You are on page 1of 15

Numerical investigation of the unsteady

aerodynamics of NACA 0012 with suction


surface protrusion
Aslesha Bodavula
Department of Aerospace Engineering, University of Petroleum and Energy Studies, Dehradun, India
Rajesh Yadav
Department of Aerospace Engineering, University of Petroleum and Energy Studies, Dehradun, India, and
Ugur Guven
Department of Aerospace Engineering, University of Petroleum and Energy Studies, Dehradun, India

Abstract
Purpose – The purpose of this paper is to investigate the effect of surface protrusions on the flow unsteadiness of NACA 0012 at a Reynolds number
of 100,000.
Design/methodology/approach – Effect of protrusions is investigated through numerical simulation of two-dimensional Navier–Stokes equations
using a finite volume solver. Turbulent stresses are resolved through the transition Shear stress transport (four-equation) turbulence model.
Findings – The small protrusion located at 0.05c and 0.1c significantly improve the lift coefficient by up to 36% in the post-stall regime. It also
alleviates the leading edge stall. The larger protrusions increase the drag significantly along with significant degradation of lift characteristics in the
pre-stall regime as well. The smaller protrusions also increase the frequency of the vortex shedding.
Originality/value – The effect of macroscopic protrusions or deposits in rarely investigated. The delay in stall shown by smaller protrusions can be
beneficial to micro aerial vehicles. The smaller protrusions increase the frequency of the vortex shedding, and hence, can be used as a tool to
enhance energy production for energy harvesters based on vortex-induced vibrations and oscillating wing philosophy.
Keywords Wind energy, MAVs, Low Reynolds number, NACA 0012, Vortex shedding
Paper type Research paper

Nomenclature r = Density, kg/m3;


m = Molecular viscosity, kg/m-s;
c = Chord length, m; mt = Turbulent viscosity;
cd = Drag coefficient; t ij = Viscous stress tensor; and
cl = Lift coefficient; v = Specific dissipation rate, s1.
cl, max = Maximum value of lift coefficient;
cp = Pressure coefficient;
1. Introduction
h = Height of protrusion normal to surface, m;
k = Turbulent kinetic energy, m2/s2; The generation of renewable energy to avoid global warming
L/D = Lift to drag ratio; because of fossil fuel-based power plants is the need of the
p = Pressure, Pa; hour. For small wind turbines, which are captive power
Reu t = Momentum-thickness Reynolds number; generators, the range of Reynolds number is between 105 and
g
Re ut = Transition thickness Reynolds number; 107. Apart from conventional wind turbines, renewable energy
t = Time, second; from wind can be extracted through windmills (McKinney and
u = Cartesian components of velocity vector; DeLaurier, 1981; Davids et al., 1999; Kinsey and Dumas,
! 2008; Xiao et al., 2012; Xiao and Zhu, 2014), kites
V = Velocity vector;
x = Spatial coordinate;
y1 = Non-dimensional cell wall distance; All the simulations carried out for this investigation were carried out on
a = Angle of attack (AOA), degrees (°); computational fluid dynamics Servers with 8 cores Xeon processors and 32
g = Intermittency; giga bytes of memory provided by the University of Petroleum and Energy
Studies. We would like to thank the university management for providing
us the necessary support required for the simulations. The authors would
The current issue and full text archive of this journal is available on also like to thank the Chancellor and the Vice-Chancellor of the university
Emerald Insight at: https://www.emerald.com/insight/1748-8842.htm to have financially supported this research. We hereby declare that there is
no conflict of interest.

Aircraft Engineering and Aerospace Technology


Received 28 January 2019
92/2 (2020) 186–200 Revised 3 June 2019
© Emerald Publishing Limited [ISSN 1748-8842] 7 August 2019
[DOI 10.1108/AEAT-01-2019-0022] Accepted 16 October 2019

186
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

(Cherubini et al., 2015; Perkovic et al., 2013; Argatov et al., reattach on the surface of the airfoil (Yarusevych et al., 2006).
2009), and mechanisms based on vortex-induced vibrations This clearly indicates the deterioration of the aerodynamic
(VIVs) (Wang and Chang, 2010; Wang et al., 2012; Rostami performance of airfoils at low Reynolds numbers. Even for the
and Armandei, 2017; He and Gao, 2013; Peng and Zhu, 2009; Clark–Y airfoil, which is designed for good low Reynolds
Kumar et al., 2017; Dunnmon et al., 2011; Barrero-Gil et al., number performance, the aerodynamic efficiency deteriorates
2012). All these energy extraction mechanisms use some form dramatically as the Reynolds number is reduced to 75,000
of wing, flat plate or cylinder motion because of aerodynamic (Marchman and Werme, 1984). This degradation in
forces at low Reynolds number. Another area of industrial performance is again, caused by the failure of the shear layer to
importance, where the low Reynolds number flows find a reattach. A detailed numerical study to understand the
massive application is design of micro aerial vehicles (MAVs) aerodynamics of various airfoils at low Reynolds numbers was
and unmanned aerial vehicles (UAVs). These vehicles are done by Winslow et al. (2017). Besides reestablishing a
required to fly in dusty and hazardous places for rescue, nonlinear lift curve behavior at low Reynolds number for
inspection and delivery of payloads. The MAVs fly at low NACA 0012, they suggested a reduction of about 46 per cent in
altitudes and their sizes vary in the range of 0.1-0.5 m. The the maximum lift coefficient for Reynolds number between 105
range of the Reynolds number for these micro and small UAVs and 104 because of the inability of the prematurely separated
are 104-106 (Lissaman, 1983). Aerodynamics of MAVs, UAVs flow to reattach. Winslow et al. (2017) also suggested that an
and the energy harvesters at low Reynolds number, operating airfoil with sharper leading edge provides better reattachment
especially in the range of 104-106 is very interesting as it of the flow, and hence, a better low Reynolds number
involves complex unsteady flow phenomena (Lissaman, 1983; performance. An increase in camber also results in an increase
Carmichael, 1982; Lin and Pauley, 1996). This has attracted in lift-to-drag ratio at lower Reynolds numbers. A recent study
the interests of many researchers into examining the low on very low Reynolds number aerodynamics of NACA 0012
Reynolds number flow phenomena over airfoil profiles of airfoil also reported a highly nonlinear lift curve with abrupt
NACA, Eppler, Selig and Clark-Y series. One of the peculiar changes in lift coefficients with changes in AOA for Reynolds
behavior of airfoil at low Reynolds number is vortex shedding, number in the range between 2,500 and 5,000 (Kim et al.,
which is caused by the dominant inviscid instability wave 2011).
induced by the inflection velocity profile downstream of the The low Reynolds number characterization of airfoils
separation point (Lin and Pauley, 1996). The airfoil reported in the literature are mostly for the clean airfoil
aerodynamic behavior and coherent structures are mainly configurations. However, during actual operations, these low
dependent on the flow Reynolds number and the angle of Reynolds number configurations are susceptible to surface
attack. The formation of the separation bubble on the upper roughness of various forms, size and origin. The deformation in
surface of the airfoil decreases the coherence and length scale of geometry of the airfoil may be concave or convex. The
wake vortices (Yarusevych et al., 2009). Serhiy Yarusevych protrusions on wing surface may be because of dirt deposition,
et al. (2009) proposed an alternative scaling for fundamental insect deposits, ice formation or bird littering. At low Reynolds
frequency and the wake vortex shedding frequency. In the number, the flow is highly susceptible to surface characteristics,
separated shear layer, the fundamental frequency is scaled and hence, these protrusions can alter the aerodynamics of
along with wavelength of the fundamental disturbances, which these wings considerably. For NACA 663-018 airfoil, a
makes the Strouhal number to lie between 0.45 and 0.5. In the distributed grit roughness on the leading edge resulted in a
near wake region, the vertical distances between two vortices rapid increase in lift coefficient with angle of attack, however,
are used as characteristic length, which reduces the Strouhal with decreased cl,max at a Reynolds number of 40,000. At a
number to 0.17. As the Reynolds number increases, there is a higher Reynolds number of 130,000, the negative lift at small
reduction in Strouhal number (Rostami et al., 2018). At low angle of attack was shown to be eliminated by the use of grit
Reynolds number and low angles of attack (AOA), the roughness at the leading edge (Mueller et al., 1980). The larger
formation of laminar separation bubble occurs on both surfaces protrusions like a rough ice accretion were found to reduce the
of the airfoil (Juanmian et al., 2013). As the Reynolds number cl, max and stall angle drastically with a premature stall of NACA
increases, the length and thickness of laminar separation 0012 in the Reynolds number range of 0.36  106-3.36  106.
bubble decrease (Meara and Muellert, 1987). Vortex shedding This was also found to accompany a large rise in drag and
is also affected by the angle of attack with higher angle of attack pitching moment coefficient (Korkan et al., 1984). For larger
accounting for longer vortex shedding cycles (Juanmian et al., protrusion, however, the Reynolds number variations in the
2013). As the angle of attack increases, the laminar separation moderate range was found have insignificant effect on the
bubble length and thickness increase (Meara and Muellert, aerodynamic coefficients. As the height of these ice accretions is
1987). At low AOA, the development of vortex shedding is increased, a severe degradation in the aerodynamics of airfoils is
closely associated with the shear layer instabilities (Huang and observed, except when the protuberance is located at the
Lin, 1995). As the angle of attack increases, separation point on leading edge (Lee and Bragg, 2003). As far as the shape of these
suction surface travels in the direction of the leading edge, and icy protrusions is concerned, the circular or hemispherical
there is an alteration in the mode of separation, from trailing accretion gives the highest lift coefficient as compared to
edge separation to leading edge separation, as the separation conical and triangular protrusions. Mirzaei et al. (2009)
bubble forms near the leading edge (Kojima et al., 2013). reported a crisp understanding of the characteristics of
Yarusevych et al. (2006), in an experimental study of NACA separation bubble behind protrusion caused by icing and its
0025, found that as the Reynolds number is dropped from effect on flow unsteadiness. They found that increasing the
1.5  105 to 1.0  105, the separated shear layer fails to angle of attack for NLF-0414 airfoil results in longer separation

187
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

bubbles at moderate Reynolds numbers. They also reported a low Reynolds number, it was found that the larger and taller
reduction in vortex shedding frequency as the angle of bumps trigger vortex breakdown and delay separation and may
incidence is increased. Zhang et al. (2011) in their study of the increase the L/D ratio of airfoils (Zhou and Wang, 2011).
performance degradation of low Reynolds number airfoil At low Reynolds number, the smallest of deformations in the
because of insect debris, reported that although protuberances geometry of airfoils can produce significant alterations in its
reduced the lift coefficient significantly, small roughness could aerodynamic characteristics at low Reynolds numbers. The
delay the stalling significantly. The protuberances are not aerodynamic performance may be enhanced or degraded based
necessarily always associated with performance degradation of on size, shape and location of the protrusions on wing surface.
airfoils. Some bio-inspired wings, with sinusoidal leading edge The effects these protrusions are also dependent on the thickness,
protuberance named tubercles, may increase the aerodynamic location of maximum thickness and the angle attack of the airfoil
efficiency of wings. Low pressure pockets are formed in the configuration. The macroscopic alteration in the surface of
troughs between the protuberances, which results in an airfoils because of insect deposits can reduce the output power of
increased lift coefficient at all AOA for wide range of Reynolds wind turbines by 25 per cent (Corten and Veldkamp, 2001). A
numbers in the range of 1.8  105-3.0  106. These sinusoidal detailed study on the aerodynamic performance degradation or
tubercles with height of 0.12c and wavelength 0.5c could delay
upgradation through careful examination of the associated flow
the stalling to up to 39° for the NACA 634-021 airfoils along
phenomena is thus required. In view of this, the current paper
with decreased minimum drag coefficient and a softer stall
reports the effects of size and location of a circular protrusion on
(Dropkin et al., 2012). These sinusoidal protrusions at the
flow behavior and aerodynamic characteristics of NACA 0012 at
leading edge are beneficial only in the post-stall regime as well,
a Reynolds number of 100,000.
for conventional NACA 4-digit airfoils (Hansen et al., 2011).
Smaller the protrusion height or the gap between protrusions,
higher are the lift coefficients and stalling angles with better post- 2. Numerical methodology
stall characteristics. Larger gap between the tubercles leads to a
2.1 Geometric modeling and grid generation
drop in aerodynamic efficiency of the airfoils (Serson and
C-shaped domain with multi-block structured grids are created
Meneghini, 2015). The enhancements in post-stall aerodynamic
around the NACA 0012 airfoil with and without protrusions at
characteristics and increase in stalling angle strengthens further
various locations on the surface using ICEMCFD® as shown in
with increasing Reynolds number in the range 75,000-300,000
Figure 1. The chord length of the airfoil in each case is 0.1 m
(Peristy et al., 2016). In addition, these protrusions perform better
and protrusions are located either at the leading edge or at
on thicker airfoils as compared to thinner ones (Paula et al., 2016).
Leading edges are not the only place where the protrusions can be 0.05c, 0.1c, 0.25c, 0.5c and 0.75c on the surface of the airfoil.
placed. Two- and three-dimensional bumps located at the various Protrusions at each location on suction or pressure surface have
chordwise location also show significant changes in the three different heights of 0.005c, 0.01c and 0.02c. The
aerodynamic performance of airfoils. At a Reynolds number of C-topology computational domain extends to 20 times the
64,200, a discrete roughness of height 1.25 times the local chord length in all directions, from the airfoil. A dense grid of
boundary layer thickness near the leading edge induces an earlier O-grid topology with stretching in all directions away from
breakdown of separated laminar boundary layer, resulting in airfoil is created around the airfoil. The first cell distance from
increased lift coefficients (Gross and Fasel, 2013). Vortex the surface of the airfoil in each case is set to be 0.000019 m
shedding does not appear for roughness elements with height less such that the non-dimensional cell wall distance y1  1. Based
than 50 per cent of boundary layer thickness for NACA 643-618 on a detailed grid independence study, the multi-block grid
at a = 8°, and hence, lift coefficient is unaffected. In an generated is converted to unstructured mesh with 169,750
investigation over a wing, with two discrete roughness elements at quadrilateral elements.

Figure 1 Grid around NACA 0012 with a circular protrusion on suction surface

188
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

2.1.1 Governing equations and solver @ ð r v Þ @ ð r Uj v Þ


1 ¼ ar S2  b r v 2
At low Reynolds numbers, the viscous forces dominate the @t @xj
flow, and thus, the two-dimensional unsteady Navier–Stokes  
@ @v
equations given by equations (1) and (2) are solved using finite 1 ðm 1 s v m t Þ
@xj @xj
volume solver ANSYS FLUENT®:
 
@r ! 1 @k @ v
1r  rV ¼ 0 (1) 1 2ð1  F1 Þ r s v 2 (4)
@t v @xj @xj
   
@ ! !!
r V 1 r  r V V ¼ rp 1 r  s 1 r ! g (2) "
@t @ ð r g Þ @ ð r Uj g Þ  #
@ mt @g
The s in equation (2) is the viscous stress tensor. 1 ¼ Pg  Eg 1 m1 (5)
@t @xj @xj s g @xj
The energy equation is not solved as the flow is an adiabatic
one and at a very low Mach number with negligible     " #
compressibility effects, and thus, with negligible macroscopic g
@ r Re ut
g
@ r Uj Re ut @ g
@ Re ut
heat transfer. The continuity equation and the two momentum 1 ¼ Pu t 1 s u t ðm 1 m t Þ
@t @xj @xj @xj
equations are solved in a segregated manner using a pressure
based algorithm. The coupling of pressure and velocity fields is (6)
achieved through semi-implicit method for pressure linked g
In equations (5) and (6), g is the intermittency, Re u t is the
equations scheme, which is based on the predictor-corrector
transition thickness Reynolds number and P g , E g and Pu t
approach to enforce mass conservation.
are intermittency source term, intermittency destruction
term and momentum thickness production source term,
2.2 Spatial and temporal discretization
respectively.
All the conservation equations are discretized using a second-
order upwind scheme with higher order under-relaxation factor 2.4 Boundary conditions
of 0.75 applied to all flow variables. The viscous terms in the The airfoil surface is a no-slip boundary with u = 0 and v = 0
momentum equations, however, are discretized using second- imposed on it. The other boundaries as shown in Figure 1
order central difference scheme and the gradient at cell centers are velocity inlet and pressure outlet. At the velocity inlet, a
are evaluated using the least square cell-based formula wherein uniform velocity of 15 m/s is specified with the
the variables are assumed to vary linearly inside the cells. As the intermittency of 1.0 and turbulent intensity of 0.1 and at the
low Reynolds number flows are inherently unsteady, the pressure outlet boundary static pressure of 1.0 atm is
unsteadiness is resolved using a second-order accurate implicit specified. The values at the inlet are also used to initialize
the solution.
transient formulation. Starting with the steady state or pseudo-
steady state solutions as initial guess, the equations are solved
2.5 Grid independence and solver validation
iteratively at each time step before advancing to the next time
A detailed grid independence study is done for clean NACA
level, with a fixed time step of 1e05 s.
0012 airfoil, to assess the number of cells required for a grid-
independent solution. In Figure 2(a), the pressure distribution
2.3 Turbulence modeling
over the airfoil surface for grids of various element counts
Low Reynolds number flows are susceptible to laminar
appears to be same over the entire surface. A careful
separation and a turbulent reattachment. The conventional one
examination of the peak suction pressure in Figure 2(b) reveals
and two-equation turbulence models fail to capture the transition
that the solution becomes grid-independent only for meshes
and give erroneous results in this regime (Rumsey and Spalart, with element count of 169,750 and more. In backdrop of this
2009; Wang et al., 2012). The turbulence model thus, selected mesh independence study, all simulations were started with a
here is a four-equation transition model, namely, transition Shear minimum element count of 170,000 and element count for all
stress transport model (Menter et al., 2004). Transition SST cases were increased to 300,000 during the course of the
model can predict the detailed instantaneous and mean flow solution using dynamic mesh adaptation based on pressure
features, including laminar separation of the shear layer, gradients.
transition to turbulence, reattachment of the shear layer, vortex Besides the grid independence study, solver validation is
shedding and wake characterization (Counsil and Goni done against the experimental observations by Ohtake et al.
Boulama, 2012). This model is based on the correlation-based (2007) and Rinoie and Takemura (2004). In Figure 3(a), a
coupling of the two-equation SST k-v model (Menter, 1994) very good agreement is observed between the time-averaged
with two extra equations for the transport of intermittency (g ) lift coefficient values obtained in the current study and those
and transition-onset momentum-thickness Reynolds number obtained by Ohtake et al. In addition, the time-averaged
(Reu t). The resultant set of transport equations solved in this pressure distribution over NACA 0012 at a Reynolds
turbulence model are thus, given by equations (3)-(6): number of 130,000 obtained in the current validation study
@ ð r kÞ @ ð r Uj kÞ f in close agreement with those obtained by Rinoie and
1 ¼ Pk  b  r kv
@t @xj Takemura, as can be seen in Figure 3(b). A slight
  disagreement in the surface pressures for a = 10°, seen in
@ @k
1 ðm 1 s k m t Þ (3) Figure 3(b) is primarily due the dependence of laminar
@xj @xj
separation bubble size on the freestream turbulent intensity.

189
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Figure 2 Grid independence study

Figure 3 Solver validation study

Thus, it may safely be presumed that the numerical Figure 4 Evolution of lift coefficient at a = 8°
methodology adopted in the current study gives reasonably
accurate solutions for scientific consumption.
A solver verification is also conducted against the
numerical results obtained by Council and Boulama for the
evolution of the lift coefficient of NACA 0012 at a Reynolds
number of 105 (Counsil and Goni Boulama, 2013). As
shown in Figure 4, the lift coefficient for NACA 0012 settles
to a steady value at a = 8° and in accordance with the
findings of Council and Boulama. This verification was done
as the current research uses a similar numerical
methodology especially the turbulence model. As observed
by Council and Boulama, at a Reynolds number of 105, the
converged lift coefficient of NACA 0012 shows non-
oscillatory behavior at a = 8°. The primary reason suggested
for this lift convergence is the amount of free stream
turbulent intensity, the turbulence model and comparatively
high Reynolds number that might inhibit the vortex
shedding (Counsil and Goni Boulama, 2013).

190
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

3. Results and discussion Figure 6 Vortex structure for airfoil with leading edge protrusion of
h = 0.005c and a = 16°
3.1 Effect of protrusion at the leading edge
In this research, numerical simulations are conducted to
investigate the effect circular protrusion over on airfoil at
Reynolds number of 105 and at the AOA in between 0° and
20°. For protrusions located at the leading of NACA 0012, a
marginal increment in the lift coefficient can be observed at
very small AOAs. In the range of AOA between a = 4° and 10°,
the lift curve coincides with that for a clean configuration as can
be seen in Figure 5(a). For a protrusion of height h = 0.005c,
the peak lift coefficient is same for that for the clean
configuration, the recovery from stall, however, for airfoil with
protrusion is immediate. For a leading edge protrusions of a = 18°. However, at a = 20°, the protrusion with height h =
height h = 0.01c, the stall is slightly smoother with full stall 0.01c lowers the fundamental frequency of oscillation to 40 Hz.
occurring at AOA between a = 14° and 16°. At a = 12° and
18°, the protrusion with h = 0.01c offers significant 3.2 Effect of protrusion at 0.05c on the upper surface
improvement in time-averaged lift coefficients because of cyclic The aerodynamic characteristics of an airfoil with protrusion
vortex shedding, which results in downward flow of located at 0.05c on the suction surface is significantly different
momentum near the trailing edge as a counterclockwise vortex from those of the clean configuration and those with
is shed away. The instantaneous vortical structures that cause protrusions at other locations. The lift curve for configuration
change in lift coefficient are shown in Figure 6. As can be seen with a protrusion height of 0.005c follows the one for clean
in Figure 6(a), with two counter-rotating vortices on the configuration up to a = 12°, as can be seen in Figure 7(a). In
suction side the lift is small as the flow separated from the fact, for AOAs up to 4°, the protrusion with h = 0.005c provides
leading edge fails to curl downwards. As the trailing a marginal increment in lift. At AOAs higher than 12°, the
counterclockwise vortex is shed away, clockwise circulation is
configurations with smallest of protrusion show remarkable
imparted to the airfoil and the primary clockwise vortex
increase in lift as compared to the clean configuration. The lift
becomes stronger and pushes the flow closer to the surface as
increases linearly from a = 12° to 20°, giving increments of 21,
can be seen in Figure 6(b). This results in an increased
36, 29 and 28 per cent at a = 14°, 16°, 18° and 20°,
instantaneous lift coefficient. The smaller protrusions lead to
more severe vortex shedding in lift and drag with higher respectively. For the clean configuration, the flow is stalled at
amplitudes of fluctuations and the time-averaged lift a = 12°, however, for configurations with protrusion at 0.05c
coefficients are significantly improved at AOA higher than a = location, the reversed flow on the suction surface is tripped at
12°. The drag, however, is largely unaffected by the presence of the protrusion resulting in formation of a small separation
protrusions at the leading edge, for AOA below a = 14°, as can bubble ahead of the protrusion. The periodic washing away of
be seen in Figure 5(b). this tiny vortex leads to flow unsteadiness, causing the primary
The addition of protrusion at the leading edge adds to the vortex to be centered over the airfoil mid chord for a longer
flow unsteadiness as the airfoil undergoes high frequency vortex duration before being shed away. This enhances the suction
shedding of 100 Hz, at a = 16°. The frequency of fluctuations causing the lift coefficient values to go up along with an
in the lift coefficient is similar for all protrusion heights at increased drag as can be seen in Figure 7(b). The increment in

Figure 5 Aerodynamic coefficients for NACA 0012 with circular protrusion at leading edge

191
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Figure 7 Aerodynamic coefficients of airfoil with protrusion at 0.05c on the upper surface

the drag observed is primarily because of viscous tugging at the local suction, which falls immediately due formation of a larger
protrusion. separation bubble aft of the protrusion. At a = 8°, the suction
As the protrusion height is increased from h = 0.005c to h = provided by the small vortex ahead of the protrusion is unable
0.01c, the flow pattern and unsteadiness remain unchanged, to match the peak suction by the clean configuration as can be
especially at a high angle of attack. The lift curve from a = 12° seen in Figure 8(b). The flow pattern at a = 10° is similar with
onwards is exactly similar for both the cases. This makes the lift much reduced suction and separation of flow aft of the
curve almost linear in the entire range of AOA from a = 10° to protrusion leading to a higher loss in lift. The separation
20°, with virtually no stalling with protrusion of h = 0.01c bubbles at a = 4° and a = 8° can be seen in Figure 9, which
located at 0.05c. The drag penalty, however, is higher, at all shows the instantaneous streamlines for clean airfoil and airfoil
AOA, as can be seen in Figure 7(b). At low AOAs between a = with protrusion of height 0.01c, located at 0.05c.
4° and a = 10°, the aerodynamic performance is highly As the height of protrusion at 0.05c is increased to 0.02c, the
degraded with lift coefficient falling significantly. At a = 4°, aerodynamic performance is further degraded at small AOAs,
pressure on the lower surface is similar to that for a clean airfoil, with the highest reduction of 39.7 per cent at a = 8°. The lift
the pressure on the suction surface is, however, slightly coefficient values remain almost unaltered between a = 2° to
modified. As can be seen in Figure 8(a), the formation of small a = 8°, even as the drag coefficients continue to rise with AOA.
separation bubble ahead of the protrusion results in a strong This degradation in lift is because of the inability of the flow

Figure 8 Surface pressure distribution for airfoil with protrusion of h = 0.01c at 0.05c on the upper surface

192
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Figure 9 Instantaneous streamlines at a = 8° for (a) clean airfoil, and (b) protrusion at 0.05c

separated from the protrusion surface to reattach as the 3.3 Effect of protrusion at 0.1c on suction surface
separation bubble extends up to the trailing edge at AOAs up to As with the protrusion at 0.05c, the protrusions at 0.1c
a = 8°. From a = 10° onwards, steady increments in lift improves the lift characteristic of NACA 0012 significantly as
coefficient can be observed because of the buildup of an can be seen in Figure 10(a). For smaller protrusions,
anticlockwise vortex near the trailing edge, which makes the increments from 11 to 29 per cent can be observed in the post-
primary vortex over the suction surface more coherent. stall regime between a = 12° and a = 20°. These increments
Increments of 31.9, 57.6 and 25.6 per cent in time-averaged lift come with no or little deterioration in lift values in the pre-stall
coefficient values are observed at a = 14°, 16° and 18°, regime and with no increase in drag coefficient values as can be
respectively. The increment in vortical lift comes, however, at seen in Figure 10(b). For protrusion height of 0.02c, however,
the cost of flow unsteadiness because of periodic vortex massive degradation in both lift and drag characteristics in pre-
shedding from the trailing edge. Nevertheless, the time- stall regime is observed with 37 per cent reduction in lift at
averaged lift curve slope for airfoil with 0.02c protrusion at a = 8°. The increment in lift, observed at a = 12° and 14°, is
0.05c, behaves linearly from a = 10° onwards. As the lift because of the suction created by the separation bubble formed
provided by these configurations are entirely vortical in nature, in front of the protrusion as can be seen in Figure 11(a). The
typical airfoil stall is not observed, as can be seen in Figure 7(a). flow detached from the surface of protrusion does not reattach
The protrusions located at 0.05c on the suction surface thus on the airfoil, forming a vortical flow region aft of the
induces high frequency vortex shedding at AOA as low as a = protrusion, which is stable. The lift and drag coefficient values
12°. For a protrusion of height h = 0.005c and 0.01c, the thus, do not fluctuate for AOAs up to a = 12°. At higher AOAs,
frequency of lift oscillation is 120 Hz at a = 14°, which reduces the flow becomes unsteady and as the vortices are shed away
100 Hz at a = 16°. For largest protrusion of h = 0.02c at 0.05c from the surface, the separation bubble in front the protrusion
location, an oscillation frequency of 120 Hz is achieved at a = vanishes and the flow fails to reattach on the protrusion. This
12°, which diminishes to 100 Hz at AOAs of a = 14° and 16°, results in diminished suction both ahead and aft of the
and to 40 Hz at a = 20°. protrusion as can be seen in Figure 11(b). The vortical lift on

Figure 10 Aerodynamic coefficients of airfoil with protrusion at 0.10c on the upper surface

193
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Figure 11 Instantaneous surface pressure distribution over airfoil with protrusion of height 0.01c, located at 0.1c

the rear of the surface, however, compensates for the suction significant deteriorations in aerodynamic characteristics can
destruction ahead of protrusion. be seen for larger protrusions. The lift curve for protrusion
The flow, in general, is unsteady for AOA of 14° and above of h = 0.005c follows the non-linearity of clean airfoil up to
for all protrusion heights at 0.1c location. The protrusions a = 14°, after which an increase in lift value of about 20 per
induce a 125 Hz high frequency, low amplitude oscillations in cent is seen. The drag coefficient also increases slightly from
the lift coefficient at a = 14°. As the AOA is increased, the 16° onwards. For the protrusions with h = 0.01c, the average
frequency of vortex shedding is reduced to 100 Hz at a = 16° lift in the pre stall region is slightly less than for the clean
and to 86 Hz at a = 18° while the amplitude of fluctuation in airfoil, and marginally higher in the post-stall region beyond
the lift coefficients increase. At a = 20°, the vortex shedding a = 16°.
frequency is unaltered by the presence of protrusions as the There is a high frequency oscillation in lift coefficient values
flow fully separated, the amplitude of oscillation, however, for an airfoil with smaller protrusions and at higher AOAs
decreases as the height of protrusion is increased. because of unsteady vortex structure. The instantaneous high
lift in the post-stall regime is primarily because of increased
3.4 Effect of protrusion at 0.25c on suction surface suction on the upper surface as can be seen in Figure 13(a).
Figure 12 shows the variation of aerodynamic coefficients for The increase in suction is because of the presence of a single
an airfoil with protrusion at the 0.25c on the upper surface. For dominant clockwise vortex spanning the entire upper surface,
the smallest protrusion of h = 0.005c, the lift and drag as can be seen in Figure 14(a). The vortex adds to downward
coefficients are unaffected in the pre-stall regime, while momentum of the flow hence enhancing circulation besides

Figure 12 Variation of aerodynamic coefficients for airfoil with protrusion at 0.25c

194
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Figure 13 Instantaneous surface pressure for airfoil with protrusion at 0.25c and a = 16°

Figure 14 Instantaneous vortex structure at a = 16° for airfoil with Figure 15 Lift history for airfoil with protrusion at 0.25c and a = 16°
protrusion at 0.25c location and h = 0.005

causing an increase in pressure on the pressure side. As the flow


rolls up at the trailing edge because of pressure gradient, an
anticlockwise vortex builds up at the trailing edge. As this
counterclockwise vortex grows in size, it increases suction near
trailing edge on both top and bottom surfaces as seen in
Figure 13(b), causing the lift to diminish. When the
anticlockwise vortex gains its maximum strength, the clockwise
vortex on suction side becomes elongated and moves farther formation of laminar separation bubble occurs ahead of the
from surface as seen in Figure 14(b), and the lift is reduced to protrusion and flow attaches to the protrusion and then
its minimum. After attaining full strength the counterclockwise separates aft of the protrusion before reattaching again near
vortex is shed away from the surface imparting circulation to the trailing edge. Because of this, there is a small reduction
the airfoil and the lift increases as the shed vortex moves away in lift coefficient for these AOAs as can be seen in
from trailing edge. Figure 12(a). At a = 18°, there is an enhancement in the
For the airfoil configuration discussed in Figure 14, the lift time-averaged lift coefficient by 13 per cent, which
coefficient fluctuates between 0.71 and 0.80 at a frequency of fluctuates between 1.24, when the vortex on suction surface
100 Hz. At a = 16°, the frequency of oscillation is the same for is strongest, immediately after the vortex is shed and 0.9 as
all protrusion heights while the amplitude of oscillation the counterclockwise vortex becomes strongest at the
diminishes for larger protrusions as can be seen in Figure 15. trailing edge. This increment in time-averaged lift
The frequency of oscillations for larger protrusions of h = 0.01 coefficient falls to 6 per cent at a = 20°.
and 0.02 is reduced to 90 Hz at a = 18°. At a = 20°, the lift For a protrusion of height 0.02c, located at 0.25c, no
coefficient fluctuates with a very low fundamental frequency of significant effect on the lift coefficient is observed at a = 0°
40 Hz, the amplitude, however, goes through two intermediate and 2°. For AOAs between 4° and 10° however, there is
points of inflection during each cycle. significant decrement in lift coefficient as can be seen in
For a protrusion of height 0.01c, the time averaged lift Figure 12(a). At a = 4° and 6°, flow separates ahead of the
coefficient is increased marginally as compared to the clean protrusion forming a short separation bubble and then
airfoil at a = 0°. From a = 2° to 16°, there is no significant separates again from the protrusion forming a long separation
change in lift coefficient except at a = 8° and 10° where some bubble aft of the protrusion, which extends up to the trailing
reductions can be observed. At both, a = 8° and 10°, the edge, as can be seen in Figure 16(a). This results in a

195
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Figure 16 Instantaneous streamlines for airfoil with protrusion at For protrusion with h = 0.01c, the trend in lift variations is
0.25c with h = 0.02c similar to those with h = 0.005c, as the protrusion degrades the
performance of the airfoil only slightly in the pre-stall regime and
improves the lift in the post-stall regime. At a = 16°, the time-
averaged lift coefficient is increased by 20 per cent, owing to the
coherent vortical lift in separated flow regime. The vortex
shedding at this AOA occurs at a frequency of 100 Hz causing lift
coefficient to fluctuate between 0.715 and 0.796, as can be seen in
Figure 18. At a = 18° and 20°, increments of 11 and 5 per cent,
respectively, are seen in the time-averaged lift coefficient, with an
significant loss of lift and a small increase in drag coefficient. increased amplitude of fluctuations. These improvements in lift
For a = 8° and 10°, the reduction in lift and increase in drag coefficients post 16° AOA, comes with increments in drag, as well
as can be seen in Figure 17(b). At a = 20°, the fundamental
are both amplified as the length of the vortical structure
frequency of vortex shedding is reduced to 40 Hz for all protrusion
extends beyond the trailing edge forming a hairpin vortex.
heights located at 0.5c. The amplitude of this oscillation, however,
The coherent structure at a = 8° for airfoil with protrusion
goes through three crests and troughs in each cycle.
height of 0.02c located at 0.25c on suction surface can be
As the height of protrusion, located at 0.5c, is increased to
seen in Figure 16(b). For this configuration of protrusion, the
0.02c a drastic change in the lift coefficient observed for a =
lift curve follows the one for clean configuration up to a =
16°, along with high amplitude fluctuations. A high
16°, with slightly higher average lift coefficient as compared instantaneous lift for this configuration is caused by a large
to clean airfoil. At a = 18° the vortices are shed from the suction particularly aft of the protrusion as can be seen in
airfoil at a frequency of 80 Hz and lift coefficient value Figure 19(a). A single dominating vortex seen in the cases of
fluctuates between 1.242 and 0.96. For a = 20°, the lift
coefficient fluctuates between 1.02 to 0.975, as the flow Figure 18 Lift history for airfoil with protrusion at 0.5c and a = 16°
separated from the leading edge remains unaffected by the
presence of the protrusion.

3.5 Effect of circular protrusion at 0.50c


The trends in the lift and drag variations for an airfoil with
protrusions located at 0.5c is similar to that for protrusion at
0.25c, for AOAs up to a = 14°, as can be seen Figure 17(a) and (b).
The smallest protrusion of h = 0.005c does not seem to affect
the flow when located at mid-chord, as the lift curve for airfoil
with protrusion overlaps the clean configuration curve up to
a = 16°. The amplitude of fluctuation of lift coefficient for
airfoil with protrusion is, however, slightly higher than those
for clean airfoil. Also, at a = 18° and 20°, the time-averaged
lift coefficients for airfoil with this protrusion height is
marginally higher than that for the clean configuration.

Figure 17 Aerodynamic coefficients of airfoil with protrusion at 0.50c

196
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Figure 19 Instantaneous surface pressure distribution for airfoil with protrusion of h = 0.02c at 0.5c

smaller protrusions splits into two smaller vortices one lift. The drag coefficient curve, on the other hand, overlaps
centered ahead of the protrusion and the other aft of the the one for clean configuration up to a = 14°, for all
protrusion, as can be seen in Figure 20(a). A small protrusion heights. For a = 18°, significant reductions in the
anticlockwise vortex can also be seen in front of the lift could be observed for all protrusion heights with marginal
protrusion for this high lift vortical system. Because of suction reductions in drag for smaller protrusion heights. The
from this system of vortices, the lift coefficient is increased to anticlockwise vortex seen over the trailing at a = 18° for clean
1.23, which diminishes to 0.96, as an anticlockwise vortex configurations are not present for airfoil with smaller
gain strength near the trailing edge, as shown in Figure 20(b). protrusions at 0.75c. This reduces the flow unsteadiness, and
This also forces the two clockwise vortices on the suction the vortical lift is diminished with elongated primary vortex
surface to merge as a result, the suction ahead of the on the suction surface. This elongated vortex also results in
protrusion is enhanced. Despite this, there is a decrement in reduced drag because of thinning of the wake.
lift because of enhanced suction on the bottom surface, as With protrusions at 0.75c, the vortex shedding frequency is
well as can be seen in Figure 19(b). Similar vortex shedding similar to those for the clean configuration. Significant
behavior is also observed for a = 18° and 20°, with slightly fluctuations in lift coefficient are observed at AOAs of a = 2°, 6°
smaller increment in the lift coefficient though. and 20°, for airfoils with tallest protrusions only. The
amplitude of fluctuation in lift coefficient values is, however,
3.6 Effect of protrusion at 0.75c very small as compared to the clean NACA 0012.
The aerodynamic characteristics of NACA 0012 remain
largely unchanged by the presence of protrusion located at 4. Conclusion
0.75c, as can be seen in Figure 21. For a protrusion of height
The aerodynamics of NACA 0012 with circular protrusions
0.02c, some significant reductions lift coefficient can be
located at various positions on the suction surface is
observed at a = 4° and a = 8°, primarily because of formation
investigated at a Reynolds number of 105. The protrusions
of laminar separation bubble ahead and aft of the protrusion
of height h = 0.005c, 0.01c and 0.02c are placed at 5, 25, 50
extending up to the trailing edge, which results in reduced
and 75 per cent chordwise position on the suction surface
suction on the top surface. At a = 6°, however, the separation
and at the leading edge as well. The lift curve for clean
bubble ahead of the protrusion breaks into smaller bubbles as
NACA airfoil is highly non-linear and the airfoil stalls at a =
can be seen in Figure 22. These bubbles are flushed away
10°, which recovers from stall at a = 16°. The lift and drag
periodically creating unsteadiness in the flow and an
curves are largely unaffected by a protrusion of height
increased time-averaged lift coefficient because of vortical
0.005c, for AOAs between a = 4° and 10°, irrespective of its
location. The smaller protrusions at the leading edge,
Figure 20 Vortex system for airfoil with protrusion at 0.5c on suction
however, enhances severe vortex shedding with higher
surface with h = 0.02c
amplitudes of fluctuations in lift and drag coefficients for
AOAs higher than a = 12°. The aerodynamic characteristics
of airfoil with protrusion located at 0.05c and 0.1c on the
suction surface is significantly different from those of clean
configuration and those with protrusions at other locations.
At AOAs higher than 12°, the configurations with smallest of
protrusions at 0.05c, show remarkable increase in time-
averaged lift, enhancing it by 21, 36, 29 and 28 per cent at

197
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Figure 21 Aerodynamics characteristic of airfoil with protrusion located at 0.75c on the suction surface

Figure 22 Instantaneous streamlines for airfoil at a = 6° with when the protrusion is placed at 0.75c, which also provides
protrusion at 0.75c with h = 0.02c reductions in drag coefficients at higher AOA for smaller
protrusion height.
The improvement in lift coefficients seen in the post-stall
regime for configurations with protrusion comes along with an
associated unsteadiness in the flow because vortex shedding.
The vortex shedding for airfoils with protrusion starts at lower
AOAs as compared to the clean configurations. Because of this
vortex shedding phenomena, the lift coefficient values oscillate
periodically between its high and low values as vortex builds up
on the surface and is finally shed away. With the anticlockwise
vortex at its full strength the lift coefficient is at the trough and
a = 14°, 16°, 18° and 20°, respectively. The drop in lift
regains its peak value after the anticlockwise vortex from the
coefficient at stall is very small, thus making the curve
trailing edge and a single clockwise vortex spans the entire
virtually stall free up to a = 20°. For larger protrusion,
suction surface. As the AOA is increased, the frequency of
located at 0.05c, the lift curve is almost linear in the range of oscillation decreases while the amplitude of oscillation in the lift
AOAs between a = 10° and a = 20°. At low AOAs between value increases. A high frequency, low amplitude oscillations at
a = 4° and a = 10°, however, the aerodynamic performance 120 Hz is observed for at a = 14° for smaller protrusion at 5 per
is highly degraded with lift coefficient falling significantly, cent chord on the upper surface. With most protrusions the lift
along with increments in drag. At higher AOA, the time- oscillates at a frequency between 100-80 Hz for AOAs between
averaged lift coefficients is improved significantly with 31.9, a = 16° and a = 18°. For protrusions at 0.25c and at 0.05c, lift
57.6 and 25.6 per cent increase at a = 14°, 16° and 18°. For coefficient shows a multi-modal oscillation with a small
smaller protrusions located at 0.1c, the time-averaged lift fundamental frequency of 40 Hz. In cases with these lower
coefficients are improved only by 11-29 per cent in the post- frequencies, three troughs and crests are observed in each cycle,
stall regime with no or little deterioration in the pre-stall making highest frequency of 120. For protrusions at 0.5c and
regime. For protrusion height of 0.02c, however, massive 0.75c, however, the oscillation is similar to that of the clean
degradation in both lift and drag characteristics in pre-stall configuration.
regime is observed with 37 per cent reduction in lift at a = The estimation of degradation in aerodynamic characteristics
8°. With protrusions at downstream locations on the suction of NACA 0012 by spanwise deposition or protrusion can be used
surface, the lift and drag coefficients are unaffected in the to evaluate the performance degradation of energy harvesters
pre-stall regime for smaller protrusions of h = 0.005c and operating at lower Reynolds numbers. The current findings show
0.01c, while significant deteriorations in aerodynamic that some artificial protrusions of suitable height can enhance the
characteristics can be seen for the largest protrusions. The aerodynamics performance, if placed suitably, as with the one
smallest protrusion of h = 0.005c does not seem to affect the located at 0.05c on the suction surface. The smaller protrusions
flow up to a = 16°, when located at or aft of the mid-chord also increase the frequency of oscillation of the airfoil because of
location. All protrusions, however, show significant vortex shedding, and hence, can be used as a tool to enhance
enhancement in the lift coefficient, in the post-stall regime. energy production for energy harvester based on VIVs and
Some deterioration in post-stall lift coefficients can be seen oscillating wing philosophy. As the vortex shedding phenomenon

198
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

is very sensitive to the Reynolds number, the simulations should NACA 0012 airfoil”, Journal of Aircraft, Vol. 48 No. 4,
be repeated at further lower Reynolds number. pp. 1212-1215.
Kinsey, T. and Dumas, G. (2008), “Parametric study of an
oscillating airfoil in a Power-Extraction regime”, AIAA
References Journal, Vol. 46 No. 6, pp. 1318-1330.
Argatov, I., Rautakorpi, P. and Silvennoinen, R. (2009), Kojima, R., Nonomura, T., Oyama, A. and Fujii, K. (2013),
“Estimation of the mechanical energy output of the kite wind “Large-Eddy simulation of Low-Reynolds-Number flow
generator”, Renewable Energy, Vol. 34 No. 6, pp. 1525-1532. over thick and thin NACA airfoils”, Journal of Aircraft,
Barrero-Gil, A., Pindado, S. and Avila, S. (2012), “Extracting Vol. 50 No. 1, pp. 187-196.
energy from vortex-Induced vibrations: a parametric study”, Korkan, K.D., C.C., Jr, Cross, E.J. and Cornellj, (1984),
Applied Mathematical Modelling, Vol. 36 No. 7, pp. 3153-3160. “Experimental aerodynamic characteristics of an NACA
Carmichael, B.H. (1982), “Low Reynolds number airfoil 0012 airfoil with simulated ice”, Journal of Aircraft, Vol. 22
survey”, Vol. 1 NASA Report No. 165803. No. 2, pp. 130-134.
Cherubini, A., Papini, A., Vertechy, R. and Fontana, M. Kumar, S.K., Bose, C., Ali, S.F., Sarkar, S. and Gupta, S.
(2015), “Airborne wind energy systems: a review of the (2017), “Investigations on a vortex induced vibration based
technologies”, Renewable and Sustainable Energy Reviews, energy harvester”, Applied Physics Letters, Vol. 111 No. 24.
Lee, S. and Bragg, M.B. (2003), “Investigation of factors
Vol. 51, pp. 1461-1476.
affecting Iced-Airfoil aerodynamics”, Journal of Aircraft,
Corten, G.P. and Veldkamp, H.F. (2001), “Aerodynamics:
Vol. 40 No. 3.
insects can halve wind-turbine power”, Nature, Vol. 412
Lin, J.C.M. and Pauley, L.L. (1996), “Low-Reynolds-Number
No. 6842, pp. 41-42.
separation on an airfoil”, AIAA Journal, Vol. 34 No. 8.
Counsil, J.N.N. and Goni Boulama, K. (2013), “Low-
Lissaman, P.B.S. (1983), “Low-Reynolds-number airfoils”,
Reynolds-Number aerodynamic performances of the
Annual Review of Fluid Mechanics, Vol. 15 No. 1, pp. 223-239.
NACA0012 and Selig-Donovan 7003 airfoils”, Journal of
McKinney, W. and DeLaurier, J.D. (1981), “The windmill: an
Aircraft, Vol. 50 No. 1, pp. 204-216.
oscillating-wing windmill”, Journal of Energy, Vol. 5, pp. 109-115.
Counsil, J.N.N. and Goni Boulama, K. (2012), “Validating the
Marchman, J.F. and Werme, T.D. (1984), “Clark-Y airfoil
URANS shear stress transport g -Reu model for low-
performance at low Reynolds numbers”, AIAA 22nd
Reynolds-number external aerodynamics”, International
Aerospace Sciences Meeting.
Journal for Numerical Methods in Fluids, Vol. 69 No. 8, Meara, M.M.O. and Muellert, T.J. (1987), “Laminar
pp. 1411-1432. separation bubble characteristics on an airfoil at low
Davids, S., Platzer, M.F. and Jones, K.D. (1999), “Oscillating- Reynolds numbers”, AIAA Journal, Vol. 25.
Wing power generator”, Proceedings of the 3rd ASME/JSME Menter, F.R. (1994), “Two-equation eddy-viscosity turbulence
Joint Fluids Engineering Conference July 18-23, 1999, San models for engineering applications”, AIAA Journal, Vol. 32
Francisco, CA. No. 8, pp. 1598-1605.
Dropkin, A., Custodio, D., Henoch, C.W. and Johari, H. Menter, F.R., Langtry, R.B., Likki, S.R., Suzen, Y.B., Huang,
(2012), “Computation of flowfield around an airfoil with P.G. and VöLker, S. (2004), “A Correlation-Based
leading-edge protuberances”, Journal of Aircraft, Vol. 49 transition model using local variables: part I – model
No. 5, pp. 1345-1355. formulation”, Turbo Expo 2004, Vol. 4, pp. 57-67.
Dunnmon, J.A., Stanton, S.C., Mann, B.P. and Dowell, E.H. Mirzaei, M., Ardekani, M.A. and Doosttalab, M. (2009),
(2011), “Power extraction from aeroelastic limit cycle “Numerical and experimental study of flow field
oscillations”, Journal of Fluids and Structures, Vol. 27 No. 8, characteristics of an iced airfoil”, Aerospace Science and
pp. 1182-1198. Technology, Vol. 13 No. 6, pp. 267-276.
Gross, A. and Fasel, H.F. (2013), “Numerical investigation of Mueller, T.J., Batill, S.M. and Dame, N. (1980),
passive separation control for an airfoil at low-Reynolds- “Experimental studies of the laminar separation bubble on a
number conditions”, AIAA Journal, Vol. 51 No. 7. two-dimensional airfoil at low Reynolds numbers”, AIAA
Hansen, K.L., Kelso, R.M. and Dally, B.B. (2011), 13th Fluid & Plasma Dynamics Conference.
“Performance variations of Leading-Edge tubercles for distinct Ohtake, T., Nakae, Y. and Motohashi, T. (2007),
airfoil profiles”, AIAA Journal, Vol. 49 No. 1, pp. 185-194. “Nonlinearity of the aerodynamic characteristics of
He, X.F. and Gao, J. (2013), “Wind energy harvesting based NACA0012 aerofoil at low Reynolds number”, Journal of
on flow-induced-vibration and impact”, Microelectronic the Japan Society for Aeronautical and Space Sciences,
Engineering, Vol. 111, pp. 82-86. Vol. 55 No. 644, pp. 439-445.
Huang, R.F. and Lin, C.L. (1995), “Vortex shedding and Paula, A.A.D., Padilha, B.R.M., Meneghini, J.R. and Paulo, S.
Shear-Layer instability of wing at low Reynolds number”, (2016), “The airfoil thickness effect on wavy leading edge”,
AIAA Journal, Vol. 33 No. 8, pp. 1398-1403. pp. 1-43.
Juanmian, L., Feng, G. and Can, H. (2013), “Numerical study Peng, Z. and Zhu, Q. (2009), “Energy harvesting through flow-
of separation on the trailing edge of a symmetrical airfoil at a induced oscillations of a foil”, Physics of Fluids, Vol. 21
low Reynolds number”, Chinese Journal of Aeronautics, No. 12, pp. 1-9.
Vol. 26, pp. 918-925. Peristy, L.H., Perez, R.E., Asghar, A. and Allan, W.D.E.
Kim, D.H., Chang, J.W. and Chung, J. (2011), “Low- (2016), “Reynolds number effect of leading edge tubercles
Reynolds-number effect on aerodynamic characteristics of a on airfoil aerodynamics”, AIAA Aviation.

199
Investigation of the unsteady aerodynamics Aircraft Engineering and Aerospace Technology
Aslesha Bodavula, Rajesh Yadav and Ugur Guven Volume 92 · Number 2 · 2020 · 186–200

Perkovic, L., Silva, P., Ban, M., Kranjčevic, N. and Duic, N. Winslow, J., Otsuka, H., Govindarajan, B. and Chopra, I.
(2013), “Harvesting high altitude wind energy for power (2017), “Basic understanding of airfoil characteristics at low
production: the concept based on Magnus’ effect”, Applied Reynolds numbers”, 7th AHS Technical Meeting on VTOL
Energy, Vol. 101, pp. 151-160. Unmanned Aircraft Systems.
Rinoie, K. and Takemura, N. (2004), “Oscillating behaviour Xiao, Q., Liao, W., Yang, S. and Peng, Y. (2012), “How
of laminar separation bubble formed on an aerofoil near motion trajectory affects energy extraction performance of a
stall”, The Aeronautical Journal, Vol. 108 No. 1081, Biomimic energy generator with an oscillating foil?”,
pp. 153-163. Renewable Energy, Vol. 37 No. 1, pp. 61-75.
Rostami, A.B. and Armandei, M. (2017), “Renewable energy Xiao, Q. and Zhu, Q. (2014), “A review on flow energy
harvesting by vortex-induced motions: review and harvesters based on flapping foils”, Journal of Fluids and
benchmarking of technologies”, Renewable and Sustainable Structures, Vol. 46, pp. 174-191.
Energy Reviews, Vol. 70, pp. 193-214. Yarusevych, S., Sullivan, P.E. and Kawall, J.G. (2009), “On
Rostami, A.B., Mobasheramini, M. and Fernandes, A.C. (2018), vortex shedding from an airfoil in low-Reynolds-number
“Strouhal number of flat and flapped plates at moderate flows”, Journal of Fluid Mechanics, Vol. 632, pp. 245-271.
Reynolds number and different angles of attack: experimental Yarusevych, S., Sullivan, P.E., Kawall, J.G., Yarusevych, S.
data”, Acta Mechanica, Vol. 230 No. 1, pp. 333-349. and Sullivan, P.E. (2006), “Coherent structures in an airfoil
Rumsey, C.L. and Spalart, P.R. (2009), “Turbulence model boundary layer and wake at low Reynolds numbers”, Physics
behaviour in low Reynolds number regions of aerodynamic of Fluids, Vol. 44101.
flowfields”, AIAA Journal, Vol. 47 No. 4, pp. 982-993. Zhang, Y., Igarashi, T. and Hu, H. (2011), “Experimental
Serson, D. and Meneghini, J.R. (2015), “Numerical study of investigations on the performance degradation of a Low-
wings with wavy leading and trailing edges”, Procedia Iutam, Reynolds-Number airfoil with”, 49th AIAA Aerospace
Vol. 14, pp. 563-569. Sciences Meeting including the New Horizons Forum and
Wang, D.A. and Chang, K.H. (2010), “Electromagnetic Aerospace Exposition, pp. 1-18.
energy harvesting from flow induced vibration”, Zhou, Y. and Wang, Z.J. (2011), “Effects of surface
Microelectronics Journal, Vol. 41 No. 6, pp. 356-364. roughness on laminar separation bubble over a wing at a
Wang, D.A., Chiu, C.Y. and Pham, H.T. (2012), low-Reynolds number”, 49th AIAA Aerospace Sciences
“Electromagnetic energy harvesting from vibrations induced by Meeting including the New Horizons Forum and Aerospace
kármán vortex street”, Mechatronics, Vol. 22 No. 6, pp. 746-756. Exposition, pp. 1-18.
Wang, S., Ingham, D.B., Ma, L., Pourkashanian, M. and Tao,
Z. (2012), “Turbulence modeling of deep dynamic stall at Corresponding author
relatively low Reynolds number”, Journal of Fluids and Rajesh Yadav can be contacted at: RAJESHYADAV@ddn.
Structures, Vol. 33, pp. 191-209. upes.ac.in

For instructions on how to order reprints of this article, please visit our website:
www.emeraldgrouppublishing.com/licensing/reprints.htm
Or contact us for further details: permissions@emeraldinsight.com

200

You might also like