Radiometry

Radiometry

Stars, galaxies, planets, and even you—everything in the universe emits some sort of radiation. This radiation can be in the form of electromagnetic radiation, such as light, or particles, such as electrons and protons. The study of this radiation is called radiometry. Here, we will be studying electromagnetic radiation. Electromagnetic radiation is a wave in the electric and magnetic fields, which propagate through space, carrying energy and momentum. These waves can be characterized by a wavelength and frequency. Radiation can be classified on the basis of its wavelength or frequency:

  • Radio waves: Used by your radio and TV sets, and are the longest wavelength radiation.
  • Microwaves: Used in microwave ovens and radar systems.
  • Infrared: Used in remote controls and thermal imaging.
  • Visible light: The light which we can see with our eyes.
  • Ultraviolet: Causes sunburns.
  • X-rays: Allow X-ray imaging.
  • Gamma rays: The most energetic forms of radiation.

A surface which emits radiation isotropically (symmetrically in all directions) is called a Lambertian surface.

Flux

What is Flux?

Flux describes any effect that appears to pass or travel through a surface or substance. Here, the flux Φ\Phi is defined as the radiant power (power associated with the radiation) going through some surface, and is measured in units of W\mathrm{W}.

The term flux density FF (flux and flux density are often used interchangeably without much thought, even though they’re two different things) is the amount of radiant energy (energy carried by the radiation) passing through a unit area per unit time. The units of flux density are Wm2\mathrm{W \, m^{-2}}. Flux density is also called irradiance.

The flux emitted by a spherically symmetric body, such as a star (power emitted by the star) into a solid angle ω\omega is Φ=ωr2F\Phi = \omega r^2 F, where FF is the flux density at a distance rr. The total flux passing through a closed surface surrounding the source is known as the luminosity LL. Here, L=4πr2FL = 4\pi r^2 F.

The luminosity is independent of distance; this means that the flux density is inversely proportional to the square of the distance from the source.

In astrophysics, we often have to deal with how much energy is being received by our telescopes from a star or any other astronomical object. This is where flux and flux density become important. For example, the flux density of the Sun’s radiation on Earth is called the solar constant, which is about 1365Wm21365 \mathrm{\, W \, m^{-2}}.

📍

If a surface of area AA receives flux Φ\Phi from radiation, the flux density FF at a point is given by

F=ΦA(2.1.1)\tag{2.1.1} F = \frac{\partial \Phi}{\partial A}

If the surface is illuminated uniformly (flux at each point is the same), then

F=ΦAF = \frac{\Phi}{A}

If the cross-sectional area of the radiation beam (projected surface area) is AprojA_\text{proj}, and the radiation is hitting the surface at an angle θ\theta to the normal, then we have the relation Aproj=AcosθA_\text{proj} = A \cos \theta. Therefore, the flux density due to radiation becomes

F=ΦAprojcosθ(2.1.2)\tag{2.1.2} F = \frac{\Phi}{A_\text{proj}} \cos \theta

The factor cosθ\cos \theta accounts for the fact that the radiation is not perpendicular to the surface. For a given beam, the flux density is maximum when θ=0\theta = 0, i.e., when the radiation is perpendicular to the surface.

The spectral flux density FνF_\nu is the flux density of the radiation in the frequency range [ν\nu, ν+dν\nu + d\nu]. This tells us how much power carried by radiation is passing through unit area per unit frequency. In radio astronomy, spectral flux densities are often measured in Janskys (Jy), where 1Jy=1026Wm2Hz11 \, \mathrm{Jy} = 10^{-26} \mathrm{\, W \, m^{-2} \, Hz^{-1}}.

Now we will define some more useful quantities related to the flow of radiation.

Radiance

Radiance is the radiant flux passing through a given surface, per unit solid angle per unit projected area. Radiance is also called brightness, since this is the quantity that determines how bright an object will appear, independent of its shape, size, and other physical properties. It is useful because it indicates how much of the power emitted, reflected, transmitted, or received by a surface will be received by an optical system looking at that surface from a specified angle of view.

📍

Assume we have some radiation passing through a surface element dAdA. Some of the radiation will leave dAdA within a solid angle dωd \omega; the angle between dωd \omega (the direction of outgoing radiation) and the normal to the surface is denoted by θ\theta. If the spectral flux or power of radiation with frequency in the range [ν\nu, ν+dν\nu + d\nu] passing through the area is Φν\Phi_\nu, then the spectral radiance is defined as

Bν=2ΦνωAproj(2.1.3)\tag{2.1.3} B_\nu = \frac{\partial^2 \Phi_\nu}{\partial \omega \, \partial A_\text{proj}}

Here the quantity BνB_\nu is called the spectral radiance of the radiation at frequency ν\nu and direction θ\theta. Aproj=AcosθA_\text{proj} = A \cos \theta is the projected area on which the radiation is incident. The total radiance is given by

B=0Bνdν(2.1.4)\tag{2.1.4} B = \int_0^{\infty} B_\nu \, d\nu
The radiance B is related to the energy passing through a surface element dA into a solid angle dω, in a direction θ

The flux density is related to the spectral radiance as

Fν=ΩBνcosθdω(2.1.5)\tag{2.1.5} F_\nu = \int_\Omega B_\nu \cos \theta \, d\omega

F=ΩBcosθdω(2.1.6)\tag{2.1.6} F = \int_\Omega B \cos \theta \, d\omega

The integral is carried over the solid angle Ω\Omega through which the flux density is being calculated.

For isotropic radiation, the flux density leaving a spherical surface of radius rr is given by

F=ΩBcosθdω=0π/202πBcosθsinθdθdϕF = \int_\Omega B \cos \theta \, d\omega = \int_0^{\pi/2} \int_0^{2\pi} B \cos \theta \sin \theta \, d\theta \, d\phi

    F=πB(2.1.7)\tag{2.1.7} \implies \boxed{F = \pi B}

Surface Brightness

For extended objects, surface brightness BB is defined as the flux density per unit solid angle. Here, the observer views an extended source and is located at the vertex of the solid angle subtended by the object.

B=Fω(2.1.8)\tag{2.1.8} B = \frac{F}{\omega}

Surface brightness is independent of distance. It is equal to the radiance of the object if the object emits isotropically.

The surface brightness B is the flux received by the observer, divided by the solid angle the object is spread over as seen by the observer

A List of Radiometric Quantities

Name Symbol Unit Description
Radiant Energy QQ J\mathrm{J} Energy of electromagnetic radiation
Radiant Flux or Luminosity LL W\mathrm{W} Radiant energy emitted, reflected, transmitted or received, per unit time
Spectral Flux or Spectral Luminosity LνL_\nu WHz1\mathrm{W \, Hz^{-1}} Radiant flux per unit frequency or wavelength
Radiant Intensity II Wsr1\mathrm{W \, sr^{-1}} Radiant flux emitted, reflected, transmitted or received, per unit solid angle
Spectral Intensity IνI_\nu WHz1sr1\mathrm{W \, Hz^{-1} \, sr^{-1}} Radiant intensity per unit frequency or wavelength
Radiance BB Wm2sr1\mathrm{W \, m^{-2} \, sr^{-1}} Radiant flux emitted, reflected, transmitted or received by a surface, per unit solid angle per unit projected area. This is sometimes also confusingly called “intensity”
Spectral Radiance BνB_\nu Wm2sr1Hz1\mathrm{W \, m^{-2} \, sr^{-1} \, Hz^{-1}} Radiance of a surface per unit frequency or wavelength. This is sometimes also confusingly called “spectral intensity”
Irradiance or Flux density FF Wm2\mathrm{W \, m^{-2}} Radiant flux received by a surface per unit area
Spectral Irradiance or Spectral Flux density FνF_\nu Wm2Hz1\mathrm{W \, m^{-2} \, Hz^{-1}} Irradiance of a surface per unit frequency or wavelength

Properties of Radiation

The energy of radiation is quantized in packets called photons. The energy of a photon is given by

E=hν=hcλ(2.1.9)\tag{2.1.9} E = h \nu = \frac{h c}{\lambda}

where ν\nu is the frequency of the radiation. The momentum carried by a photon is given by

p=Ec=hνc=hλ(2.1.10)\tag{2.1.10} p = \frac{E}{c} = \frac{h \nu}{c} = \frac{h}{\lambda}

The radiation pressure (measured in units of Pa\mathrm{Pa}) is the force per unit area exerted by radiation on a surface. It is given by

P=Fccos2θ(2.1.11)\tag{2.1.11} \boxed{P = \frac{F}{c} \cos^2 \theta}

where FF is the flux density of the radiation falling on the surface, and θ\theta is the angle between the direction of the radiation and the normal to the surface. The factor cos2θ\cos^2 \theta accounts for the fact that the radiation is not perpendicular to the surface. The radiation pressure is maximum when θ=0\theta = 0, i.e., when the radiation is perpendicular to the surface. This shows that radiation can exert measurable forces, especially significant in stellar interiors or solar sails.

The energy density uu of the radiation source (measured in units of Jm3\mathrm{J \, m^{-3}}) is the amount of energy stored in it per unit volume. For an isotropic source having radiance BB,

u=4πcB(2.1.12)\tag{2.1.12} \boxed{u = \frac{4 \pi}{c} B}

For an isotropic source with energy density uu, its internal pressure PP is

P=13u=4π3cB(2.1.13)\tag{2.1.13} P = \frac{1}{3} u = \frac{4\pi}{3c} B
What is the pressure of radiation from a flux density FF falling on a surface which perfectly absorbs all the radiation falling on it? What is the pressure of radiation on a surface which perfectly reflects all the radiation falling on it? Consider the case when the direction of the incoming radiation is perpendicular to the surface.

The pressure of radiation from a flux density FF falling on a surface which perfectly absorbs all the radiation falling on it is

P=FcP = \frac{F}{c}

If however, all the radiation falling on the surface is reflected, aside from the pressure due to the incident radiation, there is also a pressure due to the reflected radiation, since the reflected radiation carries momentum away from the surface. The net pressure is thus

P=Pincident+Pemitted=Fc+Fc=2Fc P = P_\text{incident} + P_{emitted} = \frac{F}{c} + \frac{F}{c} = \frac{2F}{c}

Hence the radiation pressure on a perfectly reflecting surface is twice that on a perfectly absorbing surface.

Blackbody radiation

What is a Blackbody?

A blackbody is an object that does not reflect or scatter any radiation falling on it, but absorbs all the radiation falling on it completely. It emits radiation evenly in all directions and at all wavelengths. Kirchoff’s definition of a blackbody is an ideal body that neither reflects any light nor allows it to pass through. The emissivity ϵ\epsilon of a body is defined as the ratio of the radiation emitted by the body to the radiation emitted by a blackbody of the same shape at the same temperature. The emissivity ϵ\epsilon of a blackbody is 1, while that of a perfect reflector is 0. The emissivity of a real body lies between 0 and 1.

ϵ=BBblackbody(2.1.14)\tag{2.1.14} \epsilon = \frac{B}{B_\text{blackbody}}

where BB is the radiance of the body and BblackbodyB_\text{blackbody} is the radiance of a blackbody at the same temperature.

Kirchoff’s law states that the emissivity of a body is equal to its absorptivity aa. This means that a good absorber is also a good emitter. So, a perfect absorber (like a blackbody) is also a perfect emitter.

Examples of blackbodies include stars (which closely approximate blackbody spectra), the cosmic microwave background (a nearly perfect blackbody from the early universe), and laboratory setups like a cavity with a small hole that traps radiation.

The radiation emitted by a blackbody only depends on its temperature, and follows Planck’s law. Planck’s law describes how the light from a blackbody is spread across different wavelengths or frequencies, depending only on its temperature. The spectral radiance of a blackbody at temperature TT is given by

Bν(T)=2hν3c21ehν/kT1Bλ(T)=2hc2λ51ehc/λkT1(2.1.15) \tag{2.1.15} \boxed{ \begin{aligned} B_\nu(T) &= \frac{2 h \nu^3}{c^2} \frac{1}{e^{h\nu / kT} - 1} \\ B_\lambda(T) &= \frac{2 h c^2}{\lambda^5} \frac{1}{e^{h c / \lambda k T} - 1} \\ \end{aligned} }

These two are connected by the relation Bν(T)dν=Bλ(T)dλB_\nu(T) \, d\nu = -B_\lambda(T) \, d\lambda (the minus sign comes from the fact that as frequency increases, wavelength decreases.).

Produced by GNUPLOT 4.4 patchlevel 0, background spectrum based on http://commons.wikimedia.org/wiki/File:Linear_visible_spectrum.svg UV UV UV UV УВ UV UW УВ UV UV SYNLIGT VISIBLE SICHTBAR VISIBILE ВИДЛИВИ VISIBLE WĪDAMINS ОЧЕВИДНИй VISIBLE VISIBLE INFRARØD INFRAROIG INFRAROT INFAROSSO ИНФРАЦРВЕН INFRARROJO INFRAWURMIS ИНФРАЧЕРВОНЕ ВИПРОМІНЮВАННЯ INFRAROUGE INFRARED 0 2 4 6 8 10 12 14 0 0.5 1 1.5 2 2.5 3 Spectral intensitet (kW · sr⁻¹ · m⁻² · nm⁻¹) radiació espectral (kW · sr⁻¹ · m⁻² · nm⁻¹) Strahlungsleistung (kW · sr⁻¹ · m⁻² · nm⁻¹) Irradianza (kW · sr⁻¹ · m⁻² · nm⁻¹) Интензитет (kW · sr⁻¹ · m⁻² · nm⁻¹) Radiación espectral (kW · sr⁻¹ · m⁻² · nm⁻¹) spektriska spārtas kāmstis (kW · sr⁻¹ · m⁻² · nm⁻¹) Інтенсивність (kW · sr⁻¹ · m⁻² · nm⁻¹) Luminance spectrale (kW · sr⁻¹ · m⁻² · nm⁻¹) Spectral radiance (kW · sr⁻¹ · m⁻² · nm⁻¹) Bølgelængde (µm) longitud d'ona (μm) Wellenlänge (μm) Lunghezza d'onda (μm) Бранова должина (μm) Longitud de onda (μm) bāngas ilgan (μm) Довжина хвилі (μm) Longueur d'onde (nm) Wavelength (μm) 5000 K 4000 K 3000 K Klassisk teori (5000 K) Teoria clàssica (5000 K) Klassische Theorie (5000 K) Teoria classica (5000 K) Класична теорија (5000 K) Teoría clásica (5000 K) Klassiska fizīki (5000 K) Класична фізика (5000 K) Théorie classique (5000 K) Classical theory (5000 K) 0 2 4 6 8 10 12 14 0 0.5 1 1.5 2 2.5 3 Spectral intensitet (kW · sr⁻¹ · m⁻² · nm⁻¹) radiació espectral (kW · sr⁻¹ · m⁻² · nm⁻¹) Strahlungsleistung (kW · sr⁻¹ · m⁻² · nm⁻¹) Irradianza (kW · sr⁻¹ · m⁻² · nm⁻¹) Интензитет (kW · sr⁻¹ · m⁻² · nm⁻¹) Radiación espectral (kW · sr⁻¹ · m⁻² · nm⁻¹) spektriska spārtas kāmstis (kW · sr⁻¹ · m⁻² · nm⁻¹) Інтенсивність (kW · sr⁻¹ · m⁻² · nm⁻¹) Luminance spectrale (kW · sr⁻¹ · m⁻² · nm⁻¹) Spectral radiance (kW · sr⁻¹ · m⁻² · nm⁻¹) Bølgelængde (µm) longitud d'ona (μm) Wellenlänge (μm) Lunghezza d'onda (μm) Бранова должина (μm) Longitud de onda (μm) bāngas ilgan (μm) Довжина хвилі (μm) Longueur d'onde (nm) Wavelength (μm) 5000 K 4000 K 3000 K Klassisk teori (5000 K) Teoria clàssica (5000 K) Klassische Theorie (5000 K) Teoria classica (5000 K) Класична теорија (5000 K) Teoría clásica (5000 K) Klassiska fizīki (5000 K) Класична фізика (5000 K) Théorie classique (5000 K) Classical theory (5000 K)
Black Body Radiation Curve (source: Wikipedia)

These equations for spectral radiance can be simplified when the wavelengths in question are much larger than the peak wavelength of the spectrum or when they are approximately equal to the peak wavelength of the spectrum.

When λλpeak\lambda \gg \lambda_\text{peak} or ννpeak\nu \ll \nu_\text{peak}, we can use the Rayleigh-Jeans approximation

Bν(T)2ν2kTc2(2.1.16)\tag{2.1.16} B_\nu(T) \approx \frac{2 \nu^2 kT}{c^2}

Bλ(T)2ckTλ4(2.1.17)\tag{2.1.17} B_\lambda(T) \approx \frac{2c kT}{\lambda^4}

The Rayleigh-Jeans approximation is used often in radio astronomy, where the wavelengths are much larger than the peak wavelength of the blackbody spectrum.

✏️
Derive the Rayleigh-Jeans approximation from Planck’s law.

When λλpeak\lambda \lesssim \lambda_\text{peak} or ννpeak\nu \gtrsim \nu_\text{peak}, we can use the Wien approximation

Bν(T)2hν3c2ehν/kT(2.1.18)\tag{2.1.18} B_\nu(T) \approx \frac{2 h \nu^3}{c^2} e^{-h\nu / kT}

Bλ(T)2hc2λ5ehc/λkT(2.1.19)\tag{2.1.19} B_\lambda(T) \approx \frac{2 h c^2}{\lambda^5} e^{-h c / \lambda k T}

Stefan Boltzmann Law

The total radiance is the spectral radiance integrated over all frequencies. It is given by

B(T)=0Bν(T)dν=2π4k415h3c2T4 \begin{aligned} B(T) &= \int_0^{\infty} B_\nu(T) \, d\nu \\ &= \frac{2 \pi^4 k^4}{15 h^3 c^2} T^4 \\ \end{aligned}

This gives

B(T)=σπT4(2.1.20)\tag{2.1.20} B(T) = \frac{\sigma}{\pi} \, T^4

Here we have defined a new constant - called the Stefan-Boltzmann constant as

σ=2π5k415h3c2(2.1.21)\tag{2.1.21} \sigma = \frac{2 \pi^5 k^4}{15 h^3 c^2}

It is numerically equal to σ5.67×108Wm2K4\sigma \approx 5.67 \times 10^{-8} \, \mathrm{W \, m^{-2} \, K^{-4}}.

🧮
The integral can be carried out as follows:

0Bν(T)dν=2hc20ν3ehν/kT1dν=2hc2(kTh)40x3ex1dx \begin{aligned} \int_0^{\infty} B_\nu(T) , d\nu &= \frac{2 h}{c^2} \int_0^{\infty} \frac{\nu^3}{e^{h\nu / kT} - 1} , d\nu \ &= \frac{2 h}{c^2} \left( \frac{kT}{h} \right)^4 \int_0^{\infty} \frac{x^3}{e^x - 1} , dx \ \end{aligned}

The integral 0x3ex1dx=π415\int_0^{\infty} \frac{x^3}{e^x - 1} , dx = \frac{\pi^4}{15} is a standard result in mathematics. For the more curious readers, the result can be derived using the Riemann zeta function ζ(s)\zeta(s) and the gamma function Γ(s)\Gamma(s). We can use the following relation:

0xs1ex1dx=Γ(s)ζ(s)\int_0^{\infty} \frac{x^{s-1}}{e^x - 1} , dx = \Gamma(s) \zeta(s)

For s=4s=4, we have Γ(4)=6\Gamma(4) = 6 and ζ(4)=π490\zeta(4) = \frac{\pi^4}{90}. Hence,

0x3ex1dx=Γ(4)ζ(4)=6π490=π415\int_0^{\infty} \frac{x^3}{e^x - 1} , dx = \Gamma(4) \zeta(4) = 6 \cdot \frac{\pi^4}{90} = \frac{\pi^4}{15}

This gives the result

0Bν(T)dν=2hc2(kTh)4π415=2π4k415h3c2T4\int_0^{\infty} B_\nu(T) , d\nu = \frac{2 h}{c^2} \left( \frac{kT}{h} \right)^4 \cdot \frac{\pi^4}{15} = \frac{2 \pi^4 k^4}{15 h^3 c^2} T^4

The flux density for an isotropic source is given by (using eq 2.1.7):

F(T)=πB(T)=σT4(2.1.22)\tag{2.1.22} \boxed{F(T) = \pi B(T) = \sigma T^4}

This is known as the Stefan-Boltzmann law.

If the radius of a star is RR and its temperature is TT, then its luminosity is given by (assuming the star emits isotropically):

L=4πR2F=4π2R2B=4πR2σT4(2.1.23)\tag{2.1.23} L = 4 \pi R^2 F = 4 \pi^2 R^2 B = 4 \pi R^2 \sigma T^4

The spectral luminosity is defined as Lλ=4π2R2BλL_\lambda = 4 \pi^2 R^2 B_\lambda. The spectral flux density at a distance rr is given by

Fλ=Lλ4πr2=πBλ(Rr)2(2.1.24)\tag{2.1.24} F_\lambda = \frac{L_\lambda}{4\pi r^2} = \pi B_\lambda \left( \frac{R}{r} \right)^2

For a blackbody having surface area AA, its luminosity is given by

L=AσT4(2.1.25)\tag{2.1.25} \boxed{L = A \sigma T^4}

If the object is not a perfect blackbody, we can still use the Stefan-Boltzmann law, but we need to multiply by the emissivity ϵ\epsilon of the object. The luminosity is then given by

L=AϵσT4(2.1.26)\tag{2.1.26} L = A \epsilon \sigma T^4
The flux density of a star on Earth is observed to be F=1012Wm2F = 10^{-12} \, \mathrm{W \, m^{-2}}. The star has a parallax of 4mas4 \, \mathrm{mas}. From spectroscopic measurements, it is found that the star has an effective temperature of T=5000KT = 5000 \, \mathrm{K}. What is the radius of the star?

This distance to the star is

d=1π=250kpc d = \frac{1}{\pi} = 250 \, \mathrm{kpc}

The flux density of the star is given by

F=L4πd2=4πR2σT44πd2 F = \frac{L}{4 \pi d^2} = \frac{4\pi R^2 \sigma T^4}{4 \pi d^2}

Rearranging, we get the radius of the star to be

R=dT2Fσ1.3×109m1.9R R = \frac{d}{T^2} \sqrt{\frac{F}{\sigma}} \approx 1.3 \times 10^9 \, \mathrm{m} \approx 1.9 \, \mathrm{R_\odot}

Wien’s Displacement Law

Hotter objects emit more at higher energies (higher frequencies and lower wavelengths). This can be quantified by Wien’s law, which relates the peak wavelength and frequency (wavelength or frequency corresponding to the mode energy of the radiated photons; the wavelength or frequency at which the blackbody emits the most) of the blackbody spectrum to the temperature of the object. The peak of the blackbody spectrum shifts with temperature, and this is described by Wien’s displacement law. This is the reason why stars have different colors, and why metals glow red then white when heated.

The wavelength corresponding to the peak of the blackbody spectrum is given by

ddλBλ(T)=0\frac{d}{d\lambda} B_\lambda(T) = 0

This gives the relation

λpeakT=bλ(2.1.27)\tag{2.1.27} \boxed{\lambda_\text{peak} T = b_\lambda}

BλB_\lambda peaks at the energy

E4.965kT(2.1.28)\tag{2.1.28} E \approx 4.965 \,kT
🧮
Taking the derivative gives

Taking the derivative ddλBλ(T)\frac{d}{d\lambda} B_\lambda (T) and defining x=hcλkTx = \frac{hc}{\lambda kT}, after some algebra we get

xexex1=5\frac{x e^x}{e^x - 1} = 5

This equation can be solved using the Lambert W function, which gives the solution for xx as x=5+W0(5e5)4.965x = 5 + W_0(-5 e^{-5}) \approx 4.965. This gives the value of bλ=hckx2.898×103mKb_\lambda = \frac{hc}{kx} \approx 2.898 \times 10^{-3} \mathrm{\, m \, K}.

The peak energy can be calculated using the relation E=hcλ=xkT4.965kTE = \frac{hc}{\lambda} = xkT \approx 4.965 \,kT.

The frequency corresponding to the peak of the blackbody spectrum is given by

ddνBν(T)=0\frac{d}{d\nu} B_\nu(T) = 0

This gives the relation

νpeak/T=bν(2.1.29)\tag{2.1.29} \boxed{\nu_\text{peak} / T = b_\nu}

BνB_\nu peaks at the energy

E2.821kT(2.1.30)\tag{2.1.30} E \approx 2.821 \,kT
🧮
Taking the derivative gives

Taking the derivative ddνBν(T)\frac{d}{d\nu} B_\nu (T) and defining y=hνkTy = \frac{h\nu}{kT}, after some algebra we get

yeyey1=3\frac{y e^y}{e^y - 1} = 3

This gives y=3+W0(3e3)2.821y = 3 + W_0(-3 e^{-3}) \approx 2.821. This gives the value of bν=kyh5.879×1010HzKb_\nu = \frac{ky}{h} \approx 5.879 \times 10^{10} \mathrm{\, Hz \, K}.

The peak energy can be calculated using the relation E=hν=ykT2.821kTE = h \nu = ykT \approx 2.821 \,kT.

The reason eq 2.1.28 and 2.1.30 are different is because the peak of the spectrum is not at the same energy for both λ\lambda and ν\nu. One cannot go from BνB_\nu to BλB_\lambda by simply changing the variable. One also needs to multiply by dν/dλ=c/λ2|d\nu / d\lambda| = c /\lambda^2, which shifts the peak of the distribution to higher energies. These peaks are the mode energy of a photon.

Equation 2.1.27 and 2.1.29 are together known as Wien’s displacement law.

Using Wien’s displacement law, determine the peak wavelength of radiation emitted by stars with the following temperatures, and indicate what color they would appear to human eyes:

a) A red dwarf at 3000 K
b) A Sun-like star at 5800 K
c) A blue giant at 25,000 K

Using Wien’s displacement law: λpeak=bλT\lambda_\text{peak} = \frac{b_\lambda}{T}

a) Red dwarf (3000 K):

λpeak=2.898×1033000966 nm\lambda_\text{peak} = \frac{2.898 \times 10^{-3}}{3000} \approx 966 \text{ nm}

This is in the near-infrared range, but the visible spectrum tail extends into red wavelengths, making the star appear red.

b) Sun-like star (5800 K):

λpeak=2.898×1035800500 nm\lambda_\text{peak} = \frac{2.898 \times 10^{-3}}{5800} \approx 500 \text{ nm}

This is in the green part of the visible spectrum, but the full spectrum makes the star appear yellow-white.

c) Blue giant (25,000 K):

λpeak=2.898×10325000116 nm\lambda_\text{peak} = \frac{2.898 \times 10^{-3}}{25000} \approx 116 \text{ nm}

This is in the extreme ultraviolet range, but the visible tail makes the star appear blue.

Properties of Blackbody Radiation

The energy density of blackbody radiation is given by

u(T)=04πcBν(T)dν=4σcT4=aT4(2.1.31)\tag{2.1.31} u(T) = \int_0^{\infty} \frac{4 \pi}{c} B_\nu(T) \, d\nu = \frac{4 \sigma}{c} T^4 = aT^4

where a7.566×1016Jm3K4a \approx 7.566 \times 10^{-16} \, \mathrm{J \, m^{-3} \, K^{-4}} is the radiation constant. The radiation pressure is given by

P=13u=13aT4(2.1.32)\tag{2.1.32} P = \frac{1}{3} u = \frac{1}{3} a T^4

As stated earlier, energy comes in quantized packets called photons. The number density of photons emitted by a blackbody at temperature TT is given by

n(T)=0uν(T)hνdν=16πk3ζ(3)h3c3T3 \begin{aligned} n(T) &= \int_0^{\infty} \frac{u_\nu(T)}{h \nu} \, d\nu \\ &= \frac{16 \pi k^3 \zeta(3)}{h^3 c^3} T^3 \\ \end{aligned}

Hence,

n(T)=βT3(2.1.33)\tag{2.1.33} n(T) = \beta \, T^3

where β2.03×107m3K3\beta \approx 2.03 \times 10^7 \mathrm{m^{-3} K^{-3}}.

🧮
The integral can be carried out as follows:

0uν(T)hνdν=2c24πc0ν2ehν/kT1dν=2c24πc(kTh)30x2ex1dx \begin{aligned} \int_0^{\infty} \frac{u_\nu(T)}{h \nu} , d\nu &= \frac{2}{c^2} \frac{4 \pi}{c} \int_0^{\infty} \frac{\nu^2}{e^{h\nu / kT} - 1} , d\nu \ &= \frac{2}{c^2} \frac{4 \pi}{c} \left( \frac{kT}{h} \right)^3 \int_0^{\infty} \frac{x^2}{e^x - 1} , dx \ \end{aligned}

Using the previously stated relation

0xs1ex1dx=Γ(s)ζ(s)\int_0^{\infty} \frac{x^{s-1}}{e^x - 1} , dx = \Gamma(s) \zeta(s)

For s=3s=3, we have Γ(3)=2\Gamma(3) = 2. This gives the result

0uν(T)hνdν=2c24πc(kTh)32ζ(3)=16πk3ζ(3)h3c3T3\int_0^{\infty} \frac{u_\nu(T)}{h \nu} , d\nu = \frac{2}{c^2} \frac{4 \pi}{c} \left( \frac{kT}{h} \right)^3 \cdot 2 \zeta(3) = \frac{16 \pi k^3 \zeta(3)}{h^3 c^3} T^3

The mean energy of the photons emitted by the blackbody is

E=u(T)n(T)=π430ζ(3)kT2.70kT(2.1.34)\tag{2.1.34} \langle E \rangle = \frac{u(T)}{n(T)} = \frac{\pi^4}{30 \zeta(3)} kT \approx 2.70 \, kT
Find the average energy of the cosmic microwave background (CMB) photons. The CMB is a nearly perfect blackbody radiation from the early universe, with a blackbody temperature of 2.7 K.

The average energy of the CMB photons can be calculated using the relation

E2.70kT\langle E \rangle \approx 2.70 kT

For T=2.7KT = 2.7 \, \mathrm{K}, we have

E1.02×1022J6.36×104eV\langle E \rangle \approx 1.02 \times 10^{-22} \, \mathrm{J} \approx 6.36 \times 10^{-4} \, \mathrm{eV}

This is in the microwave range of the electromagnetic spectrum.

Temperature

The temperature of a star can be defined in several ways.

  • The effective temperature TeT_e is defined as the temperature of a blackbody which radiates the same total luminosity as the star. For example, the Sun has a luminosity of 3.846×1026W3.846 \times 10^{26} \, \mathrm{W} and a radius of 6.96×108m6.96 \times 10^8 \, \mathrm{m}. The effective temperature of the Sun is given by
Te=(L4πσR2)1/45778K T_e = \left( \frac{L}{4 \pi \sigma R^2} \right)^{1/4} \approx 5778 \mathrm{\, K}
  • If we assume at some wavelength λ\lambda the spectral flux density FλF_\lambda at the surface of the star is given by Planck’s law, we get the brightness temperature TbT_b. This is the temperature of a blackbody that matches the brightness of the star at that specific wavelength. Since stars are not ideal blackbodies, TbT_b depends on the wavelength used. This is commonly used in radio astronomy, where the Rayleigh-Jeans approximation is used to find brightness temperature at a specific wavelength/frequency.

  • The temperature giving the best fit for Planck’s law in a wavelength range is the color temperature TcT_c of the object.

Flux densities at the wavelengths 440nm440 \, \mathrm{nm} and 550nm550 \, \mathrm{nm} are 1.301.30 and 1.00Wm2m11.00 \, \mathrm{W \, m^{-2} \, m^{-1}}, respectively. Find the colour temperature.

If the flux densities at wavelengths λ1\lambda_1 and λ2\lambda_2 are F1F_1 and F2F_2 respectively, the color temperature can be solved from the equation

F1F2=Bλ(Tc,λ1)Bλ(Tc,λ2)=(λ2λ1)5ehc/λ2kTc1ehc/λ1kTc1\frac{F_1}{F_2} = \frac{B_\lambda(T_c, \lambda_1)}{B_\lambda(T_c, \lambda_2)} = \left(\frac{\lambda_2}{\lambda_1} \right)^5 \frac{e^{hc/\lambda_2 kT_c} - 1}{e^{hc/\lambda_1 kT_c} - 1}

We define

A=F1F2(λ1λ2)5=2.348B1=hcλ1k=32700KB2=hcλ2k=26160K\begin{align*} A &= \frac{F_1}{F_2} \left( \frac{\lambda_1}{\lambda_2} \right)^5 = 2.348 \\ B_1 &= \frac{hc}{\lambda_1 k} = 32700 \, \mathrm{K} \\ B_2 &= \frac{hc}{\lambda_2 k} = 26160 \, \mathrm{K} \end{align*}

Thus the equation becomes

A=eB2/Tc1eB1/Tc1 A = \frac{e^{B_2/T_c} - 1}{e^{B_1/T_c} - 1}

Solving the equation numerically, we get Tc=7545K\boxed{T_c = 7545 \, \mathrm{K}}

Problems

Estimate the effective temperature of the photosphere of the Sun using the naked eye colour of the Sun.

Using Wien’s law,

λpeakT=bλT=bλλpeak2.898×103500×109K=5800K6000K \begin{aligned} \lambda_\text{peak} T &= b_\lambda \\ T &= \frac{b_\lambda}{\lambda_\text{peak}} \\ &\approx \frac{2.898 \times 10^{-3}}{500 \times 10^{-9}} \, \mathrm{K} = 5800 \mathrm{\, K}\\ &\approx \boxed{6000 \, \mathrm{K}} \\ \end{aligned}

The temperature is rounded as the peak wavelength has only one significant digit.

The entrance cavity of a particular bolometer is a cone with an opening angle of 3030^\circ, the surface of which has an energy absorption coefficient of a=0.99a = 0.99. Assume that there is no scattering of the incident radiation on the walls of the cavity, only multiple mirror-like (specular) reflections. The bolometer is connected to a cooler which keeps the bolometer cavity surface at practically 0K0 \, \mathrm{K} temperature. The instrument is orbiting at r=2AUr = 2 \, \mathrm{AU} from the Sun and is pointed directly at the centre of the Solar disk. Calculate the temperature of a black body which would radiate the same amount of energy as the entire bolometer opening does per unit surface area.

Drawing a diagram, the light entering parallel to the axis of the bolometer reflects six times off of the bolometer walls before escaping back to space. Hence the fraction of energy leaving the bolometer is

S=(1a)6=1012 S = (1 - a)^6 = 10^{-12}

The flux density at the bolometer is

F=L4πr2 F = \frac{L_\odot}{4 \pi r^2}

The flux density absorbed by the bolometer is thus FSF \cdot S.

The bolometer must emit the same amount of energy it absorbs to remain in thermal equillibrium. Thus

FS=σT4 F \cdot S = \sigma T^4

    T=(L(1a)64πσr2)1/4=0.28K \implies T = \left( \frac{L_\odot (1-a)^6}{4 \pi \sigma r^2} \right)^{1/4} = \boxed{0.28 \, \mathrm{K}}
Find the size of a spherical grain of dust (density ρ=5000kgm3\rho = 5000 \, \mathrm{kg \, m^{-3}}) which can remain stationary in space due to radiation pressure from the Sun.

Let the radius of the grain of dust be RR. The radiation pressure of Sun at a distance rr is given by

P=Fc=L4πr2cP = \frac{F}{c} = \frac{L}{4 \pi r^2 c}

This gives a radiation force acting on the grain of dust of area A=πR2A = \pi R^2 as

Fr=PA=L4πr2cπR2F_r = PA = \frac{L }{4 \pi r^2 c} \pi R^2

The mass of the grain of dust is given by

m=ρ43πR3m = \rho \frac{4}{3} \pi R^3

Hence, the gravitational force acting on the grain of dust is given by

Fg=GMmr2=GMρ43πR3r2F_g = \frac{GMm}{r^2} = \frac{GM \rho \frac{4}{3} \pi R^3}{r^2}

For the grain of dust to remain stationary, the radiation force must be equal to the gravitational force. Hence,

L4πr2cπR2=GMρ43πR3r2\frac{L }{4 \pi r^2 c} \pi R^2 = \frac{GM \rho \frac{4}{3} \pi R^3}{r^2}

Rearranging gives

R=3L16πρcGM115nmR = \frac{3L}{16 \pi \rho c GM} \approx \boxed{115 \, \mathrm{nm}}
Consider a surface of surface area AA and reflection coefficient of α\alpha, that is it reflects a fraction α\alpha of the radiation falling on it. The rest of the radiation is absorbed. The surface is illuminated by radiation of flux density FF, at an angle θ\theta to the normal. Find the force exerted on the surface by the radiation.

Let the normal of the surface be pointing along the zz-axis, the surface being in the xyxy-plane, and the incoming radiation be in the zxzx-plane, making angle θ\theta with the zz-axis. The unit vector corresponding to the direction of the incoming radiation is thus i^=sinθx^+cosθz^\hat{i} = \sin \theta \,\hat{x} + \cos \theta \, \hat{z}. Since the radiation is incident at an angle θ\theta, the projected area of the surface in the direction of the radiation is Aproj=AcosθA_\text{proj} = A \cos \theta.

The flux density incident on the surface is FF, and that reflected away is αF\alpha F. The reflected radiation will be in the direction r^=sinθx^+cosθz^\hat{r} = -\sin \theta \,\hat{x} + \cos \theta \, \hat{z}, according to the laws of reflection (the angle between incident ray and the normal is equal to the angle between the reflected ray and the normal, and that the incident ray, reflected ray and normal all lie in a single plane).

The force due to incident radiation on the surface is

Fincident=FcAproji^F_\text{incident} = \frac{F}{c} A_\text{proj} \, \hat{i}

The force due to the reflected radiation is

Freflected=αFcAproj(r^)F_\text{reflected} = \frac{\alpha F}{c} A_\text{proj} \, (-\hat{r})

The direction of the force due to reflected radiation will be opposite to its direction of propagation.

The net force on the surface is therefore

Fnet=Fincident+Freflected=FcAproji^+αFcAproj(r^)F_\text{net} = F_\text{incident} + F_\text{reflected} = \frac{F}{c} A_\text{proj} \, \hat{i} + \frac{\alpha F}{c} A_\text{proj} \, (-\hat{r})

=FcAproj(i^αr^)= \frac{F}{c} A_\text{proj} \left( \hat{i} - \alpha \hat{r} \right)

=FcAcosθ[(1α)sinθx^+(1+α)cosθz^]= \frac{F}{c} \, A \cos \theta \, \left[ (1 - \alpha) \sin \theta \, \hat{x} + (1 + \alpha) \cos \theta \, \hat{z} \right]

We see that if the surface is perfectly reflecting (α=1\alpha = 1), the force is directed opposite to the normal (zz-axis), having a magnitude of FcAcos2θ\frac{F}{c} A \cos^2 \theta. Whereas if the surface is perfectly absorbing (α=0\alpha = 0), the force is directed along the direction of the incoming radiation, having a magnitude of FcAcosθ\frac{F}{c} A \cos \theta.

If θ=0\theta = 0, that is the radiation is falling normal to the surface, the force experienced by the surface will be opposite to the normal or the direction of incident radiation, having a magnitude of FcA(1+α)\frac{F}{c} A \, (1 + \alpha).

Consider a hypothetical solar sail that starts its trajectory just outside the surface of the Sun and aims to reach Jupiter’s orbit. This is a square sail with a size length of l=50ml = 50 \, \mathrm{m}, a mass of m=2.0kgm = 2.0 \, \mathrm{kg}, and a reflectivity of η=85%\eta = 85\%. The remaining 15%15\% of the photons are absorbed by the sail. In order to simplify the calculations, only take into account the gravity and radiation pressure from the Sun, and assume that the sail is initially stationary with respect to the Sun. Estimate how long it would take for this sail to reach Jupiter’s orbit. Feel free to make reasonable approximations if needed.

The radiation flux on the sail when it is at a distance rr from the Sun is L4πr2\frac{L_\odot}{4 \pi r^2}. Therefore the radiation force acting on the sail is

Frad=L4πr21cA(1+η) F_\text{rad} = \frac{L_\odot}{4 \pi r^2} \frac{1}{c} A \, (1 + \eta)

where A=l2A = l^2 is the area of the sail. The gravitational force acting on the sail is

Fg=GMmr2 F_g = \frac{GM_\odot m}{r^2}

The net force acting on the sail is thus

F=FradFg=L4πr21cA(1+η)GMmr2 F = F_\text{rad} - F_g = \frac{L_\odot}{4 \pi r^2} \frac{1}{c} A \, (1 + \eta) - \frac{GM_\odot m}{r^2}

Defining the constant α\alpha to be

α=Ll24πmc(1+η)GMm\alpha = \frac{L_\odot l^2}{4 \pi \, m c} (1 + \eta) - GM_\odot m

The acceleration of the sail is therefore

a=Fm=αr2 a = \frac{F}{m} = \frac{\alpha}{r^2}

Since the acceleration is only in the radial direction,

a=dvdt=dvdrdrdt=vdvdra = \frac{dv}{dt} = \frac{dv}{dr} \frac{dr}{dt} = v \frac{dv}{dr}

We can write

vdvdr=αr2v \frac{dv}{dr} = \frac{\alpha}{r^2}

Integrating and substituting the limits, we get

0vvdv=Rrαr2dr\int_0^v v dv = \int_{R_\odot}^r \frac{\alpha}{r^2} dr

v2=2α(1r1R)v^2 = -2\alpha \left( \frac{1}{r} - \frac{1}{R_\odot} \right)

v=drdt=2αR(1Rr)v = \frac{dr}{dt} = \sqrt{\frac{2\alpha}{R_\odot} \left( 1-\frac{R_\odot}{r}\right)}

Recommended integrating a second time and substituting the limits, we get

RaJdr=0T2αR(1Rr)dt\int_{R_\odot}^{a_J} dr = \int_0^T \sqrt{\frac{2\alpha}{R_\odot} \left( 1-\frac{R_\odot}{r}\right)} dt

where aJa_J is the orbital radius of Jupiter. Rearranging and defining r=(u+1)Rr = (u + 1) R_\odot and aJ=(a+1)Ra_J = (a + 1) R_\odot gives

T=T00a1+1uduT = T_0 \int_{0}^{a} \sqrt{1 + \frac{1}{u}} \, du

where T0=R32α21.4minT_0 = \sqrt{\frac{R_\odot^3}{2\alpha}} \approx 21.4 \, \mathrm{min} and a1117a \approx 1117. This integral cannot be evaluated analytically, and numerical methods need to be used. Thus, we will approximate the integral. Here, we will be working with a very crude approximating.

Notice that when u1u \gg 1, the integrand is approximately equal to 11. When u1u \ll 1, the integrand is approximately equal to u1/2u^{-1/2}. Splitting the integral at u=1u = 1 and evaluating them separately gives

TT0(01u1/2du+1a1du)T \approx T_0 \left( \int_{0}^{1} u^{-1/2} \, du + \int_{1}^{a} 1 \, du \right)

=T0(2+a1)=T0(a+1)= T_0 \left( 2 + a - 1 \right) = T_0 (a + 1)

This gives the time taken to reach Jupiter’s orbit as T16.6days\boxed{T \approx 16.6 \, \mathrm{days}}.

Numerically integrating, the time taken to reach Jupiter’s orbit is T16.7daysT \approx 16.7 \, \mathrm{days}, very close to our crude approximation.

Assuming the human body to be a perfect blackbody, estimate the number of photons emitted in a bandwidth 10% wide around wavelength 550 nm (peak emission wavelength of Sun), by a person in their entire life. Make reasonable assumptions.

Assuming the human body to be a perfect blackbody, its spectral luminosity will be

Lλ=AπBλL_\lambda = A \pi B_\lambda

where A1.7m2A \approx 1.7 \, \mathrm{m^2} is the surface area of the human body. The temperature of the body is T310KT \approx 310 \, \mathrm{K}. The number of photons emitted per unit time per unit wavelength is thus

2Ntλ=Lλhc/λ=AπhcλBλ\frac{\partial^2 N}{\partial t \,\partial \lambda} = \frac{L_\lambda}{h c / \lambda} = \frac{A \pi}{hc} \lambda B_\lambda

Therefore the number of photons emitted by the human body in the given wavelength range in its entire lifetime of τ80yrs\tau \approx 80 \, \mathrm{yrs} is

N=Aπτhcλ1λ2λBλdλN = \frac{A \pi \tau}{hc} \int_{\lambda_1}^{\lambda_2} \lambda B_\lambda d\lambda

where λ1=λ0+0.05λ0\lambda_1 = \lambda_0 + 0.05 \lambda_0 and λ2=λ00.05λ0\lambda_2 = \lambda_0 - 0.05 \lambda_0 are the limits of the bandwidth (λ0=550nm\lambda_0 = 550 \, \mathrm{nm}). Using planck’s law and definfing x=hcλkTx = \frac{hc}{\lambda kT}, the integral becomes

N=2πk3T3Aτh3c2x2x1x2dxex1N = \frac{2 \pi k^3 T^3 A \tau}{h^3 c^2} \int_{x_2}^{x_1} \frac{x^2 \, dx}{e^x - 1}

where x1=hcλ1kT89x_1 = \frac{hc}{\lambda_1 kT} \approx 89 and x2=hcλ2kT81x_2 = \frac{hc}{\lambda_2 kT} \approx 81. Since x1x \gg 1, we can approximate ex1exe^x - 1 \approx e^x and the integral becomes

N=N0x2x1x2exdxN = N_0 \int_{x_2}^{x_1} x^2 e^{-x} dx

where N0=2πk3T3Aτh3c28×1031photonsN_0 = \frac{2 \pi k^3 T^3 A \tau}{h^3 c^2} \approx 8 \times 10^{31} \, \mathrm{photons}. The integral now can be evaluated by parts to give

x2x1x2exdx=ex(x2+2x+2)x2x14.5×1032\int_{x_2}^{x_1} x^2 e^{-x} dx = e^{-x} (x^2 + 2x + 2) \bigg|_{x_2}^{x_1} \approx 4.5 \times 10^{-32}

Therefore the total number of photons emitted by a human body in a 10%10\% bandwidth centered around 550nm550 \, \mathrm{nm} in its entire lifetime is

N4photons\boxed{N \approx 4 \, \mathrm{photons}}

The Five-hundred-meter Aperture Spherical radio Telescope (FAST) is a single-dish radio telescope located in Guizhou Province, China. The physical diameter of the dish is 500 m, but during observations, the effective diameter of the collecting area is 300 m. Consider observations of the thermal radio emission from the photosphere of the Sun at 3.0 GHz with this telescope and a receiver with bandwidth 0.3 GHz.

a) Calculate the total energy (EE_\odot) that the receiver will collect during 1 hour of observation.
b) Estimate the energy needed to turn over one page of your answer sheet (EE'). Hint: the typical surface density of paper is 80gm280 \mathrm{\,g \, m^{-2}}
c) Which one is larger?

a) Since we are measuring radio frequencies, we can use the Rayleigh-Jeans approximation

Bν(T)=2ν2kTc2 B_\nu(T) = \frac{2 \nu^2 kT}{c^2}

The spectral luminosity of the Sun is

Lν=4π2R2Bν(T)L_\nu = 4 \pi^2 R_\odot^2 B_\nu(T)

The spectral flux density recieved at Earth is

Fν=Lν4πD2F_\nu = \frac{L_\nu}{4 \pi D^2}

Hence the total energy collected by FAST in 1 hour is

E=FνΔνAt=π2ν2kTc2R2D2Δνπ4d2tE_\odot = F_\nu \Delta \nu \cdot A \cdot t = \pi \frac{2 \nu^2 kT}{c^2} \frac{R_\odot^2}{D^2} \Delta \nu \cdot \frac{\pi}{4} d^2 \cdot t

This gives E8.5×105J\boxed{E_\odot \approx 8.5 \times 10^{-5} \mathrm{\,J}}

b) The total mass of an A4 paper is

m=ρA=80gm2×297mm×210mm5g m = \rho A = 80 \, \mathrm{g \, m^{-2}} \times 297 \, \mathrm{mm} \times 210 \, \mathrm{mm} \approx 5 \, \mathrm{g}

The energy required to turn the paper is the same as the energy required to lift the center of mass of the paper by h=2102,mm=105mmh = \frac{210}{2} , \mathrm{mm} = 105 \, \mathrm{mm}

E=mgh5×103JE' = mgh \approx \boxed{5 \times 10^{-3} \, \mathrm{J}}

c) Therefore, E>E\boxed{E' > E_\odot}

Since the human body is made mostly of water, it is very efficient at absorbing microwave photons. Assume that an astronaut’s body is a perfect spherical absorber with mass of m=60kgm = 60 \,\mathrm{kg}, and its average density and heat capacity are the same as for pure water, i.e. ρ=1000kgm3\rho = 1000 \, \mathrm{kg\,m^{-3}} and C=4200Jkg1K1C = 4200 \mathrm{\,J\,kg^{-1}\,K^{-1}}.

a) What is the approximate rate, in watts, at which an astronaut in intergalactic space would absorb radiative energy from the Cosmic Microwave Background (CMB)? The spectral energy distribution of CMB can be approximated by blackbody radiation of temperature TCMB=2.728KT_\text{CMB} = 2.728 \,\mathrm{K}.
b) Approximately how many CMB photons per second would the astronaut absorb?
c) Ignoring other energy inputs and outputs, how long would it take for the CMB to raise the astronaut’s temperature by ΔT=1K\Delta T = 1 \, \mathrm{K}?

The radius of the spherical astronaut is

43πr3ρ=m    r=(3m4πρ)1/3\frac{4}{3} \pi r^3 \rho = m \implies r = \left( \frac{3m}{4\pi \rho} \right)^{1/3}

The surface area of the astronaut is

A=4πr2=4π(3m4πρ)2/3A = 4 \pi r^2 = 4 \pi \left( \frac{3m}{4\pi \rho} \right)^{2/3}

The power thus absorbed by the astronaut is

P=AσT4=4π(3m4πρ)2/3σTCMB42.3×106WP = A \sigma T^4 = 4 \pi \left( \frac{3m}{4\pi \rho} \right)^{2/3} \sigma T_\text{CMB}^4 \approx \boxed{2.3 \times 10^{-6} \, \mathrm{W}}

The number of photons absorbed per second can be approximated by dividing the total energy absorbed by the mean energy of a CMB photon.The mean energy of a photon is given by

E=2.70kTCMB1×1022JE = 2.70 kT_\text{CMB} \approx 1 \times 10^{-22} \, \mathrm{J}

Thus the number of photons absorbed is

n=PE2.3×1016s1n = \frac{P}{E} \approx \boxed{2.3 \times 10^{16} \, \mathrm{s^{-1}}}

If the time taken to raise the temperature of the astronaut by ΔT\Delta T is tt, then

ΔQ=mCΔT=Pt\Delta Q = m C \Delta T = P t

Rearranging gives

t=mCΔTP3430yrt = \frac{m C \Delta T}{P} \approx \boxed{3430 \, \mathrm{yr}}