Textbook
Textbook
1
UDC 000000
BBC 00000
E00
A. I. Pushkarev
E00 Physics and Chemistry of Low-Temperature Plasma for Material
Treatment/ A. I. Pushkarev, Yu. I. Isakova, R. V. Sazonov, G. E. Kholod-
naya; Tomsk Polytechnic University. – Tomsk: TPU Publishing House,
2011. – 181 p.
The book describes some physical and technical aspects of plasma chemistry. The
general concepts of plasma physics and chemical kinetics as well as the methods and
sources for generation of nonequilibrium plasma are considered. The application of low-
temperature plasmas in the chemical industry include: the synthesis and decomposition
of inorganic compounds; synthesis of nanosized materials; conversion of hydrocarbon
gases; synthesis of hydrogen and supramolecular carbon clusters (fullerenes); treatment
of membranes and coatings, etc.
This book is recommended for students, studding on the specialty "Physics and
technology of low temperature plasmas, plasma chemistry and plasma technology"
course 140200 " Electroenergetics".
UDC 000000
BBC 00000
2
TABLE OF CONTENTS
Preface ..............................................................................................................7
Introduction .....................................................................................................8
Сhapter 1 Introduction to plasma physics .................................................12
1.1 What is a plasma ....................................................................................12
1.2 How are plasmas made ..........................................................................13
1.3 Some applications of plasmas................................................................13
1.4 Plasma parameters .................................................................................14
1.4.1 The Degree of Ionization .................................................................16
1.4.2 Plasma Temperature ........................................................................17
1.4.3 Plasma Frequency ............................................................................20
1.4.4 Debye Shielding ..............................................................................21
1.5 The plasma parameter: strongly and weakly coupled plasmas .............22
1.6 Plasma types ..........................................................................................24
1.7 Coalitional processes in plasma .............................................................26
1.7.1 Collision cross sections, mean-free paths and collision frequencies
..................................................................................................................26
1.8 Coronal equilibrium ...............................................................................28
1.9 Penetration of neutrals into plasmas ......................................................33
1.10 Collisions with neutrals and with charged particles: relative
importance ...................................................................................................36
Chapter 2. Introduction to gas discharge physics .....................................38
2.1 What Is the Subject of Gas Discharge Physics? ....................................38
2.2 Typical Discharges in a Constant Electric Field ...................................38
2.3 Brief History of Electric Discharge Research .......................................41
2.4 Mechanism of electrical discharges .......................................................43
2.4.1 Cathode processes ...........................................................................43
2.4.2 Photoelectric emission .....................................................................44
2.4.3 Thermionic emission .......................................................................45
2.4.4 Schottky effect .................................................................................45
2.4.5 Field emission ..................................................................................46
2.5 Electrical breakdown .............................................................................47
2.5.1 Electron avalanche ...........................................................................47
3
2.5.2 Formation of a streamer...................................................................49
2.5.3 Propagation of the streamer discharges ...........................................52
2.6 Electrical breakdown in very small gaps – Townsend’s breakdown
mechanism ...................................................................................................54
2.6.1 Townsend’s experiment...................................................................54
2.6.2 Townsend’s theory of electrical breakdown ...................................56
2.6.3 Townsend’s electrical breakdown criterion ....................................60
2.6.4 Townsend’s mechanism in the presence of electron attachment ....61
2.7 Paschen’s law.........................................................................................62
2.7.1 Physical interpretation of the shape of the Paschen curve ..............64
2.7.2 Validity of Paschen’s law ................................................................65
2.8 Classification of Discharges ..................................................................65
Chapter 3 Basic principles of plasma chemical processes ........................68
3.1 Chemical kinetics...................................................................................69
3.1.1 Rates of reactions ............................................................................69
3.1.2 Dependence of rates on concentration.............................................72
3.1.3 Dependence of Rate on Temperature ..............................................78
3.2 The object and the main features of the plasma chemistry ...................81
3.2.1 Quasi-equilibrium plasma-chemical processes ...............................82
3.2.2 Non-equilibrium plasma-chemical processes..................................83
3.3 Elementary Plasma-Chemical Reactions ...............................................84
3.3.1 Ionization processes .........................................................................85
3.3.2 Classification of Ionization Processes .............................................86
3.3.3 Direct Ionization by Electron Impact ..............................................87
3.3.4 Stepwise ionization by electron impact ...........................................88
3.3.5 Ionization in collisions of heavy particles: adiabatic principle and
Massey parameter .....................................................................................89
3.3.6 Penning ionization effect and associative ionization ......................90
3.4 Elementary plasma-chemical reactions of positive ions .......................91
3.4.1 Different Mechanisms of Electron–Ion Recombination in Plasma.91
3.5 Elementary plasma-chemical reactions involving negative ions ..........92
3.5.1 Dissociative electron attachment to molecules as a major
mechanism of negative ion formation in electronegative molecular gases
..................................................................................................................92
4
3.5.2 Three-body electron attachment and other mechanisms of formation
of negative ions .........................................................................................93
3.6 Special features of plasma-chemical reactions, their dynamics and
kinetics .........................................................................................................95
3.7 Initiation of plasma-chemical processes by electron-beams ...............102
Chapter 4 Chain gas-phase processes in the external action ..................105
4.1 Classification of chain processes .........................................................105
4.1.1 Unbranched chain reactions ..........................................................107
4.1.2 Chain reactions with quadratic branching .....................................108
4.1.3 Chain reactions with degenerate branching...................................108
4.1.4 Electron-induced chain reactions ..................................................109
4.2 Chain chemical processes under external action .................................109
4.2.1 Investigation of the induction period of ignition of the oxygen -
hydrogen mixture ....................................................................................109
4.2.2 Investigation of the displacement of the ignition limits of a
stoichiometric oxygen-hydrogen mixture under external action............115
4.3 Radiation-thermal cracking of methane ..............................................119
4.4 Chain oxidation of methane under external action ..............................122
4.4.1 Oxidation of methane in the equilibrium conditions at a low
pressure ...................................................................................................124
4.4.2 The initiation of methane oxidation by an external influence.......126
4.5 The conversion of carbon disulfide CS2 in the atmospheric air ..........129
Chapter 5. Applied plasma-chemistry ......................................................134
5.1 Nonequilibrium synthesis of nanosized oxide powders ......................134
5.1.1 Comparison of the available processes for production of nanosized
TiO2 particles ..........................................................................................134
5.1.2 An overview of the techniques to synthesize nanosized
(TiO2)x(SiO2)1-x .......................................................................................137
5.1.3 Non-equilibrium plasmochemical synthesis of nanosized particles
of metal oxides........................................................................................140
5.1.4 Research of composite nanosized oxides (TiO2)x(SiO2)1-x
synthesized using a non-equilibrium plasmochemical process ..............146
5.1.5 Synthesis of composite oxides Si-C-Ox.........................................153
5.2. Plasmochemical conversion of methane ............................................156
5
5.2.1 Plasma pyrolysis of methane .........................................................156
5.2.2 Partial oxidation of methane ..........................................................159
Conclusion ...................................................................................................172
References ....................................................................................................176
6
Preface
7
Introduction
8
decomposition of impurities of different compounds (NO, NO2, SO2, CO,
CS2, etc.) in the air by a pulsed electron beam showed that the energy
used to decompose one molecule of the gas is lower than its dissociation
energy. This is due to the fact that some favorable conditions for the oc-
currence of chain processes are formed when an electron beam is in-
volved [1]. At low temperatures, when thermal initiation of the reaction
does not occur, some active centers are formed due to the influence of
plasma. These active centers are free radicals, ions or excited molecules,
which can start a chain reaction. This chain reaction occur at a tempera-
ture of 150-200 degrees below the temperature of a normal thermal pro-
cess, but with the same speed, since the plasma facilitates the most energy
intensive stage - thermal initiation of the reaction. With a sufficient length
of chain an electro physical set-up provides a small part of total energy
consumption to the chemical process. The main source of energy in this
case is the thermal energy of the original gas or exothermic energy of the
chemical reactions of chain process (e.g. oxidation or polymerization). It
is important to note that a chemical process, occurring at a temperature
below the equilibrium, allows one to synthesize compounds which are un-
stable at higher temperatures and the selectivity of the synthesis of which
at high temperatures is low. The increase in the temperature of chain
chemical processes with the radioactive impact is analogues to a catalytic
effect. But the chain process can entirely occur in the gas phase, which
greatly increases the reaction rate in comparison with the heterophase cat-
alytic process. A high speed of the reaction required for industrial tech-
nologies is achieved by the use of branched chain processes. However,
their major drawback is associated with the occurrence of an explosive
process, which significantly increases the risk of production. This disad-
vantage is eliminated by initiating of the chain process outside the scope
of self-ignition under external influence. The above features gas-phase
chemical processes with the electron beam involved indicate the prospects
of their use in large-scale chemical production [2]. Most studies of chain
gas-phase processes, including those under the external action, were per-
formed by Russian scientists, who continued the works of the Nobel lau-
reate N.N. Semenov. In his Nobel speech, he explained that the penetrat-
ing radiation will be used to initiate chain processes.
In plasma-chemical processes the energy transfer from an external
source (a source of electrical energy) to the reacting molecules occur in
several steps (see Fig. 1).
9
Fig. 1. Stages of energy transfer in a gas discharge
10
equilibrium Maxwellian. A non-equilibrium concentration of charged par-
ticles and the particles which are excited by different degrees of freedom -
rotational, vibrational and electronic levels is observed.
The object of plasma chemistry is a low-temperature plasma in mo-
lecular gases. Plasma chemistry studies the kinetics and mechanism of
chemical reactions and physical-chemical processes in low-temperature
plasmas. It should be emphasized that in plasma chemistry, the term
"plasma" is understood as a substance containing not only the charged
particles, but molecules in the ground electronic state with the excess
(non -equilibrium) internal energy.
The object of plasma chemistry as a scientific discipline is a study of
the relationship between physical and chemical phenomena in chemical
reactions in the plasma, the possibility of plasma usage for various prob-
lems in applied chemistry. As a consequence, there is a conventional divi-
sion into theoretical and applied plasma chemistry.
The terms "plasma chemistry", "plasma-chemical reactions" and
"plasma-chemical processes" were introduced in Russian scientific litera-
ture since the publication of the book entitled "The kinetics and thermo-
dynamics of chemical reactions in low-temperature plasmas", edited by
L.S. Polak (Moscow: Nauka, 1965). They reflect the fact that the object
of consideration is the specific chemical objects, which feature lies in the
fact that chemical reactions are initiated or occur in plasma. It has been
more than two hundred years since the existence of electrical discharge
chemistry as a science. It has come a long way from a phenomenological
description of phenomena to the creation of a new research area with its
own apparatus, including theoretical and experimental techniques. The
mechanisms of many plasma-chemical reactions have been studied and
understood. Besides the theoretical aspects of plasma-chemistry, there are
numerous applications of plasma-chemistry for the solution of practical
problems that are grouped under the title "applied plasma chemistry.
11
Сhapter 1 Introduction to plasma physics
12
Irving Langmuir, the Nobel laureate who pioneered the scientific
study of ionized gases, gave this new state of matter the name ‘plasma’.
The reader who is more interested in plasma physics is referred to an
excellent book written by Robert J. Goldston and Paul H. Rutherford [3]
and the lectures of Professor Richard Fitzpatrick (The University of Texas
at Austin) [4].
There are all sorts of uses for plasmas. To give one example, if we
want to make a short-wavelength laser we need to generate a population
inversion in highly excited atomic states. Generally, gas lasers are
‘pumped’ into their lasing states by driving an electric current through the
gas, and using electron-atom collisions to excite the atoms. X-ray lasers
depend on collisional excitation of more energetic states of partially ion-
ized atoms in a plasma. Sometimes a magnetic field is used to hold the
plasma together long enough to create the highly ionized states. A whole
field of ‘plasma chemistry’ exists where the chemical processes that can
be accessed through highly excited atomic states are exploited. Plasma
etching and deposition in semiconductor technology is a very important
related enterprise. Plasmas used for these purposes are sometimes called
‘process plasmas’. Perhaps the most exciting application of plasmas such
is the production of power from thermonuclear fusion.
13
1.4 Plasma parameters
14
where M–mass of the heavy particle;
W–energy of the electron;
me–mass of electron.
For an elastic collision of an electron with an argon atom, the frac-
tion of transferred energy is therefore very small, about
WTr 1
(2)
W 40000
On the other hand, a significant amount of energy is transferred in a
collision between two electrons.
The electrons gain energy through acceleration by the electric field,
which sustains the plasma and transfers that energy by inelastic collisions
with the neutral gas molecules. The inelastic collisions between energetic
electrons and the heavy species of the plasma result in excitation, ioniza-
tion, or dissociation of the target if it is multiatomic. Energy transfer in an
inelastic collision is not controlled by the mass ratio of the colliding parti-
cles. In an inelastic collision between two particles, the fraction of trans-
ferred energy is given by
WTr M
(3)
W min M
where min is the mass of particle losing energy.
According to Eq. (1.3), in an inelastic collision between an electron
and a heavy particle (min = me «M), the electron can transfer almost all its
energy to the heavy particle, creating an energetic plasma species. The in-
elastic collisions therefore sustain the plasma by producing the particles
that form it and giving the plasma its special features. Inelastic collisions
involve energy transfer in amounts that vary from less than 0.1 eV (for
rotational excitation of molecules) to more than 10 eV (for ionization).
Electron-electron collisions can also play a significant role in the en-
ergy transfer processes in the plasma. Their importance depends on the
degree of ionization prevalent in the plasma. For degrees of ionization be-
low 10-10 the contribution of the electron-electron collisions to the energy
transfer is negligible.
However, in electron cyclotron resonance (ECR) plasmas, where the
degree of ionization can be above 10 -3, electron-electron collisions domi-
nate[5].
The relative contribution of each type of collision to the processes
taking place in the plasma depends on additional plasma parameters,
which will be discussed next and which derive from the previously de-
scribed parameters.
15
1.4.1 The Degree of Ionization
The parameter that defines the density of the charged particles in the
plasma is the degree of ionization of the gas. It specifies the fraction of
the particles in the gaseous phase which are ionized. The degree of ioniza-
tion, , is defined as
ni
(4)
n
There are two basic processes of ionization, satisfying the conditions
for conservation of momentum and energy: (a) impact ionization, where
an electron strikes an atom, so that an ion and two electrons come off and
(b) radiative ionization, where a photon with sufficient energy (often in
the ultraviolet range) is absorbed by an atom, dissociating it into an ion
and an electron. Ions can recombine into atoms by the reverse of these
processes: (a) three - body recombination, where two electrons and an ion
join to make a neutral atom plus a free electron; and (b) radiative recom-
bination, where an electron and an ion combine into an atom, and a pho-
ton is emitted. These processes are illustrated in Fig. 3.
16
Table 1. Ranges of Parameters for Various Low-Pressure Plasmas
17
The electrons and the heavy species in the plasma can be considered
approximately as two subsystems, each in its own thermal quasi-
equilibrium.
The ions and electrons in the plasma can therefore be characterized
by their specific different average temperatures: the ion temperature, Ti
and the electron temperature, Te. Actually in some cases additional tem-
peratures may characterize the particles in the plasma. For example, in the
presence of a magnetic field, even a single plasma species, for example,
the ions, is characterized by two different temperatures, one representing
the translation of the ions parallel to the magnetic field, T||, and one repre-
senting the translation perpendicular to the magnetic field, . This is
caused by the fact that the forces acting on the species parallel to the
magnetic field are different from those acting perpendicular to it [7].
The situation is even more complicated, as the heavy species in the
plasma can be characterized by several temperatures at the same time,
even in the absence of a magnetic field [8]: the temperature of the gas, Tg,
which characterizes the translatory energy of the gas; the excitation tem-
perature, Tex, which characterizes the energy of the excited particles in the
plasma; the ionization temperature, Tion; the dissociation temperature, Td,
which characterize the energy of ionization and dissociation; and the radi-
ation temperature, Tr which characterizes the radiation energy. Thermo-
dynamic equilibrium will exist in the plasma only if the following equa-
tion is satisfied:
Tg Tвх Tion Td Tr Tв (6)
Complete thermodynamic equilibrium cannot be achieved in the en-
tir9e plasma because the radiation temperature, Tr at the envelope of the
plasma cannot equal the temperature in the plasma bulk. However, under
certain laboratory conditions, it is possible to achieve local thermodynam-
ic equilibrium in plasma in volumes of order of the mean free path length
(The reader who is unfamiliar with the term is referred to the Sect. 2.1).
If this happens, the plasma is called a local thermodynamic equilibrium
(LTE) plasma. In low pressure plasmas, produced by direct current glow
discharge or radio frequency excitation, the LTE conditions are generally
not achieved. These plasmas are therefore called non-LTE plasmas.
In non-LTE plasmas the temperatures of the heavy particles are
normally too small to promote chemical reactions in thermodynamic equi-
librium. The electron temperature is therefore the most important temper-
ature in non-LTE plasmas, among all those different temperatures men-
tioned previously. The fraction of electrons that will cause the different
reactions in the plasma, the overall efficiency of the plasma processes,
18
and the processing rates increase with increasing electron temperature.
The electron temperature is discussed in further detail in the following
section.
19
tribution can be replaced in low-pressure plasmas by the following as-
sumptions:
1 The electric field strength in the plasma is sufficiently low such
that one can neglect the inelastic collisions, but large enough for the elec-
tron temperature to be much higher than the ion temperature, Tв >> Ti
2 The electric field is of sufficiently low frequency, that is, it is of a
frequency ω much lower than the frequency of collisions .
3 The collision frequency is independent of the electron energy.
20
scales shorter than the distance vtτ p traveled by a typical plasma particle
during a plasma period will also not detect plasma behaviour. In this case,
particles will exit the system before completing a plasma oscillation. This
distance, which is the spatial equivalent to τp, is called the Debye length,
and takes the form
D T (14)
1
m p
Note that
0T
D (15)
nв 2
is independent of mass, and therefore generally comparable for dif-
ferent species. Clearly, our idealized system can only usefully be consid-
ered to be a plasma provided that
D
<<1 (16)
L
p
<<1 (17)
21
where Φ(r) is the electrostatic potential, and n0 and T are constant.
From ei= −ee= e, it is clear that quasi-neutrality requires the equilibrium
potential to be a constant. Suppose that this equilibrium potential is per-
turbed, by an amount δΦ, by a small, localized charge density pext . The
total perturbed charge density is written
p pexp e(ni nв ) pexp 2e 2 n0Ф / T (19)
Thus, Poisson’s equation yields
p pexp 2e n0Ф / T 2
Ф
2 (20)
0
0
which reduces to
2 2 p
2 Ф exp (21)
D 0
If the perturbing charge density actually consists of a point charge q,
located at the origin, so that , δpext=qδ(r) then the solution to the above
equation is written
q 2 r / D
Ф(r ) e (22)
4 0 r
Clearly, the Coulomb potential of the perturbing point charge q is
shielded on distance scales longer than the Debye length by a shielding
cloud of approximate radius consisting of charge of the opposite sign.
Note that the above argument, by treating n as a continuous function,
implicitly assumes that there are many particles in the shielding cloud.
Actually, Debye shielding remains statistically significant, and physical,
in the opposite limit in which the cloud is barely populated. In the latter
case, it is the probability of observing charged particles within a Debye
length of the perturbing charge which is modified.
22
1 2 e2
U (r , v) mv (25)
2 4 0 r
of one charged particle in the electrostatic field of another vanishes. Thus,
U(rc,vt)= 0.
The significance of the ratio rd/rc is readily understood. When this
ratio is small, charged particles are dominated by one another’s electro-
static influence more or less continuously, and their kinetic energies are
small compared to the interaction potential energies. Such plasmas are
termed strongly coupled. On the other hand, when the ratio is large,
strong electrostatic interactions between individual particles are occasion-
al and relatively rare events. A typical particle is electrostatically influ-
enced by all of the other particles within its Debye sphere, but this inter-
action very rarely causes any sudden change in its motion. Such plasmas
are termed weakly coupled. It is possible to describe a weakly coupled
plasma using a standard Fokker-Planck equation (i.e., the same type of
equation as is conventionally used to describe a neutral gas). Understand-
ing the strongly coupled limit is far more difficult, and will not be at-
tempted in this course. Actually, a strongly coupled plasma has more in
common with a liquid than a conventional weakly coupled plasma.
Let us define the plasma parameter
4n3D (26)
This dimensionless parameter is obviously equal to the typical num-
ber of particles contained in a Debye sphere. However, Eqs. (8), (16),
(17), and (19) can be combined to give
D 1 rd 3 2 4 03 / 2 T 3 / 2
( ) 1/ 2 (27)
rC 4 rC в3 n
It can be seen that the case «1, in which the Debye sphere is
sparsely populated, corresponds to a strongly coupled plasma. Likewise,
the case » 1, in which the Debye sphere is densely populated, corre-
sponds to a weakly coupled plasma. It can also be appreciated, from Eq.
(1.28), that strongly coupled plasmas tend to be cold and dense, whereas
weakly coupled plasmas are diffuse and hot. Examples of strongly cou-
pled plasmas include soliddensity laser ablation plasmas, the very “cold”
(i.e., with kinetic temperatures similar to the ionization energy) plasmas
found in “high pressure” arc discharges, and the plasmas which constitute
the atmospheres of collapsed objects such as white dwarfs and neutron
stars. On the other hand, the hot diffuse plasmas typically encountered in
ionospheric physics, astrophysics, nuclear fusion, and space plasma phys-
ics are invariably weakly coupled.
23
Table 2. Lists the key parameters for some typical weakly coupled plas-
mas
24
Fig. 4. Plasma types by electron density and temperature
25
explosion. They have no practical importance because they do not exist in
controlled laboratory conditions.
– Plasmas in local thermodynamic equilibrium (LTE plasmas).
These are plasmas in which all temperatures, except the radiation temper-
ature, Tr, are equal in each small volume of the plasma.
– Plasmas that are not in any local thermodynamic equilibrium – non
– LTE plasmas. These plasmas, also named cold plasmas.
The plasmas produced for research or manufacturing purposes are
either LTE or non – LTE type plasmas, designated in daily use, respec-
tively, as thermal and cold plasmas.
26
changes its identity or internal energy state. In the first case, the electron
may lose any fraction of its initial momentum, depending on the angle at
which it rebounds. The probability of momentum loss can be expressed in
terms of the equivalent cross section that the atoms would have if they
were perfect absorbers of momentum. In the second case, the probability
of ionization, for example, can be expressed in terms of the equivalent
cross section that an atom would have if it were ionized by all electrons
striking within this cross sectional area.
In Fig. 5, electrons are incident upon a thin slab of thickness dx con-
taining n, neutral atoms per unit volume. The atoms are imagined to be
opaque spheres of cross sectional area a: i.e. every time an electron strikes
the area blocked by the atom, either it loses all of its momentum (elastic
collision) or it ionizes the atom (inelastic collision).
The number of atoms per unit area of the slab is nndx and the fraction
of the slab blocked by atoms is nnσdx. If a flux r of electrons is incident
on the slab, the flux emerging on the other side is Г+dГ = Г(1-nσdx), so
that the change of flux r with distance is given by
dГ
nnГ (29)
dx
which has the solution
Г Г 0 exp( nnx) Г 0exp( x / mfp ) (30)
where
27
mfp (nn ) 1 (31)
The quantity Amfp is called the mean-free path for collisions. In a dis-
tance Amfp, the flux would be decreased to 1/e of its initial value. In other
words, an electron travels a distance mfp before it has a reasonable prob-
ability of colliding with an atom. For electrons of velocity , the mean
time between collisions is given by
mfp / v (32)
The ‘collision frequency’, namely the inverse of , is usually defined
in terms of an average over all velocities in the Maxwellian distribution
(which may have different individual collision frequencies), namely
v 1 nn v (nn / n e ) f e (v) (v)vd 2 v (33)
As is implied by this formula, for more complex collisional process-
es than that illustrated in Fig. 4, the cross section σ is often itself a func-
tion of the velocity of the incident particle.
In the case where the collision of the electron with the atom results
in ionization of the atom, we may calculate the rate of production of new
electrons per unit volume simply by multiplying the ionization collision
frequency of the electrons, equation (34), by the electron density, ne. This
‘source rate’ Se of electrons is given by
S в nв nn ion vв (34)
where ion is the cross section for electron-impact ionization and
where we assume that the electron velocities e greatly exceed the neutral
velocities n so that the velocity of impact comes mainly from the elec-
tron’s motion. This cross section is definitely a strong function of electron
velocity, at least below energies of about 30 eV, so the averaging over the
Maxwellian distribution of electrons is necessary. There is, of course, an
equal and opposite ‘sink rate’ for neutral atoms, i.e. neutral atoms are lost
by ionization at the same rate per unit volume, Se.
The dependence of the ionization cross section σion for hydrogen at-
oms on the energy of the bombarding electron is shown in Fig. 6, and the
ionization rate (σion ve) averaged over a Maxwellian distribution of elec-
trons is shown in Fig. 7. The maximum cross section σion reached for elec-
trons with energies somewhat above Ei (the Rydberg ionization energy,
28
which is about 13.6eV for hydrogen) and is in the neighborhood 1020 m2,
the ‘size’ of the hydrogen atom. However, the ionization rate is signifi-
cant even for electron temperatures well below Ei, because a Maxwellian
distribution still contains a few energetic electrons that are efficient ioniz-
ers. A good approximation to the data is given by the simple formula
2.0 10 13 Tв (вV ) 12 13.6
ion vв ( ) exp( )m 3 s 1 (35)
T (вV ) 13.6 Tв (вV )
6.0 в
13.6
The source rate for neutrals (corresponding to a sink term for elec-
trons) in a plasma in coronal equilibrium is given by
S n nв ni rвв vв (36)
29
Fig. 7. Ionization rate (σion ve ) of electron-impact ionization of hydrogen
atoms averaged over a Maxwellian distribution of electrons, temperature Te
where σrec is the cross section for radiative recombination. For a neu-
tral hydrogen plasma, ni=ne. A good approximation to the data on radia-
tive recombination in the relevant temperature regime is given by the
simple formula
13.6 12 3 1
ion vв 0.7 10 19 ( ) ms (37)
Tв (вV )
The degree of ionization of a homogeneous hydrogen plasma in cor-
onal equilibrium is given by balancing the source of electrons by colli-
sional ionization against the sink of electrons by radiative recombination.
We find that, at an electron temperature of approximately the ionization
potential, i.e. 13.6 eV, the plasma is almost fully ionized so that the neu-
trals constitute only about one part in105. Only at electron temperatures
below about 1.5 eV is the plasma less than 50 % ionized. Fig. 8 shows the
degree of ionization, i.e. ne /nn, against electron temperature for the coro-
nal equilibrium model, and also for higher-density plasmas where three-
body recombination has been included.
30
Fig. 8. Ionization equilibrium for hydrogen in the coronal equilibrium
model, and at higher electron densities with three-body recombination included
31
lidity of any coronal equilibrium model depends on the time-scale for
reaching ionizationhecombination balance (for the slowest such process,
generally at the highest relevant ionization state, in the case of high-Z
ions) being much shorter than the timescale on which particles are intro-
duced into, or lost from, the plasma. If this ‘confinement’ time begins to
be comparable to the slowest atomic processes, the ionization balance
shifts towards lower charge states. If hydrogen neutrals are present, there
is also the possibility of a ‘charge-exchange’ event, in which the electron
of a neutral hydrogen atom is captured by a high-Z ion; this process also
lowers the charge-state balance of the high-Z ions. Both electrons then
decay to the ground state, emitting photons. Dielectronic recombination
has not been included in calculating the charge-state distribution shown in
Fig. 9.
32
1.9 Penetration of neutrals into plasmas
33
ciation, which produces two atoms with equal and opposite momenta and
each with energy of about 3eV; the atom with momentum directed toward
the plasma can penetrate somewhat further into the plasma.
A second atomic process-known as ‘charge exchange’-allows much
deeper penetration of hot dense plasmas by neutrals. In hydrogen charge
exchange, an energetic plasma proton captures the electron from a lower-
energy neutral. As a result, it can escape from the plasma, or move further
into the plasma, as an energetic neutral, as illustrated in Fig. 10.
34
would escape, so that the hot plasma would quickly be converted into
cold plasma.
Fig. 11. Cross section for charge exchange in hydrogen against the ener-
gy of the bombarding ion
35
about the same average energy. However, charge-exchange transport still
provides an avenue for ion energy loss from the plasma.
Fig. 12. Trajectories of two neutrals incident (thick arrows on the left)
upon a plasma of increasing density
36
comes within a distance of a singly charged (e.g. hydrogen) ion, it expe-
riences an attractive Coulomb force:
Fr e 2 / 4 0 r 2 (40)
which tends to deflect the electron orbit toward the ion. When the
angle of deflection is as much as 900, the electron’s initial momentum is
mostly lost.
Thus, from the viewpoint of momentum exchange, a ‘close encoun-
ter’ with the Coulomb force of another charged particle is essentially the
same as a ‘collision’. The angle of deflection will be large when the po-
tential energy of the Coulomb interaction equals the kinetic energy of the
colliding electron, i.e.
e 2 / 4 0 b mv 2 / 2 Tв (41)
where m and U are the mass and velocity of the electron, and where
b is the distance of closest approach of the electron to the ion. This serves
to define an effective ‘Coulomb cross section’ of the ion, namely
в 4
i b 2 10 17 / Tв (eV ) 2 m 2 (42)
(4 0 ) Tв
2 2
37
Chapter 2. Introduction to gas discharge physics
38
filled with various gases at different pressures. The quantities measured in
the experiment are the voltage between the electrodes and the current in
the circuit. This classical device served the study of discharge processes
for nearly 150 years, and still remains useful.
39
Several conditions determine how the process develops at higher voltage.
At low pressure, say 1 to 10 Torr, and high resistance of the external cir-
cuit (it prevents the current from reaching a large value), a glow dis-
charge develops. This is one of the most frequently used and important
types of discharge. It is characterized by low current, i ~ 10-6-10-1 A in
tubes of radius R ~ 1 cm, and fairly high voltage: hundreds to thousands
of volts. A beautiful radiant column, uniform along its length, is formed
in sufficiently long tubes of, say, L ~ 30 cm at P ~ 1 Torr. (This is how
glowing tubes for street advertisements are made.) The ionized gas in the
column is electrically neutral practically everywhere except in the regions
close to the electrodes; hence, this is a plasma. The glow discharge plas-
ma is very weakly ionized, to x = 10-8-10-6 (where x denotes the fraction
of ionized atoms), and is non-equilibrium in two respects. Electrons that
_
get energy directly from the field have a mean energy 1eV and a tem-
perature Te 10 4 K . The temperature T of the gas, including the ions, is not
much higher than the ambient temperature of 300 K. This state, with
widely separated electron and gas temperatures, is sustained by a low rate
of Joule heat release under conditions of relatively high specific heat of
the gas and high rate of its natural cooling. Also as a result of the high
rate of charge neutralization in a cold gas, its degree of ionization is many
orders of magnitude lower than the thermodynamic equilibrium value cor-
responding to the electron temperature.
If the pressure in the gas is high (about the atmospheric level) and
the resistance of the external circuit is low (the circuit allows the passage
of a high current), an arc discharge usually develops soon after break-
down. Arcs typically burn at a high current (i > 1 A) at a low voltage of
several tens of volts; they form a bright column. The arc releases large
thermal power that can destroy the glass tube: Arcs are often started in
open air! Atmospheric-pressure arcs usually form thermodynamic equilib-
rium plasmas (the so-called low-temperature plasma), with Te T 10 4 K
and the ionization of x 103 101 corresponding to such temperatures.
The arc discharge differs essentially from the glow discharge in the
mechanism of electron emission from the cathode, which is vital for the
flow of dc current of the arc. In the glow discharge, electrons are knocked
from the surface of the cold metal by impacts of positive ions. In the arc
discharge, the high current heats up the cathode, and thermionic emission
develops.
If P ~ 1 atm, the interelectrode gap L > 10 cm, and the voltage is suf-
ficiently high, sparking occurs. The breakdown in the gap develops by
rapid growth of the plasma channel from one electrode to another. Then
40
the electrodes are as if short - circuited by the strongly ionized spark
channel. Lightning, whose "electrodes" are a charged cloud and the
ground, is a giant variety of the spark discharge.
Finally, a corona discharge may develop in strongly nonuniform
fields that are insufficient for the breakdown of the entire gap: A radiant
corona appears at sharp ends of wires at sufficiently high voltage and also
around power transmission line conductors.
41
Townsend discharge) in a uniform electric field. Numerous experimental
results were gradually accumulated on cross sections of various electron-
atom collisions, drift velocities of electrons and ions, their recombination
coefficients, etc. This work built the foundations of the current reference
sources, without which no research in discharge physics would be possi-
ble. The concept of a plasma was introduced by I. Langmuir and L. Tonks
in 1928. Langmuir made many important contributions to the physics of
gas discharge, including probe techniques of plasma diagnostics.
As regards different frequency ranges, the development of field gen-
erators and the research into the discharges they produce followed the or-
der of increasing frequencies. Radio frequency (rf) discharges were ob-
served by N. Tesla in 1891. This kind of discharge is easily produced if
an evacuated vessel is placed inside a solenoid coil to which high-
frequency voltage is applied. The electric field induced by the oscillating
magnetic field produces breakdown in the residual gas, and discharge is
initiated. The understanding of the mechanism of discharge initiation
came much latter, in fact, after the work of J.J. Thomson in 1926-1927.
Inductively coupled rf discharges up to tens of kW in power were ob-
tained by G.I. Babat in Leningrad around 1940.
The progress in radar technology drew attention to phenomena in
microwave fields. S.S. Brown in the USA began systematic studies of mi-
crowave discharges in the late 1940s. Discharges in the optical frequency
range were realized after the advent of the laser: A spark flashed in air
when the beam of a ruby laser producing so-called giant pulses (of more
than 10 MW in power) was focused by a lens, this success being achieved
in 1963. Continuously burning optical discharges, in which dense steady-
state plasma is sustained by the energy of light radiation, were first initi-
ated in 1970 by a cw CO2 laser. Optical discharges (this term reflects a
large degree of similarity with conventional discharges) immediately at-
tracted considerable attention. Both microwave and optical discharges
have by now been studied with at least the same thoroughness that the
discharges in constant electric fields has been during nearly 100 years of
research.
The physics of the glow discharge, one of the oldest and, presuma-
bly, best- studied fields, has lived through an unparallel revival in the past
15-20 years, and numerous new aspects of this phenomenon have been
revealed. This surge of attention was stimulated by the use of glow dis-
charges in electric-discharge CO2 lasers developed for the needs of laser
technologies. Likewise, the application of plasmatrons (generators of
dense low-temperature plasma) to metallurgy, plasma chemistry, plasma
welding and cutting, etc. provided a stimulus for new extensive, detailed
42
studies of arc plasma at P ~ 1 atm, T ~ 104 K, and of similar discharges in
all frequency ranges. These, and many other practical applications of gas
discharge physics place it within the range of sciences that lie at the foun-
dation of modern engineering.
43
Note that the electric field not only reduces the height of the barrier
but it also reduces the thickness of the barrier. This makes it easier for the
energetic electrons to jump over the barrier (Schottky effect) and a certain
percentage of the electrons to tunnel through the barrier
The work functions of typical metals are tabulated in Table 3 [13].
Element Ag Al Cu Fe W
Wα(eV) 4.74 2.98- 4.07- 3.91- 4.35-
4.43 4.7 4.6 4.6
Electrons can be removed from the metal either by giving the elec-
trons sufficient kinetic energy to surmount the potential barrier, or the
work function, at the surface, by reducing the height of the barrier so that
the electrons can overcome it or by reducing the thickness of the barrier
so that the electrons can tunnel through it. The first can be achieved by the
application of heat to the electrode, through impact of photons on the sur-
face of the electrode or by the incidence of particles such as other elec-
trons, positive ions, neutral molecules, metastable atoms onthe electrode.
The reduction in the potential barrier height or its thickness can
beachieved by the application of an electric field in the correct direction
so that the electrons will experience a force directed out of the metal sur-
face (see figure 14 b).Let us consider different physical processes that can
cause emission of electrons from metals.
44
2.4.3 Thermionic emission
As one can see from the equation given in the previous section, the
thermionic current depends on the height of the barrier that the electrons
have to surmount in order to come out of the metal. The height of this
barrier can be decreased by the application of an electric field in such a
way that the electrons in the metal experience a force out of the metal.
This is illustrated in Fig. 14 b.
In the presence of such an electric field the work function is effec-
tively reduced to [15]:
e3 E
1 (45)
4 0
This reduction in the barrier height will lead to a change in the ther-
mionic emission current. Thus, if J0 is the thermionic emission current
density for zero electric field at temperature T then the current density at
the same temperature in the presence of an electric field E is given by:
J s J 0exp( 4.4 E / T ) (46)
45
where the electric field E is given in V/cm and T in Kelvins. This was
shown to be valid for electric fields up to about 106 V/cm. This process of
enhancement of the thermionic emission current due to the reduction in
the barrier height is called Schottky emission.
Calculations done with the Schottky equation show that the thermi-
onic emission current at T = 293K for values of φ about 4.5eV is negligi-
ble even when the electric field reaches values as high as106 V/cm. How-
ever, experiments show that electrodes in vacuum do emit appreciable
currents, in the range of μA, at such electric fields. The reason for this is
the quantum nature of the elementary particles. The illustration in Fig. 14
b shows that the application of the electric field not only reduces the
height of the barrier but will also decrease its thickness. In the absence of
the electric field the barrier is infinitely thick but its thickness decreases
with increasing electric field.
Electrons incident on the barrier can be represented by a wave; dur-
ing the inter-action part of the wave will be reflected and the other part of
the wave will be transmitted. The transmitted wave attenuates rapidly
when moving into the barrier, but if the thickness of the barrier is not
large a small fraction of the wave may be able to penetrate it. Of course,
in the case of electrons, the reflection and transmission coefficients have
to be regarded as probability functions, so there is a certain probability
that an electron incident on the barrier will penetrate it. If the number of
electrons incident on the barrier per unit time is known, quantum mechan-
ical calculations can be performed to evaluate the number of electrons
coming out of the barrier. Fowler and Norheim [16] analysed this process
in detail and obtained the following expression for the field emission cur-
rent density, Jf , for pure metallic surfaces in vacuum.
1/ 2 E 2
J f 61.6 10 7 exp(6.8 107 3 / 2 / 3E ) (47)
( ) 1/ 2
46
surface contamination. The reason for this is that surface contamination
causes a reduction in the width of the barrier thus enhancing the field
emission process. Moreover, if there are protrusions on the surface the
electric field at the tip of these protrusions can reach very high values
leading to field emission from them. The field emission process is very
important in providing initiatory electrons in the creation of electrical dis-
charges.
47
ionisation collisions and these two electrons in turn will give rise to two
more electrons. In this way the number of electrons increases with in-
creasing x.
Assume that the number of electrons at a distance x from the origin
is n x . Let be the number of ionising collisions per unit length made by
an electron travelling in the direction of the electric field and be the
number of attachments over the same length. Here is Townsend’s ioni-
zation coefficient and is the attachment coefficient. Consider an elemen-
tary length of width dx located at a distance x from the origin. In travel-
ling across the length dx, n x number of electrons will give rise to dn addi-
tional electrons:
dn nx ( )dx (48)
The solution of this equation is:
nx e( ) x (49)
This equation shows that the number of electrons increases exponen-
tially with dis-tance. This exponential growth of electrons with distance is
called an electron avalanche. Fig. 15 shows a photograph of an electron
avalanche obtained in a cloud chamber [17].
The equation also shows that cumulative ionisation is possible only
if (α−η) >0. The quantity (α−η) is known as the effective ionisation coef-
ficient and denoted by .
48
tistical nature of the collision process. The probability that one electron at
the origin results in an avalanche of total number n at a distance x is given
by
n 1
1 1
Р(n, х) 1 (50)
nmean nmean
with a standard deviation given by:
1/ 2
1
nmean 1 (51)
nmean
Where
nmean е ( ) x (52)
49
closer to the cathode. The process repeats itself and the positive space
charge head travels towards the cathode as a consequence. This discharge
is called a cathode directed streamer or a positive streamer.
50
The formation of a negative streamer or an anode directed streamer
is shown in Fig. 17. The electrons of the avalanche move into the gap
leaving behind positive charge close to the cathode. When the avalanche
reaches the critical size the secondary avalanches extend the positive
space charge towards the cathode (as in a cathode directed streamer).
When the positive channel reaches the cathode both the field enhance-
ment associated with the proximity of positive space charge to the cath-
ode and the collision of positive ions on the cathode lead to the emission
of electrons from the latter. These electrons will neutralize the positive
space charge creating a weakly conducting channel that connects the neg-
ative head of the electron avalanche to the cathode. The high electric field
at the head of the avalanche pushes the negative space charge further into
the gap while the positive space charge left behind is neutralized by the
electrons supplied by the cathode and travelling along the weakly con-
ducting channel connecting the streamer head and the cathode.
51
of positive ions in the avalanche head reaches a critical value of about
108. A similar conclusion is also reached independently by Meek [22].
Thus the critical avalanche length for transition to a streamer is given by
e x 10 8
c
(54)
or
axc 18 (55)
52
The mechanism of propagation of negative streamers is a little bit
more complicated than that for positive streamers. This is shown in Fig.
19.
Note that there are two main differences between the negative and
positive streamers. In the negative streamer, the electrode has to supply
the electrons necessary for the neutralisation of the positive space charge
left behind by the avalanches whereas in the positive streamers the anode
absorbs the extra electrons generated by the secondary streamers. The lat-
ter is a much easier process than the former. Second, in the positive
streamers the electrons propagate towards the positive charge head of the
streamer and therefore into an increasing electric field. In the case of neg-
ative streamers the electrons move into the low electric field region and
some of them will be captured by electronegative atoms which create an
immobile negative space charge region that will impede the streamer
propagation. Both these features make the propagation of positive stream-
ers easier than that of negative streamers.
We have seen before that for the inception of a streamer the number
of charged particles at the avalanche head should reach a critical value.
Similarly, for the continuous propagation of a streamer the number of
charged particles in the streamer head has to be larger than a critical num-
ber, Nstab. The value of Nstab partly depends on the background electric
field, Eb, and can be calculated from the following equation [25]:
53
2.6 Electrical breakdown in very small gaps – Townsend’s
breakdown mechanism
54
The discharge gap is located in the vacuum tube and the cathode is
illuminated by ultraviolet radiation.
Fig. 21. Variation of the current flowing across the discharge tube of
Townsend’s experiment as a function of the applied voltage
55
2.6.2 Townsend’s theory of electrical breakdown
nd n0ed (58)
Consequently, the current inside the tube is given by:
I d I 0ed (59)
56
positive ions. But, in reality, the energy gained by the positive ions at
electric fields encountered in Townsend’s experiment is not sufficient to
create significant ionization. However, one process that may cause this
departure from eqn. 59 is the generation of electrons from the cathode by
positive ion bombardment. As the voltage increases the positive ions gain
more and more energy and this energy is released at the cathode. With in-
creasing energy a stage will be reached in which these positive ions will
start liberating electrons from the electrode. In order to explain the varia-
tion of current with voltage one has to take this effect into account. Let us
now derive a mathematical expression for the current in the discharge
tube taking into account the electron current created by the bombardment
of positive ions on the cathode.
Fig. 22. Variation of the logarithm of the current flowing across the
discharge tube in Townsend’s experiment as a function of the logarithm of the
electrode spacing for different values of (E/p)
57
The number of positive ions created by the electrons reaching the
anode per second is equal to n (n0 n ) and at steady state the number of
positive ions reaching the cathode per second is equal to this number.
Consequently, the number of electrons released by the positive ion bom-
bardment at the cathode per second is given by:
n n (n0 n ) (61)
Where is the average number of electrons released by each posi-
tive ion striking the cathode. This parameter is called Townsend’s sec-
ondary ionisation coefficient. Substituting this expression in eqn. 60 one
obtains:
n0ed
n (62)
1 (ed 1)
The current in the discharge tube is given by:
I 0ed
I (63)
1 (ed 1)
This equation predicts a faster current growth than eqn. 60 with in-
creasing electric field (or voltage) providing an explanation for Town-
send’s experimental results.
In deriving the above equation we have considered the bombardment
of positive ions on the cathode as the only secondary ionization process.
Let us now consider the other secondary ionization processes.
(I) Ionization of gas by positive ions: Townsend in his original deri-
vation assumed that the secondary ionization mechanism is due to the ion-
ization in the gas by positive ions. However, as mentioned above, positive
ions cannot produce significant ionisation in electric fields at which elec-
trical breakdown is observed in Townsend’s experiment. Nevertheless, if
we assume that the ions will contribute to the ionization the resulting
equation for the current will take the form [12]:
e( ) d
i i0 (64)
( /( )) ( /( ))e( ) d
Where β is the ionisation coefficient of positive ions. Since electrons
ionize more readily than positive ions one can replace ( ) by and
the equation will reduce to:
e ( ) d
i i0 (65)
(1 ( / )e ( ) d
This equation has the same form as that of the expression obtained
for positive ion bombardment at the cathode (i.e. eqn. 63).
(II) Photo emission from the electrode: Another secondary emission
process that one may take into account is the interaction of photons in the
58
discharge with the electrodes. If the incident photon has an energy larger
than the work function of the electrode then the interaction may lead to
the liberation of an electron. If this process is taken in to account as a sec-
ondary mechanism then the expression for the current will take the form
[6]:
e( ) d
i i0 (66)
(1 (g / ))[e( ) d 1]
Where g is the fraction of photons emitted in the gas that are
headed towards the cathode, is the coefficient of absorption of photons
in the gas and is the probability of photoelectric emission from the pho-
tons incident on the electrode (note that only a fraction of the incident
photons will have sufficient energy to cause photo ionization).
In general α >> μ and the equation will reduce to:
e( ) d
i i0 (67)
(1 (g / )[e( ) d 1]
This again has the same form as that of the eqn. (68).
59
ondary ionisation process under consideration the final expression for the
current has the same form. Indeed one can include all of them in a single
formula as follows:
e ( ) d
i i0 (70)
(1 ( i )[e ( ) d 1]
Where
gf
i ( ) g / m ( ) (71)
The final expression for the current given in eqn. 69 shows that the
discharge is still non self-sustained. That is, the discharge current goes to
zero if the UV illumination on the cathode is removed (i.e. as I0 → 0).
However, as the voltage continues to increase a stage will be reached at
which the discharge will transform itself from a non self-sustained dis-
charge to a self-sustained discharge. At this stage the discharge will con-
tinue to burn between electrodes even after removing the background ul-
traviolet radiation (i.e. when I0 = 0). This change of state of the discharge
is accompanied by a several orders of magnitude increase in the current
(provided that the voltage supply can sustain such an increase in the cur-
rent) in the discharge gap. This is the stage of electrical breakdown in the
gap. Townsend defined the electrical breakdown condition as the condi-
tion which makes the current in the discharge gap go to infinity.
From eqn. 69 one can see that I0→ ∞ when the denominator of the
expression for the current goes to zero. That is when:
One can indeed show that this criterion has a physical significance.
Assume that is the dominant secondary process. Assume that n0 prima-
ry electrons leave the cathode per second. These electrons will give rise to
n0 ( e( ) d 1 ) positive ions in the gap, and these positive ions on incidence
on the cathode produce i n0 (e( ) d 1) secondary electrons. When Town-
send’s breakdown criterion is satisfied the number of secondary electrons
is equal to the original number of electrons which has been drawn away
from the cathode and later passed into the anode. Consequently, each ava-
lanche will give rise to another avalanche through secondary processes
60
and so cause a repetition of the avalanche process. That is, the discharge
process becomes self-sustained.
An alternative expression for Townsend’s breakdown criterion can
be obtained by rewriting the above equation as:
1
d ln(1 ) (73)
i
The value of i is greatly affected by the cathode surface and gas
pressure. However, i is a very small number (<10−2–10−3) so 1/ i is very
large. Therefore, ln(1+(1/ i )) does not change too much and is of the or-
der of eight to ten in a Townsend’s discharge.
61
i
i0
e( ) d (79)
In the presence of secondary ionisation due to bombardment of posi-
tive ions on the cathode one can show using the procedure outlined in the
section 2.6.2.2
e( ) d
i i0
e( ) d 1
(80)
This reduces to eqn.80 in the absence of attachment (i.e. =0). From
this equation the breakdown condition in the presence of electron attach-
ment is given by:
( ) d
1 [(e 1)] 0 (81)
( )
This criterion shows that if then electrical breakdown is possi-
ble irrespective of the values of , and provided that d is large
enough. That is, for a given electric field there is a particular value of d at
which the gap breaks down. For , with increasing d, the above equa-
tion approaches an asymptotic form:
1 or (82)
( ) (1 )
This defines the limiting condition at which electrical breakdown is
possible in an electronegative gas. This condition depends only on E/p.
Noting that the value of <<1,the limiting value of E/p which can cause
electrical breakdown in electronegative gases can be obtained from the
relationship (see section 2.5). This point is illustrated in the plot
given in Fig. 23.
62
Fig. 23. Variation of the logarithm of the current flowing across the dis-
chargetube in Townsend’s experiment as a function of the logarithm of the elec-
trode spacing for different values of (E/p) for electronegative gases
This is known as Paschen’s law. The Paschen curve for air is shown
in Fig. 24; the data points correspond to measurements by several authors
and the solid black dots are generated from the equation
Vs 6.72 Pd 24 .4( Pd )) (P in bar and d in mm).
Fig. 24. Paschen curve for air in log–log scale at temperature 20◦C [28]
63
case of low and high values of Pd but there is a minimum at a certain val-
ue of pd. This minimum is called the Paschen minimum. The Paschen
minimum in air is about Pd=10−2 bar.mm. One can show that the break-
down voltage estimated using either Townsend’s or Raether and Meek’s
criterion adheres to Paschen’s law. The electrical breakdown criterion of a
uniform gap of length d is given by
d K (84)
Where K is a constant. Depending on the value of K this equation
represents both Townsend’s breakdown criterion and the streamer break-
down criterion. Substituting the expression for given in eqn. 80 we ob-
tain:
K Apde Bpd / V s
(85)
Where Vs is the voltage at which electrical breakdown is observed.
Note that in deriving this equation we have used E=Vs/d. Rearranging the
above equation we find that
Bdp
Vs (86)
lnApd / K
This equation shows that Vs is a function of pd. The general shape of
this equation is in agreement with the Paschen curve.
Where A1 and B1 are constants, is the mean free path of the elec-
trons Es=Vs/d ,and we have used the relationship that the mean free path
is inversely proportional to the pressure. This equation has the same form
as eqn. 87 except that Pd is now replaced by d / . It thus predicts that the
breakdown electric field has a minimum corresponding to a certain value
of d / , say ( d / )min, and it increases when the value of d / moves away
from this minimum. The reason for the existence of this minimum can be
explained qualitatively as follows. Let Em be the electric field in the gap
corresponding to ( d / )min. At ( d / )min an electron crossing the gap will
make a certain number of ionisation collisions. Consider the case d / >
( d / )min. Now, the number of collisions made by an electron in crossing
the gap, and hence the energy lost in collisions in crossing the gap, is
higher than at ( d / )min. If the background electric field remained at Em the
64
total number of ionisation collisions made by an electron crossing the gap
would be less than the corresponding value at ( d / )min. Consequently, the
electric field should be increased in order to compensate for the losses and
to increase the probability of ionisation. For d / < ( d / )min the number of
collisions, and hence the number of ionisation collisions, made by an
electron in crossing the gap is less than the corresponding value at
( d / )min. In this case the only way to increase the number of ionisation
collisions is to increase the probability of ionisation in each collision.
This can be achieved only by increasing the energy gained by electrons
within a mean free path. This requires a higher electric field than the one
which corresponds to ( d / )min.
In section 2.7 it was shown that Paschen’s law follows directly if the
dominant collision processes, as is often the case, are such that the coef-
ficients representing them, for example α, are directly proportional to p at
a given value of E/p. When this is the case the processes are said to obey
similarity. In general, the experimental data obeys Paschen’s law and any
deviations are relatively small and arise from the existence of collision
processes in the gas which do not conform to similarity. The deviations
from Paschen’s law can occur at high pressures and at temperatures above
about 3000–4000 K. At high pressures exceeding a few atmospheres the
processes such as field emission may play a significant role in the break-
down process. The role of field emission at high pressures and its
influence on Paschen’s law is clearly demonstrated by the observation
that when very clean molybdenem electrodes are used Paschen’s law
holds up to very high pressures. The clean electrodes do not have oxide
layers that generate field emission at low electric fields. At high tempera-
tures experimental data departs from Paschen’s law partly due to the dom-
inant role of thermal ionisation and partly due to the gradual change of
chemical composition of the gas, for example, by dissociation. Paschen’s
law may also break down at low pressures because the breakdown process
is governed by preionisation processes caused by electrodes such as ther-
mionic emission and the break down phenomena has to be described by
vacuum breakdown processes.
65
more diversified, and richer in physical effects. Steady and quasi-steady
self-sustaining discharges contain A) glow and B) arc discharges. We
have already mentioned in Sect. 2.2 that the cathode processes of two
types differ in principle. A close relation of the glow discharge is C)
Townsend's dark discharge. It proceeds with a cold cathode and at very
weak current. The D) corona discharge, also self-sustaining and also at a
low current, is a special case.
Typical conditions under which each of the combinations can be ob-
served are summarized in Table 4.
Corona has common features with glow and dark discharges. Among
transient discharges, the E) spark discharge stands out sharply, among
others. Many features of purely plasma processes, characterizing break-
down in a dc electric field, as well as the glow and arc discharges, are typ-
ical for discharges in rapidly oscillating fields, where electrodes are not
necessary at all. It is therefore expedient to construct a classification
avoiding the attributes related to electrode effects, and the following two
properties will be basic for the classification: the state of the ionized gas
and the frequency range of the field. The former serves to distinguish be-
tween A) breakdown in the gas, B) sustaining non-equilibrium plasma by
the field, and C) sustaining equilibrium plasma. Frequency serves to clas-
sify fields into A) dc, low-frequency, and pulsed fields (excluding very
66
short pulses), B) radio-frequency fields (f ~ 105 - 108Hz), C) microwave
fields (f ~ 109 - 1011 Hz, λ~ 102 – 10-1 cm), and D) optical fields (far from
infrared to ultraviolet light). The field of any subrange can interact with
each type of discharge plasma. In total, we have 12 combinations. All of
them are experimentally realizable, and quite a few are widely employed
in physics and technology.
67
Chapter 3 Basic principles of plasma chemical processes
68
stimulated oxidation process in polymer processing, and biological and
medical applications.
The contribution of atoms and radicals is obviously significant. As
an example, we can point out that O atoms and OH radicals effectively
generated in atmospheric air discharges which play a key role in numer-
ous plasma-stimulated oxidation processes. Plasma-generated photons
play a key role in a wide range of applications, from plasma light sources
to UV sterilization of water.
Plasma is not only a multi-component system, but often a very non-
equilibrium one. Concentrations of the active species described earlier
can exceed those of quasi-equilibrium systems by many orders of magni-
tude at the same gas temperature.
The successful control of plasma permits chemical processes to be
directed in a desired direction, selectively, and through an optimal mech-
anism. Control of a plasma-chemical system requires detailed understand-
ing of elementary processes and the kinetics of the chemically active
plasma. The major fundamentals of plasma physics, elementary processes
in plasma was discussed in Chapters 1 and 2, and plasma kinetics are to
be discussed in Chapters 3.
69
must be those of concentration divided by time (moles=liter/sec,
moles=liter/min, etc.). A reaction that can be written as
A→B (88)
has a rate that can be expressed either in terms of the disappearance
of A or the appearance of B. Because the concentration of A is decreasing
as A is consumed, the rate is expressed as -d[A]/dt. Because the concen-
tration of B is increasing with time, the rate is expressed as +d[B]/dt. The
mathematical equation relating concentrations and time is called the rate
equation or the rate law. The relationships between the concentrations of
A and B with time are represented graphically in Fig. 25 for a first-order
reaction in which [A]o is 1.00 M and k =0:050 min-1.
If we consider a reaction that can be shown as
aA + bB → cC + dD (89)
70
sum of the exponents in the rate law is 2, 3, etc., respectively. These ex-
ponents can usually be established by studying the reaction using differ-
ent initial concentrations of A and B. When this is done, it is possible to
determine if doubling the concentration of A doubles the rate of the reac-
tion. If it does, then the reaction must be first-order in A, and the value of
x is 1. However, if doubling the concentration of A quadruples the rate, it
is clear that [A] must have an exponent of 2, and the reaction is second-
order in A. One very important point to remember is that there is no nec-
essary correlation between the balancing coefficients in the chemical
equation and the exponents in the rate law. They may be the same, but
one can not assume that they will be without studying the rate of the reac-
tion.
If a reaction takes place in a series of steps, a study of the rate of the
reaction gives information about the slowest step of the reaction. We can
see an analogy to this in the following illustration that involves the flow
of water,
71
k[A][B], and the reaction is second-order (first-order in A and first-order
in B). It should be apparent that we can write the rate law directly from
the balanced equation only if the reaction takes place in a single step. If
the reaction takes place in a series of steps, a rate study will give infor-
mation about steps up to and including the slowest step, and the rate law
will be determined by that step.
In this section, we will examine the details of some rate laws that
depend on the concentration of reactants in some simple way. Although
many complicated cases are well known, there are also a great many reac-
tions for which the dependence on concentration is first-order, second-
order, or zero-order.
3.1.2.1 First-Order
A→B (94)
and that the reaction follows a rate law of the form
d [ A]
Rate k[ A]1 (95)
dt
This equation can be rearranged to give
d [ A]
kdt (96)
[ A]
Equation (96) can be integrated but it should be integrated between
the limits of time =0 and time equal to t while the concentration varies
from the initial concentration [A]o at time zero to [A] at the later time.
This can be shown as
[ A] t
d [ A]
k dt (97)
[ A]0
[ A] o
72
ln[ A] ln[ A]0 kt (100)
y b mx (101)
It must be remembered that [A]o, the initial concentration of A, has
some fixed value so it is a constant. Therefore, Eq. (100) can be put in the
form of a linear equation where y=ln[A], m=-k, and b=ln [A]o. A graph of
ln[A] versus t will be linear with a slope of -k. In order to test this rate
law, it is necessary to have data for the reaction which consists of the
concentration of A determined as a function of time. This suggests that in
order to determine the concentration of some species, in this case A, sim-
ple, reliable, and rapid analytical methods are usually sought. Additional-
ly, one must measure time, which is not usually a problem unless the re-
action is a very rapid one.
It may be possible for the concentration of a reactant or product to be
determined directly within the reaction mixture, but in other cases a sam-
ple must be removed for the analysis to be completed. The time necessary
to remove a sample from the reaction mixture is usually negligibly short
compared to the reaction time being measured. What is usually done for a
reaction carried out in solution is to set up the reaction in a vessel that is
held in a constant temperature bath so that fluctuations in temperature will
not cause changes in the rate of the reaction. Then the reaction is started,
and the concentration of the reactant (A in this case) is determined at se-
lected times so that a graph of ln[A] versus time can be made or the data
analyzed numerically. If a linear relationship provides the best fit to the
data, it is concluded that the reaction obeys a first-order rate law. Graph-
ical representation of this rate law is shown in Fig. 26 for an initial con-
centration of A of 1.00 M and k=0:020 min-1. In this case, the slope of the
line is - k, so the kinetic data can be used to determine k graphically or by
means of linear regression using numerical methods to determine the
slope of the line.
73
Fig. 26. First-order plot for A → B with [A]o = 1:00 M and
k = 0:020 min -1
The units on k in the first-order rate law are in terms of time-1. The
left-hand side of Eq. (98) has [concentration]/[concentration], which
causes the units to cancel. However, the right-hand side of the equation
will be dimensionally correct only if k has the units of time-1, because on-
ly then will kt have no units
The equation
ln[ A] ln[ A]0 kt (102)
can also be written in the form
[ A] [ A]0 e kt (103)
From this equation, it can be seen that the concentration of A de-
creases with time in an exponential way. Such a relationship is sometimes
referred to as an exponential decay.
Radioactive decay processes follow a first-order rate law. The rate of
decay is proportional to the amount of material present, so doubling the
amount of radioactive material doubles the measured counting rate of de-
cay products. When the amount of material remaining is one-half of the
original amount, the time expired is called the half-life. We can calculate
the half-life easily using Eq. (98). At the point where the time elapsed is
equal to one half-life, t=t1/2, the concentration of A is one-half the initial
concentration or [A]o=2. Therefore, we can write
74
A0 ln
A0 kt1 / 2 ln 2 0.693 (104)
ln
A A0
2
The half-life is then given as
0.693
t1 / 2 (105)
k
and it will have units that depend on the units on k. For example, if k
is in hr-1, then the half-life will be given in hours, etc. Note that for a pro-
cess that follows a first-order rate law, the half-life is independent of the
initial concentration of the reactant. For example, in radioactive decay the
half-life is independent of the amount of starting nuclide. This means that
if a sample initially contains 1000 atoms of radioactive material, the half-
life is exactly the same as when there are 5000 atoms initially present. It
is easy to see that after one half-life the amount of material remaining is
one-half of the original; after two half-lives, the amount remaining is one-
fourth of the original; after three half-lives, the amount remaining
is one-eighth of the original, etc. This is illustrated graphically as shown
in Fig. 27.
75
life of a chemical reaction as the time necessary for the concentration of
some reactant to fall to one-half of its initial value.
3.1.2.2 Second-Order
d A
Rate k A
2
(106)
dt
Such a rate law might result from a reaction that can be written as
2 A Pr oducts (107)
However, as we have seen, the rate law cannot always be written
from the balanced equation for the reaction. If we rearrange Eq. (106), we
have
d A
kdt (108)
A2
d A
A t
A A2 k 0 dt (109)
0
76
1/concentration. If concentration is expressed in mole/liter, then
1/concentration will have units of liter/mole. From this we find that the
units on k must be liter/mole time or M-1 time-1 so that kt will have units
M-1.
Fig. 28. A second-order rate plot for A → B with [A]o = 0:50 M and k =
0.040 liter/mol min
The half-life for a reaction that follows a second-order rate law can
be easily calculated. After a reaction time equal to one half-life, the con-
centration of A will have decreased to one-half its original value. That is,
[A] =[A]o/2, so this value can be substituted for [A] in Eq. (110) to give
1 1
A0 A0 kt1/ 2
(112)
2
Removing the complex fraction gives
2 1 1
kt1 / 2 (113)
A0 A0 A0
Therefore, solving for t1/2 gives
77
1
t1 / 2
k Ao
Here we see a major difference between a reaction that follows a
second-order rate law and one that follows a first-order rate law. For a
first-order reaction, the half-life is independent of the initial concentration
of the reactant, but in the case of a second-order reaction, the half-life is
inversely proportional to the initial concentration of the reactant.
78
The example illustrated in the figure represents an exothermic reac-
tion because the overall energy change is negative since the products have
a lower energy than the reactants. When the various rate laws are inspect-
ed, we see that only k can be a function of temperature because the con-
centrations remain constant or very nearly so as the temperature changes
only a small amount. Therefore, it is the rate constant that incorporates
information about the effect of temperature on the rate of a reaction.
There are several types of behavior exhibited when the rates of re-
actions are studied as a function of temperature. Three of the most com-
mon variations in rate with temperature are shown in Fig. 30.
The first case shows the variation followed by most reactions, that
of an exponentially increasing rate as temperature increases. The second
shows the behavior of some material that becomes explosive at a certain
temperature. At temperatures below the explosive limit, the rate is essen-
tially unaffected by the temperature. Then, as the temperature is reached
at which the material becomes explosive, the rate increases enormously as
the temperature is increased only very slightly. In the third case, we see
the variation in rate of reaction that is characteristic of many biological
processes. For example, reactions involving enzymes (biological cata-
lysts) frequently increase in rate up to a certain temperature and then de-
crease in rate at higher temperatures. Enzymes are protein materials that
can change conformation or become denatured at high temperatures.
Fig. 30. Some of the ways in which reaction rates vary with temperature
Svante August Arrhenius suggested in the late 1800s that the rates
of most reactions vary with temperature (as shown in Fig. 30 a) in such a
way that
k Ae E / RT
a
(114)
where k is the rate constant, A is the frequency factor (or pre-
exponential factor), R is the molar gas constant, Ea is the activation ener-
gy, and T is the temperature (K). This equation is generally referred to as
the Arrhenius equation. If we take the natural logarithm of both sides of
Eq. (114), we obtain
79
Ea
ln k ln A (115)
RT
By rearrangement, this equation can be put in the form of a straight
line,
Ea 1
ln k ln A (116)
R T
y m x b
Therefore, a plot of ln k versus 1/T can be made or linear regression
performed after the rate constants have been determined for a reaction
carried out at several temperatures. The slope of the line is - Ea/R and the
intercept is ln A. Such a graph, like that shown in Fig. 31, is often called
an Arrhenius plot. It is from the slope, determined either numerically or
graphically, that the activation energy is determined.
For a particular reaction, the following rate constants were obtained
when the reaction was studied at a series of temperatures which yielded
the data shown on the next page.
Fig. 31. An Arrhenius plot constructed using the data in the text
80
tions, the temperature range in which the reaction can be studied is rather
limited because at low temperatures the reaction will be very slow and at
high temperatures the reaction will be very fast. Therefore, it is generally
desired to study a reaction over a range of at least 20–250C.
If the rate constant for a reaction is determined at only two tempera-
tures, it is still possible to evaluate the activation energy, but such a case
is not nearly as desirable as the procedure described earlier. Small errors
in the rate constants will cause inaccuracy in the activation energy deter-
mined, because all of the errors in placing the line are present in only two
points. More data would be needed to ‘‘average out’’ the error in the val-
ue of any one rate constant. If k1 is the rate constant at T1 and k2 is the rate
constant at T2, we can write the Arrhenius equation as
Ea
ln k1 ln A (117)
RT1
E
ln k2 ln A a (118)
RT2
Subtracting the equation for ln k2 from that giving ln k1 gives
Ea E
ln k1 ln k 2 (ln A ) (ln A a ) (119)
RT1 RT2
3.2 The object and the main features of the plasma chemistry
81
equilibrium plasma chemical processes is determined only by a high tem-
perature of the interacting particles, whereas the specific of the non-
equilibrium plasma-chemical processes is due to a large contribution of
the internal energy of the molecules into their chemical activity.
A low-temperature plasma is plasma with a temperature of 103-105 K
and the degree of ionization of 10-6-0.1, obtained from electric, high-
frequency and microwave gas discharges, in shock tubes, adiabatic com-
pression machines and by other methods.
Plasma chemistry today is a rapidly expanding area of science and
engineering, with applications widely spread from micro-fabrication in
electronics to making protective coatings for aircrafts, from treatment of
polymer fibers and films before painting to medical cauterization for
stopping blood and wound treatment, and from production of ozone to
plasma TVs.
Applications of plasma technologies today are numerous and involve
many industries. High-energy efficiency (energy cost with respect to the
minimum determined by thermodynamics), high specific productivity
(productivity per unit volume of reactor), and high selectivity may be
achieved in plasmas for a wide range of chemical processes. As an exam-
ple, for CO2 dissociation in non-equilibrium plasma under supersonic
flow conditions, it is possible to selectively introduce up to 90 % of the
total discharge power in CO production when the vibrational temperature
is about 4000 K and the translational temperature is only about 100 K.
The specific productivity of such a supersonic reactor achieves 1,000,000
L/h, with power levels up to 1 MW. This plasma process has been exam-
ined for fuel production on Mars, where the atmosphere mostly consists
of CO2. On the Earth, it was applied as a plasma stage in a two-step pro-
cess for hydrogen production from water.
82
ered to a heavy component by the electrons, the full balance in such a sys-
tem cannot be achieved. To transfer energy to the gas the electron energy
(temperature) is always greater than the temperature of heavy particles
(unlikely to the plasma which is produced in adiabatic compression).
Plasma-chemical processes carried out in such conditions, are considered
as quasi-equilibrium processes.
Despite the equilibrium character of the physical and chemical pro-
cesses in plasmas formed in the temperature range of 3000-5000 K the
chemical reactions occur with a very high speed, so that their characteris-
tic times are of the same order that the times of heat and mass transfer. As
a result, the processes can be transformed from the kinetic region of per-
colation to the diffusion. The reaction mechanisms are undergoing signif-
icant changes, in particular, can enter the reaction with the walls of a reac-
tor. The plasma in this case is considered only as an effective energy car-
rier.
83
particle diffusion to the walls of the vessel approaches the characteristic
times of chemical reactions. As a result, the role of heterogeneous (on the
walls), physic-chemical processes is increasing to such an extent that they
should be considered when analyzing the mechanisms and kinetics of
plasma chemical reactions.
In non-equilibrium plasma the distribution of velocity and the inter-
nal states are different, in general, from that in Maxwell-Boltzmann dis-
tribution. In this situation we should use a set of Boltzmann equilibrium
distributions with different temperatures for different particle states. The
internal state of the particles are divided into rotational, vibrational, and
electronic, the populations of which are described by the equilibrium
Boltzmann distribution with temperatures Tr, Tv, Te, respectively. If the
electron velocity distribution is Maxwellian with the temperature Te, and
the distribution of atoms and molecules by velocities (a translational de-
gree of freedom) is also Maxwellian with the temperature Tg (gas temper-
ature), then, for a non-equilibrium plasma
Tg ≠ Tr ≠ Tv ≠ Te
Typically, this situation is observed in the plasma at low pressures
(less than 100 Torr). It often happens that Tg ≈ Tr <Tv <Te, and if Tg and
Tr are close to room temperature (300K), the electron temperature corre-
sponds to the average energy of several electron volts (1 eV = 11600K).
Non-equilibrium thermodynamics is a consequence of "openness" of the
system. Plasma-chemical processes carried out in such conditions, are de-
fined as non-equilibrium processes.
84
in this chapter mostly on the micro-kinetics of the elementary reactions –
on their cross sections and probabilities.
85
Atoms or molecules lose their electrons in the ionization process and
form positive ions. In hot thermonuclear plasmas, the ions are multi-
charged, but in relatively cold plasmas of technological interest their
charge is usually equal to +1e (1.6 · 10−19 C). Ions are heavy particles, so
usually they cannot receive high energy directly from electric fields be-
cause of intensive collisional energy exchange with other plasma compo-
nents. The collisional nature of the energy transfer results in the fact that
the ion energy distribution function at lower pressures is often not far
from the Maxwellian one (120), with an ion temperature Ti close to neu-
tral gas temperature T0.
Electron attachment leads to formation of negative ions with
charge−1e (1.6 10−19 C). Attachment of another electron and formation of
multi-charged negative ions is actually impossible in the gas phase be-
cause of electric repulsion. Negative ions are also heavy particles, so usu-
ally their energy balance is due not to electric fields but to collisional pro-
cesses. The energy distribution functions for negative ions, similar to that
for the positive ones, are not far from the Maxwellian distributions (121)
at pressures that are not too low. Their temperatures in this case are also
close to those of a neutral gas.
86
4 Photo-ionization takes place in collisions of neutrals with photons,
which result in the formation of an electron–ion pair. Photo-ionization is
mostly important in thermal plasmas and in some mechanisms of propa-
gation of non-thermal discharges.
5 Surface ionization (electron emission) is provided by electron, ion,
and photon collisions with different surfaces or simply by surface heating.
This ionization mechanism is quite different from the first four.
87
1 e 4 1 1 2 v 1 1
i ( _ ( )) (126)
(4 0 ) 2 I 3 I2 2
The Thomson formula (124) agrees with (123), assuming the valence
electron is at rest and εv =0. An interesting variation of the Thomson for-
mula (124) can be obtained assuming a Coulomb interaction of the va-
lence electron with the rest of the atom and taking εv = I:
1 e 4 5 1 2 I
i ( ) (127)
(4 0 ) 2 3I 3 2
All the modifications of the Thomson formula can be combined us-
ing the generalized function f (ε/I) common for all atoms:
1 e 2
i Zv f ( ) (128)
(4 0 ) 2
I 2
I
88
energy of preliminary electronic excitation of neutrals can be converted in
the ionization act, which is called stepwise ionization. If the level of elec-
tronic excitation is high enough, stepwise ionization is much faster than
direct ionization, because the statistical weight of electronically excited
neutrals is greater than that of free plasma electrons. In other words, when
Te « I, the probability of obtaining high ionization energy I is much lower
for free plasma electrons (direct ionization) than for excited atoms and
molecules (stepwise ionization).
In contrast to direct ionization, the stepwise process includes several
steps to provide the ionization event. At first, electron–neutral collisions
prepare highly excited species, and then a final collision with a relatively
low-energy electron provides the actual ionization event.
In thermodynamic equilibrium, the ionization process
+
e+A→A +e+e is inverse to the three-body recombination
A++e+e→A∗+e→A+e, which proceeds through a set of excited states.
According to the principle of detailed equilibrium, the ionization process
e+A→A++e+e should go through the set of electronically excited states
as well, which means the ionization should be a stepwise process. How-
ever, this conclusion can only hold true for quasi-equilibrium or thermal
plasmas.
A description of the stepwise ionization and the calculation of rate
coefficient can be found in Fridman (2008) [1]
89
cies about ωtr =ΔE/ provide the energy transfer between the interacting
particles. The relative weigh to probability of these fast Fourier compo-
nents is very low if ωtr » ωint; numerically it is about exp(−ωtr/ωint). As a
result, the probability PEnTr of energy transfer processes (including the
ionization process under consideration) are usually proportional to the so-
called Massey parameter:
tr E
PEnTr exp( ) exp( ) exp( PMa ) (129)
int
where PMa E / is the adiabatic Massey parameter. If PMa »1,
the process of energy transfer is adiabatic and its probability is exponen-
tially low. It takes place in collisions of heavy neutrals and ions.
90
colliding particles with an electronic energy term of the molecular ion
AB+ and, as a result, the process is non-adiabatic. Such a situation takes
place only for a limited number of excited species.
91
ble to accumulate electron recombination energy fast enough in their ki-
netic energy and, therefore, are ineffective as the third- body partner.
3 Finally, the recombination energy can be converted into radiation
in the process of radiative electron–ion recombination:
e A A* A (134)
The cross sectionof this process is relatively low and can com-
petewith the three-body recombination only when the plasma density is
not high.
92
Fig. 33. Elementary process of dissoci- Fig. 34. Elementary process of disso-
ative attachment in the case of low ciative attachment in the case of high
electron affinity electron affinity
Electron collisions with two heavy particles also can result in the
formation of negative ions:
e A B A B (136)
The three-body electron attachment can be a principal channel for
electron losses when electron energies are not high enough for the disso-
ciative attachment, and when pressure is elevated (usually more than 0.1
atm) and the third-order kinetic processes are preferable.
In contrast to the dissociative attachment, the three-body process is
exothermic and its rate coefficient does not depend strongly on electron
temperature. Electrons are usually kinetically less effective as a third
body B because of a low degree of ionization. Atmospheric-pressure non-
thermal discharges in air are probably the most important systems, where
the three-body attachment plays a key role in the balance of charged par-
ticles:
e O2 M O2 M (137)
This process proceeds by the two-stage Bloch-Bradbury mechanism
(Bloch&Bradbury, 1935; Alexandrov, 1981) starting with the formation
93
of a negative ion (rate coefficient katt) in an unstable auto-ionization state
(τ is the time of collisionless detachment):
e O2 (O2 ) *
k
att ,
(138)
The second stage of the Bloch-Bradbury mechanism includes colli-
sion with the third-body particle M (density n0), leading to relaxation and
stabilization of O−2 or collisional decay of the unstable ion:
(O2 ) * M
k st
O2 M (139)
*
(O ) M O2 e M
2
kdec
(140)
Taking into account the steady-state conditions for number density
of the intermediate excited ions (O−2)∗, the rate coefficient for the total at-
tachment process (138) can be expressed as
k att k st
k 3M (141)
1
(k st k dec )n0
Usually, when the pressure is not too high, (kst +kdec)n0«τ−1, and
(141) can be simplified
k3M k att k st (142)
Thus, the three-body attachment (137) has a third kinetic order and
depends equally on the rate coefficients of formation and stabilization of
negative ions. The latter strongly depends on the type of the third particle:
the more complicated molecule a third body (M) is, the easier it stabilizes
the (O−2 )∗ and the higher the rate coefficient k3M. Numerical values of the
total rate coefficients k3M are presented in Table 5; dependence of the total
rate coefficients on electron temperature is shown in Fig.35. For simple
estimations, when Te =1eV and T0 =300K, one can take k3M≈10−30
cm6/s.
94
Fig. 35. Rate coefficients of electron attachment to different molecules in
three-body collisions as a function of electron temperature; molecular gas is
assumed to be at room temperature
The rate of the three-body process is greater than for dissociative at-
tachment (ka) when the gas number density exceeds a critical value n0
>ka(Te)/k3M. Numerically in oxygen, with Te =1eV,T0 =300K, the rate re-
quires n0 >1018 cm−3, or in pressure units p>30 torr. Three other mecha-
nisms of formation of negative ions are usually less significant. The first
process is polar dissociation
e AB A B e (143)
95
is not above 0.1 eV; the population of the first vibrational level of the
ground electronic state predominates over the others; reactions of the fol-
lowing type: e M , M M , e M etc. are practically absent; reaction prod-
ucts are not usually excited; elastic collisions predominate over non-
elastic ones. In general molecules may be represented as hard elastic
spheres.
In the field of interest (plasma chemistry, radiation chemistry, etc.)
the situation is quite different: mean molecule energies are 0.1 eV; ex-
cited molecules are frequently in collision, and chemical reaction prod-
ucts produced by collisions appear often to be in excited states; collisions
of molecules with electrons and ions (as well as collisions between mole-
cules) are essential, and there is usually wide difference between transla-
tional rates and energy (in the laboratory coordinate system) of the
charged particles of small mass (~1-50 eV) and those of heavy particles
( ~ 0.03-2.3 eV), as well as in the vibrational energy of the latter
(~0.2-0.8 eV). The model of elastic spheres is therefore inapplicable [32].
The kinetics of chemical reactions describes the particular type of
molecule ensemble behaviour, and this description has a statistical char-
acter. Because of this, the kinetic description of ensemble behaviour of
various particles is defined, assuming the dynamics of collisions (for in-
stance binary ones) to be known, by molecule quantum level populations
and the function for particle distributions by translational energy as well
as by ratios of flux probabilities over various channels, for instance
M (i) M ( j ), M (i) M ' (k ),
where i, j, k is the totality of molecule quantum state parameters.
96
time with various relaxation times. The transitions of M (i) M ( j ), type can
no longer be neglected in chemical reaction descriptions; non-elastic col-
lision frequencies are not small by comparison with elastic collision fre-
quencies.
Naturally, the chemical reactions possible in a molecule, and their
rates, depend on the structure of the molecule (and for non-
monomolecular reactions also on other reaction component structures).
This, in particular, prevents the introduction of the term 'reactivity', with
respect to molecules. The concept of molecular structure can be formulat-
ed differently. For our purpose the following definition is believed to be
the best: the structure of a molecule, comprising several atoms, is a sys-
tem of its quantum levels and space distribution of the component parti-
cles.
The problem of connecting reactivity with molecular structure is of-
ten considered of paramount importance in 'classical' chemical kinetics.
However, it cannot be solved within the limits of the molecular model
used. Automatically, according to the character of the reactions described
using generalized chemical kinetics, the problem is being solved within
the latter. This, however, takes place when the vague term 'reactivity' is
substituted for ki.
2. Plasma-chemical reactions may be of the non-equilibrium or qua-
si-equilibrium types. Various types of plasma-chemical reactions are
enumerated in Table 5.
occurrence temperatures
97
of chemical reactions Both subsystems are
non-Maxwellian (and
non-Boltzmannian)
98
lational, vibrational, and rotational temperatures differ and/or various
components of a system (electrons, ions, and neutral molecules, for in-
stance) have different temperatures, and/or a system cannot be described
at all using the concept of temperature * (non-equilibrium, stationary, and
relaxing systems).
Strictly speaking, Arrhenius-type kinetics cannot be used in these
cases and the ordinary expression for the rate of a chemical reaction is in-
applicable.
4. Multichannel processes should be considered in plasma-chemical
kinetics. From the quantum mechanical point of view (we shall consider
only two channels for simplicity), in a certain energy range two pairs of
particles (A1, B1) and (A2, B2) can exist, so that there are two independ-
ent wave functions which satisfy boundary conditions for a given system.
It is known that in a one-channel problem the S-matrix contains all the in-
formation about interaction properties of the system, while in a multi-
channel case a similar theorem has not yet been proved6. Moreover, in
order to compose the Hamiltonian using the results of scattering all com-
ponents of the S-matrix should be known for all energies.
5. The barrier-type nature of the reactions (the existence of an energy
barrier opposing the reaction), being peculiar to the Arrhenius classical
kinetics, can be completely changed in reactions under plasma conditions.
It is necessary to allow for the occurrence of chemical reactions from dif-
ferent quantum levels of the system, and for one of the subsystems as a
whole to have above-barrier energy.
6. These five special features of plasma-chemical reactions are pecu-
liar to non-equilibrium chemical processes, and chemical reactions under
low-temperature plasma conditions are particular (but still important) cas-
es of such processes. In this section some special features of reactions un-
der plasma conditions are discussed briefly. The degree of ionization is
the most important characteristic of a plasma under equilibrium condi-
tions; given the temperature and pressure the degree of plasma ionization
may be found using the Saha equation.
Ionization processes result from such factors as mutual collisions of
heavy particles at high temperatures (energies), their collisions with elec-
trons, photo-ionization, and ion-molecular reactions. A detailed descrip-
tion of all these processes under non-equilibrium conditions necessitates a
generalization of the kinetic gas theory for the plasma state. Such a gener-
alization is still under development. In low-temperature plasmas long
range collisions occur, due to electromagnetic interaction between
charged particles, in addition to ordinary short-range collisions, such as
take place in gases. Thus, very small scattering angles and consequently
99
many interactions resulting in small momentum transfer must be consid-
ered. Moreover, interaction of charged particles with external electro-
magnetic fields should also be taken into account [34,35,36]
When electron, ion, and molecular gases are not in thermal equilibri-
um there is energy exchange between electrons and neutral particles and
between electrons and ions. The latter occurs through Coulomb-type col-
lisions. Both processes may lead, either directly or through successive
steps of excitation, to chemical reactions. The mechanisms and probabili-
ties of these reactions may be studied by plasma-chemical kinetics on the
basis of a detailed analysis of all interactions occurring in a system, and
using the basic principles of non-equilibrium chemical kinetics. At pre-
sent the limited usefulness of Arrhenius kinetics, which are valid only
close to equilibrium (for example under small perturbations when a sys-
tem can still be considered in quasi-equilibrium), is evident. In other
words, this type of kinetics can be used when a single value of tempera-
ture (being the parameter of the Maxwell-Boltzmann distribution) can be
defined for the system.
Non-equilibrium distribution of reactants takes place, for instance,
when the energy is injected pulsewise into the equilibrium system, pro-
vided that the condition i Ch.r is satisfied, where i is the duration of
the pulse and Ch.r is the chemical relaxation time (pulse electrical dis-
charges, shock tubes, flash photolysis). The treatment of kinetic problems
which depend more upon the molecular interaction mechanism and mo-
lecular quantum level populations* requires far more information about
the system [37,38,39]. The importance of the effect of vibrational excita-
tion upon the rate of chemical reaction has been confirmed by the results
of experiments[40]. As the vibrational temperature of N2 molecules is in-
creased from 1000 K to 6000 K (under Ttrans = 300K) the rate constant for
the reaction O N 2 N 2 O increases by forty times (see Fig. 36). The
same change in the translational temperature (under Tvibr = const) would
lead, in accordance with the Arrhenius equation, to a rate constant in-
crease of sixty times. The effects are comparable.
100
Fig.36. Experimental rate constant dependence on vibrational level num-
bers up to 10
101
be studied for non-equilibrium, generalized chemical kinetics. The Arrhe-
nius kinetics appears to be an extreme case and thus are believed to be
rigorously founded, their applicability limits being clearly defined. The
solution of this problem must be based upon a consideration of the Pauli
equation or the Boltzmann equation (or any equation of the Liouville
type). Using one of these equations and taking into account chemical re-
actions and relaxation processes of the internal degrees of freedom of re-
acting molecules, specific solutions under various conditions can be ob-
tained, including ordinary chemical kinetic equations and expressions for
the rate factors (constants) for chemical reactions.
Three solutions to the problems under discussion may be suggested,
allowing for the possibility that a set is found of cross sections ( ) for
the processes involved; all the required information about the system can
be obtained through averaging.
102
al, vibrational and translational degrees of freedom about the ground state
of the molecule. At low energy these losses do not exceed 17% and are
reduced to 10% with increasing electron energy (curve 5). As a result, the
energy of the electron beam consumed on the decomposition of one mol-
ecule of the original gas mixture is significantly (10-100 times) higher
than its dissociation energy.
In case of high-current beams a significant role is played by the elec-
tromagnetic fields and collective electrodynamics processes which have
an affect both on the interaction of an electron beam with gases, and on
the kinetics of plasma chemical processes taking place.
Fig. 37. The energy distribution of the electron with energy Te on ioniza-
tion levels and excitation in nitrogen: 1 – total loss on excitation, 2-total loss
on ionization, 3 - loss pn excitation of electronic levels, 4 - loss on dissociative
ionization, 5 - other losses
103
Fig. 38. Redistribution of the energy spent on ionization and ohmic losses
of the electron beam at different degrees of freedom of molecular gas plasma:
ionization (1), dissociation (2), electron (3) and vibrational (4) excitation and
heating of the gas (5). Solid lines - without electric fields, dotted line - in the
presence of electric fields
104
Chapter 4 Chain gas-phase processes in the external action
Over the past 30-40 years in Russia and abroad extensive studies on
gas-phase chemical processes initiated by the gas discharge have been
carried out. It was found that some favorable conditions for the occur-
rence of chain processes are formed when the gas discharge has an effect
on gas-phase medium. With a sufficient length of the chain an electro-
physical installation makes it possible to reduce the total energy con-
sumption required for the chemical process. The primary source of energy
in this case is the thermal energy of the initial reactionary gas or the ener-
gy of the exothermic elementary reactions of chain process (e.g., reactions
of oxidation or polymerization). This can greatly reduce the power inputs
of the electrophysical installation. The reduction of the temperature of
chain chemical processes at radiation exposure is similar to the catalytic
effect. But the chain process can occur entirely in the gas phase, which
greatly increases the reaction rate compared to the hetaeraphase catalytic
process. Carrying out a chemical process at a temperature below the equi-
librium allows one to synthesize compounds which are unstable at higher
temperatures or the selectivity of their synthesis at higher temperatures is
low. These peculiarities of the behavior of chemical processes under
plasma action have shown to be a quite promising in the use of large-scale
chemical production. A high speed of the reaction which is required for
industrial technologies is achieved by using of branched chain processes.
But their major drawback is associated with the occurrence of an explo-
sive process, which significantly increases the risk of production. This
disadvantage is eliminated by initiating the chain process under external
action out of the self-ignition region.
105
at temperatures above 1200 K. The continuation of chain consists of the
interaction of free radicals with the initial molecules resulting in the for-
mation of a stable molecule of the reaction product and a new radical. The
new radical, in turn, interacts with the initial molecules. This stage of the
chain process requires much lower activation energy (0.87 eV/molecule
for CH4), therefore, its implementation requires a lower temperature. It is
important to note that the major part of chemical reactions occur at the
stage of the continuation. An active radical that is formed at the initiation
stage can initiate 103 - 105 of the reactions at the stage of the chain pro-
cess.
Fig. 39 shows the change in energy consumption for conducting the
chemical transformation at different stages of the chain process of radia-
tion-thermal cracking of methane.
Fig.39. The chart illustrating the power inputs at stages of chain process
of methane conversion
106
chain reactions with degenerate branching, chain reactions with energy
branching and others.
107
4.1.2 Chain reactions with quadratic branching
108
4.1.4 Electron-induced chain reactions
yields OH radicals, HO2 and the electron, which can repeatedly par-
ticipate in their formation, obtaining the necessary energy from the elec-
tric field.
109
Fig. 40. The curves of the oxidation kinetics of the mixture of 2H2 + O2 at
485°C and at various initial pressures of: 1 - 8.2 Torr; 2 - 7.8 Torr; 3 - 7.4
Torr; 4 - 7.1 Torr; 5 - 6.8 Torr; 6 - 6.4 Torr; 7 - 6.1 Torr; and 8 - 5.8 Torr
Fig. 41. Relationship between the induction period and the pressure of
mixture of 2H2 + O2 at different temperatures: 1 - 435 °C; 2 - 445 °C; 3 - 458
°C. Curves are calculation
110
Fig. 42 shows the calculated dependences of the induction on the ini-
tial temperature of the stoichiometric oxygen-hydrogen mixture at a pres-
sure of 1 kPa in the case of a 1 ms laser irradiation and different values of
radiation density.
Fig. 42. The induction period versus temperature with the laser radiation
energy of 0 (1), 0.017 (2), 0.082 (3) and 0.15 (4) eV /molecule. Curve 5 - The
diffusion time of hydrogen from the excitation region versus temperature
111
Fig. 43. Formation of the chain mechanism by ignition of H2 + O2 + H2O
mixture without the excitation of H2O molecules (a) and with the excitation of
asymmetric vibrations of H2O caused by the radiation with λ = 2.66 µm (b)
112
eV/molecule (absorbed dose is 260 kGy) at a temperature of 300 K the
induction period was estimated to be 0.2 according to the data in [2].
Fig. 44. The change in pressure (initial part) in the reactor filled with a
mixture of 2H2 + O2 at different initial pressures of the mixture. The initial
temperature is 300 K
Fig. 45. The dependence of the induction period on the initial pressure of
the gas mixture for different volumes of the reactors: 1.6 liters (1) and 3.2 liters
(2)
113
The dependence of the induction period on the pressure has a shape
which is peculiar to chain processes – an increase in the induction period
near the combustion region (see Figure 40.). Increasing the reactor vol-
ume reduces an inhibitory effect of the reactor walls on the development
of the chain process, which increases the speed of the chain process and
thus causes a decrease in the induction period (curve 2 in Fig. 45). This
confirms that under the influence of a pulsed electron beam on the stoi-
chiometric oxygen-hydrogen mixture the hydrogen oxidation occurs as a
chain process.
To compare the effectiveness of different types of impact on the ini-
tiation of ignition of oxygen-hydrogen mixture the values of the induction
period for similar conditions are shown in Table 6. At high temperatures,
when the ignition of oxygen-hydrogen mixture is possible even without
an external action, a laser, or ionizing radiation causes a sharp accelera-
tion of the ignition and reduction in the induction period. The initiation of
the ignition occurs at low temperatures, under the external action, with a
low-energy input, which does not exceed a few percent of the dissociation
energy of the original molecules.
Table 6. The values of the induction period under different external effects
Energy
Tem- Induc-
Type of exter- Mixture of gas- Pres- input,
pera- tion
nal effect es sure eV/moluc
ture period
ule
No effect 2H2 + O2 7.6 Torr 730К 0 0.15 s
No effect 2H2 + O2 25 Torr 730К 0 0.2 s
Laser, λ = 762
2H2 + O2 7.6 Torr 300 К 0.082 0.2 s
nm
Laser, λ = 762
2H2 + O2 7.6 Torr 600К 0.082 0.01 s
nm
Laser, λ = 762 100
2H2 + O2 400К 0.082 0.2 s
nm Torr
Laser, λ = 2H2+O2 + 5%
7.6 Torr 300К
2.66 µm H2O
Pulsed
5%H2+air 76 Torr 770К 0 103 s
discharge
Pulsed 5∙10-4
5%H2+ air 76 Torr 770К 0.23
discharge s
Pulsed 5%H2+ air 7.6 Torr 400К 2.3 0.1 s
114
discharge
380
Continuous
Hydrogen- Torr 0.003
ionizing 400К 2s
aerial mixture 0.05 (20 kGy)
radiation
Mpa
Pulsed elec- 100 0.013 (114 0.003
2H2 + O2 300К
tron beam Torr kGy) s
115
Fig. 46. The limits of ignition of a stoichiometric of hydrogen-oxygen mix-
ture in a spherical vessel with a diameter of 7.4 cm, covered with KCl. The first
and third limits are extrapolated (dashed line)
116
temperature of ignition at the first limit should reduce to 100-140 degrees
(see Figure 47).
Fig. 47. The displacement of the ignition limits at the initial mole fraction
of hydrogen atoms: solid curve is [Н]0 = 0, points is [Н]0 = 2 10-3
With the radiation output of hydrogen atoms equal to 2.14 per 100
eV, for the production of 1.2∙1021 of hydrogen atoms in 1 mole, the ener-
gy of electron beam of 1350 J is required. For the mass of gas of 2.12 g in
one mole, the absorbed dose is 110 kGy. This value of the absorbed dose
of electron beam energy is consumed only for the decomposition of hy-
drogen molecules. In a mixture of oxygen and hydrogen the injected elec-
tron beam is scattered not only by the hydrogen molecules but also the
oxygen.
To estimate the total absorbed dose let us take into account that the
energy loss of an electron with a kinetic energy of 0.5 MeV in hydrogen
is 4.2 4.2 МeV·cm2/g, and in oxygen is МeV·cm2/g [52]. Then, for a stoi-
chiometric mixture of oxygen and hydrogen the electron with an energy
of 0.5 MeV 3.4 more loses energy on 1 cm path due to the scattering by
oxygen molecules than due to the scattering by hydrogen molecules.
Therefore, the electron beam absorbed dose required for the production
the molar fraction of hydrogen atoms of 2·10-3 in the stoichiometric mix-
ture of hydrogen and oxygen is 480 kGy. In [3, 4], the experimental stud-
ies on the initiation of ignition of the stoichiometric oxygen-hydrogen
mixtures by a pulsed electron beam were presented. For this investigation
a TEU-500 high-current electron accelerator was used. The studies of the
ignition of 2H2 + O2 mixture thought the pulsed electron beam irradiation
showed that with the injection of electrons (the absorbed dose is not more
117
than 114 kGy) the ignition of the gas mixture occurs in a room tempera-
ture. The temperature bias, as a result of the pulse radiolysis, for the first
limit of ignition is much greater than the calculated values given in [13]
where it was estimated by a numerical simulation method. When the ini-
tial mole fraction of hydrogen atoms equals to 0.002 (absorbed dose of
electrons is 480 kGy), the lower temperature limit of ignition should de-
crease to 100-140 degrees and be equal to 600K.
To compare the effectiveness of different types of impact on the dis-
placement of the ignition limits both the experimental and theoretical data
for similar conditions are shown in Table 7.
118
The ionizing radiation helps reducing the limits of ignition more ef-
fectively than the laser radiation. With an initial pressure 100 Torr in the
mixture of 2H2 + O2 under the influence of resonant laser radiation (λ =
762 nm) with an energy of 0.08 eV/molecule the ignition temperature
limit is shifted to 400K. With the same initial pressure, in the case of the
pulsed electron beam radiation with an energy of 0.013 eV/molecule the
ignition of 2H2 + O2 mixture at temperatures below 300K occur.
The use of the electron beam to initiate the ignition of oxygen - hy-
drogen mixture creates a unique environment for a study of the originali-
ty, development and termination of t combustion processes. This is be-
cause the duration of the electron beam effect is much smaller than the
induction period. The change in pressure in the mixture does not affect
the absorbed dose (at a low-pressure) provided that the density of electron
beam is constant. Furthermore, the change in the conditions of chain is
possible by introducing the atomic hydrogen into the reactor volume. The
initiation of ignition of oxygen - hydrogen mixture in gas discharge pro-
vides a high uniformity of the initial concentration of the atoms and radi-
cals in the reactor volume (ideal mixing).
119
exposure the active centers which are free radicals, may occur, which can
start a chain reaction of hydrocarbon cracking. The interaction of the rad-
icals, which contain a small number of carbon atoms, with the initial hy-
drocarbon is a distinct chain process, which is initiated by the radicals
formed due to the radiation.
In a nook of V.V.Sarajeva [54] there are four temperature regions of
the radiation-thermal cracking, they are shown in Fig. 48.
The first region is radiation, at which the radicals are generated only
by the radiation. In this region, the process is not a chain, and the activa-
tion energy is only 0.04 eV/molecule, which characterizes the radiolysis
process.
The second region is the thermal radiation. Within this temperature
range the process becomes a chain with activation energy of 0.87
eV/molecule.
RH + e (3.5 эВ) → R* + H*
R* + RH + 0.87 эВ→ stable product + R1 *
The third region is the thermal radiation. It occurs when the R1 radi-
cal is thermally unstable and can decompose into two active radical due to
the reaction of
R1* → R2* + R3*
In this temperature range it is branched chain process (the activation
energy is of the order of hundreds of kilojoules per mole).
The fourth region is thermal. Here the generation rate for the active
radicals is determined by the reaction
120
RH + tº (3.5 e) → R* + H*
The generation rate increases due to the radiation. The activation en-
ergy in this temperature range is determined only by the thermal process
and is equal to 3.5 eV. The speed of the chain process depends on the rate
of radical generation, which also depends on the dose. The influence of
the temperature and dose on the development of the chain process in de-
scribed in more details in [16] as an example of radiation oxidation of hy-
drocarbons.
Based on the analysis of experimental data on the radiolysis of liq-
uid hydrocarbons of the paraffin series it was obtained that:
1 The reaction yield of radical formation in γ-radiolysis, and a tem-
perature of 300 K is 7-10 radicals. Assuming that the average energy is
3.5 eV, it is easy to see that the chemical reaction consumes no more than
30% of the absorbed energy of ionizing radiation, and the rest of the en-
ergy is spent on heating (by collisions or energy transfer).
2 The yield of the radiolysis products at 300K contains about 80%
of the products of C - H bond breaking. It results in formation of molecu-
lar hydrogen and hydrocarbon, and about 20% are the products of C - C
bond breaking.
3 With an increase of the temperature above 600 K the radiation-
chemical yields rise dramatically, reaching 1.5 104 molecules per 100 eV
at 600 K, and the product composition approaches the composition of the
products peculiar to the thermal cracking.
Chain process of the decomposition of methane (radiation-thermal
cracking) can be described by the following basic equations [55]:
CH4 + 3.5 эВ = (CH3)* +
Chain initiation due to the thermal in-
H*
fluence
Chain initiation due to the radiation in-
CH4 + e = (CH3)* + H*
fluence
Continuation of the chain CH4 + H* = (CH3)* + H2
(CH3)* + CH4 = C2H6 + H*
(CH3)* + H* = CH4
Termination of the chain
(CH3)* + (CH3)* = C2H6
H* + H* = H2
In [56, 57] the experimental data on plasma catalysis in the conver-
sion of methane into carbon and hydrogen were presented. Preheated to a
temperature of 700 - 1100K methane was treated for a short time pulse
microwave discharge (frequency 9 GHz, pulse power up to 100 kW, pulse
length 1 ms, repetition rate 1 kHz), which caused a sharp increase in its
degree of conversion. It was shown that this effect can not be explained
121
by the thermal influence of discharge and the role of plasma is to generate
the active particles, which accelerates the conversion process. Fig. 49
shows the experimental dependence of the degree of conversion of me-
thane in the microwave discharge plasma on the temperature.
Fig. 49. The dependence of the degree of methane conversion on the tem-
perature (a): 1 - decomposition of methane without the influence of plasma, 2,
3, 4 - influence of plasma on gas, preheated to 950, 850 and 750K, respectively.
(b)The dependence of energy consumption for the formation of hydrogen mole-
cule on the plasma energy
122
the initiation reaction and, for this reason, the disappearance of the rest of
the chain. Thus, if it were possible to carry out the initiation of free radi-
cals at a lower temperature than the temperature of the low boundary of a
slow oxidation region, then the oxidation and further conversion of hy-
drocarbons could be continued even in these conditions.
With this oxidative conversion of hydrocarbons at low temperatures,
the synthesis of new compounds may occur. In a complex chain process
the radicals can enter not into one reaction but also in two or more reac-
tions. The temperature influences on the outcome of the radical reactions.
The lower the temperature of the process is, the smaller activation ener-
gies is required for radical to determine the preferred direction of the re-
action. For example, at 400K the difference between the activation ener-
gies of two reactions of 0.04 eV/molecule leads to the fact that 93% of the
response will take the path with a lower activation energy (provided that
cross the sections for interaction are equal).
The oxidation at the temperatures lower than the normal tempera-
tures in oxidation, should result in a significant decrease in the quantities
of hydrocarbons which decompose and oxidize. This follows from the
fact that out of two possible reactions of alkyl radicals - the decomposi-
tion and the oxidation
1 The decomposition of R → products of decomposition, the activa-
tion energy of 1.1-1.7 eV/molecule.
2 The oxidation of R + O2 → RO2 the activation energy of 0.02-
0.09 eV/molecule at temperature of the oxidation of hydrocarbons of 450
K. We can assume that the cracking does not occur.
The second consequence of lowering the temperature of the reaction
is the formation a significant amount of peroxides as a result of conver-
sion of peroxide radicals RO2 .Here there are two ways: the disintegration
of RO2 radicals with the formation of aldehydes and alcohols, and its in-
teraction the initial hydrocarbon with the formation of the alkylhydroper-
oxides. The activation energy of decomposition is about 0.8 - 0.9
eV/molecule. The energy of activation of the bimolecular interaction of
RO2 with hydrocarbon does not exceed 0.2-0.3 eV/molecule. For a mix-
ture of 150 Torr RH + 150 Torr O2 at 450 K the ratio of the rate of the de-
composition reaction of RO2 to the reaction of alkylhydroperoxides for-
mation is 0.53. In the case of the oxidation at 400K, this ratio is equal to
0.043. Thus, the reduction in the oxidation temperature should facilitate
the formation of the alkylhydroperoxides and cause a decrease in the yield
of aldehydes and alcohols, which are, according to the mechanism, the
decomposition products of the RO2 radicals.
123
One can also assume that at a low temperature of oxidation the al-
kylhydroperoxides will have an isometric structure rather than the normal
structure. The difference in the energy of a bond breakage of the primary
and the secondary C-H bonds is about 0.17 eV/molecule. This difference
will lead to the fact that at low temperatures the separation of the hydro-
gen atom from CH2group (rather than CH3) will occur in hydrocarbon and
the peroxide formed will have an isomeric structure.
In general, it is expected that a significant reduction in the reaction
temperature causes a change in the direction of oxidative conversion. The
alkylhydroperoxides will be mainly formed. The reduction in the tem-
perature of oxidation can be achieved by facilitating the generation by
primary free radicals. This can be done either by supplying the active rad-
icals from the outside, or addition of certain substances that are able to
decompose to radicals at low temperatures (i.e. below the boundary of the
normal region slow oxidation of hydrocarbons).
Fig. 50. The dependence of the oxidation time of methane on the mixture
pressure at temperature of 650 K and the region of ignition of hydrocarbon-air
mixtures. 1 – 13% CH4, 2 – 10 % C2H6
124
Fig. 51. The kinetics of methane oxidation obtained by: changing the total
pressure, consumption of the initial products and the accumulation of the end
products of the reaction.
125
equilibrium at low temperature is characterized by a low degree of con-
version. The oxidation process is terminated long before the formation of
the original products (see Fig. 51, lines 1 and 2).
126
peroxides and aldehydes formation are similar to the curves obtained by
the oxidation of propane.
The study of photochemical oxidation of hydrocarbons at low pres-
sures showed that:
1 The facilitation of the initiation enables the oxidation at low tem-
peratures
2 The main product of the low-temperature oxidation is the corre-
sponding alkylhydroperoxides
3 If the conversion is very low, the oxidation process is slow as well.
A small chain length in these conditions makes, however, the low-
temperature oxidation of hydrocarbons ineffective with respect to the
yields of products and the degree of conversion.
In 1958, the oxidation of methane (a mixture of 4SN4 + O2 and CH4
+ O2) under the action of a fast electron beam at room temperature was
studied by B.M. Mikhailov and others [61]. The dose rate (at Pinit = 760
Torr) was 2,7·1016 eV/cm3·s (for the mixture of CH4 + O2 the dose was 4
kGy/s). The source of fast electrons was an electron tube with an acceler-
ating voltage of 120 kV and an output current of 100 µA. The experi-
ments were conducted with a steel vessel, the irradiation time was 30 min.
The reaction yield for the mixture CH4 + O2 was equal to 7 molecules of
O2 and 6.2 molecules of CH4 at 100 eV. Fig. 54 shows the change in pres-
sure in the irradiated methane-oxygen mixture during the reaction.
127
A sharp increase in pressure for 20-30 seconds in the beginning of
the curve in Fig. 16 is associated with heating. Later the pressure curve
dpops during 3-4 min, then straight-line section can be obsereved, which
lasts until one of the initial components of the mixture is used up. After
that, the pressure drop slows down. It was found that the reaction rate in-
creases with an increase of the partial pressure of methane and the total
pressure. The consumption of methane and oxygen is proportional to the
time of irradiation. It is important to note that the rate of oxidation of me-
thane in the case of the irradiation by the electron beam is much higher
than under equilibrium conditions, although the temperature of the gas
mixture was significantly lower (300 K at the irradiation, 723K under
equilibrium conditions).
At room temperature, and initial pressure of 760 Torr the methane
mixture was subjected to transformation by 32 % (a mixture of CH4 +
O2). This greatly exceeds the degree of conversion in the mercury sensi-
tized photochemical initiation (0.5 % for propane). The products of oxida-
tion found are CO and CO2, H2O, H2, HCOOH and a small amount of
CH3OH and peroxides. The formaldehyde was found only in trace
amounts (see Fig. 55).
Fig. 55. The dependence of the oxidation products of methane on the du-
ration of electron beam irradiation. CH4 + O2 mixture: Pinit = 760 Torr., T =
300 K. 1 – H2O, 2 – H2, and 3 – HCOOH, 4 – CO and CO2, 5 – peroxides and
alcohols
128
oxidation, the degree of conversion is higher (32% with the irradiation
and 0.5 % under equilibrium). The authors believe that when bombarded
with the fast electrons 50% of the exposed methane is converted to the
excited molecules, and 50 % to the ions.
In [62, 63]] the partial oxidation of methane by the two types of mi-
crowave discharge was studied. These discharges are: the pulse-periodic
discharge (streamer pseudo-corona discharge, the wavelength of 3 cm,
pulse power of 300 kW, average power of 300 W, pulse duration of 1 ms,
repetition rate of 1 kHz) and the continuous discharge (coaxial jet dis-
charge, the frequency of 2.45 GHz, the power of 1-5 kW). The initial re-
actants were heated up to 800-1200 K and supplied into the discharge
chamber, which were combined with the methane combustion zone. The
authors note that the use of the microwave discharge acts on the system in
two ways. Firstly, it effectively introduces an additional thermal energy
even in strongly heated reagents due to the high temperature of plasma.
Secondly, the plasma generates the active particles which contribute to
the oxidation of methane in the chain reactions, and play a role of the ini-
tiator of the combustion. For the reaction of partial oxidation of methane
the energy consumption of the microwave discharge were 0.25
eV/molecule when the degree of conversion methane was 70 %, and in-
creased to 0.5 eV/ molecule with the degree of conversion about 100 %.
In this case the energy input due to thermal heating (according to the de-
gree of conversion) was 2.6 -2.8 eV/molecule. The heating of the mixture
of gases has greatly increased the length of the chain reaction of oxidation
and with the inexternal initiation of the chain process significantly in-
creased the degree of methane conversion at the given temperature.
129
For the experiments the model gas mixtures of nitrogen and oxygen
were used. The content of nitrogen N2 was ranged from 89 to 99 % and
oxygen from 0.1 to 10 %. The content of the investigated impurities CS2
varied from 0.005 to 1 %. For the irradiation of the gas mixtures two elec-
tron accelerators with different parameters were used. A compact “Ra-
dan”accelerator was used in a nanosecond time scale. It generated the
electron beam with the following parameters: the electron energy was 180
kV, the half-hight pulse duration of 3 ns, the current density of 800
A/cm2, the pulse repetition rate of 10 Hz. In some experiments the non-
self-sustained discharge, initiated by a nanosecond electron beam was ob-
served. The discharge was ignited in a gap with a length of 1 cm when a
12 nF capacitor, charged to 1-25 kV was connected to the gap. In the mi-
crosecond time scale an accelerator with a plasma-cathode was used. This
accelerator formes a radially divergent electron beam with the following
parameters: the area of ~ 1.5 m2, the electron energy of 280-300 keV, the
current density raged from 0.1 to 10 mA/cm2, the pulse duration of 48 ms.
The volume of 170 liters was subjected to the irradiated. In the same vol-
ume the nonself-sustained discharge maintained by an electron beam was
ignited in the gap of 10 cm in length. The maximum field strength in the
volume discharge plasma was 5 kV/ cm.
In a mixture irradiated by a nanosecond electron beam the initial
concentration of CS2 has a little effect on the angle of the dependence of
concentration of the carbon disulfide molecules on the number of irradia-
tion pulses. The minimum value of energy consumption was equal to 0.8
eV/molecule, which is ignificantly less than the energy of dissociation of
carbon disulfide, equal to 7.6 eV. The conversion process of CS2 in the
nonself-sustained discharge initiated by a nanosecond beam was also in-
vestigated. Fig. 56 shows the dependence of the energy consumption on
the electric field strength in the volume discharge. It was found that in a
range of the electric field strength from 0 to 1.5 kV/cm, the value of ener-
gy varies slightly and is ranged from 0.6 to 0.9 eV/molecule. This is much
lower than the dissociation energy of carbon disulfide and oxygen.
130
Fig. 56. The dependence of the energy required for the removal of one
molecule of the impurity on the field strength in the column of the nonself-
sustained discharge initiated by a nanosecond electron beam
131
(152). By reducing the concentration of CS2 below a certain critical value
the chain mechanism is interrupted, the atomic oxygen ceases to turn out
in the reactions (152), (153) and the oxidation of carbon disulfide is ter-
minated. According to the calculations, the main products of decomposi-
tion of CS2 are sulfur oxides SO2 and carbon oxides CO. In small quanti-
ties NO, SO and S2 were produced, which was observed in the experi-
ments. The presence of nitrogen in a mixture leads to the loss of radicals
O in the reactions of nitrogen oxide synthesis:
N2(ν) + O → NО +N (154)
N + O2 → NО + О
Here N2(ν) is the vibrationally excited nitrogen molecules.
Another mechanism for the conversion of carbon disulfide is imple-
mented with the excitation of N2-O2 -CS2 mixture by a low-current elec-
tron beam of microsecond duration. In this case, the beam does not have
enough intensity and the density of plasma electrons is two orders of
magnitude lower. The concentration of charged particles generated by a
low-current beam is low and comparable with the concentration of free
radicals O. In this case the chain mechanism does not occur, although the
oxidation takes place: the estimated concentration of SO2 is three orders
of magnitude smaller than in the high-current irradiation regime. In the
case of the low-current electron beam the ion losses due to electron-ion
recombination are sharply reduced and the plasma conditions are more
favorable for the conversion of CS2 via the formation of the cluster ions.
The concentration of these ions is [A±] ~ 103 cm-3. The reduction in the
concentration of CS2 occurs in the fast reactions of clusters formation of
A ± (CS2), followed by the polymerization of carbon disulfide and the
formation of (-CS-)n molecules. The rate of these processes is much high-
er than the rate of the reactions involving free radicals O. It was found
that the cluster ions in these experiments are the negative ions O2-, the rate
of which is almost entirely controlled by the process of three-body at-
tachment of the electrons to molecular oxygen.
When the ignition of a volume discharge is initiated by an electron-
beam of nanosecond duration, the influence of the electric field is mani-
fested in a monotonic decrease in the number of converted molecules of Δ
[CS2], which leads to an increased energy consumption. The main reason
for the decrease in Δ [CS2] is the loss of atomic oxygen in the reaction
(12), becausee with an increase in electric field strenth in the discharge
the rate of vibrational excitation of nitrogen increases as well
е + N2 → е + N2(ν)
A more complex nature of the influence of an electric field is ob-
sereved in the discharge, maintained by a weak electron beam. [66]. Since
132
the loss of molecules of carbon disulfide in this regime is associated with
the formation of clusters which are the ions of O2-, the behavior of Δ
[CS2] should be determined by the rate of the cluster ions formation in the
reaction of the three-body attachment
е + М + О2 → М + O2–
Fig. 57 shows the scheme of basic processes of the decomposition of
CS2 caused by a pulsed electron beam
133
Chapter 5. Applied plasma-chemistry
134
However, the high synthesis temperature (1400–1700°С) compli-
cates the manufacturing process [69, 70] and requires a considerable
power input. To reduce the temperature and to decrease the powder parti-
cle size, Bin et al.[69] investigated the gas phase hydrolysis of TiCl4:
TiCl4 (gas) + 2H2O (gas) = TiO2 (solid) + 4HCl (gas)
They showed that amorphous TiO2 was produced at a synthesis tem-
perature of 260-400°С and a crystalline material (anatase) was formed at
400-600°С. The average particle size decreased with an increase in the
synthesis temperature from 120 nm (at 260°C) to 18 nm (at 525°C). The
results of investigation of the preparation of TiO2 nanoparticles by ther-
mal vapor-phase decomposition of tetrabutoxytitanium (C4H9O)4Ti are
reported in [71, 72, 73]. The degradation temperature (Td) of (C4H9O)4Ti
is below 500°C, but the onset of formation of particles with the crystalline
phase (anatase) was detected only at Td > 500°C and all particles pos-
sessed this structure at Td > 600°C. The formation of the rutile crystal
structure began at Td > 1200°С [71]. Ayllón et al. [72] carried out the de-
composition of (C4H9O)4Ti in a mixture with oxygen in an RF discharge
plasma. Despite of a high temperature in the discharge zone, the degrada-
tion of(C4H9O)4Ti yielded amorphous TiO2 particles with an average size
of 25 nm and the crystal structure was formed during subsequent anneal-
ing at 600–800°C. The rutile lattice in noticeable amounts was formed on-
ly at T > 800°C. Li et al. [73] also detected the formation of TiO2 nano-
particles with the anatase lattice at Td of tetrabutoxytitanium of 600°C.
During the annealing of these particles (>800°С), rearrangement into the
rutile lattice took place.
The investigation of the synthesis of TiO2 nanoparticles from abla-
tion plasma produced by a pulse laser showed that it is the formation con-
ditions, not the initial temperature of particles, that are important for the
formation of the crystal lattice. Upon laser ablation of titanium or titania
(rutile or anatase) rods in an He + O2 atmosphere, the formation of TiO2
(anatase) particles of 10–50 nm in size was revealed [74]. The laser abla-
tion of a TiO2 rod in an inert atmosphere generated amorphous TiO2 na-
noparticles. Seto et al. [75] reported the results of study of nanoparticles
prepared by pulsed laser ablation of a TiO2 (rutile) substrate in a helium
atmosphere. It was shown that the particles had a metallic titanium core
and a shell composed of crystalline TiO2 (rutile and anatase).
In the synthesis of TiO2 nanoparticles, the appearance of the initial
titanium-containing material has no determining effect on the crystal
structure of the particles produced. Oh et al. [76], reported the results of
investigation of TiO2 preparation by oxidation of titanium nitride (TiN)
particles in a microwave plasma. A TiN powder was carried to the dis-
135
charge zone by a stream of oxygen mixed with an inert carrier gas. At a
low oxygen concentration, the particles had a TiN core and a TiO2 (rutile)
shell. When oxygen was in excess, the particles were composed com-
pletely of TiO2 but had the anatase structure.
Unlike vapor-phase processes for the manufacture of TiO2 nanopar-
ticles, liquid-phase synthesis occurs at a lower temperature; for example,
at T < 100°C in the sulfate process [69]. However, the particles produced
are amorphous and the formation of a crystal structure requires subse-
quent annealing at T > 600°C. In addition, thermogravimetric studies [69,
70] have shown that the removal of hydroxyl groups and residual organic
components take place at T > 500°С. Heating amorphous or anatase TiO2
to this temperature leads to the formation of the rutile lattice, thus compli-
cating the synthesis of anatase TiO2 by the liquid-phase procedure.
The explosion of conducting wires, providing a non-equilibrium na-
ture of the formation of nanoparticles [77, 78], allows one to synthesize
particles with new properties, in particular with a high excess energy. In
this method, a mixture micrometer-sized and nanosized particles is
formed. Be means of special filters it is posiible to collect the synthesized
nanoparticles with a specific area of up to 40-80 m2/g, which make up 30
% of the total mass of the particles. The capacity of the plant is 100-200
g/h. The explosion of metal wires in oxygen or air leads to the formation
of nanosized metal oxides. It was found that with the oxygen concentra-
tion in the mixture of an inert gas over 20 % (volume persent) the degree
of oxidation of metal particles synthesized by the wire explosion is up to
100 %. Synthesized nanooxides have a spherical shape and relatively
smooth surface.
The data from the comparison of parameters of titanium dioxide syn-
thesized via different processes are summarized in Table 8.
136
anatase
Droplet-to-particle TTIP. + propa- 20 – 40 500–1200 anatase
method nol nm ºС
Decomposition in TTIP. + О2 Average amor-
RF-discharge plas- size 25 nm phous
ma
Laser ablation Ti (or TiO2) + 10 - 50 nm anatase
O2
Ablation by pulsed Ti + O2 4 - 45 nm rutile
ion-beam 0.3-2 µm +anatase
Oxidation in ther- TiN + O2 average 50 anatase
mal microwave- nm
discharge plasma
Sulfate process (wet TiOSO4 + H2O 100-400 hydrolysis rutile
method) nm up to 80ºС, +anatase
annealing -
above
700ºС
Non-equilibrium TiCl4 + H2 + 30 - 200 lower 600 rutile
plasmochemical O2 nm ºС +anatase
synthesis
137
further subjected to thermal treatment at the temperature above 500 °C to
remove the hydroxyl group and the precursor material left.
In [79], amorphous (TiO2)x(SiO2)1-x (x = 0.08, 0.18 and 0.41) was
investigated using X-ray phase analysis, neutron diffraction, infra-red
spectrometry, thermal gravitometry, and other analytical techniques. As
precursor materials use was made of tetraethylorthosilicate (TEOS) and
titanium tetraisopropoxide (C4H9O)4Ti. It was found out that at a compar-
atively low concentration (х = 0.08) titanium is embedded into the lattice
of titanium dioxide to form the Ti-O-Si bond. This concentration of tita-
nium is below its limit of solubility in silicon dioxide. Given a high con-
centration (х = 0.41), the phases of titanium dioxide and silicon dioxide
are separated already in the initial gel, and the general structure of the re-
sulting material is amorphous. When the composite (TiO2)x(SiO2)1-x (х =
0.41) material is heated to the temperature above 500 ºC, titanium dioxide
forms a crystal lattice with the structure of anatase. It was noted in [5] that
the presence of amorphous silicon dioxide retards rearrangement of the
crystal structure of titanium dioxide of the anatase type into that of the ru-
tile type. No rutile-type lattice was found out when the composite materi-
al (х = 0.41) was heated to 800 ºC. Initially, the composite material
(TiO2)x(SiO2)1-x with the average concentration of titanium (х = 0.18) was
characterized by a Ti-O-Ti bonding. Upon heating the gel to 500 ºC, all
the titanium atoms formed a Ti-O-Si structure. When the gel was further
heated (750 ºC and higher), this structure was decomposed to form sepa-
rate phases of silicon dioxide and titanium dioxide (anatase).
In [80] the results of investigation of the composite (TiO2)x(SiO2)1-x
(x = 0.1, 0.3 and 0.5) are presented, which was formed by the sol-gel syn-
thesis tetraethylorthosilicate and titanium tetraisopropoxide. The initial
gel was a composite of TiO2 and SiO2 with a conspicuous number of Si-
O-Ti structures. As the temperature was increased, there was an increase
in the number of Si-O-Ti bonds that decomposed at higher temperatures.
The structure of the composite after thermal treatment was a matrix made
up by pure silicon dioxide with particles of crystalline titanium dioxide
having either an anatase or rutile lattice, depending on the treatment tem-
perature. A concurrent research in [80] was made into the changes in the
structure of titanium dioxide synthesized by the same method from titani-
um tetraisopropoxide. Initially in the course of heating to 420 ºC, the
amorphous gel TiO2 acquired crystal structure with the anatase-type lat-
tice. When heated to above 800 ºC, it exhibited a rearrangement of its
crystal lattice into the rutile form. The composite material rearranged
from the amorphous structure into crystalline (anatase) at the temperature
above 600 ˚C and maintained the anatase lattice in heating to 1000 ºC.
138
The results of investigation into the formation of nanosized compo-
site material from a mixture of titanium dioxide (anatase) gel and silicon
dioxide gel with different mixture ratios are given in [81]. The mixture of
these gels was dried in air at 110 ˚C and was further annealed at 800 ºC
for 1 hour. The structure of pure TiO2 during the high-temperature anneal-
ing treatment was partially rearranged into the rutile-type lattice. An addi-
tion of silicon dioxide inhibited the rearrangement of the crystal lattice of
the former during pyrolysis. The authors conclude that an annealing
treatment for 1 hour at 800 ºC did not form any solid solution of Ti-O-Si,
and the initial oxides were present in the composite material as separate
phases.
A study of the physical characteristics of silicon dioxide aerogel with
5 % titanium dioxide was made in [82]. The composite material was syn-
thesized from nanosized SiO2 gel in ethanol with an addition of nanosized
TiO2 (anatase). The emulsion was dried in air and subjected to thermal
treatment. An IR-spectrometry analysis did not reveal any formation of
the Ti-O-Si bonding either in the initial composite material or after an-
nealing up to 1000 ºC, the initial oxides did not form any solid solution.
With the use of TEM it was found out that TiO2 particles are built into the
structure of silicon dioxide particles 5 nm in size. As in the previous pa-
pers, it was noted that the presence of silicon dioxide inhibits rearrange-
ment of the initial structure of titanium dioxide from anatase to rutile even
at 1000 ºC.
In [83], the results of investigation of a composite (TiO2)x(SiO2)1-x
material synthesized by a thermal decomposition of the metalloorganic
precursor containing silicon and titanium atoms in one molecule. In that
work, three compounds were studied [Si(OBut)2OTi(acac)2O]2,
[(ButO)3SiO]2Ti(OPri)2 and [(ButO)3SiO]3Ti(OPri), which differed by the
number of titanium atoms in the molecule of titansiloxanes. The compo-
site material was formed as a nanostructured film with the grain size from
4 to 200 nm (depending on the pyrolysis temperature). An examination of
the adsorption IR-spectra demonstrated that with increase in temperature
the intensity of the peak from the Si-O-Ti bond is decreased, while the in-
tensity of the adsorption peaks from the Si-O-Si and Ti-O-Ti bonds is in-
creased. This is indicative of the formation of separate phases of titanium
dioxide and silicon dioxide at a high-temperature pyrolysis. At high tem-
perature of decomposition of the initial metalloorganic compound, titani-
um dioxide formed a crystal anatase-type lattice. It is worth noting that
the minimum temperature required for titanium dioxide to crystallize in
the composite material under study was 600-800 ˚C for different precur-
sors. With increase in the number of titanium atoms in the precursor mol-
139
ecule, the minimum pyrolysis temperature at which the formation of crys-
tal structure in TiO2 is decreased. A crystal phase of titanium dioxide of
the rutile type was formed in minor quantities at the pyrolysis temperature
1100 ˚C only.
140
Temperature Regime of Preparation of Oxide Nanoparticles
The use of a closed plasma reactor and the pulsed character of the
process make it possible to calculate a change in temperature during the
synthesis from pressure changes in the reactor. Fig. 59 shows the time de-
pendence of the reactor temperature. The temperatures were calculated by
the ideal gas law from experimentally measured pressures.
Fig. 59 (1) Change in temperature in the plasma reactor during the manu-
facture of oxide nanoparticles and (2) the temperature change due to heating a
gas mixture by electron beam only (without ignition)
141
Fig. 60.TEM image and bar chart of grain-size distribution of the na-
nosized titanium dioxide powder. Initial mixture composition in mmol: H2 + O2
+ TiCl4 (50 : 25 : 10)
Fig. 61. (a) TEM photograph and (b) the histogram of particle size distri-
bution of a nanosized titanium dioxide powder
The investigations also showed that the geometrical size of the syn-
thesized powder depends on the operating mode of the synthesis. A de-
crease in the titanium tetrachloride concentration in the initial mixture or
an addition of a buffer gas resulted in a grain size decreasing. This is in-
dicative of a bulk character of the process of synthesis. It is noteworthy
that the nanosized particles did not have any inner voids.
The shape and structure of nanosized particles of synthesized titani-
um dioxide were also controlled by the conditions of synthesis. At a low
concentration of titanium tetrachloride in the initial reactive mixture, TiO2
particles had a hexagonal nucleus with a characteristic tubular shell (Fig.
62a). With increased concentration of titanium tetrachloride in the initial
142
mixture, TiO2 particles mostly round, with the surface covered by smaller-
sized round particles (Fig. 62b). A further increase in the concentration of
TiCl4 in the initial mixture tended to yield particles with hexadral and cu-
bic cut and clear surface (Fig. 62c).
a b c
Fig. 62. Shape of particles of synthesized titanium dioxide for different
concentration of titanium tetrachloride in the initial mixture: 5 mmol (a), 10
mmol (b), and 15 mmol (c) Size of photo 250·250 nm (a), 1000·1000 nm (b) and
2000·2000 nm (c)
143
Fig. 63. X-ray diffraction patterns of a nanosized titanium dioxide powder
(pattern numbers 1, 3, 4 correspond to sample numbering in Table 1): (0) ana-
tase and (×) rutile
144
indicate a low amount of defects in the crystal structure and the amor-
phous phase of the prepared TiO2 samples.
Despite that the temperature of the gas-phase mixture during the syn-
thesis did not reach 800°C, crystalline nanoparticles with the rutile crystal
lattice were formed, unlike the case of other TiO2 preparation methods.
This formation is due to the nonequilibrium character of the process. An
increase in TiCl4 concentration in the reactant mixture or the addition of
an inert gas (argon) primarily led to the formation of an anatase type lat-
tice.
IR Spectrometry of Nanosized TiO2
The prepared nanosized particles were studied with a Nicolet 5700
IR spectrometer over the range of 400-4000 cm–1 (resolution 4 cm–1).
Typical IR spectra of nanosized TiO2 is shown is Fig. 64.
145
Table 10. The results of X-ray luminescence analysis
146
Decomposition of mixture of halides (TiCl4+ SiCl4 or SiCl4+CCl4)
was completed within a single electron beam pulse. With the electron
beam energy 100 J, the energy consumption for decomposition of mixture
of halides was 5 kJ/mol. This is by far smaller than the TiCl4 dissociation
energy that is equal to 804 kJ/mol. The process of destruction of TiCl4+
SiCl4 with hydrogen and oxygen with an electron beam injected into this
mixture had an explosive character. This circumstance, the presence of a
lower limit (on pressure) of the reactive mixture ignition, and the low en-
ergy consumption are indicative of a cross-linked chain character of the
process of composite (TiO2)x(SiO2)1-x synthesis through decomposition of
TiCl4 + SiCl4 in a mixture with oxygen and hydrogen.
In order to understand whether it is possible to simultaneously syn-
thesize nanosized oxides of different materials, a number of experiments
have been conducted to excite a mixture of oxygen, hydrogen, titanium
tetrachloride and tetrachlorsilane by a pulsed electron beam. For the syn-
thesis of composite oxides use was made of technological tetra-
chlorsilane. The reactor was heated to 90 ºС and, before filling in the gas
mixture, evacuated to ~1 Pa. Upon injecting the electron beam into the
following mixture (mmol): H2 + O2 + SiCl4 + TiCl4 (50 : 25 : 17 : 10), na-
nosized powder was formed in the reaction chamber. Fig. 66 shows a
SEM image of the resulting powder.
Presented in fig. 67 are a TEM image of the powder and a bar chart
of grain-size distribution. Note that the mean size of the composite pow-
der particles is smaller compared to that of pure nanosized titanium diox-
ide synthesized under similar experimental. This might be attributed to a
change in the conditions of coagulation of the particles formed upon addi-
tion of a new material.
147
Fig. 67 TEM image and a bar chart of grain-size distribution from
(TiO2)x(SiO2)1-x powder. The mean size is 29 nm. The initial mixture in mmol:
H2 + O2 + SiCl4 + TiCl4 (50 : 25 : 17 : 10)
148
Fig. 68. Dark-field TEM-image of a small composite powder particle and
its diffraction pattern
149
The chemical composition of the synthesized composite powder was
determined in an Oxford ED2000 X-ray fluorescence spectrometer, and
the results are listed in Table 11. Taking into account the content of oxy-
gen in the synthesized composite powder, the impurity concentration is
less than 0.4 %.
150
Table 12. Percentage of oxygen, silicon and titanium in the composite
powder (TiO2)x(SiO2)1-x according to the EDX-spectra
151
Fig. 72. X-ray diffraction patterns from the TiO2, SiO2 and
(TiO2)x(SiO2)1-x specimens
152
Fig. 73. IR spectra from the specimens made of (TiO2)x(SiO2)1-x (1), SiO2
(2), and TiO2 (3)
153
Shown in Fig. 74 are the TEM-image of the synthesized powder and the
bar chart of grain-size distribution.
154
Fig. 76. IR-spectrum from silicon dioxide proper (1) and composite Si-C-
Ox oxide (2)
155
5.2. Plasmochemical conversion of methane
156
hydrogen synthesis does not exceed 1 eV/molecule. Similar results were
obtained in a study of methane pyrolysis in a gliding alternating-current
(50 Hz, 3 kW) arc discharge [91]. At a gas flow rate of 2 m3/h, a pressure
of 6 atm, and a temperature of 1400 K, 34 % of CH4 was converted into
hydrogen and acetylene. The arc-discharge energy consumption for me-
thane pyrolysis was 0.82 eV/molecule.
The pyrolysis of CH4 in atmospheric-pressure spark, pulsed stream-
er, and dielectric-barrier (dc and ac) discharges at room temperature re-
quires quite a considerable consumption of energy [92]. For methane deg-
radation, the energy costs (eV/molecule) are 14-25 in spark discharge, 17-
21 in streamer discharge, 28-57 in dc dielectric-barrier discharge, and
116-175 in ac barrier discharge. Pulsed spark discharge results in a high
yield of acetylene (54%) and hydrogen (51 %) at a CH4 conversion of
69 %. In the case of barrier discharge, ethane prevails in the conversion
products.
A high efficiency and selectivity of thermocatalytic processes of
conversion of methane and other hydrocarbons stimulated studies on the
plasma pyrolysis of CH4 in the presence of catalysts. Li et al. [93] report-
ed the results of study of the two-stage pyrolysis process in pulsed spark
discharge (first step) and further in a reactor with a catalyst (second step),
which was Fe-, Co-, Ni-, Cu-, or Zn-coated grains of the HZSM-5 catalyst
(SiO2/Al2O3 = 35). The catalyst temperature was 623 K. In the absence of
the catalyst (Fig. 77), the methane conversion was 41 % at an energy con-
sumption of 17 eV/molecule and the product composition was 84 % acet-
ylene and 75 % hydrogen (percent of the maximal theoretical yield).
157
Fig. 77. Methane conversion and product composition (1) in the absence
of catalyst and in the presence of the (2) HZSM-5, (3) Fe+HZSM-5, (4)
Co+HZSM-5, (5) Ni+HZSM-5, (6) Cu+HZSM-5, and (7) Zn+HZSM-5 cata-
lysts
The addition of the uncoated HZSM-5 catalyst increases the yield of
aromatic hydrocarbons (primarily, benzene, toluene, and xylene) to 15 %
with rise in the energy costs to 19 eV/molecule. The Ni coating of grains
provides for the highest yield of aromatic hydrocarbons (47 %) and a sta-
ble catalytic activity for 3 h. Methane pyrolysis in plasma is comprehen-
sively surveyed in [94, 95], and the data are collated in Table 14.
158
Spark with Fe, Co, or CH4 46% (C6H6, 40 19
Ni catalyst on HZSM-5 C7H8,
C8H10)
Gliding (50 Hz) with CH4 + H2 (83 - 100)% 18 2.4 – 7.3
Al2O3 catalyst + Ar C2*
* C2 includes ethane, ethylene and acetylene.
159
tion was 0.25 eV/molecule at a methane conversion of 70 % and in-
creased to 0.5 eV/molecule when the conversion increased to 100 %.
Data that are interesting from the viewpoint of industrial implemen-
tation were obtained in a study of the partial oxidation of methane in arc
discharges of different types. In a two-electrode arc plasma reactor, the
partial oxidation of CH4 in a stoichiometric mixture with oxygen is char-
acterized by an energy consumption of at most 1 eV/molecule [99].
Methane conversion in a mixture with oxygen (CH4 : O2 = 1 : 2) in a
rotating-arc plasma reactor also showed a low energy consumption [100].
As the discharge energy input to the gas increased from 2.6 to 4.8 kJ/l at a
gas flow rate of 15-30 l/s, the consumption for methane conversion in-
creased from 1.6 to 2.2 eV/molecule. Fig. 78 shows the results of a study
of the conversion process.
160
Fig. 79. Plasma-chemical reactor with a surface arc discharge
161
Fig. 80. Concentration of methanol in the products of plasma reforming of
the CH4 + O2 mixture in the presence of different catalysts.
Thus, the highest yield of the synthesis gas at a low energy con-
sumption was obtained in various arc discharges, which provide a high
162
methane conversion capacity per unit volume of the reactor and do not
require a high-voltage power source. The use of combined (plasma +
catalyst) reactors makes it possible to control the composition of conver-
sion products, in particular, to increase the selectivity for CH3OH from 3
to 35 %.
163
Table16. Carbon dioxide reforming of methane in plasma
The results that hold promise for industrial implementation were ob-
tained only for methane conversion in a gliding discharge. In discharges
of other types, the separate conversion of CO2 into CO and CH4 into hy-
drocarbons (C2H6, C2H4, C2H2, etc.) takes place. The electrode material
and the catalyst have a great effect on the composition of the conversion
products of the CH4 + CO2 mixture.
164
sively studied. Under equilibrium conditions, the steam reforming of me-
thane proceeds in accordance with the reaction:
CH4 + H2O +2.1 eV = CO + 3H2
at a temperature of 1300–1900 K and a pressure above l5 MPa [97].
The conversion of methane in a mixture with steam in gliding dis-
charge plasma is characterized in that the energy consumption for CH4
decomposition does not exceed 0.92 eV/molecule [107]. A specific fea-
ture of the reactor design is the rotation of the plasma string by an exter-
nal magnetic field (Fig. 81). At a discharge power of 1 kW, a feed mixture
flow rate of 30 l/min, a temperature of 150 ºC, a pressure of 0.1 MPa, and
a CH4/H2O ratio of 0.67, the degree of conversion reaches 50 %. The
main products are hydrogen (55 vol %) and CO (10 vol %); the amount of
CO2 and C2H2 does not exceed 1 vol %.
165
The results of investigation of CH4 conversion in a mixture with
steam via the chain mechanism are reported in [109]. The experiments
were conducted in a microwave plasma generator (915 MHz; pulse pow-
er, 200 kW) with preheating the reactant mixture to 500-570 ºC; the
CH4/H2O ratio was varied from 0.5 to 1. The gaseous products of the pro-
cess were syngas and CO2. Azizov et al. [109] note that the total energy
consumption of the process is reduced by 30-60 % when the discharge is
switched on, although the discharge energy does not exceed 5 % of the
thermal energy. At a cumulative energy input of 2 J/cm3, the concentra-
tion of the product hydrogen was 11 %, a value that corresponds to the
total energy consumption for H2 synthesis of 4.2 eV/molecule.
Sekine et al. [110] examined the dependence of the conversion and
the product composition on the spark-discharge pulse repetition frequency
(20-300 Hz) and the proportion of CH4 in a mixture with water vapor. In
all experiments a quartz tube of 4.0 mm inner diameter was used as a flow
type reactor as shown in Fig. 82.
166
Fig. 83. Steam reforming of various fuels using non-equilibrium pulsed
discharge.
The cited authors [110] noted that the energy utilization efficiency of
the pulse discharge for the synthesis of hydrogen and CO reached 60 %.
Zhdanok et al. [111] reported the results of study of steam and
steam–air methane reforming in a combined reactor employing a high-
voltage atmospheric-pressure discharge and catalysts (Ni or Fe2O3). Fig. 1
gives a diagram of experimental setup to make studies on partial methane
oxidation in APHVD plasma. The major components of the setup are a
plasmachemical reactor and a gas chromatograph. A cooler and a hotwell
are installed at the reactor output (Fig. 84).
The given experimental study used two types of catalysts: hydro-
genation nickel catalyst Synetix PRICAT 400 (RP) 2.5 mm (15.7 % Ni on
Al2O3) and carbonic oxide conversion catalyst СТК - TZC 3/1 (Fe2O3)
with chrome additives.
167
Fig.84. Experimental set-up
168
300 Hz H2+CO+CO2 1.6
Spark discharge with
Ni catalyst
From the data presented in the table, it is seen that steam methane re-
forming in most low-temperature discharge plasmas requires a low energy
consumption, equal to or lower than the enthalpy of the process. This
mode of plasma-assisted conversion of methane is the most promising for
industrial implementation; however, it has been little studied.
169
duction of 0.9-1.0 eV/molecule. The data on combined methane reform-
ing in plasma are presented in Table 18.
Thus, the most important parameter from the view point of commer-
cialization of the plasma reforming of methane is the energy consumption
for its decomposition. All types of discharge can be divided into two
groups in accordance with the energy consumption, degree of conversion,
and selectivity for products, namely, inhomogeneous (arc, spark, gliding)
and space (dielectric-barrier, and corona) discharges.
In inhomogeneous discharges, the methane conversion efficiency is
higher, the energy onsumption for decomposition is below
10 eV/molecule; the conversion in arc discharges exceeds 90 % and the
selectivity reaches 90 % for certain products (C2H2 in plasma pyrolysis,
H2 and CO in steam reforming). In space discharges, a high degree of
conversion (>50 %) is reached only at high energy consumptions for me-
thane decomposition (>40-50 eV/molecule) and a broad range of products
with a low selectivity are synthesized.
A substantial reduction in the discharge energy consumption for me-
thane decomposition is attained if the conversion is conducted in the
chain mode; in this case, the energy consumption does not exceed 1
eV/molecule. A promising line in the plasma reforming of methane is its
170
conversion in a mixture with steam. In this case, the energy consumption
of gliding discharge for CH4 decomposition is lower than the enthalpy of
the steam methane reforming under the equilibrium conditions. The yield
of H2 considerably exceeds the equilibrium values, and the H2/CO ratio is
5-10.
The review of the experimental works on gas-phase plasmochemical
processes has demonstrated that the conditions developed under the action
of a pulsed electron beam on gas are favorable for initiation of chain
chemical processes. In contrast to other plasma formation techniques, a
high-current pulsed electron beam provides a considerable reduction in
the energy of electrophysical installation for completing the chemical re-
action, which is a critical factor in the case of limited power supply. Un-
der non-equilibrium conditions generated by a pulsed electron beam, no
inhibiting effect of high oxygen concentration on partial methane oxida-
tion is manifested. Unlike other techniques, a pulsed electron beam irradi-
ation helps synthesize nanosized titanium dioxide particles with crystal
structure at low temperature. A considerable reduction in the temperature
threshold of particle crystal structure formation has also been achieved for
composite nanosized particles of (TiO2)x(SiO2)1-x oxides.
171
Conclusion
In conclusion we would like emphysise some other promising areas
of practical application of plasma chemistry. The studies of liquid-phase
decomposition of hydrocarbons during heating showed that the process is
implemented as a chain (thermal cracking). At low temperatures, in the
absence of radiation the thermal initiation of the reaction does not occur;
as a result of the radiation exposure the active centres are formed, which
are free radicals that can start a chain reaction of hydrocarbon cracking.
The interaction of radicals containing a small number of carbon atoms
with the initial hydrocarbon is a distinct chain process, it is initiated by
the radicals formed though the radiation. Such a chain reaction of radia-
tion-thermal cracking will take place at a temperature of 150-200 degrees
below the temperature of a normal thermal process, but with the same
specific rate, because the radiation facilitates the most energy-intensive
stage - the thermal initiation of reaction. In addition, with a sufficient
length of chain the total energy consumption for the chemical process can
be reduced. The main source of energy in this case is the thermal energy
of the original material or the exothermic energy of elementary chemical
chain reactions (e.g., oxidation or polymerization). This can significantly
reduce the energy consumed by the chemical process. Conducting a
chemical process at a low temperature allows one to synthesize com-
pounds which are unstable at higher temperatures or selectivity of synthe-
sis of which at high temperatures is low. The above features from the
chain of chemical processes which occur in plasma are particularly rele-
vant for the petrochemical industry.
But at the interaction of ionizing radiation with matter in the liquid
phase the major part of the secondary electrons produced though ioniza-
tion of the medium, have a small path, due to its relatively low energy.
Therefore the electrons produce ionization and excitation in a local area,
close to the place of their formation. This effect observed in the radiolysis
of liquid media is called "Franck-Rabinowitch effect". This effect is to
maintain a high concentration of radicals and long-lived excited products
formed in the spurs, after the completion of the primary excitation act.
The influence of this effect is particularly noticeable in the general output
of the radiolysis products, which are less formed in the liquid phase than
in the gas phase. This leads to the fact that at temperatures below 600 K,
the radiation-chemical yield does not exceed a few molecules per 100 eV.
Pulse radiolysis of liquid-phase hydrocarbons at low temperature un-
der the action of an electron beam with a high current density (200
A/cm2), is of a great interest. At the same time the conditions favorable to
multiple collisions of particles with excess energy stored in the internal
172
degrees of freedom are formed. These conditions are similar to nonequi-
librium excitation of gas-phase environment and contribute to a chain re-
action. A requirement for creating of such conditions of radiolysis is a
high absorbed dose of the radiation.The tracks are formed during the en-
tire duration of the pulse and during the radiation the relaxation of excited
molecules and radicals recombination should be negligible. The lifetime
of active radicals formed in the track during the radiolysis does not ex-
ceed 10-7-10-8 sec. Therefore, the duration of radiation exposure should
not exceed 100 ns. Radiolysis of liquid hydrocarbons in these conditions
has not been studied and is of scientific and practical interest for studing
the mechanism of radiation liquid phase cracking of hydrocarbons at low
temperatures.
Another promising area of application of the gas discharge and
pulsed electron beams is the initiation of soot formation in non-
equilibrium conditions. It is known that the conversion of carbon-
hydrogen fuel during the combustion passes through the stage of for-
mation of polycyclic aromatic hydrocarbons. They are the centers of for-
mation of the soot particles (nuclei). For the soot formation the radical-
precursors requires; therefore the process under equilibrium conditions
does not occur at low temperatures. In addition, the soot precursors are
pyrolyzed and oxidized at high temperatures, so that the process of soot
formation under equilibrium conditions (1000 - 2000 K) completes the
synthesis of pyrolytic carbon. The formation of radicals thought the action
of a pulsed electron beam on a mixture of hydrocarbon gases and oxygen
can initiate the formation of polycyclic aromatic hydrocarbons at a tem-
perature below the equilibrium sooting. Under these conditions, the pro-
cess of soot formation can be stopped at the stage of the synthesis of aro-
matic hydrocarbons, if the duration of external exposure will not exceed
the duration of this stage. This may allow the development of a new tech-
nology of the synthetic liquid fuels or the complex hydrocarbons with
isomeric structure from a natural gas.
173
Questions
1. Please give the definitions to the following concepts in
chemical kinetics: rate of chemical reaction, law of mass action, reaction
rate constant.
2. Please explain the following: Arrhenius equation, kinetic
equation of the reaction, kinetic scheme of the chemical process
3. Please give the definitions to the following terms: plasma,
Debye radius, degree of ionization, electron energy distribution function.
4. Explain the following: elementary processes in plasmas:
classification, rate elementary processes, collision cross section.
5. Direct and stepwise ionization in the plasma.
6. The relaxation of excited particles in the plasma. The practi-
cal application of the relaxation processes.
7. Treanor mechanism.
8. Mechanisms of electrical discharge. Tawsend discharge
9. Paschen curve. Limitations of the Paschen law.
10. Types of gas discharge
11. The streamer and spark forms of gas discharges.
12. Gas breakdown in an alternating voltage. Microwave
discharge.
13. Typical designs of gas-discharge sources of low-temperature
plasma.
14. The specific characteristics of plasma-chemical reactions.
Quasi-equilibrium plasma-chemical processes.
15. Nonequilibrium plasma-chemical processes. Arrhenius equa-
tion for nonequilibrium plasma chemical reactions.
16. Principles of organization of plasma chemical processes.
17. The main types of reactions in plasma chemistry.
18. The methods of nonequilibrium excitation of the molecules.
19. Chain gas-phase processes. The kinetic scheme of chain pro-
cess.
20. Classification of chain processes. The induction period and
limits of ignition.
21. Chain oxidation of hydrogen.
22. Chain-chemical processes under external action. Methods of
initiation of chain processes.
23. Conversion of methane in low-temperature plasma. Classifi-
cation of conversion methods. Conversion products.
24. The kinetic scheme and the main plasma chemical reaction of
the methane pyrolysis
174
25. The steam conversion of methane in low-temperature plas-
ma. The kinetic scheme and conversion products.
26. Combined methods of plasma-chemical conversion of me-
thane
27. Chain plasma-chemical conversion of methane. The kinetic
scheme and the energy balance of the chain process.
28. Plasma-chemical synthesis of nanosized particles. Classifica-
tion, advantages and disadvantages of synthesis methods.
29. Non-equilibrium plasma-chemical synthesis of nanosized
metal oxides.
30. Non-equilibrium plasma-chemical synthesis of nanosized
composite oxides. The application of titanium nanosized composite ox-
ides.
31. Plasma-chemical methods for the formation of carbon
nanostructures. Allotropic forms of carbon.
32. Beam-plasma technology of hardening and surface modifica-
tion of metal products
33. Applications of low-temperature plasma in chemical produc-
tion
34. Plasma-chemical processing of medical polymers.
35. Plasma-chemical methods of waste treatment.
175
References
176
20. Loeb L.B. Electrical coronas: Their basic physical mechanism. –
Los angeles: Universityof California Press, 1965. – 417 p.
21. Raether H. Avalanches and streamers // J. Phys. – 1939. –
Vol. 112. – P.464–471.
22. Meek J.M. Formation of a negative streamer // J. Phys. Rev. –
1940. – Vol. 57. – P. 722–762.
23. Dawson G.A., Winn W.P. A spherical space charge // J. Phys. –
1965. – Vol. 183. – P. 159 –163.
24. Gallimberti I.J., Bacchiege G.L., Gazzani A., Bernadi M., and
Bondiou A. // Proceedings of the 1994 international aerospace and
groundconference on ‘ Lightning and static electricity’. – 1994. – P. 23–
29.
25. Gallimberti I. The mechanism of the long spark formation // J.
Phys. – 1979. – № 40. – Р. 193–250.
26. Towsend J.S. Electricity in gases. – Oxford: Oxford Press, 1914.
– 493 p.
27. Nasser E. Fundamentals of gaseous ionization and plasma elec-
tronics. – New York:Wiley-Interscience, 1971. – 386 p.
28. Dakin T., Luxa G., Oppermann G., Vigreux J., Wind G. Paschen
curve for air // Electra – 1974. – Vol. 32. – P. 61–69.
29. House J. E. Principles of chemical kinetics // James E. House. –
2007. – P. 963–975.
30. Frinman A. Plasma Chemistry. – New York: Cambridge Univer-
sity Press, 2008. – 725 p.
31. Polak L. Elementary chemical processes and kinetics in a
nonequilibrium and quasi-equilibrium plasma // J. Applied Chem. – 1970.
–
P. 307–342.
32. Полак Л.С. Кинетика неравновесной плазмы// Первый все-
союзный симпозиум по плазмохимии. – 1971. C. 1–15.
33. Полак Л.С. Применение вычислтельной математики в физи-
ческой и химической кинетике. – М.: Наука, 1966. – 421 c.
34. Von Engel A. Ionized Gases. – Clarendon: Oxford, 1955. –
421 p.
35. Spitzer L. Physics of Fully Ionized Gases. – New York.: Inter-
science: New York, 1962. – 358 p.
36. Shih-J-Pay. Magneto Gas Dynamics and Plasma Dynamics. –
Springer: Vienna, 1962. – 412 p.
37. Wall F. T., Hitler L. A., Mazur J. Chemical reactions under low-
temperature plasma conditions // J. Chem. Phys. – 1961. – Vol. 35. –
177
P. 1284–1295.
38. Blais N. C., Bunker D. L. Chemical reactions // J. Che. Phys. –
1963. – Vol. 39. – P. 315–319.
39. Keck J. C. Chemical reactions // J. Chem. Phys. – 1958. – Vol.
29. – P. 410–426.
40. Schmeltekopf A. L., Ferguson E. E. Fesenfeld F. C. Chemical
reactions under low-temperature plasma conditions // J. Chem. Phys. –
1968. – Vol. 48. – P. 2966–2974.
41. Месяц Г.А. Импульсная энергетика и электроника. – М.:
Наука, 2004. – 704 с.
42. Рыжов В.В., Ястремский А.Г. Распределение энергии элек-
тронного пучка в плазме азота // Физика плазмы. – 1978. – Т. 4,
вып.6. – С. 1262–1266.
43. Норман Г.Э., Полак Л.С., Сопин П.И., Сорокин Г.А.
Сильноточные релятивистские электронные пучки в плазмохимии //
Синтез соединений в плазме, содержащей углеводороды. – М.:
ИНХС АН СССР, 1985. – 166 с.
44. Налбандян А.Б., Воеводский В.В. Механизм окисления и
горения водорода. – М.: Изд-во АН СССР, 1949. – 179 с.
45. Старик А.М., Титова Н.С. О кинетических механизмах
инициирования горения водородно-кислородных смесей при
возбуждении электронных степеней свободы молекулярного
кислорода лазерным излучением // Журнал технической физики. –
2003. – Т. 73. – № 3. – С. 59–68.
46. Пушкарев А.И., Ремнев Г.Е. Инициирование окисления
водорода импульсным электронным пучком // Физика горения и
взрыва. – 2005. – № 3. – С. 46–51.
47. Пушкарев А.И., Ремнев Г.Е. Колебательный характер
процесса окисления водорода при инициировании импульсным
электронным пучком // Физика горения и взрыва. – 2005. – № 4. –
C. 18–21.
48. Ремнев Г.Е., Фурман Э.Г., Пушкарев А.И. и др. Импульсный
сильноточный ускоритель с согласующим трансформатором //
Приборы и техника эксперимента. – 2004. – № 3. – С. 130–134.
49. Пушкарев А.И., Пушкарев М.А., Ремнев Г.Е. Исследование
звуковых волн, генерируемых при поглощении импульсного
электронного пучка в газе // Акустический журнал. – 2002. – Т. 48,
№2. – С. 260–265.
178
50. Pushkarev A., Isakova J., Kholodnaya G., Sazonov R. Sound
Waves Generated Due to the Absorption of a Pulsed Electron Beam
//Advances in Sound localization. – Vienna: INTECH, 2011. – 672 p.
51. Селезенев А.А., Алейников А.Ю., Ярошенко В.В. Влияние
радиолиза на смещение пределов воспламенения водород-
кислородной газовой смеси. // Химическая физика. – 1999. – Т. 18. –
№ 5. – С. 65–71.
52. Комар А.П., Круглов С.П., Лопатин И.В. Измерение полной
энергии пучков тормозного излучения от электронных ускорителей.
– Л.: Наука. – 1972. –172 с.
53. Радиолиз углеводородов. Некоторые физико-химические
проблемы / Под ред. А.В. Топчиева, Л.С. Полака.– М.: Изд-во АН
СССР. – 1962. – 208 с.
54. Сараева В.В. Радиолиз углеводородов в жидкой фазе.
Современное состояние вопроса. – М.: Изд-во Моск. ун-та. – 1986. –
256 с.
55. Химия высоких энергий / Под ред. Л.Т. Бугаенко, М.Г.
Кузьмин, Л.С. Полак. – М.: Химия, 1988. – 368 с.
56. Бабарицкий А.И., Деминский М.А., Демкин С.А., Животов
В.К. Эффект плазменного катализа при разложении метана // Химия
высоких энергий. – 1999. – Т. 33, №1. – С. 49–56.
57. Бабарицкий А.И., Баранов И.Б., Дёмкин С. А. и др.
Плазменный катализ процессов конверсии углеводородов // Химия
высоких энергий. – 1999. – Т. 33. – № 6. – С. 458–462.
58. Штерн В.Я. Механизм окисления углеводородов в газовой
фазе. – М.: Изд-во АН СССР, 1960. – 496 с.
59. Арутюнов B.C., Крылов О.В. Окислительные превращения
метана. – М.: Наука, 1998. – 361 с.
60. Кармилова Л.В., Ениколопян Н.С., Налбандян А.Б. К
вопросу о вырожденном разветвлении. Роль формальдегида при
окислении метана // Журнал физической химии. – 1957. – Т. 31. –
С. 851–864.
61. Михайлов Б.М., Куимова М.Е., Богданов В.С. Действие
ионизирующих излучений на неорганические и органические
системы. – М.: Изд. АН СССР, 1958. – 223 с.
62. Русанов В.Д., Бабарицкий А.И., Герасимов Е.Н. и др.
Стимулирование процесса парциального окисления метана в
микроволновом разряде // Доклады Академии наук. – 2003. – Т. 389.
– № 3. – С. 324–327.
179
63. Русанов В.Д., Бабарицкий А.И., Баранов И.Е. и др.
Неравновесное воздействие плазмы микроволнового разряда
атмосферного давления на процесс конверсии метана и керосина в
синтез-газ. // Доклады Академии наук. – 2004. – Т. 395. – № 5. – С.
637–640.
64. Денисов Г.В., Новоселов Ю.Н., Суслов А.И., Устер А.М.
Продукты окисления сероуглерода в ионизованном воздухе // Журн.
техн. физики. – 2001. – Т. 71. – № 1. – С. 136–138.
65. Новоселов Ю.Н., Суслов А.И., Кузнецов Д.Л. Воздействие
импульсных пучков электронов на примесь сероуглерода в воздухе //
Журн. техн. физики. – 2003. – Т. 73. – № 6. – С. 123–129.
66. Новоселов Ю.Н., Рыжов В.В., Суслов А.И. Эффект
электрического поля при конверсии сероуглерода в ионизованном
воздухе. // Письма в ЖТФ. – 2002. – Т. 28. – № 6. – С. 35–41.
67. Remnev G.E. , Pushkarev A.I., Khim. Plasma-chemical synthesis
of SiO2 nanopowders // High Energy Chem. – 2004. – Vol. 38. – №. 5. –
P. 348–354.
68. Remnev G.E., Pushkarev A.I. Plasma-chemical synthesis of TiO2
nanopowders // IEEJ Trans. Fundam. Mater. – 2004. – Vol. 124. – № 6. –
P. 483–489.
69. Bin Xia, Li. W., Zhang B. et al. Low temperature vapor-phase
preparation of TiO2 nanopowders // Journal of Materials Science. – 1999.
– Vol. 34. – P. 3505 – 3511.
70. Jang H. D., Kim S.-K. and Kim S.-J. Effect of particle size and
phase composition of titanium dioxide nanoparticles on the photocatalytic
properties // J. of Nanoparticle Res. – 2001. – Vol.3. – № 2 – 3. P. 141–
146.
71. Ahonen P.P., Moisala A., et al. Gas-phase crystallization of tita-
nium dioxide nanoparticles // J. of Nanoparticle Res. – 2002. – Vol . 4. –
№ 1– 2. – P. 43–46.
72. Ayllon J. A., Figueras A., Garelik S., et al. Preparation of TiO2
powder using titanium tetraisopropoxide decomposition in a plasma en-
hanced chemical vapor deposition (PECVD) reactor // J. of Mater. Sci.
Lett. – 1999. – Vol. 16. – № 18. – P. 13– 9.
73. Li W., Ni C., Lin H. et al. Size dependence of thermal stability of
TiO2 nanoparticles // J. of Appl. Phys. – 2004. – Vol. 96. – №. 11. –
P. 6663–6667.
74. Harano A., Shimada K. et al. Crystal phases of TiO2 ultrafine
particles prepared by laser ablation of solid rods // J. of Nanoparticle Res.
– 2002. – Vol. 4. – № 3. – P. 215–217.
180
75. Seto T., Kawakami Y. et al. Evaluation of morphology and size
distribution of silicon and titanium oxide nanoparticles generated by laser
ablation // J. of Nanoparticle Res. – 2001. – Vol. 3. – № 2–3. – P. 185–
189.
76. Oh S.-M., Park D.-W. and Ishigaki T. Plasma Synthesis of
Spherical Titanium Dioxide from Titanium Nitride // Proc. of 16th Intern.
Symp. on Plasma Chemistry. Taormina. – 2003. – P. 45–52.
77. Kotov Yu A. Electric explosion of wires as a method for prepara-
tion of nanopowders // Journal of Nanoparticle Research. – 2003. – № 5.
– P. 539–550.
78. Назаренко О.Б. Электровзрывные порошки. Получение,
свойства, применение / Под ред. А.П. Ильина. – Томск: Изд-во
Томского университета, 2005. – 148 с.
79. Wallidge G. W., Anderson R., Mountjoy G., Pickup D. M., Gun-
awidjaja P., Newport R. J., Smith M. E. Advanced physical characteriza-
tion of the structural evolution of amorphous (TiO2)x(SiO2)1-x sol-gel ma-
terials // Journal of materials science. – 2004. – Vol. 39. – P. 6743–6755.
80. Ingo G. M., Riccucci C., Bultrini G., Dire S. and Chiozzini G.
Thermal and microchemical characterization of sol-gel SiO2, TiO2 and
xSiO2-(1-x)TiO2 ceramic materials // Journal of Thermal Analysis and
Calorimetry. – 2001. – Vol. 66. – P. 37–46.
81. Machida M., Norimoto K., Watanabe T., Hashimoto K., Fuji-
shima A. (1999) The effect of SiO2 addition in super-hydrophilic property
of TiO2 photocatalyst // Journal of materials science. – 1999. – Vol. 4. –
P. 2569–2574.
82. Young-Geun Kwon, Se-Young Choi, Eul-Son Kang, Seung-Su
Baek Ambient-dried silica aerogel doped with TiO2 powder for thermal
insulation // Journal of Materials Science. – 2000. – Vol. 35. – № 24. –
P. 6075–6079.
83. Takahiro G., Takayuki K., Yoshimoto A. Crystallization Behav-
ior of SiO2-TiO2 ceramics derived from titanosiloxanes on pyrolysis //
Journal of Sol-Gel Science and Technology. – 1998. – Vol. 13. – №1–3. –
P. 975–979.
84. Remnev G.E., Isakov I.F., Pushkarev A.I. A pulsed ion beam ac-
celerator for material modification // Surf. Coat. Technol. – 1999. – Vol.
114. – P. 206–217.
85. Nakamoto K. Infrared and Raman Spectra of Inorganic and Co-
ordination Compounds. – New York: Wiley, 1986. – 326 p.
86. Kuptsov A.Kh., Zhizhin, G.N. Fourier-Transform Raman and In-
frared Absorption Spectra of Polymers. – M.: Fizmatlit, 2001 – 126 p.
181
87. Ahn W.S., Kang K.K., Kim K.Y. Synthesis of TS-1 by micro-
wave heating of template-impregnated SiO2–TiO2 xerogels // Catalysis
Letters. – 2001. – Vol. 72. – № 3–4. – P. 229–232.
88. Fincke J. R., Anderson R. P., Hyde T. et al. Plasma thermal con-
version of methane to acetylene // Plasma Chemistry and Plasma Pro-
cessing. – 2002. – Vol. 22. – № 1. – P. 107–138.
89. Babaritskii A.I., Deminskii M.A., Demkin S.A., Zhivotov, V.K.,
A low temperature thermal initiation of cracking // High Energy Chem. –
1999. – Vol. 33. – № 1. – Р. 45–49.
90. Zhivotov V.K., Potapkin B.V., Rusanov V.D. Encyclopedia of
Low-Temperature Plasma. – M.: Fizmatlit. – 2005. – 41 p.
91. Czernichowski A., Glid arc assisted production of H2 from CH4,
CO or H2S // Proc. Int. Symp. Hydrogen Power, Theoretical and Engi-
neering Solutions. – 1999. – P. 98–104.
92. Li X.S., Zhu A.M., Wang K.J., Xu Y., Song Z.M. Methane con-
version to C2 hydrocarbons and hydrogen in atmospheric non-thermal
plasmas generated by different electric discharge techniques // J. Catal. –
2004. – Vol. 98. – P. 617–625.
93. Li X.S., Shi C., Xu Y., Wang K.J., Zhu, A.M. A process for a
high yield of aromatics from the oxygen-free conversion of methane:
combining plasma with Ni/HZSM-5 catalysts // Green Chem. – 2007. –
Vol. 9. – P. 647–689.
94. Pushkarev A.I., Novoselov Yu.N., Remnev G.E., Tsepnye
protsessy v nizkotemperaturnoi plazme. – Novosibirsk: Nauka, 2006. –
123 p.
95. Pushkarev A.I, Remnev G.E., and Ponomarev D.V., Fundamen-
tal’nye problemy prilozhenii fiziki nizkotemperaturnoi plazmy: Materialy
lektsii Vserossiiskogo simpoziuma molodykh uchenykh, studentov i as-
pirantov // Proc. All-Russia Symp. of Young Scientists, Students, and As-
pirants on Fundamental Problems of Application of Low-Temperature
Plasma Physics. – 2005. – Vol. 1. – P. 26–29.
96. Shtern V.Ya., Mekhanizm okisleniya uglevodorodov v gazovoi
faze. – Moscow: AN SSSR, 1960. – 123 p.
97. Arutyunov V.S. and Krylov O.V. Okislitel’nye prevrashcheniya
metana. – M: Nauka, 1998. – 136 p.
98. Rusanov V.D., Babaritskii A.I., Gerasimov E.N., Deminskii
M.A., Demkin S.A., Zhivotov V.K, Moskovskii A.S., Potapkin B.V.,
Smirnov R.V. and Strelkova M.I. Dokl. Akad. Nauk. – 2003. – Vol. 389.
– № 3. – Р. 324–327.
182
99. Lesueur H., Czernichowski A., Chapelle A. Electrically assisted
partial oxidation of methane //. J. Hydrogen Energy. – 1994. – Vol. 19. –
№ 2. – P. 139–142.
100. Lee D.H., Kim K.T., Song Y.-H and Cha M.S. Characteristic of
Methane Processing by Rotating Arc Reactor Proc. // 18th Int. Symp. on
Plasma Chemistry. – 2007. – P. 153–159.
101. Heo J., Choi J.-W., Lee H., Sekiguchi H. and Song H.K. Syn-
thesis Gas Production from Methane and Air Mixture with a Vortex Glid-
ing Arc Reactor Proc. // 18th Int. Symp. on Plasma Chemistry. – 2007. –
P. 554–559.
102. Remnev G.E., Pushkarev A.I., Ezhov V.V., Partial methane and
nitrogen oxidation initiated by pulsed electron beam Proc // 13th Int.
Symp. on High Current Electronics. – 2004. – P. 447–451.
103. Indarto A., Choi J.W., Lee H., Song H.K. Plasma methane oxi-
dation for methanol synthesis with Cu-Zn-Al catalyst Proc. // 18th Int.
Symp. on Plasma Chemistry. – 2007. – P. 194–198.
104. Mutaf-Yardimci O., Saveliev A., Fridman A., Kennedy L. Em-
ploying plasma as catalyst in hydrogen prodaction Int. // J. Hydrogen En-
ergy. – 1998. – Vol. 23. – № 12. – Р. 1109–1114.
105. Lee H., Choi J.-W., Song H.K., Lee C.-H. The effect of the
electric pulse polarity on CO2 reforming of CH4 using dielectric-barrier
discharges Proc. // 4th Int. Symp. on Pulsed Power and Plasma Applica-
tions. – 2003. – P. 146–157.
106. Song H., Lee H., Choi J.,Na B. Effect of electrical pulse forms
on the CO2 reforming of methane using atmospheric dielectric barrier dis-
charge // Plasma Chem. Plasma Process. – 2004. – Vol. 24. – №. 1. – P.
57–61.
107. Rusu I. and Cormier J.-M. On a possible mechanism of the me-
thane steam reforming in a gliding arc reactor // Chem. Eng. J. – 2003. –
Vol. 91. – № 1. – Р. 23–29.
108. Cormier J., Rusu I. Syngas production via methane steam re-
forming with oxygen: plasma reactors versus chemical reactors // J. Phys.
D: Appl. Phys. – 2001. – Vol. 34. – P. 2798–2801.
109. Azizov R.I., Babaritskii A.I., Demkin S.A. In Gazokhimiya v
XXI veke. Problemy i perspektivy // Gas Chemistry in XXI Century:
Problems and Prospects. – 2003. – P. 45–53.
110. Sekine Y., Urasaki K., Kado S., Asai S., Matsukata M., Kikuchi
Steam reforming of hydrocarbons and alcohols using non-equilibrium
pulsed discharge E // Proc. 16th Int. Symp. on Plasma Chemistry,
Taormina. – 2003. – P. 213–216.
183
111. Zhdanok S.A., Krauklis A.V., Bouyakov I.F. Studying methane
conversion in atmospheric pressure high-voltage discharge with different
oxidizers in presence of catalysts // Proc. IV Intern. School-Seminar
Modern Problems of Combustion and its Application. – 2001. – P. 66–72.
112. Chavadej S., Rueangjitt N., Sreethawong T. Reforming of CO-
containing natural gas with partial oxidation using an AC gliding arc sys-
tem // Proc. 18th Int. Symp. on Plasma Chemistry. – 2007. – P. 552–561.
113. Bromberg L., Cohn D., Rabinovich A., O’Brien C. and
Hochgreb S. Plasma catalytic reforming of methane Energy Fuels // Proc.
18th Int. Symp. on Plasma Chemistry – 1998. – Vol. 1. – №. 1. – Р. 11–
16.
114. Bromberg L., Cohn D., Rabinovich A., Alexeev N. Partial me-
thane and nitrogen oxidation initiated by pulsed electron beam // Int. J.
Hydrogen Energy. – 1999. – Vol. 24. – P. 1131–1136.
184