Skip to main content

Numerical calculation of temperature, velocity, and species concentration distributions in a hot filament chemical vapor deposition in a reactor using a CH4/H2 mixture

Abstract

This study is a numerical modeling of transport phenomena occurring in the reaction chamber during diamond or amorphous hydrogenated carbon films growth by a hot filament chemical vapor deposition (HFCVD) technique. A two-dimensional model was adopted to study the HFCVD reactor. The equations of heat, momentum, and mass transfer were solved numerically; the simulation was performed using a program in FORTRAN language. All temperature, velocity, and species concentration distributions were similar at the filaments and they were also similar between the filaments. The results show that the gas temperature increases when the number of filaments increases from three to four filaments. We also noted an increase in the production of CH3 and C2H5 radicals near the surface; there was also an increase in the growth rate of the thin film. The concentrations of C2H6, C2H4, and C2H5 were very high. Temperature and concentrations were affected by the distance between filaments and the distance filaments-substrates.

Introduction

Thin films based on diamond or on hydrogenated amorphous carbon (a-C:H) present many technological interests: a-C:H films are commonly used, for example, in organic printable electronic devices [1]. They can be used as protective and antireflection films on optical instruments and on solar cells [2, 3]. They can also be used in optoelectronic device [4] to develop sensors (LED: Light-Emitting Diode), electrochemical detectors, bioelectronic and bioelectrochemical applications [5,6,7,8].

These thin films can be produced by chemical vapor deposition (CVD) using a CH4/H2 gas mixture, in which H2 is dissociated into atomic H and CH4 undergoes pyrolysis reactions leading to the formation of radicals such as CH3 and CH2 and stable species such as C2H2, C2H4, and C2H6 [9]. Activation of chemical reactions can occur at elevated temperatures, up to 3000 K in plasmachemical processes [10]. The radicals H and CH3 are very important for the mechanisms of deposition [11].

The hot filament chemical vapor deposition (HFCVD) technique is very suitable because it is economic and scalable to large areas [12]. The HFCVD technique is very popular as it allows coating of complex shapes and internal surfaces, with a low operating cost [13]. An advantage of hot filament CVD reactors over the other reactors is the simplicity of the system and the uniform diamond deposition over a large area. HFCVD is a simple and a reproducible way to grow diamond at low pressures. It is also the first method to achieve nucleation and continuous growth of diamond on various substrates [9].

Diamond carbon can be deposited on a substrate made of Si, Mo, or silica, etc., which is mounted at a distance of 0.5 to 2 cm from the glowing filament and kept at 700 to 1000 °C either by the radiation from the filament or by a separate substrate heater [9]. A group of parallel filaments can be arranged in single or multiple parallel planes and in horizontal or vertical orientation [14].

In the HFCVD process, the production of thin films depends on substrate and gas temperatures and the velocity and volume density of gas as well as on complex chemical reactions of the gas phase [15]. Goodwin and Gavillet [16] developed the first one-dimensional numerical model for the diamond growth. They calculated temperature, velocity, and species concentration profiles. Debroy et al. [17] developed the first two-dimensional model of the HFCVD process. The authors showed that diffusion mechanisms are mostly responsible for both heat and mass transfer inside the reaction chamber. Some studies have focused on the concentrations of different species and radicals, in particular CH3, and the growth rate of thin layers of diamonds [18,19,20]. One-dimensional models [21, 22] were developed to understand the deposition processes and the gas-phase behavior. Other numerical studies about two- and three-dimensional models [17, 19, 23,24,25,26,27] were interested in studying the effect of the parameters such as reactor geometry, the diameter, the number of the filaments, and the separation between hot filaments in the growth process of diamond films. These numerical models allow the creation of simplified models to reduce experiences that are often expensive [10].

Many other works [28,29,30,31,32,33] were focused on several aspects related to HFCVD processes (physical parameters of gas, gas-phase diffusion, chemical reaction rates, equilibrium constants, concentrations of species, filament temperature, filament geometry, filament arrangement, deposition parameters…); these factors can affect the growth of thin films.

In experimental and calculation work, Kondoh et al. [18] studied the effect of some HFCVD parameters on diamond growth in CH4/H2 gas mixture. Spatial distributions of the gas temperature, fluid velocity, and chemical species concentrations have been calculated in a two-dimensional HFCVD system for CH4/H2 gas mixture [18]. The partial differential equations used were those of conservation of mass, momentum, energy, and chemical species. May et al. [20] studied mechanism of growth in diamond films using Ar/CH4/H2 gas mixtures in the HFCVD reactor. They used a two-dimensional computer model to calculate the gas-phase composition within the reactor. Using experimental and calculated data, they showed the competition between H atoms, CH3 radicals, and other species film morphology. Eckert et al. [34] were interested in the calculation of the sticking coefficients of species CHx (x: 0–4), C2Hx (x: 0–6), C3Hx (x: 0–2), C4Hx (x: 0–2), H, and H2 on clean and hydrogenated diamond. They used the molecular dynamic simulation method.

In recent work, Tao Zhang et al. [35] presented a three-dimensional model for depositing the diamond films by a hot filament CVD. Their simulation model coupled heat transfer mechanisms by radiation, by conduction, and by convection. The authors showed that the thermal conductivity influences the temperature. Rebrov et al. [36] and Gorbachev et al. [37] studied the impact of gas-phase reactions on characteristics of the flow in a heated cylindrical channel of a gas mixture CH4/H2 reactor with using the direct simulation Monte Carlo method which is a method based on solving the Navier-Stokes equations. The study shows the complexity of the influence of the various physical and geometric parameters on the concentrations and on the deposition.

In this work, we present a two-dimensional model for HFCVD, for a gas mixture CH4/H2. This model studies distributions of temperature, velocity, and species concentration and the effect of the number of filaments and their distance from substrate. The proposed geometry of the reactor is different from several shapes cited in the literature. Numerical simulation can give extensive information about effect of several factors on gas mixture of HFCVD process.

The proposed reactor has two substrates of horizontal form. Long, equidistant and coplanar hot filaments are parallel to the substrates. The gas mixture enters the reactor with a constant flow. The fluid is considered Newtonian and the flow is laminar. We use the fluid model in the two-dimensional stationary regime to calculate spatial distributions of gas velocity and gas temperature and species concentrations.

In the second section, we present the principal equations of the mathematical model and the proposed calculation method. We use the nonlinear stationary differential equations: Navier-Stokes equations and continuity equation for velocity, heat equation for temperature, and equation of diffusion for concentrations. In the third section, we present the numerical model based on the finite volume method (FVM) with an adequate choice of boundary conditions. The results of this work are presented in the fourth section. Spatial distributions of velocities, temperature, and species concentrations are calculated for a gas mixture CH4/H2 with discussion and comments. First, we present distributions of velocity and temperature for three and four filaments. In the next step, we present species concentrations in volume, effects of number of filaments, and effects of distance filaments-substrates. In this fourth section, we also present concentrations of CH3, C2H5, and H near the substrates. As an application, an estimation of the growth rate of deposited thin film is made. The last section includes the conclusion of this work and the future prospects.

Methods and mathematical model

The principal aim is calculation of distributions of temperature, velocity, and species concentration in an HFCVD process. The considered HFCVD reactor has two openings for the inlet and outlet gas. The substrates are horizontal plates; they have a constant temperature Tsub. Let L be the length of the reactor in the x direction and let Zmax be the distance between substrates. Long, equidistant and coplanar hot filaments are parallel to the substrates. In the HFCVD reactor, there are “n” filaments which are heated to a same temperature Tfil. The distance between two filaments is Dff. A gas mixture CH4/H2 enters the reactor with a constant flow in the x direction. The initial composition of the mixture gas is 5% of methane CH4 and 95% of hydrogen H2. The given initial temperature, velocity, and pressure are respectively T0, U0, and P0. The volumetric mass density ρ0 is given according to ideal gas state equation as ρ0 = P0/RT0, where R is the gas constant. A simplified scheme of the HFCVD reactor used in the model is shown in Fig. 1. Table 1 presents values of parameters used for the reactor. A numerical study of the phenomena is required.

Fig. 1
figure 1

Simplified scheme of the HFCVD reactor

Table 1 Values of parameters used for the reactor

For the mathematical model, we consider a gas mixture in two-dimensional coordinates in stationary regime (independent of time). We assume that the substrates are wide enough (direction y) so that this space dimension does not affect the results. We use the fluid model: the fluid is considered Newtonian and the flow is laminar; the effect of radiation has not been taken into account. The flow gas in the reactor obeys the conservation laws of mass (continuity equation) Eq. (1), momentum (Navier-Stokes equations) (Eq. (2) and Eq. (3)), and energy diffusion equation Eq. (4). To calculate gas velocity and gas temperature, we use the following nonlinear differential equations:

$$\mathit{\operatorname{div}}\left(\rho \overrightarrow{\boldsymbol{V}}\right)=0$$
(1)
$$\rho U\frac{\partial U}{\partial x}+\rho W\frac{\partial U}{\partial z}=-\frac{\partial p}{\partial x}+\mu \left(\frac{\partial^2U}{\partial {x}^2}+\frac{\partial^2U}{\partial {z}^2}\right)+\rho {g}_x$$
(2)
$$\rho U\frac{\partial W}{\partial x}+\rho W\frac{\partial W}{\partial z}=-\frac{\partial p}{\partial z}+\mu \left(\frac{\partial^2W}{\partial {x}^2}+\frac{\partial^2W}{\partial {z}^2}\right)+\rho {g}_z$$
(3)
$$\rho {C}_p\left(U\frac{\partial T}{\partial x}+W\frac{\partial T}{\partial z}\right)=\lambda \left(\frac{\partial^2T}{\partial {x}^2}+\frac{\partial^2T}{\partial {z}^2}\right)$$
(4)

where U and W are components of the velocity \(\overrightarrow{\boldsymbol{V}}\) on directions x and z respectively; T is the temperature of gas. The constants ρ, μ, λ, and Cp, are the volume density, the dynamic viscosity, the heat transfer efficiency, and the specific heat respectively. In this model, the various species considered in the calculations are nine species (kmax = 9): H2, CH4, H, CH3, C2H2, C2H3, C2H4, C2H5, and C2H6. To resolve the concentration Ck of each radical or molecule (k = 1, 2, …, kmax), we consider the equation of chemical species [38, 39]:

$$\overrightarrow{\boldsymbol{V}}\overrightarrow{\boldsymbol{\operatorname{grad}}}{C}_k={D}_k{\nabla}^2{C}_k+{R}_k$$
(5)

where Dk is the diffusivity of species k and Rk is its rate production (or consumption) in volume.

Table 2 presents the gas-phase reaction mechanism; these reactions were used in works of Olivas-Martinez et al. [15] and Mankelevich et al. [23]. The gas-phase reaction mechanism consists of 13 reactions. We calculate the variations of chemical rate constants Kr with the temperature T for interactions between molecules and radicals. The general form is Kr = ArTβ exp(−Ea/RT), where Ar and β are constants, Ea is the reaction activation energy, and R is the gas constant. For the units, Ar is in mol cm−3 s−1; Ea is in kJ mol−1, T is in K, and R is in J mol−1 K−1.

Table 2 Gas-phase reaction mechanism and corresponding rate constants

The general chemical reaction is expressed as [40]:

$${a}_1{A}_1+{b}_1{B}_1+..\dots \Rightarrow {a}_2{A}_2+{b}_2{B}_2+..\dots$$
(6)

The general form of reaction rate is given by:

$${R}_k={K}_r(T){\left[{A}_i\right]}^{a_i}{\left[{B}_i\right]}^{b_i}.\dots$$

where Kr(T) is the reaction rate constant, [Ai], [Bi] are the species concentrations, and ai, bi,..., are stoichiometric coefficients of chemical reactions. The diffusion coefficient Dk of a particle k in a gas can be calculated from the coefficient Dmm’ of binary collisions between two species m and k using the formulas below [41,42,43]:

$${D}_k=\frac{\left({C}_{tot}-{C}_k\right)}{\left(\sum \limits_{m=1,m\ne k}^{k_{\mathrm{max}}}\frac{C_m}{D_{km}}\right)}$$
(7)

where Ctot is the total concentration and Dmk is the diffusion coefficients of the neutral species k in each of the background gases m.

$$D_{mk}=\frac3{16}\frac{\left(4\pi K_bT/2M_{mk}\right)^{1/2}}{C_{tot}{\pi\sigma^2}_{mk}\Omega_D\left(T^\ast\right)}$$
(8)

Dmk is in m2/s, Ctot = P/KbT in m−3, Kb is the Boltzmann constant (J K−1)

Mmk is the reduced mass \({M}_{mk}=\frac{M_m\times {M}_k}{M_m+{M}_k}\); σmk is the binary collision diameter \({\sigma}_{mk}=\frac{\sigma_m+{\sigma}_k}{2}\), and it is in Å.

\({T}^{\ast }=\frac{T}{\sigma_{mk}}\) with the binary collision diameter εmk = (εm × εk)1/2 in K. εm and εk are Lennard-Jones parameters.

$${\Omega}_D\left({T}^{\ast}\right)=\frac{A_D}{T^{\ast {B}_D}}+\frac{C_D}{e^{D_DT\ast }}+\frac{E_D}{e^{F_DT\ast }}+\frac{G_D}{e^{H_DT\ast }}$$
(9)

with AD = 1.06036, BD = 0.15610, CD = 0.19300, DD = 0.47635, ED = 1.03587, FD = 1.52996, GD = 1.76474, HD = 3.89411 [41, 42]:

The used values of Lennard-Jones parameters are cited in [44].

In the calculation, we take for the physical constants ρ, μ, λ, and Cp those of mixture gas where the composition is dominant; they are according to data used in [31]. The details of the method or the numerical modeling for the treatment of the problem are presented in the next section.

Numerical model and boundary conditions

The numerical model is based on the finite volume method (FVM) [45]. We discretize the spatial domain of the reactor at discrete nodes (points). Let be a constant step along x axis and let i be the index of the corresponding node; let be a constant step along z axis, and let j be the index of the corresponding node. As it is known in the FVM, we apply an integral around an elementary volume at the level of the node (i,j).

We apply FVM, to calculate the component U by resolving Navier-Stokes equation Eq. (2); the component W is deduced from continuity equation Eq. (1). We apply FVM, to calculate the temperature T by resolving heat equation Eq. (4). We apply FVM also, to calculate the gas-phase species concentrations by resolving diffusion equation Eq. (5). The parameters Kr and Rk are calculated according to the data of the temperature T and the concentrations Ck of the point of coordinates (x, z).

For a fixed value i and a variable value j, the development of Eqs. (2), (4) and (5) is of the form:

$${\alpha}_{i,j-1}^i{X}_{j-1}^i+{\alpha}_{i,j}^i{X}_j^i+{\alpha}_{i,j+1}^i{X}_{j+1}^i={\beta}_j^i$$
(10)

where parameters \({\boldsymbol{\upalpha}}_{\mathbf{j},\mathbf{j}-\mathbf{1}}^{\mathbf{i}}\), \({\boldsymbol{\upalpha}}_{\mathbf{j},\mathbf{j}}^{\mathbf{i}}\), \({\boldsymbol{\upalpha}}_{\mathbf{j},\mathbf{j}+\mathbf{1}}^{\mathbf{i}}\), and \({\boldsymbol{\beta}}_{\boldsymbol{j}}^{\boldsymbol{i}}\)are constants and parameters related to physical properties and numerical values of the discretized system.

It is clear to see the matrix form [A] × [X] = [B] of Eq. (10). The matrix [A] has the elements \({\boldsymbol{\upalpha}}_{\mathbf{j},\mathbf{j}-\mathbf{1}}^{\mathbf{i}}\), \({\boldsymbol{\upalpha}}_{\mathbf{j},\mathbf{j}}^{\mathbf{i}}\), \({\boldsymbol{\upalpha}}_{\mathbf{j},\mathbf{j}+\mathbf{1}}^{\mathbf{i}}\) (which are not zero), the matrix (or vector) [B] has the element \({\boldsymbol{\upbeta}}_{\mathbf{j}}^{\mathbf{i}}\), and the matrix (or vector) [X] presents the unknown variable (\({\mathbf{T}}_{\mathbf{j}}^{\mathbf{i}}\), \({\mathbf{U}}_{\mathbf{j}}^{\mathbf{i}}\), or \({\mathbf{C}}_{\mathbf{k},\mathbf{j}}^{\mathbf{i}}\)). For a fixed value i, this system of coupled equations is easily solved with the Gauss-Seidel (GS) method [46,47,48], and with the solution convergence test for values i and j, we have iterative calculation. For an iteration p, convergence is reached for absolute value ((Xp+1-Xp)/Xp+1) lower than a limit value that is too small. This general form to solve equations is the Gauss-Seidel iterative method (GSIM).

The boundary conditions for components U and W of velocity vector (\(\overrightarrow{\mathrm{V}}\)), temperature T, and concentrations Ck are as follows:

  • In flow (x = 0): U = U0, W = 0, T = T0; Ck = Ck,z = 0;

  • Out flow (x = L) : \({\left(\frac{\partial U}{\partial x}\right)}_{x=L}=0\),W = 0, \({\left(\frac{\partial T}{\partial x}\right)}_{x=L}=0,\ {\left(\frac{\partial C}{\partial x}\right)}_{x=L}=0;\)  

  • At filament \(\left(x={X}_{fil},z=\frac{Z_{\mathrm{max}}}{2}\right)\)U = 0, W= 0, T = Tfil; the filaments are considered punctual without dimensions;

  • At substrates (z = Z or z = Zmax): U = 0, W = 0, T = Tsub, \({\left(\frac{\partial C}{\partial x}\right)}_{x=L}=0 .\)

These boundary conditions give the elements of matrix [A] and matrix [B] corresponding to values of i and j.

To avoid the problems of singularities, we choose continuous functions for the initialization of the values of the components of the velocities and the initialization of the values of the temperatures. We choose a parabolic variation for U0 (as function of z) and a linear variation for Tsub (as function of x) near the entrance of the reactor. To have the initialization of concentrations, we calculate them for each temperature from 300 to 2100 K; we use GSIM to solve equations for concentrations Ck,0(i,j):

$${R}_k=\sum \left(\pm {K}_r(T){\left[{A}_m\right]}^{\alpha }{\left[{B}_{m\hbox{'}}\right]}^{\beta }..\dots \right)=0$$
(11)

We developed a numerical program using FORTRAN language; the general numerical scheme for calculation of T, U, W, and Ck is summarized in three stages as follows:

  • Stage 1: Initialization

- Initialization of T, U, W with continues values and according to boundary conditions

  • Stage 2: Calculation of T, U, W without variation of Ck

By a loop on index i, we have to:

  1. (a)

    Calculate T by GSIM

  2. (b)

    Calculate U by GSIM and correction of flux

  3. (c)

    Calculate W from continuity equation

  4. (d)

    Return to step (a) until convergence for all values i and j

  • Stage 3: Calculation of T, U, W, and Ck

The calculated values on step 2 are initial values for this step. By a loop on index i, we have to:

  1. (a)

    Initialize T, U, W, and Ck

  2. (b)

    Calculate T by GSIM

  3. (c)

    Calculate U by GSIM and correction of flux

  4. (d)

    Calculate W from continuity equation

  5. (e)

    Calculate Kreac and Dk as a function of z (or index j)

  6. (f)

    Calculate Ck for different k particles by GSIM

  7. (g)

    Return to step (b) until convergence for all values i and j

  8. (h)

    Take final results and calculate different parameters

Results and discussion

In this section, we present temperatures, velocities, and concentrations of molecules and radicals for a reactor with three filaments (n = 3) and four filaments (n = 4). Data values are P0 = 103 Pa (7.5 Torr), U0 = 10 mm/s, Tfil = 2100 K, T0 = 300 K, and Tsub = 1173 K. Most results are for dimensions of the reactor: Zmax = 20 mm, L = 240 mm, Dff = 50 mm for n =3 and Zmax = 20 mm, L = 240 mm, Dff = 33.3 mm for n = 4. The composition of the gas mixture is 05% of CH4 and 95% of H2. These values of filament temperature (2100 K) and gas mixture rate (05%, 95%) are according to data used in [15, 16, 31].

Distributions of velocity and temperature

In this subsection, we present distributions of temperature T and distributions of velocity components U and W for the cited values. A few characteristic planes are defined. The plane of the filaments is the horizontal plane carrying the filaments for z = Zmax/2; the plane of a filament is the vertical plane containing a filament (x = 7 cm, 11.99 cm, 16.98 cm or n = 3 filaments and x = 8.66 cm, 11.99 cm, 15.30 cm, 18.61 cm for n = 4 filaments). We also define the vertical planes between filaments (example x = 9.5 cm for n = 3 filaments).

For n = 3, the calculations show that the temperature distributions at the hot filaments are identical; similarly, the temperature distributions between the hot filaments are identical. For n = 4, we have very similar curves and the conclusions are also identical. Figure 2 shows the distribution of the gas temperature as a function of z between filaments and at each filament for three and four filaments. The temperature distributions are symmetrical. The temperature takes a maximum value at the center of the reactor (Tfil = 2100 K, z = Zmax/2) when the gas comes near to the hot filaments along the z direction, and then decreases gradually when the gas goes far away from the filaments (Fig. 2, at filaments). Between the filaments, temperature increases from 3 to 4 filaments by 11.46% (Fig. 2, between filaments). The temperatures in the central regions, around filaments, are more than the temperatures far from the filament region; this result confirms other results [49].

Fig. 2
figure 2

Distribution of T as a function of z between filaments and at the filaments for n = 3 and for n = 4

Figure 3 shows the distribution of the gas temperature as a function of x for horizontal filament plane (z = Zmax/2) for n = 3 and n = 4. The symmetry at the filaments and between the filaments is confirmed. The zones between filaments have a remarkable periodicity. The difference at the inlet of the reactor and the outlet is due to the initial conditions and the conditions of the fluid flow. Indeed, at the inlet of the reactor, the temperature of the gas is T0 = 300 K and that of the substrate is Tsub = 1173 K; we have assumed a linear variation as a function of x at the first zone of the substrates. This choice of continuous function avoids the problems of singularities when solving the differential equations.

Fig. 3
figure 3

Distribution of T as a function of x for z = Zmax/2 and for n = 3 and n = 4

As for the temperature distributions, the calculations show that the distributions of components u and W of velocity as function of z at the filaments are identical. These distributions between the filaments are also identical. Figure 4 shows distribution of the component U of the velocity as a function of z between filaments and at the filaments for three filaments and four filaments. The increased gradient between the filaments was much larger than that in the filaments; the maximum value of velocity Umax is at 15.23 mm/s (for n = 3) and at 14.96 mm/s (for n = 4). Velocity distribution has near parabolic form far from filaments. For exact parabolic form Umax = (3/2)U0. The velocity reaches zero values on filaments by the severe interaction between filaments and the flow of viscous gas [31]. Between filaments, the value of Umax for 3 filaments is greater than that for 4 filaments by 2.98%. These results were consistent with what was found by Goodwin and Gavillet [16]: velocity distribution has a parabolic form far from filaments.

Fig. 4
figure 4

Distribution of U as a function of z, between filaments for n=3 (a) and for n=4 (b) and at filaments for n=3 (c) and for n=4 (d)

For three filaments and four filaments, the distribution of the component W of the velocity as a function of z is also symmetrical, between filaments and at the filaments.

Distributions of concentrations

In this subsection, we mainly present the concentrations of molecules and radicals in the reactor for n = 3 and n = 4.

Concentration of CH3, H, and C2H5

Calculations show that the curve of concentration of each radical CH3, H, or C2H5 is symmetrical and perfectly identical at the filaments. Between filaments, the curves of concentrations of these radicals are also symmetrical and identical.

Figure 5 shows the distributions of CH3 radical concentration as function of z at the filaments and between the filaments for n = 3 and for n = 4. The curves are symmetrical. At the filaments, the comparison shows that the concentration for n = 3 is higher than that for n = 4; the difference varies from 1.59 to 1.79%. Between the filaments, the concentrations for n = 4 is higher than that for n = 3, and it is from 3.95 to 2.24%.

Fig. 5
figure 5

Distribution of concentration of CH3 as function of z; at the filaments and between the filaments

The calculation of concentrations of H, as a function of z at the filament for n = 3 and n = 4, shows that the concentrations have approximate values; the concentration for n = 4 is higher than that for n = 3, and the difference is from 1.31 to 0.86%. Between the filaments, the concentration for n = 3 is higher than that for n = 4; the difference is from 3.13 to 2.95%.

For C2H5 radical, the concentration for n = 3 is higher than that for n = 4; the difference is from 5.66 to 1.25% at the filaments. Between the filaments, the concentration for n = 4 is higher than that for n = 3; it varies from 61 to 17%.

Concentrations of CH4, H2, and other species C2H6, C2H4, C2H2, C2H3

Our calculations show that the distributions of concentration of C2H6, C2H4, C2H2, and C2H3, as a function of x, for filaments’ plane (z = Zmax/2) and substrates (z = 0, z = Zmax) for n = 4 have the same remarks concerning forms and the symmetry as distributions of CH3, H, and C2H5.

As it has shown before, the temperature curves are identical and superimposable at the level of the vertical planes containing the filaments (at filaments). At the median vertical planes between two successive filaments, the temperature curves or distributions are identical and superimposable. We have homogeneity or periodicity of the zones between two successive filaments. The distributions of velocity components U and W have the same shapes respectively component to component.

The distributions of the concentrations Ck of all the species (k = 1,…, 9), as a function of z, have respectively identical curves at vertical planes containing the filaments (at filaments). We note the same remark for median vertical planes between filaments; there is homogeneity or periodicity of the zones between two successive filaments. This result is also linked to the choice of entry conditions for the gas mixture (U0 and P0) relative to the dimensions of the reactor (L, Zmax, Dff). This choice gives a periodicity mainly for T, U, and W. These results have the same remarks for the concentrations which results are mainly dependent on the reaction constants Kreac and the diffusion coefficients Dk; the effect of temperature (filaments and substrates) is the most important. Debroy et al. [17] concluded that diffusion mechanisms are mostly responsible for both heat and mass transfer inside the HFCVD reaction chamber.

Figure 6 shows the concentrations Ck of the different molecules and radicals as a function of z, four n = 4, between the filaments (Fig. 6a–c) and at the filaments (Fig. 6b–d). There is an order in the presence of different species in the reactor; the molecules of H2 and CH4 are dominant. For CH4 and between the filaments, the concentration for n = 3 is higher than that for n = 4; the difference is from 4.10 to 2.31%. At the filament, for n = 3 and n = 4, the concentrations have approximate values; the difference is about 1.21 to 0.73%.

Fig. 6
figure 6

Concentrations Ck of molecules and radicals as a function of z, for n = 4, a, c between the filaments, b, d at the filaments

The third species is ethane C2H6, in the order of 1018 m−3. For the fourth and fifth order, there are C2H4 and C2H5 with close concentration values (1016 m−3 at 1015 m−3). Contrary to the plasma enhanced chemical vapor deposition (PECVD) processes where the concentration of methyl CH3 is high, here in HFCVD processes, the concentration of ethyl C2H5 is higher than that of CH3. Other works [18,19,20] give concentration of CH3 higher than those of C2H6, C2H4, and C2H5; this result is different from ours. We have shown in a conference presentation that in HFCVD processes, temperature is too high and the reaction rates which give heavy elements (C2H6, C2H4, C2H5) are more important. Indeed, chain reactions will allow consumption of CH3 radicals by the reactions R6 and R7. We have in the sixth order CH3 with a concentration less than 1015 m−3. Although the hydrogen molecule H2 is dominant, the atomic hydrogen H has a low concentration.

The initial concentrations of H2 and CH4 are respectively 2.29 × 1023 m−3 and 1.20 × 1022 m−3, so 95% and 5% respectively. Inside the reactor, at the filaments’ plane (z = Zmax/2), we note approximate rate of H2 and CH4 of 98.1% and 1.9% respectively; there is production of H2 and consumption of CH4. We note also approximate productions of C2H6 and C2H4, about 2.3 × 10−3% and 3.6 × 10−5% respectively. It is in the seventh order with a concentration of around 1015 m−3. The concentration of C2H3 varies from 1.28 × 108 to 9.48 × 108 m−3. In the last position, we have the radical C2H2 with a very low concentration (less than 10−2 m−3).

Table 3 gives the average values of the concentrations of the different species near the substrates (z = 0 and z = Zmax) and at the plane of the filaments (z = Zmax/2); the average values are taken on the x range from 50 mm to 230 mm (to avoid the effects of the edges).

Table 3 Average concentrations of species at the substrates and at the plane of the filaments

For real experiments, at the outlet of the reactor, the radicals C2H5, CH3, and C2H3 and H should give stable molecules (H2, CH4, C2H6, C2H4, and C2H2). We think that this effect will not have too much influence on the results of our calculations nor the generality of the model used.

Effects of number of filaments and distance filaments-substrates

As we mentioned, the calculation shows the homogeneity and periodicity between the regions for n = 3 and for n = 4.

Table 4 shows the effect of number of filaments on concentrations of radicals H, CH3, and C2H5. As the number of filaments increases, the distance between filaments decreases; heating energy input increases and the temperature of the gas increases. We observe that the production of CH3 and C2H5 near the surface increases and at the same time the growth rate of the thin film will increase.

Table 4 Effect of the number of filaments on concentrations

Table 5 presents the effect of distance filaments-substrates Zmax/2 on concentrations of radicals H, CH3, and C2H5 for n = 4. The values of U0 are taken to have similar flow velocities (Umax is of the order of 15 mm/s). The maximum values of the speeds and the minimum values of the temperatures are calculated at Zmax/2 and between filaments; the values of the concentrations are the average values taken for x varying from 50 to 230 mm. If the distance filaments-substrates increase, the average concentrations of CH3 and C2H5 increase and the average concentrations of H decreases. The effects of chemical reactions and temperature in volume are the most important.

Table 5 Effect of distance filaments-substrates on concentrations for n = 4

Concentrations of CH3, C2H5, and H near the substrates

In our calculations of the concentrations in volume, we did not consider the deposition and the interaction of particles with the surface. This assumption will not significantly influence concentrations in the volume, the amount deposited is too low compared to the initial amount: this is the effect of the bath in physics.

The temperature of the substrate is Tsub = 1173 K. Near the substrates, the average concentrations of H and CH3 are practically constant for n = 3 and n = 4; the concentration ratio CH/(CH + CCH3) is 20%. The concentration of C2H5 is 4.5 greater than that of H and CH3 for n = 3, and it is 5.1 greater than that of H and CH3 for n = 4.

The type of substrates plays an important role in the growth of thin film. A growth rate can be estimated by proposing a diamond layer growth with a given sticking coefficient. We consider sticking coefficients SH = 0.01, SCH3 = 0.11, and SC2H5 = 0.07 of H, CH3, and C2H5 respectively [16, 32, 50]. For particles k of mass mk at temperature near substrate TnS = Tsub, we calculate the most probable thermal velocity [51]:

$${V}_k-=\sqrt{\frac{2{K}_b{T}_{ns}}{m_k}}$$
(12)

For diamond carbon thin film, we can estimate the value of growth rate Gk resulting from the radical k. An approximate formula of Gk is:

$${G}_k=\frac{1}{d}{C}_{k, av}\times \left(\frac{1}{6}{V}_k\right)\times {S}_k\times {m}_{kd}$$
(13)

where, for radicals H, CH3, and C2H5, Ck,av is the average concentrations, Sk is the sticking coefficient, and mkdis the mass of the deposited radical (mkd is not necessarily mk, because of the hydrogen atoms which release sites from the surface). d is volumetric mass density of diamond carbon.

Table 6 presents values of growth rate G, from some values of U0 and some values of distance filaments-substrates Zmax/2.

Table 6 Values of growth rate G for some values of Zmax

For average value d = 3.51 g/cm3 and with two radicals H and CH3, for the same values Zmax = 20 mm U0 = 10 mm/s, calculated maximum values of growth rate Gmax are 0.0022 μm/h for n = 3 and 0.00221 μm/h for n = 4. For a given number of filaments (example n = 4), when Zmax decreases, there is not necessarily an increase in the growth rate. These results are smaller than other theoretical and experimental works. The values of growth rate are about 0.4 μm/h in work [18] and less than 0.2 μm/h in [20]. Taking into account the presence of C2H5, for same values Zmax = 20 mm U0 = 10 mm/s, calculated deposition rates (0.0135 μm/h for n = 3 and 0.0151 μm/h for n = 4) are lower than the values recorded in the experiments.

We can use the Hertz- Knudsen equation [52, 53] to estimate the flux of particles which collide with surface. For a particle of mass m, at temperature T, and in pressure P, the flux φ is [52]:

$$\varphi =\frac{P}{\sqrt{2\pi \times m\times {K}_b\times T}}$$
(14)

From values of flux, growth rate can be estimated for a sticking coefficient Sk and for diamond carbon density d.

The comparison of growth rate is not so reliable because of the differences in volume concentrations (gas temperature) and velocities of radicals near surfaces (substrate temperature). The expression of growth rate G used here and in other expressions do not respond to the complexity of the phenomena of deposition. Knowledge of the sticking coefficients of radicals is not sufficient, nor even the coefficients of all radicals and molecules. It would be important to know the rate of coverage of the surface with hydrogen atoms. Otherwise, it is necessary to know the free sites occupied by the hydrogen atoms. A calculation taking into account the recombination coefficients and sticking coefficients of all the radicals and molecules, as well as the occupancy rate in H, would give a more reliable result and would be close to the experimental results. Experimental measurements of the production of stable molecules (C2H6, C2H4, and C2H2) would make it possible to better understand the mechanisms and correct the different models.

Conclusions

In this work, we have presented a two-dimensional model for HFCVD for CH4/H2 gas mixture in a reactor with three and four filaments. This model describes the phenomena that occur in the reaction chamber of a reactor and the effect of the number of filaments and their position in relation to the substrate. The proposed reactor gives identical distributions of temperatures, velocities, and concentrations at the filaments and also between the filaments; the different regions between filaments show great homogeneity and remarkable periodicity. The results show that if the temperature of gas increases, the production of CH3 and C2H5 near the surface increases and the growth rate of the thin film increases. The calculations show that concentrations of C2H6, C2H4, and C2H5 are very high. The diffusion phenomena related to temperatures and the rates of chemical reactions also related to temperatures constitute the most important aspects of the dynamics of concentrations in gas.

The study of the growth of the thin film is very complex. In fact, it requires more detailed studies on the coefficients of sticking and recombination of the species and on the rate of occupation of the sites by the hydrogen atom at the deposit surface. Experimental measurements of the production of stable molecules (C2H6, C2H4 and C2H2) would make it possible to better understand the mechanisms.

Nomenclature

[Ai], [Bi] Species concentrations, m−3

ai, bi Stoichiometric coefficients of chemical reactions

A Constant, mol cm−3 s −1

Ck Concentration of each radical or molecule, m3

Ck,av Average concentrations, m3

Cp Specific heat, J kg1 K1

Ctot Total concentration, m3

d Volumetric mass density, g/cm3

Dff Distance between filaments, mm

Dk Diffusivity of species, m3

Dmk Diffusion coefficients of the neutral species k, m2/s

Ea Activation energy, KJ mol1

Gk Growth rate, μm/h

gx Gravity force component in the x direction, m s−2

gz Gravity force component in the z direction, m s−2

k Number radical or molecule

Kb Boltzmann constant, J K−1

Kr Chemical rate constants, m3/s

L Length of the reactor in the x direction, mm

mkd Mass of the deposited radical, g

Mmk Reduced mass, kg

n Number of filaments

Na Avogadro constant, mol−1

P Gas pressure, Pa

P0 Initial pressure, Pa

R Universal gas constant, J mol−1 K−1

Rk Rate production, m−3 s−1

Sk Sticking coefficients

T0 Initial temperature, K

T Temperature of gas, K

Tfil Filament temperature, K

TnS Temperature near substrate, K

Tsub Substrate temperature, K

U0 Initial velocity, mm/s

U Velocity component in the x direction, mm/s

\(\overrightarrow{\mathrm{V}}\) Gas velocity vector, mm/s

Vk Most probable thermal velocity of particle k

W Velocity component in the z direction, mm/s

x Cartesian axis directions, mm

Xfil Filament position

z Cartesian axis directions, mm

Zmax Distance between substrates, mm

Greek symbols

β Constant

μ  Dynamic viscosity, N s m−2

ρ0 Volume density, m−3

λ Heat transfer efficiency, W m−1 K−1

σmk Binary collision diameter, Å

εmk Lennard-Jones parameters K

φ Flux of the gas molecules, m−2 s−1

Special symbols

i Unitary vector in the x direction

j Unitary vector in the z direction

\(\overrightarrow{\operatorname{grad}}\) Differential operator

div Differential operator

\(\frac{\partial }{\partial x},\frac{\partial }{\partial z}\) Partial derivatives

Nabla mathematical operator

[A],[X],[B] Matrix for Gauss-Seidel calculation

\({\upalpha}_{\mathrm{j},\mathrm{j}-1}^{\mathrm{i}}\), \({\upalpha}_{\mathrm{j},\mathrm{j}}^{\mathrm{i}}\), \({\upalpha}_{\mathrm{j},\mathrm{j}+1}^{\mathrm{i}}\),\({\upbeta}_{\mathrm{j}}^{\mathrm{i}}\) Elements of matrix [A] and the matrix [B]

ΩD(T*), AD, BD,CD, DD, ED, FD, GD, HD Dimensionless collision integral function of the temperature and its parameters

Availability of data and materials

All data generated or analyzed during this study are included in this published article and are available from the corresponding author.

Abbreviations

a-C:H:

Amorphous hydrogenated carbon

HFCVD:

Hot filament chemical vapor deposition

CH4 :

Methane gas

H2 :

Hydrogen gas

LED:

Light-emitting diode

CVD:

Chemical vapor deposition

PECVD:

Plasma enhanced chemical vapor deposition

FVM:

Finite volume method

av:

Average

References

  1. Sattam A, Krishna NM, Shashi P (2018) Stability of hydrogenated amorphous carbon thin films for application in electronic devices. Diam Relat Mater 90:172–180. https://doi.org/10.1016/j.diamond.2018.10.016

    Article  Google Scholar 

  2. Mohamad F, Hanib NM, Uzer Noor UM, Rusop M (2010) Properties of amorphous carbon thin films for solar cell applications. AIP Conf Proc 1217:140–146. https://doi.org/10.1063/1.3377800

    Article  Google Scholar 

  3. Grenadyorov AS, Oskomov KV, Solovyev АА (2018) Effect of deposition conditions on optical properties of a-C:H:SiOx films prepared by plasma-assisted chemical vapor deposition method. Optik 172:107–116. https://doi.org/10.1016/j.ijleo.2018.07.024

    Article  Google Scholar 

  4. Soga T, Jimbo T, Krishna KM, Umeno M (2000) Amorphous carbon thin films for optoelectric device application. Int J Modern Phys B 14:206–217. https://doi.org/10.1142/S0217979200000200

    Article  Google Scholar 

  5. Alix G, Khaled H, François S, Jocelyn A (2001) CVD diamond films: from growth to applications. Curr Appl Phys 1:479–496. https://doi.org/10.1016/S1567-1739(01)00061-X

    Article  Google Scholar 

  6. Enric B, Markus AMH (eds) (2011) Synthetic diamond films. Book News-Inc, Canada

    Google Scholar 

  7. Lothar S, Markus H, Roland K (2006) The versatility of hot-filament activated chemical vapor deposition. Thin Solid Films 515:1017–1024. https://doi.org/10.1016/j.tsf.2006.07.073

    Article  Google Scholar 

  8. Khmelnitskiy RA (2015) Prospects for the synthesis of large single-crystal diamonds. Physics-Uspekhi 58:134–159

    Article  Google Scholar 

  9. Liu H, Dandy DS (1995) Studies on nucleation process in diamond CVD: an overview of recent developments. Diam Relat Mater 4:1173–1188. https://doi.org/10.1016/0925-9635(96)00297-2

    Article  Google Scholar 

  10. Tamilla MA (2002) Mathematical modeling of plasmachemical processes. Turk J Phys 26:415–426 https://journals.tubitak.gov.tr/physics/abstract.htm?id=5852

    Google Scholar 

  11. Waqar A, Sein H, Malcolm J, Rego C, Israr Ul H, Karthikeyan S, Javad Y (2012) Emerging nanotechnologies in dentistry, processes, materials and applications. Elsevier Inc, Norwich

    Google Scholar 

  12. Jansen F, Machonkin MA, Kuhman DE (1990) The deposition of diamond films by filament techniques. J Vac Sci Technol 8:3785–3790. https://doi.org/10.1116/1.576494

    Article  Google Scholar 

  13. Amaral M (2013) Diamond based materials for biomedical applications. Woodhead, USA

    Google Scholar 

  14. Shenderova OA, Gruen DM (2012) Ultrananocrystalline diamond: synthesis, properties and applications. Elsevier Inc, Britain

    Google Scholar 

  15. Olivas-Martinez M, Pérez-Tello M, Cabanillas-Lopez R, Contreras-Lopez O, Soto-Herrera G, Castillon-Barraza F (2007) A computational model for the hot-filament chemical vapour deposition process to produce diamond films. Simul Mater Sci Eng 15:237–261. https://doi.org/10.1088/0965-0393/15/3/004

    Article  Google Scholar 

  16. Goodwin DG, Gavillet GG (1990) Numerical modeling of the filament-assisted diamond growth environment. J Appl Phys 68:6393–6400. https://doi.org/10.1063/1.346858

    Article  Google Scholar 

  17. Debroy T, Tankala K, Yarbrough WA, Messier R (1990) Role of heat transfer and fluid flow in the chemical vapor deposition of diamond. J Appl Phys 68:2424–2432. https://doi.org/10.1063/1.346502

    Article  Google Scholar 

  18. Kondoh E, Ohta T, Mitomo T, Ohtsuka K (1992) Experimental and calculational study on diamond growth by an advanced hot filament chemical vapor deposition method. J Appl Phys 72:705–711. https://doi.org/10.1063/1.351856

    Article  Google Scholar 

  19. Kondoh E, Ohta T, Mitomo T, Ohtsuka K (1999) Surface reaction kinetics of gas-phase diamond growth. J Appl Phys 73:3041–3046. https://doi.org/10.1063/1.353011

    Article  Google Scholar 

  20. May PW, Harvey JN, Smith JA, Mankelevich YUA (2006) Reevaluation of the mechanism for ultrananocrystalline. Diamond deposition from Ar∕CH4∕H2 gas mixtures. J Appl Phys 99:104907. https://doi.org/10.1063/1.2195347

    Article  Google Scholar 

  21. Frenklach M, Wang H (1991) Detailed surface and gas-phase chemical kinetics of diamond deposition. Phys Rev B 43:1520–1545. https://doi.org/10.1103/PhysRevB.43.152

    Article  Google Scholar 

  22. McMaster MC, Hsu WL, Coltrin ME, Dandy DS (1994) Experimental measurements and numerical simulations of the gas composition in a hot-filament-assisted diamond chemical-vapor-deposition reactor. J Appl Phys 76:7567–7577. https://doi.org/10.1063/1.358525

    Article  Google Scholar 

  23. Mankelevich YA, Rakhimov AT, Suetin NV (1996) Two-dimensional simulation of a hot-filament chemical vapor deposition reactor. Diam Relat Mater 5:888–894. https://doi.org/10.1016/0925-9635(95)00493-9

    Article  Google Scholar 

  24. Mankelevich YA, Rakhimov AT, Suetin NV (1998) Three-dimensional simulation of a HFCVD reactor. Diam Relat Mater 7:1133–1137. https://doi.org/10.1016/S0925-9635(98)00163-0

    Article  Google Scholar 

  25. Wolden C, Gleason KK (1993) Heterogeneous formation of atomic hydrogen in hot-filament diamond deposition. J Appl Phys Lett 62:2329–2331. https://doi.org/10.1063/1.109407

    Article  Google Scholar 

  26. Zuo W, Shen B, Sun FH, Chen M (2008) Simulation of substrate temperature distribution in diamond films growth on cemented carbide inserts by hot filament CVD. Appl Mech Mater 10-12:864–868. https://doi.org/10.4028/www.scientific.net/AMM.10-12.864

    Article  Google Scholar 

  27. Cheng L, Zhang J, Wang X, Zhang T, Shen B, Sun F (2014) Simulation of temperature distribution in HFCVD diamond films growth on WC-Co drill tools in large quantities. Key Eng Mater 589-590:399–590. https://doi.org/10.4028/www.scientific.net/KEM.589-590.399

    Article  Google Scholar 

  28. Tankala K, DebRoy T (1993) Transport phenomena in the scale-up of hot filament-assisted chemical vapor deposition of diamond. Surf Coat Technol 62:349–355. https://doi.org/10.1016/0257-8972(93)90266-Q

    Article  Google Scholar 

  29. Song GH, Sun C, Huang RF, Wen LS (2000) Simulation of the influence of the filament arrangement on the gas phase during hot filament chemical vapor deposition of diamond films. J Vac Sci Technol A 18:860–863. https://doi.org/10.1116/1.582267

    Article  Google Scholar 

  30. Song GH, Sun C, Huang RF, Wen LS (1999) The simulation of the space fields of gas physical parameters during HFCVD diamond films. Acta Metall Sin 35(6):648–653

    Google Scholar 

  31. Wang AY, Lee KR, Sun C, Wen LS (2006) Simulations of the dependence of gas physical parameters on deposition variables during HFCVD diamond films. J Mater Sci Technol 22(5):559–604

    Article  Google Scholar 

  32. Menningen KL, Childs MA, Anderson LW, Lawler JE (1996) Gas temperature in a hot filament diamond chemical vapor deposition system. Rev Sci Instrum 67:1546–1554. https://doi.org/10.1063/1.1146886

    Article  Google Scholar 

  33. Park S (1999) Numerical modeling of diamond growth environment in HFCVD reactors. Met Mater 5:225–229. https://doi.org/10.1007/BF03026071

    Article  Google Scholar 

  34. Eckert M, Neyts E, Bogaerts A (2008) Molecular dynamics simulations of the sticking and etch behavior of various growth species of (ultra) nanocrystalline diamond films. Chem Vap Depos 14:213–223. https://doi.org/10.1002/cvde.200706657

    Article  Google Scholar 

  35. Tao Z, Yizheng Q, Shu W, Guodong H, Lijun Z, Zhe X (2019) Influence of the heat dissipation mode of long-flute cutting tools on temperature distribution during HFCVD diamond films. Crystals 9:394. https://doi.org/10.3390/cryst9080394

    Article  Google Scholar 

  36. Rebrov A, Plotnikov M, Mankelevich Y, Yudin I (2018) Analysis of flows by deposition of diamond-like structures. Phys Fluids 30:016106. https://doi.org/10.1063/1.4996067

    Article  Google Scholar 

  37. Gorbachev YE, Leshchev DV, Plotnikov MY, Rebrov AK, Yudin B (2019) Comparison of continual and molecular modeling of gas flow for diamond deposition. Conf Ser 1382:012152. https://doi.org/10.1088/1742-6596/1382/1/012152

    Article  Google Scholar 

  38. Joseph L (1998) Le génie chimique à l’usage des chimistes. Librairie Tho, Paris

    Google Scholar 

  39. Schäfer L, Höfer M, Kröger R (2006) The versatility of hot-filament activated chemical vapor deposition. Thin Solid Films 515:1017–1024. https://doi.org/10.1016/j.tsf.2006.07.073

    Article  Google Scholar 

  40. Koutlas GN, Vlachos NS (2003) Numerical modeling of pyrolytic laser-induced chemical vapor deposition. J Appl Phys 93:3049–3056. https://doi.org/10.1063/1.1543230

    Article  Google Scholar 

  41. Strunin VI, Lyakhov AA, KhudaÏbergenov GZH, Shkurkin VV (2002) Numerical simulation of silane decomposition in an RF plasma. Tech Phys 47:760–766. https://doi.org/10.1134/1.1486201

    Article  Google Scholar 

  42. Perrin J, Leroy O, Bordage MC (1996) Cross-sections, rate constants and transport coefficients in silane plasma chemistry. Contrib Plasma Phys 36:3–49. https://doi.org/10.1002/ctpp.2150360102

    Article  Google Scholar 

  43. Babahani O, Khelfaoui F (2012) Calcul des concentrations de molécules et de radicaux lors de déposition de couches minces a-Si:H par procédés PECVD. Ann Sci Technol 4(2):115–120 https://journals.univ-ouargla.dz/index.php/AST/article/view/291

    Google Scholar 

  44. Herrebout D, Bogaerts A, Yan M, Gijbels R, Goedheer W, Dekempeneer E (2001) One-dimensional fluid model for an rf methane plasma of interest in deposition of diamond-like carbon layers. J Appl Phys 90:570–579. https://doi.org/10.1063/1.1378059

    Article  Google Scholar 

  45. Versteeg HK, Malalasekera W (eds) (1995) An introduction to computational fluid dynamics the finite volume method. Longman Group Ltd, Harlow-England

    Google Scholar 

  46. Sibony M, Mardon JC (eds) (1984) Analyse numérique II, Approximations et équations différentielles. Editeurs des Sciences et des Arts, Paris

    MATH  Google Scholar 

  47. Barrand JP, Sacadura JF (eds) (1993) Initiation aux transferts thermiques. CAST, Lyon

    Google Scholar 

  48. Pu B, Yuan (2009) The alternate iterative Gauss-Seidel method for linear systems. Conf Ser 1411:012008. https://doi.org/10.1088/1742-6596/1411/1/012008

    Article  Google Scholar 

  49. Wang AY, Sun C, Cao HA, Ji AL, Huang RF, Wen LS (2004) Two-dimensional simulations of temperature fields of the reactor wall during hot-filament CVD diamond film growth over a large area. Model Simul Mater Sci Eng 12:325–335. https://doi.org/10.1088/0965-0393/12/2/012

    Article  Google Scholar 

  50. Von Keudell A, Meier M, Hopf C (2002) Growth mechanism of amorphous hydrogenated carbon. Diam Relat Mater 11:969–975. https://doi.org/10.1016/S0925-9635(01)00553-2

    Article  Google Scholar 

  51. Moisan M, Pelletier J (eds) (2006) Physique des plasmas collisionnels application aux décharges haute fréquence. EDP Sciences, France

    Google Scholar 

  52. Kolasinski KW (ed) (2012) Surface science: foundations of catalysis and nanoscience. Wiley, USA

    Google Scholar 

  53. Persad AH, Ward CA (2016) Expressions for the evaporation and condensation coefficients in the Hertz-Knudsen relation. Chem Rev 116:7727–7776. https://doi.org/10.1021/acs.chemrev.5b00511

    Article  Google Scholar 

Download references

Acknowledgements

Our acknowledgments go to the responsible in charge of the Computer Resources Center (CRI) of Kasdi Merbah Ouargla University (UKMO) where part of the calculations were made. The authors acknowledge also the Directorate General of Scientific Research and Technological Development (DGRSDT) and the Thematic Science and Technology Research Agency (ATRST) who have given their support to the Laboratory LRPPS.

Funding

No funding was obtained for this study.

Author information

Authors and Affiliations

Authors

Contributions

FK proposed the subject of this work which is the principal axis NK’s doctoral thesis under the supervision of FK. The modelization, the interpretation, and the redaction of the manuscript were done by NK, FK, and OB. The numerical methods and the calculations are realized by NK. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Fethi Khelfaoui.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Khelef, N., Khelfaoui, F. & Babahani, O. Numerical calculation of temperature, velocity, and species concentration distributions in a hot filament chemical vapor deposition in a reactor using a CH4/H2 mixture. J. Eng. Appl. Sci. 69, 73 (2022). https://doi.org/10.1186/s44147-022-00121-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s44147-022-00121-4

Keywords