ArticlePDF Available

Quantum cascade laser investigations of CH4 and C2H2 interconversion in hydrocarbon/H2 gas mixtures during microwave plasma enhanced chemical vapor deposition of diamond

Authors:
  • Cascade Technologies Ltd, United Kingdom, Stirling

Abstract and Figures

CH 4 and C2H2 molecules (and their interconversion) in hydrocarbon/rare gas/H2 gas mixtures in a microwave reactor used for plasma enhanced diamond chemical vapor deposition (CVD) have been investigated by line-of-sight infrared absorption spectroscopy in the wavenumber range of 1276.5−1273.1 cm−1 using a quantum cascade laser spectrometer. Parameters explored include process conditions [pressure, input power, source hydrocarbon, rare gas (Ar or Ne), input gas mixing ratio], height (z) above the substrate, and time (t) after addition of hydrocarbon to a pre-existing Ar/H2 plasma. The line integrated absorptions so obtained have been converted to species number densities by reference to the companion two-dimensional (r,z) modeling of the CVD reactor described in Mankelevich et al. [J. Appl. Phys. 104, 113304 (2008)]. The gas temperature distribution within the reactor ensures that the measured absorptions are dominated by CH4 and C2H2 molecules in the cool periphery of the reactor. Nonetheless, the measurements prove to be of enormous value in testing, tensioning, and confirming the model predictions. Under standard process conditions, the study confirms that all hydrocarbon source gases investigated (methane, acetylene, ethane, propyne, propane, and butane) are converted into a mixture dominated by CH4 and C2H2. The interconversion between these two species is highly dependent on the local gas temperature and the H atom number density, and thus on position within the reactor. CH4→C2H2 conversion occurs most efficiently in an annular shell around the central plasma (characterized by 1400<Tgas<2200 K), while the reverse transformation C2H2→CH4 is favored in the more distant regions where Tgas<1400 K. Analysis of the multistep interconversion mechanism reveals substantial net consumption of H atoms accompanying the CH4→C2H2 conversion, whereas the reverse C2H2→CH4 process only requires H atoms to drive the reactions; H atoms are not consumed by the overall conversion.
Content may be subject to copyright.
Quantum cascade laser investigations of CH4and C2H2interconversion
in hydrocarbon/H2gas mixtures during microwave plasma
enhanced chemical vapor deposition of diamond
Jie Ma,1Andrew Cheesman,1Michael N. R. Ashfold,1,aKenneth G. Hay,2
Stephen Wright,2Nigel Langford,2Geoffrey Duxbury,2and Yuri A. Mankelevich3
1School of Chemistry, University of Bristol, Bristol BS8 1TS, United Kingdom
2Department of Physics, University of Strathclyde, John Anderson Building, 107 Rottenrow,
Glasgow G4 0NG, United Kingdom
3Skobel’tsyn Institute of Nuclear Physics, Moscow State University, Leninskie Gory, Moscow 119991, Russia
Received 31 July 2008; accepted 19 June 2009; published online 5 August 2009
CH4and C2H2molecules and their interconversionin hydrocarbon/rare gas/H2gas mixtures in
a microwave reactor used for plasma enhanced diamond chemical vapor deposition CVDhave
been investigated by line-of-sight infrared absorption spectroscopy in the wavenumber range of
1276.5−1273.1 cm−1 using a quantum cascade laser spectrometer. Parameters explored include
process conditions pressure, input power, source hydrocarbon, rare gas Ar or Ne, input gas mixing
ratio, height zabove the substrate, and time tafter addition of hydrocarbon to a pre-existing
Ar/H2plasma. The line integrated absorptions so obtained have been converted to species number
densities by reference to the companion two-dimensional r,zmodeling of the CVD reactor
described in Mankelevich et al. J. Appl. Phys. 104, 113304 2008兲兴. The gas temperature
distribution within the reactor ensures that the measured absorptions are dominated by CH4and
C2H2molecules in the cool periphery of the reactor. Nonetheless, the measurements prove to be of
enormous value in testing, tensioning, and confirming the model predictions. Under standard
process conditions, the study confirms that all hydrocarbon source gases investigated methane,
acetylene, ethane, propyne, propane, and butaneare converted into a mixture dominated by CH4
and C2H2. The interconversion between these two species is highly dependent on the local gas
temperature and the H atom number density, and thus on position within the reactor. CH4
C2H2conversion occurs most efficiently in an annular shell around the central plasma
characterized by 1400Tgas2200 K, while the reverse transformation C2H2CH4is favored
in the more distant regions where Tgas 1400 K. Analysis of the multistep interconversion
mechanism reveals substantial net consumption of H atoms accompanying the CH4C2H2
conversion, whereas the reverse C2H2CH4process only requires H atoms to drive the reactions;
H atoms are not consumed by the overall conversion. © 2009 American Institute of Physics.
DOI: 10.1063/1.3176971
I. INTRODUCTION
Microwave MWplasma activation of dilute hydrocar-
bon in hydrogen gas mixtures finds widespread use as a route
to growing high quality diamond films by chemical vapor
deposition CVDmethods.14As discussed previously,5the
H2molecules dissociate following excitation within the
plasma ball, and the resulting H atoms diffuse throughout the
reactor volume. The lack of efficient H atom loss processes
in these gas mixtures ensures that H atom densities in the
cooler periphery of the reactor are far in excess of those
expected on the basis of local thermodynamic equilibrium. H
atoms fulfill a number of important roles in the CVD
process—both in the gas phase and at the gas-surface inter-
face. The studies described here focus on aspects of the
former, where H atoms drive radical formation via the so-
called “H-shifting” abstraction reactions 1and 2and, in
the cooler regions, third-body stabilized H addition reactions
3and 4that culminate in the interconversion between
C1CHyand C2C2Hxspecies,1,6
CHy+HCHy−1 +H
2,y=4−1, 1
C2Hx+HC2Hx−1 +H
2,x=6−1, 2
C1Hy+H+MC1Hy+1+M,y0, 3
C2Hx+H+MC2Hx+1+M,x0. 4
Interconversions between CHyand C2Hxspecies occur on
timescales that are much shorter than the transit time for
feedstock gas through most CVD reactors, leading to the
expectation that the distribution of CHy,C
2Hx, etc., species
present in an activated hydrocarbon/H2gas mixture should
be rather insensitive to the choice of hydrocarbon feedstock
gas. Such a view is supported by early studies7demonstrat-
ing growth of diamond films with similar morphologies and
at similar rates from a range of different hydrocarbon source
aAuthor to whom correspondence should be addressed. Tel.: 117-
9288312/3. FAX: 117-9250612. Electronic mail:
mike.ashfold@bris.ac.uk.
JOURNAL OF APPLIED PHYSICS 106, 033305 2009
0021-8979/2009/1063/033305/15/$25.00 © 2009 American Institute of Physics106, 033305-1
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
gases at a constant input carbon mole fraction, and by Fou-
rier transform infrared absorption8and molecular beam mass
spectrometry9,10 measurements of the densities of various of
the more abundant stable species e.g., CH4,C
2H2, and
C2H4in MW activated CH4/H2and C2H2/H2gas mixtures.
Both of these gas phase diagnostic experiments had some
limitations, however. The former gave no spatial informa-
tion, while the latter only returned relative rather than abso-
luteconcentrations.
There is a continuing need for new and improved, quan-
titative, nonintrusive, in situ probes of the gas phase chem-
istry and composition, both to refine our understanding of the
diamond CVD process and for improved process control ap-
plications. Celii et al.11,12 demonstrated the potential appli-
cability of infrared tunable diode laser absorption spectros-
copy methods for detecting CH3radicals, and C2H2and
C2H4molecules, in HF activated CH4/H2gas mixtures in the
late 1980s, but the subsequent take-up of such methods has
been slow. In part, this is because such methods provide
line-of-sight absorbances, which can only be turned into ab-
solute column or numberdensities given detailed knowl-
edge of the generally very inhomogeneousdistribution of
species concentrations and temperatures along the viewing
column. Measurements of this kind are thus of limited quan-
titative value in the absence of companion high level reactor
modeling. Such detailed studies are now starting to appear,
however—notably measurements of CH3radicals, as well as
the stable species C2H2,CH
4, and C2H6, in MW activated
CH4/H2mixtures in a quartz bell jar reactor used for dia-
mond CVD.1315
We recently demonstrated16 the use of a quantum cas-
cade QClaser1719 for probing CH4and C2H2molecules,
and their interconversion, ina2kWMWreactor operating
with both CH4/Ar/H2and C2H2/Ar/H2feedstock gas mix-
tures, as a function of process conditions e.g., input hydro-
carbon mole fraction, total gas pressure, and applied MW
power. We also exploited the rapid output frequency sweep
“chirp”rate of pulsed QC lasers to gain insights into the
time evolution of the CH4/C2H2ratio when either is intro-
duced into, or removed from, a pre-existing Ar/H2plasma.
Here we present higher sensitivity single pass QC laser ab-
sorption measurements of CH4and C2H2molecules, and
their interconversion, in this MW reactor, as functions of the
above process conditions, the H2/Ar mixing ratio, and verti-
cal distance zabove the substrate surface for a wider vari-
ety of feedstock hydrocarbon source gases not only methane
and acetylene but also n-propane, n-butane, ethene, and pro-
pyne. These data serve to demonstrate the efficiency with
which all precursor hydrocarbons are processed to locally
equilibratedmixtures dominated by CH4and C2H2under
typical CVD conditions. As noted previously,1416 gas tem-
perature and number density considerations dictate that the
bulk of the monitored CH4and C2H2molecules are localized
in the cooler periphery of the reactor, remote from the
plasma ball. Complementary, spatially resolved, column den-
sity measurements of radical species C2aand CHX兲兴 and
of electronically excited Hn=2atoms—obtained by cavity
ring down spectroscopy—are described elsewhere.20 These
transient species are concentrated in the plasma ball itself.
Both data sets serve to tension and validate the detailed two-
dimensional 2Dmodeling of the gas phase chemistry pre-
vailing in this high pressure MW CVD reactor.5
II. EXPERIMENTAL
Details of the custom-designed MW reactor 2 kW, 2.45
GHz Muegge power supply and generatorand of the QC
laser spectrometer Cascade Technologies Ltd., operating
with a 7.85
m laserhave been presented previously.16
Here we detail improvements implemented since the initial
study. One enhancement involved changes to the reactor
viewing apertures. The 4 mm diameter apertures that defined
the previous viewing column have been replaced by two
25 mmvert.5.5 mmhoriz.slot apertures, which allow
operation of the reactor in two configurations as illustrated in
Fig. 1. In configuration I, each slot is sealed by thin rectan-
gular polished CVD diamond windows Element Six Ltd.
mounted on wedged flanges to reduce etalon effects, thereby
maintaining the same l=19 cmcolumn length as before.
Alternative, matched flanges are used in configuration II.
Each of these flanges is fabricated with a tongue that projects
into the slot aperture and supports three 5.8 mm diameter,
wedged , thin 300
mdiamond windows the centers
of which are, respectively, 1, 11, and 21 mm above the sub-
strate surface. Countersinking the windows reduces the col-
umn length to l=14 cm, thereby reducing the dominant
contribution to absorption from cold gas at the ends of the
viewing column. The second improvement allows for spatial
profiling. The complete optical assembly laser, beam steer-
ing optics, and detectoris now mounted on a rigid platform
that can be translated vertically, with better than 1 mm pre-
cision, relative to the fixed MW reactor—thereby enabling
spatially resolved, line-of-sight column density measure-
ments with configuration I as a function of z, the vertical
distance above the top surface of the 30 mm diameter Mo
substrate. zvalues were read from a rigidly mounted vernier
scale, and z=0 determined by finding the platform setting at
which the substrate first impeded transmission of the probe
laser beam. The third improvement involves the detectivity
of the QC laser spectrometer. The liquid nitrogen cooled
mercury cadmium telluride MCTdetector with external
transimpedence amplifier used previously to monitor the
FIG. 1. Schematic diagram of the CVD reactor illustrating the position of
the substrate, plasma ball, and the windows for optical probing using con-
figurations I and II.
033305-2 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
transmitted IR intensity was replaced by a new photovoltaic
MCT detector with a two stage Peltier cooler and a built-in
amplifier Vigo. This has a built-in immersion lens en-
abling improved radiation collectionand a wider bandwidth
that removes most of the overshoot of the pulsed laser sig-
nal. As illustrated below, this detector returns spectra with
substantially better signal to noise ratios, allowing observa-
tion of many weak spectral features not recognized in the
original study.
The source gases used, along with suppliers and stated
purities, were as follows: CH4BOC, 99.5%,C
2H2BOC,
98.5%, ethene Argo, 99.7%,n-propane Argo, 99%,
n-butane Argo, 99.4%, propyne Argo, 96%,H
2BOC,
99.995%, and Ar and Ne BOC, 99.995%. The hydrocar-
bon, H2, and ArNegas flows were metered through three
separate mass flow controllers MFCs兲共mks, calibrated for
CH4,H
2, and Ar; flow rates of the neon and the various
nonmethane hydrocarbons were derived using the manufac-
turer correction factors where availableand, in the case of
C2H2, verified independently by IR absorption spectroscopy
see Tables Iand II.
III. RESULTS AND DISCUSSION
A. Preamble
To aid the subsequent discussion, absorption spectra of
room temperature samples of CH4and C2H2recorded over
the wavenumber range of 1276.5−1273.1 cm−1 using reac-
tor configuration I are shown in Figs. 2and 3. The pressures
in each case have been chosen deliberately so that the stron-
gest features are saturated, thereby allowing weaker features
to be seen more clearly. Lines evident in Fig. 2are attribut-
able to rovibrational transitions of the 40
1fundamental bands
of 12CH4and 13CH4and associated hot bands originating
from both the 21and 41levels of 12CH4,21 as detailed by the
assignment combs in the figure and the listing in Table I. The
most intense features in the case of C2H2Fig. 3are P
branch transitions within the 40
150
1combination band of the
12C2H2isotopomer.21 Hot band absorptions are again identi-
fiable, originating from levels carrying one and two quanta
of excitation in
4and/or
5the trans- and cis-bending
modes, respectively22—as detailed in Table II. Several weak
features remain unassigned in Fig. 3; these are most likely
attributable to H12C13CH.
Measurements of room temperature gas mixtures pro-
vide a means of testing the absolute calibration of the MFCs.
Plot bin Fig. 2shows the CH4number densities obtained
by plotting the integrated absorbances in cm−1of the
4F23 –5F12 line measured for various
a%CH4/7%Ar/balance H2mixtures, scaled by the appropri-
ate S298 Kvalue 3.6710−20 cm−1/molecule cm−2
Ref. 21兲兴 and then divided by =19 cm, against those cal-
culated from the set flow rates and pressure assuming ideal
gas behavior. The gradient of this straight-line correlation is
TABLE I. Wavenumbers, assignments, and STline strength factors Ref. 21of CH4transitions observed in
the wavenumber range of 1276.5− 1273.1 cm−1.
Wavenumber
cm−1
Line intensity
10−24 cm−1mol cm−2−1
E
cm−1
Vibrational
transition
Rotational
transition
C
isotope298 K 450 K 3000 K
1276.331 05 123 639 8.21 1433.9722 41
23F25–4F13 12
1276.262 06 23.5 55.7 0.192 949.888 40
113F14–13F21 13
1275.945 66 564 2180 17 1250.836 40
115F24−15F11 12
1275.947 76 564 2180 17 1250.837 40
115F15–15F21 12
Sum兲共1128兲共4360兲共34
1275.779 28 671 400 0.139 104.778 40
13A21–4A11 13
1275.621 60 1.51 11.0 0.248 1639.651 21
140
14F29−4F15 12
1275.386 78 24 400 15 900 6.35 157.137 40
14E2–5E112
1275.326 54 401 240 0.083 104.780 40
13F22–4F11 13
1275.041 68 36 700 23 900 9.53 157.139 40
14F23–5F12 12
1274.984 82 269 159 0.0552 104.781 40
13E2–4E113
1274.786 13 86.6 62.3 0.0295 219.937 40
16F11–6F22 12
1274.213 93 79.3 413 5.28 1432.524 40
13E3–4E212
1274.016 90 398 238 0.0823 104.785 40
13F13–4F21 13
1273.881 63 34.7 174 2.09 1409.376 41
23E2–4E112
1273.875 25 1.66 60.4 19.7 2623.659 40
122A23−22A11 12
1273.864 56 1.00 36.3 11.8 2623.677 40
122F28−22F12 12
1273.858 75 0.669 24.3 7.90 2623.687 40
122E5−22E212
Sum兲共3.329兲共121兲共39.4
1273.840 15 11.1 119 5.12 1878.424 41
210A13–10A22 12
1273.782 40 13.6 171 9.56 1975.015 40
119E4−19E112
1273.782 49 20.4 257 14.4 1975.013 40
119F17–19F22 12
1273.782 57 34.0 428 23.9 1975.008 40
119A13–19A21 12
Sum兲共68.0兲共856兲共47.86
1273.469 77 52.2 261 3.13 1409.007 41
23F14–4F22 12
1273.418 31 120 626 7.99 1432.181 41
23F15–4F23 12
033305-3 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
reassuringly close to 1, validating the ideal gas assumption
and the MFC and pressure gauge calibrations. The C2H2feed
gas used in these experiments was metered with the same
MFC i.e., calibrated for CH4. Plot bin Fig. 3compares
the corresponding measured via the P23 ,eline of 12C2H2
and assuming S298 K= 2.2410−20 cm−1/
molecule cm−2兲共Ref. 21兲兴 and calculated using the ideal
gas lawC2H2number densities. The calculated column den-
sities in this plot again assume that the various
b%C2H2/7%Ar/balance H2mixtures behave ideally and, in
addition, that the MFC calibration factor is the same for CH4
and C2H2. The gradient of the resulting plot 共⬃0.95reflects
ithe purity or otherwiseof the C2H2source gas and ii
the reduced conductance of the CH4MFCfor C2H2.
One other aspect of these spectra and line assignments
merits comment. Although narrow, the probed wavenumber
range allows both CH4and C2H2to be monitored in a spread
of energy levels. In the case of CH4, for example, this range
allows sampling of v=0 molecules with both low e.g., J
=4and high e.g., J=19rotational quantum numbers and
of molecules in levels with v2=1 and v4=1 low Jin each
case. The relative populations of these levels, and thus the
associated transition line strengths ST, are temperature de-
pendent. Under CVD conditions, the gas temperatures along
TABLE II. Wavenumbers, assignments, and ground state term values of
C2H2transitions observed in the wavenumber range of 1276.5
1273.1 cm−1 Refs. 21 and 22.
Wavenumber
cm−1a,b
Ea
cm−1
Vibrational
transition
Spin
weighting
Rotational
transitionb
1276.336 63 1262 41
250
1ug1P23, f
1276.258 06 1708 40
152
3ug1P14, e
1276.140 79 1775 41
251
2gu1P19, f
1275.958 59 1177 40
151
2gu3P19, e
1275.716 61 1743 40
152
3ug1P15, f
1275.614 95 1881 42
350
1ug1P23, f
1275.580 94 1881 42
350
1ug3P23, e
1275.566 59 1743 40
152
3ug3P15, e
1275.512 22 649 40
150
1u
+g
+3P23, e
1275.423 25 1792 41
251
2gu3P19, e
1275.374 66 1262 41
250
1ug3P23, e
1275.288 62 1792 41
251
2gu1P19, f
1274.647 40 1822 41
251
2gu3P20, f
1274.479 50 1177 40
151
2gu1P19, f
1274.362 03 1167 40
150
1u
eg
+e3P31, ec
1274.156 39 1318 41
250
1ug3P24, f
1273.974 56 1732 40
152
3ug3P15, e
1273.857 01 1822 41
251
2gu3P20, f
1273.819 72 1225 40
151
2gu1P20, f
1273.516 79 1781 40
152
3ug3P16, f
1273.452 72 1995 42
350
1u
+g
+3P25, e
1273.366 60 1937 42
350
1ug3P24, f
1273.319 71 1781 40
152
3ug1P16, e
1273.261 95 706 40
150
1u
+g
+1P24, e
1273.103 87 1319 41
250
1ug1P24, e
aApproximated using values of rotation constants and band origins in Ref.
22.
beand findicate the l-doubling components of the and vibrational
states.
cTransition is due to rotational l-resonance between u
eand u
+esublevels of
the same Jvalue, see Ref. 25.
FIG. 2. aSingle pass transmission spectra of a room temperature sample
of CH4in the wavenumber range of 1276.5− 1273.1 cm−1.bshows the
CH4number density obtained by plotting the integrated absorbances in
cm−1of the 40
14F23–5F12 line measured for various
a%CH4/7%Ar/balance H2mixture, scaled by the appropriate S298 K
value and then divided by =19 cm, against those calculated from the set
flow rates and pressure assuming ideal gas behavior.
FIG. 3. aSingle pass transmission spectra of a room temperature sample
of C2H2in the wavenumber range of 1276.5−1273.1 cm−1.bshows the
C2H2number density obtained by plotting the integrated absorbances in
cm−1of the 40
150
1P23, eline measured for various
b%C2H2/7%Ar/balance H2mixture, scaled by the appropriate S298 K
value and then divided by =19 cm, against those calculated from the set
flow rates and pressure assuming ideal gas behavior.
033305-4 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
the probed column span the range from room temperature to
3000 K.5,20 As Table Ishows, our relative sensitivity to
CH4v=0,J=19molecules rises 10-fold and subsequently
declines over this temperature range, while our sensitivity to
CH4v=0,J=4molecules drops by 5103. Even modest
temperature changes can cause significant changes in Sfac-
tor, as exemplified by the case of the 41
2P4transition on
increasing Tfrom 300 to 450 K. The Tdependence in the
case of C2H2is less dramatic, since all of the levels moni-
tored involve broadly similar Jvalues J20. However, as
Table II illustrates, the probed wavelength range allows
access to C2H2molecules with v=0 E650 cm−1,
v4/5=1 E1150–1300 cm−1, and v4/5=2 E
1700–2000 cm−1. The temperature dependence of the vi-
brational partition function will ensure greatest relative sen-
sitivity to the most vibrationally excited C2H2molecules at
higher T. The HITRAN database21 currently lists STfactors
for C2H2v=0molecules only, but STfactors for other
transitions of interest with v0can be estimated as shown
in the Appendix.
B. Spatially resolved CH4and C2H2absorption and
number density profiles
Figure 4compares and contrasts absorption spectra re-
corded in the narrower wavenumber range of 1276.2
1274.0 cm−1 with both reactor configurations, at z
=11 mm, using x%CH4/7%Ar/balance H2Fig. 4aand
y%C2H2/7%Ar/balance H2Fig. 4bgas mixtures at a to-
tal flow rate of F=565 SCCM SCCM denotes standard cu-
bic centimeter per minute at STP, pressure p= 150 Torr,
and applied MW power P= 1.5 kW, as functions of the car-
bon flow rate. For ease of display, the two spectra for each
carbon flow recorded with the different experimental con-
figurations have been offset vertically by 0.05 absorbance
units, and the pairs of spectra recorded with different carbon
flow rates have each been offset vertically by larger amounts.
The configuration I measurements are comparable to those
reported previously16 and have thus been repeated at fewer
carbon flow rates. The line at 1275.042 cm−1 is due solely to
CH4v=0molecules, whereas both CH4v=0and
C2H2v4=1molecules can contribute to the blended line at
1275.38 cm−1. The peaks at 1275.512 and 1274.156 cm−1
are both entirely due to C2H2molecules—in their v= 0 and
v4=1 levels, respectively. The more detailed analysis of the
respective parent room temperature absorption spectra re-
ported in Tables Iand II shows the remaining line, at
1275.95 cm−1, to be a blended feature, involving contribu-
tions from both C2H2v5=1and CH4v=0,J=15
molecules—rather than being wholly associated with
C2H2v5=1molecules as presumed previously.16 Figures
4cand 4dshow the line integrated absorbances LIAs
derived for the various unblended peaks, for both reactor
configurations, plotted as functions of “carbon” flow rate
FIG. 4. Absorption spectra of activated ax%CH4/7%Ar/H2and by%C2H2/7%Ar/H2gas mixtures using three different carbon flow rates, recorded over
the wavenumber range of 1276.5− 1274.0 cm−1.KeyCH
4and C2H2features are labeled; detailed assignments of these transitions are given in Tables Iand
II. The upper/lower spectrum in each pair was recorded using configuration I/II. For display purposes, the upper of the two spectra measured at each carbon
flow rate has been raised vertically by 0.05 absorbance units, and each pair of spectra in each panel has been offset vertically by larger amounts. The total flow
rate F=565 SCCM, pressure p=150 Torr, and applied MW power P= 1.5 kWwere the same for each of the spectra. cand dshow the LIAs for the
CH440
14F23–5F12 ,C
2H240
150
1P23, e兲共, and C2H241
250
1P24, f兲共lines of interest, and their variation with carbon flow rate.
033305-5 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
where CH4and C2H2flow rates of xand ySCCM equate to
carbon flow rates of, respectively, xand 2ySCCM.
Several points emerge from these data. The same peaks
are apparent in all spectra recorded at a given carbon flow
rate, irrespective of the identity of the input hydrocarbon,
again emphasizing the efficient methane acetylene inter-
conversion. The absorption measured at low 5 SCCMcar-
bon flow rate is largely attributable to CH4, but C2H2absorp-
tion features become increasingly evident at higher carbon
flow rates. The previous suggestion16 of a population inver-
sion between the v5=1 and v=0 levels of C2H2at low car-
bon flow rates is incorrect. The 1275.95 cm−1 feature mea-
sured at low carbon input fractions is actually attributable to
rotationally “hot” CH4molecules. None of the data reported
here contradicts the assumption that the rotational and vibra-
tional state populations of these stable hydrocarbon species
are in local thermodynamic equilibrium. The nonobservation
of the cold13CH4feature at 1275.7793 cm−1 in spectra
recorded at low carbon flow rate serves to reinforce the very
different STdependences of this line and of neighboring
hotCH4feature at 1275.95 cm−1—detailed in Table I.
Source gas dependent differences also become more evident
in the spectra recorded at higher carbon flow rates. The spec-
trum recorded with 40 SCCM CH4is dominated by the CH4
feature, whereas the largest peak in the spectrum recorded
with 20 SCCM C2H2is due to C2H2. These trends can be
seen more clearly in the LIA plots Figs. 4cand 4dand
will be discussed in more detail in Sec. III D.
Figures 4cand 4dalso highlight the differences in
LIAs measured using configurations I and II. Reducing the
length of the viewing column from l=19 to l= 14 cm results
in varying reductions in all LIAs by 66%in the case of
CH4and by, respectively, 50%and 33%in the case of
C2H2v=0and C2H2v4=1molecules, reflecting their dif-
fering spatial distributions but also reinforcing the model
predictions5that most of the CH4and C2H2number density
is concentrated in the cooler regions near the wall of the
reactor. Spectra taken at the lowest 5 SCCMcarbon flow
rates show no discernible C2H2v=0absorption. Thus it is
valid to attribute all of the absorption at 1275.95 cm−1 in
such spectra to the CH440
1Q15F21 and F11 transitions. Their
combined intensity, relative to that of the CH440
1P5F12 line
at 1275.04 cm−1, provides a route for estimating the gas
temperature Tgasin the columns bounded by the water
cooled reactor wall that distinguish configurations I and II.
Specifically, we calculate the difference in the LIAs LIA
measured for each line under identical process conditions
and compare this ratio with the temperature dependentratio
of the respective STline strength factors. The latter ratio
varies rapidly with Tand matches LIA when
T330 K—a sensible value for the average gas temperature
in this region. Knowing Tgas and the difference in path
lengths l=5 cm, we can use the measured LIA values
and the appropriate ST=330 Kline strength factor to esti-
mate the average CH4number density in this region. The
value so derived, NCH4 4.41016 cm−3, matches well to
within 20%with the ideal gas estimate at this Tgas, given the
assumption that CH4is the only hydrocarbon present under
these dilute carbon conditions. Equivalent calculations at
higher carbon flow rates, and at all three zpositions 1, 10.5,
and 20.5 mm, return CH4and C2H2number density esti-
mates that agree well to within factors of 1.3 for CH4and
2 for C2H2with the corresponding r= 6 cm values re-
ported in Ref. 5.
Figure 5shows absorption spectra of activated a
CH4/Ar/H2and bC2H2/Ar/H2gas mixtures recorded in
the narrow wavenumber range of 1274.1−1273.0 cm−1,
which contains an isolated line associated with C2H2,v=0,
J=24 molecules and a blended clump of lines around
1273.8 cm−1. The absence of the 1273.262 cm−1 feature in
spectra recorded at the lowest carbon flow rate accords with
the conclusion from Fig. 4that any features evident in such
spectra are likely attributable to CH4. Thus the very weak
absorption evident at 1273.78 cm−1 in the FCH4
=5 SCCM spectrum in Fig. 5ais most plausibly assigned
to a cluster of blended lines associated with CH4v=0mol-
ecules in a highly excited J=19rotational level. Additional
features are apparent at slightly higher wavenumber in spec-
tra recorded at higher carbon flow rates. Inspection of Tables
Iand II reveal several potential contributors to this clump.
FIG. 5. Absorption spectra of activated ax%CH4/7%Ar/H2and by%C2H2/7%Ar/H2gas mixtures using three different carbon flow rates, recorded over
the wavenumber range 1274.1− 1273.0 cm−1 using configuration II at z= 21 mm. Selected CH4and C2H2features are indicated; detailed assignments of these
transitions are given in Tables Iand II. For display purposes, the various spectra have been offset vertically. The total flow rate F=565 SCCM, pressure
p=150 Torr, and applied MW power P=1.5 kWwere the same for each of the spectra.
033305-6 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
Two hot band absorptions associated with CH4v4=1mol-
ecules lie in this range, with associated STfactors that are
comparable to that of the 40
1P19 line once T450 K, but
the clear showing of the 1273.262 cm−1 feature indicates
that we need to consider possible contributions from C2H2
also—specifically the two P20transitions at
1273.857 cm−1 originating from the 4151leveland at
1273.820 cm−1 from the v5=1 level. These two transitions
have respective nuclear spin weightings of 3 and 1. At room
temperature, the effect of the population difference far out-
weighs the different nuclear spin statistics, and the lower
wavenumber line is dominant see Fig. 3. Under base oper-
ating conditions, however, the two lines appear with similar
LIAs—reflecting the convergence in relative populations of
the two levels as Tincreases. As in Fig. 4, the C2H2features
are seen to grow in relative importance at higher carbon flow
rates and, at a given carbon flow rate, to be consistently
greater when using C2H2as the source gas.
Figure 6shows the calculated5rdependence of the total
number densities of CH4,C
2H2,C
2H4, and C2H6and of Tgas
for the standard process condition 25 SCCM CH4,40
SCCM Ar, 500 SCCM H2,p= 150 Torr, P= 1.5 kWat z
=0.5 mm Fig. 6aand z=10.5 mm Fig. 6b. These
clearly illustrate the way in which Tgas effects and gas phase
chemistry combine to concentrate the stable hydrocarbon
species in the cool regions at the periphery of the reactor.
Even with the recessed windows, therefore, the measured
LIAs will be dominated by gas at the ends of the probed
column and any detailed comparison between experiment
and model calculation requires that we give due consider-
ation to the differences between the model and experimental
geometries. Specifically, the probed column in configuration
II extends to r=7 cm, and it is therefore necessary to ex-
trapolate the model outputs for a further 1 cm. The recessed
window mount is in good thermal contact with the reactor
wall so, for the purpose of this exercise, we choose to retain
the model predictions out to the largest calculated rvalue
and assume that thereafter Tdeclines linearly to 300 K at r
=7 cm. We also assume that the mole fraction Xof each
species remains constant in the 1 cm extension region, but
note that the model outputs actually show the mole fractions
of all stable hydrocarbon species increasing by factors of
1.5–2 over the range of 5.5r6 cm, presumably as a re-
sult of thermodiffusion effects. This allows estimation of
nCH4,nC2H2,Tgas, the relevant STfactors, and thus the re-
spective contributions to the total column absorption from
each cell usinga1mmgrid spacing in the region of 6 r
7 cm. Summing all contributions modeled and extrapo-
latedfrom the range of −7 r7 cm yields a calculated
LIA for direct comparison with experimental measurements
at the various process conditions.
The 2D model takes account of the changes in plasma
parameters and conditions e.g., Tgas, the electron tempera-
ture Te, and concentration ne, the power density and the
plasma chemistryinduced by varying reactor parameters
such as p,P, and the mole fractions of CH4and Ar in the
process gas mixture. Detailed descriptions of this model pro-
cedure, which uses the plasma size as an external parameter,
are presented in accompanying publications.5,20,23 By way of
example, the maximal values of neand Teare predicted to
increase by 30%and 5%, respectively, upon introducing
just 5 SCCM of CH4i.e., 0.88% CH4into an Ar/H2plasma
operating under what, otherwise, would constitute base
conditions.5,23 Increasing the CH4flow rate further, up to the
base value FCH4=25 SCCM i.e., 4.4% CH4leads to less
pronounced changes in nea further 5%increaseand Tea
3%decline relative to that at FCH4=5 SCCM. The 2D
model returns the following typical values for the plasma
parameters in the plasma core: Te1.25–1.45 eV, Tgas
2800–2950 K, power densities of 20 40 W cm−3, re-
duced electric fields E/N25–30 Td, and ne3
1011 cm−3.5,20,23 The effects of heavy 90%dilution of
FIG. 6. Plots showing the rdependent CH4,C
2H2,C
2H4, and C2H6number
densities and Tgas values calculated for the standard process conditions
FCH4=25 SCCM, FAr=40 SCCM, FH2=500 SCCM, p
=150 Torr, P= 1.5 kWat az= 0.5 mm and bz= 10.5 mm. Panel c
shows a decomposition of the calculated rdependent C2H2total number
density shown in binto number densities in the following quantum states:
v=0, J=23; and individual -doublets of v5=1, J=20; v4=1, J=24; and
v4=1+v5=1, J=20. Vertical dashed lines indicate the centers of gravity of
the various state-specific number density distributions.
033305-7 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
the process gas mixture by various noble gases He, Ne, Ar,
Kron both the plasma parameters and on nanocrystalline
diamond deposition processes have also been investigated24
experimentally and theoretically using the present 2D model
for both Ar/H/C and He/H/C mixtures.
Table III compares the LIAs derived from such analyses
of the 2D model outputs for the standard operating condi-
tions at three different zvaluesand for three alternative p,
P, and FCH4conditions, with the corresponding values
determined from measurements on the 1275.042 cm−1 line
of CH4v=0and the 1275.512 cm−1 line of C2H2v=0.
The model calculations reproduce the experimentally mea-
sured CH4LIAs rather well, under all process conditions, but
consistently underestimate the corresponding experimental
LIA value for C2H2. Most of the C2H2and, particularly, the
CH4absorption is from gas at the periphery of the reactor.5
The STline strength factor for the P23line used when
monitoring C2H2v=0molecules falls with deceasing Tfor
T450 K, and we are unable to conceive of a physically
plausible alternative Tgasrdependence at large rthat would
allow the calculated XC2H2 value to give a significantly larger
absorbance. This leads us to suspect that the model calcula-
tion underestimates XC2H2 in the cooler region of the viewing
column. We can envisage two factors that may contribute to
this underestimation. First, the predicted LIAs would be
greater if XC2H2 and XCH4increased through the region 6
r7 cm, rather than remain constant as currently as-
sumed. Thermodiffusion considerations would favor the
former scenario, but we see no robust way of predicting the
necessary Xr6cmdependences. Second, the present cal-
culations return C2H4and C2H6number densities of
1016 cm−3 at large rsee Fig. 6. It is quite possible that
aspects of the gas phase chemical mechanism at large ror,
more probably, the assumed gas-surface chemistry occurring
at the reactor walls that determines the interconversion be-
tween the various hydrocarbon species at low Tgas and thus
the local C2H2number densityneeds further refinement.
Given local thermodynamic equilibrium, the partition
functions calculated in the Appendix allow estimation of the
fraction of nC2H2 in any given v,Jlevel at a given T. Figure
6cshows the relevant decomposition for probed levels
highlighted in Figs. 4and 5, namely, the v=0, J= 23 level
E=649 cm−1and for individual -doublets of the v5=1,
J=20 E=1225 cm−1,v4=1, J=24 E= 1318 cm−1, and
v4=1+v5=1, J=20 E=1822 cm−1levels—plotted as a
function of r. As expected, the centers of gravity of the pre-
dicted state-specific number density distributions show a pro-
gressive shift to smaller rwith increasing Er
¯
6.5, 5.9,
5.8, and 5.7 cm, respectivelybut the effect is quite
modest—emphasizing, once again, the localization of num-
ber density in the cooler periphery of the reactor. The last
two columns in Table III compare LIAs for the unblended
C2H241
250
1P24, ffeature at 1274.156 cm−1 measured un-
der a range of process conditions with those calculated using
the predicted state-specific number density distributions and
the appropriate STfactors derived as outlined in the Ap-
pendix. Again, the experiment returns values that are consis-
tently two or more times greater than the model calculations,
with the greatest discrepancy at small z.
Figure 7shows the LIAs for the CH440
14F23 –5F12,
C2H240
150
1P23,eand C2H241
250
1P24, ftransitions for a
25 SCCM CH4/40 SCCM Ar/500 SCCM H2and b12.5
SCCM C2H2/40 SCCM Ar/512.5 SCCM H2gas mixtures
operating at standard conditions of total pressure and input
power. The LIA of each of the monitored species is seen to
decline with increasing zand, in each case, the hydrocarbon
source gas shows relatively more strongly. Reference to Fig.
7and to Table III suggests that the 2D model is rather suc-
cessful in capturing the zdependence of the absoluteCH4
and relativeC2H2densities at large r.
C. Variation in CH4and C2H2LIAs with process
conditions
The upper two panels in Fig. 8show the ways in which
the LIAs for the same three probe transitions measured at z
=11 mm vary with input MW power under otherwise stan-
dard conditions; all three decline with increasing P, irrespec-
tive of whether the hydrocarbon source is provided by a
FCH4=25 SCCM or bFC2H2= 12.5 SCCM. The
lower two panels show corresponding measurements illus-
trating the effect of varying the total process gas pressure p
while otherwise retaining standard process conditions of c
CH4/Ar/H2and dC2H2/Ar/H2mixing ratio, flow rate,
and input MW power. The LIA of the source gas CH4is
always greatest in the former case, and the measured LIAs
all increase roughly linearly with increasing p. In the case of
the C2H2/Ar/H2input gas mixture, however, the LIAs of the
C2H2v=0and C2H2v4=1transitions both increase near
linearly with p, while that for CH4rises much less steeply.
TABLE III. Comparison of measured expand predicted modLIAs for the CH440
1P5F12,C
2H240
150
1P23, e, and C2H241
250
1P24, ftransitions as a
function of process conditions.
Process conditions CH440
1P5F12 C2H240
150
1P23, eC2H241
250
1P24, f
FCH4
SCCM
p
Tor r
P
kW
d
mm
LIAexp
cm−1
LIAmod
cm−1
LIAexp
cm−1
LIAmod
cm−1
LIAexp
cm−1
LIAmod
cm−1
25 150 1.50 1 0.0104 0.0088 0.0066 0.0024 0.000 96 0.000 27
25 150 1.50 11 0.0070 0.0076 0.0036 0.0020 0.000 48 0.000 21
25 150 1.50 21 0.0065 0.0085 0.0034 0.0019 0.000 34 0.000 21
5 150 1.50 11 0.0034 0.0030 0 0.0002 0 0.000 02
25 150 1.25 11 0.0081 0.0085 0.0040 0.0022 0.000 46 0.000 23
25 75 1.50 11 0.0026 0.0035 0.0026 0.0010 0.000 32 0.000 11
033305-8 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
Reference to Table III shows, again, that the 2D model cal-
culations for the standard CH4/Ar/H2gas mixture captures
the observed Pand pdependences of the CH4LIAs quanti-
tatively and the relative variation in the C2H2LIAs.
The interpretation of these apparently simple trends re-
quires consideration of several contributory factors. We start
by considering the Pvariations and the case of a CH4/Ar /H2
gas mixture. The 2D model calculations for P= 1.25 and 1.5
kW show that increasing Presults in a slightly hotter, larger
plasma region, some increase in Tgas at large rand a higher H
atom density nHthroughout the reactor. The increase in Tgas
leads to a steeper temperature and thus total number den-
sitygradient at large r, which has the effect of concentrating
more of the number density in the coolest periphery of the
reactor and thus reducing the effective length of the column
that contains most of the cold hydrocarbon gas. The LIA of
all stable hydrocarbons thus declines with increasing P.
Scrutiny of Fig. 8ashows that the decline in the C2H2LIA
is less steep than that for CH4. Two factors contribute to this
trend. First, higher Presults in higher nHand thus a greater
processing efficiency of the input CH4which is the source
of the C2H2under these conditions. Second, the STdepen-
dence of the 40
150
1P23,etransition used to monitor
C2H2v=0molecules increases with Tin the range of 300
T450 K, so any increase in Tgas at large ras a result of
FIG. 7. Plots showing the zdependences of the LIAs for the CH440
14F23–5F12 ,C
2H240
150
1P23, e兲共and C2H241
250
1P24, f兲共transitions measured
for aFCH4=25 SCCM, FAr= 40 SCCM, FH2= 500 SCCM, and bFC2H2= 12.5 SCCM, FAr=40 SCCM, FH2= 512.5 SCCM gas mixtures
operating at standard conditions of total pressure and input power.
FIG. 8. Illustration of the Pand pdependences of the LIAs for the CH440
14F23–5F12 ,C
2H240
150
1P23, e兲共, and C2H241
250
1P24, f兲共transitions
measured at z=11 mm for 25 SCCM CH4/40 SCCM Ar /500 SCCM H2关共aand c兲兴 and 12.5 SCCM C2H2/40 SCCM Ar /512.5 SCCM H2关共band
d兲兴 gas mixtures operating at standard conditions of total pressure 关共aand b兲兴 and input power 关共cand d兲兴. The open symbols in aand cshow the
corresponding LIAs for the CH440
14F23–5F12 and C2H240
150
1P23, etransitions returned by the 2D model calculations.
033305-9 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
increasing Pwill have a positive effect on the measured LIA.
The data for the C2H2/Ar/H2gas mixture Fig. 8bdisplay
slightly different trends. In this case, the LIA of the source
gas C2H2shows the steeper decline with increasing P. This
can be understood if the main effect of increasing Pis to
increase Tgas and nHand thus the extent of gas processing by
which the CH4is produced.
We now consider the observed variations in LIAs with p.
The 2D model calculations for the standard CH4/Ar/H2gas
mixture show that doubling pfrom 75 Torr to the standard
150 Torr results in an approximately fourfold increase in nH
at the center of the plasma ball equivalent to an approxi-
mately twofold increase in the H atom mole fraction, but a
50-fold drop in nHfrom 31010 to 6108cm−3at
r=6 cm. The same calculations show little change in Tgas at
r=6 cm as a result of this increase in p. The reduction in nH
at large ris mainly attributable to the p−1 dependence of the
diffusion coefficients and thus the H atom diffusional trans-
fer rate from the source region the plasma ball, but will be
exacerbated by the increased opportunity for reactive loss
through reaction with hydrocarbon species at higher p. This
latter point may be significant in explaining the pdepen-
dence of the LIAs measured with the C2H2/Ar/H2gas mix-
ture. In this case, the LIA of the source hydrocarbon in both
its v=0 and v4= 1 levelsincreases roughly linearly with p,
as expected, but that for CH4grows more slowly at higher p.
All of the CH4in this mixture is formed by reactions involv-
ing H atoms, principally in regions of moderate Tgas. The
reduction in nHat large rat high pwill lead to a reduction in
C2H2CH4conversion probability in the peripheral regions
where most of the monitored CH4is located, thus account-
ing for the observed negative curvature in the CH4LIA at the
highest pstudied.
Similar considerations can account for the observed
variations in the LIAs for the same three probe transitions
when varying the Ar/H2or Ne/H2ratio in
CH4/ArNe/H2and C2H2/ArNe/H2gas mixtures. Fig-
ures 9aand 9bshow sample absorption spectra measured
at z=11 mm for CH4/Ne/H2and C2H2/Ne/H2gas mix-
tures. The carbon and total flow rates were the same in all
cases 25 and 565 SCCM, respectively, as were p150 Torr
and P1.5 kW, with the only variable being the Ne/H2
ratio. Figures 9cand 9dshow how the LIAs for the three
unblended lines of interest vary with Ne flow rate. Equiva-
lent measurements involving CH4/Ar/H2and C2H2/Ar/H2
gas mixtures return virtually identical LIA versus FAr
plots. The most striking feature of these plots is the very
different behavior of the C2H2and CH4LIAs; the former are
relatively insensitive to the Ne/H2ratio or actually increase
in the case that CH4is the source gas, whereas the CH4LIA
falls more than fivefold over the investigated range of Ne
flow rates—with the result that the strongest absorption at
high FNeis due to C2H2v=0molecules. Indeed, in the
case of the C2H2/Ne/H2gas mixture, vibrationally hot
C2H2v5=1molecules contribute 50%of the LIA of the
blended feature at 1275.3 cm−1, with the result that it ac-
tually appears more intense than the neighboring CH4v=0
feature at 1275.04 cm−1. This should be contrasted with the
FIG. 9. Absorption spectra of activated a25 SCCM CH4/xSCCM Ne/540xSCCM H2and b12.5 SCCM CH4/xSCCM Ne /552.5
xSCCM H2gas mixtures using three different Ne flow rates, recorded at z=11 mm over the wavenumber range of 1276.2 1274.0 cm−1. Three unblended
CH4and C2H2features are labeled; detailed assignments of these transitions are given in Tables Iand II. The total p150 Torrand P1.5 kWwere the same
for each of the spectra. cand dshow the LIAs for the three unblended lines of interest and their variation with Ne flow rate.
033305-10 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
spectrum obtained with the 2.5 SCCM C2H2/40
SCCM Ar/522.5 SCCM H2gas mixture Fig. 4b, which
is dominated by CH4absorptions and the higher frequency of
these two lines is just 2/3 as intense as the 1275.04 cm−1
feature consistent with the respective lower state degenera-
cies see Table I兲兴. The rationale for the present observations
is similar to that used to explain the observed pdependent
variations in LIA. Substituting H2with rare gas Ne or Ar
reduces the thermal conductivity of the gas, resulting in an
increase in Tgas in the plasma region. This facilitates CH4
C2H2conversion in the hotter regions and explains the
increase in C2H2LIA with increasing FNein Fig. 9c.
Increased Tgas in the plasma region will encourage H2disso-
ciation and thus an increase in nH, but such increases will
be countered by the reduction in FH2. As in the case of
increasing p, therefore, we can anticipate that increasing
FNewill result in a progressive decrease in nHat larger r
and thus a reduced probability for CH4formation via reac-
tion 3in the cooler regions of the reactor—consistent with
the observed decline in CH4LIA.
D. Use of other hydrocarbon feedstock gases
Four other hydrocarbon source gases ethane, propyne,
n-propane, and n-butanewere investigated in addition to
CH4and C2H2; all were found to show broadly similar be-
havior. The scan range of the QC laser spans part of the 90
2
absorption band of propyne, as illustrated by the spectrum of
a 5 Torr room temperature sample of propyne shown in Fig.
10. This absorption is also clearly evident in spectra of
8.3 SCCM propyne/40 SCCM Ar/516.7 SCCM H2gas
samples at room temperature and p= 150 Torr, but com-
pletely disappears once the plasma is ignited—to be replaced
by the familiar absorption lines of CH4and C2H2. Figures
11a11dshow LIAs for the standard probe transitions for
CH4v=0and C2H2v=0measured with these four alter-
native hydrocarbons under otherwise standard conditions at
z=11 mm plotted as a function of carbon flow rate FC.
The quoted flow rates rely on use of the manufacturer cor-
rection factors where availableand may not be as reliable
as those quoted previously for CH4and C2H2. Nonetheless, a
number of common trends are readily apparent—that are
very reminiscent of those observed when using C2H2as the
source gas Fig. 4d. In all cases, the dominant LIA at low
carbon flow rate is associated with CH4, but this LIA stops
increasing once FC兲⬃15 SCCM. The LIAs of the transi-
tions used to monitor C2H2in its v= 0 and v4= 1 levels both
increase with FC. Notwithstanding the caveat regarding the
accuracies of the quoted carbon flow rates, there is a notable
and consistent difference at high FCbetween the upper
plots involving unsaturated hydrocarbon source gasesand
the lower plots involving alkanes; in the former cases, the
C2H2LIA has actually grown to exceed the CH4LIA. This
may well be a real effect, since the choice of hydrocarbon
affects the overall C/H ratio. In the case of a
40 SCCM CH4/40 SCCM Ar/485 SCCM H2flow, for
example, the overall C/H ratio in the input gas mixture is
40/1130= 0.0354. Use of C3H8or C4H10 at the same carbon
flow rate i.e., 10 SCCM in the case of C4H10, with FH2
raised to 515 SCCM to conserve the same Ftotalgives the
same input C/H ratio. For the unsaturated hydrocarbons,
however, the input C/H ratio will necessarily be higher. For
example, the equivalent calculation for Fpropyne
=13.33 SCCM and FH2= 511.67 SCCMyields an input
C/H ratio of 0.0372. Figures 11eand 11fshow, respec-
tively, the LIAs for the probed CH4v=0and C2H2v=0
probe transitions for these four source gases plotted on a
common C/H ratio scale. The detailed explanation of such
trends is again guided by the 2D modeling results, but is
reserved pending description of the time evolution of the
LIAs of interest following introduction of both CH4and
C2H2into a pre-existing Ar/H2plasma.
E. Time dependence of the LIAs following addition of
CH4C2H2to a pre-existing Ar/H2plasma
The fast spectral acquisition rate achievable with the
chirped QC laser allows study of the way in which the vari-
ous LIAs approach their asymptotic values following intro-
duction of a hydrocarbon flow to a pre-existing Ar/H2
plasma. Figure 12 shows illustrative spectra recorded at dif-
ferent times after the addition of aFCH4=25 SCCM and
bFC2H2=12.5 SCCM to an Ar/H2plasma operating
with FAr=40 SCCM, FH2appropriate to ensure that the
asymptotic Ftotal =565 SCCM, p=150 Torr, and P
=1.5 kW. As noted previously,16 the unblended and the
blendedCH4v=0features appear at earliest time, even
when using C2H2as the source gas. As shown in Figs. 12c
and 12d, the C2H2v=0and C2H2v4=1absorption fea-
tures appear later as do features associated with the v5=1
level of C2H2and take longer to attain their asymptotic LIA
values. Such data provide a further—in this case, time-
dependent—illustration of the effect of varying the C/H ra-
tio. Qualitatively at least, the data display all of the same
trends as noted previously. At low C/H ratio early time in
these experimentsthe main hydrocarbon detected is CH4,
but C2H2LIAs grow in relative intensity at later time higher
C/H ratio.
FIG. 10. Part of the 90
2absorption band of a 5 Torr room temperature sample
of propyne.
033305-11 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
F. Mechanism of CH4^C2H2interconversion and its
spatial dependence
Figure 13 displays production-lossrates of H atoms,
CH4,C
2H2, and C2H4plotted as a function of zat r= 0 re-
turned by the 2D model calculations for standard process
conditions: i.e., 25 SCCM CH4, 40 SCCM Ar , 500
SCCM H2,p= 150 Torr, P= 1.5 kW. The zdependence of
these calculated production-lossrate profiles are relatively
unaffected by reducing FCH4to 5 SCCM or pto 75 Torr,
although both such changes reduce the respective magni-
tudes of the rates. To a reasonable approximation, these pro-
files are radially symmetric in the r,zspace above the sub-
strate. As noted in Ref. 5inspection of such figures
encourages discussion of the reactor volume in terms of three
concentric volumes: A, the plasma region that is located
directly above the substrate, contains the highest Tgas and is
characterized by large H atom production rates;B, an annu-
lar shell characterized by 1400Tgas2200 K, efficient
conversion of CH4and C2H4into C2H2and the consump-
tion of H atoms required for this conversion; and C, the
cooler outer region Tgas 1400 Kcharacterized by net
conversion of C2H2into CH4and C2H4and low H atom
consumption rates. Tables IV and Vsummarize the more
important reaction sequences i.e., those involving elemen-
tary steps with calculated rates1017 cm−3 s−1that drive
these transformations in the centers of region Bz=3.5 cm,
where Tgas 1900 Kand region Cz= 4.8 cm, where Tgas
1100 K. Several of the elementary steps in the CH4
C2H2conversion scheme shown in Table IV consume H
atoms—consistent with the substantial calculated H atom
loss rate in region B共⬃41016 cm−3 s−1. The calculated H
atom production-lossrate in region C, in contrast, is
small—consistent with the reduced mechanism listed in
Table V. H atoms are needed to drive C2H2CH4conver-
sion in region C, but they are not substantiallyconsumed
by this conversion.
Such 2D model outputs provide a rationale for all of the
trends observed experimentally. Under standard process con-
ditions, CH4or C2H2feedstock gas is converted into a mix-
ture containing both of these hydrocarbons Fig. 4. The full
2D modeling predicts the presence of other stable hydrocar-
bons also C2H4,C
2H6, etc.at lower mole fraction, but the
tuning range of the available QC laser precludes the possi-
bility of observing these species. Thermodiffusion drives the
stable hydrocarbon species toward regions of lower Tgas but
chemical processing is occurring in parallel—driving net
CH4C2H2interconversion forward in region Band back-
ward in region C. The gas composition in the outermost parts
of the reactor may be further complicated by wall reactions
and the effects of local stagnation volumes, which are prob-
FIG. 11. LIAs for the standard probe transitions of CH4v=0and C2H2v=0measured at z=11 mm in four different hydrocarbon /Ar/H2gas mixtures at
p=150 Torr and P= 1.5 kW, as a function of carbon flow rate: aethane, bpropyne, cn-propane, and dn-butane. FArand Ftotal were maintained
constant at, respectively, 40 and 565 SCCM throughout, and FH2adjusted to compensate as Fhydrocarbonwas varied. The CH4v=0and C2H2v=0LIA
data for these four molecules are plotted on a common C/H ratio scale in panels eand f, respectively.
033305-12 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
ably not treated well in the model calculations. The gas com-
position in this region may be further complicated by trans-
port of unprocessedsource gas—the LIAs shown in Figs.
4cand 4dclearly indicate a relative excess of the respec-
tive source gas at higher carbon flow rates. The number den-
sity profiles dictated by Tgasand STline strength factors
mean that the present QC laser measurements studies and
any other line-of-sight IR studies of stable hydrocarbon spe-
cies in such reactorsare necessarily heavily biased toward
material in region C—the cold periphery of the reactor re-
call Fig. 6. Even when using recessed windows configura-
tion II, therefore, gas in the hot plasma region Amakes no
direct contribution to the measured LIAs. As Fig. 11 shows,
all of the alternative hydrocarbon source gases investigated
are processed, efficiently, to similar CH4/C2H2mixtures.
The time dependent studies Fig. 12provide a particu-
larly clear illustration of the role of region C.Att=0, this
region contains just Ar, H2, and H atoms. Upon adding CH4
or any other hydrocarbonthrough inlets sited at the top of
the reactor, the shortest diffusion path is directly
downwards—i.e., transport localized in the outer region C.
As Table Vshows, CH4is the most stable hydrocarbon in
this region. Hence the dominance of the CH4features at
early t—irrespective of the choice of input hydrocarbon. The
input carbon must traverse a longer path in order for C2H2to
build up in the viewing column within region C—from the
gas inlets, through C, into Bwhere the requisite CH4
C2H2occurs Table IV兲兴, and then radially outward again
into the cooler part of the viewing column. This longer dif-
fusional path is reflected by the longer observed buildup time
for the C2H2LIA Fig. 12. The efficiency of C2H2CH4
processing in region Cdepends on the local C/H ratio. In the
time dependent studies, this ratio increases with tuntil reach-
ing the asymptotic value appropriate to the chosen process
conditions.
Finally we revisit the results of the steady state experi-
ments in the context of the above discussion. At low carbon
flow rates FC=5 SCCM, the C/H ratio in region Cis
sufficiently low that all hydrocarbon species in this region
are processed to CH4—consistent with the dominance of
CH4features in the absorption spectra measured under these
conditions Figs. 4and 11. Increasing FCleads to a pro-
gressive increase in the C/H ratio in region Cand in the
FIG. 12. Absorption spectra recorded at z= 11 mm at different times tafter addition of aFCH4= 25 SCCM and bFC2H2= 12.5 SCCM to an Ar /H2
plasma operating with FAr=40 SCCM, FH2appropriate to ensure that the asymptotic Ftotal = 565 SCCM, p= 150 Torr and P= 1.5 kW. Plots cand d
show the growth of the LIAs for the standard CH4v=0,C
2H2v=0, and C2H2v4=1probe transitions as a function of time after switching on the
hydrocarbon MFC which defines t=0.
FIG. 13. Calculated production-lossrates of H atoms, CH4,C
2H2, and
C2H4and Tgasplotted as functions of zat r=0 for standard process con-
ditions: i.e., FCH4=25 SCCM, FAr=40 SCCM, FH2=500 SCCM,
p=150 Torr, P= 1.5 kW.
033305-13 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
C2H2/CH4ratio in region B. Diffusive transport ensures that
the steady state C2H2density in region Cthus increases
also—to levels that are easily detectable in the single pass
line-of-sight IR absorption spectrum. As already discussed
Sec. III C, increasing pat constant input mixing ratioor
substituting part of FH2with Ar or Nehas the effect of
reducing nHin region C, thereby reducing the probability of
C2H2CH4conversion in this region and increasing the
relative showing of C2H2as compared with CH4in the
line-of-sight IR absorption spectrum.
IV. CONCLUSION
The QCL measurements reported here serve to highlight
many of the strengths and the limitations of applying line-of-
sight absorption methods to stable gas phase species in a
reactor containing very large temperature and number den-
sitygradients. The principle limitation is that the measured
absorptions are associated with molecules in the cool periph-
ery of the reactor C; the present measurements provide no
direct measure of molecules in the central plasma region A
from whence diamond growth occurs. Notwithstanding, they
are hugely informative with regards to gas processing within
the reactor, and to testing, tensioning and refining the com-
panion 2D model calculations. The present studies show that
any chosen hydrocarbon source gas will be converted into a
mixture dominated by CH4and C2H2, and that the continu-
ing interconversion between these two species depends sen-
sitively on both Tgas and nH, and thus on process conditions
pressure, input power, input gas mixing ratios, etc.and lo-
cation within the reactor. CH4C2H2conversion occurs
most efficiently in regions characterized by 1400Tgas
2200 K, which correlate with an annular shell region B
around the central plasma region A, whereas the reverse
transformation C2H2CH4is favored in the outermost re-
gion Cwhere Tgas 1400 K.
ACKNOWLEDGMENTS
The Bristol group is grateful to EPSRC for the award of
a portfolio grant LASER, to Element Six Ltd for financial
support and the long term loan of the MW reactor, to the
University of Bristol and the Overseas Research Scholarship
ORSscheme for a postgraduate scholarship J.M., and to
colleagues J. J. Henney, K. N. Rosser and Dr. C. M. Western
and Dr. J. A. Smith for their many contributions to the work
described here. The Strathclyde group is grateful to EPSRC
for funding the initial part of their QC laser spectrometer
development program and for a research studentship to
K.G.H., to NERC for funding subsequent QC laser spec-
trometer development, and to the Leverhulme Trust for the
award of an Emeritus Fellowship to G.D.. Y.A.M. is
pleased to acknowledge support from RF Government for
Key Science Schools Grant No. 133.2008.2. The collabora-
tion between Bristol and Moscow is supported by a Royal
Society Joint Project Grant.
TABLE IV. Simplified CH4C2H2conversion mechanism highlighting key elementary reactions with rates
1017 cm−3 s−1 under standard process conditions at z=3.5 cm i.e., region B, where Tgas 1900 K.
Reaction
Reaction rate
/cm−3 s−1
Net conversion; net rate
/cm−3 s−1
1CH
4+HCH3+H27.281019 CH4CH3; 2.51017
2CH
3+H2CH4+H 7.141019
3CH
3+H+MCH4+M1.201018
4CH
3+H1CH2+H25.241018 CH31/3CH2; 1.61017
51CH2+H2CH3+H 5.111018
63CH2+H2CH3+H 2.971017
7CH
3+H3CH2+H23.271017
81CH2+M3CH2+M1.891018 1CH23CH2; 1.3 1017
93CH2+M1CH2+M1.761018
10 3CH2+CH4C2H4+H21.351017 CHx+CHyC2H4; 1.351017
11 C2H4+HC2H3+H21.531018 C2H4C2H3; 2.1 1017
12 C2H3+H2C2H4+H 1.32 1018
13 C2H3+HC2H2+H22.361017 C2H3C2H2; 2.11017
14 C2H3+MC2H2+H+M7.141017
15 C2H2+H+MC2H3+M7.401017
TABLE V. Simplified C2H2CH4conversion mechanism highlighting key
elementary reactions with rates 1017 cm−3 s−1 under standard process con-
ditions at z=4.8 cm i.e., region C, where Tgas 1100 K.
Reaction
Reaction rate
/cm−3 s−1
Net conversion; net rate
/cm−3 s−1
1C
2H2+H+MC2H3+M2.321017 C2H2C2H3; 2.11017
2C
2H3+H2C2H4+H 2.781017 C2H3C2H4; 2.11017
3C
2H4+H+MC2H5+M1.541017 C2H4C2H5; 1.31017
4C
2H5+HCH3+CH31.421017 C2H52CH3;1.41017
5CH
3+H2CH4+H 1.001018 CH3CH4; 2.11017
6CH
4+HCH3+H27.931017
033305-14 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
APPENDIX
The estimation of STline strengths for transitions from
vibrationally excited levels of C2H2involves calculation of
the internal vibration and rotationpartition function in the
rigid rotor, harmonic oscillator limit using analytical formu-
las as outlined by Herzberg.26 Such an approach is justified
in the present context given the small size of the rotational
constant B=1.1766 cm−1relative to the Tgas values of in-
terest. In this limit, the fraction of the total population in any
given level mwith energy Emis given by
nm
N=gmexpEm/kT
Qint
,A1
where gmincorporates the rotational 2J+1and nuclear spin
3 or 1, for ortho- and para-levels, respectivelydegenera-
cies. Qint is the internal rotation, vibrationpartition func-
tion, given by
Qint =kT/
hcB
i1 − exp
ihc/kT兲兴di,A2
where
iis the wavenumber of fundamental vibration iwith
degeneracy di, the product in the denominator runs over the
five fundamental modes of vibration, and
is the symmetry
factor 2 in the case of C2H2.
Figure 14 shows the calculated Tdependence of n/N
Eq. A1for the following levels of C2H2:v=0, J=23
which shows much the strongest absorption in this spectral
region;v5=1, J=20 E=1225 cm−1;v4=1, J=24 E
=1319 cm−1; and v4=1+v5=1, J=20 E=1822 cm−1.As
noted earlier Sec. III B, the two J= 20 levels have different
nuclear spin statistical weights. Also shown right hand axis
is the STcurve for the 40
150
1P23transition.21 The two
vertical axes have been scaled to highlight the expected
reasonable match between the respective Tdependences,
thereby demonstrating that the above approach offers a sen-
sible route to predicting STfactors and thus to determining
the scaling factor proportional to the Einstein B-coefficient
from the best-fit straight line between nand Sover the com-
plete range of T. All of the C2H2transitions of interest in-
volve v4=v5=1 changes in vibrational quantum number.
We therefore assume that all have the same or similarEin-
stein B-coefficients. The right hand axis in Fig. 14 thus pro-
vides estimates of the STfactors for the transitions of in-
terest involving levels with v0. We estimate that the
cumulative error in this procedure is 5%, which is well
within the error associated with LIA measurements involving
these weak lines.
1D. G. Goodwin and J. E. Butler, in Handbook of Industrial Diamonds and
Diamond Films, edited by M. A. Prelas, G. Popovici, and L. K. Bigelow
Dekker, New York, 1998, pp. 527–581 and references therein.
2J. Isberg, J. Hammersberg, E. Johansson, T. Wikstrom, D. J. Twitchen, A.
J. Whitehead, S. E. Coe, and G. A. Scarsbrook, Science 297, 1670 2002.
3C. S. Yan, Y. K. Vohra, H. K. Mao, and R. J. Hemley, Phys. Status Solidi
A201,R252004.
4G. Bogdan, K. de Corte, W. Deferme, K. Haenen, and M. Nesládek, Phys.
Status Solidi A 203,30632006.
5Yu. A. Mankelevich, M. N. R. Ashfold, and J. Ma, J. Appl. Phys. 104,
113304 2008.
6M. N. R. Ashfold, P. W. May, J. R. Petherbridge, K. N. Rosser, J. A.
Smith, Y. A. Mankelevich, and N. V. Suetin, Phys. Chem. Chem. Phys. 3,
3471 2001and references therein.
7P. K. Bachmann, D. Leers, and H. Lydtin, Diamond Relat. Mater. 1,1
1991and references therein.
8T. Mitomo, E. Kondoh, and H. Ohtsuka, J. Appl. Phys. 70,45321991.
9W. L. Hsu, M. C. McMaster, M. E. Coltrin, and D. S. Dandy, Jpn. J. Appl.
Phys., Part 1 33, 2231 1994.
10M. C. McMaster, W. L. Hsu, M. E. Coltrin, D. S. Dandy, and C. Fox,
Diamond Relat. Mater. 4, 1000 1995.
11F. G. Celii, P. E. Pehrsson, H.-t. Wang, and J. E. Butler, Appl. Phys. Lett.
52, 2043 1988.
12F. G. Celii and J. E. Butler, Annu. Rev. Phys. Chem. 42,6431991.
13G. Lombardi, G. D. Stancu, F. Hempel, A. Gicquel, and J. Röpcke, Plasma
Sources Sci. Technol. 13,272004.
14G. Lombardi, K. Hassouni, G. D. Stancu, L. Mechold, J. Röpcke, and A.
Gicquel, Plasma Sources Sci. Technol. 14, 440 2005.
15G. Lombardi, K. Hassouni, G. D. Stancu, L. Mechold, J. Röpcke, and A.
Gicquel, J. Appl. Phys. 98, 053303 2005.
16A. Cheesman, J. A. Smith, M. N. R. Ashfold, N. Langford, S. Wright, and
G. Duxbury, J. Phys. Chem. A 110, 2821 2006.
17R. F. Kazarinov and R. A. Suris, Sov. Phys. Semicond. 5,7071971.
18J. Faist, F. Capasso, D. L. Sivco, C. Sirtori, A. L. Hutchinson, and A. Y.
Cho, Science 264, 553 1994.
19G. Duxbury, N. Langford, M. T. McCulloch, and S. Wright, Chem. Soc.
Rev. 34, 921 2005.
20J. Ma, J. C. Richley, M. N. R. Ashfold, and Yu. A. Mankelevich, J. Appl.
Phys. 104, 103305 2008.
21L. S. Rothman, A. Barbe, D. C. Benner, L. R. Brown, C. Camy-Peyret, M.
R. Carleer, K. Chance, C. Clerbaux, V. Dana, V. M. Devi, J.-M. Flaud, R.
R. Gamache, A. Goldman, D. Jacquemart, K. W. Jucks, W. J. Lafferty,
J.-Y. Mandin, S. T. Massie, V. Nemtchinov, D. A. Newnham, A. Perrin, C.
P. Rinsland, J. Schroeder, K. M. Smith, M. A. H. Smith, K. Tang, R. A.
Toth, J. Vander Auwera, P. Varanasi, and K. Yoshino, J. Quant. Spectrosc.
Radiat. Transf. 82,52003.
22Y. Kabbadj, M. Herman, G. Di Lonardo, L. Fusina, and J. W. C. Johns, J.
Mol. Spectrosc. 150, 535 1991.
23J. Ma, M. N. R. Ashfold, and Yu. A. Mankelevich, J. Appl. Phys. 105,
043302 2009.
24O. J. L. Fox, J. Ma, P. W. May, M. N. R. Ashfold, and Yu. A. Man-
kelevich, Diamond Relat. Mater. 18,7502009.
25M. Herman, T. R. Huet, Y. Kabbadj, and J. Vander Auwera, Mol. Phys. 72,
75 1991.
26G. Herzberg, Molecular Spectra and Molecular Structure II. Infrared and
Raman Spectra of Polyatomic Molecules Van Nostrand Reinhold, New
York, 1945.
FIG. 14. Plot showing the calculated Tdependence of n/NEq. A1for
selected C2H2levels of interest: v=0, J=23 E= 649 cm−1,;v5=1, J
=19 E=1177 cm−1,;v5=1, J=20 E=1225 cm−1,;v4=1, J=24
E=1319 cm−1,; and v4=1+v5=1, J=20 E=1822 cm−1,. Also
shown solid curve and right hand scaleis the STline strength factor for
the 40
150
1P23transition from HITRAN Ref. 21. The two vertical axes
have been scaled to highlight the analogous Tdependences and to establish
the best-fit scaling factor linking nand S. The inset shows the various n/N
values at high Ton an expanded 25 timesvertical scale.
033305-15 Ma et al. J. Appl. Phys. 106, 033305 2009
Downloaded 06 Aug 2009 to 137.222.40.127. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp
... The focus of this article is complementary 2D(r, z) modeling and experimental studies of different input gas mixtures and spatially-dependent and process-dependent species concentrations in the Bristol MW PACVD reactor. Early studies revealed (i) that the methyl radical concentration, [CH 3 ](r, z), exhibited an annular (shell-like) structure with comparable radial and axial concentration gradients and (ii) a complex picture of hydrocarbon interconversions throughout the whole reactor with three main zones [1,19,20]. Both these findings serve to illustrate the limitations of applying 1D modeling when the species concentrations and plasma parameters are clearly of higher dimensionality and flag the need for caution when considering results and predictions derived from such 1D modeling. ...
... The 2D(r, z) modeling reported here focuses on common trends and on how variations in p and P affect the diamond deposition processes. The 2D(r, z) self-consistent model has been developed, refined and verified in many prior combined theoretical/experimental studies of the Bristol MW PACVD reactor [2,3,[10][11][12][13][14][15][16][17][18][19][20][21]24,35,[37][38][39][40][41][42]. Section 2 provides a summary of the model and reports calculated spatial distributions of key plasma parameters and important species, including CH 3 radicals and H atoms close above the substrate surface, as functions of p and P (for more than one substrate diameter (d s ) and temperature (T s )). ...
... Fig. 10) is unsurprising, given the contributions from reactions involving CH 4 and C 2 H 5 (Eq. (4) and Table 1), the concentrations of both of which also depend on many factors [19,20] in addition to [H]. ...
... NO x are also the important species for the emission control from these processes. It has been demonstrated in several applications of TDLAS for chemical vapor deposition (CVD) process monitoring [42,43]. CH 4 and C 2 H 2 have been monitored using a quantum cascade laser at 7.84 nm [42]. ...
... It has been demonstrated in several applications of TDLAS for chemical vapor deposition (CVD) process monitoring [42,43]. CH 4 and C 2 H 2 have been monitored using a quantum cascade laser at 7.84 nm [42]. HCl has also been measured in a CVD process [43]. ...
... conversion were revealed in similar microwave PACVD reactors [15,27] with close plasma parameters: the central hot plasma zone A, enveloped by shells B and C. In the hot central zone A with a characteristic radius of ~3.5 cm and gas temperature of K, the distribution of species in the С1 (CH x , -4) and С2 (C 2 H y , In contrast to the microwave discharge, in the dc discharge under study with a substantially smaller volume (the characteristic radius of the plasma column is cm), zone A was almost absent. Although the rate of the conversion decreased from the maximum in the center ( cm) of zone B with decreasing radial coordinate, this conversion took place up to the reactor axis ( ), where the gas temperature reaches K. ...
... groups was close to equilibrium and depended mainly on the local values of the densities [H] and [H 2 ], temperature T, and thermochemical parameters of the species. In the middle zone B with a temperature of K, methane was converted into C 2 H 2[15,27] via decomposition into СН x radicals in the H-shifting reactions with a significant consumption of H atoms (about four atoms per conversion). The reverse conversion, , occurred in the outer zone C with a gas temperature of atoms, which, however, played a key role in this multistep conversion, acting as activators. ...
Article
Full-text available
Two-dimensional numerical simulations of a dc discharge in a CH4/H2/N2 mixture in the regime of deposition of nanostructured carbon films are carried out with account of the cathode electron beam effects. The distributions of the gas temperature and species number densities are calculated, and the main plasmachemical kinetic processes governing the distribution of methyl radicals above the substrate are analyzed. It is shown that the number density of methyl radicals above the substrate is several orders of magnitude higher than the number densities of other hydrocarbon radicals, which indicates that the former play a dominant role in the growth of nanostructured carbon films. The model is verified by comparing the measured optical emission profiles of the H(n ≡ 3), C2*, CH*, and CN* species and the calculated number densities of excited species, as well as the measured and calculated values of the discharge voltage and heat fluxes onto the electrodes and reactor walls. The key role of ion–electron recombination and dissociative excitation of H2, C2H2, CH4, and HCN molecules in the generation of emitting species (first of all, in the cold regions adjacent to the electrodes) is revealed.
... The interconversion of the dilute hydrocarbon species in atomic and molecular hydrogen is shown schematically in Figure 20. In the hotter portions of the gas (in the plasma), the chemistry is driven toward the lower right of the figure, by the conversion of CH n to C 2 H n , while in the colder portions of the reactor, near the reactor walls and the growing surface, the chemistry is driven to the upper left by atomic hydrogen reactions (C 2 H n to CH n ) (Ma et al. 2009). ...
... The effect of adding methane to hydrogen plasma of a microwave discharge on the emission intensity of the lines of atomic hydrogen H α , H β is shown in figure 3. It can be seen that with the addition of methane, the intensities of the H α , H β lines increase sharply (∼3 times) and remain almost unchanged with a subsequent increase in the percentage of methane in the plasma. Based on the results of [26,36,37], it can be argued that the addition of methane to hydrogen plasma leads to the appearance of a large number of C 2 H 2 radicals, whose concentration may even exceed the current concentration of methane. Considering that the ionization potential of the C 2 H 2 radical (11.4 eV) is lower than the ionization potential of methane (14.25 eV) and molecular hydrogen (15.43 eV), the addition of even a small amount of methane increases significantly the ionization rate of the gas mixture supplied to the chamber, thereby increasing the concentration of electrons in the microwave discharge. ...
Article
Full-text available
The paper describes synthesis of diamonds by the method of gas-jet deposition with microwave activation of precursor gases. This method involves the use of supersonic jet for delivering the components activated in the discharge chamber to the substrate located in the deposition chamber. A series of experiments was carried out with different amounts of methane supplied at a hydrogen flow rate of 8000 sccm. The obtained samples of diamond coatings were studied by scanning electron microscopy and Raman spectroscopy. The temperature of the mixture and the intensities of H, CH, and C 2 lines in the plasma of the discharge chamber were measured by optical emission spectroscopy. The values of pressure and temperature in the discharge chamber were used to estimate the composition of the mixture. Thus, the numerical dependences of the molar concentrations of CH 3 , CH, C 2 and C 2 H 2 on the initial concentration of methane have been obtained. These dependences are in qualitative agreement with the dependences of the intensities of H, CH, and С 2 lines. The numerical-experimental study performed allows us to conclude that the optimal value of methane concentration in the supplied mixture for the gas-jet deposition method in the considered range of parameters is about 1%.
Article
Silicon is a known trace contaminant in diamond grown by chemical vapor deposition (CVD) methods. Deliberately Si-doped diamond is currently attracting great interest because of the attractive optical properties of the negatively charged silicon-vacancy (SiV-) defect. This work reports in-depth studies of microwave-activated H2 plasmas containing trace (10-100 ppm) amounts of SiH4, with and without a few % of CH4, operating at pressures and powers relevant for contemporary diamond CVD, using a combination of experiment (spatially resolved optical emission (OE) imaging) and two-dimensional plasma chemical modeling. Key features identified from analysis and modeling of the OE from electronically excited H, H2, Si, and SiH species in the dilute Si/H plasmas include the following: (i) fast H-shifting reactions ensure that Si atoms are the most abundant silicon-containing species throughout the entire reactor volume, (ii) the low ionization potentials of all SiH x (x ≤ 4) species and efficient ion conversion reactions ensure that even trace SiH4 additions cause a change in the dominant ions in the plasma volume (from H3+ to SiH x +), with consequences for electron-ion recombination rates and ambipolar diffusion coefficients, and (iii) the total silicon content in the reactor volume can be substantially perturbed by silicon deposition and H atom etching reactions at the reactor walls. The effects of adding trace amounts of SiH4 to a pre-existing C/H plasma are shown to be much less dramatic but include the following: (i) a Si substrate or fused silica components within the reactor are a ready (unintended) source of gas-phase Si-containing species, (ii) OE from electronically excited Si atoms should provide a reliable measure of the Si content in the hot plasma region, and (iii) Si atoms and/or SiC2 species are the most abundant gas-phase Si-containing species just above the growing diamond surface and thus the most likely carriers of the silicon incorporated into CVD diamond.
Article
The fine rotational-vibrational spectra of long-chain molecules like 1,3-butadiene are very closely spaced and highly overlapped. Such investigations require high-sensitive techniques to avoid overlapping and adequate resolution to observe their fine structure spectra. In this work, we have utilized a high-resolution external cavity-quantum cascade laser at ∼6.2 μm coupled with high-sensitive cavity ring-down spectroscopy (CRDS) to determine the fine ro-vibrational spectra of 1,3-butadiene in the gas phase. We experimentally probed the hybrid A/B-type band at 1596.412 cm ⁻¹ and also calculated relevant spectroscopic parameters of 1,3-butadiene using Gaussian 16 and subsequently simulated the spectra using PGOPHER simulation. Finally, we selected a unique interference-free spectral region within the 1,3-butadiene spectrum, and performed detailed ro-vibrational study of 3 spectral lines in this region. These results are a great advance over the earlier works and provide a spectroscopic window that may enable selective and high-sensitive sensing of 1,3-butadiene in the future.
Article
Diamond synthesis by chemical vapour deposition (CVD) from carbon-containing gas mixtures has by now long been an industrial reality, but commercial interest and investment into the technology has grown dramatically in the last several years. This Feature Article surveys recent advances in our understanding of the gas-phase chemistry of microwave-activated methane/hydrogen plasmas used for diamond CVD, including that of added boron-, nitrogen-, and oxygen-containing dopant species. We conclude by considering some of the remaining challenges in this important area of contemporary materials science.
Article
Full-text available
We consider the possibility of realizing a terahertz spectrometer using the radiation sources based on quantum cascade lasers and the receiving system based on hot-electron bolometers. Precision measurements of the amplitude and phase noise of the quantum cascade lasers in this frequency range are performed. The systems for stabilizing the frequency of the quantum cascade laser are described. The diagrams of the high-resolution terahertz spectrometers are presented.
Article
Full-text available
A semiconductor injection laser that differs in a fundamental way from diode lasers has been demonstrated. It is built out of quantum semiconductor structures that were grown by molecular beam epitaxy and designed by band structure engineering. Electrons streaming down a potential staircase sequentially emit photons at the steps. The steps consist of coupled quantum wells in which population inversion between discrete conduction band excited states is achieved by control of tunneling. A strong narrowing of the emission spectrum, above threshold, provides direct evidence of laser action at a wavelength of 4.2 micrometers with peak powers in excess of 8 milliwatts in pulsed operation. In quantum cascade lasers, the wavelength, entirely determined by quantum confinement, can be tailored from the mid-infrared to the submillimeter wave region in the same heterostructure material.
Article
Full-text available
We describe laser and mass spectroscopic methods, and related modelling studies, that have been used to unravel details of the gas phase chemistry involved in diamond chemical vapour deposition (CVD) using both H/C (i.e. hydrocarbon/H2) and H/C/O (e.g. CO2/CH4) gas mixtures, and comment on the relative advantages and limitations of the various approaches. In the case of the more extensively studied hydrocarbon/H2 systems we pay particular emphasis to investigations (both experimental, and 2- and 3-dimensional modelling) of transient species like H atoms and CH3 radicals, their spatial distributions within the reactor and the ways in which these distributions vary with process conditions, and the insight provided by such investigations into the chemistry underpinning the diamond CVD process. These analyses serve to highlight the rapid thermochemical cycling amongst the various hydrocarbon species in the reactor, such that the gas phase composition in the vicinity of the growing diamond surface is essentially independent of the particular hydrocarbon source gas used. Such applies even to the case of hot filament activated C2H2/H2 gas mixtures, for which we show that CH3 radical formation (hitherto often presumed to involve heterogeneous hydrogenation steps) can be fully explained in terms of gas phase chemistry. Diamond growth using H/C/O-containing gas mixtures has traditionally been discussed in terms of an empirically derived H–C–O atomic phase composition diagram (P. K. Bachmann, D. Leers, H. Lydtin and D. U. Wiechert, Diamond Relat. Mater., 1991, 1, 1). Detailed studies of microwave activated CO2/CH4 gas mixtures, accompanied by simpler zero-dimensional thermochemical modelling of this and numerous other H/C/O-containing input gas mixtures, provide a consistent rationale for the ‘no growth ’, ‘diamond growth’ and ‘non-diamond growth’ regions within the H–C–O atomic phase composition diagram.
Article
We have developed a novel molecular beam mass spectrometry technique that can quantitatively analyze the gas-phase composition in a CVD reactor. The technique simultaneously monitors a wide variety of radical and stable species, and their concentrations can be determined with sensitivities approaching 1 ppM. Measurements performed in a diamond deposition system have given us keen insights into the important phenomena that affect the growth environment. This paper first discusses the primary gas sampling design issues. In the second part, the details of the experimental results and their implications will be described.
Article
A full Hamiltonian matrix coherent with a phase convention leading to e/f labellings and to positive l-doubling q parameters is constructed to describe the rovibrational energy levels in a linear polyatomic molecule with two bending vibrations. This model is used to investigate quantitatively the effect of the rotational and vibrational l resonances on level mixings and intensity features in C2H2.
Article
Microwave discharges of H2 admixed with CH4 in a moderate-pressure quartz bell jar reactor used for diamond deposition are studied numerically. Special attention was devoted to high-power densities which provide the most effective way for producing high-quality diamond films. First, a one-dimensional radial model describing the coupled phenomena of chemistry, energy transfer, as well as species and energy transport along the reactor’s radial coordinate was developed. Species densities predicted with the model were compared with measurements with infrared tunable diode laser spectroscopy, resulting in validation of the model. Second, a one-dimensional axial model was used to describe the plasma flow along the reactor axis in a region between the reactor end wall and the substrate surface. This model was particularly useful for studying the plasma behavior in the vicinity of the substrate surface, where thermal and composition gradients are large. Both the radial and axial transport models are based on the same discharge model in which the plasma is described as a thermochemically nonequilibrium flow with different energy distributions for heavy species and electrons. The chemistry was described with a model containing 28 species and 131 reactions. The electron temperature, the gas temperature, and the species concentration were determined by solving a coupled set of equations. A wide range of experimental conditions used for diamond deposition was simulated, from low microwave power density (9 W cm−3, i.e., 600 W, 2500 Pa, and Tg ∼ 2200 K) to high-power density (30 W cm−3, i.e., 2 kW, 12 000 Pa, and Tg ∼ 3200 K). The main chemical paths were identified, and the major species, transport effects, and reaction pathways that govern diamond deposition plasmas are discussed.
Article
Absorption spectra of C2H2 have been recorded between 50 and 1450 cm-1, with a resolution always better than 0.005 cm-1, using two different Fourier transform spectrometers. Analysis of the data provided two sets of results. First, the bending levels with Σt Vt(t = 4, 5) <= 2 were characterized by a coherent set of 34 parameters derived from the simultaneous analysis of 15 bands, performed using a matrix Hamiltonian. The following main parameters were obtained (in cm-1): ω40 = 608.985196(14), ω50 = 729.157564(10); B0 = 1.17664632(18), α4 = -1.353535(86) × 10-3, α5 = -2.232075(40) × 10-3 q40 = 5.24858(12) × 10-3, and q50 = 4.66044(12) × 10-3, with the errors (1σ) on the last quoted digit. Second, a more complete set of bending levels with Σt Vt <= 4, some of which have never previously been reported, and also including V2 = 1 have been fitted to 80 parameters. This simultaneous fit involved 43 bands and used the same full Hamiltonian matrix. Some perturbations which affect the higher excited levels are discussed.
Article
Infra-red tuneable diode laser spectroscopy (IR TDLAS) has been used to detect and quantify the methyl radical and three stable carbon-containing species (CH4, C2H2 and C2H6) in a moderate pressure microwave (f = 2.45 GHz) bell-jar reactor used for diamond films deposition. A wide range of experimental conditions was investigated, with typical pressure/power required to perform diamond deposition, i.e. pressure from 2500 to 12 000 Pa and power from 600 W to 2 kW, which means gas temperatures ranging from 2200 to 3200 K, when the power density increases from 9 to 30 W cm−3. Since TDLAS is a line of sight averaged technique, the analysis of the experimental data required the use of a one-dimensional non-equilibrium transport model that provides species density and gas temperature variations along the optical beam. This model describes the plasma in terms of 28 species/131 reactions reactive flow. The thermal non-equilibrium is described by distinguishing a first energy mode for the electron and a second one for the heavy species. Parametric studies as a function of power density and methane percentage in the gas mixture are presented. The good agreement obtained between measurement and one-dimensional radial calculations allows a validation of the thermo-chemical model, which can be used as a tool to enlighten the chemistry in the spatially non-uniform H2/CH4 microwave discharge used for diamond deposition. This is especially of interest for high power density discharge conditions that remain poorly understood.
Article
Tuneable infrared diode laser absorption spectroscopy at 16.5 µm and broadband ultraviolet absorption spectroscopy at 216 nm have both been used to measure the ground state concentrations of the methyl radical in two different types of non-equilibrium microwave plasmas (f = 2.45 GHz): (i) H2–Ar plasmas of a planar reactor with small admixtures of methane or methanol, at a pressure of 1.5 mbar, and (ii) H2–CH4 plasmas of a bell jar reactor, at pressures of 25 and 32 mbar under flowing conditions. For the first time, two different optical techniques have been directly compared to verify the available data about absorption cross sections and line strengths of the methyl radical. It was found that application of the CH3 absorption cross section of the transition at 216 nm, reported by Davidson et al (1995 J. Quant. Spectrosc. Radiat. Transfer 53 581) and of the line strength of the Q(8,8) line of the ν2 fundamental band near 16.44 µm, given by Wormhoudt et al (1989 Chem. Phys. Lett. 156 47), leads to satisfactory agreement.