You are on page 1of 299

Enhanced Wind Tunnel Techniques

and Aerodynamic Force Models


for Yacht Sails



Heikki Hansen



A thesis submitted in partial fulfilment of the requirements for
the degree of Doctor of Philosophy in Mechanical Engineering,
The University of Auckland, 2006.



Supervised by
Associate Professor Peter J. Richards
&
Professor Peter S. Jackson


Department of Mechanical Engineering
The University of Auckland
Auckland, New Zealand


































Copyright Heikki Hansen 2006. All rights reserved.
Department of Mechanical Engineering
The University of Auckland
Auckland, New Zealand


The
University
YACHT
RESEARCH
of Auckland UNIT
Abstract

Accurate prediction of performance is an important aspect of modern sailing yacht design and
provides a competitive advantage on the racecourse and in the marketplace. Although wind
tunnel testing of yacht sails is a common tool for obtaining input data for Velocity Prediction
Programs, its results have not been validated against aerodynamic full-scale measurements as
quality full-scale data is rare. Wind tunnel measurements are conducted at the Twisted Flow
Wind Tunnel of The University of Auckland and are compared to the full-scale aerodynamic
force measurements from the Berlin Sail-Force-Dynamometer. To realise this comparison
wind tunnel techniques and aerodynamic force models for yacht sails are enhanced; this in
turn also improves the accuracy of Velocity Prediction Programs.
Force and surface pressure measurements were conducted demonstrating that the interaction
of the hull/deck with the sails has a significant effect on the side force and the force
perpendicular to the deck plane, and that this should be considered in aerodynamic analysis of
sails and the performance prediction of yachts.
The first Real-Time Velocity Prediction Program for wind tunnel testing has been developed
and implemented as an additional module of FRIENDSHIP-Equilibrium. Model sails can
now be trimmed based on the full-scale performance of the yacht, and at the correct heel
angle, which makes the trimming process in the wind tunnel much more similar to the real
life situation.
Improved aerodynamic force models have been developed from realistically depowered sail
trims obtained with the Real-Time Velocity Prediction Program. An empirical model that
describes the force and moment changes due to depowering in detail has been developed and
implemented. The standard semi-empirical trim parameter model, which expresses
depowering in a more generic way, has been enhanced based on aerodynamic principles and
validated against the wind tunnel results.
Utilising the enhanced wind tunnel techniques and aerodynamic force models, a generally
good qualitative and quantitative agreement with the full-scale data is achieved. Remaining
challenges associated with full-scale and wind tunnel tests are however also highlighted and,
based on this work alone, a conclusive judgement that scaling effects are negligible cannot be
made.

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails iii

The
University
YACHT
RESEARCH
of Auckland UNIT
Acknowledgements

During the course of this project, lasting several years, I had the privilege to receive guidance
and support from numerous exceptional, dedicated and inspiring people. It is very gratifying
to have the opportunity to thank some of them here now.
First and foremost I want to thank my supervisors Professor Peter Jackson and Associate
Professor Peter Richards. Peter Jackson, thank you very much for encouraging me to come to
New Zealand to take on this research, developing the idea for the project, guiding me through
the early part of this work and remaining involved after your move to the University of
Canterbury. Peter Richards, thank you very much for taking over as my main supervisor and
always providing accessible, dedicated and practical guidance. I am also extremely
appreciative of the input from my advisory committee; Burns Fallow and Professor John
Chen.
Working with the staff and postgraduate students of the Yacht Research Unit and the
Mechanical Engineering Department has been inspiring and made this a very educational and
enjoyable journey. I feel blessed to have spent time in your company and thank you all.
David Le Pelley, your patience, dedication and hands-on approach played an important role
in this project and enriched the work environment at the Yacht Research Unit immensely.
Professor Richard Flay and Angelo Tempia, our discussions about wind tunnel testing and
aerodynamic force modelling were especially enlightening. I am also very thankful to the
technicians, Ken Brich, Brian Watson and Grant Clendons team, for producing excellent
models and equipment.
An integral part of the project was the cooperation with the Technical University Berlin and
FRIENDSHIP SYSTEMS and I am extremely grateful to Dr. Karsten Hochkirch for the full-
scale data, the use of his source-code and the continual software development. Karsten, thank
you very much for being an inspiring mentor and sharing your knowledge. Without your
enthusiastic support much less would have been achieved. At the Technical University Berlin
I especially recognise Professor Gnther Clauss and Bernhardt Krger for giving me the
opportunity to take part in their research and making the full-scale data available. At both
institutions staff and students made me feel incredibly welcome and it was an honour to work
with you. I had a great time. Bardo Krebber, your full-scale sail shape analysis was very
helpful and was much appreciated.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails v
Acknowledgements
At several conferences I had interesting conversations with numerous colleagues and received
important feedback for which I am very thankful. Professor Lars Larsson and Ian Campbell,
discussions with you were of particular relevance. Invaluable data and information was also
provided by Efva Ingebrigtsen of North Sails Denmark, Nick Locke of Opus Central
Laboratories and Chris Mitchell of Applied Engineering Services. Donna, thank you very
much for proofreading the thesis.
The IPENZ Craven Scholarship and the Yacht Research Unit lessened the financial burden of
being a postgraduate student and their contribution was much appreciated. I am also grateful
to the Research Committee for providing means to present this work at international
conferences through the Graduate Research Fund.
Distractions provided by PGSA and spark* greatly enriched my postgraduate experience and
taught me a lot. I thank everybody who was part of them; friendships forged in battle are
rarely forgotten.
Finally and most importantly I thank my family for their love, unconditional support,
enduring encouragement and financial contribution. Wewett, Harald, Schura, Omi and Opa;
this is yours
vi
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
Table of Contents

Abstract..................................................................................................................................... iii
Acknowledgements ....................................................................................................................v
Table of Contents .................................................................................................................... vii
List of Figures........................................................................................................................... xi
List of Tables........................................................................................................................ xxiii
Nomenclature .........................................................................................................................xxv
Symbols .............................................................................................................................xxv
Greek Symbols...................................................................................................................xxx
Subscripts and other Notations ........................................................................................ xxxi
Abbreviations and Names............................................................................................... xxxii
1 Introduction 1
1.1 Preface..........................................................................................................................1
1.2 Velocity Prediction Programs ......................................................................................2
1.3 Large-Scale Testing......................................................................................................3
1.4 Twisted Flow Wind Tunnel..........................................................................................6
1.4.1 Wind Tunnel Testing as a Sail Design Tool ..................................................8
1.4.2 Trimming and Depowering of Sails.............................................................10
1.5 Motivation ..................................................................................................................11
1.6 Project Objectives ......................................................................................................12
1.7 Methodology ..............................................................................................................13
2 Wind Tunnel Testing of Yacht Sails 15
2.1 Introduction ................................................................................................................15
2.2 Terminology of Sailing ..............................................................................................15
2.3 Aerodynamics of Sailing............................................................................................17
2.3.1 Effective Angle Theory................................................................................21
2.3.2 Vertical Speed Gradient and Twist in Onset Flow ......................................24
2.3.3 Position of Centre of Effort..........................................................................29
2.4 Twisted Flow Wind Tunnel........................................................................................31
2.4.1 Force Measurements ....................................................................................32
2.4.2 Scaling Effects .............................................................................................36
2.5 Experimental Set-Up..................................................................................................38
2.5.1 Wind Tunnel Model .....................................................................................38
2.5.2 Vertical Speed Gradient and Twist Profiles.................................................39
2.6 Summary ....................................................................................................................41

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails vii
Table of Contents
3 Hull and Sail Force Interaction 43
3.1 Introduction................................................................................................................ 43
3.2 Force Measurements of Hull and Deck Forces.......................................................... 44
3.3 Pressure Measurements.............................................................................................. 47
3.4 Comparison of Force and Pressure Measurements.................................................... 52
3.5 Lift and Drag Forces of Hull/Deck and Rig .............................................................. 53
3.5.1 Influence of Sails on Hull/Deck Lift and Drag............................................ 59
3.6 Hull/Deck Forces Perpendicular to Deck Plane ........................................................ 63
3.6.1 Forces Perpendicular to Deck Plane at Heel ............................................... 65
3.7 Sail Forces Perpendicular to Deck Plane................................................................... 67
3.8 VPP Modelling of Forces Perpendicular to Deck Plane............................................ 68
3.9 Conclusions................................................................................................................ 69
4 Real-Time Velocity Prediction Program 71
4.1 Introduction................................................................................................................ 71
4.2 Standard Semi-Empirical VPPs ................................................................................. 75
4.2.1 Aerodynamic Sail Force Model with Trim Parameters............................... 76
4.3 Real-Time VPP Implementation................................................................................ 80
4.3.1 Real-Time VPP LabVIEW Application ...................................................... 81
4.3.2 FRIENDSHIP-Equilibrium ......................................................................... 88
4.3.3 Aerodynamic Wind Tunnel Force Model.................................................... 91
4.4 Wind Tunnel Constraints........................................................................................... 93
4.4.1 Maximising Speed for Constant Effective Wind Angle.............................. 93
4.4.2 Maximising Speed for Constant Apparent Wind Angle.............................. 96
4.4.3 Effective Angle Correction.......................................................................... 98
4.5 Real-Time VPP Applications..................................................................................... 99
4.6 Conclusions.............................................................................................................. 101
5 Aerodynamic Force Modelling 103
5.1 Introduction.............................................................................................................. 103
5.2 Testing Procedure .................................................................................................... 103
5.3 Improved Aerodynamic Moment Model ................................................................. 105
Depowering Model with Trim Parameter Power 5.4 .................................................... 115
5.4.1 Comparison to Real-Time VPP Results .................................................... 121
5.4.2 Generic Power Model ............................................................................. 125
Depowering Changes to x 5.4.3
CoE
and C ........................................................ 129
Mf
5.4.4 Direct Force and Moment Power Model ................................................ 131
5.5 Assessing the Standard Trim Parameter Model....................................................... 135
Improved Flat and Twist Depowering Model 5.6 .......................................................... 144
5.7 Conclusions.............................................................................................................. 157

viii
YACHT
RESEARCH
UNIT
The
University
of Auckland

6 Wind Tunnel and Full-Scale Comparison 159
6.1 Introduction ..............................................................................................................159
6.2 Full-Scale Aerodynamic Data Acquisition ..............................................................159
6.3 Wind Tunnel Study of Flow Above Mast ................................................................164
6.3.1 Analysis of Cobra Probe Measurements ....................................................167
6.3.2 Influence of Sails on Flow Above Mast.....................................................169
6.3.3 Model for Correcting Full-Scale Wind Measurements..............................176
6.4 Full-Scale Aerodynamic Data Analysis ...................................................................179
6.5 Wind Tunnel Data ....................................................................................................192
6.6 Comparison of Aerodynamic Forces: Mainsail & Spinnaker ..................................194
6.7 Comparison of Aerodynamic Forces: Mainsail & Jib..............................................200
6.7.1 Comparison of Sail Shapes ........................................................................209
6.8 Conclusions ..............................................................................................................212
7 Conclusions and Recommendations 215
7.1 Conclusions ..............................................................................................................215
7.2 Recommendations for Future Work.........................................................................217
Appendices 221
A Wind Tunnel Model of Dyna 223
A.1 Design Considerations..............................................................................................223
A.2 Model Rig.................................................................................................................225
A.3 Model Sails...............................................................................................................229
B Experimental Procedures 231
B.1 Corrections for Hull/Deck Force Measurements .....................................................231
B.2 Meshing of Hull/Deck for Pressure Mapping ..........................................................232
B.3 Airflow below Model in Wind Tunnel Trough........................................................234
C Velocity Prediction Program Definitions 239
C.1 Coordinate Systems..................................................................................................239
C.2 Real-Time VPP Output File Structure .....................................................................242
C.3 FRIENDSHIP-Equilibrium User Manual ................................................................243
C.3.1 WindTunnel Force Module .....................................................................243
C.3.2 HansenRig Force Module........................................................................244
C.3.3 HansenRigDirect Force Module..............................................................247
C.4 Hydrodynamic Model of Dyna in FRIENDSHIP-Equilibrium................................249


Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails ix
Table of Contents
D Mathematical Techniques 253
D.1 General Force and Moment System......................................................................... 253
D.2 Central Axis System................................................................................................ 254
D.3 Lifting-Line Theory to Model Sail Twist ................................................................ 256
D.4 Transformation of Flow Measurements to Heeled Plane ........................................ 260
References 263

x
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
List of Figures
Page
Figure 1.1: The Berlin Sail-Force-Dynamometer Dyna ..........................................................4
Figure 1.2: The original Twisted Flow Wind Tunnel (TFWT) for testing yacht sails at
The University of Auckland..................................................................................7
Figure 2.1: Diagram showing components of a yacht ...........................................................15
Figure 2.2: Pressure distribution along the chord of a sail and the resulting lift and
drag......................................................................................................................17
Figure 2.3: Apparent wind angle, angle of attack and sheeting angle relationship...............19
Figure 2.4: Lift and drag coefficient of mainsail and jib combination against apparent
wind angle ...........................................................................................................19
Figure 2.5: The wind triangle ................................................................................................21
Figure 2.6: Diagram of onset flow in the horizontal plane....................................................22
Figure 2.7: Diagram of onset flow component in the heeled plane.......................................22
Figure 2.8: Atmospheric boundary layer ...............................................................................25
Figure 2.9: Vertical speed gradient and twist in onset flow onto a sail of a yacht on
port tack ignoring the leeway angle ....................................................................25
Figure 2.10: Speed profiles for a 10m IMS cruiser/racer and a Volvo Ocean 60 for true
wind angles of 40 and 160.................................................................................27
Figure 2.11: Twist profiles for a 10m IMS cruiser/racer and a Volvo Ocean 60 for true
wind angles of 40 and 160.................................................................................27
and V /V Figure 2.12:
A
(twist) over span of sails plotted as a function of
T S T
for a
VO60 yacht .........................................................................................................28
and V /V Figure 2.13:
A
(twist) over span of sails plotted as a function of
T S T
for a
10m IMS cruiser/racer.........................................................................................28
Figure 2.14: Twisted Flow Wind Tunnel (TFWT) with and without flow twisting vanes
installed ...............................................................................................................31
Figure 2.15: Schematic diagram of the Twisted Flow Wind Tunnel (TFWT) ........................31
Figure 2.16: Engineering drawing of the Twisted Flow Wind Tunnel (TFWT) .....................32
Figure 2.17: Six-component force balance under the TFWT floor .........................................33
Figure 2.18: Origin of force balance and location of LVDTs .................................................33
Figure 2.19: Definition of the coordinate systems and sign convention .................................34
Figure 2.20: Membrane fold height ratio.................................................................................37
Figure 2.21: Computer representation of wind tunnel model of Dyna....................................39
Figure 2.22: Internal frame of model with remote control winches ........................................39
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xi
List of Figures
Figure 2.23: Vertical speed gradient and twist profiles for upwind and downwind
conditions calculated for Dyna and modelled in wind tunnel plotted
against ratio of mast height on vertical axis ....................................................... 40
Figure 3.1: Wind tunnel model with internal six-component force balance and
connection showing internal framework structure and hull shell detached
and assembled..................................................................................................... 45
Figure 3.2: Hull/deck shell of model connected to internal frame through six-
component force balance .................................................................................... 45
Figure 3.3: Mechanism to connect or disconnect hull/deck from internal frame ................. 45
Application of load for offset zero measurement of M Figure 3.4: ..................................... 46
Z
Figure 3.5: Components of pressure tapped model............................................................... 48
Figure 3.6: Inside picture of pressure tapped model ............................................................. 48
Figure 3.7: Outside picture of pressure tapped model .......................................................... 48
Figure 3.8: Location and numbering sequence of pressure taps on hull and deck of
wind tunnel model .............................................................................................. 49
Figure 3.9: Computer model for mapping the measured surface pressures with the
mesh consisting of 7856 elements and divided into five regions ....................... 50
Figure 3.10: Triangular element i of the boundary mesh........................................................ 51
Figure 3.11: Comparison of C
D
, C and C
L Fz
of hull/deck calculated from force and
pressure measurements for model without sails in upwind configuration at
0 heel ................................................................................................................. 52
Figure 3.12: Comparison of C
D
, C and C
L Fz
of total and above water hull/deck from
pressure measurements for model without sails in upwind configuration at
0 heel ................................................................................................................. 52
Figure 3.13: Windage drag coefficient in horizontal plane with upright model in upwind
configuration....................................................................................................... 54
Figure 3.14: Windage lift coefficient in horizontal plane with upright model in upwind
configuration....................................................................................................... 54
Figure 3.15: Surface pressure plots for model without sails on port tack at
A
of 90 and
0 heel showing leeward and windward side of hull .......................................... 58
Horizontal drag coefficient (C Figure 3.16:
DA
) of hull/deck for the model with and
without sails in upwind and downwind configuration........................................ 59
Horizontal lift coefficient (C Figure 3.17:
LA
) of hull/deck for the model with and
without sails in upwind and downwind configuration........................................ 59
Figure 3.18: Drag coefficient of hull/deck for the model with and without sails in
upwind configuration at different heel angles plotted in horizontal plane
(coordinate system A) as C
DA
and in the heeled plane (coordinate system
B) as C .............................................................................................................. 61
D
Figure 3.19: Lift coefficient of hull/deck for the model with and without sails in upwind
configuration at different heel angles plotted in horizontal plane
(coordinate system A) as C
LA
and in the heeled plane (coordinate system
B) as C ............................................................................................................... 61
L
xii
YACHT
RESEARCH
UNIT
The
University
of Auckland

Figure 3.20: Surface pressure plots for model on port tack at 0 heel without sails and
with mainsail and jib at two apparent wind angles showing leeward and
windward side of hull ..........................................................................................62
Drive force and side force coefficient (C and C Figure 3.21:
Fx Fy
) of total model with
mainsail and jib and of hull/deck with and without sails for the upright
model in upwind configuration ...........................................................................62
Force coefficient perpendicular to deck plane (C Figure 3.22:
Fz
) of hull/deck for model
with and without sails in upwind and downwind configuration .........................63
Figure 3.23: Surface pressure plots for model on port tack at 0 heel without sails and
with mainsail and jib at different apparent wind angles .....................................64
Figure 3.24: Surface pressure plots for model on port tack at 0 heel without sails and
with mainsail and spinnaker at different apparent wind angles ..........................65
Force coefficient perpendicular to deck plane (C Figure 3.25:
Fz
) of hull/deck for model
without sails at different heel angles...................................................................66
Figure 3.26: Surface pressure plots for model on port tack at 30 apparent wind angle
without sails and with mainsail and jib at different heel angles .........................66
Force coefficient perpendicular to deck plane (C Figure 3.27:
Fz
) of hull/deck for model
with mainsail and jib at different heel angles......................................................67
Force coefficient perpendicular to deck plane (C Figure 3.28:
Fz
) of mainsail, jib and rig
at different heel angles ........................................................................................67
Figure 3.29: Surface describing the total force coefficient perpendicular to deck plane
(C ) as a function of effective wind angle (
Fz eff
) and heel angle.........................68
Boat speed (V Figure 3.30:
S
) with and without including force perpendicular to deck
plane (F ) and true wind speeds (V ) for different true wind angles ( ) ...........68
Z T T
Figure 4.1: Traditional wind tunnel test procedure ...............................................................72
Figure 4.2: Wind tunnel test procedure with the Real-Time VPP.........................................74
Figure 4.3: Forces and moments acting on a sailing yacht that are modelled and
balanced by a three equation VPP.......................................................................75
Generic semi-empirical VPP with flat (f) and reef (r) as trim parameters Figure 4.4: ..........75
Figure 4.5: Schematic description of the implementation of the Real-Time VPP in the
TFWT..................................................................................................................81
Figure 4.6: Screen shot of the graphical user interface of the Real-Time VPP
LabVIEW application .........................................................................................82
Reduction in apparent wind speed (V Figure 4.7:
A
/V ) with height (z/z
Aref mast
) modelled
in Real-Time VPP and calculated from vertical speed profiles .........................85
Figure 4.8: Schematic description of FRIENDSHIP-Equilibrium and the Real-Time
VPP module for wind tunnel testing ...................................................................89
Figure 4.9: Screen shot of the graphical user interface of FRIENDSHIP-Equilibrium........90
V optimised using r and f for either constant Figure 4.10:
S T
or
eff
at different V ...............94
T
V at different V Figure 4.11: ratio of optimisation for constant
S eff
against constant .......94
T T
Set of achievable V Figure 4.12:
S
at
eff
=40 and 65 for V
T
=12m/s using reef and flat
between 1 and 0.6 in increments of 0.05.............................................................95
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xiii
List of Figures
Polar plot of V optimised at V Figure 4.13:
S T
=12m/s using reef and flat for either
constant (left) or
T eff
(right)............................................................................ 95
V at different V Figure 4.14: ratio of optimisation for constant
S A
against constant ....... 97
T T
Set of achievable V Figure 4.15:
S
at
A
=40 and 65 for V
T
=10m/s using r and f between
1 and 0.6 in increments of 0.05........................................................................... 97
The Real-Time VPP predictions (left value is V Figure 4.16:
S
and right value is heel
angle ()) for two sail trims with the model dynamically heeled to the
predicted ........................................................................................................ 100
Residual pure moment coefficient about y-axis (C Figure 5.1:
M0y
) for fully powered-up
sails against ................................................................................................. 106
eff
Pitching moment coefficient (C Figure 5.2:
My
) for fully powered-up sails against
effective wind angle (
eff
) ................................................................................. 106
The ratio p of the moment coefficient in line with the force (C Figure 5.3:
Mf
) and the
force coefficient (C
F
) against ...................................................................... 107
eff
Figure 5.4: Magnitude of the total moment coefficient (|C
M
|) and the components
perpendicular to and in line with the resultant force (|C
Mv
| and |C
Mf
|) vs. .. 107
eff
Figure 5.5: x and y-axis components of total moment coefficient (C
M
) and the
component perpendicular the resultant force (C
Mv
) vs. .............................. 108
eff
x-direction component of moment perpendicular to resultant force (M Figure 5.6:
Vx
)
against x-direction component of moment (M
X
) for different force types ....... 109
y-direction component of moment perpendicular to resultant force (M Figure 5.7:
Vy
)
against y-direction component of moment (M ) for different force types........ 110
Y
Figure 5.8: Force direction () in the deck plane for hull, rig and sails, and the total
model against .............................................................................................. 111
eff
x-component of the moment coefficient in line with the force (C Figure 5.9:
Mfx
) for
hull, rig and sails, and the combination of the two vs. ............................... 111
eff
y-component of the moment coefficient in line with the force (C Figure 5.10:
Mfy
) for
hull, rig and sails, and the combination of the two vs. ............................... 111
eff
Total C Figure 5.11:
Mf
of mainsail, jib, rig and hull for different heel angles vs. ........... 112
eff
Measured total C Figure 5.12:
Mf
of mainsail, jib, rig and hull and VPP B-spline fit ............ 112
Figure 5.13: Boat speed of Dyna under mainsail and jib calculated with and without
considering C
Mf
for different V vs. ............................................................ 113
T eff
Figure 5.14: Heel angle of Dyna under mainsail and jib calculated with and without
considering C
Mf
for different V vs. ............................................................ 113
T eff
Measured total C Figure 5.15:
Mf
of hull, rig and mainsail with jib, genoa and spinnaker
based on sailcloth area...................................................................................... 114
Figure 5.16: Lift coefficient (C
L
) from wind tunnel tests with Real-Time VPP for
different true wind speeds (V ) ......................................................................... 116
T
Figure 5.17: Surface describing the lift coefficient (C
L
) as a function of effective wind
angle (
eff
) and the depowering parameter power............................................. 116
xiv
YACHT
RESEARCH
UNIT
The
University
of Auckland

The optimum lift coefficient (C Figure 5.18:
Lopt
) as a function of effective wind angle
(
eff
) ...................................................................................................................117
Surface describing the lift coefficient ratio (R Figure 5.19: ) as a function of
L eff
and
power.................................................................................................................117
Lift coefficient (C ) for all tested true wind speeds (V Figure 5.20:
L T
) from Real-Time
VPP and modelled by power.............................................................................118
Surface describing the parasitic drag coefficient ratio (R Figure 5.21:
Dp
) as a function of

eff
and power ....................................................................................................119
C for all tested true wind speeds (V Figure 5.22:
Dp T
) from Real-Time VPP and modelled
by power............................................................................................................119
Surface describing the centre of effort height ratio (R Figure 5.23:
zCoE
) as a function of

eff
and power ...................................................................................................120
z Figure 5.24:
CoE
for all tested true wind speeds (V
T
) from Real-Time VPP and
modelled by power ...........................................................................................120
Measured x Figure 5.25:
CoEopt
of mainsail, jib, rig and hull and VPP B-spline curve fit .....122
Measured C Figure 5.26: of mainsail, jib, rig and hull and VPP B-spline curve fit .......122
FzOpt
Values of trim parameter power for four true wind speeds (V Figure 5.27:
T
) from Real-
Time VPP and power model VPP calculations ..............................................123
Boat speed (V ) for four true wind speeds (V Figure 5.28:
S T
) predicted by Real-Time
VPP and VPP with power parameter ................................................................123
Heel angle () for four true wind speeds (V Figure 5.29:
T
) predicted by Real-Time VPP
and VPP with power parameter.........................................................................124
Lift coefficient (C ) for four true wind speeds (V Figure 5.30:
L T
) from Real-Time VPP
and power model VPP calculations................................................................124
Centre of effort height (z Figure 5.31:
CoE
) for four true wind speeds (V
T
) from Real-
Time VPP and power model VPP calculations ..............................................124
Parasitic drag coefficient (C ) for four true wind speeds (V Figure 5.32:
Dp T
) from Real-
Time VPP and power model VPP calculations ..............................................124
C /C Figure 5.33:
L Lopt
ratio plotted against trim parameter power for different effective
wind angles (EWA) ...........................................................................................125
z Figure 5.34:
CoE
/z
CoEopt
ratio plotted against trim parameter power for different
effective wind angles (EWA)............................................................................125
C /C Figure 5.35:
Dp DpOpt
ratio plotted against trim parameter power for different
effective wind angles (EWA)............................................................................126
C Figure 5.36:
D
/C
Dopt
ratio plotted against trim parameter power for different effective
wind angles (EWA) ...........................................................................................126
Values of trim parameter power from Real-Time VPP measurements and
VPP calculations using standard and generic form of power
Figure 5.37:
............................127
Centre of effort height (z Figure 5.38:
CoE
) predicted by Real-Time VPP and VPP with
trim parameter power in standard and generic form.........................................127
Drag coefficient (C Figure 5.39:
D
) predicted by Real-Time VPP and VPP with trim
parameter power in standard and generic form.................................................128
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xv
List of Figures
Boat speed ratio of V Figure 5.40:
S
predicted by Real-Time VPP and generic power
model relative to V from standard power model .......................................... 128
S
Drag coefficient (C Figure 5.41:
D
) predicted by Real-Time VPP and VPP with trim
parameter power in standard and second generic form.................................... 129
Boat speed ratio of V Figure 5.42:
S
predicted by Real-Time VPP and second generic
power model relative to V from standard power model............................. 129
S
x Figure 5.43:
CoE
/x
CoEopt
ratio plotted against trim parameter power for different
effective wind angles (EWA) ........................................................................... 130
C Figure 5.44:
Mf
/C
MfOpt
ratio plotted against trim parameter power for different effective
wind angles (EWA) .......................................................................................... 130
C /C Figure 5.45:
Fz FzOpt
ratio plotted against trim parameter power for different effective
wind angles (EWA) .......................................................................................... 131
Boat speed ratio of V Figure 5.46:
S
predicted by Real-Time VPP and advanced power
model, which additionally adjusts x
CoE
and C
Mf
, relative to V
S
from
standard power model .................................................................................... 131
C /C Figure 5.47:
Fx FxOpt
ratio plotted against trim parameter power for different effective
wind angles (EWA) .......................................................................................... 132
C /C Figure 5.48:
Fy FyOpt
ratio plotted against trim parameter power for different effective
wind angles (EWA) .......................................................................................... 132
C Figure 5.49:
My
/C
MyOpt
ratio plotted against trim parameter power for different
effective wind angles (EWA) ........................................................................... 133
C /C Figure 5.50:
Mz MzOpt
ratio plotted against trim parameter power for different
effective wind angles (EWA) ........................................................................... 133
Values of trim parameter power from Real-Time VPP measurements and
VPP calculations using advanced and direct form of power
Figure 5.51:
............................ 134
Boat speed ratio of V Figure 5.52:
S
predicted by Real-Time VPP and VPP with direct
power model relative to V from advanced power model ........................... 134
S
Figure 5.53: Windage drag coefficient of hull and rig in upwind configuration at 0 heel
vs. ................................................................................................................ 136
eff
Parasitic drag coefficient (C Figure 5.54:
Dp
) of mainsail and jib with and without
windage compnent at 0 heel vs. ................................................................ 136
eff
Flat and reef values chosen by VPP in different V Figure 5.55: for a range of ............. 138
T eff
Flat and twist values chosen by VPP in different V Figure 5.56: for a range of
T eff
with
twist relative to boom and c =8......................................................................... 138
t
Flat and twist values chosen by VPP in different V Figure 5.57: for a range of
T eff
with
twist relative to DWL and c =8......................................................................... 139
t
Flat and twist values chosen by VPP in different V Figure 5.58: for a range of
T eff
with
twist relative to DWL and c =12....................................................................... 139
t
Centre of effort height (z Figure 5.59:
CoE
) for four true wind speeds (V
T
) from VPP
calculations using trim parameter power and different combinations and
versions of reef, flat and twist........................................................................... 140
xvi
YACHT
RESEARCH
UNIT
The
University
of Auckland

Lift coefficient (C ) for four true wind speeds (V Figure 5.60:
L T
) from VPP calculations
using trim parameter power and different combinations and versions of
reef, flat and twist ..............................................................................................141
Total drag coefficient (C Figure 5.61:
D
) for four true wind speeds (V
T
) from VPP
calculations using trim parameter power and different combinations and
versions of reef, flat and twist ...........................................................................142
Boat speed (V ) four true wind speeds (V Figure 5.62:
S T
) from VPP calculations using
different combinations and versions of reef, flat and twist as ratio of V
S

obtained using power model...........................................................................143
Base twist (t Figure 5.63: ) for fully powered-up mainsail and jib vs.
eff 0
assuming
z is 42% of mast height ..........................................................................146
CoEminDi
z Figure 5.64:
CoE
for four V
T
from VPP calculations using power model or flat and
twist with t
2
and C function of (C /C
0 Dp L Lopt
) ......................................................146
C for four V Figure 5.65:
L T
from VPP calculations using power model or flat and twist
with t
2
and C function of (C /C
0 Dp L Lopt
) ..............................................................146
C Figure 5.66:
D
for four V
T
from VPP calculations using power model or flat and twist
with t
2
and C function of (C /C
0 Dp L Lopt
) ..............................................................147
V for four V from VPP calculations using flat and twist with t and C Figure 5.67:
S T Dp 0

function of (C
2
/C plotted as ratio of V
L Lopt
) obtained with power model ......147
S
Calculated C Figure 5.68:
D
plotted against C
D
measured in the wind tunnel with the
Real-Time VPP for sails at different depowering levels for a range of
effective wind angles (EWA)............................................................................148
Parasitic drag coefficient (C Figure 5.69:
Dp
) excluding windage calculated from Real-
Time VPP measurements plotted against C /C
L Lopt
for different ..................149
eff
Parasitic drag coefficient (C Figure 5.70:
Dp
) excluding windage calculated from Real-
Time VPP measurements plotted against (C
2
/C
L Lopt
) for different ..............149
eff
2
C Figure 5.71:
D
calculated from twist with t and c =8 and C function of (C /C
0 t Dp L Lopt
)
plotted against C
D
measured in the wind tunnel with the Real-Time VPP.......151
C /C Figure 5.72:
L Lopt
from depowering sails with Real-Time VPP at different
eff

plotted against 1-z
CoE
/z ..............................................................................151
CoEopt
Flat and twist values chosen by VPP in different V Figure 5.73: for a range of
T eff
with
C a function of twist.........................................................................................154
L
z Figure 5.74:
CoE
for four V
T
from VPP calculations using power or flat and twist with
C a function of twist.........................................................................................155
L
C for four V from VPP calculations using power or flat and twist with C Figure 5.75:
L T L

a function of twist ..............................................................................................155
C Figure 5.76:
D
for four V from VPP calculations using power or flat and twist with C
T L

a function of twist ..............................................................................................155
V for four V from VPP calculations using flat and twist with C Figure 5.77:
S T L
a
function of twist plotted as ratio of V obtained with power model ...............155
S
Figure 6.1: Schematic diagram of the Berlin Sail-Force-Dynamometer Dyna with its
internal rigid aluminium frame which allows the aerodynamic and
hydrodynamic forces to be measured separately...............................................160
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xvii
List of Figures
Figure 6.2: Aluminium frame of Dyna looking forward towards mast............................... 161
Figure 6.3: Aluminium frame of Dyna inside the hull shell with arrows indicating the
position and orientation of the six force transducers that connect the
internal frame to the hull shell to make up the six-component force balance
to measure the aerodynamic rig loads .............................................................. 162
Figure 6.4: Cup anemometer with direction vane and calorimetric sensor anemometer
on Dyna............................................................................................................. 164
Figure 6.5: Flow measurements above the mast with a Cobra Probe attached to free
standing support rig .......................................................................................... 165
Figure 6.6: Cobra Probe positioned 127mm above the mast head with model heeled ....... 165
Figure 6.7: Change in wind speed for back-to-back runs with and without adjusting for
fluctuations in reference pressure (q
ref
) ............................................................ 168
Reference flow speed factor (f Figure 6.8:
Vref
) due to the presence of the model in the
tunnel for different sail sets vs. ................................................................... 168
eff
Change in wind speed (V/V Figure 6.9:

) 127mm above mast due to presence of


mainsail with jib or spinnaker........................................................................... 169
Figure 6.10: Change in wind speed (V/V

) 127mm above mast due to presence of


mainsail with genoa .......................................................................................... 169
Figure 6.11: Change in effective wind angle (
eff
) 127mm above mast due to presence
of mainsail with jib or spinnaker ...................................................................... 170
Figure 6.12: Change in effective wind angle (
eff
) 127mm above mast due to presence
of mainsail with genoa...................................................................................... 170
Figure 6.13: Illustration of change in effective wind angle (
eff
) due to upwash in z-
plane.................................................................................................................. 171
Figure 6.14: Change in inclination angle () 127mm above mast due to presence of
mainsail with jib or spinnaker........................................................................... 172
Figure 6.15: Change in inclination angle () 127mm above mast due to presence of
mainsail with genoa .......................................................................................... 172
Figure 6.16: Inclination angle () 127mm above mast without sails and for mainsail
with jib or spinnaker ......................................................................................... 173
Figure 6.17: V/V

at different positions in z-plane 127mm above mast for mainsail


with genoa at
A
of 25 and 0 heel .................................................................. 173
Figure 6.18:
eff
at different positions in z-plane 127mm above mast for mainsail with
genoa at
A
of 25 and 0 heel .......................................................................... 174
Figure 6.19: at different positions in z-plane 127mm above mast for mainsail with
genoa at
A
of 25 and 0 heel .......................................................................... 174
Figure 6.20:
eff
, and V/V

vs. height above the mast with mainsail and genoa at

A
of 25 and 0 heel......................................................................................... 174
Figure 6.21:
eff
, and V/V

vs. height above the mast with mainsail and spinnaker


at
A
of 180 and 0 heel................................................................................... 174
xviii
YACHT
RESEARCH
UNIT
The
University
of Auckland

Figure 6.22:
eff
, and V/V

vs. heeling moment 127mm above mast for mainsail


with genoa at
eff
of 25 and 0 heel .................................................................175
Figure 6.23: Change in effective wind angle (
eff
) as modelled for the full-scale data
correction...........................................................................................................177
Figure 6.24: Wind speed factor (f
V
=V/V ) as modelled for the full-scale data correction ....177

Figure 6.25: Change in inclination angle () as modelled for the cup anemometer
calibration..........................................................................................................179
Cup anemometer speed (V Figure 6.26: ) vs.
Anemometer Vane
for full-scale mast twist
measurements used in analysis..........................................................................180
Figure 6.27: Mast twist (
Mast
) as a function of side force (F
Y
) measured on Dyna and
linear regression fit............................................................................................180
Figure 6.28: Heel angle () of Dyna with mainsail and jib plotted against effective wind
angle (
eff
)..........................................................................................................182
Figure 6.29: Inclination angle of free-stream flow (

) plotted against effective wind


angle (
eff
)..........................................................................................................182
V plotted against f Figure 6.30:
Anemometer
and showing calibration measurements and
fitted response surface.......................................................................................183
Figure 6.31: Different wind speed ratios plotted against inclination angle () .....................183
Figure 6.32: V plotted against f
Anemometer
showing calibration measurements and linear
and second order regression fit .........................................................................184
Figure 6.33: True wind speed (V
T
) during selected full-scale runs with mainsail and
spinnaker plotted against .............................................................................188
eff
Figure 6.34: True wind speed (V
T
) during full-scale runs with mainsail and jib plotted
against ..........................................................................................................188
eff
Figure 6.35: Heel angle () during full-scale runs with mainsail and jib plotted against
......................................................................................................................188
eff
Figure 6.36: Heel angle () during selected full-scale runs with mainsail and spinnaker
plotted against ..............................................................................................188
eff
Figure 6.37: Effective wind angle (
eff
) offset correction for full-scale measurements
with mainsail and jib .........................................................................................190
Figure 6.38: Effective wind angle (
eff
) offset correction for full-scale measurements
with mainsail and spinnaker..............................................................................191
Figure 6.39: Wind tunnel model of Dyna in downwind configuration with mainsail and
spinnaker and in upwind configuration with mainsail and jib ..........................193
Figure 6.40: Comparison of force and moment coefficients from selected full-scale runs
and wind tunnel measurements with mainsail and spinnaker plotted against
......................................................................................................................195
eff
Figure 6.41: Effective wind speed (V
eff
) during selected full-scale runs with mainsail
and spinnaker plotted against ......................................................................197
eff
Comparison of C Figure 6.42:
L
from selected full-scale and wind tunnel runs with
mainsail and spinnaker vs. ...........................................................................197
eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xix
List of Figures
Comparison of C Figure 6.43:
D
from selected full-scale and wind tunnel runs with
mainsail and spinnaker vs. .......................................................................... 197
eff
Comparison of C Figure 6.44:
Dp
from selected full-scale and wind tunnel runs with
mainsail and spinnaker vs. .......................................................................... 197
eff
Figure 6.45: Comparison of C /C
L D
from selected full-scale and wind tunnel runs with
mainsail and spinnaker vs. .......................................................................... 198
eff
Figure 6.46: IMS C and C
L Dp
for individual sails based on cloth sail area plotted against
..................................................................................................................... 198
eff
Figure 6.47: Comparison of z
CoE
from selected full-scale and wind tunnel runs with
mainsail and spinnaker vs. .......................................................................... 199
eff
Figure 6.48: Comparison of z
CoE
from selected full-scale and wind tunnel runs with
mainsail and spinnaker vs. ............................................................................ 199
Figure 6.49: Comparison of x
CoE
from selected full-scale and wind tunnel runs with
mainsail and spinnaker vs. .......................................................................... 199
eff
Figure 6.50: Comparison of x
CoE
from selected full-scale and wind tunnel runs with
mainsail and spinnaker vs. ............................................................................ 199
Figure 6.51: Comparison of force and moment coefficients from selected full-scale runs
with V
T
<4m/s and wind tunnel measurements with mainsail and jib plotted
against ......................................................................................................... 201
eff
Figure 6.52: Comparison of C from full-scale runs with V
L T
<4m/s and wind tunnel runs
with mainsail and jib vs. ............................................................................. 202
eff
Figure 6.53: Comparison of C
D
from full-scale runs with V
T
<4m/s and wind tunnel runs
with mainsail and jib vs. ............................................................................. 203
eff
Figure 6.54: Comparison of C from full-scale runs with V
Dp T
<4m/s and wind tunnel
runs with mainsail and jib vs. ..................................................................... 203
eff
Figure 6.55: Comparison of force and moment coefficients from selected full-scale runs
and wind tunnel measurements with mainsail and jib plotted against ........ 204
eff
Figure 6.56: Comparison of C
L
from full-scale and wind tunnel runs with mainsail and
jib vs. ........................................................................................................... 205
eff
Figure 6.57: Comparison of C
D
from full-scale and wind tunnel runs with mainsail and
jib vs. ........................................................................................................... 206
eff
Figure 6.58: Comparison of C
Dp
from full-scale and wind tunnel runs with mainsail and
jib vs. ........................................................................................................... 206
eff
Figure 6.59: Comparison of z
CoE
from full-scale and wind tunnel runs with mainsail and
jib vs. ........................................................................................................... 207
eff
Figure 6.60: Comparison of x
CoE
from full-scale and wind tunnel runs with mainsail and
jib vs. ........................................................................................................... 207
eff
Figure 6.61: Comparison of C /C
L D
from full-scale and wind tunnel runs with mainsail
and jib vs. .................................................................................................... 208
eff
Figure 6.62: Boat speed (V
s
) of Dyna with mainsail and jib and from VPP using wind
tunnel data and power model vs. .............................................................. 208
eff
xx
YACHT
RESEARCH
UNIT
The
University
of Auckland

Figure 6.63: Screen shot of sail shape analysis program SailTool showing the
parameters sail twist, boom angle and sail camber obtained from wind
tunnel picture.....................................................................................................209
Figure 6.64: Range of
eff
vs. V for analysed full-scale sail shape pictures .........................210
T
Mainsail twist vs. V Figure 6.65:
T
on Dyna and in wind tunnel with Real-Time VPP for
close hauled .................................................................................................210
eff
Boom angle vs. V Figure 6.66:
T
on Dyna and in wind tunnel with Real-Time VPP for
close hauled .................................................................................................211
eff
Mainsail camber vs. V Figure 6.67:
T
on Dyna and in wind tunnel with Real-Time VPP
for close hauled ...........................................................................................211
eff
Figure A.1: Three main components of the Dyna wind tunnel model; cradle, internal
load carrying frame and hull/deck shell ............................................................224
Figure A.2: Sailpan of Dyna.................................................................................................226
Measurement errors in vertical force (F Figure B.1:
Z
) applied at different x-positions
along centreline of model ..................................................................................231
Measurement errors in pitching moment (M Figure B.2:
Y
) applied at different x-
positions along centreline of model ..................................................................231
Corrected vertical force (F Figure B.3:
Zcorrected
) applied at different x-positions along
centreline of model............................................................................................232
Measurement errors in side force (F Figure B.4:
Y
) to starboard applied at different x-
positions along the model..................................................................................232
Figure B.5: Coarsest and finest mesh used in sensitivity analysis .......................................233
Figure B.6: Interpolated pressure for coarsest and finest mesh at
A
of 30 and 0 heel.....233
Figure B.7: Hull/deck surface area for different grid densities............................................234
C , C Figure B.8:
L D
and C of hull/deck at 0 heel for
Fz A
of 30 and 90 obtained with
different grid densities.......................................................................................234
Figure B.9: Model with cradle sitting in trough of turntable so that waterline coincides
with wind tunnel floor .......................................................................................235
C Figure B.10:
L
with constant mainsail and jib trim at 0 heel for trough open, covered
with cardboard and filled with water ................................................................235
C Figure B.11:
D
with constant mainsail and jib trim at 0 heel for trough open, covered
with cardboard and filled with water ................................................................236
C Figure B.12:
Fz
with constant mainsail and jib trim at 0 heel for trough open, covered
with cardboard and filled with water ................................................................236
C , C Figure B.13:
L D
and C on hull/deck from pressures above waterline...........................237
Fz
Figure C.1: Coordinate systems used in FRIENDSHIP-Equilibrium and the wind
tunnel .................................................................................................................239
Figure C.2: Geometry of hull, keel and rudder modelled from offset tables by
FRIENDSHIP-Equilibrium...............................................................................249
Figure C.3: Residuary resistance coefficient curve vs. Froude number used by
FRIENDSHIP-Equilibrium...............................................................................249
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xxi
List of Figures
Figure D.1: General force and moment system.................................................................... 253
Schematic of lifting line with element of planar trailing vortex sheet (y) of
width ds
Figure D.2:
............................................................................................................. 256
Fourier series terms of n=1 to n=3 in span loading distribution and
downwash
Figure D.3:
......................................................................................................... 257
xxii
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
List of Tables
Page
Table 2.1: Alphabetic listing of sailing specific terms.........................................................16
Table 3.1: Hull and rig windage coefficients for equations (3.11) and (3.12) obtained
from wind tunnel tests .........................................................................................57
Table 4.1: Inputs required in the Real-Time VPP LabVIEW Application...........................83
Regression parameters a to a to model the depowering of C , z Table 5.1:
L CoE 1 6
and
C
D
as a function of p

obtained for between 25 and 90 .............................127
eff
Table 5.2: Set-up summary of the four sets of performance predictions carried out to
investigate the standard trim parameter model .................................................138
Regression parameters a to a obtained from wind tunnel tests to account
for changes in the flow field at the position of the cup anemometer of
Dyna
Table 6.1:
1 5
..................................................................................................................178
Regression parameters a , a and b obtained from cup anemometer
calibration tests on Dyna
Table 6.2:
1 2 1
...................................................................................183
Table 6.3: Summary of the test series used in the analysis of the full-scale data ..............187
Table 6.4: Wind tunnel test matrix.....................................................................................193
Table A.1: Principal dimensions of wind tunnel models in TFWT ....................................223 T
Table C.1: Definition for coordinate systems used in FRIENDSHIP-Equilibrium and
the wind tunnel ..................................................................................................240
Table C.2: List of parameters contained in the output file of the Real-Time VPP
LabVIEW application .......................................................................................242
Table C.3: Input parameters of the WindTunnel force module of FRIENDSHIP-
Equilibrium........................................................................................................243
Table C.4: Input parameters of the HansenRig force module of FRIENDSHIP-
Equilibrium........................................................................................................245
Table C.5: Input parameters of the HansenRigDirect force module of
FRIENDSHIP-Equilibrium...............................................................................248
Table C.6: Hydrodynamic Force modules used for modelling Dyna in FRIENDSHIP-
Equilibrium........................................................................................................250
Table C.7: Input data for Mass force module of Dyna.....................................................250
Table C.8: Additional input data for Displacement, GenericHull and Propeller
force modules of Dyna ......................................................................................251
Table C.9: Input data for Keel and Rudder force module of Dyna ...............................251

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xxiii

The
University
YACHT
RESEARCH
of Auckland UNIT
Nomenclature

Symbols
Page
T
a Point on central axis that is closest to origin, (a
x
, a
y
, a ) [m] ........................ 29, 255
z
A, B, C Position vectors of nodes of triangular mesh element [m]...................................... 50
a
0,
, a Free parameters obtained from regression fit [-] .................................... 55, 177, 231
n
A Area of mesh element i [m
2
] ................................................................................... 51
i
A Area of hull above DWL projected onto x-plane [m
2
]............................................ 54
PxHull
A Area of rig projected onto x-plane [m
2
] .................................................................. 54
PxRig
A Area of hull above DWL projected onto y-plane [m
2
]............................................ 54
PyHull
A Area of rig projected onto y-plane [m
2
] .................................................................. 54
PyRig
AR Geometric aspect ratio [-] ....................................................................................... 20
A
ref
Reference area [m
2
]................................................................................................. 18
A
S
Reference sail area [m
2
] .......................................................................................... 18
A Reference sail area of model [m
2
]........................................................................... 35
Sm
A
Smain
Cloth area of mainsail [m
2
] ..................................................................................... 18
B Position of coordinate system B origin in coordinate system T [m]....................... 34 B

B Position of coordinate system B origin at B
0 m
=0 in coordinate system T [m]........ 34
C Calibration matrix [N/V, Nm/V]............................................................................. 84
c Point on central axis [m] ......................................................................................... 30

C
D
Drag coefficient in coordinate system B [-]............................................................ 19
C
D0
Drag coefficient at zero lift [-] ................................................................................ 19
C
DA
Horizontal drag coefficient in coordinate system A [-] .......................................... 36
C Skin friction drag coefficient [-] ........................................................................... 207
Df
C Drag coefficient of hull [-] ...................................................................................... 54
Dhull
C
Di
Induced drag coefficient [-] .................................................................................... 20
C
Dmax
Maximum drag coefficient of hull or rig [-] ........................................................... 53
C Minimum drag coefficient of hull or rig [-] ............................................................ 53
Dmin
C Parasitic drag coefficient in coordinate system B [-]............................................ 119
Dp
C
DpOpt
Optimum parasitic drag coefficient in coordinate system B [-]............................ 119
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xxv
Nomenclature
C Drag coefficient of rig [-] ........................................................................................ 54
Drig
C
Ds
Separation drag coefficient [-]................................................................................. 19
C
Dvis
Viscous drag coefficient [-] ..................................................................................... 20
C
Dwindage
Windage drag coefficient in coordinate system B [-]............................................ 136
C Drag coefficient of rig in x-direction [-] ................................................................. 54
DxRig
C Drag coefficient of rig in y-direction [-] ................................................................. 54
DyRig
C
F
Force coefficient in coordinate system B [-] ........................................................... 18
C Force coefficient vector is coordinate system B [-] ................................................ 35
F
C Force coefficient vector in coordinate system T [-] ................................................ 84
FT
C Force coefficient in x-direction (drive force) in coordinate system B [-] ......... 62, 78
Fx
C Force coefficient in y-direction (side force) in coordinate system B [-] ........... 62, 78
Fy
C Force coefficient in z-direction in coordinate system B [-]............................... 63, 79
Fz
C Optimum force coefficient in z-direction in coordinate system B [-] ........... 121, 133
FzOpt
C Lift coefficient in coordinate system B [-] .............................................................. 19
L
C Sectional lift coefficient [-] ................................................................................... 153
l
C Horizontal lift coefficient in coordinate system A [-] ............................................. 36
LA
C
Lmax
Maximum lift coefficient of hull or rig [-] .............................................................. 56
C Sectional lift coefficient for loading that has minimum induced drag [-] ............. 153
lminDi
C Optimum lift coefficient in coordinate system B [-] ............................................... 76
Lopt
C
lopt
Optimum sectional lift coefficient [-].................................................................... 153
C Moment coefficient in coordinate system B [-]....................................................... 18
M
C Moment coefficient vector is coordinate system B [-] ............................................ 35
M
C
Mf
Coefficient of moment in line with force in coordinate system B [-] ................... 112
C
Mf
Coefficient vector of moment in line with force vector (F) in coordinate
system B [-] ........................................................................................................... 107
C
MfOpt
Coefficient of optimum moment in line with force in coordinate system B [-] .... 121
C
MT
Moment coefficient vector in coordinate system T [-]............................................ 84
C
Mv
Coefficient vector of moment perpendicular to force vector (F) in coordinate
system B [-] ........................................................................................................... 107
C
Mx
Moment coefficient about x-axis in coordinate system B [-] ................................ 116
C
MxOpt
Optimum moment coefficient about x-axis in coordinate system B [-] ................ 116
CoA Centre of area position vector [m]........................................................................... 50
T
CoE Centre of effort position vector in coordinate system B, (x
CoE
, y
CoE
, z
CoE
) [m] ..... 29
C
P
Pressure coefficient [-] ...................................................................................... 17, 50
c Dynamic pressure factor [-]..................................................................................... 35
q
c
s
Separation drag constant [-] .................................................................................... 20
c Twist weight constant [-]................................................................................. 78, 137
t
xxvi
YACHT
RESEARCH
UNIT
The
University
of Auckland
Symbols
c
tL
Twist weight constant for lift [-]........................................................................... 153
c

Heel constant [-]...................................................................................................... 84


E Distance between mast and clew of mainsail [m]................................................... 15
e Oswald efficiency factor [-] .................................................................................... 20
E Youngs modulus of elasticity [Pa]....................................................................... 227
F Force in coordinate system B [N] ........................................................................... 18
T
F Force vector in coordinate system B, (F
X
, F , F ) [N] .................................... 29, 34
Y Z
f Trim parameter flat [-] ...................................................................................... 76, 77
F

Unit vector of F [-]................................................................................................ 254


F
A
Force vector in coordinate system A [N] ................................................................ 80
f Rotational frequency of cup anemometer [Hz]..................................................... 183
Anemometer
f 'Flattening' of sail due to twist [-].......................................................................... 153
t
T
F Force vector in coordinate system T, (F , F
T XT YT
, F ) [N]................................ 34, 85
ZT
F Force and moment vector in coordinate system T, (F , F
T(6) XT YT
, F , M , M
ZT XT YT
,
M
T
) [N, Nm]......................................................................................................... 84
ZT
f
V
Flow speed factor [-] ............................................................................................. 177
f Apparent wind speed factor [-] ............................................................................... 55
VA
f Factor between reference flow speed with and without model in tunnel [-] ........ 168
Vref
F
X
Force in x-direction (drive force) in coordinate system B [N] ............................... 72
F Force in y-direction (side force) in coordinate system B [N] ............................... 180
Y
F Force in z-direction in coordinate system B [N]..................................................... 44
Z
2
g Acceleration due to gravity [m/s ] .......................................................................... 25
H Position of heel axis of model in coordinate system T [m] .................................... 34

hS Significant wave height [m].................................................................................. 187

I Height of fore triangle [m] ................................................................................ 15, 18
I Second moment of area [m
4
]................................................................................. 227
J Distance between forestay/deck attachment point and mast [m]...................... 15, 18
k Number of samples the pressure reading is delayed [-] .......................................... 85
L Length of beam in bending moment equation [m]................................................ 227
l Reference length [m]....................................................................................... 36, 254
M Moment in coordinate system B [Nm].................................................................... 18
T
M Moment vector in coordinate system B, (M
X
, M , M ) [Nm].......................... 29, 34
Y Z
T
M
0
Moment vector of pure couples in coordinate system B, (M
0X
, M , M )
0Y 0Z

[Nm] ........................................................................................................................ 29
M
A
Moment vector in coordinate system A [Nm] ........................................................ 80
M
F
Moment vector in line with force vector (F) in coordinate system B [Nm] . 106, 254
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xxvii
Nomenclature
T
M Moment vector in coordinate system T, (M , M
T XT YT
, M ) [Nm] ...................... 34, 85
ZT
M
V
Moment vector perpendicular to force vector (F) in coordinate system B
[Nm] .............................................................................................................. 106, 254
M
X
Moment about x-axis (heeling moment) in coordinate system B [Nm]............ 10, 71
M
Z
Moment about z-axis (yaw moment) in coordinate system B [Nm] ................. 10, 71
n Number of samples [-]............................................................................................. 85
N Vector normal to triangular mesh element [m
2
] ...................................................... 51
P Distance between boom and top of mast [m] .......................................................... 15
p Trim parameter power [-] ...................................................................................... 116
q Dynamic pressure [Pa] ............................................................................................ 18
q Scalar factor in central axis equation [m/N]............................................ 30, 106, 254

q
A
Apparent dynamic pressure [Pa] ............................................................................. 19
q Apparent dynamic pressure in wind tunnel [Pa] ..................................................... 35
Am
q
eff
Effective dynamic pressure [Pa].............................................................................. 24
q Effective dynamic pressure in wind tunnel [Pa]...................................................... 35
effm
q
ref
Reference dynamic pressure [Pa] ............................................................................ 35
r Trim parameter reef [-]...................................................................................... 76, 77
R Power ratio of parasitic drag [-] ............................................................................ 119
Dp
Re Reynolds Number [-]............................................................................................... 36
R Power ratio of force in x-direction [-] ................................................................... 132
Fx
R Power ratio of force in y-direction [-] ................................................................... 132
Fy
R Power ratio of force in z-direction [-] ................................................................... 130
Fz
R Power ratio of lift [-].............................................................................................. 118
L
R
Mf
Power ratio of moment in line with force [-]......................................................... 130
R
My
Power ratio of moment about y-axis [-] ................................................................ 133
R Power ratio of moment about z-axis [-]................................................................. 133
Mz
R
xCoE
Power ratio of longitudinal centre of effort position [-]........................................ 129
R
zCoE
Power ratio of centre of effort height [-] ............................................................... 121
T
S Signal vector, (S , S , S , S , S , S ) [V]........................................................ 83
X1 Y1 Y2 Z1 Z2 Z3
S Zero signal vector [V] ............................................................................................. 84
0
T Position of coordinate system T origin in coordinate system T [m] ....................... 34

t Trim parameter twist [-]........................................................................................... 78
t Base twist when sails are fully powered-up [-] ..................................................... 145
0
T Rotational transformation matrix from the absolute coordinate system
(coordinate system A) to the body fixed coordinate system (coordinate system
B) [-] ........................................................................................................................ 23
AB
xxviii
YACHT
RESEARCH
UNIT
The
University
of Auckland
Symbols
T
BA
Rotational transformation matrix from the body fixed coordinate system
(coordinate system B) absolute coordinate system (coordinate system A) [-] ...... 34
T

Transformation matrix to rotate about z-axis by angle [-]................................. 261


V Flow speed [m/s]....................................................................................... 18, 36, 167
V Free-stream flow speed [m/s]................................................................................ 167

Unit vector of flow velocity [-] ............................................................................. 261


V
A
Apparent wind speed [m/s] ..................................................................................... 19
T
V
A
Apparent wind velocity vector in coordinate system A, (V , V , V ) [m/s] ...... 23
Ax Ay Az
V Reference apparent wind speed [m/s] ..................................................................... 26
Aref
V
eff
Effective wind speed [m/s] ..................................................................................... 21
T
V
eff
Effective wind velocity vector in coordinate system B, (V
effx
, V
effy
, V ) [m/s].... 23
effz
V Boat speed [m/s]...................................................................................................... 20
S
V True wind speed [m/s]............................................................................................. 20
T
x
CoE
Longitudinal position of centre of effort in coordinate system B [m] .................... 29
x
CoEopt
Optimum longitudinal centre of effort position in coordinate system B [m].. 79, 121
y Deflection of beam in bending moment equation [m] .......................................... 227
z Height above the water in coordinate system A [m]......................................... 25, 28

z Terrain roughness length [m].................................................................... 25, 39, 192
0
z Height of the anemometer above DWL in coordinate system B [m].................... 185
Anemometer
z
boom
Boom height above DWL in coordinate system B [m]........................................... 79
z Height of the geometrical centre of sail area above DWL in coordinate system
B [m] ....................................................................................................................... 28
CoA
z
CoE
Centre of effort height in coordinate system B [m] ................................................ 29
z Centre of effort height for minimum induced drag in coordinate system B
[m]................................................................................................................... 78, 144
CoEminDi
z
CoEopt
Optimum centre of effort height in coordinate system B [m]......................... 78, 121
z Height of the sail foot above DWL in coordinate system B [m] ............................ 27
foot
z Height above wind tunnel floor in coordinate system A [m].................................. 35
m
z
mast
Height of mast above DWL in coordinate system B [m]........................................ 20
z Reference height above the water (usually taken as 10m) in coordinate system
A [m] ....................................................................................................................... 24
ref




Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xxix
Nomenclature

Greek Symbols
Page
S Change in signal vector [V]..................................................................................... 83

V Change in flow speed [m/s]................................................................................... 167

A
Change in apparent wind angle [] .............................................................. 26, 27, 55

eff
Change in effective wind angle [].................................................................. 98, 167
Change in inclination angle [].............................................................................. 167
Angle of attack [] ................................................................................................... 17
Apparent wind angle [] .......................................................................................... 19
A
Apparent wind angle of model in wind tunnel []................................................... 35
Am
Effective wind angle [] .......................................................................................... 21
eff

effm
Effective wind angle of model in wind tunnel [] ................................................... 98
Offset in measured wind direction [-] ................................................................... 190
Offset
True wind angle [].................................................................................................. 20
T

Vane
Flow direction measured by direction vane on top of mast [] ............................. 179
Inclination angle of flow from z-plane in coordinate system B []....................... 167

Rudder angle [] ................................................................................................ 10, 76
Inclination angle of free-stream flow from z-plane in coordinate system B
[]................................................................................................................... 169, 182


Angle between force vector (F) and moment vector (M) [] ................................ 254
Heel angle []........................................................................................................... 21

m
Heel angle of model in wind tunnel [] ................................................................... 34
Sheeting angle [] .................................................................................................... 19
Resultant force direction in z-plane of coordinate system B [] ............................. 30
Leeway angle [] ......................................................................................... 10, 20, 76
Viscosity [Pas]........................................................................................................ 36
Kinematic viscosity [m
2
/s] ...................................................................................... 36
Pitch (trim) angle []................................................................................................ 22

m
Pitch (trim) angle of model in wind tunnel [] ........................................................ 34
Density [kg/m
3
] ................................................................................................. 18, 36

air
Density of air [kg/m
3
].............................................................................................. 79

Mast
Mast twist [] ......................................................................................................... 180
xxx
YACHT
RESEARCH
UNIT
The
University
of Auckland
Subscripts and other Notations

Subscripts and other Notations

X Scalar X
|X| Absolute value of scalar X
X Vector X or matrix X
|X| Magnitude (length) of vector X
X

Unit vector of vector X


X
T
Transpose of vector X or matrix X
X X in coordinate system A
A
X X in coordinate system B
X X in coordinate system T
T
X
X
, X X in direction of x-axis
x
X , X X in direction of y-axis
Y y
X , X X in direction of z-axis
Z z
X
Anemometer
X measured at anemometer position above mast
X X measured by Cobra Probe
Cobra
X Effective X in z-plane
eff
X in the wind tunnel (in relation to model) X
m
X
minDi
X at loading distribution that results in minimum induced drag
X X at optimum fully powered-up sail trim
opt
X Reference X
ref
X X of free-stream flow








Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails xxxi
Nomenclature

Abbreviations and Names
Page
A/D Analogue to Digital converter ................................................................................. 81
BLWT Boundary Layer Wind Tunnel................................................................................... 6
CFD Computational Fluid Dynamics................................................................................. 1
DLL Dynamic Link Library............................................................................................. 89
DWL Design Waterline, flotation line of upright yacht as defined in design
specification................................................................................................. 16, 20, 49

Dyna Yacht name of the Berlin Sail-Force-Dynamometer................................................. 4

FOG Fibre-Optic-Gyro........................................................................................... 163, 239

F Forward Perpendicular, most forward point where hull intersects DWL ..... 162, 240
PP
GRP Glass-fibre Reinforced Plastic............................................................................... 224
GUI Graphical User Interface ......................................................................................... 82
IACC International Americas Cup Class................................................................ 2, 3, 223
IMS International Measurement System....................................................................... 2, 5
LDV Laser Doppler Velocimetry................................................................................... 162

LP Low Pass.................................................................................................................. 81
LVDT Linear Voltage Displacement Transducer ............................................................... 32

MIT Massachusetts Institute of Technology ..................................................................... 3
PBL Planetary Boundary Layer....................................................................................... 17
PPP Performance Prediction Program........................................................................ 3, 88
RANSE Reynolds Avaraged Navier-Stokes Equations ...................................................... 5, 9
TFWT Twisted Flow Wind Tunnel....................................................................................... 6
VMG Velocity Made Good to windward or leeward .................................................. 16, 73
VO60 Volvo Ocean 60......................................................................................... 26, 53, 223
VO70 Volvo Open 70 ................................................................................................ 27, 223
VPP Velocity Prediction Program........................................................................... 1, 2, 71

xxxii
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
1 Introduction
1.1 Preface
Accurate prediction of performance is an important aspect of modern yacht design. In the
quest to gain a competitive advantage a large number of different parameters need to be
evaluated to optimise the design by taking many factors, such as the rating systems or class
rules, weather conditions and tactics, into consideration. This is a complex task and requires a
number of analysis techniques to be used.
One commonly used tool in the design process of sailing yachts is a Velocity Prediction
Program (VPP). It brings the results from the analysis of different components of a yacht
together to predict the resulting performance of a sailing yacht. It balances the aerodynamic
and hydrodynamic forces and moments created by the different yacht components so that the
vessel sails in equilibrium. An overview of related research in given in section 1.2. The
accuracy of the prediction depends mainly on how well the individual forces are modelled
and on how accurate the input parameters for these models are. The input parameters required
by VPPs can be obtained in a number of ways; by theoretical calculations, numerical
simulations, model scale testing or full-scale measurements.
The accuracy of a VPP can be assessed by comparing the predictions to the performance
achieved by the yacht once it has been built and is sailing. However, only the resulting
performance of the full-scale yacht is known and not the individual force components which
lead to it. The information obtained from full-scale yachts is hence of limited value for
improving the accuracy of VPPs. Similarly it is difficult to extract information that is useful
for future VPP analyses and design studies.
It is possible to measure the individual force components in full-scale, but this requires a
purpose-built sail force dynamometer, which represents an expensive and complex task and
has only been attempted three times so far as will be shown in section 1.3. The most
comprehensive recent example is the Berlin Sail-Force-Dynamometer.
Apart from the measurements obtained with these sail force dynamometers, the aerodynamic
input parameters for VPPs are determined mainly from theoretical methods, computational
fluid dynamics (CFD) simulations and wind tunnel testing. The preferred method for
determining the aerodynamic parameters depends on the type of investigation and the
resources available and in many cases ideally more than one method is employed. The
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 1
Chapter 1 Introduction
reasons why wind tunnel testing is a commonly used and effective tool in the design process,
particularly for downwind sails, will be discussed in section 1.4.1. For many of the same
reasons wind tunnel testing is also a very applicable research tool. The Twisted Flow Wind
Tunnel of The University of Auckland, which is one of the major facilities for testing yacht
sails since it has been developed specifically for this purpose, will be described in section 1.4.
One of the fundamental differences between yacht sails and most other lifting bodies is their
lack of rigidity so that the design shape and the flying shape are different. In addition the
flying shape is deliberately altered to best suit each sailing condition which makes the
analysis and the modelling more complex than for other lifting bodies, as will be discussed in
section 1.4.2.
1.2 Velocity Prediction Programs
The first velocity prediction for a sailing yacht was conducted by Davidson (1936). The
foundation of modern velocity prediction was however presented by Kerwin (1978). Together
with the systematic hull form series tested by Gerritsma et al. (1981) and the aerodynamic
coefficients from Hazen (1980) the speed of the yacht can be calculated for a range of sailing
conditions. The aerodynamic and hydrodynamic forces and moments acting on a yacht are
balanced by solving the non-linear system of equations for the steady state equilibrium
condition. This first Velocity Prediction Program (VPP) formed the basis of a new rating
system for sailing yachts; the International Measurement System (IMS).
A number of improvements to the force modelling in relation to the design of International
Americas Cup Class (IACC) yachts were presented by Schlageter and Teeters (1993). Van
Oossanen (1993) described a different method of solving for force and moment equilibrium
which places the hydrodynamic forces into a more central role compared to other VPPs. This
method is however less often adhered to. Van Oossanen (1995) also introduced the balance of
the yaw moment in the VPP formulation. Further improvements to the hydrodynamic
modelling have also been introduced by, for example, Fargeas and Kouyoumdjian (1997).
Recent developments in the IMS VPP formulations are described by Claughton (1999), but it
is also pointed out that the aerodynamic part of the force model has not fundamentally
changed during the past 20 years. Jackson (2001) introduced a third trim parameter twist in
addition to flat and reef to model the effect of sail twist in the VPP formulation. A
computational viscous flow investigation into trim parameters and depowering has recently
been conducted and the results are presented by Krebber (2005) and Krebber and Hochkirch
(2006). The modelling and depowering of sail forces in VPPs will be discussed in more detail
in section 4.2.1 and chapter 5.
2
YACHT
RESEARCH
UNIT
The
University
of Auckland
1.3 Large-Scale Testing
Recent developments towards dynamic time-domain based performance prediction programs
(PPP) have been presented by Day et al. (2002) and Richardt et al. (2005) where not only the
steady state equilibrium is obtained but accelerations are also considered. To date the
differences to VPPs are mainly in the solution process and many hydrodynamic and certainly
the aerodynamic models still use a quasi-static description. A different approach of realising a
VPP based on purely computational techniques has also been introduced by Roux et al.
(2002). Currently an important feature of VPPs is however, their ability to combine
information from different research methods and a purely computational VPP seems very
restrictive.
Many yacht designers have in-house VPPs of varying degrees of complexity. Martin and
Beck (2001) presented PCSAIL, a simplified VPP developed in Excel
1
at the University of
Michigan that can be used for teaching and preliminary design purposes. The commercially
most widely used VPP is WinDesign
2
. The Technical University Berlin also developed a
VPP for research purposes, which has been advanced by FRIENDSHIP SYSTEMS into the
modular workbench FRIENDSHIP-Equilibrium
3
for analysing stationary and non-stationary
modes of motion (Richardt et al., 2005).
1.3 Large-Scale Testing
The first attempt to measure rig forces was made by Davidson (1936) with the yacht
GIMCRACK and resulted in the GIMCRACK-coefficients. The methods used in this early
work were very basic. The apparent wind angle for example was only estimated. Since then a
number of people has investigated the large-scale performance of yachts and deduced the
aerodynamic characteristics but did not measure the aerodynamic forces directly.
In order to measure the aerodynamic and hydrodynamic forces and moments separately, a
specially constructed sail force dynamometer is required due to the large pretension in the rig
and the comparatively small aerodynamic forces. This concept was first implemented as the
Massachusetts Institute of Technology (MIT) sailing dynamometer, a 42% model of an
International Americas Cup Class (IACC) yacht. The design, implementation and data
handling is described by Herman (1988). Klein (1990) and Peters (1992) obtained full-scale
aerodynamic force coefficients and compare them to invicid simulations. Their work
highlighted many of the challenges associated with large-scale testing and showed the
limitations of invicid methods. Nevertheless a great deal was learnt from the full-scale
measurements and the concept of a sail force dynamometer was proved to be feasible. The

1
Software developed by Microsoft Corporation, One Microsoft Way, Redmond, WA 98052-6399, USA
2
Software developed by Yacht Research International and the Wolfson Unit MTIA, University of Southampton,
Southampton SO17 1BJ, UK
3
Software developed by FRIENDSHIP SYSTEMS GmbH, Benzstrasse 2, D-14482 Potsdam, Germany
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 3
Chapter 1 Introduction
MIT sailing dynamometer project is also summarised by Milgram (1993) and Milgram et al.
(1993). The large-scale data was not compared with results from wind tunnel tests.
A second sail force dynamometer of similar size was introduced by Masuyama and Fukasawa
(1997). Fundamentally the same concept is employed as the MIT sailing dynamometer. The
work also focused on obtaining aerodynamic force coefficients of an IACC and IMS rig and
comparing them with invicid calculations using the vortex lattice method. Although the
calculated results deviated from the measured data similar trends were observed. The trends
were also similar when changing the mainsail draft and they concluded that the vortex lattice
method is sufficient for modelling trim changes (Masuyama and Fukasawa, 1997). Other trim
changes that may result in partial flow separation were however not investigated. Again no
wind tunnel tests were conducted to provide a comparison with the large-scale data.
The third and most recent project is the Berlin Sail-Force-Dynamometer named Dyna, shown
in Figure 1.1. It was initiated at the Technical University Berlin in Germany in 1996 to
investigate the relationships between model tests, numerical simulations and full-size data
(Brandt and Hochkirch, 1999). A 10-metre full-scale sail force dynamometer was designed
and built to directly measure the aerodynamic and hydrodynamic forces and moments. The

4

Figure 1.1: The Berlin Sail-Force-Dynamometer Dyna
4
Picture courtesy of the Technical University Berlin
4
YACHT
RESEARCH
UNIT
The
University
of Auckland
1.3 Large-Scale Testing
hull and rig are based on a Dehler 33, a modern 33ft International Measurement System
(IMS) cruiser/racer designed by judel/vrolijk & co., to provide suitable reference results for
contemporary yacht design issues.
In order for force measurements to be useful as input for VPPs the individual force
components acting on the yacht need to be obtained. On Dyna this is achieved with a number
of multi-component force balances to measure the force components acting on the keel,
rudder and rig. The rig force balance is based on the same concept used on the two previous
sailing dynamometers. It comprises an internal rigid aluminium framework structure in the
hull to which all rigging components are connected through openings in the deck. The
framework structure is connected to the hull shell via six force transducers to form the six-
component force balance. In addition the forces in the individual rig components are
measured by load cells at the attachment points to the framework structure and connecting
points between components. On Dyna a six-component force balance is also used to connect
the keel to the hull and a five-component force balance is integrated in the rudderstock to
measure the forces on the rudder.
The focus of the initial four-year project was to design, develop and build the sail-force-
dynamometer and to investigate hydrodynamic aspects of sailing yacht performance
prediction as described by Brandt and Hochkirch (2000) and Hochkirch (2000). The full-size
measurements were compared to the results from CFD simulations, model scale tests in the
towing tank and cavitation tunnel tests of different keel designs. The aerodynamic data was
only used to obtain the lift and drag coefficients of the rig, which are required to perform
VPP calculations. It was not compared to model tests or numerical solutions as part of the
initial project and will be analysed in detail in chapter 6.
The initial project was followed by a second project in 2001 to investigate design principles
of modern shallow draft keels with low aspect ratio winglets (Clauss and Hochkirch, 2003).
Different winglet positions on a shallow draft keel were investigated by conducting full-scale
measurements on Dyna, CFD simulations and model tests in the towing tank and cavitation
tunnel. A third project, started in 2002, developed a system to capture the sail shape on Dyna
and used the obtained flying shapes to conduct viscous CFD simulations employing a
Reynolds Averaged Navier-Stokes Equations (RANSE) solver as outlined by Clauss et al.
(2005). The main objective was the sail shape analysis and only a few simultaneous sail force
measurements were conducted.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 5
Chapter 1 Introduction
1.4 Twisted Flow Wind Tunnel
Wind tunnel testing of yacht sails is used by researchers, yacht designers and sail designers.
There are four wind tunnel facilities in the world that conduct most of the work related to
sails: the Wolfson Unit at the University of Southampton in England, the Glenn L Martin
Wind Tunnel in the USA, the Politecnico di Milan in Italy and the Twisted Flow Wind
Tunnel (TFWT) at The University of Auckland. The TFWT is unique in that is has been
developed specifically for the testing of yacht sails. It was the first wind tunnel that can
simulate the change in wind direction with height experienced by sails. More recently other
wind tunnels have incorporated the flow twisting concept, including the Politecnico di Milan
(Fossati et al., 2006) and the University Applied Science Kiel in Germany (Mller and Graf,
2005).
The change in apparent wind direction with height is called wind twist. The apparent wind is
the combination of the yacht speed and the planetary boundary layer of the true wind and
hence changes with height as discussed in section 2.3.2. The implications this has on wind
tunnel testing of yacht sails were first discussed by Flay and Jackson (1992). Generally it can
be said that the effect of twist in the apparent wind increases with increasing true wind angle,
i.e. the deeper a yacht sails the more twist is experienced by the sails. Secondly the amount of
twist also depends on the performance potential of the yacht, i.e. the faster the yacht sails
relative to the true wind speed the more twist is experienced by the sails. Wind tunnel testing
is mainly carried out for high-performance yachts, where extensive research and development
is conducted to gain a competitive advantage. Of particular interest in the wind tunnel is the
design of reaching and downwind sails, where numerical methods such as CFD are still not
sufficiently advanced to predict the large regions of separated flow. Hence, for many
situations which require wind tunnel testing of sails, twist is likely to be of interest. This led
to the development of the TFWT by the Yacht Research Unit at The University of Auckland
in association with North Sails New Zealand and Team New Zealand for the Americas Cup
Challenge in 1995. After extensive research published by Flay et al. (1996) and Flay (1996) a
system of vertical vanes, positioned at the end of a boundary layer wind tunnel (BLWT) in
front of an open test section to twist the flow upstream of the test section, was chosen.
Figure 1.2 shows the original TFWT being used to test sails for Americas Cup winner Black
Magic NZL-32. Since 1995 the concept has remained the same but improvements are
constantly being made, and the wind tunnel is now owned by The University of Auckland and
is used for research and general consultancy work. In addition to being used by Team New
Zealand in the preparation for the defences of the Americas Cup in 2000 and 2003, the
TFWT is used by Emirates Team New Zealand and several other syndicates in preparation for
the Americas Cup 2007 and was used by most syndicates for the Volvo Ocean Race 2001-
6
YACHT
RESEARCH
UNIT
The
University
of Auckland
1.4 Twisted Flow Wind Tunnel

Figure 1.2: The original Twisted Flow Wind Tunnel (TFWT) for testing yacht sails at The University of
Auckland
2002 and 2005-2006, Open 60 projects, Maxi yacht designers and record breakers like Mari
Cha III and IV.
Research of particular relevance to this project has been conducted at the TFWT since it is
one of the few facilities in the world where the twisted flow onto yacht sails can be simulated.
Test procedures for high-performance yachts with improved speed and twist profiles at the
TFWT were developed by University of Aberdeen student Benzie (2001) and Royal Institute
of Technology student Ekblom (2002) during internships at The University of Auckland.
Improvements to the TFWT force balance were made by Bonniot (2002) and a study on
blockage corrections was conducted by Kiffer (2003). Considerable work, summarised by
Richards et al. (2004), has been undertaken to study the flow in and around the TFWT and as
a result modifications have been made that considerably improved the flow quality in the test
section of the tunnel.
The TFWT was used by Hawkins (1998) to investigate the non-dimensional number effects
on downwind sail modelling, as will be discussed in section 2.4.2, and by White and Wilson
(2000) to study the effect of heel on upwind sails. A more in-depth study of the effect of heel
and twist on the aerodynamic forces has been carried out at the TFWT by Chalmers
University of Technology students Flodn and Johansson (2001) which led to the
investigation in chapter 3. Considerable work has also been conducted at the TFWT to
investigate the aerodynamic force components generated by the mainsail and spinnaker
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 7
Chapter 1 Introduction
operating together and by themselves. The key results of these investigations is summarised
by Richards et al. (2006).
Recent sail testing of interest, conducted at other wind tunnels, includes work at the
University of Southamptons low speed wind tunnel which evaluated the upwind
performance of a sloop rig (Campbell, 1997) and investigated the performance of offwind
sails (Campbell, 1998). At the Glenn L. Martin Wind Tunnel at the University of Maryland
the sail performance and blockage corrections were determined experimentally by
Ranzenbach and Mairs (1997) and offwind sails were tested by Ranzenbach and Mairs
(1999). Work on sail shape capturing during wind tunnel tests by the means of
photogrammetry and a coordinate measurement machine has been presented by Ranzenbach
and Kleene (2002). Similarly, Fossati et al. (2006) are developing a system for determining
the sail shape in the Politecnico di Milan wind tunnel based on infrared camera pictures.
1.4.1 Wind Tunnel Testing as a Sail Design Tool
When designing a sail, rig or indeed any part of a competitive sailing yacht it is of interest to
the designer to know how well each component will perform. Clearly the earlier in the design
process this can be established the better, since time and money can be saved. There are
principally three ways to determine the performance of a yacht component: by full-scale or
large-scale tests, by model scale tests or by CFD or theoretical calculations. Since it is not
crucial to determine the sail shape early in the design process, full-scale or large-scale testing
of different sail shapes on the water is a feasible and practiced option. Building a full-scale or
large-scale sail is however much more expensive than building a model sail for the wind
tunnel. In addition, testing a sail on the water is very time consuming as the natural test
environment is not controllable; many test runs are required to gain confidence in the results
and it is often difficult to pick up small differences in sail performance. Although the
performance of different sails can be compared on the basis of the achieved boat speed, is it
extremely difficult to separate the individual force components acting on a real yacht, i.e.
how much force was produced by the sails, how much side force by the keel and how much
resistance by the hull? The individual force components are however required by Velocity
Prediction Programs (VPPs) and a few purpose-built sailing dynamometers have been used to
achieve this. Apart from these few exceptions, testing of sails on the water may reveal the
fastest sail, but the information has only limited use for future VPP analysis.
The other alternative to wind tunnel testing is to conduct CFD or theoretical calculations.
Theoretical calculations are a good basic tool but due to the complex three-dimensional flow
structure around sails, their application to real problems of interest is very limited. In CFD
calculations, the governing equations of fluid flow are used and rather than solving them
8
YACHT
RESEARCH
UNIT
The
University
of Auckland
1.4 Twisted Flow Wind Tunnel
theoretically, they are solved numerically with the aid of computers. For upwind sails, where
the flow is largely attached to the sails, accurate solutions can be achieved by using potential
flow CFD simulations. However, for downwind sails the flow is separated over large parts of
the sails and since flow separation is a viscous effect, viscous flow equations, such as the
Reynolds Averaged Navier-Stokes Equations (RANSE), need to be solved. The lift and drag
forces on a sail depend strongly on the separation and reattachment points of the flow. As
discussed recently by Collie and Gerritsen (2006), predicting these points correctly is still a
complicated challenge in viscous flow CFD simulations. The simulations are hence still
computationally very expensive and require extensive validation work so that the use for
practical design applications is very resource intense.
In addition to being more cost effective than full-scale or large-scale testing, wind tunnel
testing provides a steady and controllable test environment where the individual forces acting
on the sails can be measured. Wind tunnel testing is much less time consuming than testing
on the water and can be conducted much earlier in the design process to evaluate not only sail
shapes but also rig options. However, care must be taken to simulate the full-scale wind
conditions correctly in the wind tunnel. Since soft sails with similar properties to full-scale
sails are used in the wind tunnel they can be trimmed as in real life. The wind tunnel models
are equipped with remote control winches so that the primary sail controls can be adjusted
while the forces are measured. Modelling a soft sail in viscous flow CFD simulations
increases the computational time immensely as there must be an iterative process between the
calculated sail forces and the corresponding new shape adopted by the sail until force
equilibrium is satisfied. For example, Graf and Renzsch (2006) conducted research in this
area, but trimming the sail by looking at the shape and the resulting performance, as one does
on the water, is not possible with viscous flow CFD simulations at present.
For all these reasons wind tunnel testing is an effective tool for designing rigs and sails. It is
used by researchers to develop and improve semi-empirical aerodynamic force models and to
obtain the empirical input required for these models so that the sail forces can be calculated
by VPPs. It is also useful in analysing and understanding the flow behaviour around sails by
using flow visualisation techniques and surface pressure measurements. Yacht designers use
the wind tunnel for comparative testing of different rig configurations and to obtain empirical
input for VPPs that is specific to their designs. Sail designers come to the wind tunnel to
compare different sails to determine which is the fastest for each sailing condition. From this
the best sails can be chosen and a sail selection chart developed, which helps the crew to
decide which sail to use for a given true wind speed and direction.
Researchers and designers are generally satisfied that wind tunnel testing of yacht sails and
components gives reliable qualitative results but are unsure of the quantitative accuracy. In
practical terms this means that a sail that performs better in the wind tunnel will perform
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 9
Chapter 1 Introduction
better in real life as well, but it is not known how well the difference in sail performance
measured in the wind tunnel translates to full-scale.
1.4.2 Trimming and Depowering of Sails
Unlike most other lifting bodies sails are not rigid and their shape can be adjusted. In the
wind tunnel soft sails of similar material properties to those of the full-scale sails are
therefore usually preferred so that their flying shape can be altered as in real life to give the
optimum shape for a given sailing condition. This process of trimming the sails is very
important both at full-scale and in the wind tunnel. Due to the large number of possible
adjustments trimming is also complex and requires a lot of experience. During wind tunnel
tests professional sailors are often present to trim the sails. On a real yacht a good guide of
how well the sails are trimmed is the ratio between the achieved speed of the yacht and the
target speed predicted by a VPP. In order to find the aerodynamically most efficient shape of
the sail in the wind tunnel, the drive force, which is the force along the centreline of the yacht
(if leeway is ignored), needs to be maximised by trimming the sails. However, to complicate
matters further the aerodynamically most efficient shape does not necessarily result in the
highest boat speed, since all aerodynamic forces and moments need to be balanced by the
yachts hydrostatics and hydrodynamics, so that the yacht sails in equilibrium. The heeling
moment (M
X
) generated by the sails can lead to an excessive heel angle of the yacht, which
increases the hydrodynamic resistance and reduces the aerodynamic efficiency. In this
situation, departing from the aerodynamically most efficient shape by, for example, reducing
the angle of attack at the top of the sails to decrease M
X
, may increase the boat speed.
Similarly reducing the aerodynamic side force and consequently the leeway angle (), or
reducing the yaw moment (M
Z
) and hence rudder angle (), may be beneficial in certain
conditions. The process of departing from the aerodynamically most efficient sail shape is
called the depowering of a sail. This is another reason why trimming is a complex task
which requires a lot of experience. In real life the boat speed, heel angle and the feel of the
yacht can be used as trim indicators; this is not possible in the wind tunnel.
It follows that for VPP calculations to compare the performance of different sails or other
yacht components, not only the aerodynamically most efficient sail shape needs to be
considered but also the depowered sail shapes to ensure that the best boat speed is obtained.
This is however not trivial since there is a very large number of possible flying shapes a sail
can adopt. The traditional approach adopted for most upwind investigations is to obtain the
wind tunnel results for the aerodynamically most efficient shape, i.e. the optimum lift and
drag coefficients, and mathematically model the depowering by using trim parameters as
described in section 4.2.1. However, the trim parameters currently used may not provide a
good physical description of how the forces and moments change as a sail is depowered,
10
YACHT
RESEARCH
UNIT
The
University
of Auckland
1.5 Motivation
especially for offwind and downwind sailing conditions. This is difficult to confirm with
current methods since determining the realistically depowered sail shapes is complex in the
wind tunnel or CFD simulations. Methods have been introduced by Campbell (1997) and
Ranzenbach and Teeters (2002) which obtain the forces acting on the sails in the wind tunnel
not only for the aerodynamically most efficient shape but also for depowered shapes.
However, the fundamental problem that the trimming in the wind tunnel is conceptually
different from full-scale is however not solved by these methods as will be discussed in
chapter 4.
1.5 Motivation
The availability of full-scale aerodynamic data provides an interesting opportunity to
compare full-scale and wind tunnel measurements for the first time. This is important to
confirm the usefulness of wind tunnel testing and to gain confidence in aerodynamic force
models developed from wind tunnel tests. The full-scale sail force measurements from the
Technical University Berlin, the access to wind tunnel testing facilities for testing yacht sails
at the TFWT, and the expertise on hand in the field of sail force aerodynamics at the Yacht
Research Unit of The University of Auckland, provide the opportunity to conduct such a
comparison.
In the process of realising this comparison a number of other investigations are necessary to
improve the quality of the comparison. Most of the additional investigations are related to
modelling the aerodynamic forces in VPPs, which in itself is a very interesting and important
topic. Deficiencies are still present in VPPs to accurately predict the performance of a yacht
in all sailing conditions. One major reason for these deficiencies is believed to be the
aerodynamic force modelling part of the VPP description. Hence an improvement in the
aerodynamic force modelling would improve the accuracy of the performance prediction and
in turn the effectiveness of the VPP as a design tool. A more effective VPP may well give the
designer a competitive advantage in producing a winning yacht.
During the last 25 years the aerodynamic sail force modelling in VPPs has not fundamentally
changed and many people would like to see improvements being made. One reason for the
slow progress could be the fact that some fundamental differences exist in the way the sails
are trimmed in the wind tunnel compared to on the water. If the assumptions that are used to
account for this are not very accurate, this could be one reason for the inaccuracies in VPP
calculations since many aerodynamic VPP input parameters are obtained from wind tunnel
tests. This provides the impetus to develop a system to make trimming in the wind tunnel
more similar to the real life situation and thereby fundamentally change the way sail forces
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 11
Chapter 1 Introduction
are modelled by a VPP. In fact the forces do not need to be mathematically modelled
anymore, they simply act on the virtual yacht.
1.6 Project Objectives
The ultimate objective of the project is to compare the full-size aerodynamic forces measured
by the Berlin Sail-Force-Dynamometer, to forces obtained from wind tunnel tests for the first
time, and to show in which aspects the results differ. In order to achieve this objective, a
number of other novel investigations related to the aerodynamic force modelling, are
conducted.
The first additional objective is to investigate the interaction of the sails with the hull/deck
which has not been investigated in detail before. In the context of the full-scale and wind
tunnel comparison, this is important because wind tunnel tests traditionally include the force
and moment component due to the interaction whereas it is not included in the full-scale
measurements. For VPPs understanding the interaction is also important since it is
inconsistently modelled at present depending on the origin of the aerodynamic input
parameters.
Another objective is to develop a system to more realistically trim the sails in the wind
tunnel. The aerodynamically most efficient sail shape can be obtained reliably by experienced
trimmers by maximising the drive force measured by the force balance in the tunnel.
However, obtaining the best depowered sail shapes is more complex in the wind tunnel since
the sails cannot be trimmed based on the performance and behaviour of the yacht as is
possible in real life. By developing a novel system to predict the performance of the yacht
while the sails are adjusted in the wind tunnel, the process of trimming becomes much more
similar to the real life situation and depowered sail shapes in the wind tunnel are more
realistic.
The final objective is to improve the aerodynamic force modelling by utilising the more
realistically depowered sail trims and the information obtained regarding the hull/deck and
sail interaction. An improved aerodynamic force model enhances the quality and accuracy of
VPP calculations. It also enables a more realistic comparison between the full-scale and wind
tunnel measurements to be made, which is the ultimate goal of the project.
The four novel objectives of the project can now be summarised:
I. Determine the interaction of hull/deck and sails. (Chapter 3)
II. Develop a system to trim the sails more realistically in the wind tunnel. (Chapter 4)
12
YACHT
RESEARCH
UNIT
The
University
of Auckland
1.7 Methodology
III. Improve the aerodynamic sail force modelling for VPPs. (Chapter 5)
IV. Compare the wind tunnel and full-scale measurements. (Chapter 6)
1.7 Methodology
The research project is experimentally based and uses a variety of wind tunnel testing
techniques in the TFWT and analysis methods to achieve the objectives stated in section 1.6.
Firstly a wind tunnel model of Dyna, the Berlin Sail-Force-Dynamometer, for testing in the
TFWT was designed and built, which requires careful consideration of features previously
not implemented on wind tunnel models. All measurements for this work were conducted
with the model of Dyna with the wind tunnel set up in the usual way appropriate for testing
yacht sails as discussed in section 2.5.
A number of standard test procedures already implemented at the TFWT were used and new
test procedures have been developed. The aerodynamic forces acting on the model were
measured with the force balance of the TFWT. For the investigation on the interaction of
hull/deck and sails in chapter 3, a secondary force balance was utilised to additionally
measure the forces acting on the hull/deck. A second method to obtain the hull/deck forces by
measuring the surface pressures on the hull/deck was also implemented, which furthermore
yields the pressure distribution and helps to understand the flow around the yacht.
To realise the second objective of achieving more realistic sail trims in the wind tunnel, a
novel tool has been developed. The Real-Time VPP discussed, in chapter 4, calculates the
performance of the yacht based on the wind tunnel measurements while the sails are adjusted
so that the trimmer gets immediate feedback on the trim changes. It was decided to use an
existing VPP and incorporate in it the wind tunnel testing procedure rather than develop a
VPP for the wind tunnel from first principles. This way all the resources can be put towards
realising the Real-Time VPP in the wind tunnel and no resources are used to repeat work
already done before. Realising the Real-Time VPP included developing the concept,
programming software for the wind tunnel, and modifying and extending the existing VPP.
The program FRIENDSHIP-Equilibrium developed by Hochkirch (2000) at the Technical
University Berlin and FRIENDSHIP SYSTEMS is used and the author was able to spend
time with these institutions during the development process. With the new Real-Time VPP
realistic depowered trims are obtained. The results are compared to calculations with the
standard aerodynamic force models and based on the comparison improvements are made to
achieve the third objective in chapter 5.
For the comparison of the wind tunnel data to full-scale, measurements conducted by the
Technical University Berlin on Dyna are used. The analysis of the data is carried out by the
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 13
Chapter 1 Introduction
author and corrections based on wind tunnel experiments in the TFWT are applied as
discussed in chapter 6. For the data analysis it proved very helpful that the author had the
opportunity to be involved in collecting full-scale data on Dyna and appreciate the challenges
associated with full-scale testing on the water.
Some of the work presented in this thesis has previously been published. The Real-Time VPP
was introduced in a conference paper by Hansen et al. (2003b). Results of the investigation
on the hull and sail force interaction and improvements in aerodynamic force modelling were
presented in conference papers by Hansen et al. (2006) and Richards et al. (2006). A
preliminary comparison of wind tunnel and full-scale data was shown in a conference paper
by Hansen et al. (2002) which was also published in an international journal (Hansen et al.,
2003a). The enhanced comparison discussed in this thesis will be presented at the 18
th

Chesapeake Sailing Yacht Symposium. Furthermore, a comprehensive overview of advances
in wind tunnel testing techniques, many of which are related to this work, was given by
Hansen et al. (2005).
14
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
2 Wind Tunnel Testing of Yacht Sails
2.1 Introduction
Wind tunnel testing of yacht sails in general, and more specifically at the Twisted Flow Wind
Tunnel (TFWT) of The University of Auckland, is an integral part of this project. This
chapter introduces the main aerodynamic principles important in relation to yacht sails and
the mathematical methods used to describe them. The methods, wind tunnel set-up, and
procedures discussed in this chapter, form the foundation on which the developments of this
project are built.
2.2 Terminology of Sailing
There are many terms that are specific to sailing and the most important ones referred to in
this work are shown in Figure 2.1 or briefly explained in Table 2.1.
Mainsail
Headsail
(here jib)
Forestay
Hull
Jib sheet
Jib track
Shrouds
Mast
Runner
Backstay
Traveller
Spreader
Luff
Leech
Foot
Head
Tack
Clew
Boom
Main sheet
Vang
Sheerline
Stem
1
5
4
2
Transom
Cockpit 4
Legend
Coachroof 5
Deck 3
Stern 2
Bow 1
Cockpit 4
Legend
Coachroof 5
Deck 3
Stern 2
Bow 1
3
3
z
boom
z
mast
DWL
Rudder
Keel
E J
I
P

Figure 2.1: Diagram showing components of a yacht
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 15
Chapter 2 Wind Tunnel Testing of Yacht Sails
Table 2.1: Alphabetic listing of sailing specific terms
Term Description
Appendage An attachment to the hull of a yacht that predominately creates
hydrodynamic forces (for example a keel or rudder).
Depowering Departing from the aerodynamically most efficient sail shape by
changing the trim of the sail so that the sail develops less drive
force.
DWL Design Waterline, the flotation line of the upright yacht as defined
in the design specification.
Freeboard Vertical distance from the DWL to the sheerline of the upright yacht.
Leeward Away from the wind, opposite of windward.
Port Left of the centreline of the yacht when facing towards the front
(bow) of the yacht. Y-values are positive to port.
Port tack The yacht is sailing with the onset flow coming from the port side
(left side) of the yacht.
Reaching Sailing a course that does not require maximising VMG to
windward or leeward.
Roach The curved portion of a sail extending past a straight line drawn
between two corners. In a mainsail, the roach extends past the line
between the head and the clew.
Running rigging Wires and ropes that hold the mast and sails, and can be moved
readily, such as running backstays, halyards and sheets.
Sailing downwind Sailing away from the wind often at courses close to achieving the
maximum VMG to leeward.
Sailing off the wind Not sailing as close to the wind as possible.
Sailing upwind Sailing towards the wind often at courses close to achieving the
maximum VMG to windward.
Standing rigging Wires and ropes that hold the mast and sails, and cannot be moved
readily, such as shrouds, forestay and backstay.
Starboard Right of the centreline of the yacht when facing towards the front
(bow) of the yacht. Y-values are negative to starboard.
Starboard tack The yacht is sailing with the onset flow coming from the starboard
side (right side) of the yacht.
Trimming Adjusting the sail control lines and mechanisms to change the sail
shape relative to the onset flow. The main trimming controls can be
adjusted by remote control winches in the wind tunnel.
VMG Velocity Made Good Boat speed towards a target usually directly
upwind or downwind.
Windage Aerodynamic forces (usually mainly drag) on the hull and rig.
Windward Towards the wind, opposite of leeward.
Winglet Wing-type fin at the tip of a keel or on a keel bulb to reduce induced
drag and to generate forward force from the tip vortex.

16
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.3 Aerodynamics of Sailing
2.3 Aerodynamics of Sailing
Many principles of the aerodynamics of sails can be taken from thin aerofoil theory as
discussed by Abbott and von Doenhoff (1959), and the associated lift and drag is discussed in
even greater detail by Hoerner (1965) and Hoerner and Borst (1985). There are however
significant differences in certain areas. Compared to thin wing sections, the curvature of sails
and the angle of attack can be much larger and hence separated flow regions are much more
significant. Also, as a result of the heel angle of the yacht, the onset flow is not perpendicular
to the leading edge but similar to the onset flow onto a wing with sweepback, which will be
discussed in section 2.3.1. Finally due to the planetary boundary layer (PBL) the onset flow is
much less uniform along the span than in most aeronautic applications, which will be
discussed in section 2.3.2. Many aspects of sail aerodynamics are introduced in the book on
aero-hydrodynamics of sailing by Marchaj (1988).
In many ways, much like a wing, a sail is a lifting body, which creates lift from the fluid flow
around it. Due to its shape and the direction of the onset flow, circulation is created by a
starting vortex. The circulation increases the velocity on the upper side of the wing (leeward
side of the sail), and decreases the velocity on lower side of the wing (windward side of the
sail), so that the second stagnation point is situated at the trailing edge.
Following from Bernoullis equation the higher velocity on the leeward side creates a low-
pressure region and the lower velocity on the windward side leads to a high-pressure region.
The pressure distribution along the chord of a typical mainsail section at an angle of attack
() of 9.5 is shown in Figure 2.2. The pressure coefficient (C
P
) is relative to the static free-
-3
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Chord
C
P
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
C
a
m
b
e
r
Low-pressure on leeward side
High pressure on windward side
Sail section shape
Lift
Flow
Drag

Figure 2.2: Pressure distribution along the chord of a sail and the resulting lift and drag
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 17
Chapter 2 Wind Tunnel Testing of Yacht Sails
stream and normalised by the free-stream dynamic pressure. The forces resulting from the
pressure difference between the pressure and suction side are usually expressed in terms of
lift and drag, where drag is defined as the force acting in line with the free-stream onset flow
and the lift as the force acting perpendicular to the free-stream onset flow as shown in Figure
2.2.
The lift and drag are expressed in terms of non-dimensional force coefficients so that scaling
operations and comparisons are easier to perform. Any force F can be expressed as a
dimensionless force coefficient (C
F
) with
ref
F
qA
F
C = ,
(2.1)
where q is the dynamic pressure of the free-stream onset flow and A
ref
is the reference area.
The dynamic pressure is
2
2
V q

= ,
(2.2)
where is the density of the fluid, which in the case of sails is generally air, and V is the free-
stream onset speed of the flow. A
ref
is usually taken as the planform area for the lift and
induced drag. The planform of a wing is the shape of the wing when looking down onto it
from above. There is merit in taking the surface area as A
ref
for skin friction drag and viscous
pressure drag but often the planform area is used as well. Determining the planform area for
sails is difficult since it changes with sail trim. It is therefore common to calculate the
reference sail area (A ) according to the IMS formulation by Poor (1986) with
S
2 16 . 1
IJ A
A
Smain
S
+ = ,
(2.3)
where the A
Smain
is the cloth area of the mainsail, I is the height of the fore triangle and J is
the distance between the forestay/deck attachment point and the mast as shown in Figure 2.1.
Moments can also be expressed in terms of non-dimensional coefficients, where A
ref
and a
reference length or A
ref
1.5
are commonly employed reference parameters. In this work A
ref
1.5
is
used so that the dimensionless moment coefficient (C
M
) can be obtained from a Moment M
with
5 . 1
ref
M
qA
M
C = .
(2.4)
The lift and drag coefficients of a wing and the associated moment coefficients are primarily
a function of the angle of attack. However, on a sailing yacht it is not trivial to measure the
18
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.3 Aerodynamics of Sailing

A
= +

V
A
V
A
Apparent wind velocity

A
Apparent wind angle
Angle of attack
Sheeting angle
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 20 40 60 80 100 120 140 160 180

A
[]
C
L

&

C
D

[
-
]
CL
CD
Figure 2.4: Lift and drag coefficient of mainsail and jib
combination against apparent wind angle
Figure 2.3: Apparent wind angle, angle
of attack and sheeting angle
relationship
angle of attack () since the sail shape is constantly changed by trimming. Hence is usually
replaced by the apparent wind angle (
A
), which is the angle between the centreline of the
yacht and the relative free-stream onset flow (Figure 2.3). The free-stream speed of the onset
flow is usually called the apparent wind speed (V
A
) in sailing aerodynamics. Similarly q is
called the apparent dynamic pressure (q
A
).
A
is related to by the sheeting angle () as
shown in Figure 2.3. The force coefficients generated by the sails depend on many
parameters, including the sail geometry, the relative direction of the onset flow, the flow
structure, the flow speed relative to the size and the trim. The force coefficients of a given sail
design under normal sailing conditions trimmed to the most aerodynamically efficient shape
are primarily a function of
A
. A typical representation of the forces acting on a mainsail and
jib combination are shown in Figure 2.4 in terms of lift and drag coefficients (C and C
L D
)
plotted against apparent wind angle (
A
). The effects of heel are still ignored here and will be
discussed in section 2.3.1.
The drag is made up of several components. The flow over the sail surface creates the skin
friction drag. The viscous pressure drag is often combined with the skin friction drag to give
the drag at zero lift (C
D0
). For sails this is rather academic since a sail flaps at zero lift and
would have a much higher drag due to vortex shedding. It is therefore common to call this
drag component the parasitic drag coefficient (C
Dp
) in relation to sail aerodynamics instead of
C . Another component is the separation drag (C
D0 Ds
) due to the flow separating from the
surface of the foil or sail. In a general form it is expressed as a function of C
2
and C
L L
.
However, in relation to sail aerodynamics C
Ds
is typically approximated simply as a function
of C
2
by
L
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 19
Chapter 2 Wind Tunnel Testing of Yacht Sails
,
2
L s Ds
C c C =
(2.5)
where c is the separation drag constant. Kerwin (1978) determined c
s s
to be 0.019 for mainsail
with spinnaker and 0.016 for mainsail with jib. It is however not trivial to separate and
determine C and it is not uncommon to express C and C
Ds Dp Ds
together as the viscous drag
coefficient (C ).
Dvis
Apart from the viscous drag, the induced drag due to lift is the second major component of
drag. The induced drag coefficient (C
2
) is ideally a function of C
Di L
and depends on the
geometric aspect ratio (AR) and the efficiency (e) of the rig so that
AR
C
e
C
L
Di

2
1
= .
(2.6)
The aspect ratio of the rig is generally determined from the height of the mast above the
design waterline (DWL) plane (z
mast
) and the reference sail area (A ) with
S
S
mast
A
z
AR
2
= .
(2.7)
The total drag coefficient is usually expressed simply as the sum of its two major
components, the viscous and the induced drag coefficients, so that
. ) ( ) ( ) (
2 2 2
L Di L Ds Dp L Di Dvis D
C C C C C C C C C + + = + =
(2.8)
As shown in Figure 2.4 C and C
L D
are usually expressed as functions of the apparent wind
angle. The apparent wind experienced by a sail is the combination of the environmental wind
and the wind generated by the moving yacht. The environmental wind is called the true wind
by yachtsmen. At any particular height the wind triangle shown in Figure 2.5 relates the
apparent wind angle (
A
) and speed (V ) and speed (V
A
) to the true wind angle (
T T
), the boat
speed (V
S
) and the leeway angle (). From this geometric relationship
A
and V
A
can be
obtained from , , V and V with
T T S
( )
( )
[ ]
o
180 , 0 ,
cos
sin
tan
1

+ +
+
=

A
S T T
T T
A
V V
V



,
(2.9)
and
) cos( 2
2 2
+ + + =
T T S T S A
V V V V V .
(2.10)
There are two further phenomena that need to be considered in greater detail when dealing
with the aerodynamics of sails. The heel angle () of the yacht will be discussed further in
section 2.3.1 and the non-uniformity of the onset flow is discussed in section 2.3.2.
20
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.3 Aerodynamics of Sailing

A
V
A
V
T
True wind velocity

T
True wind angle
V
S
Boat velocity
Leeway angle
V
A
Apparent wind velocity

A
Apparent wind angle


T
V
T
V
s

Figure 2.5: The wind triangle
2.3.1 Effective Angle Theory
The effective angle concept is used to address the fact that a monohull sailing yacht heels
significantly when sailing. The heel angle influences the flow around the sails since the onset
flow can always be regarded as being horizontal. As the yacht heels the onset flow is no
longer close to perpendicular to the leading edge of the mainsail, as shown in Figure 2.6. Due
to the changes in the flow around the sails the resulting lift and drag forces are different for
each heel angle (). The lift and drag coefficients are hence not only a function of
A
but also
of . In order to make the analysis and the modelling more generic it is desirable to combine

A
and into a single angle. This angle is called the effective angle (
eff
) and is obtained from
the effective angle theory. Kerwin (1978) assumes that the sails are insensitive to the flow
component along their span (i.e. along the mast) and that only the flow component
perpendicular to the mast produces the lift and drag forces. This is further discussed by
Jackson (2001) who points out that there is strong evidence to support this model for aircraft
wings and, although matters are less certain for sails, the approach still appears to be the
correct starting point. The appropriateness of this assumption will also be investigated in
sections 3.7 and 6.7.
The flow component along the chord of the sails can be seen as the flow component in the
heeled plane, which is the plane perpendicular to the z-axis in the body fixed coordinate
system (coordinate system B) defined in appendix C.1. Ignoring mast rake and bend the
heeled plane can be imagined as being perpendicular to the mast. The flow component in the
heeled plane is called the effective flow and is defined by the effective angle (
eff
) and
effective speed (V
eff
) as shown in Figure 2.7.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 21
Chapter 2 Wind Tunnel Testing of Yacht Sails

eff
V
eff

A
V
A
Figure 2.6: Diagram of onset flow in the
horizontal plane
Figure 2.7: Diagram of onset flow component in
the heeled plane
The effective angle (
eff
) is geometrically related to
A
and and can be calculated from
( ) [ ]
o
180 , 0 , cos tan tan
1
=

eff A eff
.
(2.11)
Similarly the effective speed (V
eff
) can be obtained from
A
and as a fraction of the apparent
wind speed (V
A
) as

2 2
sin sin 1
A A eff
V V = .
(2.12)
In addition to heel, a sailing yacht can also be rotated about the y-axis and the flow
component along the chord of the sails is influenced by the resulting pitch angle (), which
usually is much smaller than the heel angle but should nevertheless be considered when
calculating the effective angle and speed. In the wind tunnel the model generally has a pitch
angle of zero but for full-scale considerations is not equal to zero and equations (2.11) and
(2.12) can be extended so that
[ ]
o
180 , 0 ,
cos
sin sin cos tan
tan
1


=

eff
A
eff


,
(2.13)
and
22
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.3 Aerodynamics of Sailing
( )
2 2 2
cos cos sin sin cos cos sin
A A A A eff
V V + = ,
(2.14)
where bow down trim of the yacht is a positive pitch angle (). Alternatively to extending
equations (2.11) and (2.12) it can be convenient to define the flow in vector form to include
the effects of pitch. The apparent wind velocity vector V
T
A
=(V , V , V )
Ax Ay Az
in the absolute
coordinate system (coordinate system A defined in appendix C.1) is given by
,

=
0
sin
cos
A
A
A A
V

V
(2.15)
T
and the effective wind velocity vector V
eff
=(V
effx
, V
effy
, V )
effz
in coordinate system B can then
be obtained from
,
A AB eff
V T V =
(2.16)
where T
AB
, is the rotational transformation matrix to transfer from the absolute coordinate
system (coordinate system A) to the body fixed coordinate system (coordinate system B) by
rotating through the heel angle () and the pitch angle (). T
AB
can be written as
,

=



cos cos sin sin cos
cos sin cos sin sin
sin 0 cos
AB
T
(2.17)
where the coordinate systems are defined in appendix C.1 and the sign conventions and
transformations are discussed further in section 2.4.1.
The effective angle (
eff
) is the angle between the flow component in the z-plane and the x-
axis given by
[ ]
o
180 , 0 , tan
1

=

eff
effx
effy
eff
V
V
,
(2.18)
and the effective wind speed (V
eff
) is the resulting flow component in the z-plane obtained
from
2 2
effy effx eff
V V V + = .
(2.19)
Following the assumption that the sails are insensitive to flow along their span, it is also
assumed that the sails do not produce any forces along their span and that all the forces lie in
the heeled plane. It can then be concluded that C and C
L D
in the heeled plane (coordinate
system B) plotted against
eff
describe the forces acting on the sails for any heel angle,
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 23
Chapter 2 Wind Tunnel Testing of Yacht Sails
providing that C and C
L D
are calculated using the effective pressure (q
eff
), which is calculated
from equation using V (2.2)
eff
, as the onset flow dynamic pressure in equation (2.1).
2.3.2 Vertical Speed Gradient and Twist in Onset Flow
In real life the onset flow onto a sail of a yacht is not uniform. The flow has a vertical speed
gradient and twist since the speed and direction of the flow change with height. This complex
flow structure results from the vertical speed gradient in the atmospheric wind combined with
the movement of the yacht.
At sufficiently great heights above the surface of the earth the influence of friction along the
ground becomes negligible. This region is called the free atmosphere. In unaccelerated free
atmospheric flow the horizontal motion of the air relative to the earths surface can be
obtained from the force balance of the pressure gradient, Coriolis force and centrifugal force
as shown by Simiu and Scanlan (1996). The resulting gradient wind velocity is parallel to the
isobars.
Within the atmospheric or planetary boundary layer (PBL) the wind velocity is slowed down
by the friction along the ground as shown in Figure 2.8. The atmospheric boundary layer
depth is typically between several hundred and 3000m depending on the wind intensity,
roughness of the terrain and angle of latitude as described by Simiu and Scanlan (1996). Due
to the reduced wind velocity in the boundary layer the Coriolis and centrifugal force decrease
whereas the pressure gradient force remains the same. As a result the wind velocity in the
boundary layer crosses the isobars and is increasingly directed towards the low-pressure
region as the height above the surface reduces. When considering yacht sails only the lower
portion of the boundary layer up to about 100m above the ground is of interest and the change
in direction in this region is small enough to be ignored. The atmospheric wind in relation to
sailing is called the true wind. The true wind direction and speed are usually defined at a
reference height (z
ref
) of 10m above the water.
Extensive research has been conducted over the years to measure and describe the velocity
profiles in boundary layers. In relation to the atmospheric boundary layer the first
representation of the wind profile in a horizontally homogeneous terrain was the power law
proposed in 1916 by Hellman, as mentioned in Simiu and Scanlan (1996). Initially the
assumption was made that the entire boundary layer depth can be approximated with a
constant exponent. A number of experiments and real life observations have shown that the
lower fraction (~10%) of the atmospheric boundary layer can be described by a logarithmic
velocity profile very accurately (Simiu and Scanlan, 1996). Hence empirical velocity profiles
24
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.3 Aerodynamics of Sailing
have been developed from theory and experiments, which are a combination of the
logarithmic law and the power law where the exponent varies with height.
For the application to yacht sails the influence of the non-logarithmic terms can however be
ignored since the logarithmic law is known to hold well up to a height of at least 100m.
Providing that both z and z
ref
are not larger than 100m, the true wind speed (V
T
) at the height z
can hence be calculated from
) ln(
) ln(
) ( ) (
0
0
z z
z z
z V z V
ref
ref T T
= ,
(2.20)
where z
0
is the terrain roughness length. The roughness length for open water is a function of
the wave height, which again is a function of the wind speed. Cook (1985) hence gives the
roughness length as
g
z V
z
ref T
2
5
0
) (
10 5

,
(2.21)
where z
ref
is 10m above the water and g is the acceleration due to gravity. For typical true
wind speeds between 5m/s and 15m/s this gives roughness lengths from 0.1mm to 1mm. If a
constant value is assumed Kerwin (1978) suggests using 1mm. Clearly ocean waves are much
larger then these roughness lengths, but their crests are rounded enough for the flow to follow
the curvature without separating, which reduces the friction and in turn the roughness length.
In general ocean waves travel in the direction of the wind, which reduces the speed
Earth surface
Gradient wind
level
Free atmosphere
Atmospheric
boundary layer
Boundary layer depth
Gradient wind
velocity

Yacht
velocity (V
S
)
True wind angle (
T
)
Apparent wind
angle (
A
)
True wind
velocity (V
T
)
Apparent wind
velocity (V
A
)

Figure 2.8: Atmospheric boundary layer Figure 2.9: Vertical speed gradient and
twist in onset flow onto a sail of
a yacht on port tack ignoring
the leeway angle
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 25
Chapter 2 Wind Tunnel Testing of Yacht Sails
differential, friction and roughness length. This expression for z
0
is very generalised and more
research is required for the roughness length of waves which are not fully developed.
Similarly the logarithmic law only applies for homogeneous terrain and gradient wind but not
for localised wind close to shore.
As shown in Figure 2.5 the flow relevant to the sail is the resultant of the true wind and the
wind due to the motion of the yacht and is called the apparent wind. The speed profile of the
true wind and the constant wind due to the motion of the yacht result in the apparent wind
varying not only in speed but also in angle with height above the water as illustrated in Figure
2.9. The apparent wind speed and angle generally increase with height above the water. The
change in apparent angle and speed does not only depend on the true wind speed, angle and
profile but also on the speed and size of the yacht.
Figure 2.10 shows the change in apparent wind speed (V
A
/V
Aref
) with height up the mast in
non-dimensional form (z/z
mast
). V
A
/V
Aref
is shown for two types of yachts, a 10m IMS
cruiser/racer and a Volvo Ocean 60 (VO60), sailing in a V
T
of 8m/s for two sailing
conditions, sailing upwind at of 40 and sailing downwind at
T T
of 160. The reference
apparent wind speed (V ) is calculated for each
Aref T
at a height 40% up the mast. The profiles
for both yacht types are similar in each sailing conditions. The change of V
A
/V
Aref
is
significantly larger when sailing downwind for both yacht types. From the geometric
relationship presented in Figure 2.5 it can be deduced that V
A
decreases as
T
increases. As a
result a change in V with height has a stronger effect on V
T A
/V as
Aref T
increases. For the
VO60 sailing downwind the V
A
has a minimum and increases rapidly close to the water
surface. Compared to the 10m IMS cruiser/racer, the VO60 is able to achieve a higher V /V
S T

ratio because it is a larger yacht and a more high-performance orientated design. As a result
the V is much larger than the V
S T
close to surface and hence dictates the wind direction and
speed in this region. When sailing deep downwind this can lead to an increase in V
A
close to
the water surface. The sail performance is however not significantly affected since this
increase in V
A
usually occurs below the foot of the sails.
The change in V
A
/V
Aref
is slightly larger for the 10m IMS cruiser/racer in both conditions.
Since its mast height is smaller than the VO60 and the true wind speed gradient is larger close
to the surface, the resulting change in V
A
/V
Aref
over the height of the mast is more pronounced
for the smaller yacht.
Figure 2.11 shows the change in apparent wind angle (
A
) with height up the mast for the
same two yacht types and sailing conditions as before. The reference height for
A
is again
taken as 40% of the mast height. For sailing upwind it can be seen that
A
over the span of
the sails is only 2-3 and the difference between the twist profiles of the two yacht types is
not distinguishable in Figure 2.11. For the downwind condition the twist is much more
26
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.3 Aerodynamics of Sailing
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
V
A
/V
Aref
[-]
z
/
z
m
a
s
t

[
-
]
VO 60 downwind
10m IMS downwind
VO 60 upwind
10m IMS upwind
VO 60 foot of sails
10m IMS foot of sails
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
-30 -25 -20 -15 -10 -5 0 5 10 15

A
[]
z
/
z
m
a
s
t

[
-
]
VO 60 downwind
10m IMS downwind
VO 60 upwind
10m IMS upwind
VO 60 foot of sails
10m IMS foot of sails
Figure 2.11: Twist profiles for a 10m IMS
cruiser/racer and a Volvo Ocean 60
for true wind angles of 40 and 160
Figure 2.10: Speed profiles for a 10m IMS
cruiser/racer and a Volvo Ocean 60
for true wind angles of 40 and 160
significant. For the 10m IMS cruiser/racer the
A
over the span of the sails is about 10 and
for the VO60 approximately 35.
A
is much larger for the VO60 since it is able to achieve a
higher V /V ratio, which, as described before, results in V
S T S
being the main contributor to the
apparent wind speed and direction close to the water surface.
Generally it can be said that both the vertical speed gradient and twist in the onset flow onto a
sail increase with the yacht performance capability relative to V , i.e. the V /V
T S T
ratio, and with

T.
The V /V
S T
for a high-performance yacht can be in the order of 1 when sailing upwind, 1.5
when reaching and 0.8 when sailing downwind. However, for cutting-edge designs such as
Volvo Open 70 (VO70) yachts and multihulls, V /V can be higher.
S T
can be rearranged and V Equation (2.9) substituted from equation (2.20)
T
to account for the
change in V with height to give
T
[ ]
o
180 , 0 ) ( ,
) ln(
) ln(
) (
cos
sin
tan ) (
0
0
1

+
=

z
z z
z z
z V
V
z
A
ref
ref T
S
T
T
A

,
(2.22)
where
A
(z) is the apparent wind angle at height z. The twist in the apparent wind angle (
A
)
over the span of the sails is the difference between
A
at the height of the top of the mast
(z
mast
) and
A
at the height of the foot of the sail (z
foot
) so that
. ) ( ) (
foot A mast A A
z z =
(2.23)
If the roughness length (z ) is assumed to be independent of V between z
0 T
,
A mast
and z
foot
is
only a function of and V /V and V /V . In Figure 2.12
T S T A
is plotted against for a VO60
T S T
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 27
Chapter 2 Wind Tunnel Testing of Yacht Sails
0 20 40 60 80 100 120 140 160 180
0.5
0.6
0.7
0.8
0.9
1
1.1
1.2
1.3
1.4
1.5

T
[]
V
S
/
V
T

[
-
]
2
4
6
8
1
0
1
5
2
0
2
5
3
0
4
0
5
0
0 20 40 60 80 100 120 140 160 180
0.5
0.6
0.7
0.8
0.9
1
1.1
1.2
1.3
1.4
1.5

T
[]
V
S
/
V
T

[
-
]
2
4
6
8
1
0
1
5
2
0
2
5
3
0
4
0
5
0
z =28.5m, z
yacht and in Figure 2.13 for a 10m IMS cruiser/racer. These plots are a useful reference to
quickly get an idea of the changes in the apparent wind angle that can be expected for a
certain sailing condition.
The vertical gradient in the onset flow can be simulated in standard atmospheric wind tunnels
but the twist cannot. For upwind sailing conditions the twist is not significant but as the
T

increases the change in
A
gets larger and is significant for high-performance yachts in
downwind sailing conditions. This explains why the Twisted Flow Wind Tunnel has been
developed primarily for downwind sail testing as described in sections 1.4 and 2.4.
As described in section 2.3 it is common and convenient to express aerodynamic lift and drag
forces in coefficient form with equation (2.1). However, in flow with a vertical speed gradient
V
A
is not constant along the span of a sail. The lift and drag are also usually described with
respect to
A
and in twisted flow
A
is not constant along the span of a sail either. The most
meaningful position to use as a reference point is the position of the centre of pressure (also
called the centre of effort) along the span of a sail. However, the centre of pressure changes
and cannot always be obtained easily; hence it is common to use the geometric centre of the
sail area to approximate the centre of pressure of the sail. The height of the geometric centre
of the sail area (z
CoA
) is therefore used as the reference position for V
A
and
A
. As the yacht
heels, the height above the water (z) of z
CoA
reduces. V
A
and
A
hence need to be obtained at
the height z, which can be calculated from
cos
CoA
z z = ,
(2.24)
where is the heel angle of the yacht, by assuming that the yacht, when heeling, rotates about
an axis which intersects the design waterline (DWL) and the centreline plane of the yacht.
mast foot
=2m and z z =15.5m, z
0
=0.001m
mast foot
=1.5m and z
0
=0.001m
Figure 2.12: Figure 2.13:
A
(twist) over span of sails plotted
as a function of
A
(twist) over span of sails plotted
as a function of and V
T S
/V and V
T
for a
VO60 yacht
T S
/V
T
for a
10m IMS cruiser/racer
28
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.3 Aerodynamics of Sailing
2.3.3 Position of Centre of Effort
While it is sufficient to approximate the centre of effort for the calculation of the apparent
wind as the geometric centre of sail area this is not accurate enough for calculating the
aerodynamic sail moments acting on the yacht. The position of the centre of effort changes
with the trim of the sails and depends on
eff
. In most cases the centre of effort is determined
for the optimally trimmed sails from wind tunnel measurements, CFD simulations or
theoretical calculations. Wind tunnel tests and CFD simulations usually provide aerodynamic
forces and moments from which the centre of effort needs to be calculated.
For determining the centre of effort it is most common to assume that the centre of effort
position vector CoE=(x
T
CoE
, y
CoE
, z
CoE
) lies on the centreline plane of the yacht (y-plane of
coordinate system B). The moment vector M=(M
T
X
, M , M )
Y Z
about the origin in a general
force and moment system may be written as
, F CoE M M + =
0 (2.25)
T
where F=(F is the force vector and M
X
, F , F )
Y Z 0
is the moment from pure couples, which
produce a moment but no force. Appendix D.1 illustrates that, if the centre of effort is
confined to lie on the centreline plane of the yacht, i.e. y
CoE
=0, the longitudinal position of the
centre of effort (x
CoE
) can be written as
Y
Z
CoE
F
M
x = ,
(2.26)
and the height of the centre of effort (z
CoE
) as
Y
X
CoE
F
M
z

= .
(2.27)
Appendix D.1 however also shows that there is a moment from pure couples about the y-axis
(M ) of
Y 0
CoE X CoE Z Y Y
z F x F M M + =
0
,
(2.28)
which may not be equal to zero. Although this approach is commonly used, a different
approach is preferred in this work.
In the second approach described by Hochkirch (2000) the centre of effort is defined as the
point where the central axis of a force and moment system intersects the centreline plane of
the yacht (y-plane). The central axis is aligned with the force vector (F) and applying F at any
point along the central axis results in the same moment. Appendix shows that the point a D.2
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 29
Chapter 2 Wind Tunnel Testing of Yacht Sails
on the central axis which is closest to the origin can be determined for a force and moment
system of F and M from
2
F
M F
a

= .
(2.29)
Since F is aligned with the central axis, any point c along the central axis can be seen as the
point through which the force is acting so that the equation of the central axis can be written
as
, F a c q q + = ) (
(2.30)
where q is a scalar factor. The intersection of the central axis and the y-plane can hence be
defined as the CoE and its position is determined from
X
Y
y
x CoE
F
F
a
a x = ,
(2.31)
and
Z
Y
y
z CoE
F
F
a
a z = .
(2.32)
Without a moment from pure couples this method gives the same results as equations (2.26)
and (2.27). If pure couples are present they are now expressed as the moment in line with the
force, which cannot be created by the force, as will be discussed in more detail in section 5.3.
Since in this approach the CoE is seen as the intersection of the resultant force with the
centreline plane, x
CoE
and z
CoE
are plotted as functions of the resultant force direction () by
Hochkirch (2000) where can be calculated from F
X
and F with
Y
X
Y
F
F
1
tan

= .
(2.33)
It is however equally feasible to plot x
CoE
and z against
CoE eff
. This has the advantage that all
parameters describing the forces and moments acting on the sails are a function of the same
variable.



30
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.4 Twisted Flow Wind Tunnel
2.4 Twisted Flow Wind Tunnel
The tests for this work are conducted in the open circuit Twisted Flow Wind Tunnel (TFWT),
which has been developed at The University of Auckland by Flay et al. (1996) especially for
testing yacht sails. Sails experience a change in apparent wind speed (V
A
) and angle (
A
) with
height above the water due to the speed of the yacht and vertical speed gradient in the true
wind as discussed in section 2.3.2. This vertical speed gradient and twist profile is simulated
in the TFWT by generating flow with a planetary boundary layer velocity profile that is
twisted by vertical vanes situated upstream of the test section (see Figure 2.14), which is 7
metres wide and 3.5 metres high. The air is sucked into the open circuit tunnel at one end by
two three-metre diameter four-bladed fans, each driven by a 45kW electric motor. Behind the
fans the flow is straightened and conditioned by a honeycomb grid and two screens with a
tight mesh as illustrated in Figure 2.15. In the enclosed section of the tunnel the planetary
boundary layer profile is developed. For these tests rectangular bars, which extend
horizontally across the width of the tunnel, are used to slow down the flow and increase the
turbulence intensity at the lower heights above the tunnel floor. The bars have different
projected areas and are arranged at specific heights above each other at approximately 20% of

Figure 2.14: Twisted Flow Wind Tunnel (TFWT) with and without flow twisting vanes installed

Honeycomb
23.6m
3.5m
Flow twisting
vanes
Screen 2
Horizontal bars to
stimulate vertical
speed profile
Force balance
Wind velocity
Screen 1

Figure 2.15: Schematic diagram of the Twisted Flow Wind Tunnel (TFWT)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 31
Chapter 2 Wind Tunnel Testing of Yacht Sails
FANS
VERTICAL VANES
TEST SECTION TO HOUSE
MEASURING EQUIPMENT
FANS
ALL FRAMING AT 600 CTRS
BEAMS AT 900 (MIN) CTRS
BEAM
NOT TO SCALE
DETAIL A
SEE DETAIL A
TRANSITION
TRANSITION
340
4100
600
7215 2000
2140
4200
5520
800 12500 3025
1800
FANS
VERTICAL VANES
TEST SECTION TO HOUSE
MEASURING EQUIPMENT
FANS
ALL FRAMING AT 600 CTRS
BEAMS AT 900 (MIN) CTRS
BEAM
NOT TO SCALE
DETAIL A
SEE DETAIL A
TRANSITION
TRANSITION
340
4100
600
7215 2000
2140
4200
5520
800 12500 3025
1800

Figure 2.16: Engineering drawing of the Twisted Flow Wind Tunnel (TFWT)
the length of the enclosed tunnel section downstream of screen 2 so that the desired vertical
velocity profile is created. The vertical vanes to alter the angle of the flow with height are
placed at the end of the enclosed section in front of the open test section. The dimensions and
construction details of the TFWT are shown in Figure 2.16.
2.4.1 Force Measurements
A six-component force balance is located under the wind tunnel floor in the test section as
shown in Figure 2.17. Six Linear Voltage Displacement Transducers (LVDTs) measure the
voltage due to their displacement in one direction and a calibration matrix, as described by
Bonniot (2002), relates the six voltages to three forces and three moments in the force
balance coordinate system (coordinate system F) defined in appendix C.1. The individual
LVDTs behave linearly over the calibration range and it can be shown from the
measurements of Bonniot (2002) that this force balance design also results in a close to linear
behaviour of the whole system. It is therefore adequate to use a linear system of equations,
which results in a six by six matrix of calibration coefficients, to calculate the forces and
moments from the measured voltages. The location and direction of displacement of the
LVDTs are shown in Figure 2.18. One LVDT measures the displacement in the x-direction,
two LVDTs in the y-direction and three LVDTs in the z-direction. As a consequence the
measurements in the x-direction are the most accurate and the one in the z-direction the least
accurate. From the calibration measurements a mean error in the x-direction of 0.09N is
obtained. In the y-direction the mean error is 0.11N and in the z-direction it is 0.27N. For
32
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.4 Twisted Flow Wind Tunnel
X1
Z3
Z2
Y2
Z1
Y1
X-axis
Y-axis
Y-plane
Origin of Coordinate
systems F and T
Figure 2.17: Six-component force balance under
the TFWT floor
Figure 2.18: Origin of force balance and location
of LVDTs
typical measurements this results in an accuracy of 1% in the x and y-direction and 5% in
the z-direction. The force balance is primarily designed to measure static loads and the
measurements are usually averaged over sampling periods between 30 and 120 seconds.
Three brackets of the balance protrude through the tunnel floor to hold the model. The force
balance can be turned together with the floor above (turntable) to change the apparent wind
angle (
A
), so that the force balance coordinate system (coordinate system F) is aligned with
the centreline of the model and the horizontal plane. The origin of the force balance
coordinate system is the intersection of the y-axis, which runs thought the centres of the slots
of the attachment brackets, and the central y-plane as shown in Figure 2.18. The turntable has
a recess to allow the waterline of the hull to coincide with the wind tunnel floor. The recess is
usually filled with water to prevent any airflow under the hull.
The TFWT is set up for the model yacht to be tested on port tack. A non-standard sign
convention is traditionally used in the TFWT to make the most relevant angles and forces
positive for common test conditions. For this work the standard right hand system is however
adopted since the wind tunnel is interfaced with a VPP, which uses the right hand system.
The force acting along the y-axis and the moment acting about the y-axis have their sign
reversed to obtain the measurements in the right hand system. The wind tunnel coordinate
system (coordinate system T) is the force balance coordinate system (coordinate system F)
transformed to comply with the right hand sign convention (see also appendix C.1). The x-
axis is aligned with the centreline of the yacht and values are positive in the forward direction
as shown in Figure 2.19. The y-axis is perpendicular to the x-axis and positive to port. Both
the x and y-axis lie in the horizontal plane. The z-axis is perpendicular to the x and y-axis and
positive upwards. The origin of coordinate system T is the origin of the force balance as
shown in Figure 2.18.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 33
Chapter 2 Wind Tunnel Testing of Yacht Sails
V
Ay
V
Ax
< 0
(here -ve)
V
A
V
S

X
T
Y
T
X
T
Y
T
L
D
X
Y
Plan view at
m
= 0
T
B
0
B
T
H
B Body fixed coordinate system origin
B
0
Body fixed coordinate system origin at
m
= 0
T Tunnel coordinate system origin
H Heel axis of model in tunnel
Z
Y
Z
T
Y
T

m
< 0
(here ve)
B
0

Figure 2.19: Definition of the coordinate systems and sign convention
T T
The forces F and moments M =(F , F
T XT YT
, F =(M , M
ZT
)
T XT YT
, M )
ZT
measured in the wind
tunnel in coordinate system T can be transformed to the body fixed coordinate system of the
yacht (coordinate system B) by rotating through the heel angle of the model in the tunnel
(
m
) with the rotational transformation matrix defined by equation (2.17), where the pitch
angle of the model in the wind tunnel ( BT
m
) is zero, and translating through , the distance
between the coordinate systems origins, to obtain F=(F
T T
and M=(M
X
, F , F
Y Z
)
X
, M , M ) from
Y Z
T m m AB
F T F ) 0 , ( = = ,
(2.34)
( )
T T m m AB
F BT M T M + = = ) 0 , ( .
(2.35)
The distance BT depends on the location of the heel axis of the model in the wind tunnel and

m
. The model is heeled about a fixed axis aligned with the x-axis in the wind tunnel. The
distance in model scale measured in coordinate system T is calculated with BT
( ) TH TH T B T BT + = =
0
) 0 , (
m m BA
,
(2.36)
where TH is the model scale distance in coordinate system T from the tunnel coordinate
system origin to the heel axis, T B
0
is the model scale distance in coordinate system T from
the body fixed coordinate system origin at
m
=0 to the tunnel coordinate system origin as
shown in Figure 2.19, and T
BA
is the rotational transformation matrix to transfer from
34
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.4 Twisted Flow Wind Tunnel
coordinate system B to coordinate system A obtained by transposing T
AB
defined in equation
(2.17) so that
.
T
AB BA
T T =
(2.37)
The forces and moments, which are now in the body fixed coordinate system, are best
expressed as force and moment coefficient vectors (C
F
and C ) based on equations (2.1) and
M
(2.4) with
F C
Sm effm
F
A q
1
= ,
(2.38)
M C
5 . 1
1
Sm effm
M
A q
= ,
(2.39)
where A
Sm
is the reference sail area of the model and q
effm
is the effective dynamic pressure in
the wind tunnel. Substituting ) allows calculating q (2.2) into (2.12
effm
from
, ) sin sin 1 (
2 2
m Am Am effm
q q =
(2.40)
where q is the dynamic pressure at the wind tunnel test section and
Am Am
is the apparent wind
angle of the model in the wind tunnel. If q
Am
is not constant with height it needs to be
obtained at the geometric centre of sail area height of the model above the tunnel floor (z
m
),
which depends on . The reference pressure (q
m
as shown in equation (2.24)
ref
) is usually
measured by a pitot tube upstream of the test section and the flow twisting vanes. The
pressure factor (c ) between the pitot tube position and test section at height z
q m
is usually
determined by measurements without the model in the tunnel so that
.
ref q Am
q c q =
(2.41)
If the model is tested at different heel angles and the tunnel flow has a vertical speed gradient
c is determined for every
q m
, which is more appropriate than using equation (2.20) to model
the change in flow speed with height since it is difficult to achieve an exactly logarithmic
speed profile in the wind tunnel. c
q
also depends on the vertical speed gradient, the twist
profile, the position of the pitot tube and possibly to a small degree on the tunnel speed and
other factors so that it should be determined or checked before every wind tunnel test session.
If the onset flow in the tunnel is twisted, in equation (2.40)
Am
is also dependent on height
and needs be measured at z due to the reduction in z
m
. The changes in
Am m
when heeling, are
small however in relation to the accuracy of twisting the flow, so
Am
is usually defined at the
upright geometric centre of sail area (z
CoA
).
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 35
Chapter 2 Wind Tunnel Testing of Yacht Sails
From C
F
the lift and drag coefficients in the heeled z-plane (coordinate system B) can now be
calculated from
,

Fy
Fx
eff eff
eff eff
L
D
C
C
C
C


cos sin
sin cos
(2.42)
where
eff
is calculated with equation (2.11).
Jackson (2001) presents an alternative approximate method to determine the forces on the
sails of a heeled yacht without performing a complete coordinate system transformation. It is
assumed that the lift and drag coefficients in the horizontal plane (coordinate system A) (C
LA

and C
DA
) are known. C
LA
then represents the useful component of the lift but in the effective
angle theory the total C
L
in the heeled plane is required, which in this case is obtained from
C
LA
/cos. This basic approach however makes the approximation that C
LA
is aligned with the
y-axis, which is only true for
A
of 0 and 180. In addition this approach obtains the C
L
in
the heeled plane only from the horizontal C
LA
and ignores the vertical force component. This
is only correct if the force acting along the z-axis in coordinate system B (F
Z
) is zero.
However, if F is not equal to zero then obtaining C simply from C
Z L LA
/cos is not physically
correct. While this method might be convenient in some situations, the force and moment
transformation discussed before is used in this work to obtain the lift and drag coefficients in
the heeled plane.
2.4.2 Scaling Effects
In addition to modelling a geometrically similar sail shape it is important to correctly model
the physical behaviour of the sails and the flow at model scale. A number of non-dimensional
ratios need to be considered. Ideally the Reynolds Number (Re) in full-scale and model scale
should be the same in order to ensure the same viscous flow behaviour. Re is defined as

Vl
= Re ,
(2.43)
where V is the speed of the flow, l is the reference length, the density and is the viscosity
of the fluid. The reference length reduces with the factor of the model scale. If the Re is to be
kept constant the speed in the tunnel must be increased or the kinematic viscosity (=/)
reduced. Significantly changing the kinematic viscosity in the wind tunnel is not practical.
The ability to increase the flow speed in the wind tunnel is also limited due to the material
strength of the soft sails so that the full-scale Reynolds Number cannot be matched in the
wind tunnel. Research has been conducted by Hawkins (1998) to investigate the sensitivity of
wind tunnel sail force measurements to changes in Reynolds Number. The measurements
36
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.4 Twisted Flow Wind Tunnel
showed little dependency on Reynolds Number for Re>610
5
, which is the likely limit of the
boundary layer transition flow regime. It is therefore assumed that not matching the full-scale
Reynolds Number for wind tunnel tests at Re>610
5
is not of first order significance but a
comparison to full-scale measurements is required to investigate this further.
A number of non-dimensional ratios are concerned with the physical properties of the sails.
The membrane mass ratio is only important when inertia forces are significant. Wind tunnel
tests are usually only concerned with static loading and no dynamic effects are considered so
that the membrane mass ratio does not need to be matched. The membrane elasticity ratio
allows for aero-elastic effects of the sailcloth. For most testing conditions sail stretch is small
and a comparable membrane elasticity ratio can be achieved by selecting the material
properties of the model sail to reflect the changes in flow speed between the wind tunnel and
full-scale. However, some full-scale spinnakers stretch more than is practical to model in the
wind tunnel. The sail shape is assumed to remain geometrically similar when stretching so
that the effect can be accounted for by a change in reference area.
The membrane fold height ratio is a measure of the sailcloth stiffness. It is defined simply as
the membrane fold height divided by the reference length as shown in Figure 2.20. The
sailcloth needs a sufficient amount of flexibility to allow the sail to adopt the flying shape
prescribed by the pressure distribution. The fold height ratio does not need to be matched
exactly to the full-scale sail as long as the model sail is flexible enough to maintain its natural
shape, although the size of wrinkles in the model scale may be larger than at full-scale.
The weight/pressure ratio is the ratio of the gravitational downward force due to the sail
weight and the pressure force of the wind. It is easy to imagine that a heavy spinnaker in a
light breeze will sag or collapse giving a very different flying shape to a fully inflated sail.
The research by Hawkins (1998) showed that the weight/pressure ratio should be within
certain limits but does not need to be matched exactly. It needs to be ensured that the
spinnaker inflates properly but also that it does not over-inflate. It is therefore important to
have experienced trimmers present during testing to ensure that the flying shape in the wind
tunnel resembles the flying shape on the water.
Wind tunnel testing is often used for comparative analysis so that the accurate scaling of the
results to full-scale is not strictly necessary. This makes the discussed approximations less
significant. One can be confident that a sail that performs better than another sail in the wind
Membrane fold height
Reference length

Figure 2.20: Membrane fold height ratio
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 37
Chapter 2 Wind Tunnel Testing of Yacht Sails
tunnel will also perform better on the water. How well the wind tunnel measurements relate
to full-scale is investigated as part of this project in chapter 6.
2.5 Experimental Set-Up
The wind tunnel tests for this project were carried out in the TFWT with a scale model of
Dyna. Firstly a wind tunnel model of Dyna was designed and built as described in section
2.5.1. The wind tunnel model sits in the trough of the turntable so that the waterline coincides
with the wind tunnel floor. As discussed in appendix B.3 cardboard cutouts were used
throughout this project instead of water to minimise the gap between the wind tunnel floor
and the model. Before conducting any experiments with the wind tunnel model the vertical
speed gradient and twist profiles corresponding to the performance and size of Dyna were
developed as shown in section 2.5.2. Based on the extensive experience of testing yacht sails
at the TFWT the tunnel speed is selected. A higher wind speed gives a better force resolution
but it should be limited so that the sails do not over-inflate or stretch unrealistically and
behave naturally. This results in different tunnel speeds being used for testing in upwind and
downwind configuration. With the upwind set-up the tunnel speed is set at 26Hz which
corresponds to approximately 10Pa or 4.1m/s. In downwind mode the tunnel runs at 18.2Hz
which corresponds to about 4.8Pa or 2.8m/s. The sampling period used for the force
measurements is based on previous studies of the effects of sampling period on repeatability.
For measurements with sails present the sampling period is 90 seconds and for measurements
without sails it is 60 seconds. Due to the complex flow pattern around sails and the
deformation of soft sails, periodic fluctuations are more pronounced and a longer sampling
period is required to ensure good repeatability.
2.5.1 Wind Tunnel Model
th
A 15% (1/6.67 ) scale model of Dyna was designed using the solid modelling computer
package ProEngineer
5
(Figure 2.21) and built comprising sails, rig, load carrying frame, deck
and hull. A cradle connects the model to the force balance and allows adjustment of the heel
angle (). The model can be rotated about a fixed axis on its centreline and 115mm above the
design waterline to approximate the heeling behaviour of the yacht. The model can be heeled
to any angle up to approximately 50. A similar concept to Dyna with a load carrying internal
frame and a detachable hull and deck shell was chosen as shown in Figure 2.22 to enable
measuring the rig and hull loads separately. This will be discussed further in chapter 3. The

5
Software developed by Parametric Technology Corp. (PTC), 140 Kendrick Street, Needham, MA 02494, USA
38
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.5 Experimental Set-Up
Figure 2.21: Computer representation of wind
tunnel model of Dyna
Figure 2.22: Internal frame of model with remote
control winches
model is fitted with six winches to trim the sails remotely. Depending on the type of testing
different sail and rig controls are adjusted with the winches.
The model sails were built from the same geometrical descriptions as the full-scale sails by
North Sails New Zealand
6
. The desirable properties are also similar, lightness with minimal
stretch but the elastic scaling is not accounted for. The mainsail, jib (100%) and genoa were
built from 9933 Polyester, and the symmetrical spinnaker of Contender Superlite (polyester).
The design, construction and set-up of the model, rig and sails are described in more detail in
appendix A.
2.5.2 Vertical Speed Gradient and Twist Profiles
To replicate full-scale sailing conditions, the sails are not tested in uniform flow. As
discussed in section 2.3.2 the speed and direction of the onset flow along the span of sails
changes. From equation (2.20) the changes in speed and direction with height can be
calculated based on , V and V
T T S
. Since it is not feasible to simulate the speed and twist
profiles for every possible sailing condition some assumptions need to be made. The target
profiles are calculated for a true wind speed of 7m/s and the boat speed of Dyna is obtained
from the full-scale measurements and VPP calculations by Hochkirch (2000). With equation
the terrain roughness length (z (2.21)
0
) is determined as 0.25mm. An upwind configuration
with mainsail and jib and a downwind configuration with mainsail and spinnaker are
investigated.
In Figure 2.23 the target profiles and the profiles developed in the wind tunnel are shown
plotted against the ratio of mast height (z/z
mast
) on the vertical axis. The geometric centre of

6
North Sails New Zealand Limited, 117 Pakenham Street West, Viaduct Basin, Auckland Central, New Zealand
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 39
Chapter 2 Wind Tunnel Testing of Yacht Sails
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1 1.2
V
A
/V
Aref
[-]
z
/
z
m
a
s
t
[
-
]
TWA 30
TWA 40
TWA 60
TWA 80
TWA 100
Empty tunnel
Upwind profile
0
0.2
0.4
0.6
0.8
1
1.2
-15 -10 -5 0 5

A
[]
z
/
z
m
a
s
t
[
-
]
TWA 30
TWA 40
TWA 60
TWA 80
TWA 100
Upwind profile
(a) Vertical speed gradient for upwind conditions
(b) (twist) for upwind conditions
A


0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1 1.2
V
A
/V
Aref
[-]
z
/
z
m
a
s
t
[
-
]
TWA 110
TWA 120
TWA 140
TWA 160
TWA 180
Downwind profile
0
0.2
0.4
0.6
0.8
1
1.2
-15 -10 -5 0 5

A
[]
z
/
z
m
a
s
t
[
-
]
TWA 110
TWA 120
TWA 140
TWA 160
TWA 180
Downwind profile
(c) Vertical speed gradient for downwind conditions
(twist) for downwind conditions (d)
A
Figure 2.23: Vertical speed gradient and twist profiles for upwind and downwind conditions calculated
for Dyna and modelled in wind tunnel plotted against ratio of mast height on vertical axis
sail area is taken as the reference height. The speed and twist profiles in the tunnel are
developed by using known profiles and investigating the effects of small changes in the set-
up until acceptable profiles are obtained. The wind speed and direction are measured by a
four-hole Cobra Probe (see also section 6.3) at different heights above the centre of the
turntable and 150mm to either side with the tunnel running at the speeds to be used for the
model tests. Since the twist profile is independent of the speed profile it is usually developed
first.
From the target twist profiles in Figure 2.23(b) it can be seen that the twist angles for the
upwind configuration are relatively small. For true wind angle between 30 and 100 that
might be of interest for the upwind configuration, the change in apparent wind angle over the
span of the sails ranges from 2.2 to 8.5. It should be remembered that the boom is at about
17% of the mast height and the bottom of the headsail at 9% of the mast height. In the upwind
40
YACHT
RESEARCH
UNIT
The
University
of Auckland
2.6 Summary
configuration courses close to the wind at
T
of around 40 are generally of most interest and
the twist is very small.
It was therefore decided not to conduct these tests in twisted flow since the disadvantages of
having the vanes in the flow exceed the advantages for small twist angles since the wake
behind the vanes introduces unevenly distributed turbulence in the flow. It is also seen as
unjustifiable to use a twist profile at larger angles of the upwind configuration since the
changes are still relatively small and altering the profile makes the measurements less
consistent and more complicated. Therefore only one upwind and one downwind profile are
to be developed and used. Figure 2.23(a) shows the target speed profiles for the upwind
configuration which change very little with the true wind angle. The speed change over the
span of the sails is about 20%. From the empty tunnel speed profile, the boundary layer and
the close to constant flow speed above, can clearly be seen. To approximate the target profiles
more closely, horizontal bars of different widths are placed in the tunnel downstream of the
screens at different heights above each other as shown in Figure 2.15 to create the vertical
speed gradient. Approximating this steep profile accurately at low heights is found to be
difficult with the current system of using wooden planks. The best profile that has been
achieved and is used for all model tests in upwind configuration is shown in Figure 2.23(a).
For the selected profile Figure 2.23(b) shows that the fluctuations in
A
without flow twisting
vanes are within 0.3 over the span of the sails.
The target twist profiles for the downwind configuration with mainsail and spinnaker are
shown in Figure 2.23(d). Compared to the upwind case the twist angles are larger and the
vanes are used to create a twist profile. For true wind angles from 110 to 160, the twist
angle does not vary much. However, there is a significant variation between 160 and 180,
where twist is zero. Although model measurements are conducted at 180 it is not a common
sailing condition and only one twist profile was developed based on the target profiles
between
T
of 110 and 160. The resulting profile twists the flow through approximately
9.5 over the span of the sails and matches the targets very well. Figure 2.23(c) shows the
target vertical speed profiles for the downwind sailing conditions and the profile developed
by carefully positioning horizontal bars in the tunnel as described before. The achieved
profile generally fits the targets better and is smoother than the steeper upwind profile. The
speed change over the span of the sails is about 26%.
2.6 Summary
In this chapter the most important concepts relating to sail aerodynamics and the associated
wind tunnel testing have been presented. The concepts and techniques introduced in this
chapter form the basis of the work conducted for this project.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 41
Chapter 2 Wind Tunnel Testing of Yacht Sails
The forces and moments are usually presented in terms of lift and drag coefficients and the
centre of effort position. The drag coefficient is often separated into its different components
for analysis and modelling. Different methods can be used to determine the centre of effort
position. Due to the heel angle of the yacht it is common to represent the forces and moments
in the body fixed coordinate system of the yacht, which heels (and pitches) with the yacht,
and the forces measured in the wind tunnel need to be transformed.
To account for the flow changes past the sails when the yacht heels the effective angle theory
is used. The sails are assumed to be influenced only by the flow perpendicular to the mast and
the effective angle and effective wind speed in the plane perpendicular to the mast are used as
reference parameters.
It was also discussed that the onset flow onto sails is not uniform but changes in speed and
direction with height above the water due to the planetary boundary layer profile in the true
wind and the motion of the yacht. The change in speed and direction depends on the sailing
condition and the yacht type and size. This complex flow can be modelled in the TFWT and
vertical speed and twist profiles have been developed for the wind tunnel tests of Dyna. A
wind tunnel model has been designed and built to allow the hull/deck forces to be measured
separately from the rig and sail forces.
42
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
3 Hull and Sail Force Interaction
3.1 Introduction
It is known that the presence of the hull/deck influences the flow around the sails as it
partially closes the gap between the sails and the water surface. In wind tunnel tests and CFD
simulations the hull/deck is therefore usually modelled. As a result wind tunnel
measurements of sails include the forces acting in the hull/deck and rig, also called windage.
It is common practice to measure the bare poles windage on the hull/deck and rig without
sails in the wind tunnel and subtract it from the measurements with the sails to obtain the sail
forces. In VPPs the windage is calculated with empirical expressions based on the projected
area of the component or from wind tunnel measurements.
In the same way as the sail forces are influenced by the hull/deck it can be argued that the
hull/deck forces should be influenced by the presence of the sails as the sails change the flow
field around the hull/deck. Research by Flodn and Johansson (2001) and Hansen et al.
(2002) has indicated that this might be the case. The windage obtained with the bare poles
method therefore may not represent the forces produced by the hull/deck when the sails are
present. With the bare poles method the sail forces obtained in the wind tunnel include the
force component of the difference between bare poles windage and the windage with the sails
present. For the VPP calculations this is positive since the difference would otherwise be
neglected. However, when comparing wind tunnel measurements to sail forces obtained in
other ways inconsistencies can occur. Sailing dynamometers like Dyna measure the forces of
rig and sails so that the difference between bare poles windage and windage with the sails
present is not included. Similarly sail forces in CFD simulations are usually calculated from
the forces acting on the sails and do not include this difference either. For an accurate
comparison of wind tunnel measurements to full-scale data the forces need to be obtained in
the equivalent way. Measuring the sail and rig forces in the wind tunnel would be possible
but the complexity of the experiment is not significantly increased if forces on the hull/deck
are also investigated instead of just ignoring them. Assessing the difference between the
hull/deck forces with and without the sails provides vital information on how to compare
wind tunnel, full-scale and CFD data and how to model the hull/deck forces in VPPs.
Traditionally only the bare poles windage drag is accounted for in the VPP formulation but
here the lift and the force perpendicular to the deck plane are also considered. Wind tunnel
tests by Flodn and Johansson (2001) and Hansen et al. (2002) have shown the presence of a
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 43
Chapter 3 Hull and Sail Force Interaction
force component perpendicular to the deck plane of the yacht (F
Z
). This force component is
not accounted for in VPPs at present, but depending on the magnitude could influence the
performance of the yacht. Previous work by Hochkirch (2000) did not show a significant
force perpendicular to the deck plane in full-scale measurements of the rig and sail forces.
Since the full-scale measurements did not include the hull/deck component the larger F
Z

measured in the wind tunnel could be attributed to the hull/deck. This can also be investigated
in detail by measuring the hull/deck forces in the wind tunnel.
The forces acting on the hull/deck were obtained by two methods in this investigation. Firstly
a secondary force balance inside the model was used to measure them. This is associated with
a few challenges and as a confirmation the forces were also obtained from surface pressure
measurements. The surface pressure distribution also contributes to understanding the flow
field around the hull/deck. Measurements with both methods were conducted for a series of
apparent wind angles for the model without sails, and with the tunnel in upwind and
downwind configuration running at the appropriate speeds as discussed in section 2.5. The
tests in upwind configuration were carried out at 0, 12.5 and 25 heel, whereas the tests in
downwind configuration were only carried out at 0 heel. Measurements with both methods
were also conducted for the model with mainsail and jib or genoa in upwind configuration at
the three heel angles and for the model with mainsail and spinnaker in downwind
configuration at 0 heel. The sails are trimmed to maximise the drive force for each test
condition and the trim settings are recorded so that the same trims are used for both methods.
For consistency the same trims are also used in the fully powered-up measurements presented
in chapters 5 and 6. The measurements for mainsail with genoa show the same trends as
mainsail with jib but are not included here. Apart from confirming the trends observed for
mainsail with jib and showing that the larger sail area increases the values, no additional
information would be gained from showing them. Nevertheless, it was important to conduct
the mainsail with genoa measurements to improve confidence in the mainsail with jib results
as well as the testing procedures.
3.2 Force Measurements of Hull and Deck Forces
In order for the hull/deck forces to be measured separately, the wind tunnel model was
constructed with an internal frame that carries the rig loads and an independent hull/deck
shell as shown in Figure 3.1. The internal frame is connected to the main wind tunnel force
balance through a cradle as shown in Figure 2.21. The model can be set up so that the
hull/deck shell is connected to the frame only through a secondary force balance installed
inside the model as illustrated in Figure 3.2. Figure 3.3 shows the mechanism that is used at
the bow to connect or disconnect the hull/deck from the internal frame. If the plunger is taken
out, there is a gap between the rod holding the internal frame and the tube of the hull/deck
44
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.2 Force Measurements of Hull and Deck Forces

(a) Internal frame and round
secondary force balance
(b) Hull shell and connector piece (c) Hull and frame connected
through force balance
Figure 3.1: Wind tunnel model with internal six-component force balance and connection showing
internal framework structure and hull shell detached and assembled
shell. A second such mechanism is used at the stern. With this system the frame and hull/deck
can be heeled in the connected and detached state.
Since the secondary force balance is attached to the internal frame it rotates with the model
when it is heeled and the forces are measured in the body fixed coordinate system (coordinate
system B) and the moments only need to be translated by the distance between the secondary
force balance origin and the origin of coordinate system B to obtain them in the body fixed
coordinate system. A high frequency force balance by JR3
7
was used. It consists of eight load
cells, two each in the F
X
and F direction and four in the F
Y Z
direction. The force balance is
supplied with a control box with six voltages as outputs. The voltages were recorded by an
Force balance
Connector
Socket on hull
Plunger
Figure 3.3: Mechanism to connect or disconnect
hull/deck from internal frame

Figure 3.2: Hull/deck shell of model connected to
internal frame through six-
component force balance
7
Force balance model 30E12A-I40 8N1.5S by JR3 Inc., 22 Harter Avenue, Woolland, CA 95776, USA
www.jr3.com
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 45
Chapter 3 Hull and Sail Force Interaction
analogue to digital (A/D) card with the measurement acquisition software LabVIEW by
National Instruments
8
and converted to the three force and moment components with the
calibration matrix supplied by JR3. A calibration check without the hull/deck attached
showed a mean error of 1.15% of the measured value in the x-direction, 0.61% in the y-
direction and 0.51% in the z-direction. As for measurements taken with the wind tunnel
force balance, the sampling period per run is 90 seconds for measurements with sails present
and 60 seconds for measurements without sails (see section 2.5).
The measurement range in the F
X
and F direction is 8N and 16N in the F
Y Z
direction. The
measurement range can be offset to compensate the weight of the hull/deck shell and the
moment about the x-axis when the model is heeled. The hull/deck shell is not permitted to
touch any part of the frame or rigging and sufficiently large clearances need to be allowed
everywhere for the hull/deck shell to deflect and distort when the wind pressure is applied.
Initially the hull/deck shell was not stiff enough to achieve this, especially when the model
was heeled, and two stringers and two frames were added which can be seen in Figure 3.1(b).
The attachment point of the secondary force balance on the frame was chosen to minimise the
deflection at the mounting plate. As a result the attachment point does not coincide with the
centre of gravity of the hull/deck shell. To minimise the initial moment about the y-axis and
the deflection, small weights were attached to the hull/deck shell to bring the centre of gravity
closer to the force balance attachment point. With this set-up the force balance range was
sufficient for all measurements in the downwind configuration. In the upwind configuration,
where the tunnel speed is higher than for the downwind set-up, the yawing moment (M
Z
)
range of 0.18Nm was exceeded in some cases. In these cases a side force producing a yawing


Figure 3.4: Application of load for offset zero
measurement of M

Z
8
National Instruments Corporation, 11500 N Mopac Expwy, Austin, TX 78759-3504, USA
46
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.3 Pressure Measurements
moment was applied as shown in Figure 3.4 and measured. The measurement range was then
offset and a zero force measurement was made with the side force applied. The wind tunnel
measurements were made without the extra side force applied and during the post-processing
of the data in MatLab
9
the extra side force was added to the wind tunnel measurements to
yield the forces and moments acting on the hull/deck shell.
Since the hull/deck shell and the mounting surface of the force balance on the frame are still
not very stiff, the calibration of the force balance was affected and the measurements of F
Y

and F have been corrected as described in appendix B.1
Z
. With the correction the mean error
in the y-direction is reduced to 0.95% of the measured value and 0.85% in the z-direction.
Great care needed to be taken during the experiments to ensure that the hull/deck shell did not
touch the internal frame or any rigging component. The forces and moments in coordinate
system B can be expressed in coefficient form with equations (2.38) and (2.39) where the
reference area is taken as the reference sail area and the reference pressure is determined at
the height of the centre of sail area so that the coefficients can be directly compared to the sail
force coefficients. The lift and drag coefficients can then be determined from equation (2.42).
3.3 Pressure Measurements
To confirm the hull/deck force measurements, surface pressure measurements on the hull and
deck were also conducted. These measurements are furthermore useful for gaining a better
understanding of the flow around the hull and deck. The concept of the Dyna model with the
internal structural frame and the hull/deck shell makes this feasible. A second hull/deck shell
was built, fitted with pressure taps and assembled around the internal frame.
A Scani-valve system, in conjunction with Setra pressure transducers (model 264 with inch
of water range) with a range of 128Pa, was employed in order to measure the pressures
accurately at the low wind tunnel speeds used for testing soft sails. An accuracy of typically
better than 1.5% of the measured value was achieved with the pressure transducers. The
Scani-valve system allows one pressure transducer to sequentially measure a number of
different taps. Hence fewer pressure transducers are required but the measurement process
takes longer and the dynamic flow behaviour cannot be investigated. A bank of six pressure
transducers, four for measuring the Scani-valve channels and two for measuring reference
pressures, is set up as shown in Figure 3.5(a) and is installed in the framework structure of the
model as can be seen in Figure 3.5(b). Figure 3.5(c) shows the Scani-valve system installed in
the hull/deck shell. The hull and deck were fitted with 188 pressure taps, the maximum
number supported by the Scani-valve system with four channels. The models design with its

9
Software developed by The MathWorks, Inc., 3 Apple Hill Drive, Natick, MA 01760-2098, USA
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 47
Chapter 3 Hull and Sail Force Interaction

(a) Bank of six Setra pressure transducers (b) Pressure transducers installed in frame (c) Scani-valve system
Figure 3.5: Components of pressure tapped model
internal frame and the hull/deck shell allows for all the tubing, Scani-valve system and the
pressure transducers to be installed inside the model as shown in Figure 3.6. This is
advantageous because the test equipment does not influence the flow field and only one
data/power cable for the pressure transducers and two tubes for the reference dynamic and
static pressure come from the model, which can influence the force measurements due to their
stiffness. The forces and moments acting on the model are measured as usual with the wind
tunnel force balance to ensure that the same set-up and sail trim is achieved as in other test
sessions. Theoretically the hull pressure could have been measured at the same time as the
hull/deck forces, which ideally is desirable to ensure that data is collected in the same
conditions, but in practice it is deemed as too complicated and the pressure and force
measurements on the hull/deck are conducted separately. The closed model with the pressure
tapped hull/deck shell is shown in Figure 3.7.
The location of the pressure taps and the numbering is shown in Figure 3.8. The pressure taps
are primarily arranged in nine cross-sectional rings along the length of the model. The taps
are consecutively numbered in rings starting at the bow. The first tap of each ring is the one
on the port edge of the deck and the last one is the tap closest to the sheerline on the port side
of the hull. The tap spacing is closest at the edge between the hull and the deck (sheerline)
Figure 3.6: Inside picture of pressure tapped
model
Figure 3.7: Outside picture of pressure tapped
model
48
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.3 Pressure Measurements
Figure 3.8: Location and numbering sequence of pressure taps on hull and deck of wind tunnel model
and increases further away from the sheerline to capture the peak pressures that occur close to
the edges. The tap spacing varies depending on the geometry but in most cases the first tap on
the deck is 10mm from the sheerline, the second 30mm and the third 60mm. Similarly the
distance of the taps on the hull from the sheerline is 15mm, 45mm, 90mm and 150mm. For
taps further away from the sheerline the spacing varies to fit the hull and deck geometry.
Pressure taps are also installed below the DWL so that the influence of the flow in the wind
tunnel trough can be determined and to ensure that pressure taps are still suitably distributed
when the model is heeled. The taps are symmetrical about the centreline of the yacht with the
exception of a few missing taps to starboard of the coachroof where the headsail track is
located.
A series of measurements for the sail combinations and range of apparent wind angles and
heel angles as described in section 3.1 was conducted. An acquire time at each pressure tap of
40 seconds was determined to be sufficient from a sensitivity study. This gives a run time of
about 33 minutes to acquire the pressures on all taps. It has been investigated that the zero
drift of the pressure transducers is not significant over this time period under measurement
conditions so that the whole model is mapped without taking zero measurements in-between.
The small amount of drift is assumed to be linear over time and determined for each pressure
transducer by taking a final zero measurement after the whole model is mapped. The pressure
measurements are corrected by the drift and expressed in coefficient form relative to the
reference pressure at the height of the centre of sail area determined from equation (2.41).
The surface pressures on the hull and deck were interpolated from the measured pressure
coefficients using the technical visualisation software TecPlot
10
. In TecPlot the hull and deck
geometry is described by a triangular boundary layer mesh as shown in Figure 3.9 which was

10
Software developed by Tecplot, Inc., 13920 Southeast Eastgate Way, Suite220, Bellevue, Washington,
98005, USA
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 49
Chapter 3 Hull and Sail Force Interaction
Figure 3.9: Computer model for mapping the measured surface pressures with the mesh consisting of
7856 elements and divided into five regions
generated from the three dimensional hull and deck model in ProEngineer. The pressure
coefficient at each node of the boundary mesh was then calculated by TecPlot from the
pressure coefficients at the 188 taps using the universal kriging algorithm described by Davis
(2002), which has been developed for spatial interpolation of geological data and usually
provides superior results than basic inverse distance methods. The mesh is generated with
each pressure tap coinciding with a node so that the measured values form part of the
interpolated solution. If this is not the case the interpolation smoothes out the measured peak
pressure coefficients. The interpolated pressure distributions are plotted in the following
sections to understand the flow pattern but TecPlot cannot calculate the resulting forces and
moments.
The interpolated pressure coefficient at each node, together with the node coordinates and a
list of the nodes constituting each element were exported, and MatLab was used to calculate
the forces and moments from this information. For each triangular element of the boundary
mesh the coordinates of the three nodes (A, B and C) and the pressure coefficient at each
node (C
PA
, C
PB
and C ) are known. Figure 3.10
PC
shows element i of the boundary mesh with
the three nodes and pressure coefficients. A constant pressure coefficient is assumed to act
over the area of the element at the centre of area (CoA). The pressure coefficient of element i
(C
Pi
) is calculated with an inverse distance method from the pressure coefficient at each node
of the element (C , C and C
PAi PBi PCi
) and the distances between the centre of area and the
nodes ( , and ) with
i i
C CoA
i i
A CoA
i i
B CoA
exp exp exp
exp exp exp


+ +
+ +
=
i i i i i i
PCi i i PBi i i PAi i i
Pi
C C C
C
C CoA B CoA A CoA
C CoA B CoA A CoA
,
(3.1)
50
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.3 Pressure Measurements
A
i
C
PAi
B
i
C
PBi
C
i
C
PCi
CoA
i
C
Pi
N
i

Figure 3.10: Triangular element i of the boundary mesh
where CoA of triangle i is calculated from CoA=(A+B+C
i i i
B
i i
)/3 and the exponent (exp) is the
weighting factor of the inverse distance method, which is taken here as one. The pressure
coefficient acts normal to the element surface so that the force and moment coefficient vector
can be obtained from C , the normal vector (N) and CoA, where the area of the element (A
Pi i i i
)
is related to the length of the normal vector by A =|N
i i
|/2. The sum of all elements divided by
the reference sail area of the model (A
Sm
) yields the total force and moment coefficient
vectors acting on the hull/deck. It also needs to be considered that C
Pi
is non-dimensionalised
by the apparent dynamic pressure but that the force and moment coefficients in coordinate
system B are relative to the effective dynamic pressure so that the relationship in equation
needs to be included. (2.40)
The force and moment coefficient vectors in coordinate system B are therefore obtained from
i
N
i
Pi
Sm A
F
C
A
N C

=

=
1
2 2
) sin sin 1 (
5 . 0

,
(3.2)
i i
N
i
Pi
Sm A
M
C
A
N CoA C

=

=1
5 . 1 2 2
) sin sin 1 (
5 . 0

,
(3.3)
where N is the vector normal to the plane of the triangle obtained from
i
. ( ) (
i i i i i
A B A C N = )
(3.4)
The lift and drag coefficients can then be determined from equation (2.42).
The mesh density was determined through the grid density sensitivity analysis in appendix
B.2 utilising the calculated coefficients. It showed that a mesh with approximately 8000
elements is sufficient. The final mesh chosen for the analysis of the wind tunnel
measurements consisted of 7856 elements and was divided into five independent regions as
shown in Figure 3.9 to avoid interpolation of the pressures around the most important sharp
edges of the geometry.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 51
Chapter 3 Hull and Sail Force Interaction
3.4 Comparison of Force and Pressure Measurements
Figure 3.11 shows the coefficient of drag (C
D
), lift (C
L
) and force perpendicular to the deck
plane (C
Fz
) of the hull/deck calculated from force and pressure measurements of the model
without sails in the upwind configuration. The reference sail area, obtained from equation
(2.3), and the effective wind speed at the height of the centre of sail area, obtained from
equation (2.12), are used to calculate the hull/deck force coefficients so that they can be
directly related to the sail force coefficients. Both the force and pressure measurements were
conducted with cardboard cutouts covering the turntable trough as discussed in appendix B.3
so that the experimental set-up is the same and the results are directly comparable. Figure
3.11 shows that the coefficients predicted from the pressure measurements are slightly larger
than the coefficients obtained from the force measurements but that the two methods produce
results that are very similar and show identical trends.
The cardboard does not completely close the gap between the tunnel floor and the model so
that some air flow under the hull of the model is possible. The force component resulting
from the flow under the hull is included in Figure 3.11. While force measurements with this
set-up always include this component, the forces and moments acting above the waterline can
be obtained from the pressure measurements by only summing up the mesh area above the
waterline. Figure 3.12 shows C , C
L D
and C
Fz
calculated from the pressure measurements on
the total hull/deck and the area above the waterline. C and C
L D
are not significantly
influenced by the component below the waterline but a noticeable downward force is
produced by the component below the waterline for
eff
smaller than 90 and an upward force
for angles larger than 90. At small
eff
the geometry creates a low-pressure region inside the
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0 20 40 60 80 100 120 140 160 180

eff
[]
C
o
e
f
f
i
c
i
e
n
t

[
-
]
CD (from forces)
CD (from pressures)
CL (from forces)
CL (from pressures)
CFz (from forces)
CFz (from pressures)
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0 20 40 60 80 100 120 140 160 180

eff
[]
C
o
e
f
f
i
c
i
e
n
t

[
-
]
CD above water
CD total
CL above water
CL total
CFz above water
CFz total

Figure 3.12: Comparison of C Figure 3.11: Comparison of C , C and C , C and C
D L Fz
of
hull/deck calculated from force and
pressure measurements for model
without sails in upwind configuration
at 0 heel
D L Fz
of total
and above water hull/deck from
pressure measurements for model
without sails in upwind configuration
at 0 heel
52
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.5 Lift and Drag Forces of Hull/Deck and Rig
trough by accelerating the flow close to the hull surface. At larger
eff
a high-pressure region
is created inside the trough by stagnating the flow. All hull/deck results which are influenced
by the vertical force component are therefore obtained from the pressure measurements above
the waterline and not from the force measurements directly.
3.5 Lift and Drag Forces of Hull/Deck and Rig
The windage forces on the hull/deck and rig are often estimated by conducting bare poles
measurements on the model without the sails present. In VPPs, only the windage drag is
currently estimated, usually separated into its two main components of hull/deck and rig drag.
The drag of the hull/deck is directly measured in the wind tunnel by the secondary internal
force balance or obtained from the hull/deck surface pressure measurements. The rig drag is
calculated by subtracting the hull/deck drag from the total drag of the model without sails,
measured by the wind tunnel force balance. Figure 3.13 shows the drag coefficients of
hull/deck and rig obtained from force measurements with the model upright in the upwind
configuration. In VPPs the effect of heel on windage is usually not considered and the drag is
calculated in the horizontal plane. In Figure 3.13 the drag is therefore also plotted in the
horizontal plane against the apparent wind angle (
A
). Both the hull and rig drag coefficients
have their maximum at
A
of 90, where the maximum hull drag coefficient is approximately
40% larger than the maximum rig drag coefficient. The changes in drag coefficient with
A

are significantly larger for the hull because of its elongated shape compared to most rig
components, which are circular in cross section.
Figure 3.13 also shows three methods of modelling the hull and rig drag. Richards et al.
(2006) show a model based on the maximum and minimum drag coefficient (C
Dmax
and C
Dmin
)
measured in the wind tunnel for hull and rig. The drag coefficient (C
D
) for hull or rig is then
given by
.
A Dmax A Dmin D
C C C
2 2
sin cos + =
(3.5)
Modelling the change in drag coefficient as a function of sine squared results in a good data
fit for the Volvo Ocean 60 (VO60) model used by Richards et al. (2006) and for windage data
of an IMS yacht model measured by Campbell (1997). The hull drag coefficients shown in
Figure 3.13 are also described well by this approach. The same is true for rig drag
coefficients up to
A
of 90; at larger
A
the rig drag is however under-predicted. For the
windage measurements the boom angle is adjusted for each
A
based on the recorded trim
setting so that the boom area projected onto the y-plane increases with
A
. As a result the rig
drag coefficient is not symmetrical about
A
of 90 but noticeably larger for
A
above 90.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 53
Chapter 3 Hull and Sail Force Interaction
The second method of modelling the windage drag shown in Figure 3.13 was used by
Hochkirch (2000) during of the Dyna project and was therefore implemented in the VPP
FRIENDSHIP-Equilibrium (see section 4.3.2). The windage drag is again described by a hull
and rig component. The drag coefficient of the rig (C ) is described by
Drig
S
A PyRig DyRig A PxRig DxRig
Drig
A
A C A C
C
sin cos +
= ,
(3.6)
where A is the area of the rig projected on the x-plane, A
PxRig PyRig
is the area of the rig
projected on the y-plane and A is the reference sail area defined in equation (2.3)
S
so that the
hull and rig drag coefficients are directly comparable to the sail force coefficients. C
DxRig
is
the drag coefficient of the rig in the x-direction and C
DyRig
is the drag coefficient of the rig in
the y-direction. In the calculations for Figure 3.13 following Hochkirch (2000) C
DxRig
is taken
from Claughton (1999) as 0.8 and C from Kerwin (1978) as 1.13. Figure 3.13
DyRig
shows that
the shape of the resulting curve is very different from the measurements. However, the
maximum and minimum magnitudes are comparable.
Hochkirch (2000) approximates the windage drag coefficient of the hull (C ) by
Dhull
S
A A PyHull A A PxHull
VA Dhull
A
A A
f a C
) sin( ) cos(
2
0
+ + +
= ,
(3.7)
where A
PxHull
is the hull area above the design waterline (DWL) projected onto the x-plane
and A
PyHull
is the hull area above the DWL projected onto the y-plane. For calculating the rig
drag coefficient it was feasible to use the apparent wind angle and speed at the height of the
centre of sail area as reference height. However, the hull is much closer to the water surface
so that the change in apparent wind speed between the height of the centre of sail area and the
0
0.05
0.1
0.15
0.2
0.25
0.3
0 20 40 60 80 100 120 140 160 180

A
[-]
C
D
A

[
-
]
Hull measured (upwind) Rig measured (upwind)
Hull Richards Rig Richards
Hull Hochkirch Rig Hochkirch
Hull van Oossanen Rig van Oossanen
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0 20 40 60 80 100 120 140 160 180

A
[-]
C
L
A

[
-
]
Hull measured (upwind) Rig measured (upwind)
Hull Richards Rig Richards
Figure 3.14: Windage lift coefficient in horizontal
plane with upright model in upwind
configuration
Figure 3.13: Windage drag coefficient in horizontal
plane with upright model in upwind
configuration
54
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.5 Lift and Drag Forces of Hull/Deck and Rig
reference height for the hull is specified by the factor f
VA
and the change in apparent wind
angle between the height of the centre of sail area and the reference height for the hull is
given by
A
. As shown by Claughton (1999) the IMS formulation uses 66% of the mean
freeboard height as reference height for the flow onto the hull. In wind engineering
applications it is common practice to normalise pressure and force coefficients with the free-
stream wind speed at the height of the top of the building since it has been shown that the
atmospheric boundary layer profile below the top of the building does not affect the pressure
distribution significantly. The general geometry and arrangement of a yacht hull on the water
is not too dissimilar to an elongated low rise building on the ground and the height of the
sheerline is taken as the reference height for the hull. For the calculations in Figure 3.13 f
VA
is
estimated from the wind tunnel upwind speed profile shown in Figure 2.23(a). The wind
speed measurement closest to the tunnel floor is taken at approximately the height of the
sheerline so that the value of f
VA
becomes 0.84. Taking a position below the sheerline as
reference height for the hull would require interpolating the wind tunnel data in Figure
2.23(a), which is another reason for using the height of the sheerline.
A
is zero since the
measurements were conducted with the tunnel in upwind configuration so that the flow
twisting vanes were not installed. In his work Hochkirch (2000) determined the constant a
0

from the full-scale measurements to be 0.764. The full-scale data was obtained with the sails
present so that the results will be discussed in section 3.5.1. Claughton (1999) suggests 0.68
for the hull drag coefficient normalised by the local flow characteristics, which is equivalent
to a .
0
and a Figure 3.13 shows that the hull drag coefficient curve calculated with equation (3.7)
0
of
0.68 has a very different shape to the wind tunnel measurements. However, the maximum
hull drag coefficient values are not too dissimilar. Measuring the vertical speed profile
accurately close to the floor of the wind tunnel is difficult and the repeatability of wind speed
measurements is not much better than 3% so that the accuracy of f
VA
is compromised. The
hull drag coefficient calculated with equation is very sensitive to changes in f (3.7)
VA
due to
the squared relationship and the value of f
VA
only needs to be changed from 0.84 to 0.87 to
obtain the same maximum hull drag coefficient as measured in the wind tunnel. Given the
limited accuracy of the experiment the agreement of the maximum modelled and measured
hull drag coefficient is good. The shape of the hull drag coefficient is however much better
modelled by a sine

squared relation as used by Richards et al. (2006).
The third method of modelling the windage drag shown in Figure 3.13 is given by van
Oossanen (1993). It also calculates the windage drag coefficients of rig and hull separately
and the equations can be rearranged into the form
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 55
Chapter 3 Hull and Sail Force Interaction
( )
S
A PxRig PyRig PxRig
Drig
A
A A A
C
sin ) ( 0 . 1 +
= ,
(3.8)
( )( )
S
A PxHull PyHull PxHull A
Dhull
A
A A A
C
sin ) ( sin 7 . 0 2 . 0 + +
= .
(3.9)
The rig drag coefficient is slightly over-predicted for small
A
and under-predicted by up to
approximately 20% at larger
A
. This is similar to the findings of Richards et al. (2006) for a
VO60 yacht. Richards et al. (2006) also found the hull windage drag coefficient to be
predicted well by the van Oossanen (1993) method whereas Figure 3.13 shows the hull drag
coefficient being significantly over-predicted. At
A
of 90 the drag is over-predicted by a
factor of 1.7. The measurements on the VO60 model presented by Richards et al. (2006) were
likely conducted with a close to constant vertical speed profile whereas the flow speed at the
sheerline of the hull is reduced by 16% for the measurements presented here. The fit of the
hull drag coefficient curve modelled by equation (3.9) to the VO60 and Dyna data is
therefore very different since it does not explicitly account for the reduction in wind speed
close to the water surface. Changes in the vertical speed profile do not only arise from
different testing procedures but also occur in real life for different yacht types as discussed in
detail in section 2.3.2. For a hull drag model to be accurate for different yacht types it is
hence important to explicitly model this wind speed reduction as done in equation (3.7). If the
f
VA
factor is incorporated in equation (3.9) the over-prediction in drag is reduced from 70% to
19% at
A
of 90.
Similarly, even simpler approaches to model windage drag are given by other authors such as
Larsson and Eliasson (1994), Claughton (1999) and Martin and Beck (2001). In addition to
drag the other windage components are also considered here. Figure 3.14 shows the lift
coefficient in the horizontal plane (C
LA
) generated by the hull and rig without the sails present
calculated from the same wind tunnel measurements as C
DA
in Figure 3.13. It can be seen that
the lift coefficient of the rig is quite small; the maximum is 15% of the maximum rig drag
coefficient. The lift coefficient of the hull is more significant with the maximum being 49%
of the hull drag coefficient. Of the windage models only Richards et al. (2006) suggest an
approximation for the lift component. The lift coefficient curves of rig and hull are calculated
from the maximum lift coefficient (C
Lmax
) measured in the wind tunnel with
A A Lmax L
C C cos sin 2 = .
(3.10)
The resulting lift coefficient curves for rig and hull are shown in Figure 3.14. It can be seen
that the sine curve representation fits the measurements reasonably well but it is also
noticeable that the hull lift is not close to zero at
A
of 90 but intersects the x-axis at
A
of
56
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.5 Lift and Drag Forces of Hull/Deck and Rig
around 80. Given the shape of the hull with a fine bow entry and the maximum beam well
aft of amidships this behaviour can be expected.
The rig and hull windage model by Richards et al. (2006) given in equations (3.5) and (3.10)
describes the wind tunnel measurements well but also requires maximum and minimum
coefficients as input parameters. Employing some of the concepts shown in equations (3.6)
and (3.7) the model can be made more generic by rewriting the equations as
S
A A Py Dy A A Px Dx
VA D
A
A C A C
f C
) ( sin ) ( cos
2 2
2
+ + +
= ,
(3.11)
S
A A A A Py Lmax
VA L
A
A C
f C
) cos( ) sin( 2
2
+ +
= .
(3.12)
The input parameters C , C and C
Dx Dy Lmax
are now relative to the projected area of the
component and the local flow conditions so that they are independent of size, proportions and
performance of the yacht and can be compared to other coefficients more readily. The height
of the centre of area of the rig is usually close to the height of the centre of sail area so that
f
VA
is one and
A
is zero. For the hull windage measurements discussed here,
A
is also
zero since the flow twisting vanes were not used and f
VA
is still 0.84. Table 3.1 shows C
Dx
,
C and C
Dy Lmax
for rig and hull which result in the same curves as the Richards et al. (2006)
method shown in Figure 3.13 and Figure 3.14. Based on aerodynamic data of basic geometric
shapes the coefficients seem reasonable. C
Dy
for the rig of 1.22 is 8% larger than the value of
1.13 suggested by Kerwin (1978). The current IMS formulation given by Claughton (1999)
uses 1.0 for circular rigging components. The drag coefficients for circular cylinders at sub-
critical Reynolds Numbers shown by Hoerner (1965) are between 1.1 and 1.2. Given that C
Dy

for the rig also includes the drag on the boom, which can be approximated as a flat plate of
aspect ratio 37 for which Simiu and Scanlan (1996) gives a drag coefficient of 1.6, the value
of 1.22 seems very applicable. C
Dx
for the rig of 0.91 is also very realistic considering the
drag coefficients of circular elements and the value of 0.8 suggested by Claughton (1999) for
mast sections.
For the hull drag coefficient Claughton (1999) suggests only one value of 0.68 which lies
between the values of 0.29 for C and 0.75 for C
Dx Dy
obtained here. Relating the hull to results
of basic geometries is more difficult due to its complex shape. An indication of the expected
Table 3.1: Hull and rig windage coefficients for
equations (3.11) and (3.12) obtained
from wind tunnel tests
C C [-] C [-] [-]
Dx Dy Lmax
Rig 0.91 1.22 0.19
Hull 0.29 0.75 0.36


Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 57
Chapter 3 Hull and Sail Force Interaction
C
Dx
is obtained from Hoerner (1965), who gives a drag coefficient of approximately 0.2 for a
cylindrical body in axial flow with a streamlined head. For a cube in atmospheric boundary
layer flow Hoerner (1965) shows a drag coefficient of 1.05 for the flow perpendicular onto a
face and 0.8 for the flow at 45. Pressure coefficients for a building of proportions more
similar to the hull (also obtained in atmospheric boundary layer flow) were presented by
Richards and Hoxey (1992) and Richards and Wanigaratne (1993). Pressure coefficients on
the windward side of the building were up to 0.5 and on the leeward side down to -0.25. The
pressure coefficients for the cube are 0.69 and -0.25 respectively. Figure 3.15 shows the
pressure distribution on the hull interpolated from the pressure measurements for the upright
model without sails and the flow 90 from port. The pressures shown are relative to the flow
speed at the height of the centre of sail area. Adjusted to the flow speed at the height of the
sheerline, the pressure coefficients close to the sheerline in the centre of the model are
determined as 0.75 on the windward and -0.43 on the leeward side. Although the pressure
coefficients are higher than for the cube, the drag coefficient is lower because the positive
pressure on the windward side of the hull reduces towards the bow whereas the building
pressures show little change across the windward face. The free-stream flow is not
perpendicular to the hull surface in the bow and stern regions which reduces the pressure
coefficient. Figure 3.15 also shows that there is no significant pressure differential across the
coachroof at this
A
since it is in the region of separated flow. The coachroof hence does not
contribute to the hull drag and if it is not included in the projected area calculations C
Dy
of the
hull increases from 0.75 to 0.83. Although the shapes are very different the comparison
indicates that the measured C and C are realistic.
Dx Dy
Table 3.1 also shows that C
Lmax
of the rig is small, at 0.19, as expected, since only the mast
and boom produce some lift. C
Lmax
of the hull is determined as 0.36 which agrees well with
the maximum lift coefficient of a circular cylinder inclined against the direction of flow,

Figure 3.15: Surface pressure plots for model without sails on port tack at
A
of 90 and 0 heel showing
leeward and windward side of hull
58
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.5 Lift and Drag Forces of Hull/Deck and Rig
which Hoerner (1965) gives as 0.4. It can be concluded that the coefficients C , C
Dx Dy
and
C
Lmax
obtained from the bare poles measurements (Table 3.1) relate well to coefficients of
geometrically similar standard shapes and that the bare poles windage drag and lift can be
modelled accurately with equations (3.11) and (3.12). Having reviewed and modified the
commonly used methods of modelling the bare poles windage, the influence of the sails on
the windage will now be investigated in the following sections.
3.5.1 Influence of Sails on Hull/Deck Lift and Drag
Figure 3.16 and Figure 3.17 show the horizontal coefficient of drag (C
DA
) and lift (C
LA
) of the
hull/deck for the upright model with and without sails calculated from the pressure
measurements. The measurements without sails are now shown in the upwind and downwind
configuration, which have different vertical speed profiles. In the downwind configuration the
onset flow is also twisted, i.e. the onset flow angle changes with height, and the model has the
spinnaker pole attached. It can be seen from Figure 3.16 and Figure 3.17 that the hull
windage without sails is slightly less in the downwind configuration and that the
measurements are shifted to the right. The windage is less since the flow is slowed down
more at the height of the hull with the downwind vertical speed profile. Due to the twist the
onset flow angle at the height of the hull is smaller than for the upwind configuration which
shifts the downwind measurements to the right.
From Figure 3.16 and Figure 3.17 it can be seen that the presence of the sails influences the
hull/deck forces significantly. The horizontal drag coefficient (C
DA
) is almost doubled for
many apparent wind angles (Figure 3.16) and the horizontal lift coefficient (C
LA
) is increased
0
0.05
0.1
0.15
0.2
0.25
0.3
0 20 40 60 80 100 120 140 160 180

A
[]
C
D
A

[
-
]
Hull windage no sails (upwind)
Hull windage no sails (downwind)
Hull windage main & jib (upwind)
Hull windage main & spi (downwind)
Hull windage sails Hochkirch (upwind)
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0 20 40 60 80 100 120 140 160 180

A
[]
C
L
A

[
-
]
Hull windage no sails (upwind)
Hull windage no sails (downwind)
Hull windage main & jib (upwind)
Hull windage main & spi (downwind)
Figure 3.16: Horizontal drag coefficient (C
DA
) of
hull/deck for the model with and
without sails in upwind and
downwind configuration
Figure 3.17: Horizontal lift coefficient (C
LA
) of
hull/deck for the model with and
without sails in upwind and
downwind configuration
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 59
Chapter 3 Hull and Sail Force Interaction
by up to a factor of three when sailing upwind (Figure 3.17). The general shape of the curves
is however still very similar to the measurements of hull/deck without sails. Figure 3.16 also
shows the hull windage coefficient calculated by equation (3.7) for the upwind configuration
with the a
0
value of 0.764 obtained by Hochkirch (2000) from the full-scale measurements
with the sails present. The predicted drag coefficient is much lower than the values measured
in the wind tunnel and more similar to the wind tunnel measurements without the sails.
Hochkirch (2000) determined the value of a
0
together with parameters describing the
residuary resistance and added resistance due to heel from the full-scale measurements using
multi variable regression. The windage drag of the hull is significantly smaller than the
hydrodynamic resistance components so that scatter in the data causes much greater relative
errors in a
0
. The lowest part of the atmospheric boundary layer profile has a significant
influence on the hull drag. It has not been measured as part of the Dyna projects and the
standard logarithmic profile given in equation (2.20) has been used to determine the value for
a . As will be discussed in detail in chapter 6
0
this may not be a suitable approximation for
some of the full-scale measurements which would affect the magnitude of a
0
. In addition the
level of detailing on the wind tunnel model and possible scaling effects could contribute to
the difference between wind tunnel and full-scale hull drag.
Another possible reason for the difference between the wind tunnel measurements and the
prediction based on full-scale data is the heel angle of the yacht. Wind tunnel tests therefore
were conducted at three heel angles and Figure 3.18 and Figure 3.19 show the influence heel
has on the lift and drag coefficient of the hull for the model with and without sails. The data
shown is obtained from the pressure measurements and is plotted in each figure in the
horizontal plane (coordinate system A) as C
DA
and C
LA
against
A
and in the heeled plane
(coordinate system B) as C
D
and C against the effective wind angle (
L eff
) using effective
angle theory to obtain the reference speed as described in section 2.3.1. Figure 3.19 shows
that the drag coefficient of the hull with the sails present changes less with heel angle if it is
plotted in the horizontal plane. The drag coefficient of the hull without the sails on the other
hand shows a strong dependency on heel angle when plotted in the horizontal plane and
changes much less with heel angle if plotted in the heeled plane. Figure 3.19 shows the same
trends for the lift coefficient. This behaviour is caused by the force perpendicular to the deck
plane that will be discussed in section 3.6.
From Figure 3.18 and Figure 3.19 it can be concluded that bare poles windage should be
calculated in the heeled plane using the effective angle theory and not in the horizontal plane
as shown in section 3.5, to minimise its dependency on heel angle. Generally it can be said
that there is some influence of heel on the measurements in the heeled plane but no clear
trend, which could be modelled, can be determined from the data. Describing the windage
forces with equations (3.11) and (3.12) used in the heeled plane is therefore feasible. Figure
3.18 and Figure 3.19 show that the hull windage in the horizontal plane with the sails present
60
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.5 Lift and Drag Forces of Hull/Deck and Rig
0
0.05
0.1
0.15
0.2
0.25
0.3
0 15 30 45 60 75 9

A
[]
C
D
A

[
-
]
0
Hull windage no sails 0 heel
Hull windage no sails 12.5 heel
Hull windage no sails 25 heel
Hull windage main & jib 0 heel
Hull windage main & jib 12.5 heel
Hull windage main & jib 25 heel
0
0.05
0.1
0.15
0.2
0.25
0.3
0 15 30 45 60 75

eff
[]
C
D

[
-
]
90
Hull windage no sails 0 heel
Hull windage no sails 12.5 heel
Hull windage no sails 25 heel
Hull windage main & jib 0 heel
Hull windage main & jib 12.5 heel
Hull windage main & jib 25 heel
Figure 3.18: Drag coefficient of hull/deck for the model with and without sails in upwind configuration at
different heel angles plotted in horizontal plane (coordinate system A) as C
DA
and in the
heeled plane (coordinate system B) as C
D

-0.05
0
0.05
0.1
0.15
0.2
0.25
0 15 30 45 60 75

A
[]
C
L
A

[
-
]
90
Hull windage no sails 0 heel
Hull windage no sails 12.5 heel
Hull windage no sails 25 heel
Hull windage main & jib 0 heel
Hull windage main & jib 12.5 heel
Hull windage main & jib 25 heel
-0.05
0
0.05
0.1
0.15
0.2
0.25
0 15 30 45 60 75 9

eff
[]
C
L

[
-
]
0
Hull windage no sails 0 heel
Hull windage no sails 12.5 heel
Hull windage no sails 25 heel
Hull windage main & jib 0 heel
Hull windage main & jib 12.5 heel
Hull windage main & jib 25 heel
Figure 3.19: Lift coefficient of hull/deck for the model with and without sails in upwind configuration at
different heel angles plotted in horizontal plane (coordinate system A) as C
LA
and in the
heeled plane (coordinate system B) as C
L
does not change significantly with heel angle so that the heel angle of the yacht during the
full-scale tests should not be the reason for the discrepancies between the wind tunnel
measurements and the prediction by Hochkirch (2000) discussed before and shown in Figure
3.16.
The cause of the increase in the lift and drag of the hull due to the presence of the sails is
shown in the surface pressure plots in Figure 3.20 for the upright model in upwind
configuration. At
A
of 30, Figure 3.20(a) and Figure 3.20(c) show some increase in positive
pressure on the windward side of the hull due to the presence of the sails but more
importantly a significant increase in suction on the leeward side of the hull close to the bow.
The foot of the jib is close to the deck so that the pressure differential between the windward
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 61
Chapter 3 Hull and Sail Force Interaction

of 30 of 60 (a) No sails at (b) No sails at
A A


of 30 of 60 (c) Mainsail and jib at (d) Mainsail and jib at
A A
Figure 3.20: Surface pressure plots for model on port tack at 0 heel without sails and with mainsail and
jib at two apparent wind angles showing leeward and windward side of hull
and leeward side of the jib is carried over onto the hull. Similarly at
A
of 60, Figure 3.20(b)
and Figure 3.20(d) show an increase in pressure on the windward side of the hull and a
reduction on the leeward side. However, the increase in suction less than at
A
of 30 because
of the increased angle of the onset flow and the bigger gap between jib and deck.
From Figure 3.20 it can be deduced that the increase in both lift and drag of the hull due to
the sails is mainly caused by an increase in side force (F
Y
). This can also be shown by
plotting the drive force and side force coefficient (C and C
Fx Fy
) of the total model with
mainsail and jib, and of the hull with and without sails in Figure 3.21. The hull forces are
again obtained from the pressure measurements. C
Fx
of the hull is only small compared to the
total drive force coefficient and does not change much due to the presence of the sails. At
eff

-0.2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 15 30 45 60 75

eff
[]
C
F
x

[
-
]
90
Hull windage no sails
Hull windage main & jib
Total force of main, jib, hull and rig
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 15 30 45 60 75 9

eff
[]
C
F
y

[
-
]
0
Hull windage no sails
Hull windage main & jib
Total force of main, jib, hull and rig
Figure 3.21: Drive force and side force coefficient (C and C
Fx Fy
) of total model with mainsail and jib and
of hull/deck with and without sails for the upright model in upwind configuration
62
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.6 Hull/Deck Forces Perpendicular to Deck Plane
of 30 it is about 1.5% of the total C
Fx
and at larger angles the percentage is significantly less.
C of the hull on the other hand is about 3.5% of the total C at
Fy Fy eff
of 30 without the sails
present and increases to 9% with the sails. At
eff
of 60 it is as much as 14% of the total C .
Fy
Ideally the hull windage with sails present should be modelled in VPPs but it is not feasible to
readily measure it as part of the wind tunnel testing procedure of sails. Measuring the bare
poles windage and subtracting the windage lift and drag from the total is more feasible. This
is an adequate method since the additional drag and lift of the hull due to the sails is then
included in the sail forces, to which it is related anyway. It can be expected that the sails and
their trim have at least as much influence on the additional hull forces as the actual hull
shape. It is important to remember that the wind tunnel sail forces include the additional hull
lift and drag when comparing to computational fluid dynamics (CFD) simulations where
these extra force components may not be modelled. For Dyna this would result in an under-
prediction of the side force by up to 14%.
3.6 Hull/Deck Forces Perpendicular to Deck Plane
Finally the third force component perpendicular to the deck plane is investigated with and
without the sails present. Figure 3.22 shows the coefficient of the force perpendicular to the
deck plane (C
Fz
) of the hull/deck for the model with and without sails calculated from the
pressure measurements. The measurements without sails were conducted in the upwind and
downwind configurations, which have different vertical tunnel speed profiles. In the
downwind configuration the onset flow is also twisted, i.e. the onset flow angle changes with
height, and the model has the spinnaker pole attached. Due to these differences C is slightly
Fz
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0 20 40 60 80 100 120 140 160 180

eff
[]
C
F
z

[
-
]
Hull windage no sails (upwind)
Hull windage no sails (downwind)
Hull windage main & jib (upwind)
Hull windage main & spi (downwind)

Figure 3.22: Force coefficient perpendicular to
deck plane (C
Fz
) of hull/deck for
model with and without sails in
upwind and downwind configuration
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 63
Chapter 3 Hull and Sail Force Interaction
less for the downwind configuration. The hull/deck acts like a bluff body and the accelerated
flow over the deck produces a suction force. For an apparent wind angle of 90 the aspect
ratio of the deck is largest and the maximum C is produced.
Fz
From Figure 3.22 it can also be seen that C
Fz
of the hull/deck is significantly affected by the
presence of the sails. For the hull/deck with mainsail and jib, the C
Fz
is close to zero for small
effective wind angles (
eff
) and increases up to a
eff
of 90, the largest effective angle tested.
The differences in the flow pattern around the hull/deck with and without sails can be seen
from the surface pressure plots at the apparent wind angles of 30, 60 and 90 in Figure 3.23.
For the model without sails at
A
of 30 Figure 3.23(a) shows an area of high suction close to
the bow and slightly negative pressure over the whole deck. At the larger
A
of 60 and 90
shown in Figure 3.23(b) and Figure 3.23(c) the suction peak at the bow is no longer visible
but increased suction can be noted along the windward deck edge as
A
increases, which
explains the positive C that increases with
Fz A
and peaks at 90. With mainsail and jib
present Figure 3.23(d) on the other hand shows that large areas of the deck experience
positive pressure at
A
of 30, especially in the slot between the jib and the mast.
This explains why C is close to zero for small
Fz eff
. With increasing
A
the positive pressure
exerted by the sails onto the deck reduces, as evident from Figure 3.23(e) and Figure 3.23(f),
which results in an increase in C
Fz
. The suction on the deck is however much less than
without the sails present and C is significantly reduced as was shown in Figure 3.22.
Fz

(c) No sails at of 90
(b) No sails at of 60 of 30 (a) No sails at A
A A


of 30 (e) Mainsail and jib at of 60 (f) Mainsail and jib at of 90 (d) Mainsail and jib at
A A A
Figure 3.23: Surface pressure plots for model on port tack at 0 heel without sails and with mainsail and
jib at different apparent wind angles
64
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.6 Hull/Deck Forces Perpendicular to Deck Plane
C
Fz
of the hull/deck behaves somewhat differently with the mainsail and spinnaker as can be
seen in Figure 3.22. It is positive even at small
eff
because the spinnaker does not exert as
much positive pressure onto the deck due to its significantly larger distance from the deck.
Figure 3.24(d) shows suction on the foredeck with mainsail and spinnaker present and
positive pressure is only exerted onto the deck further aft. As
eff
increases and the mainsail
and spinnaker are eased, the suction on the foredeck disappears as shown in Figure 3.24(e)
for the model at
A
of 100. The flow is also more directed downwards onto the deck
compared to the mainsail with jib case in Figure 3.23(f), which reduces C
Fz
and makes it
negative for
eff
larger than 120. The pressure exerted onto the deck by the mainsail and
spinnaker at large
A
is clearly shown in Figure 3.24(f).
3.6.1 Forces Perpendicular to Deck Plane at Heel
For upwind sailing conditions it is important to also investigate the influence of heel on the
force perpendicular to the deck plane. The force coefficient perpendicular to deck plane (C
Fz
)
of the hull/deck for the model without sails was determined at three heel angles and increases
with heel angle for all
eff
as shown in Figure 3.25. The difference in C
Fz
between 0 and 25
heel is up to 40%. Figure 3.26(a) to Figure 3.26(c) show that the area of high suction on the
windward side of the deck close to the bow gets larger as the yacht heels, which increases

of 40 (b) No sails at of 100 (c) No sails at of 160 (a) No sails at
A A A


of 40 (e) Mainsail and spi at of 100 (f) Mainsail and spi at of 160 (d) Mainsail and spi at
A A A
Figure 3.24: Surface pressure plots for model on port tack at 0 heel without sails and with mainsail and
spinnaker at different apparent wind angles
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 65
Chapter 3 Hull and Sail Force Interaction
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 15 30 45 60 75

eff
[]
C
F
z

[
-
]
90
Hull windage no sails 0 heel
Hull windage no sails 12.5 heel
Hull windage no sails 25 heel

Figure 3.25: Force coefficient perpendicular to
deck plane (C
Fz
) of hull/deck for
model without sails at different heel
angles
C
Fz
. The hull/deck is plotted in the body fixed coordinate system (coordinate system B) so
that only the water plane changes orientation with heel in Figure 3.26. Figure 3.27 shows C
Fz

of the hull/deck with the sails present for different heel angles. The increase with heel is even
more significant than without sails. Figure 3.26(d) to Figure 3.26(f) show that the positive
pressure exerted by the jib onto the foredeck reduces with heel angle, which increases C .
Fz

(b) No sails at 12.5 heel (c) No sails at 25 heel
(a) No sails at 0 heel


(d) Mainsail and jib at 0 heel (e) Mainsail and jib at 12.5 heel (f) Mainsail and jib at 25 heel
Figure 3.26: Surface pressure plots for model on port tack at 30 apparent wind angle without sails and
with mainsail and jib at different heel angles
66
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.7 Sail Forces Perpendicular to Deck Plane
This study on the forces perpendicular to the deck plane shows C
Fz
of the hull/deck to be as
much as 20% of the maximum lift generated by the sails at
eff
of 90 and 25 heel. This is,
however, an unusual sailing condition, but similar values could possibly also occur at smaller

eff
and larger heel angles. It is worth noting that the effects of small trim changes can easily
be seen in the resulting C
Fz
of the hull/deck. Figure 3.27 shows the difference in C
Fz
of the
hull/deck for two slightly different trim settings at a heel angle of 25 and
eff
of 37.3 (i.e.
A

of 40). One trim is considered to be the best at 0 heel, the second best at 25 heel.
3.7 Sail Forces Perpendicular to Deck Plane
C
Fz
of the sails and the rig can be determined from the total C
Fz
measured by the wind tunnel
force balance and subtracting the C
Fz
of the hull/deck measured by the internal force balance.
C
Fz
for the mainsail, jib and rig is shown in Figure 3.28 for different heel angles. The vertical
forces measured by the wind tunnel force balance are, however, less accurate than the
hull/deck force measurements and the results show more scatter. The vertical forces acting on
the model are very small compared to the forces measured by the individual force transducers
of the force balance, which include the forces due to the model weight and the heeling
moment, so that the accuracy is less than in the other force directions. The balance is
designed in this way because for most wind tunnel investigations on yacht sails the vertical
force is the least important. For 0 and 12.5 heel it is difficult to see a clear trend in C
Fz
with

eff
and C
Fz
has a value of around 0.1. For 25 heel C
Fz
increases with
eff
up to approximately
0.14 at
eff
of 90. With the method used it is not possible to remove the windage of the rig
with the sails present from the measurements, but it can be shown that the rig without sails
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 15 30 45 60 75 9

eff
[]
C
F
z

[
-
]
0
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 15 30 45 60 75

eff
[]
C
F
z

[
-
]
90
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
Hull windage main & jib 0 heel
Hull windage main & jib 12.5 heel
Hull windage main & jib 25 heel
Best trim at 0 heel
Best trim at 25 heel
Figure 3.28: Force coefficient perpendicular to
deck plane (C
Figure 3.27: Force coefficient perpendicular to
deck plane (C
Fz
) of hull/deck for
model with mainsail and jib at
different heel angles
Fz
) of mainsail, jib and
rig at different heel angles
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 67
Chapter 3 Hull and Sail Force Interaction
produces little C . The presence of the sails should not influence C
Fz Fz
of the rig significantly
so that C shown in Figure 3.28
Fz
should predominately be caused by the sails. It seems very
feasible that a small force component perpendicular to the deck plane, which is fairly
independent of
eff
and heel, is caused by the sails. A plane through the three corners of the
jib is not perpendicular to the deck plane because of the backwards-angled forestay and the
sheeting point being to leeward of the centreline. This can cause the small force perpendicular
to the deck.
3.8 VPP Modelling of Forces Perpendicular to Deck Plane
Currently VPPs assume that all aerodynamic forces on the yacht act in the deck plane and no
force component perpendicular to the deck plane is modelled. Sections 3.6 and 3.7 however
showed that the hull/deck and the sails and rig produce a positive force perpendicular to the
deck (F ) for almost all conditions investigated with mainsail and jib. F
Z Z
is therefore
modelled in a VPP to investigate its influence on the performance of the yacht. As will be
discussed in detail in chapter 4 FRIENDSHIP-Equilibrium is used for velocity predictions in
this project. The standard hydrodynamic model of Dyna outlined in appendix C.4 and the
aerodynamic force model in section 4.2.1 with the trim parameters reef and flat are used. The
aerodynamic force model is modified to include the total F
Z
of hull/deck, rig and sails. The
contribution from the hull/deck strongly depends on the effective wind angle and heel angle
so that the total C
Fz
is modelled as a function of these two parameters. The VPP input is
defined as a B-spline response surface from the measurements as shown in Figure 3.29.
2.9
3.1
3.3
3.5
3.7
3.9
4.1
40 50 60 70 80 90

T
[]
V
S

[
m
/
s
]
VT 8m/s VT 8m/s with Fz
VT 6m/s VT 6m/s with Fz
VT 4m/s VT 4m/s with Fz

Figure 3.30: Boat speed (V Figure 3.29: Surface describing the total force
coefficient perpendicular to deck
plane (C
S
) with and without
including force perpendicular to deck
plane (F
Fz
) as a function of effective
wind angle (
Z
) for different true wind
angles ( ) and true wind speeds (V ) and heel angle )
eff T T
68
YACHT
RESEARCH
UNIT
The
University
of Auckland
3.9 Conclusions
Performance predictions are carried out solving for all six degrees of freedom with and
without inclusion of the force component perpendicular to the deck plane. Figure 3.30 shows
the boat speed for true wind angles ( ) between 40 and 90 and true wind speeds (V
T T
) of 4,
6 and 8m/s. A clear trend can be seen that the inclusion of F decreases the boat speed (V
Z S
)
increasingly with wind speed. The vertical upward component of F
Z
is negligible compared to
the displacement forces of the yacht and the reduction in V
S
is due to the side force
component of F
Z
, which increases as the yacht heels more in stronger wind. The reduction in
V is however very small. For most conditions V
S S
is reduced by less than 1%. The difference
in time required to cover one nautical mile is less than 3s for most sailing conditions.
Nevertheless the force component perpendicular to the deck plane should be included in the
VPP formulation to more accurately model the physical behaviour of the force and moment
system of a sailing yacht.
3.9 Conclusions
Force measurements with a secondary internal force balance inside the hull/deck shell and
surface pressure measurements on the hull/deck shell have successfully been conducted and
used to investigate the forces acting on the hull/deck with and without sails present. The
forces obtained with the two methods agree very well and the pressure measurements provide
useful information to investigate the flow field around the hull/deck. It has been shown that
the side force of the hull/deck is increased by the presence of the sails at all apparent wind
angles. The increase is greatest for small apparent wind angles where high lift coefficients are
achieved and the foresail is close to the deck so that the pressure differential between the
windward and leeward sides of the sails is carried over to the hull. The side force generated
by the hull increases by up to a factor of 2.5 when the sails are present and can be as high as
14% of the total side force of the model with sails. Accurate modelling of the hull in the wind
tunnel is therefore not only important to achieve the correct flow over the sails but also to
account for the hulls contribution towards the aerodynamic forces of the yacht. Full-scale
measurements, panel methods and other CFD simulations that do not calculate the forces
acting on the hull significantly under-predict the total side force of the yacht.
Although bare poles measurements of windage do not represent the actual forces that act on
the hull/deck and rig once the sails are present, they are still an adequate method for taking
windage into account. The additional windage forces due to the presence of the sails are
strongly related to the sails and their trim, and it can be argued that they should be modelled
as part of the sail forces. It is also not feasible to readily measure the hull/deck and rig forces
with the sails present as part of the normal sail testing procedure whereas the bare poles
method is straightforward and quick to conduct.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 69
Chapter 3 Hull and Sail Force Interaction
In addition to the bare poles windage drag of hull/deck and rig, which are commonly
modelled in VPPs, the windage lift should also be modelled, in particular the hull/deck
component as it can be up to 49% of the hull/decks windage drag. An improved bare poles
windage drag and lift model has been presented. It models the windage drag as a sine squared
function and the lift as a sine function and defines the force coefficients so that they are
directly comparable to other aerodynamic and wind engineering data. The coefficients
obtained from the wind tunnel measurements relate well to coefficients for similar standard
geometric shapes taken from literature. The bare poles lift and drag are still modelled as
independent of heel angle, but should be calculated in the heeled plane unlike the current
windage models that describe the windage drag in the horizontal plane.
The third force component, the force perpendicular to the deck plane, was also investigated.
Due to suction on the deck an upward force perpendicular to the deck plane is generated by
the hull/deck without sails, which strongly depends on the apparent wind angle and heel
angle. The flow over the sails exerts positive pressure onto the deck which reduces the
upward force. The positive pressure exerted onto the deck by the sails reduces with heel angle
and the hull/deck force perpendicular to the deck plane with the sails present is still a function
of apparent wind angle and heel angle. The sail and rig force perpendicular to the deck plane
shows much less dependency on the apparent wind angle and heel angle. For large heel
angles the hull/deck contributes more than 50% of the total force perpendicular to the deck
plane.
The total aerodynamic force perpendicular to the deck plane has been modelled in a VPP
based on the wind tunnel data as a function of apparent wind angle and heel angle, since most
VPPs currently do not consider this force component. It was found to have a consistent but
small influence on the performance of the yacht. Its inclusion in the VPP formulation reduces
the yacht speed by typically less than 1%. Nevertheless it should be considered in VPPs as it
improves the physical model of the yacht and although its influence on the performance is
small for Dyna it may be more relevant for other design studies where small improvements
result in a competitive advantage.
70
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
4 Real-Time Velocity Prediction Program
4.1 Introduction
Velocity prediction programs (VPPs) are popular tools which predict the speed of a yacht at
the design stage. Instead of using fully computational methods VPPs are usually based on
semi-empirical mathematical models to describe the forces acting on the yacht. The input for
the different force components can be obtained from full-scale or model scale testing,
numerical simulations or theoretical calculations. VPPs allow the different force components
to be determined in the most appropriate way and be analysed together to predict the resulting
performance. The hull resistance might be obtained from towing tank tests, whereas the keel
and rudder characteristics are obtained from CFD simulations and the sail forces are
measured in the wind tunnel. It is important not to investigate a design change on a
component of the yacht in isolation but based on the resulting yacht performance. A reduction
in skin friction drag of the keel bulb, for example, probably increases the centre of gravity,
which reduces the stability of the yacht. The trade off between skin friction and stability
needs to be based on the yacht performance so that VPP predictions are essential.
For the design of yacht rigs and sails the use of VPP predictions is equally important. Only
for a few sailing conditions can it be assumed that solely the drive force of the sails affects
the boat speed. In most cases other force components like, for example, the side force and
heeling moment should be considered as well. In particular, the heeling moment (M
X
) should
not be ignored for the majority of sail types. The heeling moment generated by the sails will
result in a heel angle, which increases the hydrodynamic resistance and reduces the
appendage and aerodynamic efficiency. This could result in a slower boat speed even though
the drive force generated by the sail is larger than for a different sail that produces less
heeling moment. This argument is not confined to M
X
, similarly designing a sail that reduces
the aerodynamic side force and consequently the leeway angle or a sail that reduces the yaw
moment (M
Z
) and hence rudder angle could result in a better boat speed even though it
produces less drive force. VPPs are therefore also frequently used during the rig and sail
design process.
The force modelling of sails is however distinctly different from other yacht components in
that not only the design determines the generated forces and moments but also the trim of the
sails. In certain sailing conditions it pays to depart from the aerodynamically most efficient
sail shape to increase the boat speed for the same reasons as discussed before. The
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 71
Chapter 4 Real-Time Velocity Prediction Program
aerodynamically most efficient sail shape produces the maximum drive force and is called the
fully powered-up sail trim. Departing from this optimum sail shape is called depowering. The
need for considering the different possible flying shapes of a sail design increases the
complexity of analysing the performance of sails.
Wind tunnel testing is an effective and commonly used tool for research, development and
design of rigs and sails. It is used for research purposes to study different aspects of yacht
aerodynamics. Sail and yacht designers conduct wind tunnel testing to investigate different
sail shapes and rig configurations. One of the advantages of wind tunnel testing over CFD
simulations is that soft or semi-rigid sails can be used so that the flying shape of the sails can
be altered by trimming the sails similar to full-scale. When using soft sails the trimming
becomes very important, as on the real yacht, and is an integral part of the testing process for
obtaining meaningful results. The primary sail controls that are usually adjusted in the wind
tunnel with remote control winches are, for upwind testing; mainsail sheet, traveller, jib sheet
and jib car, and for downwind testing; mainsail sheet, boom vang, spinnaker sheet, brace,
downhaul and tweaker. Additional sail controls such as cunningham, outhaul, forestay and
(running) backstay can be adjusted on the model so that the sail control and behaviour is very
similar to real life. In the wind tunnel the measured driving force is used as a guide for
trimming the sails to the optimum fully powered-up shape as shown in the flow chart in
Figure 4.1. Firstly an apparent wind angle (
A
) and a heel angle are selected in the wind
tunnel. The sails are then trimmed with the tunnel running, constantly measuring and
displaying the forces as a trimming aid. Once the optimum sail trim has been achieved the
forces are averaged over a period of time. The forces normalised by the reference dynamic
pressure in the wind tunnel and possibly the sail area can then be compared to determine the
best performing sail for the tested
A
and heel angle () based on drive force (F
X
). This basic
Select Apparent wind angle (
A
) and heel angle ()
Trim sails to optimum shape (maximise F
X
)
Measure forces
Normalised forces
Non-dimensional coefficient curves
VPP with trim parameters Proper comparison to other sails For any
T
and V
T
Based on boat speed
or VMG
Initial comparison to other sails For one
A
and
Based on F
X

Figure 4.1: Traditional wind tunnel test procedure
72
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.1 Introduction
comparison however ignores other aerodynamic force components that can affect the
performance of the yacht as discussed before. For many types of sails this comparison will
therefore not necessarily yield the sail that performs best in real life on the water. The
optimum sail trim and the associated forces and moments are hence obtained for a number of

A
and possibly different by repeating the same trimming procedure. Non-dimensional
coefficient curves are obtained from the measurements as input for a semi-empirical VPP.
With a VPP the different sails can now be compared based on the resulting boat speed or the
velocity made good (VMG) for any true wind speed and angle (V and ) within the
T T A
range
tested in the wind tunnel.
This procedure so far assumes that the sails are trimmed to the aerodynamically optimum
shape in all sailing conditions and does not allow for depowering. The standard way of
modelling depowering in semi-empirical VPPs is by employing trim parameters, which
modify the fully powered-up input data to optimise the boat speed. This method is also used
in association with wind tunnel tests but the trim parameter model has potential shortfalls and
was developed for upwind sailing conditions whereas most wind tunnel testing focuses on
offwind sails where depowering can still be important.
Campbell (1998) describes an alternative method for comparing different sail designs in the
wind tunnel without the use of VPP predictions. The sails are still trimmed to maximise drive
force but the heeling moment is not allowed to exceed a certain critical value corresponding
to the optimum heeling angle of the yacht. If no VPP predictions are to be used this method is
a good alternative but it makes trimming of the sails even more complex and not all force and
moment components are considered. Ranzenbach and Teeters (2002) use wind tunnel test
results in a semi-empirical VPP without using the standard trim parameters. The sails are
optimally trimmed by maximising drive force in the wind tunnel and then depowered by a
matrix of predefined trim alterations. By looking at the ratio of heeling moment to drive force
reduction the best depowered trims are selected and the forces and moments modelled in a
VPP as surfaces instead of curves to incorporate the level of depowering. Although this
method should provide a more accurate mathematical model of the depowering, determining
the best depowered trims is not straight forward since the sail shapes can be altered in many
different ways so that defining the trim matrix can be complex and selecting the best
depowered trims from the trim matrix results is not necessarily obvious.
The existing methods introduced so far use the wind tunnel forces to trim the sails and
employ some form of semi-empirical model to describe the depowering. This is distinctly
different from trimming the sails on a real yacht where the boat speed, heel angle and even
the rudder and trim tab angle are used as indicators. Since trimming of the sails is an
important part of conducting wind tunnel tests, professional sailors often trim the sails in the
wind tunnel and it is advantageous to have a trimming process that is similar to real life
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 73
Chapter 4 Real-Time Velocity Prediction Program
Select Apparent wind angle (
A
)
Trim sails to optimise boat speed
Measure forces
VPP
Boat speed
Rotate model to correct heel angle ()
Compare to other sails For one V
T
and
T
,
A
or
eff
Based on boat speed
Figure 4.2: Wind tunnel test procedure with the Real-Time VPP
sailing. To achieve this the Real-Time VPP has been developed, which calculates the yacht
performance based on the forces and moments measured in the wind tunnel in real-time
11

while the sails are trimmed. The wind tunnel test procedure using the Real-Time VPP is
shown in Figure 4.2. For an apparent wind angle (
A
) set in the wind tunnel the sails are
trimmed for optimum boat speed by constantly measuring the forces and directly calculating
the resulting boat speed. While boat speed is the primary performance indicator the Real-
Time VPP also calculates other free variables depending on the degrees of freedom. The
calculation of the heel angle while the sails are trimmed provides useful information as well.
Without a Real-Time VPP a fixed heel angle of the wind tunnel model needs to be
approximated and often the tests are conducted at more than one heel angle. With the Real-
Time VPP the heel angle is calculated while trimming the sails so that the wind tunnel model
can be dynamically heeled to the correct angle. Dynamic heeling makes the trimming even
more realistic and measuring the forces at the correct heel angles gives a more accurate
representation of the full-scale situation.
In addition to making the trimming more similar to the real life situation, different sail
designs can immediately be compared to each other based on boat speed without post-
processing of the results and no approximations are required to model depowering (compare
Figure 4.1 and Figure 4.2). Before discussing the implementation of the Real-Time VPP a
closer look at standard semi-empirical VPPs is useful to understand the principles of
aerodynamic force modelling.

11
Real-time is a relative term and although there is a time delay in the system it is sufficiently small for the
results to be available as immediate feedback to any trim changes.
74
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.2 Standard Semi-Empirical VPPs
4.2 Standard Semi-Empirical VPPs
Semi-empirical VPPs are popular tools to predict the speed of a yacht at the design stage. In
most cases a steady state analysis approach is taken. The yacht is assumed to operate in a
constant wind environment. All forces acting on the yacht need to be in equilibrium for the
yacht to move at a constant speed, in a constant direction, at a constant orientation. In
practice often not all six degrees of freedom are considered, mainly because of difficulties in
modelling all the aerodynamic force and moment components. Usually two equations for the
force equilibrium in the plane of the water are set up. The drive force generated by the sails
must equal the resistance of the hull, rig and appendages and the side force generated by the
keel, hull und rudder must balance the side force of the sails. A third equation is taken from
the heel moment and righting moment equilibrium. Figure 4.3 shows the force and moment
components of a three degrees of freedom VPP. Some VPPs use a fourth equation to balance
Input of force
coefficients as
functions of
eff
, , ,
, r, f
Calculate drive force, aero side
force and CoE
Initial guess of
V
S
,, , , , r, f
Calculate from heeling and
righting moment
Is new equal to old ?
Change
to new
value
Calculate V
S
from drive force
and resistance
Is new V
S
equal to old V
S
?
Change
V
S
to new
value
Calculate from aero and
hydro yaw moment
No
Yes
Yes
No
Is new equal to old ?
Change
to new
value
Yes
No
Calculate from aero and
hydro side force
Is new equal to old ?
Change
to new
value
Yes
No
Change f and r to maximise V
S
Has maximum V
S
been
achieved?
Change f
& r to new
values
Yes
No
Specify V
T
&
T
Output result for V
S
, , , ,
r, f
Start
End
Aerodynamic forces
Hydrodynamic forces
VPP VPP
Righting moment
Side
force
Drive
force
Apparent wind (V
A
)
Apparent angle (
A
)
Lift
Boat speed (V
S
)
True wind (V
T
)
True angle (
T
)
Drag
Side
force
Resistance
Heeling moment
Leeway angle ()
Righting moment
Side
force
Drive
force
Apparent wind (V
A
)
Apparent angle (
A
)
Lift
Boat speed (V
S
)
True wind (V
T
)
True angle (
T
)
Drag
Side
force
Resistance
Heeling moment
Leeway angle ()



Figure 4.4: Generic semi-empirical VPP with flat
(f) and reef (r) as trim parameters
Figure 4.3: Forces and moments acting on a
sailing yacht that are modelled and
balanced by a three equation VPP
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 75
Chapter 4 Real-Time Velocity Prediction Program
the yawing moment. There are different ways to set up the mathematical and computational
model for solving the equations.
The most common approach is to determine the maximum yacht speed (V
S
) for a given true
wind direction ( ) and speed (V
T T
). The force coefficients need to be described in terms of the
free variables of ship speed (V
S
), heel angle (), leeway angle () and rudder angle (). The
aerodynamic sail forces are also functions of the trim parameters, which are traditionally flat
(f) and reef (r), to model the depowering. An initial guess for the free variables is required
before the solution process is started. As shown in Figure 4.4 the first step for each iteration
is to calculate the aerodynamic forces and the centre of effort (CoE). From this the remaining
free variables can be obtained through the iterative process outlined in Figure 4.4.
Alternatively more advanced optimisation methods can be employed to determine the
maximum speed as will be discussed in section 4.3.2. Either process is repeated for all true
wind speeds and angles of interest.
A second method, less often adhered to, is to assume a value for boat speed and leeway angle
from which the hydrodynamic side force and resistance can be calculated. The aerodynamic
forces are then determined by solving the equations for equilibrium. This in turn allows the
calculation of the true wind speed and angle. The calculations are performed for a series of
leeway angles for all points of sail. An interpolation method then yields V
S
, , and for
each true wind speed and angle. This approach is described in more detail by van Oossanen
(1993). With this method it is mathematically simpler to find the aerodynamic and
hydrodynamic force equilibrium and the method places the hydrodynamic characteristics of
the hull and appendages in a more central role.
4.2.1 Aerodynamic Sail Force Model with Trim Parameters
The aerodynamic sail force models of most semi-empirical VPPs derive the forces from the
optimum lift and parasitic drag coefficients (C
Lopt
and C
DpOpt
) obtained when the sails are
trimmed aerodynamically most efficient
12
and described as functions of the effective wind
angle (
eff
). To account for changes in the sail trim that might be necessary to depower the
sails in certain sailing conditions the trim parameters introduced by Kerwin (1978) and Hazen
(1980) are used in these models.
The effective angle theory assumes that the sails are insensitive to span-wise flow that results
from heeling as discussed in section 2.3.1. The sails are therefore only influenced by the flow
component in the heeled (and pitched) z-plane (coordinate system B), which is perpendicular

12
The optimum fully powered-up sail trim does however not mean that the optimum loading distribution of
elliptical or semi-elliptical loading is achieved. While C
LOpt
could also be referred to as the maximum C
L
,
the term optimum is chosen since this trim condition does not result in the maximum C .
Dp
76
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.2 Standard Semi-Empirical VPPs
to an idealised mast without bend or rake. Following the effective angle theory, C
Lopt
and
C
DpOpt
in the z-plane are assumed constant for any heel angle when plotted against
eff
. From
the curves of C
Lopt
and C
DpOpt
against
eff
the forces acting on the sails can hence be calculated
by a VPP in the following way.
From the VPP input parameter of the true wind speed (V
T
) at the reference height of 10
metres (z
ref
) the true wind speed at the height of the geometric centre of sail area above the
water (z) is calculated assuming an logarithmic velocity profile from equation (2.20) as
described in section 2.3.2. From V at height z the apparent angle (
T A
) at height z is obtained
with equation (2.9) which is also dependent on the VPP input parameter
T
and the free
variables of the VPP formulation , and V
S
. Since is usually small it has little influence on
the results and is often set to zero in equation (2.9) to reduce the number of free variables in
this equation and make the solution process more stable. The effective angle (
eff
) at height z
can now be expressed as a function of
A
at height z and the heel angle () by equation (2.11)
where is a free variable in the VPP formulation.
For a given
eff
the lift coefficient can be determined from C
Lopt
as
, 1 , 1 , ) (
2
= f r f r C C
eff Lopt L (4.1)
where reef (r) and flat (f) are the trim parameters introduced by Kerwin (1978). Trim
parameters are used to mathematically account for departing from the optimum sail shape to
depower the sails. They are free variables in a VPP, which can be adjusted from 0-1 to
maximise the speed of the yacht. In its original definition flat (f) <1 represents a linear
reduction in lift due to reducing the camber of the sails. The lift is however in reality not only
reduced by the reduction in camber but also, and probably more importantly, by a reduction
in angle of attack, which is also accounted for by flat so that the name is somewhat
misleading and should perhaps be changed. Originally reef (r) <1 represented a linear
reduction in span and chord of the sails. The sail area is therefore reduced by r
2
. Reef was
originally intended as a geometric rather than an aerodynamic factor to model the physical
reefing of a sail, but it has become thought of as representing any change to the sail trim that
reduces the centre of effort height.
2
As described in section 2.3 C is ideally a function of C
Di L
and depends on the aspect ratio
(AR) and the efficiency (e) of the rig. Reef is intended to model a reduction in sail area so that
it does not change the lift per unit area. C
4 2
is therefore not a function of r but r
Di
since only
the area is reduced by reef but not C defining C . Equation (2.6)
L Di
can be rewritten
incorporating the trim parameters as
( ) [ ] 5 . 0 , 0 , 1
) (
2
2 2 2
+ = t t c
e AR
f r C
C
t
eff Lopt
Di

, (4.2)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 77
Chapter 4 Real-Time Velocity Prediction Program
where t is the trim parameter twist and c
t
is the twist weight constant. The third trim
parameter twist has been introduced by Jackson (2001) to model the twisting-off of the sails
during depowering, which increases the induced drag due to departing from the optimum
loading distribution and had not been considered explicitly by reef and flat. Twist is based on
work by Jones (1950) which showed that the induced drag can be related to the centre of
effort position if the lift is constant. Twist therefore reduces the centre of effort height and
increases the induced drag but does not affect the lift. The induced drag increases by (1+c
2
t
t
)
from its minimum value where twist describes the reduction in centre of effort height so that
CoEminDi
CoE
z
z
t =1 ,
(4.3)
where z
CoE
is the centre of effort height and z
CoEminDi
is the centre of effort height at which the
minimum induced drag is generated. The increase in induced drag is therefore only modelled
accurately by twist if the fully powered-up centre of effort height (z
CoEopt
) is close to z
CoEminDi

so that twist would be close to zero. The induced drag is minimised at a loading distribution
that induces a constant downwash. For a sail far away from the water surface and the hull this
loading is elliptical and z
CoEminDi
is half way along the luff of the sail. For the idealised case
where the loading may be carried to the water surface due to a closely sheeted jib and a small
gap between the hull and boom, the loading is semi-elliptical and z
CoEminDi
is at 4/(3)=0.42 of
the distance from the water to the masthead. The theoretical minimum value of c
t
for
variations from elliptical loading is 8 and Jones (1950) also found 8 as the ideal value for
variations from semi-elliptical loading. In appendix D.3 lifting-line theory is used to confirm
that linearly varying downwash results in the minimum c
t
of 8 for variations from both semi-
elliptic and elliptical loading. In practice c and z
t CoEminDi
may however be different depending
on the actual optimum loading distribution.
The total drag coefficient is the sum of the parasitic drag coefficient (C
Dp
), the induced drag
coefficient (C ), and the separation drag coefficient (C ). C
Di Ds Ds
is also normally expressed as
a function of C
2
as shown in equation (2.5) and can be rewritten with trim parameters as
L
.
2 2 2
) ( f r C c C
eff Lopt S Ds
=
(4.4)
The parasitic drag coefficient is a function of the sail area and only dependent on reef. The
total drag coefficient with trim parameters is therefore given by
.
Di Ds eff DpOpt D
C C r C C + + =
2
) (
(4.5)
From C and C
L D
the drive force (C
Fx
) and side force (C
Fy
) coefficients in the z-plane are
calculated as
78
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.2 Standard Semi-Empirical VPPs
.

L
D
eff eff
eff eff
Fy
Fx
C
C
C
C


cos sin
sin cos
(4.6)
The force coefficient perpendicular to the deck plane (C
Fz
) has previously been assumed to be
zero in aerodynamic force models but can now be modelled as shown in chapter 3. The
aerodynamic force F=(F
T
X
, F , F )
Y Z
in the heeled (and pitched) z-plane can be calculated by
multiplying the coefficient vector C
F
by the reference sail area (A
S
) and the effective dynamic
pressure (q
eff
) as
2
2
eff
air
S F eff S F
V A q A

C C F = = ,
(4.7)
where V
eff
is the effective wind speed in the z-plane, which can be calculated by equation
(2.14) or (2.19), and
air
is the density of air.
The moments are obtained from the centre of effort position in the y-plane, which is aligned
with the centreline of the yacht (x-axis) and the z-axis. The x-position of the centre of effort
(x
CoE
) and z-position (z
CoE
) are described as functions of the resultant force direction () or
effective wind angle (
eff
) as discussed in section 2.3.3. The reduction of z
CoE
when
depowering the sails is modelled by the trim parameters reef and twist. Reef is assumed to
primarily reduce the sail area and centre of effort of the mainsail so that it is applied relative
to the boom. If an elliptical loading distribution over the span of the mainsail is assumed twist
should also be applied relative to the boom so that z
CoE
is calculated by
, ) 1 ( ) ) ( ( t r z z z z
boom eff CoEopt boom CoE
+ =
(4.8)
where z
CoEopt
is the fully powered-up centre of effort height in the body fixed coordinate
system (coordinate system B), here shown as a function of
eff
. z
boom
is the height of the boom
above the design waterline (DWL) also in coordinate system B. If on the other hand a semi-
elliptical distribution is assumed, twist should be applied relative to the water surface and
provided that the origin of coordinate system B is on the DWL the centre of effort height can
be calculated by
( ) ) 1 ( ) ) ( ( t r z z z z
boom eff CoEopt boom CoE
+ = .
(4.9)
The centre of effort position (CoE) is then given by the point
,

=
CoE
eff CoEopt
z
x
0
) (
CoE
(4.10)
where x
CoEopt
is the longitudinal centre of effort of the fully powered-up sails in the body
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 79
Chapter 4 Real-Time Velocity Prediction Program
fixed coordinate system (coordinate system B) here shown as a function of
eff
. The moments
M=(M
T
X
, M , M ) are now obtained from
Y Z
. F CoE M = (4.11)
Finally F and M are transformed from the body fixed coordinate system to the absolute
coordinate system (coordinate system A) by rotating through the heel angle () and pitch
angle () to obtain F
A
and M
A
in the absolute coordinate system, in which the z
A
-plane is
horizontal, with
F T F ) , (
BA A
= ,
(4.12)
M T M ) , (
BA A
= ,
(4.13)
where the rotational transformation matrix T
BA
is defined by equation (2.37). The forces and
moments in the absolute coordinate system can now be equated to the other force components
of the yacht.
4.3 Real-Time VPP Implementation
When developing the Real-Time VPP two main parts need to be considered. The forces and
moments need to be measured in the wind tunnel and the performance of the yacht needs to
be predicted. The performance prediction in the Real-Time VPP is in many ways very similar
to the functioning of a semi-empirical VPP. The hydrodynamics of the yacht are still
modelled by semi-empirical expressions and only the aerodynamic data is now directly
measured in the wind tunnel.
Developing a basic VPP is not very complicated but is time consuming, however with
increasing levels of sophistication a VPP becomes more complex and resource intensive to
develop. Due to the similarities between semi-empirical VPPs and the performance prediction
part required for the Real-Time VPP it was decided not to develop a completely new VPP for
the wind tunnel but to use the semi-empirical VPP called FRIENDSHIP-Equilibrium, which
is developed by FRIENDSHIP SYSTEMS and based on research conducted at the Technical
University Berlin by Hochkirch (2000). FRIENDSHIP-Equilibrium was originally intended
as a research tool and is designed in a modular way so that incorporating it into the Real-
Time VPP is feasible and the implementation was realised through a close cooperation with
FRIENDSHIP SYSTEMS. By incorporating FRIENDSHIP-Equilibrium the Real-Time VPP
is able to make use of sophisticated hydrodynamic force modelling and solving techniques.
Although this level of sophistication is not necessarily required to obtain meaningful results
with a Real-Time VPP in the wind tunnel this approach results in an integrated and flexible
80
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.3 Real-Time VPP Implementation

Computer
A/D card
LP-Filter Pressure transducer LP-Filter
Force balance
Aerodynamic forces
Pitot tube
Reference pressure
Accelerometer
Heel angle ()
Parallel port
Result file
Electric motor
Heel angle ()
Controller
FS-Equilibrium
RT-VPP
LabVIEW
application
Figure 4.5: Schematic description of the
implementation of the Real-Time VPP
in the TFWT

system. The comparison between Real-Time VPP and standard semi-empirical VPP
predictions, for example, becomes very efficient and accurate since the same hydrodynamic
model of the yacht can be used.
At the Twisted Flow Wind Tunnel (TFWT) the data acquisition environment LabVIEW by
National Instruments is used. The data acquisition and data handling part of the Real-Time
VPP is therefore programmed as a LabVIEW application. As shown in Figure 4.5 the
LabVIEW application and FRIENDSHIP-Equilibrium are the two core components of the
Real-Time VPP. The LabVIEW application acquires the wind tunnel data, most importantly
the force balance measurements, the reference dynamic pressure and the heel angle. The
relevant information is then passed to FRIENDSHIP-Equilibrium, which predicts the yacht
performance based on the wind tunnel data and the semi-empirical hydrodynamic model of
the yacht. Information of the resulting yacht performance is passed back to the LabVIEW
application for display and storage. The heel angle of the model in the wind tunnel is obtained
by measuring the orientation of an accelerometer. The accelerometer is not actually
measuring accelerations but is used to sense the effects of gravity on its active element so that
the orientation can be obtained. The measured heel angle is compared to the heel angle
predicted by FRIENDSHIP-Equilibrium and the difference adjusted by driving the electric
heel motor through a controller on the parallel port (see Figure 4.5). This way the model can
be dynamically rotated to the correct heel angle while the sails are trimmed.
4.3.1 Real-Time VPP LabVIEW Application
The Real-Time VPP LabVIEW application is designed and implemented as part of SailView,
the main data acquisition system at the TFWT. The voltages from the LVDTs of the force
balance (see section 2.4.1) pass through a low pass filter (LP-filter) with a cut-off frequency
of 10Hz before being read by an analogue to digital converter card (A/D card) as shown in
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 81
Chapter 4 Real-Time Velocity Prediction Program
Figure 4.5. The natural frequency of the force balance is between 14-18Hz depending on the
direction of the excitation force. The Real-Time VPP LabVIEW application acquires data
from the desired channels of the A/D card for the specified period of time at the specified
sampling frequency and calculates the average value for each channel. In addition to the six
LVDT channels of the force balance, the reference dynamic pressure and several other less
critical channels such as temperature and atmospheric pressure are read. The sampling time
and frequency are specified in the Real-Time VPP graphical user interface (GUI) of
LabVIEW shown in Figure 4.6. The Real-Time VPP has two operating modes, one for
trimming the sails and one for acquiring and saving data for a good sail trim, and a different
sampling time and frequency can be specified for each. Table 4.1 lists all the input parameters
of the Real-Time VPP LabVIEW application. Parameters that are associated with the wind
tunnel testing or that are readily adjusted are specified in the LabVIEW application whereas
parameters that are associated with the VPP calculations or that do not change often are
defined in FRIENDSHIP-Equilibrium as will be discussed in 4.3.3.
The LVDT and reference pressure voltages are corrected by the zero voltages measured
before each run with the wind tunnel turned off to obtain the voltages due to the applied flow
field. The zero voltages change constantly due to temperature drift and are usually re-
measured before each run. The zero voltages also change with heel angle since the centre of
gravity of the model shifts relative to the force balance origin and because the buoyancy of
Figure 4.6: Screen shot of the graphical user interface of the Real-Time VPP LabVIEW application
82
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.3 Real-Time VPP Implementation
the model changes if water is used in the turntable trough. Due to the new concept of
dynamically heeling the model while the wind tunnel is running the LVDT zero voltages
need to be modelled as functions of heel angle. In a calibration process the zero voltages are
measured for different heel angles and the change in voltage for each channel with respect to
heel is modelled as a second order polynomial so that the vector of voltage signal change
(S) is given by
,
2
2 1 0 m m
a a a S + + =
(4.14)
where
m
is the heel angle of the model and a to a
0 2
are the coefficient vectors obtained from
the zero voltage measurements at different
m
. The voltage signal vector S=(S , S , S
X Y Y 1 1 2
,
S
T
, S , S ) of the six LVDT channels, shown in Figure 2.18, can now be related to the
Z Z Z 1 2 3
Table 4.1: Inputs required in the Real-Time VPP LabVIEW Application
Category Input Description
V VPP input VT @ 10m [m/s] at the reference height of 10 metres
T
TWA [deg]

T
to be used by VPP (absolute value)
AWATunnel @ CoA
[deg]

A
of model in wind tunnel at the centre of
sail area (absolute value)
HeelTunnel [deg] Heel angle of model in wind tunnel
Sail area ref model [m^2] Reference sail area of model in wind tunnel
Pressure factor Pressure factor between reference pressure
measurement position and model position
Trim settings Manual / Automatic /
VPP
Method of defining heel angle in wind tunnel
Tack in VPP Tack for which VPP conducts calculations
Tack in wind tunnel Tack of model in wind tunnel
Manual input / Sampling
data
Define whether forces are manually entered
or acquired by force balance
Sampling frequency [Hz] Sampling frequency during trimming
Sampling time [s] Sampling time of each measurement during
trimming (update rate of GUI)
Data influence time [s] Time over which moving average is
calculated
Delay of ref. pressure [s] Time delay of pressure between reference
pressure measurement position and model
position
Acquire settings Acquire Starts the data acquisition process
Final zeros Measures and records a shorter sample
usually used to obtain final zeros
Sampling frequency [Hz] Sampling frequency during data acquisition
Sampling time [s] Length of data acquisition
Man. aero forces
model [N, Nm]
-FY FX FZ -MY MX MZ Manually enter forces and moments in wind
tunnel model scale
q [Pa] Manually enter reference pressure
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 83
Chapter 4 Real-Time Velocity Prediction Program
T
combined force and moment vector F
T(6)
=(F , F
XT YT
, F , M , M
ZT XT YT
, M )
ZT
in coordinate
system T through the calibration matrix (C) so that
( ) ) ( ) (
0 0 ) 6 , 6 ( ) 6 ( m m T
S S S S C F + = ,
(4.15)
where S is the zero voltage signal vector measured with the wind tunnel turned off,
m 0
is
the current heel angle of the model and is the heel angle at which S
m0 0
was measured.
Bonniot (2002) describes in detail how the calibration matrix of the wind tunnel force balance
is obtained. It is also possible to manually enter the forces, moments and a reference pressure
in the GUI for set-up purposes (see Table 4.1).
In the acquisition mode, with sampling times of usually between 20-180 seconds, the
averaged force and moment vectors are transformed to the coefficient vectors C and C
FT MT
,
still in coordinate system T, using equations (2.38) and (2.39) where the reference sail area
(A ) is specified in the GUI. The required effective dynamic pressure at the model (q
S effm
) is
calculated from the heel angle (
m
), the apparent wind angle (
Am
) and the apparent dynamic
pressure (q . The pressure factor (c ) with equation (2.40) ) used in equation (2.41)
Am q
to
calculate q at the geometric centre of sail area from the measured reference pressure (q
Am ref
) is
not constant if the model is dynamically heeled because of the vertical speed gradient in the
onset flow. Although the vertical speed gradient of the apparent wind is not logarithmic, it is
still a feasible approximation over the small range of height change due to heel. q
Am
at the
geometric centre of sail area can therefore be obtained from
ref
m
q Am
q
c
c
c q
2
) ln(
) cos ln(


,
(4.16)
where c is measured at the height of the upright geometric centre of sail area (z
q CoA
) and c

is
the heel constant. For the logarithmic speed profile of the true wind in equation (2.20)
c

=z
ref
/z and as a first approximation c
0
can be calculated here in the same way. Figure 4.7
shows the change in apparent wind speed with height calculated for Dyna for a typical
upwind and downwind sailing condition (see section 2.5.2) and the approximated reduction
due to lowering the centre of sail area with heel, where c

=z
ref
/z =6.67m/0.25mm. Ideally c
0

should be obtained for each wind tunnel profile rather than being calculated but this requires
closely spaced flow measurements around the height of the centre of sail area. This is not
done here but two measurements for the Dyna set-up are shown and their trend agrees well
with the model. The changes in pressure due to heel are noticeable but not significant. At a
typical heel angle of 25 the reduction in V
A
is 1% (2% in q ).
Am
In the trimming mode of the Real-Time VPP a much shorter sampling time is usually selected
since the GUI is updated for each measurement and a short response time is desirable when
trimming. A sampling time of around 0.2 seconds is usually recommended. For sampling
84
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.3 Real-Time VPP Implementation

0.3
0.35
0.4
0.45
0.5
0.95 0.975 1 1.025 1.05
V
A
/V
Aref
[-]
z
/
z
m
a
s
t
[
-
]TWA 40
TWA 160
Measured
Modelled
=40
=30
=25
=20
=35
=5
=10
=0
=15
Figure 4.7: Reduction in apparent wind speed
(V
times of this order the fluctuations in the force and moment measurements are significant
which could make the VPP unstable and the results unreliable. A moving average for a
number of measurements is therefore calculated to smooth the data. Due to the short sampling
time the reference pressure measured by the pitot tube may not correspond to the pressure at
the test section because of the time taken by the air to travel downstream. Based on the time
delay an earlier reference pressure measurement can be assigned to each force and moment
measurement when calculating the coefficients. The moving average of the force and moment
coefficient vectors for measurement j (C
FTj
and C
MTj
) in coordinate system T is calculated
from several force and moment measurements (F
Ti
and M
Ti
) with
N
F
C =

=
k n
q n A
j
n j i k i effm
Ti
Sm
FTj
, ,
1
) (
,
(4.17)
N
M
C =

=
k n
q n A
j
n j i k i effm
Ti
Sm
MTj
, ,
1
) (
5 . 1
,
(4.18)
where n is the number of samples in the moving average and k is the number of samples the
reference pressure reading is delayed. n is equal to the data influence time divided by the
sampling time rounded to the nearest integer, which are both specified in the GUI (see Table
4.1). Similarly k is the pressure delay divided by the sampling time rounded to the nearest
integer. The pressure delay needs to be shorter than the data influence time and ideally the
sampling time should be chosen by the operator to be a divisor of the pressure delay. It has
been found difficult to experimentally determine the pressure delay. With calculated values
based on the flow speed and the distance of the pitot tube from the test section, no significant
reduction in force fluctuations could be observed by including the pressure delay.
A
/V ) with height (z/z
Aref mast
) modelled
in Real-Time VPP and calculated
from vertical speed profiles
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 85
Chapter 4 Real-Time Velocity Prediction Program
For sharing the aerodynamic force and moment coefficients and other information with
FRIENDSHIP-Equilibrium shared memory is allocated that can be accessed by both
programs. The programming language C++ was used to implement the shared memory
allocation as FRIENDSHIP-Equilibrium is written in C++, and the code can be compiled into
LabVIEW applications through a Code Interface Node as described in the user manual
(National Instruments, 2003). Both programs are started separately and run independently of
each other. The shared memory is allocated by whichever application is started first. Each
time the LabVIEW application has obtained a wind tunnel measurement the data is written to
the shared memory and at the same time the result data from the VPP that is currently stored
in the shared memory is read and displayed. The LabVIEW application does not wait for the
VPP to return the solution to the current data before making the next measurement, which is
more efficient because LabVIEW acquires data at the same time as the VPP predicts the
performance. On the other hand the VPP predictions shown in the GUI do not correspond to
the current measurement but to an earlier measurement. It can also happen that no VPP
solution is calculated for a measurement if the data is replaced before the VPP reads the
shared memory. Similarly the VPP may calculate a solution for the same data twice. In the
trimming mode, with short sampling times of around 0.2 seconds, this method works well
because the wind tunnel data is constantly streamed to the VPP and the results are displayed
as soon as they are available. With the wind tunnel data acquisition computer, which has the
processing power of a standard personal computer, a Real-Time VPP solution typically takes
significantly less than 0.5 seconds to compute so that the response time is very good and not
noticeable when trimming the sails. The time delay from the moving average and due to flow
effects is much more significant. In the acquisition mode with much longer sampling times it
is however important to store the VPP prediction that corresponds to the acquired data. Each
data set that is written to the shared memory therefore has an identifier and in the acquisition
mode a data set is not replaced until the VPP has returned the solution of that measurement.
The wind tunnel forces and moments are passed to FRIENDSHIP-Equilibrium in coefficient
form so that the full-scale wind speed used in the VPP is independent of the tunnel speed.
The wind tunnel speed is selected to give a sail shape and behaviour that is similar to full-
scale. Since the full-scale sail stretch and mast defection are not modelled in the wind tunnel
and the Reynolds Number is not matched, the tunnel speed is not directly related to the full-
scale wind strength. The tunnel speed may be adjusted to reflect different inflation levels of a
spinnaker in different full-scale wind strength but it cannot be directly scaled to a
corresponding wind speed for the VPP calculations. In addition the tunnel speed
correspondents to the apparent wind speed (V
A
) and not the true wind speed (V
T
) used in
VPPs as an input parameter. V
T
is used as input because it represents the physical condition of
a yacht sailing in a constant true wind, which as an idealisation is true. Specifying V
A
as a
constant input on the other hand means that V changes when the sails are trimmed which is
T
86
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.3 Real-Time VPP Implementation
clearly not the case in reality. V
T
is therefore also used in the Real-Time VPP as an input
parameter. It is specified in the GUI of the LabVIEW application and passed to
FRIENDSHIP-Equilibrium so that it can be changed easily without switching to the
FRIENDSHIP-Equilibrium interface. The true wind angle (
T
) is also of physical importance
when predicting the yacht performance (as will be discussed in section 4.4) so that it is
specified in the GUI and passed to the VPP as well. The apparent wind angle of the model
(
Am
) is a very important parameter to specify the test conditions because the measured forces
and moments strongly depend on it. It is for example used in equation to calculate q (2.40)
effm

but it is also passed to the VPP. Since the turntable angle is currently not recorded by
LabVIEW, is manually specified in the GUI.
Am
The heel angle (
m
) is of similar importance to specify the model orientation in LabVIEW
and FRIENDSHIP-Equilibrium. Three methods have been implemented to define
m
. The
heel angle can be manually adjusted in the wind tunnel and
m
specified in the GUI (see
Table 4.1). The second method still specifies
m
in the GUI but the model is automatically
heeled to the desired angle by using the heel driver LabVIEW application, which is part of
the SailView data acquisition system developed at the TFWT. LabVIEW is multi-threading
capable and the heel driver runs simultaneously with the Real-Time VPP LabVIEW
application. The heel driver acquires low pass filtered voltages of an accelerometer attached
to the model through the A/D card as illustrated in Figure 4.5 and calculates
m
. Through the
parallel port a controller is instructed to start the electric motor that changes the heel angle.
Once the specified
m
equals the measured
m
the controller is instructed to stop the motor.
The third method does not specify
m
in the GUI but the full-scale heel angle () predicted by
FRIENDSHIP-Equilibrium becomes the target angle for the heel driver so that the model is
dynamically heeled. Due to possible inertia effects the forces and moments are not measured
while the model rotates. For which are more than 5 different from
m
, the heel driver is
programmed to only rotate the model 5 towards to reduce the time delay between
measurements. The next force and moment measurement is then acquired before the model is
heeled further towards the updated prediction of . This iterative process provides an
adequate response time but it could be improved in the future by increasing the rotational
speed or investigating the possibility of acquiring data while the model rotates.
The values for , and
T Am m
are specified in the GUI as absolute values to avoid confusion
and the tack for which the VPP calculations are to be carried out is specified by a push
button. The values used by the VPP are shown in the box to the right of the VPP input box
in the GUI (see Figure 4.6) and follow the sign convention in Figure 2.19. The tack for the
VPP calculations is arbitrary but the tack of the model needs to be specified by another push
button in accordance with the physical set-up in the wind tunnel. If the tack for the VPP
calculations is different from the tack in the wind tunnel the sign of F
YT
, M and M
XT ZT
is
reversed.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 87
Chapter 4 Real-Time Velocity Prediction Program
When writing measurement data to the shared memory, the LabVIEW application also reads
the FRIENDSHIP-Equilibrium results from the shared memory. The data include the values
of the free variables, the resulting values of and V
eff
,
T eff
, and force and moment related
data, which consists of the aerodynamic force and moment coefficients, the full-scale
aerodynamic forces and moments, and the residual forces and moments from the VPP
solution. Boat speed (V
S
) and heel angle () are the most important free variables to monitor
during the trimming process so that they are prominently displayed numerically and
graphically for easy viewing. Measurements obtained in the acquisition mode can be saved to
file. The data written to file includes the forces, moments and pressure measured in the wind
tunnel, other GUI input information and the VPP results including the residuals.
4.3.2 FRIENDSHIP-Equilibrium
FRIENDSHIP-Equilibrium is now developed by FRIENDSHIP SYSTEMS. It was initiated at
the Technical University Berlin by Hochkirch (2000) as a research tool for the Dyna project
and has evolved into a commercially available high-end performance prediction program
(PPP). Due to the close cooperation with FRIENDSHIP SYSTEMS the author had access to
large portions of the source code so that integration in the Real-Time VPP was more efficient.
FRIENDSHIP-Equilibrium is programmed in C++ and is platform independent. It is
primarily developed in Linux
13 14
but at the TFWT it is used in the Windows environment. A
command line and GUI version are available to allow for flexible use. Both versions can be
used in the Real-Time VPP but the GUI version was found to be more user-friendly for this
type of application.
In addition to solving for steady state equilibrium it is also possible to consider dynamic
effects with time-stepping so that manoeuvring can be modelled. In principle time-stepping
can be used in the Real-Time VPP but the original implementation described here and used in
this project utilises only the traditional quasi-static approach. The steady state equilibrium for
up to six degrees of freedom is solved by a Newton-Raphson solver. The six free variables
associated with the six degrees of freedom are ship speed (V
S
), heel angle (), leeway angle
(), rudder angle (), trim angle () and sinkage. In an outer loop the yacht speed is optimised
by varying the aerodynamic trim parameters and possibly other free variables, such as the
keel-canting angle, water ballast, trim tab angle, etc.. Since the trim parameters as described
in section 4.2.1 are not differentiable, especially at the limits, the Newton-Raphson algorithm
cannot be used and alternative strategies such as the Hooke-Jeeves-Algorithm or the Nelder-

13
Name of open source operating system trademarked by Linus Torvalds in the USA and other countries
14
Operating system developed by Microsoft Corporation, One Microsoft Way, Redmond, WA 98052-6399,
USA
88
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.3 Real-Time VPP Implementation
Simultaneous solution
through for example
Hooke-Jeeves-Algorithm
(not used in RT-VPP since
no need for r, f and t)
Simultaneous solution
through Newton-Raphson
Force modules
Winglets
S
e
m
i
-
e
m
p
i
r
i
c
a
l

m
a
t
h
e
m
a
t
i
c

m
o
d
e
l
s
Crew
Rudder
Keel
Hull
Rig (r, f, t)
Equilibrium condition
F
A
(F
XA
F
YA
F
ZA
)
M
A
(M
XA
M
YA
M
ZA
)
Free variables
Reef (r)
Flat (f)
Twist (t)
Input
True wind
speed (V
T
)
True wind
angle (
T
) or
Apparent wind
angle (
A
) or
Effective wind
angle (
eff
)
Wind tunnel
F
T
(F
XT
F
YT
F
ZT
)
M
T
(M
XT
M
YT
M
ZT
)
Free variables
Yacht speed (V
S
)
Heel angle ()
Leeway angle ()
Rudder angle ()
Trim angle ()
Sink
Results
Yacht speed (V
S
)
Heel angle ()
Leeway angle ()
Rudder angle ()
Trim angle ()
Sink
Reef (r)
Flat (f)
Twist (t)
Rig (WT)
Condition
Optimise
speed
or

Figure 4.8: Schematic description of FRIENDSHIP-Equilibrium and the Real-Time VPP module for
wind tunnel testing
Mead simplex method are employed (FRIENDSHIP SYSTEMS, 2005). Figure 4.8 shows the
inner and outer loops with the associated free variables and conditional constraints.
Due to its roots as a research tool, FRIENDSHIP-Equilibrium is designed in a modular way
and the forces acting on the yacht are abstracted by so-called force modules which can
easily be added or removed and combined in any number of ways so that the user has a lot of
flexibility when modelling a sailing yacht. Figure 4.8 shows how FRIENDSHIP-Equilibrium
can be used with either a semi-empirical rig force module Rig (r,f,t) that uses trim
parameters as described in section 4.2.1 or with the newly developed wind tunnel force
module Rig (WT) of the Real-Time VPP that utilises the forces and moments measured
with the LabVIEW application. All other force modules are unaffected by using
FRIENDSHIP-Equilibrium as part of the Real-Time VPP. On the top right of the GUI in
Figure 4.9 the loaded force modules are listed with the WindTunnel module in use and the
semi-empirical GenericRig force module with trim parameters inactive. The hydrodynamic
force modelling is not investigated as part of this work and therefore is not discussed here. A
brief summary of the hydrodynamic force modules and input data is given in appendix C.4
and more information on FRIENDSHIP-Equilibrium is given in the user manual
(FRIENDSHIP SYSTEMS, 2005).
Implementing the new WindTunnel force module is straight forward since force modules
can be defined as plug-ins in FRIENDSHIP-Equilibrium by compiling them as Dynamic Link
Libraries (DLLs) so that the main program does not need to be recompiled and the full source
code is not required. (The mathematical transformations used in the WindTunnel force
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 89
Chapter 4 Real-Time Velocity Prediction Program
Figure 4.9: Screen shot of the graphical user interface of FRIENDSHIP-Equilibrium
module to obtain the full-scale forces from the wind tunnel measurements will be described in
section 4.3.3). In addition to including the new force module it is also necessary to implement
a new simulation mode within FRIENDSHIP-Equilibrium. The Simulation Mode menu of
the Control Center on the left side of the GUI in Figure 4.9 shows the different simulation
modes with the new wind tunnel mode selected. The primary object calls of the new
simulation mode need to be included in the main executable program of FRIENDSHIP-
Equilibrium but all object definitions are compiled into the wind tunnel DLL so that the main
executable does not need to be re-compiled for every alteration during the development phase
and because the object definitions are only required when the WindTunnel force module is
loaded.
One of the main conceptual differences of the wind tunnel mode is that the VPP solves for
equilibrium constantly, once started until it is manually stopped, using the test condition input
specified in the LabVIEW application, instead of conducting a set number of calculations
based on the cycling range of V and
T T
. The second main difference of the wind tunnel
mode, and one of the main advantages of the Real-Time VPP, is that the WindTunnel force
module does not require trim parameters as the trimming of the sails is conducted by the
operator in the wind tunnel. Since trimming the sails with the Real-Time VPP is intended to
replicate full-scale it is also assumed that the operator decides the values of possible
90
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.3 Real-Time VPP Implementation
additional optimisation variables, such as canting keel angle or crew position, so that the
outer optimisation loop shown in Figure 4.8 is disabled. The maximum number of iterations
of the Newton-Raphson algorithm in the inner loop is usually set to one, which in most cases
is sufficient to find a solution and if not, a solution is obtained for the next set of
measurements since the results are similar. In the wind tunnel mode the predicted yacht
performance, in the form of the free variables and the test condition parameters, are written to
the shared memory so that they can be displayed and stored in the LabVIEW application.
A new general feature has also been added to all steady state simulation modes of
FRIENDSHIP-Equilibrium as part of the Real-Time VPP development. As with most VPPs,
FRIENDSHIP-Equilibrium uses the true wind angle (
T
) as an input parameter to define the
test condition because of its physical importance. The option of specifying either
A
or
eff

instead of as a constant input parameter has been added (see Figure 4.8
T
), which is
important to address the wind tunnel constraints that will be discussed in section 4.4. For the
implementation it needs to be remembered that
A
and
eff
should be defined at the centre of
sail area rather than at the typical reference height of 10 metres and both heights can now be
defined under Optional Settings in FRIENDSHIP-Equilibrium. For programming reasons
and due to the general nature of these parameters, they are not defined in the aerodynamic
force modules.
4.3.3 Aerodynamic Wind Tunnel Force Model
The mathematical model of the WindTunnel force module is very basic as the forces and
moments are not semi-empirically modelled but simply calculated from the wind tunnel
measurements. The force and moment coefficient vectors (C and C
FT MT
) in the tunnel
coordinate system (coordinate system T) are calculated by the LabVIEW application and are
read from the shared memory. Coordinate system T is aligned with the absolute coordinate
system (coordinate system A) so that the transformation matrix T
AB
defined in equation (2.17)
can be used to transform C and C
FT MT
to the body fixed coordinate system of the yacht
(coordinate system B) by rotating through the heel angle of the model (
m
) and translating
through the distance between the coordinate systems origins ( ) so that BT
FT m m AB F
C T C ) 0 , ( = = ,
(4.19)

+ = =
FT
Sm
MT m m AB M
A
C
BT
C T C
5 . 0
) 0 , ( ,
(4.20)
where C
F
and C
M
are the force and moment coefficient vectors in the heeled z-plane
(coordinate system B) and A
Sm
is the reference sail area of the model. Both
m
and A
Sm
are
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 91
Chapter 4 Real-Time Velocity Prediction Program
specified in the LabVIEW application and read from the shared memory. BT is calculated
with equation (2.36) based on the distance between the wind tunnel and upright body fixed
coordinate system origins and the height of the heel axis of the model, which are specified in
the WindTunnel force module. The input data required for the WindTunnel force module
is shown in the pop-up window in the centre of the GUI in Figure 4.9 and discussed in
appendix C.3.1.
C and C
L D
are calculated from C and C with equation (2.42)
Fx Fy
so that they can be displayed.
C and C
L D
are equivalent to the coefficients obtained by the semi-empirical model from
equations (4.1) and (4.5). Assuming the effective angle theory holds, these coefficients are
independent of the heel angle. Tests can hence be conducted at any heel angle but the
influence of the effective angle theory assumptions is reduced, as the heel angle in the tunnel
(
m
) gets closer to the heel angle calculated by the Real-Time VPP (). If the wind tunnel
model is dynamically heeled to the calculated angle, the effective angle theory has no
influence on the results.
For calculating the full-scale forces C and C
L D
are transformed back to C
Fx
and C
Fy
with
equation . C (4.6)
Fz
is traditionally not modelled in semi-empirical aerodynamic force models
but chapter 3 showed that it should not be ignored, and, as described in section 3.8, is now
modelled by the semi-empirical model used here. C
Fz
measured in the wind tunnel is
therefore also used when calculating the full-scale forces. The forces (F
A
) and moments (M
A
)
in the horizontal absolute coordinate system, which are used by the solver to find the
equilibrium condition, can then be calculated with
( )
F BA eff
air
S A
V A C T F ) , (
2
2

= ,
(4.21)
( )
M BA eff
air
S A
V A C T M ) , (
2
2 5 . 1

= .
(4.22)
where A is the reference sail area of the yacht as in the semi-empirical model, V
S eff
is the
effective wind speed calculated with equation , T (2.14) or (2.19)
BA
is the transformation
matrix defined in equation (2.37) and and are the heel angle and pitch angle calculated by
FRIENDSHIP-Equilibrium.
When the model is dynamically heeled so that
m
equals , the rotations with T
AB
and T
BA

cancel out (if the pitch angle of the yacht is ignored), which means that the wind tunnel
coefficients in the absolute coordinate system could be used directly for calculating the full-
scale forces and moments in coordinate system A. The rotation to the body fixed coordinate
system in the calculations is still important since
m
does not equal in many situations while
the model is being dynamically heeled to match the angles and when the model is not
92
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.4 Wind Tunnel Constraints
dynamically heeled. In this approach the model scale is implicitly specified by entering the
model and full-scale sail areas. This way the force and moment coefficients are defined in the
common manner. It is equally possible to specify a unit area of one in the LabVIEW
application as the model sail area and enter the full-scale sail area in the WindTunnel force
module as the model scale squared. This changes the coefficient values but not the full-scale
forces and the performance of the yacht.
4.4 Wind Tunnel Constraints
In the wind tunnel, measurements are traditionally made for different apparent wind angles
and heel angles. Combining these angles means that measurements are made for different
effective angles. Semi-empirical VPPs define the aerodynamic forces in terms of effective
angles but on the other hand maximise the speed of the yacht for true wind angles. For a
constant true wind speed and angle, the effective angle changes as the speed of the yacht and
the heel angle vary while the solution is optimised. For a constant true wind speed and angle,
the effective angle in the Real-Time VPP therefore also changes while the sails are trimmed
in the wind tunnel.
Ideally the apparent wind angle in the wind tunnel should be adjusted dynamically together
with the heel angle so that the correct effective angle can be achieved when trimming the
sails to maximise the boat speed for a specified true wind angle. Implementing the automatic
control of the turntable is however not seen as a feasible part of this project. In addition to the
challenges associated with the automation, inertia effects due to rotating the force balance
and changes in the zero voltages would need to be considered. For the time being an
alternative method to dynamically changing the apparent wind angle is therefore required.
4.4.1 Maximising Speed for Constant Effective Wind Angle
It is important to maximise the speed of a yacht with respect to the true wind speed because it
is the physical condition that the yacht operates in. Usually the speed is also maximised with
respect to the true wind angle since this is also an important real life parameter when sailing.
A yacht is to sail from one point to another as fast as possible in a certain wind strength.
From a sail design point of view it can however be argued that sails are designed for specific
effective angles since they relate best to the aerodynamic behaviour. Hence optimising the
yachts speed with respect to the true wind speed and the effective angle (
eff
) might be
possible for the Real-Time VPP, if the model is not dynamically heeled and
eff
is constant in
the wind tunnel. In order for this method to be meaningful the results must translate to the
maximum speed with respect to the true wind angle so that they describe the real life
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 93
Chapter 4 Real-Time Velocity Prediction Program
situation. This question cannot easily be answered and FRIENDSHIP-Equilibrium has been
modified to also predict the performance for constant effective wind angles instead of
constant true wind angle so that this can be assessed further by sample calculations.
The validity of using the effective angle to optimise the speed of the yacht is investigated
using the semi-empirical aerodynamic model where trim parameters provide the effect of
trimming the sails. This allows a more systematic variation of the sail trim compared to
using the Real-Time VPP. The VPP calculations are carried out for Dyna with the standard
hydrodynamic model described in appendix C.4 and the semi-empirical aerodynamic force
model with trim parameters shown in section 4.2.1. The force equilibrium is solved in six
degrees of freedom and the boat speed optimised using reef and flat as trim parameters for
four true wind speeds (V
T
) and a range of wind angles. Figure 4.10 shows V
S
optimised using
reef (r) and flat (f) and plotted against between 25 and 135. One set of curves shows V
T S

optimised for a range of constant . The second set of curves is generated by optimising V
T S

for a series of constant . For V
eff
and plotting the results against
T T
of 4m/s the sails are not
depowered by either optimisation method and the same boat speed is therefore calculated by
both methods since the same aerodynamic forces are used. For the higher true wind speeds
the sails are depowered at some wind angles. As long as the same values for the trim
parameters are chosen by the optimisation methods the two sets of curves still lie on top of
each other. Figure 4.10 however shows that the method optimising for constant
eff
predicts a
lower boat speed for below about 70. The reduction in boat speed is shown again in
T
Figure 4.11 as the ratio of V
S
from constant
eff
over V from constant
S T
. It can be seen that
V less than 45 V is reduced by up to 3% for between 45 and 70. For reduces
S T T S
0.92
0.94
0.96
0.98
1
1.02
1.04
0 15 30 45 60 75 90 105 120 135

T
[]
V
S

/
V
S
(
T
W
A
)

[
-
]
VT=10m/s EWA VT=6m/s EWA
VT=8m/s EWA VT=4m/s EWA
2
2.5
3
3.5
4
4.5
5
0 15 30 45 60 75 90 105 120 135

T
[]
V
S

[
m
/
s
]
VT=10m/s TWA VT=6m/s TWA
VT=10m/s EWA VT=6m/s EWA
VT=8m/s TWA VT=4m/s TWA
VT=8m/s EWA VT=4m/s EWA
Figure 4.10: V
S
optimised using r and f for either
constant
Figure 4.11: V
S
ratio of optimisation for constant
at different V at different V or against constant
T eff T eff T T
94
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.4 Wind Tunnel Constraints
significantly by up to 8%. The reduction in V
S
is generally similar for the different wind
speeds. The reduction in V at V is however less severe for very small
S T T
of 6m/s compared
to the higher true wind speeds but it should be remembered that
T
less than 40 corresponds
to unrealistic sailing conditions for this yacht.
For a true wind speed of 12m/s Figure 4.12 shows in more detail the range of true wind
angles where the speed optimised for
eff
is lower than the maximum speed for
T
. For

eff
=40 and 65 a set of achievable V
S
is shown. The points are obtained by varying the trim
parameters reef and flat systematically in increments of 0.05 and plotting the resulting boat
speeds against
T
. Reef is adjusted between 0.6-1 and flat between 0.75-1. One point in each
set results in the maximum V
s
for that
eff
and represents one point of the V
S
curve optimised
for constant
eff
. Similarly sets of achievable speeds for constant
T
could be plotted. In the
graph of V against
S T
the points in each set would simply form a vertical line and are
therefore not shown. Only the resulting curve of V optimised for is shown in Figure 4.12
S T
.
It represents the best performance and therefore best sail trim for this wind strength. At

eff
=65 the boat speed optimised at constant
eff
falls onto the curve of V
S
optimised at
constant . At this angle the best sail trim can therefore be found by optimising V
T S
at
constant
eff
. On the other hand at
eff
=40, the maximum V
S
obtained from optimising at
constant
eff
is lower than the boat speed obtained when optimising at constant
T
. It can be
seen in Figure 4.12 that the VPP correctly picks the maximum boat speed for
eff
=40 as a
point on the curve optimised at constant
eff
. However, a different sail trim should be picked
to give the best performance. The maximum V at
S eff
=40 therefore does not represent the
best sail trim for this condition. It can be envisioned why this is the case. With the effective
angle theory the lift and drag coefficients stay the same for any heel angle at a constant
3.8
3.9
4
4.1
4.2
4.3
4.4
55 60 65 70 75 80 85 90 95

T
[]
V
S

[
m
/
s
]
Optimised for TWA
Optimised for EWA
Vs at EWA 40
Vs at EWA 65
Best V
s
at
eff
= 40
Best trim and V
s
at
eff
= 65
Best trim at
eff
= 40
Figure 4.13: Polar plot of V Figure 4.12: Set of achievable V
S
at
eff
=40 and
65 for V
S
optimised at
V
T
=12m/s using reef and flat
between 1 and 0.6 in increments of
0.05
T
=12m/s using reef and flat for either
constant (left) or
T eff
(right)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 95
Chapter 4 Real-Time Velocity Prediction Program
effective angle. As the yacht heels more the lift and drag reduce since the effective wind
speed reduces so that the drive force and side force are reduced. The side force in the
horizontal plane further reduces with heel angle since a smaller portion of the side force acts
in the horizontal plane. The reduction in side force is therefore much more significant than
the reduction in drive force so that the yacht speed can increase with heel angle. The
reduction in side force leads to less induced drag from the appendages and hull, which can
outweigh the drag increase due to heel and the reduction in drive force. The maximum boat
speed for a constant
eff
is reached at a large heel angle. The yacht however sails at a larger
T

now because of the large heel angle and constant
eff
. If the yacht speed is optimised for this
a higher boat speed can be achieved by reducing the heel angle and thereby increasing
T eff
,
which for small wind angles increases the drive force.
This shows why maximising V at constant
S eff
does not necessarily result in the optimum
performance. The range of for which maximising V for constant
T S eff
does not result in the
best trim depends on the aerodynamic, hydrodynamic and stability characteristics and V
T
, but
is most strongly influenced by . At
T T
where the heeling moment is important and changes
in result in large changes of
eff
, an under-prediction of V
S
can be expected. The left side of
the polar plot in Figure 4.13 shows curves of achievable speeds for constant
T
. The curves
are developed by systematically varying reef and flat as before for constant
T
from 10 to
180 in 5 increments. The curves are straight lines and if extended run through the origin
since each curve represents achievable V for a constant
S T
. The best performance curve is
obtained from the maximum V points from each constant
S T
line. Equivalently the right side
of the polar plot shows curves developed for constant
eff
from 10 to 150 in 5 increments
and plotted against . The curves are close to tangential to the best performance curve for
T T

between 30 and 85. It follows that the maximum V does not represent the best trim. For
S T

larger than 90 the curves are at a larger angle to the optimum performance curve and the
maximum V results in the best trim.
S
For this test configuration, it seems feasible to maximise V with respect to a constant
S eff
for
downwind sail testing. On the other hand for many upwind and reaching conditions this
method incurs errors. It should be clarified again that the differences in predicted boat speed
at some angles result from the fact that the aerodynamic forces are changed when trimming
the sails to optimise the performance. For a fixed set of aerodynamic forces the results from
both methods yield the same yacht performance.
4.4.2 Maximising Speed for Constant Apparent Wind Angle
As discussed earlier in section 4.4 it is important that the effective wind angle (
eff
) used by
the VPP and in the wind tunnel are identical. With the Real-Time VPP it is possible to
96
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.4 Wind Tunnel Constraints
dynamically heel the model to the heel angle calculated by the VPP so that the same
eff

results in the same apparent wind angle (
A
) in the VPP and the wind tunnel. The pitch angle
calculated in the VPP is ignored here since its influence on the results is negligible. It is
therefore feasible to maximise the boat speed for a true wind speed and a constant
A
. As
before, for the results to be meaningful the optimised boat speed needs to correspond to the
performance predicted by the standard method of optimising at a constant
T
. FRIENDSHIP-
Equilibrium is again modified so that the performance can also be optimised for constant
A
.
The same VPP calculations, as in section 4.4.1, are carried out but this time optimising the
yacht speed for constant
A
. Unlike Figure 4.11, the results in Figure 4.14 now show an
almost identical boat speed compared to the optimisation at constant
T
. The maximum
difference in boat speed is about 0.3%, but typically it is much less. Compared to the
significant differences shown in Figure 4.11 this method of optimising for constant
A

provides much better results.
Like Figure 4.12, Figure 4.15 shows in detail two sets of achievable V at
S A
of 40 and 65
obtained by systematically varying reef (r) and flat (f) for V of 10m/s. For both
T A
the
maximum boat speed corresponds to the best sail trim and the curves of V
S
optimised for
constant are therefore almost identical. Compared to the achievable V
A
and constant
T S
at

eff
=40 in Figure 4.12, V at
S A
=40 are spread over a much narrower band of
T
. For a
constant V is only a function of V the relationship between
T A
and
T S
and the leeway angle
() as shown in equation . is generally small and V (2.9)
S
does not vary more than about
15% when trimming the sails in this example so that the changes in
T
are not very large. The
relationship between
eff
and
T
is additionally also a function of heel angle (and pitch angle,
which is however insignificant) as shown in equation (2.13) so that the changes in
T
can be
much larger when trimming the sails. The case for
eff
=40 in Figure 4.12 is a good example
3.4
3.5
3.6
3.7
3.8
3.9
4
4.1
4.2
50 55 60 65 70 75 80 85 90

T
[]
V
S

[
m
/
s
]
Optimsed for TWA
Optimsed for AWA
Vs at AWA 40
Vs at AWA 65
Best trim and V
s
at
A
= 65
Best trim and V
s
at
A
= 40
0.96
0.97
0.98
0.99
1
1.01
1.02
1.03
1.04
0 15 30 45 60 75 90 105 120 135

T
[]
V
S
/
V
S
(
T
W
A
)

[
-
]
VT=10m/s AWA
VT=8m/s AWA
VT=6m/s AWA
VT=4m/s AWA
Figure 4.14: V
S
ratio of optimisation for constant

Figure 4.15: Set of achievable V


S
at
A
=40 and 65
for V at different V
A
against constant
T T T
=10m/s using r and f between 1
and 0.6 in increments of 0.05
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 97
Chapter 4 Real-Time Velocity Prediction Program
and the resulting effects when optimising V for large changes in
T S
. For this type of boat the
changes in are sufficiently small when optimising V at a constant
T S A
to obtain the best sail
trim with this method. Generally it can be expected that this is also true for other boat types
and trimming the sails for maximum speed at a constant apparent wind angle is a feasible
method. If the model is dynamically heeled so that the apparent wind angle in the VPP and
the wind tunnel are identical, this method should be used.
4.4.3 Effective Angle Correction
In situations where the model is not dynamically heeled an alternative method is needed since
section 4.4.1 showed that optimising V for constant
S eff
does not produce accurate results in
certain conditions. The only other alternative is to maximise Vs with respect to the true wind
speed and the true wind angle. With these two parameters fixed the effective wind angle
changes as the sails are trimmed. The apparent wind angle in the wind tunnel (
Am
) and the
model heel angle (
m
) can be selected so that the resulting effective angle in the tunnel (
effm
)
is close to the
eff
calculated by the Real-Time VPP for this condition. C and C
L D
are directly
dependent on the effective angle and need to be corrected for the difference between
effm
and

eff
. If the corrections are small, theoretical methods or generic C
Lopt
and C versus
Dopt eff
curves can be used to approximate the lift and drag curve slope at
eff
. The corrected lift
coefficient is then obtained from the C measured in the tunnel at
L effm
, the optimum lift
coefficient curve slope and the change in effective angle (
eff
) such that
eff
eff
Lopt
at L at L
C
C C
effm eff



d
d
+ = .
(4.23)
For larger and more accurate corrections the C
Lopt
and C versus
Dopt eff
curves should be
developed first for each sail combination and then used to make corrections. Equation (4.23)
is then rewritten to obtain the corrected C from the C
L L
measured in the tunnel and the
optimum lift coefficient curve C
Lopt
(
eff
) as
) ( ) (
effm Lopt eff Lopt at L at L
C C C C
effm eff


+ = .
(4.24)
C
D
can be corrected in a similar manner. A correction for C and C
L D
of the form shown in
equation is implemented in the WindTunnel force module. C and C (4.24)
L D
versus
eff

curves become input parameters as shown in the pop-up window in the centre of Figure 4.9,
the GUI of FRIENDSHIP-Equilibrium, and as listed in appendix C.3.1. At present corrections
are only made for C and C
L D
since they are most influenced by
eff
and because adjustments
to the data are to be kept to a minimum. However, in principle similar corrections can be
applied to the moment coefficients and C as well.
Fz
98
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.5 Real-Time VPP Applications
The drawback of this method is that corrections need to be applied to the measurements. The
corrections can be kept relatively small but do not account for changes of the C and C
L D

curves slopes due to depowering and changes in the moments. One of the main aims of the
Real-Time VPP is to investigate the aerodynamic sail forces without the need for semi-
empirical descriptions. Additional testing may also be required if C
Lopt
and C
Dopt
curves need
to be developed to ensure an accurate correction. Nevertheless for upwind testing where the
model is not dynamically heeled this method is currently the most feasible solution.
4.5 Real-Time VPP Applications
There are two main applications for the Real-Time VPP. It is an effective tool for the
comparative testing of sails since it makes the process of trimming and depowering the sails
in the wind tunnel much more similar to the real life situation, especially when the model is
dynamically heeled. When testing sails in the wind tunnel good trimming is crucial and
designers often bring professional sailors to the wind tunnel to help with trimming the sails.
For the sailors it is a great advantage if the process of trimming the sails in the wind tunnel is
similar to the real life process to which they are familiar. The sails can be tested at the correct
heel angle with the Real-Time VPP, which previously was not possible. In an investigation
into the heel effects on downwind sails by Le Pelley and Hansen (2003), the Real-Time VPP
has been shown to significantly improve the ease and efficiency of trimming the sails and
highlighted the importance of considering the heel angle of the model in the wind tunnel.
When using the Real-Time VPP, different sails can be directly compared with each other
without any post-processing of the data. Additionally, depowering is accounted for without
having to use trim parameters or conducting a systematic depowering exercise. This has
proven particularly useful in recent sail testing for the Volvo Open 70 class. Since the testing
was for the first generation of this stability-driven class and the expected speeds were
significantly higher than for existing monohulls, the design space for the sails was largely
unknown and the Real-Time VPP allowed an efficient comparison of a large number of
design ideas. With the tight restriction placed on the number of sails that can be carried on
board the yacht for each leg it was also of particular importance to assess the performance of
the sails for depowered trims to determine the efficient operational range and crossover
points. The Real-Time VPP proved invaluable in this respect. The use of the Real-Time VPP
requires more preparation and data from the designer before the wind tunnel testing in order
to set-up the hydrodynamic model of the yacht in FRIENDSHIP-Equilibrium. However, less
post-processing is necessary, which would otherwise also require modelling the yacht in a
VPP. Ideally FRIENDSHIP-Equilibrium is not only used in the wind tunnel as part of the
Real-Time VPP but also for standard VPP calculations so that the same hydrodynamic model
of the yacht can be used.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 99
Chapter 4 Real-Time Velocity Prediction Program

(a) Model with sail twisted off, less V (b) Model with sail fully powered-up, more V
Figure 4.16 shows an example of the Real-Time VPP being used as a trimming tool. The
wind tunnel model of Dyna is set up in an upwind sailing condition with mainsail and jib.
Figure 4.16(a) shows the model with the mainsail twisted off and the Real-Time VPP
displaying a predicted V
S
of 5.33m/s and heel angle () of 8.9 for the full-scale yacht. The
model is dynamically heeled to the predicted . The mainsail is then trimmed on and Figure
4.16(b) shows the model with the mainsail fully powered-up. The predicted V
S
is now
6.12m/s and is 19.8, to which the model has been dynamically heeled. This example shows
how the Real-Time VPP clearly relates the wind tunnel measurements to the full-scale
performance and creates a more realistic wind tunnel test experience.
The second application of the Real-Time VPP is as a research tool to investigate the ways
aerodynamic forces are modelled in VPPs and in the context of this project it was the main
reason for developing the Real-Time VPP. Since the Real-Time VPP does not rely on trim
parameters it can be used to trim the sails as efficiently as possible and compare the results to
trim parameter model predictions. Initial wind tunnel tests using the Real-Time VPP have
been carried out with the wind tunnel model of Dyna and published by Hansen et al. (2003b)
to show how Real-Time VPP results compare to predictions using a semi-empirical VPP with
the trim parameters reef and flat. This preliminary study showed that using the Real-Time
VPP for this application is feasible. The Real-Time VPP is therefore used in chapter 5 to
assess the existing trim parameter model, suggest improvements and develop a new
depowering model.
S
&
S
&
Figure 4.16: The Real-Time VPP predictions (left value is V
S
and right value is heel angle ()) for two sail
trims with the model dynamically heeled to the predicted
100
YACHT
RESEARCH
UNIT
The
University
of Auckland
4.6 Conclusions
4.6 Conclusions
The Real-Time VPP for wind tunnel testing of sailing yachts has been developed and
successfully implemented. The integration in the modular structure of FRIENDSHIP-
Equilibrium results in a flexible application package and the use of LabVIEW as the data
acquisition software and GUI produces a user-friendly implementation that is part of a whole
suite of data acquisition programs developed at the TFWT. The Real-Time VPP calculates the
yacht performance based on the aerodynamic forces measured in the wind tunnel while the
sails are trimmed and the model can be dynamically heeled to the predicted heel angle so that
wind tunnel tests are now conducted at the correct heel angle. This allows the sails in the
wind tunnel to be trimmed based on boat speed instead of the measured forces, which makes
the process of trimming and depowering much more similar to the real life situation on the
water, especially if the model is also dynamically heeled.
Since the apparent wind angle cannot be dynamically adjusted yet in the wind tunnel,
alternative solutions to overcome this constraint have been developed. For tests where the
model is not dynamically heeled the effective wind angle in the wind tunnel is constant. The
concept of optimising boat speed for constant effective angles seems feasible for downwind
testing but introduces errors for tests in upwind and reaching conditions. The boat speed in
upwind and reaching tests should therefore be optimised for constant true wind angles. The
constraint to a fixed effective wind angle in the wind tunnel requires an effective angle
correction to allow the optimisation at constant true wind angles. If dynamically heeled the
apparent wind angle in the wind tunnel is constant and it has been shown that optimising the
boat speed for constant apparent wind angles is feasible and does not introduce errors. It is
therefore recommended that the model be dynamically heeled where possible so that the
effective angle correction does not need to be applied. In addition dynamically heeling the
model has the important advantage of conducting the wind tunnel tests at the correct heel
angle.
The Real-Time VPP has proven to be a valuable trimming tool when evaluating different sail
designs. It allows the sails to be trimmed and compared based on boat speed without further
post-processing the results. The wind tunnel testing process is now much more similar to the
real life situation on the water. With the Real-Time VPP it is also possible to investigate
aerodynamic forces in a more direct way than was previously feasible since trim parameters
are no longer required. In chapter 5 the Real-Time VPP will be used to assess and improve
the aerodynamic sail force modelling in semi-empirical VPPs. In the future Real-Time VPP
measurements without dynamically heeling the model could potentially be improved by
dynamically adjusting the turntable angle, but due to some challenges this may prove
unfeasible. Another future development of the Real-Time VPP could be dynamic
investigations of aerodynamic sail forces in the wind tunnel. FRIENDSHIP-Equilibrium has
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 101
Chapter 4 Real-Time Velocity Prediction Program
the capability of calculating non-stationary solutions and this functionality can already be
used in association with the Real-Time VPP. The main remaining challenge is to measure
meaningful dynamic forces in the wind tunnel, as the force balance was originally developed
to measure static loads.
102
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
5 Aerodynamic Force Modelling
5.1 Introduction
In velocity prediction programs (VPPs) the semi-empirical modelling of the forces acting on
the yacht is crucial for an accurate performance prediction. The standard aerodynamic force
model as described in section 4.2.1 is based on fully powered-up lift and drag coefficients and
the centre of effort position, which are changed by trim parameters to account for depowering
of the sails. Although many people agree that the commonly used trim parameters reef and
flat do not model the physics of depowering very well and hence do not produce results
which model the changes in the forces and moments accurately, they are still the most readily
used approach for general upwind VPP analysis. When investigating this model by Kerwin
(1978) and Hazen (1980), and developing an improved model there are two main aspects to
consider. Firstly it is important to look at how the forces and moments are to be calculated
from the input parameters and secondly how the depowering of the sails is to be modelled.
Previously one of the main challenges in assessing how well the trim parameters model the
physical behaviour of the sails is determining the best depowered trim without using trim
parameters. Achieving this by simply looking at the forces when trimming the sails in the
wind tunnel is very complex. Better results can be achieved by systematically changing sail
controls, post-processing the results in a VPP, and selecting the trim that gives the best
performance for each sailing condition. Due to the large number of possible sail controls this
is however a difficult and time consuming task. To make this process more efficient the Real-
Time VPP has been developed as described in chapter 4. The Real-Time VPP is used in this
chapter to investigate the depowering of sails in the wind tunnel, compare the results to VPP
predictions with trim parameters, and develop new aerodynamic force models.
5.2 Testing Procedure
Wind tunnel tests were conducted with the model of Dyna with the mainsail and jib at 0 heel
and the tunnel in upwind configuration, i.e. the upwind velocity profile and no vanes to twist
the flow. The model was not dynamically heeled in this study to reduce the number of
variables in the experimental set-up and because it is thought that the process of depowering
is not influenced significantly by changes in the heel angle. In the VPP calculations effective
angle theory (see section 2.3.1) was used as usual to account for the heeling of the yacht.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 103
Chapter 5 Aerodynamic Force Modelling
Tests were conducted for six apparent wind angles of 20, 25, 30, 40, 60 and 90. For
each apparent wind angle the sails were trimmed using the Real-Time VPP by maximising
the boat speed for the selected true wind angle. For each apparent wind angle the sails were
trimmed for true wind speeds between 2m/s and 14m/s. The trim for the true wind speeds of
2m/s and 4m/s represents the fully powered-up trim where the sails operate most efficiently.
For higher wind speeds a higher boat speed was achieved by depowering the sails at certain
apparent wind angles. With the Real-Time VPP these depowered trims can be obtained very
efficiently. The force equilibrium was solved in all six degrees of freedom when trimming the
sails in the wind tunnel. The VPP solved for a constant true wind angle and the effective
angle correction as described in section 4.4.3 was used. Solving for a constant apparent wind
angle would also require an effective angle correction in this study since the model is not
dynamically heeled. Therefore solving for a constant true wind angle was chosen since it best
represents the typical real life sailing situation. Once the best trim for a sailing condition had
been determined the forces were acquired for 90 seconds. The data is then time averaged and
used to predict the resulting yacht performance. After the test session the measurements and
the performance prediction for each run were adjusted for temperature drift and errors (due to
the effective angle correction) before the data was used for further analysis. The final zero
measurement taken after each set of runs was proportionally subtracted from each run based
on the sequence of runs assuming a linear temperature drift over time and that each run took
approximately the same amount of time. The yacht performance was then calculated again
from the adjusted forces, this time solving for a constant effective wind angle so that the
effective angle correction was no longer required. Since the aerodynamic force and moment
coefficients are already known, solving for a constant effective wind angle does not
encounter the problems discussed in section 4.4.1. It is assumed that the small force and
moment corrections due to temperature drift and inaccuracies in the effective angle correction
do not change the best sail trim obtained for a sailing condition. This procedure is the most
accurate approach for the current set-up in which the apparent wind angle cannot be adjusted
dynamically while the sails are trimmed and the model is not dynamically heeled.
The windage of the model is not subtracted from the wind tunnel measurements when using
the Real-Time VPP. As shown in chapter 3 modelling windage is not trivial and ultimately
windage components subtracted from wind tunnel measurements need to modelled by a VPP
force module. As long as the measurements are not to be used as generic input for different
yacht types, there is little to be gained from separating out the windage and adding it in again
later. Hence windage is initially not subtracted from the wind tunnel measurements in this
study and the Windage force module in FRIENDSHIP-Equilibrium is turned off.
104
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.3 Improved Aerodynamic Moment Model
5.3 Improved Aerodynamic Moment Model
Before even considering how to model the depowering of the sails, it is important to look at
how the forces and moments are to be calculated by the force module. The usual way of
achieving this is to calculate the forces from coefficient curves of lift and viscous or parasitic
drag against effective wind angle as described in section 2.3. The forces could similarly be
modelled by coefficient curves of C and C
Fx Fy
against effective wind angle, but the expression
in terms of C and C or C
L Dvis Dp
is aerodynamically more meaningful and easier to visualise
since it is more commonly used. In addition it is worth considering modelling the force acting
perpendicular to the deck plane as a function on
eff
and heel angle as discussed in detail in
chapter 3.
From the calculated forces and the centre of effort position the moments are then calculated.
However, determining the centre of effort from measurements requires some careful
consideration. As shown in section 2.3.3 and appendix D.1 simply defining the centre of
effort in the centreline plane of the yacht (y
CoE
=0) so that the longitudinal centre of effort
(x
CoE
) is
Y
Z
CoE
F
M
x = ,
(5.1)
and the height of the centre of effort (z
CoE
) is
Y
X
CoE
F
M
z

= ,
(5.2)
results in a residual pure moment about the y-axis (M ) of
Y 0
CoE X CoE Z Y Y
z F x F M M + =
0
.
(5.3)
From the fully powered-up tests conducted for this study it can be seen that M
0Y
increases
with effective wind angle as shown in Figure 5.1. The sail trims for the Real-Time VPP tests
at true wind speeds of 2 and 4m/s are the same since the sails are not depowered.
Nevertheless the results for both sets of the measurements are shown to indicate the
repeatability of the experimental set-up. The pitching moment (M ), shown in Figure 5.2
Y
, also
increases with effective wind angle, as is well known among sailors, and shows a similar
trend to M , which also increases with effective wind angle.
0Y
Most VPPs do not account for the pitching moment in the force equilibrium calculations so
that ignoring M
Y 0
does not affect the performance predictions. However, FRIENDSHIP-
Equilibrium and hence the Real-Time VPP are capable of including the pitching moment
balance in the equilibrium calculations and ignoring M affects the solution. For small
Y 0
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 105
Chapter 5 Aerodynamic Force Modelling
0
0.05
0.1
0.15
0.2
0.25
0.3
0 15 30 45 60 75 9

eff
[]
C
M
0
y

[
-
]
0
RT-VPP VT 4m/s
RT-VPP VT 2m/s
0
0.2
0.4
0.6
0.8
1
1.2
0 15 30 45 60 75

eff
[]
C
M
y

[
-
]
90
RT-VPP VT 4m/s
RT-VPP VT 2m/s
Figure 5.2: Pitching moment coefficient (C Figure 5.1: Residual pure moment coefficient
about y-axis (C
My
) for
fully powered-up sails against
effective wind angle (
M0y
) for fully powered-
up sails against )
eff eff
effective wind angles this approach may produce satisfactory results but for larger apparent
wind angles M should be considered. As described in section 2.3.3 and appendix D.2
Y 0
,
Hochkirch (2000) uses a central axis system approach to obtain the centre of effort. The
moment is represented by its component aligned with the force direction (M
F
) and its
component perpendicular to the force direction (M
V
). M
V
is used to determine the central
axis, which is aligned with the force direction and along which the force can be applied at any
point without affecting the moment. The centre of effort is then defined as the intersection of
the central axis with the centreline plane of the yacht (y-plane). This method correctly models
the moment perpendicular to the force direction but does not account for the moment aligned
with the force. M
F
cannot be expressed by applying F at a point in space because it is aligned
with F. In this method the moment from pure couples is now expressed in terms of M
F

instead of M in the traditional approach.
Y 0
For this definition of the centre of effort to produce accurate results M
F
needs to be
sufficiently small to be negligible. For a single point force there is no M
F
but if F is the
resultant force of distributed forces a M
F
component might be present. Sail forces generated
by a flow field result in a distributed force and a M
F
component can be present. Since M
F
is
aligned with F, its magnitude can be represented as a scalar fraction of F so that
, F M
5 . 0
S F
pA =
(5.4)
where p is the scalar fraction in non-dimensional form. Equation (D.12) defines p and in non-
dimensional form it can be written as
2
F
M F
F
Mf
p
C
C C
C
C

= = ,
(5.5)
106
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.3 Improved Aerodynamic Moment Model
-0.12
-0.1
-0.08
-0.06
-0.04
-0.02
0
0 15 30 45 60 75

eff
[]
p

[
-
]
where C
Mf
is the coefficient of the moment in line with the force. For the fully powered-up
measurements Figure 5.3 shows p against effective wind angle. It can be seen that p increases
in magnitude with
eff
. At
eff
of 90 p equals -0.106, which means that an anti clockwise
moment coefficient about the resultant force direction with a magnitude of 10.6% of the
resultant force coefficient is measured. This representation does however not put M
F
in
context with M. Figure 5.4 hence shows the magnitude of the total moment coefficient (|C
M
|)
and the components perpendicular to and in line with the resultant force (|C
Mv
| and |C
Mf
|). Due
to their geometrical relation that

2
2 2
Mf Mv M
C C C + = ,
(5.6)
even |C
Mf
| of 12% of |C
M
| at
eff
=90 only results in a reduction of |C
Mv
| compared to |C
M
| by
0.7%. It can hence be concluded that the magnitude of the moment is not affected
significantly by ignoring C
Mf
.
Ignoring M
F
however also changes the direction of the resultant moment. The true effect of
neglecting the M
F
component is hence best assessed by looking at its influence on the
moment coefficients about the x and y-axis, since the heeling and pitching moment are
important parameters when modelling the performance of a yacht. The M
F
component about
the z-axis will be small since the force component in the z-direction is relatively small
compared to the other two components as investigated in detail in chapter 3.


90
RT-VPP VT 4m/s
RT-VPP VT 2m/s
0
0.4
0.8
1.2
1.6
2
0 15 30 45 60 75 9

eff
[]
|
C
M
|

[
-
]
0
|CM|
|CMv|
|CMf|
Figure 5.3: The ratio p of the moment coefficient
in line with the force (C
Figure 5.4: Magnitude of the total moment
coefficient (|C
Mf
) and the
force coefficient (C
M
|) and the components
perpendicular to and in line with the
resultant force (|C
F
) against
eff
| and |C |) vs.
Mv Mf eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 107
Chapter 5 Aerodynamic Force Modelling
0
0.4
0.8
1.2
1.6
2
0 15 30 45 60 75 9

eff
[]
C
M

[
-
]
0
CMx
CMvx
CMy
CMvy

Figure 5.5: x and y-axis components of total
moment coefficient (C
Figure 5.5 shows the x and y-axis components of the total moment coefficient and the
moment coefficient perpendicular to the resultant force. It can be seen that the difference
between the total moment coefficient and the moment coefficient perpendicular to the
resultant force about the x-axis is insignificant for small effective wind angles (
eff
), but that
the difference gets bigger with increasing
eff
. The moment coefficient about the x-axis
(heeling moment coefficient) is up to 17% larger when C
Mfx
is ignored and only C
Mvx
is
modelled. At the same time the moment coefficient about the y-axis (pitching moment
coefficient) is up to 14% smaller when C
Mfy
is not modelled. The percentage change in
pitching moment seems to be less dependent on
eff
than the heeling moment coefficient
change. Figure 5.5 shows that the moments acting on the yacht can be quite different when
C
Mf
is not accounted for although the magnitude of the resulting moment is very similar as
shown in Figure 5.4. The difference is largely due to the change in moment direction when
C
Mf
is ignored.
To check whether the C
Mf
component is genuine or a result of deficiencies in the force
balance set-up, a series of force balance measurements of point loads was compared to
distributed loads in the form of sail force measurements. For the point load measurements C
Mf

should be zero. The point loads were applied via a string which was attached to the model at
one point. Figure 5.6 shows the x-component of the moment perpendicular to the force (M
Vx
)
plotted against x-component of the total moment (M
X
). If the moment in line with the force
(M
Fx
) is zero, the measurements lie on a straight line through the origin with a slope of one.
One set of point load measurements shown in Figure 5.6 is the calibration data -
measurements of loads applied in numerous directions at a number of points on a calibration
frame. It can be seen that the agreement between M
Vx
and M
X
is very good for moments with
M
) and the
component perpendicular the
resultant force (C

) vs.
eff Mv
108
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.3 Improved Aerodynamic Moment Model
-15
-10
-5
0
5
10
15
20
25
-15 -10 -5 0 5 10 15 20 25
M
X
[Nm]
M
V
x

[
N
m
]
Mx=Mvx , Mfx=0
Calibration point force
IACC point force
Dyna point force
Dyna sail force
Figure 5.6: x-direction component of moment perpendicular to resultant force (M
Vx
) against x-direction
component of moment (M
X
) for different force types
a magnitude larger the 1Nm. Only for very small moments are errors introduced due to the
accuracy limitations of the force balance. Point loads applied to an International Americas
Cup Class (IACC) model and the Dyna model are also shown in Figure 5.6. They also show
that M and M
Vx X
match very well if point loads are applied. Sail force measurements with the
Dyna model, also plotted in Figure 5.6, show a M noticeably larger than M
Vx X
for many
measurements. A very similar plot can be produced for the y-component of the moments.
Figure 5.7 shows that M and M match very well for point loads as well and that M
Vy Y Vy
is
smaller than M for many distributed loads. This confirms that M
Y F
does not arise due to
measurement inaccuracies but from the distributed loads generated by the flow around the
model.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 109
Chapter 5 Aerodynamic Force Modelling
-15
-10
-5
0
5
10
15
20
25
-15 -10 -5 0 5 10 15 20 25
M
Y
[Nm]
M
V
y

[
N
m
]
My=Mvy , Mfy=0
Calibration point force
IACC point force
Dyna sail force
Figure 5.7: y-direction component of moment perpendicular to resultant force (M
Vy
) against y-direction
component of moment (M ) for different force types
Y

A good example of factors that contribute to a moment in line with the force is to show what
happens when the two force and moment systems of the hull and the sails are combined to
one resultant force and moment system. From the measurements of the total force acting on
the model and the hull only discussed in chapter 3, the resultant force direction () in the
deck plane, defined by equation (2.33), can be deduced and is shown in Figure 5.8. It can be
seen that the hull force direction is relatively independent of the effective wind angle at 95 to
the centreline, i.e. 5 aft of perpendicular to the centreline. The rig and sails force direction on
the other hand decreases with effective wind angle. Although the hull forces are much smaller
than the rig and sail forces, the large different in force direction between them causes a
noticeable change in force direction in the resultant force of hull, rig and sails as shown in
Figure 5.8. Figure 5.9 and Figure 5.10 show the x and y-component of the moment
coefficient in line with the force for hull, rig and sails, and the combination of the two. The
110
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.3 Improved Aerodynamic Moment Model
0
15
30
45
60
75
90
105
0 15 30 45 60 75

eff
[]

]
90
Hull, rig and sails
Rig and sails
Hull

Figure 5.8: Force direction () in the deck plane
for hull, rig and sails, and the total
model against
eff

-0.12
-0.1
-0.08
-0.06
-0.04
-0.02
0
0.02
0 15 30 45 60 75

eff
[]
C
M
f
x

[
-
]
rig and sails produce a significant amount of C
Mf
whereas the hull generates much less. If the
rig and sail force is combined with the hull force, C
Mf
increases in magnitude by a factor of
between 2-4 for the x-direction and a factor between 2.5-4.5 in the y-direction compared to
the rig and sails values. This shows that a large portion of C
Mf
measured in the wind tunnel, is
due to measuring the forces on the rig and sails, and the hull together, which act in a very
different direction and thereby produce the moment about the resultant force direction. C
Mf

can be reduced significantly by measuring the rig and sail, and hull forces separately, but
some level of C
Mf
still remains due to the distributed force acting on the rig and sails.
90
Hull, rig and sails
Rig and sails
Hull
Hull, rig and sails - windage
0
0.04
0.08
0.12
0.16
0.2
0 15 30 45 60 75

eff
[]
C
M
f
y

[
-
]
90
Hull, rig and sails
Rig and sails
Hull
Hull, rig and sails - windage
Figure 5.9: x-component of the moment
coefficient in line with the force (C
Figure 5.10: y-component of the moment
coefficient in line with the force (C
Mfx
)
for hull, rig and sails, and the
combination of the two vs.
Mfy
)
for hull, rig and sails, and the
combination of the two vs.
eff eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 111
Chapter 5 Aerodynamic Force Modelling
Although successfully performed in the study discussed in chapter 3, it is not feasible to
measure the forces acting on the hull separately during all wind tunnel testing of yacht sails.
Chapter 3 also concludes that the forces on the hull and rig are significantly influenced by the
presence of the sails, so that simply subtracting the bare poles measurements does not reduce
C
Mf
by a similarly large amount as shown in Figure 5.9 and Figure 5.10. The C
Mf
should
therefore not be ignored in an aerodynamic force model.
If the common approach of using a centre of effort position to calculate the moments is
adhered to, the neglected moment component should be added. For the basic definition of the
centre of effort given in equations , the additional moment M (5.1) and (5.2)
0Y
, defined in
equation (5.3), should be included. For the physically more meaningful definition of the
centre of effort in the central axis system, the additional moment in line with the force (M
F
)
should be modelled. Instead of defining the scalar factor p as an input parameter in the
aerodynamic force module, the scalar moment coefficient C
Mf
is chosen as an input parameter
since it is seen as a more meaningful quantity and it is only dependent on the direction but not
the magnitude of the force. The total moment is hence obtained from the force, the centre of
effort position defined by equations and C (2.31) and (2.32)
Mf
so that
, F F CoE M M M

5 . 1
Mf eff S F V
C q A + = + =
(5.7)
where F is the unit vector of F and the scalar C

Mf
is defined as
F
M F
F Mf
p C
C
C C
C

= = .
(5.8)
From the heeled tests conducted for the study described in chapter 3, the influence of heel on
C
Mf
can be assessed. Figure 5.11 shows that there is no clear trend of C
Mf
with heel angle and
-0.25
-0.2
-0.15
-0.1
-0.05
0
0 15 30 45 60 75 9

eff
[]
C
M
f
[
-
]
0
Total 0 heel
Total 12.5 heel
Total 25 heel
0
0.05
0.1
0.15
0.2
0.25
0 15 30 45 60 75 9

eff
[]
C
M
f

[
-
]
0
Measurements
FS-Equi B-spline fit
Figure 5.11: Total C Figure 5.12: Measured total C
Mf
of mainsail, jib, rig and hull
for different heel angles vs.
Mf
of mainsail, jib,
rig and hull and VPP B-spline fit
eff
112
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.3 Improved Aerodynamic Moment Model
C
Mf
is therefore modelled as a function of
eff
in FRIENDSHIP-Equilibrium. Measurements in
the wind tunnel are conducted with the model on port tack, which are negative apparent wind
angles in FRIENDSHIP-Equilibrium. The C
Mf
curve is point-symmetrical about the origin so
that it can be mirrored and positive values for
eff
and C
Mf
are used as VPP input. Figure 5.12
shows the C
Mf
measurements of the fully powered-up trims and the B-spline curve fitted to
the data in FRIENDSHIP-Equilibrium. In the aerodynamic force model it is defined whether
an input curve is point or axis symmetrical so that the input data is only required for positive
(starboard tack) or negative (port tack) effective wind angles. The M
F
contribution to the total
moment is implemented in the VPP by replacing equation (4.11) in the standard aerodynamic
force model with equation which uses the C (5.7)
Mf
input curve from Figure 5.12.
VPP solutions are produced with and without modelling M
F
to show its influence on the
performance of the yacht. The VPP predictions for this comparison use the standard
hydrodynamic set-up for Dyna (used throughout the project and outlined in appendix C.4),
the fully powered-up aerodynamic coefficients from this study, and the trim parameters reef,
flat and twist. The force component perpendicular to the deck plane discussed in section 3.8 is
also modelled and the equilibrium is solved in all six degrees of freedom. The VPP
calculations are carried out for Dyna under mainsail and jib for a range of true wind speeds
and true wind angles. Figure 5.13 shows the predicted boat speed (V
S
) for different true wind
speeds (V ) plotted against the resulting effective wind angle (
T eff
). For yacht design purposes
it is more useful to plot V
S
against the true wind angle, but to make the comparison with the
coefficient graphs easier the VPP results are plotted against the effective wind angle here. In
Figure 5.13, the predictions shown in thick lines model the C
Mf
component whereas the thin
lines do not. It can be seen that the predicted boat speed increases noticeably for
eff
between
40 and 80 at the higher true wind speeds of 8 and 10m/s. For the results shown, the boat
2.8
3
3.2
3.4
3.6
3.8
4
4.2
4.4
4.6
0 15 30 45 60 75 9

eff
[]
V
S

[
m
/
s
]
0
VT=10m/s with CMf
VT=10m/s no CMf
VT=8m/s with CMf
VT=8m/s no CMf
VT=6m/s with CMf
VT=6m/s no CMf
VT=4m/s with CMf
VT=4m/s no CMf
0
5
10
15
20
25
30
35
40
45
0 15 30 45 60 75 9

eff
[]

]
0
VT=10m/s with CMf VT=10m/s no CMf
VT=8m/s with CMf VT=8m/s no CMf
VT=6m/s with CMf VT=6m/s no CMf
VT=4m/s with CMf VT=4m/s no CMf
Figure 5.13: Boat speed of Dyna under mainsail
and jib calculated with and without
considering C
Figure 5.14: Heel angle of Dyna under mainsail
and jib calculated with and without
considering C for different V for different V
Mf T
vs.

Mf T
vs.

eff eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 113
Chapter 5 Aerodynamic Force Modelling
speed increases by up to 1.5% when modelling C
Mf
at V of 10m/s and
T eff
of around 68. The
boat speed is increased because C
Mf
reduces the heeling moment, which, for upwind and
reaching conditions, enhances the performance of the yacht noticeably. For effective wind
angles larger than 80, the increase in V
S
becomes smaller since the heeling moment is of less
importance. For small
eff
the increase in V is very little since C
S Mf
is small (see Figure 5.12).
The effect of C
Mf
in reducing the heeling moment can be seen in Figure 5.14, which shows
the heel angle () obtained from the VPP calculations plotted against
eff
. For heel angles
below 25 the yacht is predicted to heel less when C
Mf
is modelled due to the reduction in
heeling moment. For larger than 25 this effect is not noticeable anymore since the sails are
depowered by the trim parameters. It should be noted that not all effects of pitch angle (trim
angle) on drag are considered by the hydrodynamic model currently used. However, the
change in pitch angle due to modelling C
Mf
is less than 0.1 for almost all calculated cases so
that this deficiency is not expected to have a significant influence on the trends shown.
So far the sail combination of mainsail and jib has been considered in detail. It is also
interesting to compare C
Mf
of other sail combinations. As part of the study described in
chapter 3, measurements were conducted not only for mainsail with jib but also for mainsail
with genoa or spinnaker. Figure 5.15 shows the total measured C
Mf
of the hull and rig with the
different sail combinations. Instead of using the rated IMS sail area as the reference area
when calculating the coefficients, the cloth area of the sails is used to account for their size
difference. It can be seen that C
Mf
for mainsail with jib and genoa generally agree well and
show a very similar trend. Mainsail and spinnaker measurements also have a similar shape to
the upwind sails but the C
Mf
values for
eff
below 80 are smaller. This could be due to the
large gap between the spinnaker and the foredeck compared to the upwind sails. As shown in
chapter 3 a high-pressure region is created on the foredeck by the upwind sails, which is not
-0.16
-0.14
-0.12
-0.1
-0.08
-0.06
-0.04
-0.02
0
0 30 60 90 120 150 180

eff
[]
C
M
f
[
-
]
Hull, rig, main & jib
Hull, rig, main & genoa
Hull, rig, main & spi

Figure 5.15: Measured total C
Mf
of hull, rig and
mainsail with jib, genoa and
spinnaker based on sailcloth area

114
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
seen with the spinnaker. This high-pressure creates a pitching moment which contributes to
M
F
since for small
eff
the resultant force has a similar direction to the pitching moment,
which has a direction of =90 in Figure 5.8. The pitching moment due to the high-pressure
on the foredeck in not created by the spinnaker and M
F
is therefore less for small
eff
.
It has been shown that the moment in line with the force, which decreases the heeling
moment by up to 17%, has a clear effect on the predicted heel angle and speed of the yacht.
For an accurate performance prediction with the central axis method it should therefore be
accounted for. This has been achieved by modelling the M
F
component in FRIENDSHIP-
Equilibrium. The central axis method is used here because it gives a physically more
meaningful location of the CoE, and it provides consistency throughout the project and with
the work conducted by Hochkirch (2000). However, in general it can be argued that using the
traditional approach of determining the centre of effort by dividing the moment by the
corresponding force as shown in equations (5.1) and (5.2) is favourable. If the CoE is
determined using the central axis method, M
F
influences the results regardless of whether the
pitching moment is considered or not. With the traditional approach M
0Y
only needs to be
modelled if the pitching moment is included in the equilibrium calculations. In addition the
calculations involved in obtaining the CoE are much simpler. Reverting back to the
traditional method was not considered here but should be investigated in the future.
5.4 Depowering Model with Trim Parameter Power
Having considered the force and moment modelling in VPPs in section 5.3 the depowering is
now investigated. The set of realistically depowered trims from the Real-Time VPP
measurements provides the opportunity to develop a new empirical aerodynamic force model
and compare its performance predication to the standard trim parameter model introduced in
section 4.2.1. The main conceptual difference to the standard trim parameter model is that the
new model is not attempting to model the different trim adjustments that can be made to a
sail, like reef, flat and twist do. The new model simply describes a set of good depowered
trims that have been obtained by trimming the sails in the wind tunnel using the Real-Time
VPP. The new description will therefore not be as generic as the standard trim parameter
model but aims to model the complexity of depowering for a particular rig in more detail. A
diverse and customisable model is of particular interest for high-performance offshore racing
classes, like the Volvo Open 70 class, where depowering of sails is necessary not only when
sailing upwind but also in reaching conditions. The standard trim parameter model is
intended for upwind sailing conditions and cannot be expected to work equally well in
reaching conditions. All results shown in this chapter are for the model with mainsail and jib.
For this study depowered measurements with a spinnaker have not been carried out, which
would make the force and moment changes due to depowering even more complex so that the
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 115
Chapter 5 Aerodynamic Force Modelling
standard trim parameters should not be used. The more specific depowering model will be
very applicable to offwind sails but is introduced here for the upwind configuration to enable
a meaningful comparison to the standard trim parameter model later in section 5.5.
Since the new depowering model does not aim to describe the physical changes to the sails
but aims to describe the force and moment changes that occur for a realistically depowered
sail, only one trim parameter is required that identifies the level of depowering. The new trim
parameter is called power (p) since it generally describes the level of power produced by the
sails. For p=1 the sails are fully powered-up and for p=0 the sails are totally depowered. The
value of power needs to be linked to the wind tunnel measurements in some way. Since the
main reason for depowering a sail is to reduce the heeling moment, the level of depowering
can be quantified by the reduction in heeling moment. Ranzenbach and Teeters (2002)
similarly use the heeling moment as a measure of depowering for their offwind sails
investigation. The trim parameter power (p) is hence defined as the ratio between the heeling
moment coefficient (C
Mx
) and the optimum heeling moment coefficient (C
MxOpt
) for the fully
powered-up sails so that
MxOpt
Mx
C
C
p = ,
(5.9)
where optimum here, as always in this work, refers to the optimum fully powered-up sail
trim that results in the aerodynamically most efficient shape and produces the highest drive
force coefficient.
15
0
0.4
0.8
1.2
1.6
2
0 15 30 45 60 75

eff
[]
C
L

[
-
]
90
VT=14m/s RT-VPP
VT=12m/s RT-VPP
VT=10m/s RT-VPP
VT=8m/s RT-VPP
VT=6m/s RT-VPP
VT=4m/s RT-VPP
Figure 5.16: Lift coefficient (C

15
The optimum fully powered-up sail trim does however not mean that the optimum loading distribution of
elliptical or semi-elliptical loading is achieved. While C
MxOpt
could also be referred to as the maximum C
Mx
,
the term optimum is chosen since this trim condition does not result in the maximum for other parameters
such as drag.
L
) from wind tunnel
tests with Real-Time VPP for
different true wind speeds (V
Figure 5.17: Surface describing the lift coefficient
(C
L
) as a function of effective wind
angle ( )
T eff
) and the depowering
parameter power
116
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
In the standard trim parameter model the lift, drag and centre of effort height are adjusted
when the sails are depowered and the Real-Time VPP measurements confirm that these
quantities are by far the most affected during depowering. Figure 5.16 shows the lift
coefficient (C ) for different true wind speeds (V
L T
) obtained in the wind tunnel with the Real-
Time VPP. C is increasingly reduced for V
L T
above 4m/s and the reduction is greatest for
small effective wind angles (
eff
). At V of 6m/s the C is reduced for
T L eff
below about 45
whereas it is reduced for all
eff
up to 90 in V
T
of 14m/s. The measurements show the
expected trend that the sails are depowered more with increasing V and decreasing
T eff
.
C can be modelled as a response surfaces dependent on the effective wind angle (
L eff
) and
the newly defined trim parameter power so that
. ) , ( p C C
eff L L
=
(5.10)
The response surface modelling the dependency of C
L
on
eff
and p is shown in Figure 5.17.
For the boundary conditions it is assumed that C is zero at
L eff
=0 for all values of power
and that C is also zero for p=0 for all
L eff
. The wind tunnel measurement points in Figure
5.17 confirm that the sails are depowered more for small
eff
since the range of measured
power values is largest for small
eff
and decreases with increasing
eff
.
The response surface in Figure 5.17 can model C
L
quite well but its shape is relatively
complex which makes accurate fitting to the wind tunnel data difficult. A different approach
that is more similar to the standard aerodynamic model is selected where the fully powered-
up coefficient (C
Lopt
) is modelled separately from the effects of depowering so that
, ) , ( ) ( p R C C
eff L eff Lopt L
=
(5.11)
Figure 5.18: The optimum lift coefficient (C Figure 5.19: Surface describing the lift coefficient
ratio (R
Lopt
) as
a function of effective wind angle (
eff
) ) as a function of
L eff
and
power
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 117
Chapter 5 Aerodynamic Force Modelling
where C
Lopt
is a function of
eff
as in the standard aerodynamic force model and R
L
is the
power ratio of C
L
which is a function
eff
and power (p). Figure 5.18 shows C
Lopt
as a function
of
eff
which is multiplied by the power ratio of C (R ) shown in Figure 5.19
L L
to model the
depowered lift coefficient. The power ratio of C is mainly dependant on p and less on
L eff
so
that the response surface shape is less complex and easier to define accurately. The new
power model uses the same C
Lopt
as the standard trim parameter model shown in equation
(4.1) so that the only difference between the models is in the depowering. This is important
for the comparison between the depowering models because the data fitting can have a
significant influence on the results. It needs to be ensured that any differences originate from
the depowering models and not the data fitting. Modelling the depowering separately from
the optimum parameter curves also allows the description of depowering to be transferred to
fully powered-up measurements of other sails. Since the power parameter is not based in
aerodynamic principles but solely on experimental results, the power ratios may vary for
different rig and boat types. Without further research the power ratios should only be
transferred to rigs and boats of very similar type.
Figure 5.20 shows C measured in the wind tunnel with the Real-Time VPP for different V
L T

and C modelled from the Real-Time VPP measurements using equation (5.11)
L
. The
modelled C
L
is only calculated for each wind tunnel measurement and the points are
connected by straight lines since the power values at other
eff
are not known. C
L
is modelled
in good detail but some of the discontinuities due to measurement inaccuracies are smoothed
out. The average absolute error between the C measurements and the response surface is
L
0
0.4
0.8
1.2
1.6
2
2.4
2.8
0 15 30 45 60 75

eff
[]
C
L

[
-
]
90
VT=14m/s RT-VPP VT=8m/s RT-VPP
VT=14m/s modelled VT=8m/s modelled
VT=12m/s RT-VPP VT=6m/s RT-VPP
VT=12m/s modelled VT=6m/s modelled
VT=10m/s RT-VPP VT=4m/s RT-VPP
VT=10m/s modelled VT=4m/s modelled

Figure 5.20: Lift coefficient (C
L
) for all tested true
wind speeds (V
T
) from Real-Time
VPP and modelled by power
118
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
1.2% and the maximum absolute error is 4.5%. The error is calculated relative to the fully
powered-up trim at each
eff
.
The effect of depowering on the parasitic drag coefficient (C
Dp
) is modelled in the equivalent
way so that
, ) , ( ) ( p R C C
eff Dp eff DpOpt Dp
=
(5.12)
where C
DpOpt
is the optimum drag coefficient for the fully powered-up sail and R
Dp
is the
power ratio of C describing the changes due to depowering as a function of
Dp eff
and power
(p). The parasitic drag coefficient is modelled following the standard approach of dividing the
drag into components as shown in equations (2.5) to (2.8). Due to the detailed modelling of
the depowering it would be equally feasible to model the total drag coefficient (C
D
) as a
function of
eff
and power. But since the individual drag components are to be modelled here,
C is calculated from the total measured drag coefficient (C
Dp D
) by subtracting the induced and
separation drag coefficients (C and C ) as defined in equations (2.5) and (2.6) so that
Di Ds
2
2
1
L s
L
D Ds Di D Dp
C c
AR
C
e
C C C C C = =

.
(5.13)
In this description C
Dp
still includes the windage component of hull and rig. The power ratio
of C (R
Dp Dp
) obtained from the wind tunnel measurements and the corresponding B-spline
surface modelled in FRIENDSHIP-Equilibrium are shown in Figure 5.21. The boundary
conditions for the response surface are not as clearly definable as for C since C is not zero
L Dp
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 15 30 45 60 75 9

eff
[]
C
D
p

[
-
]
0
VT=14m/s RT-VPP VT=8m/s RT-VPP
VT=14m/s modelled VT=8m/s modelled
VT=12m/s RT-VPP VT=6m/s RT-VPP
VT=12m/s modelled VT=6m/s modelled
VT=10m/s RT-VPP VT=4m/s RT-VPP
VT=10m/s modelled VT=4m/s modelled


Figure 5.22: C Figure 5.21: Surface describing the parasitic drag
coefficient ratio (R
Dp
for all tested true wind speeds
(V
Dp
) as a function of

T
) from Real-Time VPP and
modelled by power and power
eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 119
Chapter 5 Aerodynamic Force Modelling
at
eff
=0 nor p=0. A data point at
eff
=90 and p=0 is added and its R
Dp
value is selected to
constrain the surface to a sensible shape by visual inspection. For
eff
close to zero the
behaviour of the surface might not be very realistic but since such small
eff
do not constitute
feasible sailing conditions it does not affect the VPP results. More care has to be taken when
modelling the surface at small power values since power is a free variable optimised by the
VPP and any value between 0-1 is feasible.
R in Figure 5.21 is larger than one when depowering the sails at small
Dp eff
. When
depowering the sails the induced drag per unit lift increases due to changes in the loading
distribution as discussed in relation to the trim parameter twist in section 4.2.1. Twist is
intended to model this in the standard trim parameter model but in equation C (5.13)
Di

changes only through reduction in C when depowering. The increase in C
L Di
per unit lift is
therefore included in C . For larger
Dp eff
this increase is outweighed by other factors
decreasing C
Dp
when depowering the sails such as, for example, delayed separation on the
leeward side of the sails due to a reduction in angle of attack. Figure 5.22 shows C
Dp

measured with the Real-Time VPP in the wind tunnel for different V and C
T Dp
modelled from
the wind tunnel tests with equation . Although the changes in C (5.12)
Dp
are complex, they are
modelled reasonably well by the power model. The average absolute error between the C
Dp

measurements and the response surface is 3.5% while the maximum absolute error is 9.5%.
The error is again calculated relative to the fully powered-up trim at each
eff
.
The third important parameter that changes during depowering is the centre of effort height
(z
CoE
). As discussed before in section 5.3 z
CoE
is here defined using the central axis system
3
3.5
4
4.5
5
5.5
6
6.5
7
7.5
8
0 15 30 45 60 75 9

eff
[]
z
C
o
E

[
m
]
0
VT=14m/s RT-VPP VT=8m/s RT-VPP
VT=14m/s modelled VT=8m/s modelled
VT=12m/s RT-VPP VT=6m/s RT-VPP
VT=12m/s modelled VT=6m/s modelled
VT=10m/s RT-VPP VT=4m/s RT-VPP
VT=10m/s modelled VT=4m/s modelled


Figure 5.24: z Figure 5.23: Surface describing the centre of effort
height ratio (R
CoE
for all tested true wind speeds
(V
zCoE
) as a function of
eff

and power
T
) from Real-Time VPP and
modelled by power
120
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
approach and is plotted against
eff
. Equivalent to C and C
L Dp
it is expressed in terms of the
fully powered-up centre of effort height (z
CoEopt
) and the power ratio of z
CoE
(R
zCoE
). Figure
5.23 shows R
zCoE
as a function of
eff
and power. Three data points are added at p=0 to
constrain the B-spline surface and form a smooth shape that fits the measured data well. R
zCoE

reduces when depowering the sails. The reduction is most severe for small
eff
and reduces for
larger
eff
. Figure 5.24 shows z
CoE
obtained by the Real-Time VPP for different V
T
and the
corresponding modelled z
CoE
. A clear trend of reducing the centre of effort height when
depowering the sails can be seen although it should be noted that z
CoE
is plotted against a
false origin in the graph. The reduction in z
CoE
is accurately modelled while smoothing out
some of the measurement inaccuracies. The average absolute error relative to the fully
powered-up trims is 1.2% with a maximum difference of 4.7%.
5.4.1 Comparison to Real-Time VPP Results
The new power model for the depowering of sails has been introduced in section 5.4 and it
has been shown to fit the wind tunnel data well. The next step is to implement the power
model in a VPP and investigate how the predictions compare to the Real-Time VPP results
when the trim parameter power is optimised for maximum yacht speed.
A new force module called HansenRig has been implemented in FRIENDSHIP-
Equilibrium, which uses the new power model to simulate the depowering of the sails. The
new force model uses data of fully powered-up and optimally trimmed sails, described in
terms of C
Lopt
, C
DpOpt
, C , x
FzOpt CoEopt
, z
CoEopt
and C
MfOpt
, as functions of
eff
, to model the
aerodynamic behaviour in all six degrees of freedom. The fully powered-up measurements
from the Real-Time VPP tests for a true wind speed of 4m/s are used here as input; C , C
L Dp

and z
CoE
are shown in Figure 5.16, Figure 5.22 and Figure 5.24, and C
Mf
for the fully
powered-up condition is shown in Figure 5.12. The fully powered-up longitudinal centre of
effort (x
CoEopt
) is plotted in Figure 5.25 as a function of
eff
. All VPP input is in the body fixed
coordinate system which has its origin close to the waterline just in front of the mast and x-
values are positive forward (see also appendix C.1 for coordinate system definitions). Figure
5.25 also shows the longitudinal centre of area of the sails to illustrate the relative location of
x
CoEopt
.
In contrast to that suggested in chapter 3, C is only modelled as a function of
Fz eff
in this
study. Its dependency on the heel angle is ignored since the wind tunnel tests were only
conducted at a constant heel angle of 0. The aim of this study is to investigate depowering of
sails and it is therefore important to model the wind tunnel measurements as accurately as
possible to ensure that differences in the comparison originate from the depowering and not
from other aspects of the force modelling. Introducing a dependency of C on heel angle
Fz
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 121
Chapter 5 Aerodynamic Force Modelling
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
0 15 30 45 60 75 9

eff
[]
x
C
o
E
o
p
t

[
m
]
0
Main, jib, hull & rig (wind tunnel)
Main, jib, hull & rig (modelled)
Geometric centre main & jib
Geometric centre main
Geometric centre jib
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0 15 30 45 60 75 9

eff
[]
C
F
z
O
p
t

[
-
]
0
Main, jib, hull & rig (wind tunnel)
Main, jib, hull & rig (modelled)
Figure 5.25: Measured x Figure 5.26: Measured C
would describe the physical behaviour more accurately but would correspond less well to the
wind tunnel measurements conducted for this comparison. The HansenRig force module has
been implemented with the capability to model C
FzOpt
as a function of
eff
and heel angle but
for this study C
FzOpt
is only assumed dependent on
eff
as shown in Figure 5.26. More
information on the input parameters required for the new force module is given in appendix
C.3.2.
The depowering is modelled by the power ratios of C
L
, C
Dp
and z
CoE
shown in Figure 5.19,
Figure 5.21 and Figure 5.23, by which C
Lopt
, C
DpOpt
and z
CoEopt
are multiplied. C
D
is defined
by rearranging equation (5.13) and the forces in the absolute coordinate system used by the
VPP to solve for equilibrium are then calculated following the procedure shown in equations
(4.6) to (4.13) with the differences that the centre of effort position (CoE) is now given by the
point
,

=
) , ( ) (
0
) (
p R z
x
eff zCoE eff CoEopt
eff CoEopt

CoE
(5.14)
and that the moments are calculated from equation (5.7) instead of equation (4.11) to include
the moment component in line with the force as discussed in section 5.3.
For the comparison, performance predictions for true wind angles from 25 to 135 and true
wind speeds of 4, 6, 8 and 10m/s are computed solving for equilibrium in all six degrees of
freedom and using the newly defined trim parameter power to optimise the boat speed for
each sailing condition. The hydrodynamic input data is identical to the Real-Time VPP
calculations and any other predictions shown in this work. Figure 5.27 shows the power
values of the Real-Time VPP measurements that have been used to model the power ratios
CoEopt
of mainsail, jib, rig
and hull and VPP B-spline curve fit
FzOpt
of mainsail, jib, rig
and hull and VPP B-spline curve fit
122
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 15 30 45 60 75

eff
[]
p
o
w
e
r

[
-
]
90
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s power VT=4m/s power
2
2.5
3
3.5
4
4.5
5
0 15 30 45 60 75 90 105 120 135

T
[]
V
S

[
m
/
s
]
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s power VT=4m/s power
Figure 5.28: Boat speed (V Figure 5.27: Values of trim parameter power for
four true wind speeds (V
S
) for four true wind
speeds (V
and the power values optimised by FRIENDSHIP-Equilibrium with the HansenRig force
module to maximise the yacht speed. The HansenRig force model predictions agree with the
wind tunnel measurements that the sails should not be depowered in 4m/s true wind speed at
any
eff
. For stronger V
T
the HansenRig force model depowers the sails more than in the
wind tunnel but the trends are very similar. This indicates that the sails were not quite
depowered enough in the wind tunnel to give the best performance. The biggest difference
can be seen at
eff
of 60 in V
T
of 8 and 10m/s. Figure 5.28 shows that the boat speed (V
S
)
from the wind tunnel measurements and power model predictions agree well for V
T
from 4-
8m/s. The average absolute error in V
S
is 0.3% with a maximum error of 0.9%. At V
T
of
10m/s the differences are somewhat larger with an average absolute error of 0.6% and a
maximum of 1.2% at
eff
of 60 (
T
87). The maximum difference in boat speed is however
not more than 0.05m/s. The difference in power does not seem to result in a significant
difference in boat speed. When trimming the sails in the wind tunnel with the Real-Time VPP
such small differences in V
S
are difficult to detect because of the fluctuations in V
S
caused by
unsteadiness in the force measurements so that the differences in power are not surprising.
Figure 5.29 shows the lower predicted heel angle () by the power model which follows
from the reduced values of power. It seems reasonable that a reduction in increases V
S
but
that the speed increase is not significant. In conditions where the sails are not depowered the
heel angles are very similar which confirms an accurate modelling of the fully powered-up
forces and moments.
T
) from Real-
Time VPP and power model VPP
calculations
T
) predicted by Real-Time
VPP and VPP with power parameter
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 123
Chapter 5 Aerodynamic Force Modelling
0
5
10
15
20
25
30
35
40
45
50
0 15 30 45 60 75 9

eff
[]

]
0
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s power VT=4m/s power
0
0.4
0.8
1.2
1.6
2
2.4
2.8
0 15 30 45 60 75 9

eff
[]
C
L

[
-
]
0
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s power VT=4m/s power
Figure 5.30: Lift coefficient (C Figure 5.29: Heel angle () for four true wind
speeds (V
L
) for four true wind
speeds (V
T
) predicted by Real-Time
VPP and VPP with power parameter
T
) from Real-Time VPP and
power model VPP calculations

3
3.5
4
4.5
5
5.5
6
6.5
7
7.5
8
0 15 30 45 60 75 9

eff
[]
z
C
o
E

[
m
]
0
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s power VT=4m/s power
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 15 30 45 60 75 9

eff
[]
C
D
p

[
-
]
0
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s power VT=4m/s power
Figure 5.31: Centre of effort height (z
CoE
) for four
true wind speeds (V
T
) from Real-Time
VPP and power model VPP
calculations
Figure 5.32: Parasitic drag coefficient (C
Dp
) for
four true wind speeds (V
T
) from Real-
Time VPP and power model VPP
calculations

The increased level of depowering chosen by the power model leads to a generally greater
reduction in C , z
L CoE
and C compared to the wind tunnel measurements as shown in
Dp
Figure
5.30 to Figure 5.32. For C and C
L Dp
the greatest difference occurs at the measurement of

eff
=60 and V =10m/s which also results in the greatest V difference. The depowering trends
T S
124
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
are however modelled well in all cases and the good agreement of the parameters at V
T
of
4m/s confirms that the fully powered-up condition is modelled well. The fact that the sails
were not depowered quite enough in the wind tunnel does not affect the validity of the study
since the characteristics of how the forces and moments change relative to each other when
depowering should not be influenced significantly. The Real-Time VPP should simply be
seen as a tool to depower the sails efficiently and realistically in the wind tunnel. The power
model is then used to fine-tune the wind tunnel trims to find the best level of depowering
more accurately.
The power model VPP predictions are therefore used to assess the standard trim parameter
model in section 5.5 and as a benchmark for the improved twist parameter introduced in
section 5.6 instead of using the Real-Time VPP measurements directly.
5.4.2 Generic Power Model
Although the main purpose for developing the power model is a detailed description of the
depowering a more generic expression can have advantages when the representation by
response surfaces is not desired. The power ratio of C shown in Figure 5.19
L
is plotted as a
function of only power in Figure 5.33 and apart from
eff
of 90 the trend in the power ratio of
C
L
is very similar for the different
eff
. R
L
can therefore be modelled as a function of power.
The relationship is close to linear but a better data fit is obtained by a power expression so
that
,
1
) (
a
eff Lopt L
p C C =
(5.15)
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
power [-]
C
L
/
C
L
o
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
Power^a1
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
power [-]
z
C
o
E
/
z
C
o
E
o
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
Power^a2
Figure 5.33: C
L
/C
Lopt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
Figure 5.34: z
CoE
/z
CoEopt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 125
Chapter 5 Aerodynamic Force Modelling
which is shown in Figure 5.33, with the constant a given in Table 5.1
1
. Similarly the power
ratio of z
CoE
from Figure 5.23 is not strongly dependent on
eff
as shown in Figure 5.34. For

eff
between 20-40 no trend with
eff
can be seen, whereas R
zCoE
reduces less with increasing

eff
for angles of 60 and 90. The reduction in z
CoE
with power is not quite linear and can also
be modelled by a power expression so that
,
2
) (
a
eff CoEopt CoE
p z z =
(5.16)
where the value of constant a is given in Table 5.1
2
. The power ratio of the parasitic drag
from Figure 5.21 is however strongly dependent on
eff
as already discussed in section 5.4
and shown clearly in Figure 5.35. However, Figure 5.36 shows that the ratio of the total drag
is fairly independent of
eff
and can be modelled as a function of power so that
( )
4
) 1 (
3 3
a
Dopt D
p a a C C + = ,
(5.17)
where the values of constants a and a are given in Table 5.1
3 4
and the fully powered-up drag
coefficient (C
Dopt
) is calculated from the optimum lift and parasitic drag coefficient curves
with
) (
) (
1
) (
2
2
eff DpOpt
eff Lopt
eff Lopt S Dopt
C
AR
C
e
C c C


+ + = , (5.18)
so that the same input data as in the response surface version is still used.
The generic version of the power model is also implemented in the HansenRig force
module of FRIENDSHIP-Equilibrium so that the regression parameters a to a
1 4
can be
specified instead of the power ratio response surfaces. Further detail on the HansenRig
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 0.2 0.4 0.6 0.8
power [-]
C
D
p
/
C
D
p
O
p
t

[
-
]
1
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
power [-]
C
D
/
C
D
o
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
a3+(1-a3)*Power^a4
Figure 5.35: C
Dp
/C
DpOpt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
Figure 5.36: C /C
D Dopt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
126
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 15 30 45 60 75 9

eff
[]
p
o
w
e
r

[
-
]
0
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s gen power VT=6m/s gen power
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s gen power VT=4m/s gen power
VT=8m/s power VT=4m/s power
3
3.5
4
4.5
5
5.5
6
6.5
7
7.5
8
0 15 30 45 60 75

eff
[]
z
C
o
E

[
m
]
90
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s gen power VT=6m/s gen power
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s gen power VT=4m/s gen power
VT=8m/s power VT=4m/s power
Figure 5.38: Centre of effort height (z Figure 5.37: Values of trim parameter power from
Real-Time VPP measurements and
VPP calculations using standard and
generic form of power
CoE
)
predicted by Real-Time VPP and VPP
with trim parameter power in
standard and generic form
force module is given in appendix C.3.2. Figure 5.37 shows the power values selected by the
two power model versions and the original wind tunnel values. The power values selected
by the two model versions are very similar apart from at large
eff
, where the generic power
model depowers less since the depowering changes are not modelled very accurately at these
angles. Due to the good data fit of R against power in Figure 5.33, C
L L
for the generic model
is very similar to the response surface version and is hence not shown here. Figure 5.38
shows the centre of effort height (z
CoE
) which agrees well for small
eff
but is reduced more by
the generic model at medium
eff
because the data fit in Figure 5.34 is not as good as for R .
L
The drag coefficient predicted by the two model versions is very similar for
eff
between 20
and 55 as shown in Figure 5.39. For
eff
larger than 55, C
D
is larger for the generic version
because power is reduced less at these angles. For
eff
smaller than 20, C
D
is larger for the
generic version which can be attributed to the response surface modelling since no wind
tunnel measurements were conducted for
eff
smaller than 20. The difference in boat speed
(V ) between the generic and response surface version is shown in Figure 5.40
S
as a ratio
plotted against the true wind angle ( ). Also shown are the Real-Time VPP measurements
T

Table 5.1: Regression parameters a
1
to a
6
to model the depowering of C , z
L CoE
and C as a function of p
D
obtained for
eff
between 25 and 90
a a a a a a Parameter
1 2 3 4 5 6
Values for equations (5.15), (5.16) and (5.17) 0.7 0.3 0.45 1.63 - -
Values for equations (5.15), (5.16) and (5.19) 0.7 0.3 0.42 1.82 3.21 -1.10
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 127
Chapter 5 Aerodynamic Force Modelling
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 15 30 45 60 75

eff
[]
C
D

[
-
]
90
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s gen power VT=6m/s gen power
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s gen power VT=4m/s gen power
VT=8m/s power VT=4m/s power
0.92
0.96
1
1.04
1.08
1.12
1.16
1.2
0 15 30 45 60 75 90 105 120 135

T
[]
V
S

/
V
S
(
p
o
w
e
r
)

[
-
]
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s gen power VT=6m/s gen power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s gen power VT=4m/s gen power
Figure 5.39: Drag coefficient (C Figure 5.40: Boat speed ratio of V
relative to the response surface power model. Both the generic version and the Real-Time
VPP results are within 1% of the boat speed predicted by the response surface power
model for most
T
. The difference is larger for
T
below the range tested in the wind tunnel
due to uncertainties in projecting the response surfaces outside the measurement range and
since changes in C
D
have a much greater influence on the performance of the yacht at small

eff
compared to at larger
eff
.
A second method to generalise the drag coefficient is implemented to investigate the
sensitivity of the results to changes in the input data. The drag coefficient is now calculated
from the separation drag, induced drag and parasitic drag where each component is a function
of power so that
( )
4 6 5
) 1 (
1
3 3
2
2 a
DpOpt
a L a
L S Dp Di Ds D
p a a C p
AR
C
e
p C c C C C C + + + = + + =

,
(5.19)
where C
L
is obtained from equation (5.15). The constants a
3
to a
6
are obtained from the wind
tunnel data by least squares regression and the values are given in Table 5.1. The power
values chosen by FRIENDSHIP-Equilibrium for the second generic model version are almost
identical to the original generic version. Therefore C
L
and z
CoE
are also almost identical. The
drag coefficient for the second generic version is shown in Figure 5.41 and is also very
similar to C
D
for the original generic version in Figure 5.39. As a result the difference in boat
speed in Figure 5.42 is very similar to before. Both versions of the generic model produce
D
) predicted by
Real-Time VPP and VPP with trim
parameter power in standard and
generic form
S
predicted by
Real-Time VPP and generic power
model relative to V
S
from standard
power model

128
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 15 30 45 60 75

eff
[]
C
D

[
-
]
90
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s gen2 power VT=6m/s gen2 power
VT=10m/s power VT=6m/s power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s gen2 power VT=4m/s gen2 power
VT=8m/s power VT=4m/s power
0.92
0.96
1
1.04
1.08
1.12
1.16
1.2
0 15 30 45 60 75 90 105 120 135

T
[]
V
S

/
V
S
(
p
o
w
e
r
)

[
-
]
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s gen2 power VT=6m/s gen2 power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s gen2 power VT=4m/s gen2 power

Figure 5.42: Boat speed ratio of V Figure 5.41: Drag coefficient (C
very similar results that agree well with the original response surface version of the power
model.
5.4.3 Depowering Changes to x
CoE
and C
Mf

So far the power model accounts for depowering changes in C
L
, C
Dp
and z
CoE
, which are the
variables that most significantly affect the yachts performance when depowering. Since the
power model is intended to simulate the depowering in detail the effect of depowering on
the other variables, C
Fz
, x
CoE
and C
Mf
, is also investigated. The power ratio of the longitudinal
centre of effort (R
xCoE
) for different
eff
is plotted against power in Figure 5.43. Although
there is some scatter in the data a clear trend with power can be seen. For all
eff
apart from
90 R
xCoE
reduces, whereas it increases slightly for
eff
of 90. As before the power ratio is
multiplied by the fully powered-up values, which are given in Figure 5.25. x
CoEopt
is between
about 0.8 and 2.2 metres behind the body fixed coordinate system origin for the range of
eff

tested so that a R
xCoE
value of 0.5 at
eff
of 30 means a forward shift in x
CoE
of about 0.4
metres in full-scale. As expected x
CoE
moves forward for
eff
less than 90 but the shift is not
very significant. R
xCoE
is modelled in the HansenRig force module as a B-spline surface
dependent on
eff
and power (p) so that equation (5.14) to obtain the centre of effort point
(CoE) now becomes
D
) predicted by
Real-Time VPP and VPP with trim
parameter power in standard and
second generic form
S
predicted by
Real-Time VPP and second generic
power model relative to V
S
from
standard power model
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 129
Chapter 5 Aerodynamic Force Modelling
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
power [-]
x
C
o
E
/
x
C
o
E
o
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
0 0.2 0.4 0.6 0.8 1
power [-]
C
M
f
/
C
M
f
O
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
.

=
) , ( ) (
0
) , ( ) (
p R z
p R x
eff zCoE eff CoEopt
eff xCoE eff CoEopt


CoE
(5.20)
The power ratio of the moment coefficient in line with the force (R
Mf
) for different
eff
is
plotted against power in Figure 5.44. The scatter in the data is significant but a general
reduction with power can be seen. The scatter is larger at small
eff
, and R
Mf
becomes negative
for small power values which means that C
Mf
changes from decreasing heeling moment (and
increasing bow down pitching moment) to increasing heeling moment (and decreasing bow
down pitching moment). C
Mf
is very small at small
eff
(as shown in Figure 5.12) which
explains the scatter and these trends. For larger
eff
where C
Mf
is more significant the
reduction and scatter is less. R
Mf
is also modelled by a response surface as a function of
eff

and power and equation (5.7) to calculate the moment vector (M) is now written as
. F F CoE M

) , ( ) (
5 . 1
p R C q A
eff Mf eff Mf eff S
+ =
(5.21)
The power ratio of the force perpendicular to the deck plane (R
Fz
) is plotted in Figure 5.45
against power for different
eff
. A general reduction can be guessed but it is difficult to see a
trend. As discussed in chapter 3 measuring the vertical force component accurately in the
wind tunnel is difficult and requires special attention. During the Real-Time VPP
measurements this was not possible so that the large scatter in the data is not surprising. The
changes in C
Fz
due to depowering can be expected to have a very small effect on the
performance prediction of the yacht and are therefore not modelled here because of the
uncertainty in the data. If R
Fz
were to be modelled, its dependency on heel should also be
investigated since chapter 3 showed a strong dependency of C
Fz
on heel. The number of
Figure 5.43: x
CoE
/x
CoEopt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
Figure 5.44: C /C
Mf MfOpt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
130
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8
dependencies of a parameter can be arbitrarily increased to form response volumes in
FRIENDSHIP-Equilibrium so that this is possible.
Figure 5.46 shows the difference in boat speed between the advanced power model, which
additionally depowers x
CoE
and C
Mf
, and the original power model plotted against
T
. The
only noticeable difference in boat speed can be seen at small
T
in V
T
of 10m/s where the
advanced power model predicts up to 1.2% less boat speed. The power values chosen by the
VPP optimiser are very similar to the values from the standard power model solution so that
the speed difference must result directly from depowering x
CoE
and C
Mf
. By depowering C
Mf

the heeling moment is increased, which is reflected by an increase in the heel angle of up to
2.5 compared to the standard power model. Due to modelling the forward shift of x
CoE
, the
weather helm is reduced on average between 1-2 when depowering. For most investigated
sailing conditions the decrease in resistance due to the reduction in weather helm is negated
by the increase in resistance and reduction in drive force due to the increased heel angle. As a
result the boat speed does not change much. Depowering x
CoE
and C
Mf
show the expected
effects but do not significantly influence the yacht performance. The maximum change in
boat speed is 1.2% but typically much less.
5.4.4 Direct Force and Moment Power Model
The different versions of the power model described so far are all based on the most
commonly used method of defining the aerodynamic forces in terms of lift and drag
coefficients and the moments from the centre of effort position. Since the power model is
1
0.92
0.94
0.96
0.98
1
1.02
1.04
1.06
1.08
1.1
0 15 30 45 60 75 90 105 120 135

T
[]
V
S

/
V
S
(
p
o
w
e
r
)

[
-
]
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s advan power VT=6m/s advan power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s advan power VT=4m/s advan power
power [-]
C
F
z
/
C
F
z
O
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
Figure 5.45: C
Fz
/C
FzOpt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
Figure 5.46: Boat speed ratio of V
S
predicted by
Real-Time VPP and advanced
power model, which additionally
adjusts x
CoE
and C , relative to V
Mf S

from standard power model
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 131
Chapter 5 Aerodynamic Force Modelling
empirical and not strongly based on aerodynamic principles, one is not confined to this
traditional approach. In the wind tunnel the forces and moments are measured so that directly
modelling the force and moment coefficients provides a much simpler model. The force and
moment coefficients are still modelled in the body fixed coordinate system and divided into
the optimum force and moment coefficients for the fully powered-up sail trim and the power
ratios to simulate the depowering. In principle this version of the power model is very
similar to the enhanced depowering model for offwind sails described by Ranzenbach and
Teeters (2002) but more degrees of freedom are considered here.
The power ratio of the drive force coefficient (R ) is shown in Figure 5.47 for different
Fx eff

plotted against power. The reduction in R with power does not show a clear trend with
Fx eff

and the data forms a nice curve. Only the measurements for
eff
of 90 reduce less than at
smaller angles. Figure 5.48 shows the power ratio of the side force coefficient (R
Fy
) for
different
eff
plotted against power. The reduction in R
Fy
with power is linear with very little
scatter and effectively no dependence on
eff
. Since power (p) is defined in equation (5.9) as
the ratio of heeling moment to optimum heeling moment, this means that the side force
reduces linearly with heeling moment. If the power ratio of the centre of effort height (R
zCoE
)
is approximated as the heeling moment ratio divided by the side force ratio, this linear
relationship can be used to describe R
zCoE
in the generic power model with
p a a
p
R
p
R
Fy
zCoE
1 0
+
= = ,
(5.22)
where a is the y-axis intercept and a is the slope of the linear regression to the data in
0 1
Figure 5.48. This gives a better fit than the power expression in equation (5.16) but requires
one more parameter. As discussed in section 5.4.3 the power ratio of the force coefficient
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
power [-]
C
F
y
/
C
F
y
O
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
power [-]
C
F
x
/
C
F
x
O
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
Figure 5.47: C
Fx
/C
FxOpt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
Figure 5.48: C /C
Fy FyOpt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
132
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.4 Depowering Model with Trim Parameter Power
perpendicular to the deck plane (R
Fz
) is not modelled due to the inaccuracies in the wind
tunnel data.
Figure 5.49 shows the power ratio of the pitching moment coefficient (R
My
) for different
eff
plotted against power. A clear trend of reducing R
My
with power is seen. The reduction is less
for
eff
of 90 but for the other angles no trend with
eff
is noticeable. The power ratio of the
yawing moment coefficient (R ) shown in Figure 5.50
Mz
has a similar trend of reducing with
power and being almost independent of
eff
. For R
Mz
the reduction with power is however
close to linear again.
The direct version of the power model is implemented in FRIENDSHIP-Equilibrium as a
new force module called HansenRigDirect. Although the power ratios are all effectively
independent of
eff
they are still modelled as response surfaces, which are functions of
eff
and
power (p), so that the force module is more flexible and accurate. The force vector (F) in the
body fixed coordinate system in equation (4.7) is now calculated from
,

=
) (
) , ( ) (
) , ( ) (
eff FzOpt
eff Fy eff FyOpt
eff Fx eff FxOpt
eff S
C
p R C
p R C
q A



F
(5.23)
where C is the optimum drive force coefficient for the fully powered-up sails and C
FxOpt FyOpt

is the optimum side force coefficient. C is given in Figure 5.26
FzOpt
. The moment vector (M)
in the body fixed coordinate system in equation (4.11) is, for the direct power model,
calculated from
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
power [-]
C
M
y
/
C
M
y
O
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
power [-]
C
M
z
/
C
M
z
O
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
Figure 5.49: C
My
/C
MyOpt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
Figure 5.50: C /C
Mz MzOpt
ratio plotted against trim
parameter power for different
effective wind angles (EWA)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 133
Chapter 5 Aerodynamic Force Modelling
,

=
) , ( ) (
) , ( ) (
) (
5 . 1
p R C
p R C
p C
q A
eff Mz eff MzOpt
eff My eff MyOpt
eff MxOpt
eff S



M
(5.24)
where C is the optimum heeling moment coefficient, C
MxOpt MyOpt
is the optimum pitching
moment coefficient and C
MzOpt
is the optimum yawing moment coefficient. The input
parameters required for the HansenRigDirect force module are described in more detail in
appendix C.3.3. As shown in Figure 5.51 the VPP chooses very similar power values with the
HansenRigDirect module compared to the advanced power model version. The direct
power model version is best compared to the advanced version from section 5.4.3 since it
includes the effects of depowering on all force and moment coefficients apart from the
changes to C
Fz
. The boat speed ratio of the direct power model to the advanced power
model in Figure 5.52 shows that very similar predictions can be achieved. Apart from at very
small the difference in predicted boat speed is less than 1%. Figure 5.52
T
also shows the
Real-Time VPP boat speed predictions and the differences are of similar magnitude. Speed
differences up to about 1% seem to be caused by data fitting and not by the way the
depowering is modelled. The boat speed for the fully powered-up condition in V
T
=4m/s from
the direct power model is not identical to the advanced power model because the fully
powered-up input curves are not identical. The direct model uses force and moment
coefficients whereas the advanced power model is based on lift and drag coefficients and
the centre of effort position.
The large increase in predicted boat speed at small
T
when depowering with the direct
model, compared to the advanced model, occurs outside the wind tunnel measurement range
so that it is most likely related to the extension of power ratio surfaces past the measurement
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 15 30 45 60 75

eff
[]
p
o
w
e
r

[
-
]
90
VT=10m/s RT-VPP VT=6m/s RT-VPP
VT=10m/s direct power VT=6m/s direct power
VT=10m/s advan power VT=6m/s advan power
VT=8m/s RT-VPP VT=4m/s RT-VPP
VT=8m/s direct power VT=4m/s direct power
VT=8m/s advan power VT=4m/s advan power
0.98
1
1.02
1.04
1.06
1.08
1.1
1.12
1.14
0 15 30 45 60 75 90 105 120 135

T
[]
V
S

/
V
S
(
a
d
v
a
n
c
e
d

p
o
w
e
r
)

[
-
]
VT=10m/s RT-VPP
VT=10m/s direct power
VT=8m/s RT-VPP
VT=8m/s direct power
VT=6m/s RT-VPP
VT=6m/s direct power
VT=4m/s RT-VPP
VT=4m/s direct power
Figure 5.52: Boat speed ratio of V Figure 5.51: Values of trim parameter power from
Real-Time VPP measurements and
VPP calculations using advanced and
direct form of power
S
predicted by
Real-Time VPP and VPP with direct
power model relative to V
S
from
advanced power model
134
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.5 Assessing the Standard Trim Parameter Model
range. The power ratio surfaces for the direct power model have little dependency on
eff

and are not complex so that they are well constrained. The power ratio surface of the parasitic
drag coefficient in the standard and advanced power models is however strongly dependent
on
eff
and changes rapidly at small
eff
when depowering as shown in Figure 5.21. Although
the surface fits the measurements well it is questionable if the parasitic drag power ratio
continues to increase this rapidly for smaller
eff
where the sails produce much less lift. It is
likely that the standard and advanced models over-predict the parasitic drag at small
eff
when
the sails are depowered which results in a reduced boat speed compared to the direct and
generic power models as is evident in Figure 5.52, Figure 5.40 and Figure 5.42.
Modelling in terms of force and moment coefficients gives a much simpler model and the
changes in the coefficients show clearer and easier to model trends. The coefficients are not
strongly dependent on
eff
so that the response surfaces are less complex to define and easier
to extrapolate. Since the power model is empirically based and requires experimental input
data there is no conceptual reason for adhering to the lift, drag and centre of effort based
description to calculate the aerodynamic forces and moments so that the direct power model
offers a good alternative approach. Although all aerodynamic force and moment components
influence the performance of the yacht, the drive force has by far the greatest influence. The
explicit modelling of the drive force in the direct power model therefore results in a more
accurate performance prediction.
5.5 Assessing the Standard Trim Parameter Model
The newly developed power force model can now be used as a benchmark to assess the
accuracy of the standard trim parameter model described in section 4.2.1. The power force
model simulates the real life depowering accurately, as it is based on Real-Time VPP
measurements. The standard trim parameter force model uses the same fully powered-up
input curves as the power model described in section 5.4.1, to model the aerodynamic
forces in six degrees of freedom. The only difference between the standard trim parameter
force model and the power model is in the depowering. The power model uses the trim
parameter power and power ratio surfaces whereas the standard trim parameter model uses
the trim parameters reef, flat and twist to simulate the depowering. The standard trim
parameters adjust C , C
L D
and z
CoE
as shown in equations (4.1) to (4.5) and (4.8) or (4.9) so
that it is most appropriate to use the standard version of the power model introduced in the
beginning of section 5.4 as a basis for the comparison since it also depowers C , C
L D
and z
CoE
.
However, the difference compared to using another version of the power model can be
expected to be small, since section 5.4 showed that all versions produce results with boat
speeds within typically 1%. The standard trim parameter model is also implemented in the
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 135
Chapter 5 Aerodynamic Force Modelling
HansenRig force module of FRIENDSHIP-Equilibrium so that it is easily possible to
change between the trim parameter power or any combination of the trim parameters reef,
flat and twist with the same fully powered-up input data. For the VPP predictions the
equilibrium is solved in six degrees of freedom as usual and the standard hydrodynamic
model described in appendix C.4 is used.
In the power model the windage forces and moments have not been initially subtracted from
the wind tunnel measurements because of their strong interdependence with the sails as
discussed in chapter 3. The parasitic drag coefficient (C
Dp
) therefore contains a component of
windage drag. In equation of the standard trim parameter model C (4.5)
Dp
is multiplied by
reef squared to model the reduction in sail area. For this operation C
Dp
should not include the
windage drag coefficient (C
Dwindage
) since it cannot be reduced to zero by reefing the sails.
Although chapter 3 showed that the hull drag is affected by the presence of the sails it can be
assumed that a base windage drag is present that cannot be reduced by depowering the sails.
A good approximation of this base windage is the drag of the model without the sails. The
C
Dwindage
obtained from bare poles wind tunnel measurements of the model without sails in
upwind configuration at 0 heel is shown in Figure 5.53. For this study where C
Dp
includes
the windage component equation (4.5) is therefore rewritten as
( )
Di Ds eff Dwindage eff Dwindage eff DpOpt D
C C C r C C C + + + = ) ( ) ( ) (
2
,
(5.25)
to account for the fact that C
Dwindage
cannot be reduced by depowering. Figure 5.54 shows the
optimum parasitic drag coefficient for the model with mainsail and jib measured in the wind
tunnel at 0 heel including and excluding the bare poles windage drag component. The B-
spline curves representing the data in FRIENDSHIP-Equilibrium are also shown.
0
0.05
0.1
0.15
0.2
0.25
0.3
0 30 60 90 120 150 180

eff
[]
C
D
w
i
n
d
a
g
e

[
-
]
Windage hull & rig (wind tunnel)
Windage hull & rig (modelled)
0
0.2
0.4
0.6
0.8
1
1.2
0 15 30 45 60 75 9

eff
[]
C
D

[
-
]
0
CDp (wind tunnel)
CDp (modelled)
CDp-CDwindage (wind tunnel)
CDp-CDwindage (modelled)

Figure 5.54: Parasitic drag coefficient (C Figure 5.53: Windage drag coefficient of hull and
rig in upwind configuration at 0 heel
vs.
Dp
) of
mainsail and jib with and without
windage compnent at 0 heel vs.
eff eff
136
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.5 Assessing the Standard Trim Parameter Model
Four sets of VPP predictions with the standard trim parameter model are presented. Firstly
the VPP is allowed to adjust the trim parameters reef and flat originally introduced by Hazen
(1980). The second set of predictions introduces the twist parameter, which is used together
with flat while the optimisation of reef is disabled. Originally reef was intended to model the
physical reduction of the sail area by reefing but nowadays it is accepted that it is used as a
general measure for depowering a sail. With the introduction of the new trim parameter twist,
reef should not be required as a general depowering parameter anymore and should be used
according to its original intention. Since the sails are not reefed during the wind tunnel tests
reef is not employed together with twist in this study to show how well flat and twist by
themselves perform in modelling the depowering of sails. The theoretical minimum value of
8 is used for the twist weight constant (c ) in equation (4.2).
t
The second set uses the original FRIENDSHIP-Equilibrium implementation where twist is
applied relative to the boom for reducing z
CoE
as shown in equation (4.8). This however
assumes that the twist is applied to a single mainsail that is elliptically loaded and not
influenced by the water surface or hull. It also means that the fully powered-up centre of
effort should be close to 50% up the mainsail luff for the twist model to describe the
increase in induced drag accurately. In this case the yacht has a mainsail and a jib, which
extends to the deck, and chapter 3 has shown that part of the pressure differential across the
sails extends across the hull to the waterline. Krebber and Hochkirch (2006) conducted
Reynolds Avaraged Navier-Stokes Equations (RANSE) calculations on the rig of Dyna for an
upwind sailing condition which show that the loading distribution is of some shape between
triangular and semi-elliptical. The assumption that the sails are semi-elliptically loaded with
the water surface being a mirror plane seems therefore more justifiable so that twist is applied
relative to the DWL as shown in equation (4.9). The fully powered-up centre of effort height
(z
CoEopt
) should then be close to 42% of the mast height to give realistic predictions of the
increase in induced drag. z
CoEopt
in this study is between 42% of the mast height at
eff
of 15
and 37% at
eff
of 60 which also indicates that assuming semi-elliptical loading is the better
approximation here. Set three and set four of the VPP predictions apply twist relative to the
DWL with equation . Set three uses the theoretical minimum value of 8 for c (4.9)
t
as
calculated by Jones (1950) whereas for the fourth set of predictions c
t
is changed to a value of
12 as suggested in unpublished work by Tempia
16
. Table 5.2 summarised the set-up of the
four sets of performance predictions carried out.
Figure 5.55 shows the trim parameter values selected by the VPP for the first set of
predictions in which reef and flat are adjusted. Both reef and flat are adjusted by the VPP but
flat is used predominately at small
eff
whereas reef is reduced over a larger range of
eff
. In
comparison Figure 5.56 shows the values of the trim parameters for the second set of

16
Angelo Tempia, PhD candidate at The University of Auckland, 2001-2003
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 137
Chapter 5 Aerodynamic Force Modelling
Table 5.2: Set-up summary of the four sets of performance predictions carried out to investigate the
standard trim parameter model
Twist weight (c Set Trim parameters Reef application Twist application )
t
Flat, reef 1 Relative to boom - -
Flat, twist 2 - Relative to boom 8
Flat, twist 3 - Relative to DWL 8
Flat, twist 4 - Relative to DWL 12
predictions where flat and twist are free variables. One of the features of the twist parameter
is that it assumes that the reduction in centre of effort height, due to twisting the sails,
increases the induced drag, but that the lift is not affected as discussed in 4.2.1. Flat is
therefore significantly more reduced in V of 8 and 10m/s compared to Figure 5.55
T
since twist
unlike reef does not contribute to reducing C .
L
When optimising twist the reduction in centre of effort height is traded off against the
increase in induced drag. At small
eff
this results in realistic twist values but as
eff
gets larger
an increase in drag has a less significant influence on the performance of the yacht and for
eff

above 90 it is even beneficial. The VPP therefore selects increasing values of twist as
eff

increases and for
eff
larger than 90 the VPP chooses the maximum limit of twist. In the
current implementation twist does not seem to exhibit sensible behaviour for large
eff
. For

eff
where the other trim parameters are not depowered twist is therefore constrained to zero.
Figure 5.56 shows the twist values used in the VPP when applying the constraint. It can be
0
0.2
0.4
0.6
0.8
1
1.2
0 15 30 45 60 75 9

eff
[]
f
l
a
t

[
-
]
0
VT=4m/s
VT=6m/s
VT=8m/s
VT=10m/s
0
0.2
0.4
0.6
0.8
1
1.2
0 15 30 45 60 75 9

eff
[]
f
l
a
t

[
-
]
0
VT=4m/s
VT=6m/s
VT=8m/s
VT=10m/s
0
0.2
0.4
0.6
0.8
1
1.2
0 15 30 45 60 75 9

eff
[]
r
e
e
f

[
-
]
0
VT=4m/s
VT=6m/s
VT=8m/s
VT=10m/s
0
0.1
0.2
0.3
0.4
0.5
0.6
0 15 30 45 60 75 9

eff
[]
t
w
i
s
t

[
-
]
0
VT=4m/s VT=6m/s
VT=8m/s VT=10m/s
Figure 5.56: Flat and twist values chosen by VPP
in different V
Figure 5.55: Flat and reef values chosen by VPP in
different V for a range of
T eff
for a range of
T eff
with
twist relative to boom and c
t
=8
138
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.5 Assessing the Standard Trim Parameter Model
seen that twist increases at small
eff
and then remains at a relatively constant level until it is it
is forced to zero. Without the constraint twist would increase rapidly as
eff
approaches 90.
Figure 5.57 shows the values of the trim parameters for the third set of results where flat and
twist are again the free variables but twist is applied relative to the water surface when
reducing z
CoE
. Twist is again constrained to zero for
eff
where other predictions and trim
parameters indicate that the sails do not need to be depowered. Compared to the previous data
set, where twist was relative to the boom, significantly more twist is applied. The same
amount of twist now reduces z
CoE
less whereas the increase in C
Di
is the same. The larger
twist value increases C but also reduces z
Di CoE
which allows C
L
to be larger. The increase in
C
L
is shown by the smaller reduction in flat. The flat and twist values for the fourth set of
predictions where the twist weight constant (c ) is increased from 8 to 12 are shown in
t
Figure
5.58. Compared to the previous case less twist in applied due to the larger increase in C
Di
at
any given twist value. z
CoE
is hence less reduced than before so that more flat is applied,
especially at larger
eff
where the reduction in twist is also more significant.
The VPP predictions result in a large number of parameters that can be compared. For the
depowering study the three variables that are adjusted by the trim parameters are of most
interest together with the resulting performance, which is best indicated by the boat speed
(V ) plotted against true wind direction (
S T
). Probably the most important parameter in this
study is the centre of effort height (z
CoE
) so that it is investigated first. Figure 5.59 shows z
CoE

0
0.2
0.4
0.6
0.8
1
1.2
0 15 30 45 60 75 9

eff
[]
f
l
a
t

[
-
]
0
VT=4m/s
VT=6m/s
VT=8m/s
VT=10m/s
0
0.2
0.4
0.6
0.8
1
1.2
0 15 30 45 60 75 9

eff
[]
f
l
a
t

[
-
]
0
VT=4m/s
VT=6m/s
VT=8m/s
VT=10m/s
0
0.1
0.2
0.3
0.4
0.5
0.6
0 15 30 45 60 75 9

eff
[]
t
w
i
s
t

[
-
]
0
VT=4m/s VT=6m/s
VT=8m/s VT=10m/s
0
0.1
0.2
0.3
0.4
0.5
0.6
0 15 30 45 60 75 9

eff
[]
t
w
i
s
t

[
-
]
0
VT=4m/s VT=6m/s
VT=8m/s VT=10m/s
Figure 5.58: Flat and twist values chosen by VPP
in different V
Figure 5.57: Flat and twist values chosen by VPP
in different V for a range of
T eff
with
twist relative to DWL and c
for a range of
T eff
with
twist relative to DWL and c
t
=8
t
=12
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 139
Chapter 5 Aerodynamic Force Modelling
plotted against
eff
for different true wind speeds. The symbols show z
CoE
obtained from VPP
predictions with the power model that serves as the benchmark for this study and is
developed based on the Real-Time VPP measurements. The lines represent the values
calculated with the standard trim parameter model for the four data sets. The sails are fully
powered-up at V of 4m/s for all effective wind angles so that the results for V
T T
of 2m/s are
not shown since the results are identical. The VPP solutions for V
T
above 10m/s do not
represent a fair comparison since, for many effective wind angles, the sails in the wind tunnel
were depowered much more than is typical in real life where the sails would be physically
reefed in these conditions.
The comparison results for V
T
between 4m/s and 10m/s are therefore discussed because they
represent a realistic wind range in which sails are depowered without physically reefing them.
Figure 5.59(a) shows that the traditional parameters reef and flat reduce z
CoE
significantly less
than the power model at small
eff
. Only at larger
eff
where the sails are only depowered by
a small amount do the result in V of 8 and 10m/s agree well. For V of 6m/s z
T T CoE
is not
reduced by the standard trim parameter model with reef and flat whereas the power model
reduces z
CoE
for
eff
up to 45. Figure 5.59(a) also shows the results from the second data set,
which optimise flat and twist relative to the boom. z
CoE
is reduced more compared to using
reef and flat but for small
eff
the reduction is still significantly less than predicted by the
power model. For larger
eff
the reduction in z
CoE
is however too large when twist is used
instead of reef. As discussed before twist is used excessively at larger
eff
as the increase in
induced drag is less important to the performance of the yacht.
3
3.5
4
4.5
5
5.5
6
6.5
7
7.5
8
0 15 30 45 60 75 9

eff
[]
z
C
o
E

[
m
]
0
VT=10m/s power VT=6m/s power
VT=10m/s flat, reef VT=6m/s flat, reef
VT=10m/s flat, twist VT=6m/s flat, twist
VT=8m/s power VT=4m/s power
VT=8m/s flat, reef VT=4m/s flat, reef
VT=8m/s flat, twist VT=4m/s flat, twist
3
3.5
4
4.5
5
5.5
6
6.5
7
7.5
8
0 15 30 45 60 75 9

eff
[]
z
C
o
E

[
m
]
0
VT=10m/s power VT=6m/s power
VT=10m/s flat, twist ct=8 VT=6m/s flat, twist ct=8
VT=10m/s flat, twist ct=12 VT=6m/s flat, twist ct=12
VT=8m/s power VT=4m/s power
VT=8m/s flat, twist ct=8 VT=4m/s flat, twist ct=8
VT=8m/s flat, twist ct=12 VT=4m/s flat, twist ct=12
(a) Twist applied relative to boom, c
t
=8 (b) Twist applied relative to DWL, c
t
=8 or 12
Figure 5.59: Centre of effort height (z
CoE
) for four true wind speeds (V
T
) from VPP calculations using
trim parameter power and different combinations and versions of reef, flat and twist
140
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.5 Assessing the Standard Trim Parameter Model
Although not shown here calculations with all three trim parameters have also been
conducted and z
CoE
is very similar to the flat and twist results. Figure 5.59(b) shows z
CoE

when optimising twist relative to the DWL instead of the boom. For the predictions with c
t

kept at 8, z
CoE
is now significantly more reduced in most cases compared to the benchmark
calculations. Only at small
eff
in V of 6 and 8m/s is the fit reasonable. At larger
T eff
the
reduction in z
CoE
is excessive. In V of 10m/s z
T CoE
is almost reduced to the boom height of
2.45 metres. Increasing c to 12, increases z
t CoE
evenly for all
eff
, so that agreement with the
benchmark results is good for small
eff
but at large
eff
the centre of effort is still reduced too
much. The amount of centre of effort reduction can be adjusted by varying c
t
and the
reference height for the twist parameter, which are not interchangeable operations since C
Di
is
a function of twist squared as shown in equation (4.2), but the general shape of the reduced
z
CoE
curves remains the same and does not fit the wind tunnel measurements well.
Figure 5.60 shows the lift coefficient (C
L
) for the four data sets presented in the same way as
z
CoE
before. As expected the depowered C curves in Figure 5.60
L
show that the sails need to
be depowered more as V increases and as
T eff
decreases. Figure 5.60(a) shows that
optimising the two traditional trim parameters reef and flat under-predicts C
L
very slightly at
small
eff
and fits the benchmark data reasonably well at larger
eff
. Optimising twist instead
of reef increases C
L
so that the fit is better at small
eff
but worse at larger
eff
. In general C
L

from the first two data sets agree well with the power model benchmark predictions. The
predictions not shown that use all three trim parameters produce C
L
that lies in between the
two data sets. C obtained when optimising twist relative to the DWL instead of the boom is
L
0
0.4
0.8
1.2
1.6
2
2.4
2.8
0 15 30 45 60 75

eff
[]
C
L

[
-
]
90
VT=10m/s power VT=6m/s power
VT=10m/s flat, reef VT=6m/s flat, reef
VT=10m/s flat, twist VT=6m/s flat, twist
VT=8m/s power VT=4m/s power
VT=8m/s flat, reef VT=4m/s flat, reef
VT=8m/s flat, twist VT=4m/s flat, twist
0
0.4
0.8
1.2
1.6
2
2.4
2.8
0 15 30 45 60 75 9

eff
[]
C
L

[
-
]
0
VT=10m/s power VT=6m/s power
VT=10m/s flat, twist ct=8 VT=6m/s flat, twist ct=8
VT=10m/s flat, twist ct=12 VT=6m/s flat, twist ct=12
VT=8m/s power VT=4m/s power
VT=8m/s flat, twist ct=8 VT=4m/s flat, twist ct=8
VT=8m/s flat, twist ct=12 VT=4m/s flat, twist ct=12
(a) Twist applied relative to boom, c
t
=8 (b) Twist applied relative to DWL, c
t
=8 or 12
Figure 5.60: Lift coefficient (C ) for four true wind speeds (V
L T
) from VPP calculations using trim
parameter power and different combinations and versions of reef, flat and twist
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 141
Chapter 5 Aerodynamic Force Modelling
given in Figure 5.60(b). It can be seen that C is modelled well only at the smallest
L eff
and is
over-predicted at other
eff
due to the excessive use of twist. As discussed before in relation to
the twist values shown in Figure 5.57 and Figure 5.58 increasing the twist weight constant (c
t
)
increases the induced drag penalty so that less twist is applied which leads to an increase in
z
CoE
and a reduction in C . This reduction in C due to changing c from 8 to 12 can be seen in
L L t
Figure 5.60(b).
Figure 5.61 shows the total drag coefficient (C
D
) presented in the same way as z
CoE
and C
L

before. The comparison of C
D
is most applicable since the individual drag components are
depowered in different ways by the different models and trim parameters. The power model
for example does not account for changes in induced drag due to twist explicitly, as the twist
parameter does, but implicitly it is considered by the parasitic drag component which is a
function of power. Figure 5.61(a) shows that the traditional reef and flat model generally
under-predicts the total drag coefficient at small
eff
and over-predicts it at larger
eff

compared to the power model. Replacing reef with twist increases the drag coefficient but at
small
eff
it is still less than predicted by the power model. For V
T
of 8 and 10m/s the drag
coefficient becomes larger than for the fully powered-up condition due to the large twist
values that are chosen by the VPP. Large twist values increase the induced drag and but also
reduce the centre of effort so that C has to be reduced less for a given heeling moment.
L
Figure 5.60(a) shows that larger C
L
values are chosen by the VPP when the twist parameter is
used instead of reef. The larger C leads to a further increase in C and C
L Di Ds
because of their
dependency on C
2
L
. Comparing Figure 5.61(a) and Figure 5.61(b) shows that optimising twist
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 15 30 45 60 75 9

eff
[]
C
D

[
-
]
0
VT=10m/s power VT=6m/s power
VT=10m/s flat, reef VT=6m/s flat, reef
VT=10m/s flat, twist VT=6m/s flat, twist
VT=8m/s power VT=4m/s power
VT=8m/s flat, reef VT=4m/s flat, reef
VT=8m/s flat, twist VT=4m/s flat, twist
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 15 30 45 60 75 9

eff
[]
C
D

[
-
]
0
VT=10m/s power VT=6m/s power
VT=10m/s flat, twist ct=8 VT=6m/s flat, twist ct=8
VT=10m/s flat, twist ct=12 VT=6m/s flat, twist ct=12
VT=8m/s power VT=4m/s power
VT=8m/s flat, twist ct=8 VT=4m/s flat, twist ct=8
VT=8m/s flat, twist ct=12 VT=4m/s flat, twist ct=12
(a) Twist applied relative to boom, c
t
=8 (b) Twist applied relative to DWL, c
t
=8 or 12
Figure 5.61: Total drag coefficient (C ) for four true wind speeds (V
D T
) from VPP calculations using trim
parameter power and different combinations and versions of reef, flat and twist
142
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.5 Assessing the Standard Trim Parameter Model
relative to DWL further increases C
D
. From Figure 5.61(b) it can be seen that by reducing c
t

from 12 to 8 the drag coefficient also increases. The increase in C
D
is again caused by the
increase in twist and the associated increase in C as shown in Figure 5.60(b). For most
L eff
,
the traditional reef and flat model provides the better C
D
predictions because the C
L

predictions are lower and the parasitic drag coefficient (C
Dp
) is reduced by reef as shown in
equation (5.25). Only at small
eff
is C
D
modelled more accurately by flat and twist since twist
models the increase in induced drag, which is most important at small
eff
, and z
CoE
is
predicted reasonably well so that C is close to the benchmark predictions.
L
Finally the predicted performance of the different depowering methods is compared. Figure
5.62(a) shows the boat speed (V
S
) for the first and second set of predictions as a fraction of
the benchmark performance obtained with the power model. When optimising reef and flat
V is increasingly under-predicted as V
S T
increases and more depowering is required. The
results for V
T
of 4m/s are only shown as a check to confirm that the fully powered-up speeds
are identical to the benchmark predictions. In V
T
of 10m/s the boat speed is under-predicted
by up to 4% with the reef and flat model. When optimising twist instead of reef the boat
speed is under-predicted by a smaller amount in most conditions. V
S
is however under-
predicted more severely at large true wind angles ( ) in V
T T
of 10m/s with a maximum speed
difference of 5%. As described before in relation to Figure 5.52 increases in V
S
compared to
the benchmark V can be expected at small
S eff
due to the modelling of the power ratio
surfaces in the standard power model outside the wind tunnel measurement range. The
smallest
eff
investigated in the wind tunnel is 20, which corresponds to of between 30 in
T
0.94
0.96
0.98
1
1.02
1.04
1.06
1.08
1.1
0 15 30 45 60 75 90 105 120 135

T
[]
V
S
/
V
S
(
p
o
w
e
r
)

[
-
]
VT=4m/s flat, reef
VT=6m/s flat, reef
VT=8m/s flat, reef
VT=10m/s flat, reef
VT=4m/s flat, twist
VT=6m/s flat, twist
VT=8m/s flat, twist
VT=10m/s flat, twist
0.94
0.96
0.98
1
1.02
1.04
1.06
1.08
1.1
0 15 30 45 60 75 90 105 120 135

T
[]
V
S
/
V
S
(
p
o
w
e
r
)

[
-
]
VT=4m/s flat, twist ct=8
VT=6m/s flat, twist ct=8
VT=8m/s flat, twist ct=8
VT=10m/s flat, twist ct=8
VT=4m/s flat, twist ct=12
VT=6m/s flat, twist ct=12
VT=8m/s flat, twist ct=12
VT=10m/s flat, twist ct=12
(a) Twist applied relative to boom, c
t
=8 (b) Twist applied relative to DWL, c
t
=8 or 12
Figure 5.62: Boat speed (V
S
) four true wind speeds (V
T
) from VPP calculations using different
combinations and versions of reef, flat and twist as ratio of V
S
obtained using power model
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 143
Chapter 5 Aerodynamic Force Modelling
V of 10m/s and 40 in V
T T
of 4m/s, and inaccuracies in the predictions at smaller angles are
not disconcerting.
Figure 5.62(a) gives a good indication of how well flat performs together with reef or twist in
modelling V . It is however important to consider the predicted V
S S
only in conjunction with
z
CoE
, C and C
L D
as it should be of primary importance to model the changes in the forces and
moments correctly with a depowering model. If this is achieved the V
S
predictions should also
be accurate. The boat speed predictions in Figure 5.62(b) for twist optimised relative to the
DWL should be considered with caution since z
CoE
, C and C
L D
are inadequately modelled in
many conditions. It can for example be seen that the V predictions are good for
S T
between
45 and 65 with c
t
=8 but that, for the corresponding
eff
of approximately 25-40, z
CoE
, C
L

and C
D
do not necessarily agree well with the power model data.
Based on this comparison it can be concluded that the traditional reef and flat model provides
the better results for VPP predictions over a large range of
eff
. Twist only provides good
predictions at small
eff
since excessive twist is applied at larger
eff
. In the current form twist
could be used for a VPP optimisation for velocity made good (VMG) or other predictions
where also only small
eff
are considered. At larger
eff
the twist parameter unfortunately does
not give realistic results and should not be used in this implementation.
5.6 Improved Flat and Twist Depowering Model
The comparison of the standard trim parameter model to the power model based on the
Real-Time VPP measurements presented in section 5.5 provides important information on
how to improve the standard trim parameter model. The reef parameter should only be used
according to its original propose of reducing the sail area while flat and twist should model
the aerodynamic aspects of depowering. In the current model twist however only provides
sensible results for small
eff
and at larger angles excessive twist is applied. The first
enhancement to the twist parameter is based on unpublished work by Tempia
17
and accounts
for the fact that the fully powered-up sails may not have a loading distribution that induces
constant downwash and minimises the induced drag. In other words that z
CoEopt
does not equal
z and twist should not be zero when the sails are fully powered-up.
CoEminDi
The total sail twist is therefore divided into two components, the twist due to depowering of
the sails (t), which is defined as
CoEopt
CoE
z
z
t = 1 ,
(5.26)

17
Angelo Tempia, PhD candidate at The University of Auckland, 2001-2003
144
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.6 Improved Flat and Twist Depowering Model
and the base twist when the sail is fully powered-up (t ), which is given by
0
CoEminDi
CoEopt
z
z
t = 1
0
.
(5.27)
is denoted here as t and by substituting z The total twist defined in equation (4.3)
CoE
and
z with equations (5.26) and (5.27) can be written as
CoEminDi
0 0
1 tt t t
z
z
t
CoEminDi
CoE
+ = = .
(5.28)
The induced drag coefficient (C ) definition in equation (4.2)
Di
can be rewritten to include the
new definition for the total twist so that
( )
2
0 0
2 2 2
) ( 1
) (
tt t t c
AR
f r C
C
t
eff Lopt
Di
+ + =

,
(5.29)
where the efficiency (e) is now included in the t expression. 1/e describes the increase in C
Di 0

compared to the ideal value of C
2
L
/(AR) due to the sails not having a loading distribution
that induces constant downwash when the sails are fully powered-up. The same effect is now
modelled by t
0
in terms of a difference in centre of effort height. When expanding the squared
bracket in equation
2 2 2 2
the higher order terms t t -2t t -2tt (5.29)
0 0 0
can be ignored since the sum
is small for small values of t and t and the equation simplifies to
0
( ) ) 2 ( 1
) (
0
2
0
2
2 2 2
tt t t c
AR
f r C
C
t
eff Lopt
Di
+ + + =

.
(5.30)
When the sails are fully powered-up t is zero and the increase in C
Di
compared to the ideal
value is 1+c
2
t . This increase compared to ideal value is also defined as 1/e so that equation
t 0
(5.30) can be written as

+ + = ) 2 (
1
) (
0
2
2 2 2
tt t c
e AR
f r C
C
t
eff Lopt
Di

.
(5.31)
This extended expression for C
Di
is implemented in the FRIENDSHIP-Equilibrium force
module HansenRig instead of the original C definition in equation (4.2)
Di
and performance
predictions are carried out. An additional input parameter for the centre of effort height at
which the induced drag is minimal (z ) is required to define t
CoEminDi 0
. As discussed before in
section 5.5 assuming semi-elliptical loading with the water surface as the mirror plane seems
feasible so that t and now also t are defined relative to the DWL and z
CoEminDi 0
is taken as 42%
of the rig height. Figure 5.63 shows t obtained from z
CoEopt 0
measured in the wind tunnel and
from the B-spline curve of z
CoEopt
modelled in FRIENDSHIP-Equilibrium. For the VPP
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 145
Chapter 5 Aerodynamic Force Modelling
-0.1
0
0.1
0.2
0.3
0 15 30 45 60 75 9

eff
[]
t
0

[
-
]
0
Mainsail & jib (wind tunnel)
Mainsail & jib (modelled)

Figure 5.63: Base twist (t
0
) for fully powered-up
mainsail and jib vs.
eff
assuming
z
CoEminDi
is 42% of mast height
predictions the c value of 8 determined by Jones (1950) is again used. The thicker lines in
t
Figure 5.64 show z
CoE
obtained by the VPP optimising flat and twist with the new expression
of C . As before twist needs to be constrained to zero at large
Di eff
. Comparing to Figure
5.59(b) it can be seen that the shape of the z
CoE
curves does not significantly change by
introducing the t expression. For the same c of 8, z
0 t CoE
is less reduced when the t
0
expression
is used since the increase in induced drag is larger for a given twist value so that the VPP
chooses less twist. The slope of the z
CoE
curves is a little bit reduced when the t
0
expression is
included since t increases for
eff 0
between 15 and 60 and with it the induced drag penalty;
but z
CoE
is still excessively reduced at larger
eff
and eventually needs to be constrained to
zero.
3
3.5
4
4.5
5
5.5
6
6.5
7
7.5
8
0 15 30 45 60 75 9

eff
[]
z
C
o
E

[
m
]
0
VT=10m/s power VT=6m/s power
VT=10m/s t0 VT=6m/s t0
VT=10m/s t0, CDp VT=6m/s t0, CDp
VT=8m/s power VT=4m/s power
VT=8m/s t0 VT=4m/s t0
VT=8m/s t0, CDp VT=4m/s t0, CDp
0
0.4
0.8
1.2
1.6
2
2.4
2.8
0 15 30 45 60 75 9

eff
[]
C
L

[
-
]
0
VT=10m/s power VT=6m/s power
VT=10m/s t0 VT=6m/s t0
VT=10m/s t0, CDp VT=6m/s t0, CDp
VT=8m/s power VT=4m/s power
VT=8m/s t0 VT=4m/s t0
VT=8m/s t0, CDp VT=4m/s t0, CDp
Figure 5.64: z
CoE
for four V
T
from VPP
calculations using power model or
flat and twist with t
Figure 5.65: C for four V
L T
from VPP calculations
using power model or flat and twist
with t
2
0
and C
Dp
function
of (C
L
/C
Lopt
)
2
0
and C function of (C /C
Dp L Lopt
)
146
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.6 Improved Flat and Twist Depowering Model
Comparing the thicker lines in Figure 5.65 to Figure 5.60(b) shows that C
L
is slightly reduced
by the introduction of the t expression but is still significantly higher than C
L 0
predicted by
the power model, which is still regarded as the benchmark. Similarly the comparison of C
D

in Figure 5.66 and Figure 5.61(b) shows that the introduction of the t
0
expression does not
influence the predicted C
D
much. The combination of these small changes due to the t
0

expression leads to a reduction in V shown by the comparison of Figure 5.67 and
S
Figure
5.62(b). For most sailing conditions this means that V
S
is even more under-predicted now.
The introduction of the t
0
expression by itself cannot fix the deficiencies of the current model
but t
0
should nevertheless be employed since it addresses an over simplification in the
original twist model theory.
Figure 5.68 shows the modelled total drag coefficient (C
D
) calculated with equation (5.25)
plotted against C
D
measured in the wind tunnel with the Real-Time VPP for different
eff
and
sail trims. For each wind tunnel measurement the values of flat and twist are calculated by
rearranging equations to model C (4.1) and (5.26) and then used in equation (5.25)
D
. Figure
5.68(a) and Figure 5.68(b) show C
D
calculated with the original definition of C in equation
Di
with c (4.2)
t
of 8 and 12 respectively. Figure 5.68(c) shows C
D
calculated from the extended
expression of C with t and c
Di t 0
of 8. It can be seen that the calculated and measured values
agree well for
eff
of 20 and 25 but from 30 to 90 the model increasingly over-predicts C
D

when the sails are depowered. C
D
slightly increases when c
t
changes from 8 to 12 and
including the t expression with c
t 0
=8 gives similar results to the original twist model with
c
t
=12. Finally Figure 5.68(d) shows C
D
calculated with C not being increased by twist (i.e.
Di
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 15 30 45 60 75

eff
[]
C
D

[
-
]
90
VT=10m/s power VT=6m/s power
VT=10m/s t0 VT=6m/s t0
VT=10m/s t0, CDp VT=6m/s t0, CDp
VT=8m/s power VT=4m/s power
VT=8m/s t0 VT=4m/s t0
VT=8m/s t0, CDp VT=4m/s t0, CDp
0.94
0.96
0.98
1
1.02
1.04
1.06
1.08
1.1
0 15 30 45 60 75 90 105 120 135

T
[]
V
S
/
V
S
(
p
o
w
e
r
)

[
-
]
VT=4m/s t0
VT=6m/s t0
VT=8m/s t0
VT=10m/s t0
VT=4m/s t0, CDp
VT=6m/s t0, CDp
VT=8m/s t0, CDp
VT=10m/s t0, CDp
Figure 5.66: C for four V Figure 5.67: V
D T
from VPP calculations
using power model or flat and twist
with t
S
for four V
T
from VPP calculations
using flat and twist with t
0
and C
Dp
function of (C
L
/C
Lopt
)
2
0
and C
Dp

function of (C
2
/C
L Lopt
) plotted as ratio
of V
S
obtained with power model
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 147
Chapter 5 Aerodynamic Force Modelling
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1 1.2
C
D
(measured) [-]
C
D

(
c
a
l
c
u
l
a
t
e
d
)

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
No error
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1 1.2
C
D
(measured) [-]
C
D

(
c
a
l
c
u
l
a
t
e
d
)

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
No error
(a) Original twist expression with c (b) Original twist expression with c
c
t
=0). C
D
at
eff
of 20 and 25 is now under-predicted and consequently the modelling at
eff

of 30 and 40 is more accurate. From the previous discussion this is the expected behaviour
but the changes in C
D
due to twist are very small compared to the modelling errors at larger

eff
.
From the total drag coefficient (C
D
) measured in the wind tunnel with the Real-Time VPP the
parasitic drag coefficient excluding the windage component can be calculated by rearranging
equation (5.25). The values for flat and twist are calculated from the wind tunnel
measurements as before and the C
Di
expression including t
0
with c
t
=8 is used. The current
trim parameter model assumes that C
Dp
is independent of sail trim as long as reef is not
optimised. The ratio C
L
/C
Lopt
, which is equivalent to flat, reduces as the sails are depowered
so that plotting C
Dp
against it in Figure 5.69 yields whether C
Dp
is independent of
depowering. At
eff
of 20 and 25 the parasitic drag coefficient seems independent of the sail
t
=8
t
=12

0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1 1.2
C
D
(measured) [-]
C
D

(
c
a
l
c
u
l
a
t
e
d
)

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
No error
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1 1.2
C
D
(measured) [-]
C
D

(
c
a
l
c
u
l
a
t
e
d
)

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
No error
(c) Twist expression with t
0
and c
t
=8 (d) C
Di
not increased by twist (i.e. c
t
=0)
Figure 5.68: Calculated C plotted against C
D D
measured in the wind tunnel with the Real-Time VPP for
sails at different depowering levels for a range of effective wind angles (EWA)
148
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.6 Improved Flat and Twist Depowering Model
trim, but at 60 and 90, and to some extent also at 30 and 40, C
Dp
reduces when the sails
are depowered. C increases with
DpOpt eff
and the reduction at
eff
of 60 and 90 is significant
and needs to be considered to improve the modelling of C
D
at large
eff
. This dependency of
C
Dp
on trim is another reason why the traditional model with reef and flat provides reasonable
results. Although for the wrong reasons, reef models the reduction of C
Dp
. Clearly the
reduction of C
Dp
is not caused by a reduction in sail area since the sails are not physically
reefed in the wind tunnel. The concept of a dependency of C
Dp
on reef is not completely new
as C
Dp
is defined as a function of reef squared and flat for upwind sailing cases by van
Oossanen (1993) but no explanation for the reasons behind it is given.
2
In the current model the separation drag coefficient (C ) is defined as a function of C
Ds L
and
the separation constant (c
s
), which is usually taken as 0.0016 for upwind sailing conditions
and 0.0019 for downwind cases (Kerwin, 1978). This expression is taken from wing
aerodynamics where it is established that C
2
can be described as functions of C and C
Ds L L
for
small angles of attack up to the maximum C . For small
L eff
this translates well to sails but at
larger
eff
where C
Lopt
is well below the maximum C
L
the separation drag reduces for a
constant c . However, the flow over the sails at larger
s eff
is partially separated so that a
significant increase in drag is to be expected and therefore C increases with
Dp eff
. Although
this drag component is also caused by separation, it is not directly a function of C
2
. C
L Dp

usually increases monotonically up to
eff
of 180, whereas C
L
decreases toward zero at 180.
C
Dp
can be seen as a form of viscous pressure drag that is caused by an obstruction to the
flow. The sail area projected perpendicular to the onset flow increases with
eff
so that the
flow is more obstructed and more separation occurs. When depowering by easing the sails
and thereby reducing C
L
the obstruction to the flow and the separation also reduces. If the
sails were eased completely the obstruction to the flow and the separation reduce to zero. If
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0 0.2 0.4 0.6 0.8 1
C
L
/C
Lopt
[-]
C
D
-
C
D
w
i
n
d
a
g
e

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0 0.2 0.4 0.6 0.8 1
(C
L
/C
Lopt
)
2
[-]
C
D
p
-

(
C
D
w
i
n
d
a
g
e
+
0
.
0
2
8
9
)

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
EWA 30 modelled
EWA 40 modelled
EWA 60 modelled
EWA 90 modelled
Figure 5.69: Parasitic drag coefficient (C Figure 5.70: Parasitic drag coefficient (C
Dp
)
excluding windage calculated from
Real-Time VPP measurements plotted
against C
Dp
)
excluding windage calculated from
Real-Time VPP measurements plotted
against (C
2
/C /C
L Lopt
for different
eff L Lopt
) for different
eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 149
Chapter 5 Aerodynamic Force Modelling
the reduction is proportional to the area projected perpendicular to the onset flow C
Dp
is a
function of flat as introduced by van Oossanen (1993). Figure 5.69 shows however that a
linear reduction with flat does not fit the wind tunnel data very well. It may be more
appropriate to assume that the reduction is proportional to the separation so that C
Dp
is a
function of the reduction in lift squared. Figure 5.70 shows C
Dp
plotted against the lift ratio
squared, which equals flat squared. For
eff
of 60 this model fits very well whereas the
agreement is less good at the other angles. But considering the large range of
eff
and
extensive variations in sail trim included in the data, the model provides reasonable results
that are certainly much more accurate than keeping C constant.
Dp
C in Figure 5.69
Dp
includes a small component that is independent of sail trim since the
measurements at
eff
of 20 and 25 do not show a trend with C /C
L Lopt
. The constant
component is estimated as 0.0289 so that C at
DpOpt eff
of 20 and 25 is close to zero. One
contributor to the constant component is the skin friction of the sails, which can be estimated
from the skin friction on a flat plate to be about 0.0165. The strings used to attach the sails
add further drag so that most of the constant component is accounted for. The skin friction of
the sails should be independent of
eff
so that 0.0289 is assumed constant not only for changes
in trim but also for different
eff
. In Figure 5.70 a constant value of 0.0289 has been
subtracted from C
Dp
in addition to the hull and rig windage, which improves the fit of the data
to the model at small
2
eff
. The dependency of C on (C /C
Dp L Lopt
) can be included in equation
(5.25) so that it becomes
( )
Di Ds eff Dwindage eff Dwindage eff DpOpt D
C C C f r C C C + + + = ) ( ) ( ) (
2 2
,
(5.32)
where C
Dwindage
now also includes an additional component of 0.0289 to account for the skin
friction of the sails. Figure 5.71 shows the modelled total drag coefficient (C
D
) calculated
with equation plotted against C (5.32)
D
measured in the wind tunnel with the Real-Time VPP
for different
eff
and sail trims. Compared to Figure 5.68(c) the calculation of C
D
is
significantly improved for
eff
from 30 to 90. At
eff
of 90 some errors are still present but
because C
D
can vary significantly when trimming the sails since the flow is partially
separated and C
D
has a very small influence on the performance of the yacht, the model can
be considered to give very good results.
Equation (5.32) is also implemented in the HansenRig force module and VPP calculations
conducted. Looking back at Figure 5.66 it can be seen that C
D
is more reduced for many
eff

when C
2
is modelled as a function of (C /C
Dp L Lopt
) so that the agreement with the benchmark
results is generally improved. C in Figure 5.65
L
is also reduced more because a reduction in
C is now associated with a larger reduction in C
L D
. The agreement with the benchmark results
is very good for the improved drag model. z
CoE
in Figure 5.64 is less reduced at some
eff

because less heeling moment is generated as C
L
is smaller. At most
eff
, with in V of 8 and
T
150
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.6 Improved Flat and Twist Depowering Model
10m/s, z
CoE
is however still reduced excessively. Figure 5.67 shows that V
S
is now within
1% of the benchmark performance, whereas it was under-predicted by up to 3.5% before.
With the improved drag model V is under-predicted by up to 1% at
S T
of around 45. At
larger the boat speed is over-predicted by up to 1% in V
T T
of 10m/s. The development of the
power model in section 5.4 showed that errors in V
S
of 1% need to be expected due to
measurement inaccuracies and the associated curve fitting. The performance prediction of the
improved drag model is as good as can be expected but the modelling of z
CoE
at larger
eff
and
the resulting C
D
is still not accurate.
The main reason for the excessive use of twist at larger
eff
is the fact that changes in drag
have less influence on the performance as
eff
increases and that twist influences only the
induced drag component, which can be a relatively small compared to the total drag.
Although a general trim parameter model cannot necessarily be expected to work for all
eff
,
it should be useable over a range where depowering is important in normal sailing conditions,
which in this study is up to a
eff
of around 75. The use of twist at larger
eff
could be better
controlled by making c or z
t CoEminDi
functions of
eff
. As
eff
increases the choice of c
t
and
z becomes more important to control twist and for
CoEminDi eff
close to 90, large changes are
necessary. c for example would need to be increased to about 100 at
t eff
of 90 for the VPP to
chose a sensible twist value. In reality such a significant increase in C with
Di eff
when
depowering cannot be expected and varying c or z
t CoEminDi
would not describe the physics well
but merely aid the data fitting. Another approach to handle the excessive use of twist could be
to limit the maximum value of twist and
eff
up to which twist is applied. As a maximum
value of twist Jackson (2001) suggests 0.25 since it corresponds to the z
CoE
where lifting-line
theory predicts the reversal of the loading at the tip of the sail. From the centre of effort
height reduction shown in Figure 5.24 it can be calculated that twist values up to 0.42 have
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1 1.2
C
D
(measured) [-]
C
D

(
c
a
l
c
u
l
a
t
e
d
)

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
No error
0
0.2
0.4
0.6
0.8
1
1.2
0 0.2 0.4 0.6 0.8 1
1-z
CoE
/z
CoEopt
[-]
C
L
/
C
L
o
p
t

[
-
]
EWA 20
EWA 25
EWA 30
EWA 40
EWA 60
EWA 90
CL(t,t0=0.02,ctL)
CL(t,t0=0.12,ctL)
1-1*twist
1-2*twist
Reduction by flat
due to twist (f
t
)
Reduction
by flat (f )
Figure 5.72: C Figure 5.71: C calculated from twist with t
D 0
and
c
/C
L Lopt
from depowering sails with
Real-Time VPP at different
2
t
=8 and C function of (C /C
Dp L Lopt
)
eff

plotted against 1-z plotted against C
D
measured in the
wind tunnel with the Real-Time VPP
CoE
/z
CoEopt

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 151
Chapter 5 Aerodynamic Force Modelling
been achieved in the wind tunnel. At this trim the sails were severely back-winding and it
represents the extreme that is possible before the sails begin flapping. However, it is a
feasible sail trim and it shows that twist values greater than 0.25 are possible. It is therefore
not straightforward to determine the maximum limit for twist. Similarly it is not obvious how
to define the maximum
eff
for which twist may be applied as it depends on V
T
and the force
modelling. In addition to not being easy to obtain, defining limits for twist only stops
unrealistic behaviour but does not fundamentally improve the agreement with the benchmark
results obtained from the wind tunnel data. A better description of the physics seems to be
necessary to improve the use of twist.
One of the fundamental assumptions of Jones (1950) and therefore also Jackson (2001) when
defining the increase in C with the twist parameter is that C
Di L
remains constant. It is assumed
that, as the angles of attack at the head of the sail is reduced, the angle of attack at the foot of
the sail is increased so that the twisted sail produces the same C with a lower z
L CoE
and an
increased C
Di
. Whereas this is feasible for aerofoils at small angles of attack () by varying
the washout since C changes linearly with , an optimally trimmed sail at larger
L eff
operates
at close to or above the maximum C so that the sectional C
L L
at the foot of the sail cannot
be increased by increasing and C
L
cannot be kept constant. Even at small
eff
where C
L

could be kept constant by further sheeting in the foot of the sail this would reduce C
Fx
since
the fully powered-up sail is trimmed to have the best lift to drag ratio. In reality a reduction in
C
L
due to twisting off the sails is therefore likely. Figure 5.72 shows the reduction in C
L
when
depowering the sails in the wind tunnel with the Real-Time VPP for different
eff
plotted
against the reduction in z
CoE
. In the flat and twist model C /C
L Lopt
equals flat and 1-z
CoE
/z
CoEopt

is defined as twist. It can be seen that C is reduced together with z
L CoE
in a similar and close to
linear fashion for all
eff
. However, from the wind tunnel tests it is not known how much of
the reduction in C
L
is caused by twisting off the sails and how much by other trim
adjustments, including most importantly the reduction in by increasing the boom angle.
Different types of wind tunnel tests that systematically vary the sail shape are required to
assess this. It is however possible to use lifting-line theory to make a prediction.
The conditions that the sectional lift coefficient at the bottom of the sails cannot increase
when applying twist and that C is not affected by twist can be satisfied if C
L L
is lowered
before twist is applied by reducing or the camber along the whole span so that the loading
distribution remains approximately constant. Applying twist then increases the sectional lift
coefficient at the bottom of the sails back to the maximum while keeping C
L
constant. The
relationship between the reduction in z
CoE
and the increase in sectional lift coefficient at the
bottom of the sails is investigated using lifting-line theory assuming an initially semi-elliptic
load distribution with the water surface as the symmetry plane (as shown in appendix D.3). In
this idealisation the maximum increase in sectional lift coefficient occurs at the symmetry
plane so that the water surface is used as a reference point instead of the bottom of the sails.
152
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.6 Improved Flat and Twist Depowering Model
The increase in sectional lift coefficient at the symmetry plane (C(0)/C
l lminDi
(0)) is found to be
related to the reduction in the centre of effort height (z
CoE
/z ) by
CoEminDi
( )
0
1 1 1
) 0 (
) 0 (
t t c
z
z
c
C
C
tL
CoEminDi
CoE
tL
lminDi
l
+ + =

+ = ,
(5.33)
where c
tL
is the twist weight constant for lift and has the value of 5/31.667. From this
relationship it can also be deduced that C (0)/C (0)=1+c
lopt lminDi tL
t
0
so that the change in
sectional lift coefficient relative to the fully powered-up sectional lift coefficient (C
lopt
) is
given by
0
1
1
) 0 (
) 0 (
t c
t c
C
C
tL
tL
lopt
l
+
+ = .
(5.34)
For the sectional lift coefficient not to increase when twisting off the sails, C
L
needs to be
reduced by easing the sails or reducing the camber. The trim parameter flat currently models
a linear reduction in lift due to the reduction in and camber so that flat due to twist (f
t
) can
be defined as the inverse of equation (5.34) to give
0
1
1
1
) 0 (
) 0 (
t c
t c
C
C
f
tL
tL
l
lopt
t
+
+
= =

.
(5.35)
Figure 5.72 shows the curves of flat applied due to twist (f ) for the minimum and maximum t
t 0

measured in the wind tunnel. The minimum t of 0.02 was measured at
eff 0
of 20 and the
maximum of 0.12 at 60. Apart from a few exceptions at very small twist values, f reduces C
t L

less than the C
L
reduction measured in the wind tunnel. This should be the case and shows
that only part of the reduction in C is related to twisting off the sails as is indicated in
L
Figure
5.72. It should also be pointed out that for twist values above 0.25 the loading at the tip of the
sail inverts so that a righting moment is produced instead of a heeling moment and z
CoE
is
reduced more than is possible for soft sails. Twist values significantly larger than 0.25 are
however rare and this deficiency in the model is ignored. The expression for f
t
is implemented
in the HansenRig force module so that c
tL
becomes a new input parameter and equation
(4.1) is rewritten as
.
t eff Lopt L
f f r C C
2
) ( =
(5.36)
2 2
C , C and C are all functions of C and f is therefore included in equations (4.4), (5.31)
Ds Di Dp L t

and (5.32) so that they can be rewritten as
,
2 2 2 2
) (
t eff Lopt S Ds
f f r C c C =
(5.37)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 153
Chapter 5 Aerodynamic Force Modelling

+ + = ) 2 (
1
) (
0
2
2 2 2 2
tt t c
e AR
f f r C
C
t
t eff Lopt
Di

,
(5.38)
( )
Di Ds Dwindage t Dwindage eff DpOpt D
C C C f f r C C C + + + =
2 2 2
) ( .
(5.39)
VPP calculations are carried out as usual and the flat and twist values chosen by the VPP are
shown in Figure 5.73. Compared with, for example Figure 5.58, it can be seen that twist now
decreases smoothly to zero as
eff
increases and therefore does not need to be constrained
anymore. Flat is reduced less because of the additional reduction of C through f
L t
but is still
applied in a consistent manner and shows a similar trend as before. The thick lines in Figure
5.74 show the predictions of z
CoE
using the f expression and compared to the thin lines in
t
Figure 5.64 the results are almost identical at small
eff
but for larger
eff
above approximately
45, the predictions are improved. At small
eff
the choice of twist is governed by C so that f
Di t

has little effect, whereas the reduction in C becomes more important at larger
L eff
. Even with
the f expression z
t CoE
is reduced more than in the benchmark results, especially at larger
eff
.
C
L
in Figure 5.75 compared to Figure 5.65 also shows that f
t
mainly affects the predictions at
larger
eff
. For most cases the results are almost identical and a difference can only be noted
at
eff
where small amounts of depowering are necessary. For these cases the predictions
using f agree better with the benchmark results. It follows that C
t D
is also very similar at small

eff
and that the agreement with the benchmark predictions is improved by the f
t
expression at
larger
eff
as shown by comparing Figure 5.76 and Figure 5.66. Finally comparing the
predicted V in Figure 5.77 with Figure 5.67
S
shows that a speed difference can only be
detected in conditions where small amounts of depowering are applied.
The VPP results confirm that the f expression does not affect the results at small
t eff
where
twist works well anyway, but that it limits twist to more feasible values at larger
eff
. The
improved flat and twist model is still very generic since the three non-geometric input
parameters are aerodynamically meaningful quantities for which values can be obtained
analytically using, for example, lifting-line theory.
0
0.2
0.4
0.6
0.8
1
1.2
0 15 30 45 60 75 9

eff
[]
f
l
a
t

[
-
]
0
VT=4m/s
VT=6m/s
VT=8m/s
VT=10m/s
0
0.1
0.2
0.3
0.4
0.5
0.6
0 15 30 45 60 75 9

eff
[]
t
w
i
s
t

[
-
]
0
VT=4m/s
VT=6m/s
VT=8m/s
VT=10m/s
Figure 5.73: Flat and twist values chosen by VPP in different V
T
for a range of
eff
with C
L
a function of
twist
154
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.6 Improved Flat and Twist Depowering Model
3
3.5
4
4.5
5
5.5
6
6.5
7
7.5
8
0 15 30 45 60 75 9

eff
[]
z
C
o
E

[
m
]
0
VT=10m/s power VT=6m/s power
VT=10m/s CL(t,t0,ctL) VT=6m/s CL(t,t0,ctL)
VT=10m/s CL(1-2*t) VT=6m/s CL(1-2*t)
VT=8m/s power VT=4m/s power
VT=8m/s CL(t,t0,ctL) VT=4m/s CL(t,t0,ctL)
VT=8m/s CL(1-2*t) VT=4m/s CL(1-2*t)
0
0.4
0.8
1.2
1.6
2
2.4
2.8
0 15 30 45 60 75

eff
[]
C
L

[
-
]
90
VT=10m/s power VT=6m/s power
VT=10m/s CL(t,t0,ctL) VT=6m/s CL(t,t0,ctL)
VT=10m/s CL(1-2*t) VT=6m/s CL(1-2*t)
VT=8m/s power VT=4m/s power
VT=8m/s CL(t,t0,ctL) VT=4m/s CL(t,t0,ctL)
VT=8m/s CL(1-2*t) VT=4m/s CL(1-2*t)
Figure 5.74: z
CoE
for four V
T
from VPP
calculations using power or flat and
twist with C
Figure 5.75: C for four V
L T
from VPP calculations
using power or flat and twist with C
L
a
function of twist a function of twist
L

0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 15 30 45 60 75 9

eff
[]
C
D

[
-
]
0
VT=10m/s power VT=6m/s power
VT=10m/s CL(t,t0,ctL) VT=6m/s CL(t,t0,ctL)
VT=10m/s CL(1-2*t) VT=6m/s CL(1-2*t)
VT=8m/s power VT=4m/s power
VT=8m/s CL(t,t0,ctL) VT=4m/s CL(t,t0,ctL)
VT=8m/s CL(1-2*t) VT=4m/s CL(1-2*t)
0.94
0.96
0.98
1
1.02
1.04
1.06
1.08
1.1
0 15 30 45 60 75 90 105 120 135

T
[]
V
S
/
V
S
(
p
o
w
e
r
)

[
-
]
VT=4m/s CL(t,t0,ctL)
VT=6m/s CL(t,t0,ctL)
VT=8m/s CL(t,t0,ctL)
VT=10m/s CL(t,t0,ctL)
VT=4m/s CL(1-2*t)
VT=6m/s CL(1-2*t)
VT=8m/s CL(1-2*t)
VT=10m/s CL(1-2*t)
Figure 5.77: V Figure 5.76: C for four V
D T
from VPP calculations
using power or flat and twist with C
S
for four V
T
from VPP calculations
using flat and twist with C
L
a
function of twist
L
a function
of twist plotted as ratio of V
S
obtained
with power model

In this study the analytical values for semi-elliptic loading of c =8, c
t tL
=5/3 and
z =0.42z
CoEminDi mast
are used and produce predictions that agree significantly better with the
benchmark results compared to the original reef and flat model assessed in section 5.5. The fit
with the benchmark results can be further improved by modifying c , c
t tL
or z based on
CoEminDi
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 155
Chapter 5 Aerodynamic Force Modelling
the wind tunnel data but the trend that z
CoE
is reduced relatively too much at large
eff

compared to small
eff
remains.
As an alternative approach the reduction in C associated with the reduction in z
L CoE
is
modelled as linear since Figure 5.72 showed that the total reduction in C
L
is close to linear
with z
CoE
. Flat associated with twist (f ) is then given by
t
t c f
tL t
=1 .
(5.40)
Figure 5.72 shows the lines of linear reduction in C with twist for c
L tL
values of one and two.
The value of one is selected so that all wind tunnel measurements lie below the line and it
gives results not too dissimilar to the original f expression in equation (5.35)
t
for twist values
up to 0.4. The VPP predictions are very similar to the results for the original f
t
expression,
which confirms that small changes in f do not critically influence the solutions because f
t t

only affects the solution in certain conditions. VPP predictions are also carried out for c
tL
=2,
which is close to the least squares fit of the wind tunnel data (see Figure 5.72). It is now
assumed that on average all reduction of C
L
is linked to twist. As a consequence only small
amounts of flat are used in certain conditions in an incoherent way without showing clear
trends. VPP predictions only optimising twist produce very similar results but with fairer
curves since only one trim parameter is optimised so that the calculations are more
constrained. The thin lines in Figure 5.74 show that z
CoE
fits the benchmark results better if
the linear model with c
tL
=2 and only twist is used but still under-predicts at larger
eff
.
Figure 5.75 shows that C is not affected much and consequently C
L D
fits the benchmark
results better as can be seen from Figure 5.76. It is not at all surprising that the best data fit is
achieved with this model since it uses the wind tunnel measurements to obtain the value for
c
tL
and with only one trim parameter is constrained more so that the sails can only be
depowered in one way. However, the improved modelling of z
CoE
and C
D
only marginally
influences the boat speed as shown in Figure 5.77. The generally big differences in z
CoE
at
larger
eff
between the wind tunnel measurements, the power model predictions and the
various model versions using reef, flat and twist, are related to these small changes in V
S
.
When trimming the sails in the wind tunnel it is difficult even with the Real-Time VPP to
find the trim that truly maximises V if V
S S
changes little over a range of sail trims. Similarly
small deficiencies in the data modelling in the trim parameter models can lead to big changes
in z
CoE
if V changes little over a large range of C
S D
and z
CoE
. Although the linear model with
c
tL
=2 and only twist matches the benchmark results better, the original f
t
expression in
equation (5.35) is recommended since the parameters can be obtained from aerodynamic
principles and both twisting and reducing the lift are modelled as separate depowering
actions.
156
YACHT
RESEARCH
UNIT
The
University
of Auckland
5.7 Conclusions
5.7 Conclusions
The Real-Time VPP has successfully been used to develop depowered sail trims in the wind
tunnel for an upwind sailing configuration at effective wind angles between 25 and 90. The
trimming process with the Real-Time VPP is efficient and yields the best trim for a given
sailing condition without the need to conduct a systematic series of trim variations. The wind
tunnel data has then been used to investigate the depowering of upwind sails.
Before investigating the depowering it was realised that the current model of calculating the
moments from the centre of effort position neglects the moment component in line with the
force. For advanced VPP predictions with six degrees of freedom the moment component in
line with the force makes a noticeable difference that results in an increase in predicted boat
speed of up to 1.5% in reaching conditions, and is therefore modelled as a scalar moment
coefficient that acts in line with the force and is a function of the effective wind angle.
The power force model has been developed to model the depowering of the sails in more
detail compared to the standard trim parameter model. The depowering is simulated
empirically based on the Real-Time VPP measurements with the trim parameter power,
which is defined as the reduction in heeling moment. The common approach of modelling
trim changes due to depowering implicitly with trim parameters and applying them to fully
powered-up sail input data is adhered to. The effect of depowering is modelled by response
surfaces which are functions of power and the effective wind angle but can be increased in
dimensions arbitrarily. It is shown that the most important parameters for which changes due
to depowering need to be modelled are lift, drag, and the centre of effort height. For the
upwind and reaching conditions with mainsail and jib investigated here, the changes in
longitudinal centre of effort and moment in line with the force were found to have a very
small influence on the performance of the yacht. However, for different yacht types and sails
this may be different. For a large reaching spinnaker on a high-performance yacht with a
small rudder, the shift in longitudinal centre of effort can be expected to be larger and have a
greater influence on the performance. The smoothest and least complex response surfaces
describing depowering were obtained when modelling the force and moment coefficients
directly instead of using the more common method of modelling the lift and drag coefficient
and the centre of effort position. None of the response surfaces modelling the depowering of
the forces and moments were found to have a significant dependency on the effective wind
angle. If this empirical approach of depowering based on wind tunnel tests is used, modelling
the forces and moments directly is a good alternative since it results in a much simpler force
model and not only the depowering response surfaces but also the fully powered-up force and
moment coefficient curves are well defined and smooth. With the power model it is possible
to accurately model the depowering in greater detail based on wind tunnel tests, which is
particularly important for reaching and running sails of high-performance yachts. More wind
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 157
Chapter 5 Aerodynamic Force Modelling
tunnel tests with the Real-Time VPP are required to determine to what extent the depowering
characteristics change for different sail types and how far the depowering response surfaces
are transferable so that not all sails need to be tested at depowered trims. With the power
model implemented in FRIENDSHIP-Equilibrium and the Real-Time VPP, tools have been
developed to achieve this.
Since wind tunnel testing is not always feasible, generic trim parameter models that are based
on aerodynamic principles rather than experimental results are still desirable. The
performance of the traditional reef and flat parameters and the more recently introduced twist
parameter is therefore assessed against the power model results. With reef and flat the boat
speed prediction is reasonable, with an under-prediction of up to 4%, but the changes in lift,
drag and centre of effort height are not modelled well. Although the twist parameter is based
on sound aerodynamic principles, the original implementation agrees even less well with the
power model results. One reason for this is that previous research exclusively focused on
small apparent wind angles close to VMG sailing conditions. However, this study was
conducted over a wider range of apparent wind angles up to 90. Although a generic
depowering model cannot necessarily be expected to perform well at all points of sail, it
should be useable over a range of wind angles where upwind sails may be depowered. With
the help of the wind tunnel data three improvements to the twist parameter implementation
are developed. The enhanced model now accounts for the facts that the fully powered-up sails
may not have a semi-elliptic loading distribution, that the so-called parasitic drag is
dependent on the reduction in lift squared and that twisting off the sails is associated with a
reduction in lift. The additional non-geometric input parameters can all be determined
analytically based on aerodynamic principles. With these enhancements the fit with the
power model results is much better and the predicted boat speed is within 1%. The study
did not only result in an improved generic trim parameter model but it also highlighted why
the traditional reef and flat parameters have not been superseded for such a long time,
although many people agreed that they do not model the physics well. Reef and flat actually
predict the performance of the yacht reasonably well because, although for the wrong
conceptual reasons, reef implicitly models a relative increase in induced drag and a reduction
in parasitic drag as the sails are depowered. These phenomena are now modelled more
elegantly by flat and twist and the improved generic depowering model has been
implemented in FRIENDSHIP-Equilibrium. Realistically depowered trims can be developed
efficiently with the Real-Time VPP, which made it feasible to investigate depowering in
sufficient detail to make these improvements.
158
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
6 Wind Tunnel and Full-Scale Comparison
6.1 Introduction
The final part of this work sets out to compare wind tunnel measurements to the full-scale
data collected by the Berlin Sailing-Dynamometer Dyna by employing some of the methods
and results described in the previous chapters. A similar comparison has not been conducted
previously, mainly due to the lack of reliable full-scale data and the difficulties associated
with analysing the data as shown in this chapter. The full-scale data collected with Dyna is
the most accurate and reliable data collected to date, and the results from this comparison can
give valuable insight into the quantitative accuracy of wind tunnel testing. Researchers and
designers are generally satisfied that wind tunnel testing of yacht sails and components give
reliable qualitative results but are unsure of the quantitative accuracy. In practical terms this
means that a sail which performs better in the wind tunnel, will perform better in real life as
well, but it is not known how well the difference in sail performance measured in the wind
tunnel translates to full-scale.
6.2 Full-Scale Aerodynamic Data Acquisition
The most recent project to determine full-scale forces acting on a sailing yacht was initiated
at the Technical University Berlin in Germany in 1996, in order to investigate the relation
between model tests, numerical simulations and full-scale data (Brandt and Hochkirch, 1999).
A brief introduction into the Dyna project and the measurement techniques used to collect the
data is given here (although the author did not carry out the work), to highlight some of the
complexities involved in conducting full-scale measurements. Due to the close collaboration
with the Technical University Berlin the author was able to participate in some of the testing
with Dyna and gain first hand experience of the challenges involved in conducting full-scale
measurements on the water.
A 10-metre full-scale sail force dynamometer named Dyna (Figure 6.1) was specifically
designed and built for this project. The hull and rig are based on a Dehler 33, a modern 33ft
IMS cruiser/racer designed by judel/vrolijk & co., to provide suitable reference results for
contemporary yacht design issues. In addition to measuring the aerodynamic rig forces,
multi-component force balances also measured the hydrodynamic loads on keel and rudder.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 159
Chapter 6 Wind Tunnel and Full-Scale Comparison
Laser Doppler
Velocimetry
(LDV) system Keel force balance
Internal rigid aluminium frame
which forms the rig force balance

Figure 6.1: Schematic diagram of the Berlin Sail-Force-Dynamometer Dyna with its internal rigid
aluminium frame which allows the aerodynamic and hydrodynamic forces to be measured
separately
18
This full-scale aerodynamic data is of interest to this project and is used here for the
comparison to wind tunnel tests.
Measuring the full-scale aerodynamic forces acting on a sailing yacht is complex and requires
a purpose-built sailing dynamometer. Theoretically it is possible on any sailing yacht to
measure all forces acting on the standing and running rigging, and on the mast and their
direction, to calculate the resulting aerodynamic force acting on the rig. This process however
does not yield meaningful results since the rig loads in the individual components are very
large compared to the resulting aerodynamic forces. With this method the resulting
aerodynamic force is calculated from all the individual forces in the components and the
accuracy is greatly compromised since the resulting aerodynamic force is only a fraction of
most individual loads applied. In addition it is difficult to measure the directions of the forces
accurately since many of them cannot be treated as being constant. Obviously the line of
action of many running rigging components, like for example sheets, changes constantly
when trimming the sails and would need to be measured continuously. However, the line of
action of standing rigging components can also vary due to deformation of the rig and
possibly even of the hull.

18
Diagram courtesy of the Technical University Berlin
160
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.2 Full-Scale Aerodynamic Data Acquisition
Frame passing through hole in deck
to hold jib track above. The gap is
covered with flexible cloth so that no
forces are transmitted but to stop
water penetrating the hull.

19
Figure 6.2: Aluminium frame of Dyna looking forward towards mast
A purpose-built sailing dynamometer can provide more meaningful results by measuring the
resulting aerodynamic forces directly. For this to be possible all the rigging components need
to form an entity that is separated from the remainder of the yacht and only connected to it
via force transducers. In the case of Dyna and the other two previous sailing yacht
dynamometers (see section 1.3) this is achieved by having a framework structure to which all
rigging components are attached. Most of the framework structure is located inside the hull
and deck shell and the mast, forestay and shrouds pass through holes in the deck to connect to
the frame. At the stern of Dyna the frame extends through holes in the deck so that the
mainsheet traveller, the backstay and running backstays can be attached to it. Similarly it
protrudes through the deck on both sides of the coach roof to hold the headsail tracks as
shown in Figure 6.2. The gaps between the deck and frame/rigging are covered with flexible
cloth so that no forces are transmitted between frame/rigging and the deck but to stop water
penetrating the hull.
The frame is connected to the hull shell by six force transducers to form a six-component
force balance to measure the aerodynamic rig loads. The aluminium frame of Dyna and the
position and orientation of the six force transducers are shown in Figure 6.3. The voltages
measured by the force transducers are related to forces and moments in coordinate system B
by a calibration matrix. The calibration matrix also eliminates the influence of the
gravitational forces and moments, which are a function of the weight and centre of gravity of
the rig force balance, the pitch (trim) angle () and heel angle (). From the calibration
measurements Hochkirch (2000) calculated the mean absolute error to be 46N, which is a

19
Picture courtesy of the Technical University Berlin
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 161
Chapter 6 Wind Tunnel and Full-Scale Comparison
Figure 6.3: Aluminium frame of Dyna inside the hull shell with arrows indicating the position and
orientation of the six force transducers that connect the internal frame to the hull shell to
make up the six-component force balance to measure the aerodynamic rig loads
20
relative error of 2.3% of the calibrated range in the x and y-direction and 1.5% in the z-
direction.
In order to analyse the measured aerodynamic rig loads it is also important to obtain the wind
condition and the yacht performance accurately. The most important parameters of the yacht
performance that are needed for the analysis are the boat speed, leeway angle and heel angle.
In addition to a standard paddle wheel speedometer, which measures the yacht speed through
the water, a two component Laser Doppler Velocimetry (LDV) system is fitted to Dyna to
measure the boat speed and leeway angle. As indicated in Figure 6.1 the LDV system is
installed at approximately of the DWL aft of the forward perpendicular (F
PP
), above a
window in the bottom of the hull. It measures the Doppler-frequency of particles in the water
along the x-axis with a red laser and along the y-axis with an infrared laser. The target
volume for the measurements is approximately 38mm below the hull. The boundary layer
thickness can be calculated to be approximately 38mm at of DWL at a boat speed of 4m/s
so that the influence from the boundary layer is minimal. However, some influence of the
global flow field needs to be adjusted for, and potential flow calculations have been
conducted by Hochkirch (2000) with the CFD-code SHIPFLOW
21
to develop a mathematical
model to account for changes in the global flow field at the position of the LDV target
volume.
The data rate of measurements of the LDV system depends on the number of particles in the
water suitable for reflecting the laser signal. The data rate was sufficient only in certain water
conditions so that the LDV system was primarily used to calibrate the paddle wheel
speedometer. The accuracy of the off the shelf paddle wheel speedometer was initially
insufficient. It consists of six paddles mounted with the rotation axis perpendicular to the
centreline so that about 40% of the paddle wheel is exposed to the flow. A magnet is located
on every other paddle so that a sensor registers three impulses per revolution. The standard

20
Diagram courtesy of the Technical University Berlin
21
Software developed by Flowtech International AB, Chalmers Tvrgata10, P.O. Box 24001, SE-400 22
Gothenburg, Sweden
162
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.2 Full-Scale Aerodynamic Data Acquisition
electronics of the speedometer count the impulses in a certain time window and calculate the
boat speed from it. Due to the low turning rate of the paddle wheel, integration over several
seconds is necessary to produce a meaningful signal. This is sufficient for navigation
purposes but not for analysing full-scale data where a faster response time is required. The
resolution for shorter averaging times can be greatly increased by measuring the phase
directly. Hochkirch (2000) implemented a programmable counter/timer device to achieve this
which takes into consideration the possible uneven spacing of the magnets due to
manufacturing tolerances. The minimum velocity that can be measured by the modified
paddle wheel speedometer electronics is about 0.12m/s. The resolution of the paddle wheel
signal for a boat speed of 5m/s is better than 0.1%. In the calibration the boat speed is a
function of the paddle frequency and heel angle. The flow field in the region of the paddle
wheel changes as the yacht heels so that flow speed relative to the free-stream flow is not
constant.
Since the LDV system does not obtain results in many water conditions, an alternative system
was needed to obtain the leeway angle. A fin attached to a force transducer that measures the
force along the y-axis was installed on the centreline below the hull in front of the LDV
system window. This arrangement was calibrated in a cavitation tunnel and once installed on
Dyna the leeway angles were checked against the LDV system results to ensure the correct
alignment of the fin. It is also important to know the orientation of the yacht (heel and trim
angle) and a fibre-optic-gyro (FOG) reference system is used on Dyna. In addition to the
angular positions, the velocities and linear accelerations are also recorded by the FOG
reference system.
One of the main challenges associated with the analysis of full-scale aerodynamic data is
determining the wind condition (direction and speed) accurately during each measurement on
the water. On Dyna the wind direction and speed are determined from a standard yacht cup
anemometer with a direction vane and a calorimetric sensor anemometer as shown in Figure
6.4. The calorimetric sensor anemometer works on the principal that heat dissipation into the
flow medium increases with increasing flow speed so that the current required to keep the
heated thermistor at a constant temperature is a measure of the flow speed. Because of its
measuring principle it is however influenced by the heat from the sun and by rain particles so
that it is mainly used to calibrate the cup anemometer in suitable weather conditions.
The off the shelf cup anemometer works similarly to the paddle wheel speedometer. The
cup anemometer consists of three cups mounted on a ball bore shaft. On the shaft three
magnets are positioned so that a sensor registers three impulses per revolution. When
counting impulses, the resolution is approximately one knot for typical wind speeds when
sailing, which is only sufficient for navigation purposes but not for analysing full-scale
aerodynamic data. As for the paddle wheel speedometer, Hochkirch (2000) implemented a
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 163
Chapter 6 Wind Tunnel and Full-Scale Comparison
Calorimetric sensor
anemometer
Cup anemometer with direction vane

22
Figure 6.4: Cup anemometer with direction vane and calorimetric sensor anemometer on Dyna
programmable counter/timer device to measure the time taken for one revolution instead of
counting impulses, which increased the accuracy of the system significantly. With the device
wind speeds from about 0.95m/s can be measured and the error of processing the signal for a
wind speed of 10m/s is less than 0.1%.
6.3 Wind Tunnel Study of Flow Above Mast
The wind direction and speed for the full-scale tests on Dyna are measured with the cup
anemometer and direction vane situated 850mm above the top of the mast as discussed in
section 6.2. In his data analysis Hochkirch (2000) assumed that the sails do not significantly
influence the flow at the measurement position and therefore treated the measured speed and
direction as the free-stream flow. This assumption is based on experiments carried out with
the similar sized MIT sailing dynamometer by Peters (1992), where the wind direction and
speed were measured above the masthead and by a support vessel. In this study no influence
of the sails on the measurements above the mast could be detected. The anemometer was
however positioned 3 metres above the masthead compared to the 850mm on Dyna. It is also
questionable if the experimental procedure was accurate enough to notice small changes in
the flow. The support vessel measured the wind speed and direction some distance away from
the sailing dynamometer and at a different height over a time period of several minutes. Since

22
Pictures courtesy of the Technical University Berlin
164
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.3 Wind Tunnel Study of Flow Above Mast
the wind direction and speed constantly fluctuate with respect to time and position in three-
dimensional space, the accuracy of the experiment is limited. Although this method showed
that the flow above the mast is not hugely altered by the presence of the sails, it is unlikely
that wind speed changes of a few percent and wind angle changes of a few degrees would
have been picked up.
Wind tunnel studies have therefore been conducted as part of this work to investigate the
influence of the sails on the flow above the masthead in more detail, so that the full-scale
measurements can be corrected if necessary. The influence of the sails and rig on the free-
stream flow at the position of the cup anemometer was determined in the wind tunnel by
measuring the change in flow direction and speed at the corresponding model scale height of
127mm above the top of the mast. The probe was suspended from the wind tunnel ceiling or
attached to a freestanding support rig as shown in Figure 6.5. A thin rod, which simulates the
full-scale anemometer pole, was attached to the top of the mast as shown in Figure 6.6, so
that the probe could be positioned above the top of the rod easily and accurately. The distance
between the tip of the rod and the probe was varied to check for any influence of the rod on
the measurements. With a 3mm gap a small but consistent and repeatable upwards deflection
of the flow of about 0.3 due to the presence of the rod was observed. With an increased gap
of 5mm, no influence of the rod on the measurements could be measured. A rod length of
120mm from the top of the mast was therefore chosen so that with a 5mm gap, the centre of
the probe head is 127mm above the mast since the probe head radius is approximately 2mm.
Figure 6.6: Cobra Probe positioned 127mm above
the mast head with model heeled
Figure 6.5: Flow measurements above the mast
with a Cobra Probe attached to free
standing support rig
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 165
Chapter 6 Wind Tunnel and Full-Scale Comparison
The probe was adjusted by hand to the correct position with the tunnel running and the sails
trimmed to take into account mast defection. The probe position needs adjusting for each test
run since changes in the apparent wind angle, heel angle and sail trim influence the masthead
position. Instead of developing a complicated system to keep the probe aligned with the flow
when its position is changed, a reference free-stream flow measurement is conducted for each
probe position. The sequence of tasks for each test run is therefore as follows: firstly the
apparent wind angle and heel angle are selected and the probe is zeroed. The tunnel is then
turned on so that the sails can be trimmed. Once the sails are trimmed the probe is adjusted to
the correct position and approximately aligned with the flow, and a measurement is acquired.
The model is then removed from the tunnel and a reference free-stream flow measurement is
taken without changing the location and orientation of the probe. The difference of the two
measurements is the flow change due to the presence of the sails and rig. Although this
method requires a probe position adjustment and a reference free-stream flow measurement
for each test run, it has many advantages over attaching the probe to the top of the mast. If the
probe is attached to the model, alignment of the probe becomes very important, the effect of
mast twist needs to be considered and the flow angles become too large for the probes used in
these studies, unless the probe orientation is changed as the apparent wind angle increases.
The first study was carried out using a five-hole, hemispherical head, pitot tube probe, which
measures flow speed and angles of up to 30 from its centre line. The experimental set-up
and data acquisition software was developed based on previous work by Locke (1994). A
Setra 239 pressure transducer sequentially measured the pressures at the five pitot tube
openings and a second pressure transducer measured the reference pressure to account for
fluctuations in the free-stream flow. The sampling period at each opening was 30 seconds.
Due to the small dynamic pressures, which are feasible in the wind tunnel due to the soft sails
and the range of the pressure transducers available, the accuracy of the resultant dynamic
pressure was limited to approximately 5% of the measured value, while flow angles can be
measured with an accuracy of 0.5. The results of the first study are presented by Hansen et
al. (2003a) but are not shown here since the trends are very similar to the more accurate
results obtained from the second study. The better results in the second study were achieved
by using a more accurate probe and more care was taken when trimming the sails.
The second study used a Series 100 Cobra Probe, which is distributed by Turbulent Flow
Instrumentation
23
as a ready to use system including the control software. It became
available to the author only after the first study had been competed and it was decided to
repeat the investigation to verify the results and increase the accuracy. The Cobra Probe,
shown in Figure 6.6, is a four-hole pitot tube with the pressure transducers located in the
body of the probe so that the tube length between the holes and the pressure transducers is

23
Turbulent Flow Instrumentation Pty. Ltd., Tallangatta, Victoria, Australia, www.turbulentflow.com.au
166
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.3 Wind Tunnel Study of Flow Above Mast
kept as short as possible. This reduces the response time and frequencies of up to 2kHz can
be measured. This study focused on time-averaged results so that the frequency response is
not important. The flow speed range of the Cobra Probe is up to a pressure of 1kPa and the
measurement accuracy at the low wind tunnel speeds in this study is estimated at 0.5% of
the measured value. Angle measurement accuracy is estimated as typically better than 0.2
within the flow angle measurement range of 45 in yaw and pitch.
Measurements are conducted with the Cobra Probe positioned 127mm above the masthead
for mainsail with jib or genoa for six apparent wind angles between 20 and 90, and three
heel angles of 0, 12.5 and 25 with the tunnel in upwind configuration, and for mainsail and
spinnaker for apparent wind angles between 60 and 180 at 0 heel with the tunnel in
downwind configuration. In addition measurements with the Cobra Probe in different
positions are conducted for selected sailing conditions. The sampling period for all
measurements was 90 seconds. Based on the anemometer and direction vane characteristics
that will be discussed in section 6.4, the flow parameters of most interest to this investigation
are defined. The change in total measured wind speed (V) and the change in effective wind
angle (
eff
), are the primary parameters that need to be considered in a correction model but
the inclination angle of the flow (), relative to the deck plane (z-plane in coordinate system
B) and its change () due to the sails are also of interest.
6.3.1 Analysis of Cobra Probe Measurements
The change in total wind speed (V) due to the presence of the sails is obtained from the
wind speed measurement (V) with the model in the tunnel by subtracting the measurement
without the model in the tunnel, which is regarded here as the free-stream flow speed (V

). A
better repeatability is achieved if the wind speed measurements with and without the model
are not adjusted for by changes in the reference dynamic pressure. This indicates that the
wind tunnel speed for back-to-back runs is very similar and that the reference pressure pitot
tube upstream of the test section picks up fluctuations that are not measured to the same
degree by the Cobra Probe above the mast. Figure 6.7 uses 18 tests with the mainsail and
genoa where a second measurement was recorded after the model was put back into the wind
tunnel after the free-stream flow measurement was made. Comparing the second
measurement to the first measurement (before the free-stream flow measurement) gives a
good indication of the repeatability of the experiment. Figure 6.7 shows the percentage
change in flow speed between the first and second run, with and without considering the
fluctuations in reference dynamic pressure (q
ref
) measured by the pitot tube. It can clearly be
observed that the difference between the first and second run is less if the measurements are
not adjusted for fluctuations in q
ref
. Without correcting for q
ref
fluctuations the standard
deviation is 0.33% with a maximum difference of -0.72%, whereas correcting for fluctuations
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 167
Chapter 6 Wind Tunnel and Full-Scale Comparison
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
0 3 6 9 12 15 18
Test number [-]
(
V
R
u
n
2
/
V
R
u
n
1
)
-
1

[
%
]

Adjustment of qref fluctuations
No adjustment for qref fluctuations
0.98
0.985
0.99
0.995
1
1.005
1.01
1.015
1.02
0 30 60 90 120 150 180

eff
[]
f
V
r
e
f

[
-
]

Main & genoa
Main & jib
Main & spi
1-4.2*10^-5*EWA
Figure 6.8: Reference flow speed factor (f Figure 6.7: Change in wind speed for back-to-
back runs with and without adjusting
for fluctuations in reference pressure
(q
Vref
) due
to the presence of the model in the
tunnel for different sail sets vs.
eff
)
ref
in q
ref
increases the standard deviation to 0.77% with a maximum difference of -2.05%. When
not correcting the measurements for fluctuations in reference pressure it is important to check
that the presence of the model in the tunnel does not change the reference pressure due to
blockage. Figure 6.8 shows the factor (f
Vref
) between the reference flow speed measured by
the pitot tube with and without the model in the tunnel. f
Vref
is shown for the model with
different sail configurations and is plotted against
eff
since the projected area of the sails
increases with
eff
which could affect blockage. The significant scatter in the data is due to the
pressure fluctuations measured by the pitot tube as discussed before, and makes determining
any trends difficult. Assuming that there is no blockage at
eff
of zero since the projected area
of the model is very small, linear regression shows a small reduction in reference speed with
increasing
eff
. From the data f can be modelled by
Vref
eff Vref
f
5
10 2 . 4 1

= ,
(6.1)
to give a basic approximation of the blockage. The flow speed (V) measured with the sails
present is now adjusted by dividing by the modelled f to account for the model blockage.
Vref
The yaw and pitch angle measured without the model in the wind tunnel are subtracted from
the angles measured with the model in the tunnel to obtain the change in yaw angle (
Cobra
)
and change in pitch angle ( ) due to the presence of the sails. and
Cobra Cobra Cobra
are
relative to the free-stream flow in the tunnel, which is horizontal and, for the upwind testing
configuration, is aligned with an apparent wind angle of zero. The downwind configuration
uses the flow twisting vanes so that the free-stream flow is not necessarily aligned with an
apparent wind angle of zero at the height of the Cobra Probe. For test cases where the model
is not heeled, and correspond to
Cobra Cobra eff
and . For test cases where the model is
heeled, and are transformed into the body fixed coordinate system as
Cobra Cobra
168
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.3 Wind Tunnel Study of Flow Above Mast
described in appendix D.4 to obtain
eff
and . The change in inclination angle () relative
to the deck plane due to the presence of the sails is the difference between and the
inclination angle of the free-stream flow ( ) relative to the deck plane so that

= . (6.2)
Since the free-stream flow is horizontal and the wind tunnel model does not pitch,

can be
calculated from
, ) sin sin ( sin
1
m A
=

(6.3)
where
A
is negative again for the wind tunnel tests.
6.3.2 Influence of Sails on Flow Above Mast
Figure 6.9 shows the non-dimensional change in wind speed (V/V

) 127mm above the mast,


calculated as discussed in section 6.3.1, plotted against effective wind angle for the model
with mainsail and jib or spinnaker. The wind speed increases for effective wind angles up to
approximately 150 because the flow is accelerated by the flow around the sails and the tip
vortex. With mainsail and jib the increase is up to 5% in realistic sailing conditions. It
increases by up to about 7.5% at
eff
of 90 and heel of 25, which however is not a common
sailing condition. There is no clear trend with heel angle but V/V

appears to increase with

eff
to a local maximum at
eff
of about 30 where the maximum C
L
is produced and the tip
vortex is strongest. For larger
eff
the positive V/V reduces slightly as C reduces but for
L eff

of 90 it increases again as the increasing angle of attack on the mainsail causes a larger
obstruction to the flow, and the flow is accelerated to flow above the sail. Figure 6.10 shows
V/V plotted against
eff
for mainsail and genoa. The trend is very similar to that for

-10
-5
0
5
10
15
20
0 20 40 60 80 100 120 140 160 180

eff
[]

V
/
V


[
%
]

Main & jib heel 0
Main & jib heel 12.5
Main & jib heel 25
Main & spi heel 0
-10
-5
0
5
10
15
20
0 20 40 60 80 100 120 140 160 180

eff
[]

V
/
V


[
%
]

Main & genoa heel 0
Main & genoa heel 0 2nd run
Main & genoa heel 12.5
Main & genoa heel 12.5 2nd run
Main & genoa heel 25
Main & genoa heel 25 2nd run

Figure 6.10: Change in wind speed (V/V Figure 6.9: Change in wind speed (V/V

)
127mm above mast due to presence of
mainsail with jib or spinnaker

)
127mm above mast due to presence of
mainsail with genoa
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 169
Chapter 6 Wind Tunnel and Full-Scale Comparison
mainsail with jib but the values are slightly higher which is to be expected since the genoa is
a larger sail than the jib. Again there is no clear trend with heel angle. For each
eff
and heel
angle Figure 6.10 shows two measurements. One is the initial measurement before the model
is removed from the tunnel to obtain the free-stream flow and the second is the measurement
after the model is placed in the tunnel again. This shows that a good repeatability can be
achieved that is of the order of the estimated measurement accuracy of 0.5%. V/V

with
mainsail and spinnaker shown in Figure 6.9 is up to 17% at the smallest
eff
of 40 and
reduces with increasing
eff
. The low-pressure side of the spinnaker head points upward
accelerating the flow above the spinnaker. For small
eff
this accelerated flow passes over the
top of the mast. As
eff
increases the mast top is not downstream of the spinnaker head
anymore so that V is less positive and eventually becomes negative for
eff
above 150. For
effective wind angles of up to 150 the mainsail has some degree of attached flow so that the
flow above the mast is accelerated whereas the flow is almost completely separated for angles
above 150 and the mainsail acts as a bluff body that reduces flow speed around it.
Figure 6.11 and Figure 6.12 show the change in effective wind angle (
eff
) as defined by
equation (D.43) for mainsail with jib, genoa or spinnaker plotted against
eff
. For all points of
sail the effective angle was found to increase through the presence of the sails. The shape of
the
eff
curve is very similar to the lift coefficient curve, which indicates that C
L
is an
important factor causing
eff
. Lift causes a tip vortex and the air flowing over the sail head
from the windward to the leeward side increases
eff
at the position of the anemometer. In
addition lift causes upwash upstream of the sails as illustrated in Figure 6.13. A portion of the
upwash can be expected to be present above the mast and contribute to the positive
eff
. The
change in effective angle increases with heel angle and the increase gets more significant as

eff
gets larger.
eff
for mainsail with geona is larger than
eff
for mainsail with jib but the
trend for both configurations is very similar. As before two measurements with mainsail and
genoa are shown for each
eff
and heel angle in Figure 6.12. It can be seen that most
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
0 20 40 60 80 100 120 140 160 180

eff
[]

e
f
f

[

]

Main & jib heel 0
Main & jib heel 12.5
Main & jib heel 25
Main & spi heel 0
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
0 20 40 60 80 100 120 140 160 180

eff
[]

e
f
f

[

]

Main & genoa heel 0
Main & genoa heel 0 2nd run
Main & genoa heel 12.5
Main & genoa heel 12.5 2nd run
Main & genoa heel 25
Main & genoa heel 25 2nd run

Figure 6.12: Change in effective wind angle ( Figure 6.11: Change in effective wind angle (
eff
)
127mm above mast due to presence of
mainsail with jib or spinnaker
eff
)
127mm above mast due to presence of
mainsail with genoa
170
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.3 Wind Tunnel Study of Flow Above Mast

eff
V
eff
V
Anemometer
= Anemometer wind
velocity
V
eff
= Effective free-stream
wind velocity

eff
= Effective wind angle

eff
= Change in effective
wind angle
V
Anemometer

eff
z-plane

Figure 6.13: Illustration of change in effective wind
angle ( ) due to upwash in z-plane
eff
corresponding runs are within 0.1 with the maximum difference being 0.3. The trend of

eff
for mainsail with spinnaker is similar to the upwind configurations. Although the tests
are conducted at zero heel the magnitude of
eff
is similar to the mainsail with jib
measurements at 25 heel. Compared to the mainsail with jib measurements at 0 heel,
eff
is
significantly larger because the spinnaker has more area high up and forward of the mast
which increases the upwash at the position of the Cobra Probe.
The change in effective wind angle (
eff
) at a position close to the direction vane was also
obtained by Krebber (2005) from a RANSE simulation of Dyna sailing upwind under
mainsail and jib. At small
eff
and 30 heel,
eff
is between 3.6 and 6.3 depending on the
sail trim. Since
eff
is strongly related to the mainsail lift, it can be assumed that the
optimally trimmed sails produce
eff
close to the maximum of 6.3. This is noticeable larger
than the maximum of 3.9 measured in the wind tunnel for the fully powered-up sails.
Without a detailed analysis it is difficult to assess the possible reasons for the discrepancy.
However, the comparison shows that the general behaviour of the flow above the mast seems
comparable.
A direct comparison to other experimental work is difficult since many factors influence this
effect. As described by Marchaj (1988) wind tunnel tests of a 1/6 scale One Tonner model by
Kamman at an effective angle of 25 and heel angle of 20 with mainsail and genoa show an
increase in effective wind angle of about 5 in a similar position above the mast. This is about
0.7 more than determined in this study (see Figure 6.12), but the One Tonner has a masthead
rig, which increases the upwash compared to the fractional rig of Dyna.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 171
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
2
4
6
8
10
12
14
16
18
0 20 40 60 80 100 120 140 160 180

eff
[]

]

Main & jib heel 0
Main & jib heel 12.5
Main & jib heel 25
Main & spi heel 0
0
2
4
6
8
10
12
14
16
18
0 20 40 60 80 100 120 140 160 180

eff
[]

]

Main & genoa heel 0
Main & genoa heel 0 2nd run
Main & genoa heel 12.5
Main & genoa heel 12.5 2nd run
Main & genoa heel 25
Main & genoa heel 25 2nd run

Figure 6.15: Change in inclination angle ()
127mm above mast due to presence of
mainsail with genoa
Figure 6.14: Change in inclination angle ()
127mm above mast due to presence of
mainsail with jib or spinnaker
Figure 6.14 and Figure 6.15 show the change in inclination angle () relative to the z-plane
as defined by equation (6.2) plotted against
eff
. For all measurements is positive which
means that the flow is deflected up the mast. is about 1 for small
eff
and increases with

eff
up to 16.5 at
eff
of 180 for mainsail and spinnaker. The mainsail with spinnaker
measurements for small
eff
of 40 and 60 do not follow this trend. increases because the
spinnaker head deflects the flow upwards, which influences the flow above the mast for small

eff
. For slightly larger
eff
the spinnaker head is downstream of the mast and its influence is
less. As
eff
gets larger the mainsail obstructs the flow progressively more and the flow is
deflected over the mainsail resulting in a constantly increasing up to
eff
of 180. There is
no significant trend with heel angle and, as before, mainsail with genoa measurements in
Figure 6.15 are slightly bigger than the mainsail with jib results. The repeat measurements in
Figure 6.15 are all within 0.1.
Having looked at the change in inclination angle () it is also important to consider the total
inclination angle of the flow () relative to the z-plane defined by equation (D.44). can be
seen as the sum of the inclination angle of the free-stream flow, relative to the z-plane due to
heeling and pitching of the yacht, and the change in inclination angle () due to the presence
of the sails. is the flow angle relevant for the cup anemometer measurements. The cup
anemometer rotates approximately about the z-axis so that a flow inclination relative to the z-
plane may affect the measured wind speed. Figure 6.16 shows for mainsail with jib or
spinnaker, and the free-stream flow inclination angle for heeled tests calculated by equation
(6.3). For tests at zero heel, angle the free-stream inclination angle is zero for all
eff
since the
wind tunnel model does not pitch. of up to 20 possibly even 25 can be expected in
normal sailing conditions with mainsail and jib where heel angles are typically smaller than
20 for
eff
over 60. of similar magnitude can also be expected for mainsail with spinnaker
when running very deep and if large heel angles occur in reaching conditions.
172
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.3 Wind Tunnel Study of Flow Above Mast
0
5
10
15
20
25
30
0 20 40 60 80 100 120 140 160 180

eff
[]

]

Main & jib heel 0
Main & jib heel 12.5
Main & jib heel 25
Main & spi heel 0
Free-stream flow heel 12.5
Free-stream flow heel 25
0
1
2
3
4
5
-15 -10 -5 0 5 10 15
Position [mm]

V
/
V


[
%
]

x-position (+ve forward), y=0
y-position (+ve windward), x=0

Figure 6.17: V/V Figure 6.16: Inclination angle () 127mm above
mast without sails and for mainsail
with jib or spinnaker

at different positions in z-plane


127mm above mast for mainsail with
genoa at
A
of 25 and 0 heel
The influence of the sails on the flow above the mast can be expected to strongly depend on
the rig layout. In particular a mainsail with a larger roach should significantly increase the
discussed effects. A masthead rig would also increase the effects, especially with a masthead
spinnaker at small
eff
large increases in V,
eff
and can be expected since the low-
pressure side of the spinnaker head is much closer to the anemometer compared to the tests
conducted in this study which already showed significant changes to the flow above the mast.
The accurate placing of the Cobra Probe in the correct position is difficult. Measurements
were therefore taken at difference positions around the main measurement point to assess the
gradient of flow changes and the influence a slight misplacement of the Cobra Probe has on
the results. Figure 6.17 shows V/V

measured at different positions in the z-plane 127mm


above the mast with mainsail and genoa at
eff
of 25 and 0 heel. Measurements are made at
the correct point directly above the mast and 10mm forward, aft, to starboard and to port of
the mast. 10mm in the wind tunnel equate to approximately 67mm in full-scale. V/V


increases when moving from the windward side to the leeward side from 3% to 3.95%. This
is to be expected since the flow speed should be faster on the low-pressure side compared to
the high-pressure side of the sails. There is also a slight increase from 3.25% to 3.48% when
moving aft. Figure 6.18 shows
eff
for the same measurements. A small increase can be
noted when moving aft (3.3 to 3.6) and leeward (3.4 to 3.6). presented in Figure 6.19
decreases when moving aft (1.1 to 0.6) and to leeward (1.3 to 0.7) because of the tip
vortex spiralling off the mainsail head. Figure 6.17 to Figure 6.19 show the expected trends
and demonstrate that the flow changes within 10mm of the masthead do not vary dramatically
so that variations of a few millimetres in Cobra Probe placement do not significantly affect
the results.
The flow change with height for the same upwind sailing condition with mainsail and genoa
at
eff
of 25 and 0 heel is shown in Figure 6.20.
eff
, and V/V are plotted against the

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 173
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
1
2
3
4
5
-15 -10 -5 0 5 10 15
Position [mm]

e
f
f

[

]

x-position (+ve forward), y=0
y-position (+ve windward), x=0
0
1
2
3
4
5
-15 -10 -5 0 5 10 15
Position [mm]

]

x-position (+ve forward), y=0
y-position (+ve windward), x=0

Figure 6.19: at different positions in z-plane
127mm above mast for mainsail with
genoa at
Figure 6.18:
eff
at different positions in z-plane
127mm above mast for mainsail with
genoa at
height above the mast relative to the mast height. The regular measurement point 127mm
above the mast is 5.3% of the mast height above the mast. A series of measurements is shown
with the Cobra Probe positioned between 5.3% and 37.5% of the mast height above the mast.
The wind tunnel ceiling at about 47% of the mast height above the mast as indicated in Figure
6.20.
eff
and decrease monotonically with increasing height. has a maximum of only
1 and decreases to 0 at a height of 37.5%.
eff
decreases from 3.5 to 1.1 at a height of
37.5% but the trend suggests that
eff
may not decay back to zero by the time the wind
tunnel ceiling is reached. V/V

reduces from 3.2% at the lowest height to almost zero at a


height of 24% but increases again for the last measurement at a height of 37.5%. The increase
is however only 0.7% and the measurement accuracy is estimated at typically within 0.5%.
The flow change with height above the mast is also investigated for a downwind
configuration with mainsail and spinnaker at
eff
of 180 and 0 heel. In this condition the
A
of 25 and 0 heel
A
of 25 and 0 heel
-1
0
1
2
3
4
5
0 5 10 15 20 25 30 35 40 45 50
Height above mast relative to mast height [%]
Change in effective angle []
Change in inclination angle []
(Change in V) / Vinf [%]
Wind tunnel ceiling
-10
-5
0
5
10
15
20
0 5 10 15 20 25 30 35 40 45 50
Height above mast relative to mast height [%]
Change in effective angle []
Change in inclination angle []
(Change in V) / Vinf [%]
Wind tunnel ceiling
Figure 6.20:
eff
, and V/V

vs. height above


the mast with mainsail and genoa at

A
of 25 and 0 heel
Figure 6.21:
eff
, and V/V

vs. height above


the mast with mainsail and spinnaker
at
A
of 180 and 0 heel
174
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.3 Wind Tunnel Study of Flow Above Mast
model has the greatest projected area perpendicular to the flow and a large pitch angle and
reduction in flow speed have been observed in Figure 6.14 and Figure 6.9. Figure 6.21 shows

eff
, and V/V

plotted against height over the mast relative to the mast height in an
equivalent way to Figure 6.20, but with a y-axis range five times larger. reduces with
height above the mast and tends towards zero at the wind tunnel ceiling to satisfy the
boundary conditions.
eff
is very close to zero at a height of 5.3% but becomes slightly
negative up to -1.3, reducing
eff
, for heights between 6% and 15%. V/V

shows no clear
trends but seems to fluctuate between -4% and -10% for heights up to 31%. In contrast, the
highest measurement at 37.5% shows an increase in V/V

of almost 4%. The absence of a


clear trend could be caused by vortex shedding off the top of the mainsail and spinnaker
which affects the repeatability and could also cause a non-monotonic trend. It is interesting to
note that in both the upwind and downwind case presented in Figure 6.20 and Figure 6.21 the
wind speed at the highest measurement point is greater than the free-stream flow without the
model in the tunnel, presumably due to blockage. Both cases confirm that the influence of the
sails on the flow generally decreases with height above the mast but also show that the
influence requires considerable height to decay to a negligible level. The flow changes above
the mast are mainly due to inviscid flow effects so that the results can be expected to scale
well to full-scale. On a full-scale yacht placing the anemometer high above the mast will
reduce the errors due to the flow around the sails but it is not feasible to eliminate them
completely.
Finally, another aspect that should ideally be considered is the effect of sail trim on the flow
above the mast. Figure 6.22 shows
eff
, and V/V

measured in the regular position


127mm above the mast for mainsail with genoa at
eff
of 25 and 0 heel plotted against
heeling moment. To obtain the data the sails are firstly trimmed to the aerodynamically
optimum shape representing the fully powered-up condition producing 100% heeling
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
0 20 40 60 80 100
Heeling moment relative to fully powered-up condition [%]
Change in effective angle []
Change in inclination angle []
(Change in V) / Vinf [%]

Figure 6.22:
eff
, and V/V

vs. heeling
moment 127mm above mast for
mainsail with genoa at
eff
of 25 and
0 heel
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 175
Chapter 6 Wind Tunnel and Full-Scale Comparison
moment. After the fully powered-up run the sails are depowered more and more following
common trimming practice, and using the heeling moment as a guide. The reduction in
heeling moment is also used as a measure of depowering in section 5.4 by defining the trim
parameter power. For simplicity the Real-Time VPP is not used for these tests. Figure 6.22
shows that
eff
, and V/V decrease monotonically where
eff
and have an almost
linear relationship with heeling moment and V/V

is better described by a squared


relationship. As discussed before, the lift and the resulting tip vortex are mainly responsible
for the flow changes above the mast. As the sails are depowered less lift is produced,
particularly close to the top of the mast, and the flow above the mast is affected less.
Ideally the sail trim should be considered when correcting the anemometer measurements but
from the full-scale data it is not feasible to accurately and confidently determine a
benchmark, fully powered-up sailing condition that could be used as a base for the correction.
It would be feasible to use depowering data determined in chapter 5 to predict the full-scale
level of depowering for a given sailing condition. However, the wind tunnel force
measurements are compared to the full-scale measurements and therefore should not be used
for a correction to ensure the data sets remain independent. Only wind tunnel data describing
the flow field around the yacht can justifiably be used for correcting the full-scale data where
scaling effects are less significant. The full-scale measurements were obtained in light to
medium wind conditions and changes to the flow field above the mast due to depowering are
considered to be of secondary importance. Depowering is therefore not accounted for when
correcting the full-scale anemometer measurements.
6.3.3 Model for Correcting Full-Scale Wind Measurements
From the results presented in section 6.3.2 a semi-empirical model to remove the influence of
the sails from the full-scale wind speed and direction measurements can be developed by
fitting polynomials to the relevant data. As will be discussed in detail in section 6.4, the
parameters measured in full-scale are the effective wind angle (
eff
) and the wind speed (V) so
that mathematic descriptions of
eff
and V/V

are required. In addition is also of interest


in section 6.4 when investigating the cup anemometer sensitivity to flow inclination. Full-
scale measurements of mainsail with jib or spinnaker are used for the comparison, and Figure
6.23 shows
eff
for mainsail with jib or spinnaker presented before in Figure 6.11. The data
is now plotted against the effective wind angle measured above the masthead (
effCobra
) and
not
eff
since the actual
eff
is not known when correcting the full-scale data but is determined
by the correction.
eff
for mainsail with jib has a clear trend with heel angle and it therefore
modelled as a function of
effCobra
and heel angle by a B-spline surface. In addition to the
measured data, Figure 6.23 also shows the modelled response surface values at the relevant
heel angles. When modelling the response surface it is assumed that
eff
is zero at of
effCobra
176
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.3 Wind Tunnel Study of Flow Above Mast
0
0.5
1
1.5
2
2.5
3
3.5
4
0 20 40 60 80 100 120 140 160 180

effCobra
[]

e
f
f

[

]

Main & jib heel 0
Main & jib heel 0 spline surface
Main & jib heel 12.5
Main & jib heel 12.5 spline surface
Main & jib heel 25
Main & jib heel 25 spline surface
Main & spi heel 0
Main & spi 5th order polynomial
0.9
0.95
1
1.05
1.1
1.15
1.2
1.25
0 20 40 60 80 100 120 140 160 180

eff
[]
f
V

[
-
]

Main & jib heel 0
Main & jib heel 12.5
Main & jib heel 25
Main & jib 5th order polynomial
Main & spi heel 0
Main & spi 5th order polynomial
Figure 6.24: Wind speed factor (f
zero since the sails would not produce any lift and therefore should not cause upwash. For

effCobra
larger than 90 some approximations have to be made as well since no wind tunnel
tests were conducted. It is assumed that the behaviour is similar to that observed with
mainsail and spinnaker but that values at 0 heel would generally be less since the jib has
significantly less area high up compared to the spinnaker that can influence the flow.
Mainsail and jib data for downwind sailing conditions is of limited interest in this study so
that the accurate modelling of
eff
at large
effCobra
is not critical.
The model with mainsail and spinnaker has only been tested upright since heel angles were
generally small when sailing with Dyna in this configuration.
eff
for mainsail with
spinnaker is therefore only modelled as a function of
effCobra
and a sufficiently good data fit
can be achieved by a fifth order polynomial of the form

5
5
4
4
3
3
2
2 1 effCobra effCobra effCobra effCobra effCobra eff
a a a a a + + + + = ,
(6.4)
where a
1
to a
5
are the free parameters obtained from the regression fit of the wind tunnel data.
As for mainsail and jib the y-axis intersect (a
0
) is assumed to be zero for mainsail and
spinnaker as well. The fifth order polynomial fit is shown in Figure 6.23 and the values
chosen for the free parameters are given in Table 6.1.
Figure 6.24 shows the change in wind speed presented in Figure 6.9 plotted as the wind speed
factor f
V
=V/V

. Unlike
eff
, f
V
is still plotted against
eff
since the actual
eff
can be obtained
by the above-described correction before calculating f
V
. Since f
V
for mainsail and jib does not
show a clear trend with heel angle, both f
V
for mainsail with jib and spinnaker are modelled as
functions of only
eff
by a fifth order polynomial of the form
Figure 6.23: Change in effective wind angle (
eff
)
as modelled for the full-scale data
correction
V
=V/V

) as
modelled for the full-scale data
correction
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 177
Chapter 6 Wind Tunnel and Full-Scale Comparison
5
5
4
4
3
3
2
2 1
1
eff eff eff eff eff V
a a a a a
V
V
f + + + + + = =

,
(6.5)
where the free parameters a to a
1 5
are obtained from the wind tunnel data. It is assumed that
flow speed is not affected at
eff
of zero since the projected area of the model blocking the
flow would be very small and a
0
is therefore given a fixed value of one. Since no wind tunnel
measurements were conducted for mainsail and jib at
eff
above 90 it is assumed that f
V

behaves similar for mainsail with jib and spinnaker for large
eff
but that the reduction in
wind speed is less with the jib since the projected area blocking the flow is significantly less
compared to the spinnaker. As mentioned before downwind sailing conditions with mainsail
and jib are not the focus of the investigation and an accurate correction is therefore not
critical. For small
eff
with mainsail and spinnaker a weighting point is added so that the fifth
order polynomial fits the measured data more closely. Sailing on Dyna with mainsail and
spinnaker at
eff
below 60 is unusual and accurate modelling of f
V
at small
eff
is therefore
not important. The resulting fifth order polynomial regressions are shown in Figure 6.24 and
the values chosen for the free parameters are given in Table 6.1.
Although the change in inclination angle () is not required for correcting the full-scale
data, it is useful for the cup anemometer calibration discussed in section 6.4. Figure 6.25
shows for mainsail with jib and spinnaker plotted against
eff
from Figure 6.14. does
not show a clear trend with heel angle and the results for mainsail with jib and spinnaker are
very similar so that all data is modelled by one fourth order polynomial as a function of
eff

with
4
4
3
3
2
2 1 eff eff eff eff
a a a a + + + = ,
(6.6)
where the free parameters a to a are obtained from the wind tunnel data and are given in
1 4
Table 6.1. As with
eff
, is also assumed to be zero at
eff
of 0 and the resulting fourth
order polynomial is shown in Figure 6.25.
Table 6.1: Regression parameters a
1
to a
5
obtained from wind tunnel tests to account for changes in the
flow field at the position of the cup anemometer of Dyna
a a a a a Parameter Equ. Sail set
1 2 3 4 5
-5

eff
(6.4) Main &
spi
0.2202 -0.0046 4.5510 -2.2310
-7 -10
4.1810
f
V
=V/V

(6.5) Main &


jib
0.0032 -1.3110
-4 -6
2.1310 -1.3910
-8 -11
3.0510
f
V
=V/V

(6.5) Main &


spi
0.0184 -6.0010
-4 -6
7.6910 -4.2910
-8 -11
8.5810

(6.6) Main &
jib or spi
0.0541 -9.3510
-4 -5
1.1410 -2.8010
-8
-

178
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.4 Full-Scale Aerodynamic Data Analysis
0
2
4
6
8
10
12
14
16
18
0 20 40 60 80 100 120 140 160 180

eff
[]

]

Main & jib heel 0
Main & jib heel 12.5
Main & jib heel 25
Main & spi heel 0
All data 4th order polynomial

Figure 6.25: Change in inclination angle () as
modelled for the cup anemometer
calibration
6.4 Full-Scale Aerodynamic Data Analysis
The full-scale measurements are kindly made available by the Technical University Berlin for
this study and the data analysis and selection are conducted by the author. The wind direction
is measured by the direction vane above the top of the mast. Following Hochkirch (2000) it is
assumed that the direction vane is only influenced by the flow component perpendicular to its
rotational axis and that it is insensitive of the flow component parallel to its rotational axis.
The rotation axis is assumed to be perpendicular to the deck plane (z-plane) so that the angle
of the flow component in the z-plane is measured, i.e. the effective angle. The flow angle
measured by the direction vane (
Vane
) is relative to the top of the mast and the mast twist
(
Mast
) needs to be considered to determine the flow angle above the mast relative to the yacht
centreline ( ).
Anemometer
The instrumentation to measure
Mast
did not work reliably and data was only recorded for a
few runs. The recorded twist data is therefore used to model
Mast
so that a mast twist
correction can be applied to all full-scale measurements. Figure 6.26 shows the wind speed
measured by the cup anemometer (V ) plotted against
Anemometer Vane
for the full-scale runs
where
Mast
was recorded. It can be seen that the main change in V
Anemometer
is due to changes
in
Vane
and that V for a given
Anemometer Vane
varies comparatively little. All
Mast

measurements are made on the same day over a period of about three hours and the true wind
speed does not change a lot. After all the corrections that will be described in this section
have been applied, the true wind speed can be calculated to be between 6-8m/s for most
measurements. When a side force is applied to the mast the rig tends to pivot about the
windward spreader causing the mast to twist. The amount of twist is driven by the
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 179
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
2
4
6
8
10
12
0 20 40 60 80 100 120 140 160 180
|
Vane
| []
V
A
n
e
m
o
m
e
t
e
r

[
m
/
s
]
Starboard tack
Port tack
-1
-0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
0 500 1000 1500 2000 2500 3000
F
Y
[N]

M
a
s
t

[

]
Starboard tack
Port tack
0.001072*x
Figure 6.26: Cup anemometer speed (V
longitudinal bend and rake of the mast and spreader angle
24
. The mast twist characteristics
not only depend on the mast design but also on the rig set-up. The force couple caused by the
mainsail side force being applied at the trailing edge of the mast and the shroud reaction force
being applied close to the longitudinal centre of the mast also contributes to the mast twist.
This occurs particularly at the unsupported top section of a fractional rig such as Dynas.
Mast twist (
Mast
) is driven to a large extend by side force (F
Y
) and Figure 6.27 shows
Mast

plotted against F
Y
for measurements taken on starboard and port tack, where measurements
on port tack have been mirrored about the origin. The measurements on port tack show more
scatter, and on starboard tack the values are generally somewhat higher. A general trend of
increasing mast twist with increasing F
Y
can be seen. It is difficult to assess if the
instrumentation operated correctly during these runs, but the measurements are the only data
available to develop a correction and according to the mast designer twist angles of up to 3.5
are very normal for this rig in these sailing conditions. Theoretically the mast twist should be
the same on starboard and port tack, and due to the large scatter and limited number of
measurements, a symmetrical correction model is used which assumes that the mast twist is
zero if no side force is generated. The data is not accurate enough to fit a higher order
polynomial and a linear regression is applied to all the data. The regression line is shown in
Figure 6.27 and is given by

Y Mast
F 001072 . 0 = .
(6.7)
Since torsion has a linear relationship to the applied force, a relationship close to linear
between mast twist and side force can be expected and the first order regression line is
adequate.

24
Personal communication with Chris Mitchell, Applied Engineering Services Ltd., P.O. Box 8445, Auckland
New Zealand, May 2006
Anemometer
)
vs.
Figure 6.27: Mast twist (
Mast
) as a function of side
force (F
Vane
for full-scale mast twist
measurements used in analysis
Y
) measured on Dyna and
linear regression fit
180
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.4 Full-Scale Aerodynamic Data Analysis
The flow angle relative to the centreline of the yacht (
Anemometer
) above the mast is calculated
from the direction vane (
Vane
) measurement and the mast twist correction in equation (6.7) so
that
(
Y Mast Vane Anemometer
F + = ) ,
(6.8)
where
Mast
is a point-symmetric function. The free-stream effective angle (
eff
) can now be
calculated by subtracting the change in effective wind angle (
eff
) at the position of the
anemometer due to the presence of the sails with
,
eff Anemometer eff
=
(6.9)
where
eff
for mainsail with jib is modelled by a B-spline response surface dependent on
and as discussed in section 6.3.3.
Anemometer eff
for mainsail with spinnaker is
approximated by the polynomial given in equation (6.4) and the coefficients listed in Table
6.1 as a function of .
Anemometer eff
is always defined as point-symmetrical. Unlike the model
in the wind tunnel, Dyna not only sails at a heel angle but also at a pitch angle. Although the
pitch angle () is small and does not have a significant effect, it is considered here when
transforming
eff
into the horizontal plane to obtain
A
. This can be achieved by describing
the flow direction by the unit vector and following the inverse process shown in equations
(2.15) to (2.18) or through
[ ]
o
180 , 0 ,
cos
sin sin cos tan
tan
1

+
=

A
eff
A


,
(6.10)
where is the heel angle and a positive pitch angles () means bow down trim of the yacht.
The wind speed is measured above the top of the mast by a cup and calorimetric sensor
anemometer. As mentioned in section 6.2 the calorimetric sensor anemometer is only used to
calibrate the cup anemometer due to its sensitivity to the heat radiation of the sun. The
calorimetric sensor anemometer (SCHMIDT SS 20.01) is independent of the flow direction in
the z-plane and compensated to measure the total flow at inclination angles from the z-plane
of up to 45 and the accuracy within this range is 2% (Schmidt, 1998). Hochkirch (2000)
calibrated the cup anemometer by conducting measurements with Dyna at night for a range of
apparent wind and heel angles using the mainsail and jib. It is important to conduct such a
calibration to check whether the cup anemometer measures the total wind speed (as does the
calorimetric sensor) or only the wind speed component perpendicular to its ration axis, i.e. the
flow component in the z-plane which is of course the effective wind speed. The inclination
angle () of the flow relative to the z-plane primarily depends on the heel angle and the
effective wind angle, provided that the influence of the sails on the flow is ignored and the
onset flow is assumed to be horizontal. Figure 6.28 shows the heel angle () of Dyna plotted
against the effective wind angle (
eff
) for the calibration tests and the measurements which are
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 181
Chapter 6 Wind Tunnel and Full-Scale Comparison
selected for the comparison to the wind tunnel tests later in this section. It can be seen that the
calibration tests cover the range of and
eff
encountered in the typical sailing conditions
selected for the comparison.
The actual inclination angle () relative to the z-plane for the calibration tests is not known,
but assuming horizontal free-stream onset flow the inclination angle ( ) can be calculated as

, ) sin sin sin cos cos ( sin


1

A A
=

(6.11)
and is plotted against
eff
in Figure 6.29. The actual inclination angle is estimated by adding
the change in inclination angle () due to the presence of the sails, determined from the
wind tunnel tests and defined in equation (6.6), to so that

+ =

.
(6.12)
With the effective wind speed (V
eff
) can be obtained from the total wind speed (V) as
cos V V
eff
= .
(6.13)
For the likely inclination angles of up to approximately 25, a difference between V
eff
and V
of up to 9% can be expected, but for a more common inclination angles of about 15 the
difference is only 3.5%, which is of the same order as the measurement accuracy of the
calorimetric and cup anemometer, 2% and 1% respectively. Using a higher order
regression fit to evaluate the anemometer dependency on inclination angle and develop a
correction model is therefore not feasible, and the primary goal can only be to determine
whether the cup anemometer is more likely to measure the effective wind speed (V
eff
) or the
total wind speed (V). To constrain the problem as much as possible the wind speed measured
by the cup anemometer (V ) is assumed to proportionately follow the relationship in
Anemometer
0
5
10
15
20
25
30
35
40
0 30 60 90 120 150 1
|
eff
| []
|

|

[

]
80
Measurements used in sail
force comparison investigation
Calibration measurements
-5
0
5
10
15
20
25
0 30 60 90 120 150 180
|
eff
| []

]
Calibration measurements
Figure 6.29: Inclination angle of free-stream flow
(
Figure 6.28: Heel angle () of Dyna with mainsail
and jib plotted against effective wind
angle (

) plotted against effective wind
angle ( ) )
eff eff
182
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.4 Full-Scale Aerodynamic Data Analysis
Table 6.2: Regression parameters a
equation (6.13) so that

( )

2
1
2
2 1
2
1
sin 1 sin 1 b
f a f a
b
V
V
Anemometer Anemometer Anemometer

+
=

= ,
(6.14)
where V
Anemometer
is obtained from the rotational frequency (f
Anemometer
) of the cup anemometer
through a second order polynomial and a
1
, a
2
and b
1
are the regression parameters. If b
1

obtained from the regression fit is close to zero the cup anemometer measures V, but if b
1
on
the other hand is close to one V
eff
is measured. The total wind speed (V) is assumed to be
measured by the calorimetric sensor anemometer during the calibration tests and, in Figure
6.30, is plotted against the measured f
Anemometer
and , which is calculated from equation (6.12)
and includes the estimated change in inclination angle due to the presence of the sails. In
addition to the calibration measurement points Figure 6.30 also shows the least squares fit
response surface obtained from the data with equation (6.14). The resulting values of the
regression parameters are shown in Table 6.2 and it can be seen that the relationship between
f
Anemometer
and V is close to linear, and that there is some dependency of f
Anemometer
on , as b
1
is
determined to be 0.31. Figure 6.31 shows different wind speed ratios as a function of to
clarify how the value of b
1
=0.31 is to be interpreted. If the anemometer measures the total
wind speed the ratio is one for all , and if the anemometer measures the effective speed the
0.8
0.85
0.9
0.95
1
1.05
0 5 10 15 20 25
[]
W
i
n
d

s
p
e
e
d

r
a
t
i
o

[
-
]
V/V=1
Veff/V=cos(delta)
Vanemometer/V=(1-b1*sin(delta)^2)^0.5
Vanemometer/V (measured)
Figure 6.30: V plotted against f
Anemometer
and
showing calibration measurements
and fitted response surface
Figure 6.31: Different wind speed ratios plotted
against inclination angle ()
1
, a and b
2 1
obtained from cup anemometer calibration tests on
Dyna
Description a a b Parameter Equ.
1 2 1
V (6.14) Data regression to assess dependency
on
0.97534 -0.007321 0.31221
V (6.14) Data regression by Hochkirch (2000)
assuming no dependence on
0.98221 -0.006992 0

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 183
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
4
8
12
16
20
0 4 8 12 16
f
Anemometer
[Hz]
V

[
m
/
s
]
20
Calibration measurements
Poly. (Calibration measurements)
Linear (Calibration measurements)

Figure 6.32: V plotted against f
ratio reduces by about 9% at a of 25. The modelled ratio of the anemometer measurement
over the total wind speed reduces by about 31% of the effective speed reduction due to the
value of b
1
=0.31. Figure 6.31 also includes the individual measurements of anemometer
speed over total wind speed which shows that the scatter in the data is of similar magnitude to
the changes investigated. The average absolute error between the measured and modelled
ratios is 1.1%. It is therefore not advisable to model the anemometer measurements as a
function of but from the regression fit it is concluded that it is more accurate to assume that
the cup anemometer reads the total wind speed rather than the effective wind speed.
The total wind speed (V) measured by the cup anemometer above the mast is therefore
modelled as a function of only f
Anemometer
as shown in Figure 6.32. It can be seen that a linear
interpolation gives a reasonable result and that a good fit can be obtained by using a second
order polynomial, which is still of the form shown in equation (6.14) but with b
1
equal to
zero. Hochkirch (2000) used the same mathematical description in his analysis of the
calibration data so that for consistency his original regression parameters, as shown in Table
6.2, are retained.
From the total wind speed (V) measured above the mast the free-stream flow is calculated by
applying the wind speed factor (f
V
) due to the presence of the sails given in equation (6.5) so
that

) ( ) (
eff V
Anemometer
eff V
A
f
V
f
V
V

= = .
(6.15)
Anemometer
showing
calibration measurements and linear
and second order regression fit


184
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.4 Full-Scale Aerodynamic Data Analysis
The two important flow characteristics of the horizontal free-stream onset flow at the height
of the anemometer are now determined, the apparent wind angle (
A
) and speed (V
A
) with
equations (6.10) and (6.15) respectively.
From the geometric relationship of the wind triangle shown in Figure 2.5 the true wind angle
( ) and the true wind speed (V ) at the height of the anemometer can be calculated with
T T
( )
( )
[ ]
o
180 , 0 ,
cos
sin
tan
1

+
+
=

T
S A A
A A
T
V V
V



,
(6.16)
and
) cos( 2
2 2
+ + =
A A S A S T
V V V V V ,
(6.17)
where V
S
is the boat speed and is the leeway angle measured on Dyna as described in
section 6.2. From V at the height of the anemometer, V
T T
at the reference height, usually 10
metres above the water, can be calculated with the inverse of equation (2.20), which assumes
a logarithmic velocity profile for the lowest part of the planetary boundary layer. The height z
of the anemometer above the water is obtained from
cos
Anemometer
z z = ,
(6.18)
where z
Anemometer
is the height of the anemometer above the DWL in coordinate system B.
Defining V
T
at the reference height of 10 metres is common in meteorological data and is
hence also used in most VPPs as the reference height for V .
T
From an aerodynamic point of view the wind speed and direction at the height of the centre of
effort of the sails is more important since it is the most appropriate height for calculating V
eff

and
eff
for determining the force and moment coefficients. For calculating V
eff
and
eff
the
height of the centre of effort is usually approximated as being the height of the geometric
centre of area of the sails (z
CoA
). Since z
CoA
is defined in coordinate system B as the height
above the DWL, it reduces as the yacht heels and the height above the water z is obtained
from equation is used to calculate V (2.24). Equation (2.20)
T
at the height z again assuming a
logarithmic velocity profile and a roughness length (z
0
) of 0.25mm. An average true wind
speed of 7m/s is assumed to obtain z =0.25mm from equation (2.21), equivalent to section
0
2.5.2 where the same is assumed for the velocity profile in the wind tunnel. The change in
T

with height can be ignored at these low heights above the surface as mentioned in section
2.3.2. Now
A
and V
A
at height z can be calculated from and V at height z, and V
T T S
and
with equations (2.9) and (2.10). Applying the effective angle theory,
eff
and V
eff
at height z
are finally obtained from
A
and V
A
at height z, and and with equations (2.13) and (2.14).
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 185
Chapter 6 Wind Tunnel and Full-Scale Comparison
As described in section 6.2 the aerodynamic forces measured by the rig force balance are
given in coordinate system B and no further transformation is required as following the
effective angle theory described in section 2.3.1, the force and moment coefficients of the
sails are usually calculated in the heeled (and pitched) plane (z-plane). The force and moment
coefficient vectors C
F
and C
M
are calculated from the measured forces and moments with
equations (2.38) and (2.39) where equation (2.2) is used to obtain the effective dynamic
pressure (q
eff
) from V
eff
. C and C from C , C
L D
are then calculated with equation (2.42)
Fx Fy
and

eff
. The centre of effort height and longitudinal position are obtained from the intersection of
the central axis with the y-plane using equations (2.31) and (2.32).
The aerodynamic full-scale forces measured by the rig force balance include the rig (mast and
rigging) and the sail components. Rig windage measurements are carried out in full-scale but
the data are not accurate enough to deduce meaningful values for windage. Since the wind
tunnel measurements also include the rig windage component, the rig windage is not
subtracted for the comparison. Since the rig windage approximations are obtained from wind
tunnel measurements in the first place, a possible Reynolds Number scaling effect between
wind tunnel and full-scale windage could not be corrected for by subtracting the
approximated windage. Therefore the comparison is carried out based on the aerodynamic
forces on the sails and rig.
For the comparison the most suitable data is selected from the full-scale measurements
conducted on Dyna in 1999 as part of the initial project of realising a full-scale sailing
dynamometer by Hochkirch (2000). Measurements were also conducted on Dyna between
2002 and 2004 for the follow-up projects mentioned in section 1.3, but for various reasons the
original sail force data from 1999 is seen as being the most reliable. In 2002 tests on Dyna
were conducted to investigate shallow draft keels. The sail forces were measured but the
focus was on the keel force balance measurements so that the correct operation of the rig
force balance was not always closely monitored. The third project in 2003 and 2004 focused
on capturing the sail shape and a lot of time was spent implementing the system so that the
number of good test runs is limited and due to the added measuring equipment in the mast,
the rig force balance calibration might be affected. As in 2002 the focus was not on
measuring the rig forces accurately and very few downwind measurements with spinnaker
were made.
The original data from 1999 is firstly filtered by dates checking whether sensible values have
been recorded for all relevant variables and whether the true wind speed and direction were
consistent during the day. Large variations in V and
T T
indicate either problems with some
measuring device or unstable wind conditions, which make the sail trim and force
measurements less reliable. Only a limited number of full-scale measurements were
conducted with the mainsail and genoa in light airs. Due to the lack of data and the inherent
186
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.4 Full-Scale Aerodynamic Data Analysis
reduction in accuracy when measuring small forces and small wind speeds, this sail
configuration is not considered in the comparison. Full-scale measurements have been carried
out with the reefed mainsail, but these are not considered for the comparison because at
higher wind speeds inaccuracies due to dynamic effects and trimming are more likely, and
corresponding reefed wind tunnel measurements would need to be carried out. Only the two
most frequently used sail combinations are therefore investigated, mainsail with spinnaker
and mainsail with jib. The data series from which measurements are chosen for the
comparison are shown in Table 6.3. The sampling time for each run is either 25.6 seconds
(256 samples at 10Hz) or 32 seconds (256 samples at 8Hz) so that several periods of dynamic
effects such as pitching and rolling of the yacht are included in the averaged data.
Measurement runs were aborted and not considered during the analysis when non-continuous
periodic motion, due to for example the wake of another vessel, was encountered or when
sudden changes in apparent wind direction or speed, due to either changes in the true wind or
inaccurate steering by the helmsman, occurred. Hochkirch (2000) analysed the raw data and
excluded measurement runs containing obvious errors by identifying variables with a larger
than usual standard deviation and with values outside the realistic range. Only the averaged
data of each run in Table 6.3, which is not excluded by the raw data analysis of Hochkirch
(2000), is used for the analysis in this work.
Figure 6.33 shows the true wind speed (V
T
) for the series with mainsail and spinnaker listed
in Table 6.3. For effective wind angles between 80-180 the majority of measurements were
obtained in V
T
between 4-8m/s. Measurements at
eff
smaller than 80 were taken in V
T

between 2-4m/s because carrying a spinnaker at very shy angles is only feasible in light airs.
Figure 6.34 shows V
T
for the full-scale series with mainsail and jib. For the comparison in
upwind configuration small effective wind angles are of most interest. The upwind data can
be grouped into a light wind data set of V less than 4m/s and a moderate breeze data set of V
T T
Table 6.3: Summary of the test series used in the analysis of the full-scale data
Test Test series Sail sets used in the
analysis
Average Average wind
direction []
Sea state Measure-
25
V phase (date) [m/s] ment runs
T
Baltic B 990707 Mainsail + jib 6.8 316 Smooth 76
Baltic C 990714 Mainsail + jib 8.1 257 Slight 28
59 Mainsail + jib 4.0 74
Slight 990718
24 Mainsail + spinnaker 5.1 67

25
Not all measurement runs are used in the comparison.
990719 Mainsail + spinnaker 7.0 91 Slight 14
Baltic D
990720 Mainsail + spinnaker 5.6 205 Slight 27
hS ca. 1m 990814b Mainsail + spinnaker 7.3 179 33
Mainsail + jib 7.7 267 Slight to
moderate
26
990816
Mainsail + spinnaker 9.9 263 4
990817
Mainsail + jib
Mainsail + spinnaker
3.5
2.6
174
166
Slight
33
14
Baltic E
Mainsail + jib 7.0 151 Slight to
moderate
27
990818
Mainsail + spinnaker 7.3 144 14
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 187
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
2
4
6
8
10
12
14
0 20 40 60 80 100 120 140 160 180
|
eff
| []
V
T

[
m
/
s
]
Full-scale
Full-scale VT<4m/s
Full-scale VT=6-8m/s
0
2
4
6
8
10
12
14
0 20 40 60 80 100 120 140 160 180
|
eff
| []
V
T

[
m
/
s
]
Full scale
Figure 6.33: True wind speed (V Figure 6.34: True wind speed (V
between 6-8m/s as shown in Figure 6.34. This confines the true wind speed range for each
data set but still gives sufficient samples in each set. The true wind speed during the full-scale
measurements needs to be considered in case the sails are depowered. In chapter 5 the wind
tunnel study using the Real-Time VPP showed that the sails are fully powered-up in 4m/s true
wind speed but should be depowered in 6-8m/s true wind speed when sailing upwind.
The full-scale heel angles with mainsail and jib in V
T
between 6-8m/s are regularly between
25 and 30 as shown in Figure 6.35, which indicates that the sails may have been depowered
to avoid excessive heel angles that would slow the yacht down. The heel measurements in V
T

less than 4m/s are, on the other hand, too small to expect depowering of the sails. For the
downwind configuration with mainsail and spinnaker V
T
is very low for small
eff
and Figure
T
) during selected
full-scale runs with mainsail and
spinnaker plotted against
T
) during full-scale
runs with mainsail and jib plotted
against
eff eff
-5
0
5
10
15
20
25
30
35
40
0 20 40 60 80 100 120 140 160 180
-
eff
[]

]
Full-scale
Full-scale VT<4m/s
Full-scale VT=6-8m/s
-5
0
5
10
15
20
25
30
35
40
0 20 40 60 80 100 120 140 160 180
-
eff
[]

]
Full scale
Figure 6.35: Heel angle () during full-scale runs
with mainsail and jib plotted against

Figure 6.36: Heel angle () during selected full-


scale runs with mainsail and
spinnaker plotted against
eff eff
188
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.4 Full-Scale Aerodynamic Data Analysis
6.36 shows that most heel angles are below 15 so that depowering is not to be expected for
most measurements. Therefore, only one data set for all V
T
is used for the mainsail and
spinnaker measurements.
The full-scale results have a lot of scatter due to the constantly changing test environment and
the constantly varying flying shape of the sails caused by trim changes as well as dynamic
effects due to the yachts motion and the fluctuations in the onset flow. The large amount of
scatter indicates that the sail trim was not always optimal for the instantaneous sailing
condition during a measurement. A least squares data fit is therefore inappropriate. In
addition data fitting can have a great influence on the resulting curves with such a high
degree of scatter in the data. The full-scale measurements are therefore not described by a
fitted curve but the individual measurements are shown in the comparison.
One contributor to the scatter in the data is asymmetry in the measurements on starboard and
port tack. For small apparent wind angles the drive force coefficient (C
Fx
) has a high gradient
with respect to
eff
and compared to other quantities has little scatter so that it shows the
asymmetry well. Figure 6.37(a) shows C for mainsail with jib in V
Fx T
less than 4m/s plotted
against
eff
. The starboard and port measurements do not fall on one curve; the starboard
measurements are lower for a given
eff
. A similar trend can be seen in Figure 6.37(b) for the
mainsail with jib measurements in V
T
between 6-8m/s, but the difference between the
starboard and port measurements is much smaller. The amount of asymmetry in the data
seems to be related to the wind strength. Both sets of data contain measurements obtained on
several days. If the data for each day is analysed separately, a larger offset is observed on
days with light wind and a smaller offset on windier days.
There is a number of possible causes for an asymmetry performance of the yacht. Any
component that is not symmetrical about the centreline and thereby influences the
performance of the yacht to be different on one tack compared to the other causes asymmetry.
This could be the hull, keel or the mast set-up. However, on Dyna this is unlikely since the
hull and keel are manufactured under quality control and as a research vessel, considerable
time and effort was spent installing the keel and setting up the rig utilising the load cells in
each rigging component. A difference in the way the yacht is sailed on starboard and port
tack could also be responsible for asymmetric performance, but achieving such a consistent
difference is unlikely, especially as the yacht was helmed and the sails were trimmed by a
number of different people.
The most likely cause for a constant offset is the misalignment of the direction vane on top of
the mast relative to the yacht centreline. This means that the wind angle measured (
Vane
) on
starboard and port tack differs by twice the vane misalignment. Figure 6.37(d) shows the data
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 189
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 20 40 60 80 100 120 140
|
eff
| []
C
F
x

[
-
]
VT<4m/s starboard
VT<4m/s port
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 20 40 60 80 100 120 140
|
eff
| []
C
F
x

[
-
]
VT<4m/s starboard
VT<4m/s port
(a) C
for V
T
between 6-8m/s, corrected for a -1.5 offset in the measured wind direction (
Offset
).
The scatter in the data is reduced and a direction vane misalignment of 1.5 is imaginable.
One possible reason of the increased asymmetry on the light wind days could be the
metrological phenomenon of wind shear where the true wind direction changes with height
close to the water surface. Batt (2006) explains that wind shear usually occurs in a stable
atmosphere when the air temperature is warmer than the sea temperature and is generally
associated with a developing sea breeze. Wind shear is mainly caused by the Coriolis force so
that the wind direction generally changes clockwise with increasing height above the water in
the Northern Hemisphere. Wind shear therefore generally increases the measured wind angle
on starboard tack and requires the sails to be trimmed with more twist. On port tack the
measured wind angle is decreased and the sails need to be trimmed with less twist. Wind
shear does not only result in an offset in effective wind angle, but also changes the flow
Fx
vs.
eff
for V (c) C
T
<4m/s

Fx
vs.
eff
with
Offset
=-5.25, V
T
<4m/s

0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 20 40 60 80 100 120 140
|
eff
| []
C
F
x

[
-
]
VT=6-8m/s starboard
VT=6-8m/s port
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 20 40 60 80 100 120 140
|
eff
| []
C
F
x

[
-
]
VT=6-8m/s starboard
VT=6-8m/s port
(b) C
Fx
vs.
eff
for V
T
=6-8m/s (d) C
Fx
vs.
eff
with
Offset
=-1.5, V
T
=6-8m/s
Figure 6.37: Effective wind angle ( ) offset correction for full-scale measurements with mainsail and jib
eff
190
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.4 Full-Scale Aerodynamic Data Analysis
characteristics over the sails on each tack and hence the forces. However, in this case it is
most practical to adjust the measured wind angle (
Vane
) to make the yacht performance as
symmetrical as possible, but it should be noted that the offset is not physically the wind shear
angle between the centre of sail area and the top of the mast, rather the angle change required
for the different flying sail shapes to produce the same drive force.
The trend of the data in Figure 6.37(a) agrees with the effect expected from wind shear but it
is not possible to conclude with certainty that wind shear is the cause of the offset without
having more detailed weather information. Although the cause for the asymmetry in the data
cannot be determined with certainty it should be corrected for, to reduce the scatter in the
data. Figure 6.37(b) shows C for mainsail and jib for V less than 4m/s with
Fx T Offset
of -5.25
so that starboard and port measurements fall onto one curve.
From the starboard and port measurements of C
Fx
versus
eff
with mainsail and spinnaker
shown in Figure 6.38(a), it is not possible to determine any asymmetric trend in the data due
to the large scatter and limited number of data points for small effective wind angles. The
measurements in upwind configuration indicate that the asymmetry is related to the wind
strength on a given day, which could be due to direction vane misalignment and wind shear.
Therefore it is assumed that the same asymmetry applies for measurements in downwind
configuration. An offset to the wind angle measurements (
Offset
) for the upwind and
downwind configuration is applied based on the average true wind speed during the
measurements on a given day. For light wind days up to 5m/s true wind speed,
Offset
of -5.25
is applied and on days with average true wind speeds greater than 5m/s,
Offset
is taken as
-1.5. Figure 6.38(b) shows the mainsail with spinnaker measurements adjusted accordingly.
0
0.5
1
1.5
2
2.5
3
3.5
0 20 40 60 80 100 120 140 160 180
|
eff
| []
C
F
x

[
-
]
Starboard
Port
0
0.5
1
1.5
2
2.5
3
3.5
0 20 40 60 80 100 120 140 160 180
|
eff
| []
C
F
x

[
-
]
Starboard
Port
(a) C (b) C
Fx
vs.
eff Fx
vs.
eff
with offset dependent on day
Figure 6.38: Effective wind angle (
eff
) offset correction for full-scale measurements with mainsail and
spinnaker
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 191
Chapter 6 Wind Tunnel and Full-Scale Comparison
is included in equation (6.8)
Offset
so that the flow angle relative to the centreline of the
yacht ( ) above the mast is now calculated from
Anemometer
,
Offset Y Mast Vane Anemometer
F + + = ) (
(6.19)
where
Mast
is a point-symmetric function.
Although the distinction is made here between two sets of wind conditions, a constant
roughness length of 0.25mm as mentioned before is used for all measurements. Following
equation , 0.25mm corresponds to a V (2.21)
T
of 7m/s which is very applicable for the data
collected in moderate winds. Values for z as low as 0.05mm are predicted by equation (2.21)
0

for the light wind days. This seems a very small value. Friction lengths are about 30 times
smaller than the objects causing them so that the water distortions on top of the wave surface
would only be 1.5mm. A distortion height of 7.5mm corresponding to z
0
=0.25mm seems
more realistic and smaller values are unlikely. The resulting difference for z
0
between 0.05-
1mm has been investigated and although it is noticeable it does not fundamentally change the
comparison to the wind tunnel measurements, which assume z =0.25mm. A z
0 0
of 0.25mm is
therefore assumed for all full-scale measurements. In addition the light wind is most likely
caused by thermal effects and is not a gradient wind so that equation (2.21) might not be
accurate anymore. Due to the uncertainly a constant value in between the extremes is the
most sensible choice.
6.5 Wind Tunnel Data
The wind tunnel tests for the comparison with the full-scale data were carried out together
with the tests conducted for the study described in chapter 3. The total forces acting on the
model were measured by the wind tunnel force balance described in section 2.4.1 and
transformed into coefficient form in coordinate system B with equations (2.34) to (2.42). The
full-scale measurements do not include the hull/deck force component, which is included in
the wind tunnel measurements. As discussed in chapter 3 there is a strong interaction between
the hull/deck and the sails so that simply subtracting the tare hull/deck forces without the
sails present is not accurate. The hull/deck forces were therefore simultaneously measured
with the sails present by a secondary force balance inside the hull as discussed in section 3.2.
The hull/deck forces were then subtracted from the total forces to yield the aerodynamic
forces acting on the rig and sails. For the comparison, wind tunnel tests were conducted for a
downwind sailing configuration with mainsail and spinnaker, and an upwind configuration
with mainsail and jib as shown in Figure 6.39. The downwind and upwind speed and twist
profiles developed for Dyna as shown in section 2.5.2 were used as for any other tests and the
cardboard cutouts discussed in appendix B.3 covered the trough. All tests with mainsail and
192
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.5 Wind Tunnel Data
(b) Upwind configuration: (a) Downwind configuration:
Mainsail and jib (100%) Mainsail and spinnaker
= 40 to 180 = 20 to 90
A A
= 0 = 0, 12.5, 25
Flow speed change over span of sails = 26% Flow speed change over span of sails = 20%
Twist over span of sails = 9.5 No twist
Figure 6.39: Wind tunnel model of Dyna in downwind configuration with mainsail and spinnaker and in
upwind configuration with mainsail and jib
spinnaker were conducted upright whereas upwind tests with mainsail and jib were carried
out at 0, 12.5 and 25 heel. The wind tunnel test matrix is shown in Table 6.4. The sails
were trimmed to the aerodynamically most efficient shape by maximising the drive force in
the tunnel. The trim settings were the same as those used in chapter 3 and as the fully
powered-up trims for mainsail and jib in chapter 5.
Table 6.4: Wind tunnel test matrix

A
Sail set

20 25 30 40 50 60 80 90 100 120 140 160 180
Mainsail &
0 x x x x x x x x x
spinnaker
0 x x x x x x
12.5 x x x x x x
Mainsail &
jib
25 x x x x x x
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 193
Chapter 6 Wind Tunnel and Full-Scale Comparison
6.6 Comparison of Aerodynamic Forces: Mainsail & Spinnaker
Figure 6.40 shows the force and the moment coefficients obtained from the full-scale
measurements and the wind tunnel tests with mainsail and spinnaker. Point-symmetric full-
scale measurements on starboard tack are mirrored about the origin and plotted as one data
set together with the port measurements.
eff
on port tack is negative so that point-symmetric
data is plotted against -
eff
. Axis-symmetric data is plotted against |
eff
|. Wind tunnel results
for the total model including the hull/deck and results for the rig and sails only are shown.
The difference between the wind tunnel data sets is the hull/deck contribution towards the
forces and moments including the interaction with the sails, which has been discussed in
detail in chapter 3. It can be seen from Figure 6.40(b) that the hull/deck makes a significant
contribution to the side force coefficient at small
eff
, but that it does not greatly affect the
other force and moment coefficients.
The data for rig and sails are equivalent to the full-scale measurements since the rig force
balance on Dyna did not measure the aerodynamic forces on the hull/deck. These data points
are highlighted by joining them with straight lines. The range of the force coefficient plots in
Figure 6.40(a) to Figure 6.40(c) is the same and the amount of scatter in the full-scale data is
similar for all three coefficients. Similarly the range of the moment coefficient plots in Figure
6.40(d) to Figure 6.40(f) is the same. The scatter in the yawing moment coefficient shown in
Figure 6.40(f) is significantly less than for the other two moment coefficients. As a first
general assessment it can be said that the wind tunnel and full-scale measurements clearly
show very similar trends for all force and moment coefficients and that the magnitude of the
coefficients is very comparable; a clear scaling effect cannot be noticed.
The scatter in the full-scale data is probably caused, to a large degree, by differences in the
flying shape of the sails. This is not only caused by differences in sail trim but also due to
higher frequency fluctuations in the wind speed and direction to which the sail trimmer and
the helmsman cannot react but which change the flying shape of the sails relative to the wind
direction. In the controlled test environment of the wind tunnel the sails are trimmed to the
aerodynamically optimum shape, which is not always achieved in real life. It can therefore be
expected that the driving force measured in the wind tunnel forms an upper bound for the
full-scale measurements. Figure 6.40(a) shows that this is the case, where for
eff
larger than
120 only very few full-scale measurements are as high as the wind tunnel measurements.
Between
eff
of 90-110 a few full-scale measurements are significantly higher than the wind
tunnel results and do not form a smooth curve with the other measurements. It could therefore
be argued that these measurements are not reliable.
194
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.6 Comparison of Aerodynamic Forces: Mainsail & Spinnaker
-0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
0 20 40 60 80 100 120 140 160 180
-
eff
[]
C
M
x

[
-
]
Full scale
Total
Rig, main & spi
0
0.5
1
1.5
2
2.5
3
3.5
0 20 40 60 80 100 120 140 160 180
|
eff
| []
C
F
x

[
-
]
Full scale
Total
Rig, main & spi
(a) Drive force coefficient (C (d) Heeling moment coefficient (C
Fx
) )
Mx

0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
0 20 40 60 80 100 120 140 160 180
|
eff
| []
C
M
y

[
-
]
Full scale
Total
Rig, main & spi
-3
-2.5
-2
-1.5
-1
-0.5
0
0.5
0 20 40 60 80 100 120 140 160 180
-
eff
[]
C
F
y

[
-
]
Full scale
Total
Rig, main & spi
(b) Side force coefficient (C (e) Pitching moment coefficient (C
Fy
) )
My

-1
-0.5
0
0.5
1
1.5
2
2.5
3
3.5
0 20 40 60 80 100 120 140 160 180
-
eff
[]
C
M
z

[
-
]
Full scale
Total
Rig, main & spi
-1
-0.5
0
0.5
1
1.5
2
2.5
0 20 40 60 80 100 120 140 160 180
|
eff
| []
C
F
z

[
-
]
Full scale
Total
Rig, main & spi
(c) Force coefficient perpendicular to deck plane (C (f) Yawing moment coefficient (C
Fz
)
Mz
)
Figure 6.40: Comparison of force and moment coefficients from selected full-scale runs and wind tunnel
measurements with mainsail and spinnaker plotted against
eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 195
Chapter 6 Wind Tunnel and Full-Scale Comparison
The pitching moment is mainly caused by driving force so that Figure 6.40(e) has a very
similar trend. To some degree the force perpendicular to the deck plane is also related to C
Fx
,
a properly trimmed spinnaker that is fully inflated has a higher C
Fz
. It is therefore not
surprising that the wind tunnel measurements of C in Figure 6.40
Fz
(c) form an upper bound
for the full-scale measurements.
The side force coefficient (C
Fy
) for a poorly trimmed sail can be higher or lower than for the
optimally trimmed sail so that the full-scale measurements could be expected to fluctuate
about the wind tunnel measurements. For small
eff
and
eff
close to 180 non-optimum trim
of the sails is unlikely to result in |C
Fy
| much larger than the optimum trim and the wind
tunnel measurements should be close to the maximum full-scale data. For
eff
above this is the
case for C
Fy
in Figure 6.40(b) and for the heeling moment coefficient (C
Mx
) in Figure 6.40(d),
which is mainly driven by C
Fy
. For
eff
less than 80 the full-scale C
Fy
is however
significantly smaller than measured in the wind tunnel. C
Mx
for small
eff
on the other hand
shows good agreement between wind tunnel and full-scale measurements. The yawing
moment coefficient (C
Mz
) shown in Figure 6.40(f) has a similar trend to C
Fy
. It agrees well for
larger
eff
but for
eff
less than 100 the full-scale measurements of C
Mz
are much lower. It
should be remembered that all the coefficients are in the body fixed coordinate system
(coordinate system B) as defined in section C.1. The origin is close to the waterline and about
35% of the DWL aft of the forward perpendicular.
Although the full-scale measurements at small
eff
are conducted in very light true wind
speeds the effective wind speed is not significantly lower than at some larger
eff
as shown in
Figure 6.41. It is therefore unlikely that the spinnaker had a less inflated shape for small
eff

that could cause the lower C . As discussed before the full-scale
Fy eff
could be inaccurate due
to mast twist, direction vane misalignment, wind shear or other factors, which would shift the
full-scale coefficients along the y-axis but would only change the coefficient value by a small
amount. Therefore the discrepancy of C at small
Fy eff
cannot be caused simply by an error in

eff
. However, at small
eff
it is possible that the spinnaker sheet or brace may have touched
the railing, which is not part of the rig force balance.
The difference in C for small
Fy eff
is significant and is the source of discrepancies when the
coefficient measurements are processed further. The lift coefficients (C ) shown in
L
Figure
6.42 agree well for
eff
larger than 80 where the wind tunnel measurements form an upper
bound for the full-scale data. For
eff
below 80 the full-scale data are smaller than the wind
tunnel measurements because of the smaller C component. Figure 6.42
Fy
also shows the lift
coefficient calculated from the IMS coefficients for mainsail and spinnaker following the
method by Poor (1986), but assuming no blanketing effect between the sails. For large
eff
the
IMS coefficients fall within the scatter of the full-scale data and are smaller than the wind
tunnel measurements. For small
eff
they lie in between the full-scale and wind tunnel data.
196
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.6 Comparison of Aerodynamic Forces: Mainsail & Spinnaker
0
1
2
3
4
5
6
7
0 20 40 60 80 100 120 140 160 180
|
eff
| []
V
e
f
f

[
m
/
s
]
Full scale
-0.5
0
0.5
1
1.5
2
2.5
3
3.5
0 20 40 60 80 100 120 140 160 180
-
eff
[]
-
C
L

[
-
]
Full scale
Total
Rig, main & spi
IMS coefficients
Figure 6.41: Effective wind speed (V Figure 6.42: Comparison of C
The drag coefficients (C
D
) shown in Figure 6.43 also agree well for larger
eff
but at smaller
angles C
D
is significantly less in full-scale because of the smaller C
Fy
component. The
parasitic drag coefficient can be approximated using equation (2.8) and assuming the same
efficiency in the induced drag calculation as obtained by Hochkirch (2000) for mainsail with
jib. Figure 6.44 shows that C
Dp
follows the same trend as C
D
; for larger
eff
agreement
between wind tunnel and full-scale measurements is good but for small
eff
the full-scale
measurements are much lower. Since C
D
and C
Dp
cannot be negative the full-scale values
seem unrealistically small. Also shown in Figure 6.44 is C
Dp
, calculated according to the IMS
formulation by Poor (1986). It shows a similar trend to the measurements but it is lower than
the wind tunnel measurements and higher than the full-scale data for small
eff
.
eff
) during
selected full-scale runs with mainsail
and spinnaker plotted against
L
from selected full-
scale and wind tunnel runs with
mainsail and spinnaker vs.
eff eff

-0.5
0
0.5
1
1.5
2
2.5
0 20 40 60 80 100 120 140 160 180
|
eff
| []
C
D

[
-
]
Full scale
Total
Rig, main & spi
-0.5
0
0.5
1
1.5
2
2.5
0 20 40 60 80 100 120 140 160 180
|
eff
| []
C
D
p

[
-
]
Full scale
Total
Rig, main & spi
IMS coefficients
Figure 6.43: Comparison of C Figure 6.44: Comparison of C
D
from selected full-
scale and wind tunnel runs with
mainsail and spinnaker vs.
Dp
from selected full-
scale and wind tunnel runs with
mainsail and spinnaker vs.
eff eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 197
Chapter 6 Wind Tunnel and Full-Scale Comparison
Looking at the lift to drag ratio in Figure 6.45 eliminates any differences in measured force
magnitudes between full-scale and wind tunnel tests. The lift to drag ratio is much greater for
the full-scale measurements at small
eff
because of the big difference in C
D
, which is caused
experimentally by the low side force measurements in full-scale. For some measurements the
full-scale lift to drag ratio is twice the ratio obtained in the wind tunnel. The lift to drag ratio
predicted by the IMS formulation is higher than the wind tunnel results at small
eff
but fits
the wind tunnel data much better than the full-scale measurements. It needs to be remembered
that
eff
below 40 represent unrealistic sailing conditions under spinnaker. The IMS data is
calculated from coefficients for each sail shown in Figure 6.46 The spinnaker does not
produce any lift for
eff
less than 30 because it collapses, but due to the mainsail C
L
, the lift
to drag ratio in Figure 6.45 only reduces for
eff
less than 18.
Figure 6.47 and Figure 6.48 show the centre of effort height (z
CoE
) calculated from the central
axis approach described in section 2.3.3. Figure 6.47 shows z
CoE
plotted against
eff
in the
usual way but due to the asymptotic behaviour when the resultant force direction () is in line
with the yacht centreline, the plot becomes easier to interpret when z
CoE
is plotted against as
shown in Figure 6.48. For small (which corresponds to large
eff
) z
CoE
from wind tunnel and
full-scale measurements agree well, but for large (small
eff
) z
CoE
from the full-scale
measurements increases while z
CoE
from the wind tunnel tests stays fairly constant around the
height of the geometric centre of the mainsail. At large (small
eff
) z
CoE
is largely driven by
the side force and the heeling moment coefficients (C and C
Fy Mx
), so that the reduced C
Fy
for
the full-scale measurements at similar C
Mx
results in an increased z
CoE
.
Figure 6.49 and Figure 6.50 show the longitudinal centre of effort position (x
CoE
) calculated
from the central axis approach described in section 2.3.3. As before, x
CoE
plotted against in
Figure 6.50 shows the asymptotic behaviour more clearly. The wind tunnel and full-scale data
-1
0
1
2
3
4
5
6
7
8
0 20 40 60 80 100 120 140 160 180
-
eff
[]
-
C
L
/
C
D

[
-
]
Full scale
Total
Rig, main & spi
IMS coefficients
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 20 40 60 80 100 120 140 160 180

eff
[]
CL main [-]
CL jib [-]
CL spi [-]
CDp main [-]
CDp jib [-]
CDp spi [-]
Figure 6.45: Comparison of C /C Figure 6.46: IMS C and C
L D
from selected
full-scale and wind tunnel runs with
mainsail and spinnaker vs.
L Dp
for individual sails
based on cloth sail area plotted
against
eff eff
198
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.6 Comparison of Aerodynamic Forces: Mainsail & Spinnaker
-2
0
2
4
6
8
10
12
14
-20 -10 0 10 20 30 40 50 60 70 80 90
- []
z
C
o
E

[
m
]
Full scale
Total
Rig, main & spi
Geometric centre main & jib
Geometric centre main
Geometric centre jib
-2
0
2
4
6
8
10
12
14
0 20 40 60 80 100 120 140 160 180
|
eff
| []
z
C
o
E

[
m
]
Full scale
Total
Rig, main & spi
Geometric centre main & jib
Geometric centre main
Geometric centre jib
Figure 6.47: Comparison of z Figure 6.48: Comparison of z
exhibit the same trend of x
CoE
moving aft with decreasing (increasing
eff
), however the
wind tunnel x
CoE
is generally further aft. For large (small
eff
) x
CoE
in the wind tunnel is
close to the longitudinal position of the geometric centre of mainsail and jib whereas the full-
scale x
CoE
is closer to the longitudinal position of the geometric centre of the jib
26
. The offset
is mainly caused by the difference in yawing moment shown in Figure 6.40(f). Before
assessing the differences between the wind tunnel and full-scale data further the comparison
between the upwind data with mainsail and jib needs to be considered to see if similar trends
exist.

26
The geometric centre of the spinnaker cannot be defined in a convenient way because it is very dependent on

eff
and the trim so that it is not shown.
CoE
from selected full-
scale and wind tunnel runs with
mainsail and spinnaker vs.
CoE
from selected full-
scale and wind tunnel runs with
mainsail and spinnaker vs.
eff
-20
-15
-10
-5
0
5
10
15
20
0 20 40 60 80 100 120 140 160 180
|
eff
|[]
x
C
o
E

[
m
]
Full scale
Total
Rig, main & spi
Geometric centre main & jib
Geometric centre main
Geometric centre jib
-20
-15
-10
-5
0
5
10
15
20
-20 -10 0 10 20 30 40 50 60 70 80 90
- []
x
C
o
E

[
m
]
Full scale
Total
Rig, main & spi
Geometric centre main & jib
Geometric centre main
Geometric centre jib
Figure 6.49: Comparison of x Figure 6.50: Comparison of x
CoE
from selected full-
scale and wind tunnel runs with
mainsail and spinnaker vs.
CoE
from selected full-
scale and wind tunnel runs with
mainsail and spinnaker vs.
eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 199
Chapter 6 Wind Tunnel and Full-Scale Comparison
6.7 Comparison of Aerodynamic Forces: Mainsail & Jib
Figure 6.51 shows the force and the moment coefficients obtained for mainsail and jib from
the wind tunnel tests and the full-scale measurements in V
T
less than 4m/s. As described in
section 6.6 point-symmetric full-scale measurement on starboard tack are mirrored about the
origin and plotted as one data set together with the port measurements. The number of full-
scale measurements is not sufficient at larger
eff
to define a clear trend, and the accuracy is
uncertain due to the small V . Because of the small V
T T
it can however be assumed that the
sails were fully powered-up. Depowering only needs to be considered for the measurements
in higher V
T
. Wind tunnel results for the total model including the hull/deck and results for
only the rig and sails are shown for three different heel angles.
From the wind tunnel measurements of side force and heeling moment coefficient in Figure
6.51(b) and Figure 6.51(d) a clear trend with heel angle can be seen although the effective
angle theory has been used to calculate the coefficients as described in section 2.3.1. The
coefficient magnitude is reduced when the model is heeled which shows that using the
effective angle approach in VPPs to account for heel has shortfalls, and that using alternative
methods like the Real-Time VPP with a dynamically heeling model can improve the accuracy
of performance predictions. As discussed in detail in chapter 3, Figure 6.51(b) shows again
that a significant amount of the side force is generated by the hull/deck. Accounting for the
fact that the hull/deck component should not be included when comparing against the full-
scale measurements improves the similarity. The wind tunnel side force and heeling moment
coefficients at 25 heel form an upper bound for the full-scale measurements. However,
Figure 6.35 and VPP predictions show that heel angles above approximately 13 are not to be
expected in V
T
up to 4m/s so that the full/scale data should be compared against the wind
tunnel tests at 12.5 heel, which are highlighted by joining the measurement points with
straight lines. Compared against the wind tunnel data at 12.5 heel, the full-scale side force
and heeling moment coefficient magnitudes are smaller for small
eff
.
The drive force and pitching moment coefficients in Figure 6.51(a) and Figure 6.51(e) agree
well for small
eff
when sailing hard on the wind but the wind tunnel measurements are higher
for
eff
above 30 and 25 respectively. In the wind tunnel it has been observed that trimming
the sails at
eff
of 40 and 60 is particularly complex since more than one stable flow regime
can be achieved for the same nominal sail trim
27
and small changes in the surrounding flow
field can trigger a shift from one regime to the other.

27
The same nominal sail trim means that nothing on the model or the sail controls is adjusted. The flying shape
of the sail can change however due to changes in the surrounding pressure field.
200
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.7 Comparison of Aerodynamic Forces: Mainsail & Jib
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 10 20 30 40 50 60 70 80 90
-
eff
[]
C
M
x

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 10 20 30 40 50 60 70 80 90
|
eff
| []
C
F
x

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
(a) Drive force coefficient (C (d) Heeling moment coefficient (C
Fx
) )
Mx

0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 10 20 30 40 50 60 70 80 90
|
eff
| []
C
M
y

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
-2
-1.8
-1.6
-1.4
-1.2
-1
-0.8
-0.6
-0.4
-0.2
0
0 10 20 30 40 50 60 70 80 90
-
eff
[]
C
F
y

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
(b) Side force coefficient (C (e) Pitching moment coefficient (C
Fy
) )
My

0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 10 20 30 40 50 60 70 80 90
-
eff
[]
C
M
z

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 10 20 30 40 50 60 70 80 90
|
eff
| []
C
F
z

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
(c) Force coefficient perpendicular to deck plane (C (f) Yawing moment coefficient (C
Fz
)
Mz
)
Figure 6.51: Comparison of force and moment coefficients from selected full-scale runs with V
T
<4m/s
and wind tunnel measurements with mainsail and jib plotted against
eff

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 201
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 10 20 30 40 50 60 70 80 90
-
eff
[]
-
C
L

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
IMS coefficients

Figure 6.52: Comparison of C
The yawing moment coefficient in Figure 6.51(f) is slightly larger in the wind tunnel but the
trends agree very well. Figure 6.51(c) shows the force coefficients perpendicular to deck
plane (C
Fz
) which agree well if the contribution of the hull/deck is not included. It is shown,
as was investigated in chapter 3, that the sails and rig do not produce much C
Fz
and that most
C
Fz
results from suction on the deck for the model at heel.
As for mainsail and spinnaker in section 6.6, the full-scale side force coefficient is also
smaller for mainsail and jib at small
eff
, which affects the data derived from the
measurements. The lift coefficient (C
L
) shown in Figure 6.52 is smaller for most full-scale
measurements. The magnitude and trend of the wind tunnel C
L
excluding hull/deck and the
IMS C
L
agree well but the IMS curve is shifted to the left. In the IMS formulation the lift
generated by the hull/deck is not considered separately so that the total C
L
is smaller than in
the wind tunnel. Similarly the drag coefficient (C
D
) from the full-scale measurements in
Figure 6.53 is also less than in the wind tunnel at small
eff
due to the low side force
measurements from which it is calculated. The parasitic drag coefficient (C
Dp
) is again
calculated by rearranging equation (2.8) and using the efficiency obtained by Hochkirch
(2000) for mainsail with jib in the induced drag calculations. As in section 6.6 the resulting
C
Dp
presented in Figure 6.54 is unrealistically small and even becomes negative which is
physically impossible. The IMS C
Dp
agrees well with the wind tunnel data for small
eff
but
increases less rapidly compared to the wind tunnel and full-scale measurements at larger
eff
.
Having compared the light wind full-scale measurements to the wind tunnel tests, the
moderate wind full-scale measurements of V
T
between 6-8m/s are now added to the
comparison in Figure 6.55. The moderate wind data is potentially more accurate due to the
larger forces measured and a larger number of data points were collected at larger
eff
so that
L
from full-scale runs
with V

T
<4m/s and wind tunnel runs
with mainsail and jib vs.
eff
202
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.7 Comparison of Aerodynamic Forces: Mainsail & Jib
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 10 20 30 40 50 60 70 80 90
|
eff
| []
C
D
p

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
IMS coefficients
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 10 20 30 40 50 60 70 80 90
|
eff
| []
C
D

[
-
]
Full-scale VT<4m/s
Total 0 heel
Total 12.5 heel
Total 25 heel
Rig, main & jib 0 heel
Rig, main & jib 12.5 heel
Rig, main & jib 25 heel
Figure 6.53: Comparison of C Figure 6.54: Comparison of C
trends are shown more clearly. From the VPP study in chapter 5 it is expected that the sails
are somewhat depowered at small
eff
in moderate wind speeds of V
T
between 6-8m/s, so that
a comparison to the fully powered-up wind tunnel measurements is not accurate. Therefore
Figure 6.55 does not only show the fully powered-up wind tunnel measurements obtained for
this study but also the predicted depowered values calculated using the power model
developed in section 5.4.
Three true wind speeds are considered, V
T
<4m/s, V
T
=6m/s and V
T
=8m/s. For each V
T
and
each tested
A
the fully powered-up wind tunnel measurement (excluding the hull/deck
component) at the appropriate heel angle is selected and multiplied by the power ratio
expression. The form of the power ratio expression is different for each coefficient as
discussed in section 5.4 but is always dependent on the trim parameter power. For this study
the generalised relationships introduced in section 5.4.2 are employed, which are only a
function of power as shown for two examples in Figure 5.33 and Figure 5.34. The values of
the trim parameter power are obtained from the VPP results and are shown in Figure 5.27.
For V
T
<4m/s power is equal to one for all
eff
since the sails should not be depowered. For
V
T
=6m/s power is less than one for
eff
smaller than about 50, and for V
T
=8m/s the sails are
depowered for
eff
less than approximately 70. It should be noted that the power ratio
expressions and the values of power were obtained from the total forces acting on the model
including the hull/deck component, whereas they are used here to adjust the coefficients
excluding the hull/deck component. This assumes that the hull/deck forces change in the
same way as the sail forces when depowering.

D
from full-scale
runs with V
Dp
from full-scale
runs with V
T
<4m/s and wind tunnel
runs with mainsail and jib vs.
T
<4m/s and wind tunnel
runs with mainsail and jib vs.
eff eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 203
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 20 40 60 80 100 120 140
-
eff
[]
C
M
x

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 20 40 60 80 100 120 140
|
eff
| []
C
F
x

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
(a) Drive force coefficient (C (d) Heeling moment coefficient (C
Fx
) )
Mx

-1.8
-1.6
-1.4
-1.2
-1
-0.8
-0.6
-0.4
-0.2
0
0 20 40 60 80 100 120 140
-
eff
[]
C
F
y

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 20 40 60 80 100 120 140
|
eff
| []
C
M
y

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
(b) Side force coefficient (C (e) Pitching moment coefficient (C
Fy
) )
My

-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
0 20 40 60 80 100 120 140
|
eff
| []
C
F
z

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 20 40 60 80 100 120 140
-
eff
[]
C
M
z

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
(c) Force coefficient perpendicular to deck plane (C (f) Yawing moment coefficient (C
Fz
)
Mz
)
Figure 6.55: Comparison of force and moment coefficients from selected full-scale runs and wind tunnel
measurements with mainsail and jib plotted against
eff
204
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.7 Comparison of Aerodynamic Forces: Mainsail & Jib
Chapter 3 showed a strong interaction between the hull/deck and the sails so that this
assumption is more appropriate than treating the hull/deck forces as independent of sail trim.
Unfortunately it was not feasible to measure the hull/deck forces separately while conducting
the depowering study in chapter 5 but the resulting differences in this study are expected to
be small.
The drive force coefficient in Figure 6.55(a) shows that the full-scale measurements in 6-8m/s
V are smaller than in V <4m/s for small
T T eff
. For larger
eff
the maximum values are not very
different for the two data sets. This agrees well with the wind tunnel measurements which
also show the drive force coefficient reduced in 6 and 8m/s V
T
due to depowering. On
account of their interdependence, the pitching moment coefficient in Figure 6.55(e) shows the
same trends. The side force and heeling moment coefficients in Figure 6.55(b) and Figure
6.55(d) show a significant reduction in magnitude at small
eff
for both full-scale and wind
tunnel data. It needs to be remembered that the side force is negative and that a reduction in
magnitude makes C
Fy
less negative. The largest magnitudes of the full-scale measurements in
V between 6-8m/s are still smaller than the wind tunnel prediction for V
T T
of 8m/s, but many
full-scale measurements in 6-8m/s V fall in between the wind tunnel predictions for V
T T
of 6
and 8m/s. A trend of slightly reduced force coefficient perpendicular to the deck plane with
increasing V at small
T eff
can be seen in Figure 6.55(c) from both full-scale and wind tunnel
measurements. Similarly the yawing moment in Figure 6.55(f) is reduced with increasing V
T

for small
eff
in the wind tunnel whereas this trend is not clearly noticeable in the full-scale
data.
The lift coefficient in Figure 6.56 shows a number of trends discussed before. The wind
tunnel and full-scale measurements agree well for larger
eff
whereas the full-scale lift
coefficient is less at small
eff
, particularly near
eff
of maximum lift. The level of depowering
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 20 40 60 80 100 120 140
-
eff
[]
-
C
L

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s

Figure 6.56: Comparison of C
L
from full-scale and
wind tunnel runs with mainsail and
jib vs.

eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 205
Chapter 6 Wind Tunnel and Full-Scale Comparison
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 20 40 60 80 100 120 140
|
eff
| []
C
D
p

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 20 40 60 80 100 120 140
|
eff
| []
C
D

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
Figure 6.57: Comparison of C Figure 6.58: Comparison of C
seems to be very similar in the wind tunnel and full-scale. The drag coefficient plotted in
Figure 6.57 is again lower for the full-scale measurements at the smallest
eff
but the same
trend in depowering is seen between full-scale and wind tunnel data. Similarly the full-scale
parasitic drag coefficient is unrealistically small in Figure 6.58, but both the full-scale and
wind tunnel results agree that there is little change in parasitic drag due to depowering at
small
eff
.
Figure 6.59 shows the centre of effort height (z
CoE
) defined as the intersection of the resultant
force in the central axis system with the centreline plane of the yacht (y-plane) as described in
section 2.3.3. The fully powered-up wind tunnel and full-scale results show z
CoE
close to the
geometric centre of sail area. The wind tunnel measurements show an increase in z
CoE
at
small
eff
which cannot be seen in the full-scale measurements. This indicates that the sails in
full-scale may have been trimmed with more twist even in light winds, or that the top of the
mainsail generated less lift in light wind at small
eff
for some other reason. The medium
wind strength results from full-scale and wind tunnel measurements agree very well. The z
CoE

is reduced by a very similar amount. In V
T
of 8m/s the height is reduced by up to 22%. The
fully powered-up z
CoE
and the geometric centre of sail area are about 42% of the mast height
above the water. In V
T
of 8m/s the z
CoE
is therefore as low as 33% of the mast height. The
longitudinal centre of effort position (x
CoE
) moves aft with increasing
eff
as shown in Figure
6.60 by both the wind tunnel and full-scale measurements. In the wind tunnel the x
CoE
moves
slightly forward when the sails are depowered since the mainsail is predominately adjusted
during this trimming process, which reduces its side force contribution and moves the x
CoE

towards the jib. This relatively small shift cannot clearly be identified in the full-scale data
due to the amount of scatter.
D
from full-scale and
wind tunnel runs with mainsail and
jib vs.
Dp
from full-scale
and wind tunnel runs with mainsail
and jib vs.
eff eff
206
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.7 Comparison of Aerodynamic Forces: Mainsail & Jib
-4
-3.5
-3
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
0 20 40 60 80 100 120 140
|
eff
| []
x
C
o
E

[
m
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
Geometric centre main & jib
Geometric centre main
Geometric centre jib
0
1
2
3
4
5
6
7
8
9
0 20 40 60 80 100 120 140
|
eff
| []
z
C
o
E

[
m
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
Geometric centre main & jib
Geometric centre main
Geometric centre jib
Figure 6.59: Comparison of z Figure 6.60: Comparison of x
One of the main discrepancies between the full-scale and wind tunnel data is again in the side
force coefficient at small
eff
. As for mainsail and spinnaker in section 6.6 the side force
coefficient is less for the full-scale measurements for mainsail with jib as well, which again
results in a too low C
L
and C
D
at small
eff
. Plotting the lift to drag ratio in Figure 6.61 shows
that it is much larger for the full-scale measurement compared to the wind tunnel data and
IMS predictions, which agree well. The graph is very similar in trend to the mainsail with
spinnaker case in Figure 6.45. The full-scale lift to drag ratios seem unrealistically high and a
basic sample calculation for
eff
of 20 also shows this. The fully powered-up C
L
at
eff
of 20
in Figure 6.56 is 1.47 and C
L
/C
Di
can be calculated using equation (2.6) as
C
L
/C
Di
=eAR/C
L
=10. Also including the separation drag prediction from equation (2.5) of
C
Ds
=0.0016C
L
2
gives C
L
/(C
Di
+C
Ds
)=8.1. The parasitic drag coefficient can be approximated
based on the cross sectional area of the mast and rigging assuming C
D
of 1.2 for circular
sections and 0.8 for mast and forestay. C
Dp
of the sails is taken as the skin friction (C
Df
) on a
flat plate in turbulent flow given by Abbott and von Doenhoff (1959) as

( )
58 . 2
Re log
472 . 0
=
Df
C ,
(6.20)
where Re is the Reynolds Number defined in equation (2.43). The lift to drag ratio from
C
L
/(C
Di
+C
Ds
+C
Dp
) then becomes 6.1, which is very similar to the lift to drag ratios achieved
in the wind tunnel and predicted by the IMS formulation. The full-scale ratios on the other
hand are much larger. This seems to be a systematic error as it has been observed for both
sailing configurations. It is very unlikely that sails in full-scale produce much greater lift to
drag ratios than predicted using standard aerodynamic principles, since extensive research has
been carried out in aeronautics and the principles of wing theory are well understood. The
CoE
from full-scale
and wind tunnel runs with mainsail
and jib vs.
CoE
from full-scale
and wind tunnel runs with mainsail
and jib vs.
eff eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 207
Chapter 6 Wind Tunnel and Full-Scale Comparison
0
2
4
6
8
10
12
14
16
0 20 40 60 80 100 120 140
-
eff
[]
-
C
L
/
C
D

[
-
]
Full-scale VT<4m/s
Full-scale VT=6-8m/s
Rig, main & jib VT<4m/s
Rig, main & jib VT=6m/s
Rig, main & jib VT=8m/s
IMS coefficients
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
0 20 40 60 80 100 120 140 160
|
T
| []
V
S

[
m
/
s
]
Full-scale VT=3-4m/s
Full-scale VT=6-8m/s
FS-Equi VT=3m/s
FS-Equi VT=4m/s
FS-Equi VT=6m/s
FS-Equi VT=8m/s
Figure 6.61: Comparison of C /C Figure 6.62: Boat speed (V
parasitic drag coefficient has a great influence on C
L
/C
D
as it is the only component not
dependent on C
L
. The value of C
Dp
influences the C
L
value for which the maximum C
L
/C
D

occurs. If C
Dp
is under-predicted the maximum C
L
/C
D
occurs at a smaller C
L
. For the full-
scale data C
L
/C
D
therefore increases when C
L
is reduced due to depowering or deficient trim.
This combination of low C
Dp
measurements and reduced C
L
causes the unrealistically high
C
L
/C
D
. The low C
Dp
could be caused by an error in
eff
due to, for example, errors in the mast
twist measurements or the
eff
approximation. An offset to increase
eff
makes C
Dp
and
C
L
/C
D
more realistic but the trends of other variables such as C
Fx
and C
Fy
agree less well with
the wind tunnel data. Another possible reason for the low C
Dp
is the low measurement of C
Fy

relative to C
Fx
for small
eff
, due to, for example, inaccuracies in the calibration matrix for
these load cases. In addition, for small
eff
, inaccuracies in the force measurements have a
greater influence on the resulting C
L
and C
D
since the drive force is very small compared to
the much larger side force. It is difficult to measure a small drive force accurately when a
much larger side force and a significant heeling moment are present due to cross talk between
the force transducers of the multi-component force balance. Unfortunately from the available
full-scale data it is not possible to determine with certainty what caused the unrealistically
small C
Dp
at small
eff
. In spite of this likely problem, the wind tunnel and full-scale data
agree very well in general. As far as possible with the scatter in the full-scale data, no clear
scaling effect can be seen between the wind tunnel and full-scale measurements, and the same
trends are observed in the different data sets.
Finally Figure 6.62 shows the boat speed measured on Dyna and predicted by FRIENDSHIP-
Equilibrium using the wind tunnel sail force coefficients and the generalised power model
presented in section 5.4.2 for mainsail with jib plotted against the true wind angle (
T
). The
VPP results are shown up to a
T
that corresponds to
A
of 90, the largest angle measured in
L D
from full-scale
and wind tunnel runs with mainsail
and jib vs.
s
) of Dyna with mainsail
and jib and from VPP using wind
tunnel data and power model vs.
eff eff
208
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.7 Comparison of Aerodynamic Forces: Mainsail & Jib
the wind tunnel. Some full-scale boat speeds smaller than the VPP predictions are expected
because the sails were not trimmed adequately for every full-scale measurement and because
other effects like waves can slow down the yacht. The top boat speeds agree well for small

eff
but are over-predicted by the VPP at larger
eff
. This could be due to the upwind sails in
full-scale not operating as efficiently as in the wind tunnel at larger
eff
or deficiencies in the
hydrodynamic model. Figure 6.62 does not say much about the accuracy of wind tunnel
testing but shows to what degree the yacht performance can be predicted by using
aerodynamic coefficients obtained through wind tunnel testing together with a hydrodynamic
model customised to the yacht by means of CFD simulations, towing tank testing, and full-
scale measurements.
6.7.1 Comparison of Sail Shapes
The data presented in section 6.7 and the VPP predictions in chapter 5 strongly suggest that
the sails were to some degree depowered during full-scale measurements in true wind speeds
between 6-8m/s. This can be confirmed by conducting a basic sail shape analysis on the
mainsail of Dyna and in the wind tunnel for different wind strength. Krebber (2005) and
Krebber and Hochkirch (2006) conducted a basic preliminary sail shape study of the Dyna
Sail camber [%]
Sail twist []
Boom angle []
Figure 6.63: Screen shot of sail shape analysis program SailTool showing the parameters sail twist, boom
angle and sail camber obtained from wind tunnel picture
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 209
Chapter 6 Wind Tunnel and Full-Scale Comparison
mainsail and jib for 37 close hauled measurements to obtain input parameters for their
numerical investigation on depowering. The full-scale pictures were taken in 2003 but
unfortunately no rig force measurements were conducted simultaneously so that the sail
shapes cannot be linked directly to the sail forces. It was therefore decided that recreating the
full-scale sail shapes in the wind tunnel is of limited use and should not be attempted because
of the considerable effort required to realise it, but that analysing the already obtained wind
tunnel sail trims and comparing them to the full-scale sail shapes is more appropriate. From a
top view picture of the sail three parameters were measured. At a height up the sail luff the
sail camber and twist angle, which is the angle between the chord and the boom, are
measured together with the boom angle. Figure 6.63 shows how sail camber, sail twist and
boom angle are obtained in the same way from a wind tunnel picture using the sail shape
analysis program SailTool, developed at Curtin University of Technology
28
. The standard top
view picture taken in the wind tunnel does not show the necessary details of the jib so that
only the mainsail shapes are compared here. Even on the mainsail the camber is hard to
identify but for this simple study to show depowering trends the accuracy is sufficient. Figure
6.64 shows the range of V and
T eff
for the sail shapes analysed by Krebber (2005) divided
into two data sets of
eff
smaller and larger than 22.5. To cover the range of full-scale
conditions, pictures taken during the depowering study with the Real-Time VPP in chapter 5
at
eff
of 20, 25 and 30 for V
T
of 4,6,8 and 10m/s are analysed to obtain the sail shapes in
the wind tunnel.
Figure 6.65 shows the sail twist from full-scale and wind tunnel pictures plotted against V
T
. In
both full-scale and in the wind tunnel the sail twist increases with wind strength. The wind

28
Centre of Marine Science and Technology, Curtin University of Technology, GPO Box U1987, Perth WA
6845, Australia
0
5
10
15
20
25
30
0 2 4 6 8 10 1
V
T
[m/s]
|

e
f
f
|

[

]

2
0
2
4
6
8
10
12
14
16
18
0 2 4 6 8 10 12
V
T
[m/s]
S
a
i
l

t
w
i
s
t

[

]

Full-scale EWA 18.7-22.5
Full-scale EWA 22.5-27.9
RT-VPP EWA 20
RT-VPP EWA 25
RT-VPP EWA 30
Full-scale EWA 18.7-22.5
Full-scale EWA 22.5-27.9
Figure 6.65: Mainsail twist vs. V Figure 6.64: Range of
eff
vs. V
T
for analysed full-
scale sail shape pictures
T
on Dyna and in
wind tunnel with Real-Time VPP for
close hauled
eff
210
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.7 Comparison of Aerodynamic Forces: Mainsail & Jib
tunnel data shows less sail twist at
eff
of 20 compared to the larger
eff
since C
L
is smaller at

eff
of 20 and less depowering is required. The full-scale data has more scatter but a similar
trend can still be seen to some degree. Sail twist is the dominant sail alteration when
depowering the mainsail and Figure 6.65 clearly confirms that the sails have been depowered
on Dyna when sailing in moderate wind of V between 6-8m/s.
T
Figure 6.66 shows the boom angle plotted against V
T
for full-scale and wind tunnel
measurements. The wind tunnel data show clear trends of increasing boom angle with
eff
and
V
T
. As discussed in chapter 5 the sails are fully powered-up in the wind tunnel at V
T
of 4m/s
so that the sail shape at V
T
of 2m/s would be the same within the measuring tolerance. For the
fully powered-up mainsail at the smallest
eff
the boom is almost in the centreline (as can be
expected) and slightly moves to leeward as
eff
is increased. For the full-scale data the scatter
is significant and a clear trend cannot be seen but in general the boom angle increases with
V . For low V the mainsail camber shown in Figure 6.67
T T
is very similar for the full-scale and
wind tunnel measurements at around 11-13%. The full-scale camber remains constant for
higher V
T
whereas the mainsail camber in the wind tunnel reduces. The reduction is unlikely
due to measurement inaccuracies, since the same clear trend is shown for all three
eff
. The
sail camber is most likely not adjusted directly during the measurements but is influenced by
other sail control adjustments. More extensive use of the backstay in the wind tunnel could
for example cause a flatter sail. Sail stretch could also contribute to the difference in trend
since the wind tunnel flow speed is close to constant for all measurements as only the V
T
is
the Real-Time VPP is adjusted, and sail stretch is assumed to be small, whereas the pressure
on sails increases with V
T
in full-scale and sail stretch is not negligible. This simple study is
not designed to investigate this further but Figure 6.66 and Figure 6.67 still show that
comparable sail shapes have been achieved in full-scale and in the wind tunnel. Figure 6.65
-2
0
2
4
6
8
10
0 2 4 6 8 10 1
V
T
[m/s]
B
o
o
m

a
n
g
l
e

[

]

2
0
2
4
6
8
10
12
14
0 2 4 6 8 10
V
T
[m/s]
S
a
i
l

c
a
m
b
e
r

[
%
]

Full-scale EWA 18.7-22.5
Full-scale EWA 22.5-27.9
RT-VPP EWA 20
RT-VPP EWA 25
RT-VPP EWA 30
12
Full-scale EWA 18.7-22.5
Full-scale EWA 22.5-27.9
RT-VPP EWA 20
RT-VPP EWA 25
RT-VPP EWA 30
Figure 6.66: Boom angle vs. V Figure 6.67: Mainsail camber vs. V
T
on Dyna and in
wind tunnel with Real-Time VPP for
close hauled
T
on Dyna and
in wind tunnel with Real-Time VPP
for close hauled
eff eff
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 211
Chapter 6 Wind Tunnel and Full-Scale Comparison
confirms that the sails are depowered similarly in both cases as was discussed in section 6.7
based on the force measurements.
6.8 Conclusions
Both full-scale and wind tunnel testing have challenges associated with them. In the wind
tunnel care must be taken to model the flow correctly and good trimming of the sails is
crucial. The full-scale measurements on the other hand are conducted in an unsteady
environment and it is difficult to measure all relevant data accurately. As a result the scatter
in the full-scale data is significant. For sailing conditions close to the wind
eff
has a great
influence on the resulting C and C
L D
and factors like direction vane misalignment, mast twist
and flow deflection due to the presence of the sails, are accounted for in the analysis. On days
with light winds, the full-scale measurements are still not symmetrical on both tacks. This
could possibly be due to wind shear and this highlights that little is known about the velocity
profiles close to the water surface for different wind environments. The asymmetry in the
data is corrected although the cause cannot be determined with certainty. The flow deflection
due to the presence of the sails not only affects
eff
but also changes the wind speed measured
by the cup anemometer, particularly under spinnaker, and is corrected. The inclination angle
of the flow relative to the rotation axis of the cup anemometer can be up to about 25 in
certain sailing conditions but it was found not to significantly influence the measurements.
The wind tunnel tests showed that much more repeatable measurements can be obtained due
to the controlled test environment. Careful set-up of the wind tunnel and model is important
and good trimming of the sails is essential to produce meaningful results. C and C
L D
show a
noticeable reduction with increasing heel angle although effective angle theory is used in the
calculations. Heeling should therefore generally be considered during wind tunnel testing of
sails by, for example, employing the Real-Time VPP. The sails themselves do not produce a
significant force perpendicular to the deck plane but due to the interaction with the hull/deck
a force, which increases with heel, is generated by suction over the deck.
In the comparison between the full-scale and wind tunnel data the interaction between the
sails and hull/deck needs to be considered. The hull/deck creates a significant amount of lift
and drag due to the interaction with the sails, which is not included in the full-scale
measurements and needs to be subtracted from the standard wind tunnel measurements. The
full-scale forces and a basic sail shape analysis showed that the sails are depowered when
sailing upwind in moderate wind and the wind tunnel measurements need to allow for this. In
addition not all full-scale sail trims were optimal for the given conditions which further
contributes to the large amount of scatter in the full-scale data and makes the comparison
more difficult. Nonetheless all force and moment coefficients show the same trends and no
212
YACHT
RESEARCH
UNIT
The
University
of Auckland
6.8 Conclusions
obvious scaling effect can be noticed. The most noticeable difference is the low measurement
of drag for small
eff
in the full-scale data for both upwind and downwind configurations.
Possible reasons are an error in
eff
or the low measurement of the side force relative to the
drive force. All justifiable corrections have been applied to
eff
and no obvious problems with
the side force measurements have been detected so that the origin of the error has not been
determined.
From this comparison wind tunnel testing is shown to be a good quantitative tool to predict
full-scale aerodynamic forces. Although no scaling effects were observed here more
comparisons of other sails and yacht types are required to exclude the possibly of scaling
effects occurring. Questions also remain over the full-scale velocity profiles which influence
the coefficient magnitudes. Therefore, more full-scale research is required before these issues
can be addressed in the wind tunnel. Ultimately it is desirable to measure the sail shape
simultaneously with the sail forces in full-scale and in the wind tunnel to make a more direct
comparison. Until this is realised the Real-Time VPP provides a good platform to achieve
realistic sail shapes for different sailing conditions in the wind tunnel. Since all the trends in
the forces and moments are very similar in full-scale and the wind tunnel, this comparison
confirms that wind tunnel testing is a reliable qualitative tool for replicating full-scale sail
behaviour.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 213

The
University
YACHT
RESEARCH
of Auckland UNIT
7 Conclusions and Recommendations
7.1 Conclusions
In this work a number of existing and novel wind tunnel techniques have been used to
enhance the wind tunnel testing process and the aerodynamic force models for yacht sails
with the main aim to compare wind tunnel to full-scale measurements for the first time.
The interaction of the hull/deck with the sails has been investigated by conducting force and
pressure measurements on the hull/deck. The hull/deck forces are significantly affected by the
presence of the sails and bare poles windage measurements do not represent the forces which
act on the hull/deck when sailing. The hull/deck side force increases by a factor of up to 2.5
due to the presence of the sails and is up to 14% of the total side force. The accurate
modelling of the hull/deck and the inclusion of its forces is therefore important. The force
perpendicular to the deck plane is also investigated since it is currently not modelled in VPPs.
The hull/deck component is significantly affected by the presence of the sails and at large
heel angles it contributes more than half of the total force perpendicular to the deck plane. At
heel this force increases the side force, which can be noticed in VPP calculations as a small
reduction in boat speed. The findings of this study have improved the physical description of
the forces acting on a yacht and allow a better comparison of full-scale, wind tunnel and
computational fluid dynamics data.
A new wind tunnel testing system, the Real-Time VPP, has been developed which makes the
trimming of model sails more similar to the real life situation on the water. Traditionally the
model sails in the wind tunnel are trimmed based on the measured forces but with the Real-
Time VPP they can now be trimmed based on the performance of the full-scale yacht. The
Real-Time VPP solution also provides the full-scale heel angle and a system has been created
to dynamically heel the wind tunnel model to the correct heel angle while the sails are being
trimmed. The Real-Time VPP is implemented as an additional feature of FRIENDSHIP-
Equilibrium which results in a flexible application package for the performance prediction of
sailing yachts. The Real-Time VPP is an efficient tool for the comparative testing of sails
because it makes trimming in the wind tunnel more realistic and allows different sail designs
to be compared directly without the need of post-processing the data. In addition it is a useful
research tool to determine realistically depowered sail trims in a more direct way than was
previously possible.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 215
Chapter 7 Conclusions and Recommendations
Realistically depowered sail trims obtained with the Real-Time VPP have been used to
investigate the depowering of sails. An empirical power model has been developed to model
the forces of the realistically depowered sails by response surfaces which are functions of the
power parameter and the effective wind angle. This describes the depowering in more detail
compared to the standard trim parameter model, which is particularly important for reaching
and running sails of high-performance yachts. It was confirmed that the most important
parameters for which changes due to depowering need to be modelled are lift, drag and the
centre of effort height. The smoothest and least complex depowering surfaces with little
dependency on effective wind angle were obtained when modelling the force and moment
coefficients directly, instead of using the more common method of modelling the lift and drag
coefficient and the centre of effort position. For an empirical approach on depowering based
on Real-Time VPP measurements this is a good alternative since it results in a much simpler
and more elegant model where the data fitting is much easier.
Wind tunnel testing is however not always feasible and generic trim parameter models based
on aerodynamic principles are still important. The changes in lift, drag and centre of effort
height predicted by the traditional reef, flat and twist model are very different to the power
model predictions based on the wind tunnel tests. Three improvements to the flat and twist
implementation are developed based on aerodynamic principles and the enhanced model now
accounts for the facts that the fully powered-up sails may not have a semi-elliptic loading
distribution; the so-called parasitic drag is dependent on the reduction in lift squared; and
twisting off the sails is associated with a reduction in lift. With these enhancements the match
with the power model results is much better for the investigated effective wind angle range
between 20 and 90. The predicted boat speed is now within 1% compared to the under-
prediction by up to 4% of the original trim parameter model.
Both full-scale and wind tunnel testing have various associated challenges. In the wind tunnel
care must be taken to model the flow correctly and good trimming of the sails is crucial,
however the test environment is controllable and measurements are repeatable. The lift and
drag coefficients in the heeled plane show a noticeable reduction with heel angle and wind
tunnel tests should therefore be conducted at appropriate heel angles, for example by
employing the Real-Time VPP. The full-scale measurements are conducted in an unsteady
environment and it is difficult to measure all relevant data accurately which results in
significant scatter. Direction vane misalignment, mast twist and flow deflection due to the
presence of the sails are accounted for in the analysis, but an asymmetry in the full-scale data
is still noticeable on light wind days, which can possibly be attributed to wind shear. For a
meaningful comparison it is important to consider the forces generated by the hull/deck with
the sails present and depowering. In true wind speeds between 6 and 8m/s there is evidence
from the full-scale force measurements, and the sail shape analysis, that the sails were
depowered. The scatter in the full-scale data makes the comparison difficult but nonetheless
216
YACHT
RESEARCH
UNIT
The
University
of Auckland
7.2 Recommendations for Future Work
all force and moment coefficients show the same trends and no obvious scaling effect can be
detected. The most noticeable difference is the unrealistically low drag at small effective
wind angles in full-scale for which no conclusive cause could be determined.
This comparison shows wind tunnel testing to be a good quantitative tool to predict full-scale
aerodynamic forces. Although no scaling effects were observed here, these results are not
conclusive due to a number of uncertainties still associated with wind tunnel and full-scale
testing, each of which could influence the results significantly. All the trends in the forces and
moments are very similar in full-scale and the wind tunnel. The comparison confirms that
wind tunnel testing is a reliable qualitative tool for replicating full-scale sail behaviour.
The four novel objectives set out at the beginning of the project have therefore been achieved
and can again be summarised:
I. The interaction of the hull/deck with the sails has a significant effect on the side
force and the force perpendicular to the deck plane and it should be considered in the
aerodynamic analysis of sails and the performance prediction of yachts.
II. A Real-Time Velocity Prediction Program for wind tunnel testing has been
developed and implemented so that the model sails in the wind tunnel can be
trimmed based on the full-scale performance of the yacht, which makes the trimming
process in the wind tunnel much more similar to the real life situation.
III. Improved aerodynamic force models have been developed from realistically
depowered sail trims in the form of an empirical model based on the Real-Time VPP
measurements that describes the depowering of the sails in detail and, in the form of
an improved semi-empirical model based on aerodynamic principles that describes
the depowering with trim parameters in a more generic way.
IV. The wind tunnel measurements have successfully been compared to the full-scale
data by taking into account the interaction of hull/deck with the sails and the sail
depowering, and by applying a number of corrections to the full-scale data, which
highlighted challenges and uncertainties still associated with full-scale testing of
yachts.
7.2 Recommendations for Future Work
For this project considerable time has been spent in the wind tunnel testing yacht sails and
some recommendations on possible improvements to the testing procedures, which have not
been implemented yet, can be made. The limiting factor for a number of the measurements
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 217
Chapter 7 Conclusions and Recommendations
conducted for the project was the accuracy of the wind tunnel force balance in the vertical
direction. Not only are the vertical force measurements of interest in some situations, but a
component of the vertical force influences the side force in the heeled plane when the model
is not upright. With the Real-Time VPP the model can be dynamically heeled so that this is
particularly relevant. Improvements to the force balance and a closer investigation into the
induced vertical forces due to water in the trough could enhance the accuracy of the vertical
force measurements. Another possible improvement to the wind tunnel is the automation of
the turntable or the implementation of another means to change the apparent wind angle
dynamically so that Real-Time VPP measurements can be conducted at a constant true wind
angle.
Care must be taken in the wind tunnel to achieve the correct flow conditions and
improvements are certainly possible in this area, especially in regards to repeatability and
turbulence modelling. More importantly these issues should be investigated in full-scale as
little detailed information is available about vertical speed profiles close to the sea surface in
typical sailing conditions, and the associated turbulence intensities and length scales, and how
these are perceived by a moving yacht. Since the flow conditions depend on the area and
weather pattern, these should ideally be measured as part of full-scale studies in the future. In
the comparison with the Dyna data more information about the vertical speed profile and
wind shear could have eliminated some of the remaining uncertainties.
Another factor that made the comparison difficult is the scatter in the full-scale data.
Fluctuations in the measurements is however inevitable in full-scale testing due to the
unsteady test environment and changes to the sail shape caused by trim alterations or
dynamic effects. The comparison could be improved by considering individual force
measurements and replicating the sail shape in the wind tunnel. It is therefore desirable to
measure the sail shape simultaneously with the sail forces in full-scale, for which a system
has already been implemented on Dyna, and in the wind tunnel. In the wind tunnel the sail
shape capturing would ideally be in real-time while the sails are trimmed. Only a comparison
based on the same sail shape can provide accurate insight into possible scaling effects and the
comparison to CFD simulations can then easily be integrated.
Unsteady flow over the sails may occur in full-scale due to wind fluctuations, course changes
or yacht motion. A better understanding of the dynamic behaviour of sails would help to
quantify how relevant it is for full-scale testing. To advance performance prediction programs
modelling the dynamic behaviour of sails is also important and wind tunnel testing could be
an applicable tool for research in this field. With the Real-Time VPP a platform has already
been created which can possibly be used as part of such an investigation since FRIENDSHIP-
Equilibrium is capable of analysing non-stationary modes of motion.
218
YACHT
RESEARCH
UNIT
The
University
of Auckland
7.2 Recommendations for Future Work
A better understanding of the flow around the hull/deck and the influence of the sails has
been achieved. The next step could be to use this knowledge and possibly parts of the
experimental set-up to investigate if design changes to the hull and deck can reduce the
windage forces as both the side force and force perpendicular to the deck plane reduce the
performance of the yacht.
With the power model, an empirical aerodynamic force model that describes the force
changes due to depowering in detail has been developed, and input data has been obtained
with the Real-Time VPP for one sail configuration of an IMS cruiser/racer yacht. Similar
tests could be carried out for different sail combinations, sail types and yachts to determine
how sensitive the process of depowering is to these alterations. This would yield to what
extent the depowering information is transferable and if depowered measurements should to
be carried out for every sail design.
The depowered sail force measurements showed a clear dependency of the so-called parasitic
drag coefficient on depowering at larger apparent wind angles, which was modelled here as a
function of the reduction in lift squared. This has not been examined before since depowering
was previously only investigated in detail at small apparent wind angles. A better
fundamental understanding of the flow around sails at large apparent wind angles and the
associated drag components is important to further enhance generic semi-empirical
depowering models.

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 219

The
University
YACHT
RESEARCH
of Auckland UNIT
Appendices

A Wind Tunnel Model of Dyna 223
A.1 Design Considerations..............................................................................................223
A.2 Model Rig.................................................................................................................225
A.3 Model Sails...............................................................................................................229
B Experimental Procedures 231
B.1 Corrections for Hull/Deck Force Measurements .....................................................231
B.2 Meshing of Hull/Deck for Pressure Mapping ..........................................................232
B.3 Airflow below Model in Wind Tunnel Trough........................................................234
C Velocity Prediction Program Definitions 239
C.1 Coordinate Systems..................................................................................................239
C.2 Real-Time VPP Output File Structure .....................................................................242
C.3 FRIENDSHIP-Equilibrium User Manual ................................................................243
C.3.1 WindTunnel Force Module .....................................................................243
C.3.2 HansenRig Force Module........................................................................244
C.3.3 HansenRigDirect Force Module..............................................................247
C.4 Hydrodynamic Model of Dyna in FRIENDSHIP-Equilibrium................................249
D Mathematical Techniques 253
D.1 General Force and Moment System.........................................................................253
D.2 Central Axis System.................................................................................................254
D.3 Lifting-Line Theory to Model Sail Twist.................................................................256
D.4 Transformation of Flow Measurements to Heeled Plane.........................................260
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 221

The
University
YACHT
RESEARCH
of Auckland UNIT
A Wind Tunnel Model of Dyna
A.1 Design Considerations
For the wind tunnel testing of Dyna a new scale-model was designed and built. The scale of
models for the TFWT is based on the model mast height and sail area. A mast height of
around 2.2 metres is usually chosen for which it has been shown that the flow above the mast
is not significantly affected by the tunnel ceiling. For single mast yachts this mast height
gives an upwind sail area of around 1.3-1.4m
2
. If the sails are perpendicular to the flow the
model area is about 5% of the cross sectional area of the tunnel, which is a typical model size
for wind tunnel testing and for an open jet tunnel should result in a negligible amount of
blockage (Barlow et al., 1999). A scale of 1:6.67 (15%) was selected for the model of Dyna
so that the mast height above deck is 2.15 metres. Table A.1 shows the principal dimensions
of different models for the TFWT. For a very similar mast height, the sail area of the Dyna
model is slightly smaller than of the models of International Americas Cup Class (IACC),
Volvo Ocean 60 (VO60) and Volvo Open 70 (VO70) yachts. However, the beam and
freeboard of the Dyna model are much larger due to the different proportions of these
dissimilar yachts. Dyna is a much smaller yacht and has proportionally more freeboard and
beam.
Due to the novel investigation of the hull/deck forces the design concept of the Dyna model is
different to that of other sailing yacht models at the TFWT. Instead of a structural hull that
forms the base for the rig, a similar concept to that on the full-scale Dyna is employed with a
load carrying internal frame to which all rig components are attached and a hull/deck shell
that can be disconnected from the frame. The frame sits in a custom designed cradle through
which it is attached to the force balance under the wind tunnel floor and which allows it to
heel. Figure A.1 shows the computer representation of these three main components of the
Table A.1: Principal dimensions of wind tunnel models in TFWT
Model type Scale Mast
above
deck
[mm]
Typical Length Beam Freeboard
amidships
Canoe
body
draft
[mm]
sail area
upwind
[m


2
] [mm] [mm] [mm]
Tunnel trough 1780 540 100
IACC 1:15 0.067 2170 1.38 1710 235 85 60
VO70 1:14 0.071 2130 1.41 1535 365 100 40
VO60 1:12 0.083 2100 1.34 1570 350 110 40
Dyna 1:6.67 0.15 2149 1.30 1500 447 160 54
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 223
Appendix A Wind Tunnel Model of Dyna
Dyna model. The hull and deck shape is a 15% (1/6.67
th
) scale version of the full-scale
geometry. Two separate female moulds were constructed for the hull and deck geometry.
From the female moulds the hull and deck were built in glass-fibre reinforced plastic (GRP).
Building the plugs and female moulds is a lavish process but it results in thin and light shells
with a smooth surface finish, which provides advantages for both the force and pressure
measurements on the hull/deck. It is also comparatively easy to produce multiple copies, and
for this project two sets of hull and deck are used. The shells require little reinforcing since
they only have to support their own weight and the wind load.
As on other wind tunnel models a design concept with a cradle that connects the model to the
force balance was chosen since it provides an efficient method of heeling the model.
Compared to other concepts that also allow the model to heel, a cradle works well with water
in the trough since no openings in the hull are required; the model can be heeled to any angle
within its operational range and the same cradle can be used for different models. Although
the cradle is specifically designed for the Dyna model, the possible usage with other models
is kept in mind by, for example, making the model easily removable from the frame. The
frame is attached to the cradle through a rod at the bow and stern about which it can rotate.
The rotation axis is horizontal for simplicity and the axis height is selected to approximate the
real life heeling behaviour, which is of course not about a constant axis. A third connection
between the cradle and the frame transmits the heeling moment through a threaded rod in
tension, which can be turned to adjust the heel angle. This arrangement can for example be
seen in Figure 2.21 and Figure 2.22. In association with the Real-Time VPP the rod is turned
by an electric motor to dynamically change the heel angle. The hull shell is connected to the
frame at the bow and stern on the rotation axis so that it can easily be connected and
disconnected from the frame at any heel angle as discussed in section 3.2.

(a) Cradle (b) Internal load carrying frame (c) Hull/deck shell
Figure A.1: Three main components of the Dyna wind tunnel model; cradle, internal load carrying
frame and hull/deck shell
224
YACHT
RESEARCH
UNIT
The
University
of Auckland
A.2 Model Rig
For the design of the internal frame the key criteria are sufficient clearance to the hull and
deck shell and the capability to support all the necessary sail controls. Due to ideas early in
the project to measure the rig loads with load cells, the frame was designed with this in mind.
It was also ensured that that frame can be removed from the hull/deck shell without taking the
rig off. To preserve structural integrity and water tightness the hull and transom are not split
up into several components. The deck is therefore cut into seven pieces that attach to the hull
and can be removed without the need to make any changes to the rig or sail set-up. This is
important to allow access to the inside of the model without changing the sail trim. With all
deck components taken off the frame can be removed from the hull shell by unscrewing the
support bracket on the transom and then pushing the frame forward as far as possible to free
its aft end from the transom. It is therefore possible to change hull/deck shells without
compromising the rig or sail set-up.
Structural calculations were carried out during the design process of the cradle and frame to
ensure adequate strength and minimise deflections. In the final design the construction
material of the frame was steel and the cradle was built in aluminium. The complex geometry
of the frame is easier to weld in steel and by using aluminium for the cradle the weight is
reduced. The resulting total model weight including the cradle in normal testing configuration
without the pressure measurement system installed is approximately 15.3kg, of which the
hull/deck shell weighs about 4.3kg. The calculated maximum deflection at any point of the
design relative to the attachment points to the force balance is less 1mm. However, in reality
the deflection is significantly larger due to flexibility in some joints of the cradle which was
not considered carefully enough during the design and build process. As discussed in
appendix B.3, an additional aluminium plate is fitted to the bottom of the cradle to increase
the stiffness for measurements with water in the trough. This however limits the maximum
heel angle to approximately 27 compared to 50 of the original design. The additional plate
does not affect the measurements conducted as part of this project since cardboard instead of
water was used to cover the trough and the model was only tested at heel angles up to 25.
A.2 Model Rig
The model rig was constructed based on the sail plan of Dyna designed by judel/vrolijk & co.
and shown in Figure A.2. Unlike previous model masts, an attempt was made to achieve
stiffness characteristics similar to the full-scale mast to enable a more realistic modelling of
the full-scale rig and sail behaviour. Considerations in the early stages of the project such as
measuring the rig loads in the wind tunnel and the unstayed upper section of the fractional rig
made this even more relevant.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 225
Appendix A Wind Tunnel Model of Dyna

29
Figure A.2: Sailpan of Dyna

29
Drawing courtesy of judel/vrolijk & co. - engineering gmbh
226
YACHT
RESEARCH
UNIT
The
University
of Auckland
A.2 Model Rig
When scaling the stiffness characteristics of the mast the deflection (y) relative to the length
(L) should be the same in model and full-scale. The deflection of the mast (y) can be
described as a function of the applied force (F), the length (L), Youngs modulus of elasticity
(E) and the second moment of area (I) with

=
EI
FL
f y
3
,
(A.1)
where the applied force (F) is a function of the sail area (A
S
) and the apparent wind speed
(V
A
) such that
( )
2
A S
V A f F = .
(A.2)
The sail area (A ) scales with length squared (L
2
S
) so that for a constant y/L the product of
Youngs modulus of elasticity (E) and the second moment of area (I) scales with
fs
A
m
A
fs
m
L V
L V
EI
EI
4 2
4 2
= ,
(A.3)
where subscript
m
stands for model and
fs
for full-scale. The ratio between model length
and full-scale length is simply the model scale factor. The approximate wind tunnel speed
during upwind testing is 4.2m/s. The majority of upwind data on Dyna was collected in 6-
8m/s true wind speed and for 7m/s true wind speed the apparent wind speed for the best
VMG sailing condition was determined from VPP predictions as 8.9m/s. The ratio of the
apparent wind speed between model scale and full-scale was therefore estimated to be 0.47.
From this information and the design specification of the full-scale mast the target EI for the
model mast in the longitudinal and transverse direction was obtained. Based on the calculated
stiffness requirements a solid mast of soft wood was constructed. During the manufacturing
process the transverse stiffness of the mast was felt to be too low to be useable in the wind
tunnel and a thin layer of uni-directional carbon was added on both sides to increase the
transverse stiffness. The stiffness of the finished mast was determined by four point bending
tests over the constant cross section portion. The longitudinal IE is 12% larger than the target
and the transverse IE is 60% larger. This shows that the wood selection was appropriate and
that the addition of the carbon provided significant extra stiffness in the transverse direction.
Although the stiffness has not been matched accurately, the order of stiffness between full-
scale and model scale is now comparable. The model mast is already significantly more
flexible than other model masts and therefore a more careful rig set-up process is required.
The design mast rake for Dyna prior to bending was initially specified at 3. This was found
to be unrealistic during the full-scale measurements in normal sailing conditions and was
hence reduced to 1. Similar observations were made in the wind tunnel and the mast rake
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 227
Appendix A Wind Tunnel Model of Dyna
was also set to 1. The mast support under the coachroof, which is part of the internal frame,
constrains the mast transversely and longitudinally with the mast rake being adjustable.
Although this arrangement is not commonly seen on full-scale yachts it is used on Dyna and
hence also fitted on the wind tunnel model. Figure A.1(b) shows the mast support as part of
the internal frame of the Dyna model.
The rig set-up plays an important part in the sail performance and should be equivalent in
full-scale and in the wind tunnel to allow a meaningful comparison. On Dyna load cells are
installed on all rig components which provide important information for the rig set-up. This
information is not available from the wind tunnel model so that a more traditional method
needs to be adopted to ensure that the flexible mast maintains a controlled shape when sail
loads are applied. The mast rake is set to 1 as on the full-size boat. The forestay and
backstay are adjusted to the correct design length to produce an initial approximate mass
bend. The model is positioned in the wind tunnel at 90 to the flow direction with the
mainsail and jib trimmed hard on as flat as possible. The wind tunnel is run at a wind speed of
approximately 5.2m/s, a higher speed than expected during testing. This produces the
maximum heeling moment the model will experience. The leeward shrouds (V1, D1, D2) are
set to be slack and the windward shrouds are adjusted until the mast looks transversely
straight and vertical by visual inspection. Pictures of the model are taken from in front and
from behind, and the transverse mast bend is analysed with the software package SailTool
(also see section 6.7.1). With the flow turned off the leeward shrouds are tensioned by visual
inspection to bring the mast back to a transversely straight and vertical position. During this
process minor adjustments are made to the windward shrouds. Pictures are taken again and
the transverse mast bend analysed with SailTool. The maximum out of column bending of the
mast is 3mm.
The fore and aft bend of the mast is then adjusted by varying the forestay and backstay
tension. After each change a picture is taken and analysed in SailTool. The target is to match
the mainsail luff shape; an arc segment with a maximum draft at the luff midpoint of 1.6% of
the luff length. Without using the runner it was not possible to approximate the bend
adequately. The maximum draft was located well above the luff midpoint. With the runner
tensioned the target bend can be achieved accurately. The runner is therefore fixed at the
appropriate length and not adjusted during any wind tunnel tests to ensure that the rig set-up
is consistent. The target mast bend is achieved with the backstay slack and tensioning the
backstay efficiently increases the mast bend when depowering the sails. Releasing the
backstay returns the mast to its target curvature.


228
YACHT
RESEARCH
UNIT
The
University
of Auckland
A.3 Model Sails
A.3 Model Sails
The sails are built by North Sails New Zealand from the design data made available by North
Diamond Sail Makers A/S, Copenhagen, Denmark. The whole sail wardrobe of four sails
were constructed; mainsail, genoa (genoa I), jib (genoa III) and spinnaker. The leading
perpendicular of the genoa is 142% of the J measurement and for the jib it is 100%. The
design data of the jib is not available and the data for the same type of sail, rig and yacht had
to be used. Most likely the full-scale jib was constructed from the same design data but this is
not certain. If different design information was used the discrepancies are likely to be very
small. The sail design data for mainsail, genoa and spinnaker of Dyna are available so that the
exact design shapes can be reproduced. On the mainsail a minor alteration is made to the
leech curvature. Keeping the head, clew and midpoint up the mast in the specified position
the leech is smoothed out to give a visually more adequate leech shape at model scale. The
luff of the mainsail is not straight but designed to follow an arc segment with a maximum
draft of 1.6% of the luff length at the luff midpoint. The mast bend is set-up to accommodate
this curvature. The mainsail is tied onto the mast with thin strings at several points along the
luff. With wind pressure in the sail and adequate luff tension the luff follows the mast bend
tightly and no gaps are present. Placing tape along the whole length of the luff is not seen as
necessary and it would make applying and changing the luff tension more difficult. Similarly
the headsail is attached to the forestay by thin tape strips rather than applying tape along the
whole forestay. This way the luff can be tensioned more easily and it still touches the forestay
over its entire length.
The mainsail, genoa and jib are made of 9933 polyester and the spinnaker of Contender
Superlite (polyester). For the mainsail battens pattern Mylar is used. Initially one set of sails
was constructed and later in the project a second set of mainsail, genoa and jib was built with
stronger seams between panels, which are not only glued but also stitched. The original sails
used for upwind testing wore out over time mainly due to distortions on the glued overlaps of
the sail panels. All the results presented in this work are obtained with the second set of sails.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 229

The
University
YACHT
RESEARCH
of Auckland UNIT
B Experimental Procedures
B.1 Corrections for Hull/Deck Force Measurements
The secondary force balance inside the model to measure the hull/deck forces is mounted on
an aluminium plate attached to the internal frame. The plate and the hull/deck shell
connection to the force balance are not very stiff. In addition the hull/deck shell represents a
large mass with a large radius of gyration in the x-direction. For these reasons the calibration
of the force balance is affected and the measurements should be corrected. Figure B.1 shows
the measured forces for a range of known downward-acting vertical forces (F
Z
) applied at
different x-positions along the centreline of the model. The measurement error seems to be
directly proportional to the x-position where the force is applied and hence to the moment.
Figure B.2 shows the measured pitching moment (M
Y
) for a range of known moments and it
can be seen that the error in F
Z
does not translate into the pitching moment. Hence only the
vertical force needs to be adjusted and the correction can simply be a proportion of the
pitching moment so that
,
Y Z Zcorrected
M a F F
1
+ =
(B.1)
where a is the correction factor. For the data shown in Figure B.1
1
the correction factor can
be determined as 0.151 by minimising the squared errors. The corrected measurements are
plotted against the applied force in Figure B.3.
-2.5
-2
-1.5
-1
-0.5
0
-2.5 -2 -1.5 -1 -0.5 0
Applied force [N]
M
e
a
s
u
r
e
d

f
o
r
c
e

[
N
]

Fz applied at x=-0.45m
Fz applied at x=-0.25m
Fz applied at x=0m
Fz applied at x=0.25m
Fz applied at x=0.45m
Fz applied at x=0.75m
No error
-1.5
-1
-0.5
0
0.5
1
1.5
-1.5 -1 -0.5 0 0.5 1 1.5
Applied moment [Nm]
M
e
a
s
u
r
e
d

m
o
m
e
n
t

[
N
m
]
My applied at x=-0.45m
My applied at x=-0.25m
My applied at x=0m
My applied at x=0.25m
My applied at x=0.45m
My applied at x=0.75m
No error
Figure B.2: Measurement errors in pitching
moment (M
Figure B.1: Measurement errors in vertical force
(F
Z
) applied at different x-positions
along centreline of model
Y
) applied at different x-
positions along centreline of model
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 231
Appendix B Experimental Procedures
-2.5
-2
-1.5
-1
-0.5
0
-2.5 -2 -1.5 -1 -0.5 0
Applied force [N]
M
e
a
s
u
r
e
d

f
o
r
c
e

c
o
r
r
e
c
t
e
d

[
N
]

Fz applied at x=-0.45m
Fz applied at x=-0.25m
Fz applied at x=0m
Fz applied at x=0.25m
Fz applied at x=0.45m
Fz applied at x=0.75m
No error
-1.5
-1
-0.5
0
-1.5 -1 -0.5 0
Applied force [N]
M
e
a
s
u
r
e
d

f
o
r
c
e

[
N
]

Fy applied at x=-0.555m
Fy applied at x=-0.148m
Fy applied at x=0.24m
Fy applied at x=0.71m
No error
Figure B.4: Measurement errors in side force (F Figure B.3: Corrected vertical force (F
A calibration check is also conducted by applying a side force (F
Y
) to starboard at different x-
positions along the model. As shown in Figure B.4 the measurement errors are much smaller
compared to the errors in the vertical force. Nevertheless the errors seem to be related to the
applied moment as well. In this case a yawing moment (M
Z
) is applied and a correction
equivalent to that shown in equation (B.1) is employed in the form of
,
Z Y Ycorrected
M a F F
1
+ =
(B.2)
where a
1
is determined to be -0.052.
The force along the centreline (F
X
) does not require a calibration correction since it does not
produce significant moments due to its small magnitude and the location of the force balance
being close to the vertical centre of pressure on the hull/deck. The results and correction
shown in Figure B.1 to Figure B.4 are for the first test session in which the internal force
balance was used. The calibration check was repeated for the second test session since the
model was dismantled in between the sessions and the installation of the force balance can
influence the calibration errors. The behaviour of the set-up was found to be very similar so
that the calibration correction factors did not need to be adjusted for the second test session.
B.2 Meshing of Hull/Deck for Pressure Mapping
A sensitive study was carried out to determine what grid density should be used for
interpolating the surface pressure on hull and deck from the pressure measurements. For the
interpolation the universal kriging algorithm described by Davis (2002) was employed in
TecPlot. Nine meshing options with between 958 and 15736 elements were compared. All
Zcorrected
)
applied at different x-positions along
centreline of model
Y
)
to starboard applied at different x-
positions along the model
232
YACHT
RESEARCH
UNIT
The
University
of Auckland
B.2 Meshing of Hull/Deck for Pressure Mapping
meshes consist of triangular elements and each pressure tap coincides with a node of the
mesh. The coarsest and finest mesh, together with the pressure tap locations are shown in
Figure B.5. Figure B.6 visually shows how the choice of mesh density affects the
interpolation of the pressures for the coarsest and finest mesh.
As a first indication of the quality of the mesh, the surface area of the mesh is compared to
the actual surface area of the hull and deck. Figure B.7 shows the ratio of mesh area over the
actual area plotted against the number of elements. As expected the mesh area asymptotically
approaches the actual area with increasing its number of elements. Figure B.7 also shows that
the mesh area below the waterline is smaller than the actual area because of the curvature of
the hull. The mesh area above the waterline on the other hand is somewhat larger than the
actual area so that the total mesh area shows comparatively little discrepancy from the actual
area. When looking at the mesh density the influence on the forces is the most important
(a) Coarsest mesh with 958 elements (b) Finest mesh with 15736 elements
Figure B.5: Coarsest and finest mesh used in sensitivity analysis


(a) Coarsest mesh with 958 elements (b) Finest mesh with 15736 elements
Figure B.6: Interpolated pressure for coarsest and finest mesh at
A
of 30 and 0 heel
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 233
Appendix B Experimental Procedures
0.9
0.92
0.94
0.96
0.98
1
1.02
0 4000 8000 12000 16000
Number of elements [-]
M
e
s
h

a
r
e
a

/

A
c
t
u
a
l

a
r
e
a

[
-
]
Total area
Area above water at 0 heel
Area below water at 0 heel
0.75
0.8
0.85
0.9
0.95
1
1.05
0 4000 8000 12000 16000
Number of elements [-]
C
o
e
f
f
i
c
i
e
n
t
/
C
o
e
f
f
i
c
i
e
n
t
(
1
5
7
3
6

e
l
e
m
e
n
t
s
)

[
-
]
CL 30 AWA
CL 90 AWA
CD 30 AWA
CD 90 AWA
CFz 30 AWA
CFz 90 AWA
Figure B.8: C , C and C Figure B.7: Hull/deck surface area for different
grid densities
L D Fz
of hull/deck at 0 heel
for
A
of 30 and 90 obtained with
different grid densities
parameter. The coefficients of lift, drag and the force perpendicular to the deck plane
calculated from the measured surface pressures as described in section 3.3 are plotted against
the number of elements in Figure B.8 for two apparent wind angles, 30 and 90. The
coefficients are normalised by the respective coefficient for the finest mesh. For small
numbers of elements the forces are under-predicted and the coefficients asymptotically
approach one with increasing number of elements. From Figure B.7 and Figure B.8 it is
determined that a mesh with around 8000 elements is sufficient for calculating the forces and
moments from the measured surface pressures.
B.3 Airflow below Model in Wind Tunnel Trough
The wind tunnel turntable is designed with a trough as shown in Figure B.9 so that the
waterline of the model coincides with the wind tunnel floor. In real life the yacht floats on the
waterline and airflow under the hull in not possible. The flow field around the yacht
influences the sail forces and to realistically replicate it in the wind tunnel the gap between
the edge of the trough and the hull should be closed so that airflow under the hull is also
prevented in the wind tunnel. In this section the significance of airflow under the hull is
investigated and methods of preventing it are discussed. At the Twisted Flow Wind Tunnel
(TFWT) it is common practice to fill the trough with water to seal the gap. Using water has
the advantages that the gap is sealed without forces being transmitted and that the seal is
maintained for different heel angles. At other wind tunnels flexible membranes are used
which require a calibration correction and adjustments for different heel angles. Alternatively
cardboard cutouts which cover the trough around the model can be used to minimise the gap.
The gap however cannot be sealed to prevent any force transmission. Since the water plane
234
YACHT
RESEARCH
UNIT
The
University
of Auckland
B.3 Airflow below Model in Wind Tunnel Trough
-1
-0.5
0
0.5
1
1.5
2
2.5
0 20 40 60 80 100 120 140

eff
[]
C
L

[
-
]
Open trough
Open trough (spline)
Cardboard
Cardboard (spline)
Water
Water (spline)
Figure B.10: C Figure B.9: Model with cradle sitting in trough of
turntable so that waterline coincides
with wind tunnel floor
L
with constant mainsail and jib trim
at 0 heel for trough open, covered
with cardboard and filled with water
shape of the hull changes with heel angle a different cutout shape is required for each heel
angle to achieve an accurate fit and care needs to be taken during set-up to ensure that the
cardboard does not touch the model.
Figure B.10 and Figure B.11 show the lift and drag coefficients (C and C
L D
) for the upright
model with mainsail and jib measured in the wind tunnel at different effective wind angles
with the trough open, covered with cardboard and filled with water. For the comparison it is
important that the same sail trim is used for the different set-ups. Even when recording the
trim settings it cannot be ensured that the exact same trim is achieved again. It is therefore
decided to use the same sail trim for all apparent wind angles in this study to ensure that it is
identical for the three set-ups. The best sail trim developed at an apparent wind angle of 25
is used for all angles. As a result the C and C
L D
curves are not realistic but accurately show
the differences due to the different set-ups. From Figure B.10 and Figure B.11 it can be seen
that C and C
L D
measured with the cardboard cutouts and water in the trough are almost
identical whereas there are noticeable differences when the trough is left open. The
differences are most noticeable in C
D
. It can be concluded that the trough should not be left
open and that both sealing the gap with water and minimising the gap with cardboard cutouts
produce almost identical results for C and C
L D
.
For parts of this research project the force coefficient perpendicular to the deck plane (C
Fz
) is
of particular interest. C in Figure B.12
Fz
is very different for the three set-ups. Not only do the
open trough results differ from the other set-ups, the cardboard results are also very different
from the measurements with water in the trough. It was discovered that the heeling moment,
caused by the side force applied at a point up the mast, causes the cradle to flex. In addition
the life part of the force balance with the model moves slightly because of linear
displacement transducers being used as measuring devices. The movement and flex lowers
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 235
Appendix B Experimental Procedures
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
0 20 40 60 80 100 120 140

eff
[]
C
D

[
-
]
Open trough
Open trough (spline)
Cardboard
Cardboard (spline)
Water
Water (spline)
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 20 40 60 80 100 120 140

eff
[]
C
F
z

[
-
]
Open trough
Open trough (spline)
Cardboard
Cardboard (spline)
Water
Water (spline)
Water corrected
Water corrected (spline)
Figure B.12: C Figure B.11: C
the hull of the model further into the trough and if the trough is filled with water the
buoyancy of the model is increased, which results in an upwards force being measured. It can
be shown that by applying a side force at the height of the centre of sail area a vertical force
of approximately 12.7% of the applied force is induced due to the change in buoyancy. The
induced vertical force was reduced by 39% through stiffening the cradle by adding a plate to
the bottom. The induced vertical force could also be reduced by 64% through flooding the
hull of the model so that the changes in buoyancy are significantly reduced. By flooding the
model and stiffening the cradle the induced vertical force was therefore reduced by 78%,
from 12.7% to 2.8% of the applied side force.
As a basic approximation for the original measurements with water in the trough, the induced
vertical force is modelled as 12.7% of the heeling moment and the corrected C
Fz
is shown in
Figure B.12. The difference in C
Fz
between cardboard and water is now much less which
suggests that the induced vertical force was the main cause of the initial difference. A more
detailed study that develops a calibration correction from a number of load cases would
however be necessary to determine with certainty whether the difference in C
Fz
is solely
caused by the induced vertical force or also by the fact that the gap is not sealed when
cardboard is used instead of water.
Some more information regarding the influence of having a small gap between the tunnel
floor and model can be obtained from the surface pressure measurements on hull and deck
described in section 3.3. For three apparent wind angles the hull/deck surface pressures were
measured for the model upright with mainsail and jib with the cardboard cutouts in place to
minimise the gap and with the gap between the cardboard and the model sealed by tape. From
the pressures above the waterline the force coefficients are calculated and shown in Figure
D
with constant mainsail and jib
trim at 0 heel for trough open,
covered with cardboard and filled
with water
Fz
with constant mainsail and jib
trim at 0 heel for trough open,
covered with cardboard and filled
with water
236
YACHT
RESEARCH
UNIT
The
University
of Auckland
B.3 Airflow below Model in Wind Tunnel Trough
-0.05
0
0.05
0.1
0.15
0.2
0 20 40 60 80 100 120 140

eff
[]
C
o
e
f
f
i
c
i
e
n
t

[
-
]
CL
CL gap sealed
CD
CD gap sealed
CFz
CFz gap sealed

Figure B.13: C , C and C
B.13. C
L
and C
D
of hull/deck are not affected much by a small gap between the cardboard and
the model. The increase in C
Fz
with effective wind angle is up to 20% with the gap sealed as a
higher suction can be generated over the deck. Nevertheless it can be concluded that the flow
behaviour around the model is still represented well when using cardboard cutouts. For this
research project cardboard cutouts are advantageous because the force measurements are not
influenced by the induced vertical forces, and they can be used for both the force and surface
pressure measurements. The same set-up for the force and pressure measurements allows a
better comparison between the two techniques. Using water in the trough during the hull/deck
pressure measurements however, would be prohibitively difficult. The wind tunnel tests for
this project were conducted at three fixed wind angles of 0, 12.5 and 25 so that three sets
of cardboard cutouts were manufactured and used.
For other types of wind tunnel testing, water in the trough is still the preferred method since
no cardboard cutouts need to be manufactured and less effort is required before each test run
to ensure the correct positioning of the cardboard. Furthermore, cardboard cannot be
employed when the model is dynamically heeled. When using water it is important to design
a stiff model and cradle, and, if possible, to design the model in such a way that the change in
buoyancy is minimal when the model changes position in the trough. A calibration correction
to account for the induced vertical force could also be developed but in addition to the model
properties it is also dependent on the water level and the apparent wind angle. The turntable
is not perfectly horizontal so that the water level changes when the turntable is rotated to
adjust the apparent wind angle.
L D Fz
on hull/deck from
pressures above waterline

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 237

The
University
YACHT
RESEARCH
of Auckland UNIT
C Velocity Prediction Program Definitions
C.1 Coordinate Systems
In FRIENDSHIP-Equilibrium two main reference coordinate systems are used. The absolute
coordinate system (coordinate system A) is defined relative to the horizontal plane. The z-
plane is horizontal and the y-plane is parallel to the centreline of the yacht. The x-axis is
pointing towards the bow, the y-axis is pointing to port and the z-axis is pointing vertically
upwards as shown in Figure C.1. For Dyna the origin of coordinate system A is at the
position of the Fibre-Optic-Gyro (FOG), which is installed on the centreline in front of the
mast and slightly below the waterline. The exact position is described in Table C.1. The force
equilibrium is calculated in coordinate system A and all forces and moments must be
transformed into this system.
X
A
Y
A
Plan view at
m
= 0
A
Absolute
coordinate system
X
Y
Plan view at
m
= 0
Z
Y
X
F
Y
F
Plan view at
m
= 0
X
T
Y
T
X
T
Y
T
Plan view at
m
= 0
B
Body fixed
coordinate system
Z
A
Y
A
F
Force balance
coordinate system
T
Tunnel
coordinate system
Z
F
Y
F
Z
T
Y
T
Figure C.1: Coordinate systems used in FRIENDSHIP-Equilibrium and the wind tunnel
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 239
Appendix C Velocity Prediction Program Definitions
The second main reference coordinate system of FRIENDSHIP-Equilibrium is the body fixed
coordinate system (coordinate system B). It shares the same origin and sign convention (right
hand system) with coordinate system A, but is defined relative to the yacht so that it heels and
pitches with the yacht as shown in Figure C.1 and explained in Table C.1. The z-plane is
parallel to the DWL plane and the y-plane is parallel to the centreline plane of the yacht. The
x-axis is pointing towards the bow, the y-axis is pointing to port and the z-axis is pointing
roughly up the mast. The aerodynamic forces and moments acting on the yacht are defined in
coordinate system B in FRIENDSHIP-Equilibrium and it is the main coordinate system for
expressing aerodynamic forces in this work.
In the wind tunnel the forces are measured in the force balance coordinate system (coordinate
system F). The x-axis is usually aligned with the centreline of the model, the y-axis is
Table C.1: Definition for coordinate systems used in FRIENDSHIP-Equilibrium and the wind tunnel
Symbol Name and description Page
A Absolute coordinate system 23, 35
Origin for Dyna in centre of FOG reference system
X
0
= F
PP
3330mm , Y
0
= centreline , Z
0
= DWL

179mm
X horizontal and aligned with centreline of yacht, positive forward
Y also horizontal and perpendicular to X, positive to port
Z perpendicular to X and Y, positive upwards
Moments positive clockwise

B Body fixed coordinate system of yacht 21, 34
Origin for Dyna in centre of FOG reference system
X
0
= F
PP
3330mm , Y
0
= centreline , Z
0
= DWL 179mm
X in DWL plane and aligned with centreline of yacht, positive
forward
Y also in DWL plane and perpendicular to X, positive to port
Z perpendicular to X and Y, positive upwards
Moments positive clockwise

F Force balance coordinate system 33
Origin of force balance
X horizontal and aligned with centreline of model yacht, positive
forward
Y also horizontal and perpendicular to X, positive to starboard
Z perpendicular to X and Y, positive upwards
Moments positive clockwise

T Tunnel coordinate system 33
Origin of force balance
X horizontal and aligned with centreline of model yacht, positive
forward
Y also horizontal and perpendicular to X, positive to port
Z perpendicular to X and Y, positive upwards
Moments positive clockwise

240
YACHT
RESEARCH
UNIT
The
University
of Auckland
C.1 Coordinate Systems
horizontal and positive to starboard and the z-axis is positive vertically upwards as shown in
Figure C.1 and explained in Table C.1. The origin of the force balance coordinate system is
shown in Figure 2.18. Coordinate system F uses a non-standard sign convention to make the
most relevant angles and forces positive for common test conditions. However, for this work
a standard right hand system is more convenient for interfacing with FRIENDSHIP-
Equilibrium so that the tunnel coordinate system (coordinate system T) is also defined, which
is the right hand system version of coordinate system F. In coordinate system T the positive
direction of the y-axis is simply reversed as shown in Figure C.1. To convert from coordinate
system F to coordinate system T, the force along the y-axis (F
Y
) and the moment about the y-
axis (M
Y
) have their sign reversed. The transformation from the tunnel system T to the body
fixed system B is discussed in section 2.4.1.
















Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 241
Appendix C Velocity Prediction Program Definitions
C.2 Real-Time VPP Output File Structure
The Real-Time VPP results can be saved to an output file by the Real-Time VPP LabVIEW
application. Table C.2 lists and explains the information contained in this output file.
Table C.2: List of parameters contained in the output file of the Real-Time VPP LabVIEW application
Parameter name Unit Description

Comment - Comment entered in GUI
Run - Run number specified in GUI
VT m/s Full-scale true wind speed specified in GUI
TWA True wind angle specified in GUI or calculated by VPP
AWAm Apparent wind angle of model in tunnel, specified or
calculated
EWAm Calculated effective wind angle of model in tunnel
Heelm Heel angle of model in tunnel, specified or measured
SArefm m
2
Reference sail area of wind tunnel model specified in GUI
Presfac - Pressure factor between position of pitot tube and reference
height above turntable (usually height of centre of sail area)
Time s Sampling time specified in GUI
Freq Hz Sampling frequency
-FYb, FXb, FZb N Forces measured by force balance in force balance
coordinate system (coordinate system F)
-MYb, MXb, MZb Nm Moments measured by force balance in force balance
coordinate system (coordinate system F)
qPitot Pa Reference dynamic pressure measured by pitot tube
CDcor, CLcor - Correction of drag and lift coefficient calculated from the
effective angle correction (see section 4.4.3)
FXres, FYres, FZres N Residual full-scale forces of VPP equilibrium calculations in
absolute coordinate system (coordinate system A)
MXres, MYres,
MZres
Nm Residual full-scale moments of VPP equilibrium calculations
in absolute coordinate system (coordinate system A)
VS m/s Boat speed of full-scale yacht calculated by VPP
Heelf Heel angle of full-scale yacht calculated by VPP
Leeway Leeway (or yaw) angle of full-scale yacht calculated by VPP
Delta Rudder angle of full-scale yacht calculated by VPP
Trim Trim (or pitch) angle of full-scale yacht calculated by VPP
Sink m Sinkage of full-scale yacht calculated by VPP
EWAf Effective wind angle on full-scale yacht calculated by VPP
VAeff m/s Effective wind speed on full-scale yacht calculated by VPP
FN - Froude number of full-scale yacht calculated by VPP
CFx, CFy, CFz - Force coefficients of wind tunnel measurements in body
fixed coordinate system (coordinate system B)
CMx, CMy, CMz - Moment coefficients of wind tunnel measurements in body
fixed coordinate system (coordinate system B)
CD, CL - Drag and lift coefficient of wind tunnel measurements in
body fixed coordinate system (coordinate system B)

242
YACHT
RESEARCH
UNIT
The
University
of Auckland
C.3 FRIENDSHIP-Equilibrium User Manual
C.3 FRIENDSHIP-Equilibrium User Manual
In this section, only the aerodynamic modules developed by the author as part of this project
are described. A brief summary of the hydrodynamic force modules and their input data is
given in appendix C.4 and a more detailed description of FRIENDSHIP-Equilibrium and the
standard force modules can be found in the user manual (FRIENDSHIP SYSTEMS, 2005).
Three newly developed force modules are described here. The WindTunnel force module is
used in conjunction with the Real-Time VPP and enables the measured wind tunnel forces to
be used by FRIENDSHIP-Equilibrium to calculate the performance of the yacht. The
HansenRig and HansenRigDirect force modules are improved trim parameter models.
C.3.1 WindTunnel Force Module
The WindTunnel force module is used in conjunction with the Real-Time VPP and receives
the forces measured in the wind tunnel as input to calculate the full-scale forces. Information
regarding the wind tunnel set-up and wind conditions are specified in the Real-Time VPP
LabVIEW application discussed in section 4.3.1 and only the data relating to the full-scale
yacht, the coordinate systems origins and the effective angle correction are specified in the
WindTunnel force module shown in Table C.3. The wind tunnel measurements are passed
to FRIENDSHIP-Equilibrium in coefficient form. The full-scale sail area and the height of
the geometric centre of sail area, for calculating the effective wind speed, are also required.
The location of the wind tunnel coordinate system (coordinate system T) origin in relation to
the body fixed coordinate system (coordinate system B) at 0 heel at model scale and the
location of the heel axis of the model relative to the wind tunnel coordinate system
Table C.3: Input parameters of the WindTunnel force module of FRIENDSHIP-Equilibrium
Parameter name Example input Description

Name mySailForces Descriptive name
Reference sail area (A SailArea 50.87 ) of the full-scale yacht
S
CenterOfArea 6.63 Full-scale height of geometric centre of sail area
above the waterline at =0. Is only used for
calculating the onset flow onto the sails
Origin (-0.397;0.0;0.039) Origin of wind tunnel coordinate system T in
body fixed coordinate system B at
m
=0 (model
upright in wind tunnel) at model scale
HeelAxis (0.0;0.0;0.103) Location of model heel axis in wind tunnel
coordinate system T at model scale
CL
Function of Lift coefficient as function of
eff eff
for applying an
effective angle correction (see section 4.4.3)
CD
Function of Drag coefficient as function of
eff eff
for applying
an effective angle correction (see section 4.4.3)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 243
Appendix C Velocity Prediction Program Definitions
(coordinate system T) origin at model scale are also essential. In order to apply an effective
angle correction between the effective angle in the wind tunnel and the effective angle
calculated by FRIENDSHIP-Equilibrium, the lift and drag coefficient curves against effective
wind angle (
eff
) can be specified. An effective angle correction is applied if FRIENDSHIP-
Equilibrium solves for a constant true wind angle or for a constant apparent wind angle
without dynamically heeling the model. The equations used in association with these input
parameters are shown in section 4.3.3 and the wind tunnel constraints relating to the effective
angle correction are discussed in section 4.4.3.
C.3.2 HansenRig Force Module
The HansenRig force module is an experimental module that was gradually enhanced
during the process of developing improved aerodynamic force models in chapter 5. In order
to allow an easier and more accurate comparison, the different depowering methods are all
implemented in the same force module. In its final form the HansenRig force module
includes four depowering models. All methods use the same description of the fully powered-
up sail forces and only differ in the way the depowering is described. The common approach
of calculating the sail forces from the lift and parasitic drag coefficient and the height and
longitudinal position of the centre of effort as functions of the effective wind angle (
eff
) is
adhered to. In addition to existing force models, the force perpendicular to the deck plane is
modelled as a function of the
eff
and heel (see section 3.8) and the moment in line with the
force is described as a function of
eff
(see section 5.3). The windage drag coefficient can also
be specified as a function of
eff
. The difference between specifying a drag component as
windage drag instead of parasitic drag is that the windage drag remains constant when
depowering whereas the parasitic drag changes. The separation drag is calculated from
equation (2.5) and the induced drag from equation (2.6). The required input parameters of
separation constant (c
s
), aspect ratio (AR) and efficiency (e) are specified in the HansenRig
force module. For calculating the forces from the coefficients the sail area (A
S
) and the height
of the geometric centre of sail area, for obtaining the reference effective wind speed, are also
required. Table C.4 shows all the input parameters of the HansenRig force module and
indicates which parameters are used for which of the four depowering models. It can be seen
that the parameters described so far are relevant for all methods. The maximum true wind
angle for which the trim parameters are optimised can also be specified for all models.
The first depowering model is the advanced power model introduced in sections 5.4 and
5.4.3. It describes the changes in the coefficient of lift, parasitic drag and moment in line with
the force as well as the centre of effort location as ratios (R) of the fully powered-up values.
The ratios are functions of
eff
and the trim parameter power. The trim parameter power
represents the reduction in heeling moment when the sails are depowered and is adjusted
244
YACHT
RESEARCH
UNIT
The
University
of Auckland
C.3 FRIENDSHIP-Equilibrium User Manual
between one and zero to optimise the yacht performance. The response surface values for p=1
are still relevant when the other depowering models are used so that meaningful input for the
ratios is always required. The second depowering model is the generic power model
described in section 5.4.2. The depowering is still modelled with the trim parameter power
but the dependencies are not modelled by response surfaces but by a set of constant
coefficients. Only the lift, drag and centre of effort height are adjusted when depowering the
sails with equations and the coefficients a to a (5.15),(5.16) and (5.19)
1 6
are specific in the
force module.
The third depowering model is the standard semi-empirical depowering model with the trim
parameters reef, flat and twist developed by Kerwin (1978) and Jackson (2001) as described
in section 4.2.1. As input to describe the depowering the boom height (z
boom
) is required for
equation and the twist weight constant (c (4.8) ) for equation (4.2)
t
. The fourth depowering
model is the final version of the enhanced flat and twist model developed in section 5.6. In
addition to the twist weight constant (c ), the twist weight constant for lift (c
t tL
) and the centre
of effort height for minimum induced drag (z ) are required. z
CoEminDi CoEminDi
can be specified
as a function of
eff
and is given as a fraction of the mast height, so that the mast height (z
mast
)
is also an input.
The different depowering models are chosen by activating the appropriate trim parameters
and by two tick boxes called GenericPower and AdvancedTwist. Activating the power
parameter means that the power model is used. If the GenericPower box is ticked the
generic version of the power model is selected. The power parameter must not be used in
combination with the other trim parameters. Reef, flat and twist can be used in any
combination as the standard semi-empirical depowering model and ticking the
AdvancedTwist box activates the enhanced formulation developed as part of this project.

Table C.4: Input parameters of the HansenRig force module of FRIENDSHIP-Equilibrium
Model Parameter
name
Example input Description

1 2 3 4
Name mySailForces Descriptive name
Active Tick / untick Active modules are considered for
equilibrium calculations

Reference sail area (A SailArea 50.87 )
S
CenterOfArea (0 ; 0 ; 6.63) Height of geometric centre of sail
area above the waterline at =0. Is
only used for calculating the onset
flow onto the sails.

HeightMast 15.36 Height of mast above water for
upright yacht (z

mast
)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 245
Appendix C Velocity Prediction Program Definitions
HeightBoom 2.56 Height of boom above water for
upright yacht (z

boom
)
Aspect ratio (AR) of the rig RigAspectRatio 4.24

Efficiency (e) of the rig Efficiency 1.1

Separation constant (c SeparationConst 0.0016
s
) usually taken
as 0.016 for upwind and 0.019 for
downwind sails
Twist weight constant (c TwistWeight 8 )
t
Twist weight constant for lift (c TwistWeightLift 1.67
tL
)
Optimum position of centre of effort
for minimum induced drag (z
Sopt
Function of
eff

CoEminDi
)
as fraction of mast height vs.
eff

(default 0.42 for all
eff
)
Optimum lift coefficient (C CL
Function of
eff
Lopt
) in
coordinate system B vs.
eff
Optimum parasitic drag coefficient
(C
CDp
Function of
eff
DpOpt
) in coordinate system B vs.

eff
Windage drag coefficient (C CDwindage
Function of
eff
Dwindage
)
in coordinate system B vs.
eff
CFz Optimum force coefficient in the z-
direction (C
Function of
eff
FzOpt
) in coordinate
system B as function of
eff
XCE Optimum longitudinal centre of effort
position (x
Function of
eff
CoEopt
) in coordinate system
B vs.
eff
ZCE Optimum centre of effort height
(z
Function of
eff
CoEopt
) in coordinate system B vs.

eff
CMf Optimum moment coefficient in line
with force (C
Function of
eff
MfOpt
) in coordinate
system B vs.
eff
RatioL
Function of
eff
and
power

Ratio of C /C
L LOpt
(R
L
) in coordinate
system B as function of
eff
and trim
parameter power
RatioDp
Function of
eff
and
power

Ratio of C /C
Dp DpOpt
(R
Dp
) in
coordinate system B as function of

eff
and trim parameter power
RatioXCE
Function of
eff
and
power

Ratio of x
CoE
/x
CoEopt
(R
xCoE
) in
coordinate system B as function of

eff
and trim parameter power
RatioZCE
Function of
eff
and
power

Ratio of z
CoE
/z
CoEopt
(R
zCoE
) in
coordinate system B as function of

eff
and trim parameter power
RatioMf
Function of
eff
and
power

Ratio of the C
Mf
/C
MfOpt
(R
Mf
) in
coordinate system B as function of

eff
and trim parameter power
Array of constants (a ,a ,,a PowerConstants Array of constants
1 2 6
)
describing depowering for the
generic version of the power model
246
YACHT
RESEARCH
UNIT
The
University
of Auckland
C.3 FRIENDSHIP-Equilibrium User Manual
GenericPower Tick / untick If ticked the generic version of the
power model with constants instead
of response surfaces is used

If ticked the improved twist and flat


model is used instead of the original
reef, flat and twist model
AvancedTwist Tick / untick

UpwindTrim 120 Maximum true wind angle for which


trim parameters are optimised

Settings for the trim parameter reef REEF initial=1;active /
fixed;range=[0:1]

Settings for the trim parameter flat FLAT initial=1;active /


fixed;range=[0:1]


Settings for the trim parameter twist TWIST initial=0;active /
fixed;range=[0:0.5]


Settings for the trim parameter power POWER initial=1;active /
fixed;range=[0:1]


Key: 1: Advanced power model 2: Generic power model 3: Original reef, flat and twist
model 4: Enhanced flat and twist model

C.3.3 HansenRigDirect Force Module
The HansenRigDirect force module uses the direct power model developed in section
5.4.4. It is not included in the HansenRig module together with the other depowering
models described in appendix C.3.2 because it employs a different approach of modelling the
fully powered-up forces. Instead of describing the sail forces in terms of lift and drag
coefficients and the centre of effort position, the forces and moments are directly modelled by
the three force and three moment coefficients in coordinate system B. The six coefficients are
all expressed as functions of the effective wind angle (
eff
) as shown in Table C.5, which lists
all input parameters of the HansenRigDirect force module. The sail area (A
S
) and the height
of the geometric centre of sail area, for obtaining the reference effective wind speed, are
required to calculate the forces and moments. Equivalent to the advanced power model the
depowered forces and moments are expressed as ratios (R) of the fully powered-up values.
The ratios are again functions of
eff
and the trim parameter power and modelled by response
surfaces. Since the trim parameter power represents the reduction in heeling moment (M
X
),
the ratio R
Mx
simply equals power and does not require modelling by a response surface so
that only five response surfaces are necessary to fully describe the changes to due depowering
the sails. As before the maximum true wind angle for which the trim parameter power is
optimised can also be specified.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 247
Appendix C Velocity Prediction Program Definitions
Table C.5: Input parameters of the HansenRigDirect force module of FRIENDSHIP-Equilibrium
Parameter name Example input Description

Name mySailForces Descriptive name
Active Tick / untick Active modules are considered for equilibrium
calculations
Reference sail area (A SailArea 50.87 )
S
CenterOfArea (0 ; 0 ; 6.63) Height of geometric centre of sail area above the
waterline at =0. Is only used for calculating
the onset flow onto the sails
CFx Optimum force coefficient in the x-direction
(C
Function of
eff
FxOpt
) in coordinate system B as function of

eff
CFy Optimum force coefficient in the y-direction
(C
Function of
eff
FyOpt
) in coordinate system B as function of

eff
CFz Optimum force coefficient in the z-direction
(C
Function of
eff
) in coordinate system B as function of
FzOpt eff
CMx Optimum moment coefficient in the x-direction
(C
Function of
eff
MxOpt
) in coordinate system B as function of

eff
CMy Optimum moment coefficient in the y-direction
(C
Function of
eff
MyOpt
) in coordinate system B as function of

eff
CMz Optimum moment coefficient in the z-direction
(C
Function of
eff
MzOpt
) in coordinate system B as function of

eff
Ratio of C /C (R RatioFx




Function of
eff
and
power
Fx FxOpt Fx
) in coordinate system B
as function of
eff
and trim parameter power
Ratio of C /C (R RatioFy
Function of
eff
and
power
Fy FyOpt Fy
) in coordinate system B
as function of
eff
and trim parameter power
Ratio of C /C (R RatioFz
Function of
eff
and
power
Fz FzOpt Fz
) in coordinate system B
as function of
eff
and trim parameter power
RatioMy
Function of
eff
and
power
Ratio of C
My
/C
MyOpt
(R
My
) in coordinate system B
as function of
eff
and trim parameter power
Ratio of C /C RatioMz
Function of
eff
and
power
Mz MzOpt
(R
Mz
) in coordinate system B
as function of
eff
and trim parameter power
UpwindTrim 120 Maximum true wind angle for which trim
parameters are optimised
POWER initial=1;active /
fixed;range=[0:1]
Settings for the trim parameter power

248
YACHT
RESEARCH
UNIT
The
University
of Auckland
C.4 Hydrodynamic Model of Dyna in FRIENDSHIP-Equilibrium
C.4 Hydrodynamic Model of Dyna in FRIENDSHIP-Equilibrium
For all velocity predictions carried out as part of this work the same hydrodynamic models
and input data are used. The data is fundamentally unchanged from the data obtained and
used by Hochkirch (2000). In FRIENDSHIP-Equilibrium the individual force components of
the yacht are described by force modules. The hydrodynamic force modules used to describe
Dyna are listed in Table C.6. A total of eleven hydrodynamic force modules are used, four of
which define the mass of Dyna, three the displacement and four the lift and drag. The
functioning of the different types of force modules is described in the FRIENDSHIP-
Equilibrium user manual (FRIENDSHIP SYSTEMS, 2005) and only the input data for the
force modules is listed in this section. Table C.7 lists the input data for the four mass
modules, which are the mass and the centre of gravity of each component. The IMS-Crew
module shifts the specified mass to the windward side by 45% of the maximum beam. For the
calculations in this work the crew is assumed to remain on the centreline of the yacht since
this was often the case during the full-scale measurements. The maximum beam is therefore
set to zero as shown in Table C.7. The displacement modules calculate the displacement of
hull, keel and rudder from the geometries specified by offset tables. These tables cannot be
listed here but Figure C.2 shows the geometries generated by FRIENDSHIP-Equilibrium
from the input information and used to calculate the hydrostatics. The displacement module
of the hull also calculates the frictional resistance and the pressure drag from the wetted
surface area and the specified form factor, which is given in Table C.8. The residual
resistance and added resistance due to heel are calculated by the GenericHull force module
from the residuary resistance coefficient curve as a function of Froude number shown in
Figure C.3 and the heel coefficient, the reference waterline length and the centre of resistance
listed in Table C.8. For heel angles larger than 30 additional resistance due to the water
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0 0.1 0.2 0.3 0.4 0.5
Froude number [-]
R
e
s
i
d
u
a
r
y

r
e
s
i
s
t
a
n
c
e

c
o
e
f
f
i
c
i
e
n
t

[
-
]
Figure C.2: Geometry of hull, keel and rudder
modelled from offset tables by
FRIENDSHIP-Equilibrium
Figure C.3: Residuary resistance coefficient curve
vs. Froude number used by
FRIENDSHIP-Equilibrium
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 249
Appendix C Velocity Prediction Program Definitions
immersion of the railing is included as well. The input for calculating the propeller drag with
the Propeller force module is also given in Table C.8.
The drag and lift on the keel and rudder are calculated by the Keel and Rudder force
modules based on the information shown in Table C.9. The skin friction is calculated from
the ITTC friction coefficient and modified by a form factor to include the pressure drag. The
form factor is calculated in relation to the maximum thickness and its position based on a
formulation by Hoerner (1965). The induced drag is calculated with equation (2.6) from the
effective aspect ratio and the Oswald efficiency factor (1/e). Frhlich (1997) conducted
towing tank tests of Dyna to measure the keel and rudder forces. From this data Hochkirch
(2000) did not determine a clear dependency of the effective aspect ratio and the Oswald
efficiency factor on boat speed so that they are only modelled as functions of heel angle from
the values shown in Table C.9. The slope of the lift curve is also calculated from the effective
aspect ratio according to Whicker and Fehlner (1958).

Table C.6: Hydrodynamic Force modules used for modelling Dyna in FRIENDSHIP-Equilibrium
Force module
type
Descriptive
name
Description

Mass TK-Bulb-Mass Mass of the standard keel bulb
Mass TK-Root-Mass Mass of the standard keel fin
Mass DynaHull-Mass Mass of Dyna ready to sail excluding keel bulb, keel
fin and crew
IMS-Crew Crew-Mass Mass of the crew
Displacement DynaHull-Disp Displacement of the hull of Dyna, also calculates the
frictional resistance and pressure drag of the hull
Displacement Tiefkiel-Disp Displacement of the standard keel fin and bulb
Displacement Rudder-Disp Displacement of the rudder
GenericHull Hull-Drag Residual resistance of the hull and added resistance
due to heel
Propeller Propeller-Drag Drag of the propeller
Keel Tiefkiel-LiftDrag Lift and drag (induced, skin friction and pressure) of
the standard keel
Rudder Rudder-LiftDrag Lift and drag (induced, skin friction and pressure) of
the rudder

Table C.7: Input data for Mass force module of Dyna
Force module Mass [kg] Centre of mass [m]

MaxBeam [m]
TK-Bulb-Mass 749.5 (-1.453;0.000;-1.445) -
TK-Root-Mass 410 (-1.226;0.000;-0.584) -
DynaHull-Mass 3147 (-2.320;0.000;0.949) -
Crew-Mass 300 (-4.670;0.000;1.579) 0

250
YACHT
RESEARCH
UNIT
The
University
of Auckland
C.4 Hydrodynamic Model of Dyna in FRIENDSHIP-Equilibrium
Table C.8: Additional input data for Displacement, GenericHull and Propeller force modules of
Dyna
Force module Force module type Parameter name

Input
DynaHull-Disp Displacement FormFactor 1.06
CenterOfEffort [m] (1.000;0.000;0.000)
LwlRef [m] 9.5
Hull-Drag GenericHull
HeelCoeff 389.3
CD 1.2
Wake 0.1
Propeller-Drag Propeller
Area [m
2
] 0.01
Center [m] (0.000;0.000;0.000)

Table C.9: Input data for Keel and Rudder force module of Dyna
Parameter Tiefkiel-LiftDrag Rudder-LiftDrag
Force module type Keel Rudder
Lateral [m
2
] 1.55 0.5665
2
WS [m ] 3.1 1.133
Chord [m] 0.91 0.395
Center [m] (-1.132;0.000;-1.037) (-5.750;0.000;-0.455)
ThicknessRatio 0.12 0.065
MaximumThickness 0.4 0.4
Efficiency (1/e) 1.00530 1.39063
=0
1.07125 1.45797
=10
1.26910 1.65998
=20
EffectiveAspectRatio 3.98667 2.17000
=0
3.84315 2.07972
=10
3.41259 1.80889
=20
Wake - 0
ProfileDrag 0 0
StallAngle [] 99 99
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 251

The
University
YACHT
RESEARCH
of Auckland UNIT
D Mathematical Techniques
D.1 General Force and Moment System
A general force and moment system may be represented by the combination of a resultant
force vector F, which intersects the y-plane at point R=(x, 0, z)
T
and a moment M
0
resulting
from pure couples as shown in Figure D.1. The moments from pure couples (M
0
) produce a
moment but have no resultant force.
The moment M about the origin is
, F R M M + =
0 (D.1)
which can be rewritten as
( ) ( ) k j i k j i k j i M
Z Y X Z Y X
F F F z x M M M + + + + + + + = 0
0 0 0

(D.2)
( ) ( ) ( )k j i x F M x F z F M z F M
Y Z Z X Y Y X
+ + + + =
0 0 0
,
so that
z F M M
Y X X
=
0
,
(D.3)
x F z F M M
Z X Y Y
+ =
0
,
(D.4)
x F M M
Y Z Z
+ =
0
.
(D.5)

Z
X
Y
F
x
z
R
M
0


Figure D.1: General force and moment system
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 253
Appendix D Mathematical Techniques
x and z may always be chosen so that M
0X
and M equal zero and hence be written as
Z 0
y
Z
F
M
x = ,
(D.6)
y
X
F
M
z = ,
(D.7)
but M will always remain as
Y 0
z F x F M M
X Z Y Y
+ =
0
.
(D.8)
D.2 Central Axis System
As described by Hochkirch (2000) six-component force balance measurements are usually
represented by a force vector (F) and moment vector (M) about a reference point, which in
this case is the origin of the body fixed coordinate system (coordinate system B). In order to
determine a point in space through which the resultant force is acting it may be convenient to
transform the moment and force system into a central axis system.
The moment M can be represented by a component acting in the resultant force direction
(M
F
) and a component perpendicular to the resultant force (M
V
). M
F
can be determined from
F M M

cos =
F
,
(D.9)
where F is the unit force vector of F obtained from

F
F
F =

,
(D.10)
and is the angle between F and M. From the definition of the dot product it is known that
cos M F M F = .
(D.11)
) M Rearranging equation (D.11) and together with (D.10) substituting into (D.9
F
can be
written as
l
p
2
F
M F
= F F
F
M F
M pl
F
=

=
2
, where .
(D.12)
M
F
is hence related to F through the non-dimensional scalar factor p and the reference length
l, which is here taken as A
0.5
.
S
254
YACHT
RESEARCH
UNIT
The
University
of Auckland
D.2 Central Axis System
The magnitude of the moment component perpendicular to F (|M
V
|) is obtained from
sin M M =
V
.
(D.13)
The definition of the cross product states that
sin M F M F = ,
(D.14)
so that |M
V
| can be rewritten as
F
M F
M

=
V
.
(D.15)
If M
V
is combined with F to form a couple in the plane of F with a magnitude of |M
V
|=|F||a|,
a single force translated parallel through distance |a| is obtained, where |a| can be written as
2
sin
F
M F
F
M
F
M
a

= = =

V
.
(D.16)
The axis a distance |a| from the origin where F now acts is called the central axis. Since F is
parallel to the central axis and also parallel to M
F
, it can be applied anywhere along the
central axis without affecting the moments. The point (a) on the central axis closest to the
origin is hence given by
2
F
M F
a

= ,
(D.17)
and the equation of the central axis can be written as
. ( ) F a c q q + =
(D.18)
M
V
can be obtained by applying F at any position along the central axis. For simplicity q=0 is
chosen so that M
V
is given by
F M F F
F
M F
F a M

2
=

= =
V
.
(D.19)
For sailing yachts it is convenient to define the centre of effort (CoE) on the centreline of the
yacht. With equation (D.17) the CoE can hence be defined as the intersection of the central
axis with the y-plane so that y
CoE
=0 and for |F |>0
Y
X
Y
y
x CoE
F
F
a
a x = ,
(D.20)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 255
Appendix D Mathematical Techniques
and
Z
Y
y
z CoE
F
F
a
a z = .
(D.21)
D.3 Lifting-Line Theory to Model Sail Twist
Lifting-line theory can be used to investigate sail twist by modelling the changes when
departing from the optimum loading that produces constant downwash along the span. The
relationship between the reduction of the centre of effort height and the increase in section lift
coefficient at the bottom of the sail and the increase in induced drag is investigated here.
Notation typically used in lifting-line theory is shown here which is not always consistent
with the notation adhered to in the main body of this work.
Using lifting-line theory it is possible to determine the circulation distribution and the
associated downwash. As shown in Figure D.2 an element of width ds at a position y=s along
the span of a foil sheds an element of vortex sheet (s) which can be treated as a concentrated
vortex line of strength (s)ds. If this vortex line extends to infinity in both directions it
becomes a two-dimensional line vortex, but since the line is only semi-infinite, it induces
exactly half this velocity on the lifting line so that it induces a down-flow at distance y along
the span that is given by
) ( 4
) (
y s
ds s
dw
i

.
(D.22)
The total induced downwash (w
i
) at point y is then obtained by integrating over the span (b)
to give



=

=
2
2
2
2
) ( 4
1
) (
) (
4
1
b
b
b
b
ds
y s
ds dK
ds
y s
s
w
i

.
(D.23)
b/2
(y)ds
-b/2
ds
y
s

Figure D.2: Schematic of lifting line with element of planar trailing vortex sheet (y) of width ds
256
YACHT
RESEARCH
UNIT
The
University
of Auckland
D.3 Lifting-Line Theory to Model Sail Twist
The integral can be evaluated exactly if the loading distribution (circulation) K is written as a
Fourier series so that

=
N
n
n
n B V
b
y K
1
sin
2
) ( cos
2
b
y = , where ,
(D.24)
then

=
N
n
n
i
n n B
V y w
1
sin 4
sin
) (


,
(D.25)
where n is the index of the Fourier series terms, BB
n
is the weighing of each Fourier series term
and V

is the free-stream flow speed. shows the terms n=1 to n=3 of the Fourier
series in the span loading distribution and the downwash. In this idealisation of a sail only the
first few terms are important. When assuming an elliptic loading distribution of the optimally
trimmed sail the terms n=1 and n=2 are used so that C
Figure D.3
L
is affected only by the n=1 term since
the sum of the loading distribution of n=2 between -b/2 and b/2 is zero. Similarly the rolling
moment coefficient (C
MR
) is defined by the n=2 term since the moment of the loading
distribution of n=1 between b/2 and b/2 is zero. Here the optimum loading distribution is
assumed to be semi-elliptical so that the span of the sail is from 0 to b/2 and the terms n=1
and n=3 are used. C is then still given by the n=1 term but C
L MR
is now a function of n=1 and
n=3. C
MR
is also influenced by higher odd terms which are ignored here since their
contribution is relatively small.
The loading on a small span-wise section (dL) of the sail is calculated from the loading
distribution (K) as
dy y K V dL ) (

= ,
(D.26)
-1.5
-1
-0.5
0
0.5
1
1.5
-0.6 -0.4 -0.2 0 0.2 0.4 0.6
y /b [-]
L
o
a
d
i
n
g

d
i
s
t
r
i
b
u
t
i
o
n

[
-
]
n=1
n=2
n=3
-4.5
-3
-1.5
0
1.5
3
4.5
-0.6 -0.4 -0.2 0 0.2 0.4 0.6
y /b [-]
D
o
w
n
w
a
s
h

[
-
]
n=1
n=2
n=3
(b) Fourier series terms in downwash (a) Fourier series terms in span loading distribution
Figure D.3: Fourier series terms of n=1 to n=3 in span loading distribution and downwash
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 257
Appendix D Mathematical Techniques
so that the overall lift (L) over the span b/2 is
4 2
sin
2
) (
1
2
2
0 0
2
1
2
2
2 2



B
b
V d B
b
V dy y K V L
b

= =


.
(D.27)
The life coefficient (C ) can then be calculated in the usual way as
L
AR B
V A
L
C
S
L
2
1
2
2
1

= =

,
(D.28)
2
where A is the area of the sail and the aspect ratio (AR) is defined as (b/2) /A
S S
. Similarly the
rolling moment (M
R
) can be calculated from the loading on a small span-wise section of the
sail as

= =

2 2
0 0
3 1
3
2
cos sin ) 3 sin sin (
2
) (
b
d B B
b
V dy y yK V M
R

=

5 3 2
3 1
3
2
B B b
V M
R
.
(D.29)
The rolling moment coefficient (C
MR
) with the span of the sail as reference length is given by
AR
B B
V A
M
C
S
b
R
MR
2
5 3
3 1
2
2 2
1

+ = =

.
(D.30)
The centre of effort (z
CoE
) as a ratio of the span is now defined in this idealised model as the
ratio of the rolling moment coefficient over the lift coefficient so that
1
3
2
5
4
3
4
B
B
C
C z
L
MR
b
CoE

+ = = .
(D.31)
For semi-elliptical loading the BB
3
term is zero and the centre of effort height at minimum
induced drag is z =4/(3)(b/2). To reduce C
CoEminDi MR
, B
3
B needs to be negative and the
reduction in centre of effort height can then be expressed as
1
3
5
3
1
B
B
z
z
CoEminDi
CoE
+ = .
(D.32)



258
YACHT
RESEARCH
UNIT
The
University
of Auckland
D.3 Lifting-Line Theory to Model Sail Twist
The sectional lift coefficient (C
l
) at the position y along the span can be calculated from dL in
equation (D.26) so that
) 3 sin sin (
) ( ) (
) (
3 1
2
2
1

B B
y c
b
dyV y c
dL
y C
l
+ = =

,
(D.33)
where c(y) is the local chord at position y so that c(y)dy is the reference area. At y=0 the ratio
of the sectional lift coefficient to the sectional lift coefficient at minimum induced drag
(C ) reduces to
lminDi
1
3
1
) 0 (
) 0 (
B
B
C
C
lminDi
l
= .
(D.34)
The term B /B
1
can be replaced with rearranged equation (D.32) B
3
B to yield the relationship
between the reduction in centre of effort height and the increase in sectional lift coefficient at
the symmetry planes as

+ =
CoEminDi
CoE
lminDi
l
z
z
C
C
1
3
5
1
) 0 (
) 0 (
.
(D.35)
Secondly the increase in induced drag due to the reduction in the centre of effort height can
also be calculated. In ideal flow the overall force on an airfoil is the lift force perpendicular to
the local incident flow. Due to the downwash (w
i
) the incident flow does not equal the free-
stream flow so that a force component in the direction of motion is present. This is the
induced drag (D) and it can be calculated from the angle
i i
, at which the lift is inclined to the
free-stream flow at each section along the span, with equations (D.24) and (D.25) as

= = =

2 2 2
0 0 0
) ( ) (
b b b
dy y K y w dL
V
w
dL D
i
i
i i
,

=

2
0
3 1
3 1
2
2
sin ) 3 sin sin (
sin 4
3 sin 3
4 2


d B B
B B b
V D
i
,


( )
2
3
2
1
2
2
3
16 2
B B
b
V D
i
+

=


.
(D.36)
The induced drag coefficient is then calculated by substituting C from equation (D.28) as
L
( )

+ = + = =

2
1
3
2
2
3
2
1
2
2
1
3 1
2
3
8 B
B
AR
C
B B AR
V A
D
C
L
S
i
Di

.
(D.37)
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 259
Appendix D Mathematical Techniques
Finally the term B /B
1
is replaced with rearranged equation (D.32) B
3
B to yield the relationship
between the reduction in centre of effort height and the increase in induced drag coefficient
relative to the minimum induced drag coefficient (C ) as
DiMin
2
1
3
25
1

+ =
CoEminDi
CoE
DiMin
Di
z
z
C
C
.
(D.38)
The value of 25/38.33 is similar but slightly larger than the twist weight constant (c
t
) of 8
determined by Jones (1950). Jones (1950) showed that the increase in C when reducing z
Di CoE

is minimal when w
i
reduces linearly from its maximum at the symmetry plane. Accordingly,
he modelled w as a linear function of y and thereby obtained the optimum c
i t
value of 8. In the
previous calculations the first and third term of the Fourier series are used which do not
model w as a linear function of y, providing B
i
B
3
is unequal zero, as seen from (b) so
that the resulting c
Figure D.3
value is larger than the minimum.
t
In the elliptic case the first two terms of the Fourier series can be used to model a linear
variation in w which results in c of 8. Since only the first term influences C
i t L
and only the
second term influences z
CoE
it can be shown that C
Di
has to increase as more Fourier series
terms are added so that c
t
of 8 is the theoretical minimum. However, in the semi-elliptic case
C is only influenced by the first term if only the odd terms of the Fourier series are used. z
L CoE

is then dependent mainly on the first and third term but higher odd terms have some
influence. As a result c
t
calculated with the first and third term is not the theoretical
minimum. It has been checked analytically and numerically that by adding higher odd terms
with the appropriate B values, c approaches 8, and w B
n t i
approaches a linear variation. This
agrees with the solution by Jones (1950).
D.4 Transformation of Flow Measurements to Heeled Plane
As discussed in section 6.3.1 the change in yaw angle (
Cobra
) and change in pitch angle
(
Cobra
) due to the presence of the sails are measured by a Cobra Probe in the wind tunnel.
For test cases where the model is heeled, and
Cobra Cobra
need to be transformed into the
body fixed coordinate system to obtain the change in effective angle (
eff
) and the
inclination angle of the flow relative to the heeled plane (z-plane in coordinate system B).



260
YACHT
RESEARCH
UNIT
The
University
of Auckland
D.4 Transformation of Flow Measurements to Heeled Plane
For the transformation it is convenient to define the direction of the influenced flow relative
to the free-stream tunnel flow from
Cobra
and as the unit vector with
Cobra
V

Cobra
,



=
Cobra
Cobra Cobra
Cobra Cobra
Cobra



sin
sin cos
cos cos

V
(D.39)
where the sign of the x-component and y-component are reversed to transform the
measurements into a right hand system where the positive x-direction is into the flow. The
free-stream flow can be assumed to be horizontal so that the wind direction vector in the
absolute coordinate system ( ) is obtained by rotating through the apparent
wind angle. From the wind direction vector in the body fixed coordinate system
( ) is then obtained by
CobraA
V

Cobra
V

CobraA
V

CobraB
V

,
Cobra A m m AB CobraA m m AB CobraB
V T T V T V

) ( ) 0 , (

) 0 , (

= = = =
(D.40)
where T
AB
is the transformation matrix defined in equation (2.17) to convert from the
absolute coordinate system to the body fixed coordinate system by rotating through the model
heel angle (
m
) and pitch angle (
m
), which is zero in the wind tunnel. T

is the
transformation matrix to rotate about the z-axis by an angle and is defined by
,

=
1 0 0
0 cos sin
0 sin cos

T
(D.41)
where is the apparent wind angle (
A
) in this case. All heeled tests are conducted with the
upwind tunnel configuration so that no correction for twist is required and
A
is simply the
apparent wind angle set in the wind tunnel. It needs to be remembered however that all wind
tunnel tests are conducted on port tack and all
A
are negative. The coordinate systems
definitions are shown in Figure 2.19. The effective wind angle measured by the Cobra Probe
( ) is obtained using equation (2.18) so that
effCobra
[ ]
o
180 , 0 ,

tan
1

=

effCobra
CobraBx
CobraBy
effCobra
V
V
,
(D.42)
where is the x-component and the y-component of the unit velocity vector
in equation
CobraBx
V

CobraBy
V

CobraB
V

(D.40). The change in effective wind angle (


eff
) due to the presence of
the sails is the difference between and the free-stream effective wind angle (
effCobra eff
).

Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 261
Appendix D Mathematical Techniques
Since the free-stream flow is horizontal and the wind tunnel model does not pitch,
eff
can be
calculated using equation (2.11) so that
eff
can be written as
( ) [ ]
o
180 , 0 , cos tan tan
1
= =

eff m A effCobra eff effCobra eff
,
(D.43)
where
A
is negative for all wind tunnel tests since the model is tested on port tack. Similarly
the inclination angle of the flow measured by the Cobra Probe () relative to the deck plane
(z-plane in coordinate system B) is given by
,
CobraBz
V

sin
1
=
(D.44)
where is the z-component of the unit velocity vector in equation
CobraB
V

CobraBz
V

(D.40).
262
YACHT
RESEARCH
UNIT
The
University
of Auckland

The
University
YACHT
RESEARCH
of Auckland UNIT
References

Abbott, I. H., and von Doenhoff, A. E. (1959). Theory of Wing Sections, Dover Publications
Inc., New York.
Barlow, J. B., Rae, W. H., and Pope, A. (1999). Low-Speed Wind Tunnel Testing, 3rd Ed.,
John Wiley & Sons, New York.
Batt, K. (2006). "Wind Shear." Autralian Government, Bureau of Meteorology.
Benzie, A. D. (2001). "The Development of Test Procedures for the Wind Tunnel Testing of
High Performance Yachts," MEng Thesis, University of Aberdeen, Aberdeen.
Bonniot, G. (2002). "Work Experience at the Twisted Flow Wind Tunnel." The University of
Auckland, Auckland.
Brandt, H., and Hochkirch, K. (1999). "Fullscale Hydrodynamic Force Measurements on the
Berlin Sail-Force-Dynamometer." The 14th Chesapeake Sailing Yacht Symposium,
SNAME, Annapolis.
Brandt, H., and Hochkirch, K. (2000). "Entwicklung eines Segeldynamometers zur Erfassung
der aero- und hydrodynamischen Kraefte." Technical University Berlin, Berlin.
Campbell, I. M. C. (1997). "Optimisation of a Sailing Rig using Wind Tunnel Data." The 13th
Chesapeake Sailing Yacht Symposium, SNAME, Annapolis.
Campbell, I. M. C. (1998). "The Performance of Offwind Sails Obtained from Wind Tunnel
Tests." The Modern Yacht Conference, RINA, Portsmouth.
Claughton, A. (1999). "Developments in the IMS VPP Formulations." The 14th Chesapeake
Sailing Yacht Symposium, SNAME, Annapolis.
Clauss, G., Heisen, W., and Fliege, F. (2005). "CFD Analysis On The Flying Shape of
Modern Yacht Sails." International Congress of International Maritime Association
of the Mediterranean, IMAM, Lisbon.
Clauss, G., and Hochkirch, K. (2003). "Entwicklung hydrodynamischer Grundlagen fuer den
Entwurf con Hochleistungsflachkielen (Development of Hydrodynamic Principles for
the Design of Shallow Draft Keels)." Technical University Berlin, Berlin.
Collie, S., and Gerritsen, M. (2006). "The Challenging Turbulent Flows Past Downwind
Yacht Sails and the Practical Application of CFD to them." 2nd High Performance
Yacht Design Conference, RINA, Auckland.
Cook, N. J. (1985). The Designer's Guide to Wind Loading of Building Structures, Part 1,
Butterworths, London.
Davidson, K. S. M. (1936). "Some Experimental Studies of the Sailing Yacht." Society of
Naval Architects and Marine Engineers, Vol. 44.
Davis, J. C. (2002). Statistics and Data Analysis in Geology, 3rd Ed., John Wiley & Sons,
New York.
Day, S., Letizia, L., and Stuart, A. (2002). "VPP vs PPP: Challenges in the Time-Domain
Prediction of Sailing Yacht Performance." High Performance Yacht Design
Conference, RINA, Auckland.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 263
References
Ekblom, P. (2002). "Improvement of the Twisted Flow Wind Tunnel Yacht Testing
Techniques," Royal Institute of Technology, Stockholm.
Fargeas, S., and Kouyoumdjian, J. (1997). "BATOPERF, A Performance Prediciton Software
and Its Influence on Modern Yacht Design." The 13th Chesapeake Sailing Yacht
Symposium, SNAME, Annapolis.
Flay, R. G. J. (1996). "A twisted flow wind tunnel for testing yacht sails." Journal of Wind
Engineering and Industrial Aerodynamics, Vol. 63, pp. 171-182.
Flay, R. G. J., and Jackson, P. S. (1992). "Flow Simulation for Wind-Tunnel Studies of Sail
Aerodynamics." Journal of Wind Engineering and Industrial Aerodynamics, Vol.
44(Part 4), pp. 2703-2714.
Flay, R. G. J., Locke, N. J., and Mallinson, G. D. (1996). "Model Tests of Twisted Flow
Wind Tunnel Designs for Yacht Testing." Journal of Wind Engineering and Industrial
Aerodynamics, Vol. 63, pp. 155-169.
Flodn, D., and Johansson, P. (2001). "Effects of Heel and Twist in Velocity Prediction
Programs," ME Thesis, Chalmers University of Technology, Gteborg, Sweden.
Fossati, F., Muggiasca, S., Viola, I. M., and Zasso, A. (2006). "Wind Tunnel Techniques for
Investigation and Optimization of Sailing Yachts Aerodynamics." 2nd High
Performance Yacht Design Conference, RINA, Auckland.
FRIENDSHIP SYSTEMS. (2005). "FRIENDSHIP-Equilibrium User Manual."
FRIENDSHIP-Systems GmbH, Potsdam, Germany.
Frhlich, M. (1997). "Optimierung von Kielen einschlielich Rumpf fr Segelyachten auf der
Basis eines numerischen Rechenverfahrens fr viskose und instationre Strmung."
Project 1022, Schiffbau-Versuchsanstalt Potsdam GmbH, Potsdam, Germany.
Gerritsma, J., Omnink, R., and Verslius, A. (1981). "Geometry, Resistance and Stability of
the Delft Systematic Yacht Hull Series." International Shipbuilding Progress.
Graf, K., and Renzsch, H. (2006). "RANSE Investigation of Downwind Sails and Integration
into Sailing Yacht Design Processes." 2nd High Performance Yacht Design
Conference, RINA, Auckland.
Hansen, H., Jackson, P. S., and Hochkirch, K. (2002). "Comparison of Wind Tunnel and Full-
Scale Aerodynamic Sail Force Measurements." High Performance Yacht Design
Conference, RINA, Auckland.
Hansen, H., Jackson, P. S., and Hochkirch, K. (2003a). "Comparison of Wind Tunnel and
Full-Scale Aerodynamic Sail Force." International Journal of Small Craft Technology
(IJSCT), Vol. 145 Part B1.
Hansen, H., Jackson, P. S., and Hochkirch, K. (2003b). "Real-Time Velocity Prediction
Program for Wind Tunnel Testing of Sailing Yachts." The Modern Yacht Conference,
RINA, Southampton.
Hansen, H., Richards, P. J., and Hochkirch, K. (2005). "Advances in the Wind Tunnel
Analysis of Yacht Sails." 26th Symposium on Yacht Design and Construction,
Deutscher Boots- und Schiffbauer-Verband, Hamburg.
Hansen, H., Richards, P. J., and Jackson, P. S. (2006). "An Investigation of Aerodynamic
Force Modelling for Yacht Sails using Wind Tunnel Techniques." 2nd High
Performance Yacht Design Conference, RINA, Auckland.
Hawkins, P. (1998). "Non-Dimensional Number Effects on Downwind Sail Modelling," ME
Thesis, The University of Auckland, Auckland.
264
YACHT
RESEARCH
UNIT
The
University
of Auckland

Hazen, G. S. (1980). "A Model of Sail Aerodynamics for Diverse Rig Types." New England
Sailing Yacht Symposium, SNAME, Lew London, Connecticut.
Herman, J. S. (1988). "A Sail Force Dynamometer: Design, Implementation and Data
Handling," MSc Thesis, Massachusetts Institute of Technology, Cambridge.
Hochkirch, K. (2000). "Entwicklung einer Meyacht zur Analyse der Segelleistung im
Orginalmastab (Design and Construction of a Full-Scale Measurement System for
the Analysis of Sailing Performance)," PhD Thesis, Technical University Berlin,
Berlin.
Hoerner, S. F. (1965). Fluid-Dynamic Drag, Hoerner Fluid Dynamics, Albuquerque.
Hoerner, S. F., and Borst, H. V. (1985). Fluid-Dynamic Lift, Hoerner Fluid Dynamics,
Albuquerque.
Jackson, P. S. (2001). "An Improved Upwind Sail Model for VPPs." The 15th Chesapeake
Sailing Yacht Symposium, SNAME, Annapolis.
Jones, R. T. (1950). "The Spanwise Distribution of Lift for Minimum Induced Drag of Wings
Having a Given Lift and a Given Bending Moment." Report No. 2249, National
Advisory Committee for Aeronautics (NACA), Washington.
Kerwin, J. E. (1978). "A Velocity Prediction Program for Ocean Racing Yachts revised to
February 1978." Report No. 78-11, Massachusetts Institute of Technology,
Cambridge.
Kiffer, O. (2003). "Blockage Correction of the Twisted Flow Wind Tunnel," Project Report,
The University of Auckland, Auckland.
Klein, A. K. (1990). "A Comparison of Experimental and Theoretical Sail Forces," MSc
Thesis, Massachusetts Institute of Technology, Cambridge.
Krebber, B. (2005). "Numerische Untersuchungen des Einflusses von Trimmparametern an
einem Yachtrigg," Studienarbeit, Technical University Berlin, Berlin.
Krebber, B., and Hochkirch, K. (2006). "Numerical Investigation on the Effects of Trim for a
Yacht Rig." 2nd High Performance Yacht Design Conference, RINA, Auckland.
Larsson, L., and Eliasson, R. E. (1994). Principles of Yacht Design, International Marine,
Camden.
Le Pelley, D. J., and Hansen, H. (2003). "An Investigation into the Effects of Heel on
Downwind Sails." The Modern Yacht Conference, RINA, Southampton.
Locke, N. J. (1994). "Lift and Drag Distributions of Sails from Wind Tunnel Wake Surveys,"
ME Thesis, The University of Auckland, Auckland.
Marchaj, C. A. (1988). Aero-Hydrodynamics of Sailing, 2nd Ed., Adlard Coles, London.
Martin, D. E., and Beck, R. F. (2001). "PCSAIL, A Velocity Prediction Program for a Home
Computer." The 15th Chesapeake Sailing Yacht Symposium, SNAME, Annapolis.
Masuyama, Y., and Fukasawa, T. (1997). "Full Scale Measurement of Sail Force and the
Validation of Numerical Calculation Method." The 13th Chesapeake Sailing Yacht
Symposium, SNAME, Annapolis.
Milgram, J. H. (1993). "Naval Architecture Used in Winning the 1992 America's Cup
Match." Centennial Meeting, SNAME, New York.
Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 265
References
Milgram, J. H., Peters, D. B., and Eckhouse, D. N. (1993). "Modeling IACC Sail Forces by
Combining Measurements with CFD." The 11th Chesapeake Sailing Yacht
Symposium, SNAME, Annapolis.
Mller, O., and Graf, K. (2005). "Der Twist-Flow-Windkanal der Yacht Research Unit Kiel."
26th Symposium on Yacht Design and Construction, Deutscher Boots- und
Schiffbauer-Verband, Hamburg.
National Instruments. (2003). "LabVIEW User Manual - Using External Code in LabVIEW."
National Instruments, Austin, Texas.
Peters, D. B. (1992). "Determination of Force Coefficients For Racing Yacht Sails Using
Actual Force Measurements," MSc Thesis, Massachusetts Institute of Technology,
Cambridge.
Poor, C. L. (1986). "A Description of the New International Rating System." Publications of
the United States Yacht Racing Union, USYRU, Washington D.C.
Ranzenbach, R., and Kleene, J. (2002). "Utility of Flying Shapes in the Development of
Offwind Sail Design Databases." High Performance Yacht Design Conference, RINA,
Auckland.
Ranzenbach, R., and Mairs, C. (1997). "Experimental Determination of Sail Performance and
Blockage Corrections." The 13th Chesapeake Sailing Yacht Symposium, SNAME,
Annapolis.
Ranzenbach, R., and Mairs, C. (1999). "Wind Tunnel Testing of Offwind Sails." The 14th
Chesapeake Sailing Yacht Symposium, SNAME, Annapolis.
Ranzenbach, R., and Teeters, J. (2002). "Enhanced Depowering Model for Offwind Sails."
High Performance Yacht Design Conference, RINA, Auckland.
Richards, P. J., and Hoxey, R. P. (1992). "Computational and Wind Tunnel Modelling of
Mean Wind Loads on the Silsoe Structures Building." Journal of Wind Engineering
and Industrial Aerodynamics, Vol. 41(1-3), pp 1641-1652.
Richards, P. J., Le Pelley, D. J., Cazala, A., McCarty, M., Hansen, H., and Moore, W. (2006).
"The Use of Independent Supports and Semi-Rigid Sails in Wind Tunnel Studies."
2nd High Performance Yacht Design Conference, RINA, Auckland.
Richards, P. J., Orso Fiet, G., Breteau, X., and Jorritsma, M. (2004). "Wind tunnel modelling
- why move a wind tunnel?" 11th Australasian Wind Engineering Society Workshop,
Australasian Wind Engineering Society, Darwin, Australia.
Richards, P. J., and Wanigaratne, B. S. (1993). "A Comparison of Computer and Wind-
Tunnel Models of Turbulence around the Silsoe Structures Building." Journal of Wind
Engineering and Industrial Aerodynamics, Vol. 46-47, pp 439-447.
Richardt, T., Harries, S., and Hochkirch, K. (2005). "Maneuvering Simulations for Ships and
Sailing Yachts using FRIENDSHIP-Equilibrium as an Open Modular Workbench."
International Euro Conference on Computer Applications and Information
Technology in the Marine Industries (COMPIT), Hamburg.
Roux, Y., Huberson, S., Hauville, F., Boin, J. P., Guilbaud, M., and Ba, M. (2002). "Yacht
Performance Prediction: Towards A Numerical VPP." High Performance Yacht
Design Conference, RINA, Auckland.
Schlageter, E. C., and Teeters, J. R. (1993). "Performance Prediction Software for IACC
Yachts." The 11th Chesapeake Sailing Yacht Symposium, SNAME, Annapolis.
266
YACHT
RESEARCH
UNIT
The
University
of Auckland

Schmidt. (1998). "Schmidt Flow Sensor SS 20.01 Information." Schmidt Feintechnik GmbH,
St. Georgen, Germany.
Simiu, E., and Scanlan, R. H. (1996). Wind Effects on Structures, 3rd Ed., John Wiley &
Sons, New York.
van Oossanen, P. (1993). "Predicting the Speed of Sailing Yachts." Report No. 93-004, Van
Oossanen & Associates, Wageningen, The Netherlands.
van Oossanen, P. (1995). "Improvements of Sailing Yacht Performance Prediction by
Including Force-Moment Equilibrium for the Calculation of Helm Angle in a Velocity
Prediction Program." The 12th Chesapeake Sailing Yacht Symposium, SNAME,
Annapolis.
Whicker, L. F., and Fehlner, L. F. (1958). "Free-Stream Characteristics of a Family of Low-
Aspect Ratio, all Moveable Control Surfaces for Application to Ship Design." Report
No. 933, David Taylor Model Basin, Washington D.C.
White, J., and Wilson, J. (2000). "A Study of the Effect of Heel on Upwind Sails," Project
Reports ME63 and ME65, The University of Auckland, Auckland.


Hansen: Enhanced Wind Tunnel Techniques and Aerodynamic Force Models for Yacht Sails 267

You might also like