Anda di halaman 1dari 24

Plasma Chemistry and Plasma Processing, Vol. 17, No.

4, 1997

Modeling of a DC Plasma Torch in Laminar and


Turbulent Flow
J. M. Bauchire,1 J. J. Gonzalez,1 and A. Gleizes1
Received December 2, 1996; revised March 11, 1997

A mathematical 2D representation is developed describing the temperature and the


velocity profiles in a DC plasma torch and in the resulting plume. It is based on
the resolution of conservation equations using the Simple method after Patankar.
In the first part, we illustrate the effects of the turbulence, using, on the one hand,
two Prandtl's mixing length models and, on the other hand, a standard k - e model.
We also show the influence of physical parameters like the inlet mass flow rate, the
current intensity, and the kind of gas (argon or air) on the characteristics of the
plasma. The second part of this study presents a comparison of the model with
experimental results encountered in the literature. The profiles obtained at the exit
of the torch are compared to the mathematical formulation used as boundary condi-
tion by the models taking into account only the plasma jet.

KEY WORDS: Plasma torch; modeling; laminar and turbulent flow.

1. INTRODUCTION
In order to provide high quality for specialized applications and to aid
in the development of the technology, many studies of electric arcs, especially
nontransferred arcs, have been made. Comparison between experimental
results and the models leads to refinement of the calculation by taking
increasingly sophisticated phenomena into account.
Many of these studies were only made on the plasma jet. We can men-
tion for example the measurements of Capetti and Pfender(1) and Dilawari
et al.(2) Other works by McKelliget et al.(3) and Chang and Ramshaw(4) have
treated numerical simulations of argon plasma jets flowing into cold air and
nitrogen respectively. For all these applications the behavior of the plasma
is essentially studied through turbulence like in the paper of Huang et al.(5)
concerning a two-fluid model of turbulence which can predict phenomena
that escape more conventional models, e.g., unmixing calculations. These
models of the free plasma jet require adjustable data for the temperature
1
Centre de Physique des Plasmas et de leurs Applications de Toulouse, E.S.A. No. 5002,
Universite Paul Sabatier, 118 route de Narbonne, 31062 Toulouse Cedex 4, France.
409
0272-4324/97/1200-0409$12.50/0 © 1997 Plenum Publishing Corporation
410 Bauchire, Gonzalez, and Gleizes

and the velocity profiles, such as upstream boundary conditions at the exit
of the torch.
In order to avoid having to make these assumptions some authors
included the arc region in their modeling. Westhoff and Szekely,(6) for
example, computed a nontransferred arc plasma torch and the resultant
plume in laminar flow, assuming that the cathode tip was flat rather than
pointed. More recently, Murphy and Kovitya(7) published a model consider-
ing gas mixing and the effects of turbulence by a k — e model, omitting from
the energy equation the term used to account for energy transfer due to
electron flow. Paik et al.(8) studied the determination of the arc-root position
in a DC plasma torch in laminar flow, but this model was not coupled with
modeling of the free plasma jet.
The present paper has two main objectives:
First, to calculate the influence of turbulence models on the theoretical
predictions. We present a study consisting of the application of three turbu-
lence models using two Prandtl's mixing length models and also a standard
k— e model. The work shows the turbulence effects and gives a better idea
of the utility of each model. It brings out the difficulty of using a turbulence
model, without adjustment, on experimental measurements.
Second, to study the influence of physical parameters on the tempera-
ture and the velocity profiles in laminar or turbulent flow. In particular, a
comparison is made between the properties of the plasma with argon and
air as plasma gases. These two points were pursued studying a simplified
geometry corresponding to our experimental configuration so as to compare,
in future works, predicted and measured values.
In order to validate our model and to show its level of predictivity, we
will present, in the last part, a comparison between our calculated results
and the experimental data previously obtained by Westhoff and Szekely.(6)
For this comparison numerous changes were made to our model in order
to exactly respect both the experimental geometry given by the authors and
their flow conditions.

2. MODELING
2.1. Assumptions and Governing Equations
The model adopted is based on the following assumptions:
The governing equations are written in an axisymmetric system of coor-
dinates and the operation of the torch is assumed to be in a steady state
with negligible gravity effects.
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 411

Table I. Terms of the Governing Equations


Conservation of P rp sp
Mass 1 0 0

Axial momentum u ue
Radial momentum V ue
Energy h

Turbulent kinetic energy k G-pe

Turbulent dissipation rate e

The plasma is assumed to be in local thermodynamic equilibrium (LTE)


at atmospheric pressure and the radiation effects are considered through the
net emission coefficient.
This two-dimensional model does not take into account fluctuations of
the arc. We impose a zero potential on a section of the nozzle, so we consider
the arc centered on the axis.
Based on these assumptions, solutions are sought for a set of elliptical
partial differential transport equations that all have the same form:

where P is the general variable, GP the corresponding diffusion coefficient,


and SP the source term; u and v are the axial and radial velocity components,
p is the mass density, and x and r are the distances in the axial and radial
directions respectively. The equations were solved using Patankar's
algorithms(9); the different terms are given in Table I with p as the local
pressure, h the specific enthalpy, k the turbulent kinetic energy, and s the
turbulent dissipation rate. ue, u, ut, Cp, K, and a are respectively the
effective, laminar and turbulent viscosities, the specific heat, and the thermal
and electrical conductivities. No turbulent fluctuations in the density or
electromagnetic parameters are considered in this paper. This hypothesis
avoids introducing a separate equation for the temperature variance and its
dissipation rate. So, for the calculation of the electromagnetic parameters
we used the mean temperature of the plasma. kB is Boltzmann's constant, e
is the elementary charge of the electron, jr and jx are the radial and axial
current density components, BT is the azimuthal component of the magnetic
412 Bauchire, Gonzalez, and Gleizes

field, and eN is the net emission coefficient of the plasma. The self-induced
magnetic field BT may be written as

where u0 is the permeability of vacuum. Current continuity is solved in terms


of the electric potential V:

2.2. Calculation Domain and Boundary Conditions


Here we present only the calculation domain and the boundary condi-
tions used in Sections 3 and 4 for the study of the influence of the choice
of turbulence model and the parametric study. The calculation region is
sketched in Fig. 1 and the corresponding boundary conditions are given in
Table II. Axisymmetric conditions are set on the centerline. The boundary
conditions used for the conservation equation of the electric potential are
set on lines EE' and AA'. On line EE', the electric potential is set to zero
which is a boundary condition for a porous or a fictitious anode. The line
EE' is represented at the exit of the torch in Fig. 1 but it can be anywhere
(perpendicular to the axis) in the nozzle in order to adapt the known length

Fig. 1. Computation domain.


Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 413

Table II. Boundary Conditions


u V T k E

AB v =0

BC

CD u=0 T= 500 K

DE u=0 v =0 T=1000K k =0 e =0
EF u=0 v=0 T=1000K k = kw £=£w

FG u = u(r) v =0 T= 300 K — —
GA u=0 v=0 T= 3000 K k=0 e =0

of the arc or the voltage between the electrodes. On line AA', the current
density is given by an exponential function after Hsu et al.(10) The computa-
tion domain used for potential resolution is AA'EE'. Knowing the current
density distribution on line AA' and the zero potential (on line EE'), we
can use an iterative scheme to obtain the potential field. The total voltage
calculated does not take into account the potential fall in front of the
cathode. This latter value, corresponding to the distance between line AA'
and the cathode tip, has to be added to calculate the total input power. This
point will be discussed in Section 5. For the inlet velocity profile u(r) three
initial profiles were used on line FG: a parabolic, a constant, and a Couette's
profile. Our calculations showed that the difference between the three profiles
diminishes as the power of the arc becomes greater. In all cases the maximum
difference for the velocity in the free plasma jet is about 5% over the range
50-400 A for the current intensity and 0.1-1.2 g/s for the inlet mass flow
rate. In the following results u(r) was chosen as a parabolic profile.
The operating and the ambient gases were both argon or both air.
Thermodynamic and transport properties for the argon plasma are taken
from Mostaghimi(11) and Bacri and Raffanel(12) for air. The values of the
net emission coefficient assume a homogeneous plasma column radius of
5 mm and are taken from the works of Gleizes et al.(13) The net emission
coefficient for air is assumed to be equal to that of pure nitrogen.
A preliminary study was made of the influence of the radial and axial
dimensions over the computational domain. For this comparison mesh grids
35 x 85 (5 cm) and 35 x 167 (10 cm) were used as the axial distance and grids
35 x 85 (1.5 cm) and 42 x 85 (3 cm) for the effect of the radius. The good
agreement between the values obtained with the different grid sizes led us
to use the 35 x 85 grid.
414 Bauchire, Gonzalez, and Gleizcs

2.3. Turbulence Models


The turbulence of the flow is represented through the effective viscosity
which is the sum of the molecular and the turbulent viscosities. It is very
difficult to determine the most realistic turbulence model. The complexity
of the phenomenon and the fact that few experimental data on the develop-
ment of turbulence in a plasma jet are available led to the use of models of
varying complexity. The time calculation, the mathematical formulation,
and the lack of knowledge concerning the choice of boundary conditions
are often the only criteria. In order to study the turbulence phenomena and
their application in the model, we present and compare three turbulence
models generally used in plasma jet studies. The temperature and velocity
profiles obtained were compared assuming laminar flow.

2.3.1. The Zero-Equation Model


This model is based on the Prandtl's mixing length hypothesis and gives
the turbulent viscosity as

Based on Nicolet's work,(14) we adopted the following expressions of mixing


length lm and turbulent Prandtl's number Pr T :

where R is the pipe radius, y is the distance from the pipe wall as y =
R-r + rg, rg is the wall roughness equal to 0.01 mm; yc is obtained from
the continuity of lm and quantities with an asterisk denote values taken at
the wall. In the plume region, two cases are considered:
—Case I, the mixing length and the Prandtl's number expressions are
taken from the work of Gonzalez,(15) i.e., Eq. (6) for the mixing length and
Eq. (7) if r<R and Pr T =0.95 if r>R.
—Case II, a standard expression of the mixing length for a free jet is
used(16):
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 415

r1 and r2 are two radial positions chosen such that the velocity is equal to
0.15 and 0.85 times the centerline velocity respectively. C is a constant that
can be determined by experimental studies. Here C is made equal to 0.8 to
be in agreement with case I.

2.3.2. The Two-Equation Model


The turbulent viscosity can be obtained by a k— e model(7,17-19) with

The turbulent kinetic energy and the dissipation rate are calculated via the
transport equations shown in Table I with

The values of the constants are given in Table III with modifications for
free jets as suggested by Rodi(20):

where 8 is the boundary layer thickness, Au is the velocity difference between


the centerline and the free stream velocities, and u0 is the centerline velocity.
We used a wall function from Rodi(20) to calculate the values kw and ew at
the points neighboring the anode along the line EF. The wall-function
method was used for all the results presented with the k— e model. This
method was not applied along the line DE because of the small values of
the radial velocity. If y is the distance from the wall, K the von Karman
constant (K= 0.435), and E the friction parameter (E f =9), the resulting
friction velocity ur is determined by the wall law:

Table III. Constant Values of the k- e Model

Ce1 Cc2 cu PrT Prk Pr£


1.44 1.92 0.09 0.7 1.0 1.3
416 Bauchire, Gonzalez, and Gleizes

uw is the axial velocity at the first grid points away from the wall and the
values of kw and £w are deduced from the following expressions:

3. INFLUENCE OF TURBULENCE MODELS


3.1. Aims
The studies of free plasma jets indicate the location of a mixing zone,
that is, a boundary layer separating the plasma jet and the ambient gas. In
this region, the transfer of matter is not only caused by diffusion but also
by vortex rings which create a pumping of surrounding cold gas into the
center of the jet(21) leading to a decrease of the temperature. Taking turbu-
lence effects into account is now widespread in the majority of mathematical
models of plasma torches. Although the most used of these models is the
standard k— e one (which gives good results in a lot of industrial cases),
simpler analytical models like those using Prandtl's mixing length also exist.
The aim of this study was not to show the well-known turbulence effects
but to make a qualitative analysis of three turbulence models in order to
determine the advantages of each one and to see their influence on the
plasma parameters. The following calculations were done with the arbitrary
computational domain sketched in Fig. 1.

3.2. Turbulence in the Plume


Most plasma torch models only consider the plasma jet, taking arbitrary
temperature and velocity profiles as boundary conditions at the nozzle exit.
First, we assumed that flow is laminar in the nozzle and turbulent in the jet.
We have plotted, in Figs. 2 and 3, the temperature and axial velocity
variations on the centerline for an inlet mass flow rate of 1.0 g/s and a
current intensity of 200 A. The position x = 0 means the inlet section, i.e.,
0<x<9.2mm for the arc region and x>9.2mm for the jet region. The
boundary conditions on k and £ are fixed on line EE'(22) as
kEE' = 0.002u 2 EE' (r) and E EE' (r) = CuK3/2/0.03R. Turbulence increases the
exchanges between the plasma jet and the ambient gas, so an increase results
in the total mass flow rate and a decrease in the axial velocity and tempera-
ture profiles.
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 417

Fig. 2. Comparison of laminar and turbulence models in the plume: temperature profiles on
centerline, D0= 1.0 g/s and I=200 A.

The three turbulence models give very similar results for temperature
and axial velocity. We note that velocities are different in the downstream
part of the nozzle not far from the tip, although the flow is supposed to be
laminar in this zone. This is due to the elliptical character of the model.
Even if the flow has a main direction, local properties can be modified by
the downstream values. We can also note that the k — e model and case II
of the mixing-length model lead, in comparison to case I, to a faster decrease
of velocity and temperature in the last part of the jet (for x>3 cm). This
comes from the fact that for case I, the mixing length lm is not proportional
to the width of the mixing boundary layer.
The calculation times for the k- e model are longer than those for the
mixing length models: for the 35 x 85 mesh grid and 3500 iterations, the
calculation takes 1231 CPU seconds for a laminar flow, 1265 s for case I,
1266 s for case II, and 1565 s for k— s on a HP 750. So, these three models
are quite equivalent under these operating conditions although the k- E
model requires the solution of two additional conservation equations but
gives more information on turbulence in comparison with the mixing length
models. This detailed information may be necessary for other purposes. For
example, values of the local turbulence intensity or the kinetic energy are
418 Bauchire, Gonzalez, and Gleizes

Fig. 3. Comparison of laminar and turbulence models in the plume: axial velocity profiles on
centerline, D0= 1-0 g/s and I=200 A.

necessary to determine the turbulent dispersion of particles, which is impor-


tant in plasma spraying. Comparable conclusions were obtained by
Proulx(16) in previous works on plasma jets.

3.3. Turbulence in the Whole Domain

For the nozzle region, we used the mixing length model described in
Section 2.3.1 for case II in the plume region. We also made the calculation
with the k - s model over the whole computation domain with a wall func-
tion (cf. Section 2.3.2) to take into account the influence of the wall.
The temperature and axial velocity variations on the centerline are plot-
ted in Figs. 4 and 5. The boundary conditions on k and e are fixed on line
FG as kFG(r) = 0.003u2G(r) and eFG(r) = C u k 3/2 /0.03R.
We note a critical difference between the velocity profiles (Fig. 5), which
can be explained in two ways. First, the Prandtl's mixing length model was
proposed by Nicolet et al. for a high-enthalpy constricted arc heater with a
Reynolds number of 2 • 106. Turbulence is surely not as great in our case
(Re = 3000 at the nozzle exit), so the use of this model may strongly over-
estimate the turbulence intensity and thus underestimate the velocity.
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 419

Fig. 4. Influence of turbulence models in the torch and in the jet: temperature profiles on
centerline, D0= 1.0 g/s and I=200 A.

The second point is the strong dependence of the k — s model on the


boundary conditions of k and e. This problem, discussed by Dilawari and
Szekely,(23) led us to test different values of k and e. The results obtained
were intermediate between those of the k — e model and the mixing length
model shown in Figs. 4 and 5. Thus, care must be taken when using a
turbulence model within the nozzle region and it must be coupled with an
experimental study giving information on the turbulence intensity.

4. PARAMETRIC STUDY
The influence of the inlet mass flow rate and of the arc intensity on the
temperature and velocity profiles in the nozzle and in the free plasma jet
were studied considering laminar and turbulent (k— e model with the wall
function) flows.

4.1. Influence of the Inlet Mass Flow Rate


In Fig. 6 we have plotted the isotherms for a current intensity of 100 A.
In the left part, the inlet mass flow rate is D0 = 0.5 g/s, and on the right
420 Bauchire, Gonzalez, and Gleizes

Fig. 5. Influence of turbulence models in the torch and in the jet: axial velocity profiles on
centerline, D0= 1.0 g/s and I=200 A.

D0 = 1.0 g/s. The notation 2 -> 18 kK means that the minimum isotherm is
2000 K, the maximum isotherm is 18,000 K, and the interval is 2000 K. This
figure shows that an increase of the inlet mass flow rate leads to a constriction
of the plasma column and to an increase of the axial temperature in the
nozzle and in the upstream part of the jet. The properties of the plume are
determined by those of the arc, in particular by those at the exit of the torch.
Therefore, when D0 rises, energy losses increase (because of convection and
turbulence) and hence, the arc voltage rises: for I— 100 A and a turbulent
flow (case I), U= 16.78V when D0 = 0.5 g/s and U= 18.06V when D0 =
1.0 g/s. For a given section, the mean electric field increases with the losses.
So, the conductance G decreases leading to the contraction of the plasma
column and to the rise of the axis temperature. The downstream part of the
jet is under the influence of mixing with cold ambient gas. In this region
turbulence effects are strong and lead to a temperature decrease.

4.2. Influence of the Current Intensity


Figure 7 plots the total mass flow rate vs. axial position for two current
intensities: 100 and 200 A. We can see in the plume region, 1 cm from the
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 421

Fig. 6. Influence of the inlet mass flow rate on isotherms, /= 100 A and turbulent flow
(k— e model).

nozzle tip, a rise in the mass flow rate due to the cold gas drag. The quantita-
tive results are in good agreement with the works of Pateyron,(24) i.e., for
an axial distance of four times the diameter, the total mass flow rate D has
a value defined by the relationship:

where D0 is the inlet mass flow rate and k3 is a coefficient in the range 0.25-
0.45. We found the values k3 = 0.25 (laminar flow) and k3 = 0.33 (turbulent
422 Bauchire, Gonzalez, and Gleizes

Fig. 7. Mass flow rate vs. axial position for laminar and turbulent flow (k — e model).

flow) for a current intensity of 200 A. We can see that pumping of the
ambient gas rises with the turbulence and with the current intensity. Under
the effects of turbulence, the exchanges between the plasma jet and the
ambient gas rise, leading to a greater pumping of surrounding gas. The drag
also depends on the value of the exit velocity of the jet which rises with the
current intensity. So, the mass flow rate increases with the current intensity.
This behavior of the plasma with the current intensity was noted in experi-
mental works by Vardelle.(25)
In this paper we do not present the influence of geometric or mathemati-
cal parameters, but a preliminary study has shown that taking the position
of the open boundary CD (see Fig. 1) at R1 = 3.2 or R2 = 6.38 times the
nozzle diameter away from the axis gives rise to the same results. So, the
results are presented using R1 = 3.2. Other works on the modeling of plasma
jets(15,26) have shown that the boundary BC (see Fig. 1) should be located
at R1 equal to six times the nozzle diameter. In our case, for values of R1
of about 6 or 3 times the diameter, there are no significant differences in the
temperature profile. The main difference is encountered in the velocity pro-
files, with a difference of about 10% for the value on the centerline where
the mass density is very low.
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 423

Fig. 8. Influence of the length of the arc on the temperature field, D0 = 0.5 g/s and I=200 A.

4.3. Arc Root Position


In this paper, we assumed that the arc root position is fixed and we
imposed the zero-potential on line EE'. But, according to the works of Paik
et al.(8) and Betoule,(27) an increase of the current intensity may lead to a
decrease of the length of the arc. In order to study the influence of this
length we have plotted in Fig. 8 the isotherms of the plasma for an arc
length of l1 = 3.25 mm (zero potential in the middle of the axis cathode tip-
424 Bauchire, Gonzalez, and Gleizes

exit of the torch) and for an arc length of l2 = 2l1 (zero potential at the exit
of the torch), the operating parameters being D0 = 0.5 g/s and I= 200 A with
a laminar flow. Although the current intensity is the same, the lengthening
of the arc leads to an increase of the arc voltage and hence of the power:
Pl1 = 2.5 kW and Pl2 = 3.6 kW. This results in the important difference
between the temperature fields seen in Fig. 8.
On the other hand, Fig. 9 shows the isotherms of the plasma for current
intensities of 100 and 200 A with arc lengths of respectively l2 and l1, consid-
ering the approximate relationship(8)

where xs is the arc root position and A a constant depending on the upstream
mass density and axial velocity. The differences between the temperature
fields in Fig. 9 are lower than those obtained in Fig. 8 because the powers
were P=1.6 kW for I=100 A (l 2 ) and P = 2.5 kW for I=200 A (l 1 ). Arc
restrike phenomena are complex and difficult to compute but we can deduce
from these results that plasma properties depend more on the power value
than on the current intensity.

4.4. Comparison of Argon and Air


Figure 10 plots the temperature fields in a laminar flow with air on the
right and argon on the left. The boundary conditions are the same for the
two gases, the inlet mass flow rate is D0 = 0.5 g/s, and the power of the torch
is 1.9kW.
It can be noted that the temperature is higher using argon. The differ-
ence between the net emission coefficients of argon and air cannot explain
the difference of the temperature profiles even if the net emission coefficient
of air is four times greater than that of argon for temperatures around
10,000 K. The calculation using air with the net emission coefficient of argon
leads to the same result. This can be explained by the specific heat which is
much higher for air than for argon. Indeed, with a molecular gas, the energy
can reside in the vibrational or the rotational states. So, for a given energy,
the temperature is lower for a molecular gas which stores the energy differ-
ently. Nevertheless, the thermal conductivity is greater in air (10-fold that
of argon at around 5000 K) and favors heat transfers leading to a diminution
of the axial temperature. A view of these losses in the plasma is given by
the voltage drop in the arc column between the two electrodes. In our model
we did not consider the voltage drop in the electrodes or electrode regions.
We obtained U= 23.74 V (100 A) and U= 17.69V when I = 145 A for air
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 425

Fig. 9. Influence of the product arc length x current intensity on isotherms, D0 = 0.5 g/s.

and argon respectively. It can also be noted that using air leads to a constric-
tion of the plasma jet and to an increase of the mass flow rate: k3 is equal
to 0.23 for argon and 0.25 for air because of the pumping of the surrounding
cold gas.
Figure 11 shows the temperature fields for the same current intensity
I= 100 A. Although powers are different, Pair = 2.3 kW and Pargon = 1.6 kW,
we find the same kind of results than those of Fig. 10 on the temperature
fields because of the strong difference between argon and air properties.
426 Bauchire, Gonzalez, and Gleizes

Fig. 10. Isotherms of argon and air plasmas, D0 = 0.5 g/s and P= 1.9 kW.

5. COMPARISON WITH EXPERIMENTAL DATA


In the literature a good effort was made in the work of Westhoff and
Szekely(6) in the discussion and in the comparison between experimental
results and their model. The experimental system, the length of the computa-
tional domain, and the assumptions are clearly given by the authors. Thus,
in this section we have modified our model, and a comparison is made with
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 427

Fig. 11. Isotherms of argon and air plasmas, D0 = 0.5 g/s and I= 100 A.

their experimental results. Two kinds of changes have been made for this
comparison:

—changes to adapt the dimensions and the geometry. The computa-


tional domain and the geometry are those described by Westhoff and
Szekely(6) and are not reported here.
—changes on the parametric values and the boundary conditions.
428 Bauchire, Gonzalez, and Gleizes

5.1. Assumptions and Boundary Conditions


—At the inlet boundary FG the axial velocity is assumed to have a
parabolic profile with its maximum halfway between the cathode and the
anode wall.
—The temperature of the anode is T= 1000 K.
—We use argon as inlet gas.
—In our model injection was not azimuthal, so the comparison will be
made with Sw = 0 (experimental case No. B23(6)), where Sw is the ratio of
the axial flux of the azimuthal momentum to the axial flux of the axial
momentum.
—Westhoff and Szekely(6) assumed that the flow is laminar considering
the mean Reynolds number. The values at the nozzle exit are in the range
100-200. So, our calculation was performed without using a turbulence
model.
—They assumed a constant value of the potential on the anode wall.
This results in a distribution of the current density at the anode surface and
in the definition of an effective arc length L by taking the axial position of
the centroid of the current distribution curves. In our model arc attachment
is done on a fictitious anode, i.e., on line EE' (Fig. 1). So, we do not have
to deal with the difficult problem of the current transition to the anode
which would require a 3D approach and a two-temperature plasma descrip-
tion. This hypothesis allows us to avoid the influence of the cold flow
temperature near the wall leading to a very small electrical conductivity. In
our case, the input power is the sum of the power lost at the cathode and
the power required for heating the gas. The power lost at the cathode
is calculated assuming a cathode voltage fall of 4.3 V.(6) As we know the
experimental input power, we can deduce the power required for gas heating.
Then the arc length (line AA'-line EE') is adjusted to obtain this required
power.

5.2. Comparison
For this given geometry, we compare the experimental axis temperature
at the exit of the torch with our model, for /= 250 A and an inlet mass flow
rate D0 = 0.29 g/s (0.59 scmh) of argon. Following the procedure described
at the end of Section 5.1, we obtain an arc length Larc= 12.25 mm, for which
the computed total torch power is approximately equal to the experimental
one (case No. B23). Figure 12 plots the temperature profile on the centerline
in the jet as obtained from Westhoff and Szekely's measurements and from
our calculations with the arc length Larc. The figure shows that the values
predicted by our model are higher than the experimental ones over a distance
of roughly 2 cm. This difference can be explained by our conditions at the
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 429

Fig. 12. Comparison between the measurements of Westhoff and Szekely (B23) and our calcula-
tion of the temperature profile on the centerline in the jet. x = 0 is the nozzle tip location.

fictitious anode which lead to a source term that is greater on the axis than
in Westhoff's conditions. In Fig. 13 we have plotted the radial temperature
profile at 37 mm from the nozzle tip obtained from Westhoff and Szekely's
measurements and from our calculations with the arc length Larc. The previ-
ous difference is seen to vanish after an axial position of the order of 2 cm
from the exit, and a good agreement occurs between the measurements and
our calculation. Near the exit of the nozzle, the overestimation of the axis
temperature leads to a slight overestimation of the axial velocity, as can be
seen in Fig. 14 which shows the axial velocity profiles at 1 mm from the
nozzle tip.
We tried to compare this temperature and the velocity profiles with
the mathematical formulation used as boundary conditions in the models
simulating only the free jet region.
More specifically, the radial temperature profile is often given by(2)

where Tmax is the maximum temperature, R0 the internal radius of the exit
430 Bauchire, Gonzalez, and Gleizes

Fig. 13. Comparison between the measurements of Westhoff and Szekely (B23) and our calcula-
tion of the radial temperature profile at 37 mm from the nozzle tip.

of the torch, and Tw the wall temperature. In our case, T max = 12,062 K,
T w =1000K, and R0 = 6.35 mm. There is a very good agreement with
2.1 <n<2.2 between Eq. (17) and the temperature profile obtained by the
model in laminar flow. The comparison is not so good for the velocity profile
using the following relationship:

with a maximum difference of about 30% (umax= 124 m/s). As can be seen
in Fig. 14, the radial velocity profile from Eq. (18) does not correspond to
our profiles or to the predictions of Westhoff et al. In conclusion, Eq. (17)
appears to be a good boundary condition approximation and is in agreement
with prediction models.

6. CONCLUSION
In this paper we present the study of a model of a DC plasma torch in
the nozzle and in the free jet regions, with turbulent and laminar flows.
Modeling of a DC Plasma Torch in Laminar and Turbulent Flow 431

Fig. 14. Axial velocity vs. radial position at 1 mm from the nozzle tip. Comparison between
the measurements of Westhoff and Szekely (B23) and our calculation.

We discuss the utility and the results obtained with three turbulence
models. The most important problems encountered in turbulence are deter-
mining the boundary conditions for the k — s model and adjusting the con-
stants for the Prandtl's mixing length model. Measurements are always
necessary for these calibrations. The study of these turbulence models does
not allow just one to be chosen but it leads to a better understanding of
their behavior and their advantages.
The study of physical and geometrical parameters shows that the
behavior of the plasma is coherent with the experimental and theoretical
results in the literature. The evolution of the mass flow rate is in agreement
with Pateyron's works(24) where turbulence effects or a current intensity
increase lead to greater pumping of the surrounding gas. We compared the
temperature fields for two gases, argon and air. A difference was noted in
the voltage drop and a decrease seen in temperature using air.
Using another geometry, we compared the model with experimental
measurements and theoretical results. Good agreement was encountered
with laminar flow. The temperature and velocity profiles were extracted at
the exit of the nozzle in order to compare them with the mathematical
formulation used as boundary conditions by the free jet model. The usual
432 Bauchire, Gonzalez, and Gleizes

temperature law presents a very good agreement with the model, but this is
not the case with the usual velocity law which presents a difference of about
30%.
This model will allow us to consider various perspectives including
mixing of dissimilar gases and the influence of particles injected in the jet.
We are in the process of adapting the model to other experimental conditions
now under study in our laboratory.

ACKNOWLEDGMENTS
The authors wish to thank the Conseil Regional Midi-Pyrenees for
partial support of this study under projet RECH/9300395.

REFERENCES
1. A. Capetti and E. Pfender, Plasma Chem. Plasma Process. 9, 329 (1989).
2. A. H. Dilawari, J. Szekely, J. Batdorf, R. Detering, and C. B. Shaw, Plasma Chem. Plasma
Process. 10, 321 (1990).
3. J. McKelliget, J. Szekely, M. Vardelle, and P. Fauchais, Plasma Chem. Plasma Process. 2,
31 (1982).
4. C. H. Chang and J. D. Ramshaw, Plasma Chem. Plasma Process. 13, 189 (1993).
5. P. C. Huang, J. Heberlein, and E. Pfender, Plasma Chem. Plasma Process. 15, 25 (1995).
6. R. Westhoff and J. Szekely, J. Appl. Phys. 70, 3455 (1991).
7. A. B. Murphy and P. Kovitya, J. Appl. Phys. 73, 4759 (1993).
8. S. Paik, P. C. Huang, J. Heberlein, and E. Pfender, Plasma Chem. Plasma Process. 13, 379
(1993).
9. S. V. Patankar, Numerical Heat Transfer and Fluid Flow, McGraw-Hill, New York (1980).
10. K. C. Hsu, K. Etemadi, and E. Pfender, J. Appl. Phys. 54, 3, 1293 (1983).
11. J. Mostaghimi, PhD Thesis, University of Minnesota (1982).
12. J. Bacri and S. Raffanel, Rep. 49277-86-1 and 40277-86-2, CPAT, Universite Paul Sabatier,
France (1986) and Plasma Chem. Plasma Process. 7, 59 (1987).
13. A. Gleizes, J. J. Gonzalez, B. Liani, and G. Raynal, J. Phys. D: Appl. Phys. 26, 1921
(1993).
14. W. E. Nicolet, C. E. Shepard, K. J. Clark, A. Balakrishnan, J. P. Kesselring, K. E.
Suchsland, and J. J. Reese, Acurex Corporation, AEDC.TR.75.47.
15. J. J. Gonzalez, PhD Thesis, Universite Paul Sabatier, France, No. 1100 (1992).
16. P. Proulx, PhD Thesis, Sherbrooke University, Canada (1987).
17. B. E. Launder and D. B. Spalding, Comput. Meth. Appl. Mech. Eng. 3, 269 (1974).
18. Y. P. Chyou and E. Pfender, Plasma Chem. Plasma Process. 9, 291 (1989).
19. D. A. Scott, P. Kovitya, and G. N. Haddad, J. Appl. Phys. 66, 5232 (1989).
20. W. Rodi, PhD Thesis, University of London (1972).
21. E. Pfender, J. Fincke, and R. Spores, Plasma Chem. Plasma Process. 11, 529 (1991).
22. W. M. Pun and D. B. Spalding, Rep. No. HTS/76/2, Imperial College, London (1976).
23. A. H. Dilawari and J. Szekely, Plasma Chem. Plasma Process. 7, 317 (1987).
24. B. Pateyron, PhD Thesis, No. 21-1987, Limoges University, France (1987).
25. M. Vardelle, PhD Thesis, No. 28-87, Limoges University, France, (1987).
26. J. J. Gonzalez, A. Gleizes, S. Vacquie, and P. Brunelot, Plasma Chem. Plasma Process.
13, 2, 237 (1993).
27. O. Betoule, PhD Thesis, No. 24-1994, Limoges University, France, (1994).

Anda mungkin juga menyukai