Computationalhydrodynamicsofatypical3-finsurfboardsetup

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Journal of Fluids and Structures 90 (2019) 297–314

Contents lists available at ScienceDirect

Journal of Fluids and Structures


journal homepage: www.elsevier.com/locate/jfs

Computational hydrodynamics of a typical 3-fin surfboard


setup

S. Falk a , , S. Kniesburges a , R. Janka b , R. Grosso c , S. Becker d , M. Semmler a ,
M. Döllinger a
a
Division of Phoniatrics and Pediatric Audiology at the Department of Otorhinolaryngology, Head & Neck Surgery, University
Hospital Erlangen, Friedrich–Alexander-University Erlangen–Nürnberg, Germany
b
Department of Radiology, University Hospital Erlangen, Friedrich–Alexander-University Erlangen–Nürnberg, Germany
c
Computer Graphics Group, Friedrich–Alexander-University Erlangen–Nürnberg, Germany
d
Institute of Process Machinery and Systems Engineering, Friedrich–Alexander-University Erlangen–Nürnberg, Germany

graphical abstract

article info a b s t r a c t

Article history: Surfing is becoming increasingly popular, and the economic market for surfing products
Received 5 October 2018 is growing. The performance of the surfboards and the fins has evolved continuously
Received in revised form 9 April 2019 through the experience and innovative spirit of surfers and shapers. However, provided
Accepted 13 July 2019
performance characteristics are only subjective and of descriptive nature. Therefore,
Available online xxxx
we apply and analyze objective performance indicators as (1) lift and drag coefficient
Keywords: for each fin and the entire configuration and (2) lift-to-drag ratio for the entire fin
Computational fluid dynamics (CFD) configuration.
Surfboard This numerical study investigated a commercial 3-fin configuration, mounted into a
Fins rectangular simulation region, considering the flow at several different angles of attack
Surfing
between 0◦ and 45◦ . RANS and URANS simulations were performed with the SST k − ω
Hydrodynamics
turbulence model at an inflow velocity of 5 m/s.
STAR-CCM+

∗ Correspondence to: Waldstraße 1, D-91054 Erlangen, Germany.


E-mail address: sebastian.falk@uk-erlangen.de (S. Falk).

https://doi.org/10.1016/j.jfluidstructs.2019.07.006
0889-9746/© 2019 Elsevier Ltd. All rights reserved.
298 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

Simulation results showed that the fins have an optimal range of attack angles where
they produce the most lift. The lift forces generate a turning moment which forces the
board to be turned in direction to the shore favoring the controllability and stability of
the surfboard. The higher the angle of attack, the higher is the drag coefficient. Lift-to-
drag ratio values showed that for speeding up small attack angles with low drag forces
are preferable. Furthermore, unsteady effects as flow separation combined with vortex
shedding occur at high angles of attack above 20◦ which can only be resolved by URANS
simulations. These unsteady effects have high negative influences on stability due to high
fluctuation amplitudes for lift and drag forces.
In summary, this study presents, to the best of the authors’ knowledge, for the
first time, steady and unsteady forces on surfboard fins and discusses their potential
influences on the surfer’s controllability and stability of the surfboard during typical
surfing maneuvers.
© 2019 Elsevier Ltd. All rights reserved.

1. Introduction

The popularity of surf sport and especially surf-riding are rapidly growing. In 2002 there were over 10 million surfers
worldwide with increasing percentages around 12%–16% per year (Buckley, 2002). The surfing market is growing, and the
surfing industry became a multi-billion dollar industry. An overall, conservative, approximation of US$ 10 billion p.a. for
the current economic scale of surf-riding was estimated in 2002 (Buckley, 2002). From a technical point of view, surf-
riding is continually evolving. The board and the mounted fins have been developing over the years in a steady process
because of the close collaboration between surfers and surfboard shapers. Even today the shapers often build and finish
their boards by hand based on their long-lasting experience (Mc Cagh, 2013). Even well-known surfboard manufacturers
like Firewire Surfboards, which operate with computer and machine shaping using highly advanced materials, are still
dependent on the experienced shapers (Mc Cagh, 2013).
Computational fluid dynamics (CFD) have been established as standard tools in the development of new products in
every field of engineering applications. Examples of the maritime CFD applications were the investigations in sail shapes
(Knudsen, 2013), hydrodynamic analysis of super and mega yachts (Azcueta and Rousselon, 2009) and the CFD supported
design of lifeboats (Mørch et al., 2009). The significant advantages of CFD software tools are promptness and comparable
cheap costs compared to physical test models that help to speed up the analyzing process of prototypes and innovative
ideas. Unconfined access to the simulation region and high spatial and temporal resolutions are further advantages of CFD
software tools.
Although computer-aided design (CAD) and computer-aided manufacturing (CAM) techniques prevail as standard tools
in surfboard and fin design (Mc Cagh, 2013), CFD support is relatively new in the surfboard design process, and only a
few studies have been made starting with (Brandner and Walker, 2004) in 2004. The existing studies can be categorized
generally into two groups: (1) fin models without a surfboard (FIN) and (2) surfboard models without fins (BOARD). To the
best of our knowledge, studies including both elements (surfboard and fins) have not yet been performed to this point.
The FIN studies can be subdivided into those with application of laminar and turbulent flow models. Both groups
are using the steady-state Reynolds-Averaged Navier–Stokes equations (RANS) to simulate the fluid flow. Laminar flow
conditions were assumed by Lavery et al. (2009), who investigated the impact of fillet and un-fillet fins on lift and drag
forces at different flow rates and angles of attack. Turbulent flow models based on RANS were applied by Gudimetla et al.
(2009), Livanos (0000), Sakellariou et al. (2017), Macneill (2015) and Carswell (2007). Three of these five studies use a
k − ε turbulence model and just one uses a k − ω SST (shear stress transport) turbulence model (Menter, 1993), that
combines the good performances of the k − ω model in the boundary layers and the k − ε model in regions far from the
wall.
Gudimetla et al. (2009) focused on finding an adequate turbulence model and the entire performance parameters of lift,
drag and the lift-to-drag ratio at the 3- and 4-fin configurations. Similar studies were presented in Sakellariou et al. (2017)
and Macneill (2015), including a bio-inspired single fin design in (Macneill, 2015). Parameters as lift-to-drag coefficient
were evaluated and maximized and aspects of the efficiency and maneuverability have been estimated. Beside numerical
investigations, few studies also presented experimental results on fin hydrodynamics. Carswell (2007) investigated in his
bachelor’s thesis the flow around RedX surfboard fins, with focus on the influence of the flow to the lift coefficient, the
drag coefficient and the stiffness of the fins comparing different fin shapes and planform, experimentally and numerically
in a single or 3-fin arrangement.
In the second BOARD group, 3D surfboard models on a free water surface were considered in numerical simula-
tions (Oggiano, 2017; Oggiano and Pierella, 2018). A 2D flat plate that can be compared with a simplified surfboard was
studied in Barnett and Miravete (2009). Oggiano (2017) did a comparison between a modern surfboard and an Alaia that is
a historic type of surfboard used by the ancient people of Hawaii. Three different modern surfboard designs with different
tails and rockers were investigated in Oggiano and Pierella (2018).
S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314 299

Beside studies of numerical fin and board models, few studies presented experimental results (Brandner and Walker,
2004; Macneill, 2015; Carswell, 2007; Barnett and Miravete, 2009; Beggs-French, 2009). Macneill (2015) carried out lift
and drag force measurements in a water channel for validating their numerical results. Beggs-French (2009) analyzed
videos of the flow field around the fins during real surfing with the aim to get the prevailing angles of attack. Brandner
and Walker (2004) performed investigations on the lift and drag forces generated by fins in a cavitation tunnel and Dally
(2001) calculated the maximum surfing velocity with a photogrammetric analysis of surfers on the basis of two surfing
videos and correlated them to the wave height.
To the best of our knowledge, only steady-state RANS simulations have been performed, neglecting all unsteady flow
phenomena as flow separation from the fins, vortex shedding, etc. All those unsteady effects potentially produce so far
un-investigated, unsteady force excitation of the surfboard.
Therefore, our primary study goal is to provide flow field data around a 3-fin configuration with both steady-state
(RANS) and unsteady flow simulations using unsteady RANS (URANS). Catalano and Tognaccini (2011) showed that URANS
provides a very good consistency to Large Eddy Simulations (LES) of wing profiles. Thus, RANS and URANS turbulence
models are widely applied and appropriate for the simulation of wing or wing-like profiles to analyze lift and drag
forces (Catalano and Amato, 2003; Rezaei and Pasandideh-Fard, 2013; Peng et al., 2007).
The position and alignment of the fins are varied by changing the angle of attack (AoA) in the range between 0◦ and
45◦ . The geometry of the fins was obtained from commercial fins of a conventional 3-fin (so-called thruster) configuration.
Our hypotheses for the investigation of our 3-fin configuration are:

(1) Unsteady flow phenomena have to be taken into account for the evaluation of the track and roll stability of the
surfboard.
(2) The numerical results of the simulations show hydrodynamic effects that potentially influence surfing stability
during basic maneuvers.

Our results may also cross-fertilize fields of application like sailboats, kite-surfing or other domains, where fins or lifting
bodies of similar shape like surfboard fins, are used.

2. Surfing hydrodynamics

The very dynamic process of surfing is fundamentally based on the relationship between the wave height, the wave
peel velocity, and the surf-speed. To stay ahead of the breakpoint of the wave, a surfer and the surfboard have always an
equal or higher velocity than the waves peel velocity. Fig. 1 shows the relationship between the wave peel velocity upeel ,
the wave velocity uwav e and the velocity of the surfer usurf . The main goal for the surfer is to stay in the region where
the wave starts to break. This region is located between the unbroken and the broken part of the wave with the white
water region. For maneuvering, the surfer’s velocity can be faster or slower than usurf . But to stay in the wave, the mean
velocity of the surfer should be the sum of the vectors of the wave peel velocity and wave velocity (Walker and James,
1974).
For describing waves, the most important parameters are wave height hwav e , wave peel angle β , wave breaking intensity
and wave section length (Scarfe et al., 2003). The breaking intensity of a wave is influenced by strength and direction of
the wind and the shape of the seabed (Mead and Black, 2001). The wave section length represents the length of the
different conditions within a wave. Surfers desire waves where the breakpoint peels along the wave crest (Scarfe et al.,
2003). The wave peel angle β is the angle between the wave peel velocity and the surfer’s velocity. It varies between 0◦
(usurf = upeel ) and 90◦ (usurf = uwav e ); see Fig. 1. For most surfers, a peel angle of 35◦ to 65◦ is ideal (Beggs-French, 2009).
A wave is called right-hand wave when the wave peels to the right from a viewing perspective of the surfer looking to
the shore and a wave is called left-hand wave when the wave peels to the left (Scarfe et al., 2003).
The typical surfer’s velocity usurf is approximately 6 m/s for a wave height hwav e of 1.5 m and a wave velocity uwav e of
4 m/s (Paine, 1974). Dally (2001) reported that on the one hand the board speed increases with increasing wave height
and on the other hand the form drag of the board is limiting the speed of the surfboard.
A surfboard fin represents a streamlined body in a fluid flow; see Fig. 2(A). Because of the similarity to airfoils, lift
and drag forces prevail at fins in the same way. The lift force FL appears perpendicular and the drag force FD in direction
of the main flow (Durst, 2007). The lift forces are associated with surfboard maneuverability and stability, whereas drag
forces have a great influence on the surfing velocity (Gudimetla et al., 2009). Surfers want to keep their velocity during
maneuvers to stay ahead of the wave’s breakpoint in regular surfing and to be able to ride through the wave’s pocket up
to the lip in high performance surfing; see Fig. 3. Therefore, the fins should not generate too much drag. Furthermore, we
suggest that surfers want an easy to handle and controllable surfboard for all surfing situations. In the best case, the fins’
lift forces stabilize the surfer’s movement during maneuvers providing a better maneuverability. Here, stability must be
subdivided into roll and track stability. Track stability is the degree of control the surfer has over the board to follow his
desired path through the wave and roll stability is for balancing along the longitudinal axis of the board.
In Fig. 2(A), the lift and drag forces are shown at the center fin for an AoA (α ) of 10◦ . In this study, forces were
determined as non-dimensional lift and drag coefficients as defined by Durst (2007):
FL
cL = 1
2
· ρ · usurf 2 · A
300 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

Fig. 1. Realistic expansion of the left-hand wave’s whitewater region from t1 to t2 . The three velocities which prevail for the wave and the surfer
are the wave peel velocity Upeel , the wave velocity Uwav e , which is perpendicular to Upeel , and the desired reference velocity of the surfer Usurf . The
peel angle of the wave β describes the angle between Upeel and Usurf .

FD
cD = 1
2
· ρ · usurf 2 · A
FL and FD are the lift and drag forces, A is the reference area (the planform area of the fins), usurf is the surfers velocity
and ρ is the fluid density.
For each fin, the lift and drag coefficients were calculated separately, and the mean coefficients for all fins together
were calculated by the sum of the three fin forces relating to the sum of the three fin areas (Gudimetla et al., 2009).
Transferring the simulation results to the real surfing case, a positive lift force at the fins represents a turn to the shore,
and a negative lift force presents a turn into the wave. Therefore, the axis of rotation is perpendicular to the surfboard
surface, and its position is assumed to be at the midpoint between the leading edges of the front fins; see Fig. 5(A). A
positive drag force is oriented parallel to the main flow direction and therefore mainly lowers the surfboards speed. These
correlations depend on the definition of the coordinate system at the fin configuration.
Additionally, the lift-to-drag ratio (L/D) shows the maneuverability and stability of the surfboard during maneuvers. In
aircraft aerodynamics, the L/D ratio is a measure for the fuel efficiency, i.e. a high lift and a low drag is desired resulting
in a large L/D ratio (Anderson et al., 2010). In surfing situations, surfers ride the wave up and down by performing bottom
turns or carvings and need, therefore, good maneuverability with nearly no loss of speed.
The AoA is defined as the angle between the surfboard’s longitudinal axis and the direction of the fluid flow. An AoA of
0◦ corresponds to a fluid flow parallel to the surfboard’s longitudinal axis. For this, AoAs in the range of −45◦ to 45◦ are
possible (Beggs-French, 2009). Although large AoAs in performance and professional surfing will be reached, maximum
AoAs of up to 10◦ – 15◦ are usually applicable for regular surfing (Lavery et al., 2009).
Surfing is generally structured in the following steps. Before the wave reaches the surfer, he or she has to increase the
speed of the board by paddling to match the wave velocity uwav e . Usually, the surfer lies on his/her belly and paddle with
the hands. After dropping into the wave, the surfer stands up and surfs down the waves pocket as displayed in Fig. 3.
Finally, the surfer performs the bottom turn which is a common maneuver. Reaching the wave pocket, the surfer surfs
up and down the wave in alternating order. Thereby, the absolute velocity of the surfer will vary around uwav e traveling
with the wave. In our study, a bottom turn is defined by positive AoAs.

3. Fin model and numerical methods

3.1. Geometry dimensions

The numerical model and its geometric dimensions are based on the original THE DREAMCATCHER surfboard model
(ROBERTS, Ventura, CA/USA) with plugged FCS Accelerator fins (FCS, Newport Beach, NSW/Australia). The length of the
surfboard is 6′ 6′′ (1.98 m). The surfboard with the mounted fins was scanned by computer tomography (CT) using a CT
system of type SOMATOM Definition AS 64 (SIEMENS Healthineers, Forchheim, BY/GER); see Fig. 4.
In order to reconstruct the surface of the surfboard and the fins, we first recorded two high-resolution CT images with
512 × 512 × 937 (width x thickness x length) sample points each. The images were segmented into binary volumes using
MATLAB and subsequently smoothed with a Gaussian filter. The contours of the surfboard and the fins are represented
S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314 301

Fig. 2. (A) Lift FL , drag FD and resultant forces FR applying on the fins under an AoA of α = 10◦ (Bottom view, left-hand wave). (B) Surfboard, its
longitudinal axis, and the fluid flow with an AoA of α = 10◦ (Bottom view, left-hand wave).

Fig. 3. M. Döllinger surfs a bottom turn in a (left-hand) wave. (1) The Lip: Top of the wave. (2) Falling Lip: Wave zone is thrown forward as the
wave breaks. (3) The Pocket: Optimal energy zone for getting the maximum surfing velocity. (4) Flat Zone: Zone in front of the broken wave, will
slow down the surfing velocity.

by iso-surfaces, which were extracted from the volume using a modified version of the Marching Cubes algorithms which
generates watertight and manifold triangle meshes (Grosso, 2016).
302 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

Fig. 4. Exemplary proceeding of CT scanning and post-processing: (A) A NSP Surfboard positioned in the CT for scanning; (B) 3D geometry of a CT
scanned surfboard and fins; (C) Post-processed 3D data: Surfboard tail cut away and fins imported in STAR-CCM+.

Fig. 5. (A) Dimensions of our thruster configuration in the axial and transverse direction with fin depth and base. (B) The position of the axis of
rotation to realize the AoA.

After extracting the board contour, the whole 3D geometry (board and fins) was imported in STAR-CCM+. For the
CFD simulations, only the fins in the thruster configuration were considered; see Fig. 5(A). The two side fins have an
asymmetric and the center fin a symmetric profile; see Fig. 6(B). The cant angles of the two side fins are inclined to the
rails amounting approximately 8.5◦ ; see Fig. 6(A). The toe-in angles of the two side fins are approximately 3.5◦ and are
inclined to the longitudinal axis of the surfboard; see Fig. 6(A).
To get a distinct and consistent denotation for the fins in the thruster configuration, independent from surfing a
right-hand or left-hand wave, they are named as follows.

• Center fin (CF): the rear fin in the middle of the configuration.
• Outside fin (OF): the fin next to the shore.
• Inside fin (IF): the fin next to the wave crest.

3.2. Simulation region and boundary conditions

The numerical fluid domain is represented by a rectangular simulation region which is chosen similar to Gudimetla
et al. (2009); see Fig. 7. Its dimensions are 2.5 m in length, 1.0 m in width and 0.5 m in height. The bottom wall of the
rectangular box represents a simplified surfboard the fins are mounted on. They were set in a distance of three times the
base length of the fins from the inlet boundary; see Fig. 7.
S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314 303

Fig. 6. (A) Cant and Toe In angle of the surfboard fins. (B) The asymmetric and symmetrical profiles of the rail fins and the center fin are visible
at the bottom view of the surfboard.

Fig. 7. Geometry of the numerical simulation region with the inserted fins positioned in the thruster configuration at an AoA of 45◦ . The leading
edges of the two side fins have a distance of three times the base length of the fins from the inlet boundary at an AoA of 0◦ .

The fins and the bottom wall of the simulation region were defined as an impermeable wall with a no-slip condition. At
the inlet boundary of the fluid region, a homogeneous fluid velocity profile of 5 m/s was defined. At the outlet boundary,
downstream of the fins, the pressure was set to be constant at 0 Pa. No reversed flow or diversion of the fluid flow occurs
at the outlet boundary. The two side walls and the top wall were defined as symmetry planes.
To realize the range of AoA, the fins were rotated around an axis of rotation in the simulation region, which is located
at the midpoint between the two leading edges of the side fins, so that the thruster configuration is riding a left-hand
wave for 0◦ < AoA < 45◦ ; see Fig. 5(B).
2
The density and the kinematic viscosity of water were specified as ρ = 997 m3 and ν = 0.8926 · 10−6 ms corresponding
kg

to a water temperature of 25 C. The density was assumed to be constant (Mach number (Ma) < 0.3, vinlet ,max = 5 m

2
,
Ma5 m = 0.003), resulting in an incompressible fluid flow.
s

3.3. Mesh generation and independence study

For modeling the boundary layers at the fins in a proper manner and getting accurate results of lift and drag forces,
we specified five prism layers with an overall thickness of 2.8 mm added to the surfaces. Prism layers are hexahedral cell
layers that are aligned with the fin contour that help to reduce numerical diffusion in the wall boundary layers (Siemens,
2016).
304 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

Fig. 8. Drag coefficients as a function of the number of control cells. Six different meshes with an increasing number of cells and a decreasing target
cell volume. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

To identify the required grid resolution, a mesh independence study was performed. Six different mesh configurations
(Mi ) with an increasing mesh density were analyzed; see Fig. 8. It shows the trend of the fin drag coefficients as a function
of the number of cells. The deviations of the drag coefficients for meshes M5 and M6 ranges between 0.0% and 0.2%
compared to M4 . It yielded that the mesh configuration M4 with approximately 1.18 million cells is fine enough for the
simulations with a three-dimensional unstructured polyhedral grid.

3.4. Numerical methods

The commercial CFD software STAR-CCM+ (Siemens PLM software, Plano, TX/USA) was used to simulate the fluid
flow around the fins in the simulation region. To solve the Navier–Stokes equations, a cell-center Finite-Volume-Method
(FVM) was used. A second-order accuracy central difference scheme was applied to discretize the convective and diffusive
flux. The resulting linear system was solved iteratively by an algebraic multigrid method with a Gauss–Seidel relaxation
scheme. For modeling turbulence, the Reynolds-Averaged Navier–Stokes (RANS) and Unsteady Reynolds-Averaged Navier–
Stokes (URANS) equations were solved using the SST k − ω two-equation eddy-viscosity model (Menter et al., 2003). This
model combines the performance advantages of the k − ε and the k − ω turbulence models. To resolve the boundary
layers the k − ω model and for the free flow field, the k − ε model is used.
While the RANS simulations were performed for the whole range of the AoAs from 0◦ to 45◦ , the URANS simulations
were only performed for AoA ≥ 25◦ to resolve the unsteady flow separation from the fins at those large AoAs.
For the near-wall flow, STAR-CCM+ provides an all-y+ model which can handle both fine and coarse meshes (Siemens,
2016). According to this all-y+ model, y+ was specified to be in the range of 25 to 60 at the fin surfaces, depending on
the AoA. For the transient URANS simulations, the time step size was selected as 1 · 10−4 s which satisfies the Courant–
Friedrichs–Lewy (CFL) condition CFL = 1.0 ± 0.5 at the fins surfaces where the mesh is finest. The CFL number describes
the number of mesh cells that are transited by a fluid element during one time step.
While the steady-state simulations were performed at a local workstation with 16 GB RAM and an Intel⃝ R
CoreTM
i7-6700 (8 MB Cache, 3.4 GHz), the transient URANS simulations were performed on up to 16 computer nodes of the
Erlangen Regional Computing Center (RRZE) Emmy cluster (Erlangen Regional Computing Center, 0000). Each node of
the Emmy cluster consists of 20 Intel⃝R
Xeon⃝ R
E5-2660v2 cores (25 MB Cache, 2.20 GHz). 1500 iterations of the steady-
state simulations took around 2.5 h at the local workstation and the 35 000 time-steps of the URANS simulations, which
correspond to 7 runs of a flow particle through the simulation region, took in average 15 h on the high-performance
cluster.
S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314 305

Fig. 9. Lift coefficient versus the AoA of every single fin (blue, green and purple line). The red line represents the mean lift coefficient for all fins,
and the small figures of the 3-fin configuration represent the fluid flow around the fins at specific AoA.. (For interpretation of the references to
color in this figure legend, the reader is referred to the web version of this article.)

4. Simulation results and discussion

(1) Lift, drag, and lift-to-drag ratio - Maneuverability, stability, and speed (RANS)
The results in this section were obtained from the steady-state RANS simulations. Fig. 9 shows the lift coefficients
versus the AoA. For increasing AoA, the lift coefficient increases in a linear fashion from zero until the maximum is reached.
After the maximum, stall occurs and the lift coefficient decreases. The inside fin produces the largest lift with a maximum
of 0.99 at an AoA of 20◦ , whereas the maximum of the mean lift coefficient for the entire 3-fin configuration is 0.74 at an
AoA of 20◦ . Although the two side fins have the same geometry, the behavior of the lift coefficients is different compared
to each other. This appears primarily because of the asymmetric profile and the varying orientation, as a result of the
Toe In angle, of the side fins; see Fig. 6(B). The stall point of the outside fin is reached at an AoA of 12.5◦ with a local
maximum of 0.56. After the stall point, the lift coefficient first decreases followed by a further increase for AoA > 25◦ to
reach a total maximum of 0.61 at an AoA of 35◦ . This deviation of the lift trend as function of the AoA originates from
the interaction between the outside and the center fin. By moving the center fin in the wake of the outside fin, the flow
in the gap between the two fins is forced downward to an increasing degree for increasing AoA. This counteracts to the
effect of flow separation from the outside fin on the lift coefficient. The maximum values of the lift coefficients of the two
side fins at the stall points differ by around 38%.
Like the inside fin, the lift coefficient of the center fin increases up to its maximum of 0.74 at AoA = 20◦ . As can be
seen in Fig. 10(A)–(D) the effective AoA for the center fin is decreased by the impact of the outside fin upstream.
The obtained results of the lift coefficient suggest that the fins have an optimal range of AoAs where they produce
the most lift. Considering a center of rotation upstream of the fins, the lift forces the board tip to turn to the shore (AoA
of 0◦ ). If the surfer wants to ride a bottom turn, he has to exert a turning moment that is oriented against the turning
moment produced by the lift forces of the fins. Thus, the lift moment has to be in the same range as the turning moment
implied by the surfer, resulting in a dynamic equilibrium of the surfer-board system.
Relating to the interpretation of the lift force it is known that surf maneuvers like bottom turns, and carving do not
only depend on the fin forces but also on the buoyant forces when the board is inclined along its longitudinal axis and
on gravity forces exerted by the surfer by changing his center of mass (Mc Cagh, 2013).
Our results are related to a full fin-moistening neglecting the tipping of the board by the surfer (Carswell, 2007).
The drag coefficient is produced by the pressure and shear forces that act parallel to the main flow direction. Fig. 11
presents the drag coefficients of the three single fins and the coefficient of the entire 3-fin configuration as a mean value.
The drag shows a steady increase up to an AoA of 45◦ with a varying gradient. Sudden changes of the inclination occur
at AoA = 20◦ for the inside and the center fin and at AoA = 15◦ and 25◦ for the outside fin. Similar gradient devolution
306 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

Fig. 10. Governing velocity field around the outside and the center fin. (A)–(D) show the development of a reverse flow region in the slipstream of
the outside fin with an increasing AoA. (C) and (D) present the interaction of the outside fin’s recirculation region with the center fin and the free
stream.

can be observed for the mean drag coefficient of the entire thruster configuration. The inclination changes at AoA = 20◦
(inside and center fin) occur at the stall point positions of the lift coefficient devolution, shown in Fig. 9.
With increasing drag, the velocity of the surfer decreases. Therefore, the surfer slows down in high ranges of the AoA.
For bottom turns, where the fins are passing through a large range of AoA, high drag forces might represent a huge loss
of speed. However, as mentioned before, not only the fin forces are influencing surf maneuvers.
Analysis of the lift-to-drag ratio as a function of the AoA was only conducted for the mean coefficients; see Fig. 12.
The lift-to-drag ratio shows a steep increase up to a peak value of 7.71 at an AoA of 6.25◦ . After this maximum, the curve
decreases to a lift-to-drag ratio of 1.0 at an AoA of 45◦ .
This maximum L/D ratio at a relatively small AoA shows the large initial stability of the board in its initial position
at AoA = 0◦ . To initialize a maneuver, the surfer has to raise a large turning moment to deflect the board in contrast to
larger AoAs, at which smaller moments are sufficient.
(2) Validation and Discussion (RANS)
As mentioned in the introduction, there are a few numerical and experimental studies of surfboard fin that are
compared to our results, although different fin geometries were used in these studies.
Gudimetla et al. (2009) presented a computational study of a commercial FCS k2.1 fin in a 3-fin configuration. The
qualitative trend of their mean lift coefficient of the entire configuration is similar to the results presented in our study,
see Fig. 13(A) and (B). Accordingly, Fig. 13(B) displays a qualitatively similar trend of the drag coefficient. Moreover, the
trends of lift and drag coefficients exhibit characteristic devolutions and features compared to airfoils as especially the
extensive stall point and the succeeding decrease of the lift at larger AoAs Durst (2007).
However, quantitatively, we found an earlier stall point at an AoA of 20◦ , a slower decrease after the stall point and
a 18.7% lower maximum mean lift coefficient compared to Gudimetla et al. (2009). Furthermore, the drag coefficient in
our study is smaller over the entire AoA range. These quantitative deviations of both coefficients obviously are the result
from the different fin shapes and profiles of the fins as well as a different geometric arrangement of the fins on the board
used in Gudimetla et al. (2009).
As a consequence, the qualitative trend of the lift-to-drag ratio is also similar compared to Gudimetla et al. (2009).
The peaks are located at similar low AoA of approximately 6.25◦ , but the peak value was found to be 7.71 in our study
whereas (Gudimetla et al., 2009) reported a value of 4.25.
As the IF is located in the free stream for all AoA, Fig. 14 displays the lift and drag coefficients of the IF in
comparison to experimental studies reported by Brandner and Walker (2004) and Carswell (2007) who investigated
single-fin configurations. Brandner and Walker (2004) used a generic fin model, and Carswell used a RedX fin of type
X7. Additionally, Carswell also provided simulated data from the RedX X7 fin. In general, the qualitative trends of the lift
and drag coefficients matches quite well with the coefficients of the inside fin that is located in the free stream over the
S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314 307

Fig. 11. Drag coefficient versus the AoA of every single fin (blue, green and purple line). The red line represents the mean drag coefficient for all
fins, and the small figures of the 3-fin configuration represent the fluid flow around the fins at a specific AoA.. (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 12. The red line represents the mean lift-to-drag coefficient versus the AoA of the entire 3-fin configuration. The mean lift coefficient and the
mean drag coefficient are the blue and the green line, respectively. The small figures of the 3-fin configuration represent the fluid flow around the
fins at a specific AoA.. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

entire AoA range. The quantitative differences are again mainly based on different shapes and profile geometries of the
investigated fins. The RedX X7 fin possesses a different fin profile, and its planform area is approximately 7% smaller than
the FCS fin used in our study. Brandner and Walker (2004) used a generic fin with elliptical shape/planform and a half
profile of a NACA-4 airfoil, i.e., the NACA-4 curvature at the suction side and a flat profile at the pressure side.
308 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

Fig. 13. Mean lift-to-drag ratio versus the AoA. The solid red line represents the mean lift-to-drag ratio of the 3-fin configuration using the FCS
Accelerator fins, and the dashed blue line represents the lift-to-drag ratio of the 3-fin configuration using the FCS k2.1 fins (Gudimetla et al., 2009).
The small figures of the 3-fin configuration represent the fluid flow around the fins at a specific AoA.. (For interpretation of the references to colour
in this figure legend, the reader is referred to the web version of this article.)

Fig. 14. (A) Lift coefficient versus the AoA. The solid red line represents the lift coefficient of the IF of the study, the dashed purple line presents the
lift coefficient of the fin of Brandner and Walker (2004) and the solid green, and blue lines presents the numerical and experimental lift coefficients
of Carswell (2007) (B) Drag coefficient versus the AoA. The solid red line represents the drag coefficient of the IF of the study, the dashed purple line
presents the drag coefficient of the fin of Brandner and Walker (2004) and the solid green, and blue lines presents the numerical and experimental
drag coefficients of Carswell (2007). The small figures represent the fluid flow around the IF of the study at a specific AoA.. (For interpretation of
the references to colour in this figure legend, the reader is referred to the web version of this article.)

Even though a validation with equal fin shapes, profiles and arrangements are not available, the comparison to other
surfboard experimental and numerical fin studies shows the validity of the results presented here.
(3) Time-dependent effects and the deviation between RANS and URANS
The analysis of the lift and drag coefficients in part (2) were conducted for both, the steady-state and unsteady
simulations. As shown by the lift and drag coefficients, stall and fin interactions occur at a high AoA. According to Fig. 15,
stall occurs at an AoA = 15◦ for the outside fin and at an AoA = 25◦ for the center fin resulting in a detached flow region
at the suction side of the particular fin. To dissolve all appearing unsteady effects as a result of the fin interaction in this
high AoA range, transient URANS simulations are carried out.
The deviation of the URANS compared to the RANS simulations for lift coefficients of the thruster at high AoA shows
Fig. 16(A). Whereas the lift coefficient of the inside has nearly the same behavior for RANS and URANS in the AoA range
S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314 309

Fig. 15. (From left to right and from top to bottom) A chronological sequence of the development of the outside fin’s recirculation area starting
from an AoA of 5◦ up to an AoA of 45◦ . Instantaneous snapshots of the flow field within a plane parallel with a distance of 0.03 m to the wall
boundary.

of 20◦ to 45◦ , the outside fin and center fin show great fluctuations for the transient URANS values. The transient lift
coefficient deviations for the inside fin ranges between - 4.2% and + 4.5%, the outside fin fluctuates between - 18.9% and
+ 0.8% and the center fin fluctuates between - 28.3% and + 14.6%.
Similar to the lift coefficient deviations the drag coefficient deviations show the same behavior, see Fig. 16(B). The drag
coefficient deviation of the inside fin ranges between - 5.9% and - 1.2%. Again, fluctuations for the outside and center fin
are much larger. The fluctuations for the outside fin ranges between −17.6% and - 0.1% and for the center fin between -
28.9% and + 9.7%.
310 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

cL,URANS
Fig. 16. (A) Mean lift coefficient deviation of the URANS results compared to the RANS results at high AoA given in cL,RANS
%. Dashed lines represent
the maximum absolute deviation of each single fin (blue, green and purple) and the scope of the deviation of the entire configuration (red). (B) Mean
c
drag coefficient deviation of the URANS results compared to the RANS results at high AoA given in cD,URANS . (For interpretation of the references to
D,RANS
color in this figure legend, the reader is referred to the web version of this article.)

Both lift and drag deviations show that the time-dependent results for the outside fin can differ by up to 28.9%. The
results for both coefficients show mostly an overestimation of the mean value calculated by RANS compared to URANS
for the whole range of AoAs.
Fig. 17 displays instantaneous snapshots of the flow field within a plane parallel to the wall boundary for AoA = 40◦ ,
see also supplement video Outside&CenterFin_AoA40.avi. They show vortex shedding and interaction between the outside
and center fin in a time interval from 5.500 s to 5.562 s of the simulation time. The process of vortex shedding can be
described as follows:

1. The fluid flow detaches from the suction side (right side) of the outside fin and center fin.
2. Outside fin:
a. A steady recirculation area arises between the leading edges of the outside and center fin.
b. Every 63 ms, a vortex separates from the (lower) trailing edge of the outside fin and is convectively
transported in the direction to the center fin. This vortex separation occurs periodically at a frequency of
15.8 Hz at this AoA = 40◦ .
3. Center fin:
a. Two recirculation areas detach alternately from the leading and trailing edge.
b. The frequency of this alternating vortices separation is 14.3 Hz.

Whereas the center fin shows the typical alternating vortex shedding process of a Karman vortex street, the outside fin
exhibits a steady vortex originating from the leading edge and a periodically separating vortex from the trailing edge.
The reason for the steady position of the leading edge vortex is that it is stabilized in its position by the center fin that is
located in the wake of the outside fin. In contrast, the vortex shedding from the outside and center fin induces periodically
oscillations in the drag and lift coefficients as shown in the embedded diagrams in Fig. 18.
For an increasing AoA between 20◦ and 45◦ , Fig. 18(A) and (B) additionally show the maximum amplitude of the lift
and drag coefficients. Both coefficients exhibit a massive amplitude increase at the center fin for an increasing AoA up to
40◦ followed by a decrease at AoA = 45◦ . Thereby, a maximum amplitude of ± 46.8% of the mean lift and ± 46.0% of the
mean drag coefficient were detected.
The oscillation of the lift coefficient incorporates a higher frequency of 15.05 Hz and a lower beat frequency of 1.5 Hz.
This is the result of the vortex shedding processes of the outside and center fin with the slightly different shedding
frequencies (outside fin: 15.8 Hz; center fin: 14.3 Hz), see Fig. 19. The lift and drag coefficients at the outside fin exhibit
the same higher frequency and lower beat frequency components but with much smaller amplitudes.
The reason for the largest amplitudes at AoA = 40◦ at the center fin is that in this configuration the center fin is
excited by both the shedding vortices from the outside fin and the shedding vortices from the center fin itself (see Fig. 17
and supplement video file Outside&CenterFin_AoA40.avi). Thus, at AoA = 40◦ the interaction between outside and center
fin is maximized owing to the special in-series orientation of both fins. Furthermore, this special orientation potentially
S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314 311

Fig. 17. (From left to right and from top to bottom) A chronological sequence of the outside fin’s vortex shedding within an overall time interval
of 0.063 s at an AoA of 40◦ . Instantaneous snapshots of the flow field within a plane parallel with a distance of 0.03 m to the wall boundary.

results in a decrease of the deviation of the mean lift and drag coefficients resulting between RANS and URANS as shown
in Fig. 16 whereas the oscillation amplitudes of lift and drag are maximum. In contrast, at AoA = 35◦ , the center fin is
majorly excited by its vortex shedding process (see Fig. 15 and supplement video file Outside&CenterFin_AoA35.avi). At
AoA = 45◦ , the center fin excitation is mainly produced by the shedding vortices originating from the outside fin whereas
the vortex shedding from the center find almost totally attenuated owing to its leading edge position deep in the wake
of the outside fin (see Fig. 15 and supplement video file Outside&CenterFin_AoA45.avi). These fluctuations of the lift and
drag coefficient of the center fin exerted a fluctuating force on the center fin that may have a negative effect on the
surfboard stability.
Strong time-dependent forces will be induced to the fin from that the flow detaches and to the fin where the vortices
are impinging because of the flow detachment and the resulting vortex shedding. These effects may have a great impact
on the surfboards maneuverability and stability.

5. Shortcomings

Throughout the process of preparing the numerical model, the surfboard was approximated as a flat wall. Thus, no
influences of the shape and planform of the surfboard on the performance parameters were considered. However, the
312 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

Fig. 18. (A) Increase of the maximum lift coefficient amplitude versus the AoA. Fluctuation of the lift coefficient is caused by the eddy shedding
of the time-dependent case. The small figure inside shows the time-dependent lift coefficient at an AoA of 40◦ . (B) Increase of the maximum drag
coefficient amplitude versus the AoA. The small figure inside shows the time-dependent drag coefficient at an AoA of 40◦ . The amplitudes of the
lift and drag coefficient fluctuations were calculated with URANS. (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

Fig. 19. The lift coefficient oscillation frequencies and the vortex shedding frequencies of the OF (red) and CF (green) versus the AoA. (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

bottom wall can be seen as a very simplified board. No surfboard-fin-fluid interaction, including surfboard geometry
details like the nose, the tail, and the rocker, was reproduced and no free surface was considered in the simulation region.
Furthermore, surfer–surfboard interaction was not included here.
For further investigations extensions of this work should be:
1 Investigation of other fin configurations like the 4-fin configuration called the Quad with steady-state and time-
dependent simulations for receiving the overall forces and the unsteady effects. The results of the investigation of the
Quad configuration can also be compared to the results of Gudimetla et al. (2009).
2 Furthermore, parameter studies within the 3- and 4-fin configuration by simulating different symmetric configura-
tions, in which the rear fins are positioned along the axial and transverse direction, should be performed.
3 A surfboard, a surfer, and the free surface, which is represented by the Volume of Fluid (VOF) method provided by
STAR-CCM+, should be included within the simulations to get a more realistic modeling.
S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314 313

4 Experimental investigations in surfboard fins should be executed to validate the numerical results.

6. Conclusion

This study investigated the hydrodynamic performance of the 3-fin configuration with both steady-state and time-
dependent simulations. To the best of the authors’ knowledge, for the first time, the fluid flow around surfboard fins
was solved with the time-dependent URANS equations. The CFD modeling followed (Lavery et al., 2009; Gudimetla et al.,
2009; Livanos, 0000; Oggiano and Pierella, 2018) which demonstrated that computational fluid dynamics is an adequate
tool to investigate single parts of the surfboard-fin system.
Furthermore, not only the mean lift and drag coefficients for the entire 3-fin configuration were calculated, but also
the lift and drag coefficients relating to each single fin were detected.
The results from this investigation of our thruster configuration show that for a complete dissolution of all unsteady
effects that occur during fin interaction a time-dependent simulation case is necessary. Hypothesis 1 was confirmed since
these unsteady flow phenomena have a great impact on the track and roll stability of the surfboard. The overall results
for RANS and URANS simulations are following qualitatively the results of Gudimetla et al. (2009).
The numerical results of the simulations showed hydrodynamic effects that are influencing the surfing stability during
basic maneuvers, confirming our hypothesis 2. Our interpretation of the results showed that the link between numerical
simulations and real (basic) surfing situations and surf maneuvers is possible in the future.
The new findings of the investigation for our thruster configuration study relating to the surfboards maneuverability
and stability can be summarized as follows:
1 The lift forces generated from our 3-fin configuration produce a turning moment which forces the board to be turned
in direction to the shore (AoA = 0◦ ). This helps to improve the surfboard to be controllable and stable.
2 Our 3-fin configuration mostly tends to rotate the surfboard back to an AoA of 0◦ at an AoA of 20◦ ; see Fig. 9.
3 To speed up surfing velocity the surfer should choose small AoAs with low drag forces; see Fig. 11.
4 High negative influences on stability can cause of the high fluctuation amplitudes for lift and drag in the AoA range
between 25◦ and 45◦ ; see Fig. 18.
To the best of our knowledge, no such deep discussion about the influences of the fin forces on realistic surf maneuvers
was presented before.

Acknowledgments

The authors acknowledge support from the Central Institute for Scientific Computing (ZISC), Germany and computa-
tional resources and support provided by the Erlangen Regional Computing Center (RRZE), Germany. The authors also
gratefully acknowledge Tom O’Keefe (President at Daum Tooling Inc., San Clemente, CA/USA) for support in discussion
and interpretation of the simulation results. We acknowledge Felix Böschen for his preliminary work in his master thesis
who set up first fin simulations at our department.

Declaration of competing interest

None

Appendix A. Supplementary data

Supplementary material related to this article can be found online at https://doi.org/10.1016/j.jfluidstructs.2019.07.006.



Six videos of the outside and center fin’s vortex shedding in a range of AoA from 20◦ to 45 :

• Outside&CenterFin_AoA20.avi: Video of the outside and center fin’s vortex shedding at an AoA 20◦
• Outside&CenterFin_AoA25.avi: Video of the outside and center fin’s vortex shedding at an AoA 25◦
• Outside&CenterFin_AoA30.avi: Video of the outside and center fin’s vortex shedding at an AoA 30◦
• Outside&CenterFin_AoA35.avi: Video of the outside and center fin’s vortex shedding at an AoA 35◦
• Outside&CenterFin_AoA40.avi: Video of the outside and center fin’s vortex shedding at an AoA 40◦
• Outside&CenterFin_AoA45.avi: Video of the outside and center fin’s vortex shedding at an AoA 45◦

The flow field is presented within a plane parallel with a distance of 0.03 m to the wall boundary.

References

Anderson, J.D., Aerodynamics, Fundamentals.of., ed, 5th., 2010. Fundamentals of Aerodynamics, fifth ed. McGraw-Hill Education, New York.
Azcueta, R., Rousselon, N., 2009. CFD Applied to super and mega yacht design, in design. In: Construction and Operation of Super and Mega Yachts
Conference.
Barnett, N.D., Miravete, E.G.-., 2009. A study of planning hydrodynamics. In: Proceedings of the COMSOL Conference 2009 Boiston.
Beggs-French, R., 2009. Surfboard Hydrodynamics. The University of New South Wales at the Australian Defence Force Academy.
Brandner, P.A., Walker, G.J., 2004. Hydrodynamic performance of a surfboard fin. In: 15th Australasian Fluid Mechanics Conference.
314 S. Falk, S. Kniesburges, R. Janka et al. / Journal of Fluids and Structures 90 (2019) 297–314

Buckley, R., 2002. Surf tourism and sustainable development in indo-pacific Islands. 1. the industry and the Islands. J. Sustain. Tour. 10 (5), 405–424.
Carswell, D.J., 2007. Hydrodynamics of Surfboard Fins. Swansea University.
Catalano, P., Amato, M., 2003. An evaluation of RANS turbulence modelling for aerodynamic applications. Aerosp. Sci. Technol. 7 (7), 493–509.
Catalano, P., Tognaccini, R., 2011. RANS analysis of the low-Reynolds number flow around the SD7003 airfoil. Aerosp. Sci. Technol. 15 (8), 615–626.
Dally, W., 2001. The maximum speed of surfers. J. Coast. Res. (29), 33–40.
Durst, F., 2007. Fluid Mechanics: An Introduction to the Theory of Fluid Flows. Springer, Berlin Heidelberg.
Erlangen Regional Computing Center (RRZE) HPC. [Online]. Available: https://www.anleitungen.rrze.fau.de/hpc/emmy-cluster/ [Accessed: 21-Jun-2018].
Grosso, R., 2016. Construction of topologically correct and manifold isosurfaces. Comput. Graph. Forum 35, 187–196.
Gudimetla, P., Kelson, N., El-Atm, B., 2009. Analysis of the hydrodynamic performance of three- and four-fin surfboards using computational fluid
dynamics. Aust. J. Mech. Eng. 7, 61–67.
Knudsen, S.S., 2013. Sail Shape Optimization with CFD.
Lavery, N., Foster, G., Carswell, D., Brown, S., 2009. CFD Modelling of the effect of fillets on fin drag. Reef J. 1 (1), 93–111.
Livanos, A., Computational Fluid Dynamics Investigation of Two Surfboard Fin Configurations, University of Western Australia.
Macneill, M.S., 2015. Bio-Inspired Optimal Fin Shape and Angle for Maximum Surfboard Stability. Michigan Technological University.
Mc Cagh, S., 2013. The Surfboard Book: How Design Drives Performance. McCagh O’Neill Pty Ltd.
Mead, S., Black, K., 2001. Predicting the breaking intensity of surfing waves. Spec. Issue J. Coast. Res. Surf. (103), 51–65.
Menter, F.R., 1993. Zonal two equation k-w turbulence models for aerodynamic flows. AIAA J..
Menter, F.R., Kuntz, M., Langtry, R., 2003. Ten years of industrial experience with the SST turbulence model. Turbul. Heat Mass Transf. 4 (4), 625–632.
Mørch, H.J., Perić, M., Röper, J., Schreck, E., 2009. CFD-Supported Design of Lifeboats. Wiesbaden.
Oggiano, L., 2017. Numerical comparison between a modern surfboard and an alaia board using computational fluid dynamics (CFD). In: Proceedings
of the 5th International Congress on Sport Sciences Research and Technology Support. pp. 75–82.
Oggiano, L., Pierella, F., 2018. CFD For surfboards: Comparison between three different designs in static and maneuvering conditions. Proceedings 2
(6), 309.
Paine, M., 1974. Hydrodynamics of Surfboards. University of Sydney.
Peng, S.-H., Eliasson, P., Davidson, L., 2007. Examination of the Shear Stress Transport Assumption with a Low-Reynolds Number K - ω Model for
Aerodynamic Flows.
Rezaei, F., Pasandideh-Fard, M., 2013. Stall simulation of flow around an airfoil USing les model and comparison of rans models at low angles of
attack. In: 15th Conference on FLuid Dynamics.
Sakellariou, K., Rana, Z.A., Jenkins, K.W., 2017. Optimisation of the surfboard fin shape using computational fluid dynamics and genetic algorithms.
Proc. Inst. Mech. Eng. P 231 (4), 344–354.
Scarfe, B.E., Elwany, M.H.S., Mead, S.T., Black, K.P., 2003. The Science of Surfing Waves and Surfing Breaks - a Review.
Siemens_PLM_Software, STAR-CCM+ Documentation Version 11.06, 2016.
Walker, J.R., James, K.K., 1974. Recreational Surf Parameters. Look Lab. Oceanogr. Eng. Dep. Ocean Eng. Univ. Hawaii, Honolulu.

You might also like