The aim of this work is to build a numerical model of hydrogen plasma inside a microwave plasma chemical vapor deposition system. This model will help in understanding and optimizing the conditions for the growth of carbon nanostructures. A 2D axisymmetric model of the system is implemented using the finite element high frequency Maxwell solver and the heat transfer solver in COMSOL Multiphysics. The system is modeled to study variation in parameters with reactor geometry, microwave power, and gas pressure. The results are compared with experimental measurements from the Q-branch of the H2 Fulcher band of hydrogen using an optical emission spectroscopy technique. The parameter γ in Füner's model is calibrated to match experimental observations at a power of 500 W and 30 Torr. Good agreement is found between the modeling and experimental results for a wide range of powers and pressures. The gas temperature exhibits a weak dependence on power and a strong dependence on gas pressure. The inclusion of a vertical dielectric pillar that concentrates the plasma increases the maximum electron temperature by 70%, the maximum gas temperature by 50%, and the maximum electron number density by 70% when compared to conditions without the pillar at 500 W and 30 Torr. Experimental observations also indicate intensified plasma with the inclusion of a pillar.

Carbon nanostructures such as carbon nanotubes (CNTs), graphitic sheets and nano-petals, and graphene find a wide variety of applications in the fields of engineering and bio-sciences owing to their exceptional mechanical,1,2 thermal,3 and electrical4 properties. Due to the rapid increase in their utilization over the last decade, there is a need for synthesizing high quality carbon nanostructures at reduced cost. Chemical vapor deposition (CVD) is a method for producing such nanostructures on a large scale by depositing gaseous reactants onto a substrate that is placed in a reactor at ambient temperature. The precursor gas molecules dissociate in the reactor and recombine to create a material film on the substrate. The substrate temperature is a key parameter for the growth of the material. The products of CVD are of high quality, hardness, purity, and density5 when compared to materials produced by other processes.

Microwave plasma chemical vapor deposition (MPCVD) is one of the commonly used CVD processes. The first step in this process is the introduction of pure hydrogen into the reactor. The electric field generated by microwaves ionizes the hydrogen, igniting plasma in the reactor. The next step is the introduction of methane after the reactor reaches a steady state. Methane dissociates in the presence of plasma resulting in the deposition of carbon on the substrate for the growth of various carbon nanostructures. A schematic of a methane based plasma assisted CVD process is as shown in Fig. 1. The MPCVD technique is energy efficient, allows for low temperature synthesis6 of carbon nanostructures, and produces stable plasma that can be sustained over a long period of time. MPCVD systems used for diamond deposition have been modeled previously to study pure hydrogen discharges7–14 as well as hydrocarbon discharges.7,15,16 Fluid models have also been used to investigate radio frequency discharges used for plasma assisted CVD processes.17,18

FIG. 1.

Schematic of plasma assisted CVD.

FIG. 1.

Schematic of plasma assisted CVD.

Close modal

The growth rate and quality of the carbon nanostructures depend on the number density of atomic hydrogen and the concentration of certain precursors on the surface of the substrate, which in turn depend on the discharge and operating conditions.19 H2 is the major species in the H2/CH4 plasma for the MPCVD process. Although the introduction of small amounts of hydrocarbons in the feed gas may affect the plasma interaction with the substrate, the main features of the plasma discharge configuration can be obtained by studying pure hydrogen plasma inside the reactor.7 In the present work, we model such pure hydrogen plasma and its interaction with the gas, because information on the electric field, electron number density, electron temperature, and gas temperature are indispensable for understanding and optimizing the operation of MPCVD systems. The electric field is modeled by a high frequency Maxwell solver. This electric field is used to determine the electron number density by Füner's model10,11,20 and electron temperature based on the model by Yamada et al.11 The gas temperature is calculated by a heat transfer solver. The various solvers and models are coupled using COMSOL Multiphysics. A similar model has previously been demonstrated by Huang.21 A Boltzmann equation solver for electrons in weakly ionized gases under given local electric field, called BOLSIG+,22 is used in conjunction with these solvers. The aim of this work is to determine hydrogen plasma characteristics using simple modeling techniques, and the aforementioned solvers and models are chosen because they can be easily coupled without detailed chemistry modeling. The modeled gas temperatures are compared to experimentally measured temperatures of hydrogen using optical emission spectroscopy (OES) techniques.

The MPCVD reactor, a Seki Technotron Corp. AX5200S model, has been used extensively to grow carbon nanostructures such as CNTs,23 graphene,24 and contiguous graphitic fins25 over a variety of substrates. The system is shown in Fig. 2. Microwaves are generated by a 1.5 kW (2.45 GHz) ASTeX AX2100 microwave generator and a 3.5 kW radio frequency power supply. The microwave field passes through a rectangular waveguide, a TE-TM mode converter and into the MPCVD reactor which is isolated from the rest of the system by a quartz plate. The reactor or plasma chamber includes a graphitic susceptor that can move along a vertical axis and a molybdenum puck which is used to concentrate the plasma near the substrate. In the case of graphitic nanopetal growth, a dielectric pillar is placed on the puck to concentrate the plasma.

FIG. 2.

The MPCVD experimental system with spectrometer interface.

FIG. 2.

The MPCVD experimental system with spectrometer interface.

Close modal

The rotational temperatures of the Q-branch of the H2d3Πua3Σu+ (0,0) Fulcher (α) band are measured using OES. The experimental setup is shown in Fig. 3, and an example of the measured H2 spectrum is shown in Fig. 4. The emission from the plasma is collected and transmitted using an optic fiber cable to a spectrometer (Acton SP-2756 Princeton Instruments). A CCD camera (PIXIS 256E Princeton Instruments) is used to detect the light. The spectrometer is calibrated using dual mercury and neon-argon lamps in the wavelength range from 300 to 1000 nm.

FIG. 3.

Top view of optical emission spectroscopy (OES) setup in the MPCVD system.

FIG. 3.

Top view of optical emission spectroscopy (OES) setup in the MPCVD system.

Close modal
FIG. 4.

H2 emission spectrum at 500 W, 10 Torr. The inset shows the Boltzmann plot.

FIG. 4.

H2 emission spectrum at 500 W, 10 Torr. The inset shows the Boltzmann plot.

Close modal

The intensity (I) of the emission from an upper rotational level (EJ) to a lower rotational level (EJ) is given by

(1)

where Nu is the population of the excited state, Aul is the transition probability (or Einstein coefficient), ν̃ is the wavenumber, and h and c are the Planck's constant and speed of light, respectively. For a rotational transition with a particular electronic state and vibrational band that follows Boltzmann distribution, the emission intensity can be rewritten as

(2)

where SJJ is the Hönl-London factor which is taken from Qing et al.,26 λ is the wavelength of the transition obtained from Crosswhite and Dieke,27 kB is the Boltzmann constant, EJ is the rotational energy of the upper excited rotational state, and Trot is the rotational temperature of the upper state (d3Πu). The rotational energy (EJ) is calculated assuming rigid rotor model approximation

(3)

where J is the rotational quantum number for the upper energy level and Bv is the rotational constant obtained from Qing et al.26 Table I shows the Hönl-London factors, rotational energies, and wavelengths27 for Q1, Q2, Q4, Q6, and Q8. These wavelengths are selected in the Boltzmann plot since they have no overlap with other transitions.28 

TABLE I.

Transition wavelengths, upper state rotational energies, and Hönl-London factors of the selected H2 rotational lines for OES.

Transition wavelength,27 λ (nm)Upper state rotational energy, EJ (cm−1)Hönl-London factor,26SJJ
601.83 (Q159.18 4.5 
602.38 (Q2177.55 2.5 
604.27 (Q3591.83 4.5 
607.20 (Q41242.84 6.5 
611.11 (Q52130.59 8.5 
Transition wavelength,27 λ (nm)Upper state rotational energy, EJ (cm−1)Hönl-London factor,26SJJ
601.83 (Q159.18 4.5 
602.38 (Q2177.55 2.5 
604.27 (Q3591.83 4.5 
607.20 (Q41242.84 6.5 
611.11 (Q52130.59 8.5 

An example of the measured H2 spectrum with indication to the selected rotational lines is shown in Fig. 4. The Boltzmann plot is presented as an inset for this spectrum where ln(Iλ4/SJJ) is plotted as a function of the corresponding rotational energy (EJ). The rotational temperature can be deduced from the slope of the plot (1/kBTrot). In the current study, the rotational temperature of the upper state of H2 (d3Πu) is assumed to be the same as the H2 translational state. Comparison with previous temperature measurements using CARS in the same system supports the above assumption within experimental uncertainty of 10%.29 

A Princeton PI-MAX 4 iCCD camera is used to obtain optical emission images of the plasma for qualitative comparison with the modeling results.

The first step in modeling the plasma in the MPCVD reactor is to determine the electric field because the electron number density and electron temperature depend directly on it. The electromagnetic (EM) simulations treat Maxwell's equations in the form

(4)

where k02=ω2ϵ0μ0, ω is the angular frequency of the microwave field, E is the electric field vector which consists of the amplitude as well as the time component, ϵ0 and ϵr are the vacuum and relative permittivities, μ0 and μr are the vacuum and relative permeabilities, and σ is the conductivity. For the first step, vacuum conditions are assumed and a solution is obtained for the electromagnetic field. However, the presence of plasma in the domain changes the material properties of the system by increasing the conductivity of the gas. The plasma properties are evaluated as11,20

(5)
(6)

where ωp is the plasma frequency given by

(7)

Here, e represents the charge of an electron, m is the mass of an electron, ne is the number density of electrons, and νm is the collision frequency between electrons and other species evaluated as

(8)

where N is the number density of the gas, v is the mean electron velocity, and σs is the collisional cross section. For pure hydrogen plasmas, this collision frequency can be approximated as10,11,20

(9)

where pg is the gas pressure in Pa and Tg is the gas temperature in K. These plasma properties are updated in the following iterations by the EM solver.

The electron number density is evaluated based on Füner's model9,10,14 as

(10)

where Em is the minimum electric field strength needed to sustain a discharge, which is taken to be half of the breakdown voltage.30 From MacDonald et al.,31,32 we find that its value is

(11)

Because the plasma is generated by microwaves, the radiation must be able to penetrate into the plasma in order to ionize molecules. This happens only when ωp2ω2+νm2. However, the skin depth for the plasma is on the order of the size of the plasma region. Therefore, the maximum electron number density constraint is not imposed, but is used to determine the minimum electron number density. The value of ne,min that we use in the model must be lower than the value of ne,max. The maximum electron number density in the plasma is thus given by

(12)

where νm is given by Eq. (9). At the two pressures, we have

(13)

Based on this constraint as well as experimental electron number density measurements on similar systems,12–14,33 we use ne,min=1×1017m3. The value of γ is calibrated to match the gas temperature predicted by the model with experimental measurements at 500 W and 30 Torr. The variation of predicted gas temperatures with γ is shown in Table II. Since the value at γ=3×1013V1m2 matches best with the experimental measurement of 1441 K, we choose this value of γ for all computations.

TABLE II.

Variation of gas temperatures predicted by the model with the parameter γ in Füner's model.

γ (V−1 m−2)Gas Temperature (K)
3 × 1012 1326.53 
8 × 1012 1363.07 
3 × 1013 1477.60 
8 × 1013 1538.55 
γ (V−1 m−2)Gas Temperature (K)
3 × 1012 1326.53 
8 × 1012 1363.07 
3 × 1013 1477.60 
8 × 1013 1538.55 

We use the model given by Yamada et al.11 to determine the electron temperature. Simplifying the conservation of energy for electrons generated by microwaves, the rate of energy transferred from the microwaves to the electrons per unit volume, QMW, is shown to equal the rate of energy loss of electrons during collisions with heavy species, Qc. The collisional power loss term is given by

(14)

where Es, ns, and ks are the threshold energy, species number density, and reaction rate coefficient for reaction s. The term Qc/ne is approximated by a function given by

(15)

where EL, pL, and αL are obtained by fitting a curve over a range of electron temperatures to match the two sides of the following equation:

(16)

For the present study, the reactions considered are as shown in Table III.

TABLE III.

Reactions of hydrogen considered in the plasma simulations.

ReactionEquationEs (eV)
Excitation e+H2e+H2* 8.9 
Ionization e+H2e+H2+ 15.4 
Dissociation e+H2e+2H 10 
ReactionEquationEs (eV)
Excitation e+H2e+H2* 8.9 
Ionization e+H2e+H2+ 15.4 
Dissociation e+H2e+2H 10 

The rate coefficients for excitation and ionization are obtained from the Boltzmann solver22 with the cross section data taken from the LXCat database.34–36 The rate coefficient for dissociation is obtained from the Arrhenius equation given by37 

(17)

where Ad is the pre-exponential factor equal to 1×1014m3s1. Hydrogen undergoes a number of excitation reactions. The variation of Qc/ne with electron temperature for the excitation reactions with the highest values of Qc/ne in the range of electron temperatures that are of importance to the current simulations is as shown in Fig. 5. We choose the excitation reaction with Es of 8.9 eV because at the maximum of electron temperatures seen in the reactor, this reaction has the highest contribution to the collision term. The reaction with Es of 0.52 eV has a greater contribution at low electron temperatures. The error in the estimation of Qc/ne due to the omission of this reaction at Te = 2 eV is 60% at 10 Torr and 25% at 30 Torr.

FIG. 5.

Variation of collisional power loss of electrons (Qc/ne = Esnsks) per unit volume with electron temperature for different excitation reactions at (a) 10 Torr and (b) 30 Torr.

FIG. 5.

Variation of collisional power loss of electrons (Qc/ne = Esnsks) per unit volume with electron temperature for different excitation reactions at (a) 10 Torr and (b) 30 Torr.

Close modal

The curve fitting parameters used to determine the electron temperature are given in Table IV. The approximation, (Qc/ne)app, agrees well with the actual collision term, Qc/ne, as indicated by Fig. 6.

TABLE IV.

Parameters used in the electron temperature models.

Pg (Torr)pL (eV/s)EL[eVαL]αL
10 1.875×1014×(1000[K]/Tg[K]) 18.5327 0.36757 
30 5.625×1014×(1000[K]/Tg[K]) 18.5327 0.36757 
Pg (Torr)pL (eV/s)EL[eVαL]αL
10 1.875×1014×(1000[K]/Tg[K]) 18.5327 0.36757 
30 5.625×1014×(1000[K]/Tg[K]) 18.5327 0.36757 
FIG. 6.

Comparison of the collision power loss term with its approximation at (a) 10 Torr and (b) 30 Torr.

FIG. 6.

Comparison of the collision power loss term with its approximation at (a) 10 Torr and (b) 30 Torr.

Close modal

By approximating the collision term and setting it equal to the microwave power density absorbed, the expression obtained for electron temperature is given by11 

(18)

where

A heat transfer solver is used to determine the gas temperature, Tg, using the 2D heat conduction equation with a volumetric heat source term, Qg given by

(19)

where k is the thermal conductivity of the gas. The term Qg represents the power density imparted to the neutral gas molecules when they collide with the free electrons in the plasma chamber. Here, we consider only hydrogen molecules. The chamber is a cylindrical shell. For a gas temperature of 2500 K and ambient temperature of 300 K, the Rayleigh number for a vertical cylinder of height 130 mm is found to be 250, which is much lower than the critical value of 3800 for a similar configuration.38 Therefore, thermal convection is not taken into account in our model. The heat source term39 is given by

(20)

where δ is a fractional energy loss term which is a measure of the fraction of energy transferred from electrons to neutral gas during collisions and kB is the Boltzmann constant. Using the approximate electric field obtained from the EM solver and the gas temperature, the total power loss is calculated by the Boltzmann solver22 for various reactions of hydrogen. The term δ is then found using the following equation:

(21)

where dϵe/dt is the total power loss which is the sum of elastic and inelastic power losses and ϵe is the mean energy of electrons. The variation of δ with the reduced electric field for rotational, vibrational, and electronic excitation reactions of H2, ionization of H2, and elastic collisions between electrons and H2 molecules at a temperature of 1000 K is shown in Fig. 7. The δ value with all the processes included is used to calculate the gas temperature.

FIG. 7.

Variation of fractional energy loss with reduced electric field at 1000 K.

FIG. 7.

Variation of fractional energy loss with reduced electric field at 1000 K.

Close modal

The various solvers are coupled as shown in Fig. 8. The Maxwell solver determines the electric field, which is used by the user defined functions (UDFs) along with the properties from the Boltzmann solver to determine the electron number density, plasma material properties, and electron temperature. They are used to calculate the heat source term for the heat transfer solver, which determines the gas temperature. Iterations are carried out till convergence of residuals is obtained.

FIG. 8.

Schematic of the numerical model.

FIG. 8.

Schematic of the numerical model.

Close modal

The plasma region can be modeled as a 2D axisymmetric region owing to the fact that it is cylindrical. A part of the transmission line, the converter, the quartz plate, and the plasma chamber are modeled axisymmetrically. The susceptor is modeled at the bottom of the plasma chamber, with the puck placed concentrically above its top surface as shown in Fig. 9. The grey region is the part modeled for the EM simulations and the green is modeled for both EM and heat transfer simulations. In some simulations, the susceptor and puck assembly is raised or lowered by a few mm in the reactor to study the effect of the susceptor position on the plasma characteristics. For other simulations, an axisymmetric pillar is modeled above the puck in order to study the effect of change in the geometry of the susceptor stage.

FIG. 9.

2D model of the MPCVD system with boundary conditions.

FIG. 9.

2D model of the MPCVD system with boundary conditions.

Close modal

The boundary conditions applied are shown in Fig. 9. For the pure EM simulations without the plasma effects included, vacuum is assigned to the plasma reactor. When the EM solver is coupled to the plasma and heat transfer models, the reactor is composed of an inlet, outlet, walls, and susceptor-puck assembly. For the heat transfer solver, the temperature of the susceptor-puck assembly is maintained at experimentally observed values. A thermocouple placed below the susceptor surface is used to obtain these measurements while running the system with hydrogen plasma under various conditions. The initial temperature of gas in the plasma chamber is taken to be 293 K. The pillar is modeled as a thermal insulator.

A physics controlled mesh is used in COMSOL to discretize the domain with greater refinement in corner regions. A grid convergence study is conducted with different grid sizes and input conditions to ensure convergence of gas temperature within the plasma chamber. The results of the study for an input power of 500 W and gas pressure of 30 Torr are shown in Fig. 10. When the number of elements is increased beyond 9304, it does not change the temperature by a significant amount and so this grid has been adopted for the present calculations.

FIG. 10.

Grid convergence without inclusion of pillar.

FIG. 10.

Grid convergence without inclusion of pillar.

Close modal

When microwave input power is increased at constant pressure, the electric field inside the plasma chamber increases because of an increase in the energy density of the system. An increase in power always leads to the expansion of the high field region as seen in Fig. 11(a). At 10 Torr, when the power is increased, the maximum electric field reached in the plasma chamber reduces, but the average field increases due to this expansion. At 500 W and 10 Torr, a residual high field region appears below the quartz plate which may lead to the formation of a two-ball plasma if the power is increased further. Such a configuration is not desirable because it not only takes away power from the region of interest but may also cause structural damage to the system. An increase in the pressure of the gas in the chamber leads to an increase in the collision rate, thereby increasing the energy loss inside the system. This reduces the number density of electrons and thus increases plasma permittivity which in turn is fed into the EM solver and results in a decrease in the electric field. Higher pressure also causes the high electric field to be concentrated in a smaller region as seen in Fig. 11(b). Higher values of maximum field are reached at higher pressure, but the average field in the chamber is lower because the high field region shrinks.

FIG. 11.

Electric field variation with (a) power variation and (b) pressure variation.

FIG. 11.

Electric field variation with (a) power variation and (b) pressure variation.

Close modal

The electron number density depends directly on the electric field as given by Eq. (10). As the power is increased at constant pressure, the average electron number density increases, and as the pressure is increased, it decreases. The effects of input power and gas pressure on electron number density inside the plasma chamber are shown in Fig. 12. The increase in power goes into increasing the volume of the plasma region. At 10 Torr, the plasma region is slightly detached from the puck. Previous work reported by Hassouni et al.7 on similar systems demonstrated the change in electron number density while keeping the plasma volume a constant. This requires a simultaneous change in the power and pressure. However, in the present work, the power is kept constant as pressure is varied and increasing the pressure is seen to concentrate the plasma in a smaller region, thereby reducing the average electron number density in the chamber. The electron number density values stay fairly uniform both axially as well as radially within the plasma region.

FIG. 12.

Electron number density variation with (a) power variation and (b) pressure variation.

FIG. 12.

Electron number density variation with (a) power variation and (b) pressure variation.

Close modal

The electron temperature depends directly on the electric field and inversely on the gas pressure. Therefore, it is seen to increase with an increase in the power and decrease with an increase in pressure which is seen in Fig. 13. The electron temperature stays fairly constant within the plasma region axially as well as radially.

FIG. 13.

Electron temperature variation with (a) power variation and (b) pressure variation.

FIG. 13.

Electron temperature variation with (a) power variation and (b) pressure variation.

Close modal

The gas temperature depends on the energy transferred to gas molecules by electrons during collisions. The electron temperature increases as the input power is increased at constant pressure. This makes more energy available to be transferred thereby increasing the gas temperature. An increase in the gas pressure reduces the electron temperature, but increases the number of collisions. Hence, more energy is transferred and so gas temperature increases. The average gas temperature, however, reduces with increasing pressure on account of the plasma region being smaller. The effect of gas pressure and input power on gas temperature can be seen in Fig. 14.

FIG. 14.

Gas temperature variation with (a) power variation and (b) pressure variation.

FIG. 14.

Gas temperature variation with (a) power variation and (b) pressure variation.

Close modal

The experimentally obtained temperatures have been compared to the modeling results for the case without the pillar. Fig. 15 shows the comparison of these temperatures at a point 5 mm vertically above the center of the puck. This point is indicated by a star in Fig. 9. The comparison indicates that the computational values agree well with the experimental observations, especially at higher input powers and gas pressures. At 10 Torr, the numerical results are higher than the experimental measurements by approximately 150 K. For both sets of results, the gas temperature has a weak dependence on power and a very strong dependence on pressure. On an average, the gas temperature at 5 mm above the center of the puck increases from about 1100 K at 10 Torr to about 1500 K at 30 Torr.

FIG. 15.

Comparison of computational (empty symbols) and experimental (filled in symbols) gas temperatures 5 mm above the center of the puck surface at 10 Torr (blue triangles) and 30 Torr (red squares).

FIG. 15.

Comparison of computational (empty symbols) and experimental (filled in symbols) gas temperatures 5 mm above the center of the puck surface at 10 Torr (blue triangles) and 30 Torr (red squares).

Close modal

The contours of electric field at 500 W and 30 Torr are shown in Fig. 16(a). When the pillar is included, the electric field reaches very high values close to the top of the pillar. These values drop rapidly through the high field region both axially as well as radially. Consequently, even though maximum field values reached are very high with the pillar, the average field values in the chamber are very similar to when the pillar is not included. The contours of electron number density at 500 W and 30 Torr are shown in Fig. 16(b). The inclusion of the pillar concentrates the plasma region close to the top of the pillar. Very high values of ne are reached close to the top of the pillar and the values fall rapidly through the plasma region both axially as well as radially. The contours of electron temperature at 500 W and 30 Torr are shown in Fig. 16(c). When the pillar is included, very high values of electron temperature are reached close to the top of the pillar, and the value decreases throughout the plasma in the axial and radial directions.

FIG. 16.

Effect of inclusion of pillar on (a) electric field, (b) electron number density, (c) electron temperature, and (d) gas temperature at power = 500 W and pressure = 30 Torr.

FIG. 16.

Effect of inclusion of pillar on (a) electric field, (b) electron number density, (c) electron temperature, and (d) gas temperature at power = 500 W and pressure = 30 Torr.

Close modal

Fig. 16(d) shows the gas temperature in the plasma chamber at 500 W and 30 Torr. High electron number density close to the top of the pillar increases the number of collisions and high electron temperature in this region causes more energy to be transferred during each collision. Consequently, the gas temperature reaches high values close to the top of the pillar when it is included. This value decreases throughout the plasma region in the axial and radial directions as seen in Figs. 17 and 18. When the pillar is not included, the maximum value of gas temperature is reached at about 20 mm above the puck surface and reduces beyond the maximum point in the axial direction as shown in Fig. 17. Radially, the gas temperature stays fairly constant through the plasma as shown in Fig. 18.

FIG. 17.

Comparison of gas temperature profile along the plasma chamber axis with and without pillar at (a) 10 Torr and (b) 30 Torr.

FIG. 17.

Comparison of gas temperature profile along the plasma chamber axis with and without pillar at (a) 10 Torr and (b) 30 Torr.

Close modal
FIG. 18.

Comparison of gas temperature profile along a radial line 3 mm above the top of the pillar/puck surface with and without pillar at (a) 10 Torr and (b) 30 Torr.

FIG. 18.

Comparison of gas temperature profile along a radial line 3 mm above the top of the pillar/puck surface with and without pillar at (a) 10 Torr and (b) 30 Torr.

Close modal

Experimental images of plasma's optical emission in the range 200 to 800 nm (Princeton, PI-MAX 4 iCCD with 500 ms gate setting) have been captured at 500 W and 30 Torr with and without a pillar of 9.6 mm height. This pillar height is by necessity slightly less than that discussed above to allow full visual access through the available viewport. The resulting images shown in Fig. 19 are qualitatively similar to those of the model (cf. the predicted cross sections of various plasma-affected fields in Fig. 16). The plasma intensity is clearly higher and more concentrated around the pillar, which also elevates the plasma region vertically (axially) while contracting it horizontally (radially). Many species in the plasma, such as H2 and H, exist in excited electronic states due to electron impact excitation40,41 and consequently, the measured emission intensities corroborate the shapes of electron number densities predicted by the model with and without the pillar.

FIG. 19.

Optical emission images of the plasma at 500 W and 30 Torr in the range 200 to 800 nm (Princeton, PI-MAX 4 iCCD with 500 ms gate setting).

FIG. 19.

Optical emission images of the plasma at 500 W and 30 Torr in the range 200 to 800 nm (Princeton, PI-MAX 4 iCCD with 500 ms gate setting).

Close modal

In our model for electron temperature, we assume that all the power absorbed from the microwaves by the electrons is transferred to the neutral gas during collisions. The power absorbed from the microwaves may be calculated as QMW=e22mνm(ω2+νm2)ne|E|2. The efficiency of power absorption is given as (QMW/input microwave power) × 100. Fig. 20 indicates that both these quantities are higher at low pressure and the inclusion of the pillar improves them at high pressure. This is because the pillar concentrates plasma of high electron number density and electron temperature close to the top, and the power is utilized in reaching higher values of electron number density and temperatures than in expanding the plasma.

FIG. 20.

Comparison of (a) microwave power absorbed and (b) efficiency of microwave power absorption with pillar (filled in symbols) and without pillar (empty symbols) at 10 Torr (blue triangles) and 30 Torr (red squares).

FIG. 20.

Comparison of (a) microwave power absorbed and (b) efficiency of microwave power absorption with pillar (filled in symbols) and without pillar (empty symbols) at 10 Torr (blue triangles) and 30 Torr (red squares).

Close modal

The susceptor stage is moved without the inclusion of the pillar at 500 W and 30 Torr. The condition used thus far is used as the reference position (Ysus = 0 mm). When the stage is lowered by 20 mm (Ysus = −20 mm), the region of high electric field moves to the center of the plasma chamber by detaching from the puck. This causes the plasma to move away from the region of interest as can be seen in Fig. 21. Even though very high values of gas temperatures are reached in this case as shown in Fig. 22, they are not useful for the growth of nanostructures.

FIG. 21.

Electron number density variation with susceptor position when (a) susceptor is lowered and maintained at the same level and (b) susceptor is raised.

FIG. 21.

Electron number density variation with susceptor position when (a) susceptor is lowered and maintained at the same level and (b) susceptor is raised.

Close modal
FIG. 22.

Gas temperature variation with susceptor position when (a) susceptor is lowered and maintained at the same level and (b) susceptor is raised.

FIG. 22.

Gas temperature variation with susceptor position when (a) susceptor is lowered and maintained at the same level and (b) susceptor is raised.

Close modal

When the susceptor is raised by 25 mm (Ysus = 25 mm), the plasma reaches higher values of electric field, electron number density, and electron temperature close to the puck. The gas temperature reduces by 3%. The plasma region stays attached to the puck. These conditions are favorable for deposition. However, a residual field appears to be forming near the top of the chamber which may lead to the formation of two-ball plasma if the input power is increased. When the susceptor is raised even more to Ysus = 50 mm, we see the formation of secondary high field and plasma regions on the sides of the susceptor walls. This takes away power from the plasma and inhibits growth near the substrate. The secondary plasma may also damage the sides of the susceptor due to which this configuration is not desirable.

The numerical results with changes in susceptor position and inclusion of the pillar are not compared to experimental measurements quantitatively because the beam optics involved make it challenging to conduct such experiments.

A numerical model has been built to analyze hydrogen plasmas in an MPCVD system using the high frequency Maxwell solver and heat transfer solver in COMSOL Multiphysics. These solvers have been coupled with user defined functions as well as the Boltzmann equation solver, BOLSIG+, to determine plasma properties. Variation of discharge characteristics with change in input properties, susceptor stage geometry, and susceptor position has been studied. The gas temperatures obtained from modeling have been compared to experimental measurements of rotational temperatures of hydrogen in the same system, and the two show good agreement, especially at higher powers and pressures.

The simulation results indicate that an increase in the input microwave power leads to an increase in the electric field, electron number density, electron temperature, as well as gas temperature in the plasma chamber, and is accompanied by an expansion of the plasma. An increase in gas pressure reduces the electric field, electron number density, and electron temperature, and increases the gas temperature. The microwave power absorbed is higher at lower pressures. When the pillar is not included, the electric field, electron number density, and electron temperature are found to stay almost constant within the plasma region both axially and radially. The gas temperature almost stays constant radially, but in the axial direction, it peaks at about 20 mm above the puck surface. Both the numerical and experimental results show a weak dependence on input power and a strong dependence on gas pressure.

The inclusion of the pillar above the surface of the puck causes high electric field to be concentrated near its top and thereby, at 500 W and 30 Torr, the maximum electric field increases by 50%. This increases the maximum value of electron temperature by 70%, the maximum gas temperature by 50%, and the maximum electron number density by about 70%. These high values are seen right above the top of the pillar's topmost surface and these values decrease gradually through the plasma both in the axial as well as the radial directions. The electron number density contours show qualitative agreement with optical emission images of the plasma at similar operating conditions with and without the pillar. The inclusion of the pillar is favorable for growth of nanostructures because it concentrates the plasma of high strength in a smaller region near the substrate and improves the efficiency of microwave power absorption.

When the susceptor position is lowered, the plasma moves from the puck surface to the center of the plasma chamber, away from where the substrate would be placed which is not favorable to nanostructure growth. When the susceptor is raised by a small value (25 mm), it causes the plasma to be concentrated in a smaller region above the puck surface, increasing the values of electric field, electron number density, and electron temperature which is favorable. However, if the power were to increase further, the configuration may produce two ball plasma which is highly undesirable. When the susceptor is raised by a large value (50 mm), it causes concentration of plasma on the sides of the stage which causes a large part of the power to be taken away from the main plasma region above the puck surface, which is also not desirable. This indicates that the best position for the susceptor would be to raise it by less than 25 mm from its initial position.

This work was supported by NSF CMMI Grant No. 1344654. We thank Di Huang for his work on the initial modeling of the system and Dr. Abbas Semnani for his insight on the TE-TM mode conversion. We also thank Prof. Sergey Macheret for his suggestion of the maximum electron number density.

1.
M.
Shokrieh
and
R.
Rafiee
, “
A review of the mechanical properties of isolated carbon nanotubes and carbon nanotube composites
,”
Mech. Compos. Mater.
46
,
155
172
(
2010
).
2.
M.
Shokrieh
and
R.
Rafiee
, “
Prediction of Young's modulus of graphene sheets and carbon nanotubes using nanoscale continuum mechanics approach
,”
Mater. Des.
31
,
790
795
(
2010
).
3.
J.
Che
,
T.
Cagin
, and
W.
Goddard
 III
, “
Thermal conductivity of carbon nanotubes
,”
Nanotechnology
11
,
65
(
2000
).
4.
J.
Hone
,
M.
Llaguno
,
N.
Nemes
,
A.
Johnson
,
J.
Fischer
,
D.
Walters
,
M.
Casavant
,
J.
Schmidt
, and
R.
Smalley
, “
Electrical and thermal transport properties of magnetically aligned single wall carbon nanotube films
,”
Appl. Phys. Lett.
77
,
666
668
(
2000
).
5.
See http://www.engineershandbook.com/MfgMethods/cvd.htm for “Engineer's Handbook” (last accessed September 16,
2015
).
6.
M.
Meyyappan
, “
A review of plasma enhanced chemical vapour deposition of carbon nanotubes
,”
J. Phys. D: Appl. Phys.
42
,
213001
(
2009
).
7.
K.
Hassouni
,
F.
Silva
, and
A.
Gicquel
, “
Modelling of diamond deposition microwave cavity generated plasmas
,”
J. Phys. D: Appl. Phys.
43
,
153001
(
2010
).
8.
K.
Hassouni
,
T.
Grotjohn
, and
A.
Gicquel
, “
Self-consistent microwave field and plasma discharge simulations for a moderate pressure hydrogen discharge reactor
,”
J. Appl. Phys.
86
,
134
151
(
1999
).
9.
M.
Füner
,
C.
Wild
, and
P.
Koidl
, “
Numerical simulations of microwave plasma reactors for diamond CVD
,”
Surf. Coat. Technol.
74–75
,
221
226
(
1995
).
10.
M.
Füner
,
C.
Wild
, and
P.
Koidl
, “
Simulation and development of optimized microwave plasma reactors for diamond deposition
,”
Surf. Coat. Technol.
116–119
,
853
862
(
1999
).
11.
H.
Yamada
,
A.
Chayahara
, and
Y.
Mokuno
, “
Simplified description of microwave plasma discharge for chemical vapor deposition of diamond
,”
J. Appl. Phys.
101
,
063302
(
2007
).
12.
H.
Yamada
,
A.
Chayahara
,
Y.
Mokuno
,
Y.
Horino
, and
S.
Shikata
, “
Simulation of microwave plasmas concentrated on the top surface of a diamond substrate with finite thickness
,”
Diamond Relat. Mater.
15
,
1383
1388
(
2006
).
13.
H.
Yamada
,
A.
Chayahara
,
Y.
Mokuno
,
Y.
Soda
,
Y.
Horino
, and
N.
Fujimori
, “
Numerical analysis of power absorption and gas pressure dependence of microwave plasma using a tractable plasma description
,”
Diamond Relat. Mater.
15
,
1395
1399
(
2006
).
14.
H.
Yamada
,
A.
Chayahara
,
Y.
Mokuno
,
Y.
Horino
, and
S.
Shikata
, “
Numerical analyses of a microwave plasma chemical vapor deposition reactor for thick diamond syntheses
,”
Diamond Relat. Mater.
15
,
1389
1394
(
2006
).
15.
P.
Bruno
,
F.
Bénédic
,
F.
Mohasseb
,
G.
Lombardi
,
F.
Silva
, and
K.
Hassouni
, “
Improvement of nanocrystalline diamond film growth process using pulsed Ar/H2/CH4 microwave discharges
,”
J. Phys. D: Appl. Phys.
37
,
L35
(
2004
).
16.
K.
Hassouni
,
F.
Mohasseb
,
F.
Bénédic
,
G.
Lombardi
, and
A.
Gicquel
, “
Formation of soot particles in Ar/H2/CH4 microwave discharges during nanocrystalline diamond deposition: A modeling approach
,”
Pure Appl. Chem.
78
,
1127
1145
(
2006
).
17.
G.
Nienhuis
,
W.
Goedheer
,
E.
Hamers
,
W.
Van Sark
, and
J.
Bezemer
, “
A self-consistent fluid model for radio-frequency discharges in SiH4–H2 compared to experiments
,”
J. Appl. Phys.
82
,
2060
2071
(
1997
).
18.
D.
Herrebout
,
A.
Bogaerts
,
M.
Yan
,
R.
Gijbels
,
W.
Goedheer
, and
A.
Vanhulsel
, “
Modeling of a capacitively coupled radio-frequency methane plasma: Comparison between a one-dimensional and a two-dimensional fluid model
,”
J. Appl. Phys.
92
,
2290
2295
(
2002
).
19.
A.
Gorbachev
,
V.
Koldanov
, and
A.
Vikharev
, “
Numerical modeling of a microwave plasma CVD reactor
,”
Diamond Relat. Mater.
10
,
342
346
(
2001
).
20.
M.
Lieberman
and
A.
Lichtenberg
,
Principles of plasma discharges and materials processing
(
Wiley
,
2005
).
21.
D.
Huang
, “
Numerical simulation of hydrogen plasma in MPCVD reactor
,” M.S. thesis (
Purdue University
, School of AAE,
2014
).
22.
G.
Hagelaar
and
L.
Pitchford
, “
Solving the Boltzmann equation to obtain electron transport coefficients and rate coefficients for fluid models
,”
Plasma Sources Sci. Technol.
14
,
722
(
2005
).
23.
M.
Maschmann
,
P.
Amama
,
A.
Goyal
,
Z.
Iqbal
, and
T.
Fisher
, “
Freestanding vertically oriented single-walled carbon nanotubes synthesized using microwave plasma-enhanced CVD
,”
Carbon
44
,
2758
2763
(
2006
).
24.
A.
Kumar
,
A.
Voevodin
,
D.
Zemlyanov
,
D.
Zakharov
, and
T.
Fisher
, “
Rapid synthesis of few-layer graphene over Cu foil
,”
Carbon
50
,
1546
1553
(
2012
).
25.
R.
Gerzeski
,
A.
Sprague
,
J.
Hu
, and
T.
Fisher
, “
Growth of contiguous graphite fins from thermally conductive graphite fibers
,”
Carbon
69
,
424
436
(
2014
).
26.
Z.
Qing
,
D.
Otorbaev
,
G.
Brussaard
,
M.
Van de Sanden
, and
D.
Schram
, “
Diagnostics of the magnetized low-pressure hydrogen plasma jet: Molecular regime
,”
J. Appl. Phys.
80
,
1312
1324
(
1996
).
27.
H.
Crosswhite
and
G.
Dieke
,
The hydrogen Molecule Wavelength Tables of Gerhard Heinrich Dieke
(
Wiley
,
1972
).
28.
R.
Garg
,
T.
Anderson
,
R.
Lucht
,
T.
Fisher
, and
J.
Gore
, “
Gas temperature measurements in a microwave plasma by optical emission spectroscopy under single-wall carbon nanotube growth conditions
,”
J. Phys. D: Appl. Phys.
41
,
095206
(
2008
).
29.
A.
Tuesta
,
A.
Bhuiyan
,
R.
Lucht
, and
T.
Fisher
, “
Laser diagnostics of plasma in synthesis of graphene-based materials
,”
J. Micro Nano-Manuf.
2
,
031002
(
2014
).
30.
G.
Janzen
,
Plasmatechnik
(
Hüthig
,
Heidelberg
,
1992
).
31.
A.
MacDonald
,
Microwave Breakdown in Gases
(
Wiley
,
1966
).
32.
A.
MacDonald
and
S.
Brown
, “
High frequency gas discharge breakdown in hydrogen
,”
Phys. Rev.
76
,
1634
(
1949
).
33.
N.
Derkaoui
,
C.
Rond
,
T.
Gries
,
G.
Henrion
, and
A.
Gicquel
, “
Determining electron temperature and electron density in moderate pressure H2/CH4 microwave plasma
,”
J. Phys. D: Appl. Phys.
47
,
205201
(
2014
).
34.
See http://fr.lxcat.net/home/ for “Plasma Data Exchange project” (last accessed April 28,
2015
).
35.
See http://jilawww.colorado.edu/∼avp/collision_data/ for “PHELPS database” (lat accessed April 28,
2015
).
36.
D.
Rapp
and
P.
Englander-Golden
, “
Total cross sections for ionization and attachment in gases by electron impact. I. Positive ionization
,”
J. Chem. Phys.
43
,
1464
1479
(
1965
).
37.
W.
Tan
and
T.
Grotjohn
, “
Modelling the electromagnetic field and plasma discharge in a microwave plasma diamond deposition reactor
,”
Diamond Relat. Mater.
4
,
1145
1154
(
1995
).
38.
V.
Kurian
,
M.
Varma
, and
A.
Kannan
, “
Numerical studies on laminar natural convection inside inclined cylinders of unity aspect ratio
,”
Int. J. Heat Mass Transfer
52
,
822
838
(
2009
).
39.
D.
Lymberopoulos
and
D.
Economou
, “
Spatiotemporal electron dynamics in radio-frequency glow discharges: Fluid versus dynamic Monte Carlo simulations
,”
J. Phys. D: Appl. Phys.
28
,
727
(
1995
).
40.
A.
Gicquel
,
M.
Chenevier
,
K.
Hassouni
,
A.
Tserepi
, and
M.
Dubus
, “
Validation of actinometry for estimating relative hydrogen atom densities and electron energy evolution in plasma assisted diamond deposition reactors
,”
J. Appl. Phys.
83
,
7504
7521
(
1998
).
41.
P.
Bruggeman
,
N.
Sadeghi
,
D.
Schram
, and
V.
Linss
, “
Gas temperature determination from rotational lines in non-equilibrium plasmas: a review
,”
Plasma Sources Sci. Technol.
23
,
023001
(
2014
).