0% found this document useful (0 votes)
10 views194 pages

Research-Enlighten@glasgow Ac Uk

This thesis evaluates the behavior of vortical flows and vortex breakdown over slender delta wings, highlighting the importance of understanding these phenomena for aircraft stability. It employs computational fluid dynamics (CFD) to assess transonic and unsteady flow conditions, comparing different turbulence models, including URANS and DES. The findings indicate that URANS models can effectively predict unsteady flow behavior at a lower computational cost than DES, despite some discrepancies with experimental data.

Uploaded by

Fernando
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
10 views194 pages

Research-Enlighten@glasgow Ac Uk

This thesis evaluates the behavior of vortical flows and vortex breakdown over slender delta wings, highlighting the importance of understanding these phenomena for aircraft stability. It employs computational fluid dynamics (CFD) to assess transonic and unsteady flow conditions, comparing different turbulence models, including URANS and DES. The findings indicate that URANS models can effectively predict unsteady flow behavior at a lower computational cost than DES, despite some discrepancies with experimental data.

Uploaded by

Fernando
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 194

https://theses.gla.ac.

uk/

Theses Digitisation:
https://www.gla.ac.uk/myglasgow/research/enlighten/theses/digitisation/
This is a digitised version of the original print thesis.

Copyright and moral rights for this work are retained by the author
A copy can be downloaded for personal non-commercial research or study,
without prior permission or charge
This work cannot be reproduced or quoted extensively from without first
obtaining permission in writing from the author
The content must not be changed in any way or sold commercially in any
format or medium without the formal permission of the author
When referring to this work, full bibliographic details including the author,
title, awarding institution and date of the thesis must be given

Enlighten: Theses
https://theses.gla.ac.uk/
research-enlighten@glasgow.ac.uk
Evaluation of URANS and DES Predictions
of Vortical Flows over Slender Delta Wings

Lucy Schiavetta M.Eng

Department of Aerospace Engineering


University of Glasgow
February 2007

Thesis submitted to the Faculty of Engineering


in fulfillment of the requirements for the
degree of Doctor of Philosophy

©Lucy Schiavetta
2007

^ MIC
ProQuest N um ber: 10390580

All rights reserved

INFORMATION TO ALL USERS


The q u a lity of this re p ro d u c tio n is d e p e n d e n t u p o n the q u a lity of the co p y su b m itte d .

In the unlikely e v e n t that the a u th o r did not send a c o m p le te m a n u scrip t


and there are missing p a g e s, these will be n o te d . Also, if m a te ria l had to be re m o v e d ,
a n o te will in d ic a te the d e le tio n .

uest
P roQ uest 10390580

Published by ProQuest LLO (2017). C o p y rig h t of the Dissertation is held by the A uthor.

All rights reserved.


This work is p ro te cte d a g a in s t u n a u th o rize d co p yin g under Title 17, United States C o d e
M icroform Edition © ProQuest LLO.

ProQuest LLO.
789 East Eisenhower Parkway
P.Q. Box 1346
Ann Arbor, Ml 4 8 1 0 6 - 1346
fG L A ^G O W ^
UNIVERSITY
LIBRARY:
Abstract

This thesis presents a study of vortical flows and vortex breakdown over slender delta wings. The formation of
leading edge vortices over a slender delta wing provide advantageous aerodynamic characteristics at low velocities
and angles of incidence. However, as the incidence of the planform is increased these vortices are affected by
changes in the flow behaviour, which causes tliem to become unstable and breakdown into an incoherent form.
This vortex breakdown is detrimental to the aerodynamic characteristics of the wing and can cause instability of
the aircraft. Due to this adverse effect, it is important to understand the behaviour of such flows. Computational
fluid dynamics is one method which, due to the development of numerical methods and the rapid advances in
computer technology, is becoming increasingly valuable for the analysis of vortical flows and vortex breakdown
and this thesis assesses the use of CFD to predict these types of flows. To perform this assessment two test cases
are considered under different flow regimes. The first test case considers transonic flow and is investigated using
steady state calculations and the second test case considers the unsteady subsonic behaviour of the flow.

The behaviour of the flow over slender delta wings under transonic conditions is highly complex. With the occur­
rence of a number of shocks in the flow, vortex breakdown is abrupt and the overall behaviour is quite different to
that for subsonic flow. To consider this, the flow over a 65" sharp leading edge delta wing is analysed for a transonic
Mach number of M = 0.85 at two incidences: a = 18.5" and 23". A Boussinesq based RANS turbulence model
is used which has been modified for vortical flows. The flow solutions are compared to existing experimental data
and show good agreement for the cases considered and a number of shock systems within the flow are identified.
However, a discrepancy with the experimental data is shown where the critical incidence for the onset of vortex
breakdown on the wing is under-predicted. A sensitivity study of the flow to a number of computational factors,
such as turbulence model and time accuracy, is undertaken. However, it is found that these parameters have little
effect on the overall behaviour of the transonic flow and the occurrence of vortex breakdown. From analysis of the
solutions, it is determined that the onset of vortex breakdown is highly dependent on the vortex strength and the
strength and location of the shocks in the flow. The occurrence of a critical relationship between these parameters
is suggested for vortex breakdown to occur and is used to explain the discrepancies between the computational and
experimental results based on the under-prediction of the vortex core axial velocity.

The second test case considers the unsteady behaviour of the vortex flow and vortex breakdown. Downstream of
the vortex breakdown the flow is highly unsteady with many different phenomena occurring for varying frequen­
cies. This unsteady behaviour can interact with the surface of the wing or with other surfaces of an aircraft, which
can cause fatigue or stability issues. However, this behaviour is still a subject, which is a challenge for numerical
methods, such as CFD. New approaches to turbulence modelling, such as detached eddy simulation (DES) have
been proposed which allow for greater realism of the numerical predictions. However, this increase in accuracy
comes with a considerable increase in computational expense compared to traditional turbulence modelling. Both
DES and Unsteady Reynolds Averaged Navier-Stokes (URANS) turbulence methods aie considered to predict the
unsteady vortical behaviour. Calculations using DES are initially considered and the predicted behaviour and res­
olution of the flow are analysed. Both temporal and spatial refinement are considered and the effect that each of
these factors have on the flow behaviour is examined. A number of unsteady flow features and their corresponding
frequencies are identified from the solutions. From comparison with existing DES calculations and experimental
data it is determined that the DES solutions aie spatially under-resolved and are not able to capture the majority
of the turbulent scales in the flow. However, it is also noted that the flow immediately downstream of breakdown
is not dominated by turbulence and many low frequency coherent structures are found to occur. Therefore, it is
concluded that URANS methods may be capable of accurately predicting this flow behaviour.

To consider this proposal two URANS turbulence models were used for the prediction of tire flow for the same test
case. The models chosen use two different approaches to model the turbulence in the flow. A linear Boussinesq
based model with a modification for vortex flow and a non-linear model based on an explicit algebraic Reynolds
stress formulation are used. Again, the effects of both temporal and spatial refinement on the flow behaviour
were considered. The relative behaviour of each model in predicting the unsteady flow behaviour was analysed
with respect to the treatment of the turbulence in the vortex flow. It was found that each model predicted very
similar behaviour. The URANS results were then compared to the DES solutions of the previous investigation to
evaluate the relative behaviour. From this comparison it was determined that the URANS turbulence models were
able to predict the dominant features of the low frequency phenomenon present in the vortex system and in the
post-breakdown region. It was then concluded that the URANS models were suitable for predicting the unsteady
behaviour of the flow at a considerable reduction in computational cost.

m
Declaration

The author hereby declares that this dissertation is a record of work carried out in the Department of Aerospace
Engineering at the University of Glasgow during the period of October 2003 to Februaiy 2007. This dissertation
is original in content except where otherwise stated.

Lucy Schiavetta

February 2007
Acknowledgements

Firstly, I would like to thank my supervisor Prof. Ken Badcock for all the guidance and advice throughout the
course of this project and for providing me with many opportunities to develop my research skills outwith the
CFD Lab. One of these opportunities allowed me to contribute to the NATO RTO AVT-113 Task Group, which
was a very enjoyable, interesting and useful experience and I would like to thank all those involved in the group.
Particularly, I would like to thank Okko Boelens, Willy Fritz, Russ Cummings and Scott Morton for tlie data they
provided me with, all their help and for answering my many emails, and Prof. Hummel for his encouragement and
enlightening discussions.

I must also thank BAE Systems who have sponsored this project and particularly, Mike Henshaw and Dominic
Amor for their help and support during my industrial placement.

I would also like to express my gratitude to all the researchers past and present I have been fortunate to work and
socialise with. It has made this project all the easier working in an enjoyable environment. I’d like to thank past
CFD lab members Punit, Abdul, Dave L and Romy, particularly for helping me so much when I was the ’’new girl”
and for their continuing help and support when I wasn’t. I’d also like to thank Joe, Dave S, Yoh and Giangi for
their friendship and ability to keep me laughing.

I would like to thank my family for their support and encouragement throughout the years, for never doubting I
could do it and for telling me so when I needed to hear it. Their unswerving faith in me has been a source of
strength again and again. Finally, I would like to thank Adriano, for his support and patience, despite being in the
same situation and without whom I’d probably never have gotten this far.
Publications
Conference Papers
L.A Schiavetta, O J Boelens and W. Fritz. Analysis of Transonic Flow on a Slender Delta Wing Using CFD
24th AlAA Applied Aerodynamics Conference, San Francisco, June 5-8th 2006, Paper AIAA-2006-3171

L.A Schiavetta, K.J Badcock and R.M Cummings. Comparison of DES and URANS for Unsteady Vortical
Flows over Delta Wings 45th AIAA Aerospace Sciences Meeting and Exhibit, Reno, 8-11 January 2007, Paper
AIAA-2007-1085

Presentations
L.A Schiavetta. VFE-2 Facet: Progress Report
Presented at NATO RTO AVT-113, "Understanding and Modelling Vortical Flows to Improve the Technical Readi­
ness Level fo r Military Aircraft”, Task Group Techinical Meeting, Budapest, Hungary, April 2005

L.A Schiavetta. VFE-2 Facet - Progress Report and CFD Results Comparison
Presented at NATO RTO AVT-113, "Understanding and Modelling Vortical Flows to Improve the Technical Readi­
ness Level fo r Military Aircraft”, Task Group Techinical Meeting, Granada, Spain, October 2005

L.A Schiavetta, O.J Boelens and W. Fritz. Analysis of Transonic Flow on a Slender Delta Wing using CFD
Presented at NATO RTO AVT-113, "Understanding and Modelling Vortical Flows to Improve the Technical Readi­
ness Level fo r Military Aircraft”, Task Group Techinical Meeting, Amsterdam, The Netherlands, May 2006

L.A Schiavetta, and R. Cummings. VFE2 Facet: Further Transonic Analysis and Comparisons
Presented at NATO RTO AVT-113, "Understanding and Modelling Vortical Flows to Improve the Technical Readi­
ness Level fo r Military Aircraft”, Task Group Techinical Meeting, Vilnius, Lithuania, Oct 2006

L.A Schiavetta. Unsteady Vortical Flows using CFD


Presented to BAE Systems, Advanced Technology Center, Sowerby, Filton, Bristol, UK, April 2006
Also Presented to Qinetiq, Farnborough, UK, April 2006
Contents

Abstract iii

Declaration iv

Acknowledgements v

Publications vi

Table of Contents ix

List of Figures xiv

List of Tables xv

Nomenclature xx

1 Introduction 1
1.1 O v e rv ie w ................................................................................................................................................. 1
1.2 Vortical Flows over Slender Delta W ings.............................................................................................. 2
1.2.1 Leading Edge V o rtic e s.............................................................................................................. 2
1.2.2 Vortex B re a k d o w n .................................................................................................................... 5
1.3 Unsteady Aspects of Delta Wing Vortical F lo w s .................................................................................. 8
1.3.1 Vortex and Vortex Breakdown Instabilities............................................................................... 8
1.3.2 Shear Layer In stab ilities........................................................................................................... 12
1.3.3 Unsteady Flow Topology........................................................................................................... 16
1.4 Transonic Effects on Vortical F l o w s ..................................................................................................... 17
1.5 Application of CFD to Delta Wing Vortical F lo w s ............................................................................... 28
1.5.1 Inviscid M e th o d s....................................................................................................................... 28
1.5.2 Laminar M ethods....................................................................................................................... 30
1.5.3 Reynolds Averaged Navier-Stokes (RANS) M e th o d s............................................................ 31
1.5.4 LES and DNS M ethods.............................................................................................................. 34
1.5.5 Detached Eddy Simulation (D E S ) ........................................................................................... 36
1.6 O bjectives................................................................................................................................................. 37
1.7 Thesis O utline.......................................................................................................................................... 38

2 Methodology of CFD Investigations 39


2.1 Introduction: The Navier-Stokes Equations ........................................................................................ 39
2.2 Turbulence ............................................................................................................................................. 40
2.3 Turbulence M odelling.............................................................................................................................. 41
2.3.1 Reynolds Averaging A pproach................................................................................................. 42
2.3.2 Favre-Averaging for Compressible F lo w s ............................................................................... 43
2.3.3 Large Eddy Sim ulation.............................................................................................................. 44
2.4 Application of Turbulence models to Delta Wing Vortical F l o w s ...................................................... 46
2.4.1 Wilcox &—0) M o d e l................................................................................................................. 47
2.4.2 k — a with Po, E n h a n c e r........................................................................................................... 47
2.4.3 Non-Linear Eddy Viscosity M o d e l........................................................................................... 48
2.4.4 Spalart Allmaras M o d e l ........................................................................................................... 49
2.4.5 Detached Eddy Simulation ( D E S ) ........................................................................................... 49
2.4.6 Computational Cost of Turbulent Calculations......................................................................... 50

vii
2.5 PMB s o lv e r ............................................................................................................................................ 50
2.5.1 Steady State S o lv er..................................................................................................................... 51
2.5.2 Unsteady S o l v e r ........................................................................................................................ 52
2.6 Unsteady Time Step C alcu latio n ......................................................................................................... 53
2.7 Post-processing Techniques................................................................................................................... 54
2.7.1 Domain Solution data .............................................................................................................. 54
2.7.2 Integrated Loads and Probe Analysis.......................................................................................... 55

Transonic Vortical Flow on a Slender Delta Wing 57


3.1 Introduction............................................................................................................................................ 57
3.2 Summary of Test Case ......................................................................................................................... 58
3.2.1 Grid Generation ........................................................................................................................ 59
3.2.2 Transition T reatm en t.................................................................................................................. 60
3.3 Subsonic Vortical Flow: R e s u lts .......................................................................................................... 60
3.4 Transonic Vortical Flow: R e s u lts......................................................................................................... 63
3.4.1 Pre-Breakdown Flow - M = 0.85, a ~ 18.5" .......................................................................... 63
3.4.2 Post-Breakdown Flow - M = 0.85, a = 2 3 " ............................................................................. 65
3.5 Occurrence of Shocks in the F l o w ...................................................................................................... 66
3.5.1 Cross-Flow S h o c k s..................................................................................................................... 67
3.5.2 Normal S h o c k s ........................................................................................................................... 68
3.6 CFD Sensitivity Study ......................................................................................................................... 69
3.6.1 Effect of Grid R e fin e m e n t........................................................................................................ 70
3.6.2 Effect of Turbulence Model ..................................................................................................... 72
3.6.3 Effect of Solution C o n v erg en ce............................................................................................... 74
3.6.4 Comparison with Other Structured GridR esults....................................................................... 76
3.6.5 Influence of Time A c c u ra c y ..................................................................................................... 79
3.7 Shock-Vortex Interaction and Vortex Breakdown .............................................................................. 81
3.8 Discussion............................................................................................................................................... 87
3.9 C onclusions............................................................................................................................................ 88

Application of DES to Delta Wing Vortical Flows 90


4.1 Introduction............................................................................................................................................ 90
4.2 Summary of Test C a s e ......................................................................................................................... 91
4.2.1 Grid Generation ........................................................................................................................ 92
4.2.2 Transition T reatm en t.................................................................................................................. 95
4.2.3 Probe Application ..................................................................................................................... 95
4.3 Effect of Time Step Refinement .......................................................................................................... 96
4.4 Effect of Grid Refinement...................................................................................................................... 99
4.5 Evaluation of LES Region ....................................................................................................................... 103
4.5.1 Unsteady Behaviour of DES S o lu tio n ..........................................................................................103
4.5.2 Decomposition of LES S o lu tio n ................................................................................................... 106
4.5.3 Resolution of DES S o lu tio n ......................................................................................................... 107
4.6 Qualitative Comparison with Cobalt R e s u lts ........................................................................................... 109
4.6.1 Comparison with 70" ONERA SA-DES R esults..........................................................................109
4.6.2 Comparison with 65" VFE-2 SA-DES Results .......................................................................... 113
4.7 Validation of DES Results ....................................................................................................................... 116
4.7.1 Comparison with Mitchell’s Experim ent.......................................................................................116
4.7.2 Comparison to Other Unsteady Experimental R e su lts.................................................................123
4.8 Discussion................................................................................................................................................... 126
4.8.1 The Role of Turbulence in Vortical F lo w s ................................................................................... 126
4.8.2 The Role of jLisGj m the DES C alculations................................................................................... 126
4.9 C onclusions................................................................................................................................................ 127

Assessment of URANS for Predicting Vortex Breakdown 129


5.1 Introduction................................................................................................................................................ 129
5.2 Effect of Grid Refinement.......................................................................................................................... 130
5.2.1 k —(û model with 7^ Enhancer...................................................................................................... 130
5.2.2 Non-Linear Eddy Viscosity M o d e l................................................................................................132
5.3 Effect of Time Step Refinement ..............................................................................................................135

vm
5.4 Comparison between Non-Linear Eddy Viscosity Model and k —a with P® Enhancer Model . . . 138
5.4.1 Unsteady Behaviour predicted by URANS S o lu tio n s.................................................... 140
5.4.2 Effect of Eddy Viscosity T re a tm e n t...........................................................................................143
5.5 Comparison of URANS and D E S ............................................................................................................. 146
5.5.1 Comparison of Unsteady Flow Behaviour P re d ic tio n ................................................................147
5.5.2 Effect of Eddy Viscosity T re a tm e n t............................................................................................149
5.6 Discussion...................................................................................................................................................150
5.7 C onclusions................................................................................................................................................152

6 Conclusions and Future Work 153


6.1 C onclusions................................................................................................................................................153
6.2 Future W o rk ................................................................................................................................................154

A Tlirbulence Models 156


A .l W ilcox7:—£ 0 .............................................................................................................................................156
A.2 N L E V M ...................................................................................................................................................... 156
A.3 Spalart-A llm aras...................................................................................................................................... 158

B Probe Analyser Tool 159


B .l Probe A n a ly s e r..........................................................................................................................................159
B.2 Application of StatisticalM e th o d s........................................................................................................... 160
B.2.1 Mean and Root Mean Square V alues............................................................................................ 160
B.2.2 Power Spectral D ensity ..................................................................................................................161
B.2.3 Time Averaging............................................................................................................................. 162
B.2.4 Turbulent Kinetic E nergy..............................................................................................................162

References 163
List of Figures

1.1 NASAF-18 HARV at a = 20" with smoke visualisation showing vortex created by leading edge
extension and vortex breakdown interacting with aircraft ta ilp la n e ............................................... 2
1.2 Schematic of the subsonic behaviour of the flow over a delta wing at incidence (from Ref. [1]) . 3
1.3 Structure of a leading edge vortex (from Ref. [ 2 ]) ............................................................................ 3
1.4 Profiles of swirl and axial velocity through a vortex core, detailing the three main vortex regions
(adapted from Ref. [2], originally from Ref. [ 3 ] ) ........................................................................... 4
1.5 Leading edge vortices and types of vortex breakdown over a 65" delta wing. The upper vortex
exhibits the spiral form of breakdown and the lower vortex shows the bubble type (from Ref. [4]) 5
1.6 Spiral vortex breakdown occurring over delta wings (adapted in Ref. [5] from Ref. [4 ]).............. 5
1.7 Instantaneous evidence of the spiral nature of breakdown T = 0.037, shown on a plane through
the vortex core (from Ref. [6] ) .......................................................................................................... 6
1.8 Effect of sweep angle on the occurrence of vortex breakdown at the trailing edge (from Ref. [7]) 7
1.9 Movement of vortex breakdown over a 65" delta wing with increasing incidence (data from Ref.
[8] ) ..................................... 7
1.10 Lift coefficient vs angle of incidence for different sweep angles [9] 8
1.11 Streamwise fluctuations of vortex breakdown location at a: = 32" (from Ref. [ 1 0 ] ) .................... 10
1.12 Variation of non-dimensional frequency for unsteady phenomena as a function of angle of inci­
dence (from Ref. [11]) ....................................................................................................................... 12
1.13 Kelvin-Helmlioltz shear layer instabilities (from Ref. [6]) ............................................................. 12
1.14 Shear layer instabilities and the effect of Reynolds number (from Ref. [1 2 ])................................. 13
1.15 Stationary shear layer sub-structures shown using contours of x vorticity in planes perpendicular
to the wing surface (from Ref. [ 1 3 ] ) ................................................................................................. 15
1.16 Instantaneous shear layer stincture shown by iso-surface of axial vorticity for two Reynolds
numbers (From Ref. [ 1 4 ] ) ................................................................................................................. 15
1.17 Time-averaged shear layer structure shown by iso-surface of axial vorticity for two Reynolds
numbers (from Ref. [14]) ................................................................................................................. 16
1.18 Spectrum of unsteady flow phenomena as a function of Strouhal number (from Ref. [15]) . . . 17
1.19 Classification of flow behaviour over delta wings by Miller and Wood [16] ................................. 20
1.20 Schematic diagrams showing proposed positions and shapes of shock systems over transonic
delta wings [17] 21
1.21 Chordwise pressure coefficient distribution at the symmetry plane for a range of angles of inci­
dence pre- and post-breakdown [ 1 8 ] ................................................................................................. 22
1.22 Snapshots of pressure distribution on the surface of the wing using PSP techniques (from Ref.
[19]) ................................................................................................................................................... 23
1.23 Experimental results from NASA NTF wind tunnel tests for conditions: M = 0.85 and Re —
6 X 10^, for the sharp leading edge, at a range of angles of incidence 18.5" - 26.7". (data taken
from Ref. [2 0 ] ) ................................................................................................................................... 25
1.24 Experimental results from NASA NTF wind tunnel tests a range of angles of incidence 18.5" -
26.7". Legend as shown for Figure 1.23 (data taken from Ref. [ 2 0 ] ) ............................................ 25
1.25 PSP surface pressure coefficient contours for M = 0.8, Re = 2 x 10^, a = 18.4" - 25.9" [21]. . 27
1.26 PIV results showing contours of non-dimensional u velocity for a = 25.9" at M = 0.80 (data
from Ref. [2 1 ] ) ................................................................................................................................... 27
1.27 Surface pressure contours for Euler, laminar and turbulent computations, for M = 0.3, ce = 30"
and R e ^ l x 10^^ [22] ....................................................................................................................... 29
1.28 Sample of time histories and PSD analysis of normal force coefficient from Gortz’s unsteady
inviscid calculations [2 3 ].................................................................................................................... 30
1.29 PSD comparisons of normal force coefficient for five turbulence models [ 2 4 ] ............................. 33
1.30 Q-criterion isosurfaces of vortex behaviour for two different near-wall treatments (from Ref. [25]) 35
2.1 Energy Spectrum for a turbulent flow - log-log scales (From Ref. [26]) ...................................... 41
2.2 Schematic of DES formulation on a structured g r i d ......................................................................... 50
2.3 Shock detection test quantity (Adapted from Ref. [2 7 ])................................................................... 55

3.1 A summary of the flow features for various Mach numbers and angles of incidence (from Ref.
[18]) ................................................................................................................................................... 57
3.2 VFE-2 65" delta wing geometr y used in investigation...................................................................... 59
3.3 Surface mesh and ogrid topology around sting region for 65" VFE-2 delta w in g .......................... 60
3.4 Comparison of grid refinement at x/c,. = 0,5 .................................................................................. 60
3.5 Computational results compared to experimental data, M = 0.4, Re = 6 x 1 0 ^ ............................. 61
3.6 Contours of x vorticity and u velocity on slices through the vortex core for a = 18.5" and 23" -
M = 0.4 andRe = 6 X 1 0 ^ ................................................................................................................. 62
3.7 Comparison of computational results and experimental data, M — 0.4 and Re = 6 x 10^............. 62
3.8 Computational results compared to experimental data, a — 18.5", M = 0.85 and Re —6 x 10^ . 63
3.9 Axial velocity through primary and secondary vortex cores, a — 18.5", M —0.85 and Re — 6 x 10^ 64
3.10 Contours of x vorticity and u velocity on slices through the vortex core for 18.5", M = 0.85 and
Re = 6 x 1 0 ^ ....................................................................................................................................... 64
3.11 Contours of X vorticity at a position x/c,. = 0.4 for ct = 18.5", M = 0.85 and Re = 6 X 10*^ . . . 64
3.12 Computational results compared to experimental data, a = 23", M = 0.85 and Re = 6 x 10*^ . . 65
3.13 Axial velocity tlirough vortex core for post-breakdown flow, a — 23", M = 0.85 and Re = 6 x 10^ 66
3.14 Flow structure for a — 23", M —0.85 and Re = 6 x 10*^............................................................... 66
3.15 Plots for x/cr — 0.4 showing contours of flow variables to highlight locations of cross flow
shocks for a — 18.5", M = 0.85 and Re = 6 x 1 0 ^ ........................................................................ 66
3.16 Velocity vectors and contours of Mach number at chordwise station x/c,- = 0.2 showing sec- !
ondary separation for a — 18.5", M — 0.85 and Re = 6 x 10^ ..................................................... 67 !
3.17 Contours of x vorticity on a x/c,- — 0.4 plane, highlighting locations of cross flow shocks for I
a = 23", M = 0.85 and Re = 6 X 1 0 ^ 68 i
3.18 Pressure coefficient distribution at the symmetry plane on the wing for both angles of incidence 68 ;
3.19 Isosurface of x vorticity coloured by pressure coefficient showing primary vortex shear layer and
normal shock shape for both angles of in c id e n c e ........................................................................... 69
3.20 USAFA grid at symmetry p l a n e ........................................................................................................ 70
3.21 Comparison between the H-H grids for transonic conditions at cc = 18.5" and 2 3 " ....................... 70
3.22 Surface contours of pressure coefficient for comparison between the H-H g r i d s .......................... 71
3.23 Comparison of axial velocity through the vortex cores for coarse and fine grid solutions . . . . 71
3.24 Contours of x vorticity at chordwise station x/c,- = 0.4 at 18.5" ................................................... 72
3.25 Effect of turbulence model on flow solution with comparison to experiment for M — 0.85, Re =
6 X 10^ and ct = 2 3 " .......................................................................................................................... 73
3.26 Contours of surface pressure coefficient showing effect of turbulence model on flow solution
with comparison to experiment for a = 23", M = 0.85 and R e ~ 6 x 10^ .................................. 74
3.27 Convergence history of residuals for k - co with Ro, Enhancer model; a = 23", M = 0.85 and
Re = 6 X 1 0 ^ ....................................................................................................................................... 74
3.28 Effect of turbulence model on flow solution with comparison to experiment for a = 23", M =
0.85 and Re = 6 X 1 0 ^ ....................................................................................................................... 75
3.29 Contours of surface pressure coefficient showing effect of turbulence model on flow solution
with comparison to experiment for M = 0.85, Re = 6 x 10^ and a — 2 3 " ; .................................. 75
3.30 Comparisons between computational results and experiment for all codes for M = 0.85, Re =
6 X 10<^................................................................................................................................................ 76
3.31 Surface pressure coefficient contours for all codes, M = 0.85, Re = 6 x 1 0 ^ ................................ 77
3.32 Axial velocity through primary vortex core for all codes a = 23", M = 0.85, Re = 6 x 10^ . . . 77
3.33 Contours of x vorticity at x/c,- = 0.4 for all results, a = 23", M = 0.85, Re = 6 x 1 0 ^ ................. 77
3.34 Pressure coefficient distribution at the symmetry plane on the wing ............................................. 78
3.35 Effect of grid on flow solution with comparison to experiment for a — 23", M = 0.85 and
Re = 6 X 10^; a) Comparison between results from Glasgow and TSfLR grids (NLR Results); b)
Comparison between Glasgow results and NLR results on common grid using similar turbulence
model..................................................................................................................................................... 78
3.36 Contours of surface pressure coefficient showing effect of grid on flow solution with comparison
to experiment for M = 0.85, Re = 6 x 10^ and a = 23"; a) Comparison between results from
Glasgow and NLR grids (NLR Results); b) Comparison between Glasgow and NLR results on
common grid using similar turbulence models................................................................................... 79

XI
3.37 Comparisons between computational results and experiment for current results and US AFA time
accurate solutions for a — 23", M = 0.85, Re = 6 x 1 0 ^ ............................................................... 80
3.38 Surface pressure coefficient contours comparing US AFA time averaged results with steady state
cunent results, a = 23", M = 0.85, Re = 6 x 10*^........................................................................... 80
3.39 Comparisons between computational results and experiment for current steady state results and
USAFA time accurate solutions for a — 23", M = 0.85, Re = 6 x 1 0 ^ ......................................... 81
3.40 Isosurface of x vorticity coloured by pressure coefficient showing primary vortex shear layer
and normal shock shape for current results and USAFA time accurate solutions; M = 0.85 and
Re = 6 x 1 0 ^ ...................................................................................................................................... 81
3.41 Pressure distribution through vortex cores for both angles of incidence; The numbers on the plot
signify the magnitudes of the pressure ratios through the intersecting shocks............................... 82
3.42 Contours of Mach number on slice through vortex core at a constant y /s = 0.56 for a = 23",
M = 0.85 and Re = 6 X 10® ............................................................................................................. 82
3.43 Rossby number distribution against root chord location for pre- and post-breakdown cases . . . 83
3.44 Theoretical limit curve for normal shock vortex interactions, where r is the swirl ratio = l/R o
(adapted from Ref. [ 2 8 ] ) .................................................................................................................... 83
3.45 Experimental surface pressure data on conical ray at constant y /s = 0.3 to show experimental
shock strength for a = 23.6" and 24.6", M = 0.85, Re = 6 x 10® from NASA NTF data . . . . 84
3.46 Pressure distribution through vortex cores for EADS and NLR solutions....................................... 85
3.47 Comparison between u velocity contours for experimental PIV and computational results for
M = 0.80 on a slice at x/c,. = 0,5........................................................................................................ 86
3.48 u velocity through vortex core for computational results compared to exper imental PIV data for
M = 0.80, cc = 2 6 " ............................................................................................................................. 87
3.49 Vortex breakdown location for both computational and experimental re s u lts ................................. 87

4.1 Schematic diagrams showing fiow topology upstream and downstream of vortex breakdown . . 91
4.2 70" ONERA geometry (all distances marked are in mm) [13].......................................................... 92
4.3 Grid topology of H-H grids used in investigation............................................................................. 93
4.4 Comparison of grid refinement at x/c,- = 0.63 plane ..................................................................... 94
4.5 Comparison of overall grid refinement at symmetry plane for coarse and fine g r i d s ................... 94
4.6 Comparison of grid refinement in trailing edge region for fine grid and refined TE grids............. 94
4.7 DES active area for fine grid; red denotes LES region and green shows URANS region............... 95
4.8 Positions of probes for 70" ONERA w in g ......................................................................................... 96
4.9 Isosurfaces of x vorticity coloured by pressure coefficient showing instantaneous vortex core
behaviour at t = 50 with core probe locations marked for At = 0 .0 1 ............................................ 96
4.10 Mean and RMS behaviour of non-dimensional velocity components at five probe locations
through vortex core region for each solution of the time step s tu d y ............................................... 97
4.11 Instantaneous contours of y vorticity on a slice tlirough the vortex core region at T == 50 .... 98
4.12 Unsteady behaviour of non-dimensional velocity components shown by time histories and PSD
frequency plots for a probe at x/c,- = 1.00, y /s = 0 .7 and z / c, = 0 .1 ; fine grid solutions for time
steps of A t = 0.01,0.005 and 0.0025 .............................................................................................. 98
4.13 A,„„,v of fine grid on slice at trailing edge, x/cr = 1 .0 0 ....................................................................... 99
4.14 Contours of A,„„,y on a slice through the vortex core for both grids used in grid refinement study 99
4.15 Contours of instantaneous y vorticity on a slice through the vortex core for the fine and refined
TE grid at T = 5 0 .................................................................................................................................... 100
4.16 Mean and RMS behaviour of non-dimensional velocity components at eight probe locations
through vortex core region for the fine and refined TE grids ............................................................ 100
4.17 Unsteady behaviour of non-dimensional velocity components shown by time histories and PSD
frequency plots for the probe at x/c,- = 1.00, z/c,- = 0.1 for the fine and refined TE grids . . . . 101
4.18 Unsteady behaviour of non-dimensional velocity components shown by time histories and PSD
frequency plots for the probe at x/c,-= 1.2, z/c,-= 0.2 for the fine and refined TE g rid s 101
4.19 Location of probes though vortex core region compared to u velocity contours at each stream-
wise lo c a tio n .......................................................................................................................................... 102
4.20 Mean and RMS behaviour of non-dimensional velocity components through vortex core . . . . 103
4.21 Unsteady behaviour of non-dimensional velocity components at probes tlmough vortex core re­
gion shown by time histories and PSD frequency plots.........................................................................103
4.22 PSD of u velocity with x/c,- = 0.84 probe signal removed for clarity of frequency content of
remaining p ro b e s .................................................................................................................................... 104
4.23 Non-dimensional u velocity time history and PSD fora probe on the vortex axis, downsbeam of
vortex breakdown, from a highly under-resolved DESsolution, coarse grid. At = 0 . 0 1 ..................105
4.24 PSD against St and non-dimensional period for non-dimensional u velocity for a probe at x / c,- =
1.0 on the vortex core axis....................................................................................................................106
4.25 Non-dimensional u velocity time histories and PSD plots for a probe on the vortex axis, down­
stream of vortex breakdown showing both tlie stationary and non-stationary mean at different
sample rates .......................................................................................................................................... 107
4.26 Analysis of u velocity behaviour from spatial slice through vortex core at z/c,- = 0.1 to determine
resolution of g r i d .................................................................................................................................... 108
4.27 Turbulent spectrum for both spatial and temporal scales to show accuracy of energy cascade
within computational re s u lts ............................................................................................................. 108
4.28 Isosurfaces of vorticity for various USAFA unstructured grids compared to current results on
refined bailing edge structured grid. The number of cell volumes for each grid are given for
comparison. [ 2 9 ] .................................................................................................................................... 110
4.29 PSD plots of normal force coefficient for current results compared to USAFA results from literature 111
4.30 Comparison of turbulent kinetic energy through vortex core between current results and USAFA
results from literature ...........................................................................................................................112
4.31 Comparison of surface streamlines between current results and USAFA results from literature . 112
4.32 Location of probes through vortex core with reference to isosurface of x vorticity for 65" VFE-2
USAFA DES calculation........................................................................................................................113
4.33 Comparison between time histories at similar probe locations on the vortex core in the pre- and
post-breakdown flow. Cuixent results on the left hand side and USAFA results on the right. . . 114
4.34 Comparison between PSD frequencies at similar probe locations on the vortex core in the pre-
and post-breakdown flow. Pre-breakdown results on the left hand side and post-breakdown
results on the right.................................................................................................................................... 114
4.35 Analysis of u velocity behaviour from spatial slice through vortex core at z/c, = 0 .1 to determine
resolution of grid for both cuixent results and USAFA 65" VFE-2 DES results .............................115
4.36 Spatial and temporal comparisons of current and USAFA DES r e s u lts ..............................................116
4.37 Time averaged velocity results from Mitchell’s experiment compared to mean computational
resu lts.......................................................................................................................................................117
4.38 Time averaged x vorticity results from Mitchell’s experiment compared to mean computational
results ....................................................................................................................................................118
4.39 Surface pressure coefficient data for both experimental and computational results [13, 30, 31] . 119
4.40 Location of unsteady probes used for comparison with Mitchell’s d a ta ..............................................119
4.41 Time histories of unsteady pressure probe data [1 3 ] .............................................................................120
4.42 Power specb al density plots of unsteady pressure probe data [13] .................................................... 121
4.43 Comparison of unsteady vortex breakdown r e s u l t s ............................................................................. 122
4.44 PSD plot of unsteady vortex breakdown results from Mitchell’s experiment for left hand side
[13] a = 27" C/co = 0.24 (Scanned - Poor Quality) ............................................................................122
4.45 Instantaneous vortex breakdown regions for experimental and computational results. Also shown
are the locations of the data points from which the time histories of u velocity were taken. . . . 123
4.46 Time histories and PSD analyses of u velocity for Point A ................................................................ 124
4.47 Time histories and PSD analyses of u velocity for Point B ................................................................ 124
4.48 Time histories and PSD analyses of u velocity USAFA results for 65" delta wing at location on
vortex core plane, x/c,- = 1.0, z/c,- = 0.1 ........................................................................................... 125
4.49 Slice through vortex core at y /s = 0.7 showing contours of sub-grid eddy viscosity relative to
laminar viscosity created by the DES calc u la tio n ...............................................................................127

5.1 Location of probes though vortex core region compared to instantaneous u velocity contours at
each streamwise location and an isosurface of entropy at T = 50, coarse and fine grid compar­
isons for A:—£ü with P(o Enhancer m odel........................ 130
5.2 Mean and RMS velocity components through vortex core; coarse and fine grid comparisons for
k - (Û with Pfo Enhancer m odel.............................................................................................................. 131
5.3 Time history and PSD analysis of u velocity signal situated above the trailing edge on the vortex
axis at z/c,- = 0.1; coarse and fine grid comparisons for A:—co with Ry Enhancer m o d el................ 132
5.4 Slice through vortex breakdown region, on a plane tlnough vortex core, y /s = 0,7 showing in­
stantaneous contours of y vorticity, coaise and fine grid comparisons fo rk —co with Pa Enhancer
m o d e l.......................................................................................................................................................132
5.5 Location of probes though vortex core region compaied to instantaneous u velocity contours at
each streamwise location and an isosurface of entropy at r = 50, coarse and line grid compar­
isons for Non-Linear Eddy Viscosity m o d e l.........................................................................................133
5.6 Mean and RMS velocity components through vortex core; coarse and fine grid comparisons for
Non-Linear Eddy Viscosity m o d e l........................................................................................................ 134
5.7 Time history and PSD analysis of u velocity signal situated above the ti'ailing edge on the vortex
axis at z/c,- = 0.1; coarse and fine grid comparisons for Non-Linear Eddy Viscosity model . . . 134
5.8 Slice through vortex breakdown region, on a plane through vortex core, y /s = 0.7 showing
instantaneous contours of y vorticity, coarse and fine grid comparisons for Non-Linear Eddy
Viscosity model .....................................................................................................................................135
5.9 Location of probes though vortex core region compared to instantaneous u velocity contours
at each streamwise location and an isosuiface of entropy at r = 50, time step comparisons for
Non-Linear Eddy Viscosity m o d e l........................................................................................................ 135
5.10 Mean and RMS velocity components through vortex core; time step comparisons for Non-Linear
Eddy Viscosity m o d e l ........................................................................................................................... 136
5.11 Time history and PSD analysis of u velocity signals for two probes situated on the vortex axis at
z/c- = 0.1 above the wing surface; time step comparisons for Non-Linear Eddy Viscosity model 137
5.12 Slice through vortex breakdown region, on a plane through vortex core, y /s = 0.7 showing
instantaneous contours of y vorticity; time step comparisons for Non-Linear Eddy Viscosity
model .................................................................................................................................................... 138
5.13 Mean and RMS velocity components through vortex core; comparison of k —to with Pa En­
hancer, A t = 0.01, and Non-Linear Eddy Viscosity model, A t = 0.01, for the fine g r i d 139
5.14 Time history and PSD analysis of u velocity signals for two probes situated on the vortex axis at
z/c,- = 0.1 above the wing surface; Comparison of Æ- O) with Pa Enhancer and NLEVM models
for the Fine grid at At = 0 . 0 1 ...............................................................................................................139
5.15 Unsteady behaviour of non-dimensional velocity components at probes through vortex core re­
gion shown by time histories and PSD frequency plots for k — O) with Pa Enhancer model.
At = 0 .0 1 ................................................................................................................................................. 141
5.16 Unsteady behaviour of non-dimensional velocity components at probes through vortex core re­
gion shown by time histories and PSD frequency plots for NLEVM, At = 0 . 0 1 ........................ 141
5.17 Slice tlirough vortex breakdown region, on a plane through vortex core, y /s = 0.7 showing
instantaneous contours of fir /p for Wilcox k —co m o d e l...................................................................143
5.18 Slice through vortex breakdown region, on a plane through vortex core, y /s = 0.7 showing
instantaneous contours of P r /p and turbulent kinetic energy for both URANS models, fine grid.
At = 0.01 144
5.19 Slice through vortex region at x/c,- = 0.84 showing instantaneous contours of P r/li for both
URANS models, fine grid, At = 0 . 0 1 ..................................................................................................144
5.20 Surface streamlines showing comparable behaviour of secondary separation line after transition
to turbulence at x/c,- 0.36 for both URANS models, fine grid, At = 0 . 0 1 ................................... 145
5.21 Mean and RMS velocity components through vortex core; URANS k — co with Pa Enhancer
model compared to DES solutions ..................................................................................................... 147
5.22 Slice through vortex breakdown region, on a plane through vortex core, y /s — 0.7 showing
instantaneous contours of y vorticity at t = 50, for URANS k —co with Pa Enhancer model and
D E S ...........................................................................................................................................................147
5.23 Unsteady behaviour of non-dimensional velocity components at probes through vortex core re­
gion shown by time histories and PSD frequency plots for k — co with Pa Enhancer model.
At = 0 .0 1 ................................................................................................................................................. 148
5.24 Unsteady behaviour of non-dimensional velocity components at probes through vortex core re­
gion shown by time histories and PSD frequency plots for DES, At = 0,0025 ............................ 148
5.25 Time history and PSD analysis of u velocity signals for a probe situated on the vortex axis at
z/c,- = 0.1 above the wing surface for URANS k - co with Pa Enhancer and DES solutions . . . 149
5.26 Slice through vortex breakdown region, on a plane through vortex core, y /s = 0.7 showing
instantaneous contours of P r /p for URANS k —co with Pa Enhancer model and DES ................ 150
5.27 Distribution of pr through vortex core for all turbulence models used in this investigation . . . 151

B .l Graphical User Interface for probe analyser p ro g ra m ...........................................................................160


List of Tables

1.2 Non-dimensional frequencies corresponding to important unsteady features of vortical flows from
literature................................................................................................................................................... 18
1.2 Non-dimensional frequencies corresponding to important unsteady features of vortical flows from
literature (co n tin u ed )............................................................................................................................. 19
1.3 Time steps used in DES time accuracy study [ 3 2 ] ............................................................................... 37

2.1 Summary of computational costs of various turbulence methods according to Spalart (adapted from
Ref. [ 3 3 ] ) ................................................................................................................................................ 50

3.1 Summary of main features of grids used for VFE-2 investigation...................................................... 59


3.2 Summary of grids and turbulence models used for VFE-2 structured grid comparisons.................... 69
3.3 Summary of shock strength on surface conical ray at constant y /s = 0.3 for all solutions at M =
0.85, Re = 6 x 10® and a = 23" compared to NASA NTF data........................................................... 84
3.4 Summary of shock and vortex core data for all steady state calculations at a = 18.5" —26", M —
0.85 and Re = 6 x 10® ^ indicates further converged solution results................................................ 85
3.5 Summary of maximum axial velocity, shock strength and breakdown location for all solutions at
a = 23", M ~ 0.85 and Re = 6 x 10®.................................................................................................... 86
3.6 Critical incidence for transonic vortex breakdown to be found on 65" delta w i n g s .......................... 88

4.1 Frequencies corresponding to important unsteady features of vortical flows ................................... 90


4.2 Summary of main features of grids used in DES and URANS investigations.................................... 93
4.3 Experimental unsteady pressure probe lo catio n s.................................................................................. 95
4.4 Frequencies corresponding to important unsteady features of vortical flows from unsteady DES
re s u lts .......................................................................................................................................................... 105
4.5 Details of grid features for USAFA grid study and comparison with current results..............................I l l
4.6 Details of grid and time step for USAFA VFE-2 calculation and current results .................................113

5.1 Location of mean vortex breakdown for each unsteady calculation performed in this investigation . 151
5.2 Length of calculations for each turbulence treatment used in this investigation to predict a total
time of one second. Work units in brackets denote unit for one time s t e p ............................................ 152

A.l Model constants for the Wilcox k —co turbulence model [ 3 4 ] ................................................................ 156
A.2 Model constants for the Spallait-Allmaras turbulence model [35], where Cwi = ^ H- , 158

XV
Nomenclature

Symbols
üij Anisotropic component of Reynolds stress tensor
Extra anisotropic component of Reynolds Stress tensor for NLEVM model
n™' Subgrid scale anisotropic component of Reynolds stress tensor
a Anisotropic 2nd order tensor
c Chord length/characteristic length
c,- Root chord length
C£,i, ci,2 , cy 1 Model constants for Spalart-Allmaras model
Cwi >C\v2 , Cw3 Model constants for Spalart-Allmaras model
Cv Specific heat at constant volume
Cp Specific heat at constant pressure
d Distance from wall
d Variable used in Spalart-Allmaras model for DES formulation
d,nin Minimum distance from wall
e Specific internal energy
/ Frequency, Hz
fi Body forces
fp* Closure coefficents for k —co model
/ u i ,fv iifw Model functions for Spalart-Allmaras model
g, r Model functions for Spalart-Allmaras model
I Characteristic turbulent eddy length scale
lo Characteristic length of largest turbulent scales
k Turbulent kinetic energy
kr Heat transfer coefficient
kr Residual kinietic energy
p Pressure
qi Heat flux vector
çf Turbulent heat flux vector
r Ratio of the magnitude of the strain rate tensor to the magnitude
of the vorticity r = S/Q.
I'c Radius of vortex core at point of maximum swirl velocity
s Local semi-span
t Time
t Pseudo time
Ui Instantaneous velocity vector
«, Time averaged velocity vector
Ui Mass averaged velocity vector
ûi Filtered velocity vector
«■ Turbulent fluctuating velocity
u" Mean flow fluctuating velocity
u'l" Fluctuating velocity from Favre averaging
n Subgrid scale velocity
u{£) Characteristic velocity scale
U(j Characteristic velocity of largest turbulent scales

XVI
U^j Kolmogorov velocity scale
u Velocity in x-direction
V Velocity in y-direction
w Velocity in z-direction
First cell distance from wall
Xi Position vector
x,y,z Cartesian coordinates
x/c,. Non-dimensional streamwise location
Xly Location of forced transition line
y /s Non-dimensional spanwise location
z/cr Non-dimensional normal location

Co Drag coefficient
Coiïs DES model coefficient
CL Lift coefficient
Cfi Normal force coefficient
< Effective eddy viscosity coefficient for NLEVM model
& Kolmogorov constant
Cr Model constant for NLEVM model
Cp Pressure coefficent
Q Smagorinsky constant
E Total energy
E ( K) Energy spectral density
F Flux vector in the x-direction
Fourier Transform
G Flux vector in the y-direction
G Filter function
H Flux vector in the z-direction
H Total enthalpy
I Identity matrix
lia . Ils, Ills, IV, V Invariants of strain-rate and rotation tensors
M Mach number
Ml, M2 Mach numbers defined upstream and downstream of shock surface
Mn Mach number normal to shock surface
N Number of samples in signal
N ,N c,P i,P 2 ,D , c [ , c i , f i Functions for NLEVM formulation
S ons Required grid size for DNS calculations based in turbulent Reynolds number
P Pressure
Pu Pi Pressures defined upstream and downstream of shock surface
Pœ Production term in co equation
PcJ/iew Modified production terra in co equation
Pk Production term in k equation
Pr Prandtl number
Q 'j4 Turbulent Flux residual
Flux residual
Unsteady Flux residual
R Specific gas constant
Auto-correlation function of general variable <I>(f)
Re Reynolds number Re = LUc/ v
Rcr Turbulent Reynolds number Re, = E ^ i / v
Ro Rossby Number Ro = U„xiai/Co
S Magnitude of vorticity; Magnitude of strain-rate tensor (ISijSijŸ^^
SiJ Strain-rate tensor
Strain-rate tensor for compressible flow
s Local deformation rate
SiJ Filtered strain-rate tensor
SiJ Mean strain-rate tensor
S Normalised strain-rate tensor
St Strouhal number, non-dimensional frequency St = fc/U ,,
StfL Nquist non-dimensional frequency
T Temperature; Non-dimensional sample rate
To Reference Temperature
t/oo Freestream velocity
m Stationary mean velocity vector
Ui Mean flow velocity vector
Ginax Maximum flow velocity
Uq Swirl/Tangential velocity
ITnxial Vortex core axial velocity
Glocal Local velocity magnitude
Cell volume
W Vector of conserved variables

Greek Symbols
a Angle of incidence
a Model constant for k —co model
CCcr Critical incidence for vortex breakdown
Pn Model coefficients for NLEVM model
Closure coefficents for k —co model
r Ratio of specific heats, y = Cp/cv
Kronecker delta, 5,7 = 1 if i = j and Ry = 0 if i 7^ j
E Dissipation per unit mass
^iJÀ Alternating symbol
ri Kolmogorov length scale
K Non-dimensional wavenumber; Von Karman constant, k
À Wavelength
R Molecular viscosity
Po Reference molecular viscosity
Pt Turbulent eddy viscosity
PSGS Subgrid scale (or Smagorinsky) eddy viscosity
V Kinematic viscosity, v = p /p
V Spalart-Allmaras equation working variable, v = V r/fvi
Vj. Turbulent kinematic eddy viscosity, V/- = p r /p
^SGS Subgrid scale kinematic eddy viscosity
P Density
P Time-averaged density
a,G* k —co model constants
T Non-dimensional time; Swirl ratio; Turbulent time scale
To Characteristic time of largest turbulent scales
Kolmogorov time scale
T(0 Characteristic time scale
At Non-dimensional time step
Tu Viscous stress tensor
^5 Reynolds stress tensor
Subgrid scale stress tensor
Level of turbulence resolved by LES
X Model function for Spalart-Allmaras model
(0 Dissipation per unit turbulence kinetic energy
co Angular frequency
COi Vorticity vector

A Filter Width
Aa-) Ay, Aj Cell dimensions in x,y,z directions
Ao Optimal cell dimension

XVm
Maximum cell dimension
A Sweep angle
0 General variable
Q Magnitude of rotation tensor {IQijQij) ^
Q Normalised rotation tensor
Q; Rotation tensor
Ù.! Mean rotation tensor

Acronyms
AMR Adaptive Mesh Refinement
BILU Block Incomplete Lower Upper
CAD Computer Aided Design
CEO Computational Fluid Dynamics
CFL Courant-Fredrich-Lewy number
DES Detached Eddy Simulation
DPT Discrete Fourier Transform
DNS Direct Numerical Simulation
EFT Fast Fourier Transform
GUI Graphical User Interface
JSE Joint Strike Fighter
LDV Laser Doppler Velocimetry
LE Leading Edge
LES Large Eddy Simulation
NASA National Aeronautics and Space Agency
NLEVM Non-Linear Eddy Viscosity Model
NTF National Transonic Facility
PIV Particle Image Velocimetr y
PSD Power Spectral Density
PSP Pressure Sensitive Paint
RANS Reynolds Averaged Navier Stokes
RLE Round Leading Edge
RMS Root Mean Square
RSM Reynolds Stress Model
SA-DES Spalart-Allmaras Detached Eddy Simulation
SARC Spalart-Allmaras with Rotation Correction
SGS Sub-Grid Scale
SEE Sharp Leading Edge
SST Shear Stress Transport
SSTDES Shear Stress Transport Detached Eddy Simulation
TE Trailing Edge
TNT Turbulent Non-Turbulent
UAV Unmanned Aerial Vehicle
UCAV Unmanned Combat Aerial Vehicle
URANS Unsteady Reynolds Averaged Navier-Stokes
VBD Vortex Breakdown
VFE-1 1st international Vortex Flow Experiment
VFE-2 2nd international Vortex Flow Experiment
XLES eXtra Lai'gc Eddy Simulation

Subscripts
Freestream conditions
n Current time-step
n -h I Next time-step
rms Root mean square value

X IX
T Turbulent quantities

Superscripts
i Inviscid component
k,iu kl, l,„, h Time levels of variables in spatial discretisation
V Viscous component

XX
Chapter 1

Introduction

1.1 Overview
Delta wings were inti'oduced to reduce the effects of compressibility such as the transonic drag rise. However, it
was also found that these planforms were suitable for low speed flight and could produce lift at angles of incidence
much lai-ger than traditional swept wings [1]. This increased lift and improved aerodynamic characteristics allows
greater agility, particularly at low speeds, and the design of many complex configurations have been centred around
the use of such planforms.

The additional lift is due to the flow separating at the leading edges of tliese wings at low incidences and being
convected downstream by the freestream velocity. The convection of this flow results in the production of coher­
ent vortical structures, which exist above the leading edge and contribute to the lift force generated by the wing.
However, as incidence is increased these structures become unstable and breakdown into an incoherent form. This
reduction in the coherence of the flow stincture over the wing has been found to have detrimental effects on the
aerodynamic characteristics of the wing and may cause a sudden and unexpected aerodynamic instability of tlie i
aircraft. This is particularly true for transonic velocities where interactions between shockwaves and the flow over
the wing can cause sudden changes in behaviour. Due to the ability of these structures to have both beneficial and
detrimental effects on the aerodynamics of aircraft, a great deal of research and development has been earned out
over the years to consider and improve the behaviour of the wings for a range of flight conditions.

From early research, it was noted that the flow over the delta wing was not steady in nature, particularly when the V
leading edge vortices suffered breakdown into a turbulent state [4]. This unsteadiness has been found to further
complicate an already complex flow behaviour. Through interactions with the wing surface and other aircraft ap-
pennages such as the tailplane, this results in aeroelastic effects occurring on the aircraft, which can cause fatigue
and, in the most severe cases, results in catastrophic failure. This type of behaviour was found, quite recently,
to occur on the F-18A, where the vortices created from the leading edge extensions of the wings were found to
breakdown upstream of the vertical fins of the aircraft at moderate angles of incidence as shown in Figure 1.1. This
unsteady flow was then found to interact with the tailplane structure causing a buffeting effect. This was found to
cause premature fatigue of the vertical fins and may affect the control surface of the aircraft, both of which will
have serious effects on the stability and performance of the aircraft during rapid manoeuvres.

Therefore, it is clear that understanding the behaviour of this unsteady forcing is crucial to enable the alleviation
of any structural responses which may exist. This type of unsteady flow is known as buffet and is a issue for all
aircraft configurations which incorporate swept edges in their design. This is particularly important for complex
fighter configurations such as the F-16XL, EuroFighter and for the design of future configurations. This need is
further compounded by the emergence of new UAV and UCAV technologies, which are tending toward planforms
where vortical behaviour will play a large role. This means that the need for a more complete understanding of the
unsteady behaviour of vortical flows is becoming increasingly important.

The increased flight envelope of many of these ahcraft has resulted in many manoeuvres being executed at tran­
sonic velocities. As mentioned above, this introduces the presence of shockwaves, which interact with the leading
edge vortices. Generally, shocks appear due to localised supersonic regions and as the flow behaviour changes,
say with an increase in incidence, the location and strength of the shocks present in the flow will change. This can
have a significant effect on the overall flow behaviour and, as a result effect the performance of the aircraft whilst
canying out manoeuvres in this flight regime. Due to this, it is important to understand the overall behaviour of the
CHAPTER L INTRODUCTION

Figure 1.1: N ASA F-18 HARV at a = 20^ with smoke visualisation showing vortex created by leading edge
extension and vortex breakdown interacting with aircraft tailplane

flow and any interactions which occur between the shock and the vortical flow, in order to avoid significant loss o f
lift due to a sudden breakdown o f the coherent vortex caused by the presence o f shocks on the wing.

Traditionally, these issues were considered using experimental techniques, with large scale wind tunnel tests being
carried out to determine the behaviour o f the flow for various flight regimes. However, with the development o f
computational methods and the rapid advances in computer technology, computational fluid dynamics (CFD) has
emerged as an increasingly useful tool in the understanding o f aerodynamic flow behaviours. The use o f CFD to
compliment experimental testing in both research and industrial design processes is being increasingly realised
by scientists and engineers in both fields. CFD reduces the need for expensive, large scale testing programs by
allowing indication o f regions o f interest in the flow regimes before testing com m ences. CFD also allows many
situations which cannot be examined in wind tunnels, such as more realistic flight conditions and scenarios, to be
considered.

However, it is very unlikely that CFD will ever replace experimental testing, particularly as the nature o f turbulence
is still not fully understood. A great deal o f research has been carried out in recent times to create models to allow
the accurate simulation o f a turbulent flow. However, many o f these methods are based on empiricism or statistical
methods and have limits to their accuracy. Many high fidelity methods have also been proposed, which directly
solve the governing equations o f fluid flow. However, these methods are particularly expensive in computational
resources and are not in widespread use, particularly for realistic engineering flows. Therefore, the application o f
turbulence models and treatments within CFD and their ability to accurately predict interesting aerodynamic flow
behaviours is a factor which needs to be addressed for all types o f flows.

Before outlining the main objectives o f this thesis and the work carried out during this project, it is necessary to
provide an introduction to the behaviour o f vortical flows over slender, sharp-edged delta wings and to consider
the application o f CFD to resolve this behaviour.

1.2 Vortical Flows over Slender Delta Wings


1.2.1 Leading Edge Vortices
A s fluid passes over a sharp-edged delta w ing, set at an incidence, a , the flow separates along the sharp leading
edge and together with the separating boundary layer from the lower surface forms free-shear layers. These shear
layers curve upward and curl in on them selves to create two counter-rotating vortices over the upper surface o f
the wing. As the shear layer returns to the surface o f the wing, it induces a span-wise flow in the direction o f
the leading edge. If and when this flow meets an adverse pressure gradient it w ill separate again from the surface
creating secondary vortices. In som e cases this process can occur again below the secondary vortices creating
tertiary separations. The overall behaviour is shown in Figure 1.2. The primary and secondary separation and
attachment lines on the w ing surface, which are created by the stream surfaces o f the flow impinging on the
surface, and the surface streamlines, which flow from attachment lines toward separation lines are shown.
CHAPTER 1. INTRODUCTION

C rossflow plane

Prim ary v o n e x core

S econdary vortex

S econdary attachm ent line (A jl

' ' A xially attached flow

A dachm enl stream line^


P rim ary attachm ent line (A ,)

P rim ary separation line (S |)i

S econdary separation line (S2)

Figure 1.2: Schematic o f the subsonic behaviour o f the flow over a delta w ing at incidence (from Ref. [1])

A lso clear from Figure 1.2 is the increasing diameter o f the vortex cores with distance from the apex. There have
been many experimental and numerical investigations into the behaviour o f vortices. Eamshaw [36] proposed that
the vortex may be split into three main regions; the shear layer, the inviscid rotational core and the viscous subcore.
The shear layer, as mentioned above, is created at the leading edge o f the wing and feeds vorticity into the inviscid
rotational region by curling in on itself. As the shear layer moves upward and over the wing, smaller substructures
are found to occur within its structure [37, 38, 39, 40], which cause an increase in thickness as the distance from
the leading edge increases. These substructures w ill be detailed in a later section. The inviscid rotational region
which makes up the bulk o f the vortex, contains the viscous subcore at its centre. A schematic o f this behaviour is
given in Figure 1.3.

Free Shear Layer

Rotational Core

Viscous Subcore

Figure 1.3: Structure o f a leading edge vortex (from Ref. [2])

Other investigations have determined that there are a number o f parameters which are important in describing the
behaviour o f the primary vortex, these include the circulation, vorticity and both swirl (tangential) and axial ve­
locities [13, 41, 42]. Each o f these variables have been found to vary with distance from the vortex core and, as
such, the regions mentioned above can be defined from their behaviour. For example, the viscous-subcore may
essentially be defined as the region in between the two extremes o f swirl velocity, i.e the region where the swirl
velocity changes sign. Figure 1.4 shows the distributions o f swirl and axial velocities through the vortex core taken
from experiments by Pagan and Solignac [3], with the three regions defined.

It is clear that as the viscous-subcore is approached, the axial velocity increases. This profile is similar to that o f
a swirling jet. The flow is in fact accelerating along the core and it is clear from the profile that the maximum
axial velocity is approximately 2.5 times the freestream velocity. It has been found in other experiments by Payne
et al. [41] and Mitchell [13] that the axial velocity can reach velocities up to three times the freestream velocity
conditions. The maximum value reached within the vortex core is dependent on the incidence as shown in Figure
1.4 and also on the sweep angle o f the wing. It was found by Wentz and Kohlman [7] that the vortex strength,
which is related to circulation and vorticity amongst other parameters, also increases with increasing incidence.
CH APTER!. INTRODUCTION

<p = 75“ x/o = 0,27

Potential S h e a r L a y e r Kcgi<

In v isc id R o ta tio n a l C o te R egion

-25

V /U o
2

Figure 1.4: Profiles of swirl and axial velocity through a vortex core, detailing the three main vortex regions
(adapted from Ref. [2], originally from Ref. [3])

Wentz and Kohlman also showed that the strength of the vortex is dependent on the sweep angle of the delta wing,
such that as the sweep angle is increased the vortex strength decreases. This was further considered by Hemsch
and Luckring [43], who determined an analytical relationship to relate the sweep angle and vortex strength. The
location of the vortices has also been found to be dependent on the angle of incidence and the sweep angle of the
wing. From a number of experiments [44, 45, 46], it has been found that with increasing incidence, the vortex core
moves inboard and further away from the wing surface. Increasing the sweep angle was also found to move the
primary vortices inboard and closer to the surface of the wing.

Due to the majority of the vortex behaving in an inviscid or potential manner and the fact that the primary separa­
tion is fixed to the sharp leading edge, it is generally accepted that the effect of Reynolds number on the behaviour
of the primary vortex is negligible. This was confirmed in an experiment carried out by O’Neill et al. [44] on a
series of 60" and 70" delta wings where the vortex trajectory was investigated for a range of Reynolds numbers,
with only a very small difference being observed. However, other aspects of the flow are affected by Reynolds
number. In particular the secondai y and tertiary separations and the behaviour of the shear layer.

The secondary and tertiary vortices are less intense than the primary leading edge vortices and occur due to the
separation of the crossflow as described previously. The location of the secondary separation is determined by the
type of boundary layer, either laminar or turbulent, over the wing, which in turn is determined by the Reynolds
number. Due to a greater susceptibility to adverse pressure gradients, the laminar separation occurs earlier and
therefore further inboard on the delta wing surface. This means that a transition from a laminar to turbulent flow
on a delta wing may be indicated by an outboard inflection of the secondary separation line. Generally, these
smaller vortices affect the primary vortex by moving its position inboard and lifting it up off the suiface of the
wing [39, 47]. The size and strength of the secondary and tertiary vortices are also determined by the behaviour
of the boundary layer. In work carried out by Hummel [48], it was found that for a laminar boundary layer and
separation, the formation of the secondary and subsequent vortices, due to the spanwise pressure gradients, causes
a peak in the suiface pressure distribution greater than that of the primary vortex. This peak occurred in the vicinity
of the separations outboard of the primary vortex. It was also found, in comparison, that for turbulent boundary
conditions and separation that this peak is relatively flat and thus, less than the peak caused by the primary separa­
tion.

Due to the highly rotational nature of the flow within the vortex core which causes a region of high vorticity, the
local static pressure is relatively low. As the primary vortex is situated in relatively close proximity to the wing
surface, the impingement of the vortex on the surface results in a region of low pressure on the wing [49, 50, 51].
This suction force was investigated in the work by Polhamus [52], who split the lift of a sharp edged delta wing
into two components, potential and vortex lift. The vortex lift is the contribution to the overall lift created by the
suction of the leading edge vortices. Due to this extra component of lift, the presence of the leading edge vortices
are generally beneflcial to the performance of delta wings.
CHAPTER 1. INTRODUCTION

1.2.2 Vortex Breakdown


At som e point, under the influence o f external and internal instabilities, an abrupt change in the the vortex core
occurs. The vortex core expands, the axial flow stagnates and downstream, the flow becom es highly unsteady and
turbulent in nature. This process is known as vortex breakdown. Initially, at a low incidence, breakdown does
not occur over the wing and may occur downstream o f the trailing edge. However, with increasing incidence, the
breakdown position m oves upstream and crosses the trailing edge at a critical angle. Further increases in incidence,
cause the location o f breakdown to m ove further upstream on the wing until it reaches the apex, where the wing is
said to be stalled.

From vortex tube experiments, Faler and Leibovich [53] detailed many different forms o f vortex breakdown. How­
ever, Lamboume and Bryer [4], who studied the vortex breakdown process in detail, noted that two main types
occur in flows over delta wings: bubble and spiral breakdown. Both types o f breakdown are shown over a delta
wing in Figure 1.5. Bubble breakdown exhibits an axisymmetric behaviour and is generally characterised by the
occurrence o f a stagnation point on the vortex core axis with a region o f reversed flow downstream. The remaining
outer flow passes over this bubble as a bluff body before being entrained into a region o f turbulent flow down­
stream. For spiral breakdown, Lamboume and Bryer [4] suggested that three stages occurred. A deceleration
o f the vortex core, an abrupt “kink” in the vortex core, where the vortex filament spirals around a stagnant flow
region and, finally, a further breakdown into large-scale turbulent flow downstream. A schematic diagram o f this
behaviour is shown in Figure 1.6 The spiral form o f breakdown is generally much more common over slender
delta wings. However, both these types o f breakdown behaviour were found to occur in the experiments carried
out more recently by Payne et al. [4 1 ,5 4 ] for a series o f delta wings with various sw eep angles at a range o f angles
o f incidence. Bubble bursts have also been found to occur occasionally in computational investigations [5].

Figure 1.5: Leading edge vortices and types o f vortex breakdown over a 65" delta wing. The upper vortex exhibits
the spiral form o f breakdown and the lower vortex shows the bubble type (from Ref. [4])

Direction ol rotation
of tfie spiral

S e n s e of
tfie spiral

Original axis
of vortex

Spiraling

Direction of rotation
of tfie upstream vortex

Figure 1.6: Spiral vortex breakdown occurring over delta wings (adapted in Ref. [5] from Ref. [4])
CHAPTER I. INTRODUCTION

Evidence o f the spiral structure was found from the experimental investigation by Klute et al. [6 , 55] for a 75"
delta wing at an incidence o f 40". The visualisation was performed in a water tunnel facility through the use o f
Digital PIV techniques on a 2-dimensional plane through the vortex core. The helical form o f breakdown was
w itnessed by considering streamlines on a plane through the vortex core region as shown in Figure 1.7. The foci o f
the streamlines, calculated from the velocity field on the measurement plane, indicate the location where the helical
structure intersects the plane and were found to occur in a staggered pattern. This view o f the helical nature o f
vortex breakdown was also found in the instantaneous vorticity PIV results o f Ôzgôren et al. [56] and by Towfighi
and Rockwell [57], again using PIV techniques.

Similar details o f the spiral breakdown behaviour were witnessed in the computational investigation by Visbal [10]
on a 75" delta wing at a Mach number o f M = 0.2, Reynolds number o f 9.2 x 10^ and for a range o f angles o f
incidence, 17" < a < 34". It was found, for this low Reynolds number, that the time-averaged view o f the spiral
structure displayed characteristics o f an axisymmetric bubble type breakdown. However, in the instantaneous
results and through streakline visualisations, the spiral form was clear. The behaviour o f the spiral was also
captured on a plane through the vortex core and a number o f flow variables were plotted to visualise its structure.
In a plot using contours o f vorticity, the spiral structure was observed, suggested by small staggered regions o f
opposite sign vorticity. This was confirmed by the use o f streamlines on the same plane which exhibited clear foci
in the regions o f the concentrated vorticity in a similar manner to the results o f Klute et al. [6 ,5 5 ] described above.

D elt# W in g V o n « x B ra a k d o w n

Figure 1.7: Instantaneous evidence o f the spiral nature o f breakdown t = 0.0 3 7 , shown on a plane through the
vortex core (from Ref. [6 ])

There have been many investigations into the movement and sensitivity o f vortex breakdown to internal and ex­
ternal parameters both over delta wings and within vortex tubes. From these investigations, many theories have
been proposed to explain the cause o f vortex breakdown, which include an analogy to a 2-dimensional boundary
layer, hydrodynamic instability and critical state (wave) theories which are explained in detail in Délery [49], Hall
[50] and Escudier [58]. Much work has also been carried out on the theory that a critical parameter or relationship
exists at which stagnation and mass disorganisation occurs in the flow. These criteria are generally based on inter­
nal parameters such as the swirl velocity, axial velocity and adverse pressure gradient and include, a critical value
o f swirl ratio (or Rossby number), based on critical states theory and stability o f the vortex [46, 59, 60], a critical
value o f helical angle [61, 62], a switch in sign o f the azimuthal vorticity [5 7 ,6 3 ] and a critical value o f circulation
[64].

It has been found that there are a large number o f external factors which also have an important effect on the
behaviour o f breakdown. These include, an external adverse pressure gradient [65], which was found to move
the position o f breakdown upstream, geometric effects such as the inclusion o f centre-bodies or sting geometries
[66 , 67], sw eep angle and leading edge properties [68 , 69] and the proximity o f wind tunnel walls [70, 71]. In the
investigation by Wentz and Kohlman [7], the effect o f the sw eep angle on the critical incidence at which breakdown
occurs over the wing was determined. It was found that with increasing sw eep angle the onset o f vortex breakdown
could be delayed for sw eep angles lower than approximately 75". This is shown in Figure 1.8.
CHAPTER 1. INTRODUCTION

Figure 1.8: Effect o f sweep angle on the occurrence o f vortex breakdown at the trailing edge (from Ref. [7])

In a recent review by Jobe [72], the position and movement o f vortex breakdown over 65® delta wings is studied.
The study details and collates the results gathered from many different investigations carried out over the years
on various delta wing geometries for incompressible flow. Experimental, computational and empirical data was
considered for a range o f flow conditions. There is a large scatter o f data for the range o f angles o f incidence
tested. Som e o f this scatter is attributed to differences in leading edge geometry and centre-bodies. However, even
among the investigations carried out on very similar geometries there are differences, which may be attributed
to the unsteadiness o f the vortex position (which can oscillate with an amplitude o f 20 %Cr [ 8]) or to the way in
which the point o f breakdown is determined. However, a general trend was obvious from all the data for increasing
incidence. Examples o f this can be seen in Figure 1.9.

Huang and B anff


Lamboume and Bi>er
0.8 Weniz and Kohlman

0.6-

0.4

0.2

Figure 1.9: M ovement o f vortex breakdown over a 65® delta wing with increasing incidence (data from Ref. [ 8])

The data shown is from three well-known experiments by Huang and Han If [8 ] on a w ing with sharp, symmetric,
10® bevelled leading edges and a wing centre-body, Lamboume and Bryer [4] on a wing with no centre-body and
a flat leeward surface and Wentz and Kohlman [7] on a wing with 15® symmetrically bevelled leading edges and
again no centre-body. From these results, it can be seen that for subsonic, incom pressible flows that the movement
o f vortex breakdown upstream is relatively gradual for the 65® delta wing. These three experiments also help to
demonstrate the relevant conclusions o f the review, namely that the position o f breakdown is delayed downstream
by the addition o f a symmetric centre-body and also by an increase in lower surface bevel angle. A number o f
attempts have been made to correlate breakdown results and to determine an analytical relationship which will
quantify the position o f vortex breakdown considering a number o f the influential parameters, mostly geometric.
These were briefly summarised and tested on a large database o f results by Gursul [73]. Unfortunately, none o f the
relationships allowed for a collapse o f the data to a single line and thus the relationships were not deemed to be
useful.

The occurrence o f vortex breakdown has been found to have a pronounced influence on the aerodynamic charac­
teristics o f swept wings. In particular, there is a significant effect on the creation o f lift on the wing. As mentioned
CHAPTER 1. INTRODUCTION

previously, the leading edge vortices cause an increased suction on the upper surface of the wing, which creates
vortex lift. The theory which describes this contribution was proposed by Polhamus [52] and is based on a leading
edge suction analogy which does not require knowledge of the behaviour of the leading edge vortex system, only
assuming that the flow reattaches on the upper surface. However, as pointed out in the review of Lee and Ho [9],
Polhamus’ theory is not valid when vortex breakdown occurs over the wing and so is limited in the complete analy­
sis of lift generation over delta wings. It was discovered in investigations by Wentz and Kohlman [7], O’Neill et al.
[44] and Johari and Moreira [74] that for wings with sweep of more than 70^, that the point at which breakdown
passes over the trailing edge coincides with the occurrence of maximum lift. This implies that vortex breakdown
is detrimental to the production of lift. However, for wings with sweep angles below 65", maximum lift does not
occur until breakdown is almost at the apex of the wing and the wing is close to stall. Therefore, it can be assumed
that the relationship between the occurrence of vortex breakdown and the generation of lift is highly complex and
that other factors are important, such as sweep angle and vortex strength.

It has been generally thought that with increasing sweep angle, that the contribution of vortex lift increases [7, 9].
However, there is also compelling evidence that in fact the opposite is true. In a paper by Hemsch and Luckring
[43], Polhamus’ theory is considered and manipulated to provide a relationship between the change in sweep angle
and the vortex lift. This was achieved by considering the non-linear part of the vortex lift and its relationship to a
change in sweep angle. It was found from this investigation that with increasing sweep angle that the contribution
of vortex lift decreases. It also showed that the overall contribution of non-linear lift increases with increasing
sweep angle, therefore although the overall contribution of the vortex lift has decreased with increasing the sweep,
the amount of non-linear lift generated has increased. This may help to explain the differences in relationship
between the point of maximum lift and the point at which breakdown crosses the trailing edges for the wings
described above, as it may be suggested that the linear production of lift is not as susceptible to the effects of
breakdown.

In an investigation by Earnshaw and Lawford [75], the lift coefficient against incidence for a number of delta
wings with various sweep angles was plotted. The results are shown in Figure 1.10. It is apparent from this
graph that the lift characteristics of the 65" and 70" swept wings are the most favourable. It is also clear that with
decreasing sweep angle the magnitude of C,, is reduced along with the angle at which it occurs. At first this may
appear to dispute the results of Hemsch aud Luckring [43], however, with consideration of the vortex breakdown
characteristics over lower swept wings, this may be due to the theory not accurately predicting the full lift generated
over the wings.

X 76° DELTA 1.4


• 70° DELTA 1,2
+ 65° delta
' 60“ DELTA 1.0
- 55“ DELTA
- 45“ DELTA
Cu 0.6
0 .4
0.2

“ 60 -4 0 -2 0 20 40 60

-0 J 5
- 0.8

Figure 1.10: Lift coefficient vs angle of incidence for different sweep angles [9]

1.3 Unsteady Aspects of Delta Wing Vortical Flows


1.3.1 Vortex and Vortex Breakdown Instabilities
The occuixence of vortex breakdown causes an increase in the unsteadiness in the flow over delta wings. In the
investigation carried out by Earnshaw and Lawford [75], it was found that as the breakdown location crossed the
^11

CHAPTER 1. INTRODUCTION 9

trailing edge of the wing, there was a significant increase in the fluctuations of the measured normal force coef­
ficient. Coherent fluctuations due to breakdown have also been witnessed in surface pressure readings [76] over
delta wings and from vortex tube experiments [77, 78]. In the study conducted by Gursul [79], it is concluded that
the fluctuations downstream of the breakdown location are caused by a hydrodynamic instability which manifests
itself in the first helical mode. This was determined from the measurements taken by two pressure transducers
situated in the flow downstream of breakdown. The helical mode instability, which is determined to occur over
delta wings with various leading edge sweep angles, is described as a helix of the rotating vortex core filament. The
sense of this helix is found to be in the opposite direction to the vortex rotation upstream of breakdown. However,
the whole structure also rotates, with the same sense as the vortex core. This is, therefore, a description of the spiral
mode of breakdown, which was described in Section 1.2.2 and shown in Figure 1.6. As stated, this is the most
commouly witnessed mode of breakdown over delta wings. Through the analysis of the unsteady measurements
it was found that a dominant frequency could be associated with the helical mode instability, which reduced with
increasing incidence and decreasing sweep angle. For all delta wings tested, this frequency was found to occur
in the range St ^ 0.5 — 2. It was also determined that the frequency of the instability decreased with increasing
streamwise location i.e. along the vortex axis, suggesting that the pitch of the helix is increasing, therefore, the
spiral is being stretched downstream.

In the computational investigation by Visbal [10] on a 75" delta wing, discussed in Section 1.2.2, the unsteady
nature of the spiral breakdown is also considered and analysed. As mentioned, a plane tlrrough the vortex core
was considered and the instantaneous spiral structure was indicated by staggered regions of opposite sign vor­
ticity. From this plane, the increase in radius and pitch of the helical configuration, mentioned by Gursul [79],
was found. The development and movement of the helix was also witnessed through a number of consecutive
instants, as the regions of vorticity were found to move downstream. Spectral analyses were carried out on the
pressure signals measured under the vortex breakdown region. From the power spectral density (PSD) plots of
the data, it was found that a number of dominant peaks occurred. The largest peak for the majority of the data,
taken at different positions on the wing and at different angles of incidence, was centred around a non-dimensional
frequency of approximately St — 3.2. This was suggested to correspond to the frequency of the rotation of the
spiral structure. It was noted, however, that this frequency increased slightly with increasing incidence, due to the
upstream progression of the breakdown location. Other peaks which were noted to occur in many of the results
had non-dimensional frequencies of approximately St — 1.3 and 2.0. No suggestions were made to the cause of
these peaks. From further consideration of the overall behaviour of the spectral data with increasing incidence,
it was observed that with increasing incidence, and therefore stronger breakdown, the frequency response broadens.

This unsteady structure was also considered experimentally by Klute et al. [55], discussed in the previous section.
From the instantaneous velocity calculated on a plane through the vortex core, the downstream progression of
the streamline foci was witnessed, indicating the rotation of the spiral breakdown, as mentioned above. The foci
were also found to be accelerating downstream at different rates, suggesting that the radius and wavelength of the
spiral increases. This is in agreement with the findings of both Gursul [79] and Visbal [10]. The time histories of
streamwise velocity, taken from the DPIV data at a number of points downstream of the breakdown location, were
analysed using PSD techniques. From the analyses at a constant streamwise location from the breakdown location,
a number of dominant frequencies were found at approximately St — 0.44, 1.72 and 2.78, with the St — 1.72 con­
sistently exhibiting the highest energy at all points considered. Further analysis at varying streamwise locations
downstream of breakdown showed that with increasing distance from breakdown, the dominant frequency of the
flow decreased. This was attributed to the increase of the radius and wavelength of the spiral mentioned above.
This dominant frequency for the helical mode instability (St 1.7) was also captured in the inviscid computa­
tional results of Gortz [23]. The calculations were carried out on a 70" delta wing at 27" incidence and with Mach
number M = 0.2. This frequency was calculated from flow visualisations where the period of one rotation of the
breakdown spiral was observed to be approximately 0.008 seconds.

As well as the unsteadiness within the vortex breakdown region there is also unsteadiness in the location of vortex
breakdown. In the computational study by Visbal [10], described above, a high amplitude, low frequency oscilla­
tion was found to occur due to the motion of the breakdown location. This is shown in Figure 1.11. It is clear that
the large scale amplitude of the breakdown oscillation is approximately 9%c,-. The conesponding non-dimensional
frequency of this oscillation was found to be St % 0.075. A higher frequency, low amplitude oscillation was also
found, however the resolution of this was not sufficient to allow the frequency to be determined. This unsteady
behaviour has also been found to occur in experimental investigations, such as Huang and Hanff [8], Garg and
Leibovich [77], Payne et al. [41], Lowson [80] and Mitchell et a l [81]. From these investigations it has been
found that the breakdown location can oscillate with an amplitude of as much as 20%c,- [8]. In the results of

I
Payne et al. [41], it was found that these oscillations had an amplitude of approximately 2%c,- and occuixed in an
CHAPTER 1. INTRODUCTION 10

antisymmetric manner over the left and right hand side of a full span wing.

0.95-)

a so -

m o 130.0 mo mo

Figure 1.11: Streamwise fluctuations of vortex breakdown location at a — 32" (from Ref. [10])
In the experiments carried out by Mitchell et al. [81] on a 70" delta wing, this behaviour was also considered and
the frequencies associated with this phenomenon were determined. The tests were carried out in a wind tunnel
for a range of angles of incidence at Reynolds numbers of 9.75 x 10^, 1.56 x 10*^ and 2.6 x 10^. Laser sheet flow
field visualisation techniques were used to determine the breakdown location and a time history of the behaviour
was obtained for both leading edge vortices. From this study the asymmetiy of the breakdown location was also
witnessed and an interaction between the vortices was assumed. For all flow conditions tested, the frequencies of
the breakdown oscillations were found to be in the range St — 0.0443 —0.0697. The amplitude of the oscillations
was found to be as much as 20%c,- depending on the incidence and freestream velocity. The amplitude of the
oscillation appeared to increase with increasing Reynolds number and decrease with increasing incidence.

Further work to consider the sensitivity of this phenomenon to Reynolds number was reported by Lambert and
Gursul [82]. In tlris study wind tunnel tests were carried out at a Reynolds number of 1.6 x 10^ on a 80" delta wing
at a = 50". The unsteady behaviour of the flow was measured using surface pressure transducers downstream of
the location of breakdown and the dominant frequencies were determined from the resulting analysis. The vortex
breakdown oscillation was determined to correspond to a peak frequency of St = 0.15 and two further frequencies
at A 1.5 and St — 2 were attributed to the helical mode instability. Comparison witli the results of other similar
investigations carried out at relatively low Reynolds numbers [83, 84, 85] showed that the behaviour of the break­
down was insensitive to Reynolds number.

In the investigation canied out by Gursul and Yang [86] on a 70" delta wing an attempt was made to determine the
cause of these fluctuations and whether the helical mode instability could have an effect on the behaviour. From
consideration of the frequency domain data gathered from LDV analysis of the flow downstream of breakdown it
was determined that the non-dimensional frequency of the helical mode instability was dependent on the incidence
and ranged between approximately St — 1.72 and 3.5. Similarly, the dominant peaks associated with the oscilla­
tion of breakdown location was defined to occur in the range St = 0.07 —0.12. It is clear that the non-dimensional
frequency response of the breakdown fluctuations is an order of magnitude lower than the response associated with
the spiral breakdown. Dne to this, it was determined that the helical mode instability was not directly responsible
for the breakdown fluctuations. It was stated that, by this argument, the Kelvin-Helmholtz instability of the shear
layer was also not responsible for the fluctuations of vortex breakdown position as its dominant frequency is known
to be an order of magnitude higher than that found for the helical mode instability.

Further investigation into the behaviour and origins of the breakdown oscillation was carried out by Menke et al.
[15] for a series of delta wings with varying sweep angle and for a range of angles of incidence, a = 25" —42"
using flow visualisation and LDV techniques. Again, this study found that the specific dominant frequency and
amplitude associated with the fluctuation of the breakdown location was dependent on the angle of incidence.
However, all dominant frequencies occurred in the range St — 0.04 - 0.12. A dependence on the sweep angle of
the wing was also determined. Evidence of the asymmetry of the breakdown oscillations was found from consider­
ation of the full wing. It was shown that the oscillations occur at the same frequency, however they are out of phase
by approximately 180". It was determined that this asymmetry was due to an interaction between the vortices,
caused by a streamwise instability of the two breakdown regions and is non-linear in behaviour. However, with
the inclusion of a splitter plate, it was found that the oseillation of breakdown was still present although witli a
significant reduction in amplitude and RMS behaviour.
CHAPTER 1. INTRODUCTION 11

Jumper et al. [64] suggested a simple criterion for vortex breakdown based on a critical value of the circulation
of the vortex. From consideration of their model, it was suggested that the oscillation of breakdown was due to a
fluctuation of the circulation within the vortex core upstream of breakdown. As the circulation changed, the point
of breakdown would move, either upstream or downstream, to a stable location in response to this change in up­
stream flow conditions. This was concluded from consideration of the rotational direction of the spiral breakdown
in relation to the rotation of the vortex core and to the resulting induced velocity, which is in the opposite direction
to the axial flow of the vortex upstream.

A consequence of vortex breakdown asymmetry and fluctuation is vortex interaction at high angles of incidence,
where the vortices move inboard and become much closer. In an experimental investigation by Menke and Gursul
[87], the overall unsteady nature of the leading edge vortices over a 75" sharp leading-edge delta wing was con­
sidered. The experiment was carried ont in a water tunnel at a Reynolds number of 4.1 x 104. It was found from
LDV measurements and consideration of probability data that large amplitude velocity fluctuations occurred in the
core of leading edge vortices upstream of breakdown position and even in cases where breakdown was not present.
Frequency spectra were considered of the velocity time histories from three span-wise locations at x/c,- — 0.6.
Each of the three positions gave noise-like, broad-band responses, with no discernible peaks. As the frequency
range under consideration covered a number of unsteady phenomenon, such as the fluctuation of breakdown lo­
cation, helical mode instability and the Kelvin-Helmholtz instability, these could be ruled out as the cause of the
fluctuations. Another factor considered was vortex interaction, which was also discounted. It was consequently
suggested that tliese velocity fluctuations were caused by a random “wandering” of the vortex core. Suggestions
were also made as to the cause of the vortex wandering, such as the upstream influence of the turbulent unsteady
flow in the wake and the effect of three-dimensional instabilities in the shear layer, however no conclusions were
reached. In another investigation by Gursul and Xie [88], a link between this wandering behaviour and the pres­
ence of the Kelvin-Helmholtz instability in the separated shear layer was determined. It was suggested that this
interaetion between the shear layer instability and vortex wandering was due to the small scale vortices being con-
vected around the primary vortex, therefore displacing the vortex core through the process of Biot-Savart induction.

This vortex interaction and unsteadiness can, at high angles of incidence beyond the stall angle of the wing, result
in vortex shedding from the wing. This was found for a 76" delta wing in the investigations by Rediniotis et al.
[89, 90]. In the investigation the wing was tested at angles of incidence between 35" and 90" at Reynolds numbers
ranging between 3.9 x 10"^ and 9.02 x 10^. At angles of incidence greater than approximately 36" periodic shed­
ding was found to dominate the wake region of the wing. The behaviour of the periodic shedding was not found
to be influenced by Reynolds number, however the angle of onset was sensitive. For higher Reynolds numbers
the angle of onset was found to be slightly lower than for low Reynolds number. For lower angles of incidence,
in-phase shedding was witnessed on the wing, with vortices being shed at the same time from both leading edges,
however with a further increase in incidence above 70" a second shedding mode was discovered. This mode was
found to be an alternate shedding of vortices which occurs with the in-phase shedding. It was also found that the
non-dimensional frequency (Strouhal number) of the periodic vortex shedding decreased with increasing incidence
and occurred in the range St = 0.05 —0.4. It was proposed that this shedding was due to the shear layer separating
and no longer being able to create the swirling flow with a significant axial motion. This would result in vortical
structures being shed to the freestream.

In the experiments carried out by Gursul and Xie [11, 91], the transition from the helical mode instability of
breakdown to vortex shedding was investigated. The experiments were carried out in a water channel, nsing LDV
and flow visualisation techniques on a 75" sharp leading edged delta wing. LDV data was obtained from a plane
perpendicular to the wing surface, situated at the trailing edge. From this data, at different angles of incidence
between 31" and 70", the changing behaviour of the flow was observed. This is shown in Figure 1.12. It is clear
from this figure that with increasing incidence, the dominant frequency of the helical mode instability is found to
decrease, whereas the frequency of the vortex breakdown oscillation and interaction is virtually constant. However,
a change in the behaviour was found at an incidence of approximately 60", where the characteristic swirling flow
disappears at the trailing edge and a separated shear layer region appears. These results were compared to the
results from Rediniotis et al. [89] and Gnrsul [79] described above. The RMS velocities are also considered, which
show the highest velocity fluctuations occuiTing initially within the vortex core region, and then with increasing
incidence, in the shear layer itself. Spectral analyses of these velocity fluctuations were carried out for all angles
of incidence. It was found that for the swirling flow, that two frequencies were dominant, which were virtually
constant over the wing, corresponding to non-dimensional values of St 0.07 and % 1.0. The lower frequency
is consistent with fluctuations of vortex breakdown location as discussed above whereas the larger frequency was
attributed to the helical mode instability which is still present over the wing. These two peaks are found to occur in
the flow up until the critical incidence of approximately 60" mentioned before. Above a — 60" where the swirling
CHAPTER 1. INTRODUCTION 12

flow disappears, the frequency response changes and there is only one dominant peak. The non-dimensional
frequency o f this dominant peak is found to be St % 0.3 which corresponds to the frequencies determined for
vortex shedding in previous measurements o f a delta wing wake [79]. As the change in dominant frequency o f the
flow is relatively sudden it is concluded that the transition in flow behaviour to vortex shedding is also abrupt.

. p raasu ra (x/oa0.89)(Quraul 1*94)


A velocity (x/o-1.5)(RadinioUs at al.1990)
o velocity (x /c -1 .0) preaont axparim ants

1.5

8 1 .0
Helical mode
instability

0.5 Vortex shedding

Vortex interaction
\
20 30 40 50 60 70 80 90
A ngle of attack

Figure 1.12: Variation o f non-dimensional frequency for unsteady phenomena as a function o f angle o f incidence
(from Ref. [11])

1.3.2 Shear Layer Instabilities


As mentioned previously, a shear layer is formed as the fluid flows over the leading edge o f the delta wing, which
rolls up and creates the leading edge vortices. However, the shear layer is subject to a number o f instability
phenomena. These instabilities have been seen to cause the occurrence o f vortical sub-structures within the shear
layer, which can be divided into three main forms [14]:

1. An unsteady form where the discrete vortical sub structures m ove with time through the shear layer.

2. A steady laminar form, where the sub-structures are spatially fixed.

3. A mean stationary form observed in time averaged solutions o f transitional/turbulent shear layers.

The first o f these, an unsteady, time-varying instability, was first found by Gad-El-Hak and Blackwelder [38, 40]
for both 60" and 45" delta wings through flow visualisation techniques. It was found that within the shear layer,
discrete vortical sub structures occurred which pair up and rotate around each other, as shown in Figure 1.13.
These sub-structures could not be seen with the naked eye and only through the flow visualisation. They were
found to exist all along and parallel to the leading edge and occurred at a frequency which was dependent on the
freestream velocity. This type o f behaviour is w ell documented for the developm ent o f two-dimensional shear
layers between two streams o f differing freestream velocity [92] and is known as the Kelvin-Helmholtz instability.
Therefore, it was proposed that the behaviour was caused by a similar Kelvin-Helm holtz instability arising on the
shear layer.

Figure 1.13: Kelvin-Helmholtz shear layer instabilities (from Ref. [6 ])


CHAPTER L INTRODUCTION 13

lUfj At' 10HXÎÜ

• - »
’ .Q - - 4 r >

A , I INQIXI. A -, IK .^ X )

[dt A , I 7J 4 ( k i . > i:.ln o

Figure 1.14: Shear layer instabilities and the effect o f Reynolds number (from Ref. [12])

Further evidence o f these unsteady structures was found in the time-accurate computational study carried out by
Gordnier and Visbal [93]. In this investigation a 76" delta wing was used at an incidence o f 20.5". The flow con­
ditions applied were a Mach number o f 0.2 and Reynolds numbers o f 0.5 x 10^ and 9 x 10^. The higher Reynolds
number case was used in a previous study where the results were found to be unsteady [94]. This unsteadiness was
attributed to the Kelvin-Helmholtz instability in the shear layer. A time step study was carried out, resulting in a
time step o f A t = 1.25 x 10 For improved agreement with the experiments o f Gad-El-Hak and Blackwelder
[38] the lower Reynolds number was used for the majority o f the calculations. U sing instantaneous plots o f the
flow, the unsteady structures were shown to occur in the shear layer region and the roll up o f the sub-structures
were clearly seen. Time-histories o f the pressure at different streamwise positions were measured and the dominant
frequencies were considered. It was determined that the dominant frequency was almost linearly dependent on the
streamwise position, with the Strouhal number decreasing with increasing chordwise position. In the experimental
findings o f both Gad-El-Hak and Blackwelder [38] and Lowson [39] the shedding frequencies o f these structures
are found to be uniform, only dependent on Reynolds number. However, the frequencies which occurred at the
trailing edge were found to be consistent with the experimental values found. The effect o f these structures on the
surface pressure coefficient distributions was shown to be small for this Reynolds number however, it was stated
that for the higher Reynolds number used the temporal effect was greater.

In the study by Riley and Lowson [12], the developm ent o f the shear layer was investigated using flow visualisation
and LDA techniques. An 85" delta wing was used, set at an incidence o f 12.5". In this investigation both steady and
unsteady instabilities were found to occur and it was determined that this occurrence was dependent on Reynolds
number. For Reynolds numbers less than approximately 3 x 1 0 “* the shear layer appeared to be fully laminar with
no clear structures being witnessed. However, with increasing Reynolds number streamwise structures first appear
then becom e more distinct. An example o f this effect and the shear layer structure is shown in Figure 1.14. This
also clearly shows the occurrence o f turbulent disturbances in the flow at the trailing edge which move upstream
with increasing Reynolds number.

This dependence on Reynolds number was also found in the investigation carried out by Lowson [39]. After con­
sideration o f the onset o f the unsteady instabilities, it was suggested that their appearance was dependent on tunnel
velocity and therefore, the unsteady structures witnessed were a result o f external instabilities in the tunnel and not
a generic part o f the flow. However, as mentioned before, in the original investigation by Gad-El-Hak and Black­
welder [38], it was noted that frequencies o f the unsteady instabilities were dependent on the freestream velocities.
This would, therefore, suggest that the instability would develop at the m ost unstable frequency in the flow, which
may coincide with external disturbances [95]. This Kelvin-Helmholtz instability has also been witnessed in many
CHAPTER 1. INTRODUCTION 14

numerical calculations [14, 96, 97] which do not contain any external disturbances. Therefore, it must be assumed
that this type of instability is an inherent part of the shear layer behaviour and not a result of external influences.

LDA measurements were taken at a number of positions within the flow, and from the results, laminar steady
sub-structures were observed. These conespond to the second type of instability mentioned above which were
also found from the flow visualisation results to be visible to the naked eye and very sensitive to external distur­
bances. In considering the behaviour of these stationary structures, it was found that their path was helical around
the main vortex. It was also discovered that the velocity profiles were wake like when near to the leading edge
but as the structures moved away, the velocity deficit with respect to the freestream disappeared. Their strength
also increased with distance from the leading edge due to the diffusion of vorticity to the flow. The origin of these
sub-structures at the leading edge was also investigated and it was determined that the presence of the secondary
vortex was important. In a discussion of the cause of these sub-vortices being due to a cross-flow instability as
suggested by Washburn and Visser [97], described below, it was stated that the secondary vortex may be the cause
of the necessary spanwise gradient. In conclusion it was stated that the theory of a cross-flow instability was in
essence, identical to that of the three-dimensional Kelvin-Helmholtz instability, therefore these co-rotating vor­
tices must be due to a local Kelvin-Helmholtz mechanism occurring in the streamwise vortex feeding sheet. This
process was also observed in tlie DNS calculation by Shan et a l [98] on the same 85" delta wing. The results
obtained from the calculation were very similar to the experiment, however the laminar steady structures were not
witnessed. Unfortunately, no further experimental or computational investigations have been reported which have
also observed these laminar, steady structures for highly swept wings.

Gordnier and Visbal [96] carried out an investigation into the origin of the unsteady shear layer phenomenon,
which was inspired by the findings of Riley and Lowson [12], detailed above, concerning the effect of external
disturbances. Calculations were performed for three different wings with sweep angles of 70", 75" and 85", at a
range of angles of incidence from 10" to 25" with two Reynolds numbers being considered: I x 10^ and 5 x 10'^.
The grid used was based on one detailed for a previous investigation [93]. A post-processing technique was utilised
which allowed a simulated laser light sheet to be created in order to compare the results with experimental flow
visualisations. Validation was cairied out for a 70" delta wing and very good agreement was found for the flow
visualisation, with similar sub-structures observed in the computational results. Further study was carried out to
attempt to explain the mechanism which causes these unsteady structures and to determine if this could be same
mechanism which occurs in experiments. It was determined from a grid refinement study that the axial grid reso­
lution effects the shedding frequency of the sub-structures, however, no noticeable change in the overall behaviour
of the flow was found. From consideration of all the results it was suggested that the shear layer unsteadiness
was due to a boundary layer eruptive behaviour caused by the interaction of the primary vortex with the snrface of
the wing. The disappearance of these structures for low Reynolds number was also witnessed and was explained
through the elimination of the eruptive behaviour. The effect of the incidence and sweep angles detailed before,
was attributed to the increased strength of the primary vortex, thus amplifying the behaviour and causing more
unsteadiness in the flow. It was also determined that the shedding frequency behaviour with chordwise position
becomes increasingly linear with greater unsteadiness.

The third type of structure was found in the investigation by Washburn and Visser [97]. In this study, three delta
wings with sweep angles of 70", 76" and 80" were used to investigate the behaviour of steady sub-vortices. A
five-hole probe was used to obtain velocity and pressure data to allow the measurement of the conditions for the
sub-structures to occur and for their helical paths to be defined. The experiments were caiTied out for a range
of angles of incidence between 10" and 25" and Reynolds numbers between 0.5 x 10*^ and 2 x 10^ for each of
the wings and sub-structures were found for almost all cases. Due to the limitations of the five-hole probe data,
temporal instabilities could not be measured. Therefore, the structures observed were mean and steady in nature.

It was determined from the study that the size and rate of production of these vortices was dependent on the
incidence and sweep angle. An increase in the incidence or a decrease in the sweep angle resulted in an increase in
frequency and a decrease in the size of the sub-vortices. An increase in stiength of the shear layer was attributed
to the rise in frequency. It was also noted that with the same change in parameters that the structures formed
closer to the leading edge. The paths of the sub-structures were shown to be helical, as initially assumed, however
no evidence was found to support the theory that they are entrained into the vortex core downstream. From
consideration of the vorticity behaviour, it was determined that these vortices were co-rotating with the primary
vortex core and based on this, a theory to their cause was suggested. It was believed that the structures were due
to an inviscid instability in the shear layer which was based on a cross-flow instability, similar to that found in
three-dimensional boundary layers, where the resulting vortices are also found to be co-rotating. These steady
state, mean structures were also found in the experimental investigations canied out by Payne [42], Mitchell et al.
CHAPTER 1. INTRODUCTION 15

[99, 100] and Honkan and Andreopoulos [101]. Figure 1.15 shows the helical paths o f these structures from the
results o f Mitchell e ta l. [99, 100] captured using LDV techniques on a 70" delta wing.

Figure 1.15: Stationary shear layer sub-structures shown using contours o f x vorticity in planes perpendicular to
the wing surface (from Ref. [13])

In a D N S investigation by Visbal and Gordnier [14], a 75" semi-infinite delta w ing is considered to determine the
behaviour o f the shear layer without any external influences such as the presence o f vortex breakdown or trailing
edge effects. The flow conditions used correspond to an incidence o f 25" and a Mach number o f 0.1. The Reynolds
numbers considered are dependent on the length o f the region o f interest and range between 6 x 1 0 ^ and 5 x 10"*.
The effect o f Reynolds number was considered, in order to determine the effect without the presence o f external
forcing such as those witnessed in experiments. The behaviour computed was similar to that o f the experiments
with the present o f a steady laminar vortical system at low Reynolds numbers and the increase in unsteadiness o f
the shear layer as the Reynolds number is increased. It was noted that the unsteady behaviour began toward the
trailing edge o f the wing and moved toward the apex with increasing Reynolds number as also found in the results
o f Riley and Lowson [12]. It is also determined that within a small range o f R eynolds number, the complexity o f
the flow increases dramatically. A greater appreciation o f the three-dimensional com plexity o f the flow and the
differences caused by change in Reynolds number can be gained from Figure 1.16.

Repon / R e g i o n II Rcsum /// Region JJI


Region I Region II

stdc

(a ) Re = 2.5 x ! ( / (b ) Re = 5 x \ ( f

Figure 1.16: Instantaneous shear layer structure shown by iso-surface o f axial vorticity for two Reynolds numbers
(From Ref. [14])

It is clear from these diagrams that the flow is easily split into three streamwise regions, as suggested by the au­
thors. Region I corresponds to a region where no sub-structures occur and the flow is found to be essentially steady.
Region II refers to a region where the shear layer structures are evident, and appear to be well organised with a
helical path around the vortex which, as they start close to the wing, are parallel to the leading edge then become
further inclined toward the apex. In animations, it is stated that these structures are clearly seen to rotate around
CHAPTER 1. INTRODUCTION 16

the vortex. Further downstream, in Region III, the shear layer appears to be affected by further instabilities and
therefore the flow becom es more com plex. This further complexity may be caused by the sub-structures breaking
down into further discrete concentrations o f vorticity which continue in the helical path around the vortex. It is
suggested that this is due to a secondary instability o f the Kelvin-Helmholtz instability. Again, as in their previous
investigation [96], it is suggested that the instabilities in the shear layer occur due to the interaction between the
upper surface laminar boundary layer flow and the primary vortex. These results rule out the occurrence o f un­
steady separation from the trailing edge as a cause for the unsteadiness within the shear layer.

Consideration was given to the time-averaged or mean representation o f the flow, which would allow closer val­
idation with the experimental results gained by Washburn and Visser [97] and M itchell et al. [99, 100]. Figure
1.17 shows the resulting plots. These plots show the shear layer, characterised by stationary helical sub-structures,
which are co-rotating with the primary vortex as found in the experiments. From further consideration it is found
that there are regions o f high RMS velocity fluctuations within the shear layer, which appear to correspond to the
positions o f the helical sub-structures.

It is noted that the time-averaged structures only appear on the aft section o f the w ing (corresponding to Region III
from before) with the rest o f the time averaged shear layer appearing smooth. It is suggested that this behaviour
is explained by the secondary instability occurring in Region III mentioned in the discussion o f the instantaneous
results. If this secondary instability occurs with a sufficient periodicity and wavelength, then it is suggested that
it would be viewed in the time-averaged results. Therefore, the conclusion is made that the “stationary” and
“unsteady” shear layer structures are not necessarily two separate phenomena but may in fact be different view s o f
the same physical behaviour. It is noted that the laminar steady structures as shown in the investigation by Riley
and Lowson [12] for a highly slender delta wing, have not been witnessed in this investigation.

(a ) Re = 2.5 x 10* (b) = 5x 10*

Figure 1.17: Time-averaged shear layer structure shown by iso-surface o f axial vorticity for two Reynolds numbers
(from Ref. [14])

1.3.3 Unsteady Flow Topology


From consideration o f the literature summarised in the previous sections, an overall picture o f the unsteady behav­
iour o f delta wing vortical flows can be obtained. It is clear that there are many unsteady phenomena which exist
in the flow, these include (in no particular order):

• Helical mode instability

• Shear layer instabilities

• Vortex Shedding - both from the trailing edge and at high angles o f incidence

• Vortex core rotation

• Shear layer reattachment

• Vortex breakdown oscillation

• Turbulence downstream o f breakdown


CH APTER!. INTRODUCTION 17

It has been found that characteristic frequencies exist in the flow which can be associated with each of these phe­
nomena. From the literature review it is clear that there have been many investigations which consider the unsteady
behaviour of vortical flows. Table 1.2 summarises these investigations, and details the frequencies assigned to each
phenomenon. The column marked “OtheT' contains the frequencies which appeared in the investigations but were
not assigned to any phenomenon. Some of the computational results detailed will be considered with respect to
the numerical methods used in a later section.

This table allows a general appreciation of the frequency content to be obtained and patterns emerge relating to
the order and size of the dominant frequencies. For example it would appear that the majority of the frequencies
assigned to the helical mode instability fall between St = 1 —2 and similarly for the oscillation of vortex breakdown
location the majority of the investigations show this to occur with a Strouhal number between St = 0.04 —0.2.
Menke et al. [15] performed a similar analysis of the flow behaviour from work carried out by Gursul [79], Gursul
and Yang [86], Gad-el-Hak and Blackwelder [38] and Gordnier and Visbal [93]. From this, a schematic of the
frequency spectrum was created in an attempt to classify the unsteady frequencies. This is shown in Figure 1.18.
Further consideration of the unsteady behaviour of the flow over delta wings can be obtained from the reviews by
Gursul [95, 102].

,o

E oj "o o

.s §
ÎE
< l l
S i

0.001 0.01 0.1 1 10 100


Dimensionless frequency fc/Uco

Figure 1.18: Spectium of unsteady flow phenomena as a function of Strouhal number (from Ref. [15])

1.4 Transonic Effects on Vortical Flows


As the freestream Mach number is increased to transonic levels, M > 0.7, the vortical flow changes and behaves
differently to that for the subsonic regime. In the experiments carried out by Erickson et al. [117, 118], it was
found that increasing the Mach number from 0.4 to 0.95, changed the shape of the leading edge vortex. The vortex
was found to become flat and elliptical in appearance and sat progressively closer to the surface of the wing. It
was also found that with the increase in Mach number through the transonic regime, that the suction induced on
the surface of the wing by the leading edge vortices was decreased due to a fall in the upwash created by the
leading edges. From the experimental study by Elsenaar and Hoeijmakers [18] on a 65" cropped delta wing it was
also found that with increasing Mach number, the magnitude of the primary suction peak decreases, broadens and
moves inboard for a given chordwise station. The secondary peak was also found to move inboard.

As the Mach number increases, it is found that the flow becomes locally supersonic and as a result shockwaves
will appear in the flow, further altering the behaviour of the leading edge vortices. Stanbrook and Squire [119]
determined that the change in behaviour of the leading edge separated flow could be correlated by considering the
Mach number and incidence normal to the leading edge, defined respectively by
fta n a \
Mat = M «.\/1 - sin^ A costa and aN — tan ( 1. 1)
VcosA 7
Using these flow parameters, the flow behaviour could be split into two main types of flow, separated and attached.
These flow behaviours were separated by what is known as the Stanbrook-Squire boundary. Miller and Wood
[16] gave further consideration to the types of flow over delta wings for transonic and supersonic regimes from
experimental results on a number of delta wings with varying sweep angles for a range of Mach numbers and
angles of incidence. From the analysis of the results, they classified the flow into six types of behaviour, including
classical vortex, vortex with shock and shock-induced separation. These flow behaviours were also defined by
CHAPTER 1. INTRODUCTION 18

?
d
\o
d

s
O 2 § |
o
q O o 0 0 = d \o

ill
O VO d d
Io

I
s
s °

I
d d

a
<êi
m
d d
I
o
1

in m

I o
in CN
CN
?
R a
CO rd
in
I
a

I
VO
^ ‘
o

'd

o d> X o o VO m' ' o o


: X
in
o o O 't O O —I
s d)
X X X iri X X OV CO
X X VO X X t VO m x x o
VO oo o\ I—
VO
in X '— 1 v q in o vq in t> I CN
(N T-i CO v d '—1 r—I O v CN CO 'if

d d
m
gON oin
I
in r- CO I CO
m & o o
CO in <N

o
<
oo
VO Q Q
b o o" ^
I
o in in o in o o o m oo 00
< c- r- C-- oo 00 !> VO in N CO CO

1
If
p y
' 0Û

I
g | o

Ti *d
I
la
ti 'a
<o Q

I
<U & ^

iJ
«U rO
a
(2
CH APTER!. INTRODUCTION 19

in o
(N
oi r-T
CN
CO
Î o
Tj- o\ cn
^ ^ in

1
O '- 2
VO

i
d

s
\q
d
d ê
r""
o

II I
in

in o
^ M "i
' I o
VO 2 in

if I
VO
Ov

T*-

I
CN
CO

d
d o o
X d
X in d X X
c~ X VO VO VO
g m VO X in g in
CN CN in

d
Si I
Si Si
CO CN in
d
CN
I
o o o VO
1> in m 1>

_
iS] w ov
CO r-1 m
I

1
13 13 13 13
1 O' rO O) rO
g
TO 'O 'd

Î
8 8 8 g
„ CO
4J CN

'ë "§ 'ë Cj


§ 4D :
d o o o o o - -,

U Ü O Ü O Ü ^ i>
CHAPTER 1. INTRODUCTION 20

the normal Mach number and incidence used by Stanbrook and Squire and thus the classification diagram was
redefined. This is shown in Figure 1.19.

CLASSICAL
VORTEX
REOION

SEPARATION
BUBBLE WITH
NO SHOOK >
SHOCK WITH NO
SEPARATION

0 1 2
Mr

Figure 1.19: Classification of flow behaviour over delta wings by Miller and Wood [16]

In the detailed review by Narayan and Seshadri [120] on tlie transonic and supersonic behaviour of delta wings,
further classification of the flow behaviour is considered. This takes into account the individual behaviour of the
shocks in the flow and their location relative to the leading edge vortices. This provides a further three types of
behaviour. However, all these behaviours can be considered as sub-types to the classification as defined by the
Stanbrook-Squire boundary - leading edge attached flow and leading edge separated flows. Transonic flow over
delta wings generally falls into the leading edge separated category with leading edge vortices being formed. How­
ever, depending on the Mach number, shocks are found to be present.

A large number of investigations, both experimental and numerical have been canied out, which have looked at
the occurrence and behaviour of shockwaves in vortical flows for varying transonic conditions [17, 18, 117, 121,
122, 123, 124, 125]. From these investigations, a number of shockwave systems have been observed and detailed
in the literature. In the investigation of the transonic behaviour of delta wing flows carried out by Elsenaar and
Hoeij makers [18], the presence of two main shockwave systems on the upper surface of the wing is discussed
based on conjecture and experimental data. These are:

1. Underneath the primary vortex, at an approximately constant spanwise position, just outboard of the primary
suction peak;

2. On the aft section of the wing, close to the trailing edge perpendicular to the plane of symmetry.

These shocks are termed cross-flow and rear/terminating shocks respectively. Using theoretical reasoning, it is
stated that the cross-flow shock causes the secondary separation under the primary vortex. As the incidence is
increased for a particular Mach number, the shock forms under the vortex creating a large adverse pressure gradi­
ent, which results in the separation of the boundary layer. This shock is the reason for the inboard movement of
the secondary separation, mentioned above. It was determined that for a Mach number of 0.85 at the chordwise
position x/cr = 0 .6, that the switch from pressure gradient separation to shock-induced separation occurs at an
incidence of 14.5^ for this configuration. These cross-flow shocks may be apparent from the surface isobar plots
as tight contours of pressure coefficient, however in the spanwise pressure coefficient distributions the position of
these cross-flow shocks are not clear.

The occurrence of the rear shock is found from consideration of the chordwise pressure coefficient distribution at
the plane of symmetry. At low angles of incidence and at low Mach numbers the distribution along the plane of
symmetry gradually decreases toward the trailing edge as the flow conditions return toward freestieam conditions,
however at moderate angles of incidence with increasing Mach number there is a sharp change in distribution near
to the trailing edge. This sharp change characterises the occurrence of a shock-wave in this region. It was found
in this investigation that for an incidence of 15^ that the flow becomes supersonic over the wing at a Mach number
between 0.8 and 0.85 and there is clear evidence of a rear shock-wave for these conditions. This rear shock was
found to move downstream with increasing Mach number.
CHAPTER 1. INTRODUCTION 21

The occurrence of the two main shock systems was also determined from the experiments of Erickson et al. [117,
118], mentioned above, using surface pressure measurements and from surface reflective visualisation techniques
used by Donohoe and B annink [17,122]. The surface reflective technique is a type of Schlieren visualisation which
allows three-dimensional aspects of a flow to be captured and observed clearly. The experiments were carried out
on a 65" delta wing at M = 0.8 and from the results, schematic diagrams were produced showing the proposed
behaviour and shape of the shocks in the flow. These are shown in Figure 1.20.

Terminating shock surface -

Undefined
region —,

cross-flow
shock wave

Embedded cross-
flow shock wave

(a) Embedded cross-flow shock (b) Rear/terminating shock surface

Figure 1.20: Schematic diagrams showing proposed positions and shapes of shock systems over transonic delta
wings [17]

It is clear that Figure 1.20(b) shows the rear shock as being perpendicular to the symmetiy plane at the centreline
of the wing, but then arcing downstream toward the primary vortices and appearing to intersect the vortex region.
From the side view visualisations, it was also noted that at the symmetry plane the shock extends initially perpen­
dicular from the wing siuface, then curves upward toward the apex before returning to a perpendicular direction
until it disappears. This shock was witnessed at an incidence of 15" where breakdown did not occur and it was
noted that it did not appear to disrupt the vortical flow. Due to this, it was proposed by the authors that the shock
moves above the vortices as it curves downstream. These cross-flow and rear/trailing edge shocks have also been
found in many computational investigations [121, 123, 124, 125, 126].

In the computational investigation canied out by Visbal and Gordnier [125], the effects of compressibility were
considered for a 75" delta wing for a range of Mach numbers at a constant Reynolds number. From the results of
the calculations a number of shocks were witnessed in the flow for each Mach number. As well as the cross-flow
shock underneath the primary vortex and the rear/terminating shocks described above, a two further shocks were
witnessed on tlie wing. These were an upper cross flow shock, which sat above the primary vortex and a centreline
shock, which sits parallel to the wing suiface above the symmetry plane. The upper cross-flow shock has also been
found experimentally for transonic delta wings [117, 118, 127].

The occurrence of these shockwave systems in the flow introduces the complex behaviour of shock/vortex and
shock/boundary layer interactions [49, 128]. This is particularly important when considering the behaviour of
vortex breakdown for transonic flows as the breakdown behaviour is quite different to that witnessed for subsonic
vortical flows where the onset of breakdown is relatively gradual with increasing incidence [72]. The behaviour of
vortex breakdown was also detailed in the investigation by Elsenaar and Hoeij makers [18] mentioned before. The
differences in the flow behaviour pre- and post-breakdown are shown using surface isobars and chordwise distrib­
utions of pressure coefficient at the symmetry plane. The pre-breakdown flow is shown at an incidence of 22" and
post-breakdown, at an incidence of 24". It is highlighted that within this relatively small incidence increment the
position of breakdown is noted to jump from x/c,- = 0.8 to x/c,- = 0.5. The presence of the shock systems detailed
above ai'e apparent from the results, with the cross-flow and rear shock being clear for the pre-breakdown case.
At moderate incidences, the location of this shock moves downstream toward the trailing edge with an increase
in Mach number and incidence suggesting tliat its strength increases with an increase in these parameters. For
the post-breakdown case, the cross-flow shock is still witnessed upstream of the breakdown position, however,
there are now two rear shocks on the wing. The position of these shocks is clearly seen from consideration of the
chordwise pressure coefficient distributions, which are shown in Figure 1.21.
CHAPTER 1. INTRODUCTION 22

X 20 NO
- 1. 0 - A 22 I
VORTEX BREAK-DOWN
SHOCK

YES

-0 .5 - '

SHOCK

0,5
SHOCK
P U N E OF SYMMETRY
M ,^«O ,05. Ra =9x10®

Figure 1.21: Chordwise pressure coefficient distribution at the symmetry plane for a range of angles of incidence
pre- and post-breakdown [18]

Considering these results, it was found that the first rear shock is situated approximately at the point of breakdown
on the wing with the second shock appearing at roughly x/c,- = 0.9. It is proposed that the first rear shock men­
tioned is actually the terminating shock from the pre-breakdown flow shifted upstream, however it is uncertain as
to whether this upstream shift causes or was caused by vortex breakdown. It is conjectured, based on transonic
flow over an airfoil, that the rear shock is weak and thus a small change in conditions downstream, caused by
vortex breakdown, could force the shockwave to jump upstream to a new equilibrium position within the flow. The
presence of the second shock is explained by the flow reaching supersonic conditions downstream of breakdown
and thus returning to subsonic conditions before reaching the trailing edge.

In the investigation by Donohoe and Bannink [17], the presence and cause of vortex breakdown is also considered,
and similar visualisations to that mentioned earlier, were carried out at higher angles of incidence. At an incidence
of IB'^, asymmetric breakdown was witnessed over the wing. This phenomenon was found to occur on either
the port or starboard side of the wing, for the same conditions with the breakdown position rapidly fluctuating as
much as 0.4c,- on either side. It was found that on the side on which breakdown occurs the terminating shock also
moves with the position of breakdown and is thus also highly unsteady. Therefore, it is noted that this shock must
interact with the breakdown in some way. Similarly, at 20" incidence, symmetrical breakdown is witnessed over
the wing and the initial terminating vortex is seen to move upstream with the breakdown position, but retain its
bowed appearance. This confirms the observations made by Elsenaar and Hoeijmakers [18] detailed previously. It
was also found, for the M = 0.8 case, that a double terminating shock system appeared, with a second similar rear
shock appearing at the trailing edge. Donohoe and Bannink suggest that this second shock may be caused by the
acceleration of the flow in the centre region of the wing due to the symmetrical breakdown causing an effective
nozzle about the wing centreline [17]. The position of breakdown is also noted to oscillate with time for this case
but the magnitude of these fluctuations is much less than that for the asymmetric case. However, the frequency
appears to be higher.

An experimental and numerical investigation was carried out by Houtman and Bannink [129], on a 65" sharp edged
delta wing at high subsonic and transonic speeds. In the experiment, at a Mach number of 0.85 and an incidence
of 20", it was found that vortex breakdown occurred over the wing and that the flow exhibited an “irregular” be­
haviour which was not found for lower Mach numbers. This irregularity was observed in the spanwise pressure
distribution at x/c,- = 0.7, where the suction peak was seen to collapse for these flow conditions. This collapse was
attributed to vortex breakdown occurring over the wing and was shown in surface flow visualisation pictures oc­
curring at approximately x/c,- = 0.65. Therefore, it is clear that the onset of transonic breakdown causes a sudden
and complete loss of suction on the wing, characterised by the collapse of the surface pressure distribution suction
peak. Again, as before, two shocks were noted above the wing surface normal to the wing surface and the symme­
try plane. These were located between x/c,- = 0.5 and 0.6 and at approximately x/c,- = 0.825 from the chordwise
pressure distribution at the root chord and from Schlieren pictures. It was proposed that the downstream shock was
created by the vortex breakdown phenomena as the upstream shock was witnessed prior to breakdown occurring,
although the position of this shock is not stated. Consideration was given to the effect of varying the Mach number
for a fixed incidence and it was found that for slightly lower Mach numbers of 0.75 and 0.8 there was a similar
pattern within the pressure coefficient distributions. However, no evidence was found of locally supersonic flow
or rear shocks at these speeds. Due to this, the conclusion was made that the shockwaves occurring in the flow at
high subsonic freestream Mach numbers do not have a large influence on the location of vortex breakdown.
CHAPTER 1. INTRODUCTION 23

From the flow visualisations o f D onohoe and Bannink [17, 122] it was determined that the rear/terminating shock
could exist for low to moderate angles o f incidence and that vortex breakdown did not occur. It was proposed, due
to this, that the shock sat above the vortex region and did not interact with the vortex core. However, what happens
at this point is not w ell understood, and whether interaction occurs for lower angles o f incidence is not known
conclusively. From the study o f the interaction between longitudinal vortices and normal shocks in supersonic
flow [28, 130, 131] it has been found that it is possible for a vortex to pass through a normal shock without being
weakened sufficiently to cause breakdown. However, the flow over slender delta w ings is more com plex as the
shock does not appear to be normal to the freestream in the vortex core region. Therefore, it may be concluded
that it is possible for a terminating shock system to exist without the breakdown o f the vortical system, particularly
at lower angles o f incidence. The presence o f the embedded cross-flow shock was also witnessed in D onohoe and
Bannink’s experiments [17, 122] and was found to occur almost to the trailing edge. This suggests that for this
incidence the vortex is strong and thus, undisturbed by the presence o f the trailing edge or indeed the rear shock.

The behaviour o f shock/vortex interaction and transonic vortex breakdown was considered in the computational
investigation o f Kandil et al. [123, 124] using inviscid and laminar methods. In this study a 65" delta wing
was considered at Mach numbers o f M = 0.85 and 0.9 0 , Reynolds number. Re = 3.23 x 10^ and angles o f in­
cidence o f a = 2Œ^ and 24". From the results o f the calculations, the cross-flow and rear/terminating shocks
were determined, with the cross-flow shock causing the separation o f the boundary layer to form the secondary
vortex. Upstream o f the rear/terminating shock, strong leading edge vortices were noted to occur, but im medi­
ately downstream o f the shock location, the bubble form o f vortex breakdown was found. With the increase in
incidence, the rear/terminating shock and the breakdown location m ove upstream by 20% Cr. It was also found
that this shock location moved downstream with increasing Mach number, in agreement with the observations o f
Elsenaar and Hoeijmakers [18]. Unsteady calculations were also carried out which showed that downstream o f the
rear/terminating shock the flow is highly unsteady with periodic fluctuations being present. An oscillation o f the
shock location and therefore breakdown location is also witnessed. The flow upsteam o f the shock was found to
be steady in nature.

The unsteady nature o f the shock and breakdown location was considered by Kameda et al. [19] using PSP tech­
niques on a 65" flat plate delta wing. The wing was tested at a Mach number o f M = 0 .9 0 and incidence o f a = 20".
Vortex breakdown was found to occur on the wing for this incidence and was noted to be caused by an interaction
with the rear/terminating shock. The presence o f the rear/terminating shock was noted from a significant increase
in surface pressure detected by the PSP method and is clearly shown in Figure 1.22. The suction created by the
leading edge vortices is clear close to the apex o f the w ing, and it is evident that this suction disappears at the same
chordwise location as the region o f high pressure. The sequence o f PSP results indicates the unsteady nature o f the
shock and breakdown location.

10 nOkPa

Figure 1.22: Snapshots o f pressure distribution on the surface o f the wing using PSP techniques (from Ref. [19])

An inviscid numerical investigation to consider the behaviour o f breakdown location with increasing incidence was
undertaken by Longo [121]. Three delta w ings with varying sweep angles were used and the effect o f increasing
Mach number on the forces and moments o f each w ing was considered. From the calculations it was found that
as the sw eep angle is decreased the effect o f Mach number increases. This was particularly evident for the 60"
w ing where a sudden drop in lift coefficient occurs at the point at which vortex breakdown crosses the trailing edge
follow ed by a flat recovery. This sharp change was also seen in the moment coefficient and has been attributed
to a fast upstream shift in the vortex bursting location causing a large loss in vortex lift. However, this sudden
behaviour in the force and moment curves did not appear for the higher swept w ings. Further investigation into the
movement o f breakdown position with increasing incidence for the 60" w ing show s that the point o f breakdown
CHAPTER 1. INTRODUCTION 24

moves from the trailing edge to a position of approximately x/c,- = 0.35 in 5".

The behaviour at M = 0.8 was also detailed considering the presence of vortex breakdown on the flow. For the 60"
wing, it is found that breakdown occurs downstream of the trailing edge at 14" incidence. At this incidence, the
flow under the vortex was found to be fully supersonic and a small region of sonic flow appeared at approximately
x/c,- = 0.4 near the symmetry plane. A cross-flow shock was predicted close to the trailing edge but no rear shock
was found above the wing. With increasing incidence, the cross-flow shock was found to move upstream with the
position of vortex breakdown and a terminating shock appeared at the trailing edge due to the flow downstieam
of breakdown becoming supersonic. However, the location of this shock was not found to move with increasing
incidence as described by Elsenaar and Hoeijmakers [18]. For the 70" delta wing, the cross-flow shock behaved
more like the experiments detailed previously and was situated under the vortex, upstream of breakdown, for all
angles of incidence. However, it was noted that the shock was relatively weak for this case. There was no termi­
nating shock predicted for any incidence over the 70" wing, which is in agreement with the results of Houtman
and Bannink [129]. It was proposed from the results, that the flow between the vortex axis and the surface of the
wing may be considered as a convergent-divergentduct, where the flow is channelled and accelerates to supersonic
speeds. This nozzle effect causes the cross-flow shocks to appear for relatively low freestream Mach numbers. It
was concluded from the analysis that the decrease in suction peak with increasing Mach number could be attributed
to the flow in this region becoming supersonic and that the increased rate of upstream progression of the vortex
breakdown position could be attributed to supersonic core velocities within the vortex upstream.

Further consideration of the sudden change in flow behaviour due to vortex breakdown can be obtained by from
detailed analysis of the results from the experimental database of Chu and Luckring [20, 132, 133, 134]. A large
series of experiments were carried out in the National Transonic Facility (NTF) at NASA Langely for various Mach
numbers and Reynolds numbers for a large range of angles of incidence. These results also form the basis for the
papers by Luckring, which consider Reynolds number effects [135, 136] and compressibility effects [137, 138]
for both sharp and rounded leading edged wings. The experiments were carried out using a 65" delta wing with
various leading edge profiles, which was instrumented with a series of 183 static-pressure ports on the starboard
side of the wing. These ports were placed at spanwise intervals along five streamwise locations, x/c,- = 0.2, 0.4,
0.6, 0.8 and 0.95, with most of the ports being concentrated on the outboard section of the wing. Pressure ports
were also placed on both (port and starboard) leading edges at the same streamwise locations to consider the sym­
metry of the flow behaviour. A sample of these results for the sharp leading edge wing under transonic conditions,
M = 0.85, Reynolds number 6 x 10^ and angles of incidence in the range, a = 18.6" —26.7" is shown in Figure
1.23. This case was chosen for analysis as it corresponds to a test case used for the 2nd International Vortex Flow
Experiment (VFE-2), which uses this configuration. Further details of the VFE-2 are given in Chapter 3.

Within the apex region, upstream of x/c,- = 0.4, it is evident that with increasing incidence, the suction peak
generally increases in magnitude, broadens and moves inboard. This inboard movement is more pronounced
between a = 19.6" and 20.6", where the suction peak also reduces slightly. Above an angle of incidence of about
20.6", the secondary peaks, which are almost as strong as the primary peaks, also increase in size. However, they
do not move inboard. Below a — 20.6", strong secondary peaks are not obvious near the apex. At the x/c,- = 0.6
position, there is a clear difference in the flow structure with increasing incidence. For the incidence range of
a — 18.6" to 23.6" there is still clear evidence of the primary and secondary vortices, which for the higher angles
of incidence have maintained their suction from the previous chordwise station. There is also evidence of a cross-
flow shock system with a sudden jump in pressure coefficient being observed just outboard of the primary suction
peak. It is also believed that this cross-flow shock will also occur upstream of this location, however the pressure
jumps are less obvious for the streamwise locations close to the apex. It is likely that this is due to the use of
the non-dimensional scale on the spanwise axis which was used for clarity between the results at each incidence.
This proposed cross-flow shock behaviour is much clearer for the downstream streamwise locations. For angles of
incidence of 24.6" and above, there appears to be a collapse in the primary vortex suction peak. It is also found that
the secondary peak and cross-flow shock disappear altogether. This is similar to the behaviour noted by Houtman
and Bannink [129] mentioned in Section 1.4, for the case when breakdown was found over the wing. Indeed, from
considering the pressuie coefficient plots at x/c,- = 0.8 and 0.95 for angles of incidence above 24.6", it is clear
that breakdown has occurred due to the relatively flat distributions. Based on the these results and observations of
Houtman and Bannink [129], it can be suggested that the location of breakdown is just upstream of the x/c,- = 0.6
position with the drop in suction peak being a direct consequence of this.
CHAPTER 1. INTRODUCTION 25

x/c. = 0.2 x/c. = 0.4

x/c. = 0.6
x/c. = 0.8

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9


x/c. = 0.95 y /s

■ 18.6" " 21.6" ■ 24.6"

• 19.6" 22.6" ■ 25.6"

" 20.6" " 23.6" " 26.7"

Figure 1.23: Experimental results from N ASA NTF wind tunnel tests for conditions: M = 0.85 and Re = 6 x 10^,
for the sharp leading edge, at a range o f angles o f incidence 18.5" - 26.7". (data taken from Ref. [20])

- 1.6

- 1.4 - 1.4

- 1.2 - 1.2

- 0.6 - 0.6

- 0.4 - 0.4
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
X /C , x/c.

(a ) M = 0 .8 5 1 , /îe = 6 X 10* (b ) M = 0 .7 9 9 , Re = 6 x 10*

Figure 1.24: Experimental results from N A SA NTF wind tunnel tests a range o f angles o f incidence 18.5" - 26.7".
Legend as shown for Figure 1.23 (data taken from Ref. [20])

Downstream at x / c r = 0.8, for the lower angles o f incidence there is still clear evidence o f the primary and sec­
ondary vortices. The magnitude o f the primary suction peak has also not decreased significantly. However, the
cross-flow shock appears to have increased in strength, suggested by the greater magnitude and gradient o f the
jump in pressure coefficient. At the higher angles o f incidence, the pressure coefficient profile is flat and uniform
over the whole span with only a very slight change in magnitude with increasing incidence. Finally, at the trailing
edge, it is found for the lower angles o f incidence that the primary peak is still evident with the cross-flow shock,
however the secondary separation appears to have disappeared. At this position it is apparent that the position o f
the vortex has moved inboard with increasing incidence and also that the magnitude o f the peak has reduced. For
the post-breakdown angles o f incidence, the pressure coefficient distribution is similar to that for the x / c r = 0.8
position, with a relatively flat profile. It should also be noted that on the lower surface for all chordwise stations
that with increasing incidence, the average pressure coefficient values increase.

From these results, it is clear that by increasing the incidence from 18.6", the vortex moves inboard and be­
com es stronger until at a certain point vortex breakdown suddenly occurs quite far upstream on the wing, close the
x / c r = 0.6 chordwise position. This is in agreement with all the results discussed previously. Further consideration
o f the vortex breakdown behaviour can be obtained by considering the pressure coefficient distributions along the
leading edges o f the wing as shown in Figure 1.24. It is clear from Figure 1.24(a) for M = 0.85 that the pressure
CHAPTER 1. INTRODUCTION 26

coefficient distributions along both port and starboard leading edges are in good agreement and that they confirm
the sudden upstream motion of vortex breakdown at cc = 24.6", which is shown by the increase in pressure co­
efficient. Interestingly, this shows that there is a change in distribution for the 23" incidence, with an increase in
pressure coefficient at x/cr — 0.8, however breakdown is not present. Looking at the spanwise distributions at this
station shows this reduction but also shows the presence of the primary and secondary suction peaks. Further flow
field data would be needed in order to be able to comment on the cause of this increase.

Looking at the same distributions for M = 0.8, it is clear that a similar pattern emerges, however at a — 24.6" it
appears that the pressure distribution on the port side exhibits signs of vortex breakdown occurring suddenly on the
wing, but that the starboard side does not. This is also the case at 25.6". However, the behaviour of both port and
starboard is the same at a = 26.6". From this data, it is evident that vortex breakdown occurs asymmetrically for
this Mach number at a critical angle of 24.6" before occurring symmetrically at 26.6" as suggested by the spanwise
distributions. Thus, from these results it is clear that asymmetric vortex breakdown occurs for a lower Mach num­
ber case but not for M = 0.85. From consideration of other datasets within the NASA results for the same Reynolds
number but differing Mach numbers, it is found that this behaviour does not form a trend based on Mach number
as asymmetric breakdown is also witnessed for M = 0.9, but not for M — 0.831 or 0.872. Thus, this behaviour
must be caused by other factors. It is worth noting that the wing was only instrumented (with pressure taps) on the
port side and it is on this side where breakdown appears first. This may suggest a sensitivity to surface disturbances.

The asymmetry of transonic vortex breakdown was also witnessed by Schrader et al. [139] for a 63" sharp edged
delta wing at angles of incidence at which breakdown first occurred on the wing. However, with a further increase
in incidence, the asymmetry was found to disappear and symmetric breakdown was found. Further evidence of
this asymmetrical flow behaviour was found from experiments carried out by Konrath et at. [140,141,142] within
the framework of the VFE-2 mentioned above. These tests involved PSI pressure transducers, Pressure Sensitive
Paint (PSP) and Particle Image Velocimetry (PIV) to further study the flow behaviour. These results were obtained
for the 65" configuration used in the Chu and Luckring experiments detailed above at a Mach number of M = 0.80,
Reynolds number Re = 3 x 10^ and at a range of angles of incidence, a = 18.4" —25.9". The PSP data from the
experimental tests are shown in Figure 1.25, showing surface pressure coefficient for all angles of incidence tested.

From this data the change in vortex flow with increasing incidence is apparent. It is clear that as the incidence
increases, the magnitude of the suction peaks increases, as witnessed for the transducer data discussed previously.
The inboard motion of the vortical system is also evident. However, the most striking feature of these plots is the
sudden asymmetric breakdown at 25.6", where vortex breakdown suddenly appears on the right hand side of the
wing close to the x/c,- = 0.6 streamwise station. This asymmetry of breakdown is in agreement with the NASA
data for a similar Mach number, however the critical onset angle is slightly higher. Unfortunately no data was
obtained for higher angles of incidence, thus it is not possible to say if this behaviour changes to a symmetric
breakdown with a further increase in incidence. It was found from the experimental tests that this behaviour was
exhibited for all transonic conditions and that the asymmetry was consistently appearing on the same side of the
wing - which coincided with the instrumentation as before [21]. Thus, this may be further indication of the sensi­
tivity of transonic vortex breakdown to surface disturbances within experimental tests. It is also clear from these
surface pressure plots, that none of the shock systems expected to occur on the wing are apparent. This is due
to the time averaged nature of the PSP technique, which, due to the highly unsteady nature of the flow, causes
the locations of shockwaves to become smeared [140]. It was also found that the PSP suction peak heights were
underestimated due to temperature effects. However, in their analysis of the flow, Konrath et al. [140] witnessed
two terminating shocks, one close to the sting tip curving downstream and a second located over the sting between
the x/cy — 0.8 and 0.9 streamwise locations. A curving of the vortex core trajectory was witnessed in the vicinity
of the sting shock from these tests.

Consideration of the cross-flow behaviour upstream and downstream of breakdown can be obtained from the PIV
results. Figure 1.26 shows the results for the post-breakdown case, a — 25.9" at six chordwise stations. From
these planes, the elongated shape of the leading edge vortex upstream of breakdown is clear and it is found that the
axial velocity at the vortex core has a magnitude of approximately 1.96L. A secondary vortex is also clear under
the vortex, close to the leading edge. Breakdown appeals to occur between x/c,- = 0.6 and 0.7 and the behaviour
of the flow changes to a large region of reversed flow, which expands downstream and is relatively circular in
nature. Inboard of this breakdown region, it is clear that the flow is still supersonic and appears to be accelerating.
This may explain the occurrence of the second rear shock witnessed by Elsenaar and Hoeijmakers [18] mentioned
above, as the flow accelerates between the breakdown regions on the wing.
CHAPTER 1. INTRODUCTION 27

a = 18.4' a = 20.5

a = 2 1 .5 0 = 22.5 a = 2 3 .6 °

0 = 24.6 0 = 2 5 .9

0.175

0.525

0.875

1.575

Figure 1.25: PSP surface pressure coefficient contours for M = 0.8, Re = 2 x 10^, a = 18.4" —25.9" [21].

(a ) x / c r = 0.5 (b ) x / c r = 0 .5 5 (c) x / C r = 0 . 6

(d ) x/cr = 0 .7 (e) x/cr = 0 .7 5 (f) x/cy = 0 .8

Figure 1.26: PIV results showing contours o f non-dimensional u velocity for a = 25.9" at M = 0.8 0 (data from
Ref. [21])
CHAPTER 1. INTRODUCTION 28

1.5 Application of CFD to Delta Wing Vortical Flows


One of the most important issues in the use of CFD, for delta wings and in general, is the choice of turbulence
model used [ 143]. There are a number of different approaches and methods to turbulence modelling, which range in
complexity, accuracy and computational expense. These techniques, in order of complexity, are inviscid, laminar,
RANS, DES, LES and DNS methods. The choice of turbulence model is normally a trade-off between computa­
tional expense and solution realism. Each turbulence treatment however, can be applied and used for the prediction
of delta wing vortical flows, with varying degrees of realism. This section will discuss the available literature of
each method applied to delta wing flows and consider the advantages and disadvantages of each method.

1.5.1 Inviscid Methods


Inviscid methods have been extensively used for the solution of the flow over delta wings. This is mainly due
to their low computational cost when compared to Navier-Stokes calculations. However, as discussed above, this
reduction in computational cost also means a reduction in the realism of the solutions. Nonetheless, the Euler
equations can give reasonable approximations. By their nature, inviscid solutions do not predict boundary layers
and therefore cannot predict separation. However, for sharp leading edge delta wings this is not a problem as the
separation is fixed at the leading edge. Once the separation occurs, it has been found that the Euler equations can
accurately predict the transport of vorticity and enti'opy within the leading edge separation and vortices [144]. Vor­
tex breakdown can also be predicted by this model although it is evident from some calculations that the strength
of the leading edge vortices is highly dependent on the grid used [145]. This will have a considerable effect on the
vortex breakdown location and behaviour.

In a review of their earlier work, Murmann and Rizzi [146], state that they found that the most important features
of the primary vortex and the vortex-wing interaction were modelled well by the Euler equations and that the
results compared reasonably with experimental data such as surface pressure and flow visualisation. The calcula­
tions performed used sharp-edged wings, where the leading edge separation point is fixed and not dependent on
viscosity. It was noted from their work that the Euler equations could not resolve the secondary separations of
the flow. This has been found to be a limitation of the use of Euler equations for delta wings, but provided this is
taken into account, the results may be assumed to be valid. It was also found for the inviscid computations, that the
level of total pressure losses predicted in the vortex cores are realistic. This suggests that the mechanism for vortex
breakdown may be driven by an inviscid phenomenon and that the fundamental structure of the primary vortices is
insensitive to the level of viscosity, as long as it is present.

The inability of inviscid solutions to accurately predict the behaviour over rounded leading edged delta wings was
considered in an investigation by Rizzi and Müller [147]. In this investigation the differences in solution between
Euler and Navier-Stokes computations on a 65° rounded leading edge wing at M = 0.85 and a = 10° were con­
sidered. It was found for the Euler computations that the position of the formation of the leading edge vortices
was delayed to approximately 25% chord of the wing. Whereas for the Navier-Stokes computations the vortices
were formed from the apex region as for shaip leading edges. It was suggested that this difference is due to the
mechanisms for the primary separations being different for the two computations, i.e. physical and computational
viscosity. The surface pressure comparison with experiment shows good agreement for the Navier-Stokes calcu­
lations but not for the inviscid solutions, A secondary separation was found to occur close to the trailing edge for
the inviscid solution, however it was determined that this was caused by the presence of a cross-flow shock under
the primary vortex and did not result in the formation of a secondary vortex. It was found that the position of
the primary vortex and the corresponding surface pressures did not agree with the experiment and, as mentioned
before, were modelled more accurately using the Navier-Stokes equations. It was concluded from this study that
the Navier-Stokes equations were needed to determine a realistic simulation of the flow over a rounded leading
edge wing, due to the dependence of the separation points on viscosity. The same wing geometry was used by Tsai
et a l [148], who came to the same conclusions in their study.

In a numerical investigation into vortex breakdown by Agrawal et a l [22, 149], both the Euler and Navier-Stokes
equations are used to simulate the flow behaviour. The purpose of the study was to determine the effects of
viscosity on the pre- and post-breakdown regions. The geometry used was the 70° wing used in the experimental
investigations by Kegelman et a l [44, 150, 151] and the results gained from these experiments are used for
validation purposes in the study. The calculations were performed for a Mach number of M = 0.3 at various
angles of incidence. The Reynolds number for the viscous calculations was Re = I x 10^. It was found from the
investigation that the Euler equations predict the position of the vortex cores outboard of the viscous solutions
and experimental results. As mentioned before, this is expected due to the secondary vortices not being resolved
CHAPTER 1. INTRODUCTION 29

by the Euler computations and therefore their effect on the primary vortex is ignored. From consideration of
the surface pressure distributions compared to the experimental results, it was determined that the Euler results
predicted tire magnitude of tire suction peak more accurately that the viscous models. It was also found that the
Euler solution consistently predicted breakdown downstream of the experimental and viscous results. This is
shown for the surface pressure contours at a = 30*^ shown in Figure 1.27 for each solution. These differences in
the flow solutions may have been due to the grid refinement in the near wing region not being sufficient to capture
the boundary layer properties for the two viscous cases. This is not an issue for the Euler solutions, which do
not model the boundary layer. The experimental breakdown location under the same conditions was found to be
approximately 50%c,-.

Figure 1.27: Surface pressure contours for Euler, laminar and turbulent computations, for M = 0.3, a = 3O'' and
Re = l x 10^ [22]
The structure of vortex breakdown using the Euler equations was investigated for a 70^ delta wing by Kumar
[152, 153, 154] using an embedded conical grid. The wing has an incidence of a ~ 30“ and a freestream Mach
number of M = 0.3. Three grids were used in the investigation and the effect of grid refinement on the solutions
was considered. From this, it was found that with an increase in grid refinement the resolution of the subcore of the
vortex improved, with an increase in axial and swirl velocities. Thus, this also confirms the dependence of the vor­
tex strength on the grid refinement as mentioned above. From consideration of the breakdown region, it was found
that the inviscid calculations predicted the spiral form of breakdown with a clear stagnation region and widening
of the core at the breakdown location. Despite the simulations not being time accurate, an oscillation of the vortex
breakdown location was witnessed and attributed a non-dimensional frequency of = 1.6. However, this result
should be considered with cai'e, particularly as it is much higher than the frequencies found for this phenomenon In
experiments - it is closer to the frequencies attiibuted to the helical mode instability discussed in a previous section.

A similar investigation was carried out by Strohmeyer et al. [155] on a 65“ cropped delta wing, to determine
the ability of the Euler equations to describe the behaviour of breakdown over the wing. The investigation was
carried out at two angles of incidence a — 10“ and 20“ at a Mach number of M = 0.2. As in the investigation by
Kumar [152] described above, the calculations performed were steady state. However, it was also determined that
the flow was unsteady downstream of breakdown with a periodic oscillation of the aerodynamic forces occurring.
From analysis of the results the breakdown behaviour was also found to exhibit a spiral structure, which gave
good agreement with the experimental data. Therefore, it was concluded that the Euler equations were sufficient
to qualitatively resolve the salient features of the flow.

To consider the ability of the Euler equations to predict tire unsteady behaviour of the vortical flow, an inviscid
investigation was carried out by Gbrtz [23] using the 70“ full span wing used in Mitchell’s experiments at ONER A.
Time accurate calculations were carried out at three angles of incidence, 27“, 30“ and 35“ for a Mach number of
M — 0.2. The time steps used for this investigation were dependent on the angle of incidence and ranged from
3.37 —3.406 X 10"^ seconds, which corresponds to non-dimensional time steps of At —2.41 —2.44 x 10"4. The
results were considered using flow visualisation methods, such as streaklines and isosuifaces of entropy. From the
flow visualisation, the spiral form of breakdown was witnessed for all three angles of incidence, and behaved as
CHAPTER 1. INTRODUCTION 30

described in Figure 1.6 . One rotation o f the breakdown region at a = 30“ was found to have a period o f T = O.OO85,
which corresponds to a dimensional frequency o f / = 1 2 5 //z and a Strouhal number o f A = 1.75. The bubble form
o f breakdown was also witnessed on the wing at intermittent intervals. The unsteady behaviour o f the flow was
also analysed from consideration o f the unsteady aerodynamic load time histories. Figure 1.28 shows normal force
coefficient time histories and power spectral density (PSD) analyses for each o f the three angles o f incidence used
in the investigation. The PSD analysis provides details o f the dominant frequencies and power o f the fluctuations
within the unsteady time histories.

0 .3 0
0 .3 5

0 .3 0

2 04 ISO 200
t|4)

(a) Time history ( b ) P S D vs. freq u en c y

Figure 1.28: Sample o f time histories and PSD analysis o f normal force coefficient from Gortz’s unsteady inviscid
calculations [23]

From the C n time history for a = 27“, it was determined that the behaviour was periodic in nature with harmonic
oscillations occurring in the signal. A s the incidence was increased to a = 30", the fluctuations were found to
becom e irregular, with an increase in magnitude. This was also the case for the a = 35" incidence, however, a
low frequency oscillation is clear from Figure 1.28. The harmonic behaviour was not found to occur for the larger
angles o f incidence. Considering the PSD results, the frequency content o f each signal is clear. At a = 27", it
is clear from Figure 1.28 that there are three clear individual peaks, which occur at frequencies o f approximately
47, 93 and 140Hz, corresponding to Strouhal numbers o f 0.6546, 1.2954 and 1.95. The dominant peak occurs at
St = 1.95 and is assigned to the helical mode instability. It was noted that this is the third harmonic o f the low­
est frequency in the spectrum. The harmonic behaviour was attributed to an asymmetry o f the location o f vortex
breakdown on the wing.

At a = 3 fr \ it is clear that more frequencies exist and the dominant peak occurs at 1 .1 8 //z (St = 0.017). Other
peaks exist in the flow at 24, 50, 118 and l6 6 H z which correspond to non-dimensional frequencies o f St = 0.34,
0.70, 1.65 and 2.32. The dominant peak is assigned to the oscillation o f breakdown location as it is o f the order
o f the results found in previous experiments [13]. The higher frequencies are all attributed to the helical mode
instability, particularly the peak at l l S H z (St = 1.65) which is in agreement with the frequency o f the rotation de­
scribed from the flow visualisation. At the largest incidence, the frequency content has again increased, with peaks
being visible at 2.87 and l8 H z (St = 0.0 4 and 0.25). A similar higher frequency content, 74 - l3 0 H z (St = 1.03
- 1.82) is also found, attributed to the upstream movement o f the breakdown location with increase o f incidence.
It was noted that the frequency o f the helical mode instability phenomenon was found to decrease with increasing
incidence, which is in agreement with the results o f Gursul [79] described in Section 1.3.1.

From these studies it is clear that despite its limitations in predicting separation and therefore the secondary vortex,
the solutions o f vortical flow over slender delta w ings are reasonable. This suggests that for sharp leading edged
delta wings, inviscid methods are sufficient to qualitatively evaluate the behaviour o f the leading edge vortices and
vortex breakdown at a much reduced computational cost compared to viscous methods.

1.5.2 Laminar Methods


The next level o f m odelling is to consider the flow as fully laminar, where the calculations are viscous but tur­
bulence is not considered. A s discussed in Section 1.2.1, a laminar boundary layer is particularly sensitive to an
adverse pressure gradient and, thus, the separations predicted by this method are larger than for turbulent flow.
This results in an over-prediction o f the secondary vortex and as a result m oves the location o f the vortex core
CHAPTER 1. INTRODUCTION 31

inboard and away from the wing surface [22]. Fully laminar calculations are generally only considered when the
Reynolds number is sufficiently low.

Gordnier [156] considered the unsteady laminar behaviour of delta wing vortices for a 65“ delta wing at an in­
cidence of a = 30“. The flow conditions corresponded to a Mach number of M = 0.2 and Reynolds number of
3.2 X lO^. Two structured grids were used in the investigation, which had a C -0 topology and approximately
3.3 X 10^ and 4.2 x 10*' grid points. A short grid study was carried out and the effect of time step on the flow
solution was also considered. For the time step study, three time steps were used, A t = 0.001,0.0005 and 0.00025.
It was found from this study that with a reduction in time step size the solution improves, with less distortion
of higher frequencies and a less diffused solution found for the smaller time steps. It was concluded from this
study that a time step of A t = 0,0005 was sufficient to resolve the main flow features. The results were compared
to experimental data and were found to exhibit good agreement for the location of breakdown, which occurs at
approximately x/c,- = 0.288 for the finer grid and at x/c,- = 0.287 in the experiments.

With the wing set to a —4“ roll angle, the unsteady behaviour of breakdown was considered, with the spiral form of
breakdown being clearly seen in the results. This breakdown location was found to be unsteady, however the length
of the calculation was not sufficient to determine its frequency. The spiral breakdown was also observed to rotate,
but again no frequency was determined. It was concluded that the behaviour of breakdown was in agreement with
experiments. Further consideration was given to tlie laminar behaviour of vortex breakdown over delta wings by
Visbal [10], which was discussed in detail in Sections 1.2.2 and 1.3.1. The investigation by Gordnier and Visbal
[96], which considers tlie cause of shear layer instabilities, discussed in Section 1.3.2 was also carried out using a
laminar flow solver and again shows the ability of this method to accurately predict low Reynolds number flows.

The unsteady behaviour of laminar predictions may be further considered from the investigation by Cummings et
a i [111, 112]. In this investigation, the flow over a semi-span 70“ delta wing is considered at an incidence of
a = 35“, Mach number of M = 0.1 and Reynolds number of Re = 4.07 x lOF, to consider the effects of periodic
suction and blowing to control the breakdown of the leading edge vortices. The calculations were carried out on
two unstructured grids with approximately 5 x 10^ and 1.24 x 10^ cell volumes. Consideration was given to the
optimal size of the time step for each grid by considering the frequency content of the normal force coefficient
time histories, from this it was determined tliat for the fine grid the time step should be At = 0.00005^ which
coiTesponds to a non-dimensional time step of approximately A t = 0.0025. Two dominant frequencies were found
within the fine grid solutions for the various time steps, which for the optimum time step coixesponded to & = 1.3
and 6.0.

The unsteady flow behaviour, without the flow control, was considered by applying pressure taps within tire com­
putational flow field at the primary and secondary vortex cores. From these taps, the time histories of the pressure
was obtained which were analysed using a PSD to consider the frequency content. In the primary vortex, prior
to breakdown, the dominant frequency of the flow was found to be St = 8.5, however downstream of the break­
down location, this dominant frequency reduced significantly to approximately St = 1.35. These frequencies were
witnessed from flow visualisations to correspond to the shedding of the shear layer from the leading edge and
the helical mode instability, respectively. The shedding frequency was also found to be dominant in the pressure
signals from the secondary vortex core region and was found to be witliin the range of frequencies predicted by
Gad-El-Hak and Blackwelder [38]. From consideration of the breakdown behaviour of the flow, it was found that
the secondary vortex broke down upstream of the primary vortex and was effecting the breakdown of the primary
vortex. Further investigation determined that this behaviour was a result of the interaction between the secondary
vortex flow and the shear layer instability mentioned. It was concluded that the three phenomenon occurring on
the wing (primary and secondary vortex breakdown and the shear layer instability) must be directly linked.

Therefore, it is clear that for low Reynolds number flows, laminar flow solutions are reliable, without the need
of large computational expense. However, as the Reynolds number is increased, transition to turbulent flow will
occur on the wing and the validity of these solution would reduce.

1.5.3 Reynolds Averaged Navier-Stokes (RANS) Methods


To obtain further realism in the computational solutions of delta wing flows, the turbulent behaviour of the flow
needs to be considered and modelled. One of the most common methods of treating the turbulence is to use
Reynolds averaging. This method effectively simplifies the instantaneous Navier-Stokes equations into mean flow
equations and deals with the contiibution of the turbulence and the resulting extra terra, known as the Reynolds
stress tensor, through separate numerical models, known as turbulence models. There are a number of different
CHAPTER 1. INTRODUCTION 32

ways in which the Reynolds stress tensor can be modelled, using linear, non-linear and algebraic formulations. The
methodology and formulation of Reynolds averaging, along with a number of turbulence models, are explained in
more detail in Chapter 2.

A large number of turbulence models exist, which are generally classified by the number of additional equations
needed to solve the complete turbulent flow behaviour. These include, Zero equation models, such as the Mixing
Length model, one equation models such as the Spalart-Allmaras model and two equation models, for example the
k — (Û model. A further set of models can be thought of as Reynolds Stress models or Algebraic Stress models.
However, not all turbulence models are directly suitable for use in considering delta wings and vortical flows.
Therefore, a great deal of research has been done into determining what models are suitable and modifying models
to give the most accurate results when compared to experimental data.

In the investigations by Gordnier [157], the linear two equation k — e turbulence model is used to calculate the
turbulent flow over a 65“ delta wing at a Mach number o f M — 0.37 and Reynolds number of Re = 3.67 x 10^.
Two angles of incidence are used, a = 15“ and 30“, to consider the flow with and without vortex breakdown
in order to examine the suitability of this model in predicting the breakdown behaviour. A structured grid was
used for this investigation which had a C -0 topology and approximately 1.1 x lO'' grid points. Results using the
Baldwin-Lomax model with Degani-Schiff correction for vortical flows were also considered and compared to the
results from the k —e model.

From the results of the investigation for a = 15“, it was found that the standard k — e model was unable to ac­
curately predict the behaviour of the flow due to unphysical and excessive amounts of eddy viscosity, which had
adverse effects on the vortex, were predicted. Therefore, modifications of the k — e model were proposed in or­
der to reduce the build-up of eddy viscosity around the primary vortex core where the eddy viscosity should be
negligible. This is a common problem for linear Boussinesq based turbulence models in predicting vortical flows
and is a result of inaccurate prediction of the normal stresses in regions of high rotation, such as the vortex core.
These modifications are based on limiting the production of the turbulence within the vortex core regions by taking
the rotation of the vortex into account. The results using these corrections applied to the standard k —e model,
showed a great deal of improvement in the resolution of the turbulence within the flow. The levels of eddy viscosity
predicted were reduced, the vortices became stronger and the results obtained became more comparable with the
experimental results. It was concluded from analysis of the pre-breakdown results, that the k —e model with a
vorticity based correction provided the best solution and thus, only this model was used for the post-breakdown
computations.

At q: = 30“, this model predicted a bubble form of breakdown, which was not in agreement with the experimental
results. In the experiment a spiral form was noted. The breakdown location was also further downstream in the
computations than for the experimental results. The discrepancy of the form of breakdown was explained by con­
sidering the RANS formulation, which calculates the mean-flow equations with the turbulence model considering
the turbulent fluctuations and considering the notion that the bubble form of breakdown is the time average of the
spiral type, as discussed previously. Thus, it was proposed that the solution was exhibiting only the mean-flow,
which would result in a bubble form of breakdown being predicted. In conclusion, Gordnier proposed that the
RANS formulation could only predict mean-flow characteristics, even if an unsteady calculation was performed.

A similar investigation was carried out by Brandsma et al. [158], which considered the effects of two similar
rotation corrections for the Wilcox k — (û linear turbulence model. The calculations were performed on a 65“
cropped delta wing at an incidence of a = 10“ at Af = 0.85 and with a Reynolds number of Re = 9 x 10^. Again,
a structured grid was used with a C-O topology and approximately 1.8 x 10® grid points. The rotation corrections
applied to the standard k - (O model were similar to those used by Gordnier [157]. One limited the production of
the turbulent kinetic energy, P^, and the other enhanced the production of the dissipation rate, 7^, in order to reduce
the eddy viscosity in regions of high rotation. The conclusions of this investigation were very similar to those from
Gordnier’s work, where it was found that the standard model over-predicted the turbulence within the vortex core,
which resulted in a weak vortex being predicted. With the modifications applied to the model, the results improved
significantly with improved agreement with experimental data. However, the model which limited P^ was found
to be more diffusive than the P^ enhancing modification and did not adequately reduce the turbulence in the core.
Therefore, it was concluded that the modification which utilised the enhancement of Peu gave the best agreement
with the experimental data and thus was best suited to the prediction of vortical flows.

Further consideration was given to the use of turbulence models for vortical flows over delta wings in the inves­
tigation by Dol et al. [159]. In this study the ability of a non-linear eddy viscosity model to predict the flow
CHAPTER 1. INTRODUCTION 33

behaviour is considered, in comparison with experiment and with the results of a standard linear two equation
model with and without rotation correction. The standard model is tlie A:—to model. The non-linear model is an
extension of this model, formulated from an explicit algebraic Reynolds sh'ess model, which incorporates an extra
anisotropic Reynolds stress term into the Boussinesq approximation resulting in an increased dependence of the
model behaviour on the mean rotation tensor. The rotation correction used is the P(o enhanced modification pro­
posed by Brandsma et al. [158]. Further details of the application and formulation of both of these models is given
in Chapter 2. The test case used for this investigation is the cropped 65^ delta wing used by Brandsma et at. [158],
described above, with the same flow conditions for angles of incidence of a = 10*, 15* and 20*. The structured
grid used is similar to that described above, with a C-O topology and approximately 1.8 x 10® grid points.

It was found from the results of the investigation that both the k —CDmodel with rotation correction and the non­
linear version of the model provided a significant improvement on the results of the standard model. Both models
reduced the predicted eddy viscosity levels in the vortex core by different means, which resulted in stronger vor­
tices being formed over the wing surface. At Oü — 15*, it was found that the results from the rotation correction
model were over-predicting the suction peak on the surface of the wing, however the non-linear model was show­
ing very good agreement with the experimental results. This was also the case at a = 20*, where breakdown was
also found to occur over the wing for the model with rotation conection depending on the initial conditions of
the calculations. From these results, it was concluded that the non-linear model performed better in capturing the
vortex flow than the linear model with the rotation correction. A similar conclusion was reached by Bartels and
Gatski [160] for a delta wing at Mach numbers of M = 0.6 and M = 0.9. The linear Spalart-Allmaras and SST
models were used and compared to results obtained using a non-linear explicit algebraic stress model. It was found
from this study that the non-linear model gave much improved results compared to the linear turbulence models
used.

In an investigation by Morton et al. [24, 161], the effect of turbulence modelling on the unsteady behaviour of the
flow is shown. A 70* semi-span delta wing is considered at a Mach number of M = 0.069, incidence of 27* and a
Reynolds number of 1.56 x 10®, which corresponds to the experimental results by Mitchell [13]. Five turbulence
models are used in this study, three RANS models, the one equation Spalart-Allmaras (SA) model, the Spalart-
Allmaras model with a rotation correction (S ARC), Menter’s Shear Stress Transport model (SST) and two versions
of a hybrid RANS/LES approach, DES, which will be discussed in more detail in the next section, are used. The
DES models are based on the Spalart-Allmaras (SADES) and Menter’s Shear Stress Transport (SSTDES) models.
All calculations were run on an unstructured grid with 2.7 x 10® cells and used a non-dimensional time step of
A t —0.005. Figure 1.29 shows a comparison of the PSD analysis of the normal force coefficient signals for each
model used.

- SA
- 8ST
- ASARC
- SACes
- SSTDES

5 10^
a.

Figure 1.29: PSD comparisons of normal force coefficient for five turbulence models [24]

From consideration of the frequency content of the unsteady results for each of the RANS turbulence models it
was found that the S-A and SST models are unable to resolve the majority of the frequencies in the spectrum.
However, the SARC model had an improved spectrum, which was attributed to the coirection eliminating turbu­
lence dissipation within the vortex core. However, this model was still found to struggle with some mid to high
frequencies associated with the post-breakdown turbulence scales. When comparing the results to the experimental
data of Mitchell et a l [162], it was found that all the turbulence models except the SST model produced break­
down positions which were comparable to the experimental data. The SST breakdown position was approximately
CHAPTER 1. INTRODUCTION 34

10% upstream. Otlier results showed that although the general position of breakdown was predicted by the RANS
models, they failed to resolve the basic characteristics of the breakdown region as shown in the experiments. It was
determined from analysis of the flow behaviour and unsteady frequencies that the DES models, allowed a more
accurate simulation of the vortex breakdown behaviour.

The unsteady behaviour of this test case was also considered by Soemarwoto and Boelens [31], with the same
flow conditions using the Po modification of the k ~ co turbulence model, proposed by Brandsma et al. [158].
As before, only the semi-span wing is considered and a structured multi-block grid with approximately 3.7 x 10®
cells is used. After a brief time step study, a non-dimensional time step of A t = 0.0025 was used for the time-
accurate simulations. From the unsteady results, it was found that the instantaneous vortex breakdown oscillated
in the range x/c,- = 0.67 —0.75. However, the time-averaged solution showed breakdown to occur at approxi­
mately x/cr — 0.74. This is downstream of the mean location found in the experiment which was found to be
approximately x/a,- = 0.65. To consider the unsteady behaviour further, the frequency content of the normal force
coefficient was analysed using PSD methods. From this analysis, it was found that a dominant frequency occurred
in the signal at approximately 200% , which corresponds to a Strouhal number of A —9. Other peaks of significant
power occurred at Strouhal numbers of approximately 11 and 20. The difference between the time-averaged and
instantaneous flow structure was shown clearly using isosurfaces of total pressure loss. This showed that the spiral
form of vortex breakdown is an unsteady phenomenon, which is instantaneous and not found in the time averaged
flow.

It is quite clear that the ability of RANS methods to predict the behaviour, both steady and unsteady, of delta wing
flows is highly dependent on the turbulence model chosen. It is evident that standard models predict unphysical
levels of turbulence within the vortex core regions, resulting in poor predictions of the vortex behaviour. A number
of rotation corrections for various models have been proposed, to sensitise the turbulence prediction to the highly
rotational flow behaviour, with varying success. However, these are essentially “fixes” specific to vortex flows
and are not based on general physical behaviour. A more general approach is the use of non-linear models, which
by their nature are dependent on both rotation and strain-rate and so add more realism to the model. However,
these models also have limitations, as the turbulent behaviour is still dictated by a numerical relationship between
the eddy viscosity and the Reynolds stresses, which may not be physical for all flow regimes. However, for the
majority of calculations RANS and UR ANS methods are relatively inexpensive and while being dependent on grid
refinement for accuracy, do not require significantly large grids or small time steps to reach solutions which may
be reasonable.

1.5.4 LES and DNS Methods


The most general methods of predicting turbulent flow are large eddy simulation (LES) and direct numerical sim­
ulation (DNS). DNS is a method which allows the full Navier-Stokes equations to be solved directly for all scales.
The size of the grids required is highly dependent on the Reynolds number of the flow and thus DNS is only real­
istically used for low-Reynolds number flows. LES is a method which uses the size of the grid spacings as a filter
to reduce the range of scales being fully resolved. The scales which are too small to be resolved on the grid are
modelled using a sub-grid model. This allows much coarser grids to be used in comparison with the DNS method,
meaning that higher Reynolds numbers can be used. However, the grid refinement required for LES calculations
is still significantly larger than that needed for RANS calculations.

Despite the complexity of these methods and the grid refinement issues in the regions of interest, there have been a
number of investigations which use these techniques to consider vortical flows. In the investigation by Mary [25],
the use of large eddy simulation was considered for the resolution of vortex breakdown behaviour over a delta wing.
The test case used was the 70* delta wing from Mitchell [13] discussed previously, with an angle of incidence of
a = 27*, Mach number of M = 0.069 and Reynolds number of 1.6 x 10®. This Reynolds number is relatively high
for a LES calculation, therefore the grid requirements are substantial. To reduce the computational expense of the
calculations a localised structured mesh refinement method was used to refine the grid sufficiently in the regions
of interest without creating an overall computational grid which was too large for reasonable calculation times.
Three grids were created with varying refinement. However, it was accepted that the grid refinement would be
relatively coarse and as such the reliability of the LES calculation for this type of flow was considered. The LES
implementation used a laminar subgrid model and two different near-wall treatments to allow a further reduction
in grid requirements. The first treatment applied a no-slip condition with a logarithmic law and the second simply
applied a slip condition to the wall.

The results and resolution of the simulations appear to be dependent on the near-wall treatment. A Q-criterion
CHAPTER 1. INTRODUCTION 35

isosurface is used to highlight the flow behaviour, where the Q-criterion is defined as the second invariant o f
Vu. The resulting plots are shown in Figure 1.30 for both near-wall treatments. With the slip condition applied
the vortex appears to be smooth and well-organised in structure, until the point o f breakdown, which fluctuates
about a mean location o f jc/c^ = 0.7. However, with the no-slip condition applied, the flow exhibits shear layer
structures upstream o f the breakdown, which wind around the vortex core region. The breakdown location was
found not to be influenced by the presence o f these structures and occurs at the same location. Comparisons o f
the mean flow results to experimental data show that the results from the slip condition provide closer agreement
with the experiment than the no-slip results. A fact which was surprising to the authors, however, neither solution
was able to predict the behaviour accurately in the post-breakdown region. D espite the slip condition giving
reasonable agreement with the experiment, by its nature no boundary layer was formed and therefore no secondary
separations occurred, meaning that the solution was not physical. It was concluded that the wall functions used
were not suitable for delta wing flow and that the grids used were insufficiently refined to capture the flow behaviour
accurately.

(a) No-Slip (b) Slip

Figure 1.30: Q-criterion isosurfaces o f vortex behaviour for two different near-wall treatments (from Ref. [25])

A D N S calculation was carried out by Gordnier and Visbal [14], which was discussed in detail in Section 1.3.2 as
it deals with the formation and behaviour o f shear layer instabilities. This calculation was performed on a sem i­
infinite delta wing with a sw eep angle o f 75* at Reynolds numbers between 6 x 10^ and 5 x 10^. These Reynolds
numbers are very low in comparison with real configurations, however are similar to experiments carried out in
water tunnels. Three levels o f grid refinement are used, with sizes between 4.8 x 10® and 6.5 x 10® grid points.
The calculations are carried out using a non-dimensional time step o f A T = 1.25 x lO” '*, which for the highest
Reynolds number provides a resolution o f 125 time steps to a single period o f the highest frequency observed.
From these values it is clear that to scale this calculation up to Reynolds numbers, even o f the order o f 10® would
mean extremely expensive calculations. However, the results obtained have a high level o f accuracy and resolution,
as shown in Figure 1.16.

D N S was also used by Shan et al. [98] to consider the behaviour o f vortex shedding from the sharp leading edge
o f a delta wing and the formation o f shear layer structures. The investigation was carried out for a relatively high
Reynolds number o f 1.96 x 10® at a Mach number o f M = 0.1. The wing had a sw eep angle o f 85* and incidence
o f a = 12.5*. For this study, the grid used had approximately 1.9 x 10® grid points. However, this appears to be
very coarse in comparison with the grids used by Gordnier and Visbal [14], discussed above, however the grid
topology is C-H, which may allow improved refinement over the wing. N o mention o f the size o f time step used is
made in the paper. The results show vortex shedding from the leading edge o f the wing, caused by an interaction
between the secondary vortex and the primary shear layer. This appears to be similar to the phenomenon described
by Gordnier and Visbal [96] using a laminar flow solver at a lower Reynolds number, mentioned in Section 1.3.2.
The unsteady nature o f this vortex shedding was considered and a shedding frequency o f approximately A = 1.086
was determined. Unfortunately, no direct comparisons with experimental results are made.

The same technique has also been applied to non-slender delta wings by Gordnier and Visbal [113, 1 14]. A 50"
delta wing was considered for a range o f angles o f incidence, 5* < a < 15* at low Reynolds numbers and Mach
numbers and the results were compared to experiments carried out under the same conditions. The grid used for
this calculation had approximately 4.5 x 10® grid points and a time step o f A T = 0.0005 was used. These were
shown to be sufficient for the temporal and spatial scales resolved at such low R eynolds numbers. As before, the
agreement with the experimental results was shown to be very good, however for the highest incidence, the loca­
tion o f breakdown was predicted to be slightly further downstream than in the experimental results. No reasons
for this discrepancy were given. The unsteady behaviour o f vortex breakdown was also considered and dominant
frequencies were found to occur from time histories o f pressure coefficient. Upstream o f breakdown these were
CHAPTER 1. INTRODUCTION 36

found to be St — 0.63 and 10.7, with further peaks at A = 2 and A = 8 found from the PSD analysis. The higher
frequency was attributed to the shear layer suuctures while the lower frequency was assigned to a fluctuation of the
vortex breakdown location. Downstream of breakdown, the frequency response was found to be more broadband
in nature, between A = 0 and St — 5. It was also noted that the spiral form of breakdown was not found for these
cases, however this is likely to be a feature of the sweep angle and not due to the computational solution.

From these studies, it is evident that the spatial and temporal resolution needed to fully resolve delta wing flows
is prohibitive to the solution of flow at full scale Reynolds numbers. Particularly, for industrial application. How­
ever, as computational power rapidly increases, it may be possible that techniques like these may be used more
extensively for vortical flows to validate and improve the accuracy of turbulence models and treatments in the
future.

1.5.5 Detached Eddy Simulation (DES)


In order to reduce the spatial and temporal requirements of LES, particularly in the boundary layer region, hybrid
URANS/LES methods have been proposed. These methods allow the advantages of each method to be utilised
by applying the UR ANS turbulence models to the boundary layer regions and LES to the remainder of the flow
domain. This has the advantage of considerably reducing the computational requirements of LES as the boundary
layer region is not required to be as well refined. Due to the application of LES, the solutions are still heavily
dependent on the spatial and temporal resolution and will therefore require substantially more computational re­
sources than RANS methods. However, by using LES in the majority of the flow domain, the resolution of the
overall flow behaviour will be much improved compared to standard RANS methods. Generally, these hybrid
methods are known as Detached Eddy Simulation (DES). Further discussion and details of the formulation of such
methods will be given in Chapter 2.

In the investigation by Morton et al. [24, 161], mentioned in Section 1.5.3, the RANS results for the flow over a
70* wing was compared with that of two DES formulations. The DES formulations used were based on the RANS
SA and SST models and as such were refened to as SADES and SSTDES. In order to run the DES calculations
and due to their inherent sensitivity to time step size and grid resolution, both a time accuracy study and a grid
dependence study were carried out. From a PSD analysis of the time history of Q for various time steps, it was
found that with decreasing time step size the dominant frequency of the signal tended toward an asymptote. Based
on the final value of this asymptote, the optimum time step for the calculation was chosen as T —0.005. Similarly,
a detailed grid resolution study was carried out, which is detailed in [163]. The results of this study showed that
both the medium, baseline and fine grids could capture the dominant frequencies. Thus, the baseline grid, with
2.45 X 10® cells was chosen to perform the comparison with the RANS models. As stated previously, it was found
from the comparison shown in Figure 1.29, that the DES methods were able to capture the full range of frequencies
which occur in the flows over delta wings. From analysis of tlie vortex breakdown behaviour, it was also found that
the breakdown was more clearly resolved in the DES solutions. Overall it was concluded that the DES methods
more accurately predicted the behaviour of the flow.

In a parallel study conducted by Mitchell et al. [29, 164], the presence of vortical substructures was investigated
using DES based on the Spalart-Allmaras turbulence model for the same test case. As before time step and grid
resolution studies were earned out. However, for this investigation a method known as “Adaptive Mesh Refine­
ment” (AMR) was used to refine the grid in the regions of interest, namely the vortex core region. Five grids were
compared, which were all unstructured and consisted of 1.2 x 10®,2.7 x 10®, 6.7 x 10®, 10.7 x 10®grid points and
an AMR grid of 3.2 x 10® grid points. The results agreed well with the experiments, with small spatially stationary
sub-vortices observed in the shear layer of the very fine grid and the adapted grid. However, it was found that
the structures observed in the adapted grid solution were closer to the experimental results than the very fine grid.
Based on this it was suggested that the occunence of tlie structures was extremely grid dependent. With the degra­
dation of the results on the very fine grid being due to an increased refinement of the trailing edge vortices which
seem to have an upstream effect on the shear layer. It was concluded from this study that further work was needed
on a time-accurate simulation of these structures to determine their cause. The use of DES was recommended due
to the accuracy of the solutions, however, it was noted that care must be taken over the creation of the grid as the
results are heavily dependent on the resolution of the grid in the vortex region.

DES investigations have also been earned out by Gortz [5, 32,165] using the same 70* wing geometry at ct = 27*
and with a Reynolds number of Re — 1.56 x 10®. This study differs from the previous investigations as it considers
a full span wing and uses a slightly higher Mach number of M = 0.2. A structured grid was also used, with a
H-C-H topology and approximately 4.23 x 10® cells. As before, a time step study was conducted. This study was
CHAPTER 1. INTRODUCTION 37

carried out on a semi-span wing in order to reduce length of the calculations. The time steps used and the resulting
dominant frequencies from the PSD analysis of the Q signals are summarised in Table 1.3.

Dominant St o f C^
Ar, s At Time History
0.0005 0.0446 1.95
0.00025 0.0223 2.59
0.000125 0.0112 2.73
0.0000625 0.0056 2.72

Table 1.3: Time steps used in DES time accuracy study [32]

It was evident that with decreasing time step, tliat the dominant frequency increased. However, as in the investi­
gation of Morton et al. [24], this value approached an asymptote. Therefore, a time step value of At = 0.000065j
was chosen as it appeared to provide reasonable temporal accuracy. Due to the expense of the calculation, the DES
simulation was stopped after 89 non-dimensional time units, which corresponds to a total time of 0.0056 seconds.
Due to this short time duration, the results gained may be susceptible to transients within the solution and as a
result, were treated with caution. From the PSD analysis of the unsteady Q signal two dominant frequencies were
found, at 182 and 121Hz, which correspond to Strouhal numbers of St ~ 2,5 and St — 10.2. The lower frequency
was attributed to the helical mode instability and the higher frequency was determined to be related to frequencies
found for the Kelvin-Helmholtz instability, and was assumed to be characteristic of vortical substructures in the
flow. It should be noted that a Strouhal number of approximately 10 was also found in the investigation by Mitchell
et al. [29, 164] detailed above for vortical substructures. However, from analysis of the flow solution, no vortical
substructures were observed. It was proposed that this was due to the relative coarseness of the grid in the vortical
region. The signal was not sufficiently long to detect any frequencies associated with vortex breakdown oscillation,
however this behaviour was witnessed from the flow visualisations and the location and amplitude of the oscilla­
tion was found to agree well with experimental observations. Further analyses of the flow were considered and
compared to experimental data and it was found, overall that the agreement was good. In conclusion, it was stated
that DES is capable of predicting the unsteady behaviour of the vortex breakdown location accurately, however,
that further work was needed to determine grid refinement issues.

The UR ANS investigation of this test case by Soemarwoto and Boelens [31] discussed in Section 1.5.3, was ex­
tended by de Cock et al. [166] using an alternative hybrid RANS/LES turbulence treatment called extra large
eddy simulation (X-LES). This model uses the k —o) turbulence model within the boundary layer and LES for the
remainder of the flow domain. The same grid and time step are used as in Ref. [31]. From consideration of the
PSD analysis of the normal force coefficient signal, in comparison to the UR ANS solution detailed previously, it
was found that the peak at A fa 9 was not as dominant in the X-LES solution. However, more power was found
in the higher frequencies, which indicates that this method is capturing more scales than the URANS calcula­
tion. Further examination of the flow structure showed that vortex breakdown was predicted further upstream at
x/cr = 0.71 which was in slightly better agreement with the experiment. It was concluded from this study that
in comparison with the URANS results, the X-LES solutions exhibited a clear qualitative improvement due to an
increase in resolution of the details of the flow.

From these investigations, it appears that DES can provide an increase in accuracy in comparison with URANS
methods. However, as stated, this accuracy is highly dependent on temporal and spatial resolution, which results
in large computational resources being required. However, these resources are not as considerable as those needed
for LES or DNS as stated previously.

1.6 Objectives
From the literature review, it is clear that the flow over slender delta wings is complex, with the presence of break­
down and many other instabilities existing in the flow and the possibility of interactions with shockwaves occurring
at transonic velocities. It is also clear that although much progress has been made in understanding this flow behav­
iour, there are still many aspects which are not well understood. One of these aspects is the nature of breakdown in
transonic flow and the possible interactions which occur between the vortices and shocks. The sudden appearance
of breakdown in transonic flow can have significant effects on the aerodynamic performance of an aircraft. The
ability of CFD to predict this type of behaviour has also been considered and it is clear that this tool could provide
more insight into the mechanisms which drive the abrupt nature of breakdown. Due to this, one of the aims of
this project was to consider the transonic flow over a slender delta wing, with a view to considering the ability
CHAPTER 1. INTRODUCTION 38

of RANS methods to predict the flow behaviour and to examine the causes and behaviour of vortex breakdown
within such a regime. This section of the project was carried out within the framework of a NATO AVT Task group.

The second aim of this work was to consider the unsteady behaviour of vortex flows at moderate angles of inci­
dence where breakdown occurs on tlie wing. It is clear that the unsteady nature of the flow can have a significant
effect on the overall flow behaviour and can interact with the surface of the wing or with other aircraft structures
as buffet. It is evident from the literature review that much work has been carried out to consider this type of flow,
particularly using CFD techniques and a number of high fidelity turbulence treatments have been proposed which
allow further accuracy in the numerical solutions of such flows. It is clear that the use of DES allows a greater
resolution of the unsteady flow behaviour, however this improvement in resolution come with a significant increase
in computational expense over statistical methods such as URANS. It is interesting to consider the ability of DES
and URANS methods to capture the main unsteady features. To assess the capabilities of tlie DES solution, com­
parison was made with experimental data and with existing DES calculations. The ability of URANS to predict
the flow behaviour is then compared with the validated DES results.

1,7 Thesis Outline


Chapter 2 considers the numerical methods, turbulence models and other computational techniques used in the in­
vestigations. Chapter 3 considers the behaviour of vortical flows and vortex breakdown under transonic conditions.
Chapter 4 considers the ability of the DES turbulence treatment to predict the unsteady behaviour of a subsonic
delta wing vortex, including breakdown. Chapter 5 then considers the same test case, using URANS to consider
whether this approach may be used to predict the main unsteady features of the flow compared to the DES solutions
of the previous chapter. Finally, overall conclusions are drawn and extensions to the work will be considered.
Chapter 2

Methodology of CFD Investigations

2.1 Introduction: The Navier-Stokes Equations


Computational Fluid Dynamics (CED) uses numerical methods to solve the conservation equations of fluid flow,
known as the Navier-Stokes equations. These are a set of Partial Dflferential Equations (PDEs) which describe the
conservation laws for mass, momentum and energy, given by.
Mass - the continuity equation.
+ ^ = 0 (2.1)
dt dxi
• Momentum
dpui , d{puiUj) dp , d' Ui j

where Xtj is the viscous stress tensor, which for a Newtonian fluid is proportional to the strain-rate tensor,
Xij —Ip Sij (2.3)
where the viscous strain-rate tensor is.
1 f dui duj
+ (2.4)

and the laminar viscosity is defined by Sutherland’s law,

Mo \T o J T + llO
where is the reference viseosity at tlie reference temperature, To, where po = 1.7894 x I0~^kg/tns and
ro = 288.16X.
• Energy
= + (2.6)
dt dxj dxi dxj
where E is the total energy of the fluid, defined as

E — p \ e + -UiUi j (2.7)

The heat flux vector, qt is given by Fourier’s Law and is defined as,

These equations, along with the equations of state for a perfect gas

H ^E + ^, p = {y ~ l)p e , ^ = (2.9)

provide a complete description of tlie flow and heat tiansfer of the three-dimensional, Newtonian fluid flows con­
sidered in this thesis.

39
CHAPTER 2. M ETH ODOLOGY OF CFD INVESTIGATIONS 40

2.2 Turbulence
At low Reynolds numbers, the Navier-Stokes equations predict behaviour known as Laminar flow. Viscous stresses
and the viscous diffusion of vorticity and momentum damp out any small scale instabilities. However, as the
Reynolds number is increased, inertial effects become increasingly important and the viscosity of the flow is no
longer able to maintain the smooth behaviour. With the growth of instabilities, the flow becomes highly inegular
and three-dimensional and the flow transitions from a laminar state to a turbulent one. Turbulenee can be defined
as an irregular flow where the physical properties fluctuate rapidly in time and space. These fluctuations occur over
a large range of scales and are associated with structures in the flow, known as turbulent eddies. The size of these
eddies can be described by their characteristic length i. Associated with this length are characteristic velocity and
time scales, u{É) and r{ê) respectively.

The behaviour of the turbulent eddies in the flow is highly non-linear with interactions occurring between fluctua­
tions with a wide range of wavelengths. The physical process, whieh allows these interactions, is known as vortex
stretching. Vortex stretching is an inherently three-dimensional property of the flow which means that turbulence,
by its nature is three-dimensional. The turbulence gains energy from the vortex elements in the flow, if they are
aligned in such a way that the velocity gradients can stretch them. This process is known as the production of
turbulence. As a result, the largest turbulent eddies within the flow carry most of the energy. However, the larger
eddies also act on the smaller eddies, in turn stretching theii- associated vortex elements and transferring energy to
them. This transfer of energy is then continued to even smaller eddies until the viscous forces become dominant
and the energy is dissipated. This is known as the energy cascade. As well as transferring energy to the smaller
eddies, the larger eddies also migrate across the flow carrying tire smaller eddies with them. This movement of the
turbulent structures results in an increase in the mixing and therefore, diffusion of mass, momentum and energy
contained in the fluid. This is known as turbulent mixing.

The energy which is associated with turbulence and the cascade process is known as the turbulent kinetic energy,
k and the rate at which this energy is transfered is termed the rate of dissipation, defined as,

e= (2.10)

The rate of dissipation is set by the largest eddies within the flow and the smallest eddies adjust to this value. The
range of the scales in a turbulent flow, from the smallest to the lar gest eddies, is dependent on and increases in
extent with Reynolds number as,
(2 . 11)

with Ï) and io being the characteristic lengths of the smallest and largest eddies respectively and Rep being the
Reynolds number based on the turbulent characteristics of the flow. In a similar manner, the ratio between the
largest and snrallest velocity and time scales can also be defined as,

^ ^r^R er^^^ (2.12)


Uo Xq
From these equations, it is clearthat the smallest scales of the flow can be many ordersof magnitude smaller than
the largest turbulent scales. It is also evident thatas the Reynolds number increases the range of scales increases.
As turbulence is a continuum phenomenon, the smallest scales are far larger than any molecular length scale. The
size of the smallest scales was determined by Kolmogorov based on dimensional analysis and are dependent on
the rate of dissipation and the kinematic viscosity, given by,

u ^ = z { e v f/^ , T,, = (v /e )^ /^ (2.13)

These are known as the Kolmogorov scales, where

R e~ i]u ^/\> = \ (2.14)

In comparison, the largest scales in the flow can have the same order of magnitude the width of the flow, such as
the boundary layer thickness, for example. During his study of turbulent scales, Kolmogorov made a number of
hypotheses based on his observations. These included that at sufficiently high Reynolds numbers, the small scale
turbulent motions were universal in that they were independent of the flow geometry and statistically isotropic.
Anisotropic turbulent scales exist only for the largest length scales of the flow. According to Pope [167] this corre­
sponds to the largest 16% of the eddies. As stated before, the largest eddies contain most of the energy of the flow
CHAPTER 2. M ETHODOLOGY OF CFD INVESTIGATIONS 41

and therefore have the largest influence on the mean flow. As a result the anisotropic turbulence is known as the
“energy containing range”. In the isotropic turbulence range, which extends to the smallest turbulent scales, the
energy cascade continues until at the smallest eddies, the molecular viscosity is sufficient to dissipate the energy to
heat. This range of turbulent eddies is known as the “dissipative range” or “viscous range”. The region in between
the energy containing eddies and the dissipative range is known as the “inertial sub-range”, where the behaviour
of the energy cascade is dominated by inertial effects. The behavioui* of this region can be uniquely described by
a relation based on the rate of dissipation, e.

The time and length scales of a flow are generally represented by frequencies and wavelengths (or wavenumbers,
K ) which are obtained from the use of Fourier analyses of the temporal and spatial histories of the flow, respectively.

Fourier analyses and statistical methods are discussed in more detail in Appendix B. It is more usual to consider
turbulence as a range of wavenumbers, which are associated with the length scales of the turbulent eddies. Using
this convention, the turbulent kinetic energy for a range of wavenumbers can be defined as

= ^ E{K)dK (2.15)

where E { k ) i s the energy spectral density related to the Fourier analysis of k . From this relation, it is evident that
the energy within a turbulent flow may be described from the energy spectrum, E ( k ) vs. k . Figure 2.1 shows a
representation of this spectrum on a log-log scale showing each of the ranges discussed above. This plot shows
that the inertial sub-range is described by a straight line, which has a gradient of -5/3 and is only dependent on the
rate of dissipation as described above. This slope was defined by Kolmogorov, as a result of his hypotheses and
dimensional analysis and is known as the Kolmogorov -5/3 Slope (or Spectrum), given by,

(2.16)

where Q is the Kolmogorov constant.

Energy Inertial
C ontaining'
Eddies I range Range

V
1
K

Figure 2.1: Energy Spectrum for a turbulent flow - log-log scales (From Ref. [26])

2.3 l\irbuleiice Modelling


As mentioned before, the Navier-Stokes equations completely describe the behaviour of all the fluid flows consid­
ered in this thesis. For turbulent flows there are a large number of temporal and spatial turbulent scales, which need
to be resolved. When the Navier-Stokes equations are solved to resolve all scales this is called Direct Numerical
Simulation (DNS). However, this requires hugely refined grids to capture all the length scales of the flow and
makes this method unrealistic for all but the most simple, low Reynolds number flows. Therefore, to reduce the
computational expense in resolving all the scales of turbulence, mathematical modelling is introduced to account
for a proportion of the small scale turbulent fluctuations. These mathematical techniques are known as Turbu­
lence modelling. In this investigation two approaches to the simplification of the Navier-Stokes equations and the
treatment of turbulence are used. These are the Reynolds averaging approach and Large Eddy Simulation.
CHAPTER 2. METHODOLOGY OF CFD INVESTIGATIONS 42

2.3.1 Reynolds Averaging Approach


One method used to simplify the Navier-Stokes equations and to reduce the range of turbulent scales is Reynolds
averaging. This involves decomposing the instantaneous flow into a mean flow and turbulent fluctuations, which
is known as the Reynolds decomposition. It shonld be clarified at this point that “mean flow” is taken to mean
the slowly varying flow behaviour, and is not necessarily constant with time. This decomposition is then substi­
tuted into the Navier-Stokes equations, before an averaging process is applied. There are three main methods for
averaging the flow: time averaging, spatial averaging and ensemble averaging. The most common method within
engineering flows, however, is time averaging, which will be detailed in this section. Details of other methods are
given in Wilcox [26].

The velocity component, Uj, will be used to explain Reynolds averaging. This is applied to other flow variables in
a similar way. Generally, the Reynold’s decomposition takes the form,

Ui = Ui + u'i (2.17)

where U{ is the mean flow velocity and is the fluctuating velocity due to turbulence.
For statistically stationary turbulence, that is a turbulent flow where the mean flow does not vary with time, the
time average is calculated from the instantaneous variable by using,
1 f ‘+T
Hi = lim — / Uidt (2.18)
T I Jt

In practice, the term, T —><», means that the integration time T should be of a sufficient length in comparison to
the maximum period of the fluctuations. The time average of the instantaneous velocity decomposes to the time
average of the mean flow, which due to its stationary nature is equivalent to the mean, Üi — ( Ui ), and the time
average of the turbulent fluctuations, which are zero, u\ = 0.

However, it has been discussed in Chapter 1 that delta wing vortical flows have an inherently unsteady behaviour,
where the mean flow will vary with time. Therefore, any turbulence which occurs in the flow will fluctuate about
an unsteady mean flow. This is known as non-stationary turbulence. For this case, the Reynolds decomposition
as given in Equation 2,17 is still applicable, however, the mean flow velocity may be further decomposed into a
stationary mean, {Ui), and unsteady component, u'/.

Ui = {Ui) + u'f (2.19)

Thus, the instantaneous velocity can be decomposed into the form,

U i^{U i)+ u 'f + u'i (2.20)

The application of the time average is also slightly different and takes the form,

(2 .21)

where the sample time, T, should be chosen to be much larger that the small scale fluctuations of the turbulence,
but also much smaller than the scales of the mean flow oscillations. This is due to the time averaging only being
appropriate if the period of the oscillations of the mean flow are large in comparison to the time scales of the
turbulent motion. This is an inherent complication of turbulence as it is not always possible to assume that the
turbulent fluctuations will be small. However, from consideration of the discussion given in Section 1.3.3 which
shows that the majority of characteristic flow features occur at very low non-dimensional frequencies of less than
St = 10, it may be assumed for vortical flows that this is the case. Flowever, care should be taken when analysing
the results. Therefore, applying Equation 2.21 yields, as before, the time average of the mean flow, Oi and the time
average of the turbulent fluctuations, which again will be zero, = 0. However, this time, the time average of the
mean flow does not equal the mean, {Uj)

To form the Reynolds Averaged Navier Stokes (RANS) and Unsteady RANS (URANS) equations, the Reynolds
decomposition, Equation 2.17, for each flow variable, is substituted into the Navier-Stokes equations. The time
averaging process described above is then peiformed. Many of the new terms created from the Reynolds decom­
position disappear due to the time averaging of the turbulent fluctuations, «• = 0 and the momentum equations
become,

+ + (2.22)
ot dx( ^
CHAPTER 2. METHODOLOGY OF CED INVESTIGATIONS 43

However, two new non-linear terms arise from tire convective terra as the time average of the product, Uiiij, yields,

UiUj — (Ui 4- mJ)(t/j 4- u'j) UiUj 4“ u'jU'j (2.23)


For stationaryturbulence, it was stated before that the time average of the mean flow isequal to the mean of
the flow,therefore theterm UiUj simplifies to {Ui){Uj). The second term, «-Wy, cannot beapproximated and is
consistent for both stationary and non-stationary turbulence. This creates a set of six new unknowns,—p
which are known as Reynolds stresses and denoted by tfj, the Reynolds stress tensor. This term is usually included
with the viscous stress tensor and the (U)RANS equations, for incompressible flow, become,

A similar process is carried out for the energy equation, which becomes,

A ( ( Î ,. + T,") Ui - (% + g f ) ) (2,25)

where q f is the turbnlent contribution of the heat flux vector as defined in Equation 2.8, using the turbulent eddy
viscosity and turbulent Prandtl number Pr-r- The presence of these unknowns creates a closure problem, which
requires a new set of equations to calculate the Reynolds stresses from the known mean quantities. One connnon
approach is based on Boussinesq’s approximation.

The Boussinesq approximation is based on an analogy which likens the influence of the Reynolds stresses to the
viscous stresses as defined in Equation 2.3. In this, way tlie anisotropic Reynolds stresses, (a,7 = ‘! 4 - l are
assumed proportional to tlie mean strain rate and can be expressed as,
Uij = —2firSij (2.26)
This inti'oduces a viscosity parameter, known as the turbulent eddy viscosity, fir- As the Reynolds stresses also
include an isotropic part, Boussinesq’s eddy viscosity hypothesis becomes.

ij ~ ................
—2jirSij — —2 pkdij (2.27)

where k is the specific turbnlent kinetic energy of the turbulent fluctuations, given by:
u'ju[
k = ^ (2.28)

which can also be thought of as half the trace of the Reynolds stress tensor. In reality, there are two assumptions
being made in this approximation, 1. that the anisotropic Reynolds stresses can be defined at each point in space
and time by the mean velocity gradients and 2. that the turbulent eddy viscosity is a scalar property of the flow
meaning that the relationship between the anisotropy and the velocity gradients is linear. However, this method
has the advantage that tlie Reynolds averaged equations have the same form as the Navier-Stokes equations as
shown above and that the number of unknowns in the system of equations is reduced to one, greatly simplifying
the closure problem. The turbulent eddy viscosity may be defined as the product of a velocity scale and a length
scale. It is in the dhect or indirect calculation of the eddy viscosity where the majority of (U)RANS turbulence
models are applied and it is in the specification of these scales where the majority of models differ.

2,3.2 Favre-Averaging for Compressible Flows


In addition to fluctuations of velocity and pressure, the density and temperature will also fluctuate for compress­
ible flows. This means that density is also now included in the partial differentials of the convection terms of the
Navier-Stokes equations. If the Reynolds averaging procedure outlined in the previous section is applied to the
resulting Navier-Stokes equations, witli the density also defined as the sum of its mean and fluctuating parts, the
complexity of the equations increases considerably. This is due to the introduction of correlations containing the
fluctuating density, p'. This will also increase the required complexity of turbulence closure models.

This problem can be overcome by using the density-weighted averaging procedure proposed by Favre [168] (this
procedure is also known as Favre-averaging). In this method, the mass average is introduced, in a similar way to
the time average given in Equation 2.18, again using the velocity components, n, as an example,
1 1 rf+î*
üi = — lim — / puidt (2.29)
P T — *0O T Jt
CHAPTER 2. M ETH ODOLOGY OF CFD INVESTIGATIONS 44

where p is the time averaged density. This method is similar to Reynolds averaging in that the instantaneous flow
variables can be decomposed into mass averaged, t7/ and fluctuating parts, u'/'.

Ui = üt + u'/' (2.30)

The velocity decomposition is then multiplied by the density and time averaged to form the Favre average decom­
position. ___
pTTi = püi + p u f (2.31)
However, the Favre average of a fluctuating variable is equal to zero, therefore, the time average of the density
correlation is equal to the time averaged density multiplied by the mass average of the variable,

pui ~ püi (2.32)

or alternatively the mass average of the variables may be defined as,

Ui - ^ (2.33)
P
It is important to note that Favre averaging is only a mathematical simplification which eliminates the density
fluctuations from the averaged equations. It does not, however, eliminate the effects of the density fluctuations on
the turbulence in the flow.
Using the Favre averaging technique, the Navier-Stokes equations take the form,

^ + ^ = - ^ + ^ (7 " + (2.35)
It is clear that these equations are analogous to the incompressible RANS equations given in Equations 2.24 and
2.25. However, in this case the Reynolds stress tensor is given by,

tg = -pu'l'u'j' (2.36)

As a result the Boussinesq approximation is slightly altered such that the strain-rate tensor used for the calculation
of the anisotropic part of the tensor becomes,

Due to the similar form of the Favre-averaged equations to the RANS equations, the details of the turbulent closure
and applicable turbulence model are the same. Therefore for ease of presentation, all turbulence models are written
in incompressible form.

2.3.3 Large Eddy Simulation


An alternative approach to simplifying the Navier-Stokes equations, is Large Eddy Simulation (LES). LES was
initially proposed as a means to reduce the required computational expense of DNS. It does this by essentially
applying DNS to much coarser grids and therefore only resolves the larger turbulent eddies in the flow. However,
due to interactions between all scales in a turbulent flow the smaller scales are important to determine a complete
and accurate turbulent solution. Thus, the influence of the scales smaller than the grid cell dimensions, known
as Snbgrid Scales (SGS), are modelled using what is known as “Subgrid Scale modelling”. In resolving only the
larger scales, much coarser grids and time steps can be used, compared to DNS, as the smallest resolvable scales of
the flow are now much larger. This also has the consequence that LES is feasible at significantly higher Reynolds
numbers.

To apply LES, the flow variables are again split into two parts: the resolvable (or filtered) component and the
subgrid (or residual) component. As before, a velocity component will be used to illustrate. The decomposition is
analogous to the Reynolds decomposition discussed above and takes the form

Ui = Ui + Ui^^^ (2.38)
CHAPTER 2. M ETH ODOLOGY OF CFD INVESTIGATIONS 45

However, in this case the resolved scales may exhibit unsteady behaviour and the filtered residual does not become
zero. These components of the instantaneous velocity can be further decomposed, to show what is being captured.
Using similar terminology as for the URANS decompositions above, the filtered variables may be considered as the
sum of the mean velocity, unsteady mean flow fluctuations and a proportion of the turbulent fluctuations resolved
on the grid, defined by 0 .
i?:=(Ui) + u''+((>u'i (2.39)
The subgrid component constitutes the remaining turbulent fluctuations which are too small to be resolved,

= (2.40)

In the LES method, the two components of the decomposition, the resolvable and subgrid scales, are separated by
applying filtering techniques. These techniques are applied in such a way that the maximum cell dimensions of the
grid, given by Equation 2.41, are used as the filter width,

A — w<ax(Aï, Ay, Az) (2.41)

The filter usually takes the form of a convolution integral,

iJi{x) — J G(x —x')u{V)dV (2.42)

where G is the filter function, which for a box filter, such as a computational grid, takes the form.

Using these techniques, the Navier-Stokes equations can be obtained for the filtered component of the flow. These
equations take the form,
dpui dpuiUj dp dtij

As with the RANS method, this introduces a non-linear convective term into the equations. In this case, in a
similar way to Reynolds averaging, the filtered product is not equal to the product of two filtered variables with the
difference being the Sub Grid Scale (SGS) or residual stress tensor,

upTj = UiUj F (2.45)

The sub grid scale stress tensor is the sum of a number of viscous stresses created from the filtering process
due to the inequality ui ^ tq. These stresses are known as the Leonard Stresses, cross-stresses and SGS stresses
which describe interactions between the resolved turbulence and the small scale turbulence. More detail of these
stresses and their significance can be found in Pope [167] and in the lecture notes of Ferziger [169]. It can also be
decomposed into anisotropic and isoti'opic pai ts.

= 4 T + ^krôi j (2.46)

where k,- is the residual kinetic energy given by.

k,- = (2.47)

The filtered Navier-Stokes equations, now take the form.

Comparing Equations 2.24 and 2.27 with 2.44 and 2.46 it is clear that the filtered equations and subsequent stresses
are analogous to the Reynolds averaged equations. Thus, the additional sfiess tensor can he treated in a similar way
to the Reynolds stress tensor in the (U)RANS formulations. This is the basis for the simplest and most widely used
approach for the subgrid scale modelling, proposed by Smagorinsky. In this model, the anisotropic stress tensor is
assumed to be proportional to the filtered strain rate tensor m a similar manner to the Boussinesq approximation.
As before, an eddy viscosity is associated with this relationship, known as the Smagorinsky eddy viscosity, psos-

= —2psGsSij (2.49)
CHAPTER 2.M ETH ODOLOGY OF CFD INVESTIGATIONS 46

The Smagorinsky eddy viscosity is taken to be a scalar quantity defined as,

PsGs = P (CsA)^ \JSijSij (2.50)

where Q is theSmagorinsky constant taken as 0.18. As for the Boussinesq approximation, this provides a linear
relation for the sub-grid scale stress tensor

r f f = 2psGs^- - ^pkrSij (2.51)

It is important to note at this stage that although the sub grid scale stress tensor is analogous to the RANS Reynolds
stress tensor and the role it plays is similar, the physics which is being modelled are quite different. Due to the
formulation of LES discussed, the energy contained in the sub grid scales is a much smaller proportion of the
total flow energy than the RANS turbulent energy. This suggests that only the smallest, Isotiopic scales need to
be modelled and therefore the model accuracy may be less crucial for sub grid scale modelling than for (U)RANS
computations, which need to consider the anisofiopic scales. For a more detailed explanation of LES please refer
to Pope [167].

2.4 Application of Itirbulence models to Delta Wing Vortical Flows


To understand how turbulence models predict the behaviour of delta wing flows it is important to understand the
behaviour of the velocity gradients and the production of turbulence in a turbulent flow and the mathematics which
is used to describe these phenomenon.

The velocity gradients of the flow are the components of a second-order tensor and as such can be decomposed
into isotropic, symmetric-deviatoric and anti-symmehic parts. A useful and detailed explanation of the properties
of second-order tensors can be found in Appendix B of Pope [167]. The decomposition is shown in Equation 2.52
where the symmetric-deviatoric pait corresponds to the strain rate tensor and the anti-symmetric-deviatoric part is
tlie rotation tensor,
dui _ I duk
■Sij + Sij + Qij (2.52)
dxj 3 dxfc
The strain-rate tensor was defined in Equation 2.4 and the rotation tensor is defined as,
1 / dui duj
Q,ij — —( -Tz------- ^— I (2.53)
2 \ dxj dxi

The rotation tensor is related to the vorticity of the flow,

( û i ~ —£ijkCljk (2.54)

where is the alternating symbol defined as,

( 1, if (i, j, k) are cyclic i.e 123, 231 or 312


£ijk = \ - 1 , if (i, j, k) aie anticyclic i.e 321, 132 or 213 (2.55)
[ 0 , otherwise

Generally, withinshear layers it is found that the velocity gradients aie dominated by thenormal gradients, there­
fore, the strain-rateand rotational tensors will be roughly equal. However, as the vortex core is approached the
flow tends to a purely rotational state and the rotational tensor will be laiger.

The production of turbulent kinetic energy is defined as the product of tlie Reynolds stress tensor and the velocity
gradient.

This equation mathematically defines tlie transfer of energy from the mean flow to the fluctuating velocity field.
This is caused by the mean velocity gradients interacting with the Reynolds stresses due to the process of vortex
stretching discussed previously. An important feature of this behaviour is that only the symmetric part of the
velocity gradient, the strain-rate tensor, and the anisotropic part of the Reynolds stress tensor contribute to the
production of the turbulent kinetic energy, such that Equation 2.56 can be written

Pk — (2.57)

1
CH APTER!. M ETH ODOLOGY OF CFD INVESTIGATIONS 47

From this relationship, it is clear that the production of the kinetic energy is proportional to the strain-rate tensor.
Considering the behaviour of delta wing vortical flows, Equation 2.57 implies that the production of turbulence
will mostly occur within the shear layer and surrounding flow and not within the vortex core where the flow is
highly rotational. Therefore, it would be expected that the turbulence within this region would be low and the core
may be thought of as approaching a laminai' state. A successful turbulence model for this type of flow should be
able to predict this behaviour.

2.4.1 Wilcox k —(0 Model


The Æ—ca model is a two equation Boussinesq based turbulence model proposed by Wilcox [34]. This model
uses the flow parameters, k, specific turbulent kinetic energy and, co, the specific dissipation rate per unit turbulent
kinetic energy to calculate the eddy viscosity and to close the Reynolds averaged Navier-Stokes equations. The
kinematic eddy viscosity for this model is given by,
fc
Pt ^ p ~ (2.58)

To calculate the turbulent kinetic energy and dissipation rate, two transport equations are added to the Navier-
Stokes equations in the solution of the flow. The transport equation for the turbulent kinetic energy is given as.

dk dkUi
PU7 4- ^ - P*pkw (2.59)
dt + P dxj dx
Production D issipation
Convection D iffu sio n

This equation is similar in form to the momentum equations given in by Equation 2.2 and includes convection,
diffusion, production and destruction terms as indicated. The transport equation for the dissipation, co, takes a
similar form and is given, along with all the definitions of the coefficients used in this model in Appendix A. To
understand how this model applies to delta wing vortical flows, it is necessary to consider the production terms.
The production of the turbulent kinetic energy was defined in Equation 2.56 and the corresponding term for the
dissipation rate is given as,
Pco = ajP k (2.60)

As mentioned, this model uses the Boussinesq approximation to calculate the Reynolds stresses and, thus, the
production term is expanded to become
Pk — I p f S i j S i j (2.61)
It is clear from these definitions that the production of k and m within this model, aie only dependent on the mean
strain-rate of the flow and does not take the rotation rate into account. This is an over-simplification and results
in a large over-production of turbulence within the vortex core as the regions of high vorticity are not accounted
for by the model. This over-production of turbulence causes the model to predict exaggerated levels of vorticity
diffusion and, thus, results in the prediction of a weak vortex which cannot be sustained and quickly diffuses.

2.4.2 k —(T> with Enhancer


To rectify the inability of the “standard” Wilcox k — O) io accurately predict the turbulent structure of the vortex
core, two different methods of rotation coiTection were proposed by Brandsma et al. [158]. These models were
suggested to control the production of turbulent kinetic energy and hence the levels of turbulent eddy viscosity
in the core region. The first method directly limits the production of k by using the dissipation terra as a limiter.
Whereas the second method, and the one used in this investigation, increases the production of the dissipation rate
(fu) within regions of highly rotational flow. In order to apply this rotation correction to the appropriate regions
within the flow, a suitable sensor was defined to distinguish between shear layers and vortex cores. This sensor
considers the ratio of the magnitude of the zero-trace mean strain-rate tensor to the magnitude of the mean rotation
tensor defined below as,
_ 5 _
1 /2
(2.62)
^ (2QyQy)
As mentioned before, in shear layers, the strain-rate and rotational tensors are roughly equal, such that r 1,
whereas in the core of the vortex the flow approaches a rotational state, which implies r C 1. The correction for
the dissipation production term is defined as,

Pa}„e,, = cCYmax{Q?',S'^} (2.63)


CHAPTER 2. M ETH ODOLOGY OF CFD INVESTIGATIONS 48

which is equivalent to dividing the production term of co by min (r^, l) , thus,

Using this con ection, the model now enhances the production of the dissipation and, thus, increases the dissipation
of k, which, in turn, decreases the turbulent eddy viscosity and turbulent production within the core regions.

2.4.3 Non-Linear Eddy Viscosity Model


The non-linear eddy viscosity model (NLEVM) is based on the Wilcox k —co model and uses the formulation of an
explicit algebraic Reynolds stress model proposed by Wallin and Johansson [170] to model the Reynolds stresses.
Essentially, this model introduces an extra term to the calculation of the anisotropic Reynolds stresses as defined
by the Boussinesq approximation, such that,

a ij^-2 p rS ij + a \ f (2.65)

The addition of this extra term, creates a non-linear relationship for the Reynolds stresses due to its depen­
dence on both the mean strain-rate and rotational tensors. The equation for the Reynolds stresses then becomes,

u[u'j = k % —2pfSij F afj (2.66)

In this model, the mean strain-rate and rotation tensors are normalised by the turbulent time scale, t, i.e. S =
and £2 = t£2,7 , where

The extra anisotropy term is a reduction of the general form of a,y used in explicit Reynolds stress models, which
contains ten tensorially independent terms and is defined in Equation A.6. The reduced form, with tensor subscripts
omitted, is.

= A - \u ç, +136 - //aS - + j3p (S2SS2^ - (2, 68)

where I is the identity matrix, equivalent to 5/y and Uçi and IV are two of the independent invariants of S and S2.
The j3„ coefficients are defined by these invariants and their definitions and other model constants are detailed in
Appendix A.

In addition to introducing this new anisotropic term, the calculation of the turbulent eddy viscosity is also modified
from the k — ca model and takes the form,
Pr = C f p k T (2.69)
where
~2 (2.70)
In this definition of the turbulent eddy viscosity it is clear that the behaviour of the rotation tensor is also taken into
account.

To consider the behaviour of this model in the prediction of vortical flows, again the production of turbulence
should be considered. This will also now have an additional term and takes the form,

Pk=(2prSij~a^^'^"jSij (2.71)

From this relationship, it is clear that the extra term acts to reduce the turbulent production from the original
model. The relationship between the strain-rate and rotation rate tensors and the extra anisotropy are difficult to
quantify due to the complexity of the model. However, it is evident from the formulation of the model that the
rotational tensor and its invariants appear frequently. Therefore, it may be supposed that as the flow tends to a
purely rotational state, the value of the extra anisotropy will increase and therefore reduce the turbulence within
the vortex core region. The levels of turbulent eddy viscosity will also reduce in this region, further reducing the
levels of turbulence in the flow.
CHAPTER 2. METHODOLOGY OF CFD INVESTIGATIONS 49

2.4.4 Spalart Allmaras Model


The one-equation Spalart-Allmaras model [35] Is another Boussinesq based model, which solves a single equation
for a working variable v, which is related to the turbulent eddy viscosity of the flow such that,

P T = p vfu i (2.72)
The single differential equation which defines this model was proposed for application in aerodynamic flows,
snch as transonic flow over airfoils, and was derived empirically using arguments based on dimensional analysis,
invariance and molecular viscosity. The origin of each term is described in detailin Ref. [35]. The transport
equation for the undamped eddy viscosity, v is given as,

d v , d{vuj) _ . f v V , 1 d r,.. , d v ] , Ck 2 d v d v
+ cr 3- 3-
d x k o x /c
(2.73)
Production
Convection D estruction D iffu sio n

As before, the general form of the equation is similai' to the momentum equations given in by Equation 2.2and
includes convection, diffusion, production and destruction terms as indicated. The wall destruction term is derived
to reduce the turbulent viscosity in the laminar sublayer. All model coefficients and definitions are detailed in
Appendix A. In the production term, it is important to note that S denotes the modified magnitude of vorticity,
defined in Equation A.22 and is not related to the strain-rate tensor.

After calculating the tnrbulent eddy viscosity using the transport equation, the Boussinesq approximation is used to
determine the Reynolds stresses and close the Navier-Stokes equations. As a Boussinesq based model, the Spalart
Allmaras model suffers from the same problems as tlie Wilcox k —a model discussed previously. Due to the use
of the sti ain-rate tensor in the calculation of the Reynolds shesses, the model may predict unrealistic contributions
of the Reynolds stresses in regions of high rotational flow, such as the vortex core.

2.4.5 Detached Eddy Simulation (DES)


As detailed in the previous section, EES was proposed as a means to reduce the computational costs of DNS. How­
ever, limitations still exist on its use and in practice the method only increases the applicable Reynolds numbers
by about a factor of 10 compared with DNS. These limitations are due to the application of LES on grids within
the wall region of a domain. As the wall is approached, the turbulent length scales decrease in size and become of
the order of the boundary layer thickness. Therefore, to accurately simulate the flow behaviour close to the wall,
the grid refinement needed approaches DNS levels. This is a significant problem for the applicationof LES to
practical engineering flows, such as full aircraft, as the problem of computational expense returns.

To overcome these issues, without compromising the solution accuracy anymore than is necessary, a new approach
was proposed by Spalart et al. [171]. This approach utilises both LES and RANS within the solution domain, to
take advantage of the benefits of each method and to gain an accurate solution, at moderate computational expense.
This hybrid method is known as Detached Eddy Simulation (DES) and works by applying a RANS model within
the boundary layer region to utilise its near-wall modelling approach to avoid computational cost and applying LES
to the remainder of the flow domain, where larger turbulent eddies dominate. The model used in this investigation
uses the Spalart-Allmaras model as the URANS model in the implementation as initially proposed by Spalart et
al. [171].

The use of the Spalart-Allmaras model within the DES formulation is based on the connection between the pro­
duction and destruction terms of the partial differential equation defined in Equation 2.73. Balancing these two
terms gives the relationship,
V - Sd^ (2.74)
From consideration of theSmagorinsky SGS model in Section 2.3.3 and in particular the expression for the
Smagorinsky eddy viscosity given byEquation 2.50, it is clear that a similarproportionalityexists, with the term
A, based on the grid spacing, (see Equation 2.41) replacing the distance d, such tliat

VsGS ^ 5A^ (2.75)


Based on this similarity, it was suggested that if d is replaced with A in the wall destruction term, the Spalart-
Allmaras model will act as a Smagorinsky LES model. Therefore, in order for the Spalart-Allmaras model to
exhibit both URANS and LES behaviour, d in the Spalart-Allmaras model is replaced by,
d = min {<7, Cdes^ } (2.76)
CHAPTER 2. METHODOLOGY OF CFD INVESTIGATIONS 50

where Cqes is a constant, which has been set to 0.65 as suggested by Shur et al. [172] for homogeneous turbulence.
From this formulation, the model acts as a RANS model for cl < A and as a subgrid scale model for af » A. Thus,
this method restricts the use of the URANS model to near wall regions and allows LES to be used elsewhere based
on these parameters. This is shown schematically by Figure 2.2.

LE S E êm »n_
,4

Ï-T R eg
RA

/S u r f il

^ m a x {A ^, A y, A J

Figure 2.2: Schematic of DES formulation on a structured grid

Using DES, allows the application of LES to the vortical region above the wing. In using LES, the Boussinesq
approximation is not applied and thus all large scale rotations, stresses and strains are resolved fully on the grid.
This will result in a more accurate prediction of the production of turbulence and the overall behaviour of the
leading edge vortex system. Generally, the subgrid contribution to the turbulence will be small, therefore keeping
the turbulence levels within the vortex core region low and more realistic than for the Boussinesq based models.
However, to keep this contribution small, the grid should be adequately refined such that the majority of the flow
scales are resolved on the grid. Thus, the computational expense of such calculations is much larger than for more
traditional turbulence models as described previously.

2.4.6 Computational Cost of Hirbuient Calculations


In a review by Spalart [33], the relative computational costs of DNS, LES, DES and URANS were compared in a
similar manner for a target flow of a full aircraft or a car. The numerical background of each method was considered
and the suggested grid size and number of time steps needed to advance a sample calculation by roughly six “spans
of travel”, meaning the time taken for an air particle to travel six times the length of the model. This may only
result in a calculation total time of a fraction of a second in real time. The results of this analysis of each method
is shown in Table 2.1.

Type E m piricism G rid Size Time Steps R eady?


3D URANS Strong W jq S .5
1995
DES Strong 10* 10*^ 2000
LES Weak I Q ] 1.5 1 q 6 .7
2045
DNS None IQi^ 107-7 2080

Table 2.1: Summary of computational costs of various turbulence methods according to Spalart (adapted from Ref.
[33])

Using the proposed computational cost of the methods, Spalart also attempted to define a readiness date for each
method, based on that method becoming a “Grand Challenge” to general CFD practitioners as opposed to everyday
industrial use. These dates are also included in Table 2.1. This data provides a measure of the computational costs
described above and the prospect of widespread use in the future. However, it remains to be seen whether these
predicted values are accurate or if complex geometries and modelling strategies will increase these requirements.

2.5 PMB solver


All computations were performed using the Parallel Multi-Block (PMB) flow solver developed at the University of
Glasgow, which has been continually revised and updated over a number of years. The solver has been successfully
CHAPTER 2. M ETH ODOLOGY OF CFD INVESTIGATIONS 51

applied to a variety of problems including cavity flows, hypersonic film cooling, spiked bodies, flutter and delta
wing flows amongst others. The PMB code is a multi-block structured solver which solvesthe unsteady RANS
equations in a global 3DCartesian frame of reference [173]. The conservationlaws described in Section 2.1 can
be converted into vector form to simplify their use in the computational metliod, the Navier-Stokes equations now
take the form,
aw a ( F '+ F ') S l G '+ G ”) ^ ( H ' + H ' )
+ — â ï— + — ai— °
where W is the vector of conserved variables, defined by
W = (p,p n ,p v ,p w ,E )^. (2.78)
The superscripts i and v in Equation 2.77 denote the inviscid and viscous components of the flux vectors, F, G and
H, respectively. The inviscid flux components are given by

F' = { p tfp u ^ F p ,p u v ,p u w ,u {p E + p))^ ,


G '= {pv,puv,pv^ + p,pvw ,v{pE + p ) ) ^ , (2.79)
H' = [pw,puw,pvw,pw^ + ppv{pE + p ) ) ^ .
The viscous flux vectors contain terms for the heat flux and viscous forces exerted on the body are

F — (0, Txv, Ttxy)T\z > T F WTvz 3“ Tv) >

G — (0, Txy,Tyy, Ty^, liTjy F VTyy F F Ty) , (2.80)


1 7-
H — (0, Uz t tyz, Tzz ) UTvz T F F qz) •

The terms % in Equation 2.80 represent the viscous stress tensor components while qt denotes the heat flux vector.
These equations can be transformed into the averaged equations simply by substituting -I- for T,y and
for qi and taking the flow variables as averaged quantities as defined in Section 2.3. All quantities are
non-dimensionalised using the relations.
X* t*

y 2,*' ^ L *’ ^ L */U *’
p* iT V* w*
(2 .81)
"" U*'

-1
p‘ r c*
:’ 7^’
where the superscript * denotes the dimensional variables. For this investigation, the non-dimensional characteristic
length, L*, is taken to be the root chord length of the delta wing, c, .

2.5.1 Steady State Solver


The PMB flow solver uses a cell-centred finite volume approach to discretise the governing equations described
above. According to this method, the spatial discretisation of the RANS equations for each cell results in the
equation,
^ F R'ljf = 0. (2,82)
where denotes the cell volume, which due to using a fixed grid is constant for the current work, and where
represents the flux residual. The convective fluxes are discretised using Osher’s upwind scheme [174], how­
ever Roe’s flux-splitting scheme [175] is also available. The MUSCL variable extrapolation method is employed
to provide second-order accuracy with the van Albada limiter being applied to remove any spurious oscillations
across shock waves. The central differencing spatial discretisation method is approximate to solve the viscous
terms.

The system of equations defined in Equation 2.82 are integrated in time to reach a steady state solution using an
implicit time-marching scheme, defined by
vyn+ l _ YUH
"" = y y R w (2.83)
CHAPTER 2. M ETH ODOLOGY OF CFD INVESTIGATIONS 52

where the flux residual is linearised in time as,

- K m + ( w ^ - F ja ) (2 M)

By substituting Equation 2.84 into Equation 2.83, the non-linear system of equations can be approximated as

where the subscripts i j , k haye been dropped for clarity and AW = - yV'T .

To solve this linear system of equations using a dhect method is prohibitive as the number of equations becomes
large. Therefore, an iterative Generalised Conjugate Gradient method is used as it is capable of solving sparse
equations efficiently in terms of time and memory requirements. This is used in conjunction with a Block Incom­
plete Lower-Upper (BILU) factorisation method used as a pre-conditioner to solve the system of equations.

Implicit schemes require particulai- treatment during the early stages of the iterations. Generally, the traditional ap­
proach is to initially use a low CEL number and increase this as the solution progresses. However, it has been found
that smoothing out the initial flow by using an explicit method for a number of initial iterations before switching to
the implicit method was just as efficient. Therefore, in all calculations performed, a number of explicit iterations
are specified before the implicit scheme is then used.

Due to the fact that the formulation of most turbulence models can also be represented in vector form, similar to
Equation 2.77, the steady state solver for the turbulence model equations are formulated and solved in a similar
manner to the mean flow as described, with the vector W replaced by the equivalent turbulent vector Q and an
equivalent substitution for the flux residual. For the turbulence model equations the flux residual also contains the
dissipation sonrce term, however the production term is solved explicitly. The eddy viscosity is calculated from
the turbulent quantities as specified by the model and is used to advance the mean flow solution. This new mean
flow solution is then used to update the turbulence solution, freezing the mean flow values.

2.5.2 Unsteady Solver


The implicit dual-time method proposed by Jameson [176] is used for time-accurate calculations. The residual is
redefined to obtain a steady-state equation which can be solved using acceleration techniques. Using a three-level
discretisation of the time derivative, the updated flow solution is calculated by solving

W , ^ .0 (2.86)

where R i j f is the spatial discretisation as described above, with wij^k and being the vector form
of the values of W and Q, the turbulent residual, in the surrounding cells. Similarly for tlie turbulence model.

+ = 0 (2.87)

These equations represent a coupled non-linear system of equations. The superscripts, k,n, k , l,n and I; determine
the time levels of the variables used in the spatial discretisation and determine the behaviour of the coupling be­
tween the systems of equations. If km ~ k( ~ l„t ^ It ~ n + I then the mean and turbulent quantities are advanced
in real time in a fully coupled and implicit manner. However, if km = l,n = lt = n + I and kt = ii then the equations
are advanced in sequence in real time i.e. the mean flow is updated using frozen turbulence values and then the
turbulent values are updated using a frozen mean flow solution. This has the advantage that the only modification,
when compared to the laminar case, to the discretisation of the mean flow equations is the addition of the eddy
viscosity from the previous time step. The turbulence model only influences the mean flow solution through the
eddy viscosity therefore any two eqnation model can be used without modifying the mean flow solver. Hence the
implementation is simplified by nshig a sequenced solution in real time. However, the uncoupling could adversely
affect the stability and accuracy of the real time stepping, with the likely consequence of limiting the size of the
time step that can be used.
CHAPTER 2. METHODOLOGY OF CED INVESTIGATIONS 53

This non-linear system of equations can be solved by introducing an iteration through pseudo time, t, to the steady
state. This is given by,
■yyn+ 1 ,A:+ 1 _ W '* + ^ 1

^ (2 .88)

with an equivalent form for the turbulent system of equations. It is clear that this takes the form of the steady state
solver formulation as given in Equation 2.83 such that if R y ^ is replaced with Ri,j,/t a non-time varying flow will
be solved. Using this formulation the system of equations can again be linearised and iterated to a steady state
solution in pseudo time before being advanced in real time.

The flow solver can be used in serial or in parallel mode. In order to obtain an efficient parallel method based on
domain decomposition, different methods are applied to fire flow solver. An approximate form of the flux Jaco-
bians resulting from the linearisation in pseudo-time is used which reduces the overall size of the linear system
by reducing the number of non-zero entries. Between the domains of the computational grid, the BILU factorisa­
tion is also decoupled thereby reducing the communication between processes. Each processor is also allocated a
vector that contains all the halo cells for all the blocks in the grid. Message Parallel Inteiface (MPI) is used for
the communication between the processors in parallel. All computations undertaken have been performed on the
Beowulf Pentium 4 120-processor workstation cluster of the CED Laboratory at the University of Glasgow,

2.6 Unsteady Time Step Calculation


One of the most important factors in the execution of an unsteady calculation is the choice of time step. If a time
step is too lai'ge then the high frequency detail of the flow can be missed, however with a very small time step
the computational resources and time taken for the calculation increases. Therefore, care must be taken to select a
time step which is small enough to adequately resolve the unsteady fluctuations of the flow, but lai'ge enough not to
make the required computational resources too great. This generally requires a prior knowledge of the approximate
scale of the important frequencies in the flow. For numerical calculations, the non-dimensional time step is usually
used and so the the non-dimensional frequency (or Strouhal number) should be considered. For delta wing flows,
the non-dimensional time and Strouhal number are related to the dimensional time and frequency using,

and — . (2.89)
Uao Cj‘
The unsteady behaviour of delta wing flows was considered and discussed in the previous chapter and the major
frequencies of the flow were highlighted for various investigations in Table 1.2. From this discussion, it is evident
that the majority of the frequencies associated with the dominant flow features are less tlian approximately & = 10.
A time step of A t = 0.01 is the lowest time step which can be used to capture this frequency.

To show how this value for the time step was reached it is clearer to start with the sample rate at which the unsteady
behaviour is to be sampled. To adequately capture an unsteady oscillation it may be assumed that a minimum of
flve time steps are needed per cycle. Therefore, the period of the maximum frequency captured will be 5 A T —0.05.
A frequency can then be obtained from the inverse of this value, which gives St = 20. In signal processing and
data sampling theory, it is important to avoid aliasing, where higher frequencies are superimposed onto lower
frequencies, which can distort the resulting sampled signal. In order to do this the Nyquist criterion is used which
determines the maximum frequency which can be detected for a given sample rate A T ,

S („ = jip (2.90)

This essentially reduces the maximum captured frequency for a given time step by a factor of two, therefore the
maximum frequency which can be obtained from a sample rate of A t = 0.01 is St = 10. From this analysis, it is
clear that in halving the time step, the maximum frequency is doubled.

This is generally adequate for URANS calculations as the choice of time step is independent of other calculation
parameters. However, for DES calculations the size of the time step is directly related to the size of the cells within
the computational grid and there is an optimum time step for a given grid size. Therefore, as the grid is refined
the time step is also refined. This was briefly discussed in the previous section dealing with the computational
cost of DES calculations. In a guide to creating DES grids and running calculations, Spalart [177] recommended
CHAPTER 2. M ETH ODOLOGY OF CED INVESTIGATIONS 54

calculating the required time step necessary for a given grid, based on the the minimum target grid spacing within
the region of interest and the maximum velocity in the flow (as a multiple of the freestream velocity), such that,

= ^ (2.91)

However, this relationship is only a guide and as such the effect of temporal refinement should be considered when
using the DES formulation.

2.7 Post-processing Techniques


Before presenting the results of the investigations, it is important to explain some of the techniques used to process
the solutions obtained from the CED calculations. As botir steady and unsteady calculations are performed within
this investigation it is important to consider the relevant output files and the way in which they are processed. For
an unsteady calculation, there are three main types of solution file. These are:

• Domain solution data

• Integrated loads
• Point probe data

The domain solution data file is created at the end of every specified time step calculation and provides data on
instantaneous flow variables for every point within the grid file used for the calculation. The integrated loads file
is updated for each time step and, therefore, provides the time histories of the loadings and similarly, the probe
data files provide time histories of flow variables at points witliin the flow domain specified at the start of the
calculation. For a steady state calculation, only two files are created at the end of the calculation once the solution
has converged, the steady state domain solution and tire integrated loads. Each of the files described are processed
in a different way and some details of these processes are given in this section.

2.7.1 Domain Solution data


As stated, the three-dimensional domain files contain flow variables at each grid point within the flow domain.
These variables are p, u, v, w, p and the turbulent quantities k, (o and R ct for the k - o) based models and pr
and Rer for the Spalart-Allmaras based models. The flow variables in the domain files are non-dimensionalised
by the freestream properties of the flow as described in Equation 2.81. The turbulent quantities are, therefore, also
non-dimensional. In this work, the all flow properties are non-dimensional, unless otherwise stated.

To analyse the domain files, the visualisation package Tecplot is used, which allows both single and multiple files
to be viewed and manipulated. Due to the large grid sizes used in these investigations, using the complete flow
domain for analysis was restrictive due to memory requirements and so a number of macros were written to extract
the relevant flow details for analysis. These exhacted data files allow both two and three-dimensional visualisation
techniques to be employed for either single or multiple files. The extracted details include, 1-D and 2-D slices of
the solution and 3-D isosurfaces of variables, such as velocity.

The use of macros in Tecplot, also allows the same views of each time step to be created and captured for compar­
ison and provides the means to create short movies of the unsteady behaviour. From these movies, it is possible
to pick out and track some of the unsteady features of the flow. With the knowledge of the time step size between
each frame it may be possible to resolve the frequencies of particular features and relate these back to the data
obtained from the unsteady probe and integrated loads files.

The variables provided by PMB in the output file are not always sufficient to capture specific flow features ade­
quately and other flow parameters are required. A number of flow parameters were calculated within this investi­
gation using CFD Analyzer, which is an add-on package for Tecplot, these include the components of vorticity, û),-,
the Mach number, M, entropy, s, and pressure coefficient, Cp. The relationships used to derived these variables are
given in Equations 2.92 to 2.95, respectively.

= (2,92)
oXj oXi
CHAPTER 2. M ETH ODOLOGY OF CFD INVESTIGATIONS 55

\u\ yP
where (2.93)
a Y P

s = c jn ( — I+ I (2.94)

2.7.1.1 Shock Detection and Analysis


When considering transonic flows it was also necessary to consider means to identify the locations and strengths
of Shockwaves which occur in the flow. In order to facilitate the identification of the shocks, a shock detection
algorithm was used, which was provided in the CFD Analyzer add-on. The algorithm is based on the work of
Lovely and Haimes [27] and calculates the locations of shocks by using the pressure gradient to calculate the
Mach number normal to a shock surface. Where the normal Mach number is greater than or equal to one, a shock
is identified. The pressure gradient of the flow is always normal to a shock surface and so the algorithm calculates
the pressure gradients in the flow in order to determine the orientation of the shock. The local Mach vector at each
point normal to this surface is then calculated. This is shown schematically in Figure 2.3.

^ M

TTTTTTTTVTTTTTTTTTTTTT^

Figure 2.3: Shock detection test quantity (Adapted from Ref. [27])

The normal Mach number or shock test value at each point in the flow is then created by using the dot product of
these pressure gradients and local Mach vectors,

= VP (2.96)

In the CFD analyzer version of this algorithm, the pressure gradient vector is normalised by its magnitude. Due to
the negligible thickness of shockwaves, the algorithm calculates shock surfaces which surround the region where
a shock is calculated to form and this creates a new flow variable within the domain solution. When this variable is
greater than one it is proposed that a shock may occur. To initially visualise the suggested locations of the shocks
within the flow isosurfaces of this shock variable were plotted. However, it became clear that there were regions
of spurious shock surfaces, where it was not expected that shocks would occur. Therefore, to allow validation of
this algorithm and confirm the locations of the shocks in the flow, the solutions were also analysed manually using
the variables: Mach number, entropy and pressure gradients (both magnitude and in all directions). Contours of
these variables were compared to the shock feature contours produced by the algorithm described above. Thus,
considering the distributions of all of these variables and reasoning based on previous investigations, the locations
of shocks in the flow were established.

2,7.2 Integrated Loads and Probe Analysis


The analysis of the time series of flow properties provided by the integrated loads and point probe files are carried
out using Probe Analyser. This is a custom-made program created in Matlab, which allows the manipulation and
plotting of the data, using statistical analysis and signal processing techniques. Probe Analyser is based on the
initial program by Lawrie [178], and further developed by Nayyai' [179], for cavity flows and has been further
extended in the course of this investigation, specifically for unsteady delta wing flows. Details of the program, its
current capabilities and an explanation of the techniques used in this investigation are given in Appendix B.

Before either file is analysed, a number of pre-processing techniques are necessary. Generally for the integrated
loads file this only involves deleting the initial transients of the signals created from the start of the unsteady calcu­
lation. However, for the probe files this is slightly more involved. The data written to the probe files from PMB, is
not directly usable and a number of different manipulations are needed before the results can be viewed. For this
CHAPTER 2. M ETH ODOLOGY OF CED INVESTIGATIONS 56

purpose, a number of utility programs have been written to concatenate and convert the probe data from the block
probe files into the format used by Probe Analyser. The initial transients are also removed from these files.

As stated before, the unsteady integrated loads files contain the time histories of the loadings on the wing surface,
these include the normal force coefficient, Qy, lift coefficient, Q,, drag coefficient, Co, and moment coefficient. Cm-
The analysis of the time histories of these variables can tell much about the overall unsteady nature of the flow and
through calculation of such quantities as the power spectral density, an overview of the dominant unsteady forces
on the wing and their frequencies can be obtained. Generally, the quantity which is of particular interest is the
normal force coefficient, however similar analysis can be carried out on the lift, drag or moment coefficients also.
Useful quantities which are calculated for these variables include the mean and RMS values along with the PSD
as mentioned above.

The analysis of the probe data is a little different to the integrated loads file purely due to the volume of information
which can be contained in the files. The probe files contain all the unsteady flow variable data for each specified
point in the flow. Therefore, the time histories of p, u, v, w and p are available for multiple points through the flow
domain. Probe Analyser allows for each probe to be considered separately or for a range or selection of probes to
be considered and cross-plotted together for comparison, it also allows multiple probe files from different calcu­
lations to be analysed and compared at the same time. The same analysis techniques can be applied to each flow
variable as described for the integrated loads, but there are also many more functions that can be performed on
the probe file data. These include, time averaging of a signal and the calculation of turbulent quantities and corre­
lations such as the Reynolds stresses or turbulent kinetic energy based on either a stationary or non-stationary mean.
1

Chapter 3

Transonic Vortical Flow on a Slender


Delta Wing

3.1 Introduction
As detailed in the literature review in Chapter 1, much is known about vortical flow over slender, sharp edged delta
wings and there are many reviews which detail the volume o f data available on the subject, both experimental and
computational [49, 58, 102, 126, ISO]. For the most part, this data concerns subsonic freestream flow and vortex
breakdown. However, an area o f delta wing vortical flow which is not so w ell understood is the behaviour o f the
flow under transonic conditions.

From the literature, it is evident that the behaviour o f the flow is somewhat different to vortical flow in the subsonic
regime. With an increase in Mach number, the size and shape o f the vortex system changes [118] and the primary
vortex is found to sit progressively closer to the wing surface. Despite this increased proximity to the wing, the
vortex system creates a much reduced suction peak on the wing compared to subsonic flow. The shock waves
which appear are caused by localised supersonic flow regions. A number o f investigations, both experimental and
numerical have been carried out, which have looked at the occurrence and behaviour o f shockwaves in vortical
flows for varying transonic conditions [17, 18, 117, 121, 122, 123, 124, 125]. From these investigations, a num­
ber o f shockwave systems have been observed and described in the literature. From the work o f Elsenaar and
Hoeijmakers [18], a plot was created, which detailed the onset o f various flow behaviours with Mach number and
incidence, which is shown in Figure 3.1. From this diagram, it is clear, that for transonic flow both rear/terminating
and cross-flow shocks appear for increasingly lower angles o f incidence. The critical incidence for breakdown is
also shown and indicates that the incidence at which vortex breakdown occurs decreases with increasing Mach
number. Further detail on the nature and behaviour o f the shockwaves was given in Section 1.4.

a (dag)

20

O N SURFACE

POS
SHOCK

PART SPAN SEPARATION


0

P O S. PLANE O F SYMM ETRY

Figure 3.1: A summary o f the flow features for various Mach numbers and angles o f incidence (from Ref. [18])

The occurrence o f these shockwave systems in the flow introduces complex shock/vortex interactions particularly
at moderate to high angles o f incidence. These interactions have a significant effect on vortex breakdown and the
breakdown behaviour is quite different to that witnessed for subsonic vortical flows where the onset o f breakdown
is relatively gradual with increasing incidence [72]. An interaction between the rear/terminating shock, described
in Section 1.4, and the primary vortex has been found, in som e cases, to cause breakdown [17, 123] and with

57
CHAPTER 3. TRANSONIC VORTICAL FLOW O N A SLENDER DELTA WING 58

increasing incidence this shock can jump upstream quite abruptly. The upstream shift of the shock is likely to
occur in reaction to changes in the flow behaviour [17], such as adverse pressure gradient, caused by an increase in
incidence. If the conditions are sufficient that the shock/vortex interaction causes breakdown, the sudden upstream
movement of tlie shock will also cause the breakdown location to move upstream. This can cause the location of
breakdown to shift upstream by as much as 30% of the chord in a single F incidence interval [18, 121].

From the literature it has also been noted that it is possible for a terminating shock system to exist without the
breakdown of the vortical system [17] particularly at lower angles of incidence. Whether an interaction occurs
in this case is not well understood. From the study of the interaction between longitudinal vortices and normal
shocks in supersonic flow [130] it has been found that it is possible for a vortex to pass through a normal shock
without being weakened sufficiently to cause breakdown. However, tlie flow over slender delta wings is more
complex as the shock does not appear to be normal to the freestream in the vortex core region [17]. Therefore,
further investigation is needed to consider the behaviour and onset of vortex breakdown, particularly with respect
to shock/vortex interactions.

It is clear from consideration of the literature that the change in flow behaviour with increasing Mach number is
quite considerable, with the occurrence and movement of shocks in the flow and the possibility of abrupt shock-
induced breakdown. This has obvious detrimental effects on the aerodynamic performance of the wing. Aero­
dynamic characteristics such as lift coefficient distribution, stall and pitch may all be badly affected by such flow
behaviour. Therefore, understanding this behaviour is important, particularly for fighter configurations, such as
Eurofighter and JSF, which will perform manoeuvres at transonic Mach numbers.

Therefore, to consider this behaviour, the flow over a sharp leading edged, slender delta wing was considered under
subsonic and transonic conditions. This investigation was undertaken as part of the 2nd International Vortex Flow
Experiment (VFE-2), a facet of the NATO RTO AVT-113 Task Group, which was set up to consider the flow behav­
iour both experimentally and computationally over a specified 65° delta wing geometry. The work of tlie VFE-2
continues on from the first International Vortex Flow Experiment (VFE-1) [181] earned out in the late eighties,
which was used to validate the inviscid CFD codes of the time. Much progress has been made in both experimental
and computational aerodynamics, particular ly in turbulence models since the conclusion of the VFE-1. Therefore,
it was proposed by Hummel and Redecker [182] that a second experiment should be undertaken to provide a new,
comprehensive database of results for various test conditions and flow behaviours, to further the understanding of
vortical flows. The test conditions considered under the VFE-2 framework include both subsonic and transonic
Mach numbers for low, medium and high angles of incidence at a range of Reynolds numbers [183].

For this investigation two test conditions were analysed from the cases specified by the VFE-2, at a single Reynolds
number, Re — 6 x 10^. Both subsonic, M = 0.4, and transonic flow conditions, M = 0.85 will be considered, with
emphasis on the behaviour of the transonic vortical flow. Two angles of incidence are used for consideration,
which correspond to pre- and post-breakdown flow behaviour, a — 18.5° and 23°. Furtlier details of the test case,
geometry and calculation set up will be given in the following section, before analysis of the subsonic and transonic
calculations are detailed. For the transonic conditions, from consideration of the literature, it is found that both
these cases fall within the regions where it is highly likely that cross-flow and rear shocks will occur in the flow.
Therefore, the occurrence of these shocks are analysed. Comparisons between each of the calculations and with
available experimental data are made and consideration of the sensitivity of the flow behaviour to a number of
computational factors, such as grid refinement detailed. A comparison to other numerical investigations from tlie
VFE-2 will also detailed, before consideration of shock/vortex interaction and the occunence of vortex breakdown
over the wing is undertaken. Finally the results are discussed and conclusions made with respect to the discussion
given above.

3.2 Summary of Test Case


The geometry used for the VFE-2 is originally from experiments earned out by Chu and Luckring [20, 132, 133,
134] in the National Transonic Facility (NTF) at NASA Langley. These experiments considered a 65° delta wing
with four leading edge profiles (one sharp and three rounded with small, medium and large radii) for a wide range of
conditions both subsonic and transonic and for both test and flight Reynolds numbers. This data has been compiled
into a comprehensive experimental database and forms the basis for the investigations of the VFE-2. The geometry
is analytically defined for all leading edge profiles. Both the medium radius and sharp leading edge profiles are
considered within VFE-2, however, for this investigation, only the sharp leading edge profile is considered. Figure
3.2 shows the wing situated in the NTF wind tunnel and a brief overview of tlie analytical dimensions of the wing.
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 59

Mlg» gaoni«ty
L«ad0çedg*
y<V«ooi7ooe
d-oiiaaaaewe

Ti«Miig «dtf» *#omMy

y n v .- 0 1 0 |3 d t- M t= 4 d i* I
y 'C p -0 û 1 7 0 û S d . O ITOOdOOCOWO*

Taibng #dg#ck*w*#
»m09797Cf,

(a) W ing in N T F facility a t N A S A L a n g ley [20] (b ) A n aly tical d efin itio n o f w in g d ata [ 184]

Figure 3.2: VFE-2 65° delta wing geometry used in investigation

All calculations performed were steady state and used the — co turbulence model with Enhancer [158]. This
model is detailed and discussed with reference to its use for vortical flow in Chapter 2. It has been well validated
against experiment for similar sub- and transonic steady vortical flow calculations [144, 158, 159].

3.2.1 Grid Generation


One o f the most important issues for computational flow calculations is grid generation and establishing the de­
pendence o f the solution on the grid. There have been many investigations considering various aspects o f grid
generation particularly for delta wing flows [153, 154, 156, 185, 186, 187]. From these investigations the impor­
tance o f having a grid which is suitably refined in the regions o f interest, in order to accurately capture the most
important and influential flow features over the wing, is apparent. These areas include, for a delta wing, the bound­
ary layer, the shear layer region and the leading edge vortex core. Other factors which have been highlighted are:
grid topology, cell skewness, wall spacing and overall grid refinement and distribution.

The structured multi-block grids used in this investigation were manually created using the ICEMCFD mesh gen­
eration package, Hexa. The computational model consists o f the semi-span wing, reproduced from the analytical
definition. The sting was also reproduced to approximately one chord length downstream o f the trailing edge,
based on the recommendations o f Allan et al. [71], who found that the effect o f a sting or support apparatus was
negligible beyond this location. Downstream, an approximation to the experimental sting was defined to the far
field, which was defined as 20cr in each direction from the wing apex to m inimise the effect o f the boundaries on
the results.

An H-H topology was chosen with a collapsed edge at the apex o f the wing. In order to allow for a smooth grid
point distribution and refinement o f the grid, a structured Ogrid was used around the sting. An example o f this and
the surface blocking topology is shown in Figure 3.3. Overall, the blocking structure was optimised for reduced
skewness, particularly in the sting tip region and as a result a total o f 353 blocks was used. Based on this block
topology, two grids were created for this investigation with varying refinement. These are classed as coarse and
fine with the important details o f each grid summarised in Table 3.1. The nominal value is based on the first wall
spacing and the Reynolds number o f the flow and may vary slightly over the surface o f the wing. A comparison o f
the relative refinement o f the grid on a plane upstream o f the sting blocking at x/cy = 0.5 is shown in Figure 3.4.
Each grid distribution allowed for an efficient load balance o f grid points across the optimum number o f processors
used for the calculations.

N om in al N u m ber o f Points on Wing Surface


Type G rid Size Wall Spacin g y"" Stream w ise @ L E Spanw ise @ TE N orm al
Coarse 2,451,314 - 2 X IQ-^c^ ~ 4 .4 117 171 49
Fine 6,993,522 1 X 10“^Cr 2.2 170 228 81

Table 3.1: Summary o f main features o f grids used for VFE-2 investigation
CHAPTER 3. TRANSONIC VORTICAL FLOW O N A SLENDER DELTA WING 60

Figure 3.3: Surface mesh and ogrid topology around sting region for 65° VFE-2 delta wing

mi = 1 = 1 = = = : = =

IS

(a) C o arse G rid (b ) F in e G rid

Figure 3.4: Comparison o f grid refinement at x /cy = 0.5

3.2.2 Transition Treatment


Convergence issues associated with the apex singularity o f the H-H grid, mentioned above, are dealt with by fixing
a transition from laminar to turbulent flow downstream o f the apex in the computational domain. Transition was
applied at various constant streamwise locations x = 0.1 — 0.4 to consider the effect on the flow behaviour for
both the subsonic and transonic conditions. From this analysis, it was found that the subsonic results were highly
sensitive to transition location, with the optimum solution being obtained for x = 0.1 (x/cy = 0.10154 on the wing
surface). However, the transonic results were not found to be sensitive and thus, the transition was set to x = 0.4,
which corresponds to x /cy = 0.406125 on the wing upper surface.

3.3 Subsonic Vortical Flow: Results


As stated in the introduction to this chapter, both subsonic and transonic cases were considered within the VFE-2
framework for the sharp leading edge wing. Although the main purpose o f this study is to consider the transonic
behaviour o f vortical flow and vortex breakdown on the wing, it is also important to consider the behaviour under
subsonic conditions. This will allow for further validation o f the CFD solutions and therefore greater confidence
in the the predicted flow behaviour for the more com plex transonic flow. Two angles o f incidence were considered
- 18.5° and 23° at a Mach number o f M = 0.4 and Reynolds number o f 6 million. As mentioned previously, these
conditions correspond to pre- and post-breakdown flow behaviour for this geometry. All results were obtained on
the fine grid as detailed in Table 3.1.

To allow validation o f all computational results, comparisons were made with the N A SA NTF experimental pres­
sure coefficient distributions at five streamwise locations on the wing surface, x /cy = 0.2, 0.4, 0.6, 0.8 and 0.95,
shown in Figure 3.5. It is clear that the agreement between the computational solutions and the experiments are
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 61

very good. For most streamwise locations, the position and magnitude o f the suction peaks are well predicted for
both angles o f incidence. Surface contours o f pressure coefficient are also shown, which clearly show the extent
o f the primary peak and also the existence o f secondary vortices close to the leading edge. From the spanwise dis­
tributions, the suction peak due to the secondary vortices is much clearer than shown for the experimental results.
The comparable strength o f this region is evident from slight differences in primary peak location for a = 18.5°.
This shows the computational suction peak located slightly inboard compared to the experimental data, suggesting
that the secondary vortices are larger for the computational results. However, this does not appear to be the case
for a = 23°.

At a = 23°, breakdown occurs on the wing. However, agreement with the experiments is still good in the post­
breakdown region, with only a slight under-prediction o f the suction peak magnitudes. This suggests that break­
down may be slightly more severe in the computational results than in the experiment. An interaction between the
breakdown region and the surface o f the w ing is apparent from the surface contours o f pressure coefficient shown
in Figure 3.5, where a small low pressure region is found downstream o f the breakdown location.

Figure 3.6 shows contours o f x vorticity and u velocity at streamwise slices over the wing, which allows the struc­
ture o f the flow to be seen clearly. In each o f the plots the vortex core trajectory is defined. Considering the
18.5° case first, it is clear that breakdown does not occur and that the primary vortex core is strong and relatively
straight over the wing. The contours o f x vorticity also show the presence o f the strong secondary vortex system,
described previously. From analysis o f the axial velocity o f the vortex core, it was found that the axial flow ac­
celerates up to a maximum o f 1.95t/oo at x /c r = 0.9 after which it appears to decelerate. This deceleration o f the
vortex core may be caused by the highly curved nature o f the trailing edge geometry. A lso clear is an area o f stag­
nant flow and the apparent breakdown o f the secondary vortex. It is possible that this unusual behaviour is caused
by the rounded nature o f the trailing edge and the intersection between the leading edge and trailing edge curvature.

Comparing the contour plots for the pre-breakdown flow to those for a = 23°, shows that the size and strength
o f the vortices increase with increasing incidence. It is also found that the distance between the vortex core and
wing surface increases. The spiral behaviour o f the vortex breakdown is obvious from the vortex core trajectory
with the expansion o f the vortex core and the flow reversal shown clearly by the contours o f u velocity. Upstream
o f breakdown the maximum axial velocity within the vortex core was found to be approximately 2.2Uco, which is
almost a 12% increase on the pre-breakdown case. The location o f vortex breakdown, taken as the location on the
vortex core where the axial flow stagnates, Uaxiat = 0, is approximately jc/cr = 0.7 7 5 . For the higher incidence, the
secondary vortex is also clear from the flow structure. A third vortex core trajectory is also evident, which appears
to intersect regions o f higher vorticity in the shear layer.

M = 0.4, Re = ic i M =0.4,R c = tc6

CFD RcsdUc, 18.5* CFD HccuRc, 23*


- NTF Wind lumcl tautts, 10.4* NTF Wind t m c l iccultc, 23.5"

(a ) 18.5° (b ) 2 3 °

Figure 3.5: Computational results compared to experimental data, M = 0 .4 , Re = 6 x 10^


CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 62

(a ) 18.5°

■ViMtir

(b ) 23°

Figure 3.6: Contours o f x vorticity and u velocity on slices through the vortex core for a = 18.5° and 23° - M = 0.4
and Re = 6 x 10^

To allow further validation o f the results, the subsonic solutions from this investigation were also compared to
results generated by other institutions as part o f the VFE-2. Figure 3.7 shows comparisons o f the surface pressure
coefficient distributions with results obtained by NLR and E AD S-M A S. Details o f the grids, turbulence models
and flow solvers used for these results are given in Ref. [188] and summarised in Section 3.6. It is clear from these
plots that there is close agreement between the computational results, with only slight differences in the size o f the
primary and secondary suction peaks. Further details and comparisons between the current work and the results
from these institutions w ill be given in a later section detailing the transonic flow behaviour.

M m0.4, Re m6e6 M « 0.4, Re « 6e6


- EADS 18.5* - EADS 23*
- Glasgow 18.5* - Glasgow 23*
- NLR 18.5* - NLR 23*
■ NTF Wind tunnel icsults, 18.4* ~ NTF Wind tunnel icsulb,

(a ) 18.5° (b ) 2 3 °

Figure 3.7: Comparison o f computational results and experimental data, M = 0 .4 and Re = 6 x 10^
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 63

3.4 Transonic Vortical Flow: Results


The calculations performed to consider the transonic regime correspond to conditions, M = 0.85, a Reynolds
number o f Re = 6 x 10^ at the same angles o f incidence, a = 18.5° and 23°. As before, all results were obtained
on the fine grid. However, the effect o f grid refinement will be considered and is detailed in a subsequent section.
As the main purpose o f this investigation is to consider the behaviour o f transonic vortical flow, the results will
be considered in more detail than the subsonic results. The results at each incidence will, initially, be considered
separately under the headings pre- and post-breakdown flow.

3.4.1 Pre-Breakdown Flow - M = 0.85, a — 18.5^


The computational results and corresponding N A SA NTF experimental data [20] for a = 18.5°, are shown in Fig­
ure 3.8. At this incidence, it is clear that, overall, the agreement between the results is good. For most streamwise
locations, the magnitudes and positions o f the suction peaks are well predicted. Although, as with the subsonic
results, there does seem to be a consistent over-prediction o f the secondary vortex peak, which appears to lessen
with increased distance from the apex. This is not related to the location o f forced transition as it was found from
investigation, that the overall flow behaviour was insensitive to transition location and the strength o f the secondary
vortex was relatively unchanged.

M » 0.85, R* = 6e6
CFD Results 18.5'
NTF Wind tunnel results, 18.6'

Figure 3.8; Computational results compared to experimental data, a = 18.5°, M = 0.85 and Re = 6 x 10^

Contours o f surface pressure coefficient, are also shown in Figure 3.8. These clearly show the primary and sec­
ondary suction peaks and their behaviour. Downstream o f x /c r = 0.8, it appears that the secondary vortex disap­
pears and the primary vortex curves inboard toward the sting region at the trailing edge. This is also clear from
consideration o f the pressure coefficient distributions, which show a flat distribution outboard o f the primary vor­
tex for x /c r = 0.95. The axial velocity through the vortex cores was analysed and it was found that the secondary
vortex breaks down in this region as it approaches the trailing edge, as shown in Figure 3.9. From this plot, it is
evident that the secondary vortex breaks down at approximately x /c r = 0.8 5 , however, the primary vortex does not.
This behaviour is very similar to that observed for the subsonic case and, thus, may also be due to the geometry
in this region. From Figure 3.9 it is clear that the maximum axial velocity o f the primary vortex is approximately
I J U „ which corresponds to a maximum local Mach number o f 1.8 and indeed the axial flow in the secondary
vortex is also supersonic upstream o f breakdown. Thus, it may also be suggested that this location coincides with
the presence o f a shock in the flow and that a type o f shock/vortex interaction is occurring. However, if this is the
case, the primary vortex is largely unaffected by the interaction. Further consideration o f this region and analysis
o f the flow solutions is needed to determine the causes o f this behaviour. A nalysis and discussion o f the presence
o f shocks and shock/vortex interactions w ill be given in a later section.
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 64

1.5

I
I

•9.S
9 9.1 9.4 0.6 0.7 0.8 0.9 1

Figure 3.9: Axial velocity through primary and secondary vortex cores, a = 18.5®, M = 0.85 and Re = 6 x 10^

(a) P ressu re coefficient (b ) X vo rticity (c) u v elocity

Figure 3.10: Contours o f x vorticity and u velocity on slices through the vortex core for 18.5®, M = 0.85 and
Re = 6 x 10^

The flow structure o f the leading edge vortices was considered from the plots shown in Figure 3.10 with contours
o f pressure coefficient, x vorticity and u velocity used to aid in the understanding o f the behaviour o f the three
dimensional flow. From consideration o f the contours o f pressure coefficient, it is clear that the vortex cores have
a quite uneven shape, particularly in comparison to the subsonic vortices which are quite round and uniform in
the pre-breakdown flow. The vortices are also closer to the wing surface. From examination o f the x vorticity
contours, it is found that the vortex system is relatively flat and elongated over the wing surface. Closer to the apex
o f the wing, a tertiary vortex is found under the secondary vortex. Analysis o f the contours o f u velocity and the
vortex core trajectories, also confirms the occurrence o f the secondary vortex breakdown between the streamwise
positions o f x /c r = 0.8 and 0.9 with a large region o f reversed flow occurring outboard o f the primary vortex. A
fourth vortical region, with the same sign as the the primary vortex is found outboard o f the primary vortex within
the shear layer. Its location is virtually constant at each streamwise positions, until the secondary vortex breaks
down, where it moves upward, away from the leading edge region.

Figure 3.11: Contours o f vorticity at a position jc/c,- = 0.4 for a = 18.5®, M = 0.85 and Re = 6 x 10^

The vortex core structure can be further considered by examining a single slice through the vortex core. Figure
3.11 shows contours o f x vorticity on a slice o f the domain at the streamwise location, x /c r = 0.4. At this location,
both secondary and tertiary separation regions are found and the large size and strength o f the secondary vortex
is evident. Outboard o f the secondary vortex, the fourth vortical region mentioned above is clear. Initially it was
thought that this small region o f vorticity may be evidence o f a shear layer instability. However, its behaviour is
not the same as the shear layer structures described in Chapter 1 due to the steady nature o f the solution. Further
analysis suggests that the structure is caused by an interaction o f the secondary vortex and the shear layer. This will
be caused by the close proximity o f the primary vortex to the surface o f the wing and the over-predicted strength
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 65

o f the secondary vortex, which would cause high velocities in this region. Outboard o f the vortex, on the wing
surface, a much smaller region o f vorticity is found, which suggests that the boundary layer separates again under
the influence o f this region.

3.4.2 Post-Breakdown Flow - M = 0.85, a = 23"


As before, the pressure coefficient distributions are compared in Figure 3.12. From these comparisons, it is clear
that close to the apex o f the wing i.e. x / c r = 0.2 and 0 .4 , the agreement is good, with the magnitude and location
o f the primary and secondary peaks being predicted w ell. However, downstream o f the x /c r = 0.4 location, the
computational results show large discrepancies with the a — 23.6® experimental results. From consideration o f the
behaviour o f transonic vortex breakdown described in Section 1.4 and the surface pressure coefficient contours,
it is clear that these discrepancies are due to vortex breakdown occurring on the wing. Analysis o f the NASA
NTF experimental data has shown that vortex breakdown occurs at an incidence o f 24.6®, which is the next test
point in the data set. These results are also included in Figure 3.12 and show a much improved agreement with
the computational results. Therefore, it may be concluded that the vortex breakdown behaviour is predicted well.
However, discrepancies exist in the prediction o f the critical onset angle. Further consideration o f this will be given
in a later section.

M = 0.85, Rc « 6e6
CFD Results 23*
NTF Wind tunnel results, 23.6’
• NFT Wind Tunnel results 24.6*

Figure 3.12: Computational results compared to experimental data, a = 23®, M = 0.85 and Re = 6 x 10^

From the surface pressure coefficient contours, the abrupt nature o f vortex breakdown is evident. Upstream it is
clear that the vortex system is coherent and strong, however, the vortices disappear quite suddenly. This is quite
unlike the vortex breakdown found for the subsonic case. The axial velocity through the primary vortex core
(Figure 3.13) also shows the almost immediate onset o f breakdown, which occurs at approximately x /c r = 0.57.
Comparing the axial flow to the ct = 18.5® case shows that the maximum axial velocity has increased to approxi­
mately 1. 86/00. Therefore, as expected, the vortices have increased in strength.

The three-dimensional behaviour o f the flow can be seen in the plots o f Figure 3.14, which are similar to those
shown for the pre-breakdown case in Figure 3.10 . Compared to the results for 18.5® it is clear that the size o f
the vortex core upstream o f breakdown has increased in diameter, however, the vortices still have a very elongated
shape. The region downstream o f breakdown is also relatively flat against the surface o f the wing, possibly caused
by the high freestream velocity limiting the growth o f such a structure into the flow. The fourth vortical structure
found in the pre-breakdown flow, is also found for this case upstream o f breakdown at a constant location outboard
o f the primary vortex. At the breakdown o f the secondary vortex, which also occurs slightly upstream o f the
primary vortex for this case, this vortex is swept upward away from the leading edge and entrained into the post­
breakdown flow.
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 66

Figure 3.13: Axial velocity through vortex core for post-breakdown flow, a = 23®, M= 0.85 and Re = t x 10^

MV tM d ly

(a) X vorticity (b ) u velo city

Figure 3.14: Flow structure for a = 23®, M = 0.85 and Re = 6 x 10^

3.5 Occurrence of Shocks in the Flow


As detailed in the literature review in Section 1.4, it is expected that a number o f shock systems will be present
for these conditions. Care was taken to analyse the flow solutions described in the previous section to determine
the occurrence, location and behaviour o f shockwaves in the flow. The method o f analysis used was described in
Chapter 2 and allowed the interpretation o f both shocks occurring in the cross-flow and those normal to the flow
direction and wing surface. Each o f these shocks will be considered separately in this section.

(a) M ach n u m b e r (b ) P ressu re G ra d ie n t M ag n itu d e

(c ) S h o c k feature (d ) Jt v o rticity

Figure 3.15: Plots for x/cr = 0.4 showing contours o f flow variables to highlight locations o f cross flow shocks for
a = 18.5®, M = 0.85 and /?e = 6 x 10^
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 67

3.5.1 Cross-Flow Shocks


Evidence o f a com plex cross-flow shock system for both angles o f incidence, beneath and around the primary
and secondary vortices was found from consideration o f the flow structure in a spanwise cut using the methods
described previously. An example o f this flow behaviour at x /c r = 0.4 and the plots used for determination o f the
shock locations for the pre-breakdown case is shown in Figure 3.15. Each o f the identified shock locations are
marked on the variable contour plots.

The determination o f the first o f these cross flow shocks, denoted by in Figure 3.15(d), was aided by consider­
ation o f the pressure coefficient distributions and surface contours shown in Figures 3.8 and 3.12 in the previous
sections. In these plots it was found that outboard o f the primary vortex suction peak, sharp changes in pressure
coefficient are found. These sharp changes in pressure coefficient may indicate the presence o f a shock in the flow
as described in Section 1.4 and shown in Figure 1.20(a). It is clear from the surface pressure coefficient contours
and was also indicated from analysing multiple slices through the domain (not detailed here), that this shock occurs
in the flow for a constant non-dimensional spanwise location, defining a conical ray from the apex o f the wing.
These locations are approximately y / s = 0.6 4 for the ct = 18.5® solution and y / s = 0 .62 for the a = 23® solution.
With closer inspection, it was found that coinciding with the location o f this shock close to the wing, the boundary
layer thickens and separates to form a strong secondary vortex as shown in Figure 3.16.

0.6 0.61 0.62 0.63 0.64 0.65 0.66 0.67 0.68 0.69 0.7
y /s

Figure 3.16: Velocity vectors and contours o f Mach number at chordwise station x / c r = 0.2 showing secondary
separation for a = 18.5®, M = 0.85 and Re = 6 x 10^

A second sharp increase in pressure coefficient was also found outboard o f the secondary vortex in the pressure
coefficient surface contours as described before. A gain, as before, a small shock can be found in the cross-flow
planes corresponding to this location and denoted by l a in Figure 3.15(d). This shock is likely to be caused by
a similar mechanism as shock [IJ, but occurs under the secondary vortex with the cross-flow travelling toward the
centreline o f the wing. Inboard o f this location a small tertiary vortex system is found and it is supposed that the
separation is again caused by the adverse pressure gradient associated with the shock. As with shock the shock
is conical and has a constant spanwise location o f y / s = 0.82 for both a = 18.5® and a = 23® angles o f incidence.
It should be noted that while it is proposed in this work that regions and l a correspond to the locations o f
shocks in the flow, it is difficult to confirm this conclusively. There remains a possibility that these shocks are in
fact strong compression regions, which are causing the separation o f the flow. Further work, both experimentally
and computationally are needed to confirm this.

Between the secondary separation region and the primary vortex, the spanwise flow behaves in a similar manner
to that in a convergent-divergent duct and accelerates to supersonic conditions. At som e point, the flow can no
longer maintain these high velocities and a shock appears to decelerate the flow. This is likely to be the cause
o f shocks and 2 ^ in Figure 3.15(d). Shock 12a] appears to occur due to the flow accelerating again beyond
shock It is not clear at this point whether shock 0 and shock are connected or interact. However, it
appears that they sit very close and it is possible that shock is a stronger continuation o f shock [T]. If this is
indeed the case, the resulting shock curves upward from the surface to the primary vortex, as suggested by the di­
agram o f Figure 1.20(a). From the literature, it is known that a shock sits in the region between the primary vortex
and the surface o f the wing [1 2 1 ,1 2 2 ]. However, there is little existing data which confirms the shape o f this shock.

Two other shocks were found to occur in the cross-flow. Shock is found to sit above the primary vortex and
is similar to that found in the computations o f Gordnier and Visbal [125] and Shock sits above the primary
shear layer, close to the leading edge. Both these shocks are likely to be caused by the curvature o f the shear layer
causing the flow to accelerate up to conditions which cannot be sustained. All these shocks were also found to
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 68

occur for the a = 23® case, although the majority o f the shock locations are different due to the inboard movement
and relative increase in size o f the vortical system. This is shown in Figure 3.17.

Figure 3.17: Contours o f x vorticity on a x /c r = 0.4 plane, highlighting locations o f cross flow shocks for a = 23®,
M = 0.85 and Re = 6 x 10^

3.5.2 Normal Shocks


Normal shocks are also found to occur in this flow, and are identified by plotting the pressure coefficient along
the symmetry plane as shown in Figure 3.18 for both angles o f incidence. For the 18.5® case, it is clear that two
normal shocks occur at the symmetry plane. The first occurs upstream o f the sting tip at approximately x /c r = 0.6,
which is most likely to be caused by the sting geometry. Further downstream at approximately x /c r = 0.85 a
second shock is found. This second shock is likely to corresponds to the rear/terminating shock as described in
the literature [17, 18, 129] for similar conditions. A third compression region is also found close to the trailing
edge, and a third shock is found from the surface pressure contours at this location outboard o f the symmetry plane
on the wing surface. A shock occurring at this location is likely to be caused by the high curvature o f the wing
geometry and the necessity o f the flow to return to freestream conditions at the trailing edge.

&

18.5"
23'

Figure 3.18: Pressure coefficient distribution at the symmetry plane on the wing for both angles o f incidence

As the incidence is increased and vortex breakdown occurs on the wing, the behaviour at the symmetry plane,
again, shows the shock at the sting tip at approximately x /c r = 0.6. However, another shock is also found in the
flow slightly upstream o f this location at about x / c r = 0.52. Downstream o f the sting tip, it is evident that the
rear/terminating shock described for the a = 18.5® case is no longer present. From the behaviour described in
the investigations o f Elsenaar and Hoeijmakers [18] under similar conditions, it is possible that the new shock
upstream o f the sting tip is the rear/terminating shock having undergone an upstream shift with the increase o f
incidence. However, due to the presence o f the sting and the shock caused by this geometry, it not possible to state
this conclusively. As before, it is found that three normal shocks occur at the symmetry plane and close to the
trailing edge, as also found in the experiments, a second normal shock is observed. This is likely to be the same
trailing edge shock as found for a = 18.5®.

Considering the three-dimensional behaviour o f the normal shocks, it is found that the shock occurring upstream
o f the sting tip curves downstream and intersects the rolled up shear layer o f the vortex as shown in Figure 3.19
and highlighted by the dashed lines. This is also in agreement with the observations o f D onohoe and Bannink [17]
and the schematic shown in Figure 1.20(b) for the rear/terminating shock. However, it is likely that this curvature
is caused by the sting presence for this configuration. A lso highlighted are the locations o f the other normal shocks
described above. The rear/terminating shock in the 18.5® solution is found to be normal to the freestream and wing
surface and does not appear to curve downstream outboard o f the symmetry plane. This lack o f curvature may be
due to the influence o f the sting on the flow, as previous investigations have considered a flat wing without sting
CHAPTER 3. TRANSONIC VORTICAL FLOW O N A SLENDER DELTA WING 69

support [17]. A lso clear from this plot are the two cross-flow shocks which sit above the vortex described above
(fT] and 0 from Figures 3.15 and 3.17). It is possible, for both angles o f incidence, that there is an interaction
between these cross-flow shocks and the normal sting tip shock, which w ill further increase the complexity o f the
flow in this region. However, further experimental data is needed in this region to determine this behaviour.

a = 18.5 a = 23*

Figure 3.19: Isosurface o f x vorticity coloured by pressure coefficient showing primary vortex shear layer and
normal shock shape for both angles o f incidence

3.6 CFD Sensitivity Study


A s has been shown in the previous sections, the agreement with the experimental data is good for the pre- and post
breakdown flow, however the critical incidence for vortex breakdown on the w ing is not predicted w ell. Due to the
presence o f the shocks in the flow, it is quite likely that this flow will be more sensitive to computational factors
than a subsonic flow and this must be checked in order to improve confidence in the solutions. In this section,
a number o f parameters will be considered. These include grid issues such as refinement and type, turbulence
m odelling, convergence and time accuracy issues. For all cases, with the exception o f the effect o f grid refinement,
only the post-breakdown case, a = 23®, for conditions M = 0.85 and Re = 6 x 10^ will be considered.

To allow further analysis o f various aspects o f the flow behaviour, comparisons were made with calculations
performed by other institutions as part o f the VFE-2. These institutions are EAD S Military Air Systems (EADS-
M AS) and NLR using structured, multi-block grids and the United States Air Force Academy (USAFA) using an
unstructured grid. Each institution uses its own well-validated 3D RANS flow solver; FLOWer 116.17 at EADS-
M AS [189], ENSOLV at NLR [1 9 0 ,1 9 1 ,1 9 2 ] and Cobalt at USAFA [193], respectively. Comparisons between the
structured flow solvers and PMB at Glasgow University have been made in the past [194]. Detailed descriptions
o f each o f these flow solvers, computational set-up and grids used in the structured grid comparisons can be found
in Ref. [188] and are summarised along with the current investigation details in Table 3.2.

Size No. o f Grid Points on Wing


Institution Topology xlO*^ Spanwise Streamwise N o rm al Turbulence Model
EADS C -0 ~ 10.6 129 257 129 W ilcox k-tu and
Reynolds Stress Model
NLR C -0 ~4 192 112 96 TNT k-o> with
P(o Enhancer
Current H-H ~7 170 228 81 W ilcox k-tu with
Investigation with 0-grid Poi Enhancer and NLEVM

Table 3.2: Summary o f grids and turbulence models used for VFE-2 structured grid comparisons
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 70

The effect o f time accuracy is considered by comparing the current solutions to calculations performed by USAFA
using the Spalart-Allmaras DES turbulence treatment on a unstructured grid. The grid used had approximately
7.89 X 10^ cells and an average first wall spacing o f = 0.68, created specifically for a Reynolds number o f
6 X 10^. It was refined within the vortex core region to improve the grid for the application o f DES. The grid
structure at the symmetry plane is shown in Figure 3.20. The time step was defined as A/ = 5 x 10"* seconds,
which corresponds to a non-dimensional time step o f A T w 0.0047. The calculation was allowed to run for approx­
imately 20600 time steps, which results in a total time o f approximately 0.1 seconds. For these comparisons, both
instantaneous and time averaged (mean) solutions were necessary and thus, a time averaged file was created over
a total o f 4000 time steps.

Figure 3.20: U SAFA grid at symmetry plane

3.6.1 Effect of Grid Refinement


A s stated, the effect o f grid refinement was considered for both pre- and post-breakdown flow for the transonic
conditions. In this study, the solutions detailed previously for the fine grid are compared to results obtained using
the coarse grid described in Section 3.2.1. Comparisons o f the surface pressure coefficient distributions for both
angles o f incidence with the relevant experimental data are shown in Figure 3.21.

M = 0.8S, Rc = 6 t 6 M = I.8S, Re = 6 ct
- Coaoc Grid, lt.5 *
- C o m ic Grid.23*
- Floe Grid, ll.S* - Fine Grid, 23*
- NTP Wind tunnel rcsidti, H.<* " NTF Wind tkmocl m id k , 23.6"
- NTF WIml Ttaim i RecnlM, 24.6* *

(a ) 18.5" (b ) 2 3 °

Figure 3.21: Comparison between the H-H grids for transonic conditions at ct = 18.5® and 23®
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 71

Considering both angles o f incidence, it is clear that there are a number o f differences between the solutions on the
two grids, particularly in the trailing edge region. C lose to the apex, agreement is good for both cases, with the fine
grid giving slightly higher suction peaks than the coarse grid. In this region, both primary and secondary vortex
suction peaks are clear and coincide for both grids, up to x /c r = 0 .6 for a = 18.5® and x /c r = 0 .4 for a = 23®.
Downstream o f these locations, the differences in the distributions becom e more pronounced. For a = 18.5®, at
x / c r = 0.8, the pressure coefficient distribution shows that the fine grid gives better agreement with the experi­
mental data. It is clear from the under-prediction o f the pressure gradients that the coarse grid is not resolving
the cross-flow shocks as well as the fine grid, as expected. Similar behaviour is shown at x /c r = 0.95, where the
cross-flow shock region is much further outboard and the suction peak is over-predicted. For a = 23®, downstream
o f breakdown, the agreement between the two grids is close, with a similar reduction in suction peak found at
x /c r = 0.6 and similar flat distributions obvious downstream o f this location.

Further comparisons can be made from the pressure coefficient contours on the wing surface, shown in Figure
3.22. For the pre-breakdown case, these plots emphasise the smearing o f the gradients on the coarse grid, with
the primary vortex suction peak being much broader. A lso evident is the behaviour o f the secondary vortex which
does not appear to have such an obvious breakdown location in contrast to the fine grid. Considering the post­
breakdown case, the differences between the two solutions are, again, harder to determine. The behaviour o f the
vortex breakdown is almost identical, with the location o f the normal shock upstream o f the sting tip coinciding.
It is likely that as with the pre-breakdown case, the shocks in the flow are more smeared for the coarse grid, how­
ever this does not appear to have a significant effect on the flow behaviour and the occurrence o f vortex breakdown.

Figure 3.23 shows the axial velocity through the primary vortex cores for both cases. It is clear that with the
increase in grid refinement, the axial velocity increases by approximately 30%Uoo for both cases. This increase in
axial velocity is expected and is most likely to be due to the improved refinement o f the vortex core region. For the
post-breakdown case, the onset and behaviour o f breakdown is evident from this plot. It is clear that the onset o f
breakdown occurs at roughly the same point over the wing {x/cr = 0.57 for the fine grid and x /c r = 0.58 for the
coarse grid).

(a ) 18.5“ (b ) 23“

Figure 3.22: Surface contours o f pressure coefficient for comparison between the H-H grids

J J

C oaise G rid C o a n c G rid


Fine Grid Fine G rid

(a ) 18.5“ (b ) 2 3 “

Figure 3.23: Comparison o f axial velocity through the vortex cores for coarse and fine grid solutions
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 72

The differences in vortex core resolution are also shown by analysing contours o f x vorticity on a cross-flow plane.
Figure 3.24 shows a cross-flow slice at x /c r = 0.4, for both grids at a = 18.5®. It is clear from these plots that
the fine grid predicts a much more compact vortical system than the coarse grid. Both the primary and secondary
vortices are stronger for the fine grid solutions and as a result the outboard vortical region in the shear layer is
not found on the coarse grid. Tertiary vortices are found for both cases and in general the location o f each o f the
vortices is the same for both grids. Similar comparisons were also made for the post-breakdown case, but are not
shown.

(a) Coarse grid (b) Fine grid

Figure 3.24: Contours o f jc vorticity at chordwise station x / c r = 0 .4 at 18.5®

This study has shown that the behaviour and location o f vortex breakdown within transonic flow are not greatly
affected by the grid refinement carried out. It is also evident that the critical angle for vortex breakdown onset is
independent o f grid refinement, as vortex breakdown is predicted to occur early for both grids.

3.6.2 Effect of Hirbulence Model


The effect o f turbulence model on the flow behaviour was considered by comparing the results detailed in the
previous sections, calculated using the A: — cu with Enhancer model to results obtained using the Non-Linear
Eddy Viscosity model for the post-breakdown incidence, a = 23®. The Non-Linear Eddy Viscosity model calcu­
lation was performed using the same flow conditions as before and was started from the end o f the k — to with Poj
Enhancer calculation discussed in the previous sections. It was run for the same number o f total iterations using
the same calculation parameters. Further consideration o f the effect o f turbulence model will be obtained from a
similar study carried out by E AD S-M A S for the same case, comparing the standard W ilcox k — o) and a Reynolds
Stress model (RSM).

Considering the current results first. The surface pressure coefficient distributions are compared for each turbu­
lence model and to the relevant experimental data as shown in Figure 3.25(a). It is clear that close to the apex, at
x /c r = 0.2 and 0.4, the agreement between the distributions is very good. However, downstream at x / c r = 0.6 there
is a significant difference in the pressure coefficient distributions. The Non-Linear Eddy Viscosity model predicts
behaviour which is still in good agreement with the experimental data for the 23.6® data point, which suggests that
at this location breakdown has not yet occurred. Whereas for the original results, it was found that breakdown o c­
curred at x /c r = 0.57, therefore this streamwise location is downstream o f breakdown and the agreement with the
24.6® experimental data is good, where breakdown also occurs on the wing. Further downstream, by x / c r = 0.8,
it is clear that vortex breakdown has occurred for the Non-Linear Eddy V iscosity solution, although inboard there
is still som e agreement with the experimental results for a = 23.6®. There is little agreement with the results for
the A: - Ü) with Pa, Enhancer model, which as described previously, shows a very flat distribution downstream o f
breakdown. This difference in solution behaviour continues downsteam.

Further evidence o f the differences between the two solutions can be obtained from direct comparison o f contours
o f the surface pressure coefficient for the whole w ing. These are shown in Figure 3.26(a). It is evident from this
plot that the location o f breakdown is quite different for each solution. From analysis o f the vortex core behaviour
it was found that the location o f vortex breakdown for the Non-Linear Eddy Viscosity model was I4%Cr further
downstream at x / c r = 0.71. However, upstream o f breakdown, it was found that the solution were in good agree­
ment, with the same axial velocity being predicted and a similar vortex structure, as described for the original
results in the previous section, for the Non-Linear Eddy Viscosity model results. The location o f shocks within
the flow was relatively similar for the Non-Linear Eddy Viscosity results, with cross-flow shocks appearing in the
flow as previously described. A normal shock was found to occur slightly upstream o f the sting tip, however a
second shock upstream o f this location was not found for this case. Downstream close to the trailing edge, a third
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 73

normal shock is also apparent, which is in agreement with the results predicted for the ^ — O) with Poj Enhancer
model solutions. Therefore, from these comparisons, it appears that the choice o f turbulence model influences
the location o f breakdown, but the general behaviour o f the flow is relatively unchanged, particularly upstream o f
breakdown.

A similar analysis was carried out on the results from the EADS investigation, which shows that there is little
difference in the solutions predicted by the W ilcox k — (0 and RSM turbulence m odels. From the surface pressure
distributions in Figure 3.25(b), it is clear that the predicted behaviour is similar with the main differences occurring
at x / c r = 0.6, outboard in the secondary vortex location. Vortex breakdown occurs slightly downstream o f this lo­
cation for both cases, at approximately x /c r = 0.68 for the W ilcox k — w model and at approximately x /c r = 0.7 0
for the RSM. This slight difference in location may explain the discrepancy in the secondary vortex prediction at
x /c r = 0.6, which is not found to be significant to the overall flow behaviour. Downstream o f breakdown, dif­
ferences in the pressure coefficient distributions are apparent, but the agreement with the 24.6^ experimental data
point is relatively good for both models. Further evidence o f the similarities between the flow solution is found
from the contours o f surface pressure coefficient shown in Figure 3.26(b). This highlights the slight change in
location o f the breakdown but confirms the overall agreement in the behaviour on the wing surface. It is clear that
the agreement downstream o f breakdown is much better for these results compared to the current results, however
the change in breakdown location is not as significant.

Therefore, it may be concluded that main effect o f the choice o f turbulence model is in the predicted location o f the
breakdown. The differences found in the flow solution appear to originate with this change and not in differences
o f fundamental flow behaviour. Each model still predicts breakdown to occur on the wing at an incidence which
is lower than that witnessed in the experiment, thus, it may also be concluded that the critical angle for breakdown
to occur on the wing is unaffeced by turbulence model.

M -0 .8 5 , R e-6 c 6 . 23’ M « 0.85, Re « 6e6


- NLEVM - RSM, EADS 23’
- k • w wldi P , Enhancer 0.1 - Wilcox k . w, EADS 23’
- NTF Wind tunnel results, 23.6“ “ NTF Wind tunnel results, 23.6’
" NTF Wind tunnel resulb 24.6* 0.2 » NTF Wind tunnel results 24.6’

0.3

0.4

0.5

0.6

0.7

0.8

- 0 .9

»- 1

(a ) C o m p a ris o n b etw e en N L E V M an d k - u ) w ith Pa, E n - (b ) C o m p a riso n b etw e en R S M a n d W ilco x k - w m o d el


h a n c e r m odel (C u rre n t R esu lts) (E A D S -M A S R esu lts)

Figure 3.25: Effect o f turbulence model on flow solution with comparison to experiment for M = 0.8 5 , Re = 6 x 10^
and a = 23®
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 74

k - w with NLEVM
P Enhanctr

(a) C o m p a riso n b etw e en N L E V M an d k - o w ith (b ) C o m p a riso n b etw e en R S M a n d W ilco x k-tu


P(o E n h a n c e r m o d e l (C u rren t R esu lts) m o d el (E A D S -M A S R esu lts)

Figure 3.26: Contours o f surface pressure coefficient showing effect o f turbulence model on flow solution with
comparison to experiment for a = 23®, M = 0.85 and Re = 6 x 10^

3.6.3 Effect of Solution Convergence


As stated in the previous section, the Non-Linear Eddy Viscosity model was started from the end o f the original k —
0) with Po) Enhancer model solution discussed in the previous sections. Due to this restart, it is important to consider
the effect o f convergence on the flow solutions, particularly in order to strengthen the conclusions made in the
previous section and to determine if the restart would have an effect on the solution. To perform this investigation,
the original calculation was again restarted, using the same turbulence model and run for an additional 4500 implicit
time steps. Figure 3.27 shows the convergence history o f the original and restarted calculations. The residual is
the index o f the error in the numerical computations, therefore by reducing the residual by one, the error reduces
in size by an order o f magnitude. The plot shows the residual for the mean flow computations (lower trace) and
the turbulence model computations (upper trace). It is clear that for the original calculation, the residual reduces
rapidly then begins oscillating in an irregular manner, which dies down, before reaching its final values o f
and 10“ ^^. With the restarted calculation, it is evident that the behaviour o f the residuals becom es more periodic
in nature, however the unsteadiness does not disappear. The residuals are found to oscillate about mean values o f
approximately 10“ ^^ and 10'^ ^ , which are not significantly lower than the final values o f the initial calculation.

briginal New
Calculation Calculation
End End

10'
1000 2000 3000 4000 5000 6000 7000 8000
Iteration

Figure 3.27: Convergence history o f residuals for k a) with Pcj Enhancer model; a = 23®, M = 0.85 and Re =
6 x 10^

Considering the results o f the restarted calculation and the original results shown in the previous sections. Figure
3.28 shows the surface pressure coefficient distributions for both solutions at streamwise stations compared to the
relevant experimental results. As with all other comparisons, it is clear that there is little effect on the flow close
to the apex region at x /c r = 0.2 and 0.4. Downstream at x /c r = 0 .6 , the overall behaviour o f the distribution
is similar with a reduction in suction peak compared to the pre-breakdown experimental data point ( a = 23.6®).
However, the inboard distribution has a lower pressure coefficient distribution and the suction peak is higher for
the restarted calculation results. These results give an improved agreement with the a = 24.6® experimental data.
In the original results, the breakdown location was found to be slightly upstream o f this location, at x /c r = 0.57
and it is clear that breakdown will be close to this region for the restarted results. Downstream, it is evident that the
CHAPTER 3. TRANSONIC VORTICAL FLOW O N A SLENDER DELTA WING 75

overall flow behaviour has not changed significantly with increased convergence with a relatively flat distribution
o f surface pressure coefficient found for both solutions at x /c r = 0.8 and 0.95.

From the contours o f surface pressure coefficient for each solution, shown in Figure 3.29, it is clear that the
breakdown location is further downstream for the restarted calculation solution. From analysis o f the vortex core
behaviour this location was found to correspond to approximately x /c r = 0.6 4 , which is a 7%Cr downstream shift.
As with the original calculation, the breakdown location is downstream o f a normal shock, however only one
shock occurs in this region. It is clear that downstream o f the location o f the normal shock that a suction peak
continues for both solutions, however it appears to last longer for the restarted calculation solution. Upstream o f
breakdown the flow behaviour predicted is almost identical, with the same axial velocity found in the vortex core.
The shockwaves described for the original results are also found for the restarted calculation, as expected, with
the only exception being the second normal shock upstream o f the sting tip, as mentioned. Therefore, it may be
concluded that the most obvious effect o f increasing the calculation run time and thus o f the convergence o f the
solution, is to shift the breakdown location further downstream. This may also suggest that the large difference in
breakdown location between the Non-Linear Eddy Viscosity model and the original results, detailed in the previous
section, is partly due to the effect o f turbulence model and partly due to the effect o f the convergence behaviour at
the end o f the calculation.

M « 0.85, Re * 6e6, 23*


4000 ItetaOons
ro
8500 IteraOoiis
- 0.1
NTF Wind tunnel lesutts, 23.6*
NTF Wind tunnel results 24.6*
- 0.2

- 0 .3

0.4

- 0 .5

0.6

- 0 .7

>- 0.8

- 0 .9

*- 1

Figure 3.28; Effect o f turbulence model on flow solution with comparison to experiment for a = 23°, M = 0.85
and Re = 6 x 10^

" 4000 Herndons

Figure 3.29: Contours o f surface pressure coefficient showing effect o f turbulence model on flow solution with
comparison to experiment for Af = 0 .8 5 , /?e = 6 x 10*^ and a = 23*;
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 76

From further analysis, it was found that despite the difference in breakdown location, the overall flow behaviour o f
the further converged solution was very similar to the original calculation results detailed in the previous sections.
A s this calculation has been shown to have an improved convergence behaviour, the new solution will be used for
the comparisons and analysis in the follow ing sections.

3.6.4 Comparison with Other Structured Grid Results

M « 0.85, Re = 6e6 M ■ 0.85, Re « 6e6

- EADS 18.5' - EADS 23*

- Glasgow 18.5* 0.1 - Glasgow 23*

- NLR 18.5* - NLR 23*

NTF Wind tunnel results, 18.6' 0.2 * NTF Wind tunnel results, 23.6* ^
•» NTF Wind tunnel results 24.6* ''■*
0.3 fr.i

0.4

0.5

0.6

0.7

0.8

0.9 - 0.9

(a ) PiTe-breakdown ca se, a = 18.5" (b ) P o st-b re a k d o w n ca se , a = 23"

Figure 3.30: Comparisons between computational results and experiment for all codes for M = 0.85, Re = 6 x 10^

Comparisons were made with the structured grid results o f E AD S-M A S and NLR as described previously, to fur­
ther consider the validity o f the solutions presented. Referring to the pre-breakdown case first in Figure 3.30(a),
it is clear that the agreement between the computational results and the experimental data is good. A s discussed
in Section 3.4, the current results predict a secondary vortex which is slightly too strong compared to the experi­
mental data. However, it is clear that the EA D S-M A S and NLR solutions predict vortices which are much weaker
and have suction peaks less than the experimental values. These discrepancies may be attributed to differences in
transition treatment, with both E AD S-M A S and NLR running fully turbulent calculations compared to the current
results which has a forced transition from laminar to turbulent flow at x = 0.4. Downstream close to the trailing
edge at x /c r = 0.95, the agreement between each o f the computational solutions lessens. Both the EADS-M AS
and NLR solutions predict the suction peak and sudden increase in pressure further outboard than both the exper­
iment and the current results. This is likely to be due to grid refinement and topology in this region as both the
E AD S-M A S and NLR grids use a conical C-O topology and Glasgow uses an H -H grid, which is more refined
close to the trailing edge. This behaviour is also clear from the surface pressure coefficient distributions o f Figure
3.31. For each solution, the location o f the vortical system is the same with a w ell defined primary and secondary
vortex. The secondary vortex breakdown as described before is evident for both the E AD S-M A S and NLR solu­
tions, however it occurs further downstream for both cases. This supports the suggestion that the secondary vortex
breakdown may be caused by a shock/vortex interaction.

For a = 23", it is clear from Figure 3.30(b) that close to the apex o f the wing i.e. x / c r = 0.2 and 0.4, the agreement
between all the computational solutions and the experimental data is good, with the magnitude and location o f the
primary and secondary peaks being predicted w ell. As before, for the pre-breakdown case, the secondary vortex
is slightly stronger for the current results compared to the EAD S-M A S and NLR solutions. Downstream, vortex
breakdown occurs for all solutions and the flow exhibits good agreement with the experimental post-breakdown
flow. However, it is evident from the surface pressure coefficient contour plots that the location o f vortex breakdown
is different for each solution.
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 77

a » 18.5" a~ 18.5 a«23°

(a) E A D S -M A S (b ) G lasg o w (c) N L R

Figure 3.31: Surface pressure coefficient contours for all codes, M = 0 .8 5 , Re = t x 10^

Figure 3.32 shows the behaviour o f the axial velocity through the vortex core and the location o f vortex breakdown
is clear for each o f the solutions. The locations for vortex breakdown for this case corresponds to approximately
x /c r = 0.68 for EAD S-M A S, x /c r = 0.67 for NLR and x /c r = 0.64 for the current results. Upstream o f the
breakdown location, it is clear that there is som e difference in the predicted maximum axial velocities, caused by
differences in grid resolution or turbulence models used.

Figure 3.32: Axial velocity through primary vortex core for all codes a = 23", M = 0.85, Re = 6 x 10^

As before, consideration was given to the flow on a slice through the vortex core at a constant streamwise location,
x /c r = 0.4, shown in Figure 3.33 for each solution. In each plot, the elongation o f the primary vortex is clear and
the position o f the vortex cores is almost identical. Both secondary and tertiary separation regions occur in the flow
at this location for all solutions. Outboard o f the secondary vortex, the thickening o f the shear layer region is found
in all three solutions, however the strength o f this region appears to be directly linked to the relative strength o f the
secondary vortex. With the strong secondary vortex for the current results producing a fourth vortical region, as
discussed previously.

t!i w ## k# i

(a ) E A D S -M A S (b ) G lasg o w (c) N L R

Figure 3.33: Contours o f x vorticity at x /c r = 0.4 for all results, a = 23", M = 0.85, Re = 6 x 10^
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 78

(a) 18.5" (b ) 23"

Figure 3.34: Pressure coefficient distribution at the symmetry plane on the wing

The locations o f the normal shocks in the flow solutions are also slightly different for each solution. The pressure
coefficient at the symmetry plane is shown in Figure 3.18 for each set o f results. For the pre-breakdown case,
the sting tip shock is evident for all cases at approximately x /c r = 0.6 4 , however the location and strength o f
the rear/terminating shock downstream differs between results. This shock occurs at approximately x /c r = 0.9
in the E AD S-M A S and NLR results and earlier at x / c r = 0.85 in the current results. The differences in strength
and location o f this shocks is likely to be due to the nature o f the grids in this region. At an incidence o f 23*,
the behaviour o f the solutions at the symmetry plane, again, shows the shock at the sting tip at approximately
x / c r = 0.6, but this time it appears that a second shock occurs in the flow slightly upstream o f this location.
However, the compression o f these two shocks appears to merge into one for all solutions. The difference in shock
strength is likely to be caused by variations in grid refinement, particularly in the axial direction, which will cause
varying shock resolutions. Despite the variation o f shock strength, the locations o f these shocks are very similar
with the upstream shock occurring at about x / c r = 0.52 for the NLR results, x / c r = 0 .56 for the EADS-M AS
results and slightly downstream at x / c r = 0.58 for the current results .

M » 0.85, R* » 6 t 6 M > 0.85, Re « 6e6


- Glasgow Grid, 23* - NLR-Glasgow Grid, 23*
- NLR Grid, 23* Glasgow • Glasgow Griil, 23*

- NTF Wind tunnel results, 23.6* “ NTF Wind tuiuiel results, 23.6'
•> NTF Wind tunnel results 24.6* o NTF Wind tunnel results 24.6*

(a ) E ffec t o f g rid to p o lo g y (b ) E ffec t o f co m m o n g rid

Figure 3.35: Effect o f grid on flow solution with comparison to experiment for a = 23*, M = 0.85 and Re = 6 x 10^;
a) Comparison between results from Glasgow and NLR grids (NLR Results); b) Comparison between Glasgow
results and NLR results on comm on grid using similar turbulence model.
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 79

Glaiigow NLR G rid Glasgow

(a ) E ffec t o f g rid to p o lo g y (b ) E ffec t o f c o m m o n g rid

Figure 3.36: Contours o f surface pressure coefficient showing effect o f grid on flow solution with comparison to
experiment for M = 0.85, Re = 6 x 10^ and a = 23*; a) Comparison between results from Glasgow and NLR
grids (NLR Results); b) Comparison between Glasgow and NLR results on com m on grid using similar turbulence
models.

To further aid in the comparisons between each o f the computational solutions consideration was given to the effect
o f grid topology. This was considered by running the same solver and turbulence model on two o f the grids with
differing topologies. These were the fine H-H grid as described in Section 3.2.1 and N LR ’s C -0 grid. It is clear
from Tables 3.1 and 3.2 that the overall sizes o f the grids are quite different. However, this is mostly due to the
topology and chosen far-field definitions and it is found that the number o f grid points over the wing surface is
similar for both the normal and spanwise direction. The results o f this comparison are shown in Figures 3.35 and
3.36. The pressure coefficient distributions show very little difference between the solutions, both upstream and
downstream o f breakdown. Considering the pressure coefficient contours, it is clear that the apparent strength o f
the normal shock and the suction peaks o f the vortical system in the region o f this shock are different. This is most
likely to be due to differences in axial grid refinement rather than the topology o f the grids.

A comparison between the solutions for the Glasgow and NLR CFD solvers on a com m on grid was also performed.
The turbulence models used by these two institutions are similar, with the difference mainly in the specification o f
the turbulence model diffusion coefficients [195]. It is clear that the solutions are very similar. This is also true o f
the surface pressure coefficient contour plots for this case, although a slight difference in the predicted breakdown
location is clear. This is likely to be due to the level o f convergence o f the solutions as a comparison o f the NLR
results with the original calculation described in Sections 3.4 and 3.5, show s no difference in breakdown location.

3.6.5 Influence of Time Accuracy


All the computations described so far have assumed that the flow is steady state. However, it is clear from the
literature discussed in Chapter 1, particularly for the post-breakdown case, that the flow will be highly unsteady.
Therefore, it is necessary to consider the effect o f time accuracy on the solutions and the behaviour o f the flow.
In order to consider this, comparisons were made with an unsteady calculation, for the same transonic conditions,
carried out by the United States Air Force Academy (USAFA) as part o f the VFE-2.

Figure 3.37 shows the comparison o f surface pressure coefficient distributions for the time averaged USAFA so­
lutions and the steady state Glasgow solution. From this plot it is clear that close to the apex, at x /c r — 0.2 and
0.4, the agreement between the time averaged and steady state solutions is good. However, downstream where
the leading edge vortex has broken down, large differences between the flow solutions are found. At x / c r = 0.6,
the time averaged solution shows good agreement with the post-breakdown experimental data inboard close to the
symmetry plane, but outboard o f the primary suction peak large secondary suction peak is evident at this station,
suggesting that the secondary vortex is still present. The steady state solution also displays a small peak in this
region which suggests that a very weak secondary vortex may still occur at this location. Further downstream,
the time accurate result behaves slightly different to the steady state solution and post-breakdown experimental
results, and appears to be slightly closer to the experimental results for the 23.6* experimental data point. Vortex
breakdown can be confirmed to occur in the flow by considering the surface pressure coefficient contours shown
in Figure 3.38.
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 80

M = 0.85, Re * 6e6

- USAFA, Time-averaged, 23'

- Glasgow, Steady Slate, 23*

* NTF Wind tunnel lesulls, 23.6*


= NTF Wind tumiel lesulls 24.6* -2

Figure 3.37: Comparisons between computational results and experiment for current results and USAFA time
accurate solutions for a = 23*, M = 0.85, Re = 6 x 10^

G lasgow USAFA
Steady State Tim e A veraged

Figure 3.38: Surface pressure coefficient contours comparing USAFA time averaged results with steady state
current results, a = 23*, M = 0.85, Re = 6 x 10^

From Figure 3.38, it is evident that the behaviour o f the flow upstream o f vortex breakdown is very similar. How­
ever, the steady state result predicts vortex breakdown slightly further upstream than the time averaged solution,
x /c r = 0.6 4 for the steady state solutions compared to x / c r = 0.68 for the time averaged results. Downstream o f
breakdown the solutions are again similar, however the suction peak which is found downstream o f breakdown
exists for approximately 25%Cr in the time accurate result and only about 15%Cr for the steady state result. This
behaviour is confirmed by considering the axial velocity through the vortex cores for each case, as shown in Figure
3.39. From this plot, it is again clear that the steady state solution predicts breakdown further upstream than the
time accurate solution. However, the levels o f axial velocity upstream o f vortex breakdown are similar. Further
consideration o f this behaviour will be given in the follow ing section.
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING

2
1.5

-0.5 USAFA Tim e Avcniccd


(■Imsow Steady Stete

0 0.1 0.2 0.4 0.7 0.8 0.9

Figure 3.39; Comparisons between computational results and experiment for current steady state results and U S­
AFA time accurate solutions for a = 23®, M = 0.85, Re = 6 x 10^

Considering the three-dimensional behaviour o f the flow shown in Figure 3.40, it is clear that the difference in
vortex breakdown location is clearly associated with the location o f the normal shock at the symmetry plane. For
the time averaged case, the shock at the sting tip appears to interact with the primary vortex shear layer in a similar
manner to the shock in the steady state results. Thus, it may be stated that the mechanism for breakdown is likely
to be the same, but that som e difference between the solutions is changing the location o f the impinging shock.
Further consideration o f this will be given in a later section. Also shown in Figure 3.40 is the presence o f the
cross-flow shocks d~3~| and [~4~|) described in the analysis o f Section 3.5, impinging on the shear layer.

Time Averaged Steady State


USAFA Glasgow

Figure 3.40: Isosurface o f x vorticity coloured by pressure coefficient showing primary vortex shear layer and
normal shock shape for current results and USAFA time accurate solutions; M = 0.85 and Re = 6 x 10^

Therefore, it is evident that the overall agreement between the steady state and time average solutions is reasonable,
with vortex breakdown being predicted over the wing. It is found that the vortex breakdown locations are different,
but despite these differences, the vortex core properties upstream are similar and the shape and relative locations
o f the shocks in the flow correspond w ell. It may be suggested that the effects o f time accuracy on the prediction
o f transonic vortex breakdown are not significant for the purposes o f predicting the main features o f the flow. This
further suggests that the steady state solution can be used as a useful approximation to the complex unsteady flow
behaviour. However, the discrepancies in the location o f breakdown should be kept in mind. This short study also
eliminates the effects o f time accuracy on the critical onset angle for vortex breakdown, as these solutions are also
predicting breakdown to occur early on the wing.

3.7 Shock-Vortex Interaction and Vortex Breakdown


As mentioned in Section 3.5.2 and detailed in Figures 3.19 and 3.40, it is apparent that the sting tip shock intersects
the vortex system. Therefore, it is suggested that som e shock/vortex interaction takes place, particularly for higher
angles o f incidence. To consider this, the pressure in the freestream direction through the vortex cores for both
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 82

angles o f incidence were analysed. This is shown in Figure 3.41, with the calculated pressure ratios for each pro­
posed shock/vortex interaction location marked. For a = 18.5*', the interactions occur without vortex breakdown.
It has been previously suggested that this is due to the shock sitting above the vortex core [17]. However, from
consideration o f the vortex core properties it is found that there are three regions o f adverse pressure gradient which
may suggest direct interactions. These coincide with the two normal shocks at the symmetry plane and the trailing
edge shock, detailed previously, and are clear from the three dimensional view in Figure 3.19. The pressure ratios
for all three are less than 1.5 and, as shown, the primary vortex recovers after passing through each. Therefore, it
may be suggested that these are weak interactions.
1.1
I
0.9
0.8
0.7

is ® ‘
0.5
I 0.4
0.3 il.28l ESI
12.001
0.2
0.1
0
0 0.1 •.2 B.J (.4 0.5 0.4 0.7 0.8 0.9 I

Figure 3.41: Pressure distribution through vortex cores for both angles o f incidence; The numbers on the plot
signify the magnitudes o f the pressure ratios through the intersecting shocks
At a = 23*, where breakdown occurs on the wing, it is clear that there are two regions o f high adverse pressure
gradient at the vortex core. The first coincides with the location o f the normal shock upstream o f the sting tip as
shown at the symmetry plane in Figure 3.18 and also with the onset o f vortex breakdown. Very close to this, the
second, higher pressure gradient coincides with the occurrence o f com plete vortex breakdown, which can be seen
in Figure 3.32. These pressure gradients have ratios o f 2.00 and 2.36 respectively. It is likely that the first pressure
increase is due to the effect o f the normal shock at the symmetry plane on the vortex core, in a similar manner to
the interaction at the lower incidence. The second pressure gradient is much stronger and may indicate a direct
interaction between the downstream section o f the shock and the vortex core and indeed this location corresponds
to the region where the shock intersects the vortex core as demonstrated in Figure 3.40. Further detail o f this region
is found in Figure 3.42, which shows contours o f Mach number on a plane through the vortex core. The vortex
core trajectory that the data o f Figure 3.41 is obtained from is highlighted. This show s the presence o f the shocks
prior to and at vortex breakdown, where the Mach number drops significantly and suddenly.

Figure 3.42; Contours o f Mach number on slice through vortex core at a constant y / s = 0 .56 for a — 23*, M = 0.85
and Re = 6 x 10^
From these results, it is evident that there are interactions between the shocks and vortex core for both angles o f
incidence, with weaker interaction occurring for the lower incidence. Thus, it may be suggested that there is a
limiting behaviour below which the vortex can retard the effects o f the shock and remain coherent. Above this
limit, the interaction causes a considerable weakening o f the vortex core, which results in vortex breakdown. In his
comprehensive review, Deléry [49] demonstrated the importance o f a number o f parameters for vortex breakdown
caused by shock/vortex interaction. These include the tangential or swirl velocity, U q , and the axial velocity o f
the vortex core, Uaxiai- He also proposed that the swirl ratio or the Rossby number may be used as a measure o f
the vortex intensity and, thus, the susceptibility o f the vortex to shock induced breakdown. The Rossby number
is a non-dimensional parameter, defined as the ratio o f the axial and circumferential momentum in a vortex as
defined by Equation 3.1. In this investigation, the maximum axial velocity at the vortex core and the maximum
swirl velocity o f the vortex are used. This relationship is the inverse o f the axial swirl parameter described in Ref.
[49], which is used as a breakdown criterion for a free-vortex.
^ a x ia l ^ a x ia l
Ro = (3 . 1)
rcQ, He
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 83

As a vortex passes through a normal shock, the tangential velocity is found to stay relatively constant while the
axial velocity decreases, therefore reducing the Rossby number [28]. With the reduction in the Rossby number
com es an increase in vortex intensity and, as a result, the susceptibility o f the vortex to breakdown increases. A
criterion for breakdown using the R ossby number has also been investigated by Spall et al. [59] and by Robinson
et al. [60], who applied it to computational results on slender delta w ings and determined that the limiting Rossby
number occurs between 0.9 and 1.4 for most cases, with a stable vortex core occurring for values above 1.4. To
consider this criterion, the Rossby number was calculated for both pre- and post-breakdown angles o f incidence
and the resulting graph is shown in Figure 3.43 with respect to streamwise location on the wing. A lso noted on the
plot are the critical Rossby numbers for vortex breakdown.

2.M
Vortex
2.4

- Ro z 1.4
1.2 le Vortex
- • Ro = 0.9
0.8

0.4 18.5*
23* Vortex B R akdow n
0.1 0 .3 0.4 0.4 0.7 0:8 0.9

Figure 3.43: Rossby number distribution against root chord location for pre- and post-breakdown cases

These results also show the influence o f the shocks on the vortex behaviour. At a = 18.5®, it is clear that weak
interactions occur as the Rossby number decreases. However, this reduction is not significant which shows that
the vortex is not sufficiently weakened by the shock. A recovery is witnessed downstream. At a = 23®, a similar
behaviour is noted where at x / c r = 0.58 the vortex is affected by the normal shock. However, the reduction in
Rossby number is greater than for a = 18.5® and the vortex becom es unstable. Complete vortex breakdown is then
caused by a second shock at approximately x /c r = 0.62 which has a greater effect on the already weakened vortex
flow, and breakdown is almost immediate. It is interesting to note that upstream o f vortex breakdown the value o f
the Rossby number is very similar for the two angles o f incidence. This show s that for a given set o f conditions,
the Rossby number is independent o f incidence. For this case, the mean Rossby number is approximately 1.7. This
suggests that if the Rossby number o f a vortex is constant for increasing incidence, another parameter is needed to
define the limit which causes vortex breakdown to occur on the wing.

T
0.2

2.6 3 .0 3 .5 4 .0
M ach N um ber

Figure 3.44: Theoretical limit curve for normal shock vortex interactions, where t is the swirl ratio = l / R o
(adapted from Ref. [28])

It was also suggested by Deléry [49] that the susceptibility o f a vortex to breakdown is linked to the strength o f
the impinging shock and, thus, on the upstream Mach number. In the study by Kalkhoran and Smart [28], a vortex
breakdown limit for normal shock/vortex interaction based on upstream Mach number and swirl ratio is discussed
for suf>ersonic vortices with uniform Mach number profiles. The resulting limit is shown in Figure 3.44. This
shows that for a given swirl ratio, a limiting Mach number exists above which vortex breakdown occurs. However,
this curve may not be applied to transonic delta wings as the leading edge vortices have jet-like velocity profiles
and the impinging shocks in the flow may not be normal to the vortex axis. This w ill change the behaviour o f the
interactions and, therefore, the limit for breakdown.
CHAPTER 3. TRANSONIC VORTICAL FLOW O N A SLENDER DELTA WING 84

To investigate a limit for transonic delta wing vortices, the strength of the impinging shocks should be considered,
pre- and post-breakdown. Unfortunately, little experimental data exists to allow the shock strength to be measured
through the vortex core. However, the strength of the shocks incident on the surface of the wing may be considered
to improve confidence in the computational solutions. For the NASA NTF experimental results, the pressure
distributions on the surface of the wing at a constant spanwise location of y /s = 0.3 were considered for the 23.6'^
and 24.6^ angles of incidence and are shown in Figure 3.45. Unfortunately, there are only five data points, however,
the presence of an increase in pressure between x/c,- = 0.6 and 0.8 for the 23.6" incidence and x/Cr = 0.4 and 0.6
for the 24.6" incidence is still clear. As the sting tip is located at approximately x/c r ~ 0.64, these pressure jumps
are most likely to be located close to the x/cr = 0.6 streamwise location. Using this as a guide, an approximation
to the shock strength at this location can be determined. The approximate values calculated are given in Table 3.3.

0.95 —_ 23.6"
0.9 — 2 4 .6 ’
0.85
0.8
P h 0.75
0.7
0.65
0.6
0.55
0.5
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.9
x/c^

Figure 3.45: Experimental surface pressure data on conical ray at constant y / j = 0.3 to show experimental shock
strength for a = 23.6" and 24.6", M = 0.85, R e ~ 6 x 10*^ from NASA NTF data

NASA NTF Experiment - 23.6" 1.16


NASA NTF Experiment - 24.6" 1.4673
CFD - 18.5" 1.2314
CFD - 23" 1.4695

Table 3.3: Summary of shock strength on surface conical ray at constant y/^ = 0.3 for all solutions at M 0.85,
Re — 6 x 10^ and a = 23" compared to NASA NTF data.

Using the values in Table 3.3 as a guide, it is evident that there is a considerable difference in the calculated
pressure changes at the sting tip location for the pre- and post-breakdown experimental results. The calculated
pressure ratio for the post-breakdown case is roughly 25% larger than for the pre-breakdown case. Similar distri­
butions were also obtained from the computational solutions for the pre- and post-breakdown cases and the shock
strengths calculated are also stated in Table 3.3. From a comparison with the experimental data it is clear that the
magnitude of the post-breakdown pressure ratio is very similar, however, the pre-breakdown ratio is larger. This
means that overall the increase between the pre- and post-breakdown cases for the computational results is less.
The larger pressure ratio of the computational results for the pre-breakdown case may have implications for the
onset of breakdown. If the shock strength is over-predicted in the computational results, it is likely that breakdown
would occur earlier on the wing compared to the experimental results for a given vortex strength.

To consider the incidence at which vortex breakdown first occurs on the wing and relative strength of the shocks
in the flow, additional calculations were performed for intermediate angles of incidence between 18.5" and 26"
for the same flow conditions as before (M = 0.85 and Re = 6 x 10^). A summary of the important flow details
are shown in Table 3.4. These details include whether vortex breakdown occurred, the maximum vortex core axial
velocity, Mach number and the strengths and locations of the first impinging shock at each incidence. The location
of the shocks can be taken as analogous to the vortex breakdown location, where appropriate. From the analysis,
it was found that the 23" case was the only incidence to exhibit the double shock at vortex breakdown and so
the combined shock strength is instead shown for comparison with the other results. As these calculations were
performed to the same convergence level as the original calculation, the data from the original calculation has also
been included. The further converged solution results are also shown for completeness. As discussed previously, it
was found that the convergence level only affected the location of breakdown, therefore, this should not influence
the critical angle for the onset of breakdown over the wing.
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 85

a VBD? ^WÎflUr. Uaxiai Max. Maxlal ?? Shock x/Cr


18.5" X 1.74 1.76 1.5 0.62
19" X 1.76 1.80 1.67 0.64
20" y 1.74 1.83 3.73 0.64
21" y 1.74 1.86 4.87 0.64
22" V 1.79 1.88 4.67 0.51
23" V 1.80 1.92 5.25 0.55
23" t 1.83 2.00 4.75 0.62
V
24" V 1.84 2.05 5.93 0.49
25" V 1.84 2.10 5.64 0.47
26" V 1.84 2.20 5.48 0.40

Table 3.4: Summary o f shock and vortex core data for all steady state calculations at a = 18.5" - 26", M = 0.85
and Re = 6 X 10^ ^ indicates further converged solution results.
Before considering the onset o f breakdown, it is important to note the behaviour o f the flow variables with increas­
ing incidence. It is clear from Table 3.4, that the predicted shock strength increases with incidence, which is in
agreement with the experimental data in Table 3.3. The axial velocity and Mach number are also found to increase,
however, the Rossby number was found to be consistent at % 1.7 for each incidence as described before. From the
theory o f supersonic flows, it is known that the strength o f a shock is dependent on the upstream Mach number,
thus for a higher axial flow, a stronger shock will occur. However, in this case the relationship does not appear
to be linear. This is most likely to be due to changes in the shape o f the shock in response to changes in the flow
behaviour and the equilibrium conditions as the incidence is increased. This may also suggest that the behaviour
o f the vortex breakdown is also non-linear in nature.

Vortex breakdown first appears on the wing at a = 20", which coincides with a significant increase in shock
strength. At this point it may be assumed that the strength o f the shock is high enough to cause a complete
reorganisation o f the flow behaviour. Thus, the shock strength limit for breakdown for these solutions may be
given as 3.73. This appears to confirm the proposal made previously, that the normal shock strength is over­
predicted, thus causing the breakdown to occur earlier over the wing for the vortex core behaviour predicted. To
determine a link between the vortex flow conditions, as described by the Rossby number, and the shock strength
for breakdown to occur on the wing, further data, both experimental and computational, is needed. By considering
different flow conditions and configurations a trend similar to Figure 3.44 may be determined for transonic vortex
breakdown.

Figure 3,46: Pressure distribution through vortex cores for EA D S and NLR solutions

To further consider the relation between the occurrence o f breakdown, the vortex core behaviour and the predicted
shock strength, the vortex core data for the EAD S-M A S, NLR and time averaged USAFA results are considered
in a similar manner. The pressure behaviour through the vortex core, with the pressure ratios marked, is shown
in Figure 3.46. From this plot, it is clear that a similar behaviour occurs, with shocks intersecting the vortex core
axis and vortex breakdown occurring. From the E AD S-M A S and NLR solutions, the pressure ratios through the
shocks are approximately 1.77 and 1.64, and 1.5 and 2.89, respectively. The USAFA time averaged solution has
only one shock region with a ratio o f 4.5. However, from analysis o f the instantaneous solutions, it was found that
two shocks also exist at breakdown, which for the solution at a time step o f T = 16600 correspond to 2.25 and 2.71.

While the predicted strength o f a shock can be dependent on such factors as grid refinement, turbulence model and
solver treatment, it is also apparent that there are corresponding differences in predicted maximum axial velocity
through the vortex core, as shown in Figures 3.32 and 3.39 and summarised in Table 3.5. The current solution has
predicted a maximum axial velocity which is the same as the USAFA solutions and higher than for the EADS-
M AS and NLR solutions. As a result o f this increase in axial velocity the Mach number upstream o f the shock
CHAPTER 3. TRANSONIC VORTICAL FLOW ON A SLENDER DELTA WING 86

w ill increase, and the upstream pressure will reduce, resulting in a stronger shock to maintain equilibrium o f
the flow. However, it is evident that the Rossby number in each case is similar. This suggests that the shock
strength predicted by the computational solutions is dependent on the vortex core behaviour predicted upstream.
The axial flow behaviour is also dependent on the computational parameters mentioned above. However, despite
the differences in flow solutions and computational set-up, the behaviour and effect o f the shocks on the flow are
the same.

Shock at
Vortex core Shocks y / s = 0.3:
Uaxiai Maxial Ro 1st: ^ 2nd: ^ Total: VBD x /C r
EADS 1.50 - ~ 1.67 1.77 1.64 2.55 1.4274 0.68
Glasgow 1.83 2.00 ~ 1.7 2.00 2.36 4.75 1.4695 0.64
NLR 1.60 - ~ 1.74 1.50 2.89 4.33 1.5075 0.67
USAFA (time ave.) 1.80 2.03 ~ 1.67 - - 4.50 1.4409 0.68
USAFA (instant.) - - - 2.51 2.71 4.75 - 0.66

Table 3.5: Summary o f maximum axial velocity, shock strength and breakdown location for all solutions at a =
23", M = 0.85 and Re = 6 x 10^

To consider the ability o f the computational solutions to predict the axial flow upstream o f breakdown, the PIV
results obtained at DLR and described in Konrath et al. [21] were considered. These experiments were detailed
in Section 1.4 and were carried out for a slightly different flow conditions, with a Mach number o f M = 0 .80
and Reynolds number o f Re = 3 x 10^. To compare with these results, a new set o f calculations were performed,
using th t k — 0) with Enhancer turbulence model for M = 0.80 and Re = 2 x 10^ at angles o f incidence o f
a — 18.5" — 26". Figure 3.47 shows a comparison o f the cross-flow behaviour for a nominal incidence o f a =
26". The effect o f the difference in Reynolds numbers will be negligible due to the sharp leading edge. In the
experiment, it was found that vortex breakdown occurred between the x / c r = 0 .6 and 0.7 streamwise stations.
However, the computations predict breakdown further upstream a tx /cy = 0.4. Therefore, to make a comparison o f
the pre-breakdown flow, the results were compared on planes which were a similar non-dimensional distance from
the breakdown location, this corresponds to x /c r = 0.5 for the experiment and x / c r = 0.3 for the computational
results assumming that the breakdown occurs close to the x /c r = 0.6 location.

a 0.

(a ) PIV, a = 2 5 .9 " , Re = 3 x \ ( f i (b ) C F D a = 2 6 ", Re = 2 x l ( f i

Figure 3.47: Comparison between u velocity contours for experimental PIV and computational results for M = 0.80
on a slice at x /c r = 0.5.

From the comparisons o f the non-dimensional u velocity contours, a number o f observations may be made. It
is clear that the location o f the vortex core is very different between the computational and experimental results,
however this is likely to be due to the proximity o f the computational slice to the apex o f the wing as further
downstream the vortex would lift further from the wing surface. However, the shape o f the vortical system is the
same, with a very elongated primary vortex clear for both sets o f results. Considering the vortex core properties,
from the experimental data at three pre-breakdown PIV planes, it was found that the u velocity corresponds to
1.962 at x / c r = 0.5, 1.870 at x / c r = 0.55 and 1.522 at x /c r = 0.6. Although the maximum velocity found from
the measurement planes is 1.962, it is likely that the actual maximum velocity will be larger. This is evident
from Figure 3.48, which plots these three points along side the velocity behaviour o f the computational results.
The maximum u velocity for the computational results corresponds to « = 1.88, which is slightly lower than
the maximum experimental value. Therefore, it is likely that the axial flow behaviour is under-predicted in the
computational solutions.
CHAPTER 3, TRANSONIC VORTICAL FLOW O N A SLENDER DELTA WING 87

CFD
PIV
1.5:

0.5-

0.5
o!l 0:2 ol3 0:4 0:5 0:6 0:7 0.8 0.9
x/c.

Figure 3.48: u velocity through vortex core for computational results compared to experimental PIV data for
M = 0.80, a = 2 (f

Returning to the issue of the secondary vortex breakdown which occurs for the pre-breakdown case at a = 18.5"
and was initially mention in Section 3.4. The location of this breakdown, is clear from Figure 3.9 and corresponds
to xfcr = 0.86. It is evident from the location and interaction of the shocks in the flow with the primary vortex,
shown in Figure 3.41, that this location coincides with a normal shock at the primary vortex core and with a normal
shock that the symmetry plane. Due to this, it may be suggested that a phenomenon similar to that described above
for the primary vortex breakdown is the cause of the unusual behaviour. From Figure 3.41 it is found that this
shock has a strength of P2/P 1 = 128, and although it has been found that this shock interacts with the primary
vortex, it does not cause vortex breakdown. However, the secondary vortex does not have as high an axial velocity
and therefore strength as the primary vortex. Therefore, if a shock/vortex interaction occurs, it is likely that the
secondary vortex cannot recover downstream and vortex breakdown occurs.

3.8 Discussion
Having considered the mechanisms which cause vortex breakdown to occur on the wing, it is possible to return
to the issue of the discrepancies between the CFD and experimental results. It was found from the experimental
data used in this study that vortex breakdown jumps abruptly from a location downstream of the trailing edge to a
location upstream on the wing for a small increases in incidence. Indeed from the results summarised in Table 3.4,
it is clear that the flow seems to go from full vortical flow over the whole wing surface to breakdown occurring
close to the xfcr ==0.6 location in a one degree increase.
1.2:
Downslicam of Trailing Edge
i-j
0.8:

0.4'
Glasgow CFl) M = 0.85 Ke = 6e6
EADS CFD M = 0.85 Ke = 6efi Approximate values as
0.2 <
NASA NTF M = 0.85 Re = 6e« * * taken from surface Cp
DI.R M = 0.8 Re = 2e6 distributions
0 r -r I T..T .-n —r T - | . . - r i " i I I I I I I I " I f 'T - I I I 1
19 20 21 23 24 25 26

Figure 3.49: Vortex breakdown location for both computational and experimental results

The location of vortex breakdown with incidence is plotted in Figure 3.49 which also shows similar results for
EADS-MAS solutions and comparisons to available experimental data. For the experimental data, the exact loca­
tion of vortex breakdown is not known, however from the surface pressure coefficient distributions the approximate
locations could be determined. From this plot it is clear that the behaviour of the onset of vortex breakdown is very
similar for both the CFD and experiment, however the angle at which this occurs varies. With further consideration
of the literature it was found that there is a large spread of values for this critical angle. These are detailed in Table
3.6 below. It is quite clear from all these results that the critical onset angles for vortex breakdown over the wings
for current CFD solutions are consistently earlier than for the majority of the experimental results.
CHAPTER 3. TRANSONIC VORTICAL FLOW O N A SLENDER DELTA WING

Source Type Conditions Olcr


Elsenaar and Hoeij makers [18] exp. M = 0.85, Re = 9 X 10^ 23"
Houtmann and Bannink [129] exp. M = 0.85, Re = 3.6 X 10^ 20"
Chu and Luckring [20] exp. M = 0.799,Re = 6 x 10® 26.6"
” exp. M = 0.831,Re = 6 x 10® 24.6"
” exp. M = 0.851,Re = 6 x 10® 24.6"
” exp. M = 0.871,Re = 6 x 10® 24.7"
” exp. M = 0.9, Re = 6 X 10® 22.6"
” exp. M = 0.849, Re = 11.6 X 10® 24"
Longo [121] CFD M — 0.8, Inviscid 25"
Glasgow CFD M = 0.85 Re = 6 X 10® 20"
EADS-MAS CFD M = 0.85 Re = 6 X 10® 21"

Table 3.6: Critical incidence for transonic vortex breakdown to be found on 65" delta wings

To explain this difference, further consideration is needed to the discussion given above considering a critical limit
for breakdown to occur dependent on the vortex core strength and the strength and locations of the shockwaves in
the flow. As shown, with an increase in incidence the strength of the shocks in the flow increases, most likely as a
response to the increased acceleration of the flow over the wing suiface. Similarly, the axial velocity in the vortex
core increases and it has been shown that there is a critical relationship between these quantities which results in
breakdown for a critical incidence. To change the angle at which vortex breakdown occurs, it will be necessary to
have a change in either one of these parameters. For example, with an increase in vortex intensity and therefore a
decrease in axial velocity or an increase in tangential velocity, the strength of the shock needed to cause breakdown
will decrease and breakdown will occur earlier on the wing.

From the results detailed in the previous section, it may be suggested that two factors are causing the early predic­
tion of breakdown on the wing. These are an under-prediction of the axial velocity, which results in a vortex more
susceptible to breakdown and an over-prediction of the strength of the shocks within the flow. From consideration
of the effects of a number of flow parameters, it appears that these predictions are not greatly effected by grid struc­
ture, turbulence model, convergence or time accuracy. The effect of grid refinement was also considered, which
also concluded that the overall refinement of the grid had little effect on the solution. However, this study did
not consider localised refinement, particulaiiy in the vortex core region. Despite continuing improvement in CFD
codes, turbulence models and practises, prediction of the vortex core behaviour and axial flow is still a challenge.
There have been a number of collaborations and investigations which have considered the vortical flows over delta
wings, which have also generally predicted the flow behaviour well, however the axial velocity is almost always
much lower than that found from experiments. This is also true for this case and may be attributed to the abilities
of turbulence modelling and restrictions in grid refinement for the core region. To fully resolve the vortex core
behaviour it would be necessary to have similar refinement as is applied to boundary layer regions. It is unclear at
this time whether an improvement in vortex core axial velocity would alter the predicted strength of the shocks in
the flow, however, if the shock strength remained constant, with an increase in axial velocity, it may be suggested
that the angle of incidence at which breakdown occurred would increase.

3.9 Conclusions
The behaviour of transonic delta wing flows and the ability of CFD to predict these flows was considered in this
chapter. To consider this, two angles of incidence were used which corresponded to solutions which predicted pre-
and post-breakdown flows. The initial analysis showed that the CFD solutions predicted the behaviour over the
wing very well for the pre-breakdown flow, however the high incidence showed a discrepancy with the experimen­
tal results. Where the experiment was exhibiting a full vortex to the trailing edge, the CFD solution was predicting
breakdown to occur. However, it was found that breakdown occurred on the wing for the next experimental data
point and from comparison of the CFD results with this data, it was found that the CFD solutions gave good agree­
ment. Therefore, it was concluded that the flow behaviour was predicted well but that the critical incidence for
breakdown was not well predicted.

A number of transonic flow features were determined from analysis of the solutions, particularly the occurrence
of a complex cross-flow shock system and the abrupt behaviour of vortex breakdown. However, more experimen­
tal data, particularly considering the off-surface flow behaviour, is needed to both confirm the existence of these
CHAPTER 3. TRANSONIC VORTICAL FLOW O N A SLENDER DELTA WING 89

shocks and to further validate the flow solutions.

A thorough sensitivity study was carried out to determine the effect of a number of computational factors on the
flow behaviour. These factors included turbulence model and time accuracy. However, it was found that although
the flow was affected by these factors, the influence was small and there was no effect on the onset of breakdown
on the wing.

The mechanisms which determine the behaviour of transonic vortex breakdown were shown to be highly complex
and are dependent on the vortex core strength and the strength and location of the shocks in the flow. Through
consideration of computational solutions, a means to analyse the influence of each of these parameters was estab­
lished and it has been shown that a relationship must exist, which describes the critical limit for vortex breakdown
to occur. Further research is needed, both experimental and computational, to confirm the behaviour of this rela­
tionship and to allow for fui'ther analysis of the critical limit of shock/vortex interactions for delta wing flows.

It was concluded from the discussion of the shock/vortex interaction and the presence of a limit for breakdown
that further work is needed to consider the prediction of the vortex core axial velocity and shock strengths in order
to accurately capture the onset of the breakdown behaviour in comparison to the experimental data. However, the V
predictions of the flow behaviour were found to be otherwise adequate. 1
t
j
Ï
Chapter 4

Application of DES to Delta Wing Vortical


Flows

4.1 Introduction
The flow over a delta wing is dominated by the leading edge vortices. As the angle of incidence is increased the
vortices become unstable and vortex breakdown can occur over the wing. This flow is found to be highly unsteady.
For aeroelastic behaviour, such as buffet, of existing configurations, it is clear that understanding the behaviour of
unsteady forcing is crucial to allow the alleviation of any structural response, which may exist. This is potentially ^
important for complex fighter configurations such as the EuroFighter and is compounded by the emergence of new Î
UAV and UCAV technologies, which are tending toward planforras where unsteady vortical flows play a large -f
role. This means that the need for a more complete understanding of the unsteady behaviour of vortical flows is
becoming increasingly important.

To date, there has already been a great deal of research, which has considered the unsteady behaviour of this flow
and what is generally known was discussed in a summary of the literature given in Chapter 1. From tliis research,
it is clear that the unsteady behaviour of the vortical flow is complex, with a large number of flow phenomenon
existing and interacting, over and downstream of the wing. These flow phenomenon include the helical mode
instability of vortex breakdown, vortex wandering, vortex breakdown oscillations and shear layer instabilities.
From consideration of the literature available, the frequencies associated with these phenomenon were considered
and summarised in Table 1.2. From this condensation of the available data, patterns emerge relating the order and
size of the non-dimensional frequencies for these flow features. This is summarised further in Table 4.1.

Phenomenon Strouhal Number I


Helical Mode Instability 1 -2
Shear Layer Instabilities 8 -1 0 and higher frequencies -ij:
V
Vortex Shedding - T.E ^ 8
Vortex Shedding - high a 0.2 -0 .5
Vortex Breakdown Oscillation 0.01 - 0.08

Table 4.1: Frequencies coixesponding to important unsteady features of vortical flows

It was found that other dominant frequencies also featured in the literature, which were not clearly attributed to
specific phenomenon. These are, St = 2.5 - 4, 5 —6 and the higher frequencies ~ 20. It is possible that these
frequencies also correspond to the phenomenon detailed above, however further investigation is needed. It is also
important to note, that there may be more than one dominant frequency associated with a particular phenomenon,
due to the complexity of the unsteady behaviour. For example, shear layer instabilities will have at least two
associated frequencies, this is due to the rolling up of tlie shear layer into discrete subvortices, which will have a
frequency of rotation and also due to tlie movement of these structures around the vortex core. It may be difficult to
separate these frequencies within a single solution, however, it may help to explain the spread of data and dominant
frequencies assigned to particular flow features.

To allow for further understanding, these phenomena can be split into two categories, those which occur upstream
of breakdown and those occurring downstream. This is shown in Figure 4.1. Splitting the flow features in this
way allows for an appreciation of which features will dominate, depending on where vortex breakdown occurs

90
1
t

CHAPTER 4. APPLICATION CE D E S TO DELTA WING VORTICAL FLOWS 91

on the wing. For breakdown close to the trailing edge, it is likely that the shear layer attachment and shear layer
instabilities would dominant the flow, however as the breakdown moves upstream, it is likely that the helical mode
instability may dominate the frequency content. This will be important when considering the frequency content of
the results at specific regions in the flow and looking at the flow behaviour overall - particularly when considering
the unsteady loading on the wing.

Shear Layer
Instabilities V o rte x B r e a k d o w n
L o c a tio n O s d lia tio n

Vortex Core
Rotation H e lic a l M o d e
In sta b ility

V ortex Shedding

(a) Upstream (b ) D o w n s tre a m

Figure 4.1; Schematic diagrams showing flow topology upstream and downstream of vortex breakdown

Accurately predicting tliis complex flow is a challenge for numerical methods. In recent times the capabilities
of CFD solvers have improved, with more complex turbulence modelling and treatments being utilised. One such
method is DES, which is a hybrid URANS/LES turbulence treatment. This model was proposed initially by Spalart
[171] to reduce the fine resolution of the grid in the boundary layer region needed for high Reynolds number LES
calculations and is described in detail in Chapter 2. With this treatment many of the smaller turbulent scales can
be captured, which has led to a greater ability to predict more and more complex flow behaviour accurately. This
has been shown from existing DES calculations on delta wing geometries [30, 161, 164] using unstructured grids,
mostly carried out by the United States Air Force Academy (USAFA). Therefore, the use of DES for this type of
flow using structured grids will be considered.

This chapter considers the results of an investigation applying DES to the unsteady vortical flow over a slender
delta wing. The test case will be outlined, with the computational set up and grid generation discussed. The effect
of the temporal and spatial refinement on tlie DES results will be considered before the application of DES and
the resolution of the LES region is analysed. The final results will then be compared to existing DES results and
validated against experiment before the results are discussed and conclusions made.

4.2 Summary of Test Case


The test case chosen for this investigation is a 70^ delta wing at an incidence of a = 27°. Vortex breakdown occurs
over the wing and there is an extensive database of experimental data, both time-averaged and unsteady for valida­
tion purposes. There is also a considerable database of computational results available for this configuration using
both UR ANS and DES turbulence models [23, 24, 29, 30, 70, 71, 116, 144, 161, 165, 187, 196] from the NATO
RTO Task group AVT-080 which considered "Vortex Breakdown over Slender Wings" [197]. The experimental
data is taken from the PhD thesis by Mitchell [13] and associated papers [81, 99, 100,162,197]. The experiments
were carried out in ONERA's F2 and S2Ch subsonic wind tunnels with a wide range of experimental techniques
used to elucidate the flow features and create a large database of experimental results. These techniques include:
3D Laser Doppler Aneraometry (LDY), Particle Image Velocimetry (PIV), Laser tomoscopy, Suiface oil flow vi­
sualisation and data from both steady and unsteady pressure ti’ansducers (Kulites™).

The wing used in the experiments has a root chord length of 950mm and a sweep angle of 70°. It has flat upper and
lower surfaces with a 15° bevel at the leading edge. The trailing edge is blunt with a thickness of 20mm. These
details are shown in Figure 4.2. The experimental test conditions used by Mitchell were: an incidence of cc = 27°,
f/ca = 24/n^“ ^ which conesponds to a Mach number of M = 0.069 and a Reynolds number based on the root chord
of Re = 1.56 X 10^. To help with the convergence of the compressible flow solver, tlie Mach number used for the
investigation was raised to M = 0.2, which gives a free-stream velocity of C/™= 6Sms~^. As this Mach number
is still relatively low, this should not have a significant effect on the solution as compressibility effects will be
CHAPTER 4. APPLICATION CE D ES TO DELTA WING VORTICAL FLOWS 92

negligible. All non-dimensional fi'equency data calculated for the computational results in this investigation will
use this altered free-stream velocity, to allow fair comparison with the experimental data.

fixation

691.5

Figure 4.2: 70" ONERA geometry (all distances marked are in mm) [13]

This test case is also used for the investigation into the performance of URANS turbulence models for unsteady
vortical flow prediction in the following chapter and as such the grids generated for the DES calculations were also
used for this work. All grids used and the computational set up of the calculations for both investigations will be
detailed in this chapter and will not be repeated in later sections.

4.2.1 Grid Generation


As stated in Chapter 3, there are a number of factors, which are important in the creation of grids for use in CFD
calculations. When creating a grid for use in a DES simulation, these factors are compounded by the sensitivity
of the solution to both the spatial and temporal resolution of the calculations. As stated previously, the solution
of the DES model is highly dependent on the maximum cell dimension, A,„fix = max(Ax,Ay,Az,), through its use
within the LES region as a spatial filter. For a URANS calculation, the aim of grid refinement is numerical accu­
racy, however, for LES the refinement of the grid determines the level of the sub-grid scale model, which in turn
determines the smallest resolved eddies in the flow. This means that for the LES region, the maximum cell size
determines the range of scales which are subject to modelling rather than prediction by the conservation laws. This
value also determines the size of the URALfS region close to the wall. For an optimum DES calculation, the grid
cells should be orthogonal, ideally cubic, particularly within the regions of interest and at the interface between
the URANS and LES zones [177]. Attempting to create a sufficiently refined structured grid with uniform cubic
cells in all regions of interest is impractical for delta wing geometries as the required size of the grid would be too
computationally expensive to run. However, achieving orthogonality is not difficult for this type of grid and only
requires a suitable grid topology to reduce the presence of cell skewness, and allow an even distribution of points
within the regions of interest. As the physical accuracy of the DES model is dependent on small cell sizes, it would
seem prudent to refine the grid as much as is practical.

With an increase in spatial refinement comes a need for temporal refinement. This further increases the com­
putational expense, which can be prohibitive to allowing the computation of a fully optimised solution. Further
discussion of temporal dependency will be given in a later section. Grid refinement in the spanwise direction at
the leading edge is still important for the URANS zone to allow the accurate prediction of the flow separation and
shear layer region within the boundary layer. As these grids are to be used for both DES and URANS calculations,
it is important to consider the needs of each type of turbulence treatment in the creation of the grids.

To create the structured multi-block grid, the ICEMCFD mesh generation package, Hexa was used. The trailing
edge wing geometry was altered to include a 15" bevel, similar to the leading edges. A semi-span H-H grid
topology with no sting anangement was used, which sets the incidence of the wing to 27" in the grid. The grid
also uses a “collapsed apex” blocking strategy, where the edges of the blocks in the wing apex region have been
collapsed to create a singular point. An example of the blocking topology is shown in Figure 4.3. Convergence
problems associated with the singularity were again, dealt with by using laminar flow at the apex and fixing
transition to turbulence at a constant streamwise location in the grid, which will be discussed in a later section.
As for the VFE-2 grids, the far field was defined 20c,- in each direction from the wing apex to minimise the effect
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 93

o f the boundaries on the flow. This grid topology has been used successfully in a number o f investigations using
RANS turbulence models by Allan [144, 198, 199].

Figure 4.3: Grid topology o f H-H grids used in investigation

Using one o f the grids created by Allan as a starting point, two grids were created with differing levels o f refine­
ment, coarse and fine. Both grids have a first wall spacing o f 1 x lO'^c^, which corresponds to a value o f
approximately 0.1 and a stretching ratio, within the boundary layer region, o f 1.2. The value is sufficient for the
Reynolds number used in this investigation and the stretching ratio is within the recommended range for adequate
log-layer resolution suggested for RANS calculations by Spalart [177]. Exam ples o f the grid refinement in a plane
perpendicular to the wing surface and the freestream flow direction at a location x /c r = 0.63 are shown in Figure
4.4. This shows the relative refinement, particularly close to the leading edge and in the boundary layer region. The
fine grid has a higher concentration o f points in these regions than the coarse grid and also has a much improved
orthogonality over the whole area o f interest close to the wing. This is demonstrated further by Figure 4.5 which
shows the grid distribution on the symmetry plane.

A third grid was also created for the DES calculations, to consider the effect o f refinement in the trailing edge
region on the upstream vortical flow. This grid was based on the fine grid with the same distribution o f points over
the wing. However, in the region downstream o f the trailing edge more grid points were added and the stretching
ratio was decreased to improve the refinement in this region. Figure 4.6 shows the grid refinement at the trailing
edge for the two grids. The effect o f this refinement on the DES results w ill be discussed in a later section. A
summary o f the main features, including the number o f grid points over the w ing in each direction, for each grid is
given in Table 4.2.

No. o f G rid Points on Wing


Type G rid Size Stream w ise Spanw ise N orm al
Coarse 3 969,310 102 80 89
Fine 7,767,081 167 112 107
Fine - Refined TE 8,768,970 167 112 107

Table 4.2: Summary o f main features o f grids used in DES and U R A N S investigations

As this study considers DES calculations, it is important to be able to consider the active LES and U RA NS areas
within the grid structure. As part o f the DES formulation two important grid parameters are calculated, the distance
from the wall, dmin and the maximum cell length CoEsAmax at each cell location. The relationship between these
variables in the DES implementation was used to create a flag parameter, which demonstrates the distribution of
the two turbulence treatments. Where dmin > CoEsAmax, the flag is set to 1.0 and LES is active in that region and
similarly where dmin < CoEsAmax, it is set to 0.0, and the U RA NS model is active. Based on the explanation o f the
DES model given in Section 2.4.5, it is expected that the RANS model is only active within the boundary layer
region, close to the wing surface i.e. where the value o f d„,in is less than CoEsA,nax- Figure 4.7 shows contours
o f the flag parameter on a slice at x /c r = 0.63 and through the vortex core region at a constant y / s = 0.7 for the
fine grid. These show two contour regions denoted by red and green which signify the LES and URANS regions,
respectively. It is clear from these plots that the region in which U RA NS is active is very small, close to the wing
surface. Thus, LES is active for the majority o f the region where vortical flow occurs.
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 94

y/c, y/c,

(a) C o arse g rid (b ) F in e g rid

Figure 4.4: Comparison o f grid refinement at x /c r = 0.63 plane

(a) C o a rse g rid (b ) F in e g rid

Figure 4.5: Comparison o f overall grid refinement at symmetry plane for coarse and fine grids

(a) F ine g rid (b ) F in e g rid w ith refin ed tra ilin g ed g e

Figure 4.6: Comparison o f grid refinement in trailing edge region for fine grid and refined TE grids
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 95

(a) x /cr = 0.63 (b) y / s = 0.7

Figure 4.7: DES active area for fine grid; red denotes LES region and green show s URANS region.

4.2.2 Transition IVeatment


The location o f the transition between laminar and turbulent flow must be specified for these calculations. Ideally,
this would be a location which would correspond to the natural transition line or to the line o f a forced transition
in the experiment. In the computational investigation carried out by Morton et al. [30], time-accurate calculations
were performed for this test case using the SA -DES model. The flow conditions were as described above and the
investigation was carried out on a series o f unstructured meshes with varying refinement. A grid o f approximately
2.4 X 10^ grid points, locally refined in the region o f the vortex core, was initially used and the flow was fully
turbulent over the wing at the start o f the investigation. Two transition locations were considered and compared to
the experimental surface flow visualisations from M itchell’s investigation [13]. In the experiment, Mitchell noted
that there was an inflection o f the secondary separation line at approximately x /c r = 0.4 , which would suggest the
transition from laminar to turbulent flow at this point. However, no details o f transition inboard o f this location
were given. From Morton’s study, it was determined that the location o f transition had quite a significant effect
on the flow behaviour. However, from comparison with the experimental results it was concluded that a constant
transition line situated at x /c r = 0.4 gave the most accurate results.

Based on this investigation and the experimental results, the location o f a transition line was set to x = 0 .4 for
all calculations. This value o f transition corresponds to a value o f x /c r — 0 .3 5 9 1 4 on the wing upper surface.
It was felt that this slight upstream shift o f transition compared to the D ES results o f Morton is not likely to
affect the validity solutions significantly as the actual experimental transition is still largely unknown. N o further
investigation into transition will be given in this study.

4.2.3 Probe Application


In M itchell’s investigation a series o f 17 Kulite™ unsteady pressure transducers was used to consider the unsteady
behaviour o f the flow. These sensors were situated at the same chordwise stations as the time averaged data was
obtained, at constant non-dimensional spanwise locations. Details o f the locations o f each probe, and its number,
are given in Table 4.3. The unsteady data was obtained from 10"^ samples taken at a frequency o f 5kH z over 2
seconds, which corresponds to a non-dimensional sample rate o f AT = 0.005 1 .

Probe Location Probe Location Probe L ocation P robe Location


xjCr y js xjCr y js xjCr y is x jc r yIs
I. 0.84 0.7 5. 0.74 0.75 10. 0.63 0.7 14. 0.53 0.7
2. 0.84 0.65 6. 0.74 0.7 II. 0.63 0.65 15. 0.53 0.65
3. 0.84 0.6 7. 0.74 0.65 12. 0.63 0.6 16. 0.53 0.6
4. 0.84 0.5 8. 0.74 0.6 13. 0.63 0.5 17. 0.53 0.5
9. 0.74 0.5

Table 4.3: Experimental unsteady pressure probe locations

To compare with these locations and to consider the predicted unsteady behaviour, point probes were applied to
the computational domain for these calculations. At the outset o f the investigation, it was not known where the
regions o f interest lay for this case and so a large number o f probes were applied, both upstream and downstream
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 96

o f vortex breakdown. Probes were placed on constant chordwise locations: x /c r — 0 .1 , 0.2, 0 .4 ,0 .5 3 ,0 .6 3 ,0 .7 4 ,


0.84 and 1.0. Additional probes were also added, at a later stage, to consider the behaviour o f the flow downstream
o f the trailing edge at x /c r = 1.1, 1.2 and 1.3 In the spanwise direction, the probes were placed at y / s values
o f 0 - 1.1 at 0.1 intervals and normal to the wing the points were situated at z /c r = 0, 0.001,0.005, 0.01, 0.02,
0.04, 0.08, 0.1, 0.15, 0.2 and 0.4. The spanwise and streamwise locations o f the probes were chosen to coincide
with the positions used in M itchell’s experiments [13] described above. In total, 1496 probes were placed for this
investigation. Schematics showing the majority o f the probe positions are given in Figure 4.8.

(a) S id e view (b ) P lan view

Figure 4.8: Positions o f probes for 10° ONERA wing

The resulting signals from all the probes are collected in a series o f files which are processed using a custom-made
program. Probe A nalyser created using Matlab. Details o f this process were discussed in Chapter 2 and the theory
behind the analyses are explained in Appendix B.

As discussed, there are two important numerical parameters in performing good DBS calculations, the grid refine­
ment and the time step size. Each will be considered separately before the validity and applicability o f DES to
delta wing vortical flows w ill be considered.

4.3 Effect of Time Step Refinement


The effect o f temporal refinement is initially considered using the fine grid for three non-dimensional time steps
with increasing refinement. These are A T = 0.01, 0.005 and 0.0025. Each calculation was run for a similar total
time, which meant that the number o f time steps used for the calculation increased as the time step was decreased.
From the calculations, a number o f instantaneous domain files and the time histories o f the probe files were ob­
tained. These files will allow a comparison o f the flow behaviour for each time step used as described in Section 2.7.

As mentioned in the previous section, a large number o f probe locations were specified within the flow domain
above the wing surface. To simplify and reduce the amount o f data analysed, five probes were considered through
the vortex core region. These probes sit at a constant height from the wing surface, z /c r = 0.1, on the plane at
a constant y / s station o f 0.7. The five probe locations chosen occur at chordwise stations o f x /c r = 0.53, 0.63,
0.74,0.84 and 1.00. Figure 4.9 shows these probes with instantaneous isosurfaces o f x vorticity for the AT = 0.01
solution as an example. This shows the behaviour o f the vortex relative to the probe locations. The relative
locations o f the vortex and the probes are similar for the other two solutions.

Figure 4.9: Isosurfaces o f x vorticity coloured by pressure coefficient showing instantaneous vortex core behaviour
at T = 50 with core probe locations marked for A t = 0.01

Initially, the mean location o f vortex breakdown was obtained for each solution by considering the flow domain
output files at every full non-dimensional time period, T = 1 , 2 . . . etc. This effectively allowed a sample o f the flow
behaviour at every 1 0 0 ,2 0 0 and 400 time steps for the A t = 0 .0 1 ,0 .0 0 5 and 0.0025 calculations, respectively. This
provides a low sample rate, however for the purpose o f calculating the mean location this was deemed sufficient.
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 97

From this analysis it was found that the mean breakdown location differed slightly for each case, occurring at
x /c r = 0.88 for A t = 0.0 1 , jc/cr = 0 .84 for A t = 0.005 and x /c r = 0.85 for A t = 0.00 2 5 . Initially, it was proposed
that these differences may be due to the sample rate o f the analysis as it was not expected that the time step should
have a significant effect on the mean flow behaviour.

To consider this further, the mean and RMS velocity components were calculated for each o f the probe locations.
The results o f this analysis are shown in Figure 4.10. It is evident from this figure that changing the time step does
indeed have an effect on the predicted flow. However, with the exception o f the w velocity, it is found that this
mostly occurs in the region close to breakdown. For the u and v velocities, at x /c r = 0.53 and 0.63 the predicted
values are all in very good agreement and do not appear to be affected by temporal resolution. As the region o f
breakdown is approached, the difference between the coarse time step. A t = 0.01 and the other solutions increases,
however there is little difference between the finer time step solutions. The behaviour o f the w velocity is different,
with a larger difference between all the solutions being clear. For all locations there is a decrease in the velocity in
the normal direction for the A t = 0.0025 solution compared to the other time steps. This is particularly interesting
as the agreement between the A t = 0.005 and A t = 0.0025 solutions is good for the other velocity components.
However, this may be due to an inboard shift o f the vortex core for this case, particularly as the RMS velocity
values are similar. Closer to the breakdown region at x /c r = 0.84, the differences in the mean flow increases for
all velocity components.

Considering the mean u velocities with respect to the mean location o f breakdown, mentioned above and it is clear
that the breakdown location for the A t = 0.005 and A t = 0.0025 solutions may in fact be the same. However, the
breakdown location for the coarse time step solution is clearly further downstream. Therefore, it is clear that the
temporal refinement does have an effect on the location o f the vortex breakdown in the flow. It is possible that this
is due to the relative resolution and inter dependency o f the time step and grid refinement discussed previously.
Particularly, as the RMS velocities in this region are very similar for each o f the solutions. However, it is evident
that the mean behaviour o f the A t = 0.005 and A t = 0.0025 is very similar and therefore, it may be suggested that
convergence o f the time step has occurred.

C hordw lM I^ocslkm (»/e^)

(a) u velocity

rt

(b) V velocity

(c) w velocity

Figure 4.10: Mean and RMS behaviour o f non-dimensional velocity components at five probe locations through
vortex core region for each solution o f the time step study

Figure 4.11 shows a slice through the vortex core region ^^.y/s = 0.7 with instantaneous contours o f y vorticity
for each o f the solutions. From these plots it is clear that despite some differences in the mean flow behaviour the
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 98

instantaneous behaviour is similar. The location o f the vortex breakdown is close and the winding o f the vortex
core in the breakdown region is comparable. To compare the unsteady behaviour, a single probe above the trailing
edge on the vortex core plane is considered for each solution. The time history o f each velocity component was
considered and a PSD analysis o f the signal was performed. The results o f this analysis are shown in Figure 4.12.
At this location it is found from the time trace that the amplitude o f the signals are very similar. However, it
is clear that the signal behaviour is different, particularly the signal from the A t = 0.01 solution which clearly
exhibits a lower frequency oscillation that the other two signals. From the PSD analysis o f these results, the
dominant frequencies are determined and again the differences between the coarse time step and the other results
is striking. Both the A t = 0.005 and 0.0025 results show the dominant frequency to occur at approximately
St = 3.25. However, for A t = 0.01 this is lower at approximately St = 2.25. Similarly for the lower dominant
frequency in the signals, the signal for A t = 0.01 exhibits a frequency lower than the peak at St = 0.07 found for
the other two results. The agreement between the A t = 0.005 and 0.0025 unsteady behaviour is, again, very good,
further suggesting that the time step has converged.

(.7 <L| «.* I l.l 1.1 W 1.4 4.7 t.t «.« I l.l 1.1 1.1 1.4 4.7 •.* I l.l 1.1 LI 1.4

(a) At = 0.01 (b) At = 0.005 (c) At = 0.0025

Figure 4.11: Instantaneous contours o f y vorticity on a slice through the vortex core region at t = 50

/MfcifaiBiiifhinAn rtv I
Nm DiMMWiMMl riOM

(a) u velocity

Nm IMmfwrfawl Tim#
Jla, S t r w M mumb»r

(b) Vvelocity

N## Time

(c) w velocity

Figure 4.12: Unsteady behaviour o f non-dimensional velocity components shown by time histories and PSD fre­
quency plots for a probe at x /c r = 1.00, y / s = 0.7 and z /c r = 0.1; fine grid solutions for time steps o f A t = 0.01,
0.005 and 0.0025
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 99

To consider the appropriate time step for use in the DES calculations for this grid, the time step guide recommended
by Spalart [177] was used. This uses the nominal grid size in the region o f interest to define a guideline time step
for a particular grid and was discussed in Section 2.6 and defined by Equation 2.91. From Figure 4.13, which
shows the values o f Amax on a plane at the trailing edge on the fine grid, it is clear that the nominal value in the
region o f interest is approximately 0.0055. Assuming that Umax — 2.5, the guideline time step can be calculated as
A t = 0.0022. Based on this analysis and the results o f the comparisons o f the unsteady behaviour detailed in this
section it is concluded that the most suitable time step for use with this grid is A t = 0.0025. From consideration
o f the time step calculations in Section 2.6, this should correspond to a maximum non-dimensional frequency o f
approximately St = 40.

0.1 y/c, 0.3

Figure 4.13: Amax o f fine grid on slice at trailing edge, x /c r = 1 . 0 0

4.4 Effect of Grid Refinement


From the fine grid results detailed for the time step study, it was noted that the flow behaviour and small scale flow
structures appeared to dissipated too soon beyond the trailing edge. It was supposed that this was likely to be due to
the large stretching ratio o f the grid points in this region. To investigate this, a grid was created which had greater
refinement in this region as shown in Figure 4.6 and described in Section 4.2.1. The grid was created from the fine
grid and as such the maximum cell dim ensions upstream o f the trailing edge is the same for both grids as shown
in Figure 4.14. However, for the new grid, this refined region was extended further downstream by improved grid
refinement in the streamwise direction. This grid w ill be referred to as the “Refined TE Grid”. A calculation was
performed for this new grid using a non-dimensional time step o f A T = 0.0025 and the results were compared to
the fine grid results, with the same time step, detailed in the previous section.

1.1 l.l 1.1 1.4 «.? i.l «* I w, l .l l.l 1.1 1.4

(a ) F in e G rid (b ) F in e G rid - R efin ed T E

Figure 4.14: Contours o f Amax on a slice through the vortex core for both grids used in grid refinement study

A similar analysis to that conducted for the time step study was performed to consider the effect o f the grid
refinement at the trailing edge. However, a further three probes were considered downstream o f the trailing edge to
determine any changes in the flow at this location. These probes sit along the same plane as the probes described
previously at streamwise stations o f x /c r = 1.1, 1.2 and 1.3 and were analysed in the same way. The location o f
the probes in the trailing edge region is shown in Figure 4.15, which shows instantaneous contours o f y vorticity
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 100

for each grid through the vortex core region at the trailing edge. This figure highlights the relative behaviour o f
the wake region for each o f the grids, with the vortex breakdown region being clear. It is apparent that the wake
is swept in the direction o f the freestream, which is upward from the w ing surface. Thus, a fourth probe will also
be considered which sits at a central location within the wake region at x /c r = 1.2 and z /c r — 0.2. This is also
shown in Figure 4.15 and will be compared to the probe at the trailing edge, z /c r = 0.1. From analysis o f multiple
instantaneous flow domain data files for each solution, it was found that the mean locations o f vortex breakdown
were very similar. This location is approximately x /c r = 0.8 6 for the Refined TE grid solution.

K7 (.■ «.f I l.l 1.2 I.) 1.4 *.7 *.« I i.l 1.2 1.2 1.4

(a) F in e G rid (b ) R efin ed T E G rid

Figure 4.15: Contours o f instantaneous y vorticity on a slice through the vortex core for the fine and refined TE
grid at T = 50

(a) M v elocity

l,ec#U#n

(b ) V velocity

CIwH wIm l^acalion

(c ) w velocity

Figure 4.16: Mean and RMS behaviour o f non-dimensional velocity components at eight probe locations through
vortex core region for the fine and refined TE grids

The mean and RMS velocity components for the probes through the vortex core region are shown in Figure 4.16.
From analysis o f these plots, it is clear that the mean velocity behaviour at each o f the probe locations both up­
stream and downstream o f the trailing edge are very similar for the two grids. Due to the slight difference in the
mean breakdown locations determined from the instantaneous domain files as described previously, there are some
differences between the solutions close to the breakdown location. This is the case for all velocity components.
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 101

Nm l)fa»eie*wel Tim#

(a ) u v elocity

N m n tm m im m i Th—

(b ) V v elocity

I M W T K U tM I

Nm Hwe

(c ) w v elocity

Figure 4.17: Unsteady behaviour o f non-dimensional velocity components shown by time histories and PSD fre­
quency plots for the probe at x f c r = 1.00, z / c r — OA for the fine and refined TE grids

« A rtin n mit
Nm lM n t« 4 M al Tim* NirMlMl M iwhtr

(a ) u velocity

1---- Flm G fM 1

j [
(b ) V velocity

Nm Plm*md#Ml H m t

(c) w velocity

Figure 4.18: Unsteady behaviour o f non-dimensional velocity components shown by time histories and PSD fre­
quency plots for the probe at x /c r = 1.2, z /c r = 0.2 for the fine and refined TE grids
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 102

however, these differences are not significant. Downstream in the wake region, the agreement is very good and al­
though there is a slight difference in mean w velocity a tx /cy = 1 2 overall the flow behaviour is very similar. These
slight differences may be due to differences in the location o f the probes in relation to the mean post-breakdown
flow. The RMS velocities are also in very good agreement both upstream and downstream o f breakdown. D ow n­
stream o f the trailing edge, as the flow returns to freestream conditions beyond the wake, the RMS velocities tend
to zero and the unsteady behaviour disappears.

Consideration o f the unsteady behaviour on the two grids is obtained from analysis o f the probe signals at the
trailing edge and within the wake, as specified previously. The time histories and PSD analyses o f the velocity
com ponents at the trailing edge probe are shown in Figure 4.17. From the time histories, it is evident that the mean
and amplitude o f the signals from the two solutions are very similar, which is in agreement with the mean and
RMS plots discussed above. From consideration o f the frequency content o f the signals, evidence o f the similarity
o f the two solutions is clear. All the dominant frequencies o f the flow identified in the previous section, for this
location on the fine grid, are captured in the Refined TE grid solution. The agreement is good for both frequency
and magnitude. Therefore, it may be concluded that at this location and upstream, over the wing, the trailing edge
refinement has little effect on both the mean and unsteady behaviour o f the flow.

To consider the effect o f the refinement on the unsteady behaviour further within the wake region, the time his­
tories and PSD analyses for the probe at x /c r = 1.2, z /c r = 0.2 are shown in Figure 4.18. The time histories o f
each velocity component show that the refined TE grid predicts a larger amplitude than the fine grid, however the
mean values appear to be similar. A lso clear from the time histories for the refined TE grid is that there appears
to be more fluctuations at higher frequencies. This is confirmed from the PSD analysis which shows more energy
occurring at frequencies in the range Sr = 8 — 10 for all the velocity com ponents. A lso present in the u and w
velocities for this solution, is a second dominant peak at St % 2, which has similar energy to the St = 3.25 peak.
This frequency content is suggested by the fine grid results but is not w ell defined. Therefore, it is evident that the
trailing edge refinement has an effect on the unsteady behaviour o f the flow within the wake. However, this does
not have an upstream effect on the flow over the wing and on the breakdown location. It is clear that the streamwise
refinement o f the grid does allow som e higher frequency content to be predicted, however this is still lower than
would be expected for turbulence and from Figure 4.15 it is clear that any small scale structures in the flow still
dissipate quickly downstream o f the trailing edge. This suggests that significantly more grid points are needed in
the trailing edge region to capture the frequencies associated with turbulence. It may also be suggested that the
overall refinement should be considered in this region and not just in the streamwise direction.

From this investigation, it is clear that the trailing edge refinement and resolution o f the near trailing edge wake
has little overall effect on the predicted unsteady behaviour o f the flow upstream o f the trailing edge, with only a
slight downstream shift in mean breakdown location being found. However, the resolution o f higher frequencies
within the wake has been slightly improved. Therefore, as this will improve the D ES solution overall, with more
scales being resolved, it is concluded that the refined TE grid results w ill be used for the remainder o f the DES
investigation. However, throughout, it should be noted that the fine grid results are very similar.

Figure 4.19: Location o f probes though vortex core region compared to u velocity contours at each streamwise
location
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 103

4.5 Evaluation of LES Region


As mentioned above, the best available solution from the DES calculations was that obtained using the Refined TE
grid with a time step o f A T = 0.0025. Further analysis was performed on these results to consider the unsteady
behaviour o f the flow and the ability o f DES to predict this behaviour.

4.5.1 Unsteady Behaviour of DES Solution


To consider the unsteady behaviour o f the flow solution, the same five probes as used in Sections 4.3 and 4.4
are used. These are shown in Figure 4.19 relative to the vortex system at each streamwise location. From this
Figure, it is clear that the probes at x /c r = 0 .5 3 ,0 .6 3 and 0.7 4 are upstream o f vortex breakdown, with the probe
at x /c r = 0.53 sitting above the vortex core within the shear layer and probes at x /c r = 0.63 and 0.74 close to
the vortex core. The probe at x /c r = 0.8 4 also sits within the vortex core and is found to be close to the mean
vortex breakdown location, which was found to occur at x /c r = 0.86. The probe at x /c r = 1.0 is downstream
o f breakdown, below the vortex core winding. Keeping these locations in mind, the velocity components were
analysed. This analysis was carried out by considering the mean and RMS values at each location and by evaluating
the time histories and PSD frequency content o f the signals. The results o f these analyses are shown in Figures
4.20 and 4.21 respectively.

(V>
<w>

Figure 4.20: Mean and RMS behaviour o f non-dimensional velocity components through vortex core

Strmbal mwmWf

(a) u v elocity

(b ) V v elocity

Nm Uim w J iMÜ TImm

(c) w v elocity

Figure 4.21: Unsteady behaviour o f non-dimensional velocity components at probes through vortex core region
shown by time histories and PSD frequency plots.
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 104

At x / Cr = 0.53, the time history of u exhibits a relatively large amplitude periodic oscillation with a high frequency.
Closer to the vortex core aix/cr = 0.63 and 0.74, the signal oscillation becomes more irregular and the amplitude
decreases significantly. This reduction in amplitude is consistent with the reduction of the RMS values given in
Figure 4.20 for these locations. At x/c,- = 0.84, the time history changes significantly from the upstream probes,
with a high amplitude, low frequency oscillation being clear. This also coincides with a large increase in RMS
velocity, however the mean velocity has decreased. The mean velocity at this location is positive. However, from
the time history it is evident that the flow does reverse and that breakdown crosses the probe location. Downstream
of the breakdown location the amplitude decreases and a more periodic waveform returns. The mean velocity at
this point has only increased slightly compared to the x/cr = 0.84 location, however the RMS value has decreased
and the flow does not recirculate in this region (the u velocity does not become negative).

Considering the frequency content o f the u velocity signals given by the PSD plots, a number of dominant frequen­
cies at each of the probe locations is clear. The most dominant frequency found occurs for the probe at x /c f = 0.84
at a non-dimensional frequency o f approximately St = 0.07. Two other low frequencies are also apparent at
St % 0.27 and 0.67 but there is little energy at higher frequencies at this location. As vortex breakdown is found to
oscillate across this probe location, it may be suggested that this phenomenon produces this low frequency. Due to
the energy in this low frequency being so large, the frequency content of the other probes is unclear using the scale
o f Figure 4.21, Therefore, Figure 4.22 shows the same PSD plot with the x /cr = 0.84 signal removed. From this
plot, it is clear that at x / c r = 0.53, the high frequency content mentioned above corresponds to frequencies in the
range St = 4.5 —6, with frequency content also present at St 9. Again, due to the location of this probe in the
shear layer, it may be suggested that these frequencies are due to shear layer instabilities and structures, such as the
Kelvin-Helmholtz instability. Downstream, close to the vortex core, it is apparent that there is little energy in the
signal, however, with a further change in scale (not shown) it is found that at x/c,- = 0.63 there is a weak presence
of the frequencies in the range St = 4.5 —6 and at x/c,- = 0.74 the St = 0.07 frequency is weak but dominant similar
to the downstream probe at x/c,. = 0,84, Finally, downstream o f breakdown, a new range o f dominant frequencies
is found. These occur in the range St = 3 —3.5, with a dominant peak at approximately St = 3.25. There are also
frequencies present at approximately St = 0.13 and in the range St = 5 —6. These frequencies are most likely to
be connected to the upstream phenomena causing the frequencies discussed for the other probes.

Figure 4,22: PSD of u velocity with x/c,- = •0.84 probe signal removed for clarity of frequency content of remaining
probes

To consider the behaviour in the spanwise direction, the v component of velocity is considered. From Figure 4.19,
it is evident that the probes upstream of breakdown do not sit exactly on the vortex core axis. This is confirmed
from the mean velocity plot as the v velocity at each probe location does not have a zero mean. The probes at
x/Cf — 0.53, 0.63 and 0.74 sit above the vortex axis and the probe at x/cr = 0.84 sits below. The vortex core
axis is crossed at approximately x/c, = 0.8. As the vortex core axis is approached the RMS levels decrease, with
an increase occurring close to vortex breakdown. Downstream of breakdown the RMS velocity decreases again.
This is also apparent from the relative amplitudes of the time histories shown in Figure 4,21(b). As before the
fluctuations at x/c,- = 0.84 are greater than for the other probe locations with an obvious low frequency content.
This frequency corresponds to that found for the streamwise velocity, St pa 0.07, and shows that the vortex break­
down location also oscillates in the spanwise direction. However, unlike for the streamwise velocity there is also
a higher frequency range present in the signal at = 3 —3.5. Which corresponds to the frequency identified as
being caused by the vortex core winding downstream of breakdown. Therefore, it may be suggested that the helical
winding mode causes a spanwise motion of the vortex core close to breakdown. At x/cr = 0.53, the amplitude
of the signal is less than that found for the streamwise velocity, however the frequency content is similar with the
dominant frequencies occurring at St = 4 . 5 - 6 and St 9, The frequency content of the probes at x/cr = 0.63
and 0.74 is also very similar to the streamwise velocity plots, however the power of the response at St = 4 . 5 - 6
increases at x/cr — 0.63, Thus, the shear layer instabilities appear to have a greater effect on the vortex in the
spanwise direction. Downstream of breakdown, the behaviour is again comparable to the streamwise velocity,
although the strength of the signal is lower.
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 105

Finally, the normal w velocity is considered in the same way. Again, from the mean velocity plot it is evident that
only the probe at x/c,- = 0.84 sits exactly on the vortex core axis due to its zero mean. Upstream of this location
the probes sit inboard of the axis as the normal velocity is negative at these locations. The behaviour of the RMS
velocity is very similar to the spanwise velocity as the trailing edge of the wing is approached. Considering the
time histories, it is evident that the amplitudes of the signals are comparable to those for the spanwise velocity.
However, the frequency content is, again, slightly different. For the probe at x/c,- — 0.53, it was found that dom­
inant peaks occur at St % 3 and 3.5 with the frequency at St = 4.5 becoming more prominent compared to the
other velocity components. There is also no peak found at St = 9 for this case, suggesting that the shear layer
instabilities have no higher frequency normal component at this location. Overall it appears, that the frequencies in
the w velocity signal are lower than for the other components. Downstream at x/c,- = 0.63, there is also energy at
frequencies of St = 4.5 —5. At x/c,- = 0.84, the most striking difference in the frequency content compared to the
other velocity components is the disappearance of the low frequency peak at St = 0.07. This suggests that there is
no vertical motion of the vortex breakdown location. However the dominant frequencies at St = 3 —3.5 are still
evident, caused by the rotation of tire helical mode of breakdown just downstream of this location. At the trailing
edge, this frequency is also found, with another dominant frequency occurring at St = 5.5 —6.

From this analysis, it is clear that there are a number of identifiable features, both upstieam and downstream of the
breakdown location with relatively low frequencies. Upstream, the flow is dominated by a strong vortical system,
containing both primary and secondary vortices. Close to the vortex core this flow exhibits only small fluctuations
and the influence of other flow phenomena is apparent. Within the shear layer, evidence of shear layer roll up
instabilities, such as the Kelvin-Helmholtz type instability have been found from the frequency data at & = 4,5 —6
and % 9. At breakdown, the flow is dominated by the motion of the breakdown location which oscillates in
the streamwise and spanwise direction at a very low non-dimensional frequency of St = 0.07. Downstream of
breakdown, the helical mode instability is present and the frequencies corresponding to its rotation and general
behaviour have been isolated, & — 3 —3.5, and specifically a dominant peak at St = 3.25. A summary of the
frequencies determined for each flow feature is given in Table 4.4 in a similar manner to Table 4.1.
Phenomenon Stroiihal Number
Helical Mode Instability 3 - 3 .5 ,3 .2 5
Shear Layer Instabilities 4 .5 - 6 , Rj9
Vortex Breakdown Oscillation 0.07

Table 4.4; Frequencies corresponding to important unsteady features of vortical flows from unsteady DES results
It is in the post-breakdown flow where turbulent behaviour is expected to be found as the vortex breaks down and
loses its structure. However, from these results it is clear that the helical mode winding exhibits coherent periodic
behaviour, which suggests that it is not driven by turbulent phenomenon. In fact, all the phenomena described
above have been found to occur for a range of configurations as detailed in Table 1.2, and none appear to be
dependent on turbulence within the flow. Further evidence of this may be obtained from the results of a highly
under resolved (both spatially and temporally) DES calculation which was performed on the coarse grid described
in Section 4.2.1 using a time step of A t = 0.01. Using such a coarse calculation, it is not likely that any small scale
fluctuations will be captured and indeed from the time history and PSD plot for tlie same probe location shown in
Figure 4.23 it is clear that none are found. It is clear from the PSD analysis that the frequency of the helical mode
instability is identical for tliis case at St = 3.25 and although no small scale structures were captured this had no
effect on the prediction of the vortex breakdown winding and its frequency.

SCruiihal luimUcr

Figure 4.23: Non-dimensional u velocity time history and PSD for a probe on the vortex axis, downstream of
vortex breakdown, from a highly under-resolved DES solution, coarse grid. A T = 0.01
This conclusion is also confirmed from consideration of the literature. A number of numerical investigations have
been performed using both inviscid [23] and laminar [111, 112] methods, which clearly show the helical mode
instability behaviour. These investigations and their results were discussed in detail in the literature review of
Section 1.5.
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 106

4.5.2 Decomposition of LES Solution


It seems possible that the majority of the frequencies captur ed by the DES calculation can be attributed to the large
scale, coherent, low frequency flow phenomena discussed above. As these unsteady flow phenomena are dominant
flow features and an inherent part of the unsteady flow behaviour, they may be considered as par t of the mean flow
and not part of the unsteadiness due to turbulence. Thus, in the post-breakdown flow, the turbulence should be
treated as non-stationary. Considering the discussion of non-stationary turbulence and LES given in Chapter 2, the
decomposition of the instantaneous velocity was described in Equations 2.38 - 2.40 and can be summarised as,
Ui — (Ui) P U/'F (1 — (4.1)
R esolved on g rid m odelled b y SGS

The unsteady flow phenomena described above may be said to contribute to u'/, the unsteady part of the mean
flow. The DES signals shown in Figure 4.21 can then be thought of as being made up of the sum of the mean,
unsteady mean flow component and a percentage of the turbulence resolved on the grid. To identify each of these
components and their respectivebehaviour, the u velocity signal was time averaged in a similai' manner to that
used for theUR ANS methodand given by Equation 2.21. By applying this method,the non-stationary mean
may be separated from the turbulent fluctuations of the signal. However, the non-stationary mean is determined
from the time averaging sample rate, T, and therefore care should be taken when choosing this parameter. For
non-stationary turbulence, the sample rate should be large in comparison with the turbulence time scales but small
in comparison with the mean flow fluctuation time scales. To investigate the optimum sample rate for the DES
solutions, consideration was given to the u velocity signal from the probe at the trailing edge, x/c,- = 1.0, detailed
in the previous section. Figure 4.24 shows the PSD of this signal against both the non-dimensional frequency, St,
and the non-dimensional period of the oscillation, 11St. Considering tlie analysis given in the previous section,
it is clear that there is a dominant frequency in the flow at approximately St = 3.25, which has been attributed to
the motion of the helical mode instability. At this location higher frequencies are also present in the flow, around
St = 6, but it is not clear if these are related to a coherent structure in the flow in this region. Therefore, it is taken
that the highest mean flow frequency is St = 3.25, which corresponds to a non-dimensional period of 0.31.

(a) PSD vs. St (b) PSD vs. Non dimensional period

Figure 4.24: PSD against St and non-dimensional period for non-dimensional u velocity for a probe atz/cy = 1 .0
on the vortex core axis.
Using this period as a starting point, the time average of the signal was calculated with the resulting time histories
and PSD analysis plots shown in Figure 4.25. In each plot, the calculated fluctuating mean and the original signal
are shown, with the stationary mean shown in the time histories for comparison. As the initial sample rate is equal
to the period of the expected mean flow it is clear that it is insufficient to capture all the mean flow fluctuations.
Indeed the dominant frequency of the mean flow, at this sample rate, is equal to the vortex breakdown oscillation.
Therefore, the sample rate was increased and the mean flow was recalculated for T = 0.1, 0.05 and 0.01 as shown
in Figure 4.25. It is clear from this analysis, that as the sample rate was decreased, more of the flow features were
captured withm the mean flow. Until at T = 0.01 the mean signal and the original signal coincide.

As the two signals coincide at a sample rate of T = 0.01, which is four times the time step used for the calculation
and an order of magnitude larger, it is evident that despite refining the time step, as detailed in Section 4.3, the
level of temporal resolution has not increased. For a time step of A t = 0.0025 it was expected that the maximum
resolvable frequency would be approximately St = 40 (based on a minimum of 5 samples per fluctuation, see
Section 2.6). However, it is clear that this level of resolution has not been obtained and the effective time step of
the solution is approximately A t = 0.01. This corresponds to a maximum frequency of St ~ 10 which is closer
to the maximum frequencies found in the PSD analyses. As both the spatial and temporal refinement are equally
important for DES calculations, it may be suggested that this under-resolution of the frequency content of the flow
is due to the grid not being refined enough in the post-breakdown flow region, where the turbulence will begin to
form.
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 107

(a) r = 0.31

(b) 7- = 0.1

Nm Tkm

(c ) T = 0 .05

I—«

Nm « n w w é M il TWiw

(d ) 7' = 0.01

Figure 4.25: Non-dim ensional u velocity time histories and PSD plots for a probe on the vortex axis, downstream
o f vortex breakdown showing both the stationary and non-stationary mean at different sample rates

4.5.3 Resolution of DES Solution


To consider this further it is necessary to examine the behaviour o f the unsteady flow on the grid in this region.
For an unsteady and turbulent flow it should be possible to see the fluctuations o f flow parameters on the grid.
As discussed previously, the grid is used as a spatial filter and, thus, the size o f the cells are used as the spatial
sample rate, in a similar manner to the use o f the temporal sample rate. A t. However, as also mentioned before,
it is not practical to keep this sample rate uniform throughout the regions o f interest for delta wing geometries.
Figure 4.14(b) shows contours o f Amax for the refined trailing edge grid in the vortex core region which shows the
changes in grid size over the post-breakdown region. From this plot, it is clear that in the region o f interest, close
to the trailing edge the grid size is relatively constant at approximately 0.0055cr. U sing this as the sample rate,
the maximum wavenumber o f the spatial resolution can be determined. In Spalart’s guide to DES grid generation
[177], it is suggested that the minimum wavelength o f a structure captured by a grid w ill be equal to five times the
maximum grid size i.e. 5Anuix- Using this as a guide, it can be calculated that the minimum non-dimensional wave­
length captured in this region w ill be 0.0275cr. This corresponds to a maximum non-dimensional wavenumber o f
approximately k = 18 and a minimum eddy size o f 0.055cr due to the Nyquist criterion.

To confirm this analysis, a 1-D slice through the vortex core region ( y / s = 0 .7 ) was taken at a constant height
above the wing surface (z/cr = 0.1). Treating this slice in the same way as a time trace, with x /c r being analogous
to time, the data was analysed using the probe analyser program as before. Figure 4 .2 6 shows the results o f this
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 108

analysis. In a similar manner to a time history, a plot o f u velocity against Jt/cr is plotted for the flow downstream o f
breakdown. This shows that only large scale fluctuations appear to occur downstream o f the breakdown location.
Consideration o f the PSD analysis o f this signal confirms this, as the dominant peak occurs at a wavenumber o f
approximately K — 0.5 (not shown). However, it was found that there is energy at higher wavenumbers up to
approximately = 18, although very small. M ost o f the energy on the grid is found for wavenumbers less than 10,
which is similar to the temporal analysis. This suggests that although smaller eddies are captured by the grid, they
are very weak in comparison with the larger structures.

(a) S p atial h isto ry (b ) P S D vs. W avelength

Figure 4.26: Analysis o f u velocity behaviour from spatial slice through vortex core at z /c r = 0.1 to determine
resolution o f grid

The physical size o f these eddies can be considered from analysis o f the non-dimensional wavelength o f the signal,
as shown in Figure 4.26(b). The wavelength is calculated as the inverse o f the wavenumber. This plot is very inter­
esting as it shows that the minimum wavelength captured on the grid is also close to O.OScr. However, the lowest
clear peak is 0.11. Confirmation o f the size o f the captured eddies may be taken from the contours o f y vorticity
shown in Figure 4.15(b), which clearly show structures with diameters o f approximately 0.06. It is evident from
this analysis that the minimum eddy size is still approximately 5% o f the root chord, which is relatively large,
particularly with respect to the expected size o f any small scale turbulent eddies, which would be less than l%Cf.

To consider how this spatial under-resolution would affect the temporal resolution o f the solution, it is possible
to relate the frequency o f the eddies to their wavelength, and therefore wavenumber, using the local velocity
magnitude. This relationship is defined as,
St = UiocajK (4.2)
A s the velocity at a given location will fluctuate in time, this relationship may only serve as a guide to the effect
on the temporal resolution. However, in the post-breakdown flow, the instantaneous velocity is almost always less
than the freestream velocity. Therefore, the maximum non-dimensional frequency resolved on the grid will be
less than 18. The ability o f the spatial and temporal sampling rate to capture the turbulence may be determined
by considering a log-log plot o f the PSD analysis. Figure 4.27 shows the results o f this analysis. The spatial
resolution can be compared to the Kolmogorov -5/3 slope, which describes the theoretical behaviour o f the energy
within the turbulence for the inertial subrange. It is clear that there is very good agreement with the theory, thus
the grid is capturing the energy cascade w ell, despite the maximum resolved frequency being low in comparison
to the turbulent scales. For a further refined grid, it would be expected that this gradient was maintained, however
there would be more power at higher frequencies and more o f the unsteady flow behaviour would be resolved.

(a ) S p atial - P S D vs. K (b ) T e m p o ral - P S D vs. S t

Figure 4.27: Turbulent spectrum for both spatial and temporal scales to show accuracy o f energy cascade within
computational results

The temporal resolution plot is created from the non-dimensional u velocity at the post-breakdown probe location
detailed previously and also shows the maximum resolvable frequency. It is clear that beyond this point the gra­
dient o f the plot increases and the energy within the scales reduces rapidly as the frequency content o f the flow is
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 109

modelled by the sub grid scale and so is not present in the time trace of the velocity.

Therefore, from this analysis, it has been shown tliat m the post-breakdown region, the unsteady flow exhibits
non-stationary mean flow and tuibulent fluctuations. From the decomposition of the instantaneous velocity given
in Equation 4.1, we can attribute the mean flow fluctuations to u'[ based on a sample rate of T — 0.01 for the
time average. However, this shows that there are few turbulent fluctuations captured in the DES signal and so
the level of turbulence captmed by the grid, is close to zero. This means that the majority of the turbulence
for this solution is modelled by the subgrid scale model. Further analysis on the grid shows that the maximum
eddy size captured by tlie grid is approximately 5%c,-, which is large for turbulent scales. This confirms that tlie
grid in the post-breakdown region is under-resolved and that the LES region is in fact acting in a similar way to
a UR ANS model. However, despite the under-resolution of the results, it is clear that the characteristic behaviour
of vortex breakdown is captured and so the question arises: What effect does the turbulence downstream of vortex
breakdown have on the overall flow behaviour? It may be proposed that the turbulence downstream of breakdown
and the trailing edge has a minimal effect on the mean flow behaviour, such as the helical mode instability and its
characteristic frequency. Therefore, it may also be proposed that UR ANS may be able to adequately predict this
behaviour at smaller computational expense. However, before testing this proposal, it is necessary to validate the
DES solution with existing DES and experimental data.

4.6 Qualitative Comparison with Cobalt Results


It is helpful to compare the results with other DES calculations. One of the most prominent users of DES for delta
wing vortical flows is the United States Air Force Academy (USAFA) [24, 29, 30,116, 161,163]. A great deal of
work was carried out by the USAFA, using the unstructured flow solver, Cobalt, as part of the NATO RTO Task
Group AVT-080 [197, 200]. This group considered the behaviour of the flow on the 70" ONERA test case used
here. As there is a great deal of experience at the USAFA, it was felt that a qualitative comparison with their results
would give an indication on how well the structured DES results were performing and, therefore, would allow a
benchmark of the cuiTent results.

As the majority of the unsteady results obtained from the AVT-080 Task group involved only tlie behaviour of the
normal force coefficient, further data was needed to consider the flow behaviour in the post-breakdown region.
Thus, the unsteady results from the VFE-2 case, described in Chapter 3, will be used as well. Despite this case
being transonic, the non-dimensional behaviour of the flow should generally be similai- and thus a qualitative
comparison may be made.

4.6.1 Comparison with 70" ONERA SA-DES Results


The USAFA geometry for the 70" wing differs slightly to that used in the current investigation as the trailing edge
has not been bevelled and is blunt, similar to the experimental configuration. Another difference is that the US­
AFA solutions have been obtained with the experimental Mach number of M — 0.069. This discrepancy should
not make a significant difference to the results, particularly if the non-dimensional behaviour is considered.

A time step study was earned out by Morton et al. and is detailed in Ref’s [24] and [161]. Six different time steps
of varying refinement were considered for a baseline grid with 2.7 x 10^ cell volumes. The non-dimensional time
steps were: A t = 0.00125, 0.0025, 0.005, 0.01, 0.02 and 0.04. Using the PSD data of the resulting normal force
coefficients, and plotting the PSD against the non-dimensional period (l/5 t) it was found that witli decreasing
time step the dominant frequency of the flow approached an asymptotic period of approximately 0.1 (St = 1 0 ).
From this investigation, a baseline time step of A t = 0.0025 was deemed to be the optimum for accuracy and
for reasonable computational expense for this grid size. This can be compared to the time step study detailed in
Section 4.3, where it was also found that the non-dimensional period of the dominant frequency reduced with a
reduction in time step size.

The effect of grid refinement has also been considered and is detailed in R ef’s [29,30,116,163]. In these investiga­
tions a number of grids of varying refinement were created and the solutions compared. Adaptive mesh refinement
(AMR) was also used to determine the effect of localised refinement in the areas of interest. The baseline grid
described above was initially used and a factor of \/2 was then applied to scale the initial overall refinement and to
create grids with 1.2 x lO*’, 6.7 x 10® and 10.6 x 10® cell volumes. The AMR grid was created from the baseline
grid solution with isosurfaces of vorticity being used to define the region where tlie grid would be refined. This
was performed twice, witli the distribution of grid points being doubled each time. Overall, this resulted in a grid
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 110

(a) C o a rse - 1.2 x 10^ (b ) M e d iu m - 2 .7 x 10*

*t

(c) F in e - 6 .7 x 10* (d ) R eal F in e - 10.6 x 10*

(e ) A M R G rid - 3 .2 x 10* (f) C u rre n t R esu lts - ~ 8 x 10*

Figure 4.28: Isosurfaces o f vorticity for various USAFA unstructured grids compared to current results on refined
trailing edge structured grid. The number o f cell volum es for each grid are given for comparison. [29]
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 111

with 3.2 X 10^ cells. The non-dimensional time steps used were also scaled to the size o f the grid, with a time step
o f A T = 0.0025 being used for the baseline grid and the factor o f \/ 2 being applied in a similar manner as before.
Table 4.5 details the time step and grid size for the coarse, baseline and fine grids. The table also contains details
o f the cell size in the focus region as described in Ref. [177]. These features are also included for the refined TE
grid used for the current results.

Cells ^max AT
USAFA Coarse 1.2 X 10^ 0.0065 0.00357
USAFA Medium (baseline) 2.7 X 10^ 0.0046 0.0025
USAFA Fine 6.7 X 10^ 0.0035 0.0018
Current Results ~ 8 X 10^ 0.0055 0.0025

Table 4.5: Details o f grid features for USAFA grid study and comparison with current results

It is interesting to note that the maximum cell dim ensions in the focus region for the baseline grid, A^ax, is close
to that used for the current results, however the overall size o f the USAFA grid is much smaller. This is due to the
refinement in the region o f interest having to be carried out to the far field for the structured grid. This increases
the grid size and the relative computational expense. Therefore, the structured grid is o f comparable refinement
to the baseline grid from the USAFA results. This is also clear from comparison o f the flow solutions. Figure
4.28 shows instantaneous isosiufaces o f vorticity magnitude for each o f the grids, and highlights the increasing
resolution o f the flow structure with increasing grid refinement. From this, it is evident that the level o f vortical
structures captured by the current results is between the coarse and the baseline grid solutions. It is also interesting
to note that the AMR grid is comparable in flow resolution to the fine grid with 6.7 x 10^ cell volumes.

To consider the unsteady aspects o f the flow, the time histories o f normal force coefficient were used for analysis.
Figure 4.29 shows the comparisons between a) the coarse, medium and fine grids; b) the very fine grid (G 4), an
AMR grid o f 2.4 x 10^ cells (G 9A 4) and a similar AMR grid with sting and wind tunnel walls included (G 7A 1) and
c) the current results. From these plots it is clear that as the overall refinement o f the grid increases, the unsteady
behaviour captured also increases and there is more energy in the higher frequencies. This is in agreement with
the findings o f the grid study for the current results detailed in Section 4.4. It would be expected that the highest
frequencies captured would be much higher for the finer grids, however considering the current results it appears
that the relative level o f energy in the higher frequencies is similar. It should be noted that the scale o f the PSD
for the current results differs to the USAFA results. It is possible that this may be due to the method chosen to
calculate the PSD o f the signal and may not be a reflection o f the level o f energy in the flow. What is important,
in this comparison, is the relative energy o f the signals, which appear to be very similar. It is also clear that
the dominant frequency o f the current results is lower than for the finer grids o f the USAFA study, sitting at
approximately St = 5 — 6 compared to Sr = 8 — 10 respectively.

y 10

10-*

10*
io­

ta ) U SA F A R esu lts [24] (b ) U SA F A R esu lts [30] (c) C u rre n t R esults

Figure 4.29: PSD plots o f normal force coefficient for current results compared to USAFA results from literature

In Ref. [30], the resolution o f tiu-bulent kinetic energy in the vortex core was considered for the coarse, baseline,
fine and very fine grids detailed above. This is shown in Figure 4.30, with a comparison o f the current results.
From the study it was found that as the grid resolution was increased, the value o f the turbulent kinetic energy in
the vortex core approached the experimental value o f 0.5. Thus, the turbulent properties o f the flow were concluded
to be resolved w ell for the finer grids. However, as the time averaging analysis o f vortical flow in Section 4.5.2
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 112

shows, these values are high as the mean flow fluctuations will be included in the calculation o f the turbulent kinetic
energy. The plot o f the current results show s the mean non-dimensional turbulent kinetic energy calculated, both
using the stationary mean (as in the experimental and USAFA results) and the non-stationary mean (calculated
using a sample rate o f T = 0.1), which is subtracted from the signal before the fluctuating velocities are used. It
is clear that there is a signiflcant difference between the results, particularly close to breakdown, where the mean
flow fluctuates considerably. However, the results obtained from the stationary mean do give reasonably good
agreement with the USAFA results, particularly if the location o f vortex breakdown is considered. Which occurs
for x /c r = 0.5, 0.58, 0.6 2 and 0.62 for the coarse, baseline, fine and very fine grids respectively and at a mean
location o f x /c r = 0.8 6 for the current results. Unfortunately, it is not possible to state the peak value for the
current results, as this w ill occur downstream o f breakdown and there are insufficient point probes in the vortex
core region to determine this value.

Ï
g

x/c

(a ) U S A F A R esu lts (b ) C u rre n t R esu lts

Figure 4.30: Comparison o f turbulent kinetic energy through vortex core between current results and USAFA
results from literature

(a ) U S A F A R esu lts [30] (b ) C u n e n t R esu lts

Figure 4.31: Comparison o f surface streamlines between current results and U SAFA results from literature

A s stated in Section 4.2.2, a transition study was carried out by Morton [30] to consider the influence o f a forced
transition location on the computational results and therefore on the comparison with the experiment. The con­
clusions o f this investigation stated that a forced transition line at constant x /c r = 0 .4 gave the most appropriate
agreement with the experimental results. This investigation provided the basis for the choice o f transition line
applied in the current investigations which corresponds to a constant line at approximately x /c r = 0.36. To con­
sider the behaviour o f the flow due to transition, the surface streamlines are compared in Figure 4.31. From these
two plots, it is clear that the behaviour on the surface o f the wing is very similar. As expected, the transition
occurs slightly upstream for the current results. However, this does not appear to have a significant effect on the
downstream flow. Both plots are taken from instantaneous results, however the current results do not exhibit any
clear evidence o f unsteadiness in the streamlines. In contrast, the secondary separation line for the USAFA results
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 113

shows undulations which may be associated with the unsteady nature o f the flow. Interestingly, both solutions
exhibit “whorl” patterns in the trailing edge region, close to the leading edge. Initially, for the current results, this
was attributed to recirculation o f the flow due to the bevelled trailing edge. However, the USAFA geometry was
exactly reproduced from the experiment and had a blunt trailing edge. Further investigation o f this region is needed
to determine the cause o f this phenomenon. This region is unclear from the experimental results.

4.6.2 Comparison with 65^ VFE-2 SA-DES Results


To allow further comparisons o f the unsteady behaviour o f the Cobalt DES solutions with the current results, the
unsteady data from the VFE-2 solutions described in Chapter 3 were used. Details o f the grid and time step,
detailed previously, are summarised in Table 4.6. In a similar way to the current results, a number o f point probes
were situated in the flow as described previously. To make detailed comparisons, but to restrict the amount o f
data used, two probes from these calculations were chosen to compare qualitatively with the 70" test case results.
These probes were situated on the vortex core, which for this case sits at a constant y /^ = 0.6, at a constant height,
z /c r = 0.1 from the surface o f the wing for locations pre- and post-breakdown, x /c r = 0.7 and 0.9 respectively.
A s discussed in Chapter 3, the mean location o f breakdown for this case was found to be x /c r = 0.68. These
probes are to be compared to the signals from the probes on the vortex core at x /c r = 0.8 4 and 1.0 detailed before
for the current results. Although the locations o f these probes are quite different, the non-dimensional distance
from the mean vortex breakdown location are similar, x /c r = 0.02 and x /c r = 0 .0 4 upstream and x /c r = 0.22 and
x /c r = 0.14 downstream o f breakdown for the USAFA and current results respectively. Therefore, a qualitative
comparison o f the flow behaviour may be made for these locations. Figure 4.32 shows the probe locations for the
USAFA results, with isosurfaces o f x vorticity shown to demonstrate the location with respect to the flow features.
This can be compared to Figure 4.19 for the current results.

C ells G rid Type AT


USAFA 7.89 X 10^ Unstructured 0.0047
Current Results ~ 8 X 10^ Structured 0.0025

Table 4.6: Details o f grid and time step for USAFA VFE-2 calculation and current results

(a ) P lan view (b ) S id e view

Figure 4.32: Location o f probes through vortex core with reference to isosurface o f x vorticity for 65" VFE-2
U SAFA DES calculation

Figures 4.33 and 4.3 4 show the non-dimensional velocity component time histories and PSD frequency plots re­
spectively for both cases in pre- and post-breakdown flow. In Figure 4.33 the time histories o f each case are shown
separately due to the differences in non-dimensional time and likewise, for clarity Figure 4.3 4 shows the PSD
analysis o f the pre-breakdown and post-breakdown results separately for comparison between cases. From the
time histories, particularly for the u velocity, it is clear that there are many similarities between the two results. In
both cases, the location o f vortex breakdown periodically moves upstream o f the pre-breakdown probe location.
This is evident from the low frequency, high amplitude behaviour and the magnitude o f the u velocity traces peri­
odically reducing to less than zero, indicating reversed flow. The fluctuations o f the location o f breakdown seem
to be more pronounced for the USAFA case, however this is likely to be due to the presence o f shocks in the flow,
which have been shown to m ove abruptly. It is also clear from consideration o f all three velocity components that
when the breakdown location is upstream o f the probe, there is less unsteadiness in the flow. Again, this is more
pronounced for the USAFA case, however the amplitude o f the current results also noticeably decreases at this
point. Considering the frequency content o f the solutions at this location shows that the dominant frequency o f the
flow is in good agreement at approximately Sr = 0 .1. However, the power o f the U SAFA signal at this frequency is
CHAPTER 4. APPLICATION C E D ES TO DELTA WING VORTICAL FLOWS 114

Nm IMmvnelewiI TIom

(a ) u v elocity

(b ) V velocity

N— Tim»

(c ) w v elocity

Figure 4.33: Comparison between time histories at similar probe locations on the vortex core in the pre- and
post-breakdown flow. Current results on the left hand side and USAFA results on the right.

(a ) u v elocity

(b ) V velocity

(c ) w velocity

Figure 4.34: Comparison between PSD frequencies at similar probe locations on the vortex core in the pre- and
post-breakdown flow. Pre-breakdown results on the left hand side and post-breakdown results on the right.
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 115

greater, again likely due to the occurrence o f shocks in the flow. For the current results, higher frequencies between
St = 3 and 4 are evident in the spanwise and normal velocity components, due to the influence o f the helical mode
winding, downstream o f breakdown. This is not found in the USAFA solutions, but again this is likely to be caused
by the presence o f the shocks which w ill not allow any instabilities to propagate upstream.

Downstream o f breakdown it is clear that the flow behaviour is again very similar. The amplitude o f the time
traces is in good agreement, particularly for the « and w velocities. However, mean values are different. For the
u velocity this is likely to be due to differences in the severity o f breakdown between the subsonic and transonic
test cases, as described in Chapter 3. Whereas for the v and w velocities, this is likely to be caused by differences
in the location o f the probes with respect to the vortex breakdown structures, such as the helical mode instability.
Considering the frequency content o f the two signals for each velocity component, it is clear that there are more
frequencies present for the USAFA results. However, the low frequency response o f the vortex breakdown location
is still found for both cases, as is a frequency which can be associated with the helical mode. This is close to Sr = 2
for the USAFA results and Sr = 3 - 4 for the current results, as stated previously. The occurrence o f many more
frequencies within the post-breakdown flow signal, may be attributed to the presence o f more smaller structures
occurring in the flow for the USAFA solution, as shown in Figure 4.32. However, it is interesting to note that there
is still little frequency content for frequencies close to or above a Strouhal number o f 10, which would be expected
for small scale turbulent structures.

To consider this further, analysis o f the turbulence on the grid was performed in a similar manner as shown in
Section 4.5.3. A 1-D slice is taken through the vortex core region at z /c r = 0.1 and a PSD analysis is performed
to consider the spatial behaviour o f turbulence. This analysis was then compared to the current results shown
previously. Figure 4.35(a) shows the behaviour o f the u velocity on the slices for both results, downstream o f
the breakdown location. A s the location o f breakdown is different for each solution, the relative distance from
the breakdown location is used. It is clear from this plot that there are more fluctuations o f the u velocity in the
post-breakdown region for the USAFA results. Performing a PSD analysis on this data allows the wavenumber
content to be considered and the resolved eddy sizes and wavelengths to be determined. The results o f this analysis
are also shown in Figure 4.35. Compared to the results for the current grid, it was found that the dominant peak in
the PSD analysis also occurs at a wavenumber o f K = 0.5. It was also found that there is more energy in the larger
wavenumbers for the USAFA results, however the maximum wavenumber resolved is still only approximately
K = 20. This translates to a minimum wavelength o f approximately x /c r = 0.0 3 5 , which is found from Figure
4.35(b). This is not significantly higher than the minimum wavelength o f the current grid. Despite this similarity
o f minimum scales, more energy appears in the flow for all wavenumbers. This may be a consequence o f the
resolution o f smaller scales which capture the energy transfer more accurately, due to a smaller sample rate and
therefore less turbulence modelled on the grid.

(a) S p atial h isto ry relative to vortex b reak d o w n (b ) P S D vs. W avelength

Figure 4.35: Analysis o f u velocity behaviour from spatial slice through vortex core at z /c r 0.1 to determine
resolution o f grid for both current results and USAFA 65® VFE-2 DES results

As before, further consideration o f the level o f turbulence captured in the flow solutions can be obtained by com ­
paring the log o f the PSD plot with the Kolmogorov -5/3 slope. Figure 4.3 6 show s this comparison along with the
temporal comparison for the non-dimensional u velocity at the trailing edge. It is evident from both plots that the
general frequency behaviour is very similar between the two sets o f solutions. Although the USAFA grid exhibits
slightly more energy at higher wavenumbers in the spatial comparison, the temporal comparison is nevertheless
very similar, with the same gradient to higher frequencies being present. Therefore, despite the higher grid reso­
lution o f the USAFA solution demonstrated by the smaller scale structures found in the post-breakdown flow and
the greater frequency content, it may be stated that a similar level o f turbulence is captured by each solution.
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 16

From the comparisons with the USAFA results for the 70" and 65" test cases, it has been found that the overall flow
behaviour is being captured well by the current D ES solutions. However these results appear to be under resolving
the smaller flow features due to lack o f refinement in the post-breakdown region.

StrwtkBi NumWr

(a ) S patial - P S D vs. K c o m p ared to K o lm o g o ro v -5 /3 slo p e (b ) T em p o ral - P S D vs. S t fo r u velocity fro m p ro b e in p o s t­


b reak d o w n flow a t jf /f v = 1 0 0

Figure 4.36: Spatial and temporal comparisons o f current and USAFA DBS results

4.7 Validation of DES Results


As shown in the previous section, the current results agree qualitatively with other DBS results. However, it is
also important to gauge the behaviour o f the solutions against experimental results. The experimental test data
gathered by Mitchell [13] will be used mostly for this task, however unsteady data from other experiments will
also be considered qualitatively to further assess the validity o f the results.

4.7.1 Comparison with Mitchell’s Experiment


The set up o f the experiments carried out by Mitchell was described in Section 4.2 and as such will not be discussed
here. A large proportion o f the data presented in M itchell’s work was time averaged. This time averaging process
is akin to calculating the stationary mean o f the flow and does not take the unsteady mean flow, as discussed in
Section 4.5, into account. This makes comparisons between the computational and experimental results difficult,
particularly as the turbulent quantities calculated will be considerably larger than those which may have existed.
This was also considered when considering the levels o f turbulent kinetic energy compared to the USAFA results
in Section 4.6.

Instantaneous full domain flow solutions could not be used to compare with the time-averaged experimental data.
Therefore, a stationary mean was calculated from 100 time steps, over a total time o f t = I, which gives a sample
rate o f T = 0.01. This provides a relatively small period over which to average, but the amount o f data needed to
perform a full mean calculation over the all the calculation time steps was prohibitive. Due to this, the comparisons
should be treated with caution, but should be sufficient for the purposes o f validation o f the basic flow behaviour.

Figure 4.37 show contours o f the non-dimensional velocity components for each o f the chord wise stations for
both the experiment and the mean computational flow. From the contours o f u velocity it is clear that for the
experimental data breakdown occurs upstream o f the x/Cr = 0.74 position, as at this location reversed flow is found.
Indeed, from the investigation it was found that the mean position o f breakdown occurred at approximately x/c^ =
0.65. Considering the current results, it is clear that the location o f breakdown is quite different, with reversed flow
not being predicted for any o f the slices. As stated in previous sections, the mean breakdown location was found to
occur at approximately x /c r = 0.86, which is downstream o f the slices used in the experiment. The discrepancy o f
mean vortex breakdown location may be due to many factors. It was mentioned in the highlights o f the RTO AVT-
080 task group that this set o f experimental data was affected by blockage and support interference effects, which
may have caused up to 2 — 3" o f upwash. It was stated that this upwash may have caused breakdown to occur
earlier on the wing than would have been expected for this configuration. Other factors include, imperfections
on the experimental model due to the sting fitting, the differences in freestream velocity between the experiment
and computation as it was noted by Mitchell that with an increase in freestream velocity that the breakdown
location moved downstream, or the levels o f turbulent eddy viscosity predicted in the computational results. From
both the grid study and time step study it was shown that the predicted location o f breakdown did not change
significantly with any change in grid density or time step refinement, thus the DBS calculations are consistent.
Further consideration o f the prediction o f vortex breakdown location will be given in the follow ing chapter. This
discrepancy o f location should also be kept in mind when considering the unsteady nature o f the flow, which will
be discussed later in this section.
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 117

X 'c - 0.5,< (X = 5(M)iimii X.'c “ 0 .6 3 1X “ 6*N)iiiin) 11,74 (X 7(HHnin) X c « 0.H4 iX - MMlniml

(a) u velocity

0.53 ).X “ .VM)mni) X/c * 0.63 (X “ 600m m ) X c - 0.74 <X “ 700m m ) X 'c “ 0.K4 ( X - XOOmm)

(b) V velocity

X/c = 0 .5 3 1X = 5lX)mm) Xc 0 .6 3 1X 6(X)mm) Xc 0 .7 4 1X 7(N)iih ii ) X'C 0.K4IX K(X)mm)

(c) w velocity

Figure 4.37: Time averaged velocity results from M itchell’s experiment compared to mean computational results
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 118

From the contours o f u velocity, it is also found that the maximum axial velocity o f the vortex core prior to break­
down was approximately 4(/oo. The mean predicted value from the current results, is found to be considerably
less and is given as approximately 2.2CL. This is consistent with the findings o f the AVT-080 task group, where
a number o f calculations were performed for this test case using various CFD solvers, techniques and grids. The
axial velocity was not found to be accurately predicted for any o f the cases and it was concluded that the grid
refinement at the vortex core was not sufficient. Considering the v and w velocity contours, however, it is clear that
the agreement between the experimental and computational solutions is very good with the magnitude and shape
o f the velocity contours being predicted w ell, despite the differences in breakdown location.

Contours o f x vorticity were also considered and the comparisons are shown in Figure 4.38. In the experimental
plots, it is clear that there are small vortical substructures in the shear layer. However, these structures are not
found in the computational results for these contour levels, despite the unsteady probe data providing evidence
o f oscillations associated with such phenomenon. Evidence o f these structures is found by changing the contour
levels, however this shows that the predicted behaviour is weak. As with all the other experimental contour plots,
the boundary layer region is not captured due to the experimental techniques used, which cannot resolve the flow
close to walls. However, there is a suggestion o f a secondary vortex in the bottom left com er o f each contour plot.
This is also shown in the computational results, with the location o f both the secondary and primary vortices being
predicted well.

x /c = 0.74

(a ) E x p e rim en t [13]

I I
(b ) C u rre n t R esu lts

Figure 4.38: Time averaged x vorticity results from M itchell’s experiment compared to mean computational results

Comparisons can also be made with the experimental surface pressure coefficients obtained from steady pressure
transducers on the surface o f the wing at the same chordwise locations. Figure 4.39 shows these comparisons for
the current results. A lso included is data from two investigations carried out as part o f the AVT-080 task group,
from the USAFA [30], as detailed in the previous section and from work carried out at NLR on a structured grid
using a UR A NS k —(0 model with a modification for vortical flows [31]. It is clear that although the computational
results are in good agreement there is a consistent under-prediction o f the pressure coefficient compared to the
experimental results. For the current results, this corresponds to a difference o f 24% for the x / c r = 0.53 peak. For
the USAFA and NLR results it was reported that the difference was 24% and 22.4% respectively. Other compu­
tational results from the AVT-080 task group, using both DES and U RA NS methods, were also found to exhibit
these discrepancies with the experimental data and differences at x /c r = 0.53 o f 23.8% were reported [165, 196].
As all the computational results were in good agreement and factors such as grid refinement, transition and turbu­
lence model had no effect, it was determined that the differences may have been due to a scaling issue with the
experimental data.The current results scaled by 24% are also shown in Figure 4.39 to allow a broad comparison
with the computational results. This shows that generally the agreement is good when all streamwise results are
scaled by this factor. However, due to the blockage and support interference effects mentioned previously, this
straight-forward scaling can not account for all the physical differences in the flow and should be considered with
care.
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 119

■u

•f.5

I
0.1 •.2 0.4 0.7 0.8 1 O.t 0.2 0.3 0.4 0^ 0.0 0.7 0.0 I

(a ) x/Cr = 0 .5 3 ( b ) x /cr = 0 .6 3

Ji

(c ) x / C r = 0 .7 4 (d ) x / C r = 0 .8 4

Figure 4.39: Surface pressure coefficient data for both experimental and computational results [13, 30, 31]

To consider the unsteady nature o f the flow behaviour pressure readings were taken from Kulite^*^ pressure trans­
ducers on the surface o f the wing as detailed in Table 4.3. The numbering and location o f each o f the probes to
be compared are shown schematically in Figure 4.40. The resulting unsteady pressure time histories are shown
in Figure 4.41. A lso shown are the corresponding time histories for 12 o f the transducer locations taken from the
unsteady probe data on the surface o f the wing.

0 01 oa 0.3 0.4 OS o.t 07 08 08 1

Figure 4.40: Location o f unsteady probes used for comparison with M itchell’s data

It is clear from these traces that the mean pressure decreases with outboard m ovem ent on the wing. This suggests
that the vortex core sits either above or close to the y / s = 0.7 position. In the computational results, the vortex
core is also found to be close to this location. The mean computational pressure coefficients have been scaled by
the 24% factor discussed previously to aid in the comparison and it is clear that they are in reasonably good agree­
ment. The highly unsteady nature o f the flow is obvious, both up- and downstream o f breakdown, with reasonably
large amplitude oscillations occurring at many frequencies. The length o f the corresponding signals should be
considered before any comparisons can be made, with the experimental data being captured over 2 seconds and
the total computed time being equivalent to approximately 0.2 seconds. The corresponding computational time is
marked on the experimental plots for comparison. However, despite this difference, qualitative and quantitative
comparisons can be made. It is clear that the amplitude o f the unsteady fluctuations o f almost all the probes are
in good agreement, with the most obvious exception being Probe 1 at x / c r = 0 .8 4 . In the experimental data, the
signal from this probe exhibits som e rather strange behaviour with the pressure coefficient decreasing significantly
in what appears to be a random pattern. This was noticed by Mitchell, who decided that it was the response o f a
faulty transducer, thus this signal will not be considered for comparison.

To consider the frequency content o f the signals, PSD were calculated from each signal. These are shown in Fig­
ure 4.42, again with similar plots for the current results. The plots taken from M itchell’s work have been altered
slightly to show the corresponding non-dimensional frequencies for comparison. From the experimental plots,
upstream o f breakdown at x /c r = 0.53 and 0.63, the flow behaviour is dominated by a low frequency oscillation,
which occurs at approximately 2Hz {St = 0.08). There is evidence o f som e higher frequency broadband content,
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 120

Probe 17
-1.75 Probe 16
Probe 14

-2.25

-3.25
0.58 0.6 0.62 0.64 0.66 0.68 0.7 0.72 0.74 0.76
Time (*)

(a) x /c r = 0 .5 3

— Probe 13
-1.25 Probe 12
Probe 10

0.6 0.65
Time (s)

(b ) x/Cr = 0 .6 3

^1.75

-2.25
Probe 9
— Probe 8
Probe 6
-2.75
0.65
Tempo Real Time (s)

(c ) x/C r = 0 .7 4

—1.5

0.-1.75

-2.25
— Probe 4
-2.5 — Probe 3

0.7
Tlme(s)

(d ) x/Cr = 0 .8 4

Figure 4.41: Time histories o f unsteady pressure probe data [13]


CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 121

900
Probe 17
800 — Probe 16
Probe 14
700

«600

g 500

I
«
I 400

300

! 200

100

Strouhal num ber

(a) x /cr = 0 .5 3

600
Probe 13
Probe 12
500 Probe 10

400
â 0125

I 300

200

100

Strouhal number

(b ) x/cr = 0 .6 3

600
— Probe 9
— Probe 8
500 — Probe 6

400
Î 0 125

300

200

100

.92
Strouhal number

(c ) x /C r = 0 .7 4

600

500

rt 300

J 0 075
& 200

100

3.M
Fréquence (Hz)
.02
Strouhal number

(d ) x/cr = 0 .8 4

Figure 4.42: Power spectral density plots o f unsteady pressure probe data [13]
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 122

however this has relatively low power in comparison. Downstream o f breakdown, many more frequencies occur,
although the low frequency content is still dominant. Definite peaks occur in the range 15 —50 H z (St = 0.5 —2),
around lOOHz (St % 4) and around l3 0 H z (St % 5). Due to the difference in location o f vortex breakdown it may
not be possible to make direct comparisons between the frequency responses for a given chordwise location. In­
deed, from comparing the experimental and computational results at x / c r = 0.53 it is clear that although the low
frequency content is predicted w ell at St % 0.07, the dominant behaviour o f the flow at this location in the com ­
putational results appears at approximately St = 4.5. This is also true for x / c r = 0.63, but this behaviour does not
occur in the experiments. However, if the non-dimensional distance from the breakdown location is considered,
the agreement between the results is much better. A s the mean breakdown location is x /c r = 0.65 in the experi­
ments, the x / c r = 0.53 station is 0.15cr upstream and the x /c r = 0.63 station is 0.02cr upstream o f this location.
Similarly, for the computational results, the x /c r = 0.74 station is 0.14cr upstream and the x /c r = 0.84 station is
0.04cr upstream o f breakdown. If these two locations are compared, the agreement is significantly improved, with
the low dominant frequency occurring close to St = 0.07 in both results and the higher frequency content focusing
around St = 3 —5.

In the unsteady analysis performed on the current results in previous sections, it was determined that the low
frequency response found in the flow, close to breakdown, was due to the unsteady oscillation o f the breakdown
location. As a similar frequency was found for the experimental data, the behaviour o f the unsteady location o f
breakdown should be considered. Figure 4.43 shows the time traces o f vortex breakdown location for both the ex­
perimental and computational results. The computational results shown were created by considering the location
o f breakdown in the flow domain for every 100 time steps. Due to the computational expense o f the calculation,
it was only possible to consider a total time o f 0.2 seconds. This makes a comparison with the experimental data
difficult as the lowest frequency which could be captured would be approximately St = 0.069 and the dominant
frequency captured for this phenomenon in the experiment is St = 0.043. Considering the amplitude o f the oscilla­
tions it is clear that the vortex breakdown location oscillates with an amplitude o f approximately 15% root chord.
This corresponds to a location x / c r = 0.6 —0.75 for the left hand side and x / c r = 0.65 —0.8 for the right hand side
vortex. Comparatively, the computational results predict an oscillation with an amplitude o f approximately 6%
root chord. This under-prediction o f the amplitude may be due to the symmetric assumption as in the experiment
there may be interaction between the behaviour o f the two leading edge vortices.

"T5T"
Tin*
(a ) E x p e rim en tal R esu lts, B lack - L e ft, C y a n - R igh t; (S c a n n e d - P o o r (b ) C u rre n t R esults
Q u ality )

Figure 4.43: Comparison o f unsteady vortex breakdown results

n u tm ü liMt 0.2;W ri):? it.W i 0 .475 0 554

rWqeeeoe (Hz)

Figure 4.44: PSD plot o f unsteady vortex breakdown results from M itchell’s experiment for left hand side [13]
a = 27® Uoo = 0 .24 (Scanned - Poor Quality)
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 123

From the computational signal, it appears that there are approximately two low frequency oscillations over the 0.2
seconds. This corresponds to a non-dimensional frequency o f St = 0.139. This is higher than the frequency which
was assumed to be the vortex breakdown location in previous sections. However, as the signal length is short the
lower frequency may not be detected. From a PSD analysis o f the experimental data, shown in Figure 4.44, it is
clear that there are a number o f frequencies present in the experimental signal. Thus, it may be suggested that the
frequency captured by the computation is consistent with the higher frequency content. This frequency is also in
agreement with the unsteady breakdown oscillations witnessed from a similar plot for the USAFA DES results for
the 65® VFE-2 test case, which occurred for a frequency o f approximately St = 0.14. Further higher frequency
content is suggested by the trace o f vortex breakdown location plotted at a much smaller sample rate o f 100 time
steps between non-dimensional times o f t = 50 — 51, which is shown in Figure 4.43(b ) as the dotted line.

4.7.2 Comparison to Other Unsteady Experimental Results


Further comparisons may be made by considering other experimental investigations from the literature summarised
in Chapter 1. In the investigation carried out by Klute [6] and summarised in Klute et al. [55], the unsteady flow
over a sharp leading edged, 75® delta wing at an incidence o f a = 40® was considered using digital PIV techniques
in a water tunnel. The model had a root chord o f 0 . 141m and the freestream velocity was 0 .3 2 m s' ‘ which provided
a Reynolds number o f Re = 4.5 x KX*. This is low, particularly in comparison with the current configuration, how­
ever considering the non-dimensional behaviour o f the flow, qualitative comparisons may be made. The purpose
o f the investigation was to consider the unsteady nature o f the helical mode instability o f vortex breakdown and
to consider its evolution with time. Therefore, a large database o f im ages and temporal information was gathered
in the post-breakdown flow region. The digital PIV was set up to record an image 500 times a second, which
corresponds to a sample time step o f approximately A t = 0.004 and data was gathered over a period o f 4 seconds
( T = 9.08).

In this case, due to the relatively high angle o f incidence, vortex breakdown occurred at approximately x / c r = 0.5
on the wing. From the DPIV data, the unsteady velocity signals at a number o f points on a measurement plane
30%Cr downstream o f the breakdown location à tx f c r = 0.8 were isolated and considered using a PSD analysis and
it is with this data that comparisons w ill be made with the current results. Figure 4.45(a) shows the instantaneous
post-breakdown region on a plane through the vortex core at a time t = 0.101. This plane shows the velocity vectors
and corresponding streamlines for the helical mode, with the vortical regions caused by the spiral breakdown
intersecting the plane. The locations o f the two points which correspond to probes within the computational
domain are highlighted in red, their precise locations are given as non-dimensional distances, 0.108, (which will
be referred to as point A) and 0 .158 (point B) above the wing surface. Figure 4.45(b) shows the corresponding
instantaneous vortex core streamline behaviour at r = 50. The corresponding probe locations for the computational
results are also shown.

(a ) D P IV resu lts sh o w in g velocity vectors an d co rre sp o n d in g s tre a m ­ (b ) C u rre n t D E S re su lts sh o w in g strea m lin es
lines (A d a p te d fro m R ef. [6])

Figure 4.45: Instantaneous vortex breakdown regions for experimental and computational results. A lso shown are
the locations o f the data points from which the time histories o f u velocity were taken.

The behaviour o f the vortex breakdown flow structure may be considered from the streamline plots. It is clear
that the locations where the helical mode winding pass through the analysis plane for the experimental results,
are more spread out than for the current results. This suggests that the overall pitch o f the helix is much larger
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 124

1.5
0.2

0.16
0.14

0.5

0.00

-0.5 L
(* ec 0 0.5 2.5 3.5
(non-dim) 0
Froquoncy

(a) E x p e rim en tal re su lts (A d ap ted fro m Ref. [6 ])

-0.5
.-------- ^
41 42 43 44 45 46 47 48 49 SO 10 15 20 25 30
Non Dimensional Time S trouhal num ber

(b ) C o m p u tatio n al resu lts

Figure 4.46: Time histories and PSD analyses o f u velocity for Point A

1.5

0.18

0.14
0.12
2 0.5

0.04
0.02

0.5 1.5 2.5 3.5 (Hz>


(non-dim) 0
Time Frequency

(a ) E x p e rim en tal resu lts (A d ap ted fro m R ef. [6])

1.5 X 10*

0.8

0.6
0.5

0.4

0.2

-0 .5
41 42 43 44 45 46 47 48 49 50
Non D imensional Time S trouhal num ber

(b ) C o m p u tatio n al resu lts

Figure 4.47: Time histories and PSD analyses o f u velocity for Point B
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 125

and the effect o f the breakdown is greater on the wing surface. It may also be suggested that the difference in
helical pitch is due to the proximity o f the vortex breakdown location to the trailing edge in the computational
solutions. However, further analysis at higher angles o f incidence would be needed to state this conclusively. In
the experimental results, the locations o f these intersections were tracked with time and it was found that these
structures were convected downstream toward the trailing edge. With this downstream motion, the spiral o f the
helical mode winding stretched and the diameter increased. Coupled to this increase in diameter is a reduction in
dominant non-dimensional frequency. However, considering the evolution o f the computational results over a time
T = 1 it was found that there was only a slight downstream motion o f the structures on the vortex core plane and
the diameter o f the helical structure did not increase significantly. This diameter is relatively small in comparison
with the experimental results, which would suggest a higher non-dimensional dominant frequency. The behavioiur
o f the structures in the experiments at or close to the trailing edge was not mentioned. However, for the current
results, the helical winding appeared to dissipate. A s discussed in Section 4.5 this is likely to be due to the rapidly
decreasing resolution o f the grid downstream o f the trailing edge. However, it is unclear what effect the trailing
edge has on the coherent vortex breakdown structures. Keeping all these differences in mind, the u velocity traces
at point A and B can be considered and compared to the probe locations shown in Figure 4.45(b).

It is clear from Figure 4.45 that the non-dimensional distance between the vortex breakdown location and the probe
positions is greater for the experimental results. Thus, only a qualitative comparison may be made. However, com ­
paring the results shown in Figures 4.4 6 and 4.47 it is clear that there are both similarities and differences between
the two sets o f results. Considering Point A and the corresponding computational probe at a nominal distance O.lcr
from the wing surface. It is evident that the amplitude o f the time histories o f u velocity are comparable at this
location. However, the mean velocity is much lower for the experimental results, and the flow is found to reverse
for large periods o f the time history. In the computational results the flow does not reverse at any point in the time
period shown. A lso, the level o f fluctuations o f the velocity are found to be less in the computational results, but
not significantly so. The frequency content is also quite different. The dominant peaks in the PSD analysis for the
experimental results occur for St < 2 whereas for the computational results, the main peak occurs at approximately
St = 3.5. This increase is likely to be due to the differences in the helical winding discussed before. In the ex­
periment, there is also considerably more energy in the higher frequencies. Whereas for the computational results
there is som e content at frequencies, St < 10, but this reduces rapidly with increasing frequency. The energy in
the high frequencies o f the experimental results also decays but at a much reduced rate. The presence o f this high
frequency energy relates back to the observation that there are more small scale fluctuations in the experimental
time history and suggests the presence o f smaller scale structures and a turbulent behavioiu". However, this behav­
iour is secondary to the helical model instability and so the flow has not broken down into full scale turbulence at
this location. It may be suggested that a fully turbulent flow, with the breakdown o f the helical mode instability
into smaller structures, does not occur until downstream o f the trailing edge. A similar conclusion may also be
made from the USAFA results for the 65® VFE-2 wing discussed in Section 4 .6 and shown in Figure 4.48 for a
similar probe location to Probe A. For this case, the probe is also approximately 30%Cr downstream o f the vortex
breakdown location. This shows that despite greater overall grid refinement, the results are again very similar, with
the dominant frequencies occurring for St < 10. There is little frequency content above this frequency.

0.05

- 0.5
1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 10 15 20
Non Dimensional Time Strouhal num ber

Figure 4.48: Time histories and PSD analyses o f u velocity USAFA results for 65® delta wing at location on vortex
core plane, x / c r = 1.0, z / c r = 0.1

At point B, the experimental results exhibit a similar behaviour to Point A, with many scales o f fluctuations being
evident from the time trace o f u velocity. However, the PSD analysis o f the signal show s that the frequency o f the
dominant peaks has increased and there is an overall increase in the energy o f the signal. Compared to a probe
in the computational flow domain situated at z / c r = 0.15 above the w ing surface, it is clear that the behaviour
CHAPTER 4. APPLICATION OF DBS TO DELTA WING VORTICAL FLOWS 126

is quite different. The mean velocity is much greater, in fact it becomes slightly higher than the freestream and
the amplitude of the signal is greatly reduced. Considering the frequency content, it is clear that the increase of
frequency compared to Point A has been captured, but the content is very different. These differences between tlie
experimental and computational results are likely to be due to a difference in location of the measurement point
within the post-breakdown region. This region is much larger for the experimental results and the measurement
point sits well within this region. However, for the computational result, this probe location is close to the edge of
the region and closer to the freestream flow.

From consideration of these comparisons and particularly from the experimental results it is evident that the helical
mode structure is dominant downstream of breakdown for at least 30%c,- and it is likely that this structure remains
coherent until at least the trailing edge. Although there is evidence from the experimental results of high frequency
content in the post-breakdown flow, which suggests the presence of turbulence, it is clear that this is not dominant.
At some location, the coherent structure of the helical mode instability will breakdown and the flow will become
fully turbulent, at which point the frequency content will exhibit a broadband response, however this has not been ;
found to occur in the experiments. There is also little evidence of small scale structures in the streamline plot of
Figure 4.45(a). This suggests that turbulence does not become dominant until downstream of the trailing edge,
with the flow over the wing, post-breakdown being dominated by coherent structures. The level of small scale
fluctuations within the experimental signals, does not appear to be signiflcantly greater than the computational
results (particularly in view of the under-resolution of the grid discussed previously).

4.8 Discussion
4.8.1 The Role of Thrbulenee in Vortical Flows
As shown in the previous section, it is clear that the unsteady flow immediately downstream of vortex breakdown
is not dominated by turbulence. Although small scale turbulence does exist, as is evident from the low energy
response at high frequencies in the experimental data from Figures 4.46(a) and 4.47(a). In this investigation,
the ability of the DES turbulence treatment to predict this flow behaviour was analysed and particularly the role
of turbulence in the prediction of breakdown and other dominant flow features was considered in a number of ways.

From the grid refinement study, it was shown that although the sfieamwise refinement improved the resolution
of the unsteady flow in the wake region, there was little overall effect on the mean flow behaviour, particularly
upstream of the hailing edge. It was concluded from this study that an overall refinement was needed in this re­
gion, but that it may be likely that the prediction of the turbulence downstream of the trailing edge would have
only a small effect on the upstream flow predictions. From compaiisons with existing DES calculations it was
shown that with overall refinement of the grid, smaller structures could be captured, both witliin the shear layer
and downstream of breakdown, however, this did not appeal' to have a significant effect on the dominant unsteady
flow frequencies captured. Indeed, from analysis of the turbulent behaviour on the grids, it was found that gener­
ally the behaviour was very similar. Therefore, it may be suggested that the level of grid refinement to capture the
turbulence within the wake of a delta wing is considerably greater than that used in investigations to date.

However, it was shown from validation of the results with existing unsteady data, that the DES solutions were
adequately predicting the dominant features of the flow. These included the helical mode instability of breakdown
and the wandering of the vortex core due to the motion of breakdown. Evidence of shear layer structures were
also found within the frequency data, although it is felt that further investigation on more refined grids is needed
to confirm the behaviour and frequencies of these features. Therefore, it is clear that altliough the small scale
turbulence of the post-breakdown flow is not adequately captured, tins does not appear to have a significant effect
on the ability of DES to predict the dominant flow features. Therefore, it may be concluded that the overall
behaviour of vortical flows and vortex breakdown over slender delta wings is not dominated by turbulence.

4.8.2 The Role of /Isgs hi the DES Calculations


As mentioned, the structured grid used in this investigation is not sufficiently refined to capture small scale turbu­
lence and the smallest eddy size resolved on the grid is approximately 5 %c,-. This means that the level of turbulence
captured on the grid, defined as ^ in Section 4.5.2, is close to zero. The exact value of (p is difficult to quantify as
the precise levels of turbulence in this region have not been quantified. However, as discussed above, it is found
the low energy, high frequency, broadband response of small scale turbulence is missing from the DES results.
CHAPTER 4. APPLICATION OF DES TO DELTA WING VORTICAL FLOWS 127

If <t> is close to zero, this means that the turbulent fluctuations, mJ, are modelled by the subgrid scale model with
only the mean flow being resolved on the grid. This results in the velocity decom position as given by Equation 4.3.

Ui = {Ui) + u'l + u'i (4.3)


Resolved on grid modelled by SGS

From the description o f the URA NS method in Chapter 2, it is evident that for an under-resolved DES calculation,
the behaviour o f the DES model is very similar to the URANS method with the turbulence predicted by a turbulence
model, which in the case o f the DES calculation is the subgrid scale Smagorinsky model. This means that the
subgrid eddy viscosity, /ijos. behaves in the same way as the turbulent eddy viscosity, Hr and will model the
contribution o f the turbulence to the flow. To consider this behaviour, contours o f the subgrid eddy viscosity relative
to the laminar viscosity were plotted on a plane through the vortex core region and are shown in Figure 4.49. From
this plot, it is evident that the levels o f Smagorinsky eddy viscosity predicted by the subgrid model increase in the
vicinity o f the vortex breakdown region and trailing edge. However, these values are low in comparison with values
o f turbulent eddy viscosity predicted for standard Boussinesq models, such as the W ilcox k — (0 model discussed
in Chapter 2, which can be o f the order o f I O'*. This is due to the fact that the subgrid eddy viscosity is scaled by
the spatial filter length squared, A^, as detailed in Equation 2.50, which for the D ES implementation corresponds
to the maximum cell size squared A ^ . Thus, as the grid is refined, the level o f Uses w ill decrease and the value o f
0 w ill increase. It has been shown in this investigation that the magnitude o f Anmx through the grid is insufficient
for 0 to be greater than zero and thus the turbulence is modelled.

1.1 1.2 IJ 1.4

Figure 4.49: Slice through vortex core at y / s = 0.7 showing contours o f sub-grid eddy viscosity relative to laminar
viscosity created by the DES calculation

4.9 Conclusions
From consideration o f DES calculations performed on a structured grid for a slender delta wing at moderate inci­
dence, it was found that a number o f low frequency, coherent unsteady features dominate the flow. Effects o f both
temporal and spatial refinement were examined and the ability o f DES to predict the unsteady nature o f the flow
was considered, particularly in light o f the prediction o f turbulence in the post-breakdown region. Comparisons
were made to other, similar DES calculations carried out by the USAFA and with experimental data to measure
the validity o f the results.

It is clear from this investigation that the DES calculations performed are under-resolved, with little turbulence
being resolved on the grid within the LES region o f the flow domain. From this analysis and the comparisons with
existing D ES results using unstructured grids, it is suggested that the grid requirements to capture the turbulent
behaviour o f the flow close to and downstream o f breakdown are much larger than those described in this inves­
tigation. It was found that to fully capture the turbulent scales it would be necessary to refine the grid not only
over the wing, but also in the region downstream o f the trailing edge. How far downstream may not be proposed
based on the results gained here, but based on the results o f the investigation by Allan [144] who considered the
effect o f sting fairings downstream o f the trailing edge on vortex breakdown, it is felt that a distance o f at least one
root chord length downstream is a good starting p o in t. This w ill have a direct impact on the size o f the grids used
for DES for delta w ing flows, increasing the computational expense o f an already expensive turbulence method.
This is particularly prohibitive for structured grids, which have the disadvantage compared to unstructured grids
that any refinement needs to be taken to the farfield. However, this may be overcom e by considering overset grids,
hanging nodes and hybrid grids. Therefore, it may be concluded that the computational cost o f the calculations
needed to fully resolve the turbulent scales within a delta wing flow is still too high to make these calculations
CHAPTER 4. APPLICATION CE D ES TO DELTA WING VORTICAL FLOWS 128

accessible to the majority of CFD users.

However, the results of this investigation may also show that it is possible that these calculations are not necessary.
From this study it was shown that although turbulence is present in the flow downstream of breakdown, it does
not appear to have a significant effect on the prediction of the salient flow features and the validation with the
experimental data was good, despite the under-resolution of the results. The dominant flow features were shown
to be coherent, low frequency phenomenon, which could be assumed to be part of the mean flow. Therefore, it is
suggested that traditional UR ANS models may be able to perform well and accurately predict the main features of
the flow at a significantly reduced computational cost.
Chapter 5

Assessment of URANS for Predicting


Vortex Breakdown

5.1 Introduction
From the study into the use of DES to capture the unsteady flow behaviour of the vortical flow and vortex break­
down over a slender delta wing, a number of conclusions were made. It was found that the resolution of the
grid used was not adequate to resolve the turbulent scales and that further refinement, both above the wing and
downstream of the trailing edge would be needed to improve the resolution of the flow. This would have the conse­
quence of increasing the computational expanse of an already expensive calculation, particularly as the time steps
involved may have to decrease with the increasing grid refinement. However, it was also found from comparisons
with other DES solutions and with experimental data, that the flow behaviour in the region of interest, downstream
of breakdown, was not initially highly turbulent in nature with the breakdown to turbulence not being found to
occur until much further downstream. From the unsteady analysis of the DES results and from consideration of
the literature, it is clear that the majority of the frequencies associated with the flow phenomena present above the î
wing occur for Strouhal numbers less than 20. Finally, it was concluded that these lower frequencies are within the j|
grasp of more traditional URANS calculations and tliat these methods could capture the unsteady flow behaviour
for a greatly reduced computational cost.

From the literature review in Chapter 1, it is clear that there have been many investigations into the ability of RANS |
models to predict the important features of vortex breakdown, with varying success. From work carried out for f
steady state calculations, it is clear that the standard linear Boussinesq turbulence models struggle to accurately ,C
predict vortex breakdown behaviour due to their inability to correctly model the turbulent behaviour within the
vortex core. Due to this, a number of corrections have been proposed for these linear models, to account for the
rotation of the flow and to improve the flow solutions. Some of these were discussed in Chapter 1 and have been a
found to give good agreement with experimental data. Non-linear eddy viscosity models have also been proposed
and applied to the solution of delta wing flow and again also show improvement compared to linear models for
steady state solutions. This is due to the addition of a dependence on the rotation of the flow in the calculation of
turbulence. However, to date there has been little research into applying these models to unsteady flows and their
ability to accurately predict the important flow phenomena and frequencies is largely unknown.

Therefore, to consider the ability of URANS methods to predict the unsteady behaviour of vortical flow and vor­
tex breakdown, two turbulence models were used, one a linear Boussinesq model with a rotation correction for }|
vortical flows and the otlier a non-linear model. The calculations were performed on the test case and conditions
defined in the previous chapter to allow for the relative behaviour compared to the DES solutions to be considered.
The turbulence models used are, the Æ- m with Pa Enhancer, which is the Wilcox k - O) two equation model
with rotation correction for vortical flows [158], and a Non-Linear Eddy Viscosity model (NLEVM), which is also
based on the k — O) model, but which uses an algebraic formulation for the eddy viscosity instead of the Boussinesq
approximation [170]. Both models are detailed and discussed with respect to vortical flows in Chapter 2.

In order to fully consider all aspects of the URANS solutions, the effect of grid refinement and time step refinement
are considered. The relative modelling approaches and results for each model are the considered before a full
assessment of the ability of the URANS models to predict the unsteady behaviour and dominant frequencies is
carried out and discussed with respect to the validated DES results presented in the previous chapter. Finally, the

129
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 130

results are discussed overall and conclusions made.

5.2 Effect of Grid Refinement


By their nature, U RA NS flow solutions are only dependent on grid refinement for numerical accuracy. As the the
size o f the cells in a grid decreases, the numerical accuracy o f the solution should improve. Before considering the
ability o f the U RA NS m odels to predict the unsteady vortical flow behaviour, it is necessary to quantify the effect
in the grid refinement for each o f the models used. In order to consider this, calculations were performed for the
coarse and fine grids described in Section 4.2.1 with the non-dimensional baseline time step o f AT = 0.01 using
both models. Probes were put in the flow domain at the same locations as detailed before and, as far as possible,
the resulting analyses have been kept consistent to facilitate the comparisons with the DES results.

5.2.1 k —o) model with Pq) Enhancer


A s before, five probes on a plane through the vortex core region were analysed to allow comparisons to be made
using both the mean and unsteady components o f the flow. Figure 5.1 shows the location o f these probes relative
to the flow features, shown by slices o f instantaneous u velocity and an isosurface o f entropy which shows the
winding downstream o f breakdown, for both the fine and coarse grids. From these plots it is evident that the
location o f the vortex core with respect to the wing surface and the relative locations o f the five probes is very
similar for both grids.

(a ) C o a rse G n d (b ) F in e G rid

Figure 5.1: Location o f probes though vortex core region compared to instantaneous u velocity contours at each
streamwise location and an isosurface o f entropy at t = 50, coarse and fine grid comparisons fo T k — (0 with Pw
Enhancer model

Further analysis o f the flow behaviour, shows that the location o f vortex breakdown is different for the two grids.
The mean vortex breakdown location was determined, as before, from the average o f instantaneous flow data at
every 100 time steps. From this, the location o f vortex breakdown was found to occur at approximately x /c r = 0.7 0
for the coarse grid and x /c r = 0.83 for the fine results. This difference in breakdown location, is most likely to be
due to the differences in resolution o f the vortex core behaviour. From comparison o f the contours o f u velocity,
in Figure 5.1, it is clear that the vortex core behaviour is slightly different for the fine grid, with a tighter vortex
core region and the appearance o f a shear layer structure under the vortex. A tighter, more compact vortex core
region may suggest a stronger vortex which may explain the downstream location o f breakdown. The difference in
breakdown location is also obvious both from the slice at x /c r = 0.7 4 , which clearly shows a high velocity region
for the fine grid but a region o f recirculation for the coarse grid and from the isosurface o f entropy, which shows
the differences in the winding behaviour downstream o f breakdown. It is evident from this isosurface, that the
winding for the coarse grid is more elongated, with a larger pitch angle than for the fine grid, which appears to be
relatively compact in comparison.

The mean and RMS values o f the velocity components are shown in Figure 5.2 for the five probes mentioned above.
From the mean u velocity plot, the relative locations o f breakdown are clear, with the vortex breakdown occurring
upstream o f the x / c r = 0.7 4 location for the coarse grid. Both upstream and downstream o f this location, the mean
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 131

values o f both solutions are in good agreement. The agreement is also very good for the v and w components, with
the plane o f probes crossing the rotation axis at the same location. This is downstream o f breakdown for the coarse
grid and upstream for the fine grid solutions. A s suggested from the u velocity contours before, the greatest differ­
ence in mean velocity occurs at x / c r = 0 .7 4 , however, this does not seem to affect the agreement downstream. For
the probes upstream o f breakdown for both cases, it is evident that the spanwise and normal velocities are slightly
greater for the fine grid, confirming a tighter vortex core region and suggesting a stronger vortex occurs for the fine
grid solution.

Considering the RMS velocities and the differences in the behaviour o f the two grids becomes more evident.
Upstream o f breakdown, the results are very similar for all three velocity components, however close to and
downstream o f vortex breakdown the solutions are quite different. It is clear from the RMS o f u velocity that
the level o f unsteadiness at x / c r = 0.74 is very similar for both grids, despite vortex breakdown having occurred
upstream o f this location for the coarse grid. Further downstream, the level o f unsteadiness has increased for the
fine grid solutions (as vortex breakdown has occurred), it levels o ff for the coarse grid. This is consistent for
the V and w components o f velocity, where the coarse grid predicts a higher unsteadiness than the fine grid at
x / c r = 0.74 due to breakdown. Downstream o f this the levels drop o ff and it is clear that the fine grid exhibits
greater unsteadiness in the post-breakdown region.

(a ) u v elocity

<V)

Chord wW LocaCloo )
(b ) V velocity

(c ) *v v elocity

Figure 5.2: Mean and RMS velocity components through vortex core; coarse and fine grid comparisons fo r k — (o
with Pq, Enhancer model

To further analyse the unsteadiness in the post-breakdown region, a single probe situated above the trailing edge is
considered for both cases. It is clear from Figure 5.2 that at this location, the mean velocities are virtually identical
for both grids, but that the RMS velocities and therefore the levels o f unsteadiness are quite different. Figure 5.3
shows the time history and results from a PSD analysis o f the u velocity signal. From the time history, the most
noticeable difference is that the fine grid solution gives a signal with a greater amplitude than the coarse grid, in
agreement with the RMS values discussed above. Considering the frequency content o f the signals, it is clear that
the behaviour is quite different. The coarse grid predicts two dominant frequencies at approximately St = 2.6 and
4.25 with a much smaller peak evident at St w 5.2 which is the harmonic o f the first dominant peak. The fine
grid, however, only predicts one dominant peak at approximately St = 3.4 and som e higher frequency content at
St = 4.5 - 7. For the fine grid, it may be suggested that the dominant frequency captured is associated with the
helical mode instability as this is close to the frequency determined from the unsteady analysis o f the DES results.
CHAPTER 5 . ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 132

However, the source o f the two peaks in the coarse grid results is not so obvious. It is possible that they are also
related to the rotation o f the vortex breakdown winding, but at a location much further downstream o f breakdown.
This was considered due to the results o f the grid refinement study carried out for the DES results in Section 4.4
where two dominant frequencies were found in the wake flow further downstream o f breakdown. However, the
dominant peaks occurred at frequencies slightly lower than those predicted by the coarse grid URANS. It may be
suggested that this predicted behaviour is similar, however, further experimental data in the wake downstream o f
breakdown is needed to confirm the occurrence o f these two frequencies in the unsteady flow.

N«# DimcmiMwl Time StmiiMi Minfeer

(a ) T im e histo ry (b ) P S D vs. St

Figure 5.3: Time history and PSD analysis o f u velocity signal situated above the trailing edge on the vortex axis
at z / c r = 0.1; coarse and fine grid comparisons for ^ — to with Pq, Enhancer model
However, despite appearing to have reasonably predicted the unsteady nature o f the flow, it is clear that the behav­
iour o f the post-breakdown flow for the coarse grid is quite different to the fine grid as shown by the isosurface o f
entropy in Figure 5.1. This is confirmed from considering the behaviour o f the flow on a slice through the vortex
core, as shown in Figure 5.4. For the coarse grid, the location o f vortex breakdown does not appear to be well
defined and is very elongated in appearance. The stretched appearance o f the winding is also evident and it is clear
that it does not have a strong, clear structure at the trailing edge. Looking at the results for the fine grid, it is clear
that in contrast, the location o f vortex breakdown is well defined with clear evidence o f an increase in vortex core
diameter and helix pattern downstream. Smaller structures also exist at the trailing edge, which may cause the
higher frequency content found in the u velocity signal. However, these are dissipated very quickly downstream o f
the trailing edge.

(a ) C o arse ( b ) F in e

Figure 5.4: Slice through vortex breakdown region, on a plane through vortex core, y / s = 0.7 showing instanta­
neous contours o f y vorticity, coarse and fine grid comparisons for A — tu with Pa, Enhancer model
From this analysis, it is clear that the fine grid produces results with greater resolution o f the flow featiu^s, partic­
ularly downstream o f breakdown. The unsteady behaviour downstream o f the breakdown also appears to be closer
to the behaviour expected. Therefore, the fine grid results will be further analysed and compared to the DES results
in a later section.

5.2.2 Non-Linear Eddy Viscosity Model


An identical analysis was carried out for the non-linear eddy viscosity model, using the same grids and com pu­
tational set up. Figure 5.5 show the relative locations o f the probes used for the analysis compared to the vortex
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 133

core location. From these plots the relative locations o f the vortex core and probes appear to be similar for both
solutions. The location o f vortex breakdown is also clear from the isosiufaces, due to the expansion o f the core,
and with further analysis it was found that the mean location occurs at approximately x / c r = 0.77 for the coarse
grid and x / c r = 0.87 for the fine grid results. The difference in location is similar to that found for the Â:— cu model
with P(a Enhancer discussed above, however the predicted breakdown is further downstream for both grids.

VMM#,

(a) C o a rse G rid (b ) F in e G rid

Figure 5.5: Location o f probes though vortex core region compared to instantaneous u velocity contours at each
streamwise location and an isosurface o f entropy at t = 50, coarse and fine grid comparisons for Non-Linear Eddy
Viscosity model

From the instantaneous contours o f u velocity, the improvement in vortex core resolution with grid refinement is
clear. The vortex core appears to be more compact for the fine grid and again there is more evidence o f a structure
in the shear layer under the vortex core. Again, this may be the cause o f the difference in breakdown location
for the two grids. A s with the A: — OJ model with Pa, Enhancer results discussed previously, the behaviour o f the
winding downstream o f breakdown appears to be quite different. For the coarse grid the breakdown is less clear
and the winding is elongated with a lazy helical form. For the fine grid, the behaviour is more compact and the
winding appears to have a smaller pitch angle.

The mean and RMS values o f the components o f velocity are shown in Figure 5.6 through the vortex core region.
However, unlike for the A: — o> model with Pa, Enhancer model solutions, the mean velocities for the coarse and
fine grid are quite different. From the mean u velocity, the reduction o f the velocity as breakdown is approached is
clear. However for the coarse grid this reduction starts much further upstream. Although the breakdown is further
upstream for the coarse grid, it still appears that the onset o f breakdown also occurs much earlier than for the fine
grid results. This may be related to the size o f the vortex core region in relation to the probe location. This was
suggested by the contours o f u velocity discussed above and is confirmed by comparison o f the mean w velocity
which shows that the plane o f the probes crosses the vortex core axis at a point upstream o f the fine grid results (the
change in location from inboard to outboard is indicated by the change in sign o f the mean velocity). As the vortex
core region is larger it is likely that the u velocities predicted for a given location w ill be smaller. Downstream at
the trailing edge the mean u velocity is almost identical.

Considering the RMS velocities, the unsteady behaviour on the two grids is evident. For the coarse grid, just prior
to breakdown, the RMS u velocity increases significantly to a value almost five times that for the previous probe.
This is also evident for the fine grid, but the increase in RMS velocity is not so pronounced. It is likely that this
increase is due to the presence o f the vortex breakdown oscillation in this region. Far upstream and downstream o f
the breakdown location the agreement is good between the grid results, although the fine grid consistently predicts
a higher level o f unsteadiness. For the v and w RMS velocities, it appears that the level o f unsteady behaviour is
very similar between the solutions. However, som e slight differences are clear, particularly for the w RMS velocity,
both upstream and downstream o f breakdown.

The single probe in the post-breakdown flow was also considered for these cases and the resulting time histories
and PSD analysis are shown in Figure 5.7. A s mentioned before, both the mean and RMS velocities at this location
were very similar for the two cases. Looking at the time histories, it is clear that the fine grid exhibits a signal with
a slightly larger amplitude than the coarse grid. This is also apparent from the PSD frequency plot, which shows
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 134

the fine grid signal to have more power. The frequency content o f the two signals are similar with dominant peaks
occurring for the coarse grid around 5/ « 4 and around 5r w 3.6 for the fine grid. These frequencies may both be
attributed to the helical mode instability. There is also similar low frequency and high frequency content, although
the coarse grid consistently predicts the peaks at lower frequencies than the fine grid.

<u>'
(U>l

(a ) u v elocity

<v>

(b ) V v elocity

<W)
<W)

(c) w velocity

Figure 5.6: Mean and RMS velocity components through vortex core; coarse and fine grid comparisons for Non-
Linear Eddy V iscosity model

É ld É M l I
Nm i INflBvaéMwl TIm StTMrfMl OMOWr

(a ) T im e histo ry (b ) P S D vs. St

Figure 5.7: Time history and PSD analysis o f u velocity signal situated above the trailing edge on the vortex axis
at z / c r = 0.1; coarse and fine grid comparisons for Non-Linear Eddy Viscosity model

Despite the similarities in the unsteady u velocity signals between the two grids, there are still differences in the
behaviour o f the flow downstream o f breakdown. Figure 5.8 shows the breakdown region using instantaneous
contours o f y vorticity on a plane through the vortex core. From this it is clear that the behaviour o f the Non-Linear
Eddy Viscosity model on the coarse grid is similar to that o f the k — d) model with Pw Enhancer discussed previ­
ously. The vortex winding downstream o f the breakdown location is very stretched and elongated, as was shown
by the isosurfaces o f entropy in Figure 5.5. This is particularly obvious when compared to the fine grid results
which show a defined breakdown region with a clear helical structure, upstream o f the trailing edge. Again, some
smaller structures are predicted for both cases, which w ill correspond to the higher frequencies in the signal.

A s with the It — tu model with Pw Enhancer, it is concluded that the fine grid results provide a better resolved flow
solution in comparison to the coarse grid results and will, as a results be used for the remainder o f this investigation.
CHAPTER 5 . ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 135

However, it is clear that the coarse grid gave good approximations, particularly to the unsteady frequencies present
in the flow, and this should be kept in mind when considering the relative cost o f the calculations performed.

(a) C o arse (b ) F in e

Figure 5.8: Slice through vortex breakdown region, on a plane through vortex core, y / s = 0.7 showing instanta­
neous contours o f y vorticity, coarse and fine grid comparisons for Non-Linear Eddy Viscosity model

5.3 Effect of Time Step Refinement


Unlike DES calculations, the refinement o f time step and grid size for U R A N S calculations are not inter-related.
However, just as an increase in grid refinement improves the numerical accuracy o f the solution, a refinement in
time step will increase the resolution o f the unsteady behaviour and increase the maximum flow frequency which
can be captured. To consider the effect o f this on the flow behaviour, the solutions obtained using the Non-Linear
Eddy Viscosity model for the fine grid with the baseline time step o f A t = 0.01 were compared to similar results
obtained with a time step o f A t = 0.005. An analysis similar to that used for the grid refinement study was
performed to compare the results.

■ vilodta

(a) At = 0.01 (b) At = 0.005

Figure 5.9: Location o f probes though vortex core region compared to instantaneous u velocity contours at each
streamwise location and an isosurface o f entropy at t = 50, time step comparisons for Non-Linear Eddy Viscosity
model

Figure 5.9 shows instantaneous u velocity contoiu's and an isosurface o f entropy, as before, for both solutions.
From these plots it is clear that there are a number o f differences in the flow solutions. The overall location o f the
vortex core appears to be very similar, however the size o f the core region, the behaviour o f the shear layer and the
vortex breakdown location are all quite different. With a reduction in time step, the size o f the vortex core appears
to increase as suggested by the contours o f u velocity, although the maximum axial velocity is not found to increase
CHAPTER 5 . ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 136

significantly. A lso evident is a difference in the strength o f the shear layer structure found inboard o f the vortex
core region, this is not as clearly defined for the A t = 0.005 time step results. Considering the mean location o f
breakdown, it was found that with a decrease in time step size the location o f breakdown moves upstream from
x /c r = 0.87 for A t = 0.01 to x / c r = 0.81 for A t = 0.005. This may again be attributed to the change in vortex
core behaviour as a compact vortex core suggests a stronger vortex core and thus a delay in breakdown. For the
grid refinement study, these differences were attributed to the improved grid resolution o f the vortex core region.
However, these results suggest that the level o f unsteadiness o f the flow is also important for the prediction o f the
vortex core behaviour and vortex breakdown. The difference in the location o f breakdown and the winding o f the
helical mode instability in the post-breakdown region are also shown from the isosurfaces o f entropy in Figure 5.9.
From this comparison, it is clear that the winding is more elongated for the finer time step in a similar manner to
the coarse grid results shown in the previous section, particularly in comparison to the A t = 0.01 results.

(a ) u velo city

CiM nlw iM L w sltM i (x/c^)

(b ) V v elocity

Cherdwke

(c) w v elocity

Figure 5.10: Mean and RMS velocity components through vortex core; time step comparisons for Non-Linear
Eddy Viscosity model

For all mean velocity components it is clear that far upstream o f the breakdown at x / c r — 0.53 and 0.63, the so­
lutions are in very good agreement. Closer to the breakdown region at x /c r = 0.7 4 the agreement between the
solutions reduces, due to the differences in vortex breakdown location. Further downstream, the finer time step
results exhibit a greater reduction o f velocity at x / c r = 0.84 than the coarse time step results. However, by the
trailing edge region, as found for all the other comparisons, the mean velocities are, again, in agreement. This
is also the case for the other components o f velocity. The largest differences are at x / c r = 0.8 4 for the mean v
velocity and x /c r = 0.7 4 for the mean w velocity. A s before these differences are m ost likely to be associated with
the relative difference in location o f breakdown and the size and strength o f the vortex core region.

Considering the RMS velocities, it is clear that reducing the time step has a significant effect on the unsteady nature
o f the flow, as expected. Upstream o f breakdown, there is an overall reduction in all the RMS velocity components
for the fine time step. However, as the flow approaches the vortex breakdown location there is much greater exci­
tation o f the flow than for the A t = 0.01 results. For the u RMS velocity, it is clear that the level o f unsteadiness
increases upstream o f the breakdown location at x / c r = 0.74, which may be due to the influence o f the motion o f
the vortex breakdown location. This level increases again just downstream o f breakdown to a level greater than
with A t = 0.01, despite its increased distance from the location o f breakdown. Then it reduces to a value less than
the coarse time step results at the trailing edge. The behaviour o f the spanwise and normal RMS velocities are very
similar, with an increase o f unsteadiness downstream o f breakdown, before a reduction to a level below coarse
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 137

time step results. If the location o f breakdown is considered relative to the levels o f unsteadiness o f each case, it
may be suggested that unsteadiness in the post-breakdown flow for the fine time step remains higher for a greater
distance downstream.

Before discussing the unsteady behaviour predicted from the probe data for each case, it is important to consider
the expected levels o f temporal resolution for each case. As discussed in Section 2.6, the maximum frequency
which can be captured in the flow is determined by the time step (or sample rate) o f the signal based on a number
o f criteria. The baseline time step o f A t = 0.01 was used as an example and it was determined that the maximum
non-dimensional frequency would be A = 10 for this case. Applying the same method to the A t = 0.005 time step
and the maximum frequency increases to St = 20. It is important to remember that this is grid independent for
U RA N S. It was shown for the DES results that a maximum Strouhal number o f 4 0 should have been obtainable
but that the grid refinement limited this to approximately St = 10. Therefore, the comparison between the unsteady
behaviour o f the two solutions is very interesting and it is expected that the finer time step will exhibit higher
frequency content than the coarse time step.

To consider this expectation, two probe locations were considered. These were close to breakdown for both cases at
x / c r = 0.8 4 and downstream, above the trailing edge. Figure 5.11 shows the time histories and PSD analysis o f the
u velocity signals at these two locations. The increase in unsteadiness for the A t = 0.005 solution at x /c r = 0.84
can be considered further and it is clear that the finer time step results not only contain low frequency content
associated with the oscillation o f breakdown location, but are also influenced by the helical mode winding, which
w ill cause an increase in unsteadiness in the RMS velocity. However, downstream the content o f the two signals
is very similar, and it is surprising to note that there is similar energy in the higher frequencies at this location for
both solutions. For both locations the lower frequency content is almost identical for the two cases, indicating that
the behaviour o f the vortex breakdown oscillation is unchanged with time step size. The expected higher frequency
content for the finer time step is not apparent and for both solutions there is virtually no energy in the solutions
above approximately 5r = 14 for either case. It is clear from this comparison that, as for the DES results in the
previous chapter, the majority o f the dominant flow features in the post-breakdown flow occur at Strouhal numbers
less than 10 and are not greatly affected by the decrease in time step size.

(a ) x / c r = 0 .8 4

I — A t-ill 1
1--- A t.M W 1

II
•M
M2 . j
LA
JA I

I# I) M W I# »

(b ) x / C r = 1.00

Figure 5.11: Time history and PSD analysis o f u velocity signals for two probes situated on the vortex axis at
z / c r = 0.1 above the wing surface; time step comparisons for Non-Linear Eddy V iscosity model

The structure o f the vortex breakdown region is shown in Figure 5.12 using contours o f instantaneous y vorticity
on the probe plane through the vortex core region. From this comparison, it is clear that the behaviour downstream
o f breakdown is very similar for the two cases. However, there does appear to be slightly more smaller vortical
structures in the flow for the A t = 0.005 solution. Therefore, the resolution o f the expected break up o f the flow
into smaller structures downstream o f breakdown has only been marginally improved for this case. A s with all
other results this behaviour does not appear to continue downstream o f the trailing edge and this is likely to be due
to the rapid reduction in grid resolution in this region for the fine grid as discussed for the DES results in Section
4.4.
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 138

From this analysis o f the effect o f time step on the Non-Linear Eddy V iscosity model U RA NS results, it is clear
that the flow behaviour is sensitive to the level o f predicted unsteadiness in the flow such that the strength o f the
vortex and the location o f breakdown changes. However, the frequencies o f the unsteady phenomena in the flow
do not appear to be affected. The increase in time refinement has allowed the U R A N S turbulence model to capture
a few more small structures in the flow, however, this does not correspond to an increase in the presence o f higher
flow frequencies. Therefore, it may be concluded that for this type o f flow a time step o f A t = 0.01 is adequate to
capture the important frequencies o f the flow.

1 1.1 1.2 1.3 1.4


x /c .

(a) At = 0.01 (b) At = 0.005

Figure 5.12: Slice through vortex breakdown region, on a plane through vortex core, y / s — 0.7 showing instanta­
neous contours o f y vorticity; time step comparisons for Non-Linear Eddy V iscosity model

5.4 Comparison between Non-Linear Eddy Viscosity Model and k —œ


with P(o Enhancer Model
Before making comparisons with the DES results from the previous chapter, it is necessary to consider the relative
behaviour o f the two U R A N S models. Comparisons o f the solutions from the two m odels, the — cu with En­
hancer and the Non-Linear Eddy Viscosity model on the fine grid are shown. The comparisons were made using
the baseline time step. A t = 0.01. Figure 5.13 shows the comparisons o f the mean and RMS velocity components
for the five probes in the vortex core region detailed in the previous sections. The vortex breakdown locations for
these results are x / c r = 0.83 for the it — tu with Po, Enhancer model and x / c r = 0.87 for the Non-Linear Eddy Vis­
cosity model. Comparison o f the vortex core behaviour is shown in Figures 5.1(b) and 5.5(b). From the contoiu-s
o f instantaneous u velocity it is clear that the vortex core region for the Non-Linear Eddy Viscosity model is much
more compact than the A: — to with Pa, Enhancer model. It also appears that the probes sit closer to the vortex core
axis in the spanwise direction for the /t — w with Pa, Enhancer model results. The behaviour o f the shear layer
emanating from the leading edge is also quite different and does not appear to curve upward to form the vortex,
instead an inflection point is evident outboard o f the leading edge where the shear layer suddenly changes direction
inboard to create the roll up into the primary vortex. This is accompanied by a larger and stronger secondary vortex
in this region, which is also suggested from the entropy isosurface.

Considering the behaviour o f the mean velocity components, it is clear that the location o f breakdown is the cause
o f the greatest differences. For the mean u velocity this shows that the level o f axial velocity does not decrease
as significantly for the Non-Linear Eddy V iscosity model for the probes downstream o f breakdown as for the
k — (0 with Pa, Enhancer model. However, upstream o f breakdown and at the trailing edge, the behaviour is very
similar. The mean velocity in the spanwise and normal directions also exhibit similar behaviour. Emphasis o f
the relative size and location o f the vortex cores are obtained by consideration o f the mean w velocity. The Non-
Linear Eddy V iscosity model predicts consistently lower mean values, suggesting that the probe is further from
the core axis and that the vortex is weaker. This means that the vortex core is further inboard for the Non-Linear
Eddy Viscosity model, however the locations are similar in the normal direction. A further appreciation o f the
differences between the solutions predicted by each model may be obtained from analysis o f the RMS velocities
in the vortex core region. It is evident that the velocity fluctuations in the spanwise and normal directions are
greater upstream o f breakdown for the Non-Linear Eddy Viscosity model, but much less in the downstream region.
CHAPTER 5 . ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 139

However, further downstream at the trailing edge, the level o f unsteadiness is similar for the two models. For the
streamwise velocity fluctuations, indicated by Urms, the solutions exhibit similar behaviour a t x / c r = 0.8 4 although
this location is downstream o f breakdown for the A: — û> with Pa Enhancer model and upstream for the Non-Linear
Eddy Viscosity model solution. D espite the differences in breakdown location, it is clear that the effect o f the
vortex breakdown at this location is the same for each case. This is due to this probe being within the oscillating
region o f breakdown for both cases. Upstream o f breakdown, it is again clear that the level o f unsteadiness is
greater for the Non-Linear Eddy Viscosity model solution.

<U> NLEVM
(U) k-m with Enhancer

t = : = = = , -----

(a) u velocity

..

(b ) V velo city

(c ) w v elocity

Figure 5.13: Mean and RMS velocity components through vortex core; comparison o f k — w with Pa Enhancer,
A t = 0.01, and Non-Linear Eddy Viscosity model. A t = 0.01, for the fine grid

(a ) k - ( 0 w ith Pm E n h a n cer (b ) N L E V M

Figure 5.14: Time history and PSD analysis o f u velocity signals for two probes situated on the vortex axis at
z /cr = 0.1 above the wing surface; Comparison o f A — w with Pa Enhancer and NLEVM models for the Fine grid
at A t = 0.01
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BR EA KD O W N 140

To further consider the post-breakdown flow behaviour, a slice through the vortex core region in the plane of the
probes is taken and instantaneous contours of y vorticity are analysed as before. Figure 5.14 shows these results for
both turbulence models. At this instant in time, t = 50, it appears that the location of breakdown is in agreement
for the two solution and it is evident that the winding of the sph al breakdown is clear in each case. From the three-
dimensional flow behaviour, shown by the isosiufaces of entropy, this behaviour also appeared to be very similar.
However from these plots, a number of small differences in the post-breakdown flow behaviour are evident. The
onset of breakdown and the change from a clear vortex to the breakdown spiial appears to be more pronounced
for the k —(ü with P® Enhancer model and the post-breakdown region for the Non-Linear Eddy Viscosity model
solution appears to be smaller in extent in the normal direction. However, the location above the wing is the same.
The streamwise extent of the flow behaviour is also smaller for the Non-Linear Eddy Viscosity model solution,
however both cases clearly show the effect of the decrease in grid refinement downstream of the trailing edge and
the resulting decrease in numerical accuracy in this region. Further consideration of the relative behaviour of the
two URANS models can be obtained from analysis of the unsteady behaviour in the vortex core region in a similar
manner to the DES results described previously in Section 4.5.

5.4.1 Unsteady Behaviour predicted by URANS Solutions


Figures 5.15 and 5.16 show the unsteady flow behaviour at each probe location considered above, for the k —O)
with P(o Enhancer and Non-Linear Eddy Viscosity models respectively. Considering the unsteady behaviour of the
k —03 with P(D Enhancer model in the first instance. For this case, it is found that the probes at x / c,- = 0.53,0.63 and
0.74 are upstream of breakdown, with all probes sitting above the vortex core axis. The probe at x/c,- ~ 0.53 sits
within the shear layer and the probe at x/c,- = 0.63 sits in the region between the vortex core and the shear layer.
As breakdown occurs at x/c,- = 0.83, the probes at x/c,- — 0.84 and 1.0 are within the post-breakdown region.

Considering the flow behaviour in the streamwise direction, the u velocity traces show that, upstream of break­
down, the unsteady oscillations of the velocity have relatively low amplitude, particularly in comparison with the
behaviour downstream of breakdown. At x/cr = 0.53, the trace exhibits a slightly larger amplitude and higher
frequency than for the probes closer to the vortex core, due to its location in the shear layer. This is likely to be due
to the presence of shear layer instabilities. At x/c,- = 0.84, the velocity time history shows a large amplitude, low
frequency oscillation consistent with the fluctuation of vortex breakdown location and it is evident that the break­
down location passes over this position at a number of instances in the time trace as the velocity decreases below
zero, suggesting recirculating flow. A higher frequency clearly exists in this signal also. Further downstream,
at the trailing edge, the low frequency behaviour appears to have disappeared and a higher frequency remains.
Considering the spanwise and normal velocity behaviour and it is evident that these trends are similar for each
component of velocity. However, larger amplitude oscillations are found to occur in the x/cr = 0.84 signal as the
breakdown moves upstream of the probe location.

From analysis of the frequency content of the time tr aces, a number of dominant flow frequencies can be identified.
For the sti-eamwise velocity, it is evident that there are two dominant frequencies in the probes used in this inves­
tigation. At x/c,- = 0.84, the dominant frequencies in the signal appear to be centred around St = 0.07, which has
previously been identified with the oscillation of vortex breakdown location. A second smaller peak is also evident
at approximately St = 3.25 and is associated with the helical mode instability and winding. This is the higher
frequency mentioned above. Further downstream of breakdown, this frequency is also dominant, however appears
to have more energy. With a closer look at the u velocity PSD analysis, further frequencies may be determined in
the signals upstream of breakdown. It was found that the effect of the oscillation of breakdown location was also
mildly felt upstream of breakdown at x/c,- = 0.74. At x/c,- = 0.53, the higher frequencies associated with the time
trace described before were found to correspond to A % 5 —8, which is within the possible frequency range for
shear layer instabilities.

For the spanwise velocity, the helical mode winding frequency dominates the PSD analysis occmring at St ^ 3.25
as before, but with a slightly broader frequency peak. This frequency is most dominant atx/c,- = 0.84, with the en­
ergy at this frequency close to the trailing edge being significantly less. Also present at x/c,- = 0.84 is evidence of a
spanwise oscillation of the vortex breakdown location with a frequency peak again centi'ed at St = 0.07. Upstream
of breakdown, similar low energy frequencies are present in the range & % 5 —8. This pattern is also evident for
the normal velocity, w, with the same frequencies appearing. However, the effect of the vortex breakdown location,
although present, is not as significant. It also appears that the signal at the trailing edge has some content at similar
frequencies as that found for the probe within the sheai' layer, St 5 —1.

A similar analysis can be carried out for the Non-Linear Eddy Viscosity model results. Consideration was given
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 141

---«to- L»

___ _ .. ...............................
• 1 2 3 4 5 4 7 t * l «

(a ) u velocity

l 3 J 4 5 « T t f l «
N#m m m m m à m tl T k m

(b) V velocity

(c ) w v elocity

Figure 5.15: Unsteady behaviour o f non-dimensional velocity components at probes through vortex core region
shown by time histories and PSD frequency plots for A: - to with Pa Enhancer model. A t = 0.01

rmOi Ml I Mi At ÉI Ul I

(a ) u v elocity

.U —

(b) V velocity

----- a te r a t.T 4

---- # 4 . W
! “

• ' * W, •

(c ) w velocity

Figure 5.16: Unsteady behaviour o f non-dimensional velocity components at probes through vortex core region
shown by time histories and PSD frequency plots for NLEVM , AT = 0.01
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BR EA KD O W N 142

to the location of the probes with respect to the vortex core location and it was determined that the probes at
x/cr = 0.53, 0.63 and 0.74 are all in similar locations to those for the k — co with Pa, Enhancer model, upstream
of breakdown and above the vortex core. However, as the mean location of breakdown was found to be further
downstream for this case, the probe at x/c,- = 0.84 is upstream of breakdown. The probe at x/c,. = 1.0 is in the
post-breakdown flow as before. Considering the u velocity time traces, it is evident that there are both similarities /
and differences compared to the Æ—co with Pa Enhancer model results. Upstream of breakdown, at x/c,- — 0.53
the behaviour is similar to the A:- co with Enhancer results, however the amplitude and frequency of the signal
appears to be larger. This is also true for the probes at x/c,. = 0,63 and 0.74. The trend of amplitudes between the 4
upstream probe locations is, however, the same as before. As before, at x/c,- = 0.84 the behaviour is quite different
with a high amplitude, low frequency oscillation being present, superimposed on to a smaller amplitude, higher
frequency fluctuation. This is in agreement with the k — ct) with P^ Enhancer model results. However, from the
signal, it is clear that the breakdown location does not oscillate over this probe position. This indicates the effect
of the vortex breakdown location on the vortex core properties upstream of breakdown, which is not evident from
the k — O) with Pa Enhancer model results. Downstream of breakdown, a higher frequency is again found in the
time history at the trailing edge and the amplitude appears to be of a similar size to the k - O) with Pa Enhancer
model results.

Differences in the flow behaviour are also apparent from the spanwise and normal velocity time traces. Upstream
of breakdown, the behaviour is similar to the k — O) with Pa Enhancer model results, however as before, the am­
plitudes of the signals are larger. This is particularly tiue for the w velocity traces, which have amplitudes which
appear to be 2 to 3 time larger than the corresponding signals from the A:—co with Pa Enhancer model solution. 4
Close to breakdown, however, the behaviour of the unsteady flow appears to be quite different. In the k —û) with
Pa Enhancer model results, the amplitude of the signal from the x/c,- = 0.84 probe exhibited an amplitude mod­
ulation as the breakdown location was found to move over the probe location. Clearly, as this does not occur for
the Non-Linear Eddy Viscosity results, there is no modulation and it is found that the amplitude is considerably
less. At the trailing edge, the behaviour of the time traces appear to show the best agreement with the k — O) with
Pa Enhancer model solutions, although the frequency of the fluctuations does appear to be higher.

As before, the frequency content of these signals was considered from PSD analyses of the time traces. Consider­
ing the frequency content of the u velocity signals, it is again evident that a number of dominant frequencies are
present. Low frequencies associated with the fluctuations of vortex breakdown location are evident, for the signal
at 0.84. In this case, it appears that there are two dominant frequencies, one centred at St % 0.07 and a second oc-
cuiTing at St % 0.6. There does not appear to be much energy at higher frequencies at this location, however, there
is a slight indication of frequencies in the range % 3.5 —4. Downstream of the breakdown location at x/c,- = 1 .0
this higher frequency range % 3.5 —4 is much more dominant, however it has a much reduced energy level than
that found in the k — co with Pa Enhancer model results. This is likely to be the frequency associated with the
helical mode winding as it occurs at a similar frequency as found before. Closer analysis of the probes upstream f
of breakdown, show that there is also little energy in the probes at x/c,- = 0.63 and 0.74, although evidence of a
very small upstream effect of the helical mode winding and vortex breakdown location is found at x/c,- = 0.74,
by changing tire scale of the plot. At x/c,- = 0.53, energy witliin the signal is greater with dominant frequencies
occuiTing in the range St ~ 4 —1. This is in good agreement with the k — CO with Pa Enhancer model results
however, the energy at these frequencies is slightly greater. Similar frequencies are also found in the x/c,- = 0.63
signal, but at a much reduced level. Overall it is found that, with the exception of the frequencies found witliin the
shear layer region, the frequencies predicted for this case are consistently higher than those found for the k —(0
with Pa Enhancer model,

As before, the spanwise and normal velocities show similar frequency content, however there are, again, some
differences compared to the k — co with Pa Enhancer results. The most striking difference is the omission of the
large dominant peak for the x/cr = 0.84 probe at the frequency associated with the helical mode winding. This
is again due to the location of breakdown not moving upstream of this point in the unsteady solution. For the
V velocity, it also appears that there is no evidence of a spanwise motion of the vortex breakdown oscillation at
this location. However, a small peak is clear in the w velocity PSD plot. Downstream of breakdown, however,
this frequency content is clear for the v velocity but not for the w velocity signal. The frequency content for the
helical mode winding, however, occurs for both cases and appears to have a similar level of energy compared
to the k — CO with Pa Enhancer model results. Very low energy frequency content is also found for the probes
upstream of breakdown, x/c,- = 0.53 and 0.63, at the frequencies mentioned for the u velocity, with the dominant
frequencies appear to be higher for tire v velocity PSD analysis than the w velocity, these frequencies are = 4 —7
and St = 3 —5 respectively. As before, these frequencies are likely to be associated with the shear layer behaviour.
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 143

5.4.2 Effect of Eddy Viscosity Treatment


In order to further understand the comparisons and differences between these flow solutions, it is necessary to
consider the differences in formulation o f each model. Both models used in this investigation modify the linear
Boussinesq based W ilcox k — O) model to account for the rotation present in the flow due to the leading edge
vortices. The k — o) with Pa Enhancer model applies a modification which accounts for vortical flows. This mod­
ification enhances the production o f the dissipation in order to reduce the turbulence and the eddy viscosity in the
vortex core. Whereas the Non-Linear Eddy V iscosity model uses an approach derived from a explicit algebraic
Reynolds stress model, which m odels the Reynolds stresses using both strain rate and rotation tensors. This adds
extra terms to the calculation o f the Reynolds stress tensor and results in a non-linear formulation. Further detail
o f each model was given in Chapter 2.

Before considering the relative behaviour o f each model used in this investigation, it is important to consider the
turbulent behaviour o f the baseline model, the W ilcox k — o) model. A similar unsteady calculation was performed
to allow this comparison. Instantaneous contours o f the ratio o f turbulent eddy viscosity to the laminar viscosity
are shown in Figure 5.17 for these results. In the discussion o f the W ilcox k — O) model in Section 2.4.1, it was
stated that the main issue with standard Boussinesq models for vortical flows is that there is an overproduction o f
turbulence within the vortex core region. This is due to the linear dependence o f the Reynolds stress tensor on
the strain rate tensor with no accounting for the rotation o f the flow. A s a result, the levels o f eddy viscosity are
large due to its dependence on k and o) (See Equation 2.58). It is clear from Figure 5.17 that the levels o f eddy
viscosity are indeed very high in the vortex region above the wing surface and that there is no distinction between
the core region, shear layer or breakdown region. As discussed before, this generally results in the prediction o f
a very weak vortex system, which is sensitive to instabilities in the flow. The over-prediction o f turbulence also
causes the unsteadiness o f the flow to be dissipated due to an increase in turbulent m ixing and the solution becomes
unrealistically steady in nature.

■s -

Figure 5.17: Slice through vortex breakdown region, on a plane through vortex core, y / s = 0.7 showing instanta­
neous contours o f f i r / p for W ilcox k — w model

Figure 5.18 shows similar contours o f the turbulent behaviour through the vortex core for the A: — to with Pa En­
hancer and Non-Linear Eddy Viscosity models. A lso shown are contours o f instantaneous turbulent kinetic energy
for comparison. With the Pa Enhancer applied to the A: —co model, the levels o f eddy viscosity are found to reduce
in the vortex core. This is due to the enhancement o f co in regions o f high rotation as described in Section 2.4.2,
reducing the production o f turbulence. The eddy viscosity is calculated in the same way as the W ilcox A: — co
model, thus as co is increased and k is reduced, the eddy viscosity also reduces. It is clear from Figure 5.18(a), that
comparably high regions o f turbulence still exist within the shear layer region and downstream o f the vortex break­
down location. However, even in these regions the levels o f turbulence are reduced by two orders o f magnitude
compared to the standard W ilcox k — o) m odel. This is evident from the comparison o f the eddy viscosity contours
in these regions.

Considering the turbulent contours for the Non-Linear Eddy Viscosity model, it is clear that the overall behaviour
is quite different. Upstream o f the trailing edge, there is little evidence o f turbulence in the flow, with both the
ratio o f turbulent eddy viscosity to laminar viscosity and the turbulent kinetic energy exhibiting values close to
zero in this region. It appears that the levels o f these variables only increase in the shear layer region o f the flow
downstream o f the trailing edge. A s stated previously, breakdown is found to occur at approximately x/c^ = 0.87
and it would be expected that the flow would be turbulent downstream o f this location. However, this clearly does
not occur immediately. However, despite this, the vortex core region is laminar, which is the most important factor
in the prediction o f the flow behaviour, as described previously. The reduction o f the eddy viscosity and turbulence
in the vortex region was expected from the formulation o f the model. The extra anisotropy term o f the Reynolds
CHAPTER 5 . ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 144

stress equation reduces the production o f turbulence in the flow and the eddy viscosity o f the model is reduced,
particularly in the vortex core regions, due to the dependency o f the C ^ term on the rotation o f the flow. However,
it appears that this non-linear modification o f the model provides a virtually laminar solution throughout the entire
vortex region.

(a ) k - ( a w ith Pg, E n h a n cer

1 1.1 IJ IJ lA

(b ) N L E V M

Figure 5.18: Slice through vortex breakdown region, on a plane through vortex core, y / s = 0.7 showing instanta­
neous contours o f P r / p and turbulent kinetic energy for both U RA NS m odels, fine grid. A t = 0.01

(a) k - ( 0 w ith Pg, E n h a n c e r (b ) N L E V M

Figure 5.19: Slice through vortex region a tx /cy = 0 .84 showing instantaneous contours o f p r / p for both U RA NS
m odels, fine grid. A t = 0.01

To further consider the turbulent behaviour through the vortex region and particularly to consider if there are re­
gions o f turbulence apparent in the Non-Linear Eddy Viscosity model solution, a slice was taken through the vortex
at x / c r = 0.84. Figure 5.19 shows this slices with instantaneous contours o f P r / p , as before. This plane is just
downstream o f breakdown for both solutions. For the k — o) with Pa Enhancer model, the widened laminar vortex
core region is clear with higher levels o f eddy viscosity found in the shear layer and a smaller low viscosity region
evident which corresponds to the secondary vortex core region. For the Non-Linear Eddy Viscosity model solution,
the behaviour is again very different, however it is clear that turbulence exists in the solution, in the shear layer
close to the leading edge and within the secondary vortex region. However, this does not extend around the vortex
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 145

region and the flow is virtually laminar in all other regions.

Returning to the comparisons between the two m odels discussed previously and the fact that the Non-Linear Eddy
Viscosity model is predicting a very laminar flow may explain som e o f the differences wim essed between the two
solutions. For example, it was found that the location o f the primary vortex was further inboard for the Non-Linear
Eddy Viscosity model solution than for the A: — tu with P«, Enhancer results. This is due to the larger secondary
vortex also noted in the previous section. The larger, stronger secondary vortex occurs as a result o f the laminar
behaviour within the boundary layer and the secondary separation caused by an adverse pressure gradient. As
mentioned before, Hummel [48] showed that a laminar separation causes a larger and stronger secondary vortex
than a turbulent separation. Further evidence o f this behaviour can be obtained from consideration o f the surface
streamlines, as shown in Figure 5.20. As stated in the description o f the calculations given in Chapter 4 a forced
transition was set on the grid at a streamwise location o f approximately x / c r = 0 .3 6 and therefore it is assumed
that transition w ill occur just downstream o f this location where the turbulence model is active. Considering the
surface streamlines for the A: — £U with Pa Enhancer model, it is clear that this is the case. This is indicated by the
outboard inflection o f the secondary separation line, as separation will occur earlier for a laminar boundary layer
for a given adverse pressure gradient. However, for the Non-Linear Eddy V iscosity model results, this inflection
o f the secondary separation line does not occur until approximately x /c r = 0.6 4 . This suggests that the increase
o f P t in the flow is too gradual, resulting in a delayed transition to turbulent flow. The strong secondary vortex is
also evident from the surface pressure coefficient contours shown.

(a) k - ( o w ith Pa, E n h a n c e r (b ) N L E V M

Figure 5.20: Surface streamlines showing comparable behaviour o f secondary separation line after transition to
turbulence at x /c r % 0 .36 for both URA NS m odels, fine grid. A t = 0.01

Reconsidering the unsteady behaviour o f the solutions, som e o f the differences in the predictions may also be
attributed to the levels o f turbulence within the vortex region. It was found that the vortex breakdown oscillation
predicted by the Non-Linear Eddy Viscosity model exhibits a greater upstream influence on the flow behaviour
than for the it — tu with Pa Enhancer model. This is likely to be due to the decrease in dissipation and mixing
which com es with a laminar flow, resulting in the effects o f a disturbance to be felt further upstream than for a
turbulent flow. Thus, the eddy viscosity o f the A: — o) with Pa Enhancer model dissipates these fluctuations. This
increase in influence results in a higher energy o f the peak predicted by the PSD analysis for the dominant fre­
quency o f breakdown. However, downstream o f breakdown the behaviour changes and the levels o f energy in the
dominant peaks reduce compared to the A: — tu with Pa Enhancer model solutions. This may also be due to the
levels o f turbulence in the flow solution. For low levels o f eddy viscosity and turbulence, there will be much less
turbulence mixing compared to the A: — to with Pa Enhancer model solutions. This acts to smooth the gradients o f
the mean flow fluctuations, resulting in a lower energy for these frequencies.

■4
:
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKD OW N 146

However, despite the largely laminar behaviour of the flow and the differences compared to the A: m with Pa
Enhancer model described, the Non-Linear Eddy Viscosity model exhibits reasonable results, with similar domi­
nant frequencies and behaviours. This furtlier conflrms the suggestion that the level of turbulence predicted is not
a significant factor in predicting the major flow phenomenon downstream of vortex breakdown. However, it may
serve to show that with a better resolution of the turbulence in the flow, the dominant frequencies will be lower.
However, furtlier investigation would be needed to state this conclusively.

From these comparisons and the discussion of the performance of each model it may be concluded that for this
investigation that the k ~ û ) witli Pa Enhancer model has predicted the flow behaviour more accurately. These
solutions will now be compared with the DES results detailed in the previous Chapter to asses the ability of the
model to predict the unsteady vortex flow.

5.5 Comparison of URANS and DES


Having made comparisons with other URANS models and concluded that the behaviour of the solution using the
A—CL»withPco Enhancer model on the fine grid witli a time step of A? = 0.01 is reasonable, it may now be compared
to the results of the DES investigation detailed in the previous Chapter. This comparison will provide a measure of
the applicability of this linear URANS model with rotation correction to unsteady delta wing vortical flows. The
DES calculations discussed in detail in Chapter 4 were performed using the fine grid with refined trailing edge
region, as it was found in Section 4.4 that an increase in refinement in the trailing edge region slightly improved
the resolution of the turbulence and unsteady behaviour of the flow. However, for the URANS calculations, this
refinement is not necessary, as the solutions are only numerically dependent on the grid refinement. Therefore, as
stated in the previous sections, all URANS calculations were performed using the fine grid. To allow for a fair com­
parison and to keep the grid consistent, the DES solution on the fine grid will be used in this section for comparison.

Figure 5.21 shows the comparison of the mean and RMS velocity components for the two turbulence treatments
in the same manner as before. It is clear from these plots that overall the solutions are in reasonable agreement.
Considering the mean u velocity, it is clear that the k~ co with Pa Enhancer model predicts values which are lower
than the DES results for all probe locations. It is also evident that tlie mean location of breakdown is predicted to
be slightly further upstream, which is likely to be due to the prediction of a lower core velocity upstream of break­
down. The mean breakdown location for the k —co with Pa Enhancer model solution is approximately x/cr = 0.83
and for the fine grid DES results it is approximately x/c,. = 0.85. The RMS u velocity shows good agreement for
all probe locations except the probe closest to breakdown at x/c,- = 0.84. It is clear that there is considerably more
unsteadiness in the flow for the DES solntion at this location. This may be due to greater fluctuations of the vortex
breakdown location in the streamwise direction for tire DES solution.

The mean and RMS spanwise velocity show very good agreement between the two solutions, showing that the
location above the wing is the same for each solution. However, there is a consistent difference in the mean w
velocity predictions. This shows that the A—ct) with Pa Enhancer model predicts a higher normal velocity suggest­
ing that either the core region is larger than for the DES results or that the vortex sits slightly further inboard. As
this difference is consistent both upstream and downstream of the breakdown location, it may be suggested that it
is the location of the vortex core which is different. The RMS w velocity shows that there is more unsteadiness
predicted for the DES model upstream of breakdown in this direction, but that close to breakdown the k —o) with
Pa Enhancer model results exhibit a higher unsteadiness. Downstream of breakdown, the levels appear to be tlie
same for all RMS velocity components.

To consider the relative post-breakdown behaviour, instantaneous contours of y vorticity are shown in Figure 5.22
through the vortex core region. It is clear that at the instant compared, the location of breakdown is very similar
for the two solutions. This occurs at approximately x/c,- = 0.80 for the k — O) with Pa Enhancer model solution
and just upstream of this location for the DES solution. The breakdown appears to be similar in form for both
solutions with a sudden change in the behaviour of the vortex core. Downstream of breakdown, the behaviour is
also very similai* with the vortex core winding evident in both solutions. However further downstream it is clear
that more smaller structures exist in the DES solution. Whereas the A—co with Pa Enhancer model results show
some structures, however these appear to be smeared across the grid in the trailing edge region. It is also evident
that the shear layer is clearer in the DES solntion both upstream and further downstream of the trailing edge.
However, the area covered by the breakdown region is the same for each solution.
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 147

...

(a) u velocity

<V)k-®'
<V>DES

(b ) V v elocity

- # - <W> It-n) wlUi p_ Enlw im r


(W)

CherdwW l.«catlMi (»/e_)

(c) w v elocity

Figure 5.21: Mean and RMS velocity components through vortex core; U R A N S k — d) with Enhancer model
compared to DES solutions

oiuclty

0.7 0.8 0.9

(a ) k - ( 0 w ith A , E n h a n c e r A t = 0.01 (b ) D E S A t = 0 .0 0 2 5

Figure 5.22: Slice through vortex breakdown region, on a plane through vortex core, y / s = 0.7 showing instanta­
neous contours o f y vorticity at t = 50, for U R A N S k — ct) with Pa Enhancer model and DES

5.5.1 Comparison of Unsteady Flow Behaviour Prediction


Figures 5.23 and 5.24 show the time histories and PSD analyses for the k — d) with Pa Enhancer model and DES
solutions, respectively. Considering the time histories initially, the amplitude and the unsteady behaviour o f each
component o f velocity can be considered. It is clear that upstream o f breakdown, at x /c r = 0.53, 0.63 and 0.74
probe locations that the amplitude and oscillation o f all three components o f velocity are very similar for each
model. However, at x /c r = 0.7 4 for the u velocity, it is clear that for the D ES solution, the oscillation o f break­
down appears to have a more significant effect than for the k — d) with Pa Enhancer model. The behaviour at
CHAPTER 5 . ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 148

j L .......... .......................

(a ) u velocity

'L - . j à
N m Dimewiewti JUm

(b ) V v elocity

m riidid iii—b»r

(c ) w v elocity

Figure 5.23: Unsteady behaviour o f non-dimensional velocity components at probes through vortex core region
shown by time histories and PSD frequency plots for A: — o) with Pa) Enhancer model. A t = 0.01

i Vn \i i! l! \i

(a) u v elocity

---------------------- h.---- ---- 1-----------1 _


N«n Ttam

(b ) V velocity

Nm UtaMiiiriMMl T im

(c ) w velocity

Figure 5.24: Unsteady behaviour o f non-dimensional velocity components at probes through vortex core region
shown by time histories and PSD frequency plots for DES, A T = 0.0025
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 149

x /c r = 0.8 4 is also slightly different for the DES solution with the location o f breakdown clearly sitting down­
stream o f the probe location for almost half o f the signal, then it m oves upstream and seem s to oscillate over the
probe location as the velocity appears to oscillate about zero. For the A: — to with Po) Enhancer model solution,
breakdown appears to oscillate over the probe location for the whole signal length, although it is clear that this is
not the mean location. The change in behaviour for the DES solution also appears for the v and w velocity traces,
with an increase in amplitude evident. The results from the /t —o) with Pa) Enhancer model calculation exhibits this
larger amplitude for a larger portion o f the signal, which appears to occur when breakdown is close to or upstream
o f the probe location. Downstream at the trailing edge the behaviour is again very similar.

Considering the frequency content o f the probe signals for the two models, it is clear that the magnitude o f the fre­
quencies are very similar. The PSD analyses o f the velocity components show that the oscillation o f the breakdown
location and the frequency associated with the helical mode winding are both present. These occur at St = 0.07
and St = 3.25, respectively for both models. However, the power o f these frequencies within the signals are quite
different. The energy in the oscillation o f breakdown frequency is much larger for the DES solution compared
to the A — Ü) with Pa) Enhancer model results. For the u velocity, the energy o f the DES oscillation is almost ten
times larger. However, the energy o f the helical mode instability frequency is consistently larger for the it - w
with Pa) Enhancer model solutions. Higher frequency content at St = 5 — 7 is also present in both solutions. The
agreement between the solutions can be seen more clearly by directly comparing the signal from a single probe in
the flow. Figure 5.25 shows the u velocity time histories and PSD analysis from the probe above the trailing edge
for each solution. These plots further confirm the discussion given above. The time histories show that although
the signals behave differently with time, it is clear that the amplitude and oscillation o f the signals are very similar.
Considering the PSD analysis, this highlights that the frequencies present in the signals are almost identical, with
the main differences being due to the relative energy o f each frequency. It is clear that the energy o f the higher
frequencies are the same.

From this unsteady analysis and the analysis o f the mean flow behaviour, it is evident that the URANS model is
capable o f predicting the same dominant flow features and frequencies as the DES model.

(a) (b)

Figure 5.25: Time history and PSD analysis o f u velocity signals for a probe situated on the vortex axis at z/Cr = 0.1
above the wing surface for URA NS k — o) with Enhancer and DES solutions

5.5.2 Effect of Eddy Viscosity Treatment


The relative behaviour o f the solutions can also be considered in light o f the turbulence treatment o f each model. In
order to consider the relative prediction o f the turbulent behaviour by each model, the eddy viscosity in the vortex
region was analysed. Figure 5 .26 shows instantaneous contours o f the ratio o f eddy viscosity to laminar viscosity
for both turbulence treatments. Due to the under-resolution o f the turbulence for the DES solution, as discussed
in Chapter 4, the behaviour o f the subgrid scale eddy viscosity will be very similar to the U RA NS turbulent eddy
viscosity and so a comparison is valid.

It is clear from Figure 5.2 6 that the distribution through the vortex region is quite different for the two solutions.
The behaviour o f the eddy viscosity o f the A: — to with Pa) Enhancer model was described in Section 5.4 and sim i­
larly the role o f the subgrid eddy viscosity in the DES calculations was discussed in Section 4.8.2. It is clear that
in comparison that the DES model predicts much lower eddy viscosity in the vortex region, although the pattern o f
the contours is very similar. This reduction, as discussed previously is due to the dependence o f the subgrid eddy
viscosity on the grid dim ensions. The region o f high viscosity downstream o f the trailing edge is, therefore, due to
the reduction o f the grid refinement in that region. Both m odels predict higher levels o f viscosity in the shear layer
and predict a laminar vortex core region. Downstream o f breakdown, the behaviour is also quite similar, with an
increase in eddy viscosity levels in the post-breakdown flow. This is widespread for the DES solution, however the
CHAPTER 5 . ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKDOW N 150

k — û) with P(o Enhancer model predicts more localised regions o f high viscosity.

The higher levels o f viscosity predicted by the A: — w with Po) Enhancer model may explain the increased energy
o f the helical mode frequency discussed above since an increase in eddy viscosity com es an increase in turbulent
production and therefore turbulent m ixing, which will smooth out fluctuations on the unsteady mean flow. There­
fore, more energy w ill exist for the mean flow oscillations. However, the converse is true for the vortex breakdown
oscillations, which are shown to have more energy in the DES solutions. This is likely to be due to the reduction o f
eddy viscosity in this region, which means that the effects o f the breakdown fluctuations will be felt more strongly.
However, it is important to note that despite the differences in eddy viscosity distributions and levels through the
vortex region and in the post-breakdown flow, the frequencies and behaviour predicted are very similar for both
solutions.

»A «.7 ## 1.1 IJ IJ lA

(a) DES, At = 0.0025 (b) k - w with A) Enhancer model, At = 0.01

Figure 5.26: Slice through vortex breakdown region, on a plane through vortex core, y / s = 0.7 showing instanta­
neous contours o f ^ t / P for URA NS k - o ) with Pw Enhancer model and DES

5.6 Discussion
Before discussing the ability o f U RA NS to predict the unsteady behaviour o f delta wing vortical flows it is nec­
essary to review the turbulent features o f vortical flows. In Chapter 2 the application o f each turbulence model
used in this investigation was discussed with reference to delta wing flows. It was stated that it was necessary
that each model was able to predict a laminar vortex core region with higher turbulence production occurring in
the shear layer and downstream o f the vortex breakdown. Unfortunately, limited data exists to quantify the exact
levels o f turbulence within this type o f flow, therefore it is difficult to exactly measure the ability o f each model
to accurately predict the turbulence. However, from the formulation o f each model and the predicted solution,
it is possible to determine the relative behaviour o f each model and qualitatively assess the ability to predict the
turbulent behaviour accurately. This is further aided by validation o f the predictions with available experimental
data as performed for the DES solution in Chapter 4.

From the discussion o f the formulation o f the linear Boussinesq W ilcox k — (o model and the contours o f turbulent
eddy viscosity shown in Figure 5.17, it is evident that this model over-predicts the turbulence within the vortex
region and particularly through the vortex core. This has the effect o f creating a weak vortex, which has a sig­
nificantly increased susceptibility to breakdown. A lso, with the increased turbulence, the ability to capture the
unsteady behaviour is diminished. This is due to the significant increase in dissipation o f the unsteady fluctuations
o f the mean flow, which causes the flow to becom e steady in nature. To reduce the turbulence within this model,
rotation corrections may be applied to sensitise the model to the rotation o f the vortex flow. This was explained
in Chapter 2 for the t - O) with Pa, Enhancer model. Similarly, a non-linear model can be formulated, which pro­
vides a more general improvement to the W ilcox k — O) model by including further terms to the calculation o f the
Reynolds stress anisotropy based on both the rotation and strain-rate tensors. Both methods reduce the tiu'bulence
within the flow and result in reduced dissipation o f the unsteady behaviour, improving the prediction o f the vortex
system.

Having considered all U RA NS solutions and the comparison with the validated DES results from the previous
Chapter, it is possible to discuss the ability o f URANS to predict the vortex flow system over the wing. It is
clear from the comparisons shown in the previous section, that the — cu with Pa, Enhancer model adequately
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKD OW N 151

predicts the mean flow unsteady behaviour as defined in Chapter 4 for the D ES. This includes predicting the vortex
breakdown oscillation and helical mode frequencies accurately compared to the D E S solution for the same grid. It
is also evident from the comparison between the two U RA NS models that the Non-Linear Eddy Viscosity model
also predicts these frequencies. However, it is clear from analysis o f the mean properties o f the flow that the
predicted location o f breakdown is different for each turbulence treatment. This is also true in comparison with
the D ES solutions. To consider this further, the mean breakdown locations for each calculation considered in this
investigation are summarised in Table 5.1.

T urbulence T reatm ent G rid AT M ean VBD location, x /c r


k —O) with Fo) Enhancer Fine 0.01 0.83
NLEVM Fine 0.01 0.87
NLEVM Fine 0.005 0.81
DES Fine 0.01 0.88
DES Fine 0.005 0.8 4
DES Fine 0.0025 0.85
DES Refined TE Grid 0.0025 0.86

Table 5.1: Location o f mean vortex breakdown for each unsteady calculation performed in this investigation

It is clear from Table 5.1, for both the U R A N S and D ES solutions, that with a decrease in time step size, there is
an upstream shift in mean breakdown location. This suggests that the location o f breakdown is dependent on the
resolution o f the temporal behaviour o f the flow. However, from the D ES results it is evident that the location w ill
converge to a constant value as the time step is reduced. Although this value, for the D ES solutions is dependent
on the grid refinement, it may be suggested that a similar behaviour would be exhibited by the U RA NS models
for further decreases in time step size. It was also found that an increase in grid refinement had the opposite effect
for the U RA NS solutions and moved the location o f breakdown further downstream. This is also the case for
the DES solution with refinement in the trailing edge region, although the change in the mean location is small.
This suggests that with an increase in the spatial resolution o f the flow, both upstream and downstream, the mean
breakdown location m oves downstream. The effect o f the turbulence treatment is a little harder to consider. Figure
5.27 shows the instantaneous ratio o f eddy viscosity to laminar viscosity through the vortex core for each model
at the instant T = 50. From this plot, it is clear that each model keeps the eddy viscosity close to zero through
the vortex core region upstream o f breakdown, however at the location o f breakdown the turbulence in the flow
increases. It is clear that the - ft) with 7^ Enhancer model predicts the largest eddy viscosity values downstream o f
breakdown and the furthest upstream breakdown location, similarly the Non-Linear Eddy V iscosity model predicts
the low est eddy viscosity ratio, but the furthest downstream mean breakdown location. This suggests that there
is a link between the turbulence predicted downstream o f breakdown and the location o f breakdown. However,
from the values o f mean breakdown location given in Table 5.1 it is clear that the differences in predicted mean
breakdown location between each solution in this investigation is only approximately 7%Cr. Therefore, it may be
suggested that provided the vortex core is predicted as being laminar, the levels o f turbulence predicted in the flow
downstream has som e effect on the location o f the mean location o f breakdown but little effect on the unsteady
behaviour o f the post-breakdown flow.
350
k-w w ith P , Enhancer A t = O.Ol
300 NLEVM At «0.01
DES Fine Grid Ax = 0.0025

100

Figure 5.27: Distribution o f pLj through vortex core for all turbulence m odels used in this investigation

One o f the factors driving the use o f U R A N S for unsteady vortical flows, which was mentioned in the introduc­
tion o f this chapter and has been mentioned in previous discussions is the relative computational expense o f the
calculations in comparison with D ES methods. Table 5.2 shows the length o f the calculations required to simulate
one second o f real time for each calculation carried out in this investigation. From this data, it is clear that the
URA NS calculations are at a minimum four times cheaper than the D ES calculations for the same grid on h alf as
many processors. The reason that the Non-Linear Eddy V iscosity model is more expensive than the &— ft) with 7^
Enhancer model is due to the requirement o f a reduced unsteady convergence limit, thus increasing the work unit
o f the calculation, which is defined as the non-dimensional time taken to reach convergence for one time step o f
CHAPTER 5. ASSESSMENT OF URANS FOR PREDICTING VORTEX BREAKD OW N 152

the calculation. It is important at this point to also note that it was concluded in the previous chapter, that the DES
calculations were under-resolved. Therefore, to fully resolve the turbulent scales of the flow, the computational
resources required would be significantly larger. This clearly shows the advantage of using URANS to capture the
flow details.

No. o f Total No. Work A pprox. Total


T urbulence T reatm ent G rid At Processors of At Units R un Time (hrs)
k —COwith Pco Enhancer Fine 0.01 24 7,158 40,228 (5.62) 500
NLEVM Fine 0.01 24 7,158 60,843 (8.50) 1000
NLEVM Fine 0.005 24 14,316 121,686 (8.50) 2000
DES Fine 0.0025 48 28,632 161,485 (5.64) 8000
DES Refined TE 0.0025 48 28,632 161,199 (5.63) 8000

Table 5.2: Length of calculations for each turbulence treatment used in this investigation to predict a total time of
one second. Work units in brackets denote unit for one time step

5.7 Conclusions
It may be concluded from this investigation that URANS turbulence models perform well in predicting the neces­
sary features of the unsteady vortical flow and vortex breakdown provided the turbulence in the vortex core is kept
low. It is clear from these results that the majority of the frequencies and phenomena predicted by each URANS
model is in good agreement, although the mean location of breakdown is found to change. The effect of grid and
time step resolution was considered, and it was found that the mean behaviour of the flow is more sensitive to
these issues than the unsteady phenomena. Comparing the linear model with the rotational conection to the DES
solutions from the previous chapter shows that the predicted unsteady behaviour is again very similar, with the
majority of the differences occurring in the mean location of breakdown. Therefore, due to the validation of the
DES solutions, it may be suggested that the URANS model is capturing the flow behaviour well with all the main
dominant frequencies being present in the solutions. It was shown in the previous chapter that the DES solutions
were not well resolved spatially, however it was also shown that the resolution of the turbulent scales was not
important to the prediction of the main flow features. This investigation shows that due to this the URANS models
were able to predict the main features of the flow.

It was found that the mean behaviour of the flow is more sensitive to the turbulence treatment, grid and time step
size than the frequencies of the unsteady oscillations. However, from analysis of the mean breakdown location, it
was found that this difference was limited to 7%c,-, which is relatively small. Therefore, it may be concluded that it
is more important to accurately predict the turbulent behaviour in the vortex core than downstream of breakdown.
The resolution of the post-breakdown flow does have a small effect on the mean breakdown location but does not
appear to impact the frequency of oscillation of breakdown in the flow or the frequency of the helical mode winding.

Overall, it may be concluded that URANS is suitable for use in capturing the unsteady behaviour of delta wing
flows at moderate incidence where vortex breakdown occurs, provided the core behaviour is modelled accurately.
It has also been shown that this may be performed at considerably less computational expense than DES methods
and thus Is a promising tool for industrial use in the prediction of vortical flows.
Chapter 6

Conclusions and Future Work

The main conclusions drawn from this investigation are now summarised and recommendations for future research
given.

6.1 Conclusions
Within the transonic regime it has been found that shocks interact with the leading edge vortices. Vortex breakdown
is found to occur in an abrupt nature and this can have a significant effect on the aerodynamic performance. Due to
this, one of the aims of this project was to consider tlie behaviour of vortex breakdown within a transonic flow and
to consider the predictive ability of RANS methods. Steady state calculations were compared to experimental data
which showed very good agreement for the pre-breakdown flow. However, for a larger incidence a discrepancy
between the CFD and experimental results appeared due to tlie premature occurrence of vortex breakdown in the
computational results.

Analysis of the flow behaviour resulted in the identification of a number of shocks which could be classified into
two main shock types, cross-flow and normal. The locations and behaviour of these shocks was found to agree with
observations in the literature. The normal shocks, which occuiTed normal to the wing surface and symmetry plane,
were found to interact with the vortex core and were determined to cause the sudden appearance of breakdown.
A sensitivity study was performed to consider the effect of a number of factors on the predicted behaviour. These
included, grid refinement, turbulence model, solution convergence and time accuracy. However, it was found that
breakdown was consistently predicted and was not significantly affected by any of these factors. Comparisons
were also made with other structured grid results from participants within the VFE-2, but again the solutions were
found to be comparable.

From consideration of the interaction between the normal shocks and the vortex core, it was suggested that a criti­
cal limit must exist where breakdown will occur. This limit was concluded to be dependent on the strength of the
vortex and the interacting normal shock. Using this argument of a critical limit for breakdown, it was concluded
that the premature breakdown behaviour of the computations was due to under-predicting the vortex core axial
velocity accurately most likely due to grid refinement issues in this region. However, overall it was found that the
computational results were adequately predicting the transonic behaviour of the vortex flow.

The unsteady behaviour of the flow is a second aspect of delta wing flows which requires further investigation.
At moderate angles of incidence where breakdown occurs on the wing, the flow becomes highly unsteady and
interactions between the flow and aircraft structures have been found to occur. To avoid aeroelastic issues, it is
necessary to have a greater understanding of the unsteady phenomena which occur. This is becoming increasingly
important with the emergence of UAV technologies. The second aim of tliis tliesis was to consider the ability of
CFD to predict the main unsteady behaviour of the flow. In Chapter 4 the use of DES to predict the unsteady flow
behaviour associated with the flow upstream and downstream of breakdown was considered for the ONERA 70*^
delta wing geometry at a moderate incidence within the subsonic regime. Before the predicted unsteady behaviour
was considered, the effect of time step and grid refinement at the trailing edge were analysed. This determined the
optimum time step for use in tlie calculations and that there was only a small influence of grid refinement in the
trailing edge on the mean flow behaviour.

From analysis of the unsteady flow behaviour and consideration of the level of turbulence captured in the un­
steady signals, it was determined that the solutions obtained in the investigation were spatially under-resolved.

153
CHAPTER 6. CONCLUSIONS A M ) FUTURE W ORK 154

This resulted in the conclusion that the spatial and temporal requirements of a fully resolved DES calculation of
the post-breakdown flow are significantly larger than those used in this investigation and that further refinement
downstr eam of the trailing edge would be needed to capture all the turbulent behaviour in the post-breakdown flow.
This would have the effect of increasing the computational expense of an already expensive calculation.

However, from comparison of the results with other DES solutions on finer grids, it was determined that the un­
steady vortex breakdown behaviour was not dominated by turbulence with the dominant frequency occurring for
less than St = 10, which is low. It was also found that the results were in good agreement with corresponding
experimental results. Therefore, despite the under-resolution of the turbulent flow, the salient features of the flow
were being captured well. As the main unsteady phenomena were found to occur at low frequencies, and turbu­
lence was not found to be dominant in the flow, it was concluded thatURANS turbulence models should be able
to adequately predict this behaviour for a considerably reduced computational cost.

To investigate this conclusion, the ability of UR ANS to predict the unsteady flow behaviour was evaluated in Chap­
ter 5. Two UR ANS turbulence models were considered for this investigation, a linear Boussinesq model with a
rotation correction and a Non-Linear Eddy Viscosity model, both based on the Wilcox k —a) model. The effect of
grid refinement and time step refinement were considered and it was concluded that with an overall increase in grid
refinement, the solutions were found to improve, however the coarse grid results were adequate for an approxima­
tion of the flow at low computational expense. From the time step refinement study, it was found that the increase
in temporal resolution did not have a significant effect on the unsteady behaviour predicted, despite an upstream
movement of the mean breakdown location. Higher frequencies were also not found to occur with a decrease in
time step and so it was concluded that the baseline time step of A t —0.01 was suitable for UR ANS calculations.

Comparisons were then made between the URANS models and the relative behaviour of each model in predicting
the unsteady flow frequencies was analysed. This was also considered in light of tlie formulation of the models and
the treatment of the turbulence in the vortex core region. From this comparison it was concluded that both models
were adequate in reducing the eddy viscosity in tlie vortex core and that similar unsteady behaviour was predicted.
The model with the rotation correction was then compared to tlie DES results discussed in Chapter 4 to evaluate
the mean and unsteady behaviour of the solutions. From this comparison, it was clear that the majority of the
dominant frequencies of the vortex flow were captured by the URANS model and the agreement with the DES so­
lutions was very good. From this analysis it was concluded that the URANS model had predicted all the important
unsteady features of the flow and was, tlierefore, suitable for use in predicting the unsteady nature of vortical flows.

It was also determined from this investigation that the mean behaviour of the flow, such as tlie mean location of
breakdown, is far more dependent on the turbulence treatment of the models used that the unsteady behaviour. All
turbulence treatments used predicted similar dominant frequencies and unsteady phenomenon. However, tlie mean
location of breakdown was found to be different for each case. This difference was found to be approximately
7%c,; which was considered relatively small. Therefore, it was concluded that provided the core region of the
vortex flow is modelled as laminar, the turbulent treatment of the model used does not have a significant influence
on the overall flow behaviour. If this is considered in the context of the computational expense of each model used,
it is clear that URANS can predict the main features of the flow for a significant reduction in computational cost.

Overall, from this investigation, it can be concluded that CFD is a very useful tool for the prediction of vortex
flows and vortex breakdown over slender delta wings and that it is capable of predicting complex flow behaviour,
such as tiansonic vortex breakdown and the unsteady nature of the flow. In this study both RANS/URANS and
DES methods were considered and it is clear tliat both methods can be used to predict the flow accm ately. However,
some limitations of these methods have also been highlighted.

6.2 Future Work


Throughout this project a number of potential avenues for further work have presented themselves.

From Chapter 3, the main avenue for further work would be to consider the flow for different configurations and
flow conditions to attempt to define a limit for vortex breakdown based on the Rossby number of the vortex and
the shock strength. Further experimental data is also required to validate this limit and this would require measure­
ments taken across shocks and through the vortex core for a range of flow conditions. This would also allow further
validation of the location and strengtli of the shocks in the flow, improve their prediction and therefore improve the
understanding of their behaviour. Further work is also needed to consider the conclusion that the under-prediction
CHAPTERS. CONCLUSIONS AN D FUTURE W ORK 155

of the vortex core axial velocity is the cause of the premature prediction of breakdown for this case. This could be
performed by systematically refining the vortex core region and determining any link with the location and onset
of vortex breakdown.

With regards to the use of DES to predict the unsteady behaviour of the vortex flow, a number of possibilities are
clear. Firstly, further refinement of tlie grid could be canned out to analyse the behaviour of a solution with more
turbulent scales simulated on the grid and to consider the effect of this on the predicted behaviour of the flow -
particularly on the mean breakdown location. This could also include a further refinement of the trailing edge
region to capture the breakdown of the helical mode winding into turbulence downstream of the trailing edge. This
may not be important for the flow at this incidence, however it is unclear whether this turbulent region would effect
the unsteady flow behaviour over the wing as the incidence was increased. Therefore, it would also be interesting
to consider the flow for a incidence at which breakdown is much further upstream, such as a = 35^^ —40^^ for
this wing. The use of overset or hybrid grids may also be interesting to consider to reduce the cost of structured
grid DES calculations, this would allow refinement and accuracy of the solution in tlie vortex region but without
unnecessary grid points in the farfield region of the flow domain.

Further unsteady experimental data is also greatly needed to improve the understanding of this subject and to aide in
the validation of such investigations. Unsteady point probe data, similar to that shown in this investigation through
tlie vortex region, for all components of velocity would be highly beneficial to the development and validation of
CFD in the future, and in particular for URANS models. Further work is needed to understand tlie relation between
the mean location of vortex breakdown and the turbulence downstream of the breakdown location. Finally, this
work could be extended by considering the unsteady forces which are incident on the wing siuTace as a result of
the unsteadiness, the phenomena which cause this forcing and the possible structural response that this may cause.
Appendix A

Turbulence Models

The full form of each main turbulence model used in the investigation will be detailed in this section.

A .l Wilcox k — (û

The Kinematic Eddy Viscosity is calculated from,

IXf (A .l)

Where the turbulent kinetic energy, k and specific dissipation rate, (o are calculated from the partial differential
equations.
dk dpkUj d
( f t + O'* P r ) + Pk-P*pkco (A .2)

and
dû) dpcoUi d dco
p -3 7 H 5------ — -5— (P + ctPt ) + Efl) —I3p(0'‘ (/L3)
dt dxj dxj dx7 J
In the equations above the production terms of k and a), and respectively, are defined as,
£U
Pk = and Pio
Pio — ^~^Pk
Q (A .4)
'-7
The rate of dissipation, e and the length scale of the model are given by,
&V2
e = p*kci) and i = (/L5)
(0
The closure coefficients for the model are defined in Table A .l.
a j3 j8* O' O'*
5/9 0.075 0.09 0.5 0.5

Table A .l: Model constants for the Wilcox k — û) turbulence model [34]

A.2 NLEVM
In an explicit algebraic Reynolds Stress model (EARSM), the anisotropic term of the Reynolds sti-esses is de­
scribed as a function of the normalised mean strain-rate, S and rotation, Ü, tensors. Based on the Cayley-Hamilton
theorem [167], this means that the anisotropy can be described by a series of ten independent, symmetric, devia-
toric functions of S and Q or a linear combination of these ten. For the model specified in [170] this results in the
relationship,

a = jSiS -b - ^ IIs I + + A (S A - A S ) + P5 (S^A - AS^) H- jSg ^SA^ - A^S - ~1V

+j37 ^S^A^ - A^S^ (SAS^ - AS) y -b jSg ( a SA^ - A ^SA) -b jbio ( a S^A^ - A^S^a )
(A .6)

156

r:' _
APPENDIX A. TURBULENCE MODELS 137

The coefficients fin are functions of the five independent invariants of the normalised mean strain-rate and rotation
tensors, S and A , which can be written as

IIs = tr { S ^ } , / Z a ^ t r j A ^ } , 7y = t r { s A ^ j , y = (A.7)

where tr{} is the trace of the second order tensors and the turbulent time scale is givenby.

For the non-linear eddy viscosity model, this relationship for the anisotropy is reduced to a few terras and the
Reynolds stress tensor formulation becomes,

' k ÿ j ^ k ( j S , j - 2 q l ‘S^+a^'^^'j (A .9)

where,
= Ps (^A^ - ~II q + {56 ( s £ f + A^S - IIq S - + jSg (ASA^ - A^SA) (A.IO)

The turbulent eddy viscosity is calculated from,

Px —C ^ p k T (A. 11)

where,
Cj“ = - i / l ( f t + W s 2 f t ) (A .12)

As mentioned above, the (5n coefficients are derived from the invariants of the strain-rate and rotation tensors and
are defined fort his model as,

^ = (A ,3 ,

with
2IIa) (2N^ - Ila) (A. 14)

and

^ + 2 { P f-P 2 Ÿ ^ ^ c o s (^ ^ c o s -^ (^ -^ ^ ^ y P 2< 0.

where,

A = (“ + J - f T — Z7s -b -77a ) and ci = “ ( c i- - l) (A. 16)

Therefore,
(A.17)
D
with,

h = 7F\ 4>2 = V - and D = 20A^ { nc - \ c ' ^ - 77q (lOA^ -b + 10/^77^ (A.18)

For the model implimented in PMB there is no damping or low Reynolds number correction applied, thus the
coefficients are defined / i = 1.0, Q = 6.0 and c\ = 1.8
APPENDIX A. TURBULENCE MODELS 158

A.3 Spalart-Allmaras
The kinematic turbulent eddy viscosity is calculated from

P T = P V fv l (A. 19)

where,
y V
(A.20)
/u i =

In equation A. 19, v is the working variable of the transport equation of the model, which is given by

dv
(v + v) (A.21)
dxk (7 d x k d x k

where S is the modified vorticity given by

% (A .22)
S = S+ ~ ^ f v 2 and f ^2 = 1
l + Xf v \ '
where d is the distance to the closest wall and S is the magnitude of vorticity,

5"— |w| = |V X {iaPv]-\-wk) | (A.23)

Similarly, in the destruction/near wall term, the function /,i- in Equation A.21 is given by
1/6
V
fw —, (A.24)
■+c:m'3 J SK^d^

These functions take the presence of a wall into account and satisfy the wall boundary conditions where v —0. As
r increases, tends to an asymptote, therefore values of r are generally truncated to about 10. In the freestream
region, it is also best to use v = 0, provided that numerical errors do not cause v to become negative close to
the edge of the boundary layer - the exact solution cannot become negative. Generally, values less than v/10 are
acceptable. This also applies to the initial conditions. The model coefficients are given in Table A.2.

C l)\ C[)2 C yl C((,2 (T K


0.1355 0.622 7.1 3.239 0.3 2.0 2/3 0.41

Table A.2: Model constants for the Spallart-Allmaras turbulence model [35], where Cwi = ^
Appendix B

Probe Analyser Tool

B .l Probe Analyser
In the course of the investigation into unsteady flows, the unsteady behaviour was considered through use of a se­
ries of point probes applied to the flow domain. These point probes were applied to the computational grid through
specification of coordinates at tlie outset of the calculations and the flow variables were saved at each point for
every time step of the calculation. This results in a number of files being created which contain the time histories
of each flow variable. However, these files are not immediately usable and require reorganisation into a form which
is more practical. To alter the form of these files and to allow analysis of the resulting time histories in an easy and
efficient manner, a program was created using Matlab, which combines all the analytical and statistical analyses
into one interface allowing consistent analysis and easy comparison and cross-plotting of data. This program,
Probe Analyser, was originally designed to analyse acoustical data from pressure signals for cavity flows and was
created by Lawrie [178]. Further work was carried out by Nayyar [179] who added further analyses and created a
similar program for turbulent analysis.

At the start of this project. Probe Analyser was only able to perform statistical analyses on the pressure time
histories from the point probe files. However, to fully consider the unsteady behaviour over delta wings it was
necessary to be able to consider all flow variables in a similar manner. It was also important to be able to consider
the unsteady forces acting on the wing and therefore the ability to process the integrated loads files was also added.
As mentioned, the program was originally designed for cavity flows, therefore the length scales and plots were
specific to the character of these flows, these were changed to make the program more specific to delta wings.
Further work was caiTied out to improve the ability of the program to quickly cross-plot data, this involved adding
the ability to process multiple probes and multiple probe files and to plot these on the same graphs. To reduce the
size of the data sets, in order to reduce the memory requirements of the program, a facility to split the large data
files into smaller subfiles was added. This had the effect of further increasing the ease of comparison of multiple
signals. The turbulent analyses created by Nayyar were also incorporated into the main Probe Analyser program
to create a single program which was capable of fully analysing the unsteady data. The ability to time average the
signals was also added, as was the ability to consider non-stationary turbulence.

Probe Analyser has the ability to perform many more analyses than were used in this investigation. The main
analyses which were used are the calculation of the mean and RMS values, the power spectral density (PSD),
the time average of the signals and the calculation of turbulent kinetic energy. Explanations of each of these
methods will be given in a later section. However, analyses such as the probability density function (PDF), auto­
correlation, calculation of the Reynolds stresses and further turbulent correlations are also available. These will not
be discussed in detail in this Appendix. The graphical user interface (GUI) of the program is shown in Figure B.l.
This shows all the analyses available and the overall format of the program. This GUI comes from the Windows
interface, however the program can be used on the Linux operating system and a start up command allows this to be
specified. The left hand side of the GUI mainly deals with the input of the files, the specification of the important
flow parameters and the selection of the probes to be considered. The right hand side allows the selection of tlie
analysis and the specification of the resulting plots through drop-down menus. The initial manipulation of the
probe data files, created from the CFD calculation, into a usable format, as mentioned above, is done through a
secondary program accessed through the “Locate Probes” button on the top left of the GUI.

159
APPENDIX B. PROBE ANALYSER TOOL 160

wffl'aasBa
Flow Paiaaietef* SlalMlical Analyte* Calculât# T w brdent Pwopaib#*
L oc«l« P io b e i C itia c t Piobe* C lean P io b a t
OATO.dagraeSDES
Ftae-stream Mach Numbai rôT" (Variablel 3 jAveragng Type 1 J tJ
Angle of Incidence (deg) Calculate Statisticai Properties I P Calculate AccumJated Average?
Non Dimensional Time Slap Cakidate Average Caladate|
I Plot Type 1 7 3 _ 2 slJ
P ia ttw a Data |
D E S .d cbv e EnUe Data |
Fieeslieam Temperatute (K) I 293.15
I 101000
ProbabAy Der»*y CaktbbI Calculate Turbulent Quantitiet Calculâte|

O E S.acbve.faf
m t« rta n e o u » Ce-Of<fa I
Total Pressme |PaJ
Geom etric Param eter*
CiriHiatTve Oistrbution PW I CalcuWeReynoWi Sir#** Tensor
Reynold* Stras*#*
Calciiate|

integrale4.b«d<J4et A utocorrélation
r Muftipte coord»> Root Chwd Length Im)
Pfoba^dala.fies
Wne^aveaged P D aaiadby
P
D ossCorrelation Cakibbj |Va.iable1 “3
OetaAnaVtec Sweep Anjÿe (deg)
ItGfidHardMired? P
Fid Span
incrience (deg)
|Select Plot Type
S p e c tra l A nalyte*
I
I n t L oad*
Power Spectral Density C*buW»| T uibuianca Model Proper tie*
Calculate Turbulence Modal pTopertie* |
Modal Nurrtwif
Reference Area | 1 SanvSpan
C onveiM on Factor*
JSelect Plot Type
r Plot K.olmogoto v 5 /3 Slop*
3 _ S ± J
lv«Mbtei 3
sc ale I rd Hi I 0.01 y 2 f “ idd~*^ |S«iaetfW T ype 3 ^ I
P InckidoEKpt? Conversion Factor ■Time F T
IKE Powei Spechal Oentiji Cafcubbj
E«pl P a th Conversion Factor • Pressure P BandÜmrtei | Min t Matt '

R e»el P a th | Conversion Factor - Force r~ i~ Donsnant Frequencies CalcUate|


|SitootF>WType 3 ^ I

Input Fa#*
Mot ustng non-dim quanbtias? P Cabulatal Tone Selectoi Seleca |
^lessure Dal# File Fk P robe S election Band Limited SPL
'o o td Fie |V analk1 zJ
imiioDataFi# Fk Dos* Spectral Density Caki
"la 01 Signal* Signals r Smgle W indow ed S p e c tra l A nalyses
CakuWe Tw buW Slahslcal Pfopeslie:
PW J
|
*lq Of Servie* Samples jPW T ypel 3
r Range P Enter Bandwidth | Fteq [ Fieq
ntegralod Load* Fie Fk IVanabfel
Plot Windowed Spectral Demiy pw | 3
P io b a D ata F ilet
C Selection: I S elect No. of Probes 1Plot Type 1 3 Pkl 1
SingJaFrte P MtApteFiw |N q of dala f h t 7] Sim ulation of URANS Tim# A veraging
Probe* Selected: |
Sampling Time Step PttO O l O ther A nalyses
P Plot on same axes? Calculate Time Averged Signal Calculate| CakUalel
Boundaiy Layer Height
P Output to F ie? S e le c t D ata Calculate PSD of Sknel Calculate
In te g ra la d l o a d * Shear layer Thickness C a k ik k ]
SingleFb r M uiipleFbj |N o o (W .F b t H (select Flot Type " 3 pw I Energy/Momentum Tlansfer PW 1
(• Spanwfsa r Chordwis* P Include Original Signal m plot?
P Plot mean

Figure B .l : Graphical User Interface for probe analyser program

B.2 Application of Statistical Methods


Probe A nalyser was initially created in M atlab as it allowed the use o f som e existing mathematical functions and
made the creation o f the GUI easier through use o f the M atlab add-on package G uide. This program automatically
sets up the links between the GUI and the underlying code, greatly reducing the com plexity o f the programming
task. The underlying code for the Probe A nalyser program is large with many subroutines and will not be repro­
duced here, however the initial code o f the program can be found in Ref. [178]. A s mentioned above, the main
analyses used in this investigation are the calculation o f the mean, RMS, PSD, time average and turbulent kinetic
energy o f the specified signal or signals. In this section each o f these methods w ill be described. In all cases, with
the exception o f the turbulent kinetic energy which is calculated using the velocity vector, the general variable, O,
will be used as these analyses can be carried out on the signals o f any o f the flow variables. It should be noted that
only the mean, RMS and PSD analysis are able to be carried out on the integrated loads signals.

B.2.1 Mean and Root Mean Square Values


The mean o f the time varying signals is calculated from a straight forward average o f the data points such that,

10
(<D) = (B.l)
N
where N is the number o f samples in the signal.

The RMS value is a statistical measure o f the deviance o f a signal from the mean and therefore a measure o f the
intensity o f the fluctuations o f the unsteady signal. It is calculated by

(B.2)

Within the Probe A nalyser program, both these values can be plotted against location on the wing, in the spanwise,
chordwise or normal directions depending on the probes selected for analysis. They provide an excellent way to
compare results, particularly for multiple solutions as up to four files can be entered into probe analyser for analysis
at any one time.
APPENDIX B. PROBE ANALYSER TOOL 161

B.2.2 Power Spectral Density


The behaviour of an unsteady time dependent random variable, such as the flow variables in this investigation,
can be though of as the superposition of multiple oscillations at different frequencies. This can be mathematically
described as a trigonometric series of harmonic waveforms. This series is known as the Fourier series and is
defined as
, ,, , / 2nkt . 2 n k t\
O (t ) = cto + -"jr- + h s in - y - j (B .3)

where the Fourier coefficients are given by

1 ? ,. , , 2 . 2nkt , , 2 , 2nkt
cio = f j T ‘^ ( 0 * . ^ { t) c o s -jr -d t, bk = ^ 0 (r)™ i-y -^ /r (B.4)

and
(B.5)

The Fourier series, therefore, describes the signal within the frequency domain instead of the time domain. A
similar series can be formed for fluctuations in space, i.e 0(%), which provides a description of the signal in wave
number space, k . T o transform between the time and frequency domains (or space and wave number domains) fire
Fourier Transform of tire signal is used. This is derived from the Fourier series arrd for a given time dependent
variable, 0 (t) is defined as
$ (m ) = = ^ {t)e -^ ^ ‘dt (B.6)

and its iirverse is


$ (r) = ^ - i { 0 ( m ) } ^ r 0(m )e""ym (B.7)

The power spectral density function is defined as the Fourier transform of the auto-correlation function of the time
dependent variable arrd provides information of the frequency disfiibution and power of a signal in the frequency
domain. It is defined mathematically as

PSD^if^t) — ~ 2 ^ j dx (B.8)

where the auto-correlation function of the variable is defined as the mean of the product of the variable at time t
and the variable at time t4-T, such that

(t) = + (B.9)

As the signals obtained from the CFD calculation are discrete, finite in length this form of the Fourier series cannot
be applied dhectly. A computationally efficient form of a method kirown as the discrete Fourier transform (DFT)
is used instead. The DFT merely allows fire transform to be applied to sampled signals and redefines the Fourier
tiansform as
0(m ) = ^ (B.IO)
^ ;-=0
The form of this method used, is kirown as the fast Fourier transform (FFT) which is a computatioirally efficient
method of calculating Fourier transforms for signal processing. This method is used by Matlab in a number of
available standard PSD functioirs. The PSD is calculated in Probe Analyser by using the p e r io d o g r a m function.
This function applies a rectangular window (equivalent to no window) to the signal and calculates the PSD using a
standard periodogram. The periodogram calculates the PSD by taking the square of the magnitude of the FFT and
dividing it by the number of sairrples, such that

= (B .ll)

Further detail of this function and its use can be obtained from the Matlab Documentation [201]. The function
outputs the PSD and the frequency data, which can then be plotted to determine the frequency content of the signal.
The PSD magnitude is dependent on the length of the signal and whether the variables used are dimensional or
non-dimensional, therefore in this investigation, only the relative values of the PSD have been considered. In
Probe Analyser a number of plotting options are available and the PSD can be plotted against both dimensional
and non-dimensional frequency (Strouhal number) and period. The areaunder the PSD -frequency plot should be
equal to the square of the RMS value of he signal. This has been used as a check to the validity and accuracy of
this method within the investigation.
APPENDIX B. PROBE ANALYSER TOOL 162

B.2.3 Time Averaging


The time average of the signal can be defined as,

$ = - / (B.12)

where T is defined as the sample rate of the averaging process, which is specified by the user in Piobe Analyser.
This value must be a multiple of the time step of the calculation, A-r. As with the Fourier transform, this form
cannot be applied to a discrete signal and so the integral is approximated to a summation over each interval, T in
turn,
1

0 = 1 0AT (B.13)
^ H=0
The mid point of each interval is also calculated by a similar averaging technique to determine the time at which
the new point occurs. The resulting series of new data points creates the time averaged signal for the specified
sample rate. This new signal can then be evaluated in a similar manner to the original signal by using the PSD
analysis and the results can be plotted to determine the effect of time averaging. The location of the stationary
mean can also be plotted for comparison. Currently, this can only be performed on a single probe at a time. For
non-stationary process this method also allows the non-stationary mean to be determined for a specified sample
rate. This can then be used to determine the turbulent properties of the signal.

B.2.4 Hirbulent Kinetic Energy


As mentioned, the calculation of the turbulent kinetic energy is only one of a number of turbulent properties which
can be calculated in Probe Analyser. Before any of the turbulent properties can be calculated, it is necessary to
calculate the fluctuating variables of tlie flow. This is done simply by subtracting the mean from the unsteady
signal. Either the stationary mean described in Section B.2.1 or the non-stationary mean calculated from the time
averaging process can be selected. The turbulent kinetic energy is calculated from the fluctuating velocities, u', v'
and w' by
+ + (B.14)

The resulting signal can then be considered, as before, by calculation of the mean and RMS values and by calcu­
lation of the PSD.
1

References

[1] Anderson J.D. Fundamentals o f Aerodynamics. McGraw-Hill International Editions, 2nd edition, 1991.

[2] Nelson R.C and Pelletier A. The unsteady aerodynamics of slender wings and aircraft undergoing large
amplitude manoeuvers. Progress in Aerospace Sciences, 39; 185-248,2003.

[3] Pagan D and Solignac J.L. Experimental study of the breakdown of a vortex generated by a delta wing.
Recherche Aerospatiale, 3:197-219, 1986.
[4] Larabourne N.C and Bryer D. W. The bursting of leading edge vortices - some observations and discussion
of the phenomenon. Report and Memorandum No. 3282, Aircraft Research Council, April 1961.

[5] Gortz S. Realistic Simulations o f Delta Wing Aerodynamics Using Novel CFD Methods. PhD thesis, Aero­
nautics and Vehicle Engineering, KTH, June 2005.
■4
[6] Klute S.M. The Development and Control o f Axial Vortices over Swept Wings. PhD thesis, Virginia Poly­
technic Institute and State University, August 1999.

[7] Wentz W.H and Kohlman D.L. Vortex breakdown on slender sharp edged wings. Journal o f Aircraft,
8(3): 156-161, March 1971.
[8] Huang X.Z and Hanff E.S. Unsteady behaviour of spiral leading edge vortex breakdown. In AIAA At­
mospheric Flight Mechanics Conference. AIAA Paper 1996-3408, July 1996. I
'

[9] Lee M and Ho C-M. Lift force of delta wings. Applied Mechanics Review, 43(9): 209-221, September 1990.

[10] Visbal M.R. Computed unsteady structure of spiral vortex breakdown on delta wings. In 27th AIAA Fluid
Dynamics Conference. AIAA Paper 96-2074, June 1996.

[11] Gursul I and Xie W. Physics of buffeting flows over delta wings. In 36th AIAA Aerospace Sciences Meeting
and Exhibit. AIAA Paper 98-0688, January 1998.

[12] Riley A.J and M.V Lowson. Development of a tliree-dimensional free shear layer. Journal o f Fluid Me­
chanics, 369:49-89,1998.

[13] Mitchell A.M. Caractérisation et Contrôle de L ’éclatement Tourbillonnaire sur une Aile Delta aux Hautes
Incidences. PhD thesis, LUniversité Paris 6, July 2000.

[14] Visbal M.R and Gordnier R.E. On the structure of the shear layer emanating from a swept leading edge at
angle of attack. In 33rd AIAA Fluid Dynamics Conference and Exhibit. AIAA Paper 2003-4016, June 2003.

[15] Menke M, Yang H, and Gursul I. Experiments on the unsteady natur e of vortex breakdown over delta wings.
Experiments in Fluids, 27:262-272,1999.
[16] Miller D.S and Wood R.M. Leeside flows over delta wings at sirpersonic speeds. Journal o f Aircraft,
21(9):680-686, September 1984.

[17] Donohoe S.R and Bannink W.J. Surface reflective visualisations of shock-wave/vortex interactions above a
delta wing. A7AA Journal, 35(10):1568-1573, October 1997.

[18] Elsenaar A and Hoeijmakers H.W.M. An experimental study of the flow over a sharp-edged delta wing at
subsonic and transonic speeds. In AGARD Conference Proceedings “Vortex Flow Aerodynamics”, pages
15.1-15.19. AGARD-CP-494, July 1991.

163
"T|l:

REFERENCES 164

[19] Kameda M, Tabei T, Nakakita K, Sakaue H, and Asai K. Image measurements of unsteady pressure fluctua­
tion by a pressure sensitive coating on porous anodized aluminium. Measurement Science and Technology,
16:2517-2524,2005.

[20] Chu J and Luckring J.M. Experimental surface pressure data obtained on a 65® delta wing across Reynolds
number and Mach number ranges: Volume 1 - sharp leading edge. NASA Technical Memorandum 4645,
NASA Langley Research Centi'e, February 1996.

[21] KomathR. Private communication.

[22] Agrawal S, Barnett R.M, and Robinson B.A. Numerical investigation of vortex breakdown on a delta wing.
AIAA Journal, 30(3):584-591, March 1992.

[23] Gortz S. Time accurate Euler simulations of a full span delta wing at high incidence. In 21st AIAA Applied
Aerodynamics Conference. AIAA Paper 2003-4304, June 2003.

[24] Morton S.A, Forsythe J, Mitchell A.M, and Hajek D. Detached eddy simulation and Reynolds averaged
Navier-Stokes simulations of delta wing vortical flow fields. Journal o f Fluids Engineering, 124:924-932,
December 2002.
[25] Maiy I. Large eddy simulation of vortex breakdown behind a delta wing. International Journal o f Heat and
Fluid Flow, 24(4):596-605, August 2003.

[26] Wilcox D.C. Turbulence Modelling fo r CFD. DCW Industries Inc„ 1993.

[27] Lovely D and Haimes R. Shock detection from computational fluid dynamics results. In 14th AIAA Com- ;

putational Fluid Dynamics Conference. AIAA Paper 99-3285, June/July 1999. f

[28] Kalklioran I.M and Smart M.K. Aspects of shock wave-induced vortex breakdown. Progress in Aerospace
Sciences, 36:63-95,2000.

[29] Mitchell A.M, Morton S.A, and Forsythe J.R. Analysis of delta wing vortical substructures using detached
eddy simulation. In 32nd AIAA Fluid Dynamics Conference. AIAA Paper 2002-2968, June 2002.

[30] Morton S.A. High reynolds number DES simulations of vortex breakdown over a 70® delta wing. In 21st
AIAA Applied Aerodynamics Conference. AIAA Paper 2003-4217, June 2003.

[31] Soemarwoto B.l and Boelens O.J. Simulation of vortical flow over a slender delta wing experiencing vortex
breakdown. In 21st AIAA Applied Aerodynamics Conference. AIAA Paper 2003-4215, June 2003.

[32] Gortz S and Le Moigne Y. Time accurate detached eddy simulations of a full span delta wing at high
incidence. In 21st AIAA Applied Aerodynamics Conference. AIAA Paper 2003-4216, June 2003.

[33] Spalart P.R. Strategies for turbulence nodelling and simulations. International Journal o f Heat and Fluid
Flow, 21(3):252-263, June 2000.

[34] Wilcox D.C. Reassessment of the scale determining equation for advanced turbulence models. AIAA Jour­
nal, 26(11): 1299-1310, November 1988.

[35] Spalart P.R and Allraaras S.R. A one equation turbulence model for aerodynamic flows. In 30th AIAA
Aerospace Science Meeting and Exhibit. AIAA Paper 1992-0439, January 1992.

[36] Earnshaw P.B. An experimental investigation of the structure of a leading edge vortex. RAE Technical Note
No. Aero. 2740, Royal Aeronatical Establisment, March 1961.

[37] Flynn G.A, Morrison J.F, and Mabey D.G. Buffet alleviation on swept and unswept wings at high incidence.
Journal o f Aircraft, 38(2):368 - 378, March - April 2001.

[38] Gad-el-Hak M and Blackwelder R.F. The discrete vortices from a delta wing. AIAA Journal, 23(6):961-962,
June 1985.

[39] Lowson M.V. Visualisation measurements of vortex flows. Journal o f Aircraft, 28(5):320-327, May 1991.

[40] Gad-el-Hak M and Blackwelder R.F. Control of the discrete vortices from a delta wing. AIAA Journal,
25(8): 1042-1049, August 1987.

— . Ù’- A'.Li; ^ ...


REFERENCES 165

[41] Payne RM, Ng T.T, Nelson R.C, and Schifî L.B. Visualisation and flow surveys of the leading edge vortex
structure on delta wing planforms. In 24th AIAA Aerospace Sciences Meeting. AIAA Paper 1986-0330,
January 1986.

[42] Payne RM. The Structure o f Leading Edge Vortex Flows Including Vortex Breakdown. PhD thesis, Univer­
sity of Notr e Dame, Notre Dame, IN, 1987.

[43] Hemsch M.J and Luckring J.M. Connection between leading edge sweep, vortex lift and vortex strength for
delta wings. Journal o f Aircraft, 27(5):473-475,May 1990.

[44] O’Neil PJ, Roos RW, Kegelman J.T, Barnett R.M, and Hawk J.D. Investigation of flow characteristics of
a developed vortex. NADC-89114-60, McDonnell Aircraft Company, P.O. Box 516, St. Louis, MO 63166,
May 1989.

[45] Elle B.J. An investigation at low speed of the flow near the apex of thin delta wings with sharp leading
edges. Reports and Memorandum No. 3176, Aircraft Research Council, January 1958.

[46] Rusak Z and Lamb D. Prediction of vortex breakdown in leading-edge vortices above slender delta wings.
Journal o f Aircraft, 36(4): 659-667,1999.

[47] Jupp M.L, Co ton RN, and Green R.B. A statistical analysis of the surface pressure distribution on a delta
wing. The Aeronautical Journal, 103(1025):349-357, July 1999.

[48] Hummel D. On the vortex formation over a slender wing at large incidence. In AGARD CF-247 Paper No.
15. NATO AGARD, January 1979.

[49] Délery J.M. Aspects of vortex breakdown. Progress in Aerospace Sciences, 30(1): 1-59,1994.

[50] Hall M.G. Vortex breakdown. Annual Review o f Fluid Mechanics, 4:195-218,1972.

[51] Visser K.D and Nelson R.C. Measurements of circulation and vorticity in the leading edge vortex of a delta
wing. AIAA 31(1): 104-111, January 1993.
[52] Polhamus E.C. Prediction of vortex lift characteristics by a leading edge suction analogy. Journal o f Aircraft,
8(4):193-199,1971.

[53] Paler J.H and Leibovich S. Disrupted states of vortex flow and vortex breakdown. Physics o f Fluids,
20(9): 1385-1400, September 1977.

[54] Payne RM, Ng T.T, and Nelson R.C. Experimental study of the velocity field on a delta wing. In I9th AIAA
Fluid Dynamics, Plasma Dynamics and Lasers Conference. AIAA paper 1987-1231, June 1987.

[55] Klute S.M, Vlachos P.P, and Telionis D .P High-speed digital-particle-image-velocimtery study of vortex
breakdown. AIAA Journal, 43(3):642-650, March 2005.
[56] Ozgdren M, Sahin B, and Rockwell D. Vortex structure on a delta wing at high angle of attack. AIAA
Journal, 40(2):285-292, February 2002.

[57] Towfighi J and Rockwell D. Instantaneous structure of vortex breakdown on a delta wing via particle image
velocimetry. AIAA /owma/, 31(6): 1160-1162, June 1993.

[58] EscudierM. Vortex breakdown: Observations and explanations. Progress in Aerospace Sciences, 25:1^9-
229, 1988.

[59] Spall R.E, Gatski T.B, and Grosch C.E. A criterion for vortex breakdown. Physics o f Fluids, 30(11):3434-
3440, November 1987.

[60] Robinson B.A, Barnett R.M, and Agrawal S. Simple numerical criterion for vortex breakdown. AIAA
Journal, 32(1): 116-122, January 1994.

[61] Ashley H, Katz J, Jarrah M A, and Vaneck T. Survey of research on unsteady aerodynamic loading of delta
wings. Journal o f Fluids and Structures, 5:363-390, 1991.

[62] Mitchell A.M and Delery J.M. Research into vortex breakdown control. Progress in Aerospace Sciences,
37(4):385-418,2001.
■ ........................................

■■î

REFERENCES 166 .
'i
f
[63] Brown G.L and Lopez J.M. Axisymmetric vortex breakdown, Part II: Physicalmechanisms. Journal o f |
Fluid Mechanics, 221:553-576, December 1990.
1
[64] Jumper E.J, Nelson R.C, and Cheung K. A simple criterion for vortexbreakdown.In 31st AIAA Aerospace %
Science Meeting and Exhibit. AIAA Paper 93-0866, January 1993. |
... i
[65] Sarpkaya T. Effect of the adverse pressure gradient on vortex breakdown. AIAA Journal, 12(5):602-607,
May 1974. "

[66] Ericsson L.E. Effect of fuselage geometry on delta wing vortex breakdown. Journal o f Aircraft, 35(6) :898-
904, November - December 1998.

[67] Ericsson L.E. Multifaceted influence of fuselage geometry on delta wing aerodynamics. Journal o f Aircraft,
40(l):204-206, January - February 2003.

[68] Lowson M.V and Riley A.J. Vortex breakdown control by delta wing geometry. Journal o f Aircraft,
32(4): 832-83 8, July-August 1995.

[69] Srigrarom S and Kurosaka M. Shaping of delta wing planform to suppress vortex breakdown. AIAA Journal;
Technical Notes, 38(1):183-186, January 2000.

[70] Allan M.R, Badcock K.J, Barakos G.N, and Richai'ds B.E. Wind tunnel interference effects on delta wing
aerodynamics computational fluid dynamics investigation. Journal o f Aircraft, 42(1):189~198, January-
February 2005.
[71] Allan M.R, Badcock K.J, Barakos G.N, and Richards B.E. Wind-tunnel interference effects on a 70® delta
wing. The Aeronautical Journal, pages 505-513, October 2004,

[72] Jobe C.E. Vortex breakdown location over 65® delta wings empiricism and experiment. Aeronautical
Journal, pages 475-482, September 2004.
.1
[73] Gursul I. Criteria for vortex breakdown location over delta wings. Aeronautical Journal, pages 194-196,
May 1995.

[74] Johari H and Moreira J. Direct measurement of delta wing vortex circulation. AIAA Journal, 36(12):2195-
2203, December 1998.
-I
[75] Earnshaw P.B and Lawford J.A. Low-speed wind-tunnel experiments on a series of sharp edged delta wings.
Reports and Memoranda No. 3424, Aircraft Research Council, March 1964. Ij

[76] Mabey D.G. Beyond the buffet boundary. The Aeronautical Journal, pages 201-215, April 1973. J
[77] Garg A.K and Leibovich S. Spectral characteristics of vortex breakdown flowflelds. Physics o f Fluids, j
22(11):2053-2064, November 1979.

[78] Singh P I and Uberoi M.S. Experiments on vortex stability. Physics o f Fluids, 19(12):1858-1863,December
1976.

[79] Gursul I. Unsteady flow phenomena over delta wings at high angle of attack. AIAA Journal, 32(2):225-231,
February 1994.

[80] Lowson M.V. Some experiments with vortex breakdown. Journal o f the Royal Aeronautical Society,
68:343-346, May 1964.

[81] Mitchell A.M, Barberis D, Molton P, and Délery J. Oscillation of vortex breakdown location and blowing
control of time averaged location. AIAA Journal, 38(5):793-803, May 2000.

[82] Lambert C and Gursul I. Insensitivity of unsteady vortex interactions to Reynolds number. AIAA Journal,
38(5):937-939, May 2000.
[83] Ayoub A and McLachlan B.G. Slender delta wing at high angles of attack - a flow visualisation study. In
19th AIAA Fluid Dynamics, Plasma Dynamics and Lasers Conference. AIAA Paper 87-1230, June 1987.

[84] Portnoy H. Unsteady motion of vortex breakdown positions on delta wings. In International Council o f the
Aeronautical Sciences Congress. ICAS, August-September 1988.
REFERENCES 167

[85] Helin H.E and Watry C.W. Effects of trailing edge jet entrainment on delta wing vortices. AIAA Journal,
32(4): 802-804, April 1994.
[86] Gursul I and Yang H. On fluctuations of vortex breakdown. Physics o f Fluids, 7(1):229-231, January 1995.
[87] Menke M and Gursul I. Unsteady nature of leading edge vortices. Physics o f Fluids, 9(10):2960-2969,
October 1997.
[88] Gursul I and Xie W. Origin of vortex wandering over delta wings. Journal o f Aircraft, 37(2): 348-350,2000.
[89] Rediniotis O.K, Stapountzis H, and Telionis D.P. Vortex shedding over delta wings. AIAA Journal,
28(5):944-946,May 1990.
[90] Rediniotis O.K, Stapountzis H, and Telionis D.P. Periodic vortex shedding over delta wings. AIAA Journal,
31(9):1555-1562, September 1993.
[91] Gursul I and Xie W. Buffeting flows over delta wings. AIAA Journal, 37(l):58-65, January 1999.
[92] Winant C.D and Browand F.K. Vortex pahing: the mechanism of turbulent mixing-layer growth at moderate
Reynolds number. Journal o f Fluid Mechanics, 63:237-255,1974.
[93] Gordnier R.E and Visbal M.R. Unsteady vortex structure over a delta wing. Journal o f Aircraft, 31(1):243-
248, January 1994.
[94] Gordnier R.E and Visbal M.R. Unsteady Navier-Stokes solutions for a low aspect ratio delta wing. AIAA
Paper 90-1538, June 1990.
[95] Gursul I. Review of unsteady vortex flows over slender delta wings. Journal o f Aircraft, 42(2):299-319,
March-April 2005.
[96] Gordnier R.E and Visbal M.R. Instabilities on the shear layer of delta wings. In 26th AIAA Fluid Dynamics
Conference. AIAA Paper 95-2281, June 1995.
[97] Washburn A.E and Visser K.D. Evolution of vortical shuctures in tlie shear layer of delta wings. In 25th
AIAA Fluid Dynamics Conference. AIAA Paper 94-2317, June 1994.
[98] Shan H, Jiang L, and Liu C. Direct numerical simulation of three-dimensional flow around a delta wing. In
38th AIAA Aerospace Sciences Meeting and Exhibit. AIAA Paper 2000-0402, January 2000.
[99] Mitchell A.M, Molton P, Barberis D, and Délery J. Vortical substructures in the shear layers forming leading
edge vortices. In I9th AIAA Applied Aerodynamics Conference. AIAA Paper 2001-2424, June 2001.

[100] Mitchell A.M and Molton P. Vortical substructures in the shear layers forming leading edge vortices. AIAA
Journal, 40(8): 1689-1692, September 2002.
[101] Honkan A and Andreopoulos J. Instantaneous three-dimensional vorticity measurements in vortical flow
over a delta wing. AIAA Journal, 35(10): 1612-1620, October 1997.
[102] Gursul I. Recent developments in delta wing aerodynamics. Aeronautical Journal, pages 437-452, Septem­
ber 2004.
[103] Furman A and Breitsamter C. Investigation of flow phenomena on generic delta wing. In 25th International
Congress o f the Aeronautical Sciences. ICAS, 2006.
[104] Lee Y.L and Gursul I. A i investigation of unsteady interactions of a vortex pair over delta wings. In 41st
AIAA Aerospace Sciences Meeting and Exhibit. AIAA Paper 2003-0423, January 2003.

[105] Lowson M.V. The three dimensional vortex sheet structure on delta wings. In Fluid Dynamics o f Three
Dimensional Turbulent Shear Flows and Transition, pages 11.1-11.16. AGARD CP-438, October 1988.
[106] Renac F, Barberis D, and Molton P. Control of vortical flow over a rounded leading edge delta wing. AIAA
Journal, 43(7).T409-1418, July 2005.
[107] Taylor G.S and Gursul I. Unsteady vortex flows and buffeting of a low sweep delta wing. In 42nd AIAA
Aerospace Sciences Meeting and Exhibit. AIAA Paper 2004-1066, January 2004.
[108] Woppowa S and Grosche F.R. Unsteady vortex flow over a combat aircraft configuration at high angles of
incidence. In Unsteady Aerodynamics. Royal Aeronautical Society, July 1996.
.j«
REFERENCES 168

[109] Yaniktepe B and Rockwell D. Flow sti'ucture on a delta wing of low sweep angle. AIAA Journal, 42(3):513-
523, March 2004.

[110] Yavuz M.M, Elkhoury M, and Rockwell D. Near surface topology and flow structure on a delta wing. AIAA
Journal, 42(2);332-340, February 2004.

[111] Cummings R.M, Morton S.A, and Siegel S.G. Computational simulation and experimental measurements
for a delta wing with periodic suction and blowing. Journal o f Aircraft, 40(5):923-931, September-October A
2003.
[112] Cummings R.M, Morton S.A, and Seigel S.G. Computational simulation and PIV measurements of the |
laminar vortical flowfield for a delta wing at high angle of attack. In 41st AIAA Aerospace Sciences Meeting I
and Exhibit. AIAA Paper 2003-1102, January 2003.

[113] Gordnier R.E and Visbal M.R. Higher-order compact difference scheme applied to the simulation of a low
sweep delta wing flow. In 41st AIAA Aerospace Sciences Meeting and Exhibit. AIAA-2003-0620, January
2003.

[114] Gordnier R.E and Visbal M. Compact difference scheme applied to simulation of low sweep delta wing
flow. AIAA Journal, 43(8): 1744-1752, August 2005. i

[115] Wolfe S, Lin J.C, and Rockwell D. Buffeting at the leading edge of a flat plate due to a streamwise vortex;
Flow structure and surface pressure loading. Journal o f Fluids and Structures, 9:359-370,1995.

[116] Morton S.A, Steenman M, Cummings R.M, and Forsythe J.R. DES grid resolution issues for vortical flows
on a delta wing and an F-18C. In 41st AIAA Aerospace Sciences Meeting and Exhibit. AIAA Paper 2003-
1103, January 2003.

[117] Erickson G.E and Rogers L.W. Experimental study of the vortex flow behaviour on a generic fighter wing
at subsonic and ti'ansonic speeds. In 19th AIAA Fluid Dynamics, Plasma Dynamics and Lasers Conference.
AIAA Paper 87-1262, June 1987.

[118] Erickson G.E, Schreiner J.A, and Rogers L.W. Multiple vortex and shock interactions at subsonic, transonic
and supersonic speeds. In 8th AIAA Applied Aerodynamics Conference. AIAA Paper 90-3023, August 1990.

[119] Stanbrook A and Squire L.C. Possible types of flow at swept leading edges. The Aeronatical Quarterly,
15:72-82,1964.

[120] Yenga Narayan K and Seshadri S.N. Types of flow on the lee side of delta wings. Progress in Aerospace
Sciences, 33:167-257,1997.

[121] Longo J.M.A. Compressible inviscid vortex flow of a sharp edge delta wing. AIAA Journal, 33(4):680-687,
April 1995.

[122] Donohoe S.R, Houtman E.M, and Bannink W.J. Surface reflective visualization system study to vortical
flow over delta wings. Journal o f Aircraft, 32(6): 1359-1366, November-December 1995.

[123] Kandil O.A, Kandil H.A, and Liu C.H. Shock-vortex interaction over a 65-degree delta wing in transonic
flow. In 24th AIAA Fluid Dynamics Conference. AIAA Paper 93-2973, July 1993.

[124] Kandil O.A, Kandil H.A, and Liu C.H. Supersonic vortex breakdown over a delta wing in transonic flow.
In l lth AIAA Applied Aerodynamics Conference. AIAA Paper 93-3472-CP, August 1993.

[125] Visbal M.R and Gordnier R.E. Compressibility effects on vortex breakdown onset above a 75-degree sweep
delta wing. Journal o f Aircraft, 32(5):936-942, September-October 1995.

[126] Hoeijmakers H.W.M. Modelling and numerical simulation of vortex flow in aerodynamics. In AGARD
Conference Proceedings “Vortex Flow Aerodynamics”, pages 1.1-1.46. AGARD-CP-494, July 1991.

[127] Squire L.C. Leading edge separations and cross-flow shocks on delta wings. AIAA Journal, 23(3):321-325,
1985.

[128] Muylaert J.M. Effect of compressiblity on vortex bursting on slender delta wings. Project report 1980-21,
Von Karman Institute for Fluid Dynamics, Rliode St, Genese, Belgium, July 1980.
REFERENCES 169

[129] Houtman E.M and Bannink B.J. Experimental and numerical investigation of the vortex flow over a delta
wing at transonic speeds. In AGARD Conference Proceedings “Vortex Flow Aerodynamics”, pages 5.1-
5.11. AGARD-CP-494, July 1991.
[130] Thomer O, Schroder W, and Kr ause E. Normal shock vortex interaction. In Proceedings o f the RTO-AVT
Symposium on “Advanced Flow Management: Part A - Vortex Flows and High Angle o f Attack fo r Military
Vehicles” - RTO-MP-069(I), pag&s 18.1-8.10. NATORTO, 2001.
[131] Smart M.K and Kalklioran I.M. Flow model for predicting normal shock wave induced vortex breakdown.
AIAA Journal, 35(10): 1589-1596, October 1995.

[132] Chu J and Luckring J.M. Experimental surface pressure data obtained on a 65® delta wing across Reynolds
number and Mach number ranges: Volume 2 - small radius leading edge. NASA Technical Memorandum
4645, NASA Langley Research Centr-e, February 1996.

[133] Chu J and Luckring J.M. Experimental surface pressure data obtained on a 65® delta wing across Reynolds
number and Mach number ranges: Volrrme 3 - medium radius leading edge. NASA Technical Memorandum
4645, NASA Langley Research Centre, February 1996.
[134] Chu J and Luckring J.M. Experimental surface pressure data obtained on a 65® delta wing across Reynolds
number and Mach number ranges: Volume 4 - large radius leading edge. NASA Technical Memorandum
4645, NASA Langley Research Centre, February 1996.
[135] Luckring J.M. Reynolds number and leading edge bluntness effects on a 65® delta wing. In 40th AIAA
Aerospace Sciences Meeting and Exibit. AIAA Paper 2001-0907, January 2002.

[136] Luckring J.M. Transonic Reynolds number and leading edge bluntness effects on a 65® delta wing. In 41st
Aerospace Sciences Meeting and Exhibit, AIAA-2003-0753, January 2003.

[137] Luckring J.M. Compressibility and leading edge bluntness effects for a 65® delta wing. In 42nd Aerospace
Sciences Meeting and Exhibit. AIAA-2004-0765, January 2004.
[138] Luckring J.M. Reynolds number, compressibility and leading edge bluntness effects on delta wing aerody­
namics. In 24th International Congress o f the Aeronautical Sciences, pages 1-14. ICAS, 2004.

[139] Schrader K.F, Reynolds G.A, and Novak C.J. Effects of Mach number and Reynolds number on leading
edge vortices at high angles of attack. In 26th AIAA Aerospace Sciences Meeting. AIAA Paper 88-0122,
January 1988.
[140] Konrath R, Klein C, Engler R.H, and Otter D. Analysis of PSP results obtained for the VFE-2 65® delta
wing configuration at sub- and transonic speeds. In 44th AIAA Aerospace Sciences Meeting and Exhibit.
AIAA Paper 2006-60, January 2006.
[141] Konrath R, Schroder A, and Kompenhans J. Analysis of PIV results obtained for the VFE-2 65® delta wing
configuration at sub- and transonic speeds. In 24th AIAA Applied Aerodynamics Conference. AIAA Paper
2006-3003, June 2006.

[142] Schroder A, Agocs J, Frahnert H, Otter D, Mattner H, Kompenhans J, and Konrath R. Application of
stereo PIV to the VFE-2 65® delta wing configuration at sub- and transonic speeds. In 24th AIAA Applied
Aerodynamics Conference. AIAA Paper 2006-3486, June 2006.

[143] Cummings R.M, Forsythe J.R, Morton S.A, and Squires K.D. Computational challenges in high angle of
attack flow prediction. Progress in Aerospace Sciences, 39(5):369-384, July 2003.
[144] Allan M. A CFD Investigation o f Wind Tunnel Interference on Delta Wing Aerodynamics. PhD thesis.
University of Glasgow, October 2002.

[145] Newsome R.W. Euler and Navier-Stokes solutions for flow over a conical delta wing. AIAA Journal,
24(4):552-561, April 1986.
[146] Murman E.M and Rizzi A. Applications of Euler equations to sharp edged delta wings with leading edge
vortices. In AGARD CP-412 Paper 15. Applications of Computational Fluid Dynamics in Aeronautics,
November 1986.

[147] Rizzi A and Muller B. Comparison of Euler and Navier-Stokes solutions for vortex flow over a delta wing.
Aeronautical Journal, 92(914): 145-153, April 1988.
REFERENCES 170

[148] Tsai H.M, Leek C.L, Guarino L, and Lee K.M. Euler/Navier-Stokes analysis of vortical flow about rounded
leading edge delta wing. In 13th AIAA Applied Aerodynamics Conference, pages 703 - 713. AIAA Paper
95-1847, June 1995.

[149] Agrawal S, Barnett R.M, and Robinson B.A. Numerical investigation of vortex breakdown on a delta wing.
AIAA Journal, 30(3):584-591, March 1992.

[150] Kegelman J.T and Roos RW. Effects of leading edge shape and vortex burst on the flowfield of a 70 degree
sweep delta wing. In 27th AIAA Aerospace Sciences Meeting. AIAA-1989-0086, January 1989.

[151] Roos RW and Kegelman J. An experimental investigation of sweep angle influence on delta wing flows. In
28th AIAA Aerospace Sciences Meeting. AIAA Paper 90-0383, January 1990.

[152] Kumar A. On the structure of vortex breakdown. In Proceedings o f the Royal Society o f London, Series
A (Mathematical, Physical and Engineering Sciences), pages 89-110. Royal Society of London, January
1998.
[153] Kumar A. Role of flow-consistent grid in CFD. Computers and Fluids, 28:265-280,1999.

[154] Kumai' A. Accurate development of leading edge vortex using an embedded conical grid. AIAA Journal,
34(10):2038-2046, October 1996.

[155] Sti'ohmeyer D, Orlowski M, Longo J.M.A, Hummel D, and Bergmann A. An analysis of vortex breakdown
predicted by the Euler equations. In ICAS Conference, pages 1189-1200. ICAS-96-1.6.3,1996.

[156] Gordnier R.E. Numerical simulation of a 65® delta wing flowfield. Journal o f Aircraft, 34(4):492-499,
1997.
[157] Gordnier R.E. Computational study of a turbulent delta wing flowfield using two-equation turbulence mod­
els. In 27th AIAA Fluid Dynamics Conference. AIAA Paper 96-2076, June 1996.

[158] Brandsma RJ, Kok J.C, Dol H.S, and Elsenaar A. Leading edge vortex flow computations and comparison
with DNW-HST wind tunnel data. In Proceeds o f the RTO/AVT Symposium on Vortex Flows and High Angle
o f Attack. NATO RTO/AVT, 2001.

[159] Dol H.S, Kok J.C, and Oskam B. Turbulence modelling for leading edge vortex flows. In 40th AIAA
Aerospace Sciences Meeting and Exhibit. AIAA Paper 2002-0843, January 2002.

[160] Bartels R.E and Gatski T.B. Prediction of transonic vortex flows using linear and non-linear turbulent eddy
viscosity models. NASA-TM-2000-210282, NASA Langely Research Center, May 2000.

[161] Morton S.A, Forsythe J, Mitchell A.M, and Hajek D. DES and RANS simulations of delta wing vortical
flows. In 40th AIAA Aerospace Sciences Meeting and Exhibit. AIAA Paper 2002-0587, January 2002.

[162] Mitchell A.M, Molton P, Barberis D, and Délery J. Characterization of vortex breakdown by flow field
and surface measurements. In 38th AIAA Aerospace Science Meeting and Exhibit. AIAA Paper 2000-0788,
January 2000.

[163] Morton S.A, Forsythe J.R, Squires K.D, and Wurtzler K.E. Assessment of unstructured grids for detached-
eddy simulation of high Reynolds number separated flows. In 8th International Conference on Numerical
Grid Generation in Computational Field Simulations, June 2002.

[164] Mitchell A.M, Morton S.A, Forsythe J.R, and Cuimnings R.M. Analysis of delta wing vortical substructures
using detached eddy simulation. AIAA Journal, 44(5):964-972, May 2006.

[165] Gortz S. Detached-eddy simulation of a full-span delta wing at high incidence. In 21st AIAA Applied
Aerodynamics Conference. AIAA Paper 2003-4216, June 2003.

[166] de Cock K.M.J, Kok J.C, van der Ven H, Soemarwoto B.l, Boelens O.J, and Oskam B. Extra large eddy
simulation for delta wings and space launchers. In 24th International Congress o f the Aeronautical Sciences.
ICAS, 2004.

[167] PopeS.B. Turbulent Flows. Cambridge University Press, 2000.

[168] Favre A. Equations des gaz turbulents compressibles. Journal de Mécanique, 4(3):361-390,1965.
il'
'i

REFERENCES 171

■S
[169] Ferziger J.H. Large eddy simulation. In Gatski T.B, Hussaini M.Y, and Lumley J.L, editors, Simulation B
and Modelling o f Turbulent Flows, pages 109-154. ICASE/LaRC Series in Computational Science and **
Engineering, Oxford University Press, 1996.

[170] Wallin S and Johansson A. An explicit algebraic Reynolds stress model for incompresible and compressible
turbulent flows. Journal o f Fluid Mechanics, 403:89-132, January 2000.

[171] Spalart P.R, Jou W-FI, Strelets M, and Allmaras S .R. Comments on the feasibility of LES for wings and on a
hybrid RANS/LES approach. In Advances in DNS/LES, IstAFSO R International Conference on DNS/LES.
AFOSR, August 1997.

[172] Shut M, Spalai't P.R, Strelets M, and Travin A. Detached eddy simulation of an airfoil at high angle of
attack. In 4th International Symposium Engineering Turbulence modelling and Measurements, May 1999.

[173] Badcock K.J, Richards B.E, and Woodgate M.A. Elements of computational fluid dynamics on block stru-
cured grids using implicit solvers. Progress in Aerospace Sciences, 36:351-392,2000.

[174] Roe PL. Approximate riemann solvers, parameter vectors and difference schemes. Journal o f Computa­
tional Physics, 43:357-372,1981.

[175] Osher S and Solomon F. Upwind difference schemes for hyperbolic systems of conservation laws. Mathe­
matics o f Computing, 38:339-374,1982.

[176] Jameson A. Time dependent calculations using multigrid with applications to unsteady flows past airfoils
and wings. In lOth Computational Fluid Dynamics Conference. AIAA Paper 91-1596, June 1991.

[177] Spalart PR. Young person’s guide to detached-eddy simulation grids. NASA/CR-2001-211032, NASA
Langley Research Centre, July 2001.

[178] Lawrie D. Investigation o f Cavity Flows at Low and High Reynolds Numbers using Computational Fluid
Dynamics. PhD thesis. Department of Aerospace Engineering, University of Glasgow, January 2004.

[179] Nayyai' P. CFD Analysis o f Transonic Turbulent Cavity Flows. PhD thesis. Department of Aerospace
Engineering, University of Glasgow, August 2005.

[180] Délery J.M. Physics of vortical flows. Journal o f Aircraft, 29(5):856-876, September - October 1992.

[181] Elsenaar A, Hjemberg L, Biitefisch K-A, and Bannink W.J. The international vortex flow experiment. In
Validation o f Computational Fluid Dynamics - AGARD-CP-437 Volume 1, pages 9.1-9.23. AGARD, 1988.

[182] Hummel D and Redeker G. A new vortex flow experiment for computer code validation. In Proceedings
o f the RTO-AVT Symposium on “Advanced Flow Management; Part A - Vortex Flows and High Angle o f
Attack fo r Military Vehicles” - RTO-MP-069(1), pages 8.1-8.32. NATO RTO, 2001.

[183] Hummel D. Effects of boundary layer formation on the vortical flow above slender delta wings. In RTO-
M P-AVT-lll, pages 30.1-30.22. NATO, October 2004.

[184] Boelens O.J. Private communication.

[185] Gortz S and Rizzi A. Computing the high alpha aerodynamics of delta wings - evaluation and analysis. In
39th AIAA Aerospace Sciences Meeting and Exhibit. AIAA Paper 2001-0115, January 2001.

[186] Gortz S. Steady and unsteady CFD simultations of vortex breakdown over delta wings. Master’s thesis,
Royal Institute of Technology, Stockliolm, Sweden, 2002.

[187] Gortz S, Rizzi A, and Munukka K. Computational study of vortex breakdown over swept delta wings. In
17th AIAA Applied Aerodynamics Conference. AIAA Paper 99-3118, June/July 1999.

[188] Schiavetta L.A, Boelens O.J, and Fritz W. Analysis of transonic flow on a slender delta wing using CFD. In
24th AIAA Applied Aerodynamics Conference. AIAA Paper 2006-3171, June 2006.

[189] FLOWer, Installation and User Handbook.

[190] Kok J.C. Mathematical physical model of ENSOLV version 3.20; a flow solver for 3D Euler/Navier-Stokes
equations in ai'bitrary multi-block domains. NLR-CR-2000-621, Nationaal Lucht- en Ruimtevaartlaborato-
rium, NLR, 2000.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy