You are on page 1of 9

Computers & Fluids 82 (2013) 29–37

Contents lists available at SciVerse ScienceDirect

Computers & Fluids


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c o m p fl u i d

Numerical analysis of multiple, thin-sail geometries based on Prandtl’s


lifting-line theory
Robert E. Spall ⇑, Warren F. Phillips, Brian B. Pincock
Department of Mechanical and Aerospace Engineering, Utah State University, Logan, UT 84322-4130, United States

a r t i c l e i n f o a b s t r a c t

Article history: Solutions obtained from a numerical method based on Prandtl’s lifting-line theory, valid for multiple lift-
Received 26 July 2012 ing surfaces with arbitrary sweep, are obtained for a number of rigid wing and sail geometries. The results
Received in revised form 26 March 2013 are compared against solutions obtained using established vortex-lattice methods, and computational
Accepted 17 April 2013
fluid dynamics solutions to the Euler equations. For the case of an untwisted, rectangular wing, numerical
Available online 3 May 2013
lifting-line, vortex-lattice, and Euler solutions were all in reasonable agreement. However, the numerical
lifting-line method was the only method to predict the constant ratio of induced-drag coefficient to lift
Keywords:
coefficient squared, which has been predicted from the analytic solution and confirmed by well estab-
Potential flow
Aerodynamics
lished experimental data. Results are also presented for a forward-swept, tapered wing. Additional results
Lifting line theory are presented in terms of lift and induced-drag coefficients for an isolated mainsail, and mainsail/jib com-
binations with sails representative of both a standard and tall rig Catalina 27. The influence of the non-
linear terms in the lifting-line solution appears minimal, with the exception of mainsail results when
considering jib/mainsail combinations.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction equations. These include lifting-line methods (c.f., Refs. [1–4]), pa-
nel methods (c.f., Ref. [5]), and vortex-lattice methods (c.f., Ref.
A wide range of computational tools are available for the analysis [6]). In the present work, potential flow solutions are obtained
of fluid flow about three-dimensional sails, including solutions to using numerical lifting-line method (NLLM) [4] and a vortex-lattice
the Reynolds-averaged Navier–Stokes (RANS) equations. However, method (VLM) [6]. Potential flow solutions also find utility when
for wide-ranging parameter studies and preliminary design, pre- coupled to boundary-layer equations, where overall solution times
mium is placed on ease of solution, accuracy, and minimizing the are significantly below those required to solve the full Reynolds-
required computational costs. Three-dimensional RANS solutions averaged Navier–Stokes equations.
are computationally expensive, and hence may not represent the Lifting-line theory was first developed by Prandtl [1,2] in 1918
ideal tool for preliminary design studies, particularly for sails in up- for a single lifting surface with no sweep or dihedral and has been
wind configurations for which the flow remains primarily attached. used for the analysis of isolated sails by a number of authors [7–9].
Solutions to the Euler equations, which govern the velocity and The theory was extended to a pair of parallel lifting lines based on
pressure fields for a fluid in which viscous forces are negligible, Munk’s equivalence theorem [3] and solutions were later pre-
represent a viable alternative. There are several approaches to sented in Glauert [10]. Subsequently, Milgram [11] applied the the-
obtaining solutions to the Euler equations. One involves a direct ory to the analysis of multiple sails. However, the series solution
computational fluid dynamics (CFD) solution to the equations method presented by Milgram [11] requires that the lifting lines
which is still a relatively expensive endeavor. However, if an invis- are parallel, that the sails are without sweep, and have an elliptic
cid, constant density fluid flow about a body originates from a irro- lift distribution. More recently Berbente and Maraloi [12] applied
tational flow, such as a uniform flow, then Kelvin’s theorem shows Prandtl’s lifting-line theory to multiple sails with fewer restrictions
the fluid will remain irrotational everywhere. This permits the on the configuration. However, their approach still utilizes an infi-
introduction of a velocity potential with the resulting potential nite series solution, which is strictly valid only for a single, un-
flow equation which may be solved in lieu of the Euler equations. swept lifting surface. As such, multiple sails were modeled by
There are a number of approaches to solving the potential flow Berbente and Maraloi [12] as a single hybrid lifting surface, which
does not account for three-dimensional mainsail/jib interactions.
This is particularly problematic for fractional rigs, which were
⇑ Corresponding author. Tel.: +1 435 797 2878.
not considered by Berbente and Maraloi [12]. Consequently, the
E-mail address: spall@engineering.usu.edu (R.E. Spall).

0045-7930/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compfluid.2013.04.020
30 R.E. Spall et al. / Computers & Fluids 82 (2013) 29–37

Nomenclature

A determines distribution of sail fullness V1 magnitude of freestream velocity


AR measure of camber at trailing edge vji dimensionless velocity vector induced at control point i
AV measure of camber at leading edge by horseshoe vortex j, having unit strength
e Li;a
C section lift slope for section i x coordinate direction aligned with boat centerline
ci root mean square chord length for section i y coordinate direction normal to boat centerline
es span efficiency factor ^x dimensionless coordinate direction aligned with the lo-
E foot length of mainsail cal sail chord line
Fa resultant aerodynamic force vector ^
y dimensionless coordinate direction normal to the local
Gi dimensionless vortex strength for section i sail chord line
I measured from genoa head to main deck ^max
y maximum dimensionless depth of sail at a given section
J base length of foretriangle z sail or wing spanwise coordinate direction
k drag-due-to-lift factor a angle of attack, or wind angle relative to boat centerline
LP shortest distance between clew and luff of genoa ai local angle of attack at section i
N number of horseshoe vortices aL0i local zero-lift angle of attack at section i
P luff length of mainsail Ci local vortex strength at section i
RA aspect ratio dSi planform area of spanwise increment of sail covered by
Sr global reference area horseshoe vortex i
uni normal unit vector at control point i fi dimensionless spanwise length vector
uai axial or chordwise unit vector at control point i q air density
u1 unit vector in direction of freestream / velocity potential

above lifting-line approaches are not ideal for the general analysis strengths Gi ¼ Ci =ðci V 1 Þ for each horseshoe vortex and the proper-
of multiple sails. ties of the sail given as:
Phillips and Snyder [4] developed a numerical method for  ! 
 XN 
obtaining solutions to Prandtl’s lifting-line equation for systems   e Li ¼ 0
2 u1 þ

v ji Gj  fi Gi  C

ð1Þ
of lifting surfaces of arbitrary location and orientation. However, j¼1
their application was for wings of significant thickness and rela-
tively small camber, not the geometries that characterize sails: where
very thin, swept, and with large camber. The application to sail " P #
ðu1 þ Nj¼1 v ji Gj Þ  uni
analysis is significant as this extension permits the modeling of ai ¼ tan1 P ð2Þ
multiple sails using Prandtl’s lifting-line theory without restric- ðu1 þ Nj¼1 v ji Gj Þ  uai
tions regarding sweep. In addition, the three-dimensional interac- e Li is a function of the local angle of
and the section lift coefficient C
tions between the upwind and downwind sails are properly
attack. Eq. (1) is applied at N control points located on the quarter
accounted for. The purpose of this work is to evaluate the viability
chord resulting in a nonlinear system of N equations and N un-
of the lifting-line method of Phillips and Snyder for preliminary sail
knowns which may be rapidly solved using Newton’s method.
analysis. Results are compared with solutions resulting from vor-
If a small-angle-of-attack approximation for both the geometric
tex-lattice methods and Euler CFD solutions.
and induced angles of attack is imposed, and all 2nd-order terms
Commonly, vortex-lattice methods with only a single lattice
are neglected, Eq. (1) reduces to:
element in the chordwise direction are termed lifting-line meth-
ods. For instance, Milgram [11], in addition to series solutions of X
N

Prandtl’s lifting-line theory, also utilized this approach for the anal-
e Li;a
2ju1  fi jGi  C v ji  uni Gj ¼ Ce Li;a ðu1  uni  aL0i Þ ð3Þ
ysis of jib/mainsail interactions, and termed the procedure a lift- j¼1

ing-line method. More recently Bertin and Cummings [13] stated The above equation requires the lift-slope coefficient and the
that the lifting-line method of Phillips and Snyder [4] is equivalent zero-lift angle of attack at each section, and is linear in the vortex
to vortex-lattice methods that use only a single lattice element in strength Gi. It is expected to provide good results for slightly swept,
the chordwise direction. However, calling an under-resolved vor- high aspect ratio sails at small angles of attack. The total nondi-
tex-lattice method a lifting-line method seems questionable at mensional aerodynamic force vector is computed from the
best, as the vortex-lattice method is not based on Prandtl’s lift- relation:
ing-line theory. As described in the following section, the method !
of Phillips and Snyder [4], and a single chordwise element vor- Fa XN X
N
dSi
¼2 G i u1 þ Gi Gj v ji  fi ð4Þ
tex-lattice method, are fundamentally quite distinct. 1
qV 21 Sr Sr
2 i¼1 j¼1

The vortex-lattice method represents the surface as a grid of


2. Numerical methods horseshoe vortices. The most mathematically rigorous approach
is to locate the vortex panels on the mean camber surface, and to
In developing a numerical lifting-line method for multiple sur- allow the trailing vortices to align with the free stream. In the pres-
faces, Phillips and Snyder [4] applied the vortex lifting law to a dif- ent work, the public domain vortex-lattice code AVL [6] is used.
ferential element of a lifting line, and equated the resulting The methodology implemented in AVL invokes a linear approxima-
differential expression to the force developed on a spanwise differ- tion such that the bound and trailing vortices are always assumed
ential section using a local section lift coefficient. This results in a parallel to the x-axis. The linear formulation indicates that results
nonlinear relation between the unknown dimensionless vortex should be limited to low-angle-of-attack applications. Two options
R.E. Spall et al. / Computers & Fluids 82 (2013) 29–37 31

are available in the AVL code to impose an angle of attack. The user where
may align the mean chord line with the x-axis and adjust the angle
of the free stream, or align the free stream with the x-axis and ad-
A
B¼ ð6Þ
just the angle of the chord line. In either case, both the bound and ðAV þ 2ÞðAV þ 1Þ
trailing vorticity are always oriented parallel to the x-axis. Both ap-
and
proaches give similar results at low angles of attack.
We again note that the vortex-lattice method, using a single AR
vortex-lattice element in the chordwise direction, is not equivalent C¼ B ð7Þ
6
to the lifting-line method of Phillips and Snyder [4]. In particular,
In addition, AV and AR are related to the entrance angle at the
with the VLM, using a single element in the chordwise direction,
leading and trailing edges, respectively. The constant K is deter-
the bound vortex and control points are (usually) placed at one-
mined to provide a pre-determined maximum depth y ^max at the
and three-quarter chord locations, respectively. A zero normal
location where dy ^=d^
x ¼ 0. This involves finding the root of the
velocity (Neumann) condition is specified at the control point
equation dy ^=d^
x ¼ 0, and substituting the resulting value for ^ x,
which provides the system of N equations and N unknowns. Conse-
along with the desired y ^max , into Eq. (5) to solve for K. The constant
quently, with a single vortex-lattice element in the chordwise
A alters the distribution of sail fullness fore and aft. Laine and Laine
direction the solution is dependent only on the position and slope
recommend A = 1 for sails used in light air conditions, and
of the sail at the control point. Sail curvature upstream or down-
A = 1 + AV/4, which provides for more fullness forward and a flatter
stream of this control point does not influence the solution. With
leech, for sails used in heavier air (and represents the definition
the lifting-line method of Phillips and Snyder [4], the chordwise
used in this work).
variation in the shape of the sail is taken into account through sec-
The three dimensional sail shape is developed by specifying sec-
tion lift coefficients as indicated in Eq. (1), or section lift coefficient
tion shapes at the foot, approximate vertical mid-point, and head
slopes as indicated in Eq. (3). These values may be available
of the sail. Quadratic interpolation is used to determine necessary
through experimental data, or, in the present work, computed
values of AV, AR, and the depth y ^max to define intermediate section
using either thin airfoil theory or a two-dimensional vortex panel
profiles.
method. The means by which these are acquired influences the
execution speed of the code.
Solutions to the Euler equations using the general purpose, 4. Results
unstructured mesh CFD solver STAR-CCM+ [14] are also presented.
However, we note that solutions to the Euler equations will always Results are first presented for two different thin wings, distin-
be subject to the influences of numerical diffusion, and hence will guished by their taper ratio, in which the section shapes were de-
not represent true potential flow solutions. fined by a NACA 9400 airfoil. Subsequently, we present results for
In terms of computational efficiency, Phillips and Snyder [4] an isolated mainsail, and finally two mainsail/jib combinations,
performed a careful analysis of computation time for the numerical with section shapes defined by the procedure of Laine and Laine
lifting-line method, a panel method, and a structured CFD Euler [15]. The mainsail and jib dimensions are representative of those
solver. Their conclusions were that the nonlinear lifting-line solu- for both a standard and tall rig Catalina 27. Results for all cases
tions were four and six orders of magnitude faster than the panel are presented for lifting-line, vortex-lattice, and Euler solutions.
method and CFD Euler solvers, respectively. However, we caution The ranges of angles of attack presented include levels above and
that computation time for codes is very subjective, varying be- below what would be feasible for a flexible sail. This is done simply
tween methods with respect to the size of the system of unknowns to extend the range of comparison between the different solution
(e.g., grid resolution), levels of convergence specified, compilers methodologies.
and optimization levels, and programming skills. In terms of the We first consider an isolated, thin, rectangular wing of aspect
lifting-line method as coded for the current work results for the ratio 3 with section shapes defined by a NACA 9400 airfoil. For
solutions to the linearized equations are achieved in less than the Euler solutions, the sail was located within a rectangular box
0.1 s. The nonlinear solver employing a series solution for section defining the outer boundaries of the computational domain with
lift coefficients converges in approximately 1 s. If a two-dimen- the boat centerline aligned along the x-axis. The dimensions are gi-
sional panel method is utilized to obtain the section lift coeffi- ven as 20 m 6 x 6 21 m, 20 m 6 y 6 20 m, 20 m 6 z 6 23 m,
cients, the compute time increases to several minutes, depending where the leading edge (luff) of the sail was located at x = 0 m,
on the streamwise resolution of the panel method solver. Although the lateral location of the luff at y = 0 m, and the foot of the sail
not utilized in the present work, should experimental values of the at z = 0 m. Consequently, the influence of the water surface is not
section lift coefficients be available, compute time would be on the accounted for. Symmetry boundary conditions were imposed
order of 1 s. around the perimeter, while a fixed (uniform) velocity condition
u = 1 m/s was imposed at the inlet boundary (x = 20 m). Zero-
3. Sail shapes derivative boundary conditions were imposed at the outlet bound-
ary (x = 21 m). Angles of attack were changed by rotating the wing
Two approaches to define section camber shapes were followed. about its leading edge.
Initial calculations for isolated rectangular and tapered wings are The steady Euler equations were solved using the unstructured,
presented for camber lines derived from NACA 9400 airfoil sec- general purpose CFD code STAR-CCM+ [14]. Second-order upwind-
tions. This provides a maximum camber of 9%, occurring at 40% ing interpolation was applied to the convective terms and the SIM-
chord, representing a reasonable approximation of a section profile PLE procedure was employed for pressure–velocity coupling. A
for a generic mainsail. For the Catalina 27 sails, section shapes fol- segregated multi-grid solver was employed and solutions were
lowing the work of Laine and Laine [15] were used. In particular, considered converged when all residuals were decreased to
the equation giving the sail depth ðy^Þ at any point ^
x along a section approximately 106, and lift and induced-drag coefficients no long-
was given by: er changed with further iterations. Solutions were obtained on
!
Að1  ^xÞAV þ2 AR 3 three different mesh sizes for each angle of attack studied. In gen-
^ð^xÞ ¼ K
y  ^x þ C ^x þ B ð5Þ eral, mesh sizes ranged from approximately 700,000 to 5.5 million
ðAV þ 2ÞðAV þ 1Þ 6
cells for the coarsest to finest meshes, respectively. Results pre-
32 R.E. Spall et al. / Computers & Fluids 82 (2013) 29–37

sented were obtained through a Richardson extrapolation of these


three solutions. In all cases, the mesh was significantly refined in
the regions surrounding the tip vortices.
The vortex-lattice method was implemented using the public
domain solver AVL [6]. The mesh utilized 41 points in the chord-
wise direction and 30 in the spanwise direction. Since vorticity is
shed more rapidly near the tips of a wing, cosine clustering of no-
dal points toward the tips was implemented. The angle of attack
was varied by adjusting the angle of the chord line with the x-axis.
The lifting-line results were obtained by distributing 100 vorti-
ces over the sail span. Again, cosine clustering was implemented to
distribute nodal points towards the wing tips [4,5]. Depending on
whether the nonlinear equation (Eq. (2)) or linear equation (Eq.
(3)) was being solved, either the section lift coefficient, or the sec-
tion lift-slope coefficient and local zero-lift angle of attack was re-
quired. For solutions to the linear system the section lift-slope
coefficient was taken as 2p, as dictated by thin airfoil theory, and
the zero-lift angle of attack was a function of only the shape of
the camber line. For the nonlinear equation set, section lift coeffi- Fig. 2. Drag polar for rectangular wing of aspect ratio 3.
cients were computed using a two-dimensional vortex panel
method [5]. The panel method used 61 points in the chordwise
direction. The primary advantage of the linear system is a signifi-
cant savings in computer time.
We present in Fig. 1 the computed lift coefficients as a function
of angle of attack over the range 14 6 a 6 14 . In addition to the
numerical solutions, results for the analytic series solution are also
included. Results for the linear and nonlinear NLLM are nearly
indistinguishable and agree quite well with the analytic solution.
The nonlinear NLLM predicts a lift slope of 3.64. The VLM results
are similar, predicting a slightly lower lift slope of 3.37. Euler solu-
tions were limited to the range 0 6 a 6 14 . This is due to the
influence of numerical diffusion and the separation that occurs
about the leading edge at lower and higher angles of attack, which
is indicated in Fig. 1 as a decrease in the lift coefficient relative to
the NLLM and VLM results. That is, for a stagnation point that oc-
curs either below (at high angles-of-attack) or above (at low an-
gles-of-attack) the leading edge, the flow is unable to remain
attached as it accelerates about the leading edge. The inability of
Euler CFD solutions to accurately predict inviscid flow near a sharp
corner is well known.
Fig. 3. Drag-due-to-lift factor (k) as a function of angle of attack for a rectangular
Results in terms of the induced-drag polar are presented in wing of aspect ratio 3.
Fig. 2. The numerical lifting-line method and VLM both agree quite
will with the analytic series solution, which shows that the in-
duced-drag coefficient increases as the square of the lift coefficient significant deviation from the expected results occurs at both low
for an untwisted lifting surface. In the case of the Euler solutions, a and high angles of attack. Again, the lift decreases and induced-
drag increases at higher and lower angles of attack because of flow
separation at the leading edge caused by the numerical errors.
Additional insight is revealed in Fig. 3 where the induced drag-
due-to-lift factor, k ¼ C D =C 2L , is plotted as a function of angle of at-
tack. The results from both lifting-line approaches are essentially
constant with a value of k  0.108. This is quite close to the theo-
retical value, computed using an analytical series solution to Pra-
ndtl’s lifting-line equation, and given by 1=ðp RA es Þ. The validity
of this result has been well established through comparison with
experimental data. In particular, with a section lift slope coefficient
of 2p, the efficiency factor is computed as es = 0.981482 and hence
k = 0.108105. The VLM AVL values are not constant, showing vari-
ations over the entire angle-of-attack range, but particularly near
the angle-of-attack at which the lift coefficient approaches zero.
This inaccuracy is not unexpected since the trailing vorticity is al-
ways oriented parallel to the x-axis and does not align with the free
stream. The results for the Euler solutions differ from the theoret-
ical value by as much as 6% over the range of 0 6 a 6 14 . This too
is not unexpected since Euler solutions computed using a finite-
volume procedure do not provide a true potential flow solution
Fig. 1. Lift coefficient vs. angle of attack for rectangular wing of aspect ratio 3. due to the influence of numerical viscosity.
R.E. Spall et al. / Computers & Fluids 82 (2013) 29–37 33

Fig. 6. Drag polars for tapered wing.


Fig. 4. Drag coefficient plotted as a function of the square of the lift coefficient near
the zero-lift angle of attack for a rectangular wing of aspect ratio 3.

Fig. 5. Lift coefficient vs. angle of attack for tapered wing.

We further examine the behavior of the solutions in the vicinity


of the zero-lift-coefficient angle of attack (a  9.14°). In particu- Fig. 7. Rig measurement notation.
lar, we show in Fig. 4, for LLT, VLM, and the series solution, CD as
a function of C 2L . Here, the LLT results both agree quite well with
the series solution results, with a slope of approximately 0.108. higher angles of attack due to the initiation of numerical-viscos-
However, the VLM results are displaced vertically such that the ity-induced flow separation. This becomes quite apparent in the in-
drag coefficient has a non-zero value associated with a zero lift duced-drag polar shown in Fig. 6 where induced-drag coefficients
coefficient. In addition, the slopes for CL > 0 and CL < 0 differ slightly for a given lift coefficient increase dramatically for the Euler solu-
such that there is a small discontinuity in k at the zero-lift-coeffi- tions. This phenomenon is exaggerated for the tapered wing as the
cient angle of attack. flow begins to separate near the untwisted tip region.
Shown in Figs. 5 and 6 are lift coefficients vs. angle of attack and We next present calculations for an isolated mainsail, and a
induced-drag polars, respectively, for an isolated, untwisted, ta- mainsail/jib combination. Standard notation for the sail dimen-
pered, forward swept wing. The wing sections are again defined sions is shown in Fig. 7. The rig measurements correspond to a
by a NACA 9400 airfoil. The taper ratio is 0.25, the forward sweep standard rigged Catalina 27. In particular, the P and E measure-
is 4.76°, and the aspect ratio is 4.8. This geometry begins to take ments for the mainsail were 8.74 m and 3.2 m, respectively. The I
the shape of a mainsail, with a head dimension 25% of the foot. measurement (height of forestay attachment above deck) was
However, given the forward sweep of the wing, analytic solutions 10.36 m and the J measurement (jib tack to mast distance) was
are not available. Solver parameters are similar to those for the 3.43 m. For simplicity, the planform of the jib is as shown in
rectangular wing, and hence will not be repeated here. The results Fig. 7 (a right triangle), and does not correspond to any Catalina
are qualitatively similar to those for the rectangular wing. The pre- 27 ‘‘standard’’ cut jib. The resulting aspect ratios for the mainsail
dicted lift coefficients for the linear and nonlinear lifting-line and jib are 5.3 and 6.03, respectively. The jib luff perpendicular
methods are slightly higher than those for the vortex-lattice meth- (LP) length is 4.88 m. The height of the jib tack above the water
od. The lift coefficients for the Euler solutions tend to drop off at (a symmetry plane) was 1.3 m.
34 R.E. Spall et al. / Computers & Fluids 82 (2013) 29–37

Fig. 8. Placement of jib/mainsail combination within the computational domain for


Euler solutions.

For calculations involving an isolated mainsail, values of AV and


AR, related to the entrance angle at the leading and trailing edges,
were set to 3 and 1, respectively. The maximum camber (depth) at
the foot, mid, and head sections of the sail were set to 6%, 12%, and
6%. In addition, a linear geometric twist of 10° was specified. Since Fig. 9. Lift coefficients for isolated mainsail vs. wind angle relative to centerline.
the section shapes vary, unlike the previous wings, this sail also
contains aerodynamic twist. These levels of twist (or washout) pre-
vent significant separation near the tip of the sail for the Euler cal-
culations. (Without twist, large regions of separated flow originate
near the tip region.)
Two different mesh densities were implemented for the iso-
lated mainsail Euler solutions. These ranged from approximately
800,000 to 3 million cells, depending on the specific angle of attack
and sail configuration. For the mainsail/jib combinations, three dif-
ferent meshes were used, with cell counts ranging from 700,000 to
5.5 million. Richardson extrapolation was not performed due to the
difficulty of implementing a uniform reduction in mesh size over
the domain. Consequently, for all cases results for the finest
meshes are presented. Differences between fine and coarse mesh
computed lift and induced-drag coefficients were on the order of
1–2%. A mesh refinement procedure was also implemented in
which the mesh in the wake region of the sails was refined to bet-
ter resolve the tip vortices and large gradients inherent in the solu-
tion. The dimensions of the computational domain were defined as
60 m 6 x 6 60 m, 60 m 6 y 6 60 m, 0 m 6 z 6 50 m. Angles of
attack were adjusted by altering the direction of the inlet flow on Fig. 10. Drag polars for isolated mainsail.
the xmin and ymin windward boundaries. The placement of the
mainsail within the computational domain is shown in Fig. 8
(which also includes placement of the jib). The xmin and ymin wind-
ward boundaries were specified as velocity inlets. The xmax and
ymax lee boundaries were specified as constant pressure outlets.
The water surface boundary, zmin, and the zmax upper boundary
were each specified as symmetry planes.
Results are shown in Fig. 9 in terms of lift coefficient vs. angle of
attack, and Fig. 10 in terms of the induced-drag polar. For the lift
coefficient, results for the linear and nonlinear NLLM are nearly
indistinguishable. The VLM predicts values below the NLLM at all
angles of attack. This same trend was observed in Fig. 5 for the ta-
pered wing, although to a lesser extent. The Euler results tend to
fall between the NLLM and VLM results. The ratio of lift-to-drag
coefficients as a function of wind angle is shown in Fig. 11. The re-
sults for the linear and nonlinear NLLM are nearly identical. The
VLM predicts a lower lift-to-drag ratio over the entire range of an-
gles of attack. The Euler solutions are quite close to the NLLM re-
sults, particularly in the intermediate ranges of the angle of
attack where the numerical errors in the Euler solutions are ex-
pected to be lowest. These results indicate that, for this case, the Fig. 11. Ratio of lift-to-drag coefficients as a function of wind angle for isolated
mainsail.
nonlinear terms are of minor importance in the lifting-line formu-
lation. Hence, the aspect ratio of 5.3 and the maximum 12% camber
do not significantly violate the assumptions required to linearize Results for the mainsail/jib combination are presented next. For
Eq. (1). Consequently, the linearized numerical method represents these sail definitions, AV and AR were again set to 3 and 1, respec-
a viable approach for most mainsail configurations. tively for both the mainsail and jib. For the mainsail, camber values
R.E. Spall et al. / Computers & Fluids 82 (2013) 29–37 35

were also unchanged. However, the linear twist was increased


from 10° to 18° due to the influence of the jib downwash on the
mainsail. The jib camber values were set to 6%, 15%, and 6%,
whereas the linear twist was set to 14°. The mainsail and jib sheet-
ing angles with respect to the boat centerline were 0° and 18°,
respectively. (Consequently, the effective angle of attack of the
jib is 0° at a nominal wind angle relative to the centerline of
18°.) We note that the 18° rotation of the jib as imposed in the
AVL code was a rotation of chord lines parallel to the water surface
(z constant plane) rather than a rotation about the forestay. This
has the effect of distorting the shape of the sail. The jib/mainsail
combination is shown in Fig. 12, with a (coarse level) Euler solution
mesh superimposed on the sails. The placement of the jib/mainsail
combination within the computational domain, which employs the
same dimensions and boundary conditions as the domain for the
isolated mainsail calculations discussed previously, is shown in
Fig. 7.
The lift coefficients computed using each method for the main-
sail are shown in Fig. 13. Overall, a significant decrease in the lift Fig. 13. Lift coefficients vs. wind angle for mainsail.
coefficients, relative to the isolated mainsail case, is observed. This
behavior is well known, and is due to the downwash induced on
the mainsail by the jib. In particular, the downwash results in a
movement of the mainsail stagnation point toward the leading
edge, and a concomitant deflection of streamlines around to the
lee side of the jib which reduces the speed of the flow over the
lee side of the mainsail [16]. We also note much greater differences
overall in the predicted lift coefficients for the different methods.
This is true even between the linear and nonlinear lifting-line re-
sults. The lift coefficient for the linear lifting-line solution increases
nearly linearly with angle of attack. The nonlinear lifting-line re-
sults show a positive second derivative of lift coefficient as a func-
tion of angle of attack, whereas the vortex-lattice method has a
negative second derivative. One may attribute the differences in
the linear and nonlinear lifting-line results to the linearization of
Eqs. (1) and (2). In particular, to obtain the linear form given by
Eq. (3), the small angle of attack approximation is applied to Eq.
(2), and both Eqs. (1) and (2) are linearized with respect to the vor-
tex strengths. These approximations lead to the slight differences
between linear and nonlinear lifting-line results. The nonlinear
terms become more important as difference between the wind an- Fig. 14. Lift coefficients vs. wind angle for jib.
gle and sheet angle becomes large.

Lift coefficients for the jib are shown in Fig. 14. These results
show smaller differences between the different methods relative
to the mainsail results, likely due to the fact that the jib is not in
the wake of the mainsail. The linear lifting-line method predicts
larger lift values than does the nonlinear method which is opposite
the trend for the mainsail. Higher lift values on the jib result in
greater downwash on the main, which results in lower lift values,
and vice versa. The results for low wind angles relative to the boat
centerline that are shown in Fig. 14 are not realistic for flexible
sails, because the jib would luff under the influence of the negative
lift. The Euler solutions predict lower values for the lift coefficient
than the other methods, and reveal a large drop-off at the higher
angles of attack. This is due to separation at the leading edge, a re-
sult of numerical viscosity.
The induced-drag polars for the mainsail are shown in Fig. 15.
The figure reveals that for a given lift coefficient, the predicted in-
duced-drag coefficients are highest for the Euler and vortex-lattice
solutions, and lowest for the nonlinear lifting-line method. The fig-
ure also reveals a negative induced-drag coefficient for the nonlin-
ear lifting-line method at small angles of attack. This may at first
seem untenable, however, the overall induced-drag coefficient
(mainsail and jib combined) is positive. The negative mainsail in-
duced-drag shown in Fig. 15 is a result of upwash induced on the
Fig. 12. Mainsail/jib combination with coarse Euler solution surface mesh. mainsail by negative lift on the jib.
36 R.E. Spall et al. / Computers & Fluids 82 (2013) 29–37

Fig. 15. Drag polars for mainsail. Fig. 17. Comparison of lift coefficients between tall and standard rig sails computed
using nonlinear lifting-line method.

Fig. 16. Drag polars for jib.


Fig. 18. Comparison of drag polars for tall and standard rig sails computed using
nonlinear lifting-line method.

Shown in Fig. 16 are the induced-drag polars for the jib. The re- at a given lift coefficient than does the tall rig. The opposite trend
sults for the linear and nonlinear lifting-line methods are in good is observed for the jib. Again, these results are consistent with the
agreement. The vortex-lattice method predicts a smaller induced- differing aspect ratios of the sails. Results for the combined sails
drag coefficient at a given lift coefficient, than either of the lift- are also shown. Here it is clear that overall, the tall rig has a lower
ing-line methods. The induced-drag coefficients at a given lift coef- induced-drag coefficient for a given lift coefficient, and hence one
ficient are much higher for the Euler results. This is again the would expect improved sailing performance.
influence of numerical viscosity errors on the solution, which can-
not be completely eliminated regardless of the mesh density. 5. Conclusions
The final set of results considers only the nonlinear lifting-line
method in a comparison between standard and tall rig Catalina A numerical method based on Prandtl’s lifting-line theory has
27 configurations. The aspect ratios for the tall rig mainsail and been applied to isolated and multiple thin-sail lifting surfaces
jib are 5.92 and 5.9, respectively. Consequently, the aspect ratio and the results compared with solutions obtained using a vortex-
of the mainsail is increased compared to the standard rig, but the lattice method and CFD solutions to the Euler equations. Where
aspect ratio of the jib is decreased. Results in terms of lift coeffi- solution behavior was known, such as for an untwisted, rectangu-
cients and induced-drag polars are shown in Figs. 17 and 18, lar wing, the lifting-line method performed quite well. In fact, the
respectively. In terms of lift coefficients for the mainsail, the tall numerical lifting-line method was the only method to predict a
rig values are slightly higher than the standard rig. For the jib, constant ratio of induced-drag coefficient to lift coefficient
the lift coefficients for the standard rig are slightly higher than squared. The Euler solutions suffered from the effects of numerical
the tall rig. These results are consistent with the differing aspect diffusion, resulting in artificial flow separations at angles of attack
ratios for the tall and standard rig sails. In terms of the induced- that require rapid acceleration around the sail leading edge. The
drag polars, the standard rig mainsail has a higher induced-drag vortex-lattice solutions obtained using the AVL code were limited
R.E. Spall et al. / Computers & Fluids 82 (2013) 29–37 37

to small angles of attack due to the alignment of the bound and References
trailing vorticity with the x-axis. This is not a limitation of the
vortex-lattice method in general, but is a result of a linearization [1] Prandtl L. Tragflügel Theorie. Nachrichten von der Gesellschaft der
Wisseschaften zu Göttingen, Geschäeftliche Mitteilungen, Klasse; 1918.
procedure used in the AVL code. The linear and nonlinear lifting- [2] Prandtl L. Theory of lifting surfaces. NACA TN 9; July 1920.
line results are quite comparable for isolated sails or for the lead [3] Prandtl L. Induced drag of multiplanes. NACA TN 182; July 1920.
sail when both the jib and mainsail are considered. However, for [4] Phillips WF, Snyder DO. Modern adaptation of Prandtl’s classic lifting-line
theory. J Aircraft 2000;37(4):662–70.
the jib/main combination, the results for linear and nonlinear [5] Phillips WF. Mechanics of flight. 2nd ed.. John Wiley and Sons Inc.; 2010.
lifting-line method differ significantly for the mainsail. Nonlinear [6] Drela M. <http://web.mit.edu/drela/Public/web/avl>.
lifting-line method was also applied to compare the difference in [7] Wood CJ, Tan SH. Towards an optimal yacht sail. J Fluid Mech 1978;85:459–77.
[8] Sugimoto T. A first course in optimum design of yacht sails. Fluid Dyn Res
lift and induced-drag coefficients between standard and tall rig
1993;11:153–70.
Catalina 27 sails. The results indicated an increase in the lift [9] Jackson PS. Modeling the aerodynamics of upwind sails. J Wind Eng Ind
coefficient at a given angle of attack for the mainsail, but a de- Aerodyn 1996;63:17–34.
creased value for the jib. This is consistent with the larger mainsail [10] Glauert H. The elements of airfoil and airscrew theory. 2nd ed. Cambridge
Science Classics; 1947.
aspect ratio for the tall rig relative to the standard rig, but a smaller [11] Milgram JH. The analytical design of yacht sails. Trans – Soc Naval Architects
aspect ratio for the jib. The tall rig displays a lower induced-drag Mar Eng 1968;76:118–60.
coefficient for a given lift coefficient, which should improve sailing [12] Berbente C, Maraloi C. Theoretical and experimental research regarding the
aerodynamics of a ship sail system. UPB Sci Bull Ser D 2006;68:17–32.
performance. Overall conclusions are that the accuracy realized [13] Bertin J, Cummings R. Aerodynamics for engineers. 5th ed. Prentice-Hall; 2009.
from the numerical lifting-line method is as good as or better than [14] CD-Adapco. STAR-CCM+ users manual; 2011.
that obtained from the vortex-lattice method or inviscid CFD solu- [15] Laine R, Laine J. The sailcut CAD handbook. <http://sailcut.sourceforge.net/
docs/en>; 2009.
tions, but particularly for the linearized equations, at a small frac- [16] Gentry A. The aerodynamics of sail interaction. In: Proceedings of the third
tion of the computational cost. AIAA symposium on aero/hydronautics of sailing, Redondo Beach, CA;
November 1971.

You might also like