You are on page 1of 50

ALAQS-

CFD Comparison of Buoyant Free


and Wall Turbulent Jets


EEC/SEE/2007/004


ALAQS- CFD Comparison of Buoyant Free and Wall Turbulent Jets

This report was prepared for EUROCONTROL Experimental Centre ALAQS by:
School of Engineering and Design, Brunel University, UK.


Author(s):
Syoginus S. Aloysius, Luiz C. Wrobel



School of Engineering & Design
Brunel University
Uxbridge
Middlesex
UB8 3PH
UK
Web: http://www.brunel.ac.uk/about/acad/sed


Ian Fuller, EUROCONTROL Experimental Centre

Review:
ALAQS Peer Review Group:
Brian Y Underwood AEA Energy & Environment
Dr. Ulf Janicke Janicke Consulting
.

EEC Note : EEC/SEE/2007/004

European Organisation for the Safety of Air Navigation EUROCONTROL 2007

This document is published by EUROCONTROL in the interest of the exchange of information. It may be copied in
whole or in part providing that the copyright notice and disclaimer are included. The information contained in this
document may not be modified without prior written permission from EUROCONTROL.

EUROCONTROL makes no warranty, either implied or express, for the information contained in this document,
neither does it assume any legal liability or responsibility for the accuracy, completeness or usefulness of this
information.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets



EXECUTIVE SUMMARY

This report aims at comparing the simulation of a turbulent jet in the presence of a horizontal
solid boundary with a previous ALAQS study on a similar unbounded situation. It is believed that
the information provided by these simulations can help increasing the understanding of the
source dynamics behind an airplane jet engine, which can be used for future improvements of
existing dispersion models. Before carrying out such comparison, the turbulent wall jet
simulation was validated against existing experimental and analytical results available in the
literature for buoyant and non-buoyant conditions. Then, the buoyant case is used to assess the
impact of the solid boundary on the fluid mechanics of the jet flow.
The method employed to characterise the plume dynamics employs state-of-the-art
Computational Fluid Dynamics (CFD) software packages, which represent the most advanced
mathematics that can be applied to the simulation of fluid flow. The Large Eddy Simulation
(LES) turbulence model was used to model the smaller eddies present within the grid resolution,
and to solve explicitly the large eddies contained outside it.
The results show that the presence of the wall has major influences on the fluid dynamics. The
potential core length is extended and the decay of its maximum velocity is much slower for the
buoyant wall jet than for the buoyant free jet. This validates the accepted fact that deeper
penetration of the flow occurs when a solid boundary is present. In addition, it was found that
the growth rate is faster for the free jet. Unlike the free jet, the wall jet is constantly generating
vortices due to the presence of a solid boundary. As a consequence, the flow is forced down
and clings to the ground for a longer distance; this phenomenon is also known as the Coanda
effect. As the fluid velocity reduces further down the control volume, the wall vortical intensity
decreases and buoyancy effects take over, lifting the flow from the ground.
The ability of the CFD model to reproduce experimental and theoretical results, as well as
providing an in-depth knowledge of the fluid mechanics, gave confidence to progress to a more
complex situation. The complete wing and engine of an aircraft should be the next simulation
step, to consider important properties on the interaction between both components during
different stages of the take-off and landing phases.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

iv EEC/SEE/2007/004
(This page intentionally blank)



ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 v

REPORT DOCUMENTATION PAGE

Reference:

SEE Note No. EEC/SEE/2007/004
Security Classification:

Unclassified
Originator:
School of Engineering and Design
Brunel University
Uxbridge
UB8 3PH
United Kingdom
http://www.brunel.ac.uk
For Society, Environment, Economy
Research Area
Originator (Corporate Author) Name/Location:

EUROCONTROL Experimental Centre
Centre de Bois des Bordes
B.P.15
91222 BRETIGNY SUR ORGE CEDEX
France
Telephone: +33 1 69 88 75 00
Sponsor:

EUROCONTROL EATM
Sponsor (Contract Authority) Name/Location:

EUROCONTROL Agency
Rue de la Fuse, 96
B 1130 BRUXELLES
Telephone: +32 2 729 90 11
TITLE:
ALAQS CFD Comparison of Buoyant Unbounded and Partially Closed Turbulent Jets

Authors :
Syoginus Aloysius, Luiz C.
Wrobel

EEC Contact: Ian Fuller
Date

08/07
Pages

50
Figures

27
Tables

1
Appendix

9
References

24
EATMP Task
Specification
-
Project
ALAQS

Task No. Sponsor
TRSA03PT
Period
2007
Distribution Statement:
(a) Controlled by: EUROCONTROL Project Manager
(b) Special Limitations: None
(c) Copy to NTIS: YES / NO
Descriptors (keywords):
ALAQS, CFD, Source Dynamics, Buoyancy Effects, Turbulent Wall Jets, LES, Ground Effects
Abstract:
This report investigates the effects of a solid boundary on the flow properties of a buoyant jet. A
comparison was made with the results of a previous ALAQS report on an unbounded jet to highlight the
difference jet behaviours. It is believed that the inclusion of the wall represents a further step towards a
better understanding of the source dynamics behind an airplane jet engine during the take-off and
landing phases. The information provided from these simulations can be used for future improvements
of existing dispersion models. Before carrying out such comparison, the turbulent wall jet simulation was
validated against existing experimental and analytical results available in the literature for buoyant and
non-buoyant conditions. Then, the buoyant case is used to assess the impact of the solid boundary on
the fluid mechanics of the jet flow.

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

vi EEC/SEE/2007/004
(This page intentionally blank)


ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 vii
TABLE OF CONTENTS
1 INTRODUCTION...................................................................................................1
2 ForthmannS Experiment Validation..................................................................2
2.1 CFD Simulation.........................................................................................................................2
2.2 2D Analysis of the Results ........................................................................................................3
2.2.1 Axial Velocity Profile for Non-Buoyant Wall Jet.................................................................4
2.2.2 Vertical Velocity Profile for Non-Buoyant Wall Jet ............................................................6
2.3 Buoyant Wall Jets ...................................................................................................................12
2.3.1 Axial Velocity Profile for Buoyant Wall Jet ......................................................................12
2.3.2 Vertical Velocity Profile for Buoyant Wall Jet ..................................................................13
3 Three-dimensional study of turbulent wall jet ................................................16
3.1 Geometry and Boundary Conditions.......................................................................................16
3.2 Mean Velocity Profile Comparison..........................................................................................17
3.2.1 Streamwise Direction Analysis........................................................................................17
3.2.2 Spanwise Direction Analysis ...........................................................................................22
3.3 Vorticity Profile Comparison....................................................................................................24
4 CONCLUSIONS..................................................................................................26
5 REFERENCES....................................................................................................28

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

8 EEC/SEE/2007/004
LIST OF FIGURES
FIGURE 1 GENERAL DESCRIPTION OF WALL JET............................................................................................. 3
FIGURE 2 TWO-DIMENSIONAL WALL JET VELOCITY CONTOUR ................................................................. 3
FIGURE 3 TURBULENCE KINETIC ENERGY CONTOUR PLOT......................................................................... 4
FIGURE 4 MAXIMUM VELOCITY DECAY ALONG THE AXIS OF THE WALL JET........................................ 5
FIGURE 5 FREE AND WALL JET COMPARISON FOR THE MAXIMUM VELOCITY DECAY........................ 6
FIGURE 6 CFD VERTICAL PROFILE COMPARISON AT DIFFERENT DISTANCES BEHIND THE WALL
JET............................................................................................................................................................. 7
FIGURE 7 CFD AND EXPERIMENTAL RESULTS FOR THE VERTICAL VELOCITY PROFILE ..................... 8
FIGURE 8 ANGLE OF SPREAD COMPARISON BETWEEN FREE AND WALL JETS....................................... 9
FIGURE 9 VORTICITY CONTOUR AROUND Z-AXIS WITH ISOLINES (1/S).................................................. 10
FIGURE 10 Z-VORTICITY PROFILE COMPARISON AT DIFFERENT DISTANCES BEHIND THE NON-
BUOYANT WALL JET.......................................................................................................................... 10
FIGURE 11 HEIGHT OF THE PLUME CENTRELINE AS A FUNCTION OF DOWNWIND DISTANCE......... 11
FIGURE 12 MAXIMUM AXIAL VELOCITY DECAY PROFILES WITH AND WITHOUT BUOYANCY........ 12
FIGURE 13 VELOCITY CONTOURS FOR BUOYANT WALL JET..................................................................... 13
FIGURE 14 VELOCITY PROFILES WITH BUOYANCY AT DIFFERENT DISTANCES BEHIND THE WALL
JET........................................................................................................................................................... 13
FIGURE 15 SPREAD B COMPARISON FOR WALL JET WITH AND WITHOUT BUOYANCY...................... 14
FIGURE 16 COMPARISON OF THE HEIGHT OF PLUME CENTRELINE AS A FUNCTION OF DOWNWIND
DISTANCE.............................................................................................................................................. 15
FIGURE 17 Z-VORTICITY PROFILE AT DIFFERENT DISTANCES BEHIND THE BUOYANT WALL JET.. 16
FIGURE 18 GEOMETRY AND MESH DISTRIBUTION OF THE 3D WALL JET............................................... 17
FIGURE 19 MAXIMUM VELOCITY DECAY COMPARISON BETWEEN FREE AND WALL JET................. 18
FIGURE 20 MEAN VELOCITY PROFILE EVOLUTION THROUGH TIME FOR BUOYANT FREE AND
WALL JETS............................................................................................................................................ 19
FIGURE 21 STREAMWISE VELOCITY PROFILES AT DIFFERENT DISTANCES BEHIND THE BUOYANT
WALL JET .............................................................................................................................................. 20
FIGURE 22 3D COMPARISON OF HEIGHT OF PLUME CENTRELINE AT DIFFERENT DISTANCES
BEHIND THE EXHAUST, AFTER 10S ................................................................................................ 21
FIGURE 23 SPANWISE VELOCITY PROFILE COMPARISON OF WALL JET WITH THEORETICAL
CURVES OF THE FREE JET AFTER 10S............................................................................................ 23
FIGURE 24 COMPARISON OF SPANWISE DIRECTION OF SPREAD B AS FUNCTION OF DISTANCE
AFTER 10S.............................................................................................................................................. 24
FIGURE 25 INSTANTANEOUS VELOCITY PROFILE COMPARISON AFTER 5S ........................................... 24
FIGURE 26 VORTICITY MAGNITUDE COMPARISON AFTER 5S.................................................................... 25
FIGURE 27 STREAMWISE VORTICITY COMPARISON BETWEEN BUOYANT FREE AND WALL JETS .. 26

LIST OF TABLES
TABLE 1 - MESH DISTRIBUTION............................................................................................................................ 2

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 ix
ABBREVIATIONS
CFD Computational Fluid Dynamics
BS Bell Shape
SR Successive Ratio
RMS Root Mean Square
TI Turbulence Intensity
LS Length Scale
ALAQS Airport Local Air Quality Studies
NOx Nitrogen Oxides
LES Large Eddy Simulation
PIV Particle Image Velocimetry



ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 1
1 INTRODUCTION
Wall jets have great importance in engineering problems. Automobile demister is a good
example of a wall jet application in a real world environment 0. The design improvement of such
device can only be done with a thorough understanding of the fluid mechanics involved.
Regarding local air quality, a wall jet is a good representation of an airplane engine near the
ground emitting different types of pollutants. The study of the dispersion of these pollutants will
help dispersion modellers to produce more accurate simulations and predictions to regulate
airport traffic.
Experimental studies on turbulent flows have increased in accuracy, from simple pitot tube
measurements to complex Particle Image Velocimetry (PIV) photos, but this can be very
expensive and time consuming for such complex problems. Computational Fluid Dynamics
(CFD) has become an alternative to those complicated and laborious works, especially with the
increase in computer technology. Processor speeds have increased at an unprecedented rate
and models such as Large Eddy Simulation (LES) can play a major role in understanding the
flow behaviour.
The basic concept of the LES model is to explicitly solve the larger eddies of the control volume,
whereas the smaller eddies are modelled through spatial filtering. Although this model is not
presently suited for a complete airport simulation due to extensive requirements of computer
speed and memory, it can be used efficiently in local studies related to source dynamics and
engine plume/wing tip vortex interactions.
Forthmann, as well as being the first to conduct investigations on a plane jet, was also the
earliest to produce some experimental results on wall jet. Even though this first experiment
dates some more than 70 years ago, it presents some interesting results that can be easily
compared. Since then progress has been very slow but nowadays has seen an ever-growing
interest in this field due to its applicability in different domains.
The LES model was used in this report to investigate the presence of a solid boundary on the
fluid dynamics of a buoyant jet. The first part of the report aims at validating the CFD simulation
with Forthmanns experimental work and some analytical results in a non-buoyant condition.
Then with confidence of the obtained results, buoyancy effects were added to the simulation
and comparison was made with the previous ALAQS report made on an unbounded free
buoyant jet in a similar condition.





ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

2 EEC/SEE/2007/004
2 FORTHMANNS EXPERIMENT VALIDATION
Forthmanns experiment (1934) is the earliest on turbulent wall jets recorded in the literature 0.
This study was an extension of the turbulent free jet investigation Forthmann had previously
undertaken. The simplicity of its results and its comparison with different analytical models
makes this experiment suitable to validate CFD codes.
The experiment was fully described in a previous ALAQS report 0 and, therefore, it will not be
repeated here.
2.1 CFD Simulation
Because this study is very similar to the free jet reported in 0, a similar control volume will be
used here, with the geometry of the computational domain extending horizontally by 3.2m and
vertically by 3.3m.
The boundary conditions are as follows:
The inlet jet has a velocity of 35m/s. The inlet condition is assigned right at the beginning
of the control volume, like in the previous report 0. The air coming out of the jet is
isothermal with the ambient air of the control volume, so that no buoyancy is present in
the simulation. The turbulence intensity is 5% and the length scale 0.025m, as
configured in the previous report.
The boundaries just next to the jet were set as walls.
The top and bottom boundaries were also assigned as walls.
The bottom boundary was assigned as a stationary wall with no slip condition.
The classical outflow boundary condition is assigned at the outlet.
The surrounding air is at rest and adiabatically stratified. The main change from the previous
study is that the bottom wall now touches the jet exhaust, and particular attention has to be
given to this part of the control volume where the flow will develop. As a consequence, a more
refined mesh has to be allocated near this bottom wall. The mesh distribution was set up as
follows:


1 2 3
Mesh
Density
100
S.R.=1.0
225
S.R.=1.025
500
S.R.=1.009
162,500
cells
Table 1 - Mesh Distribution

0.03m
3.2m
3.3m
1
3
2
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 3

Figure 1 General description of wall jet
2.2 2D Analysis of the Results

Figure 2 below shows the jet velocity contours. The range of this plot goes from 0 to a maximum
of 35m/s, the jet exhaust velocity. The results present some of the characteristics discussed on
the previous free jet report 0, in that the potential core can easily be recognized by a triangular
shape of the isolines, and the velocity decay as the flow progresses through the control volume
is consistent with the one found in the free jet study. Obviously, because of the presence of the
wall, the symmetrical pattern found in the free jet study is now disrupted.

Figure 2 Two-dimensional wall jet velocity contour

Before further discussion, it is advisable to define two layers of the flow in order to compare the
results with other studies. Glaubert 0 was the first to introduce these regions:
y
x
z
Flow
development
region
Fully developed flow
0
b m
u
2
m
u

b

Potential core
Vertical
direction
Streamwise
direction
Spanwise
direction
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

4 EEC/SEE/2007/004
The inner layer is the region between the wall and the point where the velocity is
maximum, also known as the normal boundary layer thickness ( ).
The outer layer is the rest of the wall jet above the inner layer.
The CFD results will be compared with theoretical and empirical formulae found in the literature,
and will also be compared with the free jet results to assess the effect of the wall on the flow
pattern. The following two sections will discuss this, in the form of axial and vertical velocities.
2.2.1 Axial Velocity Profile for Non-Buoyant Wall Jet
Figure 2 shows that, unlike the free jet, the point of maximum velocity is not situated at the
centreline axis of the jet exhaust but below the centreline of the jet axis, and is found to vary
behind it. This has to be taken into account when quantifying the velocity decay as the flow
progresses behind the jet. The reason for the change in position can be explained by the
presence of the wall, but this will be discussed in more detail later in the report.
In the case of a free jet, it was argued that the disappearance of the potential core was the
result of turbulence penetration leading to the fully developed region. The entrainment of the
ambient fluid was the factor that triggered the enhancement of turbulence.
For a wall jet, only one side can entrain ambient fluid while the other side creates a boundary
layer due to the wall. The disappearance of the potential core in the case of a wall jet, as also
explained by Rajaratnam 0, is due to the free shear layer created by the ambient fluid meeting
with the boundary layer at the bottom wall.
To illustrate this behaviour, Figure 3 presents a contour plot of turbulent kinetic energy. A large
amount of turbulence is concentrated outside the jet geometry and above it; this clearly shows
that the level of turbulence is higher when entrainment of the ambient fluid is happening. This
high level of turbulence only happens on one side of the jet because the other side is closed by
the wall. Another interesting point is the presence of lower turbulence intensity near the wall;
this detail was also observed by Tangemann & Gretler 0. In addition to this, it can be seen that
the first portion behind the jet is lower in turbulence. This statement is in line with the findings of
ODonovan & Murray 0 in their study of impinging jets, and with our free jet results 0.

Figure 3 Turbulence Kinetic Energy Contour Plot

The influence of turbulence on the flow properties was properly discussed in the previous report
0, but it is important to stress its effect on the velocity decay. Rajaratnam 0 analysed the
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 5
maximum velocity decay of different experimental studies found in the literature and presented
a unique equation that fit all the curves, in the form:
0
0
5 . 3
b
x
U
U
m
=
which is only valid up to 100 diameters behind the jet. Szorza & Herbst 0 reported the work of
Sigalla and Bradshaw & Gee, who found that the maximum velocity decay is roughly
proportional to
x
1
.

Figure 4 Maximum velocity decay along the axis of the wall jet

Figure 4 shows the maximum velocity decay along the jet axis. The non-dimensional coordinate
of the abscissa is the length behind the wall jet divided by the jet diameter

0
b
x
, and of the
ordinate is the maximum velocity at that distance divided by the jet exhaust velocity

0
U
U
m
.
The plots show a reasonable agreement between the CFD results and the empirical equation.
The maximum velocity is slightly over-predicted by the CFD simulation and stays above
Rajaratnams empirical curve. The rate of decay is similar for both plots, with the difference
between their maximum velocities being almost constant.
The different behaviour in the upstream region may be easily explained. As reported in the
previous study 0, the empirical formula does not take into account the transitional region
between the initial and the main regions of the jet. The length of the potential core has to be
corrected according to Abramovichs law to include the transition phase 0, as explained in the
previous report 0. According to the CFD results, the length of the potential core extends to
about 7.5 diameters. Correcting the empirical results, it was found that its prediction of the
potential core extends to about 8 diameters. Thus, the CFD and empirical results are very close
near the jet exhaust.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

6 EEC/SEE/2007/004
Recollecting the data used in the previous report and changing the scales to correct to the
parameters previously used, it is possible to compare the maximum velocity decays of the free
and wall jets, as shown in Figure 5.

Figure 5 Free and wall jet comparison for the maximum velocity decay

The plots show a substantial difference between the wall and free jet. First, the decay rate is
higher for the free jet than for the wall jet. The maximum velocity is still high at 60 diameters for
the wall jet, whereas the free jet is relatively low. Second, the length of the potential core is
shorter for the free jet than for the wall jet. The length of the potential core predicted by the CFD
simulation is about 5.5 diameters for the free jet and 7.5 diameters for the wall jet. Similar
results were presented by Curd 0, who compared free and wall jets in controlling air
contaminants.
The extension in the potential core length and the lower rate of decay for the wall jet as
compared to the free jet can only be a consequence of the presence of the wall, because all
other parameters are kept the same. Baydar & Ozmen 0 and ODonovan & Murray 0, while
studying impinging jets, also reported similar differences between confined and unconfined
configurations. They reported that confining the jet increases the length of the potential core,
and explained this behaviour by the wall limiting the entrainment of ambient air only on one
side. By restricting the entrainment, the rate of spreading decreases leading to an increase in
the potential core length for the wall jet.
2.2.2 Vertical Velocity Profile for Non-Buoyant Wall Jet
In the case of a free jet, it was shown that there exists a unique self-preserving flow that
represents the whole vertical velocity profile after the flow development region. The behaviour of
the wall jet is slightly different due to the presence of wall. Two types of flow can be
encountered in the wall jet, a boundary layer and a shear layer type of flow, each of them
exhibiting their respective self-similar profile.
In the literature, there is a disagreement regarding the point where the flow changes from the
boundary layer type to a shear layer type of flow. As mentioned earlier, in this report, Glauberts
demarcation of layers will be used. This refers to the point where the flow attained the maximum
velocity, which is also the boundary layer thickness. Many researchers adopted this definition,
such as Launder & Rodi 0. Kurka & Eskinazi 0, on the other hand, found that the best
separation point between the two layers is where the Reynolds shear stress vanishes, but they
also stated that this method of determination is impractical.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 7
While there is an agreement amongst researchers concerning the self-similarity of the outer
layer, the establishment of a self-preserving inner layer is unclear. Some, such as Schwarz &
Cosart 0, believe that the velocity profiles in both inner and outer layers showed self-similarity
characteristics. But others, like Wygnanski et al. 0 and Swamy & Bandyopadyay 0, partially
disagree with this statement, refuting the self-similarity in the inner layer.
Figure 6 represents the vertical velocity profile plots at different locations behind the wall.
Similar to Figure 4, the non-dimensional coordinate of the abscissa is the distance away from
the wall jet axis divided by the distance where the velocity reaches half the axial velocity

b
y
,
and of the ordinate is the ratio of the local velocity to the axial velocity
m
U
U
.

Figure 6 CFD vertical profile comparison at different distances behind the wall jet

Unlike the free jet, which represented a unique pattern throughout the control volume, the wall
jet presents some different characteristics. The profiles only start to become similar at about
0.5m behind the wall jet. This confirms the previous discussion concerning the length of the
potential core. The length taken to reach a similar state represents the portion where the flow
undergoes the potential core and intermediate regions. This statement correlates with the
findings of Wygnanski et al. 0 who showed that the self-similarity is attained at longer distances
from the nozzle for the wall jet than for the free jet.
When the profile reached a similarity pattern, the boundary layer region is very small compared
to the shear layer region. Townsend 0 found that this region occupies between 1/10 to 1/5 of the
total wall jet velocity profile. Further away from the wall, all plots converge to zero, showing the
end boundary of the mixing properties with the ambient fluid.
Glaubert 0 was the first to derive an analytical solution for the wall jet problem, based on the
Blasius empirical power friction formula and on using matching and integrating techniques.
His work offers a discussion on this problem which will not be repeated here. Other researchers
also presented empirical formulae to represent the self-similarity of both profiles, such as
Newman et al. 0 and Schwarz & Cosart 0, but all failed to do so because the inner and outer
layer have completely different self-similar profiles. This point will be discussed more thoroughly
in the next section.
The CFD results were also compared with the Forthmann experimental results given by
Rajaratnam 0, who used only data from 0.6m behind the jet to show the vertical velocity pattern
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

8 EEC/SEE/2007/004
at different distances downstream, confirming the previous discussion on the length of the
potential core. Figure 7 shows that the general velocity profile is well predicted by the CFD
simulation at different distances behind the jet. Slight differences are only found near the wall
and further away from the wall, and these will be explained separately.


Figure 7 CFD and experimental results for the vertical velocity profile

The differences near the wall come from the fact that the wall properties are unknown in
Forthmanns experiment, such as the surface roughness which is an important property
especially near the wall where its effects are significant 0.
The differences at the end of the shear layer may come from the fact that the experiment
instruments were quite basic. As explained in a previous report 0, Forthmanns experiment
failed to comply with three criteria for assessing experimental data. In addition to this, the
results were taken very near the jet exhaust where the flow is supposed to exhibit considerable
scatter at its edges, as in the results of the CFD simulation 0.
Similar to the free jet study, it is possible to analyse the spreading parameters affecting the flow
in the case of the wall jet. Figure 8 shows a comparison between the spreading parameters for
a free and a wall jet. The non-dimensional coordinate of the abscissa is the distance behind the
jet divided by the jet diameter

0
b
x
, and of the ordinate is the length where the velocity attained
half of the axial velocity divided by the jet diameter

0
b
b
. It was found in 0 that the plots are
linear and, as a consequence, can be characterised by a linear trend line in the form:

+ =
2
0
1
0
K
b
x
K
b
b

When 0
0
=
b
b
, the approximate position of the virtual point is given by
1
2
K
K
while
0
1
b
K
gives a
measure of the spreading rate the wall jet is subject to.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 9

Figure 8 Angle of spread comparison between free and wall jets

Considering first the position of the virtual point, its location is situated behind the actual origin
of the wall jet and the value
1
2
K
K
gives a magnitude of 9.75 diameters. This compares very
well with the findings of Rajaratnam 0, who predicted the location of the virtual origin to be
around 10 diameters behind the nozzle.
Although the location of the virtual origin is accurately predicted by the CFD results, the angle of
spread is slightly under predicted with a difference of about 11.5% below the value given by
Rajaratnam 0 ( ) x b 068 . 0 = . However, there are uncertainties regarding this parameter. Launder
& Rodi 0 found a growth rate in the order of 002 . 0 073 . 0 but Tachie et al. 0 found this value to
be much higher. This value can vary from 0.085 to 0.09 at low Reynolds numbers. The position
of the virtual origin is clearly different for the free and the wall jet, with the wall jet located further
behind the nozzle compared to the free jet.
The spreading angle is more pronounced in the case of the free jet than in the wall jet. Launder
& Rodi 0 predicted this behaviour and showed that the growth rate of the free jet is more than
30% above that of the wall jet. Similarly, Rajaratnam 0 stated that the wall jet growth is about
0.7 times that of the free jet. Tangemann et al. 0 explained this behaviour by highlighting the
walls influence as the main reason for the reduction in the spreading growth as compared with
the free jet. This confirms the wall effect on increasing the potential core length.
As in the previous report 0, the existence of streamwise vortices can be proved in the case of a
wall jet by both the effects of the flow encountering the surrounding irrotational fluid and the
presence of the wall. This is illustrated in Figure 9, which presents isolines for the streamwise
vorticity contours.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

10 EEC/SEE/2007/004

Figure 9 Vorticity contour around z-axis with isolines (1/s)

The figure shows that there is no symmetrical pattern as in the free jet case, only the upper side
mixing with the ambient fluid resembles slightly the figure obtained for the free jet. The points
highlighted in the previous report are still valid:
The potential core is characterised by lower vortices inside it.
The boundary of the jet spreading can be clearly seen as a separation between
rotational and irrotational flows.
The vortices are found to be counter-rotating but unlike the free jet, the negative rotation around
the z-axis is due to the wall.
Townsend 0 explained that the presence of a solid boundary restricts the vortices created by
the shear layer to sizes which, in comparison to the free jet, are smaller thus causing a
reduction in the dissipating length scale. Figure 10 shows the vortices around the z-axis at
different distances behind the wall jet.

Figure 10 Z-Vorticity profile comparison at different distances behind the non-buoyant wall jet
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 11
It can be seen that, near the wall, the magnitude of negative streamwise vortices is very high
compared to the positive vortices created by the shear layer. These negative vortices around
the z-axis remain very high at large distances behind the wall jet, whereas the effect of the
positive vortices decreases downstream. The findings of Aloysius et al. 0 on the structure of
contra-rotating vortices formation and their development, also discussed in 0, are still valid here.
Another interesting point concerns the location of the maximum positive vorticity amplitude;
which are shown to be going into the wall jet axis. The free jet, on the other hand, was found to
go away from the jet axis 0. It was explained in 0 that this behaviour is related to the widening of
the jet. This influence can clearly be seen here; this pushing down effect restricts the widening
of the jet or, in other words, the growth rate. The wall plays an important role in the creation of
high magnitude negative vortices, thus gradually influencing the positive vortices which, as a
consequence, reduce the jet growth rate.
The location of the point of maximum velocity was found to vary behind the wall jet. To show
this variation, Figure 11 represents the height of the maximum velocity at different distances
behind the wall jet. The coordinate of the abscissa is the distance behind the jet divided by the
jet diameter

0
b
x
, while that of the ordinate is obtained by subtracting the vertical distance
where the velocity is maximum by the jet radius ( ) 2 /
0
b y
m
.
The figure shows that the location of the maximum velocity starts below the centreline axis of
the jet but, at a distance of about 29 diameters, this position starts to rise. Overall, the position
of the maximum velocity is below the centreline axis whereas in the free jet the position was
found to be at the centreline axis. This clearly shows the important effect of the wall on the flow
pattern. The pushing down mechanism can be seen here and is due to the high magnitude of
negative vortices and decreasing positive vortices around the z-axis, which pushes down the
flow near the wall, confirming the reduction in the spread rate and the increase in the potential
core length when compared to the free jet.


Figure 11 Height of the plume centreline as a function of downwind distance

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

12 EEC/SEE/2007/004
2.3 Buoyant Wall Jets
This section aims to compare the results obtained in the previous non-buoyant configuration
with a similar buoyant flow. Such condition is possible by adding a temperature difference and
assigning a release of different species and density magnitude.
The control volume and the boundary conditions were kept the same, except on the jet exhaust
where the velocity magnitude was maintained at 35m/s but this time releasing a mass fraction of
NOx ( )
4
10 37 . 0

at a temperature of 690K. The simulation was run for the same time step as
the non-buoyant case.
2.3.1 Axial Velocity Profile for Buoyant Wall Jet
A comparison between the maximum velocity for the buoyant and non-buoyant wall jets can be
seen in Figure 12. The decay rate for the buoyant wall jet is greater than for its non-buoyant
counterpart. All plots share a very similar pattern near the jet exhaust, but far downwind
divergences occur. The decay of the buoyant jet follows the theoretical curve upstream but
completely diverge from it further behind. The non-buoyant wall jet, on the other hand, follows
the theoretical rate of decay as seen in Figure 4.


Figure 12 Maximum axial velocity decay profiles with and without buoyancy

The length of the potential core for the buoyant wall jet was found to be about 6.5 jet diameters,
which is about 1 diameter shorter than that of the non-buoyant wall jet. This compares well with
the free jet results of 0, where the length of the potential core was 4.9 diameters for the buoyant
free jet and 5.6 diameters for the non-buoyant free jet. This difference was attributed to the
buoyancy effect on the flow pattern. Appendix A shows a comparison between the results for
free and wall jets, with or without buoyancy, with their respective theoretical profiles.
Figure 13 presents the velocity contours for the buoyant wall jet, which can be compared with
Figure 2 for the non-buoyant wall jet. The upstream region is very comparable to the non-
buoyant simulation, as shown in Appendix B. The potential core region can be recognized by its
triangular shape right at the exit of the nozzle creating the potential core length discussed
previously. In the fully developed region, the profile has the same configuration: a boundary
layer profile created by the wall and a shear layer profile formed by the surrounding fluid.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 13


Figure 13 Velocity contours for buoyant wall jet

Unfortunately, the control volume of this simulation is not wide enough to show plume rise
results, but it will be shown later in the 3D section of the report that the plume will actually lift-off
and rise due to buoyancy.
2.3.2 Vertical Velocity Profile for Buoyant Wall Jet
The vertical velocity profiles at different distances behind the jet for the buoyant wall jet are
presented in Figure 14. All the curves have different patterns from the nozzle exit to about 0.5m
behind the jet, whereas after that distance they all display a similar profile. As discussed
previously, this is due to the flow still being in the potential core region or in the intermediate
region. After 0.6m behind the jet, the figure shows that the profiles are very similar throughout
the vertical distance.

Figure 14 Velocity profiles with buoyancy at different distances behind the wall jet
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

14 EEC/SEE/2007/004
Appendix C compares the CFD results for the buoyant and non-buoyant wall jets. The ordinate
was changed to a dimensional form to highlight the differences. The analysis can be done in
two different sections. The first section is the boundary layer; except for the 0.6m behind the jet
which shows a shorter length for the buoyant wall jet than the non-buoyant case, the other
sections show approximately the same boundary layer thickness.
Secondly, the shear layer behaves very differently for the buoyant and the non-buoyant cases.
Although the point where the velocity ratio reaches zero is shorter for the buoyant simulation
than the non-buoyant simulation, it can be seen that this position becomes closer to the non-
buoyant jet and has almost the same velocity ratio at 0.2m above the wall. This point was
discussed in the previous report 0 as a confirmation of buoyancy effects acting on the flow.
As in the case of the non-buoyant wall jet, the analysis of different locations where the velocity
reaches half the axial velocity can be made by plotting these at different distances behind the
wall jet.
Figure 15 shows a comparison between the spreading parameters for the buoyant and non-
buoyant simulations. Even though the upwind region is fairly similar for both simulations, the
downwind region is very different, with the buoyant simulation showing higher values of the
position where the velocity reached half of the maximum velocity. This happens after about 30
diameters behind the jet nozzle, and is due to the buoyancy effect on the flow. The same
behaviour has been reported in 0 and the relevant figure is presented in Appendix D for
comparison.


Figure 15 Spread b comparison for wall jet with and without buoyancy

As mentioned earlier on the report, the spread b as a function of distance is much lower for the
wall jet than for the free jet simulation without buoyancy. The addition of buoyancy does not
change this pattern; spread b is still higher for the free jet than for the wall jet at large
downwind distances.
Another interesting point concerns the location where the buoyant curve separates from the
non-buoyant curve, for free and wall jet simulations. This was found to be about 20 diameters
for the free jet and 30 diameters for the wall jet. This means that the wall jet takes longer to go
from a jet like to a plume like behaviour as compared with the free jet simulation, confirming the
potential core length discussion.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 15
Figure 16 presents a comparison of the height of the plume centreline for the buoyant and non-
buoyant simulations, at different positions behind the wall jet. The axes are kept the same as in
Figure 11.

Figure 16 Comparison of the height of plume centreline as a function of downwind distance

Figure 16 shows that both simulations follow the same pattern near the jet exhaust, whereas
further downwind differences clearly start to appear. The buoyant wall jet near the end of the
control volume exhibits a sudden rise of the height of the plume centreline, while in the non-
buoyant case this rise progresses at a lower rate. All locations of the plume centreline are below
the jet axis for the non-buoyant simulation while the buoyant simulation shows a rise above it
further downwind, confirming the lift-off of the plume due to buoyancy. This point will be
discussed in greater detail for three-dimensional simulations.
The reason why the height of the plume centreline is kept below the jet axis for such long
distance can be explained by the Coanda effect. Ramsdale & Tickle 0 explained this effect as
the tendency of the jets to attach themselves to nearby solid boundaries. Sharp & Vyas 0
added that the flow clings to the ground for large distances before buoyancy takes over and
causes the jet to lift and rise above the ground.
An analogy was made in the non-buoyant study between the roles of the vortices in the pushing
down effect of the flow near the ground. It is interesting to see whether the same observations
can be made for the buoyant wall jet simulation.
Figure 17 shows Z-Vorticity profiles at different distances downwind. Similar to the non-buoyant
case shown in Figure 10, the location of the maximum positive vorticity amplitude goes down
into the wall jet axis affecting the widening of the jet. This explains why the growth rate of the
buoyant free jet is higher than the buoyant wall jet. As a matter of fact, the high magnitude of
negative vortices, which are created due to the presence of the wall, gradually influences the
positive vortices created by the entrainment with the ambient fluid leading to a pushdown of the
flow. As a consequence, the length of the potential core is increased while, at the same time,
the growth rate is decreased.

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

16 EEC/SEE/2007/004

Figure 17 Z-Vorticity profile at different distances behind the buoyant wall jet

With the initial results and discussions from this two-dimensional study, it is now possible to
advance with confidence to three-dimensional simulations. A more realistic scenario can be
adopted, with a jet close to the ground and a co-flowing velocity to replicate the headwind flow.
3 THREE-DIMENSIONAL STUDY OF TURBULENT WALL JET
The simulation in this section involves a round jet of 0.93m diameter very close to a wall, in an
ideal co-flow headwind configuration. First an isothermal simulation is carried out; then, the
same geometry will be simulated with buoyancy effects. The CFM56-3C jet engine was selected
because it has already been studied in previous reports 0 and 0.
A transient simulation was conducted to monitor the progress of the flow throughout the control
volume. The geometry, mesh distribution and boundary conditions are discussed in the
following section.
3.1 Geometry and Boundary Conditions
The computational domain is rectangular, composed of 104 different volumes of different mesh
densities. This fine decomposition of the control volume was done to optimise the mesh
distribution. The near wall region including the jet and the first few 100 meters, where the
buoyancy effects are expected to be most significant, was set up as fine as possible. The
overall geometry is presented in Figure 18, and extends to
0
6 . 537 b ,
0
5 . 64 b and
0
86b in the x, y
and z directions, respectively, with
0
b the jet diameter. The total mesh density for this problem
comprises about 3,965,760 nodes.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 17

Figure 18 Geometry and mesh distribution of the 3D wall jet

The boundary conditions are as follows:
The jet exhaust is set up as a velocity inlet with velocity magnitude 80m/s, releasing hot
gases (690K) of NOx emitted at a mass fraction of
4
10 74 . 0

, as described in 0.
The faces adjacent to the exhaust are also defined as velocity inlets, but with a different
velocity of magnitude 2.5m/s to replicate the co-flow condition of a headwind. The co-
flow is adiabatically stratified across the domain with a temperature of 290K.
The bottom wall is setup as stationary with surface roughness of 0.003m, corresponding
to the concrete PCN 60 R/B/W/T used at Zurich airport. Werner and Wengles near wall
treatment was used to account for the boundary layer formation when the mesh is
coarse.
The external boundaries are defined as symmetry walls, as they are located far away
from the jet.
The face opposite to the jet is the outflow of the control volume.
The Navier-Stokes equations of fluid flow are solved by the Large Eddy Simulation (LES)
technique.
3.2 Mean Velocity Profile Comparison
3.2.1 Streamwise Direction Analysis
This section aims to compare the results of the co-flowing wall jet with the co-flowing free jet,
both with buoyancy, to assess the effects on the fluid mechanics when the wall is added.
Results of the two-dimensional simulations have shown that the decay rate for the free jet is
greater than the for wall jet, while the potential core is much longer for the wall jet than for the
free jet. These trends are confirmed in the three-dimensional simulations, as shown in Figure
19.
537.6b
0

64.5 b
0

86 b
0

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

18 EEC/SEE/2007/004
An analysis of the results for the first time step shows that there is a slight difference in the
length where the free and wall jet both reach 0
5 . 2
0
=
U
U
m
. The free jet simulation attained this
value faster than the wall jet simulation; on the other hand, the potential core is much shorter
and its decay rate is much greater for the free jet simulation than for the wall jet simulation. The
following time step also shows a similar behaviour upwind, but with a more elongated pattern
because the fluid is still progressing through time and has not yet attained a steady state
condition.

Figure 19 Maximum velocity decay comparison between free and wall jet

Figure 20 shows a comparison of the mean velocity profile for the simulation of the free jet (left)
and the wall jet (right), at different times.
The first time level (after 1s) shows a much further penetration for the wall jet than for the free
jet. The free jet still exhibits an almost symmetrical pattern, as in the case of a non-buoyant free
jet. The wall jet, on the other hand, shows a different pattern with the fluid rising at some
distance behind the jet exhaust.
Buoyancy effects can clearly be seen in the second time level (after 5s), when the free jet
velocity profile shows a deviation from the centreline axis. The wall jet has approximately the
same pattern as in the previous time step, with the flow rising much higher than for the free jet
simulation. The flow penetration is also much deeper for the wall jet than for the free jet
simulation, as stated earlier.
The third time level (after 10s) continues the previous trends, with greater flow penetration and
higher rise for the buoyant wall jet than for the buoyant free jet.
Another interesting point that can be observed for the wall jet is that, at large distances behind
the exhaust, the flow separates from the wall. Although not so apparent, this effect can already
be seen after 1s, at about 28m behind the exhaust, and is clearly seen after 10s at about 60m
behind the exhaust. This lift-off effect will be discussed in more detail in the next section.




ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 19
After 1 second

Free jet

Wall Jet
After 5 seconds

Free jet

Wall Jet
After 10 seconds

Free jet

Wall Jet

Figure 20 Mean velocity profile evolution through time for buoyant free and wall jets
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

20 EEC/SEE/2007/004
The streamwise velocity profile at different distances behind the exhaust for the buoyant wall jet
is shown in Figure 21. This figure is very comparable to Figure 14 in section 2.3.2. From the exit
of the nozzle to a distance of 7.5m, the wall jet is still in the potential core and intermediate
regions, whereas from this distance onwards the wall jet is in the fully developed region and
exhibits a self similarity profile.
Forthmanns experimental results are used to compare the self-similar profile. It can be seen
that all the plots after 7.5m behind the exhaust closely follow the experimental results, both in
the boundary layer and the free shear layer parts.


Figure 21 Streamwise velocity profiles at different distances behind the buoyant wall jet

It was previously argued that the profile of the free shear layer is comparable with the free jet
situation; Appendix E shows this comparison with Forthmanns experimental data for the wall jet
and the theoretical results of Tollmien, Goertler and Bradbury for the free jet. It can be seen that
they all have the same pattern; the CFD results follow very closely the experimental curve to
about
b
y
75 . 1 , whereas further away Tollmiens theoretical results are the closest to the CFD
results. A comparison between the co-flow and stagnant condition shows that the boundary
layer thickness is greater in a co-flowing situation. This is in line with the findings of Launder &
Rodi 0, who explained this phenomenon as a consequence of the presence of the external
stream weakening the relative strength of the free shear layer. This, in turn, has less impact on
the wall boundary layer, making it thicker.
A comparison between the spread b as a function of distance of the free and wall jets, at two
different time levels, is given in Appendix F. Both graphs show the same general pattern as for
the 2D simulations, section 2.2.2. The rate of spread is greater for the free jet than for the wall
jet. As mentioned earlier, several authors such as Launder & Rodi 0 had already obtained this
behaviour in a non-buoyant situation, and claimed that the growth rate of the free jet is more
than 30% higher that of the wall jet. As can be seen in Appendix F, this trend tends to grow as
the flow progresses through time. It is interesting to note that the growth rate is actually
increasing with time for the buoyant free jet simulation, whereas for the buoyant wall jet the
growth rate is actually decreasing. It will be shown later in the report that this behaviour is due
to the vortices generated by the presence of the wall; this is true for some distances behind the
exhaust but, at larger distances, the wall effects decrease and the buoyancy effects will take
over and lift jet above the ground.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 21
While studying the 2D wall jet, another influence of the wall was found on the maximum height
of the plume centreline. To understand it better in a 3D situation, Figure 22 gives a comparison
of the maximum height of the plume centreline as a function of downward distance. The
ordinate axis was changed to the non-dimensional form
m
y
b y
0

, with
m
y the maximum height of
both simulations.


Figure 22 3D comparison of height of plume centreline at different distances behind the exhaust,
after 10s

Figure 22 shows a different pattern for both simulations; the rise above the actual centreline
axis of the jet is much faster for the free jet than for the wall jet simulation. The buoyant wall jet
starts rising above the centreline at about 55 diameters; before this point, the plot never reaches
a positive value. As a matter of fact, it is either approaching the centreline value (0) or stays
negative, meaning the flow is going down. This behaviour was mentioned earlier in the 2D part
of the report, as the clinging phenomenon due to the Coanda effect.
Although in a steady-state condition, Sharp & Vyas 0 found a crude linear relation between the
relative distance
0
b
L
over which the flow stays attached to the ground and the densimetric
Froude number F :
c F
b
L
=
0

with 2 . 3 = c found experimentally and L the length of cling. For this simulation, at the time
when it reaches a steady-state condition, the length of cling should be 166m behind the
exhaust. At the time 10s, the length of cling in Figure 22 is about 55 diameters, which is in the
region of the plume lift-off shown in Figure 20.
A comprehensive literature review on the lift-off phenomenon was carried out by Ramsdale &
Tickle 0 in their study of ground-based buoyant clouds. One interesting parameter when
studying the lift-off of a buoyant gas is the Richardson number
p
L , given by:
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

22 EEC/SEE/2007/004


=
a
p
u
gH
L

2

with H the effective depth,

u the friction velocity, g the gravitational acceleration and the


density difference between the ambient fluid and the jet.
Briggs was the first to propose an approximate value for
p
L in the order of 2.5 to 10, but he
recognized himself that these values are not entirely correct. Hanna et al. 0 reviewed wind
tunnel studies and concluded that this value was raised to about 20, with an uncertainty of a
factor of two. Even these experimental results are not very reliable as the technique used to
obtain the data is very basic, Meroney 0 relied on the average optical judgement of observers
whereas Sinclair et al. 0 found this distance by linear interpolation between the neighbouring
vertical cross sections of their measured temperature field.
Sinclair et al. 0 collected some of the parameters that play an important role in the plume lift-off;
these include the jet temperature, the jet exhaust velocity and the jet orifice geometry. They
agreed with Ramsdale & Tickle 0 on the influence of the jet to ambient velocity ratio and the jet
geometry aspect ratio on the plume lift-off. A higher jet to ambient velocity with a low aspect
ratio increases the clinging length, and hence delays the lift-off, whereas a higher aspect ratio
shortens the lift-off from the ground.
Another point of interest to many researchers relates to the importance of vortical structures,
especially in the streamwise direction, because their strength governs the lift-off of the buoyant
wall jet from the ground 0. The vortical profile section investigates the influence of vortices on
the flow pattern and compares the results with the previous ones for the buoyant free jet
simulation 0. Before attempting this, it is also interesting to observe the fluid mechanics in the
spanwise direction as this is very relevant to plume dispersion.
Contours of mean concentration of NOx are presented in Appendix G.
3.2.2 Spanwise Direction Analysis
The velocity profile comparison in the spanwise direction at different time steps is given in
Appendix H; the velocity scales are kept the same for both simulations. The plane shown is cut
through the centre of the jet axis at 0.465m above the ground for the wall jet simulation.
It can be seen at the first time level (after 1s) that both simulations show very similar velocity
contour patterns. Both free and wall jets follow a somewhat symmetrical distribution, and both
display almost the same dispersion length for this time level. As mentioned in the previous
report 0, buoyancy effects at this time are not yet relevant and differences only start to appear
at later times. After 5s, the free jet simulation shows a narrower dispersion than the wall jet
simulation, and the penetration length is deeper in the wall jet simulation than the free jet.
The final time level, after 10s, is the most interesting as some of the previous discussion still
applies here. The symmetrical pattern is conserved in both simulations, and the wall jet shows a
wider spread of velocity contours than the free jet simulation. It is only at this time that the
buoyancy effects show their true influence on the flow field. As mentioned earlier, the
penetration of the velocity contours is much deeper for the wall jet simulation than the free jet
simulation. The buoyancy effects for the free jet simulation have taken place at a much shorter
distance than the wall jet simulation, about 60 to 65 m behind the wall jet. This correlates well
with the distance of lift-off from the ground found previously.
To illustrate the symmetrical pattern of the flow in the spanwise direction, Figure 23 shows the
velocity profile at different distances behind the jet. Theoretical curves of Goertler, Tollmien and
Bradbury are also compared with the CFD results; their description and derivation can be found
in the previous report 0.

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 23

Figure 23 Spanwise velocity profile comparison of wall jet with theoretical curves of the free jet
after 10s

The wall jet results show a profile similar to the free jet reported in 0. The characteristic
Gaussian distribution of the free jet simulation can be seen in this figure. The curves are
different for the first two distances behind the exhaust (2.5m and 5m) because the jets are still
in the potential core and the flow development region.
All the plots are very much alike after 5m, showing a self-similar pattern. A comparison between
the theoretical and CFD results reveals that, in the inner region, they are very much similar but
some differences occur in the outer layer. Similar to the results in 0, this was attributed to the
mixing with the ambient surrounding occurring in this region.
The spread b as a function of distance can be plotted using the data for the velocity profile.
Figure 24 shows a comparison of the spread b between the free and wall jet simulations. As
mentioned earlier in the comparison of the horizontal velocity contours, even though upstream
the rate of spread seems to be higher for the free jet, further downwind this situation is reversed
and the wall jet grows more rapidly.
Another comparison can be made concerning the wall jet simulation with the one presented in
Appendix F. It can be seen very clearly that the vertical spread of the wall jet is very much
smaller than the spanwise spread. This observation was also made by several authors, such as
Rajaratnam 0 and Swamy & Bandyopadyay 0. They all found that the growth rate of the wall jet
in the spanwise direction is much larger than in the vertical direction. This difference in rate of
spread leads to the formation of distinct virtual origins. In the vertical direction presented in
Appendix F, the virtual origin can be clearly seen to be situated upstream of the actual position
of the jet exhaust. But in the spanwise direction in Figure 24, if the graph would have started at
8 diameters (lower than this distance corresponds to the potential core and flow development
region), it can be seen that the virtual origin would have been situated this time far downstream
of the actual position of the jet exhaust. This discussion confirms the findings of Swamy &
Bandyopadyay 0, who found this wide separation to be about 12.5 diameters between the two
virtual origins; the vertical one being situated upstream of the orifice and the spanwise one
located downstream of the orifice.

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

24 EEC/SEE/2007/004

Figure 24 Comparison of spanwise direction of spread b as function of distance after 10s

3.3 Vorticity Profile Comparison
In order to better understand the role vortices play in the dispersion process of the buoyant wall
jet, contours of instantaneous velocity profile were taken and compared with the equivalent free
jet ones 0. The scales are kept the same and the addition of an iso-surface of 80m/s is
illustrated with its grid (black part).


Buoyant free jet

Buoyant wall jet
Figure 25 Instantaneous velocity profile comparison after 5s
Comments made in 0 are still valid for the buoyant wall jet simulation, as a localised area of
lower velocity is also apparent here. These, like in the buoyant free jet, are located outside the
jet domain.
The puffing phenomenon can be seen in Figure 25, and is also closely linked to the vorticity
parameter. As mentioned in 0, the puffing pattern is created by the entrained eddies breaking
the fluid. These entrainment vortices are situated in the local areas where the jet velocity is
lower than the surrounding velocity.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 25
There are several other points that can be drawn out of this instantaneous velocity profile
comparison. Firstly, it can be seen from Figure 25 that the potential core is much longer for the
buoyant wall jet than for the buoyant free jet. This statement can be made by analysing the
velocity iso-surface of 80m/s. While in the wall jet simulation there is one puff extending to about
7.5m behind the exhaust, the free jet shows its last puff to be at a distance less than 5m. This
confirms the previous discussion on the length of the potential core, and that the wall jet has
deeper penetration behind the exhaust than a similar free jet.
Finally, Figure 25 gives an insight on the spreading rate. It can be noticed that, at large
distances behind the exhaust, the velocity contours go much higher for the free jet than its wall
jet counterpart. At about 20m behind the jet exhaust, the velocity contours reach 4m in the
positive streamwise direction for the free jet whereas it stays below it for the wall jet. This
statement is also in line with the discussion on the rate of spreading, and confirms that the
growth rate is higher for the free jet than the wall jet.
The puffing phenomenon was earlier attributed to the entrainment eddies located outside the jet
domain. The vorticity magnitude contours presented in Figure 26 show exactly this, with vortices
of higher magnitude concentrated near the jet radius and the wall for the wall jet simulation. The
vortices magnitude is much higher for the wall jet than for the free jet simulation, especially near
the jet exhaust.


Buoyant free jet

Buoyant wall jet
Figure 26 Vorticity magnitude comparison after 5s

It can also be seen that the vortices are more numerous for the free jet than for the wall jet
simulation. This observation was also made by Townsend 0, who attributed this behaviour as
the consequence of the presence of a solid boundary.
Focusing on one vorticity component, the rotation around the z-axis, also known as the
streamwise vorticity, found to be the most interesting for this study, Appendix I shows the
evolution of these streamwise vortices at different time steps for the buoyant free and wall jets.
Similar to the 2D simulation, it can be seen that counter-rotating vortices are formed right at the
jet exhaust. The positive vortices are caused by the shearing mechanism of the jet flow
encountering the moving surrounding fluid but, unlike the free jet simulation, the negative
vortices present at the bottom are due to the wall. This pattern resembles the one presented by
Launder & Rodi 0 in their review paper on three-dimensional wall jets.
It can be seen, at all time steps that the merge between the positive and negative vortices
happens further away from the jet exhaust for the wall jet than for the free jet. It can be recalled
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

26 EEC/SEE/2007/004
from the previous report 0 that this merge corresponds to the disappearance of the potential
core and, from this comparison, it can be seen that the potential core is much longer for the wall
jet than the free jet. As previously mentioned in this report, this has an effect on the penetration
through the control volume and the rate of decay of the maximum velocity. The wall jet
simulation shows a deeper penetration and lower decay rate through the control volume.
The vortices evolution after the merging is somewhat similar to the free jet simulation but, as
can be seen on Appendix I, the number of vortices is lower in the wall jet simulation than the
free jet simulation, confirming the statement made by Townsend 0. The vortices first undergo a
coalescence cascade before breaking up, the only difference being that the influence of the
negative vortices created by the solid boundary still remains important. This has an effect on the
flow pattern in terms of keeping it down. As shown previously, the negative vortices force the
flow to cling to the wall, exhibiting the Coanda effect at a very large distance.
After the clinging length, the flow was found to lift-off from the ground with large gusts as found
by Sharp & Vyas 0. Sinclair et al. 0 discussed the importance of the streamwise vortices in the
role they play in lifting-off the wall jet from its solid boundary. In that sense, Figure 27 shows
that the intensity of the vortices is much higher in the case of the wall jet, and these are closer
to the ground compared to the free jet simulation.


Buoyant free jet

Buoyant wall jet
Figure 27 Streamwise vorticity comparison between buoyant free and wall jets
4 CONCLUSIONS
Following a previous ALAQS report on a CFD comparison between buoyant and non-buoyant
turbulent free jets 0, the problem of near field simulation in the presence of a horizontal solid
boundary very close to the jet exhaust is now presented. This report is first aimed at validating
the CFD technique for a plane wall jet using experimental and theoretical results available in the
literature. Then, by increasing the complexity of the computation to account for buoyancy, a
direct comparison was carried out to assess the impact of the ground and to compare the fluid
mechanics of buoyant free and wall jets.
The results of the 2D and 3D non-buoyant wall jet simulations were found to agree with
classical results, successfully replicating both the boundary layer profile created by the wall and
the free shear layer profile generated by the ambient fluid.
The comparison between the buoyant free and wall jets revealed several differences: First, it
was shown that the potential core region is much longer for the wall jet than for the free jet. This
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 27
has an effect on the flow penetration through the control volume. As shown in the report, the
wall jet simulation offers a deeper penetration than the free jet case. A correlation can be found
between this parameter and the maximum velocity decay as the penetration involves higher
velocity pushing into the control volume. It was found that the maximum velocity decay is much
faster for the free jet than for the wall jet. Finally, the spreading rate was found to be higher in
the case of the buoyant free jet simulation.
All the parameters discussed above are interconnected with the streamwise vortical structure of
the buoyant wall jet. As in the case of the buoyant free jet, counter rotating vortices are created
on one side by the surrounding fluid and on the other by the solid boundary. What is different
from the free jet situation is the presence of the wall generating vortices whereas the influence
of the vortices created by the surrounding fluid gradually decreases.
The first point of merging of the counter-rotating vortices occurs in the potential core length. As
the flow progresses, the intensity of the vortices created by the wall is much stronger than the
ones created by the surrounding fluid, causing a clinging of the flow, also known as the Coanda
effect. This pushing down phenomenon restricts the growth of the jet, hence a lower rate of
spread than for the buoyant free jet. As the velocity further away from the jet exhaust
decreases, the vortices created by the wall decrease and buoyancy takes over, with positive
vortices lifting up the flow from the ground.
The ability of the CFD model to reproduce experimental and theoretical results, as well as
providing an in-depth knowledge of the fluid mechanics, gave confidence to progress to a more
realistic configuration of the engine geometry with its wing for a typical type of airplane, to
consider important properties on the interaction between both components during different
stages of the take-off and landing phases. This type of study will help improving the knowledge
of near-field source dynamics and will provide important information for dispersion modelling
using standard Gaussian or Lagrangian techniques.

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

28 EEC/SEE/2007/004
5 REFERENCES
Launder, B.E. & Rodi, W., (1983), The Turbulent Wall Jet-Measurements and Modelling.
Annual Review of Fluid Mechanics, Vol. 15, pp. 429-459.
Forthmann, E., (1936), Turbulent Jet Expansion. English translation, NACA Technical
Memorandum TM-789.
Aloysius, S. & Wrobel, L.C., (2007), CFD Comparison of Buoyant and Non-Buoyant Turbulent
Jets, Eurocontrol Report, EUROCONTROL Experimental Centre.
Glaubert, M.B., (1956), The Wall Jet. J. Fluid Mech., Vol. 1, pp. 625-643.
Rajaratnam, N., (1976), Turbulent Jets, Series in Developments in Water Science, Volume 5,
Elsevier.
Tangemann, R. & Gretler, W., (2000), Numerical Simulation of a Two-Dimensional Turbulent
Wall Jet in an External Stream. Forschung im Engenieurwesen, Vol. 66, pp. 31-39.
ODonovan, T.S. & Murray, D.B., (2007), Jet Impingement Heat Transfer- Part 1: Mean and
Root-Mean-Square Heat Transfer and Velocity Distributions. International Journal
of Heat and Mass Transfer, Vol. 50, pp. 3291-3301.
Sforza, P.M. & Herbst, G., (1970), A Study of Three Dimensional, Incompressible, Turbulent
Wall Jets. AIAA Journal, Vol. 8 (2), pp. 276-282.
Abramovich, G.N., (1963), The Theory of Turbulent Jets, MIT Press.
Curd, E.F., (1981), Possible Applications of Wall Jets in Controlling Air Contaminants. Annal.
Occup. Hyg., Vol. 24(1), pp.133-146.
Baydar, E. & Ozmen, Y., (2006), An Experimental Investigation on Flow Structures of Confined
and Unconfined Impinging Air Jets. Heat Mass Transfer, Vol. 42, pp. 338-346.
Kruka, V. & Eskinazi, S., (1964), The Wall-Jet in a Moving Stream, J. Fluid Mech., Vol. 20, pp.
555-579.
Schwarz, W.H. & Cosart, W.P., (1961), The Two-Dimensional Turbulent Wall-Jet. J. Fluid
Mech., Vol. 10, pp.481-495.
Wygnanski, I., Katz, Y., & Horev, E., (1992), On the Applicability of Various Scaling Laws to the
Turbulent Wall Jet. J. Fluid Mech., Vol. 234, pp. 669-690.
Swamy, N.V.C. & Bandyopadyay, P., (1975), Mean and Turbulence Characteristics of Three-
Dimensional Wall Jets. J. Fluid Mech., Vol. 71, part 3, pp. 541-562.
Townsend, A.A. (1976), The Structure of Turbulent Shear Flow. Cambridge Monographs on
Mechanics and Applied Mathematics, Cambridge University Press.
Newman, B.G., Patel, R.P., Savage, S.B., Tjio, H.K., (1972), Three-Dimensional Wall Jet
Originating from a Circular Orifice. Aeronautical Quarterly, Vol. 23, pp. 188-200.
Tachie, M.F., Balachandar, R. & Berstom, D.J., (2004), Roughness Effects on Turbulent Plane
Wall Jets in Open Channel. Experiments in Fluids, Vol. 37, pp. 281-292.
Aloysius, S., Pearce, D. and Wrobel, L.C., (2006), CFD Simulations of Emission Dispersion
from an Aircraft Engine during Take-off, Eurocontrol Report, EUROCONTROL
Experimental Centre.
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 29
Ramsdale, S.A. & Tickle, G.A., (2001), Review of Lift-off Models for Ground Based Buoyant
Clouds, AEA Technology Report AEAT-4262 (2) for EC URAHFREP Project.
Sharp, J.J & Vyas, B.D., (1977), The Buoyant Wall Jet, Proc. Institution of Civil Engineers, Vol.
63, Part 2, pp. 593-611.
Hanna, S.R., Briggs, G.A., Chang, J.C., (1998), Lift-off of Ground-Based Buoyant Plumes,
Journal of Hazardous Materials, Vol. 59, pp. 123-130.
Meroney, R.N, (1979), Lift-off of Buoyant Gas Initially on the Ground, Journal of Industrial
Aerodynamics, Vol. 5, pp. 1-11.
Sinclair, J.R., Slawson, P.R., Davidson, G.A., (1990), Three-Dimensional Buoyant Wall Jets
Released into a Coflowing Turbulent Boundary Layer, Journal of Heat Transfer,
ASME, Vol.112, Issue 2, pp. 356-362.

ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

30 EEC/SEE/2007/004
Appendix A Maximum Axial Velocity Comparison between
Free and Wall Jets






Appendix B Two-dimensional Jet Velocity Contours for
Non-Buoyant Wall Jet




ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 31
Appendix C Two-dimensional Velocity Comparison
between Buoyant and Non-buoyant Wall Jets










ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

32 EEC/SEE/2007/004









ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 33

Appendix D Angle of Spread Comparison for Free Jet with
and without Buoyancy



Appendix E Shear Layer Profile Comparison with
Experimental and Theoretical Data






ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

34 EEC/SEE/2007/004
Appendix F Angle of Spread Comparison between Buoyant
Free and Wall Jets

After 5 seconds


After 10 seconds








ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 35
Appendix G Mean NOx concentration contours at different
time steps
After 1 seconds

After 5 seconds

After 10 seconds
Appendix H Spanwise Velocity Contours Comparison

After 1 second
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

36 EEC/SEE/2007/004

Free jet

Wall jet
After 5 seconds

Free jet

Wall jet
After 10 seconds

Free jet

Wall jet
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

EEC/SEE/2007/004 37
Appendix I Comparison of Vortices around the z-axis
After 1 second

Free jet

Wall jet
After 5 seconds

Free jet

Wall jet
After 10 seconds

Free jet

Wall jet
ALAQS CFD Comparison of Buoyant Free and
Wall Turbulent Jets

38 EEC/SEE/2007/004





(This page intentionally blank)




For more information about the EEC Society, Environment and Economy Research Area please contact:

Ted Elliff
SEE Research Area Manager,
EUROCONTROL Experimental Centre
BP15, Centre de Bois des Bordes
91222 BRETIGNY SUR ORGE CEDEX
France
Tel: +33 1 69 88 73 36
Fax: +33 1 69 88 72 11
E-Mail: Ted.Elliff@eurocontrol.int

or visit : http://www.eurocontrol.int/

You might also like