Anda di halaman 1dari 18

Numerical investigation into the blade and wake aerodynamics of an H-rotor vertical

axis wind turbine


H. Y. Peng, H. F. Lam, and H. J. Liu

Citation: Journal of Renewable and Sustainable Energy 10, 053305 (2018); doi: 10.1063/1.5040297
View online: https://doi.org/10.1063/1.5040297
View Table of Contents: http://aip.scitation.org/toc/rse/10/5
Published by the American Institute of Physics
JOURNAL OF RENEWABLE AND SUSTAINABLE ENERGY 10, 053305 (2018)

Numerical investigation into the blade and wake


aerodynamics of an H-rotor vertical axis wind turbine
H. Y. Peng,1,a) H. F. Lam,2 and H. J. Liu1
1
School of Civil and Environmental Engineering, Harbin Institute of Technology, Shenzhen,
China
2
Department of Architecture and Civil Engineering, City University of Hong Kong, Hong
Kong, China
(Received 16 May 2018; accepted 6 September 2018; published online 21 September 2018)

Blade aerodynamics is of critical importance in the design of an isolated vertical axis


wind turbine (VAWT). In contrast, wake aerodynamics plays a crucial role in the
optimum placement of multiple VAWTs. In this study, both the blade and the wake
aerodynamics of a straight-bladed VAWT are investigated using a three-dimensional
computational fluid dynamics (CFD) model. The algebraic wall-modeled large eddy
simulation (LES) was used for turbulence modeling. The LES predictions were
compared with and had good agreement with the published experimental data. A
semi-empirical method was proposed to estimate the convergence time of CFD cal-
culations. Further, guidelines for the LES modeling of VAWTs were provided.
Relatively thick airfoils are recommended for VAWTs to reduce fatigue loading on
blades. Torque analysis suggested that winds in built environments would benefit
the self-starting performance. Comparisons between transient and time-averaged
velocity fields indicated a statistically stable wake. Moreover, the blade speed ratio
effect on the development of wake velocities was assessed. The underlying causes
of the significant wake asymmetry were addressed. The periodical vortex-ring
structures, counter-rotating vortical motions, and wake asymmetry were analyzed
and concluded as the signature features of H-rotor VAWTs’ wake. Published by
AIP Publishing. https://doi.org/10.1063/1.5040297

I. INTRODUCTION
The exploitation of renewable and clean wind energy has increased substantially during the
past decade (Kjellin et al., 2011; Islam et al., 2013; Bedon et al., 2015). Recent studies have
suggested that vertical axis wind turbines (VAWTs) have great potential in urban (Knight,
2004; Chen et al., 2017; Lam et al., 2018) and offshore (Borg et al., 2014; Tjiu et al., 2015;
Chowdhury et al., 2016) areas. Aerodynamic research on VAWTs is classified into two basic
categories: (a) blade aerodynamics and (b) wake aerodynamics. Analysis of blade aerodynamics
plays a critical role in improving the power performance of an isolated VAWT. In contrast,
study of wake aerodynamics is of crucial importance for raising the power output of multiple
VAWTs in clusters. The wake of a VAWT is affected by velocity deficit and turbulence
increase because of: (a) blockage of the turbine and (b) power extraction of the turbine. As doz-
ens to even hundreds of turbines may be placed in clusters (Li et al., 2016), a great majority of
these are subject to wake interference and hence reduced power production (Vermeer et al.,
2003; Sanderse, 2009; Dabiri, 2011; Barthelmie and Pryor, 2013). In the literature to date, great
research efforts have been devoted to blade aerodynamics of VAWTs, most of which comprise
blades with no camber. However, research on wake aerodynamics of VAWTs, especially for
cambered-bladed ones, is somewhat limited (Tescione et al., 2014; Peng et al., 2016), and

a)
Author to whom correspondence should be addressed: penghuayi@hit.edu.cn

1941-7012/2018/10(5)/053305/17/$30.00 10, 053305-1 Published by AIP Publishing.


053305-2 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

knowledge of wake aerodynamics is far from adequate (Shamsoddin and Porte-Agel, 2016;
Araya et al., 2017).
Vallverdu (2014) improved the double-multiple streamtube model (DMST) by considering
the dynamic stall effect together with the wake interaction. Though the DMST model turned
out to be very useful to achieve fast VAWTs’ aerodynamics, it was not suitable for precise tur-
bine design. In comparison with the streamtube model, a fairly advanced aerodynamic model,
i.e., the vortex model, was developed. Li and Calisal (2010a) proposed a newly derived
discrete-vortex model named as DVM-UBC for vertical axis straight-bladed tidal turbines. The
DVM-UBC model was able to predict aerodynamic results comparable to the unsteady
Reynolds-averaged Navier-Stokes (URANS) models, with far less computational effort. Further
investigation suggests that the three-dimensional (3-D) effect was significant if the turbine
aspect ratio (AR) was less than one, whereas it became negligible for AR  3.5 (Li and Calisal,
2010b). A two-dimensional (2-D) model was recommended at the early turbine design stage
due to its economic computational cost, with acceptable accuracy. Shi et al. (2014) developed a
consistent vortex model combining a vortex-type free wake potential flow solver with viscous
corrections of loads for the analysis of VAWTs’ aerodynamics. The consistent vortex model
was validated by both wind tunnel tests and field measurements and proved to be able to accu-
rately predict the aerodynamic loading and power output. On top of these classical analytical
aerodynamic models, computational fluid dynamics (CFD) simulation models, i.e., URANS,
began to seize researchers’ interest in solving the complex aerodynamics of VAWTs.
Howell et al. (2010) analyzed the blade aerodynamics of a straight-bladed VAWT using
wind tunnel tests and 2-D and 3-D CFD simulations. The re-normalization group (RNG) k–e
(Yakhot and Orszag, 1986) was used for modeling turbulence. The CFD predictions were com-
pared with the wind tunnel test data. The 2-D CFD model considerably overestimated the
power coefficient, CP, compared with the measurements (Howell et al., 2010). In contrast, the
3-D CFD model showed close agreement with the experimental data. Lee and Lim (2015)
investigated effects of various parameters on aerodynamic performance of Darrieus-type
VAWTs using 3-D CFD simulations. The RNG k–e was used to model the turbulence, and the
predicted CP values agreed well with the wind tunnel test measurements. The studies revealed
that a VAWT with no helical angle, i.e., a straight-bladed VAWT, demonstrated the best perfor-
mance (Lee and Lim, 2015). Bedon et al. (2016) developed a method for optimizing the airfoil
section of an H-rotor VAWT based on transient 2-D CFD simulations. The URANS equations
were closed by the shear stress transport k–x (SST) model (Menter, 1994). The rated power
production of a turbine with an optimized airfoil noticeably outperformed the baseline configu-
ration with a NACA 0018 airfoil. Bachant et al. (2018) proposed an actuator line model (ALM)
to study the aerodynamics of VAWTs. The ALM is in fact a hybrid numerical model which
combines the classical blade element theory with the CFD simulation. The ALM was able to
capture some of the important features of VAWTs’ aerodynamics with a significantly reduced
computational cost. In comparison, the high-fidelity pure CFD simulation is able to finely
resolve the complex flow fields of VAWTs, with predicted results comparable to experiments
(Wong et al., 2018). It has been noticed that intensive research has been conducted on blade
aerodynamics of VAWTs with symmetrical rotors. Nevertheless, studies of blade aerodynamics
of cambered-bladed VAWTs with reliable self-starting performance are rather limited.
Tescione et al. (2014) carried out particle image velocimetry (PIV) tests to study the devel-
opment of the near wake of a straight-bladed VAWT up to two turbine diameters (2D) down-
stream. The wake had an asymmetric pattern on the blade mid-span plane. Moreover, earlier
breakdown and vertical movement of tip vortices might contribute to a faster wake recovery.
Lam and Peng (2016) studied the wake aerodynamics of a straight-bladed VAWT using 2-D
and 3-D CFD simulations. The transition SST model (Menter et al., 2004) was used. The 2-D
CFD model underestimated the velocity deficits in the near wake, and overestimated the deficits
in the far wake. The under- and overestimation was attributed to the 2-D model, unlike the 3-D
model, being unable to consider span-wise motions and tip vortices. Posa et al. (2016) investi-
gated the wake structure of a straight-bladed VAWT with PIV tests and large eddy simulations
(LESs). The LES results were in good agreement with the experimental data. The vortical
053305-3 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

structures at two blade speed ratios (BSRs) were analyzed, and BSR was found to be an impor-
tant factor in wake development. Peng et al. (2016) systematically measured the flow field of a
straight-bladed VAWT up to 10D downstream to improve understanding and knowledge of the
wake aerodynamics. Significant wake asymmetry was found, and an engineering wake model
with that asymmetry was proposed. In addition, a pair of counter-rotating (CR) vortical motions
was identified (Peng et al., 2016), which was believed to contribute to enhanced flow mixing
and fast wake recovery. Li and colleagues conducted field experiments to investigate the wake
characteristics of an H-rotor VAWT (Li et al., 2017). The recovery process of the velocity field
was observed to be asymmetric. A Gaussian function model was proposed to predict the veloc-
ity distribution in the wake. Close scrutiny of the literature suggests that systematic wake analy-
sis of straight-bladed VAWTs addressing the development of velocities and turbulent structures
downstream is still lacking.
In this study, both the blade and the wake aerodynamics of a straight-bladed VAWT were
investigated with full-scale 3-D LES. Algebraic wall-modeled LES (WM-LES) (Shur et al.,
2008), which has the potential to overcome the Reynolds number scaling limitations of LES
models, was used. Stream-wise velocities in the wake from the wind tunnel tests (Peng et al.,
2016) were used to validate the LES predictions. The validated LES model was used to assess
dynamic stall and wake structures of the VAWT. Moreover, the effect of BSR on the wake
development of the VAWT was assessed. In addition, the time-averaged and the transient
velocity contours in the wake were examined. Finally, the signature wake characteristics of
VAWTs’ wake were classified and analyzed.

II. MODELING METHODS


A. Description of the turbine
The VAWT has five blades (i.e., N ¼ 5) straight extruded from a cambered airfoil. The
cambered blade has the advantage of deflecting the approaching flow onto its upper and lower
surfaces, and hence can produce lift even at a zero angle of attack (AOA, a). The blade is
mounted onto the central shaft via a strut. The blade is pitched at the blade mid-chord with a
preset pitch angle, b ¼ 10 . The five blades, strut, and shaft function as an integrated structure
to rotate at the same angular speed, x. The diameter of the turbine is D ¼ 300 mm, and the
depth of the blade is Hb ¼ 300 mm. As a result, the aspect ratio of the turbine is Hb/D ¼ 1.0.
The chord length of the blade is c ¼ 45 mm. Consequently, the turbine has a solidity ratio of
r ¼ Nc/D ¼ 0.75, which corresponds to a high-solidity VAWT. The blade mid-span stands at
the mid-height of the wind tunnel. For detailed information about the wind tunnel, please refer
to the experiments in Peng et al. (2016).

B. Turbulence modeling
Because URANS models demonstrate unstable performance when predicting large eddies
and their effects (Sim~ao Ferreira, 2009), they are not used for resolving complex blade and
wake aerodynamics. LES models, in which large eddies are directly resolved and small eddies
are modeled (Versteeg and Malalasekera, 2007), have demonstrated superior performance in
predicting the aerodynamics of VAWTs (Elkhoury et al., 2015). Nevertheless, for boundary
layer flows, even the largest scales of turbulence are geometrically small and require very fine
spatial and temporal discretization. Moreover, the mesh resolution increases exponentially with
the Reynolds number. The algebraic WM-LES model (Shur et al., 2008) is able to well address
this kind of problem. In the WM-LES, the inner part of the log layer is handled by RANS
equations, whereas the outer part of the log layer is taken care of by LES. This means that the
WM-LES approach can be applied to flows with an ever-increasing Reynolds number by using
a moderate grid resolution. The WM-LES fits the purpose of this study where the wall-bounded
flows and the wake flows are to be resolved. The essential feature of WM-LES is its new defi-
nition of the subgrid length scale, D. The subgrid length scale considers not only the grid spac-
ing but also the wall distance (Shur et al., 2008)
053305-4 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

 
D ¼ min max½Cw dw ; Cw hmax ; hwn ; hmax ; hmax ¼ maxðhx ; hy ; hz Þ; (1)

where hx, hy, and hz are the local stream-wise, wall-normal, and cross-stream grid sizes, respec-
tively; hmax denotes the maximum local spacing; Cw is an empirical constant; dw represents the
distance to the wall; and hwn is the grid step size in the wall-normal direction.
The eddy viscosity is calculated using D through the combination of the Prandtl-van Driest
RANS and the Smagorinsky subgrid scale (SGS) models (Shur et al., 2008)
n on h  3 i o
vt ¼ min ðkdw Þ2 ; ðCSMAG DÞ2 1  exp  yþ =25 S; (2)

where k is the von Karman constant, CSMAG denotes the Smagorinsky constant, yþ is the non-
dimensional wall distance, and S represents the magnitude of the strain tensor.
To bridge the modeled and resolved flows, it is assumed that the SGS stresses are propor-
tional to the local rate of strain of the resolved flows (Boussinesq’s hypothesis)

1
sij ¼ 2qvt Sij þ sii dij ; (3)
3

where sij is the SGS stress, Sij ¼ 1=2ð@ui =@xj þ @uj =@xi Þ represents the resolved rate of strain
with ui and uj being the resolved velocity components, and dij is the Kronecker delta. By math-
ematical processing, the filtered Navier-Stokes equations for incompressible flows are rewritten
in the tensor form as follows (Jimenez et al., 2007)

@ ui
¼ 0; (4)
@xi

@ ui @ui uj 1 @ p @Sij


þ ¼ þ 2ðvk þ vt Þ ; (5)
@t @xj q @xi @xj

where vk ¼ l=q is the kinetic viscosity with l being the molecular viscosity. As a result, the
target flows can be properly resolved.

C. Computational domain and mesh


A schematic of the numerical setup in the computational domain is presented in a perspective
view in Fig. 1. In addition, the top view of the computational domain is shown in Fig. 2. A 3-D
Cartesian coordinate system was introduced with the origin at the tower base. As shown in Fig. 1,

FIG. 1. Perspective view of the schematic of the computational domain.


053305-5 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

FIG. 2. Plan view of the schematic of the computational domain.

the x-axis aligns with the free-stream wind direction pointing downstream, the y-axis is perpendic-
ular to the wind direction, and the z-axis aligns with the tower centerline pointing upward.
To make reasonable comparisons with the measurement results, the cross-section of the
computational domain, i.e., H  W (height  width), was defined to be the same as the tunnel’s
cross-section area. This induced a blockage ratio of 1.8% based on the frontal swept area. It is
recommended that a blockage ratio be below the range of 6.0%–7.5% for neglecting the tunnel
wall’s interference with the flow over the model (Howell et al., 2010); hence, no correction to
the predictions was required. Moreover, to be consistent with the experimental settings, the
VAWT was positioned at the mid-width of the tunnel, and the blade mid-span was located at
z ¼ H/2. The wind entrance was positioned at L1 ¼ 6D upstream of the turbine. The wind exit
was positioned at L2 ¼ 16D downstream of the turbine.
The computational domain was divided into two parts: (1) the rotational zone and (2) the
stationary zone; the purpose of which was to simulate the motion of the turbine. One part corre-
sponded to the rotational cylinder incorporating the five blades and the central shaft; the other
to the stationary zone. The struts were not included in the modeling, as the effect of blade–strut
interaction is negligible (Elkhoury et al., 2015). The absence of the struts in the computational
domain enables the production of a structured mesh, which benefits both computational effi-
ciency and solution accuracy.
As mentioned above, the structured mesh was produced in both the rotational and the station-
ary regions to give high-quality computational results. The mesh in the two regions was generated
independently. The mesh in the rotational region contains five blades, and the central shaft is
shown in Fig. 3. The rotational and stationary mesh zones were later linked together by a non-
conformal interface pair (see Fig. 2). The interface pair comprised two individual interface zones:
one at the rotational zone and the other at the stationary zone (see Fig. 4). An interior zone was
formed where the two interface zones overlapped. Data communication between the two regions
could be accomplished as a result. In the vicinity of the blades, a high-resolution mesh was pro-
duced to accurately calculate the boundary layer flows, which are critical for predicting the blade
aerodynamics and wake characteristics. The yþ values were set to be well below 1.0 for the first-
layer grids off blades, and the growth rate of the grid sizes off blades was set to be 1.1. Each of
the individual blades was covered by 40 layers of grids in the normal direction. Along the chord-
wise direction, each layer of mesh off a blade was covered by 140 grid points. In the span-wise
direction, 40 grid points were used to discretize an individual blade span.

D. Boundary conditions and solution methods


The velocity inlet was set for the wind entrance upstream of the VAWT. Outflow was
defined for the wind exit downstream of the turbine. In accordance with wind tunnel conditions
053305-6 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

FIG. 3. Isometric view of the mesh in the computational domain.

FIG. 4. Close-up view of the mesh near the interface zone (top view).

(Franke et al., 2007), a free-slip wall with zero wall shear stress was applied to the four side
walls. The sliding mesh technique was used to handle turbine rotation. To reproduce the experi-
mental results, the approaching flow conditions were directly applied to the LES model. The
settings of the parameters in the LES model are summarized in Table I. The approaching wind
speed was U0 ¼ 11.3 m/s, with a background turbulence intensity of I0 ¼ 2.5%. The BSR of the
VAWT was k ¼ 0.75. Accordingly, a constant rotational speed, VB ¼ 539.5 RPM, was defined
for the rotational zone, which slides over the stationary zone. Meanwhile, the non-conformal
interface was constructed to provide data communication between the two zones. For accurate
prediction, the time step size was set sufficiently small at d ¼ 0.5 , which means that the blades
rotated 0.5 per time step.
Via the discretization of space and time and the application of boundary conditions, the fil-
tered partial differential equations, i.e., Eqs. (4) and (5), were converted into algebraic equa-
tions. ANSYS Fluent was used as the CFD solver for these algebraic equations. The flow fields
governed by these equations were solved in the double-precision mode by the pressure-velocity
coupling SIMPLEC (semi-implicit method for pressure-linked equations consistent) scheme.
053305-7 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

TABLE I. Settings of the parameters in the LES model.

U0 I0 VB k d

Values 11.3 m/s 2.5% 539.5 RPM 0.75 0.5

III. VERIFICATION AND VALIDATION


A. Independence studies
In the finite volume method (FVM), the calculated results depend essentially on discretiza-
tion of the computational domain. To reduce numerical error, the effect of grid sizes on the
computational results was checked. The stream-wise velocities, U/U0, at the blade mid-span,
i.e., z ¼ H/2, in the wake predicted by LES models with different mesh densities are presented
in Fig. 5. Three kinds of mesh were generated: (1) a coarse mesh with 1.69  106 cells, (2) a
medium mesh with 2.78  106 cells, and (3) a fine mesh with 3.63  106 cells. It can be seen
that the U/U0 profile by the model with the medium mesh does not vary much from and agrees
well with that by the model with the fine mesh. In contrast, the model with the coarse mesh
does not capture the U/U0 profile well. To reduce the computational effort, the LES model with
the medium mesh was selected. The grid number of the medium mesh is at the same level as
those in previous LES studies of wind turbines (Mo et al., 2013; Xie and Archer, 2014).
In a transient CFD simulation, temporal discretization is of equal importance to spatial dis-
cretization. To assess the time step size, a systematic reduction in d was performed: d ¼ 4.0 ,
d ¼ 0.5 , and d ¼ 0.25 were chosen for the LES model. Figure 6 shows the U/U0 profiles eval-
uated using different time step sizes. The U/U0 profiles predicted by the LES models with
d ¼ 0.5 and d ¼ 0.25 almost overlap each other; the model with d ¼ 4.0 fails to accurately
predict U/U0. Hence, d ¼ 0.5 was chosen.
The initial flow conditions influence the predicted results in the transient CFD problems.
Sufficient computational time is required to reduce the initial conditions on the results. An
approach was proposed for approximating the minimum number of revolutions, NR, required.
The time for the flow to pass through the computational domain is T0 ¼ L0/U0, where L0 repre-
sents the longitudinal length of the domain, i.e., L0 ¼ L1 þ L2 as shown in Figs. 1 and 2. The
fundamental rotation period of the VAWT is T ¼ (pD)/(kU0). The approach is based on the
computation time being sufficiently large that the free-stream flow can pass through the whole
computational domain. Therefore, NR has a semi-empirical form

T0 L0
NR  ¼k : (6)
T pD

To check whether the influence of the initial conditions was indeed reduced, the number of rev-
olutions was systematically increased. Figure 7 shows U/U0 for three numbers of revolutions,

FIG. 5. Effect of the mesh size on the predicted velocity along z ¼ H/2.
053305-8 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

FIG. 6. Effect of the time step size on the predicted velocity along z ¼ H/2.

the 3rd, 7th, and 10th revolutions. It can be seen that the U/U0 profiles for the 7th and 10th rev-
olutions coincide. In contrast, the predictions for the 3rd revolution deviate from the U/U0 pro-
files for the 7th and 10th revolutions. This indicates that insufficient computational time intro-
duces numerical error. From Eq. (6), NR should be 5.3. The equation correctly predicted that
NR ¼ 3 is insufficient whereas NR ¼ 7 is satisfactory. The predicted results from the 7th revolu-
tion were thus used.
Finally, the effect of rotational zone size on the computational results was examined. The
radius of the rotational zone was systematically increased by an increment in the blade chord
length, c. Figure 8 presents U/U0 profiles for LES models with three different radii, R0 ¼ R þ c,
R0 ¼ R þ 2c, and R0 ¼ R þ 3c. The number of grids in the three models was kept the same to
ensure that the comparisons would be meaningful. Based on the results shown in Fig. 8, the
radius of the rotational zone exerts a very limited influence on predicted U/U0. The radius of
the rotational zone was selected as R0 ¼ R þ 2c according to Peng and Lam (2016).

B. Comparisons with wind tunnel tests


The stream-wise velocities measured at the blade mid-span level in the wake are adopted
to validate the LES model, which is similar to the validation measure by Abkar and Dabiri
(2017). Comparisons between the LES predictions and the wind tunnel test data at z ¼ H/2 from
x ¼ 1D to x ¼ 10D are presented in Fig. 9. The overall match is encouragingly close. As a
result, the LES model is considered to be acceptable for further analysis of the VAWT. The
wake asymmetry of the VAWT is accurately predicted by the LES model. Starting at x ¼ 2D,
the shear layers to windward and leeward begin to be bridged. Referring to analyses of the
wakes of horizontal axis wind turbines (HAWTs) (Vermeer et al., 2003), x ¼ 2D marks the end
of the near wake in this study. Careful examination of the matching results shown in Fig. 9
reveals that the velocity deficits in the near wake are underestimated by the LES model. The

FIG. 7. Effect of the revolution number on the predicted velocities along z ¼ H/2.
053305-9 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

FIG. 8. Effect of the rotational zone size on the predicted velocities along z ¼ H/2.

same kind of problem is encountered in wake predictions for HAWTs using CFD simulations
(El Kasmi and Masson, 2008; AbdelSalam and Ramalingam, 2014). A possible reason for the
underestimation is that the near wake experiences the most violent and complex turbulence,
which remains a challenging problem for turbulence modeling. Another reason may be that the
struts were not included in the LES model. The solid presence of the struts imposes a blockage
effect on the air flows, and hence contributes to larger velocity deficits than that by CFD pre-
dictions in the wake. As it approaches a large distance downstream, the wake region slightly
shifts windward. It is assumed that the WM-LES model tends to overestimate the windward
expansion as the wake distance increases.

FIG. 9. Comparisons between the LES predictions and wind tunnel test data.
053305-10 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

IV. RESULTS AND DISCUSSION


A. Blade aerodynamics
Aerodynamic forces on the blades are of fundamental importance in the operation and
power production of VAWTs. Accordingly, intensive research effort has been devoted to the
study of these aerodynamic forces. Figure 10 presents a schematic of the velocities of and
forces on a blade. Figure 10 is based on the assumption that there is no loss in flow momentum
through the turbine, and it provides the understanding of the severity of the relative angle of
attack at the upstream half-revolution (0  h < 180 ) (McLaren, 2011). When flows pass over
the upper and lower surfaces of a blade, pressure gradients cause lift, FL, and drag, FD, forces
to be produced. FL is perpendicular to the resultant velocity, VR, whereas FD is parallel to VR.
FL and FD can be further decomposed into thrust, FT, and radial, FR, forces. FT tangential to
the blade speed, VB, together with the radius forms the torque, which drives the turbine to rotate
and hence produces electrical power.
Figure 11 shows the lift and thrust coefficients, CL and CT, respectively, at four BSRs. In
Fig. 11(a), the magnitude of the CL peak increases as the BSR increases. As the BSR increases,
the occurrence of dynamic stall is delayed. The bell shape of the CL curve covers a much
broader area at a larger BSR at the upstream half-revolution (0  h < 180 ). Conversely, at the
downstream half-revolution (180 < h  360 ), almost no differences exist between these CL
curves. A thin airfoil section is normally associated with abrupt losses of CL (Lan and Jan,
2003). At k ¼ 0.25, an abrupt drop of CL occurs at an azimuthal angle of 104.5 . The loss of
CL is assumed to be caused by a deep dynamic stall, due to which, vortices are drastically shed
from both the leading and trailing edges and thereby strong flow separations take place. The
drastic vortex shedding and strong flow separations from the blade at h ¼ 104.5 can be clearly
observed in Fig. 12. Afterwards, the intensities of vortex shedding and flow separations start to
decrease, and flow attachments take place. This may help to explain the abrupt increase in CL
at an azimuthal angle between 240 and 270 . The abrupt changes of CL also occur at k ¼ 0.75.
As the BSR further increases, the maximum angle of attack (AOA, a) of a blade will decrease.
Hence, the extent of the dynamic stall will be alleviated. As a result, there are no abrupt
changes of CL when the BSR reaches higher values such as k ¼ 1.0 and k ¼ 1.25.
In Fig. 11(b), it can be seen that at a larger BSR, the CT peak occurs at a higher azimuthal
angle, which is consistent with Fig. 11(a), though phase shifts between CL and CT peaks are
observed. This phenomenon is attributed to that the occurrence of a dynamic stall is delayed to
a higher azimuthal angle at a larger BSR. Along the downstream half-revolution, CT values are
close to zero. It indicates that a blade does not produce any electrical power at its downstream
half-revolution, regardless of the BSR.
Figure 13 shows the torque coefficients of the turbine at two different BSRs. The torque
coefficients have N ¼ 5, i.e., the number of blades, periodic bins which evenly distribute over a

FIG. 10. Schematic of the velocities of and forces on a single blade.


053305-11 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

FIG. 11. Lift and thrust coefficients at different BSRs of a single blade.

FIG. 12. z-vorticity contour on the blade mid-span plane (units, 1/s).

full 360 . The phase shifts between the torque coefficients are attributed to the delayed dynamic
stall at high BSRs as discussed above. In addition to those favorable peak values, there are tor-
ques of very low magnitudes at some azimuthal positions. This kind of phenomenon reflects
the significant fluctuating feature of the torque of a VAWT as systematically investigated by Li
and Calisal (2010c). If the turbine happens to be seated at these azimuthal positions of the low
torque, it will encounter the self-starting problems. In this circumstance, winds characterized by
frequently changed directions and speeds in urban environments benefit the self-starting of
VAWTs. Moreover, it is clear that the turbine operating at a lower BSR experiences a much
severer dynamic stall as more torque losses occur at a lower BSR.

B. Wake aerodynamics
The 3-D vortical structures that developed in the wake of the straight-bladed VAWT were
examined further. To identify the characteristic pattern of these vortices, the widely used Q
053305-12 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

FIG. 13. Torque coefficients of the turbine at two BSRs.

criterion (Hunt et al., 1988), which treats vortices as connected fluid regions with a positive
second invariant of the velocity gradients, was used. In incompressible flows, the Q criterion is
a local measure of the excess of the rotation rate to the strain rate. As a result, Q > 0 suggests
the potential existence of a vortex (Chakraborty et al., 2005).
Figure 14 presents a 3-D view of the vortical structures on the isosurface plotted at Q ¼ 50
and h ¼ 67 in the wake. The contour bar shows the magnitude of the stream-wise velocity.
The vortical structures demonstrate a vortex-ring (VR) shape. The diameter of the VR structures
expands as the downstream distance increases. This finding is consistent with the experimental
results (Peng et al., 2016) that the length scales of vortices grow towards downstream. Within
the turbine-swept regions, the tip and span vortices are characterized by complex vortex shed-
ding. The tip vortices contribute to the top and bottom parts for the formation of the VR

FIG. 14. Isosurface plot of the Q criterion (Q ¼ 50) at h ¼ 67 (units, 1/s2).
053305-13 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

structures, whereas the span vortices form the mid-outer parts of the VR structures as shown in
Fig. 14. The tip vortices evolve in conjunction with the span vortices to develop the periodical
vortex-ring (VR) structures as vortices are shed periodically due to the rotating blades. The tip
and span vortices are shed predominantly towards windward due to the dynamic stall [refer to
Fig. 11(a)]. As a consequence, these VR structures evolve primarily towards windward. In con-
trast, there are no VR structures to leeward. Therefore, the vortical structures to leeward are far
smaller than their windward counterparts in terms of length scales. These characteristics result
in the strong wake asymmetry escalating with the downstream distance in the horizontal direc-
tion. Furthermore, the normal unit vector of VR structures is not parallel to the approaching
wind direction. This phenomenon is attributed to that the outer parts of the VR structures
recover much faster and hence have larger velocities than their inner counterparts. A tower
wake is also observed. The turbulent structures of the tower are moderate due to the non-
rotating property of the tower, and hence the tower wake vanishes very fast.
Figure 15 shows the z-vorticity on the blade mid-span plane and at the tower at h ¼ 67 .
The black line characterized by y ¼ 0 and z ¼ H/2 on the blade mid-span plane functions as a
reference line. With reference to this line, significant wake asymmetry in the horizontal direc-
tion can be clearly visualized. The wake asymmetry develops further as the downstream dis-
tance increases. Towards windward, the vortices are shed periodically from the blade mid-span
due to the dynamic stall and have negative values, implying clockwise vortical motions. The
length scales of these vortices increase with the downstream distance. The vortices contribute
to the formation of the periodic VR structures in Fig. 14. In contrast, to leeward the vortices
have positive values, indicating counterclockwise flow motions. The vortices are far weaker in
strength to leeward than to windward as is reflected in Fig. 14. In general, the wake region
expands linearly towards the downstream, which is consistent with the experimental results
(Peng et al., 2016). It can be seen that the z-vorticity at the tower is much weaker in strength
than on the blade mid-span plane. The vortices are shed both to windward and to leeward of
the tower. Due to its non-rotational property, the tower’s wake is approximately symmetrical.
Furthermore, the declining strength of the vortices downstream indicates wake recovery.
Figure 16 presents both the time-averaged and the transient stream-wise velocity contours on
the y–z planes downstream. The blades are included in Fig. 16 as a reference for the analysis of
wake development. The time-averaged velocity contours at x ¼ 2D and 6D are shown in Figs.
16(a) and 16(c), respectively, and their transient counterparts are presented in Figs. 16(b) and 16(d),
respectively. Comparisons between the time-averaged and the transient velocity contours suggest
that the wake at a given downstream distance approximately preserves the same pattern. This

FIG. 15. z-vorticity on the blade mid-span plane and at the tower at h ¼ 67 (units, 1/s).
053305-14 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

FIG. 16. Stream-wise velocity contours on the y–z plane: (a) time-averaged, x ¼ 2D; (b) transient (h ¼ 67 ), x ¼ 2D; (c)
time-averaged, x ¼ 6D; and (d) transient (h ¼ 67 ), x ¼ 6D (viewed along the x-axis) (units, m/s).

interesting phenomenon indicates that the wake flow characteristics are statistically stable, and
hence the time-averaged results are representative of the wake (Lam and Peng, 2017a).
The wake asymmetry of the turbine against the x–z plane is noticeably visualized compared
with the symmetrical wake of the tower. The maximum velocity deficit region migrates towards
windward echoing the results in the literature (Li et al., 2017). Conversely, the wake of the tur-
bine, especially at a large downstream distance with sufficient flow mixing, is approximately
symmetrical against the blade mid-span plane. Moreover, the wake expands and recovers
toward downstream due to entrainment and turbulent flow mixing. Lateral wake expansion is
pronounced in comparison with vertical expansion at large downstream distances, e.g., x ¼ 6D.
The velocity deficit region migrates towards windward with the increase in the downstream dis-
tance, whereas the symmetry plane remains at the blade mid-span level. The wake at the top
and bottom blade tips steadily evolves into two separate zones as the downstream distance
increases. As revealed by the wake measurements (Lam and Peng, 2017b), these two flow
regions correspond to the CR vortical motions.
The effect of the BSR on the stream-wise velocity along the y direction at z ¼ H/2 in the
wake was examined further as shown in Fig. 17. The wake velocities were examined at three
BSRs: (1) k ¼ 0.25, (2) k ¼ 0.75, and (3) k ¼ 1.25 (rated). In general, the stream-wise velocities
have similar patterns at different BSRs. U/U0 demonstrates the most severe deficits at x ¼ 2D,
after which U/U0 steadily recovers toward downstream. It can also be seen that there are, in
053305-15 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

FIG. 17. Stream-wise velocities along the y direction at different BSRs.

general, especially at a large BSR, two deficit peaks in the U/U0 profile, e.g., Fig. 17(b). This
phenomenon is assumed to be induced by vortex shedding from the blade spans toward both
windward and leeward.
The effect of the BSR on U/U0 was analyzed further. At x ¼ 2D, U/U0 demonstrates the
largest deficits at the rated BSR, i.e., k ¼ 1.25, in comparison with the smallest deficits at
k ¼ 0.25. This suggests that in the near wake, a higher BSR will cause a larger velocity deficit.
The higher power production and larger blockage effect (Chen and Liou, 2011) at a higher
BSR contributed to the larger velocity deficit. It is interesting that U/U0 at the rated BSR recov-
ers faster than that at k ¼ 0.75 despite demonstrating larger deficits. The faster wake recovery at
the rated BSR is assumed to be caused by stronger entrainment. Furthermore, the U/U0 profiles
at the rated BSR demonstrate outstandingly contracted shapes compared with the other profiles,
the expansion rates of which are quite close. This feature is beneficial to the placement of mul-
tiple VAWTs in a wind farm scenario. The Magnus effect (Van Dyke, 1982) is assumed to
account for the contracted U/U0 profiles at the rated BSR. The strong Kutta-Joukowski lift
(Clancy, 1975) induced by the Magnus effect helps to push small-sized shed vortices to leeward
toward windward, and hence causes the wake to contract.
The wake asymmetry is accounted for by two main effects. First, the far stronger vortex
shedding windward than leeward due to the dynamic stall to windward, as shown in Figs. 11(a)
and 14, contributes to the wake asymmetry. Second, the Magnus effect contributes to the lateral
transportation of the wake flows windward (Clancy, 1975). As discussed, the VR structures, CR
motions, and asymmetry are the signature characteristics of the VAWT’s wake. The vortices
shed from the blade tips and spans merge and evolve to form the periodical VR structures and
CR motions. It is assumed that the VR structures and CR motions contribute to the entrainment,
the flow mixing, and hence the fast wake recovery.

V. CONCLUSIONS
In this study, blade and wake aerodynamics of a cambered-bladed VAWT were systemati-
cally investigated using the full-scale 3-D LES modeling, of which standard procedures were
053305-16 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

given as guidelines. The algebraic WM-LES was used to model turbulence. The computational
domain was set to be the same as the wind tunnel. The boundary conditions were specified
according to the experimental settings. Independent studies of CFD parameters were imple-
mented to ensure calculation accuracy. Comparisons between the LES predictions and the mea-
sured data were made, with close agreement obtained.
A semi-empirical method was proposed to estimate the minimum revolution number, NR,
required to reach convergence of CFD calculations. CFD predictions showed that the VAWT with
a thin airfoil tended to encounter abrupt changes of CL at low BSRs. Hence, a relatively thick air-
foil is recommended for VAWTs to reduce fatigue loading on blades. At high BSRs, the occur-
rence of a dynamic stall was delayed with more power produced. It is believed that urban winds
are beneficial to the self-starting and operation of VAWTs in accordance with the torque pattern.
Comparisons between the time-averaged and the transient velocities indicated a statistically
stable wake. The underlying causes were addressed with regard to wake asymmetry in the hori-
zontal direction. First, the occurrence of a deep dynamic stall windward induced more vortex
shedding. Second, the Magnus force transported the wake flows toward windward. The strongest
Magnus force at the rated BSR was attributed to the exceptional wake contraction. The VR struc-
tures, CR vortical motions, and wake asymmetry were analyzed and concluded as the signature
characteristics of the VAWT’s wake. Moreover, the VR structures and CR motions are believed
to contribute to the fast wake recovery, beneficial to the placement of multiple VAWTs.

ACKNOWLEDGMENTS
This work was supported by a grant from the Research Grants Council of the Hong Kong
Special Administrative Region, China (Project No. CityU 11242716). The authors would like to
take this chance to express their appreciation for the financial assistance.

AbdelSalam, A. M. and Ramalingam, V., “Wake prediction of horizontal-axis wind turbine using full-rotor modeling,”
J. Wind Eng. Ind. Aerodyn. 124, 7–19 (2014).
Abkar, M. and Dabiri, J. O., “Self-similarity and flow characteristics of vertical-axis wind turbine wakes: An LES study,”
J. Turbul. 18, 373–389 (2017).
Araya1, D. B., Colonius, T., and Dabiri, J. O., “Transition to blu-body dynamics in the wake of vertical-axis wind
turbines,” J. Fluid Mech. 813, 346–381 (2017).
Bachant, P., Goude, A., and Wosnik, M., “Actuator line modeling of vertical-axis turbines,” preprint arXiv:1605.01449 (2018).
Barthelmie, R. J. and Pryor, S. C., “An overview of data for wake model evaluation in the Virtual Wakes Laboratory,”
Appl. Energy 104, 834–844 (2013).
Bedon, G., De Betta, S., and Benini, E., “A computational assessment of the aerodynamic performance of a tilted Darrieus
wind turbine,” J. Wind Eng. Ind. Aerodyn. 145, 263–269 (2015).
Bedon, G., De Betta, S., and Benini, E., “Performance-optimized airfoil for Darrieus wind turbines,” Renewable Energy
94, 328–340 (2016).
Borg, M., Shires, A., and Collu, M., “Offshore floating vertical axis wind turbines, dynamics modelling state of the art. Part
I: Aerodynamics,” Renewable Sustainable Energy Rev. 39, 1214–1225 (2014).
Chakraborty, P., Balachandar, S., and Adrian, R. J., “On the relationships between local vortex identification schemes,”
J. Fluid Mech. 535, 189–214 (2005).
Chen, T. Y. and Liou, L. R., “Blockage corrections in wind tunnel tests of small horizontal-axis wind turbines,” Exp.
Therm. Fluid Sci. 35, 565–569 (2011).
Chen, W. H., Chen, C. Y., Huang, C. Y., and Hwang, C. J., “Power output analysis and optimization of two straight-bladed
vertical-axis wind turbines,” Appl. Energy 185, 223–232 (2017).
Chowdhury, A. M., Akimoto, H., and Hara, Y., “Comparative CFD analysis of vertical axis wind turbine in upright and
tilted configuration,” Renewable Energy 85, 327–337 (2016).
Clancy, L. J., Aerodynamics (John Wiley & Sons, New Jersey, U. S., 1975).
Dabiri, J. O., “Potential order-of-magnitude enhancement of wind farm power density via counter-rotating vertical-axis
wind turbine arrays,” J. Renewable Sustainable Energy 3, 043104 (2011).
El Kasmi, A. and Masson, C., “An extended k–e model for turbulent flow through horizontal-axis wind turbines,” J. Wind
Eng. Ind. Aerodyn. 96, 103–122 (2008).
Elkhoury, M., Kiwata, T., and Aoun, E., “Experimental and numerical investigation of a three-dimensional vertical-axis
wind turbine with variable-pitch,” J. Wind Eng. Ind. Aerodyn. 139, 111–123 (2015).
unzen, H., and Carissimo, B., The COST 732 Best Practice Guideline for CFD Simulation of
Franke, J., Hellsten, A., Schl€
Flows in the Urban Environment (COST Office, 2007).
Howell, R., Qin, N., Edwards, J., and Durrani, N., “Wind tunnel and numerical study of a small vertical axis wind turbine,”
Renewable Energy 35, 412–422 (2010).
Hunt, J. C. R., Wray, A. A., and Moin, P., “Eddies, stream, and convergence zones in turbulent flows,” Center for
Turbulence Research Report No. CTR-S88 (1988), pp. 193–208.
053305-17 Peng, Lam, and Liu J. Renewable Sustainable Energy 10, 053305 (2018)

Islam, M. R., Mekhilef, S., and Saidur, R., “Progress and recent trends of wind energy technology,” Renewable Sustainable
Energy Rev. 21, 456–468 (2013).
Jimenez, A., Crespo, A., Migoya, E., and Garcia, J., “Advances in large-eddy simulation of a wind turbine wake,” J. Phys.:
Conf. Ser. 75, 012041 (2007).
Kjellin, J., B€
ulow, F., Eriksson, S., Deglaire, P., Leijon, M., and Bernhoff, H., “Power coefficient measurement on a 12 kW
straight bladed vertical axis wind turbine,” Renewable Energy 36, 3050–3053 (2011).
Knight, J., “Urban wind power: Breezing into town,” Nature 430, 12–13 (2004).
Lam, H. F. and Peng, H. Y., “Study of wake characteristics of a vertical axis wind turbine by two- and three-dimensional
computational fluid dynamics simulations,” Renewable Energy 90, 386–398 (2016).
Lam, H. F. and Peng, H. Y., “Development of a wake model for a Darrieus-type straight-bladed vertical axis wind turbine
and its application to micro-siting problems,” Renewable Energy 114, 830–842 (2017a).
Lam, H. F. and Peng, H. Y., “Measurements of wake characteristics of co-rotating and counter-rotating twin H-rotor verti-
cal-axis wind turbines,” Energy 131, 13–26 (2017b).
Lam, H. F., Liu, Y. M., Peng, H. Y., Lee, C. F., and Liu, H. J., “Assessment of solidity effect on the power performance of
H-rotor vertical axis wind turbines in turbulent flows,” J. Renewable Sustainable Energy 10, 023304 (2018).
Lan, C. T. and Jan, R., Airplane Aerodynamics and Performance (Design, Analysis and Research Corporation, Kansas,
2003).
Lee, Y. T. and Lim, H. C., “Numerical study of the aerodynamic performance of a 500 W Darrieus-type vertical-axis wind
turbine,” Renewable Energy 83, 407–415 (2015).
Li, Y. and Calisal, S. M., “A discrete vortex method for simulating a stand-alone tidal-current turbine: Modeling and vali-
dation,” J. Offshore Mech. Arct. Eng. 132, 031102-1 (2010a).
Li, Y. and Calisal, S. M., “Three-dimensional effects and arm effects on modeling a vertical axis tidal current turbine,”
Renewable Energy 35, 2325–2334 (2010b).
Li, Y. and Calisal, S. M., “Numerical analysis of the characteristics of vertical axis tidal current turbines,” Renewable
Energy 35, 435–442 (2010c).
Li, Q., Murata, J., Endo, M., Maeda, T., and Kamada, Y., “Experimental and numerical investigation of the effect of turbu-
lent inflow on a horizontal axis wind turbine (part II: Wake characteristics),” Energy 113, 1304–1315 (2016).
Li, Q., Maeda, T., Kamada, Y., Ogasawara, T., Nakai, A., and Kasuya, T., “Investigation of power performance and wake
on a straight-bladed vertical axis wind turbine with field experiments,” Energy 141, 1113–1123 (2017).
McLaren, K. W., “A numerical and experimental study of unsteady loading of high-solidity vertical axis wind turbines,”
Ph.D. thesis (McMaster University, Canada, 2011).
Menter, F. R., “Two-equation eddy-viscosity turbulence models for engineering applications,” AIAA J. 32, 1598–1605
(1994).
Menter, F. R., Langtry, R. B., Likki, S. R., Suzen, Y. B., Huang, P. G., and Volker, S., “A correlation based transition
model using local variables, Part 1—Model formulation,” J. Turbomach., ASME 128, 413–422 (2006).
Mo, J. O., Choudhry, A., Arjomandi, M., and Lee, Y. H., “Large eddy simulation of the wind turbine wake characteristics
in the numerical wind tunnel model,” J. Wind Eng. Ind. Aerodyn. 112, 11–24 (2013).
Peng, H. Y. and Lam, H. F., “Turbulence effects on the wake characteristics and aerodynamic performance of a straight-
bladed VAWT by wind tunnel tests and large eddy simulations,” Energy 109, 557–568 (2016).
Peng, H. Y., Lam, H. F., and Lee, C. F., “Investigation into the wake aerodynamics of a five-straight-bladed vertical axis
wind turbine by wind tunnel tests,” J. Wind Eng. Ind. Aerodyn. 155, 23–35 (2016).
Posa, A., Parker, C. M., Leftwich, M. C., and Balaras, E., “Wake structure of a single vertical axis wind turbine,” Int. J.
Heat Fluid Flow 61, 75–84 (2016).
Sanderse, B., “Aerodynamics of wind turbine wakes” Report No. ECN-e–09-016 (Energy Research Centre of the
Netherlands, 2009).
Shamsoddin, S. and Porte-Agel, F., “A large-Eddy simulation study of vertical axiswind turbine wakes in the atmospheric
boundary layer,” Energies 9, 366 (2016).
Shi, L., Riziotis, V. A., Voutsinas, S. G., and Wang, J., “A consistent vortex model for the aerodynamic analysis of vertical
axis wind turbines,” J. Wind Eng. Ind. Aerodyn. 135, 57–69 (2014).
Shur, M. L., Spalart, P. R., Strelets, M. K., and Travin, A. K., “A hybrid RANS-LES approach with delayed-DES and wall-
modeled LES capabilities,” Int. J. Heat Fluid Flow 29, 1638–1649 (2008).
Sim~ao Ferreira, C. J., “The near wake of the VAWT: 2D and 3D views of the VAWT aerodynamics,” Ph.D. thesis (Delft
University of Technology, Netherlands, 2009).
Tescione, G., Ragni, D., He, C., Sim~ao Ferreira, C. J., and van Bussel, G. J. W., “Near wake flow analysis of a vertical axis
wind turbine by stereoscopic particle image velocimetry,” Renewable Energy 70, 47–61 (2014).
Tjiu, W., Marnoto, T., Mat, S., Ruslan, M. H., and Sopian, K., “Darrieus vertical axis wind turbine for power generation II:
Challenges in HAWT and the opportunity of multi-megawatt Darrieus VAWT development,” Renewable Energy 75,
560–571 (2015).
Vallverdu, D., “Study on vertical-axis wind turbines using streamtube and dynamic stall models,” thesis (Illinois Institute
of Technology, US, 2014).
Van Dyke, M., An Album of Fluid Motion (Stanford University, Stanford, CA, 1982).
Vermeer, L. J., Sorensen, J. N., and Crespo, A., “Wind turbine wake aerodynamics,” Prog. Aerosp. Sci. 39, 467–510
(2003).
Versteeg, H. K. and Malalasekera, W., An Introduction to Computational Fluid Dynamics: The Finite Volume Method, 2nd
ed. (Pearson Education Limited, England, 2007).
Wong, K. H., Chong, W. T., Poh, S. C., Shiah, Y. C., Sukiman, N. L., and Wang, C. T., “3D CFD simulation and paramet-
ric study of a flat plate deflector for vertical axis wind turbine,” Renewable Energy 129, 32–55 (2018).
Xie, S. and Archer, C., “Self-similarity and turbulence characteristics of wind turbine wakes via large-eddy simulation,”
Wind Energy 18, 1815–1838 (2015).
Yakhot, V. and Orszag, S. A., “Renormalization group analysis of turbulence I. Basic theory,” J. Sci. Comput. 1, 3–51
(1986).

Anda mungkin juga menyukai