Вы находитесь на странице: 1из 17

Carbon, 2.

Diamond
OTTO VOHLER, former Sigri GmbH, Meitingen, Germany
GABRIELE NUTSCH, former Technische Universitat Ilmenau, Ilmenau, Germany
FERDINAND VON STURM, former Sigri GmbH, Meitingen, Germany
ERHARD WEGE, former Sigri GmbH, Meitingen, Germany

1.
2.
2.1.
2.2.

Properties and Structure of Natural


Diamond . . . . . . . . . . . . . . . . . . . . . . .
Diamond Synthesis. . . . . . . . . . . . . . . .
HPHT Synthesis. . . . . . . . . . . . . . . .
The Diamond Chemical Vapor
Deposition (CVD) Process . . . . . . . .

2
4
5

2.3.
2.3.1.
2.3.2.
2.3.3.

Diamond CVD Techniques. . . .


Nonthermal Plasma CVD. . . . . .
Thermal Plasma Diamond CVD .
Other CVD Techniques . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

9
10
12
13

For many centuries, natural diamond has been


appreciated as a valuable gem and investment.
The first center of mining was India, which was
replaced by Brazil in the early 1700s, South
Africa in the mid-1800s, and most recently
Kongo (formerly Zaire), Australia, Russia,
Botswana, and Namibia. Output amounted to
ca. 55  106 ct in 1998, but this does not satisfy
the increasing demand for industrially used
diamonds. Table 1 lists the diamond mining
output of some of major producing countries
and the total world mining production in
1998.
The sources of natural diamonds are probably ultrabasic rocks, especially eclogite, which
consists essentially of omphacite and pyrope.
The famous blue ground of Kimberlite slots is a
decomposed eclogite rock containing about
0.5 ct of diamond per tonne. Such primary
deposits are found in South Africa, Russia, and
Tanzania. Most of occurences of diamond are
secondary deposits. From so-called placers, of
which the L
uderitzbay deposit in Namibia is an
example, diamond is mined explosively.
Less than the half of the mined diamonds can
be used for gems; the majority are used for
industrial purposes. The prospering interest in
industrial diamonds is based on their combination of outstanding properties besides their high
hardness, which makes them highly suitable for
many applications. Synthetic diamond has been
produced commercially for over 30 years [35]
 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
10.1002/14356007.n05_n01

by high-pressure, high-temperature (HPHT)


techniques (see Fig. 1) Synthetic industrial diamond is available in far larger quantities than its
natural diamond counterpart. The estimated
capacity for synthetic diamond of 100  106 ct
in 1982 increased to 350  106 ct in 1990 and
to more than 500  106 ct in 1999 (see
Table 2). Consequently, synthetic diamond accounts for more than 90% of the industrial
diamond used. Various reports estimate the total
industrial diamond output worldwide to be more
than 600  106 ct, valued at (600800)  106
dollars. More than 30 countries produced
industrial diamond in 1998. In addition to the
countries listed in Table 2, Korea produces synthetic diamond, but specific data on their output
could not be confirmed. China and Russia also
presumably produce far more than the output
given in the table. Thus, two-thirds of industrial
diamond output is concentrated in four countries: the United States, Ireland (possibly the
worlds largest producer), Russia, and South
Africa.
Interest in industrial diamond has been further increased by the more recent discovery
[6, 7] that it is possible to produce polycrystalline diamond films or coatings by a wide variety
of chemical vapor deposition (CVD) processes
from a hydrocarbon gas in an excess of hydrogen. This CVD diamond can show mechanical,
tribological, and even electronic properties
comparable to those of natural diamond.

Carbon, 2. Diamond

Table 1. World diamond mining output in 1998, 106 ct


Country

Gem

Industrial

Total

Australia
Russia
Congo
Botswana
Rep. of South Africa
Angola
Namibia
Brazil
China
Ghana
Other
World

18.897
9.250
3.000
11.000
5.360
3.600
1.300
0.700
0.230
0.125
1.955
55.400

23.096
9.250
15.000
5.000
6.000
0.400

41.993
18.500
18.000
16.000
11.300
4.000
1.300
1.300
1.130
0.630
2.807
117.000

0.600
0.900
0.505
0.852
61.600

Figure 1. Phase diagram of different carbon modifications

Table 2. Synthetic diamond, estimated world production, 106 ct


Country

1995

1997

1999

Belarus
China
Czech Republic
France
Greece
Ireland
Japan
Poland
Romania
Russia
Slovakia
South Africa
Sweden
Ukraine
United States
Total

25.0
15.5
5.0
3.0
1.0
60.0
32.0
0.256
5.0
80.0
5.0
60.0
25.0
8.0
115.0
440.0

25.0
16.0
5.0
3.0
0.75
60.0
32.0
0.26
5.0
80.0
5.0
60.0
25.0
8.0
125.0
451.0

25.0
16.5
3.0
3.0
0.75
60.0
32.0
0.2
3.0
80.0
3.0
60.0
25.0
8.0
208.0
467.0

1. Properties and Structure of


Natural Diamond
Diamond is a colorless mineral, generally characterized by its hardness (Mohs hardness 10)
and brilliance after cutting. The density is
3.515 g/cm3 (cf. graphite 2.266 g/cm3). The
lattice of the cubic modification is face-centered
with the lattice constant a 0.3567 nm. The
atomic arrangements in diamond is shown in
Figure 2. Diamond is tetrahedrally bonded, and
all atoms have equal bond lengths of 0.1544 nm
and are linked by overlapping sp3 hybrid orbitals
(sp3sp3 s-bond). The bonding angle of the
covalent hybrid bonds is 109.28 . This contrasts
with the hexagonal structure of graphite with
trigonal bonding (Fig. 2). The sp2sp2 s-bonding is supplemented by pp p-bonds oriented
perpendicularly to the s-bonds. All s-bonds lie
in the same plane, and the bond angle is 120 .
The stacking sequence for the carbon atoms in
the diamond structure is ABCABC (Fig. 3 b). It
shows a chair configuration (hatched) of puckered hexagonal rings in both the vertical and
horizontal stacking directions. The first carbon
atom that starts nucleating a new {111} diamond layer is shown in the Figure 3 a (top). By
addition of a further carbon atom the basic unit
of the structure is obtained. This is also the
nucleation kernel for a new plane.
The color of diamonds can vary from white to
pink, yellow, light blue, and even black (carbonado), depending on the incorporated impurities. On the basis of the different characteristics
of the IR absorption spectra of natural diamond,
a distinction is made between Type I and Type
II diamond, with several subgroups. A majority
of gem-quality diamonds are of Type I a and
contain substitutional nitrogen impurities up to
0.5 atom%. These diamonds have a characteristic, slightly greenish yellow color. Only about
1% of natural diamonds contains impurities in
the form of substitutionally isolated nitrogen
atoms. These diamonds are classified as Type
I b. All synthetic diamonds belong to this group.
Their nitrogen atoms did not havesufficient time
to form aggregates, in contrast to the natural
diamonds of Type I a, most of which remain
deep in the earths mantle after their formation
and are exposed to extremely high temperatures
and pressures. During this phase, the majority of
the nitrogen atoms can form aggregates.

Carbon, 2. Diamond

Figure 2. Bonding hybridization and corresponding crystal structure of the carbon allotropes [1]

Diamonds without measurable substitutional


nitrogen atoms or aggregates are classified as
Type II a. They are very rare in nature; less than
1% of all diamonds belongs to this group. Their
absorption bands are caused only by the vibrational absorption of the lattice that is common to

all diamonds. They are often colorless or light


pink. Type II a diamonds are electrical insulators like Type I, but have excellent optical and
thermal properties. Type II b diamonds containing ca. 1 ppm of boron atoms and even
smaller amount of nitrogen impurities are the

Carbon, 2. Diamond

Figure 3. Stacking of carbon atoms in diamond crystals

purest form of naturally occuring diamonds.


They are usually blue to bluish gray in color.
The presence of boron atoms makes these diamonds to electrical semiconductors.
Diamond crystals cleave well parallel to
(111). The cutting hardness is different for the
various crystal faces, and this must be taken into
consideration when cutting. The coefficient of
friction depends on the purity of the surfaces.
With increasing degasification, the coefficient
rises from about 0.1 to 0.6 [8, 9].
Diamond is known for its combination of its
outstanding properties: it has extreme mechanical hardness (ca. 90 GPa, or Mohs hardness 10),
it is the strongest known material with the highest bulk modulus (1.2  1012 N m2) and lowest compressibility (8.3  1013 m2 N1). Its
thermal expansion coefficient at room temperature of 0.8  106 K1 is comparable with that
of low-expansion ironnickel alloys (Invar).
The specific heat is 6.2 J g1 K1 at 300 K,
13.55 J g1 K1 at 500 K and 21.6 J g1 K1
at 1000 K. Diamond Types I and II a are good
electrical insulators, with a room-temperature
resistivity of about 1016 W cm. However, diamond can be doped, for example with boron
(Type II b), to vary its resistivity over the range
10106 W cm, so that it becomes a semiconductor with a wide band gap of 5.4 eV.

Diamonds are excellent conductors of heat.


Type II a diamond has the highest known thermal conductivity at room temperature of
2  103 W m1 K1, which is five times higher than that of copper. Diamond has a broad
optical transparency from the deep - UV to the
far - IRndash OK? (Korektur nicht eindeutig
auch im Folgenden) region of the electromagnetic spectrum. Moreover, it is biologically
compatible and very resistant to chemical erosion and abrasive wear at room temperature. At
higher temperatures, it is attacked by oxygen,
oxidizing agents, and carbide-forming elements. At temperatures of about 2000 K and
above, diamond is converted to graphite, both
internally and on the crystal surfaces. The rate of
phase transformation increases rapidly with increasing temperature [10].

2. Diamond Synthesis
According to the phase diagram (Fig. 1)
there are only a few ways to synthesize diamond: by shockwaves (mostly used in Japan),
by high pressure and high temperature (HPHT
synthesis), and by chemical vapor deposition
(CVD) at relatively low pressures and moderate
temperatures.

Carbon, 2. Diamond

2.1. HPHT Synthesis


Figure 1 shows how the state of carbon phases
depends on temperature and pressure [11].
Diamond is a metastable phase at standard
temperature and pressure, and graphite is the
only thermodynamically stable carbon phase.
The transformation of graphite into diamond is
possible at high temperatures and pressures
[1214]. Above 12.5 GPa and temperatures of
about 3000 K, graphite spontaneously collapses
into polycrystalline diamonds.
The difficulty in the diamond synthesis
process has always been the simultaneous
achievement of sufficiently high pressures and
temperatures. These conditions require special
apparatus which was first developed by P. W.
BRIDGMAN [15] in the form of a self-sealing highpressure packing which operated without leaking and permitted hydraulic operation, namely,
the unsupported area seal. For the invention of
the apparatus to produce extremely high pressures, BRIDGMAN was awarded the Nobel Prize in
1946.
The General Electric process [35] uses a
modified Bridgman apparatus [16, 17]. In the
catalytic process, graphite is transformed into
diamond by first dissolving it in molten metals
and then cystallizing it above equilibrium pressure. Nickel is the most widely used metal for
dissolving carbon, for which it has a high solubility (ca. 4 wt% at 1500  C and 5.7 GPa).
Under these conditions, graphite is metastable,
and diamond, which is less soluble (ca. 3.6
wt%) crystallizes out. In the past it was assumed
that the metals used (e.g., nickel, manganese,
iron, cobalt, palladium, platinum, and alloys
thereof) act as catalysts. The term catalyst is
still frequently used, but it is now clear that the
metals serve purely as solvents [18], and that the
growth process is analogous to the travelingzone solvent technique [19].
Any form of carbon can be used as the
starting material [20]. Naphthalene initially
forms graphite; paraffin wax and sugar are
converted directly to diamond. Even peanuts
are transformed to the extent of up to 60% into
diamond at 2000 K and 10 GPa in 5 min.
The belt and multipiston diamondsynthesis
apparatus is shown in Figure 4. It consists of two
truncated conical pistons and a prestressed cylinder assembly, which forms the reaction cham-

Figure 4. Belt-type pressure apparatus [2]


a) Top piston; b) Cylinder; c) Bottom piston; d) Gasket
assembly; e) Current ring; f) Sample holder; g) Graphite; h)
Insulating pill

ber. The circular gaps between the sections are


filled with conical gaskets made of pyrophyllite,
a hydrous aluminium silicate, coated with a thin
dusting of hematite (Fe2O3) to further increase
the friction. The gaskets act as seals, pressure
transmitters, and thermal and electric insulators.
By using strongly tapered pistons (a and c in
Fig. 4) a large piston motion becomes possible
without squeezing the pyrophyllite gasket too
thin. The compressed gasket in turn provides
additional strength by supporting the pistons.
This arrangement is suitable for a large volume
of starting material, which consists of stacked
graphite and metal plates or a mixture of graphite powder and nickel or other group 810 metal
powder, and is heated directly by resistance
heating. The current is supplied via pistons,
current rings and nickel plates. Since the graphite has the highest resistance in the current path,
it is heated to the desired temperature. The
nickel melts, dissolves the graphite, and begins
to crystallize out diamond. The molten metal
film is less than 0.1 mm thick, and the movement of the carbon through this metal film is
quite rapid. In a few minutes the nickel film
moves through the graphite from both ends to
the hotter center, and all graphite is converted to
diamond. The pressure is maintained until after
current has been cut off and the sample cooled to
prevent reversion to graphite. After cooling, the
diamonds are separated by dissolving the metal
with acids. The average size of the synthetic

Carbon, 2. Diamond

diamonds is about 0.050.5 mm. The lifetime of


this apparatus with a diameter of up to 40 mm is
ca. 100 to 1000 cycles when properly operated.
Smaller diamonds are produced only by shockwaves (see Fig. 1).
Diamond crystals can be controlled in size
and mechanical properties by varying the process parameters such as the form and nature of
graphite, the metal solvent used, temperature,
pressure, and time. Grit up to 0.5 mm in size is
produced without crushing. Also the shape of
the crystals can be controlled by the temperature. Because of this size- and shape-control
ability, synthetic diamond grit is generally
superior to natural diamond. Cutting blades
made from GE diamonds have been manufactured for many years [21, 22]. For cutting rocks
or glass, blocky grit sintered in a metal matrix is
used. For shaping hard metals and tungsten
carbide, a friable grit in a resinoid matrix is
preferred. For polishing pastes, the grit is often
mixed with olive oil, lithium stearate, and some
amorphous silica. There are also diamond
compounds for polishing based on water with
polyethylene glycol, potassium hydroxide in
methanol, and amorphous silica. There is no
difference in hardness between synthetic and
natural diamond.
About 15 years after the first announcement
of industrial diamonds, single crystals in gem
quality were also grown by this method [20].
The equipment is similar to that described
above, but the large, high-quality crystals must
be grown very slowly (about one week is
necessary to grow a 5 mm, 1 carat, high-quality
diamond crystal) and under much more carefully controlled and stable conditions. This
makes gem diamond growth so difficult and
expensive.

net reaction:
CH4 gactivation energy ! Cdiamond2 H2 g

In addition to methane, a wide variety of


carbon-containing substances have been used
as precursor materials. The diamond CVD process involves chemical reactions in the gas
phase as well as at the solid surfaces which
cause depositions onto the surfaces (Fig. 5).
Usually, the CH3 radical is regarded as the
precursor. This depends on the boundary conditions at the substrate. With high-energy plasma jets, atomic carbon is the most important
precursor species.
All CVD techniques used for producing diamond films require:
*

a particular gas atmosphere and hence, a


closed reaction chamber
a means for activating the gas-phase carboncontaining precursor molecules
a substrate temperature of about 400 to
1000  C
dissolution of the precursor gas in an excess
of hydrogen

While HPHT synthesis requires special


industrial equipment used by only few

2.2. The Diamond Chemical Vapor


Deposition (CVD) Process
Chemical vapor deposition (CVD) of diamond
films has developed dramatically [2331] since
its initiation in the 1950s [6, 7]. The diamond
film deposition process utilizes temperature and
pressure conditions under which graphite is the
stable form of carbon (see Fig. 1). However, the
kinetic factors of the CVD process allow crystalline diamond to be produced by the typical

Figure 5. Diamond CVD process (after [28])

Carbon, 2. Diamond

Figure 6. Schematic depiction of the growth of diamond


(dark arrows) and removal of unwanted carbon phases (light
arrows)

companies, the CVD of diamond films can be


performed by using more or less the laboratory
equipment. Deposition methods are diverse.
The Role of Hydrogen. Hydrogen is a very
important species for diamond growth [32].
Hydrogen plays two main roles. First, molecular
hydrogen suppresses the formation of aromatics
in the gas phase; second, atomic hydrogen
activates the surface.
With pure CH4, growth of a layer also occurs,
but it consists of both diamond and unwanted
carbon phases (Fig. 6). With hydrogen alone,
only etching of both graphite and diamond
occurs. The graphitic phases are always more
readily etched by reaction with hydrogen atoms
than the diamond phase. Therefore, hydrogen
atoms should be present in higher than equilibrium concentration to obtain successful diamond growth and etching of non-diamond carbon phases.
Due to the excess of hydrogen the diamond
surface is almost fully saturated with hydrogen
during the growth process (step 1 in Fig. 7).
Desorption of bound hydrogen by incoming
atomic hydrogen creates vacant sites or surface
radicals (step 2 in Fig. 7). These vacant sites

Figure 7. C, H bounded C and H atoms,*


H ,  excited
atoms
*

can be refilled by atomic hydrogen (step 3 a in


Fig. 7), or H atoms can be added to unsaturated
carboncarbon bonds. Thereby, conversion into
sp3 bonds at low temperatures or gasification of
sp2 bonds at high temperatures takes place, and
diamond growth occurs (step 3 b in Fig. 7).
Without activation of the gas by plasma or
other energetic sources, the concentration of
atomic hydrogen, and therefore the number
density of active sites, is low. Hence, the
filmgrowth rate is very low.
The Role of Substrate Temperature. The
deposition of polycrystalline diamond films
occurs in a surface temperature range of about
8501500 K. The lower limit approaches about
700 K and is nearly independent of the deposition technique [29, 30, 33], but it depends on the
precursor used, the growth rate is very low, and
non-carbon phases are implanted into the layer [34, 35]. However, the growth rate can be
enhanced by using halogenated precursors [36].
The surface chemistry depends strongly on the
surface temperature. At low temperatures, the
rate of creation of active sites is low. Increasing
the temperature also increases this creation
rate. Above about 1500 K the number density
of active sites decreases due to thermal
decomposition and subsequent graphitization.
The sp2 bonds are rapidly etched by the hydrogen atoms.
At too low surface temperatures the condensation of aromatics can lead to amorphous carbon phases. Therefore, an efficient diamond
growth rate can be controlled by the surface
temperature, the absolute number density of the
carbon-containing species and atomic hydrogen, and the ratio of the two.
By varying the substrate temperature, the
properties of the diamond film can be tailored.
It varies from well-defined facets to irregular
forms or round shapes. Variations in the diamond morphology with the substrate temperature are shown in Figure 8 for the DCTP jet.
With increasing substrate temperature, it
changes from predominantly more {111} to
{100} facets, which was also found for an
RF-ICTP Jet [37] and for DCTP CVD [38]. The
crystallite size also increases with increasing
substrate temperature [3843], as is shown in
Figure 8. However, the mechanism of diamond
growth is very complex, and contrary results

Carbon, 2. Diamond

Figure 8. Morphology of diamond films manufactured with


a DCTP jet and 5 vol% CH4 in H2 at various substrate
temperatures
a) 780  C; b) 850  C; c) 950  C; d) 1050  C

were also published [44]. Therefore, the


substrate temperature must be kept constant
during the deposition time and must be as
uniform as possible across the entire surface
area to minimize or prevent radial variations in
the morphology of the crystallites. In the thermal plasma jet diamond CVD the substrate must
be intensively cooled. Special designs of substrate holder exist [45, 46] by which either the
different thickness or/and the varying materials
influence the heat transfer to the cooling fluid
(water is mostly used, but also ethylene glycol).
To adjust for varying heat fluxes from the
plasma jet to the substrate, special cooling
systems have been developed [47, 48].
The Role of the CH4/H2 Ratio. The morphology and therefore the quality of the diamond films depend on the ratio of the precursor
dilution in hydrogen. The morphology can be
changed in a similar way as for the substrate
temperature with an increasing methane/hydrogen ratio (Fig. 9). The ratios vary considerably
with the deposition method used. The ratios of
more than 5 vol% given in [33, 48, 49] are
relatively high compared to those reported
in [5052] with less than 5 vol%. For epitaxial
growth [53] or with nonthermal plasma techniques, concentrations of less than 1 vol% CH4
based on the H2 content are used.

Figure 9. Morphology of DCTP diamond films at varying


volume content ratios of CH4 to H2 at a substrate temperature
of 1050  C
a) 4 vol% CH4; b) 5 vol% CH4; c) 7.5 vol% CH4; d) 10 vol
% CH4

Precursors and Substrates. A wide variety


of precursor and substrate materials has been
tested. Methane has been most frequently used
as precursor material. Other carbon-containing
substances [54, 55] such as aliphatic and aromatic hydrocarbons, alcohols, amines, ethers,
are also used:
1. Gaseous
a. Methane
b. Ethane
c. Ethylene
d. Acetylene
2. Liquid
a. Methanol
b. Ethanol
c. 1-Propanol
d. 1-Butanol
e. Acetone
f. Benzene
g. Cyclohexane
3. Halogenated
a. Dichloromethane
b. Trichloromethane
c. Carbon tetrachloride
d. Trifluoromethane
e. Tetrafluoromethane

Carbon, 2. Diamond

More recently, halogenated hydrocarbons


have been recommended, mainly because of
indications of enhanced diamond nucleation
density [56] and lower substrate temperature [57]. Substrate temperatures below 300  C
are possible. An overview on halogenated precursors is given in [58, 59]. However, the use of
fluorine restricts the choice of substrate materials; for example, silicon is always damaged,
even if only a small amount is added. Chlorine-containing compounds prohibit the use of
stainless steel vessels.
The variety of substrate materials used are
listed in in the following:
1. Carbons
a. Amorphous carbon
b. Graphite
c. Diamond
2. Metals
a. Aluminum
b. Titanium
c. Molybdenum
d. Nickel
e. Copper
f. Tungsten
g. Tantalum
h. Platinum
i. Germanium
j. Iridium
k. Silicon
l. Tool steel (with interfacial stainless steel
layer)
3. Ceramics
a. Boron carbide
b. Cubic boron nitride
c. Silicon carbide
d. Silicon nitride
e. Titanium carbide
f. Zirconium carbide
g. Hafnium carbide
h. Chromium carbide
i. Molybdenum carbide
j. Cemented carbide, such as WCCo
compound
k. Alumina
l. Silica
4. Polymers
a. Polyimide (with interfacial layer)

Materials which easily form carbides are


particularly suitable for diamond deposition.
Due to the high solubility of carbon in tool steel,
interfacial layers have to be deposited, such as
tungsten or silicon carbide and silicon nitride.
Silicon single crystals, commonly taken
from the microelectronics industry, are the most
frequently used substrate materials in laboratory
reactors. However, well-adhering, high-quality
diamond films are necessary to realize coatings
on industrially important substrate materials,
such as ceramics, steels, nickel alloys, cemented
carbides, and most alloys containing transition
metals. Basically, all samples are diamondground and washed with alcohol in an ultrasonic
bath. Different pretreatments have to be
performed on these as-ground substrates to
enhance nucleation and adhesion. Chemical and
mechanic pretreatments increase the surface
roughness, which improves both the nucleation
density and the mechanical anchoring of the
film. A review on different chemical and
mechanical pretreatments is given in [60],
especially for cemented carbides.
Another means to improve the nucleation
density is a secondary discharge to the substrate.
It is used to enhance the chemical nonequilibrium in the boundary layer in front of the substrate. In this way, the substrate can be cleaned
when negative biased. The diamond growth rate
is improved by positive biasing [61] due to the
increased flux of atomic hydrogen to the substrate surface.

2.3. Diamond CVD Techniques


Today a wide range of methods exists to prepare
CVD diamond layers. The techniques can be
categorized according to the specific method of
producing chemically active species, such as
thermal decomposition, hot-filament technique,
combustion process, plasma-assisted processes, and beam technologies (Fig. 10). The
plasma-assisted processes include nonthermal
plasmas, such as the direct current (DC) glow
discharges, capacitively and inductively coupled 13.45 MHz radio frequency (CC- and
IC-RF) plasmas, microwave (mW) plasmas
excited by 915 MHz and 2.45 GHz radiation,
as well as thermal plasmas (TP), such as the
thermal inductively coupled radio frequency

10

Carbon, 2. Diamond

Figure 10. Diamond CVD technologies (from [54])

plasmas (IC-RFTPs) or thermal direct current


plasmas (DCTPs) which produce plasma jets
with high velocities and temperatures.
Beam technologies, such as electron or ion
beams, and lasers are not often used for CVD
diamond film production. Initially, thermal
decomposition was used for diamond film production. The gas and substrate temperatures
must be high enough to decompose the precursor gas molecules, and the resulting growth rates
are too low to be industrially interesting. Therefore, plasma CVD methods were developed.
Energy transfer from activated species of the
plasma to precursor molecules leads to their
decomposition and to chemical reactions already in the gas phase. In this way the deposition
rate can be enhanced without increasing the
substrate temperature.

2.3.1. Nonthermal Plasma CVD


Plasmas generated under high vacuum conditions are characterized by different temperatures of electrons and heavy particles such as
ions, atoms, molecules, and radicals [62]. The
low pressure is tantamount to low number densities. The long mean free path enables the
electrons to gain high energies, but energy
transfer to the heavy particles is reduced due
to the low collision probability. Therefore, plasmas at pressures lower than 0.01 MPa are no
longer in thermodynamic equilibrium. Such
plasmas have the advantage of a large and
homogeneous plasma volume; the disadvantages are the low gas velocities, thick boundary
layers in front of the substrate [63], and medium
growth rates (see Fig. 11).

Carbon, 2. Diamond

11

Figure 11. Linear growth rate of various plasma jet techniques in comparison to other techniques, such as HF (Hot
Filament), MW (microwave), RF and DC (radio-frequency
and d.c. nonthermal plasma processes)
Figure 12. Ellipsoidal microwave reactor AIX P6 [68]

The 2.45 GHz microwave discharge was the


first and is still most frequently used diamond
CVD technique with plasma discharge. An
overview on the different microwave reactor
types used for diamond CVD is given in [64].
The NIRIM reactor type was the first reactor
reported [65] and operates with a power of up to
1500 W and pressures typically below 1 kPa.
The carbon-containing gas mixture consists of
1% methane in hydrogen, and the growth rate is
about 0.5 mm/h for substrate temperatures
around 1000  C. The main drawback of this
reactor type is the smallness of the substrate,
which is only 23 cm in diameter. In the domeshaped bell-jar reactor [67], three-inch silicon
wafers are coated with polycrystalline diamond
at substrate temperatures of 9001000  C, pressures of about 1 kPa, and 1% methane in hydrogen gas at a growth rate of about 0.20.3 mm/h.
ASTex improved and commercialized the
bell-jar reactor in the 1990s for using fourinch-diameter substrates. The highest growth
rate reached with a high-pressure microwave
source (HPMS) reactor with 5 kW and about
10 kPa amounts to 14 mm/h. The drawbacks are
connected with the 2.45 GHz frequency: the
power density, plasma size, and therefore the
diamond deposition rate are limited. Nonuniform temperature distributions and film thickness, as well as variations in phase purity are
possible. Shifting to 915 MHz can help to overcome the power limitation and to increase the
geometric uniformity.
On the basis of mathematical simulations,
the more recently developed ellipsoid reac-

tor [66] for large-scale diamond film production


was designed. Avariety of different reactor sizes
and designs have been explored, especially for
915 MHz. The plasma has the shape of a hemisphere sitting on the substrate. The design was
commercialized by the AIXTRON Company.
By using the AIX P6 (6 kW; Fig. 12) at
2.45 GHz and, in particular, the 915 MHz AIX
P60 (60 kW) [68], free-standing polished diamond windows up to six inches in diameter
(Fig. 13) can be produced with a maximum
growth rate of 12 mm/h at a maximum pressure
of 20 kPa.
As is shown in Figure 11, the growth rate can
be increased also in the lower pressure region by
means of DC supersonic jet synthesis [69]. Here
a distinction is made between the DC plasma jet,
which is more or less a thermal plasma jet, and
the DC supersonic jet, which is also arc-heated
but has a large deviation from thermodynamic

Figure 13. Optical-window-quality free-standing diamond


films [67]

12

Carbon, 2. Diamond

Figure 14. Schematic of the supersonic DC jet torch [72]

equilibrium. Typically, the arc is generated inside a torch, similar to a DC plasma torch. The
jet leaves the converging/diverging nozzle
(Fig. 14) with a minimum diameter of about
1 mm in the reactor chamber, maintained at a
pressure of 23 kPa [70], which is much lower
than that in the torch (ca. 0.6 MPa). The high
pressure difference is responsible for the gas
dynamic conversion of thermal arc energy into
high kinetic energy of the jet, which expands
into the reactor with supersonic velocities
(110 km/s) [71]. The temperatures in the jet
are not higher than 2 kK, and reach the highest
values of up to 5 kK in the bow shock near the
substrate [72]. This kind of jet shows a large
deviation from the local kinetic and chemical
equilibria. The high flow velocities lead to small
transport timescales relative to the chemical
kinetic timescales [73]. Therefore, the recombination of atomic hydrogen is reduced and the
carbon-utilization rate is high. The growth rate
ranges between 6 [74] and 10 mm/h [75], which
is much higher compared to the other nonthermal plasma techniques operating in the same
pressure region (see Fig. 11). This results from
the thinning of the boundary layer thickness due
to the high flow velocities.
2.3.2. Thermal Plasma Diamond CVD
A promising alternative to the conventional
low-pressure gas discharges is the use of
high-pressure thermal plasma jets. Plasma jets
are generated by electric discharges at

moderate to high pressure (10100 kPa and


higher), so that the plasma leaves the nozzle
with both high temperature (> 6 kK) and high
velocity (often > 500 m/s). Due to the high
plasma temperatures, the reactant decomposition process is much more intensive. All of the
molecular hydrogen is dissociated into atomic
hydrogen, the most important precursor in the
diamond deposition process. The delivery of
the atomic hydrogen to the substrate is
enhanced because of the reduced thickness of
the boundary layer due to the relatively high
gas velocities. Thus, the highest growth rates
are obtained with thermal plasma jets, as is
shown in Figure 11.
The plasma jets can be categorized in accordance with the method by which the electric
discharge is sustained. The radio frequency
inductively coupled discharge and the direct
current discharge, shown schematically in
Figure 15, are the most frequently used plasma
jets.
The radio-frequency inductively coupled
thermal plasma (RF-ICTP) is an electrodeless
plasma and is therefore almost free of contaminants. Two separate gas flows are necessary for
maintaining the RF discharge: the sheath gas for
discharge stabilization and for preventing direct
contact between the plasma and the plasmaconfining wall, and the central gas flow for the
working gas. The features of the RF-ICTP are
the large volume, and consequently the low flow
velocities of the plasma exiting the torch, with
or without a nozzle. The diameter of the plasma
jet leaving the torch ranges from about 30
[76, 77] through 76 [78], up to 100 mm [79],
depending on the torch geometry [81] and the
reactor pressure [82]. The velocities at the nozzle exit decrease with increasing diameter and
range from 30 m/s [83] up to 250 m/s [84].
Supersonic plasma jets generated by RF-ICTP
with converging/diverging nozzles are also possible [85]. An other feature of the RF-ICTP is
the possibility for axial precursor injection by
means of a water-cooled feeding probe. Moreover, there are no restrictions on the kind of
gases used.
The direct current thermal plasma (DCTP) is
generated in a nontransferred plasma torch by
the high-current arc generated between the
cathode, usually made from tungsten, and an
anode shaped as a nozzle. The nozzle diameter

Carbon, 2. Diamond

13

Figure 15. RF-ICTP jet and DC-TP jet

ranges from 3 to 8 mm, depending on the


power level used. The argonhydrogen gas
mixture is heated by the high-current arc and
leaves the anodic nozzle as a currentless plasma
jet with a much higher velocity in comparison to
the RF-ICTP jet [76, 77]. Impinging onto the
surface, the DCTP jet also provides a higher
momentum. However, the small diameter of the
DCTP leads to small deposition areas. The
deposition on extensive or three-dimensional
substrates requires movement of the substrate.
In contrast to the RF-ICTP torch, precursor
injection must occur outside of the torch to
prevent contamination of the tungsten electrode
by carbon.
The main differences between the DCTP jet
and the RF-ICTP jet are the temperature distributions in the axial and radial directions. The
DCTP jet has steep gradients in both radial and
axial directions. In the case of the RF-ICTP jet,
the temperatures in both the axial and radial
directions decrease much less within the plasma
core than in the thermal DC plasma jet. The
most important difference between the DCTP
and ICTP jets is the thickness of the boundary
layer in front of the substrate. In general, the
boundary layer thickness of the ICTP jet is on
the same order of magnitude as in the lowpressure plasma techniques. The higher the
velocity of the jet impinging the substrate,
the thinner is the boundary layer. In the case of
the DC plasma jet, the boundary layer extends

only to few tenths of millimeters. Therefore,


chemical reactions cannot occur inside the
boundary layer (frozen chemistry). Thus, the
high temperatures at the end of the boundary
layer of about 4000 K generate a lot of atomic
hydrogen and carbon. Therefore, atomic carbon
is the most important species for diamond deposition with the DCTP Jet [63].
The experimental set-up for Plasma Jet Diamond CVD (Fig. 16) consists of the plasma
torch with its power supply, the gas supply with
mass flow controllers for exact control of the gas
flow and composition, the reactor chamber, the
control unit for cooling water, and the automatic
registration unit for plasma parameters (chamber pressure and substrate temperature, measured by a thermocouple in the substrate holder
and by a pyrometer on the substrate
surface) [86].
2.3.3. Other CVD Techniques
The hot filament deposition method is a simple
technique and is therefore widely used. The
principal task of the hot filaments is to dissociate
molecular hydrogen into atomic hydrogen.
Bare metal wires generate atomic hydrogen
more efficiently than silicon carbide filaments.
Atomic hydrogen is generated by thermal
dissociation of molecular hydrogen at the hot
solid filament surface. The upper operating

14

Carbon, 2. Diamond

Figure 16. DCTP jet diamond CVD equipment


a) Cooling water temperature measurement; b) DC power supply; c) Automatic registration unit; d) Pyrometer; e) Plasma torch;
f) Mass flow controller; g) Pressure measurement; h) Window; i) Substrate; j) Water atomizer; k) Thermocouple (substrate
temperature); l) Reactor wall, water-cooled; m) Substrate holder, movable

temperature (ca. 2200  C) of the filament restricts the quantity of atomic hydrogen produced. To overcome this, low pressures (e.g.,
< 10 kPa) are used to enhance its production
and transport to the substrate. The substrate
distance is relatively small (ca.120 mm). However, the low gas-phase concentration permits
only relatively low growth rates compared to the
plasma jet techniques (Fig. 11). Because of its
low cost and its very simple construction, the hot
filament process has been the most frequently
used method since it was first reported [87]. The
methane content based on the flow rate of
molecular hydrogen is always less than
5% [88]. A higher methane content lowers
substantially the rate of recombination of atomic hydrogen at the substrate surface. Besides the
relatively low growth rate, another drawback is
filament recrystallization due to exposure to the
carbon-containing atmosphere. The wire materials, such as tungsten and tantalum, change
their structure as a result of carburization and
hydrogen embrittlement. Cracks are formed,
and the lifetime is limited to about 20 h,
depending on the operating temperature [89].
Rhenium filaments allow carbon to diffuse

rapidly through them without forming carbides.


Carburized rhenium is ductile [90] and therefore
has a longer lifetime. Single filaments are often
used in laboratory reactors; multifilament reactors are more common for industrial applications. An overview on the hot filament process
and scale-up for industrial purposes (see
Fig. 17) is given in [91].
The oxy-acetylene combustion flame method
was first used in 1988 [93]. The mechanism of
diamond growth in combustion CVD is similar

Figure 17. Drill bits in a hot filament reactor [92]

Carbon, 2. Diamond

15

Figure 18. Regions for diamond growth in combustion CVD

to the other diamond deposition systems, even


though the composition of acetylene flames
differs much from the common CH4/H2 mixtures. The combustion flame is divided into
three regions: the inner flame, the acetylene
feather, and the outer diffusion flame. The
substrate is placed in the acetylene-rich feather
region. The ratio of acetylene to oxygen is
the most important parameter for diamond
growth. At values of unity, the acetylene burns
completely. Heat and atomic hydrogen are provided by the chemical energy of the oxidation of
acetylene. A certain supersaturation of acetylene is essential. The highest diamond quality is
obtained in a slightly rich acetylene flame with a
supersaturation of up to 15% (Fig. 18). Combustion CVD is possible at atmospheric pressure
and reduced pressure. The lower the pressure,
the lower is the growth rate. The growth rate at
atmospheric pressure is about 40 mm/h, and at
6 kPa about 4 mm/h. Different types of
burners are used. More recently, flat-flame burners with a uniform axial gas velocity have
been used to enhance the deposition area
while simultaneously improving the film
uniformity [94]. Material deposited by combustion CVD has not been tested extensively.
However, the quality is similar to that produced
by hot filaments or plasma systems. There are
not so many applications. Deposition onto steel
has been tested with nitrided chromium as
buffer layer [95]. Combustion CVD necessitates
lower investment costs compared to the plasma
CVD processes, but these saving are overridden
by the high running costs due to the high quantities of acetylene and its relatively high
costs.

References
Specific References
1 H. Marsh, F. Rodriguez-Reinoso (eds.): Science of Carbon
Materials, University of Alicante, Alicante 2000.
2 R. Berman: Physical Properties of Diamond, Clarendon Press,
Oxford 1965.
3 General Electric Comp., US 2 992 900, 1961 (H.P. Bovenkerk).
4 General Electric Comp., FR 1 341 561, 1963 (M. R. H. Wentorf).
5 ASEA, FR 1 346 568, 1963.
6 B.V. Spitsyn, B.V. Derjaguin, patent application 10.7.56.
7 US 3 0301 87 and 3 030 188, 1962 (W. Eversole)
8 J. E. Field (ed.): The Properties of Diamond, Academic Press,
London 1979.
9 R. Berman, Physical Properties of Diamond, Clarendon Press,
Oxford 1965.
10 V.R. Howe, Proc. Phys. Soc. London 80 (1962) 648.
11 F.P. Bundy, J. Geophys. Res. 85 (B12) (1980) 6930.
12 J. Abrahamson, Carbon 12 (1974) 111.
13 F.P. Bundy, J. Chem. Phys. 38 (1963) 631.
14 F.P. Bundy, H.M. Strong, R.H. Wentdorf in P. L. Walker,
P. A, Thrower (eds.): Chemistry and Physics of Carbon,
vol. 10, Marcel Dekker, New York 1973.
15 P.W. Bridgman, Sci. Am. 193 (1955) 42.
16 F.P. Bundy, H.T. Hall, H.M. Strong, R.H. Wentdorf, Jr., Nature
176 (1955) 51.
17 H.P. Bovenkerk, F.P. Bundy, H.T. Hall, H.M. Strong, R.H.
Wentdorf, Jr., Nature 184 (1959) 1094.
18 Y.A. Litvin, Inorganic Materials 4 (1968) 144.
19 K. Nassau: Synthesis of Bulk Diamond, in: R. F. Davis (ed.):
Diamond Films and Coatings, Noyes Publications, Park Ridge
1993.
20 R.H. Wentdorf, Jr., Phys. Chem. 69 (1965) 3063.
21 K. Subramanian, V.R. Shanbhag: Abrassive Applications of
Diamond, in M. A. Prelas, L. K. Bigelow (eds.): Handbook of
Industrial Diamonds and Diamond Films, Marcel Dekker, New
York 1998, chap. 28.
22 W. Morrow, Lapidary J. 17 (1963) 922.
23 R.C. DeVries, Annu. Rev. Mater. Sci. 17 (1987) 161.
24 J. C. Angus, C . C. Hayman, Science 241 (1988) 913.
25 K.E. Spear, J. Am. Ceram. Sci. 72 (1989) 171.
26 F.J. Celii, J.E. Butler, Annu. Rev. Phys. Chem. 42 (1991)
643.
27 R. E. Clausing, L. L. Horton, J. C. Angus, P. Koidl (eds.): NATO
ASI Series, Series B: Physics, Vol. 266, 1990.

16

Carbon, 2. Diamond

28 J.E. Butler, R.L. Woodin, Philos. Trans. R. Soc. Lond. A342


(1993) 209.
29 K.E. Spear, M. Frenklach, Pure Appl. Chem. 66 (1994) 1773.
30 K.E. Spear, M. Frenklach, Synthetic Diamond: Emerging CVD
Science and Technology, Wiley, New York 1994.
31 H. Liu, D.S. Dandy, Diamond Chemical Vapor Deposition:
Nucleation and Early Growth Stages, Noyes Publications, Park
Ridge, NJ, 1995.
32 B.V. Deryaguin, D.V. Fedossejev, Growth of Diamond and
Graphite from the Gas Phase, Nauka, Moscow 1977.
33 Z.P. Lu, J. Heberlein, E. Pfender, Plasma Chem. Plasma Proc. 12
(1992) 35.
34 I. Watanabe, T. Matsushita, K. Sasahara, Jpn. J. Appl. Phys. 31
(1992) 1428.
35 Y. Liou, A. Inspektor, R. Weimer, R. Messier, Appl. Phys. Lett. 55
(1989) 631.
36 I. Schmidt, C. Benndorf, Diamond and Related Materials 7
(1998) 266.
37 J.W. Lindsay, J.M. Larson, S.L. Girshick, Diamond and Related
Materials 6 (1997) 481.
38 Z.P. Lu, J. Heberlein, E. Pfender, Plasma Chem. Plasma Proc. 12
(1992) 5569.
39 Status and Applications of Diamond and Diamond-Like Materials, National Materials Advisory Board (NMAB)-445, Academy
Press, 1995.
40 A.B. Harker, J.F. Natale, Diamond Depositions, Science and
Technology, Aug. 1991, 1213.
41 K.E. Spear, M. Frenklach, Pure Appl. Chem. 66 (1984) 1773.
42 A. Boudina et al., Proc. Diamond Films, 1991, Nizza.
43 W. Zhu, R. Messier, A.R. Badzian, Proc. 1st Int. Symp. Diamond
& Diamond-like Films, Electrochem. Soc. Pennington, NJ, 1989,
66.
44 W.A. Yarbrough, R. Messier, Science 274 (1990) 688.
45 Q.D. Zhuang, H. Guo, J. Heberlein, E. Pfender, Diamond and
Related Materials 3 (1994) 319.
46 S.K. Baldwin, T. Owano, M. Zhao, C.H. Kruger, Diamond and
Related Materials 6 (1997) 202206.
47 M.T. Bieberich, S.L. Girshick, Plasma Chem. Plasma Process.
16 (1996) Suppl., pp. 157/s168/s.
48 M. Breiter, Ch. Doppleb, K.H. Weiss, G. Nutsch, Diamond and
Related Materials 9 (2000) 333336.
49 J. Heberlein, E. Pfender, 2nd Plasma-Technik-Symposium, 1992,
Vol. 1, p. 187.
50 J.M. Larson, S.L. Girshick, Proc. 14th Int. Symp. on Plasma
Chemistry, 1999, Prague, 1663.
51 A. Higa, A. Hatta, T. Ito, M. Toguchi, A. Hiraki, Diamond Films
and Technology 6 (1996) 277.
52 A. Hirata, M. Yoshikawa, Diamond and Related Materials 4
(1995) 1163.
53 A. Higa, A. Hatta, T. Ito, T. Maehama, Jpn. J. Appl. Phys. 2 35
(1996) L577.
54 P.K. Bachmann et al., Diamond and Related Materials 1 (1991)
10211034.
55 P.K. Bachmann, H. Lade, D. Leers, D.U. Wiechert, Diamond and
Related Materials 3 (1994) 79.
56 R.A. Rudder, et al., Appl. Phys. Lett. 59 (1991) 791.
57 M.S. Wong, C.H. Wu, Diamond and Related Materials 1 (1992)
369.
58 Assmann, M. Heberlein, J.E. Pfender, Diamond and Related
Materials 8 (1999) 116.
59 Assman, M. Nelson, K. Heberlein, J.E. Pfender, Proc. ISPC 14,
1999, Praha, p. 1657.
60 F. Deuerler, H. van den Berg, R. Tabersky, A. Freundlieb, M. Pies,
V. Buck, Diamond and Related Materials 5 (1996) 1478
1489.

61 S.K. Baldwin, T.G. Owano, C.H. Kruger, Appl. Phys, Lett. 67


(1995) 194.
62 M.I. Boulos, P. Fauchais, E. Pfender, Thermal Plasmas, Plenum
Press, New York 1995.
63 B.W. Yu, L. Si, S. Girshick, J. Appl. Phys. 75 (1994)
39143923.
64 P.K. Bachmann: Microwave Plasma Chemical Vapor Deposition of Diamond, in M. A. Prelas et al. (eds.): Handbook of
Industrial Diamonds and Diamond Film, Marcel Dekker, New
York 1997.
65 M. Kamo, Y. Sato, S. Matsumoto, N. Seteka, J. Cryst. Growth 62
(1983) 642.
66 M. Funer, C. Wild, P. Koidl, Surface & Coatings Techn. 7475
(1995) 221226.
67 P.K. Bachmann et al. in A. Badzian et al. (eds.): Diamond and
Diamond-like Materials, MRS Symp. Proc. EA-15 (1988)
p. 99.
68 AIXTRON AG, Aachen, Germany, Prospectus.
69 M.A. Capelli: Arcjet Synthesis of Diamond, in M. A. Prelas
et al. (eds.): Handbook of Industrial Diamonds and Diamond
Film, Marcel Dekker, New York 1997.
70 D.S. Dandy, M.E. Coltrin, Appl. Phys. Lett. 66 (1995)
391393.
71 H.M. Loh, M.A. Capelli, Surface and Coatings Techn. 54/55
(1992) 408413.
72 Juchmann, W. Luque, J. Wolfrum, J. J. B. Jeffries, Proc.
Gas Discharge Applications, 1997, Greifswald, Germany,
pp. 346349.
73 H.M. Loh, M.A. Capelli, Proc. 3rd Int. Symp. Diamond Materials, Vol. 9317, pp. 1723, Electrochem. Soc., Honolulu, HI,
1993.
74 J. Laimer, H. Pauser, H. Stori, R. Haubner, B. Lux, Diamond and
Related Materials 6 (1997) 406410.
75 M.H. Loh, J.G. Liebeskind, M.A. Capelli, AIAA 943233, 30th
Joint Propulson Conference, Indianapolis, IN, 1994.
76 G. Nutsch, B. Dzur, Elektrow
arme Int. B (1995) B201.
77 G. Nutsch: Anwendung des Thermischen Plasmas in der
Werkstofftechnik, Habilitationschrift, TU Ilmenau, 1995.
78 R. Hernberg, T. Lepisto, T. Mantyla, T. Sternberg, J. Vattulainen,
Diamond and Related Materials 1 (1992) 255.
79 M. Kohzaki, K. Uchida, K. Higuchi, S. Noda, Jpn. J. Appl. Phys.
Part 2 32 (1993) L 438.
80 M. Rahmane, G. Soucy, M.I. Boulos, J. High Temp. Chem.
Processes 3 (1996) .
81 S.L. Girshick, C. Li, B.W. Yu, H. Han, Diamond and Related
Materials 2 (1993) 1090.
82 M. Hollenstein, M. Rahmane, M.J. Boulos, Proc. 14th Int. Symp.
Plasma Chem., Prague, 1999, p. 257.
83 B. Dzur, G. Nutsch, Proc. UTSC 1999, Essen.
84 T.G. Owano, D.G. Goodwin, C.H. Kruger, M.A. Capelli, 10th Int.
Symp. Plasma Chemistry (ISPC 10) 1991, Bochum, Germany,
3. 18 pp 16.
85 E. Theophile, F. Gitzhofer, M.I. Boulos, Proc. ISPC, Prague
1999, pp. 21452148.
86 M. Breiter, G. Nutsch: Diamond and Releated Materials, Proc
10th Europ. Conf. Diamond & Diamond-like Materials, Prague
1999, pp. 333336.
87 S.P. Chauhan, J.C. Angus, N.C. Gardner, J. Appl. Phys. 47 (1976)
10311033.
88 L. Schaefer, C.P. Klages, U. Meier, K.K. Honighaus, J. Appl.
Phys. Lett. 58 (1991) 571573.
89 C. T. Lynch (ed.): Practical Handbook of Materials Science,
CRC Press, Boca Raton 1989.
90 G.B. Gaines, C.T. Sims, R.I. Jaffee, J. Electrochem. Soc. 106
(1959) 881885.

Carbon, 2. Diamond
91 C.P. Klages: Untersuchungen zu Herstellung und Eigenschaften
von Diamantschichten (Study of Production and Properties of
Diamond Coatings), Habilitationsschrift, TU Braunschweig,
1998.
92 Fraunhofer Institut Schicht- und Oberflachentechnik, Diamanttechnologie, Prospectus.
93 Y. Hirose, S. Amanuma, K. Komaki, J. Appl. Plugs. 68 (1990)
p. 6401.

17

94 C.A. Wolden, Z. Sitar, R.F. Davis: Combustion Flame Deposition of Diamond, in B. Dischler, C. Wild (eds.): Low-Pressure
Synthetic Diamond, Springer Series in Materials Processing,
Springer-Verlag, Berlin 1998, pp. 4158.
95 J.G. Buinsters, F.M. Bouwelen, J.J. Schermer, W. J. P. van
Enckevort, J.J. ter Meulen, Diamond and Related Materials
9 (2000) 341345.

Вам также может понравиться