You are on page 1of 17

Journal of Fluids and Structures 65 (2016) 455–471

Contents lists available at ScienceDirect

Journal of Fluids and Structures


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

A scaling law for thrust generating unsteady hydrofoils


Timothy C.W. Lau n, Richard M. Kelso
Centre For Energy Technology, School of Mechanical Engineering, The University of Adelaide, 5005 SA, Australia

a r t i c l e i n f o abstract

Article history: An investigation into the flow around a fully submerged fish-inspired heaving and
Received 23 September 2015 pitching hydrofoil was performed using a combination of a quasi-steady model, based on
Received in revised form classical hydrodynamics, and experimental force measurements. From the quasi-steady
13 April 2016
model, a scaling law for the time-averaged foil thrust, based on the non-dimensional term
Accepted 24 June 2016
St (St − St0) where St is the Strouhal number and St0 is the “neutral force” Strouhal number,
Available online 25 July 2016
was proposed. Experimental force measurements were conducted for a wide range of
Keywords: non-dimensional heave amplitudes, 0.25 ≤ h0 /c ≤ 1, pitch amplitudes, 0 ≤ θ0 ≤ 45° and
Unsteady foil Strouhal numbers, 0.1 ≤ St ≤ 0.95, to confirm the accuracy of the model and scaling law.
Quasi-steady model
The Reynolds number for these experiments spanned the range 1500 ≤ Re ≤ 12500 . For
Aquatic propulsion
most flow cases, the experimental results show good agreement with the quasi-steady
model and show excellent collapse when plotted against the proposed scaling term
St (St − St0). However, under conditions where unsteady flow dynamics (such as dynamic
stall) are expected to be important, i.e. when the non-dimensional heave amplitude is
large relative to the pitch amplitude, the experimental results depart from the model and
the scaling law. This shows that the quasi-steady model, although simple, is sufficient at
predicting the foil thrust development when unsteady effects are small, and that any
departure from scaling law can be used to indicate that the flow regime is dominated by a
range of unsteady flow effects.
& 2016 Elsevier Ltd. All rights reserved.

1. Introduction

This study investigates the dynamics of a fish-inspired unsteady hydrofoil. The inspiration for this investigation is the
assumed notion that the process of evolution leads to optimal designs, and therefore, the propulsion dynamics of aquatic
animals should be very effective. In the current investigation, the scope of study is limited to steadily swimming aquatic
animals that predominantly use their caudal tails to generate propulsion (i.e. carangiform fish, consistent with the classi-
fication of Breder, 1926, 1926). Carangiform swimmers are not only ubiquitous in nature, but are also associated with high
powered locomotion (Lighthill, 1975; Webb, 1984; Triantafyllou and Triantafyllou, 1995).
Traditionally, the propulsion of carangiform swimmers has been explained in terms of the forces generated by the
motion of the carangiform caudal tails. These tails are typically modeled as wing-like structures undergoing simultaneous
heaving and pitching motions (Lighthill, 1969; Blake, 1983; Yates, 1983; Videler, 1993; Vogel, 1996; Dickinson, 1996; An-
derson et al., 1998; Sfakiotakis et al., 1999; Triantafyllou et al., 2000). If the angles of attack are favorable, then intuitively, the
carangiform caudal tails may act like hydrofoils in a free-stream, generating lift forces at each instantaneous angle of attack.

n
Corresponding author.
E-mail addresses: timothy.lau@adelaide.edu.au (T.C.W. Lau), richard.kelso@adelaide.edu.au (R.M. Kelso).

http://dx.doi.org/10.1016/j.jfluidstructs.2016.06.015
0889-9746/& 2016 Elsevier Ltd. All rights reserved.
456 T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471

Across an entire tail oscillation period, the upstream facing component of this lift may produce nett thrust to propel the
actively swimming fish.
Early mathematical studies have shown that these unsteady wing-like surfaces are capable of generating thrust at high
propulsive efficiencies, ηF ≈ 80–92% (Bose and Lien, 1989; Chopra, 1976; Fish, 1993; Rohr and Fish, 2004; Webb, 1988; Yates,
1983). Here, we note that the propulsive (Froude) efficiency is defined as
FtU∞
ηF =
Pin

where Ft is the time-averaged thrust force, U∞ is the steady swimming speed of the fish (or the free-stream velocity relative
to the foil), and Pin is the power required to produce the oscillating motion.
More recently, experiments utilising fully submerged hydrofoils undergoing simultaneous heaving and pitching motion
have shown that unsteady foils can either produce large thrust coefficients, or produce low thrust coefficients very effi-
ciently (Anderson et al., 1998; Hover et al., 2004; Read et al., 2003). These experiments indicate that one of the key para-
meters that characterise the overall propulsive performance of an unsteady foil is the Strouhal number, defined as
2fh0
St =
U∞ (1)

where f is the oscillating frequency, and h0 is the heave amplitude of the foil. Crucially, these studies found that in all
examined cases, the thrust coefficient increases monotonically with the Strouhal number, while peak efficiencies can ap-
proach 64% ≲ ηF ≲ 87% at 0.2 ≲ St ≲ 0.4 . Interestingly, in a perturbation analysis of an unsteady foil, this Strouhal number
range was found to approximately correspond to the “maximum amplification of vortical structures” within the wake
(Triantafyllou et al., 1993). The importance of this Strouhal number range is also reflected in the observed swimming ki-
nematics of fish, with some surveys revealing that the majority of steadily swimming carangiform fish tended to swim in the
range 0.2 ≲ St ≲ 0.4 , regardless of body geometry and/or swimming speed (Fish, 1993; Triantafyllou et al., 1993; Fish, 1998;
Rohr and Fish, 2004; Taylor et al., 2003). However, more recently, the role of the Strouhal number as the sole parameter that
characterises the propulsive performance of an unsteady foil is debated (Paraz et al., 2016; Young and Lai, 2004, 2007). In
studies of flexible oscillating panels, Dewey et al. (2013) found that a condition for optimal thrust production is for the
Strouhal number to be in the range 0.2 ≲ St ≲ 0.4 , while Paraz et al. (2016) found that efficient thrust develops when the
forced oscillation matches the first eigenvalue of the resonance frequency of the panel-fluid system, which occurs in the
much wider Strouhal number range 0.08 ≲ St ≲ 0.5. Numerical studies have shown that peak propulsive efficiencies of
unsteady foils and steady locomotion of carangiform swimmers occur at higher Strouhal numbers when the Reynolds
number is low (Borazjani and Sotiropoulos, 2008, 2010; Dong et al., 2006). Furthermore, more recent surveys of aquatic
swimmers suggest that there is a large scatter in the measured Strouhal number of steadily swimming fish, and that this
Strouhal number is also a function of fish size and swimming velocity (Eloy, 2012; Lauder and Tytell, 2006).
Previous studies have also indicated that the resultant wake structure of an unsteady foil may be linked to the propulsive
performance of the foil (Dong et al., 2006; Hover et al., 2004; Read et al., 2003; Anderson et al., 1998; Young and Lai, 2004).
While the morphology of the flow structure in the wake of a three-dimensional unsteady foil is highly complex, consisting
of multiple interconnected vortex loops that may vary depending on the foil geometry and kinematic parameters, in general
it has been shown that during efficient propulsion the foil wake structure results from vorticity generated at the leading
edge which combines constructively with vorticity generated at the trailing edge of the foil (Blondeaux et al., 2005; Dong
et al., 2006; von Ellenrieder et al., 2003; Fenercioglu and Cetiner, 2012; Parker et al., 2007). In many instances, the leading-
edge vortex is relatively large, and is generated when the instantaneous angle of attack of the foil is large (Anderson et al.,
1998; Blondeaux et al., 2005). This phenomenon is strongly linked to the unsteady (or transient) flow mechanism of “dy-
namic stall” (Choudhry et al., 2014; McCroskey, 1982), implying that fish may use unsteady flow mechanisms to augment
force production (Platzer et al., 2008; Triantafyllou et al., 2000).
In this study, we develop a simple quasi-steady model, based on classical hydrodynamics, of an unsteady foil in a free-
stream undergoing simultaneous heaving and pitching motion. From this model, a new scaling law for the time-averaged
thrust developed by the heaving and pitching foil under quasi-steady assumptions is proposed. The accuracy of the model
and the scaling law is then examined by comparing these to extensive experimental measurements of the forces developed
by an unsteady foil under a large range of heaving and pitching amplitudes.

2. Experimental procedure

The experiments consisted of an investigation of a fully submerged NACA 0026 foil, with a chord length c ¼50 mm and a
span s ¼200 mm. This foil geometry was selected as a compromise between generating sufficient stiffness in the supporting
struts, and accurately representing the fluke geometry of carangiform fish and mammals (Fish et al., 2007; Fontanella et al.,
2011).
The flow was generated in a recirculating water channel with a working cross-section of 500  500 mm and a length of
2000 mm. Upstream of the working section, the channel contains a series of flow conditioners consisting of two perforated
T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471 457

Fig. 1. Experimental setup.

plates, a honeycomb flow straightener and three nylon mesh screens. Finally, the flow passes through a 4:1 three-dimen-
sional contraction prior to entering the working section. The turbulence intensity within the working section was measured
to be 1.7% of the maximum free-stream velocity used in the current experiments.
The foil was placed in a recirculating water tunnel such that the centreline of the foil coincides with the centre of the
tunnel (see Fig. 1), ensuring that the free surface and the channel boundaries were as far away from the foil as possible. The
foil was connected to the heave platform using a 10 mm solid aluminium round rod, penetrating the foil 0.3c from the
leading edge of the foil. The deflection of the rod was found to be negligible. During initial dye flow visualisation performed
with this experimental setup (not shown here), the effect of the rod, the channel boundaries, and the free surface on the
flow was found to be negligible.
Strain gauges were employed to measure the foil chordwise and normal force components, as well as the pitching
moment around the foil pitching axis. These three independent measurements were made using 4 strain gauges each in a
temperature-compensated Wheatstone-bridge arrangement. The strain gauges were connected to a variable-gain, 3-input
strain gauge conditioning circuit. The output was logged using a 14-bit NI-USB 6009 data logger. Sampling rate was set at
200 Hz, which is 200 times the maximum flapping frequency used in these experiments. Typical recording time for each
experimental case was approximately 10 min, therefore providing approximately 120,000 data points per experimental
case. The force measurements were also repeated without water in the tunnel to measure the forces on the foil due to its
own inertia. This allows the foil's inertia to be subtracted from the force measurements, therefore providing a measure of
the hydrodynamic forces only. All measurements were digitally filtered using an order-4 Butterworth filter with a cut-off
frequency of 5 Hz.
The unsteady foil pitch motion was produced by a cam-follower system leading to a pitch motion described by

θ (t ) = arctan⎡⎣ tanθ0 cosϕ(t )⎤⎦ ≈ θ0cosϕ(t ) (2)

where θ0 is the pitch amplitude, ϕ(t ) = 2πft is the oscillation phase, f is the oscillation frequency (in Hertz), and t is time (see
Fig. 2). This pitching profile is similar to that used in other experiments utilising unsteady foils (Triantafyllou et al., 1993;
Anderson et al., 1998; Liao et al., 2003).
The foil heave motion was generated by a scotch yoke mechanism, and is described by

h(t ) = h0cos(ϕ(t ) − Ψ ) (3)

where h0 is the heave amplitude and Ψ is the phase angle between heave and pitch, here fixed at π/2 rad. It should be noted
that Ψ = π /2 rad was selected because it leads to an unsteady foil motion that approximates the tail carangiform swimmers
(Fish and Rohr, 1999) and has been shown to be correlated to efficient thrust generation in other unsteady foils (Moore,

2014; Read et al., 2003). The heave amplitude is normalised by the foil chord length, c, such that h0 = h0 /c . For convenience,
we also define the parameter χ, which is a measure of the dominance of the heave motion relative to the pitch motion,
458 T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471

Fig. 2. Definition of foil parameters.


h0
χ= .
sinθ0 (4)

The instantaneous angle of attack of the foil can be described by

α(t ) = arctanh˜′(t ) − θ (t ) (5)

where h˜′(t ) = h′(t ) /U∞ = πSt cosϕ(t ) is the heave velocity normalised by the free-stream velocity. For low pitch amplitudes,
the maximum angle that the foil achieves during its oscillation motion, αmax, can be estimated from
αmax ≈ arctan(πSt ) − θ0. (6)

In the current experiments, the non-dimensional heave amplitude was varied between 0.25 ≤ h˜ 0 ≤ 1 in increments of
0.25 and the pitch amplitude was varied between 0° ≤ θ0 ≤ 45° in increments of 15°, as summarised in Table 1. The resultant
Strouhal numbers were in the range 0.1 ≤ St ≤ 0.95 while the Reynolds number (based on the chord) ranged between
1500 ≤ Re ≤ 12500, where
ρ U∞ c
Re = ,
μ

ρ is the density of water and μ is the kinematic viscosity of water.

Table 1
Summary of foil kinematic parameters.

Group Set h̃0 θ0(°) χ St0, a St0, Q − S St0, exp

1 0.75 0 ∞ 0 0.11 ≈0.23

I 2 0.50 0 ∞ 0 0.11 ≈0.20


3 0.25 0 ∞ 0 0.11 ≈0.14
4 0.75 15 2.90 0.09 0.10 ≈0.17

II 5 0.50 15 1.93 0.09 0.09 ≈0.15


6 0.25 15 0.96 0.09 0.10 ≈0.11
7 0.75 30 1.50 0.18 0.19 ≈0.20

III 8 0.5 30 1.00 0.18 0.19 ≈0.20


9 0.25 30 0.50 0.18 0.18 ≈0.17
10 0.75 45 1.06 0.32 0.32 ≈0.34

IV 11 0.50 45 0.71 0.32 0.32 ≈0.31


12 0.25 45 0.35 0.32 0.32 ≈0.31
T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471 459

Of interest is the Strouhal number at which the foil is expected to produce no nett thrust (or drag), which we call the
“neutral force” Strouhal number, St0. As a first-order estimation, St0 can be calculated by assuming that the foil produces no
skin friction, and therefore St0 occurs when αmax = 0. In this instance, St0, a can be calculated using Eq. (6), i.e.
tanθ0
St0, a ≈
π (7)

where the subscript a denotes that the calculation is based on analytic derivations. The results are summarised in Table 1,
and includes St0 results obtained from the Quasi-Steady model (see Section 3), denoted by the subscript Q − S , and ex-
perimental measurements (see Section 4), denoted by the subscript exp. Following this, it is expected that the foil produces
nett thrust when St > St0 , and conversely, produces nett drag when St < St0 .

3. Quasi-Steady (Q-S) model of the unsteady hydrofoil

A Quasi-Steady (Q-S) model of the unsteady foil was developed using classical steady-state hydrodynamics to predict the
forces on the foil. The Q-S model is based on the argument that the forces on an unsteady foil at any given time are identical
to the forces developed by a static foil in a steady flow with the same geometric properties as the unsteady foil. In essence,
the Q-S model takes an instantaneous “snapshot” of the flow, and applies steady-state hydrodynamic theory to that flow,
ignoring most of the history and temporal variance of the flow. In the current analysis, three separate force-producing
mechanisms are considered, namely, thrust, drag, and “virtual mass”. These will be discussed separately below.

3.1. Thrust

The lift coefficient of a static foil in a uniform, non-accelerating potential flow is (Prandtl and Tietjens, 1934; Vallentine,
1967)
CL = klsinα (8)

where kl is some constant, and α is the angle of attack of the foil. Eq. (8) is usually invoked for low angles of attack, certainly
below the stall angle of the foil. However, for the current analysis, it is assumed that Eq. (8) holds for 0 ≤ α ≤ 2π rad. This
implies that the foil never “stalls”, which is consistent with classical potential flow assumptions.
Using the Q-S assumption, Eq. (8) should be valid at any given time of the foil oscillation cycle. Thus, using the definition
of force coefficient,
F
C=
1 2
ρU A
2 (9)

where F is the force, ρ is the density of the fluid, U is an appropriate reference velocity, and A = sc is the planform area, we
can convert Eq. (8) into the instantaneous lift force
1 2
FL(t ) = ρAkl Veff (t )sinα(t )
2 (10)
2
where Veff = U∞ + h′ 2(t ) is the effective velocity of the foil (relative to the surrounding fluid). We then normalise FL by the
free-stream dynamic pressure to obtain the non-dimensional, instantaneous lift force produced by the oscillating foil
⎛ ⎞
∼ FL(t ) ⎜ 2 ⎟
FL = = kl⎜ h˜′ (t ) + 1⎟sinα(t ).
1 2 ⎜ ⎟
ρU∞A
2 ⎝ ⎠ (11)

3.1.1. Determination of kl
The success of Eq. (11) hinges on an appropriate estimation of the value of kl. For real symmetrical wings in a steady flow,
Eq. (8) has been shown experimentally and computationally to be accurate up to the stall angle of the wing (Jacobs, 1932;
Loftin and Smith, 1949; Abbott and von Doenhoff, 1959; Sheldahl and Klimas, 1981). Typical values range from kl = 2.6
(Akbari and Price, 2003) to kl = 6.1 (Abbott and von Doenhoff, 1959), depending on flow conditions and foil geometry,
including the foil aspect ratio, Λ. For foils with a rectangular platform, the aspect ratio is defined as the span-to-chord ratio,
i.e., Λ = s/c . In general, kl is higher for larger Reynolds numbers and foil aspect ratios.
In the current experiments, 1500 ≤ Re ≤ 12500, Λ = 4 and the foil cross-section geometry is based on a NACA 0026 foil
(see Section 2). To obtain an accurate estimate of kl for the current analysis, calibration experiments were performed on the
NACA 0026 foil at fixed angles of attack, under steady-state conditions at Re ≈ 7500. It should be noted that although kl
460 T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471

Fig. 3. Calibration measurements to determine kl. Error bars denote 95% confidence interval.

shows some Reynolds number dependence, within the ranges used in the present study the variation in kl was relatively
small. Forces were measured using strain gauges, as discussed in Section 4. The results are shown in Fig. 3.
These measurements indicate that the foil used in the current study stalls at αstall ≈ 8°. From α = 0 − 6°, the lift coefficient
is approximately linear with respect to α. A linear curve fit of this results in kl ≈ 3.9. This value is similar to kl ≈ 3.8 obtained
for rectangular wings with Λ = 5 (Prandtl and Tietjens, 1934), kl ≈ 3.8 obtained for NACA 0025 foils with Λ = ∞ at
Re = 8 × 104 (Sheldahl and Klimas, 1981) and kl ≈ 4.0 obtained for NACA 0012 foils with Λ = ∞ at Re = 1 × 104 (Sheldahl and
Klimas, 1981). Therefore, a fixed value of kl = 3.9 is used in the current Q-S model.

3.2. Drag

We assume that the drag on the foil can be divided into two separate and independent components,
FD = FDf + FDi (12)

where FDf is the drag force due to the combination of pressure (form) drag and skin friction, while FDi is the induced drag of
the foil. The induced drag coefficient can be stated as

kiCL2
CDi =
πΛ (13)

where ki is a correction factor for non-elliptical wings (typically k i > 1), which was set to k i = 1.1 (Clancy, 1975). Therefore,
combining Eqs. (9) and (13), we obtain

2kiFL2(t )
FDi(t ) = .
πΛρA⎡⎣ U∞
2
+ h′ 2(t )⎤⎦ (14)

Normalising this with the dynamic pressure and simplifying, we obtain

∼ k ·k 2 ⎛ 2 ⎞
FDi(t ) = i l ⎜⎜ h˜′ (t ) + 1⎟⎟sin2α(t ).
πΛ ⎝ ⎠ (15)

To obtain FDf, we have to rely on empirical measurements as there is no known analytical equation that describes FDf as a
function of α(t ). Therefore, we performed steady-state force measurements on the foil at fixed angles of attack at Re ≈ 7500.
The induced drag was estimated from Eq. (13), and subtracted from the total drag measured by the strain-gauges. The
resultant drag, which is a combination of form drag and skin friction, is shown in Fig. 4. These results are plotted together
with the results of Sheldahl and Klimas (1981), who performed their measurements with a similar foil and Reynolds number
(NACA 0025 foil at Re ≈ 10000) as the current calibration experiment (NACA 0026 at Re ≈ 7500). The force measurements
beyond the static stall angle, αstall ≈ 8° from the current study are discarded, because the foil is assumed to never stall,
regardless of the angle of attack. Therefore, using the experimentally measured force coefficients for α ≤ 8°, together with
the results of Sheldahl and Klimas (1981), we obtain the following empirical relationship using an order-3 curve fit
T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471 461

Fig. 4. Drag coefficient versus foil angle of attack under steady-state conditions. The equation of the curve fit is shown in Eq. (16).

CDf (t ) ≈ − 1.158 α 3(t ) + 2.492α 2(t ) + 0.0518 α(t ) + 0.02 (16)

The instantaneous drag force on the foil due to pressure drag and skin friction is then
1
FDf (t ) = ρAVeff (t )CDf (t )
2 (17)

and the normalised drag force due to pressure drag and skin friction is

1 2 ⎛ ⎞
ρAVeff (t )CDf (t ) ⎜ 2 ⎟
∼ 2
FDf (t ) = = ⎜ h˜′ (t ) + 1⎟CDf (t ).
1 2 ⎜ ⎟
ρAU∞
2 ⎝ ⎠ (18)

3.3. Virtual mass

As the foil moves through the fluid, it accelerates and displaces surrounding fluid, which leads to an additional side force
component on the foil, FV. This component of force is estimated using the “virtual” or “added” mass method. Although the
effect of virtual mass is not strictly a steady phenomenon, it can be estimated using classical hydrodynamics, and is
therefore included in the quasi-steady model to improve the model's accuracy. Using the virtual mass method, the force
experienced by the foil due to fluid displacement is approximated by the force it takes to accelerate a volume of fluid
equivalent to a cylinder encompassing the projected chord of the foil (Newman, 1977), as shown in Fig. 5. The force due to
the virtual mass is then
FV (t ) = ρ · ∀virtual (t )·Veff
′ (t ) (19)
2
where ∀virtual (t ) = π ⎡⎣ csinα(t ) /2⎤⎦ s is the volume of the virtual cylinder. With some mathematical manipulation, the non-
dimensional force due to virtual mass can be shown to be

∼ ρ · ∀virtual (t )·Veff
′ (t ) π 4sin2α(t )St 3 sin4πft
FV (t ) = 1
= 1
.
2
ρU∞ A ∼
2
2
(
4h0 h˜′ (t ) + 1
2
) (20)

3.4. Time-averaged force coefficients and efficiency

Finally, the total instantaneous thrust and side force on the foil, predicted using the Q-S model, can be expressed as
∼ ∼ ∼ ∼
( )
FT (t ) = FL(t )·cosγ (t ) − FDf (t ) + FDi(t ) ·sinγ (t ) (21)
462 T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471

Fig. 5. “Virtual mass” of fluid entrained by the foil.

∼ ∼ ∼ ∼ ∼
( )
FS (t ) = FL(t )·sinγ (t ) + FDf (t ) + FDi(t ) ·cosγ (t ) + FV (t ) (22)

where γ(t ) is the instantaneous angle between the foil centreline and the lift vector, that is,
π
γ (t ) = − arctanh˜′(t ).
2 (23)

The time-averaged thrust coefficient (denoted using lower-case subscript t), is


4 T /4 ∼
Ct =
T
∫0 FT (t )dt
(24)

where T = 1/f is the oscillation period of the foil. Finally, the propulsive (Froude) efficiency is defined as
T ∼
U∞ ∫ 4 FT (t )dt
0
ηF = T ∼ T
∫ 4 F (t )· h′(t )dt ˜ (t )·θ ′(t )dt
+ c∫ 4 M
0 S θ
0 (25)
˜ (t ) = 2
where M θ Mθ (t ) /(1/2ρU∞ Ac )
is the instantaneous moment coefficient acting on the foil, while θ′(t ) is the instantaneous
angular velocity of the foil. For symmetrical foils in a free-stream at an angle of attack that is lower than the static stall angle,
the aerodynamic moment about the quarter-chord (0.25c) is typically negligible (Abbott and von Doenhoff, 1959; Eppler,
1990; Jumper et al., 1987; Sheldahl and Klimas, 1981). This implies that the aerodynamic centre is located at approximately
0.25c from the leading edge. As the centre of rotation of the foil is located 0.3c from the leading edge, the moment is
calculated by summing the contributions of the drag and lift forces acting on the aerodynamic centre, i.e.,
˜ (t ) = 0.05⎡⎣ F˜ (t ) + F˜ (t )⎤⎦sinα(t ) + 0.05F˜ (t )cosα(t ).
Mθ Df Di L (26)

Eqs. (24) and (25) provide a method to estimate two key unsteady foil dynamics, namely, the time-averaged thrust
coefficient and the propulsive efficiency, based on known foil parameters. It is interesting to note that the time-averaged
thrust coefficient is a function of St and θ0 only, while propulsive efficiency is also a function of the non-dimensional heave
amplitude, h̃0 (due to the virtual mass term). If the virtual mass term is ignored or assumed to be negligible, then both key
parameters Ct and ηF are expected to be functions of St and θ0 alone.
Eqs. (24) and (25) are solved numerically to obtain values of Ct and ηF for all examined foil parameters. These numerically
T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471 463

estimated values are then compared to experimentally measured values in Section 4.


By setting Ct = 0, Eq. (24) can also be solved for St to obtain the neutral force Strouhal number based on the quasi-steady
model, St0, Q − S . The values of St0, Q − S for all tested foil parameters are shown in Table 1. As is expected, St0, Q − S ≥ St0, a for all
groups. The difference between St0, a and St0, Q − S is generally very small for most groups except Group I (where heave is the
dominant motion). The reason for the higher discrepancy at higher heave amplitudes is due to the fact that even at small
Strouhal numbers, αmax can be significantly high when χ is large. This leads to high drag (both form and induced) even at
low St. Drag is not taken into account by the analytic value of St0.

3.5. Limitations of the Quasi-Steady model

As previously noted, the two main assumptions used in the development of the quasi-steady model are that the foil's
dynamics are independent of its time history (i.e., the flow is treated as a series of instantaneous, steady flows) and that the
foil does not experience the effects associated with conventional static stall. A survey of the literature suggests that the
former assumption is inaccurate, as even a small rate of change in the effective angle of attack has been shown to lead to
dynamic stall, which is an inherently unsteady phenomenon (Choudhry et al., 2014; Jumper et al., 1987; Leishman, 1990). In
studies of airfoils pitching at a constant rotational velocity in a free-stream, it has been shown that dynamic stall can occur
when the non-dimensional rotation rate, α′ * = cα′/(2U∞), is on the order of 10  2 or higher, where, α′ is the time-averaged
rate-of-change of α (Choudhry et al., 2014; Jumper et al., 1987). For the current foil, α′ * = Stαmax/h˜ 0 , and therefore unsteady
effects are expected to become important for Stαmax/h˜ 0 ≳ 10−2. For all combinations of foil kinematic parameters investigated
in this study (see Table 1), this expression is met when St ≳ St0 , implying that unsteady flow effects are significant for all
cases where the foil is expected to produce thrust. Additionally, studies of harmonically pitching wings have shown that
dynamic stall will occur when the reduced frequency, κ = πfc /U∞ ≳ 0.05 (Mulleners and Raffel, 2012; Rival and Tropea, 2010).
Using this data, dynamic stall is expected for St ≳ 0.1h˜ 0 /π , i.e., for all the cases examined in the current study.
At first glance, it would appear that this will severely limit the applicability of the Q-S model. However, one effect of
dynamic stall is to delay the onset of flow separation and the sudden loss in lift that characterises static (or steady) stall to
much larger angles of attack (Choudhry et al., 2014; Jumper et al., 1987; Mulleners and Raffel, 2012; Rival and Tropea, 2010).
These larger dynamic stall angles, αdy, stall , are typically in the range 25° ≲ αdy, stall ≲ 30° (Jumper et al., 1987; Mulleners and
Raffel, 2012). Below the dynamic stall angle, the lift-curve is approximately linear with a similar gradient to that found in the
steady case below the static stall angle. As the quasi-steady model does not simulate steady state stall conditions, dynamic
stall effectively expands the applicability of the quasi-steady model to predict mean thrust to much larger instantaneous
angles of attack.
During dynamic stall, significant departure from the linear lift-curve and the loss in lift typically occurs if the foil is
maintained at an angle of attack α > αdy, stall for an extended period of time (Leishman, 1990; McCroskey, 1982). Although this
exact time-scale has not yet been conclusively quantified, in general the time lag between the instant α increases beyond
the dynamic stall angle and the sudden loss in lift increases as the reduced frequency increases (Choudhry et al., 2014; Rival
and Tropea, 2010). As discussed above, in the current experiments the reduced frequencies and Strouhal numbers are large
relative to the limit required for “unsteady flow”, and thus the foil is not expected to experience a sudden loss of lift if its
angle of attack intermittently exceeds αdy, stall . Therefore it can be assumed that the quasi-steady model will estimate mean
thrust inaccurately only when the average angle of attack, αave, and not the maximum angle of attack, αmax, exceeds the
dynamic stall angle. Thus, we expect the Q-S model to predict the time-averaged thrust of the unsteady foil where
αave ≲ αdy, stall ≈ 30°, that is, St ≲ 0.31, 0.46, 0.69 and 1.06 for θ0 = 0°, 15°, 30° and 45°, respectively. Beyond these conditions,
any departure between the modelled and measured mean thrust can be attributed to unsteady flow effects other than
delayed stall and aerodynamic lag discussed above. These unsteady effects may include (but are not limited to) wake capture
(Birch and Dickinson, 2003; Dickinson, 1996; Sane and Dickinson, 2001; Sane, 2003), post-dynamic-stall lift recovery
(Choudhry et al., 2014) and vorticity control (Gopalkrishnan et al., 1994; Zhu et al., 2002).
Dynamic stall also produces large magnitude, highly transient drag forces and aerodynamic moments on the foil
(Choudhry et al., 2014; Rival and Tropea, 2010) which are challenging to model. Accurate modelling of the full effects of
dynamic stall is expected to require time-resolved simulations of flow separation, particularly on the suction side of the foil,
which can only be achieved using more sophisticated models (e.g. large-eddy simulations) or through direct numerical
simulations. The moments and drag forces induced by dynamic stall are not included in the current quasi-steady model, and
hence the model is expected to over-estimate propulsive efficiencies for all cases.

4. Force measurements

Figs. 6–9 show the experimentally measured time-averaged thrust coefficients as well as propulsive efficiencies, for
Groups I–IV.
For all groups, the thrust coefficient appears to increase quadratically as a function of St. This is consistent with ob-
servations made of similar unsteady foils (Anderson et al., 1998; Read et al., 2003; Schouveiler et al., 2005; Triantafyllou
et al., 1993). In the current experiments, for all examined flow cases, there was no observed drop-off in the thrust coefficient,
even at large Strouhal numbers (where αmax is typically large; see Eq. (6)). Therefore, in theory, this allows the unsteady foil
464 T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471

Fig. 6. Force and propulsive efficiency measurements for Group I (θ0 = 0°) , Re = 2300–9500 .

to develop very large time-averaged thrust coefficients by increasing St.


The experimentally derived value of St0, exp is obtained by curve fitting a quadratic equation to the experimental data (not
shown here), and solving for Ct = 0 for each foil heave/pitch combination. The results are shown in Table 1. A comparison of
St0, exp and St0, Q − S indicates that for most cases, St0, Q − S is a good predictor of St0.
Figs. 6–9 also include data of an unsteady NACA 0012 foil of aspect ratio Λ = 6 at Re ≈ 40000, measured experimentally
by Anderson et al. (1998) and Read et al. (2003), as well as data for oscillating two-dimensional foils calculated numerically
by Guglielmini and Blondeaux (2004) and Platzer et al. (2008) at Re ¼1100 and Re¼20000, respectively. All three experi-
mental results for the thrust coefficient (including the current study) show strong agreement for all cases. Interestingly, the
numerically calculated thrust coefficients of Guglielmini and Blondeaux (2004) and Platzer et al. (2008) closely match the
quasi-steady model, which suggests that within the range of conditions within the current study, the current quasi-steady
model predicts approximately the same values of mean thrust as the more sophisticated models of Guglielmini and
Blondeaux (2004) and Platzer et al. (2008).
In general, it appears that the experimentally measured thrust coefficients show excellent agreement with the Q-S
model, except in cases where the heave amplitude and the Strouhal number are large (Group I). On close inspection, there is
a departure between the experimentally measured and modelled mean thrust coefficients for St ≳ 0.3, St ≳ 0.5 and St ≳ 0.6
for Groups I, II and III, respectively, with the departure the strongest for Group I. This is because at these large Strouhal
numbers, the average angles of attack start to exceed the dynamic stall angle, as discussed in Section 3.5. As the quasi-steady
model takes into account the unsteady effects of delayed stall and aerodynamic lag, the strong similarity between the
T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471 465

Fig. 7. Force and propulsive efficiency measurements for Group II (θ0 = 15°) , Re = 2300–12500 .

modelled and experimentally measured thrust coefficients below these Strouhal numbers also implies that these two un-
steady effects are employed in the current oscillating foil for all heave-pitch combinations.
For Group I (θ0 = 0°), at large Strouhal numbers (St ≳ 0.3), the experimentally measured thrust coefficients exceed the
predictions of the quasi-steady model. As the pitch amplitude for Group I is zero, the discrepancy between the Q-S model
and the experimental results cannot be attributed to the exclusion of rotational effects from the Q-S model. Furthermore, for
Group II (θ0 = 15°), the data also shows that the experimentally measured values of Ct are higher than that predicted from
the Q-S model, although the differences are smaller than Group I. Together, these results suggest that the oscillating foil,
under the conditions of high heave to pitch amplitudes such that the average angles of attack are larger than the dynamic
stall angle, αdy, stall , utilises unsteady flow mechanisms that are in addition to delayed stall and aerodynamic lag to augment
thrust production. The nature of these unsteady flow mechanisms cannot be deduced from the current data alone, however,
the large leading edge vortices developed under these conditions (not shown here, but available in Lau, 2010) and the large
angles of attack, in excess of the dynamic stall angle, are consistent with the unsteady flow mechanisms of wake–wing
interaction (or wake capture) and dynamic stall vortex control (Anderson et al., 1998; Birch and Dickinson, 2003; Birch et al.,
2004; Choudhry et al., 2014; Dickinson, 1996; Sane, 2003). The combined effects of these unsteady flow mechanisms result
in very large thrust coefficients, with the maximum thrust coefficient recorded, Ct = 10.79 at St¼0.71 for Group I. Under
these conditions, the maximum angle of attack, αmax is large, approaching 70°, which is significantly higher than both the
static stall (αstall ≈ 8°) and the dynamic stall (α ≈ 30°) angles.
Measured propulsive efficiencies are generally moderate to poor, particularly when the heave amplitude is large,
466 T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471

Fig. 8. Force and propulsive efficiency measurements for Group III (θ0 = 30°) , Re = 3100–12500 .

consistent previous measurements (Young and Lai, 2007). The propulsive efficiencies for Group I, where thrust coefficients
are the largest, are low, ηF ≈ 10–30% . At St¼0.71, where the highest thrust coefficient was recorded, the efficiency is only
η = 18.7% . Groups II and III generally show the highest efficiencies, an order of ηF ≈ 20–40% . The highest recorded efficiency

was ηF = 43.2% , occurring at St¼ 0.2, h0 = 0.25, and θ0 = 30°. Under these conditions, the thrust coefficient is very small,
Ct ≈ 0.20, and interestingly, the maximum angle of attack is small, αmax = 2.7°, well below the static stall angle of the foil.
This suggests that peak efficiencies occur when the foil does not employ any unsteady effects, as unsteady effects are only
expected when the maximum angle of attack exceeds the static stall angle, i.e., αmax ≳ 8°. When the pitch amplitude is large,
the results also show that the efficiencies are low, ηF ≈ 10–30% . In all cases, propulsive efficiencies show a peak in the range
of St ≈ 0.2–0.4 , where αmax and Ct are generally small. This is an interesting result, for two reasons. Firstly, it implies that
large thrust forces typically cannot be developed at high propulsive efficiencies, and secondly, it suggests that steadily
swimming carangiform fish may select to swim at this Strouhal number range to maximise propulsive efficiency.
A comparison between the propulsive efficiencies experimentally measured by Anderson et al. (1998) and Read et al.
(2003) and the current study reveal that ηF varies considerably between the three studies. In general, the propulsive effi-
ciencies of Anderson et al. (1998) and Read et al. (2003) are typically higher than the current study. This is possibly due to
the differences in the Reynolds number and/or the foil geometry. In particular, the experiments of Anderson et al. (1998) and
Read et al. (2003) utilised a foil with end plates and an aspect ratio of Λ = 6, resulting in a flow more closely resembling a
two-dimensional flow than the current experiments, which utilised a foil with no end plates and an aspect ratio of Λ = 4.0.
However, this does not explain the significant difference in the measured propulsive efficiency between Anderson et al.'s
(1998) and Read et al.'s (2003) experiments, which were performed using the same foil geometry and Reynolds number. The
T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471 467

Fig. 9. Force and propulsive efficiency measurements for Group IV (θ0 = 45°) , Re = 1500–12500 .

authors do note that in the range of St ≈ 0.2–0.4 where Ct is small, the values of ηF are very susceptible to errors and noise
due to the small forces and moments that are involved. This may have contributed to the significant differences in the
measured values of ηF across all three studies.
In almost all cases, the Q-S model over-estimates the propulsive efficiencies developed by the foil. This is because the
quasi-steady model does not accurately model the drag and aerodynamic moments caused by unsteady effects such as
dynamic stall, as previously discussed in Section 3.5. Furthermore, the Q-S model does not accurately simulate boundary
layer effects, such as skin friction and/or boundary layer thinning/thickening, which will increase the effective drag on the
foil. The discrepancy between the modelled and experimental results increases as the pitch amplitude increases, which
suggests that the primary source of error in the Q-S model is the inaccurate modelling of aerodynamic moments because it
is the only term in the propulsive efficiency equation that is dependent on the pitch amplitude (see Eq. (25)).
The numerically calculated propulsive efficiencies of Platzer et al. (2008) for Group I do not match any of the experi-
mental data (including data from Anderson et al., 1998 and Read et al., 2003) or the data from the Quasi-Steady model
(Fig. 6). This highlights the difficulty in fully modelling the forces and moments on the oscillating foil due to unsteady
effects, which are expected to be particularly significant for Group I where the instantaneous and mean angles of attack can
be very large. On the other hand, for Group III, the differences between the numerically calculated propulsive efficiencies of
Guglielmini and Blondeaux (2004) and the current experimental results are relatively small (Fig. 8). Furthermore, the data of
Guglielmini and Blondeaux more closely resembles the current experimental results than the prediction of the Q-S model,
which is not surprising considering the significant limitations of the Q-S model to predict propulsive efficiencies.
468 T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471

Fig. 10. Consolidated plot of all force measurements, Re ¼ 1500–12500.

Based on the force measurement results, two major observations can be made. Firstly, when the heave amplitude (re-
lative to the pitch amplitude) and the Strouhal number are simultaneously large (resulting in large αmax), the time-averaged
thrust coefficients developed by oscillating foil are very large, far exceeding the theoretical Q-S model predictions. It is
deduced that these large thrust coefficients develop due to a combination of unsteady effects, including wing–wake in-
teractions, leading edge vortex control, delayed stall and aerodynamic lag. The propulsive efficiencies under these condi-
tions, however, are low. Secondly, in cases where the χ is moderate, leading to a flow that is neither dominated by heaving
nor pitching motions, the unsteady foil generates thrust at peak propulsive efficiency in the range St ≈ 0.2–0.3, where the
angles of attack are generally small, typically below the static stall angle (αmax ≲ 8°). The thrust developed under these
conditions is small (Ct ≈ 0.1–0.5).

4.1. Consolidation of force measurements

A summary of the thrust measurements for all flow conditions, plotted against the Strouhal number, is shown in Fig. 10.
Again, it is evident that Group I generates the highest thrust coefficients, while Group IV produces the lowest. It is also clear
that in most flow conditions, the quasi-steady model performs admirably, except when the heave amplitude is large relative
to the pitch amplitude (Group I).
Of particular interest is that the results show excellent collapse within each group, implying that the thrust coefficient is
strongly influenced by St and θ0, as predicted from the Q-S model. In an effort to combine all the measured thrust coef-
ficients on a single curve, the physics of the foil are re-examined to determine the dominant parameters that relate to Ct.
Revisiting Eqs. (21) and (24), and assuming the drag on the foil is negligible, the time-averaged thrust coefficient should be
of the form
T
4kl
Ct ≈
T
∫0 4 ( h˜′ (t) + 1) ·h˜′·sinα(t)dt.
2
(27)

Recalling Eqs. (2) and (5), with some trigonometric manipulation, this reduces to

Ct ≈
(
4klπ 2St St − St0, a ) I(t )
T (28)

where
T
cos2ϕ(t )
I (t ) = ∫0 4 dt .
( tan θ cos ϕ(t ) + 1)
2
0
2
(29)

The integral term I(t) cannot be determined analytically, however, it can be reduced to a function of Legendre's elliptical
equations (of the first and second kind), and can be evaluated numerically. Numerical calculations and curve fitting indicate
that the integral term can be approximated by
T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471 469

Fig. 11. Thrust coefficients versus the non-dimensional term St ( St − St0) . The dashed black line is a linear curve fit to the current experimental data of
Groups II–IV only.

⎛ 1⎞
I (t ) ≈ T ⎜ −0.0479θ02 + ⎟
⎝ 8⎠ (30)

which reduces to T /8 at low pitch amplitudes. Thus, we suggest that

Ct ∝ St ( St − St0). (31)

Plotting the experimental data against the non-dimensional term St ( St − St0), we obtain Fig. 11. The results are pro-
mising, with most data collapsing on a single, straight line, with the exception of Group I, where the time-averaged thrust
coefficients are larger than expected. Therefore, for most combinations of heave and pitch amplitudes that were tested, the
time-averaged thrust coefficients developed by the foil can be universally scaled to the non-dimensional term St ( St − St0).
Furthermore, as the scaling law presented in Eq. (31) was developed based on the quasi-steady model which does not take
into consideration the history of the flow, any departure from this relationship is indicative that unsteady flow effects in
addition to delayed stall and aerodynamic lag, such as wake–wing interactions or leading edge vortex control, are sig-
nificant. For the present experiments, this departure occurs for Group I (where the pitch amplitude is zero), providing
further evidence that the foil undergoing large heave amplitudes (relative to the pitch amplitude) utilise a range of unsteady
flow mechanisms to augment thrust development. A linear curve-fit of the data from Groups II–IV only yields

Ct ≈ 11.8 St ( St − St0) − 0.06. (32)

When St = St0 , Ct = − 0.06 ≈ 0, as expected.

5. Concluding remarks

An extensive investigation into the dynamics of a hydrofoil undergoing simultaneous pitching and heaving motions
utilising a combination of a quasi-steady model and experimental force measurements has revealed that:

 for all examined flow cases, the time-averaged thrust coefficient, Ct, produced by the foil varies approximately linearly
with St2, and there is no drop-off in Ct at large St where αmax can be much larger than both the static and dynamic stall
angle of the foil. The foil is shown to develop large thrust coefficients, particularly when the heave amplitude is large
relative to the pitch amplitude (Group I), and Strouhal number is simultaneously large, leading to very large angles of
470 T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471

attack (αmax > 40°) which are typically in excess of the estimated dynamic stall angle. It is deduced that these large thrust
coefficients develop due to a combination of unsteady effects, including wing–wake interactions, leading edge vortex
control, delayed stall as well as the aerodynamic time lag between the angle of attack exceeding the dynamic stall angle

and the sudden loss in lift. The largest measured thrust coefficient was Ct ≈ 10.79, which occurred at St¼0.71, h0 = 0.5 and
θ0 = 0° (leading to αmax = 66°). However, the propulsive efficiencies under these conditions are low (ηF ≈ 19%);
 in general, the propulsive efficiencies are moderate to low, typically ranging from ηF = 10–40% , consistent with previous
measurements (Young and Lai, 2007). The highest efficiencies typically occur when both the heave amplitude and the
pitch amplitude are moderate (Groups II and III), and in the range of St = 0.2–0.4 , where αmax is generally lower than the
static stall angle (αmax ≲ 8°), implying that peak propulsive efficiencies develop when unsteady flow mechanisms are not

employed or not significant. The highest efficiency recorded was ηF = 43.2% , occurring at St¼0.2, h0 = 0.25, and θ0 = 30°
(resulting in αmax = 2.7°);
 for most of the flow parameters examined, the Q-S model satisfactorily predicts the time-averaged thrust coefficients
produced by the foil. As the Q-S model includes the unsteady effects of delayed stall and aerodynamic lag, this suggests
that the unsteady foil typically employs these two unsteady mechanisms for most flow conditions;
 the Q-S model does not accurately predict propulsive efficiencies, with the discrepancy between experimentally mea-
sured and numerically calculated results increasing as the pitch amplitude increases. This is primarily attributed to the
inaccurate modelling of the drag force and aerodynamic moments caused by unsteady flow effects;
 the thrust developed by the foil can be characterised by the non-dimensional term St ( St − St0). This is supported by
experimental evidence, which shows that Ct is linearly proportional to St ( St − St0) for all tested cases, except for the case
where the heave amplitude is much larger than the pitch amplitude, i.e., when unsteady flow mechanisms other than
delayed stall and aerodynamic lag are expected to be significant. Importantly, this implies that the departure of the linear
relationship between Ct and the term St ( St − St0) can be used as a universal indicator for the transition toward a flow
regime where these additional unsteady effects become significant.

Acknowledgements

The authors would like to acknowledge Dr. Peter Lanspeary and Mr. Eyad Hassan for their useful discussions and
technical input.

References

Abbott, I.H., von Doenhoff, A.E., 1959. Theory of Wing Sections. Dover Publications, New York.
Akbari, M.H., Price, S.J., 2003. Simulation of dynamic stall for a NACA 0012 airfoil using a vortex method. J Fluids Struct 17, 855–874.
Anderson, J.M., Streitlien, K., Barrett, D.S., Triantafyllou, M.S., 1998. Oscillating foils of high propulsive efficiency. Journal of Fluid Mechanics 360, 41–72.
Birch, J.M., Dickinson, M.H., 2003. The influence of wing-wake interactions on the production of aerodynamic forces in flapping flight. Journal of Ex-
perimental Biology 206, 2257–2272.
Birch, J.M., Dickson, W.B., Dickinson, M.H., 2004. Force production and flow structure of the leading edge vortex on flapping wings at high and low Reynolds
numbers. Journal of Experimental Biology 207, 1063–1072.
Blake, R.W., 1983. Fish Locomotion. Cambridge University Press, Cambridge.
Blondeaux, P., Fornarelli, F., Guglielmini, L., Triantafyllou, M., Verzicco, R., 2005. Numerical experiments on flapping foils mimicking fish-like motion.
Physics of Fluids 17 (113601).
Borazjani, I., Sotiropoulos, F., 2008. Numerical investigation of the hydrodynamics of carangiform swimming in the transitional and inertial flow regimes.
Journal of Experimental Biology 211, 1541–1558.
Borazjani, I., Sotiropoulos, F., 2010. On the role of form and kinematics on the hydrodynamics of self-propelled body/caudal fin swimming. Journal of
Experimental Biology 213, 89–107.
Bose, N., Lien, J., 1989. Propulsion of a fin whale (Balaenoptera physalus): why the fin whale is a fast swimmer. Proc R Soc Lond B 237, 175–200.
Breder, C.M., 1926. The locomotion of fishes. Zoologica 4, 159–256.
Chopra, M.G., 1976. Large amplitude lunate-tail theory of fish locomotion. Journal of Fluid Mechanics 74, 161–182.
Choudhry, A., Leknys, R., Arjomandi, M., Kelso, R., 2014. An insight into the dynamic stall lift characteristics. Exp Thermal Fluid Sci 58, 188–208.
Clancy, L.J., 1975. Aerodynamics. John Wiley and Sons, New York.
Dewey, P.A., Boschitsch, B.M., Moored, K.W., Stone, H.A., Smits, A.J., 2013. Scaling laws for the thrust production of flexible pitching panels. Journal of Fluid
Mechanics 732, 29–46.
Dickinson, M.H., 1996. Unsteady mechanisms of force generation in aquatic and aerial locomotion. Am Zool 36, 537–554.
Dong, H., Mittal, R., Najjar, F.M., 2006. Wake topology and hydrodynamic performance of low-aspect-ratio flapping foils. Journal of Fluid Mechanics 566,
309–343.
Eloy, C., 2012. Optimal Strouhal number for swimming animals. J Fluids Struct 30, 205–218.
Eppler, R., 1990. Airfoil Design and Data. Springer-Verlag, Berlin.
Fenercioglu, I., Cetiner, O., 2012. Categorization of flow structures around a pitching and plunging airfoil. J Fluids Struct 31, 92–102.
Fish, F.E., 1993. Power output and propulsive efficiency of swimming bottlenose dolphins (Tursiops truncatus). Journal of Experimental Biology 185, 179–193
.
Fish, F.E., 1998. Comparative kinematics and hydrodynamics of odontocete cetaceans: morphological and ecological correlates with swimming perfor-
mance. Journal of Experimental Biology 201, 2867–2877.
Fish, F.E., Beneski, J.T., Ketten, D., 2007. Examination of the three-dimensional geometry of cetacean flukes using computed tomography scans: hydro-
dynamic implications. Anatomical Record 290, 614–623.
Fish, F.E., Rohr, J.J., 1999. Review of dolphin hydrodynamics and swimming performance. Tech. rep., San Diego, August.
Fontanella, J.E., Fish, F.E., Rybczynski, N., Nweeia, M.T., Ketten, D.R., 2011. Three-dimensional geometry of the narwhal (Monodon monoceros) flukes in
relation to hydrodynamics. Marine Mammal Science 27, 889–898.
T.C.W. Lau, R.M. Kelso / Journal of Fluids and Structures 65 (2016) 455–471 471

Gopalkrishnan, R., Triantafyllou, M.S., Triantafyllou, G.S., Barrett, D., 1994. Active vorticity control in a shear flow using a flapping foil. Journal of Fluid
Mechanics 274, 1–21.
Guglielmini, L., Blondeaux, P., 2004. Propulsive efficiency of oscillating foils. Eur J Mech B 23, 255–278.
Hover, F.S., Haugsdal, O., Triantafyllou, M.S., 2004. Effect of angle of attack profiles in flapping foil propulsion. J Fluids Struct 19, 37–47.
Jacobs, E.N., 1932. Airfoil section characteristics as affected by protuberances. NACA report no. 446, National Advisory Committee for Aeronautics, Virginia,
July.
Jumper, E.J., Schreck, S.J., Dimmick, R.L., 1987. Lift-curve characteristics for an airfoil pitching at constant rate. Journal of Aircraft 24, 680–687.
Lau, T.C.W., 2010. The flow around a fish-inspired heaving and pitching hydrofoil (Ph.D. thesis), School of Mechanical Engineering, The University of
Adelaide.
Lauder, G.V., Tytell, E.D., 2006. Hydrodynamics of undulatory propulsion. Abbreviation Main Titles 23, 425–468.
Leishman, J.G., 1990. Dynamic stall experiments on the NACA 23012 aerofoil. Exp Fluids 9, 49–58.
Liao, J.C., Beal, D.N., Lauder, G.V., Triantafyllou, M.S., 2003. The Karman gait: novel body kinematics of rainbow trout swimming in a vortex street. Journal of
Experimental Biology 206, 1059–1073.
Lighthill, J., 1975. Mathematical Biofluid Dynamics. Society for Industrial and Applied Mahematics, Pennsylvania.
Lighthill, M.J., 1969. Hydromechanics of aquatic animal propulsion. Annual Review of Fluid Mechanics, 413–446.
Loftin Jr., L.K., Smith, H.A., 1949. Aerodynamic characteristics of 15 NACA airfoil sections at seven Reynolds numbers from 0.7 × 106 to 9.0 × 106 . NACA
technical report 1945, Washington.
McCroskey, W.J., 1982. Unsteady airfoils. Annu Rev Fluid Mech 14, 285–311.
Moore, M.N.J., 2014. Analytical results on the role of flexibility in flapping propulsion. Journal of Fluid Mechanics 757, 599–612.
Mulleners, K., Raffel, M., 2012. The onset of dynamic stall revisited. Exp Fluids 52, 779–793.
Newman, N., 1977. Marine Hydrodynamics. MIT Press, Cambridge.
Paraz, F., Schouveiler, L., Eloy, C., 2016. Thrust generation by a heaving flexible foil; resonance, nonlinearities, and optimality. Physics of Fluids 28 (011903).
Parker, K., von Ellenreider, K.D., Soria, J., 2007. Morphology of the forced oscillatory flow past a finite-span wing at low Reynolds number. Journal of Fluid
Mechanics 571, 327–357.
Platzer, M.F., Jones, K.D., Young, J., Lai, J.C.S., 2008. Flapping-wing aerodynamics: progress and challenges. Aiaa Journal 46, 2136–2149.
Prandtl, L., Tietjens, O.G., 1934. Applied hydro- and aeromechanics. Dover Publications Inc., New York.
Read, D.A., Hover, F.S., Triantafyllou, M.S., 2003. Forces on oscillating foils for propulsion and maneuvering. J Fluids Struct 17, 163–183.
Rival, D., Tropea, C., 2010. Characteristics of pitching and plunging airfoils under dynamic-stall conditions. Journal of Aircraft 47, 80–86.
Rohr, J.J., Fish, F.E., 2004. Strouhal number and optimization of swimming by odontocete cetaceans. Journal of Experimental Biology 207, 1633–1642.
Sane, S.P., 2003. The aerodynamics of insect flight. Journal of Experimental Biology 206, 4191–4208.
Sane, S.P., Dickinson, M.H., 2001. The control of flight force by a flapping wing: lift and drag production. Journal of Experimental Biology 204, 2607–2626.
Schouveiler, L., Hover, F.S., Triantafyllou, M.S., 2005. Performance of flapping foil propulsion. J Fluids Struct 20, 949–959.
Sfakiotakis, M., Lane, D.M., Davies, J.B.C., 1999. Review of fish swimming modes for aquatic locomotion. IEEE J Ocean Eng 24, 237–252.
Sheldahl, R.E., Klimas, P.C., 1981. Aerodynamic characteristics of seven symmetrical airfoil sections through 180-degree angle of attack for use in aero-
dynamic analysis of vertical axis wind turbines. Sand80-2114 energy report, Sandia National Laboratories, New Mexico.
Taylor, G.K., Nudds, R.L., Thomas, A.L.R., 2003. Flying and swimming animals cruise at a Strouhal number tuned for high power efficiency. Nature 425,
707–711.
Triantafyllou, G.S., Triantafyllou, M.S., Grosenbaugh, M.A., 1993. Optimal thrust development in oscillating foils with application to fish propulsion. J Fluids
Struct 7, 205–224.
Triantafyllou, M.S., Triantafyllou, G.S., 1995. An efficient swimming machine. Scientific American, 40–48.
Triantafyllou, M.S., Triantafyllou, G.S., Yue, D.K.P., 2000. Hydrodynamics of fishlike swimming. Annual Review of Fluid Mechanics 32, 33–53.
Vallentine, H.R., 1967. Applied Hydrodynamics. Butterworths, London.
Videler, J.J., 1993. Fish Swimming. Chapman and Hall, London.
Vogel, S., 1996. Life in Moving Fluids, 2nd Edition Princeton University Press, Princeton.
von Ellenrieder, K.D., Parker, K., Soria, J., 2003. Flow structures behind a heaving and pitching finite-span wing. Journal of Fluid Mechanics 490, 129–138.
Webb, P.W., 1984. Form and function in fish swimming. Scientific American 251 (1), 72–82.
Webb, P.W., 1988. “Steady” swimming kinematics of tiger musky, an esociform accelerator, and rainbow trout, a generalist cruiser. Journal of Experimental
Biology 138, 51–69.
Yates, G.T., 1983. Hydromechanics of body and caudal fin propulsion. In: Webb, P.W., Weihs, D. (Eds.), Fish Biomechanics, Praeger Publishers, New York,
pp. 177–213. (chapter 6).
Young, J., Lai, J.C.S., 2004. Oscillation frequency and amplitude effects on the wake of a plunging airfoil. Aiaa Journal 42 (10), 2042–2052.
Young, J., Lai, J.C.S., 2007. Mechanisms influencing the efficiency of oscillating airfoil propulsion. Aiaa Journal 45, 1695–1702.
Zhu, Q., Wolfgang, M.J., Yue, D.K.P., Triantafyllou, M.S., 2002. Three-dimensional flow structures and vorticity control in fish-like swimming. Journal of Fluid
Mechanics 468, 1–28.

You might also like