Sunteți pe pagina 1din 14

769

Thermophysics and Aeromechanics, 2013, Vol. 20, No. 6


Modelling of the behavior
of hollow ZrO
2
particles in plasma jet
with regard to their thermal expansion
*

I.P. Gulyaev and O.P. Solonenko
Khristianovich Institute of Theoretical and Applied Mechanics SB RAS,
Novosibirsk, Russia
E-mail: Gulyaev@itam.nsc.ru; solo@itam.nsc.ru
(Received June 6, 2013)
The effect of the expansion of hollow micro-spherical droplets due to their heating at their motion in a plasma jet is
considered by the example of ZrO
2
. A fairly simple model is proposed, which accounts for the variation of the droplets
size and shell thickness because of the thermal expansion of the gas cavity as well as their possible evaporation.
The conducted computations have enabled the assessment of the scale of the variation of diameter (1020 %)
and shell thickness (up to 50 %) of ZrO
2
particles under the conditions typical of the plasma treatment of powder
materials and producing the coatings. The influence of the given effect on the dynamics of particles heating and accele-
ration is investigated, and a comparative analysis of the behavior of hollow and dense particles in plasma jet is done.
Key words: thermal plasma, plasma treatment of powders, zirconium dioxide, hollow microsphere, thermal
expansion.
Introduction
The powders consisting of hollow spheres of size of dozens of microns (hollow powders)
are used successfully at present in various industry branches. This is, first of all, the production
of composite constructional materials, catalysts, absorbents, encapsulating and gas-sepa-
rating media, etc. The hollow zirconia powders are used in the thermal spray technology for
production of thermal barrier coatings (TBC) on the blades of aviation and rocket engines, gas
turbines of electric power plants, in which an increase in the inlet gas temperature and the work
temperature in combustion chamber are the main techniques for increasing the efficiency.
A high melting temperature, the absence of phase transitions in operating temperature range,
high adhesion and chemical inactivity, low thermal conductivity and agglomeration velo-
city of the porous microstructure [1] are the most important demands made for such coatings.
Zirconia stabilized by 78 % yttrium (YSZ) has gained by now the most widespread acceptance
owing to a combination of high material stability and strength.
To obtain the heat-resistant YSZ coatings one employs to date the powders obtained
by crashing of the melted compact, agglomeration and sintering of fine grains of the material, or
by a preliminary thermal treatment leading to the formation of hollow spheres (HOSP) [2, 3]. Many

*
The work was financially supported by the Russian Foundation for Basic Research, Project No. 12-08-31150.
I.P. Gulyaev and O.P. Solonenko, 2013
I.P. Gulyaev and O.P. Solonenko
770
authors note that the TBC obtained by using hollow powders have a lower thermal conductivity
as compared to the coatings obtained from dense particles [1, 46]. The porosity affects to
a considerable extent both the heat-protecting and strength characteristics of the coating.
It was shown in the works [2, 7] that the coatings obtained from hollow powders possess
a lower thermal conductivity at the same values of the total porosity. The reason for this is
an increased amount of micro-cavities and micro-cracks oriented in parallel with the coating
surface, which is due to a fundamentally different character of the divergence of hollow
and dense droplets at their impact onto a solid surface [8, 9]. It is these cavities mainly affect
the coating heat conduction in a direction perpendicular to the surface.
The application of hollow ceramic powders may broaden significantly the possi-
bilities of such relatively low-temperature techniques as HVOF and detonation spraying.
The difficulties of doing the YSZ coating in such technologies are due to a low temperature
of the carrying jet (of the order of 3000 ) and a low thermal conductivity of the material: one
succeeds in a complete melting of dense particles of the size of the order of 10 m [1012], and
the core of larger particles, which are usually used for coating, remains non-melted. The use of hol-
low powders with particles size of 4050 m and shell thickness of 25 m may enable a full use
of the advantages of above techniques for restoring and doing the TBC.
Despite a large number of the works devoted to the study of properties of the coatings
obtained by using hollow powders, not enough attention was paid to the questions of the nume-
rical and experimental investigation of the behavior of hollow particles in plasma jets.
The heating and acceleration of hollow ZrO
2

particles in an axisymmetric plasma jet was
modeled in the work [13]. The authors found that the hollow droplets generally accelerate
(decelerate) and heat up (cool) faster than the dense particles of the same size. The application
of hollow powders for doing coatings makes it possible, according to calculations, to increase
the coefficient of the material use at short spray distances by 50 %.
A numerical investigation of the dependence of the trajectory, velocity, and temperature
of hollow and dense ZrO
2
particles in plasma jet was conducted in the work [14]. To describe
the heat exchange inside the particle the authors have separated it into spherical layers, which
has enabled them to account for a partial melting and evaporation of the droplet. As the esti-
mates of the Weber number show, the assumption employed by the authors that the gas cavity
breaks down at a complete melting and the particle takes the shape of a dense microsphere
of the same mass, however, has a particular character. When using a model jet profile
the computations of the velocity and temperature of particles showed a fair agreement with expe-
rimental data of other authors. Significant differences in the paths of hollow and dense par-
ticles, a more fast and complete melting of hollow microspheres, and lesser non-uniformity
of temperature within their volume are also noted in the work [14].
The question of the formation of hollow spheres at the processing of agglomerated pow-
ders in plasma jet was considered in the work [15], where the examples of a practical obtaining
of hollow ceramics and metal powders were presented. These results confirm the fact that
the hollow microspheres preserve their integrity in the process of their treatment in plasma jet
and at a subsequent hardening.
The fundamental feature of hollow particles is the presence of a closed gas cavity (Fig. 1),
which is capable to change the liquid droplet size after the material melting in plasma under
the temperature variation. The given effect was mentioned for the first time in the work [16],
where it was shown that at the heating of hollow ZrO
2
particles of size 50 m up to the boiling
temperature their diameter may increase by 520 % depending on the shell initial thickness.
The present work is directed towards the study of the effect of the expansion
of hollow droplets in a plasma jet as well as towards its effect on the acceleration and heating
of particles. To this end, a sufficiently simple physical model of the variation of the size
of hollow droplets at their heating has been developed. The heating and acceleration of particles
was computed by using the known dependencies for the drag and heat transfer coefficients without
Thermophysics and Aeromechanics, 2013, Vol. 20, No. 6
771
considering the gradient heating of particles, the varia-
tion of the material density at melting, and other finer
effects, which do not affect directly the phenomenon
under study.
1. The model of the hollow droplet behavior
in plasma jet
We consider the motion of a single particle along
the plasma jet axis under the following assumptions:
the uniform temperature distribution over the droplet volume, the temperature dependence
of the droplet size, the absence of the particles feedback on plasma jet parameters.
The model operates with the mean-mass particle temperature in view of a small shell
thickness
p
, because: 1) the Biot number Bi =
p
/
p
for the ZrO
2
particles at the heating
initial stage did not exceed the values Bi 0.1, and it decreased at a further particle motion in
the jet, and the liquid microspheres were characterized by values
2
Bi 10

and less; 2) typical


times of the particle shell heating t
p
~
2
p p
/ a ~ 10
5
10
4
s and of the gas thermal relaxation
inside a hollow droplet
g
t ~ 10
7
10
6
s are much less than the typical time of the particle resi-
dence in plasma jet (
R
t ~ 10
3
s and higher).
At the interphase heat transfer computations, the radiation heat losses by the particle were
taken into account, however, under the conditions under consideration, their contribution to
the total heat exchange with gas was below 25 % (depending on the particles size and flow
characteristics), which agrees with known results [1719]. Table 1 presents the ZrO
2
properties
employed in computations.

Tabl e 1
Physical properties of ZrO
2

Density, kg/m
3

5600
Surface tension, N/m 0.43
Heat capacity (solid/liquid), J/(kgK) 755 / 811
Temperature of melting/boiling, K 2960 / 4573
Heat of melting/boiling, J/kg
8.5210
5

/ 73.6410
5

1.1. Computation of the hollow particle motion and heating in plasma jet
The computation of the particle behavior in plasma jet is based on the equations of the inter-
phase exchange by the momentum and heat. A one-dimensional problem of the particle motion
along the plasma jet symmetry axis z is considered. The particle velocity
p
, U its enthalpy
p
H ,
and temperature
p
T are determined from equations
2
p p
p d f f p f p
1
( ) ,
4 2
dU D
m C U U U U
dt

=
(1)
p p
2 2 4
p p p f p SB p p
( ) ,
dT dH
c m D T T D T
dt dt
= =

Fig. 1. Main characteristics of a hollow particle.
I.P. Gulyaev and O.P. Solonenko
772
the second item entering the heat transfer equation is responsible for particles radiation losses.
The initial conditions (t = 0) have the form:
p0
0, z =
p0
0, U =
p0 p p p0
, H c m T = where m
p
=
=
p p
3 3
p
p

[1 (1 2 ) ]
6
D
D

is the particle mass,


p0
300 K T = is the particle initial temperature.
The CarlsonHoglund [20] and RanzMarshall formulas [21] are employed, respectively, for
the drag and heat exchange coefficients:
0.687
d
24
(1 0.15Re ),
Re
C = +
( )
0.5 0.33 film
p
2 0.6Re Pr ,
D

= +
where the Reynolds number
film p f p film
Re / D U U = and the Prandtl number
Pr =
film
c
film
/
film
are calculated from gas properties at the film temperature T
film
= (T
p
+ T
f
)/2.
The temperature T
p
of the particle, its diameter D
p
, and the shell relative thickness

p
=
p p
/ D are determined uniquely from the current values of the total enthalpy H
p
of the particle
and its mass m
p
.
1.2. Determination of the particle temperature and mass
The particle temperature is calculated from the known value of the particle total enthalpy
p
. H
Four supporting values of enthalpy, which correspond to:
1 ps p pm
H c m T = is the beginning
of melting,
2 1 p pm
H H m L = + is the complete melting,
3 2 pl p pb pm
( ) H H c m T T = + is the begin-
ning of boiling (evaporation), and
4 3 p pb
H H m L = + is the complete evaporation, are first
determined for the particles, respectively. Here c
ps
and c
pl
are the heat capacities of the particle
material in solid and liquid state, T
pm
and T
pb
are, respectively, the temperatures of melting and
boiling of the material. The particle temperature T
p
is further determined by the following
formulas:
p ps p p 1
pm 1 p 2
p
pm p 2 pl p 2 p 3
pb 3 p 4
/ at ,
at ,
( ) / at ,
at .
H c m H H
T H H H
T
T H H c m H H H
T H H H
<

=

+ < <


Upon termination of each temporal iteration, the correction of the particle mass is done with
allowance for particles possible evaporation. It is assumed that at
p 3
H H the particle mass
decreases by the value
p p 3 pb
( ) / , m H H L = and the particle enthalpy decreases by H
p
=
= m
p
L
pb
. If the droplet enthalpy exceeds the H
4
value, it is assumed that the particle has evapo-
rated completely. The quantities H
1
, H
2
, H
3
, and H
4
are re-calculated after the mass correction.
1.3. Determination of the particle current size
The particle external diameter D
p
and the shell relative thickness
p
do not change until
the melting of the particle occurs in the process of its heating. A further heating of the particle
leads to the expansion of internal gas, which makes the work against the forces of surface ten-
sion and external pressure. The diameter of the gas cavity (and of the entire droplet) increases,
and the pressure therein drops. At each moment of time (at each temperature value), the bal-
ance of above pressures is satisfied. In particular, each diameter of the particle D
p0
and the shell
dimensionless thickness
p0
correspond to the balance of the Laplace pressure, the pressure
Thermophysics and Aeromechanics, 2013, Vol. 20, No. 6
773
in gas cavity P
g0
, and the ambient pressure P


at the material melting temperature T
0
= T
pm
,
that is they correspond to the conditions,
under which the hollow solid sphere was
obtained:
g0
p0 p0
4 1
1 .
1 2
P P
D


= + +


(2)
We adopt the assumption that the gas pressure and composition inside the microsphere at
the moment of its complete melting are not subjected to the variation since the moment of its
obtaining that is the solid microsphere shell remains impermeable at the powder storing.
We make use of the equations proposed in the work [2], which establish a relation be-
tween the current values of the radius R
p
and the thickness
p
of the shell of a hollow droplet
at a given gas temperature T
g

= T
p
inside the particle:
3
p g0 g p
g
3
p p p p
p g
2 3
1
1 1 0,
1 /
4
m RT
P
R R R
R M


+ + =




p p0
3
3
p
p p
3
1 1 ,
4
m
R
R

=
where m
p0
and m
g0
are the given values of the initial mass of the hollow sphere and the gas
contained inside it, M
g
is the relative molecular (atomic) mass.
To obtain the dependence of the current pressure in the gas cavity P
g
on temperature
p
T
we further use the ideal gas law and the condition for the constancy of mass of the captured
gas at the moment of the hollow particle formation:
g 0 g0 g0 0 g g p
/ / , m M P V RT MPV RT = = (3)
where
3 3
g0 p0 p0
(1 2 )
6
V D

= is the initial volume of the gas cavity,


3 3
g p p
(1 2 )
6
V D

= is
the gas cavity volume at temperature T
p
; M
0
and M are the relative molar masses of gas at
the initial and current temperature, R is the universal gas constant. On the other hand, the gas
current pressure must satisfy in the equilibrium state the equation for the balance of forces act-
ing on the liquid shell, which is similar to equation (2):
g
p p
4 1
1 .
1 2
P P
D


= + +


(4)
Using expressions (3) and (4), we obtain the following equation coupling the current
values of the shell diameter and relative thickness at a given particle temperature:
g0 p
0
g0
p p g 0
4 1
1 0.
1 2
V T
M
P P
D V T M




+ + =





(5)
Let us express the shell thickness
p
via the particle current diameter
p
D by using
the mass balance equation:
3 3
p p p p
[1 (1 2 ) ],
6
m D

=
from where
Fig. 2. Corridor of the dependence of the hollow
particle diameter with
p0
= 0.1 on temperature
at arbitrary values of surface tension.

I.P. Gulyaev and O.P. Solonenko
774
( )
3
3
p p p p
1 2 1 6 ( ). m D = (6)
Substituting (6) into (5) we obtain an algebraic equation for variable D
p
, which is solved numeri-
cally using the Newtons iteration method.
Note the interesting fact revealed as a result of conducted computations: the expansion
of particles at heating depends weakly on the surface tension coefficient of the material. To under-
stand the reason for this at a qualitative level let us consider the expansion of a hollow droplet
with a negligibly small wall thickness the soap bubble. Let the particle have the initial
diameter D
0
, the pressure in gas cavity P
0
, and temperature T
0
, which equals, for example,
the material melting temperature. Let us calculate the value of diameter
1
, D which the particle
will have at its heating up to the maximum temperature
1
T (boiling temperature) corresponding
to the pressure in cavity
1
. P According to the ideal gas law we have:
( )
1 0 0 1 1 0
. V V P T PT =
Let us express the ratio of volumes in terms of diameters: ( )
3
1 0 1 0
/ / V V D D = and write
explicitly the full gas pressure in the particle 8 . P D P

= + Then
3
0 1 1
0 0 1
1 8
.
1 8
P D D T
D T P D

+
=

+

(7)
Consider two limiting cases: an infinitesimal value of the surface tension
0
8 / 1 P D

<<
and the infinitely high value
0
8 / 1. P D

>> Then we obtain from equality (7)
( )
( )
1/ 3
1 0 0
1
1/ 2
0
1 0 0
/ , 8 / 1,
/ , 8 / 1.
T T P D
D
D
T T P D

<<

=

>>

(8)
Thus, in both limiting cases the expansion of a particle the soap bubble depends
at heating fairly weakly on the material surface tension, the dependence of the relative increase in
the size
1 0
/ D D is bounded from above and from below by the square and cubic roots of the rela-
tive temperature increase
1 0
/ . T T If one accounts for the nonzero thickness of the hollow sphere
shell, then the dependence of the relative diameter
1 0
/ D D increase on the relative temperature
increase
1 0
/ T T is much weaker in comparison with dependencies (8): small changes of the drop-
let diameter ensure larger changes of the gas cavity volume.
Figure 2 shows the range of the variation of the relative diameter of hollow particles
with the shell initial thickness
p
= 0.1 depending on temperature: at
1 0
/ 2 T T = , the variation
of the surface tension coefficient value within any limits leads to the relative expansion variation
1 0
/ D D by no more than 6 %, and by 12 % at
1 0
/ 5 T T = . Note that for real materials, the max-
imum value of the temperature ratio
1 0
/ T T is determined by the ratio of the boiling tempera-
ture to the melting temperature, which equals
pb pm
/ 1.5 T T for ZrO
2
.
Turning to the properties of gas filling the particle cavity we note that if the relative mo-
lar gas mass M does not undergo the changes in the temperature range under consideration (there
are no dissociation and ionization), then the droplet diameter change with increasing tempera-
ture does not depend on the type of filling gas. Otherwise one should account for the depen-
dence of the particle expansion on the complex T
1
M(T
0
)/T
0
M(T
1
) rather than on the ratio T
1
/T
0
.
1.4. The profiles of temperature and velocity of the plasma jet
As a basis for computing the plasma jet parameters the real conditions of the plasmatorch
operation with an inter-electrode insert (IEI) of nominal power of 50 kW developed
at the ITAM SB RAS [22] were chosen. Table 2 presents the power characteristics of operation
Thermophysics and Aeromechanics, 2013, Vol. 20, No. 6
775
regimes of the plasmatorch with the nozzle 8 mm in diameter, plasma forming gas (air) flow
rate of 0.75 g/s (40 l/min).
The mean-mass values of the velocity
f 0
U and temperature
f 0
T in the nozzle exit sec-
tion were computed for a given diameter
0
d and the flow rate of the plasma forming gas G by
using the measured values of the operation current I and arc voltage U with allowance for
the plasmatorch thermal efficiency .
T
Then the thermal power of the plasma jet ,
T T
P UI =
the specific enthalpy of plasma / ,
T
H P G = and its mean-mass temperature
f 0
T is determined
from equation
f 0
0
0
( )
T
T
H H c T dT = +

using the tabular data characterizing the temperature depen-
dence of the air heat capacity ( )
p
c T [23]. The mean-mass velocity of plasma in the nozzle exit
section equals
f 0
2
0 Air f 0
4
.
( )
G
U
d T
=
As is seen, an increase in the arc current from 100 to 350 A leads to a considerable altera-
tion of the arc main characteristics: an increase in the plasma initial velocity from 420
to 1000 m/s, and temperature from 6670 to 11200 K.
Tabl e 2
Characteristics of plasma jets in different regimes of the plasmatorch operation
G 0.75 g/s (40 l/min)
d
0
, mm 8
I, A 100 150 200 250 300 350
U, V 210 201 192 191 188 189

T
, %
77.4 70.1 67.4 65.2 62.8 62.4
P
T
, kW
16.3 21.1 25.9 31.1 35.4 41.3
H, kJ/g 21.7 28.2 34.5 41.5 47.2 55
T
f 0
, K
6670 7150 7650 8500 9840 11200
U
f 0
, m/s
420 500 580 700 840 1000

Fig. 3. Computed profiles of the gas velocity U
f
and temperature T
f
along the jet axis.
Current I = 100 A (1), 150 A (2), 200 A (3), 250 A (4), 300 A (5).
I.P. Gulyaev and O.P. Solonenko
776
The plasma velocity and temperature distributions along the jet axis, which were used in
computations, were obtained with the aid of the software package ANSYS Fluent. The statio-
nary problem was solved in the two-dimensional axisymmetric statement using Reynolds stress
model (RSM) [24]. The data of the work [23] on temperature dependent thermodynamic and
transfer properties of air was used. Figure 3 shows the distributions of gas temperature and veloc-
ity along the axis of the air plasma jet exhausting from the plasmatorch (the nozzle diameter is
8 mm, the length is 15 mm), at arc currents 100300 A and operating gas flow rate 0.75 g/s.
2. Results of numerical experiments
2.1. Analysis of computational results
Consider the motion of hollow particles of diameter
p0
D = 50 and 100 m with the shell
thickness
p0
= 0.1 as well as the dense particles of equivalent mass whose diameters are
determined by expression
3
3
p eq p p
1 (1 2 ) D D = and are equal to 39 and 79 m, respectively.
Figure 4 shows as an example of the variation of the velocity and temperature of particles
along the axis of a jet corresponding to the current 150 A (Table 2).
Table 3 presents the maximum values of the velocity and temperature, which are reached
by the hollow and dense particles in plasma as well as the corresponding time intervals since
the moment of the particle injection and when the particle reaches the jet given section. It also
presents the computations of particles mass loss at the expense of the material evaporation.
One can see that in all cases, the hollow particles reach the maximum values of the velocity
and temperature during lesser time intervals and in the jet sections lying closer to the nozzle.
An increase in the operation current leads to a growth of maximum velocities and temperatures
of particles (the spheres of a smaller size, both a hollow one and an equivalent dense one, reach
in all cases the boiling temperature). For the plasmatorch operation regime 200 A, the lower-size
particles lose the most part of their mass at the expense of their evaporation: the hollow particle los-
es 91 %, and the dense one 63 %, which leads to a sharp velocity loss at the exit from the jet.
In different jet exhaustion regimes, the maximum values of the velocity and temperature
of hollow particles with the shell thickness
p
= 0.1 are higher by about 15 % than for the dense
particles of equivalent mass. This is due to a more intense interphase exchange by the momen-
tum and heat gasparticle at the expense of a larger area of the hollow spheres surface.
Tabl e 3
Maximum values of the velocity and temperature of ZrO
2
particles.
Arc
current
Particles of ZrO
2

hollow 50 m,
p

= 0.1 dense 39 m hollow 100 m,
p

= 0.1 dense 79 m
Maximum velocity, m/s (time, ms / jet section, cm)
100 A 190 (0.9 / 10) 165 (1.2 / 12) 110 (1.8 / 12) 95 (2.0 / 12)
150 A 240 (0.8 / 10) 195 (1.0 / 12) 130 (1.6 / 12) 110 (1.9 / 12)
200 A 330 (0.7 / 10) 233 (0.9 / 12) 150 (1.5 / 12) 125 (1.7 / 12)
Maximum temperature, K (time, ms / jet section, cm)
100 A 4573 (0.5 / 3.3) 4573 (0.7 / 4.5) 3490 (1.8 / 12) 3165 (2.0 / 13)
150 A 4573 (0.4 / 2.5) 4573 (0.5 / 3.2) 4055 (1.6 / 12) 3480 (2.0 / 13)
200 A 4573 (0.3 / 1.8) 4573 (0.4 / 2.1) 4573 (1.0 / 7) 4200 (1.7 / 12)
Mass loss, %
100 A 28 15 0 0
150 A 58 37 0 0
200 A 91 63 9 0
Thermophysics and Aeromechanics, 2013, Vol. 20, No. 6
777
Consider in more detail the behavior of hollow spheres of diameter
p0
100 m D = with
a different shell thickness in a plasma jet for regime I = 150 A. Figures 5a and 5b show the vari-
ation of the velocity and temperature of hollow ZrO
2
particles with the initial shell thickness
p0
= 0.05, 0.1, 0.15, and 0.2. We also present here, for comparison, the velocity and
temperature distributions of a dense particle of the same diameter. It is seen that lighter
particles are characterized by higher velocity and temperature values. The lightest par-
ticle with shell thickness
p0
0.05 = quickly reaches the boiling temperature (0.7 ms, z = 5 cm)
and loses 18 % of its mass, at the expense of which reaches the maximum velocity (> 170 m/s),
which exceeds considerably the velocity of denser spheres. The maximum velocity of the dense
particle amounts to 93 m/s; for a particle with
p0
0.2 = it reaches 103 m/s (+11 %), with
p0
0.15 = 112 m/s (+20 %), and with
p0
0.10 = 130 m/s (+38 %).
The maximum temperatures of droplets
pmax
T also depend substantially on the shell
thickness: if the particle with
p
0.05 = reaches the boiling temperature, then the dense sphere
has the time to melt only up to 81 % of its mass. The maximum temperatures of hollow
droplets with
p0

= 0.1, 0.15, and 0.2 are equal to 4054, 3448, and 3183 K, respectively.
Figures 5 and 5d show the variation of the external diameter and the shell relative thick-
ness of the same particles. The diameter of a sphere with the initial shell thickness
p0

= 0.05
varies in the range from 98 to 118 m (after the mass loss, the diameter becomes less than
the initial one), the least value of the shell thickness for this particle reaches 0.025, i.e., 50 %
of the initial value. The variations of sizes of particles with thicker shells are less considerable,
which is related to a smaller volume of the gas cavity and lower temperatures of heating.
So a particle with
p0

= 0.1 expands to diameter 109 m, whereas the variation of the diameter
of a particle with
p0

= 0.2 does not exceed 1 %.
2.2. Influence of the expansion of hollow ZrO
2
particles on their behavior
in plasma jet
As was shown above, the hollow spheres can alter substantially their size because
of the gas cavity expansion. To determine the influence of this effect on particles behavior in
the jet the computations of the velocity and temperature of hollow ZrO
2
microspheres with
the initial diameter of 50 m and relative shell thickness
p0
= 0.05, 0.1, and 0.2 at their motion
along the axis of a plasma jet exhausting at the arc current I = 150 A (Table 2) with and without

Fig. 4. Variation of the velocity (a) and temperature (b) of hollow and dense ZrO
2
particles at their
motion along the plasma jet.
The arc current I = 150 A, the jet initial velocity U
f0
= 500 m/s, the initial temperature T
f0
= 7150 K; hollow particles:
50 m (1), 100 m (2) with
p0
= 0.1; dense particles: 39 m (3), 79 m (4).
I.P. Gulyaev and O.P. Solonenko
778
allowance for gas cavity expansion were done. The consideration of the gas cavity expansion is
shown to lead, under the same remaining conditions, to some increase in the velocity of par-
ticles and the reduction of their temperature, however, the given changes amount to several
percents: for particles with
p0

= 0.05, the velocity increases by 5 %, and temperature by 2 %.
For microspheres with a larger shell thickness (
p0
= 0.1 and 0.2), these differences are
negligibly small. Similar computations done for hollow particles of 20100 m size and dif-
ferent regimes of the plasmatorch operation showed that for a fixed distance of treatment,
the influence of the expansion of ZrO
2
droplets on their velocity and temperature at the axial
introduction of powder in the plasma jet is much less than the errors caused by the choice
of empirical dependencies for the drag and heat-exchange coefficients, consideration or neg-
lecting the film temperature, and other effects. Consequently, the engineering estimates
of the velocity and temperature of hollow ZrO
2
droplets (for example, for a given spraying distance)
using the distributions of the mean-mass velocity and temperature of plasma along the jet axis
can be done without considering the expansion of hollow droplets. At the same time, for more
detailed computations, in particular, at the radial introduction of particles in the jet, it is neces-
sary to account for substantial velocity and temperature gradients in the plasma carrying flow,
which conditions a considerable difference in the paths of hollow particles and, hence, condi-
tions for their treatment.
2.3. Efficiency of interphase heat transfer
As was shown, the temperature and velocity of hollow particles in the plasma flow depend
substantially on their diameter and shell thickness. The heating and acceleration (the increase
in temperature and velocity) of the particle are proportional to the area of its surface
surf
S and the time of residence in plasma , t and inversely proportional to the mass m
p
:

Fig. 5. Variation of the characteristics of hollow ZrO
2
particles with D
p0
= 100 m in a plasma jet.
I = 150 A, U
f0
= 500 m/s, T
f0
= 7150 K; a velocity, b temperature, external diameter, d shell relative
thickness. Hollow particles:
p
= 0.05 (1), 0.1 (2), 0.15 (3), 0.2 (4); dense particles 5 (a, b).
Thermophysics and Aeromechanics, 2013, Vol. 20, No. 6
779
surf p
~ / . t S m However, at a higher specific area
spec surf p
/ S S m = , the particles increase
their velocity faster, and the time of their residence in a high-temperature region of the jet t
becomes less. To determine which of these factors is more significant we have carried out
the computations of the acceleration and heating of hollow particles with different shell thickness,
in particular, of dense particles, in a control section of the plasma uniform flow with parame-
ters U
f0

= 580 m/s, T
f0

= 7600 K. The control section z = 2.4 cm was chosen to ensure that all
particles preserve their initial mass, and evaporation of the material does not affect the compu-
tational results. Figure 6 shows the computed dependencies of the velocity and specific enthalpy
for 36 particles (nine values of the diameter and four values of the shell thickness).

Fig. 6. Velocity (a, , ) and specific enthalpy (b, d, f) of particles vs. the diameter (a, b), mass (c, d),
and specific surface (e, f) in a control section of uniform plasma flow.
Hollow particles

= 0.05 (1), 0.1 (2), 0.2 (3); dense particle 4.


I.P. Gulyaev and O.P. Solonenko
780
Along the abscissa axis on the first series of graphs (a, b), the diameter is measured, on
the second (c, d), the mass, and on the third (e, f) the specific area of particles. The specific
enthalpy is the ratio of the heat obtained by the particle from plasma to its mass and is
a direct figure of the efficiency of heat transfer from the gaseous phase to the disperse
phase. The temperature of spheres is not used in the results to avoid the uncertainties related
to the material melting when the particles, which have obtained a different amount of heat may
have the same temperature
pm
. T
Table 4 presents the values of the mass and specific area for hollow particles of diameters
D
p
= 20, 40, 60, 80, and 100 m with shell thickness
p0
= 0.050.5. It is seen that in all three
computational runs (Fig. 6), significant differences in the velocity and specific enthalpy of par-
ticles with different shell thickness are observed, that is the dependence of these quantities on
D
p
, m
p
, and S
spec
is not invariant under the thickness of particles shell. The thinner the particle
shell the stronger the particle accelerates and heats up. The pace of the acceleration and heat-
ing of hollow particles with the shell thickness
p0

= 0.05 is higher by factors from 2 to 2.5
than in the case of dense spheres of the same diameter, and it is 1.31.5 times higher
than for the equivalent dense particles of the same mass and 1.72.1 times higher than in
the case of dense particles having the same specific area. It is to be noted that the least dif-
ferences in velocity and temperature are observed for particles of the same mass, therefore, one
can simulate the behavior of hollow particles as a crude estimate (the differences up to
3050 %) with the aid of dense particles of equivalent mass.
The remaining results presented in Fig. 6 enable the following conclusions to be drawn:
the hollow particles accelerate and heat up, firstly, more intensively than the dense particles
of the same diameter, secondly, more intensively than the equivalent dense particles of the same
mass, thirdly, less intensively than the dense particles having the same specific surface.
Conclusions
1. The one-dimensional model of the motion, heating, melting, and evaporation
of hollow particles, which has made it possible to elucidate the influence of the gas cavity
expansion on the behavior of droplets in plasma jet, has been proposed and numerically
investigated. It has been shown that at the plasma treatment, the diameter and shell thickness
of hollow droplets can undergo substantial changes up to 20 % and 50 %, respectively.
Tabl e 4
Characteristics of the ZrO
2
particles of different sizes and shell thickness

p
0.05 0.1 0.15 0.2 0.5 (dense)
D
p
= 20 m
m
p
, 10
10
kg
0.06 0.11 0.15 0.18 0.23
S
spec
, m
2
/kg 198 110 82 68 54
D
p
= 40 m
m
p
, 10
10
kg
0.51 0.92 1.23 1.47 1.88
S
spec
,m
2
/kg 99 55 41 34 27
D
p
= 60 m
m
p
, 10
10
kg
1.72 3.09 4.16 4.96 6.33
S
spec
, m
2
/kg 66 37 27 23 18
D
p
= 80 m
m
p
, 10
10
kg
4.07 7.32 9.86 11.76 15.01
S
spec
, m
2
/kg 49 27 20 17 13
D
p
= 100 m
m
p
, 10
10
kg
7.94 14.30 19.25 22.98 29.31
S
spec
, m
2
/kg 40 22 16 14 11
Thermophysics and Aeromechanics, 2013, Vol. 20, No. 6
781
2. Hollow particles generally accelerate and heat up more intensively than the dense
particles of the same diameters or the same mass and less intensively than the dense particles
of the equivalent specific area. The minimum differences in the dynamics of heating and acce-
leration are observed for particles of the same mass.
3. It has been shown that the expansion of hollow particles at their heating does not
depend on the type of the filling gas, if the relative molar mass does not undergo changes in
the temperature range under consideration.
Nomenclature
D
p

particle diameter, m,

p
thickness of the hollow particle shell, m,

p
=
p
/D
p
dimensionless thickness of the particle shell,
z
p
particle coordinate along the z axis, m,
U
p
particle velocity along the z axis, m/s,
T
p
particle mean-mass temperature, K,
H
p

particle enthalpy, J,
m
p
particle mass, kg,
m
g
gas mass in cavity, kg,

p
density of the particle material, kg/m
3
,
c
ps
, c
pl

heat capacity of the particle material in the
solid and liquid state, J/(kgK),

p
thermal conductivity of the particle material,
W/(mK),
a
p
thermal diffusivity of the particle material, m
2
/s,
U
f

local flow velocity, m/s,
T
f

local flow temperature, K,
C
d
drag coefficient of the sphere,
coefficient of heat exchange between the particle
and plasma, W/(m
2
K),

film
,
film
,
film
density, dynamic viscosity and ther-
mal conductivity of the plasma-forming gas,
computed at the film temperature
T
film
= (T
f
+ T
p
)/2,
integral emissivity of the material,

SB
StefanBoltzmann constant, W/(m
2
K
4
),
Re Reynolds number of the particle relative motion in
the flow,
Pr Prandtl number for plasma-forming gas,
coefficient of the material surface tension, J/m
2
,
P
g
gas pressure in the particle cavity, Pa,
V
g
volume of the particle gas cavity, m
3
,
M molar mass of gas in the cavity, kg/mole.
References
1. N. Markocsan, P. Nylen, J. Wigren, and X.-H. Li, Low thermal conductivity coatings for gas turbine applica-
tion, J. Therm. Spray Technol., 2007, Vol. 16, No. 4, P. 498505.
2. O.P. Solonenko, A.V. Smirnov, and I.P. Gulyaev, Spreading and solidification of hollow molten droplet under
its impact onto substrate: computer simulation and experiment, in: Complex Systems: 5th Intern. Workshop on
Complex Systems, 2528 September 2007, Sendai, Japan. AIP Conf. Proc., 2008, Vol. 982, P. 561568.
3. O.P. Solonenko, I.P. Gulyaev, and A.V. Smirnov, Plasma processes of obtaining the powders consisting of hol-
low microspheres, in: Problems and Achievements of Applied Mathematics and Mechanics. Collection of Sci.
Works, Nonparel, Novosibirsk, 2010, P. 502519.
4. K.S. Ravichandran, K. An, R.E. Dutton, and S.L. Semiatin, Thermal conductivity of plasma-sprayed monolithic and
multilayer coatings of alumina and yttria-stabilised zirconia, J. Am. Ceram. Soc., 1999, Vol. 82, No. 3, P. 673682.
5. M.R. Dorfman, M. Nonni, J. Mallon, W. Woodard, and P. Meyer, Thermal spray technology growth in gas turbine
coatings, in: Proc. Int. Thermal Spray Conf. Osaka, Japan, 2004, DVS-Germany, Dusseldorf, Germany, P. 9095.
6. H.B. Guo, S. Kuroda, and H. Murakami, Comparative study on segmented thermal barrier coatings sprayed
from different feedstocks, in: Proc. Int. Thermal Spray Conf., Basel, Switzerland, 2005, P. 935939.
7. W. Chi, S. Sampath, and H. Wang, Ambient and high-temperature thermal conductivity of thermal sprayed coat-
ings, J. Therm. Spray Technol., 2006, Vol. 15, No. 4, P. 773778.
8. O.P. Solonenko, A.A. Mikhalchenko, and E.V. Kartaev, Splat formation under YSZ hollow droplet impact onto
substrate, in: Proc. Int. Thermal Spray Conf., Basel, Switzerland, 2005, P. 14101415.
9. I.P. Gulyaev and O.P. Solonenko, Hollow droplets impacting onto a solid surface, Exp. Fluids, 2013, Vol. 54,
No. 1, P. 1432-11432-12.
10. B. Kadyrov, Y. Evdokimenko, V. Kisel, and E. Kadyrov, Calculation of the limiting parameters for oxide ce-
ramic particles during HVOF spraying, in: C.C. Berndt and S. Sampath (Eds.), Thermal Spray Industrial Applica-
tions, ASM International, Materials Park, 1994, P. 245250.
11. S.V. Joshi, Comparison of particle heat-up and acceleration during plasma and high velocity oxy-fuel spraying,
Powder Metall. Int., 1992, No. 24, P. 373378.
12. T. Dobbins, R. Knight, and M. Mayo, HVOF thermal spray deposited Y
2
O
3
-stabilized ZrO
2
coatings for thermal
barrier applications, J. Therm. Spray Technol., 2003, Vol. 12, No. 2, P. 214225.
I.P. Gulyaev and O.P. Solonenko
782
13. T. Klocker, T.W. Clyne, and M.R. Dorfman, Process modelling to optimize the structure of hollow zirconia par-
ticles for use in plasma sprayed thermal barrier coatings, in: Proc. Int. Thermal Spray Conf., Singapore, 2001,
P. 149155.
14. D. Wroblewski, O. Ghosh, A. Lum, M. VanHout, S.N. Basu, M. Gevelber, and D. Willoughby, Analysis
of plasma spray particle state distribution for deposition rate control, in: Proc. Int. Thermal Spray Conf. Maastricht,
Netherlands, 2008, P. 826831.
15. O.P. Solonenko, I.P. Gulyaev, and A.V. Smirnov, Thermal plasma processes for production of hollow spherical
powders: Theory and experiment, J. Therm. Sci. Technol., 2011, Vol. 6, No. 2, P. 219234.
16. O.P. Solonenko, I.P. Gulyaev, and A.V. Smirnov, Hollow droplets micro explosive thermal spraying: fundamen-
tals, in: Proc. Int. Thermal Spray Conf. Maastricht, Netherlands, 2008, P. 229234.
17. V.V. Kudinov, P.Yu. Pekshev, V.E. Belashchenko, O.P. Solonenko, and V.A. Safiullin, Spraying of Coatings
by Plasma, Nauka, Moscow, 1990.
18. P. Fauchais, Understanding plasma spraying, J. Phys. D: Appl. Phys., 2004, Vol. 37, No. 9, P. R86R108.
19. D. Xu, X. Wu, and X. Chen, Motion and heating of non-spherical particles in a plasma jet, Surf. Coat. Technol.,
2002, Vol. 171, P. 149156.
20. D.J. Carlson and R.F. Hoglund, Particle drag and heat transfer in rocket nozzles, AIAA J., 1964, Vol. 2, No. 11,
P. 19801984.
21. W.E. Ranz and W.R. Marshall, Evaporation from drops, Chem. Engng. Prog., 1952, Vol. 48, No. 3, P. 141146.
22. O.P. Solonenko, A.P. Alkhimov, V.V. Marusin et al., High-energy Processes of Metals Treatment, Nauka, Novosibirsk,
2000.
23. A.S. Predvoditelev, E.V. Stupochenko, A.S. Pleshanov et al., Tables of Thermodynamic Functions of Air,
Computer Center of the USSR Acad. Sci., Moscow, 1962.
24. FLUENT 6.3 Users guide. [Electronic resource].

S-ar putea să vă placă și