You are on page 1of 9

Aerospace Science and Technology 12 (2008) 150–158

www.elsevier.com/locate/aescte

Water vapor condensation in forced convection flow over an airfoil


S.J. Karabelas ∗ , N.C. Markatos
National Technical University of Athens (N.T.U.A.), Greece Department of Chemical Engineering, Computational Fluid Dynamics Unit, 157 80 Athens, Greece
Received 26 November 2006; received in revised form 9 May 2007; accepted 10 May 2007
Available online 17 May 2007

Abstract
The present paper presents an investigation of water vapor’s condensation to liquid substance in highly convective flow conditions. An airfoil
geometry was chosen to demonstrate the applicability of the model developed. The flow is considered subsonic and compressible, at high Reynolds
number. The contribution of turbulence effects is accounted for by the Spalart–Allmaras model, which is suitable for such type applications. The
study of condensation is based on a mixture two-phase model, which allows for interpenetrating substances, while following the economical single-
fluid approach. Furthermore, the phases may move at different velocities (slip velocity) in a manner that is described by conservation equations.
Results are presented for several flight conditions, concerning the angle of attack and the ambient-air humidity levels. Since the condensation
effects are investigated on an airfoil configuration, emphasis has been given to the liquid mass-fraction distribution in several critical areas, where
considerable formation of liquid droplets is observed. Results show that for humidity levels higher than 0.7, high liquid concentration areas are
indicated onto the rear part of the foil. Attention is focused on the dispersion of the condensation cloud and on the areas of mass-transfer interaction
between liquid droplets and humid air.
© 2007 Elsevier Masson SAS. All rights reserved.

Keywords: Airfoil; Humidity; Water vapor; Condensation cloud; Two-phase flow

1. Introduction For forced convection flows, there exist references in litera-


ture that present studies of heat and mass transfer for multiphase
Heat- and mass-transfer processes among phases in multi- processes involving water vapor condensation, but refer mainly
phase flow systems, usually affect considerably the overall flow to laminar flows [1,2,4,11], shock-wave flows [14,15] and in-
structure in many engineering applications. Some examples are dustrial flows [8,12,17]. Most of the studies in the aerodynamic
windy-day evaporation, vaporization of moisture and fog, distil- field, incorporate steady-state nucleation theory, accounting for
lation of a volatile component from a mixture of non-volatiles, the vapor condensation [19,20]. One of the most important the-
cryogenic pumping and gas cooling and many more. ories was developed by Wegener and Mack [21], who speci-
Humidity conditions affect flight conditions, especially at fied analytically the nucleation rate of the water droplets. They
high altitudes. It is well known that at these altitudes, there also proposed relationships for the evaluation of the ‘critical
is snow formation on the wings, which of course affects the droplet’s work’, which is required to produce one stable spher-
boundary layer flow and changes the values of the aerodynamic ical droplet of critical size reversibly and isothermally.
forces (different aero-elasticity behavior). At low and medium Most of these considerations are based on the kinetic the-
altitudes, co-trails are observed near the body of the aircraft, ory and correlations of the experimental data. P. Wegener and
usually near the root of the wing and above the cockpit, which A. Pouring performed experiments on condensation by homo-
form in high manoeuvrability turns or in direct flight under geneous nucleation in supersonic nozzles [20], analyzing the
intense humidity conditions (flight above the sea, humid day, state-of-the-art in the area of nucleation theories. They presume
etc.). that the appearance of a liquid phase in vapor depends on small
clusters formed by fluctuations in the gaseous phase. Several
* Corresponding author. Tel.: +30-210-8030879. researchers implemented growth-rate models with relative suc-
E-mail address: stkarabelas@yahoo.gr (S.J. Karabelas). cess [19–22]. They predicted quite accurately the aerodynamic
1270-9638/$ – see front matter © 2007 Elsevier Masson SAS. All rights reserved.
doi:10.1016/j.ast.2007.05.003
S.J. Karabelas, N.C. Markatos / Aerospace Science and Technology 12 (2008) 150–158 151

Nomenclature

E total energy . . . . . . . . . . . . . . . . . . . . . . . . . . Joule/m3 v̄ eddy viscosity . . . . . . . . . . . . . . . . . . . . . . . . . m2 /sec


Gv production of turbulence viscosity. . . . . . . . . N/m2 x1 x coordinate (axial) . . . . . . . . . . . . . . . . . . . . . . . . . m
M Mach number x2 y coordinate (crosswise) . . . . . . . . . . . . . . . . . . . . . m
P static pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pa α mass fraction
SM mass source . . . . . . . . . . . . . . . . . . . . . . . Kgr/m3 sec μ molecular viscosity . . . . . . . . . . . . . . . . . . Kgr/m sec
Sh latent heat . . . . . . . . . . . . . . . . . . . . . . . . . . Joule/m3 ρ fluid density . . . . . . . . . . . . . . . . . . . . . . . . . . Kgr/m3
T static temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . K τ̄ij stress tensor
Yv destruction of turbulence viscosity . . . . . . . . N/m2
Subscripts
d distance from the wall . . . . . . . . . . . . . . . . . . . . . . . m
hfg reference latent heat of condensation . . Joule/Kgr k index for phases
k thermal conductivity . . . . . . . . . . . . . . . . . . . W/m K c control volume
m fluid mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Kgr eff effective value
ṁ fluid mass rate . . . . . . . . . . . . . . . . . . . . . Kgr/m3 sec m mixture value
t time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . sec sat variable value at saturation conditions
u velocity component . . . . . . . . . . . . . . . . . . . . . m/sec w.v. water vapor
vdr drift velocity vector . . . . . . . . . . . . . . . . . . . . . m/sec 1 first phase (air and water vapor)
v velocity vector . . . . . . . . . . . . . . . . . . . . . . . . . . m/sec 2 second phase (liquid water)

performances of airfoils for various Mach numbers and inci-


dence angles.

Most of these studies analyze the effect of the water va-


por condensation process on the transonic flow region and
they consider non-equilibrium and homogeneous condensa-
tion. G.H. Schnerr and U. Dohrmann [29] investigated the
aerodynamic performance in lift and drag terms over air-
foils. They examined equilibrium and non-equilibrium ho-
mogeneous cases, for a variety of incidence angles, for spe-
cific reservoir conditions of pressure, temperature and va-
por concentration. J.C. Lee and Z. Rusak in their studies
[30,31] evaluate theoretically and numerically the effect of
the latent heat released from non-equilibrium condensation
processes.
Fig. 1. Computational domain over the Clark-y foil.

In the present work, a new methodology is developed, where


the water droplets are being treated as a separate phase, with cules’ fluctuations, with constant rate at a prescribed pressure
different thermodynamic and flow properties. Specifically, a and temperature.
two phase ‘mixture model’ was adopted, where the mixture of The geometry examined is a Clark-y airfoil, bounded by
air and water vapor (moist air) is considered as the primary two wall regions and another two fixed-pressure-type boundary
phase and the liquid water as the secondary. The phases ex-
conditions (Fig. 1). For computational cases of this type, when
change momentum and energy, but the dominant term is the
turbulence is a significant factor contributing to the flow de-
mass transfer, which determines the rate of birth for the water
velopment, the Spalart–Allmaras one-equation model is a suit-
droplets and consequently the rate of death of the vapor mass.
The condensation process is assumed to be at equilibrium, able and economical one, as it was specifically designed for
which seems to be a realistic approach for atmospheric flow aerospace applications with wall bounded areas [5,6]. However
conditions [13,19,31]. Therefore, as soon as saturation condi- the present model is general and allows for any type of turbu-
tions are reached, the condensation process starts immediately. lence model to be easily included.
The water droplets are spheres, with diameter of d = 10−5 m, Results of velocities, pressure, volume-fraction and mass
much greater than the critical size for growth. Therefore, in the diffusion of the condensation cloud (consisting of water drop-
saturated state the liquid droplets reach the critical size. Fur- lets), have been obtained for various humidity ambient condi-
thermore, the process is assumed to be homogeneous, hence tions. Particular attention is paid to the critical areas of water
the water droplets are created, by the collapsed vapor mole- vapor condensation.
152 S.J. Karabelas, N.C. Markatos / Aerospace Science and Technology 12 (2008) 150–158

2. Mathematical model Turbulence effects have been accounted for by Spalart–


Allmaras’ one-equation model, which is suitable for aerospace
2.1. Governing equations applications, and it is low cost since it is not necessary to cal-
culate a length scale related to the viscous layer’s thickness.
The mathematical model is based on the two-phase “mix- Eq. (5) is the transport equation for the eddy kinematic viscos-
ture model” conservation equations, where the primary phase ity, ῡ , as used in this model:
(moist air) is treated as a compressible fluid and the secondary ∂ ∂
as incompressible. The conservation equations of mass, mo- (ρ v̄) + (ρ v̄ui )
∂t ∂xi
mentum and energy are derived by summing the individual 

corresponding equations over all phases. The volumetric con- 1 ∂ ∂ v̄


= Gv + (μlam + ρ v̄)
straint is applied, which demands the total volume fraction to σv ∂xj ∂xj
be equal to 1.0. This model allows for slip velocity formulation,  2
∂ v̄
e.g. the two phases may move at different speeds. The flow can + Cb2 ρ − Yv (5)
∂xj
be described by the following governing equations:
where Gv = Cb1 ρS v̄, S = S1 + v̄fu2 /κ 2 d 2 and fu2 = 1 −
Continuity: χ
χfu1 , Yv = Cw1 ρfw ( d ) with χ = νlam and Cb1 , Cb2 , κ, σv ,
v̄ 2 ν̄
∂ρm
+ ∇ · (ρm υ m ) = SM (1) equal to 0.1355, 0.622, 0.41 and 0.6666, respectively.
∂t
 k=2
where υ m = ( k=2
k=1 ak ρk υ k )/ρm , ρm = k=1 ak ρk .
2.2. Inter-phase heat and mass transfer
Momentum: Concerning inter-phase mass transfer, we assume that the
∂(ρm υmi ) rate of mass transfer between the two phases is thermodynam-
+ ∇ · (ρm υmi υ m )
∂t ically related to the saturation of the water vapor, which is a
 k=2 
 process dependent upon the ambient humidity conditions. We
= −∇p + ∇ · τ̄ij + υ dr,k ∇ ak ρκ υdr,k (2) also consider that the droplets formed by condensation are of
k=1 mean diameter 10−5 m. The saturation mass is given by em-
where τ̄ij represents the viscous stress tensor given by pirical correlations as a function of temperature. The empirical
    formulation used here is:
∂ui ∂uj 2 ∂uk
τ̄ij = μeffm + − δij μeffm
∂xj ∂xi 3 ∂xk msat = 0.001 5.018 + 0.3232T + 0.0081847 · T 2
k=2
for i, j = 1, 2, 3 and μeffm = k=1 ak μeffk is the mixture effec- + 0.00031243 · T 3 Kgr/m3 (6)
tive viscosity, i.e. μeff = μlam. + μturb. , where μturb. is obtained Eq. (6) predicts quite accurately the saturation mass per
from Eq. (5) below for the dependent variable ν̄ = μturb. ρ ; and, cubic meter of air for the usual flight temperature fields. As-
υ dr,k the drift velocity for phase k, which is defined as: suming homogeneous condensation, mass transfer from phase
υ dr,k = υ k − υ m 1 to phase 2 is related to the temperature distribution. The rate
of mass exchange between water vapor and liquid can be de-
Energy: scribed by the following state equation:
 k=2   k=2  ṁ = (u2 + v2 + w2 )(m2 )Ac /Vc
∂  
ak ρ κ E k + ∇ · ak υ k (ρk Ek + p) = (u2 + v2 + w2 )(mw.v. − msat )Ac /Vc (7)
∂t
k=1 k=1
= ∇ · (keff ∇T ) + Sh (3) where mw.v. − msat gives the total water mass and the main
contribution to every computational cell is defined according to
where keff , the effective conduction coefficient is computed ac- the local velocity vector and temperature. The total mass rate is
cording to the turbulence model used. considered as a mass source SM and it contributes to the conti-
Finally the volumetric constraint has to be applied, which nuity equation.
means that the volume fraction of the second phase has to be Coupled with the condensation effect is the latent heat, re-
specified by another equation: leased to the ambient fluid. This heat rate is also related to the
∂(ak ρk ) temperature field and contributes as a heat source to the energy
+ ∇ · (ak ρk υ m ) = −∇ · (ak ρk υ dr,k ) (4) equation:
∂t
In Eqs. (1)–(4), the density of air is assumed to follow the Sh = hfg mw (8)
perfect gas law and water is treated as an incompressible fluid.
Moreover, laminar viscosity is modeled by Sutherland’s equa- 2.3. Boundary conditions
tion [18], to account for sharp temperature differences. The
properties of the primary phase, which is a mixture (air and wa- The mixture is assumed to be at a specific ambient rela-
ter vapor) have to be evaluated. The properties of moist air are tive humidity. From the properties of moist air [16], we com-
taken from [16]. pute, wherever it is needed, the volume fraction of water vapor.
S.J. Karabelas, N.C. Markatos / Aerospace Science and Technology 12 (2008) 150–158 153

At the inflow boundary (Fig. 1) we impose a dynamic pres-


sure boundary condition, with total pressure specified at Ptot =
1.3 × 105 Pa and static Pstatic = 1.01325 × 105 Pa for moist
air. The total temperature is specified at Ttotal = 315 K. At the
outlet boundary a fixed pressure of Pstatic = 1.01325 × 105 Pa
is applied and the total temperature is considered ambient at
Ttotal = 315 K. The volume fraction for the water liquid (sec-
ondary phase) is zero at both boundaries.
At the walls, adiabatic conditions (zero temperature gradi-
ent) for the energy equation and for the water mass fraction,
and no-slip conditions for velocities are assumed:
Fig. 2. Static pressure distribution across the foil’s surface.
∂T ∂wwater
= 0, = 0,
∂n ∂n
um = νm = 0 (9)
where wwater is the mass fraction of the water phase. Details for
the latter type of boundary conditions are given in [11].

3. Numerical solution

3.1. The test case considered

In the present study, calculations are performed for flow of


moist air over a Clark-y-type airfoil, that results in formation Fig. 3. Skin friction coefficient plots across the foil’s surface.
of condensation cloud of water liquid droplets. For a thorough
simulation and understanding of the process, four different am-
bient humidity values were chosen as well as four different
angles of attack, which for the purposes of this study did not
exceed 8 degrees (small angles of attack). The airfoil’s chord
length is chosen to be 1 meter for better understanding of the
results. The coordinates for all figures presented in this paper,
are expressed based on the chord’s reference length.

3.2. Solution method

The mathematical model presented above was implemented


in the well-known computational solver-package FLUENT. The
solution method for the discretization of the partial differential
equations is the finite-volume formulation. The discrete equa-
tions are solved by the pressure correction technique and specif- Fig. 4. Temperature profiles for the three grids.
ically the SIMPLE algorithm of Patankar and Spalding [10].
The equations are solved by an iterative Gauss–Seidel method 3.3. The computational grid used
in conjunction with an algebraic multigrid model. A second-
order upwind scheme was chosen for the convective fluxes,
while the classical second order central differencing scheme The computational grid adopted in this study is fully unstruc-
for the diffusion terms. More details for the above mathemat- tured and is generated based on the classical Delaunay method
ical modeling may be found in Refs. [3,7–9] and [24–27]. enhanced with a Laplace smoothing technique for the whole
The contribution of the extra terms, described above (heat grid [28].
and mass transfer rates) is included as follows: We tested several different unstructured grids for obtaining
When the mixture temperature is less or equal to the dew the optimum accuracy in our simulation and we found that
point for given relative humidity, then activate the heat and mass the choice of 87 000 cells was satisfactory, for obtaining vir-
transfer rate terms described by Eqs. (7), (8). In this case the tually grid-independent results for static pressure and volume
continuity and energy equations are adapted correspondingly. fraction. To obtain enhanced accuracy in the numerical compu-
Otherwise, set the two transfer rates to zero and do not account tations, the grid was finer near the airfoil to account for the steep
for these terms in the conservation equations. This takes place variations of the solution variables. Figs. 2, 3 present the sta-
far away from the airfoil, where the primary phase dominates tic pressure and the skin friction coefficient along the airfoil’s
and the temperature is relatively high (ambient temperature, in surface for the tested grids and Figs. 4 and 5 present mixture
our case 293 K). temperature and water volume fraction for the tested grids.
154 S.J. Karabelas, N.C. Markatos / Aerospace Science and Technology 12 (2008) 150–158

Fig. 7. Mass fraction profiles for several humidity conditions across the suction
side.
Fig. 5. Liquid volume fraction plotted against the suction surface.

Fig. 8. Mass fraction profiles for several humidity conditions, distributed on the
Fig. 6. Water mass fraction results with latent heat release (red points) and with- vertical x = 1.0, which intersects the trailing edge.
out. (For interpretation of colour in this figure legend, the reader is referred to
the web version of this article.)

Differences are observed for the temperature and skin fric-


tion plots, which are mainly restricted in the range of 2–3%
compared to the predictions of the finest grid. Accordingly,
the choice of 87 000 cells is considered adequate for under-
standing condensation cloud characteristics qualitatively, al-
though higher accuracy may still be obtained for even finer
grids.

4. Results and discussion

Some of the results obtained are presented in Figs. 6 to 16.


Fig. 9. Temperature profile across the airfoil’s upper part.
For all figures presented the airfoil chord length was 1 m. Vol-
ume fraction distribution along the upper part of the airfoil
(expansion part) is plotted in Fig. 6 for relative humidity of suppressed. For high relative humidity conditions this peak is
85%, Mach number M = 0.61 and temperature of ambient air closer to the trailing edge (position x = 1.0).
293 K. The latent heat release is included in the calculations It is of interest to examine Fig. 8. This plot shows the devel-
via Eq. (8), although its’ effect is not significant because of the opment of mass fraction profile across the vertical line, which
large values of liquid water enthalpy. intersects the trailing edge. Comparing the four curves corre-
Fig. 7 presents mass fraction distributions for various hu- sponding to four humidity values, one can conclude that there
midity values. Four values of relative humidity were chosen: is a maximum concentration for every value at about the same
50, 60, 70 and 85%. The contribution of water liquid over the point (namely at 0.1 m above the trailing edge). Of course
upper part of the airfoil is investigated. It is clearly seen that for for high humidity levels the maximum mass fraction is much
low humidity levels, the mass fraction is considerably lower. higher than in low-humidity conditions (even 10 times larger
Moreover observing the profile of mass fraction for all humid- value). Also, as seen in the same figure, for 85% relative hu-
ity levels, it should be noted that there is a peak, after which the midity the mass fraction reaches zero values at about 1 m above
water liquid’s mass fraction is reduced gradually. Finally, when the foil’s end. In the other cases this distance does not exceed
approaching the trailing edge point, the liquid concentration is 0.55 m. This plot shows that when humidity levels increase lin-
S.J. Karabelas, N.C. Markatos / Aerospace Science and Technology 12 (2008) 150–158 155

early, the relative positions of zero liquid concentrations move maximum mass fraction value is reached inside the trailing edge
far away from the airfoil exponentially. This result confirms that area and not on the suction side. Figs. 10, 11 and 12 verify that
the ambient humidity conditions strongly affect the momentum the maximum water concentration exists near the trailing edge
of water liquid and when the mass fraction of water vapor is point. From a first point of view this is not as expected, since
significant in moist air, the droplets have enough energy and there the temperatures are much higher, than those at the front
momentum to move far away from the source (here regions of side of the suction area.
low temperatures) towards all directions. Temperature distribution is plotted in Fig. 9, indicating that
Comparing the water concentrations at a fixed relative hu- the lower temperatures exist at about 0.2 m far away from the
leading edge, where there is no significant water liquid mass
midity, plotted in Figs. 7 and 8, it is easily concluded that the
fraction. Near these regions of low temperature (i.e. 0.2 m. from
leading edge), the latter falls below the dew point, and the cloud
of liquid droplets begins to form.
Although it is obvious from Fig. 9 that the dew point (for r.
humidity = 70% and T = 293 K then Tdew = 287 K) is reached
further downstream at 0.30–0.50 m from the leading edge, the
formation of the cloud is much more intense after this area,
above the trailing edge. This result may be the most signifi-
cant because it describes the domination of the inertia forces.
In this case moist air has considerable amount of momentum,
which it exchanges with water. When the secondary phase is
born (dew point area), it inherits the momentum of the primary
phase. So, although we expect high mass fraction levels at low
temperature regions, liquid layers enriched with strong momen-
tum convect the droplets near the foil’s end, where we see the
Fig. 10. Contours of water mass fraction at humidity level 70%.
maximum water liquid concentration. After the trailing edge,
there is no source of mass transfer, since temperatures become
ambient, and therefore clearly a gradual reduction of mass frac-
tion is observed (Figs. 10, 11).
Another observation which comes out of these plots is that
the liquid concentration vanishes after 10–15 m. It is worth

Fig. 11. Concentration of the water liquid droplets backwards the airfoil for Fig. 13. Volume fraction across the cross-stream direction (x = 1.0) for differ-
relative humidity 85%. ent angles of attack.

Fig. 12. Mass fraction distributions of the liquid phase for 0 and 8 incidence angles.
156 S.J. Karabelas, N.C. Markatos / Aerospace Science and Technology 12 (2008) 150–158

Fig. 14. Water velocity magnitude for 0 and 8 degrees angle of attack.

mentioning that the rarefaction of the condensation cloud oc- degrees makes inertia forces capable of moving the liquid layer
curs quite smoothly. This is due to the large amount of energy far away from the foil’s surface, where there are no sources of
and momentum that the secondary phase has already gained mass or heat to give birth to the secondary phase.
from the upstream regions, where the mass transfer rate was Wing’s aerodynamic performance is always of major con-
high. The disappearance of the cloud, is due to the loss of mo- cern for every designer. In the present study, drag, lift and mo-
mentum and energy, since no source exists beyond the airfoil. ment coefficients are investigated in humid conditions to reveal
The diffusion of the cloud for different angles of attack is any significant effects for the flight performance of the Clark-y
quite interesting, because co-trails are usually seen by the hu- foil. These coefficients are plotted against relative humidity at
man eye in high-g turns and at considerable angles of attack, zero incidence angle (Fig. 15). Based on the water-vapor data
indicating higher concentration of water droplets. Results are [16], two main areas of physical interest appear here. These are
obtained for the ambient conditions described above and three clearly indicated in drag coefficient plot in Fig. 15, where for
cases are investigated in addition to the initial case. These cor- relative humidity up to 35%, there is no generation of liquid
respond to 2, 6 and 8 degrees angle of attack. phase, hence the flow is assumed single-phase (humid air) and
Fig. 12 presents the liquid layer for 0 and 8 degrees angle the only effect in the aerodynamic performance is a very small
of attack. The boundary liquid layer is much wider for high an- change of density, because of the difference in vapor consti-
gles of attack, as expected, indicating a considerable increase tution. When humidity level exceeds 35%, then liquid phase is
of crosswise momentum. Furthermore, this plot indicates that born and the effects become more pronounced for the flight per-
the regions of maximum mass fraction spread more in all direc- formance. Specifically the mixture convection and drift-terms
tions. Specifically, in Fig. 13 we see a diagram of mass fraction in the momentum equations, affect the pressure distribution, by
values across the vertical line passing the trailing edge for all engaging the density and velocity of water-phase. According
angles of attack. Between 0 and 2 degrees there is no significant to Fig. 15, the lift and the pitch moments across the aerody-
difference in the width of the layer, which could be explained namic center (c/4, 0) increase in the two-phase region. On the
by the initial geometric configuration of the airfoil. By further contrary the drag coefficient decreases as the relative humidity
increasing the angle, there is considerable increase in liquid increases.
droplets’ volume fraction.
Flow of the secondary phase is investigated in more detail in This trend for lift and drag terms may result from the decrease
Fig. 14, where the velocity vectors of water liquid are presented in mixture static pressure, through the contribution of the mix-
for 0 and 8 degrees. The influence of the flow deviation at 8 ture convection term, expressed in the left side of Eq. (2).
S.J. Karabelas, N.C. Markatos / Aerospace Science and Technology 12 (2008) 150–158 157

Fig. 15. Variation of the drag lift and moment coefficients as a function of the humidity level. The two-phase area, affects the aerodynamic performance.

Differences are observed near the stagnation area of the air-


foil, which may be due to the fact that the grid has not been
optimized for skin friction predictions.

5. Conclusions

Turbulent flow over an airfoil with heat and mass transfer


due to phase change has been investigated. The effects of hu-
midity and angle of attack are specifically investigated. A two-
phase mixture model accounting for the exchange of mass,
momentum and energy between the two phases was used. Ad-
vanced numerical techniques were used to solve the system of
Fig. 16. CFD and experiment pressure coefficient curves.
partial differential equations, such as second order discretiza-
tion and multigrid solution algorithms implemented together
The humid aerodynamic loads deviat a few percentage with classical iterative techniques (Gauss–Seidel).
The following remarks are the major results of the study:
points of the dry values (up to 4%), but they should not be ig-
nored. In manoeuvrability flight conditions and more extreme 1. In forced convection flows with condensation, the flow is
humidity levels (very low temperatures and very high humidity influenced by the mass transfer between two phases which
levels), they could affect seriously the force-momentum equi- affects significantly the momentum of both phases.
librium. 2. When the humidity levels rise, the mass fraction of water
Despite the lack of reliable experimental data for the case liquid phase increases non-linearly.
considered, an effort was made to validate the present numerical 3. High angles of attack reduce the static pressure and temper-
predictions with experimental data [23] for Re = 300 000 and ature at the upper part and at the same time lead to a much
wider condensation cloud due to the considerable mass dif-
for a chord’s length of l = 0.3048 m. Fig. 16 presents the exper-
fusion rates.
imental pressure coefficient distribution with chord’s position
4. The strong inertia forces cause fast mass transport of the
for zero incidence angle, together with the present numerical liquid phase downstream. Thus, although the temperature
predictions. Comparison between both curves in Fig. 16, indi- levels there do not allow inter-phase mass transfer, there
cates satisfactory agreement between the wind-tunnel and the exist considerable amounts of droplets, visible far down-
present numerical results. stream of the foil.
158 S.J. Karabelas, N.C. Markatos / Aerospace Science and Technology 12 (2008) 150–158

5. Maximum water liquid concentration takes place above the [13] M. Kaviany, Principles of Heat Transfer in Porous Media, Springer, Berlin,
trailing edge approximately at 10% of the chord’s length. 1991.
6. The aerodynamic performance is affected, when significant [14] J. Lee, Z. Rusak, Transonic flow of moist air around a thin airfoil with
equilibrium condensation, J. Aircraft 38 (4) (2001) 693–702, 0021-8669.
humidity levels are reached. Lift and pitching moment in- [15] L.T. Smith, U. Yale, Experimental investigation of the expansion of moist
crease, while the drag decreases with the relative humidity air around a sharp corner, AIAA J. 9 (10) (1971) 2035–2037, 0001-1452.
level. [16] T. Fujii, Y. Kato, K. Mihara, Expressions of transport and thermodynamic
properties of air, steam and water, Sei San Ka Gaku Ken Kyu Jo, Report
Acknowledgements No. 66, Kyu Shu Dai Gaku, Kyu Shu, Japan, 1977.
[17] W.J. Chang, C.I. Weng, An analytical solution to coupled heat and mois-
ture diffusion in porous materials, Internat. J. Heat Mass Transfer 43
The authors would like to thank Assistant Professor P. Koyt- (2000) 3621–3632.
mo (University of Patras, Greece), for providing the license for [18] S. Deck, P. Duvaeu, P. d’Espincy, P. Guille, Development and application
using the commercial code Fluent for the numerical implemen- of Spalart–Allmaras one equation turbulence model to three dimensional
tation. supersonic complex configurations, Aerospace Science and Technology
(2002) 171–183.
[19] Z. Rusak, J. Lee, Transonic flow of moist air around a thin airfoil with non-
References equilibrium and homogeneous condensation, J. Fluid Mech. 403 (2000)
173–199.
[1] L.C. Chow, J.N. Chung, Evaporation of water into a laminar stream of air [20] P.P. Wegener, A.A. Pouring, Experiments on condensation of water vapour
and superheated steam, Int. J. Heat Mass Transfer 25 (1982) 499–511. by homogeneous nucleation in nozzles, Phys. Fluids 7 (1964) 352–361.
[2] C.H. Wu, D.C. Davis, J.N. Chung, Simulation of wedge-shaped product [21] P.P. Wegener, L.M. Mack, Condensation in supersonic and hypersonic
dehydration using mixtures of superheated steam and air in laminar flow, wind tunnels, Adv. Appl. Mech. 5 (1958) 307–447.
J. Numer. Heat Transfer 11 (1987) 109–123. [22] G.H. Schnerr, U. Dohrmann, Transonic flow around airfoils with relax-
[3] E.R.G. Eckert, R.M. Drake, Journal of Analysis of Heat and Mass Trans- ation and energy supply by homogeneous condensation, AIAA J. 28
fer, Chapters 20, 22, McGraw-Hill, New York, 1972. (1990) 1187–1193.
[4] J.L. Manganaro, O.T. Hanna, Simultaneous energy and mass transfer in [23] S. Ghosh, M. Muste, F. Stern, Laboratory experiment #3: Measurement of
the laminar boundary layer with large mass transfer rates toward the sur- pressure distribution and forces acting on an airfoil, 57:020 Mechanics of
face, AIChE J. 16 (1970) 204–211. Fluids and Transfer Processes.
[5] P.R. Spalart, S.R. Allmaras, A One Equation Turbulence Model for Aero- [24] B.E. Launder, D.B. Spalding, The numerical computation of turbulent
dynamic Flows, AIAA, 1999. flows, Comp. Methods Appl. Mech. Engrg. 3 (1974) 269.
[6] P.R. Spalart, S.R. Allmaras, A one-equation turbulence model for aero- [25] N.C. Markatos, Transient flow and heat transfer liquid sodium coolant in
dynamic flows, Technical Report AIAA-92-0439, American Institute of the outlet plenum of fast nuclear reactors, Internat. J. Heat Mass Trans-
Aeronautics and Astronautics, 1992. fer 21 (1978) 1565.
[7] J.C. Tannehill, D.A. Anderson, R.H. Pletcher, Computational Fluid Me- [26] D.B. Spalding, A general purpose computer program for multi-
chanics and Heat Transfer, Taylor & Francis Inc., 1997. dimensional one- and two-phase flow, in: Mathematics and Computers in
[8] N.C. Markatos, Mathematical modelling of single and two phase flow Simulation, vol. XXIII, North Holland, Amsterdam, 1981, pp. 267–276.
problems in the process industries, Revue de l’Institut Francais du Pet- [27] D.B. Spalding, Mathematical Modeling of Fluid Mechanics, Heat-
role 48 (6) (1993) 631–632. Transfer and Chemical-Reaction Processes, A Lecture Course, CFDU Re-
[9] N.C. Markatos, On numerical modelling of embedded subsonic flows, Int. port. HTS/80/1, Imperial College, London, 1980.
J. Num. Methods Fluids 6 (1986) 103–112. [28] P. Hansbo, Generalized Laplacian smoothing of unstructured grids,
[10] S.V. Patankar, D.B. Spalding, A calculation procedure for heat, mass and Comm. Numer. Meth. Engrg. 11 (5) (2005) 455–464.
momentum transfer in three-dimensional parabolic flows, Internat. J. Heat [29] G.H. Schnerr, U. Dohrmann, Drag and lift nonadiabatic transonic flow,
Mass Transfer 15 (1972) 1787–1806. AIAA J. 32 (1994) 101–107.
[11] W.M. Yan, Y.L. Tsay, T.F. Lin, Simultaneous heat and mass transfer [30] J. Lee, Z. Rusak, Parametric investigation of nonadiabatic compressible
in laminar mixed convection flows between vertical parallel plates with flow around airfoils, Phys. Fluids 13 (1) (2001) 315–323.
asymmetric heating, Internat. J. Heat Mass Transfer 10 (1989) 262–269. [31] J. Lee, Z. Rusak, Theoretical and numerical studies of transonic flow of
[12] M.K. Choudhary, K.C. Karki, S.V. Patankar, Mathematical modeling of moist air around a thin airfoil, Theoret. Comput. Fluid Dynam. 15 (6)
heat transfer, condensation, and capillary flow in porous insulation on a (2002) 359–372.
cold pipe, Internat. J. Heat Mass Transfer 47 (2004) 5629–5638.

You might also like