Plasma Electronics Applications in Microelectronic Device Fabrication
Plasma Electronics Applications in Microelectronic Device Fabrication
Plasma Electronics Applications in Microelectronic Device Fabrication
Applications in Microelectronic
Device Fabrication
Plasma Electronics:
Applications in Microelectronic
Device Fabrication
T Makabe
Keio University, Japan
Z Petrovic´
Institute of Physics Belgrade, Serbia
Makabe, T. (Toshiaki).
Plasma electronics : applications in microelectronic device fabrication / by T. Makabe and I. Petrovic
p. cm. -- (Series in plasma physics)
Includes bibliographical references and index.
ISBN 0-7503-0976-8
1. Plasma engineering. I. Title. II. Series.
TA2020.M35 2005
621.044--dc22 2005056888
Over the past three decades low-temperature plasma applications have been
extended from primarily lighting to the fabrication of microelectronic devices
and new materials, far exceeding our expectations. Radio frequency plasmas
ranging from 105 Hz to 109 Hz, in particular, are now used to process metal-
lic, semiconductor, and dielectric materials for the fabrication of ultra large-
scale integrated (ULSI) circuits and to deposit various kinds of functional
thin films and to modify the surface properties. Without plasma-produced
ions and dissociated neutral radicals for etching and deposition on wafers,
microelectronics manufacturing for ULSI circuit would simply be unfeasible.
The advent of ULSI fabrication has greatly changed how the field of plasma
science is approached and understood. Low-temperature non-equilibrium
plasmas are sustained mainly by electron impact ionization of a feed gas
driven by an external radio frequency power source. These low-temperature
plasmas acquire characteristics intrinsic to the feed gas molecules as deter-
mined by their unique electron-collision cross-section sets. This uniqueness
means that plasmas must be understood using quantum, atomic, and molecu-
lar physics. The disparate time and spatial scales involved in low-temperature
plasma processing (submeter to nanometer and seconds to nanoseconds)
makes plasma processing an inherently stiff problem. The characteristics of
low-temperature plasmas contrast markedly with highly ionized equilibrium
plasmas that are ensembles of charged particles whose behavior can be un-
derstood through their long-range Coulomb interactions and collective effects
and characterized by plasma frequency, Debye length, and electron tempera-
ture. The fundamental collision and reaction processes occurring both in gas
phase and on surface in low-temperature plasmas are the bases for under-
standing their behavior and exploiting them for practical applications.
In the emerging nanotechnology era, device design, reliability, and the
design of integrated plasma processes for their fabrication are tightly cou-
pled. Being able to predict feature profile evolution under the influence of
spatio-temporally varying plasmas is indispensable for the progress of nano-
technology. Prediction of plasma damage and its mitigation are also crucial
prediction & mitigation are 2 process and will be performed through a series
of vertically integrated numerical modeling and simulations ranging from the
reactor scale to those cognizant of device elements subtending the plasma.
Motivated by the important role of plasmas in technology and the need for
simulations to understand the associated complex processes, we emphasize
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Plasma and Its Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Application of Low-Temperature Plasma . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Academic Fusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
FIGURE 1.1
Electron trajectory in three different media. In strongly ionized plasma (a), in weakly ionized
plasma (b), and in neutral gas (c).
Pressure (Pa)
1.0 101 102 103 104 105
1013
µ -plasma
PDP
1012 ICP ICP
ECR MEMS
Plasma density (cm-3)
FIGURE 1.2
Low-temperature plasma and material processes.
Passivation
M6
M5 Global
interconnect
(up to five layers)
M4 Via
M3
Cu
Low-k
Intermediate
interconnect
M2 (up to four layers)
M1 Local
interconnect
SiO2
n+ n+ STI p+ p+
p-well n-well
nMOS pMOS
substrate
FIGURE 1.3
Typical large scale integrated circuits.
Low-Temperature
Plasma and
Its Technologies
FIGURE 1.4
Academic fusion for low-temperature plasmas.
1990s [3]. Further achievements in plasma technology at the design level will
lead to greater miniaturization of ultra-large-scale integrated (ULSI) circuits
in microelectronics and nanoelectronics and to the further functionalization of
new materials. These developments will be synchronized with the progress of
computers utilizing high-speed and high-performance ULSI chips. We stand
at the advent of the design era of sophisticated plasmas based on atomic and
molecular physics both in the gas phase and in the surface phase.
The low-temperature plasma and related technologies will join the excit-
ing and practical fields of academic fusion reconstructed by computational
science with quantum atomic and molecular physics, surface physics and
chemistry, Boltzmann equation of particles, and Maxwell’s electromagnetic
theory as is shown in Figure 1.4.
To address the design issue, it is essential to prepare a two- or three-
dimensional image of the plasma structure in a reactor and feature profile on
a substrate surface by experimental observations or computational modeling.
References
1. Goldston R.J. and Rutherford P.H. 1995. Introduction to Plasma Physics. Bristol:
IOP.
2. American Institute of Physics. 2003. 50 years of science, technology, and the AVS
(1953–2003), J. Vac. Sci. Technol. A (Special Issue) 21:5.
3. Makabe, T., Ed. 2002. Advances in Low Temperature RF Plasmas, Basis for Process
Design. Amsterdam: Elsevier.
n(r) I E
II
III z
0
r
FIGURE 2.1
Spatial profile of electron density for a swarm released from a single point in space and allowed
to develop under the influence of electric field and collisions with background molecules.
∂n
Γ = nv0 = nvd − D k, (2.1)
∂z
where m and e(> 0) are the mass and charge of the electron, respectively. In
the stationary state, dvz /dt = 0, the total differential on the left-hand side
of Equation 2.2 is separated into two parts, one of which is zero. Applying
0 = (−eE − mvd Rm ) n,
∂n
0 = mvz2 − mDRm . (2.3)
∂z
It is justified (as shown later) to assume that the z component of the velocity
is not substantially perturbed by the electric field-induced drift, and there-
fore 13 v 2 = vx2 = v 2y ∼ vz2 . In that case we may simplify the second part of
Equation 2.3 to give the approximate value of the diffusion coefficient as
eE
vd = − . (2.4)
mRm
v2
D= . (2.5)
3Rm
PROBLEM 2.1.1
Derive Equations 2.4 and 2.5.
e2 E2 2e Eεm 1 ∂n 2m
− − εm Rm = 0. (2.7)
mRm 3mRm n ∂z M
We consider solving this equation under certain conditions, particularly in
regard to the spatial profile of the electron number density n(z).
1. ∂n/∂z = 0: This corresponds to the point I in Figure 2.1, that is,
to the maximum of the density distribution. From Equation 2.7 we
then obtain
M eE 2
εm0 = εm | ∂n
∂z =0
= . (2.8)
2m2 Rm
2. ∂n/∂z =
0: This corresponds to all points except for I in Figure 2.1.
For the region with a negative density gradient (region II in
Figure 2.1) the mean energy is higher than that at the peak (point I):
εm = εm0 +
ε.
So far we have assumed that the collisional rate was independent of the
electron energy. However, this is normally not the case. Thus we use a Taylor
expansion around the mean energy at the peak I for the collision rate:
∂ Rm
Rm (εm ) = Rm0 + |0
ε,
∂ε
which in combination with Equations 2.7 gives
2ε 1 ∂n
εm = εm0 1 − .
m0
(2.9)
Rm εm0 n ∂z
3eE 1 + 2 ∂∂ε |0 Rm0
Exercise 2.1.1
Discuss the value of the mean electron energy in two cases, (a) ∂n/∂z < 0
and (b) ∂n/∂z > 0, as compared with the mean energy at the peak of the
swarm εm0 . For case (a), the gradient of the electron number density shown
in Figure 2.1 is negative as in region II, and therefore εm > εm0 . However,
for case (b), the gradient of the electron number density is positive (see region
III in Figure 2.1) and therefore εm < εm0 . Thus, unless the collision rate Rm
changes very rapidly, we may expect the mean energy of electrons to increase
toward the front of the swarm where flux due to diffusion adds to the drift-
induced component. The spatial profile of energy may have consequences on
the transport coefficients, provided that some energy-dependent processes
also exist, which is generally the case.
PROBLEM 2.1.2
Derive expression 2.9 for the mean electron energy εm .
where
εm0 ∂Rm
ξ= 0. (2.11)
Rm0 ∂ε
As a result of the influence of the electric field, the diffusion along the
z-axis (direction of the field) has special properties and is in general
different from the diffusion in perpendicular directions. The diffusion of
electrons (and charged particles in general) becomes anisotropic with two
components of the diffusion tensor D: DL (longitudinal diffusion coefficient)
and DT (transverse diffusion coefficient). These two coefficients may be
given as
ξ
DL = D0 1 − , (2.12)
1 + 2ξ
and
v2
DT = D0 = , (2.13)
3Rm
which are valid only for approximations that were involved in the derivation
of the formulae. In the case where the collision rate Rm is independent of the
mean energy, the diffusion will be isotropic. The same is true for the zero
electric field.
In this section we have seen that the transport of electrons (the same is
true for ions) may be described with the aid of transport coefficients. We
have given approximate relations for the drift velocity and components of
the diffusion tensor. In general these properties will depend on the mean
electron energy. However, it is not practical to use mean energy for tabulating
the data for transport coefficients. It is better to use an external parameter
associated with the value of the external electric field E. The reduced electric
field is defined as the ratio between the magnitude of the electric field E
and the neutral gas number density N, that is, E/N. It is shown that the
reduced field is proportional to the energy gained by electrons between two
collisions. E/N is expressed in units known as Townsend (in honor of the
founder of gaseous electronics), which is defined as 1 Td = 10−21 Vm2 =
10−17 Vcm2 .
PROBLEM 2.1.3
By substituting Equation 2.9 into Equation 2.6, derive the longitudinal DL
(Equation 2.12) and transverse DT (Equation 2.13) coefficients. Assuming analytic
dependence of the collision rate Rm on energy find the ratio DL /DT in analytic form.
Discuss the conditions when the ratio is equal to, less than, or greater than 1. (A hint:
choosing Rm ∼ (εm /εm0 )(
+1)/2 is particularly convenient.)
PROBLEM 2.1.4
Pressure p may be used to represent the gas number density N; find the relationship
between E/N and E/ p given that
m 3/2
mv 2
g M (v) = exp − . (2.14)
2π kT 2kT
At the same time, we may define the velocity distribution function g(v, r, t),
which describes swarm development with respect to both velocity and real
space and with respect to time.
g( )
vr
0
vz
FIGURE 2.2
Typical electron velocity distribution g(v) as a function of axial vz and radial vr velocities under
a DC electric field.
PROBLEM 2.2.1
The drift velocity and mean energy of electrons in argon under the uniform field,
E = 100 Vcm−1 , and at pressure of 1 Torr (at temperature of 300 K) are vd = 2.3×107
cms−1 and 8 eV, respectively. Show, then, that the mean random velocity vr and drift
velocity vd satisfy the relation (vd /vr ) 1.
vr
E
v v'
θ
vz
0 0'
∆V
FIGURE 2.3
Contour plot of electron velocity distribution g(v) as a function of axial vz and radial vr velocities.
When the electric field is turned on, a small component of velocity
V is added to all electrons
in the direction opposite to the field E.
In Figure 2.3 we show the effect of the electric field on the velocity distribu-
tion g(v) at each velocity v with and without the field. When the electric field
E is turned on, a small component of velocity
V is added to all electrons in
the direction opposite to the field. In the simplified model the circular contour
representing random (thermal) isotropic motion of electrons is slightly shifted
as if its center were moved to the position O from O. The resulting velocity
is v , which is given by: v − v ∼
=
V cos θ (where θ is the angle between v
and −E). We can now estimate the effect of the electric field on the electron
velocity distribution as
dg0 (v)
g(v ) = g(v +
V cos θ) ∼
= g0 (v) +
V cos θ , (2.15)
dv
where we applied the condition | v |
|
Vcos θ |. The distribution g(v ) is
composed of an isotropic part in all directions and having, in general, a non-
Maxwellian term and a small anisotropic term. Both terms are influenced
by the electric field. This expansion is analogous to a two-term expansion in
spherical harmonics of the velocity distribution, where the second term is the
anisotropic term (see Chapter 5).
Now we consider the momentum conservation of electrons with velocity
distribution g(v) in gases under an electric field. First, let us define Qm (v)
as the momentum transfer cross section of the electron by elastic scattering
with the background gas molecule with number density N. Then, for one
group of electrons with velocity v, the collision frequency of the momentum
exchange is NQm (v)v, and the total momentum dissipated in collisions in the
direction opposite to the field E is mv cos θ NQm (v)v. The total momentum
exchange, therefore, is obtained by averaging over all velocities by using the
This is an integral equation that is satisfied for an arbitrary field E and cross
sections Qm (v) only if
V = − [eE/mNQm (v)v]. In this case we write the
expansion for g(v) as
The second term in Equation 2.18 shows the effect of an external field on
the velocity distribution. The cosine term is the lowest-order development of
the actual distribution in the velocity space in addition to the symmetric g0 (v)
term. Usually the cross section for the momentum transfer of elastic scattering
is much greater than the sum of cross sections for other inelastic collisions.
So far we have taken into account only elastic collisions. The energy balance
of electrons will consist of energy loss in such collisions, as we mentioned
before, 2mε/M. Therefore, the total energy loss in elastic collisions in unit
time is expressed as
2m 1 2
mv NQm (v)vg(v)dv.
M2
The energy gain, on the other hand, is determined by the motion of elec-
trons along the electric field axis:
After integration over the polar angles θ and ϕ and by equating the terms
inside the integrals we obtain the following first-order differential equation:
m 4π eE dg0 (v)
4π mv 2 NQm (v)vg0 (v) = − eEv .
M 3 mNQm (v)v dv
The solution will be obtained by direct integration without making any
assumptions on the velocity dependence of the cross section Qm (v):
3m v NQm (v)v 2
g0 (v) = A exp − vdv , (2.20)
M 0 eE/m
where A should be determined from the normalization g0 (v)dv = 1. Now
we perform the conversion to electron energy distribution f (ε):
1 2
g0 (v)4πv 2 dv = f (ε)dε; mv = ε. (2.21)
2
The Maxwellian distribution in Equation 2.14 in thermal equilibrium is rewrit-
ten as a function of energy :
√ 3ε
f M (ε) = A ε exp − , (2.22)
2ε
and the mean energy is given by ε = 3kTe /2. In a more general case when
there is an electric field but for a constant cross section Qm , we obtain from
Equation 2.20 the so-called Druyvestyn distribution:
2
√ ε
f D (ε) = A ε exp −0.548 . (2.23)
ε
Exercise 2.2.1
Plot the distributions f (), f M (), and f D () for electrons in Ar at 100 Td. Here
the mean energy < > is 6.47 eV.
Figure 2.4 shows the Maxwellian, Druyvestyn, and nonequilibrium ( f 0 )
energy distributions for electrons at the same mean energy. Note the dif-
ferences among the three distributions.
f0
0.10 100 Td
fD 6.47 eV
f (ε)
0.05
fM
0.00
0 5 10 15 20 25 30
Electron energy (eV)
FIGURE 2.4
Comparison among f (), f M () , and f D () with the same mean energy.
coefficients are defined in the same way as for ionization. However, these co-
efficients do not directly enter the transport equations, although they may be
used in some corrections and energy balance equations.
Thus momentum transfer is dominated by elastic collisions, whereas
energy balance is primarily controlled by inelastic processes (with the excep-
tion of rare gases and some metallic vapors that have no inelastic processes at
low energies). However, it is not practical to use mean energy as the swarm
parameter because it cannot be measured directly, so we traditionally define
a property of electron swarms known as characteristic energy εk = e DT /µ,
where µ = vd /E is the mobility. Characteristic energy has the dimension of
energy and is usually within 30% of the mean energy (although sometimes
the differences may be up to a factor of two or more). This quantity is usually
expressed in units eV or, even more often, by the analogous quantity D/µ,
which is expressed in volts. The definition of characteristic energy is given by
combining Equations 2.26 and 2.28:
√
ε
DT DT −e E 1
3
2
m NQm (ε)
f (ε)dε
εk = e = e E =
,
µ vd 1 2
e E NQεm (ε) dε
d f√(ε)
dε
3 m ε
In the case of the thermal equilibrium where ε = 3kT/2, we obtain the
following relation:
DT kT
= . (2.32)
µ e
PROBLEM 2.2.2
Discuss the momentum transfer loss of inelastic collision, and compare the influence
of the collisional effect between the energy loss and the momentum loss.
where α is the polarizability of the ion in the dipole interaction with the
neutral molecule [2]. The governing equation of the velocity distribution G(v)
of ions is then given in the form of the relaxation equation using the BGK
approximation [3, 4],
1/2
β⊥ β
G 0 (v) = exp − β⊥ vr2 − β (vz − < vz > + q Eτ/Mp )2 , (2.34)
π π
where β⊥ = 1/(2 < vx2 >), and β = 1/2[< vz2 > − < vz >2 −(q Eτ/Mp )2 ].
Equation 2.33 is then solved as
β⊥ vz − < vz > + qEτ/Mp 1
G(v) = exp −β⊥ vr2 − +
2π(qEτ/Mp) qEτ/Mp 4β (qEτ/Mp )2
1/2 qEτ 1 vz − <vz > + qEτ/Mp
× erfc 2β − , (2.35)
Mp 4β (qEτ/Mp )2 2(qEτ/Mp )
The integral represents the mean velocity, that is, drift velocity vdp in velocity
space, which is then equal to
Mp + M qE
vd p = . (2.37)
Mp M Rm
Mp M 2
εl = Mp v p − MV 2 . (2.38)
(Mp + M) 2
E / N = 20 Td T g = 300 K
1.0 3
Energy distribution f log (log10ε)
200 Td
120 Td
Qm( γ) γ
0.5 2
3
kT
2 g
0 1
0.01 0.1 1 10
Relative kinetic energy (eV)
FIGURE 2.5
Energy distribution at three different E/N and elastic momentum transfer cross section of
O− in O2 .
PROBLEM 2.2.3
Derive the relation 2.38.
In a steady state, the energy balance of the ion swarm in unit time is given
by
Mp M 2 Mp + M 1
qE · vdp −
εl Rm =qEvdp 1− Mp v p −MV
2
=0,
(Mp + M)2 Mp M vd2 p
(2.39)
which is satisfied when
(Mp + M )vd2 p + MV 2 − Mp v 2p = 0 .
By using this relation we obtain the expression of the mean ion energy, which
is known as Wannier’s formula [2]:
1 1 1
ε p = Mp v 2p = (Mp + M)vd2 p + MV 2 . (2.40)
2 2 2
The physical meaning of Equation 2.40 is that the mean energy of ions is
separated into three parts: one due to the effect of the electric field given by
the drift velocity (first term), one due to the apparent kinetic energy of the
encounter molecule (second term), and the thermal energy equal to 3kTg /2
(third term).
PROBLEM 2.2.4
When an ion swarm has a shifted Maxwellian velocity distribution in gases with a
Maxwell distribution of Tg , derive that the relative velocity distribution between the
ion and the molecule G(vr ) also has the Maxwellian velocity distribution.
case of LTE, the temperature is merely a parameter used to apply the laws of
thermodynamic equilibrium to a plasma, but this approximation is neverthe-
less quite effective.
Γ = nv
∂n
= nvd − D∇n = nµE − D k. (2.41)
∂z
µ −µ ∂V(z) 1 ∂n
E= = . (2.42)
D D ∂z n ∂z
We substitute the Einstein relation 2.32 into Equation 2.43 and obtain
eV(z)
n(z) = n0 exp − , (2.44)
kT
gM
0
Vz
Vx
FIGURE 2.6
Equilibrium velocity distribution in vx –v y space.
where the magnitude of the velocity v is defined by v 2 = vx2 + v 2y + vz2 , and the
distribution in terms of v 2 is then defined as
Exercise 2.3.1
Show that the solution to Equation 2.47 is
g vx2 = a 2exp −αvx2 , (α > 0). (2.48)
where a = G(0) and, by the same argument, g(v 2y ) = a 2 G(v y ) and g(vz2 ) =
a 2 G(vz ). Therefore, Equation 2.47 is rewritten as
1
g(v 2 ) = g vx2 + v 2y + vz2 = 6 g vx2 g v 2y g vz2 . (2.49)
a
Note that the functional Equation 2.49 satisfies Equation 2.48 as the solution
of each component.
Therefore, we obtain
g(v 2 ) = g vx2 + v 2y + vz2 = A exp −α vx2 + v 2y + vz2 , (2.50)
and therefore
3/2
α
A= .
π
3 1 α 3/2 m
vx2 + v 2y + vz2 e −α(vx +v y +vz ) dvx dv y dvz
2 2 2
kT = mv 2 =
2 2 π 2
α 3/2 3m 1 π 3/2
= ,
π 2 2α α
= 4πv g M (v 2 )dv .
2
(2.53)
The following averaged velocities are defined for the Maxwellian speed
distribution: (i) most probable speed ṽ, (ii) root mean square velocity v 2 ,
and (iii) mean speed v. These are obtained as follows.
ii. The root mean square velocity (v 2 )1/2 is the square root of the mean
value of v 2 :
3kT
v 2 = v 2 F (v)dv = . (2.55)
m
Figure 2.7 shows the plot of the normalized Maxwell speed distribution
F (v) and gives the values of the three possible choices for averaging the ve-
locity. Different physical quantities may be averaged with different powers of
0.8
0.6
F(v)
0.4
0.2
0 v
0 0.5 1.0 1.5 2.0 ~
v
FIGURE 2.7
Normalized Maxwell speed distribution and the relationships among these averaged velocities.
speed and will be associated with different effective velocities. It is clear that
the mean speed may be used to establish the momentum transfer, that the
mean square speed is relevant for consideration of the mean energy, and that
the most probable speed is the peak of the probability distribution. Although
one may directly produce averages of physical phenomena, associating them
with mean values of speed provides a better physical insight into the phe-
nomenon under consideration.
References
1. Parker, J.H., Jr. and Lowke, J.J. 1969. Phys. Rev. 181:290–301.
2. Wannier, G.H. 1953. Bell Syst. Tech. J. 32:170.
3. Bhatnagar, P.L., Gross, E.P., and Krook, M. 1954. Phys. Rev. 94:511.
4. Whealton, J.H. and Woo, S.B. 1971. Phys. Rev. A 6:2319.
3.1 Introduction
Plasma is a compound phase with electric quasi-neutrality, generally consist-
ing of electrons, positive ions, and neutral molecules. In particular, strongly
ionized plasma consists of electrons and positive ions. The phase is called
collisionless plasma, as the electron (ion) has few short-range binary colli-
sions with neutral molecules, and the plasma system is subject to long-range
Coulomb interactions. When the short-range two-body collision is major, the
system is collisional, that is, collision dominated. Collisionless plasma shows
unique characteristics. We briefly describe the characteristics [1–3].
3.2 Quasi-Neutrality
The presence of electrons and ions with density ne and np will produce an
electric field subject to Poisson’s equation,
n p − ne
divE = e , (3.1)
0
where e(>0) is the elemental charge and 0 is the permittivity (dielectric con-
stant) of vacuum. A plasma where the density of positive ions is exactly
equal to that of electrons has no electric field. Naturally, based on the great
difference in mass between the electron and positive ion, a quasi-neutral
state with np ∼ ne is maintained in plasmas on a macroscopic scale. This
causes a finite positive plasma potential with respect to the wall earthed to the
ground.
Notice that a low-temperature plasma externally excited by direct current
(DC) or rf power source has a plasma production region and holds a structure
bifurcating into a positive ion sheath with a high field to produce an electron–
ion pair and a bulk plasma with quasi-neutrality.
27
plasma
(ne~np)
p-ion
d electron
FIGURE 3.1
Plasma fluctuation and the charge separation in a collisionless plasma with macroscopic quasi-
neutrality.
Exercise 3.3.1
Calculate the Debye length when plasma density and electron temperature
in a collisionless plasma are 1016 m−3 and 3 eV, respectively.
ε0 kTe 1/2 kT[eV] 1/2
λD = ∼ 7.43 × 10 3
× [m]. (3.4)
ne 2 n[m−3 ]
That is,
1/2
3
λ D = 7.43 × 10 × 3
= 1.29 × 10−4 [m].
1016
PROBLEM 3.3.1
Discuss that the plasma frequency ω p of the ion with charge Ze is defined in the same
way as the electron,
1/2 1/2
Z2 e 2 n p m
ωp = =Z ωe . (3.8)
Mp ε0 Mp
Also discuss that practical plasma oscillation is given by (ωe2 + ω2p )1/2 =
ωe (1 + z2 me /Mp )1/2 .
PROBLEM 3.3.2
Estimate the values of ωe and ω p in collisionless Ar plasma with density 1016 m−3 .
PROBLEM 3.3.3
A plasma is externally irradiated by an electromagnetic wave with frequency of ω.
Derive the condition that the external disturbance does not immerse deeply in the
plasma.
1 d dV (r ) e
∇ 2 V (r ) = r2 =− n p − ne . (3.9)
r 2 dr dr ε0
There are number of electrons and ions in the shielded region in front of the
sphere. The number density of electron and ions, ne (r) and n p (r), follows the
V n
np(r)
n0
q
ne(r)
4πε0r
-V(r)
0 r
λD
FIGURE 3.2
Plasma shielding of a charged sphere.
eV (r )
ne = n0 exp , (3.11)
kTe
where n0 is the quasi-neutral plasma density far from the influence of the
impurity metal. At a long distance from the metallic surface satisfying eV(r )
kTe , kTp , the exponential terms in Equations 3.10 and 3.11 will be expanded
by the Taylor series, and we obtain
1 d dV (r ) en0 eV (r ) eV (r )
r2 =− exp − − exp
r 2 dr dr ε0 kTp kTe
e 2 n0 Tp + Te
≈ V (r ) . (3.12)
ε0 kTp Te
The potential close to the metal with a surface charge q 0 is expressed by the
Coulomb potential, q 0 /4π0r . Then, the final form is
q0 r
V (r ) = exp − . (3.14)
4π ε0r λD
Noted that V(r ) is negative due to q 0 (< 0) under me
Mp . Equation 3.14
shows that the scale length of the shielding is on the order of the Debye
length λ D .
I
insulator metal a
b
(I)
plasma (II)
λD
VS
Vfl VS
d c V
Isi
(III)
FIGURE 3.3
Ideal Langmuir probe characteristics in a plasma.
When the probe has a sufficiently large surface, a saturated electron cur-
rent will be collected at V > Vs where Vs is the space potential of the plasma.
Practically due to the finite surface area, I (V) characteristics at V > Vs will
increase gradually as shown in Figure 3.3. The probe potential at c where the
current is zero is named the floating potential. The name comes from the zero
net flux at the surface of an insulator inserted into a plasma.
The probe will be given an absorbing boundary that absorbs electrons and
ions incident on the surface. At the region bc that satisfies V < Vs , the electron
sheath will be formed in front of the probe (Vs − V < 0). The electron incident
right on the probe surface will be retarded by the field, and only the electron
with kinetic energy mv 2 /2 > |e(Vs − V)| will be absorbed as the electron cur-
rent. We consider that the electron has the Maxwellian velocity distribution
with temperature Te in the plasma and has no collision in the passive sheath
in front of the probe. Then the electron probe current will be given by
∞ ∞ ∞ 3/2
m
I (V) = ene S 2e|V | vz
vx =−∞ v y =−∞ vz = m
p 2π kTe
m(vx2 + v 2y + vz2 )
× exp − dvx dv y dvz
2kTe
3/2 ∞ ∞
m mvx2 mv 2y
= ene S exp − dvx exp −
2πkTe −∞ 2kTe −∞ 2kTe
∞
mvz2
× dv y exp − dvz
vz 2kTe
3/2 √ √
m 2 π 2 π 1 eVp
= ene S
m exp − , (3.15)
2πkTe 2 2kT m
2 m
kTe
kTe
e 2kTe
where S is the effective area of the probe. By considering the mean speed
ve = (8kTe /πm)1/2 in Equation 3.15, the probe I (V) characteristics are
described by
ene ve V − Vs
I (V) = S exp . (3.16)
4 kTe
ene ve
I = S. (3.17)
4
Exercise 3.4.1
Discuss the pressure condition that the electron temperature Te is estimated
from the curve in region (II) in bc in Figure 3.3.
As a typical plasma we assume, ne = 1015 m−3 , kTe = 3.0 eV. Then, the Debye
length is
1/2 √
kTe [eV] 7.43 × 103 × 3
λe = 7.43 × 10 × 3
= √ = 4.07 × 10−4 [m].
ne [m−3 ] 10 × 107
The collision rate R is roughly approximated at 107 p[Pa s−1 ], and the flight
time of the electron in the static sheath in front of the probe is
λe 4.07 × 10−4 1 1
≈ ≈ 3.51 × 10−10
≈ .
ve 1.16 × 106 R 107 p
1 1
p
× ≈ 2.85 × 102 [Pa].
107 3.51 × 10−10
plasma wall
(ne ~ np) A+
e
Γe ~ Γp
Ea
FIGURE 3.4
Fluxes of charged particles in front of a wall.
e ≈ p (3.18)
is realized.
We deal semiquantitatively with the ambipolar diffusion phenomena
caused between the static plasma and the surrounding reactor wall. Each
of the fluxes of electrons and ions with density, ne and n p , is given,
dne
Γe = ne ve = ne vde − De , (3.19)
dr
dn p
Γ p = n p v p = n p vd p − Dp , (3.20)
dr
where v j , vd j , and D j are the mean velocity, drift velocity, and diffusion co-
efficient of electrons and positive ions, respectively. In a steady state as men-
tioned above, the electron and ion have the same velocity toward the wall,
ambipolar diffusion velocity va ,
1 dn
va = −Da , (3.21)
n dr
µ p De + µe Dp
Da = , (3.22)
µ p + µe
In the case of Te Tp ,
Te
Da = Dp . (3.24)
Tp
In particular, in Te = Tp ,
Da = 2Dp . (3.25)
PROBLEM 3.5.1
Derive the ambipolar diffusion coefficient, Equation 3.22, and show that the ambipolar
diffusion field is written as
Dp − De 1 dn
Ea = . (3.26)
µe + µ p n dr
∂
n(r, t) = D∇ 2 n(r, t). (3.27)
∂t
The diffusion Equation 3.27 implies that the spatial distribution and time
scale of the diffusion are determined by the geometry of the reactor. Here, we
divide n(r, t) into two independent functions n(r) and T(t). Then we have
1 dT (t) 1 1
= D∇ 2 n(r) ≡ − . (3.28)
T (t) dt n(r) τD
PROBLEM 3.5.2
Discuss the physical meaning that the r.h.s. of Equation 3.28 has negative value.
and
1 n(r)
∇ 2 n(r) = − n(r) = − 2 , where 2 = Dτ D . (3.30)
Dτ D
PROBLEM 3.5.3
Calculate the characteristic diffusion length in the case where the reactor geometry
is (a) an infinite parallel plate (separation of d0 ), (b) an infinite rectangle (two sides,
d0 and l0 ), (c) a cylinder (radius r0 ), and (d) a sphere (radius r0 ), respectively.
1
= (π/d0 )2 ; infinite plates,
2
= (π/d0 )2 + (π/l0 )2 ; infinite rectangle,
= (2.405/r0 )2 ; cylindrical,
= (π/r0 )2 ; spherical. (3.31)
PROBLEM 3.5.4
Neutral molecules diffuse in three-dimensional space without boundary. Calculate
the number density of molecules n(r, t) in Equation 3.27. In particular, when the
initial condition of the density is given by n(r, t = 0) = δ(r), derive the density
distribution
3/2
1 r2
n (r, t) = exp − . (3.32)
4π Dt 4Dt
The above stochastic process is known as the Wiener process and the
density distribution is the normal distribution N(0, 2Dt) with mean value of
0 and variance of 2Dt.
The characteristic diffusion length D is much influenced by the loss
mechanism of the particle. For example, we consider the electron diffusion in
electronegative gases with density N and with the electron attachment rate
coefficient ka . Then the electron continuity equation is
∂
n(r, t) = −ka n(r, t)N + D∇ 2 n(r, t). (3.33)
∂t
Exercise 3.5.1
Molecules excited to an optically forbidden level are called metastables and
have a long lifetime. The metastable state is de-excited without photo-
emission when it reacts with the wall with absorbed molecules and exhausts
its excess inner energy to the wall. Estimate the effective lifetime of metasta-
bles in a reactor of volume Vol with reflecting wall. Reflection coefficient γr at
the wall is given by γr = 1 − St , where St is the surface sticking coefficient,
2 2
2 2Vol (2 − β)
τ= + . (3.36)
π D St vβ
1. The frequency of the external power source is much higher than the
collision rate ω R (i.e., the collisionless sheath).
2. In the momentum continuity equation (see details in Chapter 5),
the pressure gradient dP(z)/dz = kTdne (z)/dz is dominant for elec-
trons, whereas other terms except the pressue are predominant for
ions.
3. The ion flux is constant, that is, n p (z)vdp (z) = const, without ion-
ization and recombination in the collisionless sheath. Under these
conditions, the momentum equations of electrons and ions in the
steady state are, respectively,
dP(z) dV(z)
me = −ene (z) , (3.37)
dz dz
dvdp dV(z)
Mp n p (z)vd p = en p (z) . (3.38)
dz dz
From the above ion flux continuity and Equation 3.38, we obtain
dV(z)2 e
= − [n p (z) − ne (z)]
dz2 ε0
In the sheath, positive ions are dominant; that is, n p (z) − ne (z) > 0. Accor-
dingly,
−1/2
2e[V(z) − V(z0 )] e[V(z) − V(z0 )]
1− − exp > 0. (3.39)
Mp vdp (z0 )2 kTe
In particular, at the position z close to the boundary between the bulk plasma
and the sheath, e[V(z) − V(z0 )]
kTe will be satisfied. By using a Taylor
expansion of the exp term in Equation 3.39,
2e[V(z) − V(z0 )] 2e[V(z) − V(z0 )]
1− < 1− .
kTe Mp vd p (z0 )2
Finally, we have the initial directional velocity of ions just entering the
sheath from the bulk plasma (i.e., Bohm velocity):
1/2
kTe
vd p (z0 ) > . (3.40)
Mp
Therefore, the electron density at z0 is
e[Vplasma − V(z0 )]
ne (z0 ) = ne0 exp − = 0.61ne0 . (3.42)
kTe
When a small metal plate isolated electrically from the reactor is immersed in
a bulk plasma with a plasma potential Vplasma and electron temperature Te , the
surface potential Vf l , called the floating potential, is given at the condition of
zero net current ee = e p as
kTe Mp
Vf l = Vplasma − ln . (3.43)
2e 2.3m
Exercise 3.6.1
Derive the above expression.
Although ee = e p is satisfied, positive ions have a strong directional flux,
and electrons show an isotropic flux with random speed < ve >= (8kTe /π m)1/2
(see Chapter 2). Therefore,
1 e[Vplasma − Vf l ] kTe 1/2
ne0 exp − < ve >(= (n p (surf )vd p (surf )) = n p (z0 ) .
4 kTe Mp
By using Equation 3.42 under ne (z0 ) ∼ n p (z0 ),
e[Vplasma − Vf l ] π m 1/2
− = ln 2.4 .
kTe 8Mp
References
1. Chen, F.F. 1984. Introduction to Plasma Physics and Controlled Fusion, Vol. 1.
New York: Plenum.
2. Golant, V.E., Zhilinsky, A.P., and Sakharov, I.E. 1980. Fundamentals of Plasma
Physics. New York: John Wiley & Sons.
3. Cherrington, B.E. 1979. Gaseous Electronics and Gas Lasers. Oxford: Pergamon.
41
arising from surface heating and surface reactions. A special group of surface
processes associated with plasma etching are also described briefly here and
in greater detail in Chapters 8 and 12.
ε = h̄ω (4.2)
p = h̄k. (4.3)
PROBLEM 4.1.1
Show that the plane wave satisfies the uncertainty principle.
Exercise 4.1.1
Calculate the de Broglie wavelength λe for the electron that has crossed a
potential drop V0 (see Figure 4.2).
wave
front
k . r = const.
at t = t0
r
p
z
m 0 k λ λ
(particle) (plane wave)
FIGURE 4.1
Classical and quantum (plane wave) description of a particle with the mass m and momentum
p(0, 0, pz ).
The energy of a particle is equal to the potential drop V0 times the elementary
charge e, so we can obtain the momentum of the particle as a function of V0 :
ε = p 2 /2m = eV0 .
C A G
λe
electron beam
emitter
V0
FIGURE 4.2
A stationary electron beam with mono-energetic energy eV0 produced in a vacuum.
Exercise 4.1.2
Obtain the dispersion relation between the frequency and the wave number
for a plane matter wave (r, t) in free space (see Equation 4.4).
If we start from the time- and space-dependent Schrödinger equation
∂
H (r, t) = jh̄ (r, t), (4.6)
∂t
h̄ 2 2
H=− ∇ + V(r, t), (4.7)
2m r
then the solution in free space (V = 0) to this system with the plane wave in
Equation 4.4 gives us
h̄ 2 k 2
(r, t) = h̄ω (r, t).
2m
Finally, we obtain the dispersion relation (between the wave number k and
the angular frequency ω)
h̄k 2
ω(k) = . (4.8)
2m
PROBLEM 4.1.2
The wave-front of a plane wave is defined as k · r = const at time t, and in that case
the wave front is perpendicular to the wave vector k. Show that a spherical wave with
a circular wave front is expressed as
A0
(r, t) = exp[ j (k · r − ωt)]. (4.9)
r
0 +k
k
(z, t) = (z, t; k)
k0 −k
k0 +k
= A(k) exp[ j (kz − ωt)]dk. (4.10)
k0 −k
λ0
FIGURE 4.3
Wave packet of a group of particles moving in the z-direction.
ω = ω(k − k0 )
dω 1 d 2 ω
= ω0 + (k − k ) + (k − k0 )2 + · · ·,
dk 0 2 dk2 0
0
In Figure 4.3, we show a plot of Equation 4.11. The locally isolated particle
group is represented by a strongly modulated wave with the fundamental fre-
quency ω0 (k0 ). There is an envelope with a maximum in the position where a
particle should be in a classical model (i.e., the center-of-mass) and the undu-
lations extend approximately over the one wavelength λ0 . The exponential
term in Equation 4.11 determines the phase of the wave.
also known as the group velocity. For the particular example given in
Equation 4.8, the result is
d h̄k 2 h̄k0
Vg = = . (4.13)
dk 2m 0 m
Exercise 4.1.3
Derive the phase velocity of the wave packet in Equation 4.11.
The phase in Equation 4.11 is
= k0 z − ω0 t = const,
dz d ω0 t ω0
Vp = = = .
dt dt k0 k0
Using Equation 4.8 allows the phase velocity to be related to the group velocity
1 h̄k02 h̄k0 Vg
Vp = = = .
k0 2m 2m 2
PROBLEM 4.1.3
A pulsed electron beam is formed in a vacuum by using the electrical shutter between
grids A and G in Figure 4.2. The half width of the group z is 1 mm, and V0 is 150 V.
Discuss the uncertainty principle using vz /vz .
incident
plane wave
v
z z
m M
incident target
particle molecule
FIGURE 4.4
Classical (a) and quantum (b) representation of a collision.
the Coulomb force has infinite range, collisions between charged particles are
described as many-body, long-range collisions.
There is also a whole group of three-body processes where the third
particle is required to satisfy the conservation laws. A three-body process
occurs, for example, when one of the species is left in an excited state after
the two-body collision and the excited species needs a subsequent collision
with a third body to transfer its extra energy for stabilization. Three-body
processes will occur only if the gas density is sufficiently high.
Here we deal primarily with two-body, short-range collisions. In clas-
sical mechanics, particles have definite identities, positions, and velocities
(momenta). The classical collision of a light projectile of mass m and veloc-
ity v on a heavy target of mass M is depicted in Figure 4.4a. After the
collision the light particle will be scattered at an angle θ with respect to its
original direction z and into the solid angle d(= sin θ dθ dφ). It will have
a definitive velocity and path, and the differential cross section σ (θ, φ, ε)d
will give the probability of scattering into the solid angle d. In the quantum
case, an incoming free particle will be represented as a plane wave described
by Figure 4.4b and Equation 4.4. Then, as a result of spherical scattering,
the outgoing wave will be given by the sum of the plane and the spherical
waves.
where (m, M) are, respectively, the masses of the charged and neutral particle,
and (v , V ) are the velocities before and (v, V) the velocities after the colli-
sion. A similar conservation law is written for the energy. However, in this
case internal excitation can occur and the inelastic process should also be
accounted for. In the energy conservation law, it is convenient to change the
frame of reference to a center-of-mass (CM) frame, and the kinetic energy is
represented as
1 2 1 1 1
εkin = mv + MV 2 = (m + M)vg2 + µr vr2 , (4.15)
2 2 2 2
where the CM velocity vg and the relative velocity vr are given as
mv + MV
vg = (4.16)
m+M
and
vr = v − V. (4.17)
The CM velocity remains constant before and after the collision under the
conservation of momentum expressed in Equation 4.14. As a result, the trans-
formation from the standard laboratory (LAB) system to the CM system effec-
tively allows us to describe a two-particle process as a single-particle process.
The reduced mass of the effective particle is therefore
mM
µr = . (4.18)
m+M
Using these transformations to the CM frame, we can write the energy
conservation law as
1 1
µr vr = µr vr2 + ε R ,
2
(4.19)
2 2
where ε R is the energy loss (reaction energy) due to the excitation of internal
energy levels of one or both particles. We define three kinds of collisions, each
defined by their value and sign of the reaction energy.
i. ε R > 0: Collisions of the first kind, in which we have loss of the total
kinetic energy to inelastic processes;
ii. ε R < 0: Collisions of the second kind, in which we have gain of the
total kinetic energy from the internal excitation energy; and
iii. ε R = 0: Elastic collisions, in which the total kinetic energy is
conserved.
PROBLEM 4.2.1
Explain that the CM energy is conserved in the case of binary collision, and that in
the case of inelastic scattering the energy to excite the molecule is provided from the
relative kinetic energy.
PROBLEM 4.2.2
Derive the energy transfer as a function of scattering angle in an elastic collision of
two hard spheres (r, m, v ) and (R, M, 0), both defined by their radius, mass, and
velocity before collision.
The concept of an “effective area” associated with target particles for any
pair of colliding particles (two-body collision) is useful even if the hard sphere
approximation is not valid. There are, however, some conditions that must
be met for it to be applicable.
i. The collisions between incident particles should be negligible; that
is, the density of the incident particles should be sufficiently low;
a2
v a1 a0
z
V
FIGURE 4.5
Binary encounter of solid spheres with radii of a 1 and a 2 , and the concept of the collision
cross section.
ii. The density of the target particles should be low in order to avoid
collisions of the incident particles with more than one target at the
same time. That is, the de Broglie wavelength of the incident particle
should be shorter than the mean distance between target molecules.
In defining the collision cross section we return to the differential cross section
that was mentioned earlier. When a certain number of projectiles n with
energy ε are directed toward a target consisting of N particles, the number of
particles dn scattered into the surface element dS on the spherical surface at
a distance r from the target center will be proportional to the density of the
projectiles and the targets and also proportional to the area of the surface dS
and r −2 , that is, the solid angle in the collision experiment:
dn ∝ nN ds/r 2 ,
In Equation 4.20, σ (θ, φ; ε)d is the probability that one projectile is scat-
tered into a small solid angle d at scattering angle (θ, φ). σ (θ, φ; ε) has the
dimensions of the area and so is termed the “differential cross section.”
The differential cross section may be expanded into the Legendre polyno-
mials Pn (cos θ ) under the azimuthal symmetry
∞
1
σ (θ ; ε) = (2n + 1)Qn (ε)Pn (cos θ), (4.21)
4π 0
where Qn (ε) is the nth integral cross section. Multiplying the equation by
Pn (cos θ ) and using the property of orthogonality of the Legendre polynomi-
als (see Section 5.6.1), we obtain the following definition of the integral cross
section Qn (ε):
π
Qn (ε) = 2π σ (θ; ε)Pn (cos θ) sin θ dθ. (4.22)
0
The coefficients Qn (ε) and the differential cross section σ (θ ; ε) are defined
for each of the many scattering processes that can occur. They are interpreted
as the scattering probability averaged over all angles and weighted by some
Legendre polynomial. For example, Q0 (ε) is the total cross section for the
process (i.e., the probability of scattering at any angle) and corresponds to
the area of the target as seen by a point projectile. The most commonly used
angular averages of the differential cross section are the following:
These integral forms of differential cross section play a special role in transport
theory. For inelastic collisions with threshold energy ε j , the definitions are the
same except that the momentum transfer cross section is defined as
π
εj
Qm (ε) = 2π 1− 1− cos θ σ (θ; ε) sin θdθ, (4.26)
0 ε
> ε j when the scattering is nearly isotropic, Qm (ε) will be very
Thus, at ε ∼
small.
Exercise 4.2.1
Explain the physical meaning of the momentum transfer cross section for
elastic scattering.
Assuming that the initial momentum is p and that particles are scattered at
an angle θ with no loss of the magnitude of the velocity, the difference of
momentum along the axis of the initial velocity is
p = p − p cos θ = p(1 − cos θ).
In other words, the fractional change of the momentum is
p
= (1 − cos θ). (4.27)
p
Therefore, the cross section Qm in Equation 4.24 is in some way a representa-
tion of the momentum transfer to the gas molecules. Qm for elastic collision
has a finite magnitude, and the elastic energy loss (2m/M)ε of the electron
with energy ε is negligibly small.
PROBLEM 4.2.3
Discuss the physical meaning of the viscosity cross section.
Exercise 4.2.2
Gas molecules with density N and temperature Tg are present in a reactor.
Estimate the limitations of the two-body approximation for the collisions
between molecules.
This criterion may be related through the relationship between the mean dis-
tance between molecules d and the de Broglie wavelength λde of the molecule,
d λde .
The mean distance between the gas molecules may be obtained from the gas
number density N as d = N−1/3 . On the other hand, the wavelength, which is
used to approximate the range over which quantum effects are appreciable,
will be calculated for the most probable speed of molecules, 2kTg /M. Hence
the de Broglie wavelength is given by
h h h
λde = ≈ = .
p Mṽ 2MkTg
The condition therefore becomes
λde h N1/3
≈ 1. (4.28)
d 2MkTg
PROBLEM 4.2.4
Calculate Q0 , Qm , and Qv for the following models of differential scattering cross
sections:
f (ε)
1 exp
λ ( λz )
0 z
ε collision
electron
beam
FIGURE 4.6
Distribution of the mean free path f (z; ε) for a beam of particles with energy ε incident on a gas
target with random distribution.
Finally, we note that the total collision probability Pc is often used to represent
the number of collisions in a gas at some pressure p and at temperature Tg [K].
The total collision probability is equal to Pc = ( p/kTg )Q. Here, the standard
value that is used is p = 1Torr at room temperature, where we have
Exercise 4.2.3
Derive the distribution function of the free path (Equation 4.30).
When we define F (z) as the integral distribution of collisions on the segment
(0, z), the number of collisions in the small segment (z, z + z) is proportional
−1 F (z + dz) − F (z)
= −c,
1 − F (z) dz
and thus the formula (Equation 4.30) for the distribution function of the
free path is obtained. The distribution f (z) can be used to determine the
probability of collision (or of the length of the free path) of a single particle as
well. One should bear in mind that the mean free path depends on the energy
of the incident particles, that is, < λ(ε) >. Thus if we have a distribution of
incident energies we perform additional averaging and obtain the mean free
path for the whole ensemble.
v'
v
π−ω c
2
ωc
vg
ωL
V
V'
FIGURE 4.7
Velocities of two particles before and after collision and the corresponding angles in the CM
frame.
Exercise 4.2.4
Obtain the relationship between differential cross sections for the laboratory
and the CM reference frames in the elastic scattering.
Consider two particles, (m) and (M), moving toward one another with
velocities (v , V ) before and (v, V) after the collision in the LAB frame. The
magnitude of the relative velocity vr is conserved before and after the elas-
tic scattering in the CM frame, (| v − V |=| v − V |=| vr |). We may thus
represent the motion of particles before and after the scattering by two straight
lines of equal lengths (|vr |) intersecting at the point vg (CM velocity) (see
Figure 4.7). In the CM frame, the velocities of the two particles before and
after the collision are located on spheres with radii of v = vr M/(m + M) and
V = vr m/(m + M), respectively.
We can identify one equilateral triangle with the CM frame scattering
angle ωc , that is, the triangle [vg , v , v]. Thus we write
v v Mvr /(m + M)
= ,
sin ωc sin[(π − ωc )/2]
where the length between points [v, v ] is denoted v v. If, however, we take
the triangle [V , v , v], the following relation is obtained by using the LAB
frame scattering angle of particle m, ω L :
v v vr
= .
sin ω L sin[π − ω L − (π − ωc )/2]
sin ωc
tan ω L = (4.32a )
(m/M) + cos ωc
and
(m/M) + cos ωc
cos ω L =
1/2 . (4.32b)
1 + 2(m/M) cos ωc + (m/M)2
Because the same number of particles is scattered through the solid angle
in both systems, we have
or
d cos ωc
σ L (ω L ; ε) = σc (ωc ; ε)
d cos ω L
3/2
1 + 2(m/M) cos ωc + (m/M)2
= σc (ωc ; ε) . (4.33)
1 + (m/M) cos ωc
It is possible to show that the energy transfer from the projectile to the target
(i.e., the energy loss to the first particle) is then equal to
2
εM MV 2 /2 M 2mv
= = cos ω L
εm mv 2 /2 mv 2 m + M
4mM
= cos2 ω L . (4.34)
(m + M)2
PROBLEM 4.2.5
Determine the scattering angles ω L and ωC M when (a) m = M and (b) m M.
PROBLEM 4.2.6
Show that the energy loss in elastic collisions is given by Equation 4.34, and that for
the electron-molecule scattering it is simplified after averaging to (2m/M)ε.
PROBLEM 4.2.7
Prove that for a head-on (ω L = 0) collision of a heavy particle with a light target
m M, the maximum velocity of the target will be V = 2v .
It is also of interest to determine the velocities after the collision in the case
of inelastic processes characterized by the inelastic energy transfer ε. After
inelastic scattering, the velocity of the target in the LAB frame is equal to
1 m
v= mv cos θ L ± [mMv 2 cos 2 θ L − 2(m + M)ε] (4.35)
m+M M
m+M
εm = ε.
M
For particles of the same mass (m = M) one needs at least εm = 2ε
to achieve an inelastic process, whereas for light projectiles (m M) the
minimum energy is εm = ε.
v0
v0 dθ
db θ
b
z
0
2πbdb
dΩ = 2πsinθdθ
FIGURE 4.8
Definition of the scattering angles and the impact parameter in classical scattering
phenomenology.
σ (θ ) = 2π bdb/d
db(θ ) dθ
= b(θ )
dθ sin θ dθ
1 db(θ )
=
b(θ ) . (4.38)
sin θ dθ
Our goal is thus to determine the dependence b(θ ) for any given potential
interaction.
v0
1 (π − θ)
2 b
1 (π − θ)
2
v0′ rmin θ
r
b′ ξ
FIGURE 4.9
The geometry of classical scattering on a target with interaction potential V(r ).
In Figure 4.9 we define the coordinates for the scattering and basically
the incoming particle with reduced mass µ follows the angle ξ . From
Equation 4.39 it is shown that
L2 µv 2 b 2
+ V(r ) = + V(r ) = Veff (r ), (4.43)
2µr 2 2r 2
and Veff (r ) is the effective potential in the radial direction. The interaction
potential usually has a minimum; as a result, the projectile is for a while
accelerated to the target, but it will not be bound as it has a positive energy.
The addition of the centrifugal term to the interaction potential leads to an
increase in the potential in the region of the minimum as the centrifugal term
increases. The potential may even become repulsive. It may also be possible
that a maximum is formed at some distance rc with a shallow minimum for
smaller distances. Near the maximum the radial motion will be very small,
but it will take several revolutions before the particle can leave after scatter-
ing as the angular momentum is large. This effect is known as “orbiting.”
Orbiting allows longer interaction times and therefore a greater probability
for processes that involve transitions with finite lifetimes.
To calculate the trajectory from Equation 4.43, we must obtain an equation
that will describe the trajectory through the angle ξ ,
dξ 1
= ± 1/2 . (4.44)
dr 2µr 4
L2
{E − V(r )} − r 2
When dξ/dr < 0, the incoming particle approaches the target and the
interaction force increases. When dξ/dr > 0, the projectile leaves the target.
Because of the symmetry of the scattering (see Figure 4.9), we may perform
the integration in two parts. First we perform it up to the point of the closest
approach rmin , from ξ = 0 (for r → ∞) up to ξ = (π − θ)/2. We obtain
(π −θ)/2 rmin
dξ = −[{2µr 4 /L 2 }{E − V(r )} − r 2 ]−1/2 dr
0
∞∞
(π − θ )/2 = [{2µr 4 /L 2 }{E − V(r )} − r 2 ]1/2 dr. (4.45)
rmin
The limit rmin (the point of the closest approach) may be obtained from
dr/dξ = 0, which in combination with Equation 4.45 leads to {2µrmin 4
/L 2 }
{E − V(rmin )} − rmin = 0. Introducing u = 1/r , we derive the following result
2
Exercise 4.3.1
Determine the dependence of the impact parameter b(θ ) for classical
hard
sphere elastic scattering, that is, for the potential V(r ) = ∞(r ≤ r0 )
0(r > r0 )
. Also deter-
mine the effective cross section for the scattering.
In Figure 4.10 we show that the hard sphere potential is a very good
approximation for many classical potentials. The point of minimum approach
∞
V(r)
hard sphere
realistic potential
r
0 r0
π−θ
2 π−θ
2
π−θ
2
b (θ)
0 r0
FIGURE 4.10
Comparison between the hard sphere interaction potential and a realistic potential (dashed line)
and geometry of the hard sphere scattering for a given impact parameter.
= no collision; (b > r0 ).
Substituting the above relation into Equation 4.38 we obtain the expression
for the differential cross section:
and from the definition of the total cross section Equation 4.23 we obtain
Q0 = σ (θ)d = (r02 /4) sin θ dθ dφ
PROBLEM 4.3.1
Derive the momentum transfer cross section Qm for hard sphere elastic scattering,
and discuss the relationship between momentum transfer and the total cross section.
Exercise 4.3.2
Derive the impact parameter dependence of the scattering angle θ(b) and of
the differential cross section σ (θ) for a light H+ (with charge e and mass m)
and massive ion (with charge Ze and mass M) in a fully ionized plasma (i.e.,
collisionless plasma) by considering the classical Rutherford scattering in a
Coulomb field.
Classical Rutherford scattering and its physical quantities are shown in
Figure 4.11. We may regard Coulomb scattering as resulting in small angle
scattering or small changes of the momentum under the long-range interac-
tion. Thus it is reasonable to assume that
sin θ ∼ θ = pT / p.
Here, the Coulomb force will induce a velocity perpendicular to the original
velocity and the resulting momentum is denoted by pT . We may calculate the
transverse momentum by integrating the effect of the force FT in the region
of interaction
∞ ∞
pT = FT dt = (Ze 2 /4π ε0r 2 ) cos βdt
−∞ −∞
∞
= {Ze 2 /4π ε0 (b 2 + v 2 t 2 )}{b/(b 2 + v 2 t 2 )1/2 }dt
−∞
= 2Ze 2 /4πε0 bv.
A
e vt
p = mv e
θ
b pT
β
r
Ze
FIGURE 4.11
Classical Rutherford scattering between a light H+ (with charge e and mass m) and a massive
ion (with charge Ze and mass M) in a Coulomb field.
Therefore,
θ (b) = pT / p
2Ze 2 1 2Ze 2
= · = , (4.50)
4π ε0 bv mv 4π ε0 mv 2 b
and the differential cross section can be obtained from Equation 4.38 as
Here, 1 − cos θ = 2/{1 + [b/(Ze 2 /4π ε0 mv 2 )]2 } in Figure 4.11, and we may
determine the cross section Qm by using
We thus obtain
Here ln is defined as
4π ε0 mv 2 λ D
ln = ln (4.54)
Ze 2
and is known as the Coulomb logarithm.
PROBLEM 4.3.2
Show that the Coulomb logarithm corresponds to the ratio b max /b min .
ii. The change of the momentum p must be much greater than the
uncertainty pd h. For an effective range of the potential V0 , the
conservation of energy gives vr p ∼ eV0 . The condition is thus
f (θ, φ) ikz
r e
(spherical wave)
π−θ
V(r)
Aeikz
(plane wave)
FIGURE 4.12
The planar wave of the incoming particle and the spherical wave of the scattered particle in
quantum theory of elastic scattering.
(c) After scattering. From the condition that scattered particles eventu-
ally escape to far distances unaffected by V(r ) and distribute them-
selves spherically around the scattering center (target), we may
conclude that the scattered particle can be described by a spheri-
cal wave. Thus, the total wave function is the sum of the scattered
spherical wave and the unaffected plane wave:
f (k; θ, φ)
(r) = exp(ikz) + exp(ikr ). (4.60)
r
The spherical term is normalized by f (k; θ, φ), which is known as the scatter-
ing amplitude. For example, if there is no interaction, the scattering amplitude
is zero and the planar wave is unaffected. The number of particles n scattered
into a small area dS normalized to the total number of the incident particles
n represents the probability of the scattering in a given direction (θ, φ) and is
therefore equal to
2
f (k; θ, φ) 1
n/n = exp(ikr ) dS ·
r exp(ikz)2
dS
= | f (k; θ, φ)|2 .
r2
We can then obtain the relation between the differential cross section and the
scattering amplitude in quantum theory as
This makes it possible to derive the differential equation for the radial part of
the wave function R(r ):
1 d 2 dR(r ) 2µE 2µ h̄ 2 l(l + 1)
r + − 2 V(r ) + R(r ) = 0. (4.63)
r 2 dr dr h̄ 2 h̄ 2µ r 2
h̄ 2 l(l + 1)
Veffl (r ) = V(r ) + . (4.64)
2µ r 2
As in the classical case, the effective radial potential is the result of
the combined effect of the interaction potential and the centrifugal term. When
we make transformation R(r ) = u(r )/r , the differential equation can be writ-
ten as
h̄ 2 d 2 u(r )
− + Veffl (r )u(r ) = Eu(r ). (4.65)
2µ dr2
We now consider the asymptotic behavior of the solutions. From Veffl → 0
when r → ∞ it follows that
u(r ) → Asin(kr ) + B cos(kr )
→ Asin(kr + ηl − lπ/2); at r → ∞.
The radial component of the wave function has the effect of scattering
through a phase-shift (ηl − lπ/2) at a long distance from the scattering cen-
ter (target). It should be noted that without scattering u(r ) is exactly equal to
the incoming plane wave, expressed below as Asin(kr −lπ/2). Consequently,
the solution for R(r ) is
and by combining Equations 4.59, 4.62, and 4.66, we obtain the general
solution for the wave function as
(r) = Al R(r )Pl (cos θ )
l
∼ Al (1/kr ) sin(kr − lπ/2 + ηl )Pl (cos θ) (at r → ∞)
l
∼ Al (1/2ikr ){exp[i(kr − lπ/2 + ηl )]
l
− exp[−i(kr − lπ/2 + ηl )]}Pl (cos θ) (at r → ∞).
We must now match the asymptotic forms of the above general solution
and Equation 4.60. For that purpose we must expand the incoming plane
wave, exp(ikz) in Equation 4.60 for r → ∞ as
(Rayleigh formula)
2l + 1 l
= Pl (cos θ ) 2i jl (kr ).
l
2
(2l + 1) 1
f (x) = a l Pl (x) and a l = f (t)Pl (t)dt, (4.67)
l
2 −1
and the asymptotic form of the spherical Bessel function in the limit x → ∞:
Thus, we may write the asymptotic expansion of the planar wave in the limit
r → ∞ as
exp(ikz)r →∞
∼ (2l + 1)i l (1/kr ) sin(kr − lπ/2)Pl (cos θ)
l
(2l + 1)i l
lπ
lπ
∼ exp i kr − − exp −i kr − Pl (cos θ)
l
2ikr 2 2
(4.69)
(r) − exp(ikz)
2l + 1 l
= Al R(r )Pl (cos θ ) − Pl (cos θ) 2i jl (kr )
l=0 l
2
1 lπ lπ
∼ Al exp i kr − + ηl − exp −i kr − + ηl Pl (cos θ)
l=0
2ikr 2 2
(2l + 1)i l
lπ
lπ
− exp i kr − − exp −i kr − Pl (cos θ)
l
2ikr 2 2
1 lπ
∼ Pl (cos θ ) exp i kr − Al exp(iηl ) − (2l + 1)i l
l=0
2ikr 2
1 lπ
− exp −i kr − [Al exp(−iηl ) − (2l + 1)i l ] (at r → ∞). (4.70)
2ikr 2
f (k; θ )
exp(ikr )
r
1 lπ
∼ exp i kr − [(2l + 1)i l exp(2iηl ) − (2l + 1)i l ]Pl (cos θ)
l=0
2ikr 2
1
ilπ l exp(ikr )
∼ exp − i (2l + 1){exp(2iηl ) − 1}Pl (cos θ) , (4.71)
l=0
2ik 2 r
and from the expression for f (k; θ), it is possible to obtain the differential
cross section σ (θ ):
1
f (k; θ ) = (2l + 1){exp(2iηl ) − 1}Pl (cos θ) (4.72)
2ik l=0
u(r)
V(r) = 0
V(r) > 0
r
kr0
ηl
FIGURE 4.13
The effect of phase-shift on the wave function.
σ (θ ) = | f (k; θ )|2
2
1 2l + 1
= 2 {exp(2iηl ) − 1}Pl (cos θ) . (4.73)
k l=0
2
Hence the only quantities that describe the effect of the scattering are the
phase-shifts ηl . The use of angular momentum quantum numbers l is analo-
gous to the use of an impact parameter in classical scattering. The scattering
cross section is obtained by summing up the different partial waves in respect
to l = 0, 1, 2, . . . , and the procedure shown here is known as the partial wave
method. This technique was developed for light scattering by Rayleigh and
applied to particle scattering by Faxen and Holtsmark.
PROBLEM 4.4.1
Show that the total cross section Q0 of the elastic scattering is given in partial wave
expansion as
Q0 (ε) = | f (k; θ)|d (4.74)
4π
= (2l + 1) sin2 ηl .
k 2 l=0
Exercise 4.4.1
Explain the phase difference ηl in Equation 4.66.
In Figure 4.13, we show two functions u(r ), one without any interaction with
a target (V(r ) = 0; solid line) and one with an interaction (V(r ) > 0; dashed
line). When the potential is repulsive, the outgoing wave function is pushed
to the outside as compared with the wave at V(r ) = 0, and then ηl > 0. On
the other hand, if it is attractive, then ηl < 0.
Exercise 4.4.2
In Figure 4.14 we show the spherical well potential:
V(r ) = −V0 r ≤ r0
= 0 r > r0 .
V(r)
0 r0
r
-V0
FIGURE 4.14
The spherical well potential (spherical attractive potential).
Calculate the total cross section and the phase-shifts for elastic scattering of
a low-energy particle.
From Equation 4.65, we have
d 2 u(r )
2µ l(l + 1)u(r )
22
+
{E − V(r )}u(r ) − = 0,
dr h̄ r2
and for the potential it follows that
<
{d 2 /dr2 + K 2 − l(l + 1)/r 2 }ul,k (r ) = 0; K 2 = (2µ/h̄ 2 )(E + V0 ); (r ≤ r0 ),
2 >
{d /dr + k − l(l + 1)/r
2 2 2
}ul,k (r ) = 0; k = (2µ/h̄ )E;
2 2
(r > r0 ).
The solution for R(r ) = u(r )/r must be finite at r = 0, and the u(r ) must be
zero at r = 0. For low-energy scattering that satisfies K 2 − l(l + 1)/r 2 > 0,
the component in the partial wave expansion will mainly be l = 0 (s-wave).
Therefore,
u0,k (r ) = u<
0,k (r ) = Asin K r
= u>
0,k (r ) = B sin(kr + η0 ).
The wave function u(r ) and its derivative satisfying the Schrödinger equation
must be continuous at r = r0 . Then,
Asin K r0 = B sin(kr0 + η0 ),
K Acos K r0 = k B cos(kr0 + η0 ),
and we obtain
(1/K ) tan(K r0 ) = (1/k) tan(kr0 + η0 ),
k
(or) η0 = tan−1 tan(K r0 ) − kr0 .
K
As a result, the scattering cross section for l = 0 (s-wave) is
4π 4π k
Q0 (k) = 2 sin2 η0 = 2 sin2 tan−1 tan K r0 − kr0 . (4.75)
k k K
V(r)
V0
r
0 r0
FIGURE 4.15
The profile of the spherical repulsive potential.
PROBLEM 4.4.2
In Figure 4.15 we show the spherical repulsive potential:
V(r ) = + V0 r ≤ r0
= 0 r > r0 .
Calculate the total cross section and the phase-shifts for elastic scattering of a low-
energy particle. Show that the total cross section is equal to Q0 = 4πr02 in the
limit V0 (r ) → ∞. Considering that the classical hard sphere cross section is equal
to Q0C = πr02 (see Exercise 4.3.1), explain why the quantum result is so different
(Q0q = 4πr02 ).
in the partial wave method. Here k is the wave number of the incoming
electron with energy ε. It follows that
4π
Qm (ε) = 2 (2l + 1) sin2 (ηl − ηl+1 ) (4.76)
k l
and also that
4π
Q0 (ε) = (2l + 1) sin2 ηl . (4.77)
k2 l
These formulae are sometimes used to analyze the differential cross section
of the electron and perform analytic extrapolations or integrations of the
experimental differential cross section.
πα 2
tan η1 = k − A1 k 3 , (4.79)
15a 0
where α and a 0 are the polarizability and Bohr radius, respectively. For
all higher-order phase-shifts (l > 1), the Born approximation is sufficiently
accurate:
π αk 2
tan ηl = .
(2l + 3)(2l + 1)(2l − 1)
The parameters A (the scattering length), A1 , D, and F should be
regarded as fitting parameters that may be obtained from transport coeffi-
cients of electrons and available cross section data. The value of parameter
A may be determined using the dimension of the molecule at zero energy
as Q0 = Qm = 4π A2 . For example, the best fit for argon is obtained for
A = −1.459a 0 , A1 = 8.69a 03 , D = 68.93a 03 , and F = −97a 04 . MERT was found
to work well for scattering of electrons on atoms at low energies where the
elastic collisions are the only scattering process. In other words, this approx-
imation is valid for electron–atom scattering below 1 eV.
PROBLEM 4.4.3
Calculate Q0 , Qm for argon from 0 eV to 1 eV. Show that the Ramsauer–Townsend
minimum (RTM) occurs when the contribution of both η0 and η1 are close to zero for
the same energy. Show that the RTM occurs at different energies for Q0 and Qm and
explain why.
1 α(r )e 2
Vpol (r ) → − , (4.80)
8π ε0 r 4
where α is the polarizability of the molecule. If there is a permanent dipole
moment or if the target is charged, then the interaction may be dominated
in the far region by electrostatic terms. Molecules with a permanent dipole
moment, although neutral, have a very long-range potential, and thus the
cross sections may be large, especially for low-incident energies. As a result
of these interactions and of the internal degrees of freedom of molecules, there
are numerous electron–molecule collisions that may occur; these are listed in
Table 4.1.
TABLE 4.1
List of Electron–Molecule Collisions
Process Reaction Scheme Energy Loss ∆ε(Typical)
ε
Elastic e + M −→ e + M ∼ 2(m/M)ε ∼0
ε>εex
Electronic excitation e + M −→ e + M j εex 10 eV
ε>εr
Rotational excitation e + M(r ) −→ e + M(r ) εr kT
ε>εv
Vibrational excitation e + M(v) −→ e + M(v ) εv 0.1 eV
ε>εd
Dissociation e + AX −→ e + A + X εd 10 eV
ε>εi
Ionization e + M −→ e + e + M+ εi 15 eV
ε>εa
Attachment e + M −→ M− εa ∼ kT
ε>εa
Dissociative attachment e + AX −→ A + X− εa ∼ eV
ε>εi p
Ion pair formation e + AX −→ e + A+ + X− εi p 20 eV
ε
Superelastic collision e + M j −→ e + M 0.01∼ –20eV
could survive long enough to suffer an additional collision with another gas
molecule and it would be stabilized, but usually the system is unstable and
it decays by autodetachment. Resonances are characterized by their energy
width ε, which may be associated with the compound’s lifetime through
the uncertainty principle.
PROBLEM 4.5.1
Calculate the energy width of the resonance that has a lifetime of 10−14 s. Describe
the elastic and the inelastic resonant scattering on a molecule.
Exercise 4.6.1
Using the selection rules, explain the optically allowed and forbidden transi-
tions in an atom.
25
Ionization +
24.6 He (2S)
33S
Energy (eV)
21 21P
2.06µ 23P
21S 1.08µ
20 He singlet
584Å 23S
19 He triplet
He
0 11S
FIGURE 4.16
Energy levels of helium. Radiative transitions between singlet and triplet states are not allowed.
Rules (iii) to (v) are only approximate, but even strict rules may be broken if
there is, for example, strong spin-orbit coupling in strong magnetic fields, due
to nuclear perturbations or for autoionizing states. Rule (iv) does not allow
transitions among different multiplicities (e.g., triplet to singlet transitions).
Transitions with L= 0 are very rare. For our purposes, it suffices to note that
transitions to the ground state of helium 11 S from 23 S and 21 S are forbidden
because of the S = 0 rule in the former case and because of the L = 0 → 0
rule in the latter case.
-15
10
Qm
-16
10
Cross section (cm2)
Qi
3 1
-17 Q2 S Q2 P
10
1
3
Q3 P
Q2 P
-18 Q others
10 1
Q2 P
1
Q2 S
3
Q3 P 1
Q3 S
-19
10 -2 -1 2 3
10 10 1 10 10 10
Electron energy (eV)
FIGURE 4.17
Set of collision cross sections of electrons in helium as a function of electron energy. The subscript
others denotes the summed effect of all other higher levels, and the subscript i indicates ionization.
All other levels are denoted by their terms.
+
Ar 2
P3/2 15.76 3V
15.76 3s2 3p5 2
P1/2
15 3p5
3p6
Energy (eV)
5
3p 5p 3p7
3p8
3p9
14 3p10
2p1 13.48 eV
2p2 13.33 eV
419.8 nm 3p5 4p 2p7 13.15 eV
13 2p9 13.08 eV
420.1 nm 2p10
772.376 nm
750.4 nm 772.421 nm
811.5 nm 1
12 1s2 P1 11.83 eV
3
5 1s3 P0 11.72 eV
3p 4s 1s4 3
P1 11.62 eV
1s5 3
P2 11.55 eV
104.8 nm
11
106.7 nm
Ar 1
0 3s2 3p6 S0
FIGURE 4.18
Energy levels of argon. The Paschen notation is also given.
Quite often Paschen notation is used to label the levels, especially for
heavier rare gases (see Figure 4.18). This is really just an assignment of levels
in their order rather than a notation associated with quantum numbers. Thus,
in this notation the ground state would be 1s1 , and the lowest metastable level
would be 1s5 . Two resonant lines from the four low-lying levels are in the
ultraviolet region, but most lines from the 4p and 5p states are in the visible
and near-infrared regions and are often used for diagnostics of argon plasmas.
Properties of the ground state, of the low-lying levels, and of the ionization
limit for the rare gases are given in Table 4.2.
Exercise 4.6.2
In the case of argon (see Figure 4.18), the ground state is 1 S0 . Explain the
allowed and forbidden transitions by using selection rules (and their possible
breakdown). Discuss the selection rules in the case of other low-lying levels
of the rare gases as given in Table 4.2.
TABLE 4.2
Energies; Configurations; and Terms of Ground States, Low-Lying Levels, and
Ionization Continuum for Rare Gases
Excited Levels
Gas Configuration Term Resonant (eV) Metastable (eV) Ionization (eV)
In the case of argon we have four low-lying levels: two of these are metastable
(3 P2 and 3 P0 ), and the other two (3 P1 and 1 P1 ) are resonant (with allowed
transitions to the ground state 1 S0 ). Transitions from the two metastables are
strongly forbidden by the rule that only J = 0, ±1 transitions, except for 0 →
0, are allowed. In this case, the 3 P1 level has a resonant transition to the ground
state in spite of the weaker rule that only S = 0 transitions are allowed.
All rare gases with the exception of helium have a similar configuration of
metastable and resonant states.
Cross sections for electron–argon collisions are shown in Figure 4.19. One
should observe that for energies below the threshold for the lowest metastable
(11.55 eV in this case) there are no inelastic losses, as in helium. However,
the elastic momentum transfer cross section has a broad minimum at around
0.23eV. This is the RTM, which is the first quantum effect observed for particles
with nonzero mass. When the RTM is present, the minimum value of the cross
section is two orders of magnitude lower than the maximum (and the value
that one would expect without the RTM). RTM is present in heavier rare gases
(Ar, Kr, Xe) and in many molecules.
One should observe that above the four lowest levels there are a large
number of excited levels, having transitions to those four levels including the
two metastables. If one wants to calculate the excitation rate to a particular
level, one needs to consider the excitation to all higher states and include the
radiative cascading by properly taking into account the branching ratios from
higher levels. In addition to excitation by electrons and radiativ ede-excitation,
there are several oher processes that provide nonradiative transitions to lower
or nearby resonant levels. These represent collisional quenching with other
gas molecules, electrons, or walls of the chamber and are discussed later. We
note, however, that some of the measurements of cross sections for excitation
suffer from the effect of cascading. All higher levels are also excited if the
-14
10
Qm
Qi
-16
10
Cross section (cm2)
Qex
Qexm
-18
10 Qex,(2p1)
Qex,(3p5)
-20
10 Qi,4p( D)
4
-2 -1 2 3
10 10 1 10 10 10
Electron energy (eV)
FIGURE 4.19
Cross sections for electron–argon scattering as a function of electron energy. Excitation is repre-
sented by effective summed cross sections and is separated into excitation of metastables and
excitation of higher levels. Three cross sections for specific transitions are added, two for ex-
citations into a neutral and the other for ionization into an excited ion from the ground state
neutral. These are employed for comparison between modeling and the optical diagnostics of a
low-temperature plasma in Ar.
Ak j
Nj = Ndirect
j + Nk ,
k> j i<k Aki
εT = εr + εv + εex . (4.83)
16
12
N(3S) + N(2D)
A3 Σu+
v' = 20
X1 Σg+
Potential (eV)
v=5
0
0.4 0.8 1.2 1.6 2.0 2.4 2.8
Internuclear distance (Å)
FIGURE 4.20
Potential energy curves for a nitrogen molecule.
1 2
T= µv + V(r ),
2 r
which can be transformed to the spherical coordinate system
2 2
1 2 dθ 1 dr
T= µr + µ + V(r ), (4.84)
2 dt 2 dt
L2 h̄ 2 d 2
H= − + V(r ). (4.85)
2µr 2 2µ dr2
The Hamiltonian may be separated into two terms, one corresponding to the
rotational θ -motion
L2
Hrot = (4.86)
2µr 2
and the other to the vibrational r -motion
h̄ 2 d 2
Hvib = − + V(r ). (4.87)
2µ dr 2
Hrot φr = εr φr , (4.88)
where
L2 L2
εr = = .
2µr 2 2I
h̄ 2l(l + 1)
εr = = Be l(l + 1), l = 0, 1, 2, · · ·. (4.89)
2I
Here, Be = h̄ 2 /2I is a rotational constant, and for most molecules it is on the
order of 10−4 eV∼ 10−3 eV, that is, meV. The transition energy between two
successive levels εr with l = 1 is
h̄ 2 h̄ 2
εrot = {l(l + 1) − (l − 1)l} = 2l = Be 2l. (4.90)
2I 2I
For molecules with a quadrupole moment, the selection rule is l =
±2 (l, l + 2), so in that case the energies of rotational transitions are εrot =
Be (4l + 6). (In both cases the transitions from the ground state are obtained by
choosing l = 0.) So far it has been assumed that the molecule is rigid; that is,
the distance between two atoms does not change as the molecule rotates. In
reality, however, the distance between molecules changes and so the moment
of inertia changes. This effect leads to the energy levels being described by
h̄ 2l(l + 1) 1
εr = = Be + α v + l(l + 1), l = 0, 1, 2, · · ·,
2I 2
where v is the quantum number of the vibrational level (discussed in the next
section) and α is a constant tabulated in books on spectroscopy.
Exercise 4.7.1
Estimate the energy of rotational transitions of an HCl molecule, and deter-
mine the ratio of the population of l = 1 to that of l = 0 at 300 K based on this
approximation. Prepare a table of exact values of Be for different molecules
and calculate exact energy losses for the first four levels of H2 , D2 , N2 , O2 ,
and CO2 .
For HCl, estimated values for the relevant quantities are
MH MCl
r ≈ 1.27 × 10−10 m, µ= ≈ 1.63 × 10−27 kg,
MH + MCl
which give the following estimate for the order of magnitude of the first
rotational level:
h̄ 2 h̄ 2 (6.63 × 10−34 /2π )2
εr (l = 1) = = ≈
I µr0 2 1.63 × 10−27 × (1.27 × 10−10 )2
≈ 0.423 × 10−21 J ≈ 2.64 × 10−3 eV.
TABLE 4.3
Rotational Transition Energy of Several Molecules with a
Quadrupole Moment
∆ε (eV)
Molecule q4 (ea20 ) Be (eV) 0–2 1–3 2–4 3–5
TABLE 4.4
Typical Wavelengths and Transition Energies for Rotational,
Vibrational, and Electronic Excitations
Transition Wave Wavelength (m) Energy(eV)
will have the selection rule l = ±1 for molecules with a permanent dipole
and l = ±2 for molecules with quadrupole moment. Gases such as hydro-
gen have very large rotational energies that are comparable with and even
larger than their mean thermal energies at room temperature. Thus, we may
expect that a only few rotational levels will have any significant population.
For this reason, it should be possible to develop an accurate theory to repre-
sent rotational excitation and perhaps to use a normalizing factor for all the
levels. For parahydrogen at 77 K, in which only the ground rotational level
was populated, it was actually possible to determine the rotational excitation
cross section (see Figure 4.21) for the l = 0 → 2 transition from the transport
data. It is then possible to extend the analysis to the l = 1 → 3 transition from
the data for normal hydrogen. The good agreement with available theories
provided justification for the use of theory to determine cross sections for a
few higher transitions that exist at room temperature.
Exercise 4.7.2
What are para, the ortho, and the normal hydrogen? Calculate the population
of rotationally excited states under thermal equilibrium at room temperature
and at 77 K for hydrogen and deuterium. The nuclear spin for hydrogen
is I = 1/2 and that for deuterium is I = 1.
The para and ortho states of diatomic molecules are those in which the
nuclear spins of the two atoms are in the same direction and in the opposite
1.0
Qr (1-3)
Cross section (10 cm )
2
Qv (0-1)
-16
0.5
Qr (0-2)
Qv (0-2)
0.0
-2 -1
10 10 1 10
Electron energy (eV)
FIGURE 4.21
Rotational and vibrational excitation cross sections for hydrogen obtained from the swarm data.
Nl pl exp(−εr /kT)
= ∞ . (4.91)
N l=0 pl exp(−εr /kT)
Here, the statistical weight is equal to pl = (2I + 1)(I + a )(2l + 1). Here a = 0
for even l and a = 1 for odd l for molecules like H2 . The values of a are exactly
the opposite for molecules like D2 and N2 , which depend on the value of the
nuclear moment. The results of the calculations are given in Table 4.5.
TABLE 4.5
Rotational-Level Populations of H2 and D2
l=0 l=1 l=2 l=3 l=4 l=5
PROBLEM 4.7.1
A good approximation of the realistic potential of diatomic molecules is the Morse
potential
2
V(r ) = D 1 − e −a (r −r0 ) .
Show that for small departures from the equilibrium position r0 , this potential may
be represented by the harmonic potential 4.93. Perform the calculations and plot
both potentials for the example of KCl where D = Equation 4.42 eV and a 2 =
0.602 × 10−20 m−2 .
PROBLEM 4.7.2
mass µ
Derive the solution of the classical vibrational motion of a particle of reduced
for force −2A r . Compare the angular frequency of the oscillation ν = 2A2 /µ to
2
that given in Equation 4.95. Relate the classical energy to that in Equation 4.96 and
explain.
Qm
Q eC
Q e1s2s
Q re02
-16
Q di
10 Q re13
Cross section (cm ) Q eB
2
Qi
-18 Q v1
10
Q v2 Q eE
Q re24
Q e1sn3
10
-20 Q re35 Q elsls
Q e2p2s
-2 -1 2 3 4
10 10 1 10 10 10 10
Electron energy (eV)
FIGURE 4.22
A set of cross sections for electron scattering of H2
PROBLEM 4.7.3
Estimate the magnitude of εv for H2 .
Exercise 4.7.3
Assuming that the potential energy of a diatomic molecule at equilibrium
distance is zero (V(r0 ) = 0), the energy of the lowest vibrational state is
εv=0 = h̄ν/2. Discuss this zero point energy from the standpoint of quantum
mechanics.
Quantum phenomena have a finite energy even at 0 K as fluctuation. It is well
known that the nonzero energy arises from the uncertainty principle.
10
Q elast. Q tot.
Cross section (10-16 cm2)
1 Q vib.
Q att.?~× 50 Q ion.
0.1 Q exc.
Q rot.
-2
10
Q vib.
-3
10 -2
0.1 1 10 10
Electron energy (eV)
FIGURE 4.23
A set of cross sections for electron scattering on NO. Vibrational excitation shows a strong reso-
nance that is also present in the total scattering cross section.
V(r)
5
4
Ec c> 0
3
Potential energy
2
1
v=0
excited state
grounded state
VD
v=0
Pr (v)
r
Transition r0
probability 0 Internuclear distance
FIGURE 4.24
Dissociative electronic excitation and Franck–Condon principle of a molecule.
Exercise 4.7.4
List the most important selection rules for transitions between excited states
of molecules.
Some of the most important selection rules are:
i. = 0, ±1, where is the orbital angular momentum;
ii. = 0, where is the spin quantum number and = S,
S − 1, · · ·, −S;
iii. = 0, ±1, where = | + |, and for 0 → 0 it follows that
J = 0; and
iv. When the vibrational quantum number v changes, the v = ±1
transitions are the strongest but other transitions are not forbidden.
PROBLEM 4.7.4
Interpret the molecular terms X1 g+ and A3 u+ of N2 and discuss whether the upper
state is metastable.
TABLE 4.6
Dissociation Energy (VD ) of Ground-State
Molecules and Their Ions
Molecule VD (eV)
H2 4.48
H+ 2 2.63
B2 2.93
C2 6.24
N2 9.76
O2 5.11
F2 1.59
Cl2 2.48
Br2 1.97
I2 1.54
Exercise 4.7.5
Describe a typical example of dissociation of molecules in thermal and
nonequilibrium plasmas.
Dissociation will occur if energy sufficient to cross the dissociation barrier VD
in Figure 4.24 is available. Data for dissociation energies for several molecules
are given in Table 4.6. In thermal plasmas, molecules will be dissociated
through a number of rotational and vibrational transitions by electron–
molecule, and even more so by molecule–molecule or ion–molecule colli-
sions. However, in nonequilibrium low-temperature plasmas, only electron–
molecule collisions to upper electronic excitation states will result in the dis-
sociation due to the low kinetic energy of heavy particles.
and the cross sections for excitation of Balmer α and β emission from a CH4
molecule are shown in Figure 4.25.
-1
10
beam experiment
swarm analysis
-3
10
Hβ
-4
10 2 3
10 10 10
Electron energy (eV)
FIGURE 4.25
Cross section for electron-induced dissociative excitation of CH4 into H(n = 3, 4) levels.
Exercise 4.7.6
Prove Equation 4.98 for a cross section of the superelastic collision based on
the principle of detailed balance.
The principle of detailed balance states that, in thermal equilibrium (at a
temperature Tg ), the number of transitions of electrons from higher energies
to lower energies due to inelastic collisions is equal to the opposite process
of superelastic transition from low to high energies. In other words, for a
σs24 superelastic
1
Cross section (10-16 cm2)
-1
10 inelastic
σv24
-2
10 -1 2
10 1 10 10
Electron energy (eV)
FIGURE 4.26
Cross section for electron-induced excitation and superelastic de-excitation of 2–4 mode of
vibration of CH4 .
8 6
6
σ (10-16 cm2)
1-1 4 1-2
4
2
2
0 0
2 3 4 5 2 3
4 4
1-3 1-4
σ (10-16 cm2)
2 2
0 0
2 3 2 3
Energy (eV) Energy (eV)
FIGURE 4.27
Cross section for the electron-induced vibrational excitation of a N2 (X1 g ; v = 1). 1–1 denotes
elastic scattering, and the excitations to higher vibrational state are indicated by 1–v.
10 2 3
P2 - 2P
3 3P1
P2 -
10
Cross section (10 -16 cm 2)
1 3P2 - ionization
10 -1 1P1
10 -2
3
P2
10 -3
10 -1 1 10 10 2 10 3
Electron energy (eV)
FIGURE 4.28
Cross section for the electron excitation of the metastable Ar(3 P2 ).
shown together with ionization. The thresholds for these processes are 1.55
eV and 4.16 eV, respectively. Hence these processes may become very impor-
tant once the metastable density is sufficiently high. For comparison we have
also shown the cross sections to the lowest metastable state Ar(3 P2 ) and the
resonant state Ar(1 P1 ) from the ground state Ar(1 S0 ).
In Table 4.7 we show the data required to determine the double differential
cross section σ (ε, εs ) for atomic ionization.
The ionization cross section is usually determined very accurately in
electron beam experiments and it is often used directly in cross section sets
for plasma modeling. In Figure 4.29 we show data for ionization of an argon
atom to several different charge states, Ar+ , Ar2+ , and so on. We should also
note that there is a possibility that excited states of an ion or even multiply
charged ions may be excited/ionized directly from the ground state of the
neutral. For most nonequilibrium plasmas for which applications currently
exist, only the ionization leading to a singly charged ion needs to be consid-
ered. The collisions of ions and electrons leading to further ionization may be
TABLE 4.7
Double Differential Cross Section Parameters for
Ionization of Several Atoms and Molecules
Molecule εi (eV) ε (eV)
He 24.6 15.8
Ne 21.6 24.2
Ar 15.8 10.0
Kr 14.0 9.6
Xe 12.1 8.7
H2 15.4 8.3
O2 12.2 17.4
N2 15.6 13.0
CH4 13.0 7.3
neglected. The ionization cross section peaks at a few hundreds of eV, and at
these energies it becomes the dominant collision process, even exceeding the
elastic scattering. For molecules, ionization is often accompanied by disso-
ciation. Simultaneous ionization and excitation often occur from the ground
state of the neutral molecule, and emission from the excited ions may be used
as a probe for high-energy electrons. Close to the ionization threshold, the
cross section increases as
Qi ∼ (ε − εi )1.127 ,
i=0
-16
10
Cross section (cm )
2
i=1
-17 i=2
10
i=3
i=4
i=5
-18
e + Ar i+ Ar (i+1)++ 2e
10
2 3
10 10 10
Electron energy (eV)
FIGURE 4.29
Ionization cross section for e + Ai+ → A(i+1)+ + 2e (i = 0–5). This cross section is different from
one for multiple ionizations from the ground state of a neutral atom.
according to Wannier’s classical theory. For very high energies the cross
section decays according to the Bethe–Born approximation as
1
Qi ∼ (Aln ε + B),
ε
where A and B are the terms for optically allowed and forbidden transitions,
respectively.
and
e + AX → (AX∗ + e) → A+ + B− + e. (4.102)
In the first step, the electron is attached to the excited state [AB]∗ . The state
may decay, through dissociation with probability pd , to form two fragments,
one being a negative ion (Equation 4.100). The other possible channel is for the
excited state to be stabilized in collisions with the third particle M, preferably
a heavy buffer gas molecule. It will then proceed to form a stable ground-state
negative ion of the parent molecule (Equation 4.101). Excess energy will be
taken away by the third particle. Another way to describe these processes
is by following the dynamics on potential energy curves V(r ), as shown in
Figure 4.30.
Potential energy
A+X A+X A+X
AX AX AX
A+X-
A+X-
AX-
r0 rc
Internuclear distance
FIGURE 4.30
Potential energy curves for dissociative and nondissociative electron attachment.
not have the energy to dissociate will oscillate in the potential well
until autodetachment occurs. Even if the minimum of the excited
state is lower than the ground state of the neutral molecule there is
a threshold for this process to occur.
iii. If the potential bottom of the excited state (AX− )∗ is below that of the
lowest level of the neutral molecule, then there is a chance that the
curves will cross close to the minimum and that even electrons with
zero energy may be attached. The excited negative ion will spend
time oscillating until it is stabilized by collision with the third body
or until autodetachment releases the bound electron.
The overall rate of attachment will strongly depend on the ratio of the
mean free time between collisions with gas molecules and the lifetime of the
excited state. A good example of the potential curves of ground-state neutrals
and negative ions is shown in Figure 4.30. An example of the potential curves
is given in Figure 4.31 for oxygen. There, one may expect nondissociative
attachment to occur due to crossing of the ground state of the molecule by
the potential curve for the negative ion O− 2 . The electron affinity E a is the
energy released if a zero-energy electron is detached from a negative ion. The
values of affinities for electrons for several atoms and molecules are shown
in Table 4.8.
TABLE 4.8
Electron Affinity
Atom Ea (eV) Molecule Ea (eV)
F 3.40 F2 3.08
Cl 3.61 Cl2 2.38
Br 3.36 Br2 2.51
I 3.07 I2 2.58
O 1.46 O2 0.44
S 2.08 SF6 0.60
C 1.26 CF2 2.10
CF3 1.92
15 (a)
Potential energy (eV)
10 • 2
-
O2 ( Π u )
(b)
• 2
- +
O2 ( Σ g ) 6
-
O ( ∆g
1 2
Πu )
0 - 0
O2 ( Σ g )
3
2 4 6 8 10
-2
1.0 2.0 3.0 Electron energy (eV)
Internuclear distance
FIGURE 4.31
Potential energy curves for oxygen showing the ground state O2 (3 g− ), metastable O2 (1 g ), and
higher excited states (a), cross sections for dissociative attachment of the O2 (3 g− ) and O2 (1 g )
states (b).
The O− ions produced may then undergo any of the following reactions:
ε>1.02eV
O− + O2 −−−−−→ O + O−
2 (charge transfer), (4.104)
ε>0.39eV
O− + O2 −−−−−→ O3 + e (associative detachment), (4.105)
or
ε>1.46eV
O− + O2 −−−−−→ O2 + O + e (electron detachment). (4.106)
-15
10
Qm Qi
B3Σu
-16
10
c1Σu+C3∆u
Cross section (cm )
vib.
2
10
-17 a1∆g others
+
A3Σu
-18
10
+
b1Σg
Qa Qp
-19
10 -2 -1 2 3
10 10 1 10 10 10
Electron energy (eV)
FIGURE 4.32
A set of cross sections for electrons in O2 .
-14 Q a1
10
Qm Qi
Cross section (cm )
2
-16
10 Qe
Q a3
Q a2 Q a4
Q a5 + +
10
-18 Q [N2 (B3 Σu )]
Qv
10
-20 Q [N2 (C3 Πu)]
-2 -1 2 3
10 10 1 10 10 10
Electron energy (eV)
FIGURE 4.33
A set of cross sections for electrons in SF6 . There is one nondissociative (a 1 ) and several disso-
ciative channels for attachment. Cross sections for excitation of two levels of N2 are also shown.
TABLE 4.9
Cross Sections for Dissociative
Attachment to H2 (1 g+ )
v r εa (eV) Qa (cm2 )
TABLE 4.10
Thermal Attachment Rate Coefficients (300 K)
Molecule ka = Ra /N (cm3 s−1 )
d
ne (t) = −krm ne (t)n p (t), (4.113)
dt
where the rate coefficient is associated with the cross section through
TABLE 4.11
Recombination Processes and Estimated Rate Coefficients (300 K)
Recombination Thermal Rate
Process Reaction Scheme Coefficient
e + NO+ → N + O + εk .
Exercise 4.8.1
Explain why radiative attachment and recombination are less effective than
other nonradiative processes.
Radiative attachment is described by
A + e → A− + hν.
The lifetime of the excited state is typically on the order of 10−8 s, and the
time that the electron spends in the neighborhood of the atom is on the order
of 10−15 s (dimension 10−7 cm and velocity 108 cm s−1 ). Thus, the probability
V(r)
AB
A+ + B
AB+
Potential energy
VD+
εk
AB
A+B
Internuclear distance
FIGURE 4.34
Potential diagram for dissociative recombination of the AB+ molecular ion.
that radiative stabilization of the transient negative ion will occur is 10−7 .
Therefore, the effective cross section will be on the order of 10−21 cm2 . For
slower electrons the cross section will be larger. For pairs of electrons and
ions, the interaction will also occur at a greater distance, which will lead to
a somewhat longer interaction time and a larger cross section. In the case
of dissociative processes we do not need photon emission to reduce the en-
ergy of the electron from a free to a bound state. Thus, the efficiency will be
defined by the collision cross section only. Radiative attachment and recom-
bination will be roughly seven to five orders of magnitude less effective than
the corresponding dissociative processes.
Exercise 4.8.2
Explain the inverse processes of radiative recombination and photo-
ionization. Define and explain the dissociative and dielectronic recombination
and their inverse processes.
It is obvious that the photo-ionization, that is,
hν + A ↔ A+ + e,
TABLE 4.12
Ion–Molecule Collisions
Process Reaction Scheme
Elastic Scattering X± + AB → X± + AB
Positive Ions
Nonresonant charge transfer X+ + AB → AB+ + X
Resonant charge transfer AB+ + AB → AB + AB+
Dissociative charge transfer X+ + AB → A+ + B + X
Projectile excitation X+ + AB → X+∗ + AB
Target excitation X+ + AB → X+ + AB∗
Target ionization X+ + AB → X+ + AB+ + e
Atom interchange X+ + AB → XA+ + B
Ion interchange XB+ + A → X + AB +
Addition X+ + AB → XAB+
Three-body association X+ + A + B → XA+ + B
Negative Ions
Heavy atom transfer X− + AB → XA + B−
Associative detachment X− + AB → XAB + e
Detachment X− + AB → X + AB + e
-14 Qe.m.t.
10
Qc.t.
Cross section (cm )
2
-15
10
-16
10
-17
10 Qiso.
(Ar+ + Ar)
-18
10 -1 2 3 4
10 1 10 10 10 10
Relative energy εr (eV)
FIGURE 4.35
A set of cross sections for the Ar+ ions in Ar.
neutral to the ion. Therefore, the fast neutral continues with almost the same
kinetic energy as the incoming ion. The ion is left with a very small energy,
equal to that of the target atom.
If we have ions in their parent gas, the charge transfer is said to be res-
onant (because of the matching energy levels) or symmetric. This process is
-14
10
Qm
-15
10
Cross section (cm )
2
-16
10
Q chtr
-17
10
Q j (B3 )
-18
10
Q j (391nm)
-19
10 -1 2 3 4
10 1 10 10 10 10
Labratory energy (eV)
FIGURE 4.36
A set of cross sections for N+ ions in N2 .
proton transfer: H2 S+ + H2 O → H S + H3 O+ ;
(hydrogen) atom transfer: C2 H4+ + C2 H4 → C2 H5+ + C2 H3 ;
hydride ion transfer: Ra+ + Rb H → Ra H + Rb+ .
H2+ + H2 → H3+ + H.
which play a role in atmospheric chemistry. Cross sections (rates) for ion–
molecule reactions are quite high and often are almost as large as the
orbiting limit. Orbiting occurs for special combinations of the potential and
centrifugal forces, when the radial component of velocity vanishes and the
projectile makes many revolutions around the target before being released
(see Section 4.5). The extended interaction time makes it possible to complete
long-lived transitions.
As an example of reactions with negative ions, typical reactions involving
oxygen ions are
O− + H2 O → OH − + OH,
O− + 2O2 → O3− + O2 and
O− + O2 → O2− + O,
but there are also reactions involving thermal detachment (including associa-
tive detachment):
O− + O → O + O + e,
O− + O2 → O + O2 + e,
O2− + O2 → O2 + O2 + e,
O2− + O → O3 + e.
Exercise 4.9.1
Determine the rate for orbiting collisions between a slow ion and neutral
with an attractive interaction potential given by V(r ) = −αe 2 /8π ε0r 4 for
r ≥ a and V(r ) = ∞ for r < a , where the interaction is due to the polarization
determined by polarizability α.
The effective radial interaction energy is equal to
b2 E
Veff (r ) = V(r ) + ,
r2
where the second term is due to the centrifugal force, E is the energy of the
incoming particle at r → ∞, and b is the impact parameter (see Equation 4.43).
The radial velocity is equal to
vradial = 2(E − Veff (r ))/M,
and for orbiting motion, the radial velocity must be equal to zero. Combining
these two conditions one can derive the value of the critical impact parameter
b c , which corresponds to the orbiting motion as
1/4
αe 2
bc = .
2π ε0 E
The orbiting cross section is thus equal to
1/2
αe 2
Qor = π b c2 =π .
2π ε0 E
For a finite probability of transition, the cross section should be multiplied by
probability ζ for the transition during orbiting.
-15
10
Qtotal
Cross section (cm )
2
-16 Qm
10
Qi
Q(811nm)
-17 Q(795nm)
10
-1 2 3 4
10 1 10 10 10 10
Labratory energy (eV)
FIGURE 4.37
A set of cross sections for fast Ar atoms in argon. Ionization and ultraviolet excitation cross
sections are identical.
Studies of Townsend discharges at very high E/N revealed that fast neutrals
are actually more effective in producing excitation (and possibly ionization)
than are ions. The energy distribution of fast neutrals upon formation is the
same as that of ions, but they do not gain energy. In Figure 4.37 we show a
set of data for cross sections for fast neutral Ar(fast)-Ar collisions [9].
It was discovered that the production of fast neutrals may also take place
on a surface. Collisions of fast ions with a surface often result in reflection and
neutralization, sometimes with small or negligible energy loss. Furthermore,
for molecular ions, the breakup of molecules occurs and molecular neutral
fragments are reflected.
TABLE 4.13
Collisions of Excited Species
Process Reaction Scheme Energy Condition
Leading to Ionization
Ne (3P0.2) + Ar (1S0)
-15 Qi (Penning)
10
Cross section (cm )
2
Qi (associative)
-16
10
Qelas × 0.01
-17
10
-3 -2 -1 2 3
10 10 10 1 10 10 10
Relative energy (eV)
FIGURE 4.38
Cross sections for the Penning ionization, chemi-ionization, and elastic scattering.
for comparison. One can see that Penning ionization is 10 times more efficient
at relatively low energies (which is relevant for applications, because standard
collisions occur at room temperature, which is equivalent to 24 meV). At
higher energies the cross section for the associative process becomes equal
to and even larger than that for the Penning ionization. The cross section
for the elastic process is another 10 times larger. Thus a small mixture of Ne
in Ar will have a very efficient ionization, and as such it is used in many
applications.
PROBLEM 4.9.1
By using the cross sections from Figure 4.38 and thermal Maxwellian distributions
for Ne(3 P0,2 ) and Ar(1 S0 ) for room temperature (300 K), calculate the rate coefficient
k(cm3 s−1 ) for Penning and chemi-ionization.
Exercise 4.9.2
Explain how the ionization rate in the mixture of Ar and Ne exceeds that of
either Ar or Ne as shown in Figure 4.39.
For the purpose of this evaluation, it is better to represent ionization through
the number of ionizations per unit time, ki = vd (α/E) · (E/N), where vd is the
drift velocity. The production of electrons is given by
dne ∗
= ki−Ar ne NAr + ki−Ne ne NNe + kPenn NAr NNe .
dt
Here, ki denotes the ionization rate coefficients for two gases and kPenn the
rate coefficient for Penning ionization. At the same time, the population of
-3 -2 -1
10 10 10 (= NAr / NNe+Ar)
-4
10
-1
3×10
-2
10
-5
10
α / E (V )
-1
-6
3×10
1
-3 (Ar 100 %)
10 -6
0
10
(Ne 100 %)
10 100 1000
E / N (Td)
FIGURE 4.39
Ionization coefficients normalized by the field (α/E[V −1 ] number of ionizations per unit of
potential difference) for neon–argon mixtures.
neon metastables is
∗
dNNe ∗ ∗
= k Ne ∗ ne NNe − kPenn NAr NNe − k Q−Ne NNe NNe .
dt
Here, k Ne ∗ is the electron excitation coefficient, k Q is the two-body quenching
coefficient of the neon metastable by the ground-state neon (quenching is
discussed in the next section), and we have safely assumed that Penning
ionization dominates the quenching of neon metastables by argon. Under
stationary conditions this leads to
∗ k Ne ∗ ne NNe
NNe = .
kPenn NAr + k Q−Ne NNe
The effective growth of ionization is therefore
dne
(= Ri ne ) = ki−Ar ne NAr + ki−Ne ne NNe (4.114)
dt
k Ne ∗ ne NNe
+ kPenn NAr .
kPenn NAr + k Q−Ne NNe
We can see that the effective ionization will be modified by a term that depends
strongly on the abundance of the constituents in the mixture. As Ar density in-
creases, the third term will increase substantially. Subsequently, Ar will affect
the energy distribution through its inelastic processes with lower thresholds
+
He
25
* 1
He 2S
20 + 2 -7
3 Cd D3/2 (τ ∼ 10 s)
2S 2
Penning D5/2
Energy (eV)
1 1
1S S
He Cd
FIGURE 4.40
Energy diagram of He and Cd.
than those for Ne. However, when the argon density becomes large enough
that the Penning term exceeds the quenching by Ne, the result will simply
be
dne
(= Ri ne ) = ne (ki−Ar NAr + ki−Ne NNe + k Ne ∗ NNe ). (4.115)
dt
For small E/N the ionization rate Ri will be equal to the excitation coefficient
for metastable Ne. Thus, a way to measure the excitation rate could be to
determine the ionization rate for the Penning mixture.
Exercise 4.9.3
Explain the principle of operation of the He–Cd laser (its energy diagram is
shown in Figure 4.40).
In collisions of metastable He(3 S1 , 1 S0 ) with the ground state of Cd(1 S0 ), it is
possible to ionize and excite Cd to high-energy levels Cd+ (2 D3/2,5/2 ), which
are then more strongly populated than the lower-energy levels Cd+ (2 P3/2,1/2 ).
Because inverse distribution of the population is achieved, it is possible to
achieve lasing on the two lines that originate from the transitions (λ = 325
nm and 441.6 nm).
TABLE 4.14
Quenching of Excited States Often Used in Plasma Diagnostics
Molecule Feed Excitation Rad-Lifetime Emission Quenching Coefficient
(Excited State) Gas Energy (eV) τrad (ns) hν (nm) kq (cm3 s−1 )
dNm
= 0 = kexc
m
ne N − Am Nm − kqm Nm N,
dt
which can be solved as
m
kexc ne N
Nm = ,
Am + kqm N
and therefore
m
kexc ne N
Nm Am = Im = kqm N
.
1+ Am
the value of intensity without the effect of quenching (which may be used to
exc
determine the excitation rate km for the state m from the intensity of radia-
tions).
Several quenching processes are listed in Table 4.14. Quenching of excited
molecules may be induced by collisions (collision–radiative decay) that transfer
molecules to a different excited state that subsequently decays radiatively.
This can happen both for short- and long-lived states. In the latter case the
most important transitions are made to nearby resonant states. Thermal col-
lisions with molecules may provide enough energy to make the transition to
nearby excited levels (which are within several times of kT from the orig-
inal level). This may happen within one multiplet. Emission of a photon,
even from a metastable state, may be induced by collisions (collision-induced
emission). During collisions a temporary dipole moment is induced (or a tran-
sient molecule is formed), which provides a mechanism for radiative decay.
Special attention should be paid to the quenching of metastable levels.
Two-body quenching consists mainly of collision-radiative and collision-
induced emission processes, as outlined above. Three-body quenching mainly
leads to the formation of eximers [A∗2 ]:
PROBLEM 4.9.2
Determine the effective lifetimes of the metastable levels of rare gases at 1 Torr and at
10 Torr. Compare them with natural radiative lifetimes.
TABLE 4.15
Radiative Lifetimes, Two- and Three-Body Quenching, and Diffusion
Coefficients of Rare Gas Metastable States in Their Parent Gases
Metastable Natural Lifetime Two-Body Three-Body Diffusion
State τrad (s) k2b 3 −1 k3b 6 −1 ND(1018 cm−1 s−1 )
q (cm s ) q (cm s )
TABLE 4.16
Rate Coefficients for Quenching of Rare Gas
Metastables by Molecular Gases (10−10 cm3 s−1 )
Metastable N2 H2 O2 NO CH4 CF4 SF6 Cl2
metastables [10].
PROBLEM 4.9.3
Calculate the values of the cross sections for quenching of rare gas metastables by
their parent molecules from the rate coefficients given in Table 4.15.
Here, δ is the factor of unharmonicity and ε01 is the energy of the first vibra-
tional level. This distribution will have a plateau and even a maximum at
moderate vibrational quantum numbers. The actual vibrational distributions
should be calculated from kinetic equations, and they will be affected by the
electron excitation from the ground and from vibrationally excited levels and
by the V–T and V–V processes. However, excitation to higher levels will be
mainly due to V–V transfers. On the other hand, highly excited molecules
may dissociate to neutral fragments or have a very large cross section for
dissociative attachment.
–
O
–
2 O2
-20
3.5×10 cm 2
0
0.5 1.0 1.5 2.0 2.5 3.0
Photon energy (eV)
FIGURE 4.41
Cross sections for photon-induced detachment for negative ions in oxygen.
is defined by coefficient A21 , and thus the total number of emitted photons is
equal to A21 N2 . If the transition 2 → 1 is the only radiative decay of the level
2, then the lifetime of the excited level is equal to τ2 = A−1 21 . In the presence
of radiation with a density (per unit frequency) of line radiation of transition
2 → 1 equal to ρ(ν21 ), it is possible that either the radiation will be absorbed
by the atoms in the lower level with probability B12 ρ(ν21 ) or a stimulated
emission by the photons of exactly the same frequency will occur from the
atoms in the upper levels, with probability B21 ρ(ν21 ).
Exercise 4.10.1
Assuming thermal equilibrium for a system with two excited levels 2 and 1,
find the relationships between Einstein’s coefficients.
In thermal equilibrium, the number of emissions (spontaneous plus stimu-
lated) must equal the number of absorptions, so we have
A21 N2 + B21 ρ(ν21 )N2 = B12 ρ(ν21 )N1 .
At the same time, the populations of two excited states can be determined
from the Boltzmann distribution
N2 g2
= e hν21 /kT .
N1 g1
Therefore, we can solve the equation for radiation density as
8π hν21 3
/c 3
ρ(ν21 ) = , (4.116)
e hν21 /kT − 1
we obtain the following relations:
3
8π hν21
A21 = 3
B21 . (4.118)
c
Stimulated emission reduces absorption, and if we have population inver-
sion, that is, if the density of the excited state exceeds the density of the
lower state, we will have overall amplification of emission. This is a possi-
bility in nonequilibrium discharges and is the fundamental principle behind
gas lasers. As discussed above, the kinetics of excited states will be more
complicated due to possible nonradiative transitions such as quenching.
g1 2π e 2 ν 2
A21 = f 21 . (4.119)
g2 ε0 mc 3
The absorption of radiation can be represented by an absorption coefficient,
and it will be associated with the attenuation of a beam of radiation of intensity
Iν = cρν as it passes through a gas of absorbers
1 dIν hν
− dν = [B=12 dN1 − B21 dN2 ] kν dν, (4.120)
Iν dx c
where dN refers to a subgroup of atoms that are absorbed within the fre-
quency interval dν, and kν is the absorption coefficient. If we integrate over
all frequencies, we have
hν hν c 3 A21 g2 g1 N2
kν dν = [B12 N1 − B21 N2 ] = 3 g
N1 1 − . (4.121)
c c 8π hν21 1 g2 N1
In addition to the integration over all frequencies covered by the line profile,
one can independently perform integration along the length of the absorbing
medium (where one can expect variation in the density of absorbers). Thus
the solution to Equation 4.121 for a uniform absorbing medium of length l is
and therefore one can determine the absorption coefficient kν from the exper-
imental measurement of absorption profiles. One can then define the optical
l
depth as τν = 0 kν dx = kν l. The frequency profiles of lines can be determined
e2 γ /4π
kν = N1 f 12 , (4.125)
(4πε0 )mc (ν21 − ν)2 + (γ /4π )2
where γ is the width of the line and (ν21 γ ). If we integrate the absorption
coefficient over all frequencies (i.e., over the line profile) we obtain [11]
∞
e2
kν dν = N1 f 21 .
0 4ε0 mc
PROBLEM 4.10.1
Evaluate the profile due to Doppler broadening, and determine k0 and the width of
the line by assuming a thermal distribution of absorbers at temperature Tg .
PROBLEM 4.10.2
Plot the Lorentzian and Doppler profiles (normalized to the same integral value) with
the same half width and compare the center of the line and the width.
• g= √1.875
k0 l πln(k0 l)
, for parallel plane geometry, and
• g= √ 1.6
k0 R πln(k0 R)
, for cylindrical geometry,
Source
B
+ M e A
+
hv
B
A
+ e e
plasma +
e A
W
H
– * e
A
H W
e
Metal
FIGURE 4.42
A schematic representation of plasma surface interactions in a typical plasma chamber.
Because production and loss of charged particles are of key importance for
the maintenance of plasmas, we primarily consider the production and the
loss process of charged particles on a surface. On the other hand, for plasma
technologies we require additional processes such as chemical reactions at the
surface, physical sputtering, and implantation. A special form of interaction
between a plasma and a surface is the formation of the sheath that protects
plasma from losing the charged particles to the conductive surfaces. Thus,
the properties of charged particles arriving at the surface and reaching the
plasma from the surface will strongly depend on the sheath potential. We
should note that inasmuch as we mainly have surface modification in mind
when we discuss plasma processing, the understanding of plasma surface
interaction is critical.
11e
V(r ) = − . (4.126)
4π ε0r
3p
3p band
3s
Potential energy of
2p 3s
2s 2p
11e
V(r) = 2s
4πε0r
1s
Na level
r 1s
Na Na Na Na
a a a
FIGURE 4.43
Energy levels of a single atom of Na (a) and of atoms in a solid (a one-dimensional array of Na atoms) (b).
125
© 2006 by Taylor & Francis Group, LLC
126 Plasma Electronics: Applications in Microelectronic Device Fabrication
11e 11e
V(r ) = − − (4.127)
4π ε0r 4π ε0 (a − r )
11ea
=− . (4.128)
4π ε0r (a − r )
When the electrons of the outermost shell (3s) with energy level εk keep the
following relation at the maximum value of V(r ),
then each of (3s) electrons is not potentially restricted in one atom but free
between N atoms (see Figure 4.43b). Here, εi is the ionization energy of Na.
When the external field is applied from both sides of the one-dimensional
lattice, (3s) electrons can move to the neighboring atoms as free electrons.
The electrons that may freely leave atoms will combine their levels into a
band because electrons are fermions. The band is known as the conduction
band, and the electron as the conduction electron (or valence electron). If no
levels populated by electrons satisfy the criterion 4.129 then the material is
nonconducting.
Now we should consider a lattice of atoms and the potential at the edge
of the solid, that is, at the surface (see Figure 4.44). The last atom facing
the surface will have a potential, which increases toward the vacuum level
unperturbed by the neighbours, and conduction electrons cannot leave the
surface potential barrier. To determine this potential more accurately we ob-
serve the force acting on one electron that is leaving the surface. The attraction
of one charged particle at a distance z from the surface may be described by
using the image charge whereby the whole surface is replaced by the opposite
charge at a distance −z.
Within the conduction band in the metal of temperature T, the electrons
will fill the lowest states according to the Fermi–Dirac distribution,
1
f (ε) = ε−ε , (4.130)
exp kT
F
+1
where ε is the energy level of the conduction band, and ε F is the Fermi level.
At zero temperature (T = 0), the maximum energy level of electrons is equal
to the Fermi energy ε F . Electrons at the top of the conduction band require
energy equal to φ to leave the surface, and this energy is known as the work-
function. In Table 4.17 we show the work-functions and the Fermi energies of
several metals. Alkaline metals have the lowest work-function, typically on
the order of 2 eV.
ε V(z)
distance from surface
W 0 z
φ
W
εF
V(z)
N(ε)
0
state density
a a a
crystal lattice vacuum
surface
FIGURE 4.44
Potential energy and state density of bound electrons in a metal.
1 2
εkinmax = mv = hν − φ. (4.131)
2 max
TABLE 4.17
Work-Function of Typical Metals
Metal εF (eV) φ(eV) Metal εF (eV) φ(eV) Metal εF (eV) φ(eV)
Al
-1
10
Sus
Au
Photo-electric yield (e / h )
1.0
-5
10
Reflectivity
-1
10
-2
10
-7 10
5 15 25
Photon energy (eV)
FIGURE 4.45
Photo-emission yield for some metals as a function of photon energy.
PROBLEM 4.11.1
Calculate the cutoff wavelengths for the materials in Table 4.17. Compare them with
the well-known wavelengths of light sources, such as a mercury lamp, nitrogen lamp,
and rare gas lamps.
e2
V(z) = − − e E z. (4.133)
16π ε0 z
√
The bound potential has a negative minimum at zm = e/16π ε0 E that is
equal to
e eE
Vmax = − (4.134)
2 π ε0
where αe and βe are constants that are best determined by fitting the ex-
periments and both depend on the work-function and Fermi energy of the
metal. Fields of the order of E ∼ 106 Vcm−1 should be required for significant
emission. However, even smaller fields are sufficient to produce the emission.
For example, the microroughness of the surface produces local fields that are
(a) V(z)
z
eEz
εF
Wb
Tt
FIGURE 4.46
Field-induced emission is a result of the changes in potential due to external field: (a) formation
of the potential barrier, (b) wave function of electrons.
very high. In addition, very thin layers of surface oxide may have deposited
charges on them, which produce very high fields and lead to enhanced emis-
sion.
PROBLEM 4.11.2
Prove that the effective reduction of the work-function under the influence of the
external field E is given by Equation 4.134.
Exercise 4.11.1
Explain the principle of operation of the scanning tunneling microscope
(STM) as shown in Figure 4.47.
The STM consists of a very sharp probe made of W or Pt that is brought close to
the surface at a distance of 0.1 to 0.5 nm. We label the quantities related to the
sample substrate (metal) with subscript s and those related to the probe with
the subscript p. A potential is applied between the probe and the sample, and
the current is measured; therefore, STM may be applied only to conductive
materials. The direction of current depends on the relative work-functions of
the probe tip and of the substrate. The principle of the STM is described in
Figure 4.47.
There are two possible modes of operation of an STM. It is possible to
operate in a constant height mode, where the tip is held at a fixed position
Ef
φp – eV
tip metal
φp
z=0
eV
Vb(z)
Tt
z = zs
φs
Ef
(b)
surface (z = zs)
tip (z = 0)
tunneling
V
current
z=0
(a)
FIGURE 4.47
Schematic diagram of the scanning tunneling microscope (STM) and the relevant potentials.
while variations in the current are recorded. Consider what happens as the
probe tip gradually approaches the sample. At a distance of ∼1 nm (1) un-
der the condition, φ p < φs , the conduction electrons tunnel from the tip to
the sample substrate until both Fermi levels become equal, and the electric
dipole with potential Vb = φs − φ p is formed between the two surfaces. When
the external potential V is applied at the probe with respect to the sample, the
tunneling current is detected at the probe. By scanning the substrate surface
in the constant height mode, the quantum state close to the Fermi level is
observed in the two-dimensional image.
(a) (b)
ε=0
φ
Excited wb
state W
εi
εex
Metal
Ground
state
Atom (or Ion) Atom (or Ion)
FIGURE 4.48
Schematics of a potential energy diagram for interaction of a particle with the surface (a) at a
distance and (b) in closest contact.
The other mode in which STMs operate is the fixed current mode, where
the position of the tip is modified to keep the current constant. As current due
to tunneling decays exponentially within the barrier, the measured current
is extremely sensitive to the distance between the tip and the substrate. That
also means that most of the current flows through the atom at the top of
the tip, and thus resolution may be very high (subatomic) provided that the
movements of the tip are controlled to within one hundredth of an atomic
diameter.
These four processes are discussed in detail in this section. We first explain
the Auger processes. As a particle with excitation and ionization energy
εex and εi approaches the surface with the work-function φ, the potential
barrier between the particle and the surface is reduced in height and espe-
cially in the width wb , becoming so small that it assumes atomic dimensions.
e
emission
ε=0 ε=0 ε=0
φ
εi
εex
+ *
A A A
(initial) (transient) (final)
FIGURE 4.49
Electron ejection in the case of (εi − εex ) ≥ φ.
There are two interesting situations for the relative magnitude of potentials as
follows:
The maximum kinetic energy of the released electron e(↑) will occur
if it originates from the top of the conduction band, that is,
εmax = εex − φ,
e
emission
ε=0 ε=0 ε=0
φ
εi
εex
* +
A A A
(initial) (transient) (final)
FIGURE 4.50
Electron ejection in the case of (εi − εex ) < φ.
PROBLEM 4.11.3
When He*(3 S1 ) is approaching the W-surface, estimate the maximum kinetic energy
of electrons ejected from W by Auger potential ejection.
0.32
0.28 +
Ne
0.24
Yield (molecule / ion) He
+
0.20
0.16
0.12 +
Ar
0.08 +
Kr
0.04 +
Xe
0
0 200 400 600 800 1000
Ion energy (eV)
FIGURE 4.51
Secondary electron yields for rare gas ions colliding with a clean tungsten surface.
The efficiency of absorbing the released energy may be limited and there
is always a chance that the electron may dissipate its energy on the way
to the surface. Ejected electrons may be trapped at the surface by adsorbed
molecules or may be reflected from gas molecules back to the metal surface.
Therefore, there is a limited efficiency of secondary electron release per in-
cident ion, which is described by the secondary electron yield. Examples of
secondary electron yields for rare gas ions colliding with a tungsten surface
are given in Figure 4.51.
One should pay attention to the fact that ions may be neutralized while ap-
proaching the surface and also while leaving it. Thus the probability will
depend on both the initial and final velocities (Vi and Vf ), which may be
quite different:
A 1 1
Pneut (V⊥ ) = 1 − exp − + . (4.138)
a Vi⊥ Vf ⊥
If there is a penetration of the solid, the formula becomes more complex and
will include probabilities for CT in binary encounters inside the solid.
(a) (b)
0 0
r r
φ εi φ
εa( εi)
A
A
Metal Metal
FIGURE 4.52
Potentials for surface ion production: (a) εi < φ: production of positive ions, (b) εa > φ: produc-
tion of negative ions.
PROBLEM 4.11.4
Explain the processes of (i) surface ionization and (ii) negative ion formation. The
potential energy curves and the corresponding transitions are shown in Figure 4.52.
Exercise 4.11.2
A hydrogen atom resides at the surface of tungsten (W; φ = 4.5 eV) in an
external field 2 × 108 V cm−1 . Determine the critical distance zc required to
achieve field ionization.
Because the work-function of tungsten is φ = 4.5 eV and the ionization energy
of the ground-state hydrogen atom H(2 S1/2 ;1s) has εi = 13.6 eV, we have
εi−φ>0. Hence the surface ionization without an external field is not possible
(see the dashed line in Figure 4.53). If we add the electric field, the potential
15 Ea = 2 × 108 V cm-1
10
VHE
5
Potential energy (eV)
zc
0 4 8 12
0 z (Å)
φ
-5
εF
VHO
W
-10
1s
-15
surface H
FIGURE 4.53
Potentials for a hydrogen atom on a W surface: (a) without external field (dashed line), (b) with
the external field (solid line).
of the bound electrons will be modified and the condition for ionization will
be given by
At that point the electron from the neutral H may tunnel through the barrier
and a positive ion H+ may emerge from the surface. Therefore the condition
is achieved for zc > (εi − φ)/eE. The probability of tunneling will strongly
depend on zc and thus ionization will occur for a narrow range of distances
as the atom approaches and leaves the surface.
4.11.4 Adsorption
Atoms and molecules may be bound to surfaces in a process known as ad-
sorption. Molecules may be physisorbed (weakly bound) or chemisorbed
(strongly bound). In the former case there is a Van der Waals interaction be-
tween the molecule and the surface, whereas in the latter case the interaction is
chemical, with an electron exchange, and therefore the value of the interaction
potential is much larger. The schematic potential curves for both processes
Potential energy
2O + M
B C A εdiss
O2 + M
0
εphys z
εchem
metal surface
FIGURE 4.54
Potential energy diagrams for adsorption of an oxygen molecule on a metal surface.
rate of adsorption
s= .
rate of collisions with the surface
The rate of adsorption should be proportional to the rate of collisions with
the surface ka , the pressure, and the fraction of vacant sites, ka p(1 − θ), and
the rate of desorption should be proportional to the surface coverage and the
specific rate of desorption kd :
kd θ.
bp
θ= . (4.142)
1 + bp
In the limit of low pressures, we obtain θ∼ p, and at high pressures the correct
limit θ ∼ 1 is obtained. At intermediate pressures one obtains θ ∼ p 1/n , which
is known as Freundlich’s isotherm. We typically have n = 2.
The lifetime of particles at the surface (i.e., the residence time) is another
way to describe the kinetics of desorption:
where τ0 ∼ 10−12 s. Thus, one may estimate the residence times for small and
large values of activation energy.
A number of processes proceed on the surface. Molecules may be dissoci-
ated upon adsorption, or atoms may recombine at the surface. In the case of
hydrogen it was found that recombination of atoms into molecules leads to a
release of excited molecules in high vibrational levels and that the properties
of these molecules are dependent on the surface temperature. The recombi-
nation of two atoms of nitrogen to form a molecule at the surface leads to
formation of a molecule in the metastable A3 state. However, at the sur-
face, the metastable molecule is able to release an electron through the Auger
process. Thus, surface recombination of atoms may provide a source of sec-
ondary electrons in gas discharge. In nitrogen discharges, it was found that
the atoms may remain in afterglow for very long periods of time, on the or-
der of hours. This is shown by the Lewis–Rayleigh afterglow, which is due
to emission of excited states formed in the gas phase. The recombination of
these atoms at the surface releases the electrons and makes it much easier to
start the discharge even hours after the previous discharge.
There are two possible mechanisms for surface reactions. In the Langmuir–
Hinshelwood mechanism, the reaction proceeds through adsorption of two
atoms on the neighboring sites. In other words, two atoms that have been
adsorbed come in proximity at the surface and react. In the Eley–Rideal mech-
anism, a single gas phase atom reacts with the atoms already adsorbed on the
surface. As a result, the two processes have a different pressure dependence.
References
1. French, A.P. and Taylor, E.F. 1978. An Introduction to Quantum Physics. Boston:
MIT Press.
2. Eisberg, R. and Resnick, R. 1974. Quantum Physics of Atoms, Molecules, Solids,
Nuclei and Particles. New York: John Wiley & Sons.
3. Massey, H.S.W., Burhop, E.H.S., and Gilbody, H.B. 1969. Electronic and Ionic
Impact Phenomena, Vol 1–4. Oxford: Oxford Clarendon Press.
4. Massey, H. 1979. Atomic and Molecular Collisions. London: Taylor and Francis.
5. McDaniel, E.W. 1989. Atomic Collisions: Electron and Photon Projectiles. New York:
John Wiley & Sons.
6. O’Malley, T.F., Rosenberg, L., and Spruch, L. 1962. Phys. Rev. 125:1300.
7. Opal, C.B., Peterson, W.K., and Beaty, E.C. 1971. J.Chem. Phys. 8:4100–4106.
8. Phelps, A.V. 1994. J. Appl. Phys. 76:747–753, J. Phys. Chem. Ref. Data 20:557.
9. Phelps, A.V. 1990, 1991. J. Phys. Chem. Ref. Data 19:653; J. Phys. Chem. Ref. Data
20:1339.
10. Golde, M.F., 1976. Gas Kinetics and Energy Transfer. Chem. Soc. London 2:123–
174.
11. Thorne, A.P., 1988. Spectrophysics. London: Chapman and Hall.
12. Thomas, E.W. 1984. In Data Compendium for Plasma-Surface Interactions, Special
Issue, Eds. R.A. Langley, J. Bohdansky, E. Eckstein, P. Mioduszevski, J. Roth, E.
Taglauer, E. W. Thomas, H. Verbeek, and K.L. Wilson, Nuclear Fusion, IAEA,
Vienna.
5.1 Introduction
The Boltzmann equation is the basis for the kinetics of charged particles and
neutral molecules in nonequilibrium and low-temperature plasmas that are
maintained and controlled under a short-range two-body collision with feed
gas molecules [1]. Of course, gas molecules have their own quantum states
through their electronic configuration in a low-temperature plasma. Each of
the atoms and molecules with a different quantum state exhibits unique char-
acteristics in collision with an electron or ion. This provides a very interest-
ing world of low-temperature plasmas with different functions reflected by
the quantum structure even in a fixed external plasma condition. Therefore,
it is essential to understand the terms of the collisions—including various
kinds of two-body collisions—between electron and neutral molecules in the
Boltzmann equation.
The Boltzmann equation is the transport equation in phase space, and
usually we transfer the Boltzmann equation to the kinetic transport equation
in real space in order to be readily able to calculate the plasma structure [2].
Accordingly, we derive the transport theory in detail. The purpose of this
chapter is to provide a foundation, that is, the kinetic transport theory, and par-
ticularly electron transport theory, for the modeling and design of nonequi-
librium and low-temperature plasmas.
143
where α is the acceleration due to the force, and m and e(> 0) are the mass
and charge of the charged particle. Here, the number density of the swarm of
the charged particle n is a function of position r and time t:
In addition, each of the particles has a different velocity v and the velocity
changes over time. The transport of the charged particle is therefore desig-
nated in phase space, that is, both in the configuration (real) space and in
the velocity space. The number of charged particles within a small volume
defined by dr around a position r and within dv around a velocity v is desig-
nated dn, which can be defined as
v-space
g(v, r, t)
r-space
Vx
n(r, t)
x E(r, t) Vz
dr
E(r, t) z
FIGURE 5.1
Example of the density distribution of charged particles n(r, t) in real space and the velocity
distribution g(r, v, t) in phase space.
Here, the small element of phase space dv dr after dt is connected to the initial
element dv dr by a Jacobian
∂(r , v )
dv dr = dv dr, (5.5)
∂(r, v)
and from Equations 5.3 and 5.4 we obtain
∂ x ∂ x ∂ x ∂ x
∂x ∂y . .
∂v y ∂vz
∂y
∂x . . . . .
∂(r , v ) . . . .
= = 1. (5.6)
∂(r, v) . . . .
∂v y . . .
∂x
∂v
z . . ∂vz
∂x ∂vz
PROBLEM 5.2.1
Confirm Equation 5.6 (or dv dr = dv dr ) by using the condition ∇v · F = 0
for external forces in Equation 5.1.
where vr = 0. The ensemble average of the velocity, often labeled as fluid
velocity, is given by
1
v(r, t) = vg(v, t)dv, (5.13)
n
and the mean (kinetic) energy is
1 1 2 1
ε(r, t) = mv g(v, r, t)dv = m vd2 + vr2 . (5.14)
n 2 2
In the case where the number density has a spatial nonuniformity, n(r, t),
the velocity distribution function g(v, r, t) may be represented in powers of
the density gradient ∇r n as
Here the coefficients in the density expansion, gk , are tensors of rank k and
indicates a kfold scalar product. The functions independent of position are
normalized as
1; k = 0
gk (v, t)dv = (5.16)
0; k =
0 .
PROBLEM 5.3.1
In the case where there is a magnetic field as well as an electric field, the drift tensor
has six components. Derive each component of the drift velocity.
The first term on the right-hand side (pressure tensor) shows the contribu-
tion of the thermal or random velocity, and the second term is due to drift
(directional) velocity. As random velocity is usually much larger than the
directional velocity, the total pressure tensor PT is almost isotropic with
equal components in the x-, y-, and z-directions, and each component (scalar
pressure p) is written according to
Px 0 0
1
P = 0 Py 0 , p = px = p y = pz = mnvr2 . (5.24)
3
0 0 Pz
The energy flux vector is defined as the energy carried by charged particles
that cross unit area per unit time
1 2
Q(r, t) = mv vg(v, r, t)dv. (5.25)
2
Q represents the third-order velocity moment and
PROBLEM 5.3.2
The diffusion coefficients, DL and DT , are obtained by the trace of the trajectory of
particles in the configuration space in accordance with
1d
Then, in an initial stage of the transport, the fluid velocity expressed by the
drift velocity and diffusion tensor in Equation 5.17 will be extended to a form
including the 3rd- and 4th-order transport coefficients.
∂g(v, r, t) ∂g(v, r, t)
A(v, r, t) dv + A(v, r, t)v · dv
∂t ∂r
∂g(v, r, t)
+ A(v, r, t)α · dv = A(v, r, t)J (g, F )dv. (5.30)
∂v
∂g ∂ ∂A
A dv = Agdv − gdv
∂t ∂t ∂t
∂ ∂A
= (n(r, t)A ) − n(r, t) , (5.31)
∂t ∂t
∂g ∂ ∂
Av · dv = · (vAg)dv − · vA gdv,
∂r ∂r ∂r
∂ ∂
= · (n(r, t)vA ) − n(r, t) · vA . (5.32)
∂r ∂r
∂g ∂ ∂
Aα · dv = · αAgdv − · αA gdv
∂v ∂v ∂v
∂ ∂
= · (n(r, t)αA ) − n(r, t) · αA
∂v ∂v
∂
= −n(r, t) α · A . (5.33)
∂v
Now we substitute Equations 5.31 to 5.33 into Equation 5.30 and obtain
the general continuity equation as
∂ ∂A(v, r, t) ∂
(n(r, t)A(v, r, t) ) − n(r, t) + · (n(r, t)vA(v, r, t) )
∂t ∂t ∂r
∂ ∂A(v, r, t)
− n(r, t) · vA(v, r, t) − n(r, t) α ·
∂r ∂v
= A(v, r, t)J (g, F )dv. (5.34)
PROBLEM 5.4.1
Show that for the number density continuity of electrons, the term R 0 on the right-
hand side of Equation 5.35 is expressed through the ionization rate Ri and electron
attachment rate Ra as
∂ ∂ ∂
· (mn(r, t)vdvd ) + · (mn(r, t)vr vr ) = vd · (mn(r, t)vd )
∂r ∂r
∂r
∂ ∂
+ mn(r, t) vd · vd + · (mn(r, t)vr vr ),
∂r ∂r
Here the third term on the left-hand side is expressed by using the pressure
tensor P (which is defined in Equation 5.24) as ∇r ·P.
PROBLEM 5.4.2
Show that the last term in the momentum conservation Equation 5.37 is approximated
by using the rate of the momentum transfer collision Rm as
Exercise 5.4.1
Consider the case that the directional velocity of charged particles is inde-
pendent of space and time under a constant E without B. Then, derive the
current envd .
From Equation 5.37 we obtain the relation under Rm R 0
∂
mnvd Rm = enE − m · (nvr vr ).
∂r
∂
J = envd = enµE − e (Dn),
∂r
where µ and D are equivalent as µ = e/mRm and D = vr vr z /Rm = vr2 /3Rm ,
where the electric field is directed along the z-axis. Here µ is the mobility
and D is the diffusion coefficient and these equations are consistent with
simple expressions 2.4 and 2.5 in Chapter 2.
The first term is the change in the mean energy ε(r, t), and the second is the
change in the energy flow. When we substitute Equation 5.26 in the second
term we obtain
∂ ∂ ∂
(n(r, t)ε(r, t) ) + · (n(r, t)ε(r, t) vd ) + · (P · vd )
∂t ∂r ∂r
∂ mv 2
+ · q − n(r, t)eE · vd = J dv. (5.40)
∂r 2
PROBLEM 5.4.3
Derive the collisional integral term in the energy conservation Equation 5.40 for
electrons that is given by
mv 2 2m
J dv = ε(r, t) Rm ne (r, t)
2 M
1/2
2ε
+ ε j R j + εi Ri − εNQa (ε) ne (r, t), (5.41)
m
properties by the differential cross section σ (θ, φ; ε) and the integral cross
sections Q(ε). The velocity distribution of neutral gas molecules is represented
by F (V , r,t), where gas molecules have a mass M and velocity V before col-
lision. The charged particles of mass m and velocity v before collision are de-
scribed by the corresponding velocity distribution function g(v , r, t). When
we consider a small element of the phase space dv dr, then the number of the
charged particles that enter this element during a short time dt is equal to
J in = F (V , r, t)dV g(v , r, t)dv vγ σ (θ , φ ; vγ )d dr dt. (5.42)
v
During the same time dt, the number of charged particles leaving the same
element dv dr is
J out = F (V, r, t)dV g(v, r, t)dvvγ σ (θ, φ; vγ )ddr dt, (5.43)
v
where vγ and vγ are the relative speed between the gas molecule and the
charged particle just before and after the collision. The resulting change in
the number of charged particles in the element dv dr in time dt is
J dv dr dt = {J in − J out }. (5.44)
where Ymne
(θ, ϕ) = Pnm (cos θ ) cos mϕ are the spherical harmonics by Morse–
Feshbach notation [4]; this expansion is discussed further in Section 5.6.1. The
spherical harmonic expansion helps us take care of the dependence on three
components of the vector v by taking advantage of the cylindrical symme-
try of the problem and turning the solution of the velocity dependence to a
v
m
m M
v' V'
M
V
(before collision) (after collision)
ω
V' v'
vg
V
M v M v
m+M r m+M r
FIGURE 5.2
The definition of a collision in (a) a laboratory frame of reference (real space) and (b) a center-of-
mass frame of reference (velocity space).
i. The relative velocity between the electron and the gas molecule is
not changed after the collision;
ii. The velocity of gas molecules is much less than the velocity of elec-
trons, |V| |v|, and we represent the velocity distribution of gas
molecules with the number density N by F (V) = Nδ(V), where δ is
Dirac’s delta function; and
iii. From the momentum and energy conservation equations before and
after collisions, v and v, we have the relation
v 2 − v 2 2m
= (1 − cos ω), (5.46)
v 2 M+m
where ω is the scattering angle as defined in Figure 5.2b.
From Equation 5.46 we obtain dv /dv = (v /v)3 , and thus we replace the small
velocity element dv by dv. As a result, we obtain the elastic term of electrons
as follows:
J elas dv = Nδ(V )g(v )vγ σ (vγ , ω)d dv dV
V
− Nδ(V)g(v)vγ σ (vγ , ω)ddv dV
V
v 3
=N g(v )v σ (v , ω)d − g(v)vσ (v, ω)d dv. (5.47)
v
3
and in order to divide the differential cross section into spherical (σ0 ) and
forward-directed (σ1 ) parts, and the like, we expand the differential cross
section σ (θ, ϕ; ε) in terms of the Legendre function Pn (cos θ) according to
σ (v, ω) = σn (v)Pn (cos ω). (5.49)
n
m=1 n
2n + 1 mn
Here we consider
the orthogonality property of the Legendre function, 4π/
(2n + 1) = Pn (cos ω)Pn (cos ω)d (see Equation 5.74). Then the elastic colli-
sion term is given as
J elas
1 4
=N e
Ymn (θ, ϕ) v gmn (v )σ (v , ω)Pn (cos ω) − v 4 gmn (v)σ (v, ω) d
mn v 3
1 4
=N e
Ymn (θ, ϕ) v gmn (v )σ (v , ω) − v 4 gmn (v)σ (v, ω) Pn (cos ω)d
mn v 3
−N e
Ymn (θ, ϕ)gmn (v)vσ (v, ω) {1 − Pn (cos ω)} d. (5.50)
mn
When we compare the magnitudes of velocities before and after the elastic
collision by using Equation 5.46, we conclude that both of the velocities have
nearly the same magnitude, and v 2 = v 2 − v 2 . Therefore we are able to
expand the first term in the right-hand side of Equation 5.50 by the Taylor
series with the variable v 2 = v 2 − v 2 as
1 2m ∂[v 4 gmn (v)σ (v, ω)]
∼
J elas = N e
Ymn (θ, ϕ) (1 − cos ω) Pn (cos ω)d
mn
v M+m ∂(v 2 )
−N e
Ymn (θ, ϕ)gmn (v)vσ (v, ω) {1 − Pn (cos ω)} d. (5.51)
mn
For thermal equilibrium, electrons have the Maxwellian distribution. For zero
field, the temperature of electrons Te are equal to Tg when the gas temperature
is not zero; that is, Tg =
0. However, under all circumstances collisions with
thermal gas molecules lead to an additional velocity of electrons that is on
the order of O(V). Thus the total velocity has two components, one for zero
temperature and an additional component due to thermal motion (Tg =
0):
v = v0 + O(V).
For thermal equilibrium the distribution function has to be a Maxwellian,
00
and therefore J elas = 0. As a result, we obtain the order of V 2 according to the
principle of equipartitioning of energy:
2
2 ∂g00 v02 2kTg
O(V ) = −g00 v0 = . (5.53)
∂(v )
2 m
Finally, we obtain the elastic collision integral for a general case of nonzero
gas temperature (Tg =
0):
m 1 ∂ kTg ∂
00
J elas =N v 4
Qm (v) g00 (v) + g00 (v) . (5.54)
M + m v 2 ∂v mv ∂v
where the integrated cross section is defined on the basis of the corresponding
differential cross section as
Q j (v) = σ j (v, ω)d.
1 1
=N Ymn (θ, ϕ) gmn (v )v 2 σi (v , ω)Pn (cos ω)d
mn
v (1 − )
1 2
+ gmn (v )v σi (v , ω)Pn (cos ω)d − Ngmn v Qi (v) , 2
(5.60)
where the integrated ionization cross section is defined as
Qi (v) = σi (v, ω)d.
where tensor gk (v, t) has components that are velocity distributions of differ-
e
ent order in density gradient expansion, and Ymn (θ, ϕ) are spherical harmonics
in Morse–Feshbach representation [4].
The dependence on two angles (θ, φ) may be separated by using Ynm (θ, ϕ) =
(θ )m (ϕ), where
1
m (ϕ) = √ e imϕ (5.66)
2π
e
Y00 (θ, φ) P0 = 1
e
Y01(θ, φ) P1 = cosθ
0
Y11(θ, φ) sinθ sinφ
e
Y11(θ, φ) sinθ cosφ
FIGURE 5.3
Four spherical harmonics of the lowest order in Morse–Feshbach notation.
and
1/2
2n + 1 (n − m)!
(θ ) = (−1)m Pnm (cos θ), −n ≤ m ≤ n. (5.67)
2 (n + m)!
Here, Pnm (cos θ ) is the associated Legendre polynomial. The spherical har-
monics is written in Morse–Feshbach representation as
e
Ymn (θ, ϕ) = Pnm (cos θ) cos mϕ (5.68)
and
0
Ymn (θ, ϕ) = Pnm (cos θ) sin mϕ. (5.69)
e 0
A few terms of Ymn and Ymn are shown in Figure 5.3 and thus the veloc-
ity distribution function is expanded as g(v, t) = gmn (v, t)Ymn
e
(θ, ϕ) (see
Equation 5.63). Normalization of the distribution function requires the fol-
lowing integrals of spherical harmonics:
2π π 2
e
Ymn (θ, ϕ) or 0
Ymn (θ, ϕ) d
0
0
4π (n + m)!
, n = 1, 2, 3, . . .
= 2(2n + 1) (n − m)! (5.70)
4π , n = 0.
and
∂
(2n + 1) cos2 θ Ye (θ, ϕ)
∂ cos θ mn
= (n + 1)(n + m)Ym(n−1)
e
(θ, ϕ) − (n − m + 1)Ym(n+1)
e
(θ, ϕ), (5.72)
and these are also useful. Because the spherical harmonics in Morse–Feshbach
notation is connected to the associated Legendre polynomials by Ymn e
=
Pn cos mϕ and Y0n = Pn , Pn and Pn also satisfy the relations 5.71 and 5.72.
m e m
TABLE 5.1
Legendre Polynomials
Pn
P0 (cos θ ) 1
P1 (cos θ ) cos θ
P2 (cos θ ) (3 cos2 θ − 1)/2
P3 (cos θ ) (5 cos3 θ − 3 cos θ )/2
P4 (cos θ ) (35 cos4 θ − 30 cos2 θ + 3)/8
TABLE 5.2
Associated Legendre Polynomials
Pnm
If we have a function of both azimuthal and polar angles f (θ, ϕ), then the
expansion is in terms of Ymne
(θ, ϕ):
f (θ, ϕ) = e
a mn Ymn (θ, ϕ), (5.78)
where
π
2(2n + 1) (n − m)! 2π
a mn = f (θ, ϕ)Ymn
e
(θ, ϕ)d. (5.79)
4π (n + m)! 0 0
∂ 1 ∂ 1
g (v, t) + α(t) · g (v, t) + R 0 (t)g1 (v, t)
∂t ∂v
= J(g 1 , F ) + vg 0 (v, t) − vd (t)g 0 (v, t), (5.81)
and
∂ 2 ∂ 2
g (v, t) + α(t) · g (v, t) + R 0 (t)g2 (v, t)
∂t ∂v
= J(g 2 , F ) + vg1 (v, t) − vd (t)g1 (v, t) + D(t)g 0 (v, t), (5.82)
Then, the second term in the right-hand side of Equation 5.80 is rearranged
by the aid of the addition theorem of the Legendre function as
∂ 0
α(t) · g (v, t)
∂v
∂g 0 sin2 θ ∂g 0
= αz (t) cos θ +
∂v v ∂ cos θ
∂gn0 (v) sin2 θ 0 ∂ Pn (θ )
= αz (t) cos θ Pn (θ ) + αz (t) gn (v) .
n
∂v n
v ∂ cos θ
In the above equation, cos θ Pn and sin2 θ∂ Pn /∂ cos θ are replaced by using
Equations 5.71 and 5.72, and the equation is simplified to
∂ 0
α(t) · g (v, t)
∂v
n+ 1 n
0
∂gn (v)
= αz (t) Pn+1 (θ ) + Pn−1 (θ )
n
2n + 1 2n + 1 ∂v
g 0 (v)
n(n + 1) n(n + 1)
+ αz (t) n
− Pn+1 (θ ) + Pn−1 (θ ) .
n
v 2n + 1 2n + 1
The above equation is further arranged after binding each term associated
with Pn (θ ) as
∂ 0
α(t) · g (v, t)
∂v
n ∂gn−1
0
(v) n + 1 ∂gn+1
0
(v)
= αz (t) +
n
2n − 1 ∂v 2n + 3 ∂v
(n − 1)n gn−1
0
(v) (n + 1)(n + 2) gn+1
0
(v)
− + Pn (θ )
2n − 1 v 2n + 3 v
n
∂ n−1
= αz (t) − 0
gn−1 (v, t)
n
2n − 1 ∂v v
n+1 ∂ n+2
+ αz (t) + 0
gn+1 (v) Pn (θ ). (5.84)
2n + 3 ∂v v
g1 (v, t) = gx1 (v, t)i + g 1y (v, t)j + gz1 (v, t)k. (5.86)
When the external force is along the z-axis, we have α(t) = −αz (t)k. From
Equation 5.81 we obtain
∂ 1 ∂ 1 J(gx1 , F )
g (v, t) αz (t) g (v, t) gx1 vx g 0
∂t x ∂vz x
∂ 1 ∂ 1
g y (v, t) + αz (t) g (v, t) + R 0 g 1y + vy g0 = J(g 1y , F ) . (5.87)
∂t ∂vz y
∂ 1 ∂ 1
g (v, t) αz (t) g (v, t) gz1 (vz − vd )g 0 J(gz1 , F )
∂t z ∂vz z
Here, the component along the external field (i.e., the z-axis), gz1 (v, t), has a
directional component of velocity vd , and the other two components vx and v y
g0 (υ )
3.3
Vr
(1
3.3
0
8
0.0 0.0
-3.3
cm
-1
Vz (1 08
cm s )
s
-1
)
g00(ε)
Distribution (arb.)
0.5
0
0 5 10 15 20 25
Electron energy (eV)
FIGURE 5.4
Lowest-order solution for the velocity distribution function for electrons at 100 Td in Ar: (a) g 0 (v)
and (b) gn0 (v) (n = 0, 1, 2, . . .).
are symmetric to each other. The longitudinal component gz1 (v, t) is symmetric
with respect to the external field direction (i.e., the z-axis) and is expanded by
spherical harmonics for m = 0:
gz1 (v, t) = gz10n (v, t)Y0n
e
(θ, ϕ) = gz1n (v, t)Pn (θ ). (5.88)
n n
(vz − vd )g 0 (v, t)
= (v cos θ − vd )gn0 Pn (θ )
n
n n+1
= v Pn−1 (θ )gn + v
0
Pn+1 (θ )gn − vd gn Pn (θ )
0 0
n
2n + 1 2n + 1
n+ 1 n
= vgn+1
0
(v, t) + vgn−1
0
(v, t) − vd gn0 (v, t) Pn (θ ), (5.89)
n
2n + 3 2n − 1
When we use Equations 5.71 and 5.72 to replace cos θ Pn1 (θ ) and sin2 θ∂ Pn1 (θ )/
∂ cos θ, we obtain
n n+1 1
∂gx1n (v, t)
= αz (t) cos ϕ P 1 (θ ) + P (θ )
n=1
2n + 1 n+1 2n + 1 n−1 ∂v
n2 (n + 1)2 1
1
gxn (v, t)
+ αz (t) cos ϕ − 1
Pn+1 (θ ) + Pn−1 (θ ) ,
n=1
2n + 1 2n + 1 v
gz1 (υ )
3.3
Vr
(1
3.3
0
0.0
8
0.0
-3.3
cm
-1
8 )
Vz (10 cm s
s
-1
)
g1z 2(ε)
0
g1z 3(ε)
Distribution (arb.)
g1z 1(ε)
-0.5
g1z 0(ε)
-1.0
0 5 10 15 20 25
Electron energy (eV)
FIGURE 5.5
The longitudinal component (z) of the first-order solution for the velocity distribution function
for electrons at 100 Td in Ar: (a) gz1 (v) and (b) gzn
1 (ε) (n = 0, 1, 2, . . .).
and after we order the right-hand side of the above equation in terms of Pn1 (θ ),
we have
n − 1 ∂gx1n−1 (v, t) n + 2 ∂gx1n+1 (v, t)
= αz (t) +
n=1
2n − 1 ∂v 2n + 3 ∂v
(n − 1)2 gx1n−1 (v, t) (n + 2)2 gx1n+1 (v, t)
− + Pn1 (θ ) cos ϕ
2n − 1 v 2n + 3 v
n−1
∂ n−1
= αz (t) − gx1n−1 (v, t)
n=1
2n − 1 ∂v v
n+2 ∂ n+2
+ αz (t) + gx1n+1 (v, t) Pn1 (θ ) cos ϕ. (5.92)
2n + 3 ∂v v
gx1 (υ )
3.3
Vr
3.3
(1
0.0 0.0
0
8
-3.3
cm
-1
8 )
Vz (10 cm s
s
-1
)
g1x 3(ε)
0
g1x 4(ε)
g1x 2(ε)
0 5 10 15 20 25
Electron energy (eV)
FIGURE 5.6
The transverse component (x) of the first-order solution for the velocity distribution function for
electrons at 100 Td in Ar: (a) gx1 (v) and (b) gx1n (v) (n = 1, 2, . . .).
reduce to five. Even at a very poor resolution of 100 points per dimension, the
number of mesh points is formidable, though it may be within the range of
modern computers. Nevertheless, expansions make the problem more com-
plex but numerically much easier to handle.
TABLE 5.3
Each of the Collision Terms Appearing in J(g, F )
Collision Type Collision Integral Expanded Collision Term
m 1 ∂ kTg ∂g00
Elastic J m g00 NQm (v)v 4 g00 +
M v 2 ∂v mv ∂v
m 2m
! 1 ∂
" 3
!
kTg
" #
= ε − kTg NQm (ε) + − NQm f 00
4π M 2 ∂ε 2 4ε
m 2m ∂
# ∂
+ (ε + kTg )NQm (ε) + kTg ε NQm (ε) f0
4π M ∂ε ∂ε 0
m 2m ∂ 0
+ kTg εNQm (ε) f
4π M ∂ε 0
m 1 ∂
J m g10 −NQm (v)vg10 + (NQv (v) − NQm (v)) v 4 g10
M v 2 ∂v
m
= NQm (ε) f 10
4π
m 2m 3
+ (NQv (ε) − NQm (ε))
4π M 2
∂
#
+ε [NQv (ε) − NQm (ε)] f 10
∂ε
2m ∂ 0
+ ε (NQv (ε) − NQm (ε)) f1
M ! 3" ∂ε
3 m
J m g20 − NQv (v)vg20 = − NQv (ε)
2 4π 2
m
J m gx011 −NQm (v)vgx111 = − NQm (ε) f x111
4π ! "
3 m 3
J m gx012 − NQv (v)vgx112 = − NQv (ε) f x112
2 4π 2
1 2
Excitation J j g00 v NQ j (v )g00 (v ) − v 2 NQ j (v)g00 (v)
v
ε+ε j
m 1 ∂ √
= √ εNQ j (ε) f 00 (ε)dε
4π ε ∂ε ε
m
J j g10 org20 −NQ j (v)vg10 org20 = − NQ j (ε) f 10 or f 20
4π
m
J j gx011 orgx012 −NQ j (v)vgx111 orgx112 = − NQ j (ε) f x111 or f x112
$ 4π %
1 1 + k 2
Ionization J i g00 v NQi (v )g00 +(1 + k)v 2 NQi (v )g00 −v 2 NQi (v)g00
v k
(1+k)ε+εi
m 1 ∂ √
= √ εNQi (ε) f 00 (ε)dε
4π ε ∂ε
1+k
ε
ε+εi
∂ k √
+ εNQi (ε) f 00 (ε)dε
∂ε 0
m
J i g10 org2 0
−NQi (v)vg10 org20 = − NQi (ε) f 10 or f 20
4π m
J i gx011 orgx012 −NQi (v)vgx111 orgx112 = − NQi (ε) f x111 or f x112
4π
m
Attachment J a g00 −NQa (v)vg00 = − NQa (ε) f 00
4π
m
J a g10 org20 −NQa (v)vg10 org20 = − NQa (ε) f 10 or f 20
4π
m
J a gx011 orgx012 −NQa (v)vgx111 orgx112 = − NQa (ε) f x111 or f x112
4π
velocity space:
g(v, r, t) = gk (v, t) ⊗ (∇r )k n(r, t)
k
= k
gmn (v, t)Ymn
e
(θ, ϕ) ⊗ (∇r )k n(r, t)
k mn
= gn0 (v, t)Pn (θ ) n(r, t)
n=0
+ gx11n (v, t)Y1n
e
(θ, ϕ)i + g 1y1n (v, t)Y1n
e
(θ, ϕ)j
n=1 n=1
∂
+ gz1n (v, t)Pn (θ )k · n(r, t) + O ∇r2 n
n=1
∂r
3 cos2 θ − 1
= g00 (v,
t) + g10 (v,
t) cos θ + t) g20 (v, + · · · n(r, t)
2
∂
+ gx11 (v, t) sin θ cos ϕ + gx12 (v, t)3 cos θ sin θ cos ϕ + · · · i · n(r, t)
∂r
∂
+ g 1y1 (v, t) sin θ cos ϕ + g 1y2 (v, t)3 cos θ sin θ cos ϕ + · · · j · n(r, t)
∂r
3 cos2 θ − 1 ∂
+ gz10 (v, t) + gz11 (v, t) cos θ + gz12 (v, t) + · · · k · n(r, t)
2 ∂r
+ O ∇r2 n . (5.95)
Exercise 5.6.1
Discuss and develop equations for two-term approximation in the case of
electron transport in gases. Discuss the order of errors and applicability.
Two-term approximation consists of taking the first two terms in the spherical
harmonic expansion in the lowest order of gradient expansion (k = 0) in a
steady state. From Equation 5.95 we directly obtain
g(v) = g00 (v) + g10 (v) cos θ + O g20 ne ,
where the order of error in spherical expansion is given by O(g20 ), and higher-
order terms in the gradient expansion are neglected. In order for this solution
to be accurate (i.e., in order that the spherical harmonic expansion converges),
we need to satisfy the condition g00 (v, t) g10 (v, t). The corresponding equa-
tions of the two-term expansion are given by
m 1 ∂ 1 eE ∂ 2
− NQm (v)v 4 g00 + R 0 g00 (v) + + g10 (v) = 0, (5.96)
M v ∂v
2 3 m ∂v v
1 eE ∂g00 (v)
g10 (v) = − & . (5.97)
NQk (v)v m ∂v
PROBLEM 5.6.1
Derive the following expression of the diffusion tensor of electrons in gases without
inelastic collisions in a direct-current field:
v(v − vd )g 0 (v) v emE · ddv g1 (v)
D= dv + dv, (5.103)
NQm (v)v(1 + (v)) NQm (v)v(1 + (v))
where (v) is a function of M and Tg of gas molecules,
kTg d
(v) = ln[NQm (v)v].
Mv dv
References
1. Chapman, S. and Cowling, T.G. 1970. The Mathematical Theory of Non-Uniform
Gases, 3rd ed., Cambridge: Cambridge University Press.
2. Shkarofsky, I.P., Johnston, T.W., and Bachynski, M.P. 1966. The Particle Kinetics of
Plasmas. Reading, MA: Addison-Wesley.
3. Kumar, K., Skullerud, H.R., and Robson, R.E. 1980. Kinetic theory of charged
particles swarms in neutral gases. Aust. J. Phys. 33:343–448.
4. Morse, P.M. and Feshbach, H. 1953. Methods of Theoretical Physics. New York:
McGraw-Hill.
6.1 Introduction
Swarm parameters are generally applied to plasma modeling and simula-
tions. At the same time, the nonequilibrium regime in discharges is well repre-
sented under a broad range of conditions by using the Boltzmann equation, as
discussed in the previous chapter. In this chapter we discuss the basic features
of transport coefficients (swarm parameters) of electrons and also, to a lesser
degree, of ions. Together with transport data we compile cross sections of the
electron for interesting gases. First we discuss the basic characteristics of the
E/N dependence of the coefficient for swarms in the hydrodynamic regime.
We consider swarms in direct current (DC) fields and also in time-varying ra-
dio frequency (rf) fields. Collective phenomena that cannot be easily described
on the basis of single electron dynamics are labeled kinetic phenomena, and
some special examples are described. Differences between the transport in
conservative and nonconservative situations (i.e., when there are number-
changing collisions such as ionization and attachment) are briefly considered.
175
103
electronic
excitation
III IV
polyatomic gas
2
vde (arb.) 10
II atomic gas
10 I ∝ (E / N)1/2
∝E/N
1
1 10 102 103
DC- E / N (arb.)
FIGURE 6.1
Typical characteristics of electron drift velocity as a function of DC-E/N.
E/N. The pulsed Townsend experiments are made between two parallel
electrodes with a swarm of electrons produced by a photo-emission. Their
accuracy is somewhat smaller, but it may still be possible to achieve an un-
certainty of less than ±3%. The measurements may be extended to higher
E/N. There is a class of the drift velocity experiment measuring the optical
emission from short-lived excited species in the electron swarm that can be
performed only at higher E/N.
In Figure 6.1 we show the typical characteristics of the drift velocity of
electrons in gases as a function of DC-E/N. At sufficiently low E/N, where
an electron loses all the kinetic energy equal to the gain from the field E at
one elastic collision, the drift velocity is proportional to E/N (region I). In
region I the mean energy of electrons is close to the thermal energy 3kTg /2
and the velocity distribution is approximated by the Maxwellian. A finite
kinetic energy of electrons is steadily maintained when the energy gain from
the field is balanced with the energy loss by elastic collisions during a unit
time. Then, the region II appears in Figure 6.1. Even at low E/N, polyatomic
molecules have a probability of vibrational (and rotational) excitations with
the electron as well as the elastic scattering. This means that there is also a
region III that mixes II with I as shown in Figure 6.1. The magnitude of the
drift velocity in the mixed region strongly depends on the magnitude of the
vibrational cross section. A rapid increase in the drift velocity at high E/N
(region IV) is caused by electronic excitations (including ionization) with a
high energy loss from several eV to 20 eV [4].
The E/N dependence of the drift velocity of electrons in He, Ar, H2 ,
HCl, CH4 , CF4 , and SiH4 is shown in Figure 6.2. As qualitatively shown in
Figure 6.1, CF4 , SiH4 and CH4 have more distinct characteristics than He and
Ar in the region of moderate E/N. That is, the drift velocity has a peak in re-
gion III in CF4 , SiH4 , and CH4 . The first significant increase is often known as
8 Ar
10 H2
He
CH4
Drift velocity (cm s )
-1
CF4 HCl
10
7 CH4
SiH4
SiH4
H2
6
10
Ar
HCl
5
10 2 3
1 10 10 10
DC- E / N (Td)
FIGURE 6.2
Electron drift velocities in various gases as a function of DC-E/N.
enhanced mobility, and the decrease of the drift velocity is known as negative
differential conductivity (NDC). The name of the conductivity originated
from the fact that in a weakly ionized plasma the conductivity σ is defined as
je = ene vd = σ E.
Observe, however, that the conductivity is proportional to both the drift ve-
locity vd and the number density ne , and thus the NDC will also occur if the
electron number density changes with E/N, which will be possible in the
case of the nonconservative collisions, that is, under the electron attachment
or ionization. The most interesting phenomena associated with the NDC arise
from the negative slope of vd with respect to E/N, and we use the term NDC
just to describe the negative slope of the drift velocity. Typical gases that have
NDC are CF4 , SiH4 , CH4 , and mixtures of Ar with molecular gases such as N2 .
Exercise 6.2.1
Derive the conditions for NDC based on the momentum and energy balance
of electrons. If we start from the energy balance between the energy gain from
the field E and dissipation in inelastic collisions with a threshold energy ε j
and rate R j in a steady state with uniform ne (here, we neglect the energy loss
in elastic collisions)
d
< εm >= e Evd (E/N) − ε j R j (E/N) = 0.
dt
d
(mvd (E/N)) = e E − mvd R j (E/N) = 0.
dt
After eliminating E from both of the balanced equations, we obtain the
following condition for NDC:
∂ eε j R j 1 ∂ Rj 1 ∂ Rm
vd (E/N) = − .
∂ E/N 2mvd Rm R j ∂ E/N Rm ∂ E/N
Here, we observe directly that if we want the slope of the drift velocity to
be negative, then the collision rate for inelastic scattering R j should decrease
as a function of E/N and the collision rate for the elastic momentum trans-
fer should increase. Although the above relation is derived for very simple
assumptions, the conclusions are generally applied for the NDC.
(a)
g0 (υ )
8.7
8.7
0.0
0.0 -8.7
Vr
(10 7
cm -1
7 cm s )
s -1 Vz (10
)
(b)
g0 (υ )
2.2
2.2
0.0
0.0 -2.2
Vr
(10 8
cm -1
8 cm s )
s -1 Vz (10
)
(c)
g0 (υ )
4.0
4.0
0.0
0.0 -4.0
Vr
(10 8
cm -1
8 cm s )
s -1
) Vz (10
FIGURE 6.3
Velocity distribution g 0 (v) as a function of DC-E/N in pure SiH4 : (a) 20 Td, (b) 100 Td, and
(c) 500 Td.
NDLL
23 NDTT
10 Ar
CF4
22
10 SiH4
SiH4
21
10 2 3
1 10 10 10
DC- E / N (Td)
FIGURE 6.4
Diffusion coefficients of electrons in various gases as a function of dc-E/N. Transverse diffusion
coefficient (· · ·), longitudinal diffusion coefficient (—).
than DT at very low E/N. In Figure 6.4 we show data for the components of
the diffusion tensor in Ar, CF4 , and SiH4 .
The variations of the DT /DL ratio attain a magnitude as large as 10 in gases
such as Ar, CH4 , CF4 , or SiH4 . In particular, the most striking feature is the
large difference between the two components of the diffusion tensor between
10 Td and 100 Td, which is almost one order of magnitude. This feature is a
direct result of the very strong Ramsauer–Townsend minimum in the elastic
scattering cross section.
PROBLEM 6.2.1
Discuss the numerical algorithm to calculate g 0 (v) as a boundary condition problem.
2
10
H2
Ar
CH 4
He
1
N2
CF4
10
-1 HCl
SiH4
2 3
1 10 10 10
DC- E / N (Td)
FIGURE 6.5
Electron mean energy in gases as a function of DC-E/N.
(a)
10
10
Ri
8
10
Rex
Rexm
Rex(2p1)
6
10
Rate (s )
-1
Rex(3p5)
4
10
2
10
0
10 2 3
1 10 10 10
DC- E / N (Td)
(b)
10
10
Rv3
8
10 Rv1
Rv4
6
10
Rate (s )
-1
4
10
Ri1
Rex
2 Rd1
10
Ra Rd2
Rd3
0
10 2 3
1 10 10 10
DC- E / N (Td)
FIGURE 6.6
Excitation rates for electrons in Ar (a) and CF4 (b) as a function of DC-E/N.
10
10
ionization
attachment
8
10 SF6-
SF5-
SF6
SF4-
F-
6
10
Rate (s )
-1
CF4 SiH4
CF4
4
10 Ar SiH4
H2 N2
2
10
SF6
0
10 2 3
1 10 10 10
DC- E / N (Td)
FIGURE 6.7
Ionization and attachment rates of electrons in Ar, H2 , N2 , CF4 , SF6 , SiH4 , and CH4 as a function
of DC-E/N.
In Figure 6.6 we show some of the excitation rates with ionization and electron
attachment rates in Ar and CF4 . One can see that for a range of E/N there is a
negligible rate, and then at some point the rate starts to increase very rapidly.
The point where this occurs is related to the magnitude of the energy loss
for that process, and the rapid increase marks the overlap of the high-energy
tail of the energy distribution with the cross section for that process. At some
point the increase slows down until a maximum is reached, and eventually
the rate starts to decay. The rate (constant) is a key to calculate the kinetics of
various plasma chemical processes and densities of excited states and is thus
used directly in plasma modeling.
The ionization rates are significant for relatively high mean energies (i.e.,
high E/N) and are essential for establishing the accuracy of the energy distri-
bution of electrons and checking the cross section data at higher energies (see
Figure 6.7). Dissociative electron attachment also has a threshold energy at
moderate energy. The electron attachment to a parent gas at very low thermal
energy may occur. Thus sometimes the electron attachment rate will have a
maximum at thermal energy. Attachment has a threshold and peak at lower
E/N than ionization, and thus there is an E/N where ionization catches up
and becomes equal to attachment, as shown in Figure 6.7. Slightly above that
point is the usual electrical breakdown point in electronegative gases. Thus
in electronegative gases, the E/N that satisfies Ri = Ra will give an ideal bulk
electric field in low-temperature plasmas.
where ω and E 0 are the angular frequency and the amplitude, respectively,
and E0 is usually taken to be −E 0 k. The rf transport is characterized as a
function of E 0 /N and ω/N in a gas with number density N, whereas the
DC-transport is only a function of E 0 /N.
∂ 0 1 0
g0 (v, t) + ζ E 0 , ω, g10 (v, t) = − g (v, t), (6.2)
∂t τe (v) 0
and for momentum relaxation
∂ 0 1
g1 (v, t) + ξ E 0 , ω, g00 (v, t) = − g 0 (v, t), (6.3)
∂t τm (v) 1
where g00 (v, t) and g10 (v, t) are the isotropic and directional parts of the velocity
distribution g 0 (v, t) expressed by the two-term approximation,
where θ is the angle between v and −E; that is, θ = −(v · E)/vE. The two time
constants τe (v) and τm (v) are defined, respectively, as
2m εv d
−1
τe (v) ∼Nv Qm (v)+ (Qv (v)v) + Q j (v)+Qi (v)+Qa (v) ,
M ε dv
(6.4)
Ar CF4
(a)
-15
10 10
-15 Qv3
Qi Qm
Cross section (cm )
2
Q1 Q
-16 -16 Qv1 d1
10 10
Qex
Qexm Qv4 Qex Q Qd3
-17 -17 d2 Qi3
10 10 Qi4
Qex(2p1) Qi2 Qi5
-18 -18
10 10 Qi6
Qex(3p5) Qi7
Qa
-19 -19
10 -2 -1 2
10 -2 -1 2
10 10 1 10 10 10 10 1 10 10
Electron energy (eV) Electron energy (eV)
(b) 10-6 10
-6
Collision frequency (cm s )
i
3 -1
3 -1
-7 -7 d1
10 10 i1
ex
v3 d3
-8 -8
10 exm 10 i3
m d2 i4
m v1 ex
-9 -9
10 ex(2p1) 10 i5
i2 i6
-10 -10 v4
10 10 i7
ex(3p5)
a
-11 -11
10 -2 -1 2
10 -2 -1 2
10 10 1 10 10 10 10 1 10 10
Electron energy (eV) Electron energy (eV)
(c) 10-2 10
-2
τe τe
-4 -4
10 10
Relaxation time (s)
-6 -6
10 10
τm
-8
10 10
-8
τm
-10 -10
10 10
-2 -1 2 -2 -1 2
10 10 1 10 10 10 10 1 10 10
Electron energy (eV) Electron energy (eV)
FIGURE 6.8
A set of cross sections of electrons in pure Ar and CF4 : (a) cross section, (b) collision frequency,
and (c) collisional relaxation time at 1 Torr.
10 MHz
6.0
Mean energy (eV)
1 GHz
2.0
0 2π 4π 6π 8π 10π
ωt
FIGURE 6.9
Examples
√ of the relaxation of the mean energy of electrons initiated at g M (v) at 2 eV at E 0 /N =
50 2 Td and 1 Torr in Ar: (a) 100 kHz, (b) 10 MHz, and (c) 1 GHz.
The properties in the second and third region are observed in the ioniza-
tion rate Ri in Ar in Figure 6.10. Periodic steady-state characteristics with
twice the frequency of the external field are realized as a function of ωt.
The time-averaged ionization rate is kept at an almost constant value up to
100 MHz. At frequencies greater than 100 MHz, the DC component rapidly
decreases, because electrons do not have enough time to get energy from the
external field before its direction changes. That is, electrons fall into a spa-
tial trapping as is shown in the term of the phase delay with respect to the
external E(t).
Exercise 6.3.1
By considering the time evolution of the electron number density ne (t) due to
ionization and electron attachment, estimate the form of ne (t) under a periodic
steady-state condition.
The time-dependent distribution function G 0 (v, t) may be separated into
the product of a time-varying number density ne (t) and a velocity distribution
function g 0 (v, t) normalized to unity under a periodic steady state
G 0 (v, t) = ne (t)g 0 (v, t) where g 0 (v, t)dv = 1. (6.6)
v
(a)
8.0
Ionization rate Ri(t) (107 s-1) 1 MHz
4.0
2.0
1 GHz
5 GHz
0.0
0 π 2π
ωt (rad.)
(b)
90
4.0
Ri(t)DC (107 s-1)
3.0 60 ϕi (deg.)
2.0
30
1.0
0.0 0
106 107 108 109 1010
Frequency (Hz)
FIGURE 6.10
√
Ionization rate Ri (t) at E 0 /N = 300 2 Td at 1 Torr in Ar as a function of external frequency:
(a) time-modulation of Ri (t), (b) DC-component and phase delay.
∂ 0 eE(t) ∂ 0
g (v, t) + · g (v, t) = J (G, F ) − RT (t)g 0 (v, t), (6.7)
∂t m ∂v
where RT (t) represents the difference between the ionization and attachment
rates, Ri (t) − Ra (t), and is related to the electron number density by
t
ne (t) = no exp RT (t)dt , (6.8)
0
E0 1 E0 1
E e f f (v) = √ 1/2 = √
2 1/2 . (6.9)
2 1 + (ωτm (v)) 2 2 ω
1+ R
m(v)
Exercise 6.3.2
Derive the effective field approximation 6.9 by using the momentum balance
equation of electrons.
The momentum balance of electrons with a uniform density in gases is
given by
d
m vd (t) = −eE(t) − Rm mvd (t),
dt
where vd is the drift velocity and Rm is the momentum transfer rate. The
solution is
eE0 1 −1 ω
vd (t)=− cos(ωt+φ) where φ= tan . (6.10)
mRm [1+( Rωm )2 ]1/2 Rm
In the limit of low frequency ((ω/Rm )2 1), the drift velocity converges to an
instantaneous DC-like solution, vd (t) = −e E 0 /mRm cos ωt, and for the high
frequency limit (ω/Rm )2 1) the solution is
eE0
vd (t) = − cos(ωt + φ). (6.11)
mω
H2 N2
(a)
10-14 10-14
Qm
Qm
Cross section (cm2)
νi νi
νm νeB 10 -7
3 + νd
10-8 B Σg
νeE νe2p2s
νre02 νe1sn3 10-8
νm ν vib. others
νdi
νre13 a Πg
1
10-4
Relaxation time (s)
τe
10-6
τe
10-6
10-8
τm 10-8 τm
10-10 10-10
10-2 10-1 1 10 102 10-2 10-1 1 10 102
Electron energy (eV) Electron energy (eV)
FIGURE 6.11
A set of cross sections of electrons in H2 and N2 : (a) cross section, (b) collision frequency, and
(c) collisional relaxation time at 1 Torr.
SiH4 SF6
(a) -14
10
Qm Qa1
10 -14
Qd Qm
Qi
Cross section (cm 2 )
10 -20
10 -20 -2
10 10 -1 1 10 10 2 10 -2 10 -1 1 10 10 2
Electron energy (eV) Electron energy (eV)
(b)
10 -6 10 -6
ν a1
νi
Collision frequency (cm 3 s -1 )
νd ν ex
νi νm
10 -7 -8
νm 10
ν a3
10 -8 ν a2 ν a4
10 -10
ν a5
100 ν a
10 -9 νv
-10
10
10 -2 10 -1 1 10 10 2 10 -2 10 -1 1 10 10 2
Electron energy (eV) Electron energy (eV)
(c) 10 -4
10 -4 τe
Relaxation time (s)
10 -6
10 -6
τe
10 -8
10 -8
τm
τm
10 -10 10 -10
10 -2 10 -1 1 10 10 2 10 -2 10 -1 1 10 10 2
Electron energy (eV) Electron energy (eV)
FIGURE 6.12
A set of cross sections of electrons in SiH4 and SF6 : (a) cross section, (b) collision frequency,
and (c) collisional relaxation time at 1 Torr.
We can see that in the limit of (Rm /ω)2 → 0 the electron motion is always in
the field direction and thus the electron cannot gain any energy from the field.
If there are collisions (represented by Rm ), electrons are able to attain some
mean energy. Thus collisions with gas or any other dissipation of momentum
will lead to a net gain of energy for electrons in rf fields. Next, the power
absorption of electrons is achieved by the drift in an rf field:
π π
PW = −eE(t) · vd (t)dωt/ dωt
0 0
2
1 e E0 e 2 E e2f f
= √ = . (6.12)
mRm 2(1 + (ω/Rm )2 )1/2 mRm
ikω(m/2)1/2 f sk ()
e ER s d s
+ √ [λ1 f s−1 () + f s−1 ()] − [λ1 f s−1 () + f s−1 ()]
k−1 k+1 k−1 k+1
2 2s − 1 d 2
e ER s + 1 d s +1
+ √ [λ1 f s+1 () + f s+1 ()] +
k−1 k+1
[λ1 f s+1 () + f s+1 ()]
k−1 k+1
2 2s + 3 d 2
m
1/2
+J [ f k (), Tg ] s=0
= Isk () 0
2 −N[Q m () + Q j () + Q i () + Q a ()] f s
k
() s= 0,
(6.15)
k k
J [ f 0k (), Tg ]
√ 2m d 2 d −1/2 k
= NQm () f o () + k B Tg NQm ()
3/2 k
[ f o ()]
M d d
d + j
+ −1/2 1/2 NQ j () f ok ()d
j
d
o
−1/2 d
+ 1/2 NQa () f ok ()d
d a
/δ+i /(1−δ)+i
d
+ −1/2 + 1/2
NQi () f ok ()d ,
d o
(6.17)
(6.18)
This is identical to the expression of Margenau and Hartman [6], except for
the presence of rotational and vibrational collisions. In atomic gases subject
only to elastic scattering, the distribution f 000 (ε) in the limit, ω−1 τm (ε), has
a Maxwellian form with an effective gas temperature
e ER 2 3k B Tg
Te f f = Tg 1 + / . (6.19)
mω M
The presence of rotational and vibrational collisions with low threshold en-
ergies in molecular gases changes the shape of the distribution substantially
in comparison with that obtained in atomic gases. That is, f 000 (ε) can be ex-
pressed in the asymptotic form of Equation 6.19 as
√ 1 [Qr ()r + Qv ()v ]
f 0 () = A exp −
00
1+ d . (6.20)
k B Te f f 0 (2m/M) Qm ()
The second term of the integral in Equation 6.20 shows the ratio of the colli-
sional energy loss between inelastic and elastic scattering. Recalling the fact
that the higher the threshold energy, the faster the energy relaxation time, one
would expect that ε/ε j should be weighted by the energy relaxation time for
inelastic scattering and that the effective relaxation time is expressed as
τee f f (ε)−1 = N[(2m/M)Qm (ε) + Q j (ε)ε j /ε]v (6.21)
k
3 0
∞
1
= Re (2/m)1/2 ei(2k+1)ωt 1/2 f 12k+1 ()d . (6.24)
k
3 0
(a)
ω t =0
g0 (υ )
5.0
Vr
(1
5.0
0
8
0.0 0.0
cm
-5.0 -1
Vz (10 8
cm s )
s
-1
)
(b)
0.5 ω t =0
g00 (ε)
Distribution gn0 (ε) (arb.)
g02 (ε)
0
g03 (ε) g10 (ε)
0 10 20
Electron energy (eV)
FIGURE 6.13
Electron velocity distribution g 0 (v, ωt √ = 0) (a) and each component of Legendre polynomials
gn0 ()(n = 0, 1, 2, 3) (b) at E 0 /N = 300 2 Td and 1 MHz at 1 Torr in Ar.
v = [eE(t)/m]t. (6.26)
Equation 6.25 suggests that it should be possible to calculate the time evolu-
tion from an arbitrary initial distribution to a final periodic steady state under
the rf field by a differential method. It is essential to consider the change both
of the incident angle and of the energy at collision in terms of a deterministic
finite probability for each type of collision. For this purpose, it is convenient to
where
NQm (v1 )v14
J (vx , v y , vz , t) = g(v1 , t)d
4πv 3
Q j (v2 )v22
+N g(v2 , t)d
4π v
NQi (v3 )v32
+ g(v3 , t)d
4πv(1 − δ)
NQi (v4 )v42
+ g(v4 , t)d
4πvδ
−N[Qm (v)+ Q j (v)+Qa (v) + Qi (v)]vg(v, t). (6.29)
A method not restricted by the relation in Equation 6.30 has been proposed [8].
In Figure 6.14 we show the √ time modulation of the velocity distributions
g 0 (vz , t) in Ar at E 0 /N = 50 2 Td and 13.56 MHz in a periodic steady state.
74 25.6
6.4
Tim 34 0
e ( 6.4 eV)
ns) gy (
0 25.6 Ener
FIGURE 6.14
√
Time modulation of the electron velocity distribution g 0 (vz , t) at E 0 /N = 50 2 Td and 13.56 MHz
at 1 Torr in Ar.
vde
1.0
0.0
-1.0
0 π 2π 0 π 2π 0 π 2π 0 π 2π
3.0
>ε >
Mean energy (eV)
2.0
1.0
0.0
0 π 2π 0 π 2π 0 π 2π 0 π 2π
(10 22 cm -1 s -1 )
Diffusion coefficients
1.5
NDT
1.0
NDL
0.5
0.0
0 π 2π 0 π 2π 0 π 2π 0 π 2π
1.0
0.8 Ri
Rate (10 6 s -1 )
0.6
0.4
0.2 Ra
0.0
0 π 2π 0 π 2π 0 π 2π 0 π 2π
ωt ωt ωt ωt
FIGURE 6.15
√
Time modulation of the swarm parameters in SiH4 at E 0 /N = 160 2 Td frequencies of 100 kHz,
13.56 MHz, 100 MHz, and 1 GHz.
PROBLEM 6.3.1
Discuss how to calculate the velocity distribution g 0 (v, t) of electrons in a periodic
steady state by using a time evolution algorithm of the DNP.
-14 Qe.m.t.
10
Qd.c.t.3 Qc.t.
-15
Cross section (cm )
10
2
-16
10
FIGURE 6.16
Cross sections for Ar+ in Ar (—) and CF4 (· · ·).
O-/ O2
+
CH3 /Ar
5
Drift velocity (cm s )
10 Cl-/ Ar
-1
+
O2 / O2
+
Ar / Ar
4
10
2 3
10 10 10
DC-E / N (Td)
FIGURE 6.17
Ion drift velocities in gases as a function of DC-E/N.
NDLL
NDTT
+
O2 / O2
18
10 2 3
10 10 10
DC-E / N (Td)
FIGURE 6.18
Diffusion coefficients of ions in gases as a function of DC-E/N. Transverse diffusion coefficient
(· · ·) and longitudinal diffusion coefficient (—).
that are possible for different ions. For example, ions that are not in their par-
ent gas (such as N+ in N2 or Cl− in Ar) may not have such a large charge
transfer cross section at high energies. This leads to a situation in which the
high energy ions may have a small cross section and may increase their energy
continuously, never coming to quasi-equilibrium with the external field (i.e.,
runaway). Fast neutrals produced in charge transfer collisions should be fol-
lowed separately, as they may participate in a number of processes, including
gas phase excitation (or even ionization), and secondary electron production
and sputtering and etching on a surface. Diffusion coefficients are classified
into longitudinal and transverse coefficients with respect to the direction of
the external electric field. Some diffusion coefficients are shown in Figure 6.18
as a function of E/N.
References
1. Huxley, L.G.H. and Crompton, R.W. 1973. The Drift and Diffusion of Electrons in
Gases. New York: Wiley Interscience.
2. Dutton, J. 1983. J. Phys. Chem. Ref. Data 4: 577; Gallagher, J.W., Beaty, E.C., Dutton,
J., and Pitchford, L.C. J. Phys. Chem. Ref. Data 12:109.
3. Tagashira, H., Sakai, Y., and Sakamoto, S. 1977. J. Phys. D 10:1051; Taniguchi, T.,
Tagashira, H., and Sakai, Y. J. Phys. D 10:2301.
4. von Engel, A. 1983. Electric Plasmas: Their Nature and Uses. London: Taylor &
Francis.
5. Goto, N. and Makabe, T. 1990. J. Phys. D 23:686.
6. Margenau, H. and Hartman, L.M. 1948. Phys. Rev. 73:309.
7. Maeda, K., Makabe, T., Nakano, N., Bzenić, S., and Petrović, Z.Lj. 1997, 1994.
Phys. Rev. E 55:5901; Maeda, K. and Makabe, T. Physica Scripta T53:61; Maeda, K.
and Makabe, T. Europhys. Conf. Abstr. 18E:151.
8. Matsui, J., Shibata, M., Nakano, N., and Makabe, T. 1998. J. Vac. Sci. Technol.
A 16:294.
9. White, R.D., Robson, R.E., and Ness, K.F. 1995. Aust. J. Phys. 48:925.
10. McDaniel, E.W. and Mason, E.A. 1973. The Mobility and Diffusion of Ions in Gases.
New York: John Wiley & Sons.
11. Mason, E.A. and McDaniel, E.W. 1988. Transport Properties of Ions in Gases.
New York: John Wiley & Sons.
12. Farrar, J.M. and Saunders, W.H. Jr. Ed. 1988. Techniques for the Study of Ion-
Molecule Reactions. New York: John Wiley & Sons.
13. Lindinger, W., Mark, T.D., and Howorka, F. Ed. 1984. Swarms of Ions and Electrons
in Gases. Wien: Springer Verlag.
14. Mason, E.H. and Viehland, L.A. 1976, 1978, 1984. Atomic Data and Nuclear Data
Tables, 17: 177; Ellis, H.W., McDaniel, E.W., Albritton, D.A., Viehland, L.A., Lin,
S.L., and Mason, E.A., ibid. 22:179; Ellis, H.W., Thackston, G., McDaniel, E.W.,
and Mason, E.A. ibid. 31:113.
7.1 Introduction
Early attempts to model plasmas were based on zero-dimensional and
phenomenological models. An example of this approach is the circuit model,
which represents plasma behavior using passive circuit elements, resistance,
capacitance, and inductance. Although such an approach is useful for repre-
senting specific plasmas and studying how they behave once connected to
the external circuit, it is beyond the scope of this chapter to specify plasma
characteristics intrinsic in a feed gas molecule. There is a rate equation model
describing zero-dimensional plasma, which has proven quite useful for the
kinetic optimization of a gas laser, mainly by utilizing a bulk plasma.
During the 1980s and early 1990s, it was recognized that plasma-enhanced
chemical vapor deposition of amorphous silicon and plasma etching for
integrated circuits could not be improved further by a trial-and-error
approach, and that a complete understanding of these processes was required
for a new generation of technologies. This was partially motivated by the
increasing cost of the development of the equipment and complexity of reac-
tion chambers.
The need for radio frequency (rf) plasmas arose from the fact that, in these
technologies, one needs also to treat metals, semiconductors, and dielectrics
at the same time. There is an important difference between modeling of direct
current (DC) and alternating current (AC) discharge plasmas, especially those
operating in an rf range. Rfs have periods on the order of nanoseconds, much
shorter than the characteristic lifetimes of many of the gas-phase processes,
and yet the discharge clearly develops and exhibits complex behavior on such
short time scales, so the characteristic time step of the numerical simulation
must be significantly shorter than the period. At the same time, one must take
into account slow processes such as the kinetics of chemically active neutrals
and the flow of gas through the system that will occur on time scales from
milliseconds to seconds. As a result of this vast difference in time scales, we
need to cover as many as 10 orders of magnitude in time development. Thus
the rf plasma for material processings belongs to the class of the stiff system,
and special attention should be paid to the numerical techniques or physical
205
FIGURE 7.1
Governing equation system of a low-temperature plasma for material processing.
momentum balance ⇐ (Boltzmann Eq.) vdv (7.2)
energy balance ⇐ (Boltzmann Eq.) v2 dv. (7.3)
These laws should be written for all particles involved and should be
completed by a set of Maxwell’s equations. This model treats particles as
fluids. Continuum models have a key and practical advantage of having a
short computation time, and the validity is kept under conditions that the
mean free path of charged particles, especially electrons, is shorter than or
equivalent to the characteristic length of a plasma reactor. In principle, these
Equations 7.1 to 7.3 are not closed, so additional equations should be added.
For example, in the case of thermal equilibrium, the equation of state is used to
complete the system. In nonequilibrium plasmas, the electron energy distri-
bution, which is dominated by the short-range electron–molecule collisions,
is non-Maxwellian, and we need a procedure to find solutions to the corre-
sponding transport (Boltzmann) equation.
Typically, nonequilibrium rf plasma consists both of a bulk plasma
with quasi-neutrality between electrons and positive ions and of ion sheaths.
Both regions are periodically modulated in time even when quasi-stationary
conditions are achieved. In order to deal with such a complex system, we
need to establish a modeling hierarchy that involves several different
models.
In the one-dimensional space along the axial field direction (z-axis), the
continuity equation for electrons has the form
Poisson’s equation for the electric field E(z, t) and the space potential V(z, t) is
∂ E(z, t) ∂ 2 V(z, t) e
=− = {n p (z, t) − ne (z, t) − nn (z, t)}. (7.7)
∂z ∂z2 ε0
Here, the number densities of electrons, positive ions, and negative ions are
ne (z, t), n p (z, t), and nn (z, t), respectively; Vdj and DL j are the drift velocity
and the longitudinal component of the diffusion tensor of the jth particles
(e, p, n); Ri and R a are the ionization and electron attachment rates; and Rr e
and Rri are rates of recombination in electron–ion and ion–ion collisions. The
data for transport coefficients Vdj , DL j , Ri , and so on, which are taken from a
database of experimental or theoretical data, are provided for the local and in-
stantaneous electric field E(z, t). One example of a set of data for the collision
rate for O2 is shown in Figure 7.2. In Equation 7.7, e(> 0) is the elementary
charge and ε0 is the permittivity (dielectric constant) of the vacuum.
The set of Equations 7.4 to 7.7 is numerically calculated for a time-varying
one-dimensional capacitively coupled plasma (CCP) with infinite parallel
plates. In a finite plasma source or axisymmetric plasma, these equations are
developed in the two- or three-dimensional configuration space.
PROBLEM 7.2.1
Derive the continuity of the current density in the one-dimensional position space
∂ JT ∂
= J e (z, t) + J p (z, t) + Jn (z, t) + JDis (z, t) = 0 (7.8)
∂z ∂z
from the Equations 7.4 to 7.7. Here Je , J p , and Jn are the current density of elec-
trons and of positive and negative ions, respectively. J Dis is the displacement current
density.
FIGURE 7.2
Collision rates of the electron in pure oxygen as a function of DC-E/N.
Faraday’s law:
1 ∂E z ∂E θ Br
−
r ∂θ ∂z
∂Er ∂E z ∂
− =− Bθ ; (7.10)
∂z ∂r ∂t
1 ∂ 1 ∂Er
(r E θ ) − Bz
r ∂r r ∂θ
Ampere’s law:
1 ∂Bz ∂Bθ Er Jer J pr
−
r ∂θ ∂z
1 ∂Br ∂Bz ∂
− = −ε0 E θ + Jeθ + J pθ ; (7.11)
µ0 ∂z ∂r ∂t
1 ∂ 1 ∂Br
(r Bθ ) − Ez Jez J pz
r ∂r r ∂θ
and Coulomb’s law in magnetics:
1 ∂ 1 ∂Bθ ∂Bz
(r Br ) + + = 0; (7.12)
r ∂t r ∂θ ∂z
where µ0 is the magnetic permeability of vacuum and where µ0 ε0 = c −2
(c is the speed of light).
PROBLEM 7.2.2
Derive the following continuity equation for a short-lived excited molecule with a
number density Nj (k)
∂ Nk
Nk (t) = j k jk ne Nj + l m klk Nl Nm − j l kk j Nk Nl − k , (7.13)
∂t τrad
Even for higher neutral density (pressure), the spatial variation of the field
in the sheath close to the electrode is very large and the application of LFA
for these regions will be inaccurate. In practical terms, the LFA algorithm
provides very fast numerical calculation and thus should be the first algorithm
to be attempted whenever there exists a database of the transport (swarm)
parameters.
where k is the Boltzmann constant and ki is the constant intrinsic to the feed
gas. The electron temperature is obtained from the energy 3 kTe (z, t)/2 con-
servation equation given as
∂ 3 ∂q e (z, t)
ne (z, t)kTe (z, t) +
∂t 2 ∂z
∂ne (z, t)
= e E(z, t) −ne (z, t)Vde (z, t) + DLe
∂z
εi
− εi ne (z, t)Nki exp −
kTe (z, t)
εa
− εa ne (z, t)Nka exp −
kTe (z, t)
εj
− ε j ne (z, t)Nk j exp − , (7.15)
kTe (z, t)
This model is based on the second-order velocity moment, that is, energy
balance. It is applicable to a plasma modeling, if one can represent data
TABLE 7.1
Low-Temperature Plasma Model and the Variable
Model Variable References
where τm is the effective time constant for momentum transfer (the inverse
of the momentum transfer collision rate and approximately equal to mVde0 /eE
by use of the DC-drift velocity Vde0 ). The first term on the right-hand side is
the momentum gain due to electric field acceleration, the second term is the
loss due to momentum transfer collisions, and the third term describes the
effect of the spatial gradient of drift velocity on relaxation.
The relaxation of the mean energy of electrons, εm (z, t), is described by the
conservation of energy,
∂ ∂ne (z, t)
ne (z, t)εm (z, t) = eE(z, t) − ne (z, t)Vde (z, t) + DLe
∂t ∂z
2m
− Rm εm (z, t) + R j ε j + Ri εi ne (z, t)
M
∂ ∂ne (z, t)
− ne (z, t)Vde (z, t) − DLe εm (z, t) . (7.18)
∂z ∂z
The r.h.s in Equation 7.18 is described by the terms of the energy gain due
to flux of charged particles in the electric field, followed by energy losses
in elastic momentum transfer, inelastic and ionization collisions, and finally
the spatial gradient of the mean energy. We introduce the effective field E eff ,
where E(z, t)2eff = εm /e µτe , which gives a field modified from the actual field
E(z, t) to take into account the energy relaxation from the proper transport
data under quasi-equilibrium. The relaxation equation for the effective field
may be written as
∂
q e (z, t) = ne (z, t)Vde (z, t)E eff (z, t)2 − DLe (z, t) ne (z, t)E eff (z, t)2 . (7.20)
∂z
-15 Qv3
10
Qm
Qil
-16
Qv1 Qdl
Cross section (cm )
10
2
Qd3
Qv4 Qex
Qd2 Qi3
-17
10
Qi4
Qi5
Qi2
-18
10 Qi6
Qi7
Qa
-19
10 -2 -1 2
10 10 1 10 10
Electron energy (eV)
FIGURE 7.3
Set of collision cross sections in CF4 as a function of electron energy.
The temporal relaxation time for electron diffusion is of the same mag-
nitude as that for the mean energy. However, the energy relaxation time of
electrons with energy greater than the excitation threshold j is more diffi-
cult to estimate than those for drift velocity and mean energy, which are the
result of the effect of the entire energy distribution of electrons. In particular,
it is important to establish the relaxation time constant of the ionization rate
τ Ri . The ionization is the result of only the electrons at the highest end of the
energy distribution, and these have kinetics quite different from the bulk of
the distribution. Therefore, for τ Ri we must make an estimate from the relax-
ation times for the mono-energetic groups of electrons (see Figure 7.3). The
RCT model is a theory that maintains all the advantages of continuum models
with respect to speed, simplicity, and ease of interpretation.
PROBLEM 7.2.3
Discuss the time behavior of the drift velocity of electrons as a function of external
field E(t) = E 0 cos ωt from the momentum relaxation Equation 7.17 in the case of
τm
ω−1 (a) and τm ω−1 (b). For simplicity, we assume that the electron number
density and the drift velocity are distributed uniformly in space.
equation (see Section 5.6), but the most difficult issues are the need for huge
CPU time and the possibility of numerical diffusion if the resolution is
reduced in order to speed up the computation. Here we discuss the two-
term approximation (TTA) as a means of solving the Boltzmann equation for
the time- and space-modulated electric field E(z, t). The equation for the en-
ergy distribution, f (ε, z, t), based on the TTA of the velocity distribution of
electrons is
∂
f (ε, z, t)
∂t
2 2m ∂ Me E(z, t)ε ∂ ∂
= NQm (ε)ε 3/2 f + eE (ε −1/2 f )+ (ε−1/2 f )
m M ∂ε 6mNQT (ε) ∂ε ∂z
2 ε ∂2 ∂2
+ eE(z, t) (ε −1/2 f ) + 2 (ε −1/2 f )
m 3NQT (ε) ∂ε∂z ∂z
ε + εj εa
2 ∂ 2 ∂
+ ε 1/2 NQ j (ε) f dε − ε 1/2 NQa (ε) f dε
m j ∂ε ε m ∂ε 0
1+k
(1 + k)ε + εi k ε + εi
2 ∂
+ + ε 1/2 NQi (ε) f dε . (7.21)
m ∂ε ε 0
λe ≥ d, (7.22)
d2 e
2
r(t) = [E(r, t) + v × B(r, t)] , (7.23)
dt m
and during the short period t, the collision occurs when the total collisional
probability NQT (ε)vt satisfies
Here, {ξr } is the random number distributed uniformly in [0, 1]. Of course,
the time step t must satisfy the condition
1
t
. (7.25)
NQT (ε)v
PROBLEM 7.3.1
The probability that the electron will have a collision in the time interval (0, t) is
given by
t
p(t) = NνT (ε(t))exp − NνT (ε(t))dt ,
0
At the same time, the scattering angle (θ, φ) after the collision must be
determined by the random number. In the case of a collision, the energy and
(a)
Qm Qex1 Qex2 … Qi
collision probability
NQT v∆t
(b) ϕ
v'
FIGURE 7.4
Determination of the collision and the kind by the random number {ξr } (a) and the scattering
angle at collision {γr } (b).
the velocity of the particle change, subject to the momentum and the energy
conservation rule between two colliding particles (see Chapter 4). The trace
of the electron with (v, r, t + t) after the collision restarts during the next
t, and the same procedure is carried out.
Exercise 7.3.1
An electron has a binary collision with a neutral molecule and is scattered
isotropically. Derive the distribution of the scattering angle (θ, φ) with respect
to the incident direction given by
1
p(θ ) = sin θ, (7.26)
2
1
p(φ) = . (7.27)
2π
The function p(θ, φ) (the probability density function) distributed uniformly
in a unit sphere (1, θ, φ) is expressed by
sin θ dθ dφ
p(θ, φ) = .
4π
The θ and φ in the spherical coordinates are independent of each other, and
we can write
Therefore,
2π
2π sin θ dθ 1
pθ (θ )dθ pφ (φ)dφ = = sin θ dθ, (7.29)
0 4π 2
π
2dφ 1
pφ (φ)dφ pθ (θ )dθ = = dφ . (7.30)
0 4π 2π
Exercise 7.3.2
Obtain the scattering angle (θ, φ) distributed isotropically by using a series
of random numbers {γr } uniformly distributed between [0, 1].
From Equation 7.29,
θ
1
sin θ dθ = γr → cos θ = 1 − 2γr . (7.31)
0 2
a eh ∆z - a
z
zi zj z i+1
∆z
FIGURE 7.5
Superparticle and the charge sharing between the nearest two grids.
at distance a from the lattice z j in Figure 7.5, the charge is divided into aeh
and (z − a )eh at z j and z j+1 , respectively. The electric field at the position of
the superparticle is given by linear interpolation, after the field is calculated
by Poisson’s equation at each of the lattices
t < ω−1
pe , (7.37)
z < λ D , (7.38)
where ω pe and λ D are the plasma frequency and Debye length in a collisionless
plasma (see Chapter 3).
When a plasma is maintained by electron impact ionization of a feed gas
molecule, we must consider the short-range interaction even in the case where
a strongly ionized plasma is mainly subject to the long-range interaction.
The binary collision is considered by MCS. Thus the model is classified as a
PIC/MCS model. PIC/MCS is widely used in low-pressure plasmas. When
the operating pressure of a plasma increases, the number of collision events
between the electron and feed gas molecule is huge, and thus the continuum
fluid model is much more appropriate for describing the system. The main
advantage of the particle model as opposed to the fluid model is that, in
the former, no assumption for the velocity or energy distribution of charged
particles is made implicitly. In compensation, the cost is high because the PIC
model is very time consuming.
R sh1 Rb R sh2
L1 Cb
IT
50 Ω Vsus
C1 C2 C sh1 Lb C sh2
FIGURE 7.6
Typical equivalent circuit of a capacitively coupled rf plasma with an external impedance match-
ing network.
ne e 2
Je (t) = ene vde (t) = Eb exp( jωt)
m(Rm + jω)
ne e 2 Rm 1 1
= −j Eb exp( jωt) .
mω ω 1 + ( Rωm )2 1 + ( Rωm )2
As a result, the first in-phase oscillation term gives a power dissipation at the
resistive impedance (plasma resistance Rb ), the second and third out-of-phase
oscillations correspond to an inductive and capacitive impedance ( jωL b and
1/jωCb ) with no power dissipation (reactance).
mω ω Rm 2 db ω 2 1 Rm 2 ε0 Sb −1
Rb = 1+ = 1+ ,
ne e 2 Rm ω Sb ω pe Rm ω db
(7.40)
2
2
−1
mω0 Rm db 1 Rm ε0 Sb
Lb = 1+ = 2 1+ , (7.41)
ne e 2 ω Sb ω pe ω db
0 Sb
Cb = , (7.42)
db
where ω2pe ( = e 2 ne /mε0 ) is the electron plasma frequency. Sb and db are the
cross sectional area and thickness of the bulk plasma. In fact, the third in
Equation 7.39, displacement component will be negligible in electropositive
plasmas. In a collision-dominated plasma with Rm ω, the impedance of the
bulk plasma is represented by a resistive component R b .
In the sheath of a typical rf discharge plasma, the circumstances change
the relation completely among the current components, Je , J p , and JDis . The
sheath region having n p ne is formed in front of electrodes or a wafer in a
discharge plasma, and the displacement current occupies a large part of J T ,
although finite currents of positive ions and electrons flow under a DC self-
bias voltage in a time-averaged fashion with (Je (t) + J p (t)) = 0. As a result,
the equivalent circuit of the sheath having a voltage drop Vsh is composed
L1 Cb
I sus
M
Rc
50 Ω Vsus Lc Lp Rp
C1 C2
FIGURE 7.7
Typical equivalent circuit of an inductively coupled rf plasma with an external impedance match-
ing network.
of the capacitance Csh (= ε0 Ssh /dsh ) and resistance R sh , where Ssh and dsh are
the cross sectional area and thickness of the sheath. The power is dissipated
at the resistance Rsh by a collisional effect between the charged particle and
feed gas molecule.
1
J T (t) ∼ JDis + J p ∼ jωCsh + Vsh . (7.43)
Rsh
We show a typical equivalent circuit of the CCP with metallic electrodes con-
nected to an external impedance matching network in Figure 7.6.
When there is a dielectric substrate on a metallic electrode, we have to
add a series capacitor between the sheath circuit and the matching network
(or grounded-metal electrode). In front of the reactor wall, a wall sheath is
formed and is modeled by the capacitance similar to the above model. The
equivalent circuit model is usually created by using the passive elements (i.e.,
resistance, capacitance, and inductance). Sometimes, active elements such as
the diode and current source may be included in the circuit to realize the
detailed system.
(a) Z/ 2 Z/ 2
Vi-1, j Vi, j
(b) Z/ 2 Z/ 2
Vi-1, j Vi, j
(c) Z/ 2 Z/ 2
Vi-1, j Vi, j
Ii-1, j ZE Y
FIGURE 7.8
Transmission line, modeled by T-type equivalent circuit (a) and the other elemental circuits (b)
and (c).
dV(x) dI(x)
= −ZI(x) and = −YV(x).
dx dx
d 2 V(x)
= ZYV(x). (7.44)
dx 2
Here note that V(x), I(x), Z, and Y are complex numbers, expressing the am-
plitude and phase shift. In the TLM, three types of T-type circuits are usually
used (see Figure 7.8). We deal with a plasma sustained between the rod and
plane with distance l. The series impedance Z and parallel admittance Y have
to be investigated. The widths of the sheath and bulk plasma are dsh and db ,
perpendicular to the electrodes, respectively; that is, l = 2dsh + db . Then, the
impedance Z is written as
Z = R + jωL ,
1+ r02 + 1
µ0
= ρ/(2πr t) + jω log − r02 + 1 + r0 , (7.45)
2π r0
where ρ is the resistivity of the rod electrode with radius r0 . The rf current
flows in a thin region r0 ≥ r ≥ (r0 − dskin ) along the rod. dskin is the skin
depth. µ0 is the permeability in vacuum. The plasma admittance including
the sheath Y per unit length is given in the form similar to Section 7.5.1,
1 Rb 1 Rb 2
= 2 + j 2 −j , (7.46)
Y ωL b ωCsh
1+ Rb
ωL b
1+ Rb
ωL b
F(r) e r
= E(r) = , (7.47)
q 4π ε0r r
2
E(t)
E(t)
I(t)
B(t)
I(t)
I 0 sinωt
FIGURE 7.9
Schematic representation of the effect of a time-varying current I0 sin ωt that induces external field
E(t). However, the external time-varying magnetic field B(t) induces a field in the conductor
(dashed lines) and affects the resulting current.
= B(r, t) · ndS. (7.52)
S
These are Maxwell’s equations, which provide a basis for understanding the
development of electric and magnetic fields.
References
1. Chua, L.O. and Lin, P.M. 1975. Computer-Aided Analysis of Electronic Circuit.
Englewood Cliffs, NJ: Prentice-Hall.
2. Makabe, T. Ed. 2002. Advances in Low Temperature RF Plasmas: Basis for Process
Design. Amsterdam: Elsevier.
3. Kim, H.C. Iza, F., Yang, S.S., Radjenovic, M.R., and Lee, J.K. 2005. J. Phys. D.
Topical Review 38: R283.
4. Boeuf, J.-P. 1987. Phys. Rev. A 36: 2782.
5. Graves, D.B. 1987. J. Appl. Phys. 62: 88.
6. Dkazaki, K., Makabe, T., and Yamaguchi, Y. 1989. Appl. Phys. Lett. 54: 1742.
Makabe, T., Nakano, N. and Yamaguchi, Y. 1992. Phys. Rev. A 45: 2520.
7. Sommerer T.J., Hitchon, W.N.G., and Lawler, J.E. 1989. Phys. Rev. Lett. 63: 2361.
8. Birdsall, C.K. and Langdon, A.B. 1985. Plasma Physics via Computer Simulation.
New York: McGraw-Hill.
9. Kushner, M.J. 1986. IEEE Trans. Plasma Sci. PS-14: 188.
10. Binder, K. and Heermann, B.W. 1988. Monte Carlo Simulation in Statistical
Physics. Berlin: Springer-Verlag.
11. Koenig, H.R. and Maissel, L.I. 1970. IBM J. Res. Develop. 14: 168.
12. Piejak, R.B., Godyak, V.A., and Alexandrovich, B.M. 1992. Plasma Sources
Sci. Technol. 1: 179.
13. Satake, K., Yamakoshi, H., and Noda, M. 2004. Plasma Sources Sci. Technol.
13:436.
1
= Rm + R j , (8.1)
τm
where Rm and R j are the rates of the elastic momentum transfer and the in-
elastic scattering with threshold energy j , respectively. The energy relaxation
time of electrons by two-body collisions in a time-varying field are obatined
from the conservation of energy in Chapter 7:
1 2m j R j
= Rm + , (8.2)
τ M m
where m and M are the mass of the electron and the molecule, respectively,
and m is the averaged energy of electrons.
231
102
etching
10-4
energy transfer
-6 (1 Torr)
10
ion transit
ωpe
-1
10 -8 τrad
momentum transfer
(1 Torr)
10-10 ωpp
-1
dielectric
~1 ~10
εe (eV)
FIGURE 8.1
Time constant and relaxation time of each of the processes in a low-temperature plasma.
where kr is the reaction rate constant, and Nr the reactant number density. For
charged particles, the characteristic drift time is given by
L
τd = , (8.5)
vd
where vd is the drift velocity and L is the characteristic drift length.
p Vvol
τres = , (8.6)
Qflow
where p, Vvol , and Qflow are the pressure, effective plasma volume, and total
mass flow rate, respectively.
Exercise 8.1.1
A plasma reactor with volume 103 cm3 is operated at pressure 1 Torr and flow
rate 50 sccm (standard cubic cm per minutes). Estimate the residence time of
the feed gas.
1 Torr
× 103 cm3
τres = 760 Torr
50 cm3
∼ 1.6 s.
60 s
PROBLEM 8.1.1
Derive Equation 8.4 from the continuity equation of the system,
∂
Nj (r, t) = −kr Nk Nj + D∇ 2 Nj . (8.7)
∂t
In Equation 8.4, kr Nr 2 /D is called the Damkohler number D̃. Show that the sys-
tem is controlled by diffusion in the case of D̃
1 and by reaction in the case of
D̃ 1.
∂ρT (t)
+ div(eΓT ) = 0 , (8.8)
∂t
where ρT (t) = e{n p (t) − ne (t)}. Under these circumstances, ΓT is a drift flux
only because the uniformity of the charge distribution and the drift flux may
be related to the electric field by apparent mobility µ:
-9
10
∆tnew
-10
10
t (s)
-11
10
τd
-12
10
-13
10 9 10 11 12
10 10 10 10
Plasma density (cm-3)
FIGURE 8.2
Dielectric relaxation time constant τd and the time steps used in the semi-empirical method as a
function of plasma density.
which forms the relaxation equation for total charge, leading to the solution
t
ρT (t) = ρT0 exp − ,
τd
and therefore we may define the dielectric relaxation time from Equation 8.10
as
ε0
τd = . (8.11)
e(µ p n p + µe ne )
τd is also known as the Maxwell relaxation time. Solutions to the system are
obtained only if the time step t is shorter than the relaxation time (during
relaxation time τd the charge may decay by a factor of 1/e):
ε0
t < τd = , (8.12)
e(µ p n p + µe ne )
Exercise 8.1.2
Estimate the dielectric relaxation time for plasmas with density
(a) 109 cm−3 and (b) 1012 cm−3 .
ε0 ε
τd = ∼ 0 µp np
e(µ p n p + µe ne ) eµe ne 1 + µe ne
8.854 × 10−12 1
∼ × (s),
1.6 × 10−19 µe [m2 s −1 V −1 ]ne [m−3 ]
where for µe ∼ 70 m2 /sV, the calculated values are (a )τd ∼ 5.5 × 107 × 1.4 ×
10−2 × 10−15 ∼ 10−10 s and (b)τd ∼ 10−13 s.
The dielectric relaxation time becomes very small with increasing plasma
density. For example, for n ∼ 1012 cm−3 , the relaxation time is τd ∼ 10−13 s,
which creates a difficult situation in the time-development modeling method
due to the excessive CPU time required.
Exact
Backward
Central
Forward
∆z ∆z
FIGURE 8.3
Example of a finite difference scheme.
∂ ni − ni−1 z ∂ 2n
n(z, t) ∼ + (BD); (8.15)
∂z z 2 ∂z2 i
∂2 ni+1 − 2ni + ni−1 (z)2 ∂ 4n
n(z, t) ∼ − (CD). (8.16)
∂z2 (z) 2 12 ∂z4 i
Exercise 8.2.1
A low-temperature rf plasma is described by a system of stiff differential
equations. It is known that Gear’s algorithms of orders 1 to 6 are stiffly stable
[2]. Discuss the algorithms.
For example, for a differential equation ∂n(z, t)/∂t = d(n(z, t), t), Gear’s
algorithms for the first three orders are
∂n ∂n ∂ 2n
= −vd + D 2. (8.17)
∂t ∂z ∂z
Here, we use the central difference for uniform spatial mesh (index i) and
calculate the spatial density nim+1 = n(tm+1 , zi ) at time tm+1 = tm + t as
nm − ni−1 m
nm + ni−1 m
− 2nim
nim+1 = nim + −vd i+1 + D i+1 t, (8.18)
2z (z)2
where z is the spatial step. The solution is rewritten as [1],
c m c m
nim+1 = (1 − 2d)nim + d − ni+1 + d + n , (8.19)
2 2 i−1
where
t
d=D (> 0). (8.20)
(z)2
d is the ratio between the time step t and the characteristic diffusion time
(z)2 /D (i.e., the time it takes to move by z due to diffusion).
t
c = vd (> 0) (8.21)
z
is the ratio between the time step and the characteristic drift time (z/vd ),
and c is also known as the Courant number. Finally, we rewrite Equation 8.19
as
change both with time and with changing conditions in the spatial profile
of the plasma. The difference between the densities obtained in subsequent
iterations is:
2
ε = nim − nim−1 = nim − nim−1 , (8.23)
i
where the number density nim is stable under |λ|2 < 1. There are three possible
situations (see Figure 8.4).
PROBLEM 8.2.1
Derive the relations in Equation 8.26 by reference to the trajectory of Equation 8.25
in the complex plane in Figure 8.4.
(z)2
t < . (8.27)
2D
This equation indicates that if we reduce the size of the spatial mash z (or
increase the spatial resolution) by a factor of 2, the time step should be reduced
by a factor of 1/4, and therefore the overall CPU time should increase by a
factor of 8. Also, by combining Equations 8.21 and 8.26 we have
z
t < . (8.28)
vd
Im
1
c
Re
0 1-2d 1
-c
FIGURE 8.4
Trajectory of the eigenvalue Equation 8.25.
uim+1 − uim
= βuim .
t
This means that at time tm+1 the solution um+1 may be represented by the
previous time tm so that
and if we start at t = 0 from the initial distribution u(0) after m steps or after
time t = mt(= T), we obtain
|1 + βt| ≤ 1, (8.34)
and therefore
−βt ≤ 2. (8.35)
From Equation 8.33 and in the limit t → 0 one obtains a continuous solution
βT
(1 + βt)m = (1 + βt)T/t = (1 + βt)1/βt → e βT ,
1
|δ| = (βT)(βt). (8.36)
2
When we choose very small t for large T = mt, the relative error will
approach 1. The relation 8.36 may be rewritten as a condition of accuracy:
2δ0
t < . (8.37)
β2T
Γi + 1
ni 2 ni + 1
i i+ 1 i+1
2
FIGURE 8.5
Staggered mesh method. Scalar quantities are defined on integer i mesh points and vector quan-
tities on mid i + 1/2 points.
∂n ∂n
= −vd . (8.38)
∂t ∂z
The drift equation should be solved from the upwind to the downwind
direction. There are a few methods that may be used to achieve stability of the
drift-diffusion Equation 8.17 when condition 8.30 is not met. One is known as
the Patankar difference method, but the practical application that is discussed
here is the so-called Scharfetter–Gummel (SG) method [3]. In this procedure
we use the staggered mesh method for spatial distribution; in this method,
all scalar quantities are defined on integer mesh points and vector quantities
are defined on midpoints, as described in Figure 8.5.
The flux at time tm is a combination of drift and diffusion fluxes:
∂nm
m = nm vdm − Dm , (8.39)
∂z
where the values of m and vdm are represented at the midpoint (i + 1/2)
between the mesh point i and i + 1. The first-order differential Equation 8.39
is analytically solved in the range of z(0 < z < z) as
vdm i+1/2 i+1/2
m
nm (z) = C exp z + , (8.40)
m
Di+1/2 vd m i+1/2
and we take into account the boundary condition nm (0) = nim at i. We derive
the value of the constant as
i+1/2
m
C = nim − . (8.41)
vdm i+1/2
As a result, nm (z) = ni+1
m
is expressed as
i+1/2
m vdm i+1/2 i+1/2
m
m
ni+1 = nim − exp z + , (8.42)
vdm i+1/2 m
Di+1/2 vdm i+1/2
vm
nim exp d i+1/2
m
Di+1/2
z − ni+1
m
i+1/2
m
= vd i+1/2
m vd i+1/2
m
. (8.43)
exp m
Di+1/2
z − 1
vm vm m
i+1/2
m
− i−1/2
m m
ni+1 − exp Ddim z + 1 nim + exp Ddim z ni−1
− =
i
vm
i
. (8.44)
z exp Ddim z − 1 z
vm i di
PROBLEM 8.2.2
Prove that Equation 8.44 reduces to: (i) for Pe → ∞, then the equation is expressed
by:
m
vdi
− nim − ni−1
m
(8.45)
z
and we have the backward difference equation for the drift flux, and (ii) for Pe → 0,
we obtain the diffusion flux
Dim m
ni+1 − 2nim + ni−1
m
, (8.46)
(z) 2
Equation 8.44 is thus able to represent the whole range of conditions occur-
ring in the sheath and in the bulk where both drift-dominated and diffusion-
dominated fluxes occur.
Exercise 8.2.2
Show that the backward difference Equation 8.45 is equivalent to a drift-
diffusion equation with an artificial coefficient D = vd z/2.
m
vdi m ni+1 − ni−1 v m z ni+1 − 2nim + ni−1
m m m m
− nim − ni−1
m
= −vdi + di , (8.47)
z 2z 2 (z)2
we obtain two terms, respectively describing the derivatives of the drift flux
and the effective diffusion flux with the artificial diffusion coefficient, vd z/2,
in the central difference scheme. The truncation error with order O(z)2 in
Equation 8.47 has the effect of an artificial diffusion.
PROBLEM 8.2.3
When the drift is along the positive z-axis, show that Equation 8.45 should be ex-
pressed as
|vd | m
− ni − ni−1
m
. (8.48)
z
On the other hand, when the drift is along the negative z-axis, show that it is given
as
|vd | m
− ni − ni+1
m
. (8.49)
z
This method, in which the artificial diffusion in the central diffusion scheme is reduced
by checking determining the flow direction, is called the upwind scheme.
t t+∆t
z
zi-1 zi
zi - vd∆t
FIGURE 8.6
Temporal development of the number density subject only to the drift equation.
where the constants h i and ki are obtained by n(zi+1 ) and dn(zi+1 )/dt in
Equation 8.52. Then,
dnm
nim+1 = nim + − h im vdi
m
t + kim vdi
m
t − i vdi
m
t (8.53)
dz
dnim+1 dnim m m m
= + 3h i vdi t − 2kim vdi t, (8.54)
dz dz
where by using the sign, that is, isgn = 1 at vd < 0 and isgn = −1 at vd > 0,
we have
m
d
dz
nim + ni+isgn
m
2 ni+isgn − nim
h im = × isgn −
(z)2
(z)3
m m
d
dz
2ni + ni+isgn
m
3 ni+isgn − nim
ki = −
m
× isgn + .
z z)2
Then, by collecting the terms in Equation 8.55 with respect to the position
grids, i − 1, i, and i + 1, we obtain
m
Di−1/2 t m
Di+1/2 t Di−1/2
m
t m+1 Di+1/2
m
t m+1 ∗ m
− m+1
ni−1 + 1+ + ni − ni+1 =ni +Si t.
z2 z 2 z 2 z 2
(8.56)
After we write the Equation 8.56 for all the spatial meshes, the simultaneous
equations are expressed in the matrix form
m+1
n1 d1
b1 c1 0 m+1
a n2 d2
2 b2 c2
.
.
.. ..
... ...
= . (8.57)
.. ..
... ... .
.
a I −2 b I −2 c I −2
m+1
n I −2 d I −2
0 b I −1 c I −1
nm+1
I −1
d I −1
∂ρT (z, t)
−ε0 ∇ 2 V(z, t + t) = ρT (z, t + t) ∼ ρT (z, t) + t. (8.59)
∂t
Here, we use ρT (z, t) = −ε0 ∇ 2 V(z, t) and ∂ρT /∂t = −div(eΓT (t)), and thus
we obtain
and, consequently,
Note that the reflection (as well the absorption) at the surface affects the
velocity distribution function close to the surface, g(v, r, t). At the perfect
absorbing boundary, the velocity distribution function satisfies
In the case of the ideal perfect reflection surface, the velocity after reflection
v only changes direction perpendicular to the surface:
where αref = 0 for the perfect absorbing boundary, and αref = 1 for specular
reflection in Equation 8.64.
In Figure 8.7 we schematically show the effect of reflection on the flux of
particles and on the velocity. In addition, this schematic shows that, due to
surface irregularities, the direction of the particles after scattering may not
have the same angles as the incoming particles.
plasma surface
nT
v'
θ
θ
v
v'
θ
v
r = r0
FIGURE 8.7
Reflection of particles from the surface: reflection of a flux of particles, velocities before v and
after reflection v against an ideally flat and against a realistic irregular surface.
Chantry introduced the effect of the reflection of the species incident on the
boundary,
∂n(r, t) n(r, t) 1 + αref
= β = λ cos θ , (8.67)
∂r r0 β 1 − αref
where λ is the mean free path, and αref is the reflection coefficient of the species.
β is defined as a distance between r0 and the point where the tangential line
of n intercepts the r-axis (see Figure 8.8). The sticking coefficient st on the
boundary surface is frequently used instead of the reflecton αref in the field
of thin film deposition. Then,
st = 1 − αref (8.68)
√
A value 2/3 or 1/ 3 is taken for cos θ [6, 7].
plasma surface
n
∼λ
r = r0
FIGURE 8.8
Boundary condition of Chantry and Phelps and the particle density close to the surface
conditions in time:
iπ
(v, r, t) = v, r, t + i = 1 or 2. (8.69)
ω
The series of structures on a wafer will come into contact with a plasma having
a spatial periodicity defined as
where the self-bias voltage Vself (t) is given by the incident fluxes of electrons,
and positive and negative ions, e , p , and n , from the plasma to the metallic
surface S, and the displacement current density J dis ,
where
t
1
Q(t)|sur = e{ p (r, t) − e (r, t) − n (r, t) + J dis (r, t)} d Sdt. (8.73)
−∞ S e
V(t)
Vrf (t)
Vp (t)
Vrf
Vsus Vp
0 t
Cb
Vdc
electrode
Vsus (t)
FIGURE 8.9
Potentials at points around an electrode with voltage source Vrf .
Usually, Vself (t) is temporally modulated and DC self-bias voltage Vdc is ex-
perimentally defined as
T T
1 1
Vdc = Vsus (t)dt = −Vself (t)dt. (8.74)
T 0 T 0
Exercise 8.3.1
There is a trench structure with a finite surface conductivity σ on an SiO2
wafer. The trench is exposed by fluxes of electrons and positive ions from a
plasma. When the surface etching and deposition are negligible, derive the
equation of the local number density of the surface charge.
By using the local surface fluxes incident from the plasma, e (r, t) and p (r, t),
a simple continuity equation of electrons and positive ions at the surface can
be given as
∂ne
+ dive (r, t) = −div[je surf (r, t)/e] − Rr ne n p , (8.76)
∂t
∂n p
+ div p (r, t) = −Rr ne n p , (8.77)
∂t
where je surf (r, t) = σ (r, t)Esuf (r, t) is the surface electron current density, and
Rr is the surface recombination coefficient. The Equations 8.76 and 8.77 are
simultaneously solved with Poisson’s equation. A self-consistent solution is
obtained with respect to the potential of the metallic electrode behind the
wafer. The Laplace equation must, of course, be used inside the dielectric
SiO2 .
plasma surface
Γp
Yetch
Γn
Ydepo
FIGURE 8.10
Boundary conditions with charge exchange and mass transfer.
Φ(r, t) = d
d Φ(r, t) > 0
Φ(r, t) = 0
t
Φ(r, t) < 0
Φ(r, t+∆t) = 0
t+∆t
FIGURE 8.11
The moving boundary in material processing.
Note that the local surface reaction rate as the local surface velocity Rreact (r, t)
is a negative value in etching and sputtering (see Equations 8.80 and 8.82)
and a positive value in deposition (see Equation 8.81).
An approach to the surface evolution based on the numerical solution of
the Hamilton–Jacobi Equation 8.83 is named the Level Set method. The Level
Set method was developed in [8] and is robust in two- and three-dimensional
evolution problems.
References
1. Ferziger, J.H. and Penc, M. 1996. Computational Methods for Fluid Dynamics. Berlin:
Springer Verlag.
2. Chua, O.L. and Lin, P.M. 1975. Computer-Aided Analysis of Electronic Circuit.
Englewood Cliffs, NJ: Prentice-Hall.
3. Sharfetter, D.L. and Gummel, H.K. 1969. IEEE Trans. on Electron Devices ED-16:64.
4. Takewaki, H., Nishiguchi, A., and Yabe, T. 1985. J. Comput. Phys. 61:261–268.
Nakamura, T. and Yabe, T. 1999. Comput. Phys. Commun. 120:125–154.
5. Ventzek, P.L.G., Hoekstra, R.T., and Kushner, M.J. 1994. J. Vac. Sci. Technol.
B12:461.
6. Chantry, P.J. 1987. J. Appl. Phys. 62:1141.
7. Phelps, A.V. 1990. J. Res. Natl. Inst. Stand. Technol. 95:407.
8. Osher, S.J. and Sethian, J.A. 1988. J. Comput. Phys. 79:129. Sethian, J.A. and
Strain, J. 1992. J. Comput. Phys. 98:231.
through a blocking capacitor Cb. Here, V0 and ω are the amplitude and angular
frequency, ω(= 2π f ), of the rf power supply, respectively. The other electrode
as well as the side metallic wall is grounded to the earth.
When an rf voltage Vrf (t) is applied to the parallel plate reactor in a
vacuum, the current IT (t) between electrodes leads the applied rf voltage
waveform Vrf (t) by π/2. The current is known as the displacement current or
charging/discharging current between the two electrodes, and the external
source has no power dissipation to the reactor; that is, V rf (t) · I T (t) = 0.
255
sheath
Cb
bulk
plasma
V0 sinωt
FIGURE 9.1
A typical CCP reactor.
t
1
Vsus (t) = Vrf (t) − IT (t)dt, (9.2)
Cb −∞
where the second term on the right-hand side is the voltage drop between
Cb to ensure a net direct current (DC) of zero in a periodic steady state. As a
result, the sustaining voltage has a finite negative bias voltage expressed by
the second term on the right-hand side of Equation 9.2.
Capacitively coupled rf plasma is generally classified into two regions
by the phase-shift between the total current IT (t) and the sustaining voltage
Vsus (t). Low-frequency CCP is defined simply when the ions in the plasma
satisfy the relation
eff 1
Vd > d, (9.3)
2f
eff
where Vd is the effective drift velocity of positive ions between parallel plates
with distance d. That is, at low-frequency plasma, the ions can follow the local
field change in time, and the current-sustaining voltage characteristics are
resistive without phase difference. Note that the sheath behaves resistively at
low-frequency CCP.
1000
SF6/N2(10%)
Minimum sustaining voltage 800
CH4(10%)/H2
600 H2
(Vpp) [V]
400 Ar
He
O2
200
0
10k 100k 1M 10M 100M
Frequency [Hz]
FIGURE 9.2
Minimum sustaining voltage (experimental) in a quasi-symmetric parallel plate CCP as a function
of applied frequency at 1 Torr and d = 2 cm.
eff 1
Vd d. (9.4)
2f
At high-frequency CCP, the total current IT (t) has a finite phase lead with
respect to the sustaining voltage waveform Vsus (t), and the sheath behaves
capacitively. Usually there exists a boundary between the low- and high-
frequency plasmas at several MHz of the external frequency (see Figure 9.2),
and the sustaining mechanism of the discharge plasma is completely different
between the two regions.
Exercise 9.2.1
Discuss the maximum-power transfer theorem in the CCP system sustained
by a sinusoidal voltage in Equation 9.1 in Figure 9.1.
The circuit current I T (t) in the series circuit with plasma impedance Z and
external variable impedance z is given by V rf (t)/(z + Z), and the average
power dissipated in the discharge plasma, Pa v , is expressed by
T
1 1 2 Re(Z)
Pav = V sus (t) · I T (t)dt = V , (9.5)
T 0 2 0 |z + Z|2
where T is the period of the external rf voltage source. The maximum power
dissipation to the discharge plasma is obtained at the condition of Z = z
(impedance matching). That is, the impedance of the discharge plasma Z
must be matched to the conjugate of the power supply, z (maximum-power
transfer theorem).
(a) (b)
Λj (1011cm-3s-1)
5.6 240 W 10 4
V
Current [mA]
2.8 120 5 2
Power [W]
Voltage [V]
0.0 0 0 0
20 0 5 10
I
D Time [µs]
is -120 -5 -2
ta 10
nc 10
e 5 -240 -10 -4
(m
m 0 (µs)
) 0 T e
im
FIGURE 9.3
Typical example of the net excitation rate (a) and voltage-current characteristics (b) during one
period in a low-frequency CCP at 100 kHz at 1 Torr in H2 .
3 (a)
(b)
Ion flux
2
0
0 200 400 600 800 1000
Energy (eV)
FIGURE 9.4
Time-averaged ion energy distribution incident on the powered electrode at a low-pressure rf
plasma. The (a) low- and (b) high-frequency cases are shown.
field is formed between the two ion sheaths. The whole system is completely
controlled by Poisson’s field.
The total current in the sheath, which consists mainly of a conduction com-
ponent of positive ions and electrons, coincides with the sustaining voltage
waveform Vsus (t) without phase difference. Even in the case where the sus-
taining voltage is a sinusoidal waveform, the total current is nonsinusoidal in
time, and the plasma potential behaves nonsinusoidally. Under the condition
of a low-frequency CCP in which ions are influenced by the instantaneous
sheath field, the energy of ions incident on the electrode through the sheath
is characterized by a saddle-shaped profile, that is, bimodal distribution (see
Figure 9.4). The energy dispersion between the maximum and minimum en-
ergy peaks decreases as the frequency of the rf source increases to the high-
frequency region.
The high degree of loss of ions from the bulk plasma to both electrodes or
the side wall during a half-period of the low-frequency plasma exhibits some
peculiar characteristics:
i. Formation of low-density plasma with a strong and thick sheath;
ii. Production of high-energy ions incident on the powered electrode;
and
iii. A considerably high value of the sustaining voltage.
PROBLEM 9.2.1
A parallel plates reactor is connected to a low-frequency voltage source by a matching
network consisting of inductance L m . Discuss the DC self-bias voltage at the powered
electrode.
the earth as well as the grounded chamber wall. Under these circumstances,
the asymmetry of the discharge plasma is enhanced as there exists a large
difference in the effective area between the two electrodes.
Exercise 9.2.2
Consider the pair of asymmetric parallel plate electrodes with the surface
areas AP and AG shown in Figure 9.1. Estimate the ratio of each of the drops
in sheath potential in front of both electrodes, Vsh P and VshG , in the case of a
space-charge-limited regime at low frequency.
In a space-charge-limited regime, the ion current density incident on the
sheath with width of d is given by the Child–Langmuir expression,
3/2
K o Vsh
Jp = 1/2
, (9.6)
Mp d 2
where K o is constant, and Mp is the mass of the ion. Provided that the ion
current is dominant in the sheath, the ion currents incident on both electrodes
during one period are equal to each other; that is, J p P AP = J pG AG . Assuming
that the time-averaged potential drops in the sheaths with capacitances C P
and C G are related to the total charge Q flowing to each of the sheaths as
(a) (b)
4.1 60 V W 24 1.2
Λj (1011cm-3s-1)
Current [mA]
Power [W]
Voltage [V]
2.1 30 12 0.6
0.0 0 37 74 -0.6
0 0
20 Time (ns)
D 0
is -30 -12
ta 10
nc 74
e I -1.2
(m 37 -60 -24
m 0 (ns)
) 0 T e
im
FIGURE 9.5
A typical example of the net excitation rate (a) and voltage-current characteristics (b) during one
period of a high-frequency CCP at 13.56 MHz in H2 .
in addition to ions will gradually enter into their trap in space. Due to the
presence of electron trapping inside the reactor, an apparent field of ambipo-
lar diffusion will drop in front of the powered electrode in a time-averaged
fashion. This will introduce a reduction of the minimum sustaining voltages
as shown in Figure 9.2. The boundary between the low- and high-frequency
plasma under conventional external conditions will appear at approximately
several MHz.
When a sinusoidal high-frequency voltage is supplied at the powered
electrode, the total current in the sheath, consisting of the predominant dis-
placement current and the conduction of electrons and ions, has a sinu-
soidal waveform. The capacitive sheath results in a sinusoidal variation of
the plasma potential. The high-frequency plasma has several proper charac-
teristics:
i. Formation of high-density plasma with a weak thin sheath;
ii. A relatively low sustaining voltage;
iii. Production of low-energy ions incident on an electrode; and
iv. A sinusoidal current waveform IT (t) leading the sustaining voltage
waveform Vsus (t) is formed, and the discharge is defined as capaci-
tive.
The electron charge flowing into the blocking capacitance Cb through a
small powered electrode during the positive potential should be equal to
the positive ions during the rest of one period in a periodic steady state.
The great difference between the mass of the electron and that of the ion will
cause an excess negative charge in the capacitor during one period. Therefore,
a negative bias voltage Vdc to the small electrode is needed in order to keep
the zero net DC current in a periodic steady state through the Cb . That is, the
surface of the powered electrode is negatively biased as
TABLE 9.1
Classification of Sheaths Formed in Discharge Plasmas
Sheath Circumstances Position Formed Physical Law
Negative Ion
Here, −Vdc (t) is the negative self-bias voltage with time variation. The time-
averaged DC self-bias voltage,
2π
1
Vdc = Vdc (t) = Vsus (t) d(ωt), (9.10)
2π 0
PROBLEM 9.2.2
At low pressure a collisionless sheath with thickness dsh , bulk electrons with average
velocity < v > diffusing to the plasma sheath boundary will interact electrically
with the moving boundary having an velocity V sh (t). This is not a binary collision
between the electron and the neutral molecule but a wavelike interaction resulting
from the long-range Coulomb interaction. Provided that the interaction is perfectly
elastic under dsh /v > 2π/ω, the reflected electrons will change their velocity from
n p ∼ (ne + nn ). (9.12)
Fully negative ion plasma is performed under Hen = 1, though the plasma is
not maintained without electrons. In an electronegative plasma with densities
ne , n p , nn , and N, the macroscopic measure of the plasma is characterized by
both the electronegativity Hen and the degree of ionization Hdi given by
np
Hdi = . (9.14)
N
Let us consider again the CCP system consisting of light electrons and
massive positive ions as the components of the charged particle. The system
TABLE 9.2
Negative Ions in Processing Plasmas
Feed Gas Negative Ion Material Processing
(a) (b)
Λj[SF5+](1016cm-3s-1)
2.8
Power density (Wcm-2)
10 74 V
st
an
(ns)
0 0 Time
(m
m
)
FIGURE 9.6
Typical example of net excitation rate in electronegative high-frequency CCP in SF6 /N2 (10%) at
1 Torr.
electrode
p-ion sheath
np ~ (ne + nn)
0
bulk plasma
FIGURE 9.7
Double layer formation in front of the instantaneous anode in an rf electronegative plasma.
attachment, because the peak of the attachment cross section is several times
higher than the the ground state O2 (3
g− ), and the threshold energy stands at
2.76 eV (see Section 4.8.2).
Exercise 9.2.3
In an rf electronegative plasma at a range of Torr, a negative charge-layer is
formed, facing the electrode in the instantaneous anode phase (see Figure 9.7).
Discuss the negative layer formation (it is usually called double layer forma-
tion).
Minority electrons in an rf plasma with Hen ∼ 1 vibrate between the two
electrodes. At the instantaneous anode phase, the total charge in front of the
anode, ρ = n p + nn + ne , is negative due to the flow of electrons from the
bulk plasma to the sheath, though the sheath is positive except for the anodic
phase.
PROBLEM 9.2.3
In an electronegative plasma with electronegativity Hen ∼ 1, negative ion density at
low power is the rate-limiting step between the production by an electron attachment
to the feed gas molecule and the loss by a recombination with positive ion. Then,
derive the simple relation between nn and ne as
1/2
Ra ne
nn N , (9.15)
Rr
where Ra and Rr are the electron attachment rate and recombination rate, respectively.
Plasma
0 100 cm 0 6 cm
(b)
Electrode voltage (V)
80
Numerical
40
Experimental
0 20 40 60 80 100
Position (cm)
FIGURE 9.9
Surface potential of a powered electrode exposed to a VHF plasma at 100 MHz at 13 Pa in Ar.
Power is supplied at z = 0, and z = 100 cm is the open end.
TABLE 9.3
Time-Averaged Power Deposition
Feed Gas Active Sheath Bulk Plasma Passive Sheath
of the element. Then, the control of the power deposition in space and in time
is of first importance to the technology. See Table 9.3.
PROBLEM 9.2.4
Scaling lows of internal plasma parameters (plasma density, sheath width, etc.) are
prepared with respect to the external parameters (frequency, amplitude, pressure,
etc.) in a simplified condition. Discuss the relationship between plasma density and
driving frequency in a high-density plasma at a sufficiently low pressure.
Exercise 9.2.4
There exists an external electromagnetic wave with frequency ω, which is
lower than the plasma (electron) frequency ω < ω pe ; then, it is difficult for
the external electromagnetic wave to penetrate the plasma. Discuss that this
does not mean none of plasma production/maintenance.
By applying the rf wave with frequency ω and amplitude V0 through an
antenna to a neutral gas molecule, the interaction with the gas molecule will
cause the electron multiplication when we increase the V0 . A periodic steady
state plasma will be formed through an active sheath in front of the antenna.
Then, inside of the formed bulk plasma a cut-off density nc (= ω2pe m
e −2 ) will
exist.
Drive (VHF)
symmetric
axis
metallic wall
electrode
z wafer insulator
Bias (LF)
FIGURE 9.10
Schematics of a 2f-CCP reactor for SiO2 etching.
(a)
VHF (100 MHz)
6.1E+16 6.7E+16
0.0 0.0
LF (1 MHz)
300 V 700 V
(b)
VHF (100 MHz)
6.1E+16 7.3E+16
0.0 0.0
RF (10 MHz)
300 V 700 V
FIGURE 9.11
Example of the functional separation in a 2f-CCP in CF4 (5%)/Ar at 50 mTorr in terms of net
ionization rate.
The values of the negative self-bias voltage Vdc at each of the electrodes
are considerably different, with the value being very high at LF and very low
at VHF. That is,
i. A high Vdc at the bias electrode is capable of generating the beamlike
ions with an anisotropic velocity distribution at the surface;
ii. High-energy secondary electrons produced by the high-energy ion
bombardment are not confined by a very low Vdc at the opposite
VHF electrode; and
iii. The contribution of the fast secondary electrons to an excess de-
gree of dissociation is significantly reduced in the 2f-CCP at low
pressure.
In particular, high Vdc is realized at the biased electrode without distur-
bance of the plasma profile in the VHF/LF-CCP system. A narrow gap 2f-CCP
is operated by three potentials, Vsur at the powered and biased electrodes, and
a wall potential usually grounded to the earth. The plasma potential in the
system is kept considerably high. One example of the sheath potential drop in
front of the wafer and the surrounding wall is shown in Figure 9.12 as a func-
tion of radial position. The sheath potential drop is defined as the difference
between the plasma potential and the surface potential at the surface. The po-
tential drop is strongly distorted between the wafer edge and the grounded
metallic wall. The phase of the maximum potential drop depends on the ra-
dial position. On the SiO2 wafer the maximum value appears at ωt = 3π/2
when the wafer surface has the lowest potential, and on the grounded metallic
side wall the maximum value appears at ωt= π/2 when the plasma potential
is highest. The magnitude and the energy of charged particles incident on the
surface are influenced by the temporal change of the radial field caused by
the edge characteristics of the potential drop.
The velocity distribution of ions incident on the wafer gives critical infor-
mation for the reactive ion etching of chemically active plasma. There are a
number of investigations of the influence of the bias frequency on the incident
ion energy distribution. It is known that at high-frequency regions the ions
can not relax their energy to the temporal change of the wafer potential; the
energy distribution exhibits a stational profile. However, at low-frequency,
ions follow the instantaneous potential change at the wafer, and a strong time
modulation arises on the energy distribution with a maximum and minimum
at low pressure (see Figures 9.4 and 9.12). Note that the radial characteristics of
the velocity distribution of the incident ions are the key to realizing a uniform
etch process on a wafer. The vicinity of the wafer edge is strongly influenced
by the potential difference between the wafer and the outer grounded wall,
whereas the central part of the wafer keeps the intrinsic characteristics of the
ion velocity distribution as a function of sheath potential drop.
In the other system of 2f-CCP, both of the power sources are supplied to
the same electrode and the second electrode is grounded. This system is less
functional for the material etching device.
1.0
4.0
ΓAr+
0.0 0.0
0.0 2.0 4.0 6.0 8.0 10.0
400 (b)
V plasma
200 Ez Ez
Potential (V)
V surface
0
FIGURE 9.12
Time-averaged plasma density and flux (a) and potential (b) in a 2f-CCP maintained under
conditions of Figure 9.11a.
Λj (1015 cm s )
6.8
-3 -1
3.4
0.0
20
10.4
10
z
(m
m) 0
0 )
t (µs
500 kHz
5.6
Λj (1014 cm s )
-3 -1
2.8
0.0
20
20
10
z
(m
m) 0 10.4 )
t (µs
500 kHz
0
Electrode potential (V)
0
-3
-200
-400 Bias
-600
0 10 20
Time (µs)
FIGURE 9.13
Pulsed-plasma source operation in 2f-CCP with VHF (100 MHz) and LF (1 MHz) at 25 mTorr in
CF4 /Ar: (a) net excitation rate during on phase and (b) during off phase and (c) driving voltage
on both electrodes.
References
1. von Engel, A. 1983. Electric Plasmas Their Nature & Uses, London: Taylor &
Francis.
2. Hirshi, M.N. and Oskam, H.J. Eds., 1978. Gaseous Electronics, Vol. 1. New York:
Academic Press.
3. Satake, K., Yamakoshi, H., and Noda, M. 2004. Plasma Sources Sci. Technol. 13:436–
445.
4. Ohmori, T., Goto, T., and Makabe, T. 2004, 2003. J. Phys. D. 37:2223, and Appl.
Phys. Lett. 83:4637.
275
(a) z (b) z
I sinωt
I sinωt
0 r 0 r
insulator insulator
wafer wafer
FIGURE 10.1
A typical example of ICP (a) and TCP (b) reactors.
In the E-mode the power deposition in the discharge is low, and electron
density is low whereas the mean electron energy is high with a nonequilib-
rium distribution. In the H-mode high power is dissipated with a low-power
reflection. High density of plasma with a lower mean energy of electrons is
sustained in the H mode. The differences of the electron energy diminish at
low pressures due to the nonlocal transport of electrons.
Figure 10.3 shows the computerized tomography (CT) images of the three-
dimensional number density of the excited Ar(2p1 ) in pure Ar at 300 mTorr
and 13.56 MHz in E-mode (a) and H-mode (b) [2]. Both images are scanned
close to the coil plane. These figures demonstrate that in the E-mode the
coupling of energy into the plasma is capacitive and a strong ion sheath is
105
(II)
300 mTorr
Net ionization rate (arb.)
104
15 mTorr
103
102 50 mTorr
(I)
100 mTorr
10
5.0 10.0 15.0 20.0
Amplitude of coil current (A)
FIGURE 10.2
Net excitation rate of Ar(2p1 ) as a function of coil-current amplitude for pressures of 15 mTorr
(), 50 mTorr (), 100 mTorr (×), and 300 mTorr (◦) in ICP at 13.56 MHz in Ar in the reactor
shown in Figure 10.1a.
1.6 5.7
Nj (106 cm-3)
Nj (107 cm-3)
0.8 2.8
0.0 0.0
-60 60 -60 60
x (m 0 0 )
m x (m 0 0 )
m
m) 60 -60 y (m m) 60 -60 y (m
FIGURE 10.3
Emission CT images of short-lived excited atom Ar(2p1 ) in Ar. External conditions are 300 mTorr,
50 sccm, and amplitude 10 A of coil current at 13.56 MHz. (a) E mode (7 W) and (b) H mode
(30 W) correspond to (I) and (II) in Figure 10.2.
formed. The voltage on the coil is divided between the capacitances of the
sheath and the glass wall, so the transfer of power is azimuthally nonuni-
form. As a result, a local peak appears near the powered terminal. At lower
pressure, the nonlocal behavior of electrons distributes the excitation over the
entire radius and the azimuthal anisotropy is lost.
PROBLEM 10.2.1
Prove that the ionization event under E × B fields leads by π/4 with respect to that
in the E θ -drift motion of electrons.
Exercise 10.2.1
Discuss the measurement method of the power deposition in an ICP sustained
by an external coil current.
When the coil is supplied from an rf source with variable power and an
appropriate matching box, the power deposited into the ICP is determined
by monitoring the coil current I (t) and subtracting Reff I (t)2 from the power
W(t) measured by the power meter, that is, W(t)− Reff I (t)2 . Here, the effective
impedance Reff is determined from the dissipated power W (t) in a vacuum
with no plasma in the same reactor at the same external coil current I (t). It is
assumed that the phase difference between W (t) and I (t) is negligible.
TABLE 10.1
Collision and Reaction Processes of Metastable Ar*(3 P2 ,3 P0 ) in Ar
Process Collision/Reaction Rate Coefficient(cm3 s−1 )
direct ionization in Ar is
si ksi ne N∗ ksi N∗
= = , (10.5)
di kdi ne Ng kdi Ng
and for the rate of the metastable pooling to the direct ionization,
2
mp kmp N∗ N∗ kmp N∗ Ng
= = . (10.6)
di kdi ne Ng kdi Ng ne
(a) (b)
4.1 4.6
Λdi (1016 cm-3 s-1)
2.0 2.3
0.0 50 0.0 50
0 0
25 25
25 25
50 0 50 0
m ) m)
r (m z (m r (m z (m
m) m )
coil coil
FIGURE 10.4
Two-dimensional net ionization rate in ICP at 13.56 MHz and 50 A at 50 mTorr in Ar.
1.0
15 mTorr
0.8
0.6 50 mTorr
Λsi + Λdi
Λdi
0.4
0.0
0.0 0.5 1.0 1.5 2.0 2.5
Power (kW)
FIGURE 10.5
Contribution of direct ionization to the maintenance of ICP at 13.56 MHz in Ar as a function of
dissipated power.
(a) (b)
4.4 1.8
N* (1011 cm-3)
N* (1011 cm-3)
2.2 0.9
0.0 34 0.0 35
0 17 0 20
25 26
50 0 52 5
r (m m) r (m m)
m) z (m m) z (m
coil coil
FIGURE 10.6
Time-averaged number density of Ar*(r, z). (a) numerical value at 386 W (0.49 Wcm−3 ) and (b)
experimetal value at 400 W (0.25 Wcm−3 ). The other external conditions are 15 mTorr in Ar at
13.56 MHz.
excited states. The number density of the metastables is typically 1011 cm−3
under the balance between the production and loss by electron quenching
and diffusion to the wall. In a high-density plasma with ne ∼ 1011 cm−3 , the
effect of ionization of the metastables by a low-energy electron impact be-
comes less important under a strong electron quenching. On the other hand,
under a low-power and low-density plasma condition, metastable species
make a great contribution to the plasma maintenance through a low-energy
electron impact ionization (stepwise ionization) or metastable-metastable col-
lision (metastable pooling).
Exercise 10.2.2
An Ar atom has two metastable states, Ar(3 P2 ,3 P0 ). Investigate the density
ratio of the two metastables in an ICP in pure Ar.
The steady-state value after production depends on the electron quenching.
The value is 4 × 10−7 cm3 s−1 for Ar(3 P2 )–Ar(3 P0 ) and 1.5 × 10−7 cm3 s−1
for Ar(3 P0 )–Ar(1 P1 ), respectively. As a result, the density ratio N[Ar (3 P2 )]/
N[Ar (3 P0 )] is measured at 100 W as 5.7 at 50 mTorr and 7.3 at 5 mTorr, respec-
tively.
6.0 15 mTorr
10 × N*
100 mTorr
ne
Number density (1012 cm-3)
4.0
5 × ne
30
Potential (V)
2.0
20
Vp 10
0 0
0 10 20 30 40 50
FIGURE 10.7
Typical plasma density and potential distribution at coil plane in ICP sustained at 13.56 MHz in
Ar: (a) 15 mTorr and 50 A, (b) 100 mTorr and 30 A.
PROBLEM 10.2.2
The plasma potential in an ICP is usually 10 V to 20 V and the ambipolar diffusion
of electrons to the reactor wall controls the plasma density in the case where there is
no extinction of electrons in the bulk plasma. Estimate the time constant of electron
loss at 50 mTorr in Ar.
Exercise 10.2.3
Estimate the typical Larmor frequency ω L in an ICP and compare the value
with the rf frequency ω and the total collision rate Rt .
The typical ICP in Figure 10.1 is driven at 100 mTorr at 13.56 MHz at a current
amplitude of 30 A. Then, the Lamor frequency (cyclotron frequency) ω L of
the electron at 1 cm inside the glass wall is
On the other hand, the external frequency ω is 8.5 × 107 s−1 , and RT is usually
108 s−1 . The relation among these three frequencies is ω L ∼ ω ∼ RT .
Here, the conduction of massive ions and the displacement current are ne-
glected in an rf wave. The wave propagation is described by the Maxewll
electromagnetic
wave (ω)
ω < ωpe
δ
plasma (ωpe)
electromagnetic
ω > ωpe wave (ω)
FIGURE 10.8
Propagation, cut-off, and decay of an electromagnetic wave in a uniform plasma.
1 ∂
rotB(t) = µJe (t) + E(t), (10.9)
c 2 ∂t
∂
rotE(t) = − B(t). (10.10)
∂t
If we take the rotation of Equation 10.10 and substitute it into Equation 10.9,
we derive the wave equation
ω2
∇ 2 E(t) + E(t) = jωµJe (t). (10.11)
c2
When the wave propagates in the z-direction in the plasma, the wave equation
is divided into the z- and r -directions:
d 2 E z (t) ω2
+ 2 E z (t) = jωµJ ez (t), (10.12)
dz2 c
ω2
Er (t) = jωµJ er (t). (10.13)
c2
We eliminate the conduction current density J e (t) in Equation 10.12 by using
Equation 10.8, and then, the axial component of the wave E z (t) satisfies
d2 ω2 ω2pe Rm ω2pe
E z (t) = − 1− 2 −j 2 E z (t), (10.14)
dz2 c2 ω + Rm
2 ω + Rm ω
2
where ω pe is the plasma electron frequency, (e 2 ne /m0 )1/2 . The solution has
the form
10 Rm / ω = 103
Skin depth (cm)
10-2
1.0
10-1 102
10
10-1
1.0
FIGURE 10.9
Skin depth of electromagnetic waves incident on a plasma.
√ 2 2 1/2 −1/2
1 2c ω pe
2
ω pe Rm
2
ω pe
2
δ= = 1− 2 + − 1−
2
,
ki ω ω + Rm 2 ω + Rm ω
2 2 ω + Rm
2
(10.17)
where δ is shown in Figure 10.9 as a function of ω pe /ω and Rm /ω. The skin
depth is a concept describing the propagation of the electromagnetic wave
in a plasma and not for activation of the plasma production of neutral gases.
See Table 10.2.
TABLE 10.2
Typical Skin Depth as a Function of ω, ω pe , and Rm
Plasma Region ω and ωpe Skin Depth
c
Collisionless (Rm ω) ω pe
ω ∼ ω pe
Rm ∼ ω ω pe
ω ∼ √ 2 c
( 2+1)1/2 ω pe
Rm ∼ ω ω pe ∼ ω ∼ √ 2 c
( 2−1)1/2 ω pe
√
2c
Rm 1/2
Collisional (Rm
ω) ∼ ω pe ω
and E θ is given by
3.83r
E θ (r, z, t) = E 0 J 1 exp[ j (ωt − kz)], (10.19)
r0
where J 1 is the first term of the Bessel function of the first kind and r0 is the
radius of the cylindrical reactor. k is the complex wavenumber and is obtained
from
3.83 2 ω2 ω2pe ω(ω + j Rm )
− − k2 + 2 = 2 .
r0 c c ω2 + Rm2
where b is
2 2
c 3.83 ω 2 2 2
b= 1+ − 1 + Rm /ω + 1.
ω pe r0 c
Exercise 10.3.1
There is a collisionless plasma that has a linear density distribution n(z) = n0 z.
Discuss the profile of the wave propagation as a function of z.
In a collisionless plasma (Rm ω) the dispersion relation is given by
c2k2 ω2pe
ξ2 ≡ 2 =1− 2 . (10.21)
ω ω
As shown in Figure 10.10, the propagation properties of electromagnetic
waves are classified into three regions with different modes as a function
of ne (z). An incident wave propagates from left to right. As the plasma fre-
quency ω pe is proportional to ne (z)1/2 , the wavelength of the electromagnetic
wave 2π/k becomes longer, and at ω pe = ω the wavelength is infinity. That
is, the wave is almost stationary in space and oscillates with ω in time. In the
region of ω pe
ω, k is complex, and the wave is attenuated with the atten-
uation constant Im{k}. The frequency at ω = ω pe gives the cut-off frequency.
The position at ω pe = ω acts as the reflective wall, and the wave is reflected
to the left. Here, ξ is known as the refractive index.
n (z) = n0 z
ωpe (z) ∝ √z
reflection wave
attenuated wave
FIGURE 10.10
Propagation of electromagnetic waves into a plasma with a spatial profile n0 z.
Exercise 10.3.2
When the intensity of electromagnetic waves is low enough not to perturb
the bulk plasma, the plasma density is measured by using the principle that
a wave propagating through a plasma has a phase shift relative to the wave
propagating in a vacuum. Practically, a microwave is used in order to obtain a
spatial resolution of ∼cm [3]. Derive a simple relation between the phase-shift
φ and and the plasma density ne .
The phase shift between two path φ(rad) is given as a function of the electron
plasma frequency, ω pe (z) = (e 2 ne (z)/m0 )1/2 :
1/2
ω pe (z)2
φ = (k0 − kplasma )dl = k0 1− 1− dz. (10.22)
l l ω2
Then,
φ = 2.82 × 10−17 λ0 ne (z)dz. (10.23)
l
In particular, when the plasma is uniform in the radial direction of the ICP
reactor, we obtain the plasma density ne (cm−3 ):
φ
ne = 1.18 f , (10.24)
L
where f (Hz) is the microwave frequency and L(cm) is the effective width of
the plasma investigated.
PROBLEM 10.3.1
The earth is surrounded with the ionospheric layer of plasma density, ∼106 cm−3 .
A shortwave broadcasting (frequency: ∼MHz-30 MHz) is operated on the ground
station. A satellite-based broadcasting is serviced at frequency greater than GHz.
Discuss the difference of the available broadcasting-frequency.
References
1. Ventzek, P.L.G., Hoekstra, R.J., and Kushner, M.J. 1994. J. Vac. Sci Technol. B
12:461.
2. Miyoshi, Y., Petrovic, Z.Lj., and Makabe, T. 2002, 2005. IEEE Trans. on Plasma
Sci. 30:130. Miyoshi, Y., Miyauchi, M., Oguni, A., and Makabe, T. IEEE Trans. on
Plasma Sci. 33:362.
3. Auciello, O. and Flamm, D.L. 1989. Plasma Diagnostics, Vol. 1. San Diego:
Academic Press.
Magnetron plasmas and electron cyclotron resonance (ECR) plasma are used
in dry plasma processing at low pressure in which an external permanent
magnetic field makes a significant contribution to the plasma maintenance
by reducing electron losses to walls and electrodes. Helicon wave plasma
and magnetic neutral loop discharge (NLD) are also magnetized plasmas.
Collisionless electron heating through an electron cyclotron or a helicon wave
is the main mechanism to sustain these plasmas at very low pressure. On the
other hand, an induced magnetic field operates on an inductively coupled
plasma (ICP) and surface wave plasma (SWP).
289
TABLE 11.1
Magnetron Plasma for Sputtering
Type DC Magnetron rf Magnetron
Exercise 11.1.1
Both the electric E(=Ek) and the magnetic field B(=Bj) are applied in a gas at
a number density of N. Derive the neutral density N estimated by apparent
electron changes to
2 1/2
eB 1
N 1+ 2
. (11.1)
m Rm
Equation 11.4 means that the gas number density changes from N to the above
expression 11.1. A decisive effect of a magnetic field on plasma production
is expressed under the condition of the electron cyclotron frequency (Larmor
frequency) ωce (= eB/m) Rm .
z
target
r
N S N
DC
yoke
(b)
40 129
Br (Gauss)
Bz (Gauss)
-150 -241
-340 -611
30 30
z( z(
m m
m N 88 m N
88
) 0 ) 0
0S 0S
r (mm) r (mm)
FIGURE 11.1
A typical DC magnetron plasma reactor describing a magnetic field line (a); radial and axial
components of B(z, r ) (b).
(a)
3.5 ne
0.0
30
z
(m
m 88
) N
0
0S
r (mm)
ne
np
(b)
5
Potential (V)
-98
-200
30
z
(m
m N 88
) 0
0S
r (mm)
FIGURE 11.2
Plasma density (a) and potential (b) distributions in a DC magnetron in Ar at 5 mTorr driven at
200 V in the reactor in Figure 11.1.
Figure 11.1a. A thin sheath region with a strong potential difference appears
in front of the target, and the potential in the bulk plasma (plasma poten-
tial) is very low when the plasma is surrounded by the metallic reactor wall
grounded to the earth. In these potential distributions, electrons are mainly
produced by the collisional ionization at the radially localized region with
Bz ∼ 0 in the sheath edge, and these electrons diffuse to the bulk plasma in
the very low electric and magnetic fields. On the other hand, ions produced
locally as the pair of the ionization are strongly accelerated to the target with
a beamlike energy, and these ions sputter the target material. It is easy to
estimate a local erosion profile by ion sputtering from the local ion flux inci-
dent on a target. Magnetron plasma is usually operated in a region of current
source. Removal of the external magnetic field, after the magnetron plasma
is formed, introduces the discharge plasma with extinction.
PROBLEM 11.1.1
Discuss the reasons why maximum ionization efficiency is realized at the point where
the magnetic field crosses at a right angle with the electric field in front of the target
(cathode) in a magnetron discharge.
PROBLEM 11.1.2
An azimuthal drift current in the plasma ring in front of the target is on the order of
few amperes. Estimate the magnetic field generated by the loop current of electrons
and compare the value with the external permanent magnetic field.
FIGURE 11.3
Comparison between a balanced and unbalanced magnetron (a) and (b).
ωt = 0 ωt = π/2
1.2 1.2
ne [1010cm-3]
ne [1010cm-3]
0.6 0.6
0.0 0.0
S 8.0 S 8.0
4.0 4.0 4.0 4.0
N N
r[cm] 8.0 0.0 z[cm] r[cm] 8.0 0.0 z[cm]
ωt = 3π/2 ωt = π
1.2 1.2
ne [1010cm-3]
ne [1010cm-3]
0.6 0.6
0.0 0.0
S 8.0 S 8.0
4.0 4.0 4.0 4.0
N N
r[cm] 8.0 0.0 z[cm] r[cm] 8.0 0.0 z[cm]
FIGURE 11.4
Typical electron density distribution in an rf magnetron sustained in Ar at 5 mTorr at 13.56 MHz
and V0 = 400 V.
(a)
0
-1
cm s )
-3 -1
-2
15
Dielectric target
Ions flux(10
-3
Metallic target
-4 PB
Target
-5
0 20 40 60 80
r (mm)
(b)
PB
Depth (arb.)
-1
Target
0 20 40 60 80
r (mm)
FIGURE 11.5
Ion flux incident on Cu target (a) and erosion profile (b) in an rf magnetron plasma.
S
Er
B
N
θp
FIGURE 11.6
Magnetically confined plasma reactor. Arrangement of multipolar magnets and net ionization
rate.
The magnetic mirror effect near the cusp of two magnetic poles suppresses
the electron diffusion in the direction normal to the magnetic field line in ad-
dition to the electron reflection in the wall sheath on the reactor. A set of
straight magnets is arranged in the axial direction on the outer surface of the
cylindrical reactor. The resulting line cusp magnetic field near the chamber
wall confines electrons and ions. Such a field configuration has the proper-
ties of a magnetic mirror and has a very small effect on the confinement of
slow electrons. That is, with increasing the number, the electron confinement
becomes stronger, and the escape of electrons from the loss cone becomes
significant [2]. The electron mean energy rises near the wall due to the pres-
ence of the high-energy electrons trapped by the multipolar magnetic field.
The net ionization rate is expected to have a local peak. The local electron
energy increases when the magnitude of the magnetic field increases due to
the increase of the trapped high-energy electron. There is an optimum for the
number of the magnetic poles. The magnetic field has a very small effect on
the confinement of slow electrons. At pressure range where the electron mean
free path is less than the reactor dimension, the magnetic field has little effect
on the electron confinement by the trapping.
(a) (b)
microwave quartz
electron-cyclotron wave guide tube
wave
quartz window
antenna rf power
coil
B
coil
parmanent
magnet
ECR layer
B
plasma
wafer plasma
wafer
rf bias rf bias
FIGURE 11.7
Representative for magnetically resonant plasma. ECR plasma source (a) and helicon source (b).
PROBLEM 11.5.1
Calculate the magnitude of the magnetic field at the ECR condition of electrons in a
collisionless plasma supplied by the microwave power at 2.45 GHz.
References
1. Kuroiwa, S., Mine, T., Yagisawa, T., and Makabe, T. 2005. J. Vac. Sci. Technol. B,
23:2218.
2. Takekida, H. and Nanbu, K. 2004. J. Phys. D (37:1800).
3. Boyd, T.J.M. and Sanderson, J.J. 2003. The Physics of Plasmas. Cambridge:
Cambridge University Press.
4. Sturrock, P.A. 1994. Plasma Physics. Cambridge: Cambridge University Press.
5. Nicholson, D.R. 1983. Introduction to Plasma Theory. New York: John Wiley &
Sons.
6. Lieberman, M.A., and Lichtenberg, A.J. 2005. Principles of Plasma Discharges
and Materials Processing (2nd Edition). Hoboken: John Wiley & Sons.
12.1 Introduction
Plasma processes (i.e., sputtering, deposition, etching, surface treatment, etc.)
require information about the surface reactions of active species and their
probability, as well as information on gas-phase collision/reaction processes
and their cross sections. Physical or chemical quantities describing the plasma
surface interaction are given by an effective surface reaction probability con-
sisting of the sticking coefficient and the yield for etching or deposition.
Historically, the collision cross section of the electron or ion in the gas
phase has been continuously accumulated theoretically and experimentally
as a function of the impact energy, in addition to the study of the quan-
tum characteristics of mono- and polyatomic molecules in the field of atomic
and molecular physics. In the same way, we expect that the surface reaction
process and the probability of neutral molecules and ions with a material
surface will be rapidly elucidated and accumulated by using first-principle
molecular dynamics and measurements.
301
Substrate
St
deposition
+
M , εp
sputtered
impact
Y(ε)
Target erosion
FIGURE 12.1
Physical sputtering system.
Exercise 12.2.1
A zero-dimensional model of a reactive sputtering process is successfully
used for compound film deposition in a DC magnetron sputtering with a
metallic target in admixture with reactive gas and Ar. Derive the governing
equations of the system in which a tantalum (Ta) target is sputtered in an
Ar/O2 mixture to deposit Ta2 O5 compound on a substrate [4].
The reaction on the surface in this system is divided into three regions: the tar-
get; substrate; and side wall with areas of AT , A S , and AW (see Figure 12.3a).
A fraction (1 − X) of the target material sputtered by the Ar+ ions with flux
p deposits on the substrate, and X arrives at the side wall. T , S , and W
give the degree of the compound (Ta2 O5 ) formation on the target, substrate,
and reactor wall, respectively. The values of T and S are a practical indi-
cator of the reactive sputtering from pure metallic target ( T = 0) resulting
in deposition of a stoichiometric compound film at the substrate ( S = 1).
st , ss , and sw are the sticking coefficients of O2 to the metallic part of the
target, substrate, and wall, respectively. The sticking of the sputtered parti-
cles on the target is assumed to be unity. YTm and YTc are the sputtering yields
of the metal and compound material at the target. The number ratio of the
reactive atom (O) between the reactive gas (O2 ) and the compound molecule
(Ta2 O5 ) is denoted by h r .
At the target surface, the balance equation of the compound is
∂
T (t)AT = h r O2 st (1 − T )AT − p YTc T AT . (12.1)
∂t
(a)
Incident ion energy (eV)
100 300 500
101
1
Sputtering yield Y(ε)
+
Ar / Cu
10-1 3.0
Y(250 eV; θ)
Y(ε; θ)
2.0
10-2
Y(100 eV; θ) 1.0
10-3 0
0 30 60 90
Incident angle (deg.)
(b)
Incident ion energy (eV)
0 50 100 150 200 250
1.2
10 eV
1.0
15 eV
Sticking probability
0.8
Sputtering yield
25 eV
0.6
+
Cu / Cu
0.4
35 eV
0.2 Y(ε)
50 eV 75 eV
0
100 eV
0 20 40 60 80 100
Incident angle (deg.)
FIGURE 12.2
Sputtering yield Y(ε) and Y(ε, θ ) of Ar+ -Cu target (a), and Y(ε) of Cu+ –Cu target and sticking
coefficient St(θ ) of Cu on Cu substrate (b).
∂
S (t)AS = b m h r O2 ss (1 − S )AS + (1 − S )b m (1 − X) p YTc T AT
∂t
− S (1 − X) p YTm (1 − T )AT , (12.2)
(a) Substrate : AS
(1-ΘS) ΘS
SS
ΘW
Wall : AW
reactive gas O2
(1-ΘW)
positive ion SW
ΓO2
Γp
ST
YTm YTc
(1-ΘT) ΘT
Target : AT
(b) 1
Deposition rate (arb.)
0.5
0
0 1 2
Reactive gas flow (arb.)
FIGURE 12.3
The 0th-order reactive sputtering model (a), and the predicted deposition rate of a compound
film as a function of partial pressure (b).
∂
W (t)AW = b m h r O2 sw (1 − W )AW + b m (1 − W )X p YTc T AT
∂t
− W X p YTm (1 − T )AT . (12.3)
The total supply q of the reactive gas into the reactor is equal to the consump-
tion on the surface and the quantity pumped out to the outside. That is,
AT
R depo = p {YTc T + YTm (1 − T )}(1 − X). (12.5)
AS
The deposition rate R depo in the reactive sputtering has a range with multival-
ues as a function of reactive gas flow, as shown in Figure 12.3b.
atoms removed
Y( p ; target) = , (12.6)
incident ion
for the system between an incident ion with energy p and target material.
The databases are widely available in the literature [1, 2]. The number of
target atoms ejected by the sputtering of ions with velocity v incident on the
target surface dS from the magnetron plasma is estimated by the velocity
distribution of the incident ion, g p ( p , θ), as
ρ dSdl = n p Y( p ; target)vg p ( p , θ, r)dθ d p dt dS, (12.7)
θ p
where ρ is the atomic number density of the target material, dl is the erosion
depth during a small time dt, and g p ( p , θ, r) is normalized to unity as
g p ( p , θ, r)dθd p = 1. (12.8)
p
The erosion depth profile of the target is estimated numerically by the time
development of the ion sputtering in a magnetron plasma.
PROBLEM 12.2.1
Even in a low-pressure magnetron plasma, ions affect a target with a finite angle of
incidence. Then, we must consider the angular dependence of the sputtering yield
Y(ε, θ ). Revise Equation 12.9 of the sputter rate by considering Y(ε, θ).
On the other hand, the energy spectrum of the sputtered atom is estimated
by the Thompson formula [5],
PROBLEM 12.2.2
Discuss the reason that the angular distribution of the ejected neutral is usually
approximated by the cos θ -law in the physical sputtering process.
The Thompson formula is derived under the condition that the sputtered
atoms come from a well-developed collision cascade in a material. The colli-
sion cascade is realized in a system between an incident heavy ion and light
atom in the material. In a system between a light incident ion with low energy
and a target consisting of a massive atom, however, the ion from the sheath
will be easily backscattered by a massive target atom and will knock off an
atom in the top layer of the target. Then, the sputtering is caused by a single
knock-on mechanism. A modified formula has been proposed for the system
between a light ion and massive target atom [6]:
γ (1 − γ ) p 2
f s (; p , rt ) = A ln , (12.13)
( + Ut )α+1 + Ut
where α = 3/5 for the H+ –Fe system. The angular distribution of the ejected
atom is described by cosine law.
We consider the arrival flux of the sputtered atom at the substrate rs . First,
the flux without collision in gas phase is given by
Γs (rs ; ) =
0
Γs (rt )exp(−NQl)dS, (12.14)
St
transport of the relaxed component Nsrel by collision with feed gas molecules.
The neutral transport is described by the diffusion equation
∂ rel
N (r, t) = Ds ∇ 2 Nsrel (r, t) + rel
s (r, t), (12.15)
∂t s
where Ds is the diffusion coefficient of the sputtered atom in the feed gas
molecule. The production rate of the relaxed component rel s is given under
the assumption that the sputtered atom is randomized in energy and direction
after one collision with feed gas molecules. Then, from the difference of the
ejected flux distribution at small distance (l, l + dl),
rel
s (r) = −∇ · s (r).
0
(12.16)
In a steady-state (∂/∂t = 0), the spatial transport of the sputtered atom under
relaxation is obtained by
The arrival flux of the sputtered atom, collisionally relaxed in the gas phase,
is given by
srel (rs ) = −Ds ∇ · Nsrel (rs ). (12.18)
As a result, the total flux arriving at the substrate at rs is
Figure 12.4 shows the spatial density distribution of Ar+ ions and sput-
tered Cu neutrals in a DC magnetron sputtering system operated between
r
N S N
FIGURE 12.4
Relations of the distribution between the ion density n p (r) and sputtered atom Ns (r) in a DC
magnetron plasma.
TABLE 12.1
Chemical Vapor Deposition of SiH4 (+H2 )
Deposited Film Process Radicals Conditions Ref.
a-Si:H, µc-Si:H Plasma process H, Si, SiH2 , SiH3 400 K, 1–103 Pa [7]
Poly-Si Thermal process SiH2 , H2 , 900 K, 10–103 Pa —
a-Si:H, µc-Si:H Catalytic process H, Si 2000 K, 0.5–10 Pa [8]
a-Si:H Photo process — Room temperature —
and wear resistant. The head and disk surfaces in magnetic disk drives are
usually coated with a thin layer of DLC to minimize wear and corrosion.
Thin films deposited by plasma polymerization (surface grafting) have
a large variety of applications. The plasma has a high ability to modify the
surface. These materials are widely used as biocompatible films with hy-
drophilicity and are suitable for various biomedical and pharmaceutical mate-
rials, for example, optical lenses, implants, and drug delivery devices. Plasma
polymerization is usually performed in a glass reactor excited by helical coils
driven at 13.56 MHz (i.e., inductively) in carrier gas Ar at several hundred
mTorr with monomer vapor. Excitation of the carrier gas takes place in radio-
frequency (rf) Ar plasma, generating active species (ions and radicals) in
the gas phase. These active species interact with the substrate and generate
reactive sites (mainly free radicals) on the wafer surface. The monomer va-
pors in the plasma chamber readily react with these radicals, yielding grafted
surfaces.
Exercise 12.3.1
Diamond thin film grows on a diamond substrate in a thermal plasma in
CH4 /H2 . The surface under deposition is divided into two sites: the diamond
lattice growth with probability D and nth layers of amorphous carbon with
probability Cn . Atomic H and C, dissociated from the feed gases CH4 /H2 , in
thermal plasma diffuse to the substrate surface. The thin film growth is highly
selective on the site and highly competitive between deposition and etching
and is modeled by the birth and death process in the stochastic process [9–11].
That is, on the substrate surface, the evolution equation is given in the form
of the 0th-order simultaneous rate equations
dD
= kc H C1 − ks C D + ketch H C1 (12.20)
dt
dC1
= −kc H C1 + ks C D − ketch H C1 − ks C C1 + ketch H C2 (12.21)
dt
dCn
= ks C Cn−1 − ketch H Cn − ks C Cn + ketch H Cn+1 (n > 1), (12.22)
dt
where ks is the sticking coefficient of the C atom. ketch is the etching probability
of a -C:H by an H atom. kc is the conversion probability of a -C:H to diamond
lattice by an H atom, and
∞
D+ Cn ≡ 1. (12.23)
n=1
D= ,
1 + (1 − χ ) kketch
c
(1 − χ )χ
C1 = ,
1 + (1 − χ ) kketch
D
(1 − χ n )χ
Cn = . (12.26)
1 + (1 − χ ) kketch
C
ks (1 − χ ) kketch
c
c
R depo ∼ kc H C1 = . (12.27)
1 + (1 − χ ) kketch
c
At kc ketch ,
ks (1 − χ )C
R depo (χ ) = ∼ kc C . (12.28)
(1 − χ ) + kketch
c
At kc ketch ,
kc
R depo (χ ) = ks (1 − χ ) C . (12.29)
ketch
TABLE 12.2
Plasma Etching of Materials
Materials Feed Gas Property
substrate. More specifically, the etching yield of ions is a function of the impact
energy ε p and the incident angle θ , Y(ε p , θ). The etching rate R etch (r) of ions
with velocity distribution g p (v, r) incident on the substrate r is related to the
etching yield Y(ε p , θ ) when we define R etch (r) as
n p (r)
R etch (r) = Y(ε p , θ)vg p (v, r)dv/ g p (v, r)dv, (12.30)
ρn
where n p (r) is the number density of ions close to the wafer. ρn is the atomic
number density of the substrate material.
Reactive plasma
resist resist
Passivation layer
(SiOClx) Passivation layer
(CFx-Polymer)
Mixed layer
poly-Si SiO2
(a) (b)
FIGURE 12.5
Typical etching profile in Si-ULSI: Si-gate etching (a) and trench or hole etching of SiO2 (b).
PROBLEM 12.4.1
Reactive ion etching (RIE) is based on a first adsorption of radical species on a
surface and successive ion impact with high energy. Discuss the necessary condition
of radicals and ions incident on the surface in the RIE.
where Vplasma and Te are the time-averaged potential and electron temperature
in the bulk plasma. Mp and m are the mass of the positive ion and the electron,
respectively.
TABLE 12.3
Reactive Product in Plasma Etching and the Boiling Point
and Vapor Pressure
Materials Reactive Product Boiling Point (C) Vapor Pressure
Al AlF3 1291 —
AlCl3 183* 0.6
AlBr3 255* —
Si SiF4 –90 ≤ 760
SiCl4 57 —
SiBr4 5.4 —
Fe FeF2 ≤ 1000 —
FeCl3 319 —
FeBr2 684 —
Ni NiCl2 1000 —
Ni(CO)4 -25 — —
W WF6 17* —
WCl6 346* —
Co CoF2 1200 —
CoCl2 1050 —
∗ Sublimation temperature.
iii. It is preferable that the feed gas, the dissociated molecules, and the
reactive products are all nontoxic and all have low global warming
potentials.
Admixture gases are usually used for plasma etching for the following
purposes:
The influence of each of the gases on surface etching is very complicated. Most
of the feed gases used for plasma etching are greenhouse gases. That is, green-
house gases absorb infrared radiation and trap energy in the atmosphere. The
atmospheric lifetime and global warming potential (GWP) characterize the
effect of the greenhouse gas. The GWP of a greenhouse gas is the ratio of
global warming from one unit mass of a greenhouse gas to that of one unit
mass of CO2 over a period of time. Hence this is a measure of the potential
for global warming per unit mass relative to CO2 (see Table 12.4).
TABLE 12.4
Global Warming Potential (GWP) and
Lifetime of Gases for Plasma Etching
Feed Gas GWP (Units of 500 yr) Lifetime (yr)
at room temperature:
4F + Si → SiF4 . (12.34)
PROBLEM 12.4.2
The volume density of neutral etching products is simply estimated without redepo-
sition in a plasma as
1 R etch ρ Seff
Netch = , (12.36)
V R j ne + kpump
where Seff is the effective area of a substrate exposed to etching, V is the plasma
volume, R j is the rate of the destruction processes of the species in gas phase, and
kpump is the pumping speed. Derive the expression 12.36.
(a)
6 +
4 Cl2 /Cl2
1 Ar+/Cl2
8
6
4
Ar+/ Ar
2
0.1
8
6
4
0 500 1000
Incident ion energy (eV)
(b)
2.0 +
CF3
Etching Yield (SiO2 /ion)
1.5 +
CF2
1.0 +
CF
Ar+
0.5
F+
0.0
0 500 1000 1500 2000
Ion energy (eV)
FIGURE 12.6
Etching yield of Si (a) and SiO2 (b) as a function of ions from plasmas.
ions and neutral radicals incident on the surface and passive ejected flux of
neutral etch products can be, in principle, extracted from a self-consistent
modeling of the total system of the dry etching. However, it may be difficult
to perform the modeling due to the lack of an available database.
12.4.4 Al Etching
Al etching is generally performed by chlorine and the compounds. However,
fluorine compound is not valid for this purpose, because the reactive etching
product has a high boiling point and is usually nonvolatile (see Table 12.3).
The surface etching is described by the direct reaction between Cl2 and Al:
As a result, the etching speed depends on the gas number density of Cl2 . On
the other hand, in BCl3 the plasma density controls the etching speed because
the surface reaction proceeds by way of the dissociated Cl2 in BCl3 plasma:
That is, the etching speed in rf plasma in Cl2 is independent of the dissipated
power, and the speed is proportional to the power in BCl3 plasma. This means
that ion bombardment from the plasma has no effect on the etching rate, and
isotropic etching is performed. The native surface oxide is removed by the
ion impact in BCl3 plasma.
PROBLEM 12.4.3
Discuss the effects of additive gases (Ar, O2 , CO, H2 ) on SiO2 etching in fluorocarbon
plasma (see Table 12.5).
2.0
N2 /H2 plasma
1.5
Sputtering yield (C / ion)
N2 plasma
1.0
Ar plasma
0.5
H2 plasma
0.0
0 400 800 1200
Ion energy (eV)
FIGURE 12.7
Etching yield of organic low-k as a function of ions from plasmas.
Here, R etch is the etching rate (speed function) of a material surface. The sur-
face as a function of Cartesian coordinates (x, z) and time t in Figure 12.9 is
TABLE 12.5
Function of Additive Gas in SiO2 Etching in Fluorocarbon
Plasma
Gas Property
1.0
0.0
0
p
+20
0
1500 -20
En
erg
y(
eV (deg)
) Angle
FIGURE 12.8
Example of the energy and angular distributions of ions incident on a wafer as a function of
radial position.
z
ion trajectory
Φ=0 Mp
Φ<0 Φ>0
l rp-1 s
r0
n (1-s)
rp m Gas
Solid
FIGURE 12.9
Schematic diagram of the Level Set method and the surface evolution.
defined as
(x, z, t) = 0. (12.40)
That is, the gas and solid phase are given at > 0 and < 0 at t, respec-
tively. The dielectric surface (r, t) = 0 is usually locally charged by ions and
electrons incident on the surface during etching, and the metallic surface is
kept at equipotential.
A brief description of the Level Set method is given below. First, we in-
vestigate the relation between the time constant τ0 for one monolayer etching
and the time step to trace an ion trajectory ttr in order to estimate a surface
evolution t under the following physical requirement:
When the ion with energy ε p is reflected at the material surface having a
reflection coefficient αref (ε p ), that is, when the condition
ξ ≤ αref (ε p )
is satisfied, the trace of the ion trajectory in gas phase is continued at Step 1.
Here, ξ is a uniform random number distributed over [0, 1].
In the case of ξ > αref (ε p ), the information of the flux velocity Γ(v0 , r0 ) at
the surface is transcribed into the adjacent two-grid nodes (see Figure 12.9):
1/2
−1
s 2 + (1 − r )2
Γ(v)i+1, j = 1 + Γ(v0 , r0 ),
r 2 + (1 − s)2
1/2
−1
r 2 + (1 − s)2
Γ(v)i, j+1 = 1 + Γ(v0 , r0 ). (12.43)
s 2 + (1 − r )2
Then, the incident angle of the transcribed flux vector θ with respect to the
normal vector n = ∇/|∇| at both nodes is obtained by (r) and the deriva-
tive of the nodes:
∂
−1 ∂x
· vx + ∂
∂y
· vy
θ = cos . (12.44)
∇φ
·
v
Then we have the component of the etching yield at the node as a function
of the ion energy ε p and incident angle θ as Y(ε p , θ). The above procedure
is carried out for the number of ions at initial position rintl by considering
g(v, rinitl ) and n p (rintl ). Then, the flux velocity Γ(v, r) at each node adjacent to
the surface = 0 is accumulated during t.
Step 3: After tracing the ion trajectories incident on a trench structure from
a plasma during t = m ttr (m ∼ 100), the accumulated flux velocity (ε p , r),
energy ε p , and angle θ at each of the grid nodes adjacent to the surface = 0
enables us to estimate the new surface = 0. That is, at each of the nodes,
the etching rate (speed function) is given as
R etch (r) = n pintl vg p (v, r)Y(ε p , θ)dε p dθ/ vg p (v, r)dε p
= Γ(ε p , r)Y(ε p , θ)d ε p d θ. (12.45)
FIGURE 12.10
A typical example of feature profile evolution by plasma etching: (a) without charging and
deposition, (b) with charging and without deposition, and (c) with charging and deposition.
PROBLEM 12.4.4
Discuss the string model of the feature profile evolution in etching in comparison
with the Level Set method.
ratio. The etching needs a high etch rate with anisotropy and selectivity to the
mask material. Fluorocarbon plasma, maintained in CF4 , C4 F8 , SF6 , and so
on, in admixture with O2 are used for the processes under the basic reactions
that F isotropically etches Si wafer with rapid chemical etching, and a fluo-
rinated silicon oxide passivating layer is formed on the sidewall to keep the
anisotropic etching.
A series of alternating processes of etching and depositing is effective in
forming MEMS construction. The time-multiplexed deep etching is known
as the Bosch process. One of the effective and rapid plasma processes consists
of three steps: isotropic etching of Si/polymer in SF6 , polymer passivation in
C4 F8 , and polymer depassivation in O2 . The resulting feature profile keeps a
high anisotropy, though the etching is chemically isotropic. The Plasma Bosch
process is widely carried out by using a high-density ICP reactor at pressure
between 1.33 Pa (10 mTorr) and 13.3 Pa (100 mTorr). The dimension of the
MEMS structure is usually comparable with the sheath thickness or larger.
It means that the bulk plasma and sheath structure exposed to the substrate
will change in time (plasma molding).
Exercise 12.4.1
In a plasma for deposition or etching, a large size polyatomic molecule may
grow up in circumstances of high degree of dissociation of the feed gas. The
generated species is a named particle, dust, or powder [15]. Discuss the tool
for detection of these large polyatomic molecules in a plasma.
The plasma potential is always a positive value during the cw operation.
Therefore, positive ions and neutrals are detected through an orifice on the
electrode or reactor wall. However, it is not easy to detect negative ions from
the electrode. Technically, a pulsed operation of the plasma is introduced to
measure negative ions in the afterglow phase. Mie scattering is an in situ,
active procedure (see Table 12.6).
TABLE 12.6
Detection of Polyatomic Molecule or Cluster in Plasmas
Method Molecular Size Mass Resolution Ref.
Wafer e
e
surface Wafer
Open area surface Resist
Resist
M+
SiO2
Poly-Si Tunneling
current
SiO 2
Metal
Damage
Si current SiO2
Damage
Si current
Cb
Cb
FIGURE 12.11
325
Schematic diagram of the local charging damage during plasma etching: gate-Si etching (a) and SiO2 trench etching (b).
plasma
quasi-conductive
layer
Γe
Γe
jesurf
Γp
SiO2 jesurf
charge
Si
FIGURE 12.12
Local surface charges and conduction in the trench (hole) of SiO2 during etching.
where Γe and Γ p are the instantaneous local fluxes of electrons and positive
ions incident on the surface, respectively. je sur f (r, t) = σ (r, t)Esur f (r, t) is a sur-
face current density of the electron conduction, σ is the electrical conductivity
of the CFx thin polymer, and R r is the surface recombination coefficient. The
time evolution of the local surface potential Vsur f (r, t) is solved at t in the sys-
tem including the continuity Equations 12.47 and 12.48 and Poisson’s equa-
tion under consideration of the incident fluxes of electrons and ions from the
plasma through the sheath as described in the previous section.
The resulting potential distribution across the gas, surface, and solid phase
is calculated by solving
ρs (r, t)
∇ 2 (r, t) = − : surface of SiO2
ε0 εr
= 0 : inside of SiO2 (12.49)
n p (r, t) − ne (r, t)
= −e : in gases,
ε0
where ρs is the surface charge density on SiO2 , and ε0 and ε0 εr are the
permittivities of a vacuum and SiO2 , respectively. That is, the overall po-
tential from the gas phase to the bulk SiO2 is simultaneously solved by
changing the mesh size in gas and solid. Then, the distribution of the sur-
face charge at the boundary between SiO2 and lower poly-Si (or metal) in
Figure 12.12 is iteratively calculated under the principle that the equipo-
tential of the surface of lower poly-Si (or metal) must be maintained. The
origin of the local charging of electrons and positive ions arises from the
significantly different velocity distribution. In a steady state, the velocity dis-
tribution of ions incident on a wafer with a beamlike component is quite
different from an isotropic distribution of electrons in the positive ion sheath,
and both of the charged fluxes incident on a flat surface exposed to the
plasma have the same magnitude in a time-averaged fashion. The great dif-
ference of the velocity distribution leads to a local accumulation of electrons
at the upper part of the trench (hole) and ions at the lower part and bottom,
that is, charging on the inside wall of a hole or trench with a high aspect
ratio.
Figure 12.11 demonstrates the schematic diagram of the charging damage
in the gate-Si etching by Cl2 plasma (a) and SiO2 trench etching by CF4 /Ar
plasma (b). Both of the chargings seriously damage the profile and lower-level
device elements (i.e., the thin gate etc.).
Exercise 12.4.2
Discuss the relationship among the time constants of charging, radical depo-
sition, and ion etching in a typical SiO2 etching (see Figure 12.13.)
Typical fluxes of radicals and ions incident on a wafer are 1018 cm−2 s−1 and
1016 cm−2 s−1 , respectively. Also, the time constant for local charging is usually
ms and the time for effective monolayer etching needs 100 ms.
1.6
AR = 2.0
0.8
10.0
0.4
ion flux
electron flux
0.0
0.0 1.0 2.0 3.0 4.0 5.0
Time (ms)
FIGURE 12.13
Time constant for local charging as a function of aspect ratio in typical trench etching of SiO2 .
e2 Vox Vox 1
J DT = 2
φB − exp −4π dox 2em∗DT φ B − , (12.50)
2π hdox 2 2 h
4π 2m∗FNT φ B
3/2
e 3 Vox2
J FNT = 2 φ
exp− , (12.51)
8π hdox B 3π eh
where dox is the thickness of the oxide film, and Vox is the applied voltage
between both sides of the thin oxide. h is the Planck constant, φ B is the barrier
height, and m∗ is the effective mass of the electron. Plasma current damage
is typical in a metal gate etching as shown in Figure 12.11a, and besides, in
a multilayer interconnect system we take care of the current damage during
trench or via etching of SiO2 (see Figure 12.14).
Γe Γp
resist
V
SiO2
metal
I
poly-Si
gate oxide LOCOS
0 5 ms 10 ms
Si
FIGURE 12.14
Current damage at the gate oxide during plasma etching.
References
1. Behrisch, R., Ed. 1981. Sputtering by Particle Bombardment I. Topics in Appl. Phys.
47. Berlin: Springer Verlag.
2. Behrisch, R., Ed. 1983. Sputtering by Particle Bombardment II. Topics in Appl. Phys.
52. Berlin: Springer Verlag.
3. Kress, J.D., Hansen, D.E., Voter, A.F., Liu, C.L., Liu, X.-Y., and Coronell, D.G.
1999. J. Vac. Sci. Technol. A 17:2819.
4. Engelmark, F., Westlinder, J., Nyberg, T., and Berg, S. 2003. J. Vac. Sci. Technol. A
21:1981.
5. Thompson, M.W. 1968. Philos. Mag. 18:377.
6. Kenmotsu, T., Yamamura, Y., and Ono, T. 2004. Jpn. Soc. Plasma. Sci. Nucl. Fusion
Res.
7. Matsuda, A. 2004. Thin-film silicon (Inv. Review paper), Jpn. J. Appl. Phys.
43:7909.
8. Matsumura, H., Umemoto, H., and Masuda, A. 2004. J. Non-Cryst. Solids
19:338–340.
9. van Kampen, N.G. 1981. Stochastic Processes in Physics and Chemistry. Amsterdam:
North-Holland.
10. Gardiner, C.W. 1983. Handbook of Stochastic Methods for Physics, Chemistry and the
Natural Sciences. Berlin: Springer Verlag.
11. Ford, I.J. 1995. J. Appl. Phys. 78:510.
12. Balooch, M., Moalem, M., Wang, W-E., and Hamza, A.V. 1996. J. Vac. Sci.
Technol. A 14:229.
13. Sethian, J.A. and Strain, J. 1992, 1993. J. Comp. Phys. 98: 231; Sethian, J.A.
and Chopp, D.L. 1993, J. Comp. Phys. 106:77.
14. Esashi, M., and Ono, T. 2005. J. Phys. D, Topical Review 38:R223.
15. Bouchoule, A., Ed. 1999. Dusty Plasmas. Chichester: John Wiley & Sons.
16. Auciello, O. and Flamm, D.L. 1989. Plasma Diagnostics, Vol. 1. San Diego:
Academic Press.
17. Bruno, G., Capezzuto, P., and Madan, A. 1995. Plasma Diagnostics of Amorphous
Silicon-Based Materials. San Diego: Academic Press.
18. Saito, N., Koyama, K., and Tanimoto, M. 2003. Jpn. J. Appl. Phys. 42:Part 1, 5306.
19. Bohren, C.F., and Huffman, D.R. 1983. Absorption and Scattering of Light by Small
Particles. New York: John Wiley & Sons.
20. Hirose, M. 1996. Mater. Sci. Eng. 41:35.