30.10.2016 Views

EVALUATION OF MULTI-ELEMENT WING SAIL

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

5 th High Performance Yacht Design Conference<br />

Auckland, 10-12 March, 2015<br />

<strong>EVALUATION</strong> <strong>OF</strong> <strong>MULTI</strong>-<strong>ELEMENT</strong> <strong>WING</strong> <strong>SAIL</strong> AERODYNAMICS FROM<br />

TWO-DIMENSIONAL WIND TUNNEL INVESTIGATIONS<br />

Alexander W. Blakeley 1 , abla089@aucklanduni.ac.nz<br />

Richard G.J. Flay 2 , r.flay@auckland.ac.nz<br />

Hiroyuki Furukawa 3 , furukawa@meijo-u.ac.jp<br />

Peter J. Richards 4 , pj.richards@auckland.ac.nz<br />

Abstract. Following the 33rd America's Cup which featured a trimaran versus a catamaran, and the recent 34th America's Cup in<br />

2013 featuring AC72 catamarans with multi-element wing sail yachts sailing at unprecedented speeds, interest in wing sail<br />

technology has increased substantially. Unfortunately there is currently very little open peer-reviewed literature available with a<br />

focus on multi-element wing design for yachts. The limited available literature focuses primarily on the structures of wings and their<br />

control, rather than on the aerodynamic design. While there is substantial available literature on the aerodynamic properties of<br />

aircraft wings, the differences in the flow domains between aeroplanes and yachts is significant. A yacht sail will operate in a<br />

Reynolds number range of 0.2 to 8 million while aircraft operate regularly in excess of 10 million. Furthermore, yachts operate in the<br />

turbulent atmospheric boundary layer and require high maximum lift coefficients at many apparent wind angles, and minimising drag<br />

is not so critical. This paper reviews the literature on wing sail design for high performance yachts and discusses the results of wind<br />

tunnel testing at the Yacht Research Unit at the University of Auckland. Two wings with different symmetrical profiles have been<br />

tested at low Reynolds numbers with surface pressure measurements to measure the effect of gap geometry, angle of attack and<br />

camber on a wing sail’s performance characteristic. It has been found that for the two element wing studied, the gap size and pivot<br />

point of the rear element have only a weak influence on the lift and drag coefficients. Reynolds number has a strong effect on<br />

separation for highly cambered foils.<br />

NOMENCLATURE<br />

AWA<br />

c<br />

C l<br />

C d<br />

D<br />

g<br />

L<br />

R<br />

S<br />

TWA<br />

V A<br />

V mg<br />

V S<br />

V T<br />

<br />

<br />

<br />

<br />

A<br />

H<br />

<br />

Apparent wind angle<br />

Chord<br />

Lift coefficient<br />

Drag coefficient<br />

Drag on wing and any sails<br />

Gap<br />

Lift on wing and any sails<br />

Resistance of hull and appendages<br />

Side force on hull and appendages<br />

True wind angle<br />

Apparent wind velocity<br />

Velocity made good<br />

Yacht velocity<br />

True wind velocity<br />

Angle of attack of front element<br />

Angle between rear and front elements<br />

Apparent wind angle<br />

True wind angle<br />

Aerodynamic drag angle<br />

Hydrodynamic drag angle<br />

Leeway angle<br />

1. INTRODUCTION<br />

Following the 33rd America's Cup in 2010 which<br />

featured a trimaran versus a catamaran, and the recent<br />

34th America's Cup in 2013 featuring AC72 catamarans<br />

with multi-element wing sail yachts sailing at<br />

unprecedented speeds, interest in wing sail technology<br />

has increased substantially. Unfortunately there is<br />

currently very little open peer-reviewed literature<br />

available with a focus on multi-element wing design for<br />

yachts. Hence a wind tunnel study of multi-element wing<br />

sails was carried out by the University of Auckland.<br />

Some preliminary results of this research are available in<br />

[1].<br />

Solid wing sails offer several advantages over flexible<br />

sails for high speed sailing. One advantage is that the<br />

wing has an internal structure which is used to give the<br />

wing its shape. The shape is not dependent on the tension<br />

in the lines at the sail corners. The shape of a flexible sail<br />

cloth mainsail is highly dependent on the tension in the<br />

mainsheet, particular the vertical component which is<br />

required to keep the leech from twisting excessively. In a<br />

wing sail, this vertical component is eliminated<br />

completely from the mainsheet, which is used simply to<br />

alter the angle of attack using a horizontal force. The<br />

twist in the solid sail is controlled by internal control<br />

lines which are under considerably less tension than a<br />

conventional mainsheet. Thus the angle of attack can be<br />

changed relatively easily and quickly, and the power of<br />

the sail can be controlled much more easily than in a<br />

conventional soft sail. This makes sailing a high<br />

performance catamaran much easier with a wing sail than<br />

a soft flexible sail.<br />

1 PhD Student, Yacht Research Unit, Department of Mechanical Engineering, University of Auckland, NZ<br />

2 Professor, Director Yacht Research Unit, Department of Mechanical Engineering, University of Auckland, NZ<br />

3 Associate Professor, Department of Mechanical Engineering, Meijo University, Japan<br />

4 Associate Professor, Department of Mechanical Engineering, University of Auckland, NZ<br />

37


Multi-hull yachts have considerably less hydrodynamic<br />

drag than mono-hulls since they have much lower<br />

displacements than monohulls of similar length. This is<br />

because they derive their righting moment from their<br />

wide beam, and not from having a heavy keel.<br />

Furthermore they heel only a small amount. A dangerous<br />

feature of multi-hulls is that once the overturning<br />

moment exceeds their maximum righting moment, they<br />

will overturn unless the overturning moment is reduced<br />

quickly. When coupled with a high performing solid sail<br />

or wing, this means that they are able to sail with small<br />

apparent wind angles both upwind and downwind as<br />

illustrated in Error! Reference source not found.. It<br />

was shown by Lanchester [2] that the minimum angle<br />

( A ) at which a boat can shape its course relative to the<br />

wind is the sum of the under and above water drag<br />

angles, namely H and A . A wing sail operating at low<br />

angles of attack has a low aerodynamic drag angle, and<br />

so the result of this is that high performance multi-hulls<br />

with wing sails operate at low apparent wind angles in<br />

both upwind and downwind sailing. These angles are<br />

illustrated in Error! Reference source not found.. This<br />

means that it is only necessary to have aerodynamic data<br />

for wing sails for low angles of attack.<br />

V A<br />

A<br />

V S<br />

Upwind<br />

V T<br />

T<br />

V mg<br />

V mg<br />

Figure 1 Velocity triangles for up and downwind sailing<br />

showing the low apparent wind angle A which occurs for<br />

fast multi-hulls with wing sails.<br />

Understanding the performance of multi-hulls can be<br />

enhanced by analysing a free-body diagram of the yacht<br />

and considering the forces in the horizontal plane. This is<br />

close to reality, since they do not heel much. The<br />

aerodynamic and hydrodynamic forces on a catamaran<br />

are shown in Error! Reference source not found.. Sail<br />

forces are lift (L) and (D) being normal and parallel to<br />

the apparent wind direction, and hull/appendage forces<br />

are sideforce (S) and resistance (R) which are normal and<br />

parallel to the yacht’s course through the water. There<br />

may be a small leeway angle (). Note that for<br />

equilibrium, the hydrodynamic and aerodynamic forces<br />

in the plane of the water must be equal and opposite, so<br />

. Furthermore, as noted above,<br />

. Reducing D and R lead to reductions in<br />

and and thus to a reduction in AWA, which will lead<br />

V S<br />

V A<br />

A<br />

Downwind<br />

V T<br />

T<br />

to increased velocity made good (V mg ) to the top or<br />

bottom mark.<br />

S<br />

H<br />

Error! Reference source not found.Figure 2 Sketch of<br />

horizontal forces acting on a catamaran showing the<br />

important directions.<br />

Ignoring the structural aspects and considering primarily<br />

the aerodynamic aspects, high speed multi-hull yachts<br />

with wing sails need wings that are symmetrical so that<br />

they can tack and gybe at will. They need to be able to<br />

develop high lift and have low drag. They need to be able<br />

to twist so that the height of the thrust can be lowered<br />

when required to reduce the overturning moment. From<br />

observations of existing vessels, it is apparent that severe<br />

twist is important so that the direction of the lift force can<br />

be reversed at the top of the wing to provide a righting<br />

moment. Generally it is apparent from the 2013<br />

America’s Cup yachts that for a two-element wing, the<br />

chords of the main wing and flap are usually of similar<br />

length, and the gap between them is relatively small. The<br />

first element is thick so that it can take the required<br />

structural loads, while the flap is thin. Typical thickness<br />

ratios are 25% and 9% for the first and second elements<br />

respectively.<br />

2. SOME THEORETICAL CONSIDERATIONS<br />

By applying the sine rule to the velocity triangles in<br />

Figure 1 it is evident that<br />

The velocity made good is given by<br />

R<br />

, so (1)<br />

(2)<br />

(3)<br />

Thus to improve performance (i.e. increase the ratios<br />

and ) at a fixed heading (i.e. fixed ) it is necessary to<br />

reduce . Figure 3 shows curves of the ratio for<br />

constant and for various , which are evidently<br />

circular arcs. Figure 3 shows clearly how smaller<br />

<br />

L<br />

D<br />

A<br />

38


gives more speed. Maximum yacht speed occurs when<br />

and it is the diameter of the circule. The<br />

maximum velocity made good is determined from the<br />

horizontal tangents to the circle, and occurs for the<br />

angles and for upwind and<br />

downwind respectively.<br />

Thus the importance of being able to determine and<br />

minimise the aerodynamic drag angle has been<br />

demonstrated. Wing performance between candidate<br />

options can therefore be compared on the basis of the<br />

aerodynamic drag angle as it influences directly.<br />

1<br />

0.5<br />

0<br />

‐0.5<br />

V T<br />

‐1<br />

‐1.5<br />

‐2<br />

T<br />

V S<br />

A<br />

V A<br />

bA = 25 deg<br />

bA = 45 deg<br />

bA = 90 deg<br />

bA = 150 deg<br />

0 0.5 1 1.5 2 2.5<br />

Figure 3 Velocity triangles (speed polars) for fixed values of<br />

A for a range of T .<br />

3. LITERATURE REVIEW<br />

The limited available literature on wing sails focuses<br />

primarily on their structure and control, rather than on<br />

their aerodynamic design. While there is substantial<br />

available literature on the aerodynamic properties of<br />

aircraft wings, the differences in the flow domains<br />

between aeroplanes and yachts is significant. A yacht sail<br />

will operate in a Reynolds number range of 0.2 to 8<br />

million while aircraft operate regularly in excess of 10<br />

million. Furthermore, yachts operate in the turbulent<br />

atmospheric boundary layer and require high maximum<br />

lift coefficients at many apparent wind angles. A brief<br />

review of some of the literature on wing sails is<br />

presented below.<br />

Elkaim [3] at the University of California undertook a<br />

large amount of work on wing sails while involved in the<br />

Atlantis Project, which focused on designing an<br />

Autonomous Marine Surface Vehicle for a wide range of<br />

applications. A rigid wing sail was used on this vehicle<br />

because it was regarded as more aerodynamically<br />

efficient than a soft sail, by pivoting it near the centre of<br />

pressure it required less force to actuate than a regular<br />

soft sail, and it could be designed to be self-trimming.<br />

Marchaj [4] reviewed some research on wing sails,<br />

pointing out their advantages in some situations. In<br />

discussing the 1976 C-Class race between Aquarius V<br />

with a soft sail, and Miss Nylex with an advanced wing<br />

sail, he pointed out the advantage of the soft sail with the<br />

lighter rig in light conditions, compared to the more<br />

aerodynamically efficient but heavier wing sail rig.<br />

Some wing sails have been made of double-surface soft<br />

sails, and are a good idea for cruising as they can be<br />

reefed and dropped when wind conditions are<br />

unfavourable, e.g. the Wally and Omer wing sail design<br />

[5]. They do not have all the benefits of rigid wing sails,<br />

e.g. the leech tension needs to be maintained, requiring a<br />

sophisticated boom and powerful mainsheet.<br />

C-Class catamaran racing has often produced very<br />

innovative ideas regarding wing sails. Examples are the<br />

slot gap on the yacht Cogito, described by MacLane [6],<br />

which could be changed onshore, but was fixed once<br />

racing. Killing and Clarke [7] describe further<br />

enhancements of C-Class wings for the yachts Alpha and<br />

Rocker, which had more advanced cross-sectional shapes<br />

than Cogito which were developed using software<br />

developed by Drela at MIT [8,9].<br />

There are maximum heeling and pitching moments that<br />

can be applied to catamarans before they tip over. Thus<br />

in strong wind conditions it is advantageous to lower the<br />

height of the line of action of the side and thrust forces.<br />

Thus it can be shown that the optimum lift distribution<br />

for a catamaran is such that it changes sign at the top of<br />

the wing, requiring large amounts of wing twist [10].<br />

Wing sails can be designed with mechanisms for<br />

producing such large amounts of twist, which is<br />

impossible for soft sails, and so this is another significant<br />

difference between wing and soft sails.<br />

One of the commonest areas of wing sail use is for highspeed<br />

land yacht sailing, more specifically ice yachts.<br />

These yachts skate on ice and are sometimes powered by<br />

large wing sails. They are capable of very high speeds.<br />

The Greenbird project is arguably one of the most<br />

successful versions of wind powered land/ice yachts. On<br />

land, the Greenbird holds the land speed record of 126.2<br />

mph, while on ice, they hope to beat the current record of<br />

84 mph by setting a speed equal or higher than the<br />

126.2mph that they set on land [11].<br />

Recently relatively small yachts have been<br />

experimenting with wing sails. The X-Wing sail project<br />

[12] is a 2-element wing designed for small sailing<br />

dinghies, specifically the “Sunfish” class. The designer<br />

was inspired by the 2010 America’s Cup and has<br />

designed a low-cost wing for those wanting to<br />

experiment with the technology.<br />

39


The literature cited above shows that work on wing sail<br />

development has taken place for a variety of reasons, but<br />

mainly to obtain high speeds or to improve control, and<br />

that developments have occurred in both small as well as<br />

larger craft. However, it is evident that there is a paucity<br />

of multi-element wing aerofoil data obtained specifically<br />

for yacht sail design. This paper discusses the results of<br />

wind tunnel testing at the Yacht Research Unit at the<br />

University of Auckland to help provide this information.<br />

4. CFD OPTIONS<br />

To analyse the flow around upwind sails, the potential<br />

flow assumption and the panel method are basic<br />

methods, but they cannot model viscous effects which<br />

are dominant in locations of flow separation. The Navier-<br />

Stokes equations can be solved directly (DNS) in cases<br />

with flow separation phenomena. When the Reynolds<br />

number is high and the flow becomes turbulent, it is<br />

difficult to use DNS because it needs an enormous<br />

number of grid points. The turbulence must be modelled<br />

with turbulence or sub-grid models. Reynolds Average<br />

Navier Stokes (RANS) turbulence models obtain a<br />

stationary solution that is representative of the mean<br />

flow. They are robust and easy to use, but cannot be<br />

adopted to unsteady flow. Large-eddy simulation (LES)<br />

is used to model the larger, three-dimensional unsteady<br />

turbulent motions. In LES, the smaller scale motions are<br />

filtered and subsequently modelled. LES can be expected<br />

to be more accurate and reliable than RANS models for<br />

flows where large-scale unsteadiness is significant.<br />

The program MSES [8] is a coupled viscid/inviscid Euler<br />

method. It solves the Euler equations on a discrete two<br />

dimensional grid, coupled with an integral boundary<br />

layer formation. MSES was created to analyse the flow<br />

around an aerofoil more accurately than inviscid codes<br />

by incorporating viscous effects and a boundary layer<br />

formation into its solutions. Therefore it theoretically<br />

offers more realistic results than a vortex lattice method<br />

(VLM), although it possibly requires more computational<br />

power. A derivation of MSES is the MISES [9] code,<br />

which has been developed to determine the flow around<br />

multi-element aerofoils.<br />

While CFD analysis of wing sails is highly advanced and<br />

capable of accurate predictions, it is not easy to predict<br />

sail performance if there is flow separation. As<br />

mentioned in the present paper, it is important for<br />

designers of wing sails to know where flow separation<br />

occurs, and physical testing such as wind tunnel testing<br />

can help in this regard. Wind tunnel testing can also give<br />

good information for checking the output from CFD<br />

codes, although they may be unable to give results for the<br />

very high Reynolds numbers appropriate for large yachts<br />

sailing very fast, e.g. America’s Cup AC72 yachts.<br />

However, the Reynolds numbers from wind tunnels<br />

would be commensurate to those for smaller C-class<br />

yacht wing sails.<br />

CFD analysis of wing sails is not considered further in<br />

the present paper.<br />

5. EXPERIMENTAL SETUP<br />

5.1 Wind Tunnel<br />

The experiments were performed in the University of<br />

Auckland open-return wind tunnel. The wind tunnel has<br />

been specifically designed for testing yacht sails and has<br />

a standard operating cross section of 7 m wide by 3.5 m<br />

high. The flow is produced by two 3-m diameter 4-<br />

bladed fans and then driven through a 1 m thick<br />

honeycomb screen and two tight mesh screens to remove<br />

swirl and to give a uniform velocity profile and relatively<br />

low turbulence flow. Permanent pitot-static probes set up<br />

well upstream of the wing recorded the dynamic and<br />

static pressures whilst another probe measured the<br />

atmospheric pressure outside the working section.<br />

The wind tunnel walls were brought inwards to give a<br />

2.5m wide by 3.5m high test section and the wing models<br />

were located near the outlet of the nozzle as shown in<br />

Figure 4.<br />

Figure 4 Photograph of wing model in wind tunnel showing<br />

base of support struts and endplate marking for angle<br />

fixing.<br />

5.2 Multi-element Wing Construction<br />

The multi-element wing used in the wind tunnel<br />

comprised a NACA 0025 main element and a NACA<br />

0009 aft element with a 50:50 chord ratio. The wing<br />

spans 2.5 m across the width of the wind tunnel and has a<br />

chord length of 1 m (with zero aft element deflection and<br />

no gap). 20 mm thick MDF boards at each end of the<br />

model contained the placement locations for the elements<br />

and tested configurations. Once attached to the wind<br />

tunnel walls the model can be rotated and then securely<br />

fixed to achieve various angles of attack. By constructing<br />

the model in this fashion the trailing vortices as<br />

experienced by a finite spanned wing are virtually<br />

eliminated.<br />

40


Each element has 64 independent pressure taps which<br />

measure pressure differentials through 1 mm diameter<br />

holes placed around the centreline of the windward and<br />

leeward surface of each wing, sufficiently far away from<br />

the wind tunnel side-wall boundary layers. They are<br />

dispersed from the leading to the trailing edge with a<br />

higher concentration around the leading edges of the<br />

elements. The pressure taps can be seen in Figure 5.<br />

They are slightly off-set from the stream-wise direction<br />

to avoid interference from wakes of upstream holes<br />

interfering with downstream taps. Tubes connecting to<br />

each individual pressure tap feed out an end of each wing<br />

to a box containing pressure transducers. The pressures<br />

were logged at a rate of 200 Hz over a 30-s period. The<br />

pressure transducer box was located downstream of the<br />

wing out of the free-stream.<br />

5.3 Experimentation Variables<br />

The variables altered during the wind tunnel testing were<br />

as follows:<br />

• Front wing model angle of attack, relative to<br />

main element chord<br />

• Aft element chord-line deflection, , relative to<br />

main element chord<br />

• Gap between main and aft element, g, defined at<br />

zero aft deflection angle , as a % of the main<br />

wing chord<br />

• Aft element pivot point position, as a % of the<br />

main wing chord.<br />

• Free stream flow velocity, V.<br />

Figure 6 shows these variables with respect to the model.<br />

It also defines the pivot location for the aft element and<br />

defines the gap between the foils.<br />

Figure 5 Close up photograph of leading edge of a wing<br />

model showing the 1 mm diameter pressure tap holes.<br />

The pressure measurements, along with the locations of<br />

the pressure taps can be used to obtain a pressure<br />

distribution across the surface of each element.<br />

Aerodynamic forces result from these pressure<br />

distributions acting over the surface of the wings. By<br />

integrating the pressure distributions in the vertical and<br />

horizontal directions, the vertical and horizontal forces<br />

can be computed, respectively. Since pressure acts<br />

normal to a surface, the surface curvature of the aerofoil<br />

between two points should not be neglected. However,<br />

the pressure taps were distributed closely together around<br />

regions of high curvature on the wing elements and<br />

therefore a straight line approximation of the surface<br />

between any two pressure taps was regarded as<br />

sufficiently accurate for the present work. The integrated<br />

x and y forces were firstly resolved into directions<br />

relative to the chord of the aerofoil. Therefore their<br />

resultants were then decomposed appropriately to find<br />

the lift and drag relative to the free-stream velocity.<br />

Finally non-dimensional representations of the<br />

coefficients of pressure (Cp), lift (Cl), and drag (Cd)<br />

were computed.<br />

Figure 6 Wing variables altered in experiment.<br />

Reynolds number Re is defined by the reference length<br />

2c (=1.0 m), the reference velocity V and the kinematic<br />

viscosity for air.<br />

41


6. RESULTS AND DISCUSSION<br />

6.2 The influence of angle of attack<br />

6.1 Pressure coefficients<br />

Figure 7 Example of pressure coefficient distribution. =<br />

10°, g = 1%, = 2°, Re = 800,000.<br />

The shapes of the main and aft elements are NACA0025<br />

and NACA0009, respectively [13]. The chord of the<br />

main element c1 is 500 mm, and it is same as the chord<br />

of the aft element c2. The total wing chord c = c1 + c2 =<br />

1 m.<br />

Figure 7 shows an example of the pressure coefficient<br />

variation over the foils. In Figure 7 the deflection angle<br />

of the aft element is 10°, the gap between main and aft<br />

elements is 1% referenced to the main element chord c1,<br />

the angle of attack, is 2° and Reynolds number based<br />

on the total wing chord c as the reference length is<br />

800,000.<br />

Figure 8 Example of stall. = 10°, g = 1%, Re = 700,000.<br />

Figure 8 shows the distributions of the lift coefficient C l<br />

and the drag coefficient C d versus angles of attack C l<br />

increases almost linearly with increase in up to 10° and<br />

C d also increases a little. C l reaches a maximum value at<br />

around = 10°. At larger angles of attack the wing stalls<br />

and C l decreases sharply and C d increases rapidly, as<br />

expected.<br />

6.3 The influence of aft element deflection<br />

The horizontal axis of the graph indicates the<br />

dimensionless distance along the chord direction. The<br />

vertical axis shows the pressure coefficient. The position<br />

x/c = 0 corresponds to the leading edge of the main<br />

element, and x/c = 0.5 indicates the trailing edge of the<br />

main element and the leading edge of the aft element.<br />

The vertical axis has been inverted with negative values<br />

at the top. Thus the red line indicates the pressure<br />

coefficient at the top of the elements, and those along the<br />

bottom are shown by the blue line. Like ordinary wings,<br />

the pressure at the top is negative, and lift occurs due to<br />

the pressure difference between the top and bottom of<br />

elements. The distance between the lines is indicative of<br />

the normal force acting on each element.<br />

= 0<br />

= 0<br />

= 15<br />

= 15<br />

= 30<br />

= 30<br />

Figure 9 C l and C d versus for varying aft element<br />

deflection. g = 2%, Re = 700,000.<br />

42


= 0<br />

= 15<br />

= 30<br />

= 25<br />

Figure 10 Pressure coefficient distribution for varying .<br />

g = 2%, Re = 700,000, = -1°.<br />

Figure 9 shows C l and C d versus for varying aft<br />

element deflections. Deflecting the aft element<br />

downwards increases C l substantially, although the<br />

values of C l for = 15° and 30° are almost the same. The<br />

gradients for each are nearly identical. The stall angle<br />

for at high aft element deflections (15°) is lower<br />

than that at = 0°. The drag coefficients are substantially<br />

higher when the aft element is deflected at angles of 15°<br />

and 30° compared to = 0°. It can be seen that<br />

increasing the flap angle from = 15° to = 30°<br />

produces almost no increase in lift and a significant<br />

increase in the drag.<br />

To show why the lift is so low for = 30°, Figure 10<br />

compares the pressure distributions for three different<br />

flap deflection angles for = -1°. At = 30°, the<br />

pressure coefficients are smaller than for = 0° and 15°,<br />

and the difference between the values at the upper and<br />

lower surface is small. In particular, it can be seen that<br />

the large aft element deflection of = 30° produces<br />

virtually no pressure difference across it, as well as<br />

virtually no pressure difference across the front element<br />

as well.<br />

Figure 11 C l and C d versus for varying aft element<br />

deflection. g = 2%, Re = 300,000, pivot = 80%.<br />

Figure 11 shows another example when the cambers are<br />

varied. C l at ° is larger than that at ° over<br />

whole rangebefore the flow stalls. C l at ° is<br />

larger than that at ° when is small, but the slope<br />

of C l -a curve for ° is lower than the other curves,<br />

and C l at ° becomes larger beyond °This<br />

means C l has a local maximum at a certain camber.<br />

6.4 The influence of gap size<br />

Figure 12 shows a comparison of wing performance for<br />

different gap sizes. As can be seen in the figure, the<br />

distributions of C l and C d for each gap are almost the<br />

same and the only small difference is that g = 1% gives a<br />

slightly higher C l at = 10°. In other tests in this study it<br />

has been observed that the influence of gap size is very<br />

small.<br />

43


Figure 14 C l and C d versus for varying gap size for<br />

negative gaps, = 15°, pivot = 90%, Re = 700,000.<br />

Figure 12 C l and C d versus for varying gap size for = 10°<br />

and Re = 700,000.<br />

Figure 14 shows lift and drag coefficient result when the<br />

gap is negative, i.e. the main and aft elements are<br />

partially overlapping. Note that this is only physically<br />

possible for non-zero values of . While C d at g = 0 and<br />

at g = -2% are amost identical, C l at g = 0 is larger than<br />

that for g = -2% near the angle where the flow stalls.<br />

Thus negative gaps do not seem to be beneficial.<br />

6.5 The influence of aft element pivot point position<br />

Figure 13 C l and C d versus for varying gap size for =<br />

10°, Re = 300,000.<br />

Figure 13 shows the effect of varying gap size at a<br />

smaller Reynolds number. While the flow stalls at<br />

° in Figure 12, C l begins to decrease at a smaller <br />

in Figure 13. Thus a smaller Reynolds number causes<br />

stall to occur at a smaller <br />

Figure 15 C l and C d versus for varying pivot points for =<br />

10°, g = 0%, Re = 700,000.<br />

The pivot point is the location about which the aft<br />

element rotates as shown in Figure 6. It is defined by the<br />

distance aft of the leading edge of the main wing in terms<br />

of % main element chord. Thus for example, when the<br />

pivot point is 0%, it means that the pivot point is at the<br />

leading edge of the main element, and 100% means that<br />

44


pivot point is at the trailing edge of the main element.<br />

Figure 15 shows results for various pivot points. g and<br />

Re are all constant. It can be seen that C l at pivot = 40%<br />

is slightly larger than other results, but otherwise there is<br />

little obvious effect, even for the very large changes of<br />

position from 40% to 90%.<br />

Figure 17 shows a comparison of the C p variation when<br />

Re is varied. For Re = 800,000, the pressure coefficient<br />

curve is smooth. On the other hand, when Re is 200,000,<br />

the pressure coefficient has a step-like distribution at<br />

x/c = 0.18. The authors speculate that this step-like<br />

distribution at this low Reynolds number may be a<br />

laminar separation bubble. Furthermore, there may also<br />

be a laminar separation bubble on the pressure surface at<br />

x/c = 0.4. When the Reynolds number is larger, flow<br />

separation does not occur in these results and the curves<br />

are smooth. The pressure coefficient distributions on the<br />

aft elements are almost the same for both Reynolds<br />

numbers.<br />

Figure 16 C l and C d versus for varying pivot points for =<br />

10°, g = 2%, Re = 300,000.<br />

Figure 16 shows the results when Re is smaller<br />

(300,000). It can be seen that C l and C d are almost<br />

identical, but that the flow separates at smaller when<br />

the pivot point is 40% compared to 90%.<br />

Figure 18 C l and C d versus Re. = 10°, = 2°, g = 1%.<br />

6.6 The influence of Reynolds number<br />

Figure 19 C l and C d versus Re. = 10°, = 10°, g = 1%<br />

Figure 17 Pressure coefficient variation over the elements<br />

for two different Re, for = 10°, g = 1%, = 2°.<br />

Figure 18 and Figure 19 show C l and C d versus Reynolds<br />

number. In many cases in this study, C l and C d do not<br />

change significantly when the Reynolds number is varied<br />

as shown in Figure 18. On the other hand, when the aft<br />

45


element deflection and the angle of attack are large, C l<br />

and C d change with Re as indicated in Figure 19.<br />

Generally, C d increases with increase in C l as shown in<br />

Figure 11 and Figure 16.But, in Figure 19, C d becomes<br />

small when C l becomes large for this value of<br />

To analyse why this reduction in C d occurs, the<br />

pressure coefficients corresponding to Re = 400,000,<br />

500,000 and 600,000 in Figure 19 are shown in Figure<br />

20.<br />

has the same effect as roughening the surface of an<br />

aerofoil, and causes early transition, and thus the<br />

emulation of higher Reynolds number behaviour. Even<br />

the turbulence from a small upstream rod [14] has been<br />

found to be sufficient to substantially alter the separation<br />

behaviour of bluff bodies.<br />

For testing aerofoils like those in the present tests,<br />

turbulence in the onset flow will affect the location of<br />

transition, as will the presence of the pressure gradients<br />

on the top and bottom surfaces. It is expected that the<br />

tested wings will simulate the behaviour of C-class<br />

catamarans, which operate at similar Reynolds Numbers.<br />

It is not known with certainty in the present tests where<br />

transition was occurring, and whether or not bypass<br />

transition was present.<br />

The effect of the transition location on the behaviour of<br />

wings can be investigated by tripping the boundary layer<br />

artificially using a wire or similar, and it is planned to<br />

carry out such tests in the future.<br />

Re = 0.4 x 10 6<br />

Figure 20 Pressure coefficient variation over the elements<br />

for three different values of Re. = 10°, = 10°, g = 1%.<br />

The pressure coefficient over the top of the wing at Re =<br />

400,000 is almost constant downstream of the leading<br />

edge. This indicates that the flow has separated at the<br />

leading edge of the main element for this small Reynolds<br />

number. The wing is stalled over its entire surface, and<br />

thus C l becomes small and C d is large. When the<br />

Reynolds number is larger (Re = 600,000), no separation,<br />

or only a very small amount of flow separation occurs,<br />

and C d is again small. The pressure coefficient near the<br />

leading edge of the main element is significantly large,<br />

and C l is large. When the angle of attack and the aft<br />

element deflection are large, it is necessary to operate the<br />

wing at Reynolds numbers in excess of 600,000 in order<br />

to avoid separation and poor performance. Note that the<br />

pressure distribution at Re = 600,000 may show small<br />

laminar separation bubbles on the suction surfaces of<br />

both the fore and aft elements near their leading edges in<br />

the strongly adverse pressure gradient regions.<br />

The results presented in the paper were obtained with the<br />

two-element, 1-m chord wing mounted between the walls<br />

of the 2.5-m wide duct near the outlet of the open jet.<br />

The turbulence intensity of the flow was about 1%,<br />

which is much lower than that in the atmosphere, but is<br />

quite large compared to 0.1% in good quality<br />

aeronautical wind tunnels. Turbulence in the onset flow<br />

7. CONCLUSIONS<br />

In this study, the influence of geometric factors such as<br />

the aft element deflection, the gap and the aft element<br />

pivot point have been investigated in a wind tunnel<br />

study. From the study, the following conclusions can be<br />

drawn.<br />

C l and C d are significantly related to the aft element<br />

deflection. As the aft element deflection is increased, C l<br />

and C d also increase. It is the same aerodynamically as<br />

adding camber to a single element wing. The addition of<br />

aft element deflection increases C l , but too high a<br />

deflection results in an increase in drag.<br />

The influences of the gap size and the pivot point<br />

location are found to be relatively small in this study.<br />

The Reynolds number has a large influence on the<br />

sensitivity of the wing to separation. As expected, a<br />

smaller Reynolds number caused the flow to separate<br />

more easily, especially with high aft element deflections<br />

and high angles of attack.<br />

Acknowledgements<br />

We would like to recognise the following people who<br />

have contributed to this project. We thank Joseph Nihotte<br />

and James Turner for carrying out a large amount of the<br />

wind tunnel testing, and for analysing some of the<br />

results. We also thank David Le Pelley for his help and<br />

advice in setting up and overseeing the wind tunnel<br />

testing.<br />

References<br />

1. Blakeley, A.W., Flay, R.G.J. and Richards, P.J.,<br />

Design and optimisation of multi-element wing sails<br />

for multihull yachts, Proceedings of the 18th<br />

Re = 0.6 x 10 6 46


Australasian Fluid Mechanics Conference 3rd - 7th<br />

December 2012 Edited by P.A. Brandner and B.W.<br />

Pearce. Published by the Australasian Fluid<br />

Mechanics Society. ISBN: 978-0-646-58373-0.<br />

2. Lanchester, F.W., Aerodynamics, Vol. 1, p 431, A.<br />

Constable and Co., London, 1907.<br />

3. Elkaim, G.H., “Autonomous Surface Vehicle Free-<br />

Rotating Wingsail Section Design and Configuration<br />

Analysis”. Journal of Aircraft, 2008. 45(6): p. 1835-<br />

1852.<br />

4. Marchaj, C.A., Sail Performance: Techniques to<br />

Maximize Sail Power. 1996: McGraw-Hill.<br />

5. Gonen, I. Omer Wing-Sail Ltd. Available from:<br />

www.omerwingsail.com.<br />

6. MacLane, D., “The Cogito Project- Design and<br />

Develop of an International C-Class Catamaran and<br />

Her Successful Challenge to Regain the Little<br />

Americas Cup”. Marine Technology and SNAME<br />

News, 2000. 37(4): p. 163-184.<br />

7. Killing, S., “Alpha and Rocker - Two Design<br />

Approaches that Led to the Successful Challenge for<br />

the 2007 International C Class Catamaran<br />

Championship”. Proceedings of the 19th<br />

Chesapeake Sailing Yacht Symposium. 2009.<br />

Annapolis, Maryland.<br />

8. Drela, M. "Newton Solution of Coupled<br />

Viscous/Inviscid Multielement Airfoil Flows,",<br />

AIAA paper 90-1470, presented at the AIAA 21st<br />

Fluid Dynamics, Plasma Dynamics and Lasers<br />

Conference, Seattle Washington, June 1990.<br />

9. Drela, M. and Youngren H., "A User's Guide to<br />

MISES 2.53", MIT Computational Sciences<br />

Laboratory, December 1998.<br />

10. Peart, C.R.C., “Numerical Modeling of a wing-sail<br />

Rigged Class in a Wind Gradient in Both Upwind<br />

and Off the Wind Conditions”. MSc Thesis, School<br />

of Engineering Science, University of Southampton,<br />

UK, 2011.<br />

11. Greenbird. Greenbird supported by ecotricity.<br />

Available from: http://www.greenbird.co.uk/.<br />

12. Sails, X.-W. X-Wing Sails - Wing Sails for Sunfish<br />

and More. Available from:<br />

http://www.solidwingsails.com/.<br />

13. Abbott, I.H. and A.E. Von Doenhoff, Theory of<br />

Wing Sections: Including a Summary of Airfoil Data.<br />

1959: Dover Publications.<br />

14. Gartshore, I.S. “Some effects of upstream turbulence<br />

on the lift forces imposed on prismatic two<br />

dimensional bodies”, J. Fluids Eng. 106(4), 418-<br />

424, 1984.<br />

47

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!