to plasma physics and controlled fusion*, + 3rd edition, Springer. +2. M. Salewski, A.H. Nielsen, + *Plasma physics: lecture notes*, + 2021, unpublished. diff --git a/content/know/concept/lorentz-force/index.pdc b/content/know/concept/lorentz-force/index.pdc index 2362766..83357d2 100644 --- a/content/know/concept/lorentz-force/index.pdc +++ b/content/know/concept/lorentz-force/index.pdc @@ -5,6 +5,7 @@ publishDate: 2021-09-08 categories: - Physics - Electromagnetism +- Plasma physics date: 2021-09-08T17:00:32+02:00 draft: false @@ -27,6 +28,44 @@ $$\begin{aligned} \end{aligned}$$ +## Uniform electric field + +Consider the simple case of an electric field $\vb{E}$ +that is uniform in all of space. +In the absence of a magnetic field $\vb{B} = 0$ +and any other forces, +Newton's second law states: + +$$\begin{aligned} + \vb{F} + = m \dv{\vb{u}}{t} + = q \vb{E} +\end{aligned}$$ + +This is straightforward to integrate in time, +for a given initial velocity vector $\vb{u}_0$: + +$$\begin{aligned} + \vb{u}(t) + = \frac{q}{m} \vb{E} t + \vb{u}_0 +\end{aligned}$$ + +And then the particle's position $\vb{x}(t)$ +is found be integrating once more, +with $\vb{x}(0) = \vb{x}_0$: + +$$\begin{aligned} + \boxed{ + \vb{x}(t) + = \frac{q}{2 m} \vb{E} t^2 + \vb{u}_0 t + \vb{x}_0 + } +\end{aligned}$$ + +In summary, unsurprisingly, a uniform electric field $\vb{E}$ +accelerates the particle with a constant force $\vb{F} = q \vb{E}$. +Note that the direction depends on the sign of $q$. + + ## Uniform magnetic field Consider the simple case of a uniform magnetic field @@ -64,12 +103,12 @@ $$\begin{aligned} \end{aligned}$$ Where we have defined the **cyclotron frequency** $\omega_c$ as follows, -which is always positive: +which may be negative: $$\begin{aligned} \boxed{ \omega_c - \equiv \frac{|q| B}{m} + \equiv \frac{q B}{m} } \end{aligned}$$ @@ -78,7 +117,7 @@ the solution for $u_x(t)$ is given by: $$\begin{aligned} u_x(t) - = - u_\perp \sin\!(\omega_c t) + = u_\perp \cos\!(\omega_c t) \end{aligned}$$ Where $u_\perp \equiv \sqrt{u_x^2 + u_y^2}$ is the constant total transverse velocity. @@ -87,13 +126,12 @@ Then $u_y(t)$ is found to be: $$\begin{aligned} u_y(t) = \frac{m}{q B} \dv{u_x}{t} - = - \frac{m \omega_c}{q B} u_\perp \cos\!(\omega_c t) - = - \mathrm{sgn}(q) \: u_\perp \cos\!(\omega_c t) + = - \frac{m \omega_c}{q B} u_\perp \sin\!(\omega_c t) + = - u_\perp \sin\!(\omega_c t) \end{aligned}$$ -Where $\mathrm{sgn}$ is the signum function. -This tells us that the particle moves in a circular orbit, -and that the direction of rotation is determined by $q$. +This means that the particle moves in a circle, +in a direction determined by the sign of $\omega_c$. Integrating the velocity yields the position, where we refer to the integration constants $x_{gc}$ and $y_{gc}$ @@ -101,10 +139,10 @@ as the **guiding center**, around which the particle orbits or **gyrates**: $$\begin{aligned} x(t) - = \frac{u_\perp}{\omega_c} \cos\!(\omega_c t) + x_{gc} + = \frac{u_\perp}{\omega_c} \sin\!(\omega_c t) + x_{gc} \qquad \quad y(t) - = - \mathrm{sgn}(q) \: \frac{u_\perp}{\omega_c} \sin\!(\omega_c t) + y_{gc} + = \frac{u_\perp}{\omega_c} \cos\!(\omega_c t) + y_{gc} \end{aligned}$$ The radius of this orbit is known as the **Larmor radius** or **gyroradius** $r_L$, given by: @@ -112,16 +150,18 @@ The radius of this orbit is known as the **Larmor radius** or **gyroradius** $r_ $$\begin{aligned} \boxed{ r_L - \equiv \frac{u_\perp}{\omega_c} + \equiv \frac{u_\perp}{|\omega_c|} = \frac{m u_\perp}{|q| B} } \end{aligned}$$ -Finally, it is trivial to integrate the equation for the $z$-direction velocity $u_z$: +Finally, it is easy to integrate the equation +for the $z$-axis velocity $u_z$, which is conserved: $$\begin{aligned} z(t) - = u_z t + z_{gc} + = z_{gc} + = u_z t + z_0 \end{aligned}$$ In conclusion, the particle's motion parallel to $\vb{B}$ @@ -129,116 +169,24 @@ is not affected by the magnetic field, while its motion perpendicular to $\vb{B}$ is circular around an imaginary guiding center. The end result is that particles follow a helical path -when moving through a uniform magnetic field. - - -## Uniform electric and magnetic field - -Let us now consider a more general case, -with constant uniform electric and magnetic fields $\vb{E}$ and $\vb{B}$, -which may or may not be perpendicular. -The equation of motion is then: - -$$\begin{aligned} - \vb{F} - = m \dv{\vb{u}}{t} - = q (\vb{E} + \vb{u} \cross \vb{B}) -\end{aligned}$$ - -If we take the dot product with the unit vector $\vu{B}$, -the cross product vanishes, leaving: +when moving through a uniform magnetic field: $$\begin{aligned} - \dv{\vb{u}_\parallel}{t} - = \frac{q}{m} \vb{E}_\parallel -\end{aligned}$$ - -Where $\vb{u}_\parallel$ and $\vb{E}_\parallel$ are -the components of $\vb{u}$ and $\vb{E}$ -that are parallel to $\vb{B}$. -This equation is easy to integrate: -the guiding center accelerates according to $(q/m) \vb{E}_\parallel$. - -Next, let us define the perpendicular component $\vb{u}_\perp$ -such that $\vb{u} = \vb{u}_\parallel \vu{B} + \vb{u}_\perp$. -Its equation of motion is found by -subtracting $\vb{u}_\parallel$'s equation from the original: - -$$\begin{aligned} - m \dv{\vb{u}_\perp}{t} - = q (\vb{E} + \vb{u} \cross \vb{B}) - q \vb{E}_\parallel - = q (\vb{E}_\perp + \vb{u}_\perp \cross \vb{B}) -\end{aligned}$$ - -To solve this, we go to a moving coordinate system -by defining $\vb{u}_\perp = \vb{v}_\perp + \vb{w}_\perp$, -where $\vb{v}_\perp$ is a constant of our choice. -The equation is now as follows: - -$$\begin{aligned} - m \dv{t} (\vb{v}_\perp + \vb{w}_\perp) - = m \dv{\vb{w}_\perp}{t} - = q (\vb{E}_\perp + \vb{v}_\perp \cross \vb{B} + \vb{w}_\perp \cross \vb{B}) -\end{aligned}$$ - -We want to choose $\vb{v}_\perp$ such that the first two terms vanish, -or in other words: - -$$\begin{aligned} - 0 - = \vb{E}_\perp + \vb{v}_\perp \cross \vb{B} -\end{aligned}$$ - -To find $\vb{v}_\perp$, we take the cross product with $\vb{B}$, -and use the fact that $\vb{B} \cross \vb{E}_\perp = \vb{B} \cross \vb{E}$: - -$$\begin{aligned} - 0 - = \vb{B} \cross (\vb{E}_\perp + \vb{v}_\perp \cross \vb{B}) - = \vb{B} \cross \vb{E} + \vb{v}_\perp B^2 - \quad \implies \quad \boxed{ - \vb{v}_\perp - = \frac{\vb{E} \cross \vb{B}}{B^2} + \vb{x}(t) + = \frac{u_\perp}{\omega_c} + \begin{pmatrix} + \sin\!(\omega_c t) \\ \cos\!(\omega_c t) \\ 0 + \end{pmatrix} + + \vb{x}_{gc}(t) } \end{aligned}$$ -When $\vb{v}_\perp$ is chosen like this, -the perpendicular equation of motion is reduced to: - -$$\begin{aligned} - m \dv{\vb{w}_\perp}{t} - = q \vb{w}_\perp \cross \vb{B} -\end{aligned}$$ - -Which is simply the case we treated previously with $\vb{E} = 0$, -with a known solution -(assuming $\vb{B}$ still points in the positive $z$-direction): - -$$\begin{aligned} - w_x(t) - = - w_\perp \sin\!(\omega_c t) - \qquad - w_y(t) - = - \mathrm{sgn}(q) \: w_\perp \cos\!(\omega_c t) -\end{aligned}$$ - -However, this result is shifted by a constant $\vb{v}_\perp$, -often called the **drift velocity** $\vb{v}_d$, -at which the guiding center moves transversely. -Curiously, $\vb{v}_d$ is independent of $q$. - -Such a drift is not specific to an electric field. -In the equations above, $\vb{E}$ can be replaced -by a general force $\vb{F}/q$ (e.g. gravity) without issues. -In that case, $\vb{v}_d$ does depend on $q$: - -$$\begin{aligned} - \boxed{ - \vb{v}_d - = \frac{\vb{F} \cross \vb{B}}{q B^2} - } -\end{aligned}$$ +Where $\vb{x}_{gc}(t) \equiv (x_{gc}, y_{gc}, z_{gc})$ +is the position of the guiding center. +For a detailed look at how $\vb{B}$ and $\vb{E}$ +can affect the guiding center's motion, +see [guiding center theory](/know/concept/guiding-center-theory/). diff --git a/content/know/concept/rabi-oscillation/index.pdc b/content/know/concept/rabi-oscillation/index.pdc new file mode 100644 index 0000000..cf393a4 --- /dev/null +++ b/content/know/concept/rabi-oscillation/index.pdc @@ -0,0 +1,218 @@ +--- +title: "Rabi oscillation" +firstLetter: "R" +publishDate: 2021-09-22 +categories: +- Physics +- Quantum mechanics +- Optics + +date: 2021-09-18T00:41:43+02:00 +draft: false +markup: pandoc +--- + +# Rabi oscillation + +In quantum mechanics, from the derivation of +[time-dependent perturbation theory](/know/concept/time-dependent-perturbation-theory/), +we know that a time-dependent term $\hat{H}_1$ in the Hamiltonian +affects the state as follows, +where $c_n(t)$ are the coefficients of the linear combination +of basis states $\ket{n} \exp\!(-i E_n t / \hbar)$: + +$$\begin{aligned} + i \hbar \dv{c_m}{t} + = \sum_{n} c_n(t) \matrixel{m}{\hat{H}_1}{n} \exp\!(i \omega_{mn} t) +\end{aligned}$$ + +Where $\omega_{mn} \equiv (E_m \!-\! E_n) / \hbar$ +for energies $E_m$ and $E_n$. +Note that this equation is exact, +despite being used for deriving perturbation theory. +Consider a two-level system where $n \in \{a, b\}$, +in which case the above equation can be expanded to the following: + +$$\begin{aligned} + \dv{c_a}{t} + &= - \frac{i}{\hbar} \matrixel{a}{\hat{H}_1}{b} \exp\!(- i \omega_0 t) \: c_b - \frac{i}{\hbar} \matrixel{a}{\hat{H}_1}{a} \: c_a + \\ + \dv{c_b}{t} + &= - \frac{i}{\hbar} \matrixel{b}{\hat{H}_1}{a} \exp\!(i \omega_0 t) \: c_a - \frac{i}{\hbar} \matrixel{b}{\hat{H}_1}{b} \: c_b +\end{aligned}$$ + +Where $\omega_0 \equiv \omega_{ba}$ is positive. +We assume that $\hat{H}_1$ has odd spatial parity, +in which case [Laporte's selection rule](/know/concept/selection-rules/) +states that the diagonal matrix elements vanish, leaving: + +$$\begin{aligned} + \dv{c_a}{t} + &= - \frac{i}{\hbar} \matrixel{a}{\hat{H}_1}{b} \exp\!(- i \omega_0 t) \: c_b + \\ + \dv{c_b}{t} + &= - \frac{i}{\hbar} \matrixel{b}{\hat{H}_1}{a} \exp\!(i \omega_0 t) \: c_a +\end{aligned}$$ + +We now choose $\hat{H}_1$ to be as follows, +sinusoidally oscillating with a spatially odd $V(\vec{r})$: + +$$\begin{aligned} + \hat{H}_1(t) + = V \cos\!(\omega t) + = \frac{V}{2} \Big( \exp\!(i \omega t) + \exp\!(-i \omega t) \Big) +\end{aligned}$$ + +We insert this into the equations for $c_a$ and $c_b$, +and define $V_{ab} \equiv \matrixel{a}{V}{b}$, leading us to: + +$$\begin{aligned} + \dv{c_a}{t} + &= - i \frac{V_{ab}}{2 \hbar} \Big( \exp\!\big(i (\omega \!-\! \omega_0) t\big) + \exp\!\big(\!-\! i (\omega \!+\! \omega_0) t\big) \Big) \: c_b + \\ + \dv{c_b}{t} + &= - i \frac{V_{ab}}{2 \hbar} \Big( \exp\!\big(i (\omega \!+\! \omega_0) t\big) + \exp\!\big(\!-\! i (\omega \!-\! \omega_0) t\big) \Big) \: c_a +\end{aligned}$$ + +Here, we make the *rotating wave approximation*: +assuming we are close to resonance $\omega \approx \omega_0$, +we decide that $\exp\!(i (\omega \!+\! \omega_0) t)$ +oscillates so much faster than $\exp\!(i (\omega \!-\! \omega_0) t)$, +that its effect turns out negligible +when the system is observed over a reasonable time interval. + +In other words, over this reasonably-sized time interval, +$\exp\!(i (\omega \!+\! \omega_0) t)$ averages to zero, +while $\exp\!(i (\omega \!-\! \omega_0) t)$ does not. +Dropping the respective terms thus leaves us with: + +$$\begin{aligned} + \dv{c_a}{t} + = - i \frac{V_{ab}}{2 \hbar} \exp\!\big(i (\omega \!-\! \omega_0) t \big) \: c_b + \qquad \quad + \dv{c_b}{t} + = - i \frac{V_{ba}}{2 \hbar} \exp\!\big(\!-\! i (\omega \!-\! \omega_0) t \big) \: c_a +\end{aligned}$$ + +Now we can solve this system of coupled equations exactly. +We differentiate the first equation with respect to $t$, +and then substitute $\dv*{c_b}{t}$ for the second equation: + +$$\begin{aligned} + \dv[2]{c_a}{t} + &= - i \frac{V_{ab}}{2 \hbar} \bigg( i (\omega - \omega_0) \: c_b + \dv{c_b}{t} \bigg) \exp\!\big(i (\omega \!-\! \omega_0) t \big) + \\ + &= - i \frac{V_{ab}}{2 \hbar} \bigg( i (\omega - \omega_0) \: c_b + - i \frac{V_{ba}}{2 \hbar} \exp\!\big(\!-\! i (\omega \!-\! \omega_0) t \big) \: c_a \bigg) + \exp\!\big(i (\omega \!-\! \omega_0) t \big) + \\ + &= \frac{V_{ab}}{2 \hbar} (\omega - \omega_0) \exp\!\big(i (\omega \!-\! \omega_0) t \big) \: c_b - \frac{|V_{ab}|^2}{(2 \hbar)^2} c_a +\end{aligned}$$ + +In the first term, we recognize $\dv*{c_a}{t}$, +which we insert to arrive at an equation for $c_a(t)$: + +$$\begin{aligned} + 0 + = \dv[2]{c_a}{t} - i (\omega - \omega_0) \dv{c_a}{t} + \frac{|V_{ab}|^2}{(2 \hbar)^2} \: c_a +\end{aligned}$$ + +To solve this, we make the ansatz $c_a(t) = \exp\!(\lambda t)$, +which, upon insertion, gives us: + +$$\begin{aligned} + 0 + = \lambda^2 - i (\omega - \omega_0) \lambda + \frac{|V_{ab}|^2}{(2 \hbar)^2} +\end{aligned}$$ + +This quadratic equation has two complex roots $\lambda_1$ and $\lambda_2$, +which are found to be: + +$$\begin{aligned} + \lambda_1 + = i \frac{\omega - \omega_0 + \tilde{\Omega}}{2} + \qquad \quad + \lambda_2 + = i \frac{\omega - \omega_0 - \tilde{\Omega}}{2} +\end{aligned}$$ + +Where we have defined the **generalized Rabi frequency** $\tilde{\Omega}$ to be given by: + +$$\begin{aligned} + \boxed{ + \tilde{\Omega} + \equiv \sqrt{(\omega - \omega_0)^2 + \frac{|V_{ab}|^2}{\hbar^2}} + } +\end{aligned}$$ + +So that the general solution $c_a(t)$ is as follows, +where $A$ and $B$ are arbitrary constants, +to be determined from initial conditions (and normalization): + +$$\begin{aligned} + \boxed{ + c_a(t) + = \Big( A \sin\!(\tilde{\Omega} t / 2) + B \cos\!(\tilde{\Omega} t / 2) \Big) \exp\!\big(i (\omega \!-\! \omega_0) t / 2 \big) + } +\end{aligned}$$ + +And then the corresponding $c_b(t)$ can be found +from the coupled equation we started at, +or, if we only care about the probability density $|c_a|^2$, +we can use $|c_b|^2 = 1 - |c_a|^2$. +For example, if $A = 0$ and $B = 1$, +we get the following probabilities + +$$\begin{aligned} + |c_a(t)|^2 + &= \cos^2(\tilde{\Omega} t / 2) + = \frac{1}{2} \Big( 1 + \cos\!(\tilde{\Omega} t) \Big) + \\ + |c_b(t)|^2 + &= \sin^2(\tilde{\Omega} t / 2) + = \frac{1}{2} \Big( 1 - \cos\!(\tilde{\Omega} t) \Big) +\end{aligned}$$ + +Note that the period was halved by squaring. +This periodic "flopping" of the particle between $\ket{a}$ and $\ket{b}$ +is known as **Rabi oscillation**, **Rabi flopping** or the **Rabi cycle**. +This is a more accurate treatment +of the flopping found from first-order perturbation theory. + +The name **generalized Rabi frequency** suggests +that there is a non-general version. +Indeed, the **Rabi frequency** $\Omega$ is based on +the special case of exact resonance $\omega = \omega_0$: + +$$\begin{aligned} + \Omega + \equiv \frac{V_{ab}}{\hbar} +\end{aligned}$$ + +As an example, Rabi oscillation arises +in the [electric dipole approximation](/know/concept/electric-dipole-approximation/), +where $\hat{H}_1$ is: + +$$\begin{aligned} + \hat{H}_1(t) + = - q \vec{r} \cdot \vec{E}_0 \cos(\omega t) +\end{aligned}$$ + +After making the rotating wave approximation, +the resulting Rabi frequency is given by: + +$$\begin{aligned} + \Omega + = \frac{\vec{d} \cdot \vec{E}_0}{\hbar} +\end{aligned}$$ + +Where $\vec{E}_0$ is the [electric field](/know/concept/electric-field/) amplitude, +and $\vec{d} \equiv q \matrixel{a}{\vec{r}}{b}$ is the transition dipole moment +of the electron between orbitals $\ket{a}$ and $\ket{b}$. + + + +## References +1. D.J. Griffiths, D.F. Schroeter, + *Introduction to quantum mechanics*, 3rd edition, + Cambridge. diff --git a/content/know/concept/time-dependent-perturbation-theory/index.pdc b/content/know/concept/time-dependent-perturbation-theory/index.pdc index b16e3ee..1fbd9ce 100644 --- a/content/know/concept/time-dependent-perturbation-theory/index.pdc +++ b/content/know/concept/time-dependent-perturbation-theory/index.pdc @@ -181,7 +181,10 @@ which can cause confusion when comparing literature. In any case, the probability oscillates as a function of $t$ with period $ |