]>

Spatial Dispersion and Warm Plasma*

10
Spatial Dispersion and Warm Plasma
*

In Chapter 9, we discussed wave propagation in cold isotropic plasma. The effects of dispersion, to be more precise, temporal dispersion, are emphasized. In literature, cold plasma is sometimes referred to as “temperate plasma” [1,2]. The permittivity of warm plasma, in addition to being the function of the frequency ω, is also a function of the wave number k. For this reason, the warm plasma is said to be spatially as well as temporally dispersive [1,2,3]. Metal at optical frequencies can be modeled as warm plasma [4], the Fermi velocity playing the role of the thermal velocity of the warm plasma. While the model based on the cold plasma is said to represent local optics, the warm plasma model represents nonlocal optics. Section 10.7 gives rudimentary explanation of the technical terms used in describing the plasma state and the references give a deeper exposure.

10.1Waves in a Compressible Gas

A compressible gas satisfies the equation of state

(10.1)P=kBNT,
P=kBNT,

where P is the pressure, N the number density, T the temperature, and kB Boltzmann’s constant (1.38 × 10−23 J/K).

Let

(10.2)P=P0+p,
P=P0+p,
(10.3)N=N0+n.
N=N0+n.

The values with subscript zero are the average values and the lower-case ones are a.c. values.

For an isothermal process (T constant), from Equation 10.1, Equation 10.2, Equation 10.3,

P0=kBN0T,
P0=kBN0T,
p=kBnT.
p=kBnT.

For acoustic waves, the adiabatic process where no heat transfer is taking place is more appropriate; P and N satisfy the relation

(10.4)PNγ=P0Nγ0,
PNγ=P0Nγ0,

where γ is the ratio of specific heats.

The value of γ depends on the type of gas (monatomic, diatomic, or polyatomic). For a one-dimensional plasma problem, γ = 3 is appropriate.

From Equation 10.2, Equation 10.3, Equation 10.4, we have

(10.5)p=γkBnT.
p=γkBnT.

The hydrodynamic equation (Navier–Stokes equation) is

(10.6)mdvdt=1Np,
mdvdt=1Np,

where m is the mass of the particle.

The continuity equation is

(10.7)[Nv]+Nt=0.
[Nv]+Nt=0.

The RHS of Equation 10.6 is nonzero in a compressible gas.

For small-signal a.c. values, after linearizing, Equations 10.6 and 10.7 are approximated as

(10.8)mdvdt=1N0p,
mdvdt=1N0p,
(10.9)N0v=nt.
N0v=nt.

Taking the divergence of Equation 10.8, we obtain

(10.10)(v)t=2pN0m.
(v)t=2pN0m.

From Equations 10.9 and 10.5, we obtain

(10.11)v=1N0γkBTpt.
v=1N0γkBTpt.

Substituting Equation 10.11 into Equation 10.10, we obtain the wave equation for the pressure:

(10.12)2p1a22pt2=0,
2p1a22pt2=0,

where the acoustic velocity a is given as

(10.13)a=(γkBTm)1/2.
a=(γkBTm)1/2.

10.2Waves in Warm Plasma

Let us consider the case where the gas has charged particles. In Chapter 9, we considered free electron gas, called cold plasma, and the force equation was based on the Lorentz force:

(10.14)mdvdt=qE.
mdvdt=qE.

In the cold plasma case, we ignored the term on the RHS of Equation 10.6 assuming that it is negligible compared to the Lorentz force. The assumption is valid as long as the temperature of the gas is low and the gas is considered as incompressible. In the warm plasma case, we consider the contributions to the force due to pressure gradients and we modify Equation 10.6 for the warm plasma case as

(10.15)mdvdt=qE1Np.
mdvdt=qE1Np.

Taking the divergence of Equation 10.15 (and approximating d/dt → ∂/∂t),

(10.16)(v)t=qm(E)1Nm2p.
(v)t=qm(E)1Nm2p.

Using Equation 10.11 in Equation 10.16, we obtain the wave equation for the pressure:

(10.17)2p1a22pt2+N0qE=0.
2p1a22pt2+N0qE=0.

The variable E in Equation 10.17 can be eliminated by using the Maxwell equation

(10.18)×H=ε0EtN0qv,
×H=ε0EtN0qv,
(10.19)0=(×H)=ε0(E)tN0q(v),0=ε0(E)t+qγkBTpt,E=qγkBTε0p,N0qE=N0q2γkBTε0p=N0q2mε0mγkBTp,N0qE=ω2pa2p.
00EN0qEN0qE=====(×H)=ε0(E)tN0q(v),ε0(E)t+qγkBTpt,qγkBTε0p,N0q2γkBTε0p=N0q2mε0mγkBTp,ω2pa2p.

Thus, the wave equation for the pressure variable p for the warm plasma case is given by

(10.20)2p1a22pt2ω2pa2p=0.
2p1a22pt2ω2pa2p=0.

To obtain the wave equation for the electric field, we start with the Maxwell equation

(10.21)ˉ×E=μ0Ht
¯¯¯×E=μ0Ht

and take the curl on both sides

××E=μ0(×H)t
××E=μ0(×H)t

and from Equation 10.18, we obtain

××E=μ0t(ε0EtN0qv),
××E=μ0t(ε0EtN0qv),

and from Equation 10.15

××Eμ0ε02Et2+N0μ0q(qmE1N0mp)=0,××E1c22Et2μ0ε0ω2pEμ0qmp=0.
××Eμ0ε02Et2+N0μ0q(qmE1N0mp)=0,××E1c22Et2μ0ε0ω2pEμ0qmp=0.

Using Equation 10.19 to eliminate p from the above,

××E1c22Et2ω2pc2E+μ0qmˉ(N0qa2ω2pE)=0.
××E1c22Et2ω2pc2E+μ0qm¯¯¯(N0qa2ω2pE)=0.

By simplifying, we obtain the wave equation for E as

(10.22)××E1c22Et2ω2pc2E+μ0ε0a2(E)=0.
××E1c22Et2ω2pc2E+μ0ε0a2(E)=0.

We should be able to recover the wave equation for E for the cold plasma case by substituting a = 0 in Equation 10.22:

××E1c22Et2ω2pc2E=0,(E)+2E1c22Et2ω2pc2E=0.
××E1c22Et2ω2pc2E=0,(E)+2E1c22Et2ω2pc2E=0.

For the cold homogeneous plasma case (incompressible gas), p = 0 and ∇ · E = 0, and hence we obtain

(10.23)2ˉE1c22ˉEt2ω2pc2ˉE=0.
2E¯¯¯1c22E¯¯¯t2ω2pc2E¯¯¯=0.

Let us solve Equation 10.22 for a harmonic wave with a phase factor et e−jk·r. Noting that ∇ can be replaced by (−jk), for such a phase factor, Equation 10.22 becomes

(10.24)(jk)×(jk)×E+ω2c2Eω2pc2E+a2c2(jk)(jkE)=0,k×k×E+ω2c2Eω2pc2Ea2c2k(kE)=0.
(jk)×(jk)×E+ω2c2Eω2pc2E+a2c2(jk)(jkE)=0,k×k×E+ω2c2Eω2pc2Ea2c2k(kE)=0.

Since k × k × E = k(k · E) − k2E, we can write Equation 10.24 as

(10.25)k(kE)(1a2c2)[k2k20(1ω2pω2)]E=0.
k(kE)(1a2c2)[k2k20(1ω2pω2)]E=0.

Equation 10.24 can be written separately for E and E, where E is a component parallel to k and E is normal to k.

For the parallel component k ⋅ E = kE, we have

(10.26)k2E(1a2c2)[k2k20(1ω2pω2)]E=0,k2a2c2+k20(1ω2pω2)=0,k2=k20(1(ω2p/ω2))a2/c2=ω2ω2pa2.
k2E(1a2c2)[k2k20(1ω2pω2)]E=0,k2a2c2+k20(1ω2pω2)=0,k2=k20(1(ω2p/ω2))a2/c2=ω2ω2pa2.

Let

(10.27)ac=δ,
ac=δ,
(10.28)ω2pω2=X,
ω2pω2=X,

Then

(10.29)k2=k201Xδ2.
k2=k201Xδ2.

From Equation 10.21, we obtain

(10.30)H=1μ0jωk×E,
H=1μ0jωk×E,

and since k and E are parallel for this case

(10.31)H=0.
H=0.

The parallel mode under discussion has an electric field component in the direction of propagation and the zero magnetic field. It is also called an electron plasma wave. It is an acoustic wave, like a sound wave, modified by the pressure of a charged compressible fluid. Its phase velocity is given by

(10.32)Vp=ωk=a1X=cδ1X.
Vp=ωk=a1X=cδ1X.

The ω−k diagram for the electron plasma wave is shown in Figure 10.1.

image

FIGURE 10.1
ω−k diagram for the electron plasma wave.

Note that the phase velocity a becomes negligible for the cold plasma approximation and the electron plasma wave becomes a plasma oscillation at the plasma frequency ωp. A simple way of interpreting the electron plasma wave is to say that the plasma oscillation becomes a longitudinal electron plasma wave when the plasma is considered as warm.

The electron plasma wave propagates for only ω > ωp. For ω < ωp, this wave is evanescent. Now let us examine the wave propagation when E is perpendicular to k.

From Equation 10.24,

(10.33)[k2k20(1ω2pω2)]E=0,
[k2k20(1ω2pω2)]E=0,
(10.34)k2=k20(1ω2pω2).
k2=k20(1ω2pω2).

Equation 10.34 is the same as the dispersion relation for the cold isotropic plasma, discussed in Chapter 9. One can obtain H from Equation 10.30, which shows that

(10.35)E=ηpH,
E=ηpH,

where

(10.36)ηp=η01X.
ηp=η01X.

While the electron plasma wave and TEM wave can exist independently in an unbounded warm plasma, the amplitudes of the two will be determined in a bounded warm plasma by the boundary conditions. Appendix 10A discusses such a problem. Appendices 10B and 10C deal with such problems for the cases of warm magnetoplasma media.

10.3Constitutive Relation for a Lossy Warm Plasma

For a lossy warm plasma, Equation 10.15 will be modified as

(10.37)mdvdt+mνv=qE1Np.
mdvdt+mνv=qE1Np.

Since

(10.38)J=N0qv,
J=N0qv,

Equation 10.37 can be written as (multiply Equation 10.37 by −N0q/m)

dJdt+νJ=ε0ω2pE+qmp,p=γkBTn.
dJdt+νJ=ε0ω2pE+qmp,p=γkBTn.

Since

ε0E=ρν=qn,n=ε0q(E),p=γkBTε0q(E),qmp=γkBTε0m(E)=ε0a2(E).
ε0Enpqmp====ρν=qn,ε0q(E),γkBTε0q(E),γkBTε0m(E)=ε0a2(E).

Thus, the constitutive relation is given by

(10.39)dJdt+νJ=ε0ω2pEε0a2(E).
dJdt+νJ=ε0ω2pEε0a2(E).

One can model the warm plasma as a dielectric by using Equation 10.39 and the Maxwell equation

(10.40)×H=J+ε0Et.
×H=J+ε0Et.

For harmonic variations in space and time ej(ωtk·r), Equation 10.39 becomes

j(ω+v)J=ε0ω2pEε0a2(jk)(jkE).
j(ω+v)J=ε0ω2pEε0a2(jk)(jkE).

Since

J=J+J(withreferencetok),
J=J+J(withreferencetok),

(10.41)E=E+E,(jω+ν)J=ε0ω2pE+ε0a2k2E,
E=E+E,(jω+ν)J=ε0ω2pE+ε0a2k2E,
(10.42)(jω+ν)J=ε0ω2pE+0.
(jω+ν)J=ε0ω2pE+0.
(10.43)(ˉ×ˉH)=J+jωε0E=(ε0ω2pjω+ν+jωε0)E,(ˉ×ˉH)=jωεpE,
(¯¯¯×H¯¯¯)=J+jωε0E=(ε0ω2pjω+ν+jωε0)E,(¯¯¯×H¯¯¯)=jωεpE,

where

(10.44)εp=1ω2pω(ωjν).
εp=1ω2pω(ωjν).

Let us derive next an expression for εp∥.

From Equation 10.41,

(10.45)J=ε0ω2pjω+νE+ε0a2k2jω+νE,(ˉ×ˉH)=J+jωε0E=ε0ω2pjω+νE+ε0a2k2jω+νE+jωε0E,(ˉ×ˉH)=jωε0εpE,
J(¯¯¯×H¯¯¯)(¯¯¯×H¯¯¯)====ε0ω2pjω+νE+ε0a2k2jω+νE,J+jωε0Eε0ω2pjω+νE+ε0a2k2jω+νE+jωε0E,jωε0εpE,

where

(10.46)εp=1ω2pω(ωjν)a2k2ω(ωjν).

If we neglect collisions, then

(10.47)εp=1ω2p+a2k2ω2.

Substituting Equation 10.26 for k, we obtain

(10.48)εp=0.

This will be true even if collisions are taken into account.

Equation 10.47 implies that E can be nonzero even if D = 0.

The dispersion relation Equation 10.26 and the ω−k diagram in Figure 10.1 can be obtained by imposing Equation 10.47.

10.4Dielectric Model of Warm Loss-Free Plasma

Maxwell’s equations and the constitutive relations for warm plasma, viewed as a dielectric medium, are given below:

(10.49)×E=Bt,
(10.50)×H=Dt,
(10.51)D=0,
(10.52)B=0,
(10.53)B=μ0H,
(10.54)D=η0ˉεpE,
(10.55)ˉεp=[εp00εp],

where

(10.56)εp=1ω2pω2,
(10.57)εp=1ω2pω2a2k2.

Note that Equations 10.47 and 10.57 are the same when εp∥ = 0.

10.5Conductor Model of Warm Lossy Plasma

Maxwell’s equations in the time domain and the constitutive relations for a warm lossy plasma, viewed as a conducting medium, are given below:

(10.58)×E=μ0Ht,
(10.59)×H=J+ε0Et,
(10.60)H=0,
(10.61)dJdt+νJ=ε0ω2pEε0a2(E).

Note that

(10.62)ε0E=ρ,
(10.63)ρ=qn,
(10.64)p=γkTn.

In Equations 10.62 and 10.63, ρ is the volume charge density.

For time-harmonic fields, Equation 10.61 can be written as

(10.65)J=ˉσE,
(10.66)ˉσ=[σ00σ],
(10.67)σ=jε0ω2pωjν,
(10.68)σ=jωε0ω2pω(ωjν)ak2.

By noting

(10.69)ˉεp=ˉI+ˉσjωε0,

we can obtain the conductivity tensor ˉσ from ˉεp and vice versa.

10.6Spatial Dispersion and Nonlocal Metal Optics

The longitudinal wave in warm plasma has the dispersion relation

(10.70)εp=0=1ω2pω(ωjν)a2k2

and the ω−k diagram is sketched in Figure 10.1. In the sense that

(10.71)εp=εp(ω,k),

the warm plasma is spatially as well as temporally dispersive.

A simple metal, such as sodium, is modeled as a lossy plasma in Section 8.3. Typical values for sodium are ωp = 8.2 × 1015 rad/s and ν/ωp = 3 × 10−3. The plasma frequency is in the optical range. Due to the high electron density in metals, it can be shown that a parameter called Fermi velocity vf is also important in modeling metal as plasma [5]. A typical value of vf for sodium is vf = 9.8 × 105 m/s. The Fermi velocity vf plays the role of the acoustic velocity in warm plasma. Thus, the modeling of metal for studying the interaction of thin metal films with optical waves with frequencies near the plasma frequency require a warm plasma model, which takes into account the spatial dispersion due to the excitation of the electron plasma wave in the metal, when the source wave is a p wave. Spatial dispersion will have important consequences and such a study is labeled as the study of nonlocal optics in contrast to the local optics (Fresnel Optics). A comprehensive account of the nonlocal optical studies using the warm plasma model is given by Forstmann and Gerhardts [4].

10.7Technical Definition of Plasma State

We have earlier given a vague definition of plasma medium as an ionized medium with positive ions and electrons with an “overall” charge neutrality. We also used the terms cold plasma and warm plasma. While reading books on plasma science, we also come across terms such as collective behavior, temperate plasma, hot plasma, nonneutral plasma, Debye length and plasma sheath, and so on. This section is used to explain the various terms quantitatively in the context of the theory discussed in Sections 8.3 and 8.4, Chapter 9, and this chapter. The one-dimensional electron–proton plasma with one degree of freedom is assumed.

10.7.1Temperate Plasma

In cold plasma, we neglect the effect of the temperature. The thermal velocity Vth is related to temperature T by

(10.72)12mV2th=12kBT.

The RHS of Equation 10.72 is one-half of an electron volt (eV) and substituting for kB, the Boltzmann constant, we obtain the equivalence

(10.73)1eV=11000°K.

In linearizing the Maxwell equations and the force equation, we assume that the electron velocity v is small compared to Vth. The cold plasma approximation neglects the spatial dispersion, which will be a valid approximation if Vth is less than the phase velocity Vph = ω/k of the wave. So the cold plasma approximation is valid, if [1]

(10.74)vVthVph.

Thus, vVth is implied by calling cold plasma as temperate plasma.

A plasma is called a warm plasma if the temperature is such that Vth is approximately equals Vph of the electron plasma wave.

10.7.2Debye Length, Collective Behavior, and Overall Charge Neutrality

A plasma medium shields out any excessive fields in the medium. A characteristic distance, called “Debye length” λD, is the length-scale parameter, which ensures the shielding of the excessive electric fields in plasma [6]. The Debye length depends on the temperature and the electron density and is given by

(10.75)λD=(ε0kBTNq2)1/2=aγωp,
(10.76)λD=69(TN)1/2.

It can be shown [6] that any excess potential Φ in the plasma will decay as

(10.77)Φ=Φ0e(|z|/λD).

For |z| > 5λD, Φ = Φ0e−5 ≈ 0, and hence the excess potential is screened out in about five times λD.

If the plasma system dimension L is much larger than λD, the bulk of the plasma has a charge neutrality, that is, the net charge in a macroscopic volume is zero:

(10.78)NeNiN.

In Equation 10.78, Ni is the ion density and N is often referred to as the plasma density. The qualifying word “overall” before the word neutrality is used to indicate that the plasma is not so neutral to blot out all interesting electromagnetic forces. This shielding becomes possible, because the electric Coulomb force is inversely proportional to the square of the distance. This force is considered long range compared to the other short-range forces. If we assume that the electron is a point, then the charges in a volume (4π/3)λ3D collectively shield the effect of this electron. It is this collective behavior, in contrast to the collision between single particles, that control the motion of the charged particles. The collective behavior can occur only when the particle density is such that there are enough particles in a Debye volume.

10.7.3Unneutralized Plasma

The phenomenon of Debye shielding in a modified form can occur in an electron stream in klystron, and a proton beam in a cyclotron. These are not strictly plasmas but because they use the concepts of collective behavior, and shielding and use the mathematical tools of plasma physics, they are referred to as unneutralized plasmas [7].

References

  1. 1.Allis, W.P., Buchsbaum, S. J., and Bens, A., Waves in Anisotropic Plasmas, MIT Press, Cambridge, MA, 1963.
  2. 2.Heald, M.A. and Wharton, C. B., Plasma Diagnostics with Microwaves, Wiley, New York, 1965.
  3. 3.Akira, I., Electromagnetic Wave Propagation, Radiation, and Scattering, Prentice Hall, Englewood Cliffs, NJ, 1991.
  4. 4.Forstmann, F. and Gerhardts, R.R., Metal Optics Near the Plasma Frequency, Springer, New York, 1986.
  5. 5.Kittel, C., Introduction to Solid State Physics, Wiley, New York, 2005.
  6. 6.Chen, F.F., Introduction to Plasma Physics and Controlled Fusion, Vol.1: Plasma Physics, Springer, New York, 1983.
  7. 7.Davidson, R.C., Physics of Nonneutral Plasmas, Addison-Wesley, New York, 1990.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.139.104.214