NowhereLog

Why is Sky Blue?

April 10, 2024

Why is sky blue? You may know the answer is Rayleigh scattering. But how do we understand Rayleigh scattering? We will try to find the answers in Maxwell’s equations and a simple dipole model.

Recap on E&M

All easy questions can be extremely hard if you keep asking the questions. “Why is sky blue”? “Because blue light, which has larger frequency, scatters more.” One might answer. “But what causes the scattering and why is there a frequency dependence?” For this post, we will find the answer from Maxwell’s equations.

E=ρϵ0    B=0\vec{\nabla} \cdot \vec{E} = \frac{\rho}{\epsilon_0} \ \ \ \ \vec{\nabla} \cdot \vec{B} = 0 ×E=Bt    ×B=μ0J+μ0ϵ0Et\vec{\nabla} \times \vec{E} = -\frac{\partial \vec{B}}{\partial t} \ \ \ \ \vec{\nabla} \times \vec{B} = \mu_0 \vec{J} + \mu_0 \epsilon_0 \frac{\partial \vec{E}}{\partial t}

Witht the right Gauge transformation (Lorentz Gauge), we have the clean potential equations

2ϕ=1ϵ0ρ    2A=μ0J\square^2 \phi = -\frac{1}{\epsilon_0} \rho \ \ \ \ \square^2 \mathbf{A} = -\mu_0 \mathbf{J}

where 2=21c22t2\square^2 = \nabla^2 - \frac{1}{c^2} \frac{\partial^2}{\partial t^2}.

If the potential is not time-dependent, we recover the Poisson equations, for which the solutions are

ϕ(r)=14πϵ0ρ(r)rrd3r\phi(\mathbf{r}) = \frac{1}{4 \pi \epsilon_0} \int \frac{\rho(\mathbf{r}')}{\|\mathbf{r} - \mathbf{r}'\|} \, d^3 \mathbf{r}' A(r)=μ014πJ(r)rrd3r\mathbf{A}(\mathbf{r}) = \mu_0 \frac{1}{4 \pi} \int \frac{\mathbf{J}(\mathbf{r}')}{\|\mathbf{r} - \mathbf{r}'\|} \, d^3 \mathbf{r}'

where ρ\rho is the charge density, J\mathbf{J} is the current density.

For the full time-dependent equations, the solutions look very much like the above, but the charge and current density are evaluated at the “retarded time” trt_r.

ϕ(r,t)=14πϵ0ρ(r,tr)rrd3r\phi(\mathbf{r}, t) = \frac{1}{4 \pi \epsilon_0} \int \frac{\rho(\mathbf{r}', t_r)}{\|\mathbf{r} - \mathbf{r}'\|} \, d^3 \mathbf{r}' A(r,t)=μ014πJ(r,tr)rrd3r\mathbf{A}(\mathbf{r}, t) = \mu_0 \frac{1}{4 \pi} \int \frac{\mathbf{J}(\mathbf{r}', t_r)}{\|\mathbf{r} - \mathbf{r}'\|} \, d^3 \mathbf{r}'

where tr=trrct_r = t - \frac{\|\mathbf{r} - \mathbf{r}'\|}{c}. This manifests the causality: the information that the densities changes cannot propagage faster than the speed of light.

For fun but not really related to our main discussion, we will digress and derive the potentials for a point charge qq with a trajectory w(t)\mathbf{w}(t). The potential equation reduces to

ϕ(r,t)=q4πϵ01ss^v(tr)/c\phi(\mathbf{r}, t) = \frac{q}{4 \pi \epsilon_0} \frac{1}{\mathbf{s} - \hat{\mathbf{s}} \cdot \mathbf{v}(t_r)/c}

where s=rw(tr)\mathbf{s} = \mathbf{r} - \mathbf{w}(t_r).

The potential looks very similar to a static charge,

ϕ(r)=q4πϵ01s\phi(\mathbf{r}) = \frac{q}{4 \pi \epsilon_0} \frac{1}{\mathbf{s}}

But again, due to causality, or the relativistic effects, we see the manifest of length contraction with the 1s^v(tr)/c1-\hat{\mathbf{s}} \cdot \mathbf{v}(t_r)/c factor. The retarded time is implicitly defined by

c(ttr)=rw(tr)c(t-t_r) = \|\mathbf{r}-\mathbf{w}(t_r)\|

Similarly we have that

A(r,t)=vc2ϕ(r,t)\mathbf{A}(\mathbf{r}, t) = \frac{\mathbf{v}}{c^2} \phi(\mathbf{r}, t)

Radiation

Now we are finally in good shape to talk about radiation. In essence, radiation is the emission of energy by moving charges. Moving charges cause a change in the fields, which causes EM waves, which propagates energy. In the case of the atmosphere, the interaction between sunlight (radiation from the nuclear reactions of the sun) and the charges in the atmosphere. As a simplified model (which turns out to succesffully explain the color of the sky), we will assumes the charges oscillate like a electric dipole, i.e., two opposing charges q(t)q(t) and q(t)-q(t) are held dd apart, where q(t)=q0cos(ωt)q(t) = q_0 \cos(\omega t). This may seem a bit unphysical, and an equivalent picture of the system is when the two opposing charges q0q_0 and q0-q_0 are connected by a spring which oscillates with frequency ω\omega, but this will make calculations a bit more complicated.

ϕ(r,t)=14πϵ0(q0cos(ω(ts+/c))s+q0cos(ω(ts/c))s)\phi(\mathbf{r}, t) = \frac{1}{4\pi \epsilon_0} \left(\frac{q_0 \cos(\omega (t - s_+/c))}{s_+} - \frac{q_0 \cos(\omega (t - s_-/c))}{s_-}\right)

where s±=r2rdcosθ+(d2)2s_{\pm} = \sqrt{r^2 \mp r d \cos \theta + \left(\frac{d}{2}\right)^2}.

dipole

Our next step is to make some physical approximations. To start with, we will assume that the separation between the charges are very small. How small? (1) d<<rd<<r, i.e. we are only worried about the field sufficient far away; (2) d<<cωd << \frac{c}{\omega} or d<<λ=2πcωd << \lambda = \frac{2 \pi c}{\omega}, i.e. the separation is small compared to the wavelength of the oscillation. Thus, the potential can be simplified into

ϕ(r,θ,t)=q0dcosθ4πϵ0r[ωcsin(ω(trc))+1rcos(ω(trc))]\phi(r, \theta, t) = \frac{q_0 d \cos \theta}{4\pi\epsilon_0 r} \left[ -\frac{\omega}{c} \sin\left(\omega(t - \frac{r}{c})\right) + \frac{1}{r} \cos\left(\omega(t - \frac{r}{c})\right) \right]

Now at the limit of ω0\omega \rightarrow 0, the second term reduces to the potential of stationary charge. However, the second term will not survive due to the 1/r1/r dependence, and thus is not of the interest for radiation. We are interested in the range of (3) r>>cωr >> \frac{c}{\omega}, in which case potential reduces to

ϕ(r,θ,t)=q0d4πϵ0cωcosθrsin(ω(trc))\phi(r, \theta, t) = - \frac{q_0 d}{4\pi\epsilon_0 c} \omega \frac{ \cos \theta}{r} \sin\left(\omega(t - \frac{r}{c})\right)

We can similarly calculate the vector potential A(r,θ,t)\mathbf A(r, \theta, t), and with Maxwell equations and the approximation at large rr, we can find the EM fields:

E(r,θ,t)=q0d4πϵ0cω2sinθrcos(ω(trc))θ^\mathbf{E}(r, \theta, t) = - \frac{q_0 d}{4\pi\epsilon_0 c} \omega^2 \frac{ \sin \theta}{r} \cos\left(\omega(t - \frac{r}{c})\right) \hat{\theta} B(r,θ,t)=q0d4πϵ0c2ω2sinθrcos(ω(trc))ϕ^\mathbf{B}(r, \theta, t) = - \frac{q_0 d}{4\pi\epsilon_0 c^2} \omega^2 \frac{ \sin \theta}{r} \cos\left(\omega(t - \frac{r}{c})\right) \hat{\phi}

The energy radiated by an oscillating electric dipole is determined by the Poynting vector:

S(r,θ,t)=1μ0(E×B)=μ0c[q0d4πω2sinθrcos(ω(trc))]2r^\mathbf{S}(r, \theta, t) = \frac{1}{\mu_0}(\mathbf{E} \times \mathbf{B}) = \frac{\mu_0}{c} \left[ \frac{q_0 d}{4\pi} \omega^2 \frac{ \sin \theta}{r} \cos\left(\omega(t - \frac{r}{c})\right) \right]^2 \hat{\mathbf r}

The intensity of the emitted light is obtained by averaging (in time) over a complete cycle:

S=(μ0q02d232πc2)ω4sin2θr2r^\langle\mathbf{S}\rangle = \left( \frac{\mu_0 q_0^2 d^2}{32\pi c^2}\right)\omega^4 \frac{\sin^2 \theta}{r^2} \hat{\mathbf r}

Now it’s easy to see that the intensity depends on the frequency ω\omega at its 4th power. Thus, the higher the frequency, the larger the intensity. And Voilà! Among the colors most sensitive to our eyes (blue, red, green), blue light has the largest frequency.

It’s curious to note also that the intensity has a θ\theta dependence. In particular, the intensity is 00 when θ=0\theta = 0. Since EM waves are transverse, the dipoles oscillate in a plane orthogonal to the sun’s rays. This means that when the eye of sight is perpendicular to the sun’s rays, θ=0\theta=0 and we should not see any radiation! But this is of course not true based on our observation. Our model accounts for a polarized EM wave which induces a particular dipole direction. However, sun’s rays are unpolarized and consist of a mixture of polarization directions. In effect, we can reduce any direction to a linear combination of two axis-aligned directions, and each polarized ray produces intensity as calculated which looks like a donut (top two plots below). And for an unpolarized light, the distribution can be obtained intuitively by rotating a polarized distribution around the axis of propogation of light (bottom plot below), results in a (1+cos2θ)(1+cos^2 \theta) dependency instead, which is correctly characterized by Rayleigh scattering.

rayleigh

Now you can always ask further where does Maxwell’s equations come from, but unfortunately the discussion will end here today.

References

[1] Zajc, William. PHYS GU4021-4022 Quantum Mechanics I-II, Department of Physics, Columbia University. 2019-2020.

[2] Griffiths, David. Introduction to Electrodynamics. 4th Edition.

[3] Petty, Grant W.. A First Course in Atmospheric Radiation. 2nd Edition.


By NowhereMan who goes nowhere.


© 2024, NowhereLog by NowhereMan