Sunday, November 2, 2008

Charged-Particle Energy Loss in Matter

Contents

  1. Introduction
  2. Alpha Radiation
  3. Modified Bessel Functions of the Second Kind
  4. Classical Energy Loss
  5. Oscillator Strengths
  6. Quantum-Mechanical Energy Loss
  7. Fields in a Dense Medium
  8. Cherenkov Radiation
  9. Stopping of Magnetic Monopoles
  10. Comments and Numerical Examples
  11. References

This article fills in the details in Jackson's excellent account of the energy loss of charged particles in matter in his Chapter 13, including omitted derivations and mathematical details. This has always been an important subject since the discovery of radioactivity, and is of great importance in experiments and practical radioactivity. Both particle radiations, classically alpha and beta rays, and electromagnetic radiation, gamma rays, cause ionization in their passage through matter. Electrons are freed from neutral atoms, producing an ion pair of electron and positive ion. Of course, this disturbs chemical bonding as well as produces free charge carriers, which is the source of the biological effects of radiation.

It is interesting to ponder the fact that these radiations traverse matter at all, showing that apparent solidity is not solid at all. To these radiations, matter is mainly empty space in which electrons wander, with the massive, but very small, atomic nuclei presenting very little obstruction. The interactions are mainly with electrons, transferring energy to them, which is necessarily lost by the radiation. In fact, in all our analysis we shall treat the transfer of energy to the electrons, and use the principle of the conservation of energy to find the energy lost by the radiation.

The different radiations behave differently in their passage through matter. Particle radiations can be divided into heavy particles--that is, those massive relative to electrons, say more than about 20me in mass--and light particles, which are usually electrons or positrons, usually called beta rays. Both create many ions in frequent interactions with electrons, perhaps thousands per mm in air. Heavy particles, however, lose little of their energy in a single collision, and are practically undeflected, unless by a rare encounter with a nucleus. Light particles can lose a large fraction of their energy in a single collision with an electron. In fact, they can lose all of it, creating a knock-on electron of about the same energy. They are much more easily deflected, so their paths are not as straight as those of heavy particles. Electromagnetic radiation, considered as particle-like quanta of energies greater than 100 keV, does not fritter away its energy in multiple collisions with electrons, but ends its life in infrequent catastrophes, such as the ejection of an electron with disappearance of the quantum, or electron-positron pair production if is energy is greater than 1.05 MeV. Energy can be partially lost in Compton scattering fender-benders as well. In general, electromagnetic radiation finds matter much more transparent than charged particles do. Incidentally, a fast neutral atom quickly loses its electrons and becomes a positive ion, while neutral particles like uncharged pions act more like electromagnetic radiation. The most important factor in producing ionization and energy loss is the electromagnetic field of the charged particle.

The paths of charged particles can be seen macroscopically by the traces they leave in a cloud chamber, a bubble chamber or a photographic emulsion. What is seen is actually the effects of the ionization. In a cloud chamber, water droplets condense on negative ions. In a bubble chamber, the liquid vaporizes for a similar reason. The energy loss in a photographic emulsion causes the crystallites to become developable. Charged particles can also be seen by the elctromagnetic shock wave of Cherenkov radiation when they travel faster than the phase velocity of light in the medium. All these things can be discussed on the basis of the theory that will be presented here.

In this article, as in Jackson, formulas and electrical quantities are expressed in Gaussian units. Therefore, the electronic charge e = 4.803 x 10-10 esu should be used. This is c/10 times the charge in coulomb.

Alpha radiation consists of fast helium nuclei, with charge z = +2 and mass 4.00 amu, or rest mass 3.724 GeV, emitted from heavy nuclei. Their energies range from a maximum of about 10.53 MeV (Po212) down to about 4.05 MeV (Th232), below which the probability of alpha-decay becomes very small. The half-life of Po212 is only 300 ns, while that of Th232 is 1.39 x 1010 y. The half-life is a very strong function of the disintegration energy. It is clear that such natural alpha particles are nonrelativistic. However, this is not true of artificially produced heavy particles, such as protons or deuterons, nor of cosmic radiation, where fast protons are the usual primary radiation.

Alpha particles make thick, straight tracks in a cloud chamber of a definite length, the range R of the alpha particle. Looking more closely, the tracks are occasionally bent through an angle, usually near the ends, and straggle a little in range. The range in air at 15°C and 760 mmHg of the 10.53 MeV alpha from Po212 is about 11.6 cm, while the alphas from Th232 have a range of 2.49 cm. The energy of an alpha particle can be deduced fairly accurately from its range. The empirical relation R = 0.318E3/2, where R is in cm and E is in MeV, seems to hold fairly well for ranges from 3 to 7 cm.

Aluminium, with a density of 2.70 g/cc, has a stopping power about 1700 times that of air at STP. That is, the ranges in aluminium are 1700 times shorter than in air. The Po212 alpha is stopped by 0.07 mm of aluminium. This illustrates the very small penetrating power of heavy charged particles in solid matter.

The speed of an alpha particle of energy E MeV is v = 6.944 x 108√E cm/s. For particles of rest energy Eo, the speed of a particle of kinetic energy T can be found from γ = 1 + T/Eo and β = √(γ2 - 1)/γ. For a 1 GeV proton, γ = 2.07 and β = 0.878. Recall that 1 amu = 931 MeV rest energy. For a 1 GeV electron, γ = 1962 and β = 1.0000.

Modified Bessel functions of the second kind, Kn(x), appear prominently in the theory, so the properties we shall need are reviewed here. Usually, such functions appear in connection with problems with cylindrical geometry, but here they arrive as Fourier transforms. They are not difficult to work with, but should be well-understood. For integral n, they are infinite at x = 0, and decrease like a negative exponential as x increases. We will use K0(x) and K1(x) here, usually multiplied by x. Then, xK0(x) → 0 and xK1(x) → 1 as x → 0.

We know that Bessel's equation, y" + (1/x)y' + (1 - n2/x2)y = 0, with n integral, has one solution y = Jn(x) that is finite at the origin and for large x behaves like y = (2/πx)1/2cos(x - nπ/2 - π/4). All other solutions that are linearly independent of this one are singular at the origin. The one that is carefully arranged to behave like y = (2/πx)1/2sin(x - nπ/2 - π/4) at large x is called Weber's function, Nn(x). Then, the complex linear combinations Jn(x) ± Nn(x) will behave like e±(x - nπ/2 - π/4), which is very convenient. These combinations are called the Hankel functions.

If we now let x = ix, Bessel's equation becomes y" + (1/x)y' + (1 + n2/x2)y = 0, whose solutions are called modified Bessel functions. The one regular at the origin, Jn(ix) = (i)nIn(x), behaves like (-i)n (1/2iπx)1/2ex for large x, the positive exponential completely dominating the negative exponential. In(x) is the modified Bessel function of the first kind. We now need a linearly independent second solution that behaves like e-x when x is large to complement this solution. To find one, we must wipe out the positive exponential completely, and we can do this by starting with the Hankel function that behaves like eix and then letting x → ix. To simplify the result, we multiply by (π/2)in+1 and call the function Kn(x), which for large x behaves like √(π/2x) e-x. When you see these various Bessel functions defined in a math text, the motivation for the strange choices of constants is seldom clear. The reason is to give certain simple behaviors for large x, or asymptotically.

For small x, K0(x) → ln (1/x) + ln2 + γ, where γ is Euler's Constant, 0.57721 56649 ... . Euler's Constant is the limit as m → ∞ of 1 + 1/2 + 1/3 + 1/4 + ... + 1/m - ln m. The series is the divergent series used as an example of the fact that a series may not converge even if its terms approach zero. If the signs alternate, its sum is ln 2. Here it appears in getting the ln (1/x) from the series. The product xK0(x) → 0 as x → 0, which can be proved by l'Hôpital's Rule and writing it as ln (1/x)/(1/x). Kn(x) goes as [(n - 1)!/2](2/x)n for small x, so xK1(x) → 1 as x → 0. These are the only two functions we will need, and it is only necessary to remember that they go to 0 and 1, respectively.

At x = 1, we have xK0(x) = 0.4210 and xK1(x) = 0.6019. At x = 2, xK0(x) = 0.2278 and xK1(x) = 0.2797. For larger values of x, both functions decrease exponentially.

If ' stands for differentiation with respect to the argument, then K0' = -K1 and K1' = -K0 - K1/x. From these, it is easy to work out (xK0)' = K0 - xK1 and (xK1)' = -xK0. Each of these relations can be turned into an indefinite integral. Furthermore, (xK1K0)' = -x(K02 + K12) and [x2(K12 - K02)]' = -2xK02. These will help us to integrate squares and products of the K's.

The Fourier cosine transform is F(y) = ∫(0,∞)f(x)cos(xy)dx. If f(x) is an even function of x, it is easy to turn the cosine transform into an exponential transform by expressing the cosine in terms of exponentials, and letting x → -x in the second integral. From Erdélyi, et. al. we find that the cosine transform of (1 + x2)-n-1/2 is (y/2)n(√π/Γ(n+1/2))Kn(y). From this, it is not hard to show that ∫(-∞,+∞)eiωxdx/(1 + x2)1/2 = 2K0(ω), and ∫(-∞,+∞)eiωxdx/(1 + x2)3/2 = 2ωK1(ω), since Γ(1/2) = √π and Γ(3/2) = (1/2)Γ(1/2). We'll see that these are the Fourier transforms of the fields of a moving charge.

Another integral yielding K1(ω) can be found by differentiating the integral for K0 with respect to ω, and using the relation K0'(ω) = -K1(ω). This integral is ∫xeiωx/(1 + x2)1/2 = 2iK1(ω). This integral can also be obtained from a sine transform.

A typical energy loss encounter between a particle of charge +ze moving with velocity v and an electron of charge -e at rest is shown at the right. The impact parameter is b. Any analysis in which we use an impact parameter is a classical one, of course. The relativistically correct electric field components E1 and E2 are shown. The field points away from the present position of the charge, but is not isotropically distributed. The field component E1 alternates in direction as the charge passes by, but E2 is constant in direction.

As a first approximation, we suppose that the charge ze moves by before the electron can move any appreciable distance. The electron receives an impulse ∫(-∞,+∞)(-eE2)dt that is equal to the momentum Δp transferred to it. The heavy particle ze also experiences an equal and opposite impulse that deflects it a little, but only a very little, which can be neglected in calculating the energy transfer. The energy transfer to the electron is ΔE = Δp2/2m, and the energy of the particle ze decreases by the same, relatively very small, amount. It would be much more difficult to estimate directly the force on the particle ze, involving the exact trajectories of both particles, but we use conservation of energy, trusting that it will give the correct answer.

The integral is very easy to perform, and the result is Δp = 2ze2/bv. From this, ΔE = 2z2e4/mv2b2. Let's suppose there are N atoms per cm3, each with Z electrons, which we assume are distributed randomly and evenly. The number of electrons in a thickness dx with impact parameters between b and b + db is then dn = 2πNZb db dx. Each of these electrons will experience the same energy transfer ΔE. The total energy transfer in a distance dx will be -dE/dx = 2πNZ ∫ΔE(b)bdb. With the expression for ΔE just derived, -dE/dx = (4πNZz2e4/mv2)∫db/b = (4πNZz2e4/mv2)ln(bmax/bmin). We already have the form of the expression for the stopping power -dE/dx that we shall see over and over.

We cannot simply let bmax = ∞ and bmin = 0, for then the stopping power would diverge. Getting a useful formula depends on making good choices of these parameters. Since they appear in a logarithmic term, order-of-magnitude estimates may be good enough. The energy transfer increases without limit as b → 0, but we know that there is a certain maximum energy transfer that occurs in a head-on collision. We should choose bmin so that this impact parameter corresponds to the maximum energy transfer.

A maximum energy transfer collision, b = 0, is shown in the figure. In the approximate CM system, the electron simply bounces off elastically, its velocity reversed. Returning to the Lab system by adding the velocity v, and using the correct relativistic formula for adding parallel velocities, we find that the electron now has a velocity v' = 2v/(1 + β2). Now, 1/γ'2 = 1 - β'2 = (1 - β2)/(1 + β2). The kinetic energy of the electron is T = (γ' - 1)mc2 = 2mβ2c2γ2 = 2γ2mv2, which is the maximum energy transfer.

Setting ΔE equal to this value, we find bmin = ze2/γmv2. A more exact derivation would show that replacement of b2 by b2 + bmin2 would have the same effect, with the lower limit of integration b = 0.

As b becomes larger and larger, the encounter becomes softer and longer. If the time of the collision is on the order of the frequencies of motion of the electron, we know that the orbits are only perturbed adiabatically, and no change takes place. If we set the period of the motions 1/ω equal to the time of the collision b/γv (the time during which the fields are appreciable), we find that this happens when b/γv = 1/ω or bmax = γv/ω. This is a very rough estimate, and we soon shall do very much better. However, this gives the stopping power -dE/dx = (4πNZz2e4/mv2)ln[(γv/ω)/(ze2/γmv2)] = (4πNZz2e4/mv2)ln(γ2mv3/ze2ω).

The argument of the logarithm is usually denoted B. The formula obtained by Bohr in 1915 is almost exactly the same, except that B is multiplied by a factor 1.123 that has negligible effect, and the meaning of the frequency ω is a bit clearer. For relatively slow, heavy particles that can be treated classically with fair accuracy, it agrees well with experiment. For lighter particles, and for fast particles, it overestimates the energy loss.

Before taking this matter up, we shall handle the energy transfer to a harmonically bound charge in a better manner, and also obtain some useful results for the fields. Let's assume, as in Lorentz's electron theory, that an electron in the atom has an equation of motion x" + Γx' + ωo2x = -(e/m)E(t). The easiest way to solve this equation is to use Fourier transforms, which will reduce it to an algebraic equation that is easily solved. Following Jackson, we'll use the transforms in the form x(t) = (2π)-1/2∫(-∞,+∞)x(ω)e-iωtdω. The inverse transform is the same, except that the sign of the exponent is changed. We transform both the position and the field.

The result is x(ω) = (-e/m)E(ω)(ωo2 - iωΓ - ω2)-1, which should be familiar. If Γ is small, the resonance is sharp. What we want is the energy transfer, which should be given by ΔE = -e∫(-∞,+∞)v·Edt, where v = -iωx, and E is the field at the electron. We now write this expression using the Fourier transforms.

ΔE = -(e/2π)∫dω∫dω'∫e-i(ω + ω')dt (-iω)x(ω)·E(ω'). All limits are from -∞ to +∞. The integral over dt gives δ(ω + ω'), and this delta function can be used to integrate over dω'. The result is ΔE = -e∫dω(-iω)x(ω)·E(-ω). Since E is real, E(-ω) = E*(ω) (the reality condition), so ΔE = -e∫dω(-iω)x·E*(ω). Now split this integral into two integrals, one from -∞ to 0, and the other from 0 to ∞. Let ω = -ω in the first integral, and use the reality conditions again on both x and E. The result can be written most simply as ΔE = 2e Re{∫(0,∞)iωx(ω)·E*(ω)dω}. This is our equation for the energy transfer in terms of the Fourier transforms.

We already have the transform x(ω). Now we must find the transforms of the fields E1(t) and E2(t) at the position of the electron. These fields were given as a function of time in the first diagram above. All we have to do is evaluate the integrals E(ω) = (2π)-1/2∫E(t)eiωtdt. For E2, we have E2(ω) = zebγ(2π)-1/2∫(b2 + γ2v2t2)-3/2eiωtdt. If u = γvt/b, this integral is easily expressed as (ze/bv)(2π)-1/2∫(1 + u2)-3/2exp[i(ωb/γv)u]du. We have already shown the value of this integral in our discussion of the Bessel functions. Therefore, E2(ω) = (ze/bv)√(2/π)[ξK1(ξ)], where ξ = ωb/γv.

The transform of E1, with the same substitution, can be expressed as E1(ω) = -(ze/γbv)(2π)-1/2∫ (1 + u2)-3/2exp[i(ωb/γv)u]udu. This can be integrated by parts, taking w = exp[i(ωb/γv)u], dv = -udu(1 + u2)-3/2. Then dw = i(ωb/γv)ex[[i(ωb/γv)u]du and v = (1 + u2)-1/2. The integrated part wv vanishes at both limits, so E1(ω) = -i(zeω/γ2v2)(2π)-1/2∫ (1 + u2)-1/2exp[i(ωb/γv)u]du. This integral was also given above, so the final result is E1(ω) = -i(ze/bvγ)√(2/π)[ξK0(ξ)], where ξ has the same meaning as in the preceding paragraph. We now have the Fourier transforms of the fields.

Now we go back and substitute x(ω) in the integral for ΔE. The result is ΔE = -(2e2/m) Re{∫ |E(ω)|2o2 - ω2 - iωΓ)-1. The real part of everything inside the integral except the square of the field is -ω2Γ/[(ωo2 - ω2)2 + ω2Γ2]. This function is strongly peaked at ω = ωo. We make the approximation that ω+ωo = 2ωo, and put x = (ω - ωo)/Γ. Then, ΔE = (2e2/m)|Eo)|2∫(-∞,+∞)dx/(1 + 4x2). The limits of the integral have been extended to infinity with little error. Its value is π/2, so finally ΔE = (πe2/m)|Eo)|2. Since we already have the Fourier transforms of the field, we can find ΔE for any value of ξ = ωb/γv; that is, as a function of the impact parameter b. This has been quite a long pull, but now we do not need to estimate bmax because we have an exact expression for ΔE(b).

Indeed, |Eo)|2 = (2/π)(ze/bv)2ξ2[K1(ξ)2 + K0(ξ)22], where ξ = ωob/γv. The stopping power is given by the same integral we used long ago, -dE/dx = 2πNZ ∫ΔE(b)bdb. However, we now introduce a new refinement. So far we have discussed only one resonant frequency ωo that corresponds to one electron. We now consider a more correct model of the atom.

Spectroscopy showed that atoms did not contain harmonically bound electrons providing a set of fundamental frequencies and their harmonics. Instead, the spectrum consisted of many discrete lines of apparently unrelated frequencies, and of a wide range of intensities. This was clarified by Bohr's recognition that the observed frequencies ν were related to differences in energy levels through hν = W(a) - W(b), where a is the initial state and b the final state of the transition. It was found that these frequencies could be thought of as corresponding to an oscillating electron with natural frequency ν, provided that the contributions of the oscillators to, for example, the emission of radiation, were those of the classical oscillator times a constant less than unity called the oscillator strength, f(a,b).

The intensity of a spectral line is proportional to the absolute value squared of a matrix element of the dipole moment, -e(a|x|b), where x is a vector and a,b are the initial and final states. A state is characterized by its total angular momentum j, and there are 2j+1 states of the same energy that correspond to the same values of the other quantum numbers. The rate of radiation in erg/s is (64π4ce2σ4/3) Σ|(a|x|b)|2, where σ is the wavenumber, 1/λ, in cm-1. The sum is over all the final states and an average over all the initial states making up the line; it is called the strength S(a,b) of the line.

The oscillator strength of the line is defined as f(a,b) = (8π2m/3e2h)ν(a,b)S(a,b)/(2j+1), where ν(a,b) = [W(a) - W(b)]/h, so that ν(b,a) = -ν(a,b). The reason for the choice of the constant factor will be clear shortly. The momentum and position operators satisfy the commutation relation pr - rp = -3ih', where h' = h/2π, and p, r are vectors, and their product is the scalar product. The Hamiltonian for one electron is H = p2/2m + V(r), so the commutation relation gives us p = (-im/h')(rH - Hr). Taking matrix elements between states a and b, and summing over third states c in evaluating the products rH and Hr, while remembering that H is diagonal, we get (a|p|b) = -2πimν(a,b)(a|x|b). Then, the diagonal matrix components of pr - rp are 4πmΣ(b)ν(a,b)|(a|x|b)|2, and also 3h', since the commutator is a pure number and (a|a) = 1. If we sum over final states by multiplying by 2j+1, we have 4πm(3e2)h/8π2m)Σf(a,b) = 3h/2π This simplifies, miraculously, to Σf(a,b) = 1. On the other hand, if we did not know the constant factor in the definition of f(a,b), we could determine it to make this true (which, of course, was how it was originally done).

The condition Σf(a,b) = 1 is called the Thomas-Kuhn Sum Rule (1925), and means that the effects of all the spectral frequencies characteristic of the atom do in fact add up to one electron overall. Jackson says it is "obvious," but I think it is a rather clever and non-obvious result that allows us to use the classical electron theory with quantum-mechanical atoms. If the atom contains Z electrons in all, then Σf(a,b) = Z, which is probably obvious.

In the preceding section we found the energy transfer ΔE for a particular frequency ωo. Now we consider the atom in the light of oscillator strengths, and conclude that the total energy transfer is Σf(a,b)ΔE(a,b), where we sum over the transitions a,b. Now we have -dE/dx = 2πNΣf(j)∫ΔE(j,b)bdb, where we have replaced a,b with j to avoid confusion with the impact parameter b.

Putting things together, we now have ΔE(j,b) = (πe2/m)|E(b,ωj)|2 = (2e2/m)(ze2/bv)ξ2[K12 + (1 - β2)K0(ξ)], where ξ = ωjb/γv, and 1/γ2 has been replaced by 1 - β2. When this is inserted into the integral, and the variable of integration changed to ξ, one power of ξ disappears and the integrals can be done as shown in the section on Bessel functions. This is quite straightforward, and the result can be written down at once.

-dE/dx = (4πNz2e4/mv2)Σ{ξK1(ξ)K0(ξ) - (β2/2)ξ2[K12(ξ) - K02(ξ)]}, where ξ = ωjbmin/γv. We see our familiar expression for the stopping power reappearing once more. If ξ << dx =" (4πNz2e4/mv2)Σf(j)[ln(γv/ωjbmin) - ln 2 + 0.5772... - (β2/2)]. Let Σf(j)ln ωj = Z ln Ω, where Ω is an average frequency. then, remembering that Σf(j) = 1, we have -dE/dx = (4πNZz2e4/mv2)[ln (1.123γ2mv3/ze2Ω) - β2/2. The factor 1.123 is 2/e0.5772... = 2/1.7810. This is Bohr's famous 1915 result, which here appears as a limit of the more general result using oscillator strengths. The constant Ω was usually determined empirically in using Bohr's formula practically.

Quantum considerations may affect our results both at large and small impact parameters. We have defined bmax as corresponding to ξ = 1, or bmax = γv/ω. The energy loss at this impact parameter is ΔE = (2z2e4ω2 / γ2mv4)[K12(1) + (1 - β2)K02(1)]. We can express this in a more transparent form by using vo = e2/h' = 2.187 x 108 cm/s, the velocity in the first Bohr orbit, and I = me4/2h'2, the ionization energy of hydrogen, 13.6 eV. The result is ΔE = (z/γ)2(vo/v)4 [(h'ω)2/I][.5395 - 0.17724β2]. Since usually v >> vo, and h'ω <>

We know that energy transfers with an atom or molecule are probable only when the energy transfer approximates the energy difference between two stationary states, so an arbitrarily small amount of energy cannot be absorbed. In spite of this, it is found that the classical result is correct on the average. Most encounters result in no energy transfer, but occasionally one occurs with a large energy transfer, so the average over collisions is the small energy calculated classically. This is encouraged by the use of the oscillator strengths defined quantum-mechanically, which we have explained above. It should be clear that the energy transfer is not necessarily anything like h'ω, but can be much larger at small impact parameters. Therefore, our energy loss formula needs no significant correction at large impact parameters.

Conditions are quite different at small impact parameters. We have used a bmin that gave a ΔE equal to the maximum energy transfer. In quantum mechanics, we cannot specify an impact parameter that is less than the distance given by the uncertainty relation, Δx = h'/p. This is certainly a rough and unsubstantiated estimate, but it turns out to give the correct answer. We have a classical bmin,c = ze2/γmv2, and a quantum bmin,q = h'/γmv, and must use whichever one is the larger. The ratio η = bmin,c/bmin,q = ze2/h'v can tell us which to use. For η > 1, the classical value should be used; for η <>

To treat this problem quantum-mechanically, we consider the collision in the CM system. That's the reason we used the electron mass m in the preceding paragraph. Partial-wave analysis can be used to solve the problem in the CM system, and then the energy transfer can be found by transforming back into the lab system. Bethe did this in 1930, obtaining the most-used stopping power formula, -dE/dx = (4πNZz2e4/mv2) [ln(2γ2mv2/h'Ω) - β2]. If we simply take the ratio bmax/bmin,q as the argument of the logarithm, we find that it is γ2mv2/h'Ω, the same except for the factor 2.

For electrons, the CM system is different from that for a massive incident particle. Then, bmin = (h'/mc)√[2/(γ - 1)], and the argument of the logarithm becomes (γ - 1)√[(γ + 1)/2](mc2/h'Ω) instead.

For either Bohr's or Bethe's formula, the stopping power decreases rapidly as v increases, then passes through a minimum and afterwards increases slowly for ultrarelativistic particles, due to the γ2 in the argument of the logarithm. One factor of γ comes from the increased maximum energy transfer in a hard collision, while the other comes from the greater effective range of the compressed fields of a fast particle. However, it is found that the energy loss of very fast particles is less than that predicted by the stopping-power formulas. The reason is in the shielding due to the dielectric constant of the medium, which reduces the fields and so the energy loss. This is called the density effect, because it is largest in dense media.

Now we make another digression, but one that will not only allow us to explain the density effect in reducing the stopping power, but also to consider some other phenomena associated with the passage of fast charged particles through matter, such as Cherenkov radiation. In another article, Relativistic Electrodynamics, we used Maxwell's equations to find the field of a moving charged particle in a fully covariant form. In that case, the medium was vacuum. Now we wish to find the electromagnetic fields of a charged particle in a polarizable medium. Our result will be relativistically correct, of course, since it will be based on Maxwell's equations, but not manifestly covariant.

Here is what we are going to do. First, Maxwell's equations will be written with D, and Fourier transformed from functions of x,t to functions of k,ω, eliminating the space and time derivatives. Then D will be set equal to ε(ω)E. Allowing ε to be a function of ω allows a more realistic characterization of the medium. This will be enough for our purposes; we are not attempting to characterize the medium accurately, and shall be satisfied with the general behavior. This is not the same thing as Fourier transforming ε(t)E(t), incidentally. We shall also set H = B, since they are the same in a nonpermeable medium.

Now we will introduce the scalar potential φ and the vector potential A, from which the fields can be found by differentiation. As is well-known, this identically satisfies the homogeneous Maxwell equations, and uncouples the others, so that we will have equations for φ and A individually. We do this in the frequency domain, although it can also be done in the time domain, of course. The equations satisfied in the frequency domain by φ and A are algebraic equations equivalent to the inhomogeneous wave equations of the time domain. The potentials must be related by the Lorentz condition for this to occur. This is a Lorentz-invariant condition that is equivalent to requiring the fields to be transverse.

The sources proper for a moving point charge will then be introduced into the equations for the Fourier transforms of the potentials, which will give us the Fourier transforms of the field. We already know how to find the stopping power from the field transforms, so all we have to do is carry out the algebra as we have done above with the simpler fields of the vacuum. Now let's do the work.

The figure at the right shows how the Fourier transformed inhomogeneous wave equations are derived. This is a straightforward procedure, and everything needed is in the diagram. In the time domain, the moving point charge is represented by ρ(x,t) = zeδ(x - vt) and J = vρ(x,t). Of course, x, v and J are vectors. In the frequency domain, ρ(k,ω) = (ze/2π)δ(ω - k·v), and A(k,ω) = vρ(k,ω). To find these transforms, simply use the delta function to do the time integration, and remember that the transform of the delta function is unity. We can use these source transforms to find the transforms of the potentials. The transforms of the fields can be expressed in terms of φ(k,ω) alone.

Now we can invert the space part of the transform to get back into ordinary space, while leaving us in the time domain for the frequency part. We assume the particle is moving on the x-axis, and we want the field at (0,b,0), where the electron is. The electric field is then E(x,ω) = (2π)-3/2∫d3kE (k,ω)eibk2. This integral can be evaluated, but the algebra is tedious. However, it is important to see how it is done.

Writing it out in detail, E1(ω) = (2ize/ε)(2π)-3/2∫d3k exp(ibk2)[(ωεv/c2) - k1] [k-2 - ε(ω/c)2]-1δ(ω - vk1). The delta function can be used to integrate over k1, remembering to divide by v by the rule for a delta function of a function. The wave equation denominator becomes (ω/v)2 + k22 + k32 - ε(ω/c) 2. If we introduce λ2 = (ω/v)2 - ε(ω/c)2 = (ω/v)2(1 - β2ε). If βε <> 1, then λ will be pure imaginary, with important effects on the fields. When ε = 1, λ = ω/γv = 1/bmax. Think of λ as the reciprocal of the range of the fields in the polarizable medium.

It is now possible to integrate over k3. The integral is of the form ∫dx/(a2 + x2), and has the value π/(λ2 + k32)-1/2. Finally, the integral over k2 is the one that gives us K0. The final expression is E1(ω) = (-izεω/v2) (2π)-1/2[1/ε - β2]K0(λb). In all this, remember that ε is a function of ω. When ε → 1, we find the same field as before.

The y-component is E2(k,ω) = (-2izek2/ε)δ(ω - k·v)/[k2 - (ω/c)2ε]. The integrals over k1 and k3 are done exactly as for E1. If we let x = k2/λ, then E2(b,ω) = (-izeλ/vε)√(2π) ∫xeibλx/(1 + x2)1/2. The integral is the one given in the last paragraph of the section on Bessel functions, and has the value 2iK1(bλ). The final result is E2(b,ω) = (ze/v)√(2/π)(λ/ε)K1(λb).

Our earlier formula for the energy loss to a single electron oscillator can be generalized using oscillator strengths, which will be more accurate than applying oscillator strengths to an approximation at the end. The generalized formula is ΔE(b) = 2eΣf(j)Re∫(0,∞) iωxj(ω)·E(ω). Instead of using x(ω) from the equation of motion, we can find the equivalent using ε(ω). Now ε(ω)E = E + 4πNp, where p is the dipole moment induced in the atom or molecule by the electric field. This moment is -eΣf(j)xj = (ε - 1)E/4πN, which we can put into the integral for the energy loss. The result is ΔE(b) = (1/2πN) Re∫(0,∞)-iωε(ω) |E(ω)|2 dω. The contribution due to the 1 in ε - 1 vanished because it is pure imaginary. Energy transfer depends on there being an imaginary part of ε, so the integral has a real part.

The stopping power is, as before -dE/dx = 2πN∫(a,∞)ΔE(b)bdb, where a is a cutoff distance like bmin. Combining the energy loss and this equation, we find Fermi's result: (-dE/dx)(b > a) = (ze/v)2(2/π) Re∫(0,∞) iωλ* a K1(λ*a)K0(λa) (1/ε - β2)dω. In order to evaluate this formula, the behavior of ε(ω) must be known. With reasonable assumptions, it gives the same results as we have already obtained, but it also includes the density effect that was not present in the previous formulas.

We won't go into the rather complicated details of evaluating the density effect in general, but only state the results for the extreme relativistic case, where it is most important. The relativistic limit of Fermi's equation is -dE/dx = (zeωp/c)2ln(1.123c/aωp), where ωp is the plasma frequency, [4πNZe2/m]1/2. Note that there is no power of γ in the argument of the logarithm, so the energy loss does not increase as rapidly with energy. Also, the energy loss does not depend on the details of atomic structure, only on the total number of electrons present. The stopping power without the density effect can be obtained from the usual classical expression (Bohr's) by putting v = c. The result is -dE/dx = (zeωp/c)2[ln(1.123γc/aΩ) - 1/2]. The difference, for very relativistic particles is Δ(dE/dx) = (zeωp/c)2[ln(γωp/Ω) - 1/2], which increases as ln γ.

Cherenkov (usually spelled Cerenkov in English, but I agree with Jackson in preserving the sound in the spelling) radiation is named after P. A. Cherenkov, who reported it in 1934. It is radiation emitted by a charge moving faster than the local speed of light c/n in a medium. The explanation of it was given by I. Frank and I. Tamm in 1937, who calculated its spectrum. It is a weak radiation, but of considerable interest. We can give a good account of it from the fields we found in the density effect.

Cherenkov radiation is easily observed in the cooling ponds of nuclear reactors. Spent fuel rods are surrounded by a bluish-white glow in the water, giving visual evidence of their strong activity. These days it is difficult to get near enough to a reactor to see this, but I can assure you it is impressive and eerie. The glow comes from fast electrons and positrons created by energetic gamma rays absorbed in the water. Homer Simpson's uranium rods exhibit a greenish glow, but nuclear materials do not glow at all normally, and the reputation for luminosity is unfounded.

Cherenkov radiation also occurs in nature: it comes from the sky. Cosmic rays of TeV energies (1012 eV) produce showers of fast particles in the atmosphere, and beneath these showers there is a nanosecond burst of Cherenkov radiation a hundred yards or so in diameter. The chief problem is separating the Cherenkov radiation from the background. TeV gamma rays are not influenced by electric fields, so they preserve their direction of approach. Cherenkov radiation from gamma showers can locate the direction to within a degree or so. One source was found to be the Crab Nebula pulsar, a neutron star with a rotational period of 33 ms that emits energetic gamma rays. Cherenkov counters are useful in nuclear and particle physics.

We are now interested in the fields at large distances, for bλ large instead of small, as in the density effect. The fields we need are easily obtained from our previous expressions by using the asymptotic forms of the Bessel functions. They are: E1(b,ω) = (izeω/c2)(1 - 1/β2ε)e-λb/√(λb), E2(b,ω) = (ze/εv)√(λ/b)e-λb), and B3(b,λ) = βεE2. The energy loss to distances greater than a can be calculated by Poynting's Theorem as well as by our previous method. The formula that results is -dE/dx = -ca Re∫(0,∞)B3*E1 dω. If we insert the Fourier transforms for the fields, the energy loss turns out to be -dE/dx = (ze/c)2 Re[-i√(λ*/λ)]∫ω (1 - 1/β2ε)e-(λ + λ*)adω.

Now, λ = (ω/v)√(1 - β2ε), so if β2ε > 1, λ is pure imaginary, say λ = -iκ. In this case, λ*/λ = -1 and λ + λ* = 0, so we have -dE/dx = (ze/c)2∫ω(1 - 1/β2ε)dω, where the integral is over the frequencies for which β2ε is greater than unity. This expression no longer contains the distance a, so the energy is lost to infinite distance; that is, it is radiated. The integrand is the spectrum of the Cherenkov radiation, Frank and Tamm's result. The shaded area in the diagram at the left illustrates the spectrum in the case of a resonance in the ultraviolet, the usual case in most transparent media. It is easy to see why Cherenkov radiation in water is blue.

The electric field is transverse to the distance from a point on the path to the observer. This distance makes an angle θ with the path, and tan θ = -E1/E2. Using the expression for λ, it is not hard to show that tan θ = √(β2ε - 1), or cos θ = 1/β√ε = c/nv, where n is the index of refraction. The wavefronts are cones whose normals make this angle with the path of the particle. The radiation is polarized in the plane defined by the path and the observation point. For water, n = 4/3, so Cherenkov radiation is emitted by particles moving faster than 0.75c, or 2.25 x 1010 cm/s. For electrons, the kinetic energy must be greater than 261 keV for the emission of Cherenkov radiation. Of course, electrons of this energy will not get far in water. Fast particles emit the radiation more and more perpendicularly to their paths. Geometrically, it's like the shock wave from a supersonic projectile.

A Cherenkov telescope is shown in the figure. Three photomultipliers and three plane mirrors are arranged at 120° around the axis. The baffle stops radiation that is not emitted around the angle θ. When the three photomultiplers have an equal output, the axis of the telescope is directed toward the source. Therefore, the telescope is sensitive to both particle energy and direction.

P. A. M. Dirac showed in 1931 that the existence of a magnetic monopole would imply the quantization of electric charge. A magnetic monopole is a free magnetic charge, that probably would be of two kinds, N and S, and is the source of a magnetic field H just as electric charge is a source of E. Such monopoles have not been found (alas) but speculation about them is interesting.

If there were only magnetic charge, Maxwell's equations would appear as at the left. We can get these equations from the duality transformation E' = -E, B' = B, and similarly for D and H, and changing electric charge to magnetic. The magnetic field H produced by a magnetic charge g is the same as the electric field produced by an electrical charge q. Incidentally, we shall consider H = B and E = D for simplicity, so our monopole exists in a vacuum. The H of a point monopole is the same as the -E of a point charge, even relativistically, so -E1 becomes H1, and -E2 becomes H2. B3 is then βE2, so that βg replaces ze in the expression for the field. When a monopole zips by an electron, the electron will be knocked to the side, not toward the path of the monopole, as for an electric charge.

Dirac showed that consistency required (ge/h'c) = n/2, where n could be ±1, ±2, ... . For a given g, this means that electric charge is quantized in units of e. Taking the known value for e, we can solve the equation for the pole strength g, g = h'c/2e, where we have taken n = 1 to get the value analogous to the electronic charge. From this expression, we find that g = 3.2913 x 10-8 emu. This is quite large compared to e = 4.803 x 1010 esu, and shows that elementary monopoles will interact strongly.

We can use all our equations for the stopping of charged particles if we make the replacement ze = βg. There is a small approximation in that we have no B1 field, but it contributes very little. The charge equivalent to a monopole is then z = βg/e = 1/2α, where α is the fine-structure constant e2/h'c. Numerically, for β = 1, this is z = 68.5! A fast elementary monopole would be stopped like a heavy nucleus. The Bohr or Bethe formulas (let's choose the Bethe formula) would give us -dE/dx = (4πNZe2/m)(ge/c)2[ln(2γ2mv2/h'Ω) - β2). The main thing to notice is that the β in &geta;g has wiped out the velocity from the coefficient and replaced it by c. The only velocity dependence is now in the logarithm, so the variation is small. Energy loss for a monopole increases slowly with γ, and does not show the rapid rise in energy loss at small velocities, which is exactly cancelled by the decrease in the electric field at small velocities. This is certainly one way to distinguish monopoles from charges.

The formulas for Cherenkov radiation cannot be used, however, because they depend on the dielectric constant ε which becomes the permeability in the case of monopoles, which is unity. In a permeable material, there would, of course, be rather strong radiation. The dielectric constant does affect the electric field E3, however. This would require working through the theory again with this assumption, which I have not done. However, it surely has been done, but I do not know the results at present.

Some general characteristics can, however, be predicted. There will be the same Cherenkov cone, but the radiation will be polarized with the magnetic field in the plane containing the path of the monopole and the observer, and the electric field perpendicular to that. This difference in polarization should be easy to detect, and will be a distinct signature of the monopole. Because of the large size of g, the radiation should not be very weak, but quite prominent.

Air is an important stopping medium. The number density N/V at 15°C and 760 mmHg is pNA/RT = 2.5482 x 1019 per cc. N2, with Z = 14, makes up a fraction 0.78, O2, with Z = 16, makes up a fraction 0.21, and A, with Z = 18, makes up 0.01. Therefore, the value of NZ for air is 3.792 x 1020 per cc. The plasma frequency for air is ωp = 1.0985 x 1015 s-1. Hence the factor ωpze/v = 1.0552 x 106 cm/s / v. If we use B = 2γ2mv2/h'Ω for the argument of the logarithm, and estimate h'Ω = 35 eV, γ = 1, we find B = 1.143 x 105E, where E is in MeV. For the fast alphas from Po212, energy 8.78 MeV, and v = 2.05 x 109 cm/s. Hence, ln B = 13.82. Putting this all together, we find that -dE/dx = 3.66 x 10-6 erg/cm, or 2.29 MeV/cm, which is certainly of the correct order of magnitude.

The value of 35 eV that we used for h'Ω is only a very rough guess. It is the average energy loss per ion pair created in air. Not all interactions create an ion pair, of course; most transfer much less energy. The ionization potential of N2 is 15.576 eV, and that of O2 is 12.2 eV. Both molecules dissociate more easily than they ionize, with dissociation energies of 7.373 eV and 5.080 eV, respectively. High-energy particles hardly notice the molecular associations of the atoms.

Electrons have a further energy loss mechanism, the emission of radiation when they are scattered by nuclei. This is called Bremsstrahlung, brake radiation, and is an important energy loss for fast electrons. The formulas for bremsstrahlung loss look very much like the energy-loss formulas in this article, except that they involve the classical electron radius ro, and the maximum impact parameter is determined by shielding of the nuclear charge by the atomic electrons. In fact, it is used in electron synchrotron X-ray sources. Jackson treats bremsstrahlung in his Chapter 15. Cherenkov radiation, incidentally, has negligible stopping effect.

J. D. Jackson, Classical Electrodynamics, 2nd ed. (New York: John Wiley & Sons, 1975). Chapter 13.

I. Kaplan, Nuclear Physics (Cambridge, MA: Addison-Wesley, 1955). Chapter 13.

E. U. Condon and G. H. Shortley, The Theory of Atomic Spectra (Cambridge: Cambridge University Press, 1935). p. 108f. Oscillator strengths and the Thomas-Kuhn sum rule.

M. Abramowitz and I. A. Stegun, eds., Handbook of Mathematical Functions (Washington, DC: USGPO, 1964). pp. 374-379 and pp. 416-429 (numerical tables).

A. Erdélyi, et. al., Tables of Integral Transforms, Vol. I (New York: McGraw-Hill, 1954). p. 11, 1.3(7).

Relativistic Electrodynamics: Field of a Point Charge

Contents

1. Introduction
2. A Uniformly Moving Charge
3. A Charge in Arbitrary Motion
4. Radiation
5. The Electron Synchrotron
6. References

Introduction

Maxwellian electrodynamics is naturally relativistic; that is, it conforms to Einstein's requirement of invariance under transformations between coordinate systems moving with constant relative velocity. It has exactly the same form in any inertial system, and light travels with the same velocity c in any system. The Lorentz transformation of space and time coordinates was established for electromagnetism before relativity was formulated and it was known that this transformation applies to all physical processes. In particular, the time transformation is real, and not just a substitution that preserves the form of the equations. The formalism of relativity makes working with electromagnetism much easier. In this article, I will try to explain as clearly as I can how the electromagnetic field of a point charge in arbitrary motion can be determined at any point in space. A very important application of this theory is the radiation by moving charges.

This material is excellently explained in chapters 11, 12 and 14 of Jackson's classic text, to which the reader is referred. On first sight, the material seems very difficult and unclear, even if it has been mastered before and is now being reviewed after forgetting it. For the beginner, it may be truly frightening and formidable. The beauty of Jackson's exposition is that when one digs in and tries to understand each statement, the subject becomes continually clearer and in a few hours its elegance shines out, and it is seen to be logical and satisfying. This is the evidence of the author's comprehensive understanding achieved through labor and thought, that he communicates to you. In any case where this does not happen, the fault is more often with the author than with the reader, and particularly if the subject is a false one. Relativistic electrodynamics is one of the best-known and verified physical theories, the work of giants. Even the classical theory, as presented here for the model of a point charge, is of wide applicability. I hope I can make some points clearer in this article, and help the reader to understand.

The reader will need some preparation in relativity, but that is not difficult to acquire, even with Jackson alone. I will use the index notation for tensors with Greek indices taking the values 0,1,2,3 for the coordinates ct, x, y, z. A sum is implied over the same index in contravariant (high) and covariant (low) positions. This notation is very much like that of Euclidean tensors in 3-dimensional space, except that the distinction between contravariant and covariant is necessary because the metric tensor gαβ is diagonal with g00 = 1, g11 = g22 = g33 = -1. Changing from contravariant to covariant means only changing the sign of the space components. Then, if Aα = (A0, A) [In expressions like this, subscripts are not tensor indices, but simply labels.] Then, A · B = AαBα = A0B0 - A·B.

If xα is the coordinate four-vector (ct,r), then x · x = (ct)2 - r2 = s2 is a scalar, and hence an invariant, under Lorentz transformation. It is the invariant interval between the event X described by the vector, and the event O at the origin, t = 0 and r = 0. If s2 > 0, then a Lorentz transformation to a coordinate system where X and O occur at the same point, but at a later or earlier time, is possible. If s2 < 2 =" r2," r =" ±ct." dt =" γdτ," t =" 0." t =" γt'" t =" t'/γ." t =" γt'" r2 =" b2" 1 =" β2γ2," r =" γ[(b2" n =" (-β)[(vt" r =" (b2" ex =" Ex'," ey =" γEy'," bz =" βEy'," ex =" -γqvt/(b2" ey =" qb/(b2" bz =" βEy." v2t2 =" b2" v2t2 =" r2" v2t2 =" r2[1" e =" qr/r3γ2(1" e =" γq/r2" bz =" βEy," t =" 2b/γv." dl =" 4πI/c," x =" 0" i =" q/(2b/γv)" x0 =" ct." z =" x" k =" -(2π)-4∫d3k" ko =" ±κ,"> 0, and zo = cto <> 0, then we can close the contour below. The large semicircle contributes nothing to the integral, so the integral will be equal to 2πi time the sum of the residues at the two poles, which are exp(-iκzo)/2κ and -exp(iκzo)/2κ. This gives D(z) = θ(zo)(2π)-3∫d3k exp(ik·z)sin κzo/κ. The theta function θ(zo) is zero when its argument is negative, +1 when its argument is positive. d3k = κ2dk sinθdθdφ, so we can do the integral over the angles right away. After a little manipulation, we find D(z) = [θ(zo)/2π2]∫(0,∞)sinκr sinκzodκ. Expressing the product of the sines as the sum of the cosines of the sum and difference angles, writing them as exponentials and combining terms to extend the limits of integration from -∞ to +∞, and using the Fourier transform of the delta function, we have D(z) = [θ(zo)/4πr]δ(r - zo), where r is the distance between source and field points. The delta function forces evaluation at the retarded time t' = t - r/c. This solution has all we could desire for the effects of a disturbance from a impulse at the origin at t = 0 at the point z. The Green's function with these properties is called the retarded Green's function, or propagator. It is, of course, of general utility, and represents another theoretical triumph.

We can express the propagator in manifestly invariant terms by considering δ[(x - x')2]. As we know, δ[f(x)] = δ(x - x')/|f'(x')|, where x' is a root of f(x). Now δ[(x - x')2] = δ[(xo - xo')2 - |x - x'|2] = δ[(xo - xo' - r)(xo - xo' + r)] = (1/2r)[d(xo - xo' - r) + δ(xo - xo' + r)]. Since the theta functions selects only one of the delta functions in the sum, we can write D(x - x') = (1/2π)θ(xo - xo')δ[(x - x')2]. The theta function is a Lorentz invariant, since the past and future light cones are invariantly separated, and the argument of the delta function is also a Lorentz invariant. Therefore, the propagator has been expressed in a manifestly invariant form independent of a coordinate system.

The solution to our problem is now easily obtained. Omitting any solution of the homogeneous wave equation (which would be fields not due to our moving charge), we have Aα(x) = (4π/c)∫d4x'D(x - x')Jα(x'). If you substitute this expression in the inhomogeneous wave equation, you will find that the equation is satisfied. This is the promised solution.

Now we must find Jα(x'), the source current. In some coordinate system, suppose the path of the charge is r = r(t). If the point charge is q, then the charge density is ρ = qδ[x - r(t)] and the current density is ρv = qvδ[x - r(t)]. The current density Jα = (ρc, ρv) will not be expressed in a relativistically invariant way, and so our potential Aα will not be relativistically invariant. To escape from this inconvenience, we can start by parametrizing the trajectory in terms of the proper time of the charge τ instead of t. Since dτ = dt/γ, this is not hard to do by integration. Now, r(τ) specifies an invariant world line that does not depend on the coordinate system. Of course, we need a coordinate system to specify it in any particular case, but we can imagine a world line drawn in 4-dimensional space, labelled with the proper time, that is invariant, and which determines the functions r(τ) in any coordinate system.

For Jα(ct,x), we want to pick out the time t that corresponds to any point x(τ) on the world line. This can be done with a delta function δ[t - ct(τ)], where t(τ) returns the time corresponding to r(τ), and is the 0-component of the position 4-vector. Allowing a choice from any value of τ, we are led to the integral Jα = ec∫dτUα(τ)δ[x - r(τ)], where Uα = (γc, γv) is the tensor 4-velocity of the charge. The delta function is a 4-dimensional one, and r(t) is the tensor 4-position of the charge. This expression is relativistically invariant, since it contains only Lorentz tensors. If it reduces to the correct charge density and current in a particular coordinate system, then it is what we want.

Choose a coordinate system, then, and do the integral over τ using the functions that result. From the rule for the delta function of a function, integration over the time delta function sets t equal to t' and divides by γc. Then, J0 = qc(γc)(1/γc)δ[x - r(t)] = qδ[x - r(t)], exactly what is wanted. The space components of Jα also are easily seen to give the desired results. We are now confident about our invariant expression for Jα.

Now all we have to do is substitute D(x - x') and Jα in the integral for Aα to find an invariant expression for Aα. The integrals are done using the delta functions, and we find that Aα(x) = qUα(τ)/Vβ[x - r(t)]β, evaluated at τ = τo, the retarded proper time. The most important property of this expression is that it is manifestly covariant, constructed only of Lorentz tensors. The fields can now be found by differentiation, but this is not a simple process, and we shall not do it here, but only quote the results. It is simpler, in fact, to differentiate before the final integration so the retardation is easier to handle, but we won't do that either. See Jackson if you need the expressions.

The time τo comes from the condition [x - rτo]2 = 0 enforced by the space part of the 4-dimensional delta function in the integral. This means that xo - ro(τo) = |x - r(τo)| = R, the retarded distance. Further, U·(x - r) = γcR - γv·nR = γcR(1 - β·n), with the symbols meaning the same as in the section on the uniformly moving charge, and where this expression will be recognized. From the expression for Aα, we then find the components φ = [q/(1 - β·n)]ret and A = [qβ/(1 - β·n)]ret. These potentials are expressed in terms of the retarded position and time, as we defined in connection with the uniformly moving charge. They are called the Liénard-Wiechert potentials, which were found in 1898 on the basis of Maxwell's equations alone, without the help of relativity.

The electric field is E = q{(n-β)/[γ2R2(1 - β·n)3]}ret + (q/c){n x [(n - β)x(dβ/dt)]/[R(1 - β·n)3]}, and the magnetic field is B = [n x E]ret. The first term is independent of the acceleration, and falls off as R-2, so it resembles a static field. The fields of the second term fall off as R-1, so they are radiation fields that transport energy to large distances, and are smaller by a factor of c-1.

Let's check our previous results for a charge in uniform motion. The y component of the field will be Ey = q cosθ/γ2R2(1 - β·n)3. Multiplying top and bottom by R, and using γR(1 - β·n) = (b2 + γ2v2t2)1/2, we find our previous result at once, Ey = qγb/(b2 + γ2v2t2)3/2. Ex = -q(cosθ + β)/γ 2R2(1 - β·n)3. Again multiplying top and bottom by R, and using R cosθ + βR = vt, we find Ex = -qγvt/(b2 + γ2v2t2)3/2. This confirms our earlier procedure of Lorentz-transforming the static fields.
Radiation

When β is much less than unity, the fields are very closely E = qn/R2 + n x (n x dβ/dt)(e/cR), where n is a unit vector from the retarded position at retarded distance R. If distances are not large, R is approximately the present distance r. This is quite a different limit from the unrealistic c → ∞, where the radiation field also vanishes. In the radiation field, the electric vector lies in the plane of dv/dt = a and n and is perpendicular to the radius vector. If the angle between a and n is θ then the magnitude of the electric field is E = (qa/c2r)sinθ. The energy flux per unit solid angle in the direction of n will be dP/dΩ = r2c|E|2/4π = (q2/4πc3)a2sin2θ. The total power radiated is found by integration over 4π steradians. ∫2πsin3θdθ = 8π/3, so P = (2e2/3c3)a2. This is Larmor's result for the radiation from an accelerated charge.

Jackson shows how to find the power radiated by a relativistic electron by a clever extension of Larmor's formula. The direct use of the fields is rather laborious. The result is P = (2e2/c)γ6[(dβ/dt)2 - (β x dβ/dt)2], which Liénard found in 1898. This is a very useful formula for finding the power radiated by fast particles, as in accelerators, or in astrophysics.
The Electron Synchrotron

The first particle accelerators were linear accelerators using high-voltage DC power supplies, such as the Van de Graaff and Cockroft-Walton machines, which could attain about 1 MeV. Next came the cyclotron (1934), where the circular orbit made periodic acceleration possible with lower RF voltages on the dees, the accelerating electrodes. The angular velocity of revolution depended only on the e/m ratio of the particles, not on their speed, so a constant RF frequency could be used. Protons and other heavy particles could be accelerated to about 25 MeV. These accelerators were very useful for nuclear studies, but the energies were too low for the creation of new particles. Cyclotrons cannot accelerate electrons because of the rapid increase in mass that brings them out of step with a constant RF frequency. The betatron (1940), which could accelerate electrons to similar energies by using transformer action, was then devised. Great attention had to be paid to designing the magnetic field so that the electron orbit would remain constant in radius and focussed, but the frequency problem did not arise. Betatrons provide a pulse of fast electrons at the end of each cycle of increasing magnetic field. For more on betatrons, see The Particle Electron.

Cyclotrons could be used to accelerate heavy particles even when relativistic effects entered by changing the accelerating frequency. Such FM cyclotrons or synchrocyclotrons were indeed developed, but the frequency modulation is very troublesome. Betatrons had the disadvantage of very heavy magnetic cores, and so their size was limited. In 1947, the first electron synchrotron was constructed in California, following the ideas of McMillen (1947) in the U.S. and Veksler (1946) in the U.S.S.R., who introduced the idea of synchronous acceleration. There were some smaller experimental synchrotrons, but the electron synchrotron was born full-grown, the first one providing 300 MeV electrons, and a dozen others following soon after. These electrons are generally used to produce high-energy photons in collisions with a target, which then can cause photonuclear reactions, such as the production of π mesons.

The electron synchrotron takes advantage of two consequences of the small mass of the electron. First, the electron is easily accelerated to speeds very close to c, when their orbital period in a magnetic field will not change significantly with energy. Secondly, the electron is easily deflected by a magnetic field, even at high energies, so that accelerators can be a convenient size. A diagram of an electron synchrotron is shown at the right. The toroidal vacuum chamber, and the C-shaped magnet cores, are not shown. A magnetic field B, shown directed into the page, deflects electrons into a circular orbit of radius r in which they move anticlockwise, approximately at speed c. The kinetic energy of the electrons is T = mc2(γ - 1). The electrons are accelerated on each revolution by the electric field in a resonant cavity, shown at the left. In order for the radius to remain constant, the magnetic field B must be increased synchronously with the increase in energy of the electrons, which gives the name to the machine. At the end of an accelerating cycle, the accelerating voltage is turned off, while the magnetic field continues to increase, causing the beam to spiral in and strike the target. This happens about 60 times a second, so the beam appears continuous to the human observer.

The magnet windings are usually connected across a large capacitor to form a resonant circuit, so the magnet power supply need only furnish the losses, and the power factor is acceptable. At the beginning of a cycle, a large pulse of electrons is injected by a thermionic electron gun in the direction of the orbit. Most of these electrons are lost, but a sufficient number assume orbits that do not strike the vacuum chamber walls. The magnetic field B must be carefully shaped to focus the beam. In general, B bows outward like the normal fringing field at an air gap, and can be described by B = Bo(ro/r)n. If n > 0, the electron orbits will oscillate stably about the orbital plane, and if n < 1 the orbits will oscillate stably in a radial direction. As B increases, these oscillations will be damped so that the beam will become narrow and well-defined. The electrons must also form a bunch along the orbit so that they will have phase stability. An electron that receives a little too much energy on one pass will forge ahead and receive less on the next pass, and the same sort of compensation will result for an electron that falls behind. The focussing and phase stability are necessary for achieving a useful beam current, which is usually in the region of microamperes.

Electrons are injected with energies of around 100 keV. They are then accelerated to about 2 MeV by betatron action. The necessary large field within the orbit is provided by flux bars of relatively small cross section that pass between N and S poles of the magnets. Initially, much of the flux passes by this route because of the much smaller air gap, but at the end of betatron acceleration the flux bars saturate and thereafter act like an air gap, most of the flux passing through the main pole faces to steer the electrons. At 2 MeV, the electron speed has become constant enough near c that constant frequency RF acceleration can take over. With an acceleration of 10 keV per pass, and a frequency of 47.7 MHz (appropriate for r = 1 m), the electrons will gain 477 MHz in 10 ms, neglecting losses. The synchrotron principle works very well for electrons.

A large number of electron synchrotrons have been built, of which the largest seems to be the 10 GeV machine at Brookhaven National Laboratories, which has r = 100 m. The maximum magnetic field is only 3300 gauss, and the RF accelerating voltage is 10.5 MV. Note that the electrons from this machine have γ = 20,000, approximately, and they are heavier than protons! Most electron synchrotrons, however, are relatively small machines. If 10,000 gauss is taken as a convenient upper limit for the magnetic field, then for 100 MeV electrons, r is only about 33 cm. Therefore, an electron synchrotron is a convenient source of high-energy radiation.

The relation between B, r and T can be found by equating the magnetic force to the mass times acceleration, or (evB/c) = (γmv2/r), since the effective mass of the electron is γm, and it does not change in circular motion. This then gives Br = γmvc/e = βγmc2/e = βE/e ≈ E/e. Note that the expression found in some references, [T(T + 2mc2]1/2 is just βγmc2. The acceleration frequency is f = v/2πr = eB/2πγmc. If we express this in terms of r instead, using the value of Br, we find f = (c/2πr)(1 - 1/γ). If r is in metres, then f = 47.7 MHz/r. These relations are shown in the diagram of the synchrotron above.

There are losses in the synchrotron, of which the most important is the radiation loss due to the circular orbit of the electrons. Since β is at right angles to the centripetal acceleration, the formula for radiated power becomes P = (2e2/3c3)γ4|a|2, and a = v2/r, so P = (2e2c/3r2)β4γ4. Multiplying by the time per revolution, we have the energy loss per revolution due to radiation of δE = (4πe2/3r)β3γ4. In most cases, β can be set equal to 1. In evaluating these expressions, note that Gaussian units must be used, in which e = 4.803 x 10-10 esu. The result is δE (erg) = 2.125 x 106 (E4 erg/r cm). This works out to δE (MeV) = 0.0884 E4 GeV/r m. For the Brookhaven synchrotron, this is 8.84 MeV, not far below the 10.5 MV acceleration per turn.

The total power lost by radiation will be NδE/T = (I/e)δE, which works out to 106IδE W, where I is in ampere and δE in MeV. With a beam current of 1 μA, the loss in the Brookhaven machine will only be about 8.84 W, and much less in smaller machines. Nevertheless, this radiation is easily detected and quite interesting. Each electron makes a sharp pulse when it is at the point tangent to the line of sight, so the frequency spectrum is broad, extending from the orbital frequency up to the maximum energy of the electrons. The radiation is in a very narrow forward cone, that shrinks as the energy increases. It was first observed visually in 1948, looking at the radiation through a mirror, since it is not advisable to stand directly in the beam! At 60 MeV it is a red glow, then becoming white and most intense at 200 MeV, after which the maximum passes into the ultraviolet. This is synchrotron radiation, which can be seen in the Crab Nebula, among other places. This should not be confused with the radiation produced by the beam itself on impact with the target, which can be thousands of Röngen per minute. Jackson gives details on the angular distribution and spectrum of synchrotron radiation.
References

J. D. Jackson, Classical Electrodynamics, 2nd ed. (New York: John Wiley & Sons, 1975). Chapters 6, 11, 12 and 14.

M. S. Livingston, High-Energy Accelerators (New York: Interscience, 1959). Chapters 2 and 3.

The Particle Electron and Thomas Precession

Contents

  1. Introduction
  2. Larmor's Theorem
  3. Magnetic Interaction Energy
  4. Bohr Atom
  5. Zeeman Effect
  6. Spin-Orbit Interaction
  7. Thomson Scattering
  8. Compton Scattering
  9. Beta Decay
  10. The Betatron
  11. References

From J. J. Thomson's demonstration of the particle nature of the electron in 1896 until the creation of modern quantum mechanics by Heisenberg and Schrödinger in 1926, the electron was considered to be a small classical particle described by classical mechanics and relativity, as modified by Planck's quantum theory and other ad hoc quantum properties, as seemed necessary. This theory, which may be called Lorentz's electron theory, was more successful than it had any right to be, explaining many of the phenomena of what has become known as "modern physics." It is still used quite generally for practical purposes, and still forms the basis of most physicist's concepts of the electron. The limitations of this method were becoming obvious just as quantum mechanics was introduced. Quantum mechanics corrects and extends the theory, giving it a firm foundation as a model of microscopic behavior of matter. This article will explain those aspects of the particle model of the electron that are not treated elsewhere on this site, up to the explanation of the Thomas precession, a relativistic effect of some mystery that is a triumph of the theory. In many cases, the particle theory is much easier to comprehend, while confirmed by better but much more complicated fully quantum-mechanical analysis.

The particle electron was conceived as a small material particle of mass 9.10939 x 10-28 g carrying a negative electrical charge of magnitude e = 4.803207 x 10-10 esu. The electromagnetic force on a particle of charge q and mass m is given by the Lorentz force, f = q[E + (1/c)v x B], where E is the electric field in statvolt/cm, B the magnetic field in gauss, and v the velocity in cm/s. I will use Gaussian (cgs) units here and in what follows, but will give the equivalents in MKS units in most cases. The use of MKS units in atomic physics can be inconvenient, especially when magnetic fields are considered. Otherwise simple equations sprout meaningless conversion factors with εo and μo and 4π's in them, and dimensional analysis is confused. I will avoid this whole mess by using Gaussian units. The conversions 10,000 gauss = 1 T and 300 V = 1 statvolt are useful in expressing results in practical units. The electronic charge in MKS is e = 1.602177 x 10-19 C. (Multiply by c/10, where c is in cm/s, to get esu). The electron-volt (eV) unit of energy is 1.602 x 10-19 J or 1.602 x 10-12 erg. For more information on units, see Monopoles.

In MKS units, the Lorentz force drops the factor c (making relativistic considerations a little more inconvenient), and the force is in newton instead of dyne. This is a fully relativistically-correct force equation (unlike the others commonly appearing in nonrelativistic electrodynamics) that is Lorentz invariant. However, it can be used with Newton's equations of motion dv/dt = f/m to describe the motion of the particle electron in arbitrary electric and magnetic fields. This is done in the article Charged Particle Dynamics, to which the reader is referred for this very important application.

The equations of motion dv/dt = (q/m)[E + (1/c)v x B] can be expressed in a very simple form by transformation to a frame of reference rotating with a constant angular velocity ω. If v' is the velocity in this rotating frame, then in the inertial frame v = v + ω x r'. Taking the time derivative, dv/dt = dv'/dt + 2ω x v' + ω x (ω x r'). We recognize the "fictitious" Coriolis and centrifugal forces in this expression. When this is put into the equations of motion, we find dv'/dt = (q/m)E + (e/c)v' x [B + (2mc/q)ω] + (ω x r')[(q/c)B + mω] after a little manipulation.

If the angular velocity of rotation is chosen to be ω = -(q/2mc)B, the Coriolis term disappears, and the centrifugal term is zero from the properties of the cross product. All that remains is dv'/dt = (q/m)E. In the rotating system, the electron moves as if it were under the influence of E alone! This also holds for an assembly of more than one particle, so long as they all have the same (q/m). All electrons, in fact, have exactly the same mass and charge, which really follows from quantum mechanics and the concept of identical particles. Indeed, this is verified experimentally to an extremely high precision. The angular velocity ωL = -(q/2mc)B is called the Larmor angular velocity, and the movement itself is called Larmor precession. Note that the axis of rotation is the direction of the magnetic field, and that the senses are opposite for a positive charge.

An electron has q = -e, so ωL = (e/2mc)B for electrons. A little confusion may be introduced because the charge on the electron is negative. It should be remembered that for a positive particle, the rotation is in the opposite direction. The quantity e/2mc = 8.7942 x 106 esu-s/gm-cm is called the gyromagnetic ratio. To see why this name is used, consider an electron moving with velocity v in a circle of radius r. The period of rotation is 2πr/v, so the average current is ev/2πr. The magnetic moment of a current loop is the product of the current, i/c and the area of the loop, cm2, or μ = (i/c)A. The current i is in esu/s, and the c converts it to emu/s. Our revolving electron then has a magnetic moment of μ = (ev/2πrc)(πr2) = evr/2c. The angular momentum of the revolving electron is l = mvr, so μ = emvr/2mc = (e/2mc)l. Therefore, e/2mc is the ratio of the magnetic moment to the angular momentum of the particle. For an electron, it is easy to see that they are opposite, so that really μ = -(e/2mc)l. An angular momentum is always accompanied by a magnetic moment.

The magnetic moment just mentioned can be realized as a current loop, with the direction of the moment normal to the area in the direction a screw would advance when rotated in the direction of a positive revolving charge, and of magnitude (i/c)A. Alternatively, it is the limit of equal and opposite magnetic poles keeping the product μ = md constant as d → 0. Magnetic poles are an idealization, since no particles with magnetic charge have yet been discovered.

A uniform magnetic field B exerts a torque N = μ x B tending to turn the moment in the direction of the magnetic field. Since the magnitude of this torque is μB sin θ, the energy U = ∫Ndθ as a function of the angle can be found by integration, U = -μB cos θ, or U = -μ · B, where the zero is taken as the state in which the moment is normal to the field.

The equation of motion for a body of angular momentum L acted upon by a torque N is dL/dt = N. Here, μ = -(q/2mc)L, so dL/dt = -(q/2mc)L x B = (q/2mc)B x B. This means that L precesses with angular velocity (q/2mc)B about B. For a positive charge, the precession would rotate a screw in the direction of B. For an electron, the precession is in the opposite direction. This is, of course, just the Larmor precession that was described above. Since the magnitude of the torque is N = μB sin θ, and dL/dt = ωL sin θ, the result is not difficult to derive, showing the close connection between the gyromagnetic ratio and the Larmor angular frequency.

The interaction energy can also be interpreted as rotational energy. Using index notation, the rotational energy of a body with a symmetrical inertia tensor Ijk is W = (1/2)Ijkωjωk. If ω j → ωj + ωj', then W' = (1/2) Ijkωjωk + Ijkωjωk' + (1/2) Ijkωjk', making use of the symmetry of the inertia tensor. If ω' is much less than ω, the final term is negligible, so the change in energy is U = W' - W = Ijkωjωk'. The angular momentum of the body is Ijkωj = Lk, so U = Lkωk' = - (q/2mc)L · B = -μ · B, as above.

The most celebrated success of the particle electron model was the Bohr atom of 1913, coming just after Rutherford established that the atomic nucleus was small and massive. The idea of electrons orbiting this sun-like nucleus as planets was a persuasive one. It probably would have been decisively rejected on stability grounds if the theory did not give, astoundingly, the correct numbers for the hydrogen spectrum. We now know that the planetary analogy is quite incorrect, and in atoms even a wave-packet electron soon spreads uniformly, but the correctness of the results commanded respect.

Suppose an electron revolves in a circular orbit of radius r around a nucleus of charge +Ze with speed v. Newton's second law gives mv2/r = Ze2/r2, so mv2r = Ze2. Bohr then made the quantum ansatz that the angular momentum was an integral multiple of h/2π = h': mvr = nh'. Dividing these two expressions yields v = Ze2/nh'. Then, we easily find that r = nh'/(mZe2/nh') = n2h'2/mZe2. The ground state of hydrogen has Z = 1 and n = 1, so v = e2/h' and r = h'2/me2. The first number, divided by c, is the ratio v/c for the ground state of hydrogen, known better as the fine structure constant α = e2/ch' = 1/137.04. The second number is the Bohr radius 0.052918 nm.

In fact, the ground state of hydrogen is a minimum-uncertainty state of zero orbital angular momentum absolutely nothing like an orbiting electron. The Bohr radius is, however, rather close to the "size" of the hydrogen atom as found from scattering experiments. The energy of the electron in a state of quantum number n is the sum of its kinetic energy and potential energy, E = mv2/2 - Ze2/r. Substituting our values for r and v, we find E = -mZ2e4/2n2h'2. These energies are exactly correct (except for some very small corrections for things unknown at the time). In theoretical spectroscopy, wave numbers σ = 1/λ in cm-1 are generally used in place of frequencies as more convenient in size. In terms of wave numbers and Planck's quantum of radiation, E = hν = hcσ, or σ = E/hc. We then have, taking the absolute value of the energy, σ = Z2e4m/4πch'3n2 = RZ2/n2, where R is the Rydberg constant, which had been determined experimentally with great precision.

Bohr then assumed that the observed spectrum was the consequence of transitions between states of different n. While in any state, the atom did not radiate. This was an astonishing assertion, but suggested from the empirical expression of spectral lines as differences in "term values" which up to then had no good explanation. This gave σ = RZ2(1/n2 - 1/n'1), in emission if n1 > n2, and in absorption if n2 > n1. Efforts to find harmonic relationships between spectral lines, as in vibrations of material bodies, had been wholly and completely unsuccessful. For large quantum numbers, however, the energy differences between levels correspond to orbital frequencies. Bohr laid the search for harmonics to rest once and for all. The Rydberg constant was now expressed in terms of elementary constants, and gave the precise value 109,737.315 cm-1. This is the greatest success of the Bohr model, and is astonishing.

The Bohr model was refined by Sommerfeld to include elliptical orbits and relativity, and even then it gave excellent results for hydrogen-like spectra. The Bohr model failed rather completely when used for more than one valence electron, or for molecules, however. This is characteristic of particle electron analyses, which can give very good numbers in simple cases, but cannot be extended to more complex situations, since the analogy is very imperfect. Quantum mechanics gives a much more satisfactory account of the hydrogen atom, and the theory can be extended to more complex atoms and to molecules as well.

Quantum mechanics introduces a better description of the angular momentum, introducing the quantum number l, which for a state of principal quantum number n goes from 0 up to n-1. The orbital angular momentum is space-quantized so that its value along any direction can take only the values lh', (l-1)h', ..., -lh' = mh', where m is the magnetic quantum number corresponding to l. It is called magnetic because these states are separated in energy when in a magnetic field, as will be discussed in the next section. Any state can be specified uniquely by n, l, m. We shall see later that electrons have a fourth degree of freedom, spin, which has s = 1/2 and ms = ±1/2. The letters s,p,d,f,g,h,... are used for l = 0,1,2,3,... for historical reasons. Therefore, in hydrogen, n = 1 corresponds to an s state, n = 2 to s and p states, n = 3 to s, p and d states, and so on. There are 2l+1 states for any value of l, labelled by their m values. For any value of n the total number of states is n2, where an s is one state, a p 3 states, a d 5 states, and so on. This degeneracy (several states with exactly the same energy) occurs only for hydrogen-like spectra, and is a result of a 4-dimensional symmetry depending on the 1/r potential. For all other spectra, the states of different l are separated in energy.

Whatever the complicated electronic motions there were that produced observed spectral lines, the application of a uniform magnetic field should produce changes because of the Larmor precession. A general vibration can be resolved into three harmonic components at right angles. Suppose the magnetic field is in the +z-direction. Then the vibration in the z-direction will be unaffected by the magnetic field, while the x and y motions will precess with the Larmor angular velocity (e/2mc)B in a right-handed sense about the +z-axis. An x-vibration x = a cos ωt can be expressed as the sum of two circular motions (a/2)eiωt and (a/2)e-iωt rotating in opposite directions, and similarly for a y-vibration. The clockwise motion will be speeded up by the Larmor precession to ω + ωL, and the anticlockwise motion will be slowed down by the same amount.

When we look at the light emitted in a direction normal to the magnetic field, we should see one component polarized in the direction of the magnetic field ("p" polarization) of unchanged frequency, and two components polarized perpendicularly to the direction of the field ("s" polarization), one at a smaller frequency (or energy) and one at a larger, from what we know of the radiation from a linear dipole. If we look along the magnetic field, say with B pointing towards us, we should see two circularly polarized lines, the right-circularly polarized one of greater frequency, the left-circularly polarized line of smaller frequency. There will be no line of unchanged frequency, since an electron does not radiate in its direction of motion. The change in wave number from the original frequency will be Δσ = ωL/2πc = eB/4πmc2 = 4.6686 x 10-5B cm-1. We will denote this Zeeman shift by Z for short.

The pattern of the central p-polarized line flanked by the two s-polarized lines is called a normal Zeeman triplet. Even for a field of 10,000 gauss, approaching the largest that can conveniently be produced by an iron-core electromagnet, the width of the triplet will be only 0.93 cm-1. For comparison, a green line of wavelength 555 nm has σ = 18,018 cm-1, so it requires great spectral resolution to see a Zeeman triplet. At first, Zeeman was only able to note a broadening of a line, and the polarization of its edges, to indicate that something was happening. Since then, increased resolution has allowed the examination of many Zeeman patterns.

It was great puzzlement when not the normal triplet, but more complex symmetrical patterns were found. Only in the case of singlet lines, for which the spin quantum number S was zero, was a normal triplet observed. The overall size of the effect was, however, correctly predicted by the classical analysis. These patterns were eventually completely explained by a semiclassical analysis due to Landé, called the vector model of the atom, and based on the newly-postulated spin of the electron. It happened that the electron was not just a charged particle, but that it had an intrinsic angular momentum whose component along any direction could only be ±h'/2. This resembled the angular momentum of a rotating body, hence the term "spin," but any explanation of it as a classical rotation of the electron fails. It was shown by Dirac to be a consequence of a proper relativistic treatment of the electron.

We would expect the magnetic moment corresponding to the spin to be opposite in direction, and of magnitude (e/2mc)(h'/2), but Dirac showed that it was actually twice this, eh'/2mc, which is the magnetic moment corresponding to an orbital angular momentum of one full unit, h'. The magnetic moment eh'/2mc erg/gauss is called the Bohr magneton, a convenient unit for atomic magnetic moments. The gyromagnetic ratio of the electron can be written g(e/2mc), where the g-factor is 2. The complication in the Zeeman effect is that the orbital angular momentum and the spin angular momentum tend to precess at different rates, since they have different g-factors.

We'll consider the case of one electron, which will include all the essential features. The case of two or more electrons is similar, and can be found in the References. The total angular momentum j is the sum of the orbital angular momentum l and the spin angular momentum s. It is a constant of the motion, and its magnitude is specified by the quantum number j. Its vector magnitude is √l(l+1) h' = j*h', and its component along any axis can have the values mh' = jh', (j-1)h', (j-2)h', ..., -jh', where m is called the magnetic quantum number. There is actually a set of 2j+1 states of equal energy in the absence of an external field, and the atom may be in any one of them, or in a linear combination. The orbital and spin angular momenta are described by quantum numbers l and s = 1/2 in a similar way. These quantum numbers are constants, but their projections on any direction in space are not. The total angular momentum can be viewed as the vector sum j = l + s, where j can take the values l + 1/2 and l - 1/2 only, if l > 0. When l = 0, j = 1/2 only, and the angular momentum is due entirely to the spin. The projections of l and s on j are constant.

What happens is that only the projections of the magnetic moments on the direction of j are constant, and the sum of the projections is the total magnetic moment corresponding to j. The orbital angular momentum and the spin angular momentum will not precess independently about B, but the total angular momentum and its associated magnetic moment will. The actual ratio of the resultant magnetic moment to the total angular momentum can be expressed as g(e/2mc), where g will be between 1 and 2. By considering a vector diagram for the angular momenta and the magnetic moments, the Landé g-factor can be shown to be given in terms of the quantum numbers by g = 1 + [j(j+1) + s(s+1) - l(l+1)]/2j(j+1). When s = 0, g = 1, while if l = 0, g = 2. s = 0 does not occur for one electron, of course, but it does for two electrons, in which the states are either singlets with S = 0 or triplets with S = 1. Capital letters are used for more than one electron. In LS-coupling, the individual orbits couple to L and the spins to S, and then the two resultants couple to J. The g-factor is given by the same formula.

For example, a 2P3/2 state has s = 1/2, l = 1 and j = 3/2 in the usual notation. In labelling states, S,P,D and F mean l = 0,1,2 and 3, the superscript is 2S + 1, and the subscript is J. The g-factor for this state is g = 1 + [(3/2)(5/2) + (1/2)(3/2) - (1)(2)]/2(3/2)(5/2) = 4/3. In a magnetic field B, each of the 4 states now has a different energy (4Z/3)m, which makes a symmetrical set of equally-spaced states. Similarly, a 2S1/2 state has g = 2 since l = 0, and splits into two states of energies 2Zm = +Z and -Z. Each spectral line in the observed Zeeman pattern of the line that is a transition between the P state and the S state starts on one of the P levels and ends on one of the S levels, obeying the selection rule Δm = ±1 for s-polarized lines, and 0 for p-polarized lines. The selection rules are given by quantum mechanics, and are conditions that the transition matrix elements not vanish from symmetry. The consequent pattern is composed of six lines instead of three, as shown in the diagram. Relative intensities of the lines are shown above the arrows representing the transitions.

All of the many Zeeman patterns that are observed can be calculated by this method, and the results agree very well with the observations. This is excellent verification of the spin and angular momentum of the electron, and its anomalous g value of 2, as predicted by the Dirac theory. The magnetic interaction energy is, in practice, much less than the dependence of the energy on the relative orientations of l and s, which is called the spin-orbit interaction energy. A very strong magnetic field can succeed in decoupling the orbital and spin motions so that they precess independently in the field. In this case, the two precessional energies simply add with their individual g-factors, so the wavenumber shift is Zm + 2Zms, where ms = ±(1/2). For one electron, the levels are simply Zm ± Z, plus a correction for the spin-orbit interaction. This is called the Paschen-Back effect, and is rather hard to observe. A field of 32,000 gauss was required to see the effect in the hydrogen line at 656.3 nm.

Spin-Orbit Interaction

Hydrogen-like single-electron states can be labelled by the quantum numbers n, l, j and m, where m is now the magnetic quantum number associated with j, not l. These states are often called orbitals, and linear combinations of them are used to construct approximate wave functions for many-electron systems (many can mean 2). Energy levels in more complex spectra can be labelled by L and J, as well as by the value of S that is coupled with L to form J. L and S are thought of as vector sums of the orbital and spin angular momenta of the individual electrons, which is a reasonable approximation in many cases. The letters S, P, D, F and so on are used to signify the L value, 2S+1 is placed as a superscript before this letter, and J as a subscript following.

The energies of the two states 2P3/2 and 2P1/2 of the same principal quantum number n for one electron differ in the relative orientation of the spin and orbital angular momenta. In the first state, they are more or less parallel, and in the second state more or less opposed. In more complicated atoms, the energies of states with different L and S have different electrostatic energies, but in hydrogen the difference in the energies of the states j = l ± 1/2 can be expected to be small, and due principally to the different energies of the spin in the magnetic field that appears in the rest frame of the electron that is moving in the electrostatic field of the nucleus. This magnetic field is E x (v/c). Since E = (Ze/r3)r, the magnetic field is seen to be proportional to the orbital angular momentum of the electron. In fact, B = l(Zeh'/mcr3). The Larmor precession frequency in the rest frame of the electron is then ωL = - g(e/2mc)(Zeh'/mcr3)l. We have put g in this equation for more generality. For an electron, g = 2.

When the energy due to this precession is calculated, it turns out to be twice as large as the observed value, as if the electron had g = 1. Since this is definitely not the case, there must be something wrong. What this is has to do simply with relativistic kinematics. The rest system of the electron is obtained from the fixed inertial system by using a relative velocity v that is continually changing in direction. Let the rest system and the inertial system both have the same origin for simplicity. Since Lorentz transformations do not commute, the effect is that the successive Lorentz transformations deriving the electron rest system from the fixed inertial system do not leave the coordinate axes parallel. That is, the electron rest system is continuously rotating as the electron revolves about the nucleus (in our particle picture), and the rate of rotation depends on the acceleration of the electron. This rotation is called the Thomas precession, after L. H. Thomas, who first recognized it in 1926. Since the acceleration is also due to the nuclear electric field, this kinematic precession has exactly the same form as the spin-orbit precession, and happens to be half as large and in the opposite direction. The precession angular velocity in the inertial system is then the sum of these two angular velocities, which gives the result ω = - (e/2mc)(Zeh'/mcr3)l. This value of the spin-orbit interaction agrees with observations.

The interaction energy is the product of this precessional angular velocity and the spin angular momentum h's, or ΔU = (Z/2)(g - 1)(eh'/mc)2r-3l·s. To apply this result, we need an expression for l·s in terms of the quantum numbers j,l and s, and a suitable value of r must be used. Squaring j = l + s, we find j(j+1) = l(l+1) + 2l·s + s(s+1), using the proper quantum-mechanical expressions for the squares of the momenta. Then, the dot product is seen to be (1/2)[j(j+1) - l(l+1) - s(s+1)]. The best way to handle r-3 is to average it over the orbital electron density. The result for hydrogen-like atoms is avg(r-3) = (Z/a1)3[n3l(l+1/2)(l+1)]-1.

Putting this all together, the spin-orbit interaction energy Γ = (a/2)[j(j+1) - l(l+1) - s(s+1)], where the constant a = Rα2Z4/n3l(l+1/2)(l+1) cm-1, in terms of the Rydberg constant R (109737 cm-1) and the fine-structure constant α (1/137). For Z = 1, the numerator is 5.85 cm-1. For the level n = 2, l = 1 we have the states 2P3/2 and 2P1/2. From our formula, Γ = 0.122[j(j+1) - 11/4], which gives +0.122 for j = 3/2 and -0.244 for j = 1/2. The "doublet" is shown at the right. The numbers to the left of the states give the degeneracy. The level of smallest j lies lowest, and the "center of gravity" is at the level without spin-orbit coupling, since 4(0.122) + 2(-0.244) = 0. This is called a "normal doublet." The total splitting is 0.366 cm-1, a very small amount. Careful experiments give 0.364 cm-1, so the agreement is excellent.

In more complex atoms where the orbital angular momenta are couped to L and the spins to S, and these coupled to a total angular momentum J (LS-coupling), there is also a spin-orbit coupling depending on L·S, but the splitting factor a may be a result of several effects. It can be described, at least empirically, as we have done for the magnetic interaction of a single electron. The states arising from one L and S are called a "multiplet," and the multiplicity is 2S + 1, the number of possible spin orientations, and so of J values. Each state of given J = L+S, L+S-1, ..., |L-S| is a set of 2J+1 states of the same energy when B = 0 that can be observed in the Zeeman effect. LS-coupling occurs only when states of different S are well-separated in energy. S affects the symmetry of the state, which in turn affects the electrostatic repulsion between the electrons. States of high S have the electrons well separated, and so have lower energies. A similar effect occurs for L, so states with the highest value of S lie lowest, and of these those with the highest value of L lies lowest. This is called Hund's Rule, and was observed experimentally long before it was explained by theory.

Since matter is composed largely of electrons, the interaction of electromagnetic radiation with electrons is a subject of considerable interest and utility. Electromagnetic radiation covers a wide range of frequency f, or what is equivalent, of quantum energy hf. The spectrum includes radio wave, microwave, infrared, visible, ultraviolet, X-ray and gamma-ray regions, with which the reader is generally familiar. Wave models for lower frequencies give way to particle models at higher frequencies, but radiation is the same thing whatever the frequency, and is described accurately by quantum mechanics.

The most notable consequence of the quantum nature is that energy transfers are in units of hf or h'ω, which seem to be the creation and annihilation of "quanta" or photons. This "particle model" is useful, but only represents a few properties of radiation, leaving out the essential role of phase, and cannot be carried far. A classical wave model is much more useful, but it must always be used with an eye to quantum behavior. In the photon model, the photon has an energy h'ω and a momentum h'ω/c, so its rest mass is zero, from the relativistic relation E2 - (pc)2 = (mc2)2 satisfied by the energy-momentum 4-vector. It also happens that the photon has an angular momentum h' ("spin 1") that takes only the values ±h' with respect to the direction of propagation.

In the wave model, the electric field of a monochromatic wave can be represented by εEeik·r - iωt. k is the wave vector, in the direction of propagation, and of magnitude k = 2π/λ, where λ is the wavelength, where kω = c, the speed of propagation, about 3 x 1010 cm/s. ε is a unit vector called the polarization vector, which must be normal to k. Two such polarization vectors at right angles are sufficient to describe any state of polarization of the radiation in a unique way. The energy density in the wave is |E|2/8π, and the energy flux is c|E|2/8π erg/cm2s. There is a magnetic field B perpendicular to E that oscillates in phase with it, but the magnetic field has a negligible influence on the motion of an electron on which the wave falls.

Linear polarization results when the x and y vibrations are in phase. Circular polarization results when the x and y vibrations are equal in amplitude but differ in phase by 90°. The polarization vector εL = (1/√2)(εx + iεy) represents left-circular polarization with time dependence e-iωt, and (1/√2)(εR = εx - iεy) represents right-circular polarization. Right-circular polarization means that the electric vector rotates clockwise as you face the oncoming radiation. Ordinary unpolarized radiation can be considered to be an incoherent superposition (that is, with an arbitrary phase difference) of linear polarizations at right angles, or of right- and left-circularly polarized components, with half the power in each of the polarizations.

Now we look at the radiation from an accelerated charge. The power radiated into solid angle dΩ by a charge with acceleration a = dv/dt is dP/dΩ = (e2/4πc3)|ε·a|2 where ε is the polarization vector of the radiation. For a derivation of this formula, see Jackson. To make the meaning of the formula clearer, consider an electron moving along the z-axis with z = Ae-iωt. The magnitude of the acceleration is a = ω2A, so dP/dΩ = [(Aeω)2/4πc3] |ε·εz|2.

Consider radiation in a direction making an angle θ with the z-axis. One polarization vector can lie in the xy-plane normal to the direction of propagation, the other can lie in the plane defined by the direction of propagation and the z-axis, as shown in the diagram. We note that there will be no radiation of the first polarization, since the dot product is zero. The other polarization vector makes an angle 90° - θ with the z-axis, so the dot product gives sin2θ. Therefore, dP/dΩ = C sin2θ. We see that there is no radiation along the axis of the vibration, and that the radiation is a maximum in the xy-plane. It is easy to integrate over all directions to find that P = (8π/3)C. These results are probably familiar.

If the plane wave ε'Eeik·r - iωt falls on an electron, the acceleration a = -(eE/m)ε', omitting the exponential factor. From the radiation formula, we then have dP/dΩ = (e4E2/4πm2c3) |ε·ε'|2 = (cE2/8π)(e2/mc2)2 |ε·ε'|2. The first factor is the incident energy flux. If we divide by this, we get the cross-section for scattering into solid angle dΩ, so dσ/dΩ = (e2/mc2)2 |ε·ε'|2.

The quantity e2/mc2 has the dimensions of a length. Suppose the electron was constructed by starting with a conducting sphere of radius r, and bringing in its charge e from infinity in small increments dq. This will require more and more work as the charge builds up. In fact, the increment in energy is dE = (q/r)dq. Integrating from q = 0 to q = e, we find that E = e2/2r. Einstein tells us that E = mc2, so we have r = e2/2mc2. If all the mass of the electron were due to the electrostatic energy in assembling it, then this would be roughly the size of the electron. Of course, this is not true, and the electron's mass is not totally electromagnetic, but it is a good story. Therefore, e2/mc2 is called the classical electron radius ro, and has the value 2.82 x 10-13 cm. This is roughly a nuclear size, but we know that actual electrons cannot be confined in such a small volume. It is tempting to think of an electron as a little hard sphere rattling around in the region where its wave function is nonzero, but the question of what the electron looks like is unlikely to be answered by any such picture, and our confusion has no practical effect.

The scattering cross section is dσ/dΩ = ro2 |ε·ε'|2. Let's suppose that the wave vector of the unpolarized incident wave is directed along the +z-axis, so that the electric fields lie in the xy-plane, and their polarizations are defined by the unit vectors in the coordinate directions, with half the power in each. Polarization directions can be chosen for radiation scattered into θ,φ with one in the xy-plane, and the other in the plane of the scattered wave vector and the incident wave vector. These will be, respectively, ε1 = -εx sinφ + εy cosφ, and ε2 = cosθ(εx cosφ + εy sinφ) + εz cosφ. It is now easy to pick out the contributions from the two incident polarizations. The result is cos2θcos2φ + sin2φ and cos2θsin2φ + cos2φ. On adding, φ drops out, and we have just 1 + cos2θ, which must be divided by two, since half the power is in each incident polarization.

Our final result is dσ/dΩ = (ro2/2)(1 + cos2θ), called the Thomson scattering formula. The total cross section is σ = (8π/3)ro2 = 0.665 x 10-24 cm2. It is a small cross-section, of the size of nuclear cross-sections, independent of frequency, and back scattering is as important as forward scattering. For light, Rayleigh scattering from density fluctuations of a size comparable to the wavelength is far more probable, with its λ-4 wavelength dependence, which gives the blue of the sky.

We also note that the scattered radiation is unchanged in frequency. If there were a significant momentum transfer, we would expect the electron to recoil and take some energy with it, so that the scattered radiation would be of smaller frequency. This, in fact, becomes important when the photon momentum h'ω/c is not negligible compared to mc. The condition for the validity of the Thomson formula is also h'ω/mc2 <<>

The Thomson formula applies to a single free electron. If we have scattering from Z electrons, we can add the scattered powers to obtain a cross section Zσ provided the scattered amplitudes are incoherent, which means that in the average value of the square of the amplitudes the cross terms average to zero due to random phases, so that |Σa|2 = Z|a|2. If the phases are not random, then |Σa|2 can be as large as |Za|2 = Z2|a|2. This is coherent scattering, where the intensity may be gathered into sharp peaks, as in X-ray scattering by crystals. In Thomson scattering, the cross section may peak sharply in the forward direction when the scattered amplitudes add coherently, as from the Z electrons in a single atom.

In 1922, Arthur Holley Compton studied the energy distribution of X-rays scattered at 90° by a graphite target. In addition to X-rays of unchanged wavelength, he found a peak due to X-rays of slightly larger wavelength, whose position varied with the scattering angle. This became known as the Compton Effect, which Compton explained as an elastic collision of an X-ray photon with an atomic electron. The recoiling electron carried away some of the energy of the incident photon, leaving a photon of smaller energy. A simple calculation on the basis of the conservation of energy and momentum gave the wavelength of the scattered X-ray as a function of scattering angle, which was fully confirmed by experiment.

A diagram of such a collision in the laboratory system is shown at the right. The electron of rest mass m is initially at rest, while the incident X-ray photon approaches from the left. The scattered photon leaves at an angle φ, the recoil electron at an angle θ. For each particle, the energy and momentum are shown in the diagram. These quantities are the relativistic values, with β = v/c and γ = (1 - β2)2, where v is the recoil velocity of the electron. Conservation of energy gives h(f - f') = mc2(γ - 1), where f is the frequency of the X-ray before collision, and f' the frequency after. Conservation of momentum is expressed by two equations. One is (f - f'cosφ) = (γmc2β/h) cos θ = K cosθ, and the other is f'sinφ = K sinθ. Squaring the two equations and adding, we get f2 -2ff'cosφ + f'2 = K2, which eliminates the angle of recoil θ. Subtracting the result of squaring the equation resulting from the conservation of energy gives 2ff'(1 - cosφ) = K2 - [mc2(γ - 1)/h]2. The right-hand side reduces to 2(mc2/h)2(γ - 1) = 2(mc2/h)(f - f'), using γ22 - 1) = -1, and the conservation of energy again. The result is f - f' = ff'(h/mc2)(1 - cosφ). Since f = c/λ, (f - f')/ff' = (1/c)(λ' - λ), so finally we get λ' - λ = (h/mc)(1 - cosφ), a quite elegant result.

Consulting several references, I found that the authors did not present this algebra, merely waving their arms and usually saying that it was complicated, which probably means that they had trouble doing it. Therefore, I have given it here to show that it is simple (provided you do it in the order shown). It is also easy to use nonrelativistic mechanics, where the kinetic energy of the recoil electron is mv2/2 and its momentum is mv. The manipulations are much like those above, except that near the end you get (f - f')[1 - h(f - f')/2mc2] instead of just (f - f'). Then you argue that the second term in the brackets (half the energy transfer divided by the rest energy of the electron) is very small and can be neglected whenever nonrelativistic mechanics is a good approximation, and get the same result as the relativistic theory. It is easier just to use relativistic mechanics and get an exact result!

The quantity (h/mc) is called the Compton wavelength, with the value 2.4263 x 10-3 nm. It is the increase in wavelength at a scattering angle of 90°, such as Compton used, and is typical. For visible radiation, with λ = 500 nm, it amounts to about 5 parts per million, so is generally lost in the line width. In scattering from bound electrons, the mass in most scattering events is the atom mass, so there will be no Compton effect because of negligible recoil. For both these reasons, the Compton effect is not observed with visible light, but it still occurs whenever there is a recoil.

Radiation with photon energies of from about 1 keV to 100 keV, or wavelengths from around 1 nm to 0.01 nm, are generally called X-rays, and those of greater energies γ-rays. The Compton effect is easily detected for X-rays, since the Compton wavelength is a considerable fraction of the wavelength (from 0.002 to 0.2). The Compton wavelength itself corresponds to an energy of 0.517 MeV, near the rest energy of the electron, mc2 = 0.511 MeV. The ratio (hf/mc2) varies from 0.002 to 0.2 for X-rays, so the effect of electron recoil can be important.

A competing process is the complete transfer of the energy of the photon to an electron, which is called the photoelectric effect. It is easy to show that energy and momentum cannot be conserved in this process, so it occurs only when the excess momentum can be simultaneously transferred to the atom. The theory is involved, so we will not consider it here. Another competing process, for photons of energy greater than 1.022 MeV, is pair production, in which the photon disappears and is replaced by an electron and a positron, its antiparticle. Considering these things would take us too far from our aims, but the reader can find them discussed where the absorption of gamma rays in matter is discussed. X-rays can eject electrons from atoms, creating positive ions, and tracks of recoil electrons can even be seen in a cloud chamber.

The cross-section for Compton scattering is very closely related to the Thomson formula. In fact, dσ/dΩ = ro2(k'/k) |ε·ε|2, where k' and k are the magnitudes of the wave vectors for the scattered and incident radiation. This formula holds for spinless particles. The theory is rather involved, finally leading to the famous Klein-Nishina formula for the general case, so a description will not be attempted here.

Now, k/k' = λ'/λ = (λ' + λ - λ)/λ = 1 + (λ' - λ) = 1 + (hf/mc2)(1 - cosφ). Therefore, the factor k/k' = 1/[1 + (hf/mc2)(1 - cosφ)]. At φ = 180°, the factor is 1/[1 + 2hf/mc2]. This shows that the cross section is reduced for backscattering. If hf = mc2 (0.5 MeV gamma-rays), the cross section is smaller by a factor (1/3)2 = 0.11.

We now should have a good picture of what happens when electromagnetic radiation falls on an electron. Compton scattering supplements Thomson scattering by allowing for recoil of the electron. It is very interesting to see how far classical concepts can give reasonable results, when adjusted here and there to account for quantum effects.

Electrons are more than just electromagnetic particles, and run in a mysterious company of their own. One of the astounding properties of radioactivity is that atoms emit electrons of high energy, which were named "beta rays" early on. Some are orbital electrons that were already present, but others are new electrons produced in the nucleus, and these include not only negative electrons, but their positive antiparticles, positrons. It is impossible for electrons to exist as such in the nucleus, so any electrons coming from the nucleus must be newborn.

Although electrons cannot exist completely inside the nucleus, the wave functions of s electrons are not zero there, so there is some "handle" on them for nuclear processes. Sometimes a nucleus can make a transition from an initial state to a more stable state by emitting a photon, just as an excited atom emits radiation when the electrons rearrange themselves more comfortably. The energy differences are, however, much larger, up to many keV instead of eV, so the photons are gamma rays. It is usually easier for the nucleus to grasp the handle on an orbital electron and fling it away than to emit a photon, though the processes compete. These internal conversion electrons have definite energies, and leave the atom in an excited state from which it usually emits characteristic X-rays. The process is a nuclear one, not the emission of a gamma ray followed by the photoelectric effect on its exit from the atom.

In some cases, the nucleus would be more stable if a neutron were replaced by a proton, or a proton by a neutron. For example, if there were an unpaired proton and an upaired neutron in the nucleus, such a change would result in all the nucleons being paired up, probably a more stable configuration. In most naturally radioactive isotopes, the nucleus has too many neutrons as the result of a previous alpha-decay, and would like to change one into a proton to even up the neutron-proton ratio. Fermi showed that there was a force acting between neutrons, protons and electrons that would allow a neutron to create a positive and negative charge within itself, then hold on to the positive charge while kicking the negative charge out as an electron. This force had to be of very short range, so it was called the weak force because it had no effect except in nuclear close quarters, where, of course, it is locally powerful. It is strong enough to accelerate an electron to MeV energies over nuclear distances, it must be remembered, in opposition to electrical forces that would tend to attract the electron.

Although the change in nuclear energy in beta-decay was quite definite, the emitted electrons had a puzzling continuous distribution. The maximum energy agreed with the change in nuclear energy levels, but the electron energy went down to quite small values, as if energy were being lost. The only reasonable answer was that another particle was emitted that carried off the missing energy, but that this particle could not be detected by the means then available. Since then, this particle has indeed been observed, though its existence was accepted as a fact because of overwhelming circumstantial evidence. It was called the neutrino, of zero mass and spin 1/2, like the electron's. There is a whole family of neutrinos and antineutrinos, related to electrons and their fatter relatives the muon and tauon. The neutrino emitted in negative beta-decay is an electron antineutrino. Electrons and electron neutrinos both have the propery of "electron-ness," the electron of +1, the electron antineutrino of -1. If an electron antineutrino is created simultaneously with an electron in beta decay, then the net change in electron-ness is 0, or electron-ness is conserved, which appears to be a natural rule. In beta decay, nucleon number is also conserved, 1 neutron to 1 proton.

Negative beta decay is, symbolically, n0 → p+ + e- + ν'0, which conserves charge, neutron number, electron number, energy and momentum. The prime distinguishes an antiparticle, and the charge is written as an exponent. Fermi calculated the electron energy spectrum to be expected under many different condtions, and his results agreed with experiment. This amounted to all but conclusive proof of the existence of the neutrino. This alone would have secured Fermi's place as one of the greatest physicists of all time, but it is only one remarkable contribution of many.

Some artificially created isotopes have too many protons, and would like to see fewer. This can be accomplished by positive beta decay, p+ → n0 + e'+ + ν0, where an antielectron and a neutrino are emitted. Except that the positron is helped out by the positive nuclear charge, it is the same as negative beta decay in principle. In this case, a related process is possible, p+ + e- → n0 + ν0. This is called K-capture, since the electron is usually one of the s electrons in the K shell, but can be any s electron. Now, only the neutrino comes out, and the atom emits the usual X-rays to put its electrons in order. If the half-life for electron emission is long, K-capture is the preferred method of decay.

These reactions can be reversed, such as ν0 + n0 → p+ + e-, but the probability of such reactions is extremely small. There is a lot of nuclear activity in the sun, and a huge number of neutrinos are emitted which bathe the earth. It was expected that this large flux would permit the neutrino to be detected, and would even throw light on nuclear processes in the sun. Disappointingly, very few such inverse neutrino reactions were observed, and it followed that the neutrino flux from the sun was much less than expected. It is now thought that neutrinos from the sun are not pure electron neutrinos, which would cause the reactions, but in a state that mixes them with mu or tau neutrinos, which would not. As the neutrinos move, they oscillate back and forth between being electron neutrinos and mu neutrinos. When they get to the earth, they are in their mu phases, and so do not cause reactions. This was a great relief to solar theorists.

A few decades ago, the electromagnetic and weak forces were shown to be aspects of the same force, the unified electro-weak force, mediated by a heavy vector boson (particle of integral spin) as well as the massless photon. All the fermions (particles of half-integral spin) subscribe to the electro-weak force, which is responsible for beta decay as well as the electromagnetic field. This seems to me to be the last significant progress in physical theory, which has now gone off into speculations that are interesting but not conclusive, and some of which are not even provable. It would be very satisfying if some of this activity explained even the least of the mysteries that still surround the matters considered in this article, but it seems totally detached from reality.

Particle accelerators are an important application of the classical point electron model. The fields of a charge in arbitrary motion are studied in Relativistic Electrodynamics and the Field of a Point Charge, and the electron synchrotron is described there. The first machine accelerating electrons to high energies was the betatron, proposed by D. W. Kerst in 1940, and put into operation at General Electric in 1945, producing 100 MeV electrons.

The betatron accelerates electrons without the use of a static electric field. Instead, the accelerating field is produced by a changing magnetic field that also serves to maintain the electrons in a circular orbit of fixed radius as they are accelerated. The electric field can be found from Faraday's Law, curl E = -(1/c)dB/dt, which in integral form is ∫E·ds = -(1/c)∫B·dA = -(1/c)dΦ/dt. With cylindrical symmetry, this is 2πrE = -(1/c)dΦ/dt. The tangential force on the electron is -eE = (e/2πrc)dΦ/dt, which equals the rate of change of tangential momentum p = γmv, where γ = (1 - v2/c2)-1/2. Integrating from t = 0, p = 0, Φ = 0 to t, we find that eΦ/2πrc = γmv. This is the first expression for the momentum, coming from Faraday's Law.

The magnetic field also holds the electron in a circular orbit, and equating the magnetic force to the centripetal acceleration times the mass gives γmv2/r = eBv/c, or γmv = reB/c. Note that γ does not change in circular motion, since v is constant, so we can do this. This is a second expression for the momentum, coming from Newton's Law. If we equate our two expressions for the momentum γmv, we find eΦ/2πrc = reB/c, or B = Φ/2πr2. This is the condition that the magnetic field at the orbit, B, will be just right to keep the accelerating electrons in an orbit of fixed radius. Since the average value of the magnetic field inside the orbit is Φ/πr2 = , we have = 2B, the famous 1:2 condition. The average field inside the orbit must be twice the value of the field at the orbit. This can be arranged by properly shaping the pole faces, as shown in the diagram. It is a good exercise to verfy that the sense of revolution in the orbit is consistent with the magnetic field, and that the induced electric field is in the correct direction to accelerate the electrons when the flux increases. Don't forget that the electronic charge is negative.

Electrons injected into the vacuum chamber may have orbits that rise above and below the midplane, and which oscillate radially as well. The magnetic field must be shaped so that these betatron oscillations are stable. The betatron introduced these considerations into accelerator design. As the magnetic field increases during acceleration, the oscillations are damped and a well-defined electron beam is produced. The maximum energy for a betatron is about 500 MeV. The electrons are allowed to strike a target at the end of each cycle of acceleration, where they produce powerful X-rays. In fact, the main use of betatrons is as an X-ray source. The electrons also radiate as they revolve in their orbits, producing synchrotron radiation. This radiation is treated at some length in the article on Relativistic Electrodynamics.

The magnet core of the General Electric betatron of 1945 weighed 130 tons. The large amount of iron required is one of the principal disadvantages of betatrons, especially in large sizes. The electron synchrotron overcame this disadvantage, since its magnets have only a guidance function.

While considering the magnificent contributions to science made in the past by organizations like General Electric, Bell Labs and Radio Corporation of America, it is sad to realize that the present-day successors will never again have the skills and intellectual resources to be able to do the like.

H. E. White, Introduction to Atomic Spectra (New York: McGraw-Hill, 1934). A classic introduction to the vector model and atomic spectra. See especially chapters 8 and 10.

J. D. Jackson, Classical Electrodynamics, 2nd ed. (New York: John Wiley & Sons, 1975). Section 11.8, pp. 541-547, explains the Thomas precession thoroughly. There is also an excellent account of Thomson scattering.

I. Kaplan, Nuclear Physics (Cambridge, MA: Addison-Wesley, 1955). A very accessible account of beta-decay is given in Chapter 14. Chapter 15 treats the passage of gamma-rays through matter, including Compton scattering. The algebra is given in Chapter 6. This excellent text, full of basic information, holds its value even today.