You are on page 1of 12

Renewable Energy 121 (2018) 632e643

Contents lists available at ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

Dynamic modeling of vortex induced vibration wind turbines


A. Chizfahm, E. Azadi Yazdi*, M. Eghtesad 1
School of Mechanical Engineering, Shiraz University, Shiraz, Iran

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

Article history: This paper studies the dynamic modeling of four configurations of vortex-induced vibrations of a
Received 26 October 2017 bladeless wind turbine (BWT). The BWTs consist of a bluff body mounted on a flexible structure in the
Received in revised form flow field. The shape of the bluff body and its mounting structure are different among the proposed
18 December 2017
BWTs. The Euler-Bernoulli beam theory and the Galerkin procedure are used to derive a nonlinear
Accepted 13 January 2018
distributed-parameter model for the BWTs under a fluctuating lift force due to periodically shedding
Available online 16 January 2018
vortices. The derived dynamic model is validated through comparison with a 3D CFD-FEM numerical
simulation. The effects of the wind speed on the induced lift force, turbine deflection, and generated
Keywords:
Bladeless wind turbine
power of four BWTs are investigated. It is verified that the amplitude of the vibrations of the BWT in-
Vortex induced vibration creases significantly when the vortex shedding is synchronized with the structural oscillations. The
Dynamic modeling results show that, while conic BWTs have a higher performance at post-synchronization region (i.e. high
Lock-in phenomenon wind speeds), the right circular cylinder BWTs exhibits a better performance at pre-synchronization
region (i.e. low wind speeds).
© 2018 Elsevier Ltd. All rights reserved.

1. Introduction the required support structure. In contrast, recently a BWT is


studied in Ref. [13] that eliminates the need of the linear support by
Submerged bodies may experience periodically shedding mounting the cylinder on the tip of a flexible beam.
vortices in the flow fields close to their surface. The periodic The proposed VIV bladeless wind turbine (BWT) consists of a
shedding vortices in the wind or water currents result in fluctua- bluff body (often a relatively long cylinder) mounted on a flexible
tions in the pressures around the submerged body which yields to structure in the flow field. The absence of rigidly moving parts is the
vortex-induced vibrations (VIV) in flexible structures. The struc- main difference between VIV bladeless and conventional wind
tural VIV is often a detrimental phenomenon due to the fatigue turbines. The flow power transmitted to the VIV bladeless wind
damage consequences (for example in power transmission lines, turbine is considerable when the lock-in phenomenon occurs, i.e.
towing cables, and mooring lines); however, it can be beneficial in the frequency of vortex shedding is close to the structural natural
wind power [1e3] and hydropower utilities [4,5]. The practical frequency. Hence, to study and improve the energy production rate
significance of VIV energy in renewable energy and energy har- of VIV bladeless wind turbines, the effects of the turbine design
vesting has led to a large number of fundamental studies, for parameters on the lock-in phenomenon should be carefully
instance VIV of circular cylinders [6,7], flutter of airfoils [8,9], and analyzed through dynamic modeling.
galloping of sliding structures [10,11]. Most of the previous research An accurate model of VIV bladeless wind turbines requires a
studies have considered a VIV harvester similar to the so-called model for the forced-vibrations of the structure subjected to fluid
VIVACE (Vortex Induced Vibration Aquatic Clean Energy) device flow forces attained by the NaviereStokes equations in the pres-
presented in Ref. [12]. The VIVACE is a spring mounted cylinder that ence of moving boundaries [14]. However, due to computational
exhibits transverse oscillations in the flow field. VIVACE and similar complexity, the numerical model is not useful for structural design
devices cannot be up-scaled easily due to the relatively high cost of and controller synthesis purposes. Moreover, computational diffi-
culties arise for simulating three-dimensional (3D) flexible VIV
wind turbines that often have large aspect ratios [15].
Alternatively, the computational complexity associated with the
* Corresponding author.
fluid flow modeling can be reduced by considering a two-
E-mail address: ehsanazadi@shirazu.ac.ir (E.A. Yazdi).
1
Current address: Department of Mechanical Engineering, The University of dimensional (2D) flow model that exerts a crosswise flow-
British Columbia, Vancouver, Canada. induced force on each slice of the structure. In this approach the

https://doi.org/10.1016/j.renene.2018.01.038
0960-1481/© 2018 Elsevier Ltd. All rights reserved.
A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643 633

sophisticated CFD calculations are replaced by reduced order


models [16], semiempirical models [17], or analytical models [18].
In general, the traditional semiempirical models are the most
favorable, because they have neither the conservative simplifica-
tions of the analytical models nor the computational complexity
associated with the reduced order models. The most proper semi-
empirical model of the crosswise flow-induced force follows the
idea of a wake oscillator with a single flow variable that models the
harmonic nature of the vortex shedding [19]. The flow variable
satisfies van der Pol or Rayleigh equation which results in a stable
and nearly harmonic oscillation [20]. The wake oscillator semi-
empirical model has been shown to have the potential to demon-
strate basic features of VIV, such as limit cycle oscillations and lock-
in phenomenon [21,22].
Moreover, the modeling complexity of the VIV bladeless wind
turbine caused by the structural forced-vibration model, which
leads to partial differential equations (PDEs) of motion, can be
reduced by various methods. The PDE can be reduced to an ordi-
nary differential equation (ODE) via Galerkin approach [14]. Several
other methods have also been proposed to simplify the structural
vibrations modeling such as the method of lumped parameters
[23,24], Rayleigh-Ritz method [25], assumed modes method [26],
collocation method [27] and least squares method [28].
The purpose of this paper is to compare and contrast various
aspects of four different concepts for VIV bladeless wind turbines.
The proposed BWTs are different in the shape of their bluff body
and their mounting structure. Specifically, we aim to determine the
power levels that can be transmitted to BWTs and variations of
these levels with the free-stream air speed. To this end, simplified
analytical dynamic models have been derived for the transverse
displacement degree of freedom (DOF) of BWTs. The dynamic
models have been used to analyze the lock-in phenomenon and
investigate its dependence on the free-stream velocity for the
proposed BWTs.
The paper is organized as follows. Section 2 presents the dy-
namic models of the proposed BWTs. The wake oscillator semi-
empirical model is used to derive a simplified expression for the Fig. 1. The bladeless wind turbines; (a) BWT1 (b) BWT2 (c) BWT3 (d) BWT4.
crosswise flow-induced force on the BWTs. The wake oscillator
model has been combined with the transverse vibrations model of
the structure for each BWT. The proposed dynamic models have
dynamic models of the proposed BWTs are determined in this
been validated through comparison with 3D CFD and FEM nu-
section. Note that the computational complexity associated with
merical simulations. Then, we determine the power level that can
the CFD and FEM models are avoided by deriving simplified dy-
be transmitted to BWTs through numerical simulations of the dy-
namic models. Also note that a power take-off unit should be
namic models. We also study variations of the power level with the
installed near the bottom end of the BWT which is not considered
air speed. Finally, conclusions are presented in Section 4.
in this paper for the sake of modeling simplicity.

2. Dynamic modeling of VIV bladeless wind turbines

The VIV bladeless wind turbines consist of a relatively long 2.1. Aerodynamic force model
cylinder that is either flexible or mounted on a flexible structure
exposed to air flow. We consider four BWTs that differ in the shape The wake oscillator semiempirical model is used to obtain a
of the cylinder and its mounting structure (Fig. 1): simplified expression for the crosswise flow-induced force on the
BWTs. The nonlinear wake oscillator model was proposed for the
 BWT1: A right circular flexible cylinder first time by Skop and Griffin [21]. The model has been further
 BWT2: A conic flexible cylinder developed by Skop and Balasubramanian [14] to be able to accu-
 BWT3: A right circular rigid cylinder mounted on a flexible rately capture the asymptotic, self-limiting structural response for
beam flexible cylinders. In the model a stall term is incorporated to
 BWT4: A conic rigid cylinder mounted on a flexible beam describe the dependence of the cross-flow force on the transverse
velocity of the structure. The velocity coupling enables the model to
In BWTs the periodic shedding vortices in the air flow along the capture the lock-in phenomenon. Moreover, the displacement
z-direction induce vibrations in the y-direction. The amplitude of coupling and the acceleration coupling were also introduced in [20]
the vibrations (and consequently the flow power transmitted to the to achieve a more accurate model of the lock-in phenomenon.
BWT) is considerable when the frequency of vortex shedding is Based on the wake oscillator semiempirical model, the fluctuating
close to the structural natural frequency that is known as lock-in lift coefficient CL(x,t) at the time t on a slice of the BWT located at a
phenomenon. Hence, to study the lock-in phenomenon, the distance x from the origin is defined as:
634 A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643

2a _  
CL ðx; tÞ ¼ Q ðx; tÞ  Yðx; tÞ (1) v4 v2 Yðx; tÞ 1 2 2a _
Dus ðEIYðx; tÞÞ þ r ¼ r V DQ ðx; tÞ  Yðx; tÞ (6)
vx4 s
vt 2 2 us
where Q ðx; tÞ is the excitation component of the fluctuating lift
_ tÞ is the transverse velocity of the slice of the
coefficient and Yðx;   _ tÞ
Yðx;
BWT. The second term on the right side is called the stall term, Q€ ðx; tÞ  us G CL0
2
 4Q 2 ðx; tÞ Q_ ðx; tÞ þ u2s Q ðx; tÞ ¼ us F
where a is an empirical constant, us is the vortex shedding fre- D
quency that is defined as us ¼ 2pS VD, S is the Strouhal number [14], (7)
V is the wind velocity and D is the mean diameter of the slice of the A simplified form of the partial differential equation (PDE) of
 
motion can be developed using the Galerkin method, where the
BWT. In this model, the stall parameter, D2uas , is defined to limit
solution to the PDE is assumed to be of the form:
the fluctuating lift coefficient for large deflections based on the
observations of Triantafyllou [29]. The van der Pol equation is X
satisfied by using the excitation component of the fluctuating lift Yðx; tÞ ¼ fi ðxÞyi ðtÞ (8)
coefficient Q(x,t) as:
where fi ðxÞ and yi ðtÞ are called the mode shape and the modal
  _ tÞ
Yðx; response factor, respectively. Furthermore, based on the results of
Q€ ðx; tÞ  us G CL0
2
 4Q 2 ðx; tÞ Q_ ðx; tÞ þ u2s Q ðx; tÞ ¼ us F Skop and Balasubramanian [14], mode shapes of Q ðx; tÞ can be
D
approximated to be the same as the mode shapes of Yðx; tÞ. There
(2)
are two possible explanations for the above statement. First, the
response of the BWT is dominated by the lock-in phenomenon.
where, CL0, G and F are empirical parameters [14]. For stationary
Hence, only mode shapes of the fluctuating lift coefficient which
cylinders, the right-hand side of the (2) is equal to zero, so that the
are close to structural mode shapes have a considerable contribu-
above equation has a self-excited, self-limited solution given by,
tion to the response of the system near lock-in region. Second, if van
Q ¼ CL0 sinðus tÞ (3) der Pol equation (1) is linearized, it can be easily shown that the
mode shapes of Q ðx; tÞ exactly match the mode shapes of Yðx; tÞ
This means that the fluctuating lift coefficient on stationary [14]. Therefore, forus zun;i, Q ðx; tÞ may be assumed as:
cylinder has the amplitude equal to CL0, where based on Protos et al.
[30] CL0 ≪1 for circular sections. X
Q ðx; tÞ ¼ fi ðxÞ qi ðtÞ (9)

where qi ðtÞ is the modal response factor for the fluctuating lift
2.2. Structural VIV models excitation component. To comply with the boundary conditions (5),
the ith mode shape of the clamped-free cantilever is chosen as,
Since the structure of the proposed BWTs differs, the equations
of motion have been derived separately for each BWT. In all BWT
configurations, the wind flow is assumed to be uniformly steady ðcos bi l þ cosh bi lÞ
fi ðxÞ ¼ ðcos bi x  cosh bi xÞ þ
with a constant velocity and the boundary layer effect on the ðsin bi l þ sinh bi lÞ (10)
clamped end is neglected. ðsin bi x  sinh bi xÞ

The corresponding natural frequency is given by:

2.2.1. BWT1: A right circular flexible cylinder sffiffiffiffiffiffiffiffiffiffi


In the BWT1 configuration presented in Fig. 1(a), the turbine is a EI
uniform right cylinder that can be modeled as a clamped-free uni ¼ ðbi lÞ 2
(11)
ms l4
Euler-Bernoulli beam exposed to air flow. Considering the distrib-
uted aerodynamic force expression on the beam element as The largest amount of wind energy transmitted to the BWT is in
1 rV 2 DC ðx; tÞ, the equation of motion is expressed by Euler-
2 L the first mode, where we have ðb1 lÞ2 ¼ 3:52 [23]. Hence, a dynamic
Bernoulli equation as: model that considers only the first mode has an acceptable error in
the prediction of the energy production rate of the BWT. By
v4 v2 Yðx; tÞ 1 2 considering only the first mode shape, substituting (8) and (9) into
ðEIYðx; tÞÞ þ rs ¼ rV DCL ðx; tÞ (4) (6) and (7) and multiplying the mode shape of the system by both
vx4 vt 2 2
sides of the equation and integrating on x, the reduced-order
where E is the Young modulus of elasticity of the cylinder, I is the equations (ODEs) of the motion are written as:
second moment of area of the cross section, rs is mass per unit
length of the cylinder, and r is the air density. Let us consider the
ZL
boundary conditions of the clamped-free beam as,
EIyðtÞ fðxÞfð4Þ ðxÞdx
vY 0
Yð0; tÞ ¼ 0; ð0; tÞ ¼ 0
vx ZL   ZL
(5) 1 2a
v2 Y v3 Y þ rs yðtÞ
€ f ðxÞdx ¼ r V DqðtÞ 
2 2 _
yðtÞ f2 ðxÞdx
ðL; tÞ ¼ 0; ðL; tÞ ¼ 0 2 us
2 3
vx vx 0 0
(12)
Substituting (1) into (4), the two coupled dynamic equations of
motion are obtained as,
A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643 635

resultant point moment on the tip of the stand as shown in Fig. 2.


  ZL Hence, using (8), the kinetic energy and the potential energy of the
€  us GCL0
qðtÞ 2 _
qðtÞ þ u2s qðtÞ f2 ðxÞ dx system are given by,
0

  ZL ZLs ZLs
XX
þ 4us Gq2 ðtÞqðtÞ
_ f4 ðxÞdx T¼
1 2
Y_ ðx; tÞmdx ¼
1
y_ i ðtÞy_ j ðtÞ fi ðxÞfj ðxÞmdx
0
2 2 i j
0 0
  ZL 1X
_
yðtÞ ¼ Mi y_2i ðtÞ
¼ us F f2 ðxÞdx (13) 2
D
0 (17)

where, f0 and fðnÞ represent df=dx and dn f=dxn respectively, and


for simplicity of notation fðxÞ :¼ f1 ðxÞ, yðtÞ :¼ y1 ðtÞ, and ZLs Z Ls
qðtÞ :¼ q1 ðtÞ. The coupled equations of motion are linear with 1 00 2 1XX 00 00
U¼ EI Y ðx; tÞdx ¼ yi ðtÞyj ðtÞ EIf i ðxÞf j ðxÞdx
respect to yðtÞ and its derivatives, but nonlinear with respect to qðtÞ 2 2 i j
0 0
due to the appearance of the term q2 ðtÞqðtÞ in (13).
1X
¼ Ki y2i ðtÞ
2
2.2.2. BWT2: A conic flexible cylinder (18)
Considering BWT2 (Fig. 1(b)), the turbine is a tapered conical
flexible cylinder attached to the ground. Similar to the previous Z Ls
section, the equation of motion is derived based on the Euler- where Mi ¼ f2i ðxÞms dx is the generalized mass and
0
Bernoulli beam theory as: Z Ls
" # Ki ¼ EI½fi "ðxÞ2 dx is the generalized stiffness. The work done by
2 2 2 0
v v v Yðx; tÞ 1 2
EIðxÞ 2 Yðx; tÞ þ rs AðxÞ ¼ rV DðxÞCL ðx; tÞ the distributed fluctuating lift force and the mass and inertia of the
vx2 vx vt 2 2 mast is given by,
(14)
The following coupled ODEs of motion are derived for BWT2 dW ¼ FM dYðLs ; tÞ þ MM dqðLs ; tÞ
(Please see Appendix for the proof), X  X 
¼ FM fi ðLs Þdyi ðtÞ þ MM f0 i ðLs Þdyi ðtÞ (19)
ZL  00   0  
00
EyðtÞ fðxÞ I ðxÞf ðxÞ þ 2I ðxÞfð3Þ ðxÞ þ ðIðxÞfð4Þ ðxÞ Þ dx FM and MM are respectively the force and moment exerted from the
0
mast on the tip of the stand, defined by,

ZL
þ rs yðtÞ
€ AðxÞf2 ðxÞdx LZ
s þLm
1 2
0 FM ¼ rV DCL ðx; tÞdx  mm y€m (20)
2 3 2
ZL ZL Ls
1 24 2a
¼ rV q DðxÞf2 ðxÞdx  _
yðtÞ f2 ðxÞdx 5
2 us
0 0
(15)

 Z L
€  us GCL0
qðtÞ 2 _
qðtÞ þ u2s qðtÞ f2 ðxÞdx
0

n oZL ZL 2
f ðxÞ
þ us G4q ðtÞqðtÞ
2 _ f ðxÞdx ¼ us F yðtÞ
4 _ dx (16)
DðxÞ
0 0

2.2.3. BWT3: A right circular rigid cylinder mounted on a flexible


beam
The third configuration, BWT3 (Fig. 1(c)), consists of a flexible
stand and a rigid mast. The flexible stand enables the BWT to
vibrate, while the mast causes the vortex shedding and generates
an oscillatory aerodynamic force. In this configuration, the aero-
dynamic force on the stand can be neglected because it is relatively
narrow. Hence, to study the lock-in phenomenon, the stand can be
modeled as a clamped-free Euler-Bernoulli beam. The effect of
mass and inertia of the mast and the distributed fluctuating lift
force on the mast can be replaced by a resultant point force and a Fig. 2. BWT3 and BWT4 (a) The schematic model (b) Free-body diagram.
636 A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643

LZ
s þLm
1 2
MM ¼ rV DCL ðx; tÞðx  Ls Þdx  mm Lm y€m  Im €qm (21)
2
Ls

where for the center of mass of the mast;, Lm is the distance to the
tip of the stand, and I m is the moment of inertia about z-axis. Also
ym and qm are written as

X
qm ¼ Y 0 ðx; tÞjx¼Ls ¼ f0 i ðLs Þyi ðtÞ (22)

Lm X Lm X 0
ym zYðx; tÞjx¼Ls þ qm ¼ fi ðLs Þyi ðtÞ þ f i ðLs Þyi ðtÞ
2 2
X Lm

¼ fi ðLs Þ þ f0 i ðLs Þ yi ðtÞ
2
(23)
Now, substituting (17)e(19) into Lagrange's equation
0 1
d @ vT A  vT þ vU ¼ Q ), the ODEs of motion are given as,
(dt vq_ i vqi vqi i

Mi y€i ðtÞ þ Ki yi ðtÞ ¼ FM fi ðLs Þ þ MM f0 i ðLs Þ (24)


Since the mast is assumed to be a rigid body, the velocity of the Fig. 3. (a) Fluid flow problem configuration and boundary conditions. (b) Details of the
mast (that is required for the lift coefficient equation (1)) is written constructed mesh.
as,

X "
_ tÞzYðx;
Yðx; _ tÞ þ ðx  Ls Þq_ m ¼ fi ðLs Þy_i ðtÞ þ ðx  Ls Þ ZLs 
Lm

x¼Ls
X M¼ f ðxÞms dx þ mm f2 ðLs Þ þ Lm þ
2
ff0 ðLs Þ
0
f i ðLs Þy_i ðtÞ 2
0
! # (29)
Ls  x  Ls þ Lm Lm Lm Im 02
þ þ f ðLs Þ
(25) 2 mm
As it has been explained earlier, the most significant contribu-
tors among the mode shapes of the excitation component of the ZLs h 00 i2
fluctuating lift coefficient are the ones that match the structural K¼ EI f ðxÞ dx (30)
mode shapes, therefore,
0

Q ðx; tÞ ¼ qi ðtÞ½fi ðLs Þ þ ðx  Ls Þf0 i ðLs Þ (26) LZ


s þLm
h 
1
Considering only the first mode of the vibrations of the stand, G ¼ rV 2 D f2 ðLs Þ þ ðx  Ls Þ 2fðLs Þf0 ðLs Þ
2
from (1) and (20)e(26) the reduced-order model is obtained as: Ls
i
2
þ 2ðx  Ls Þf0 ðLs Þ dx (31)
 
2a
€ þ
M yðtÞ _ þ KyðtÞ ¼ GqðtÞ
GyðtÞ (27)
Dus hðxÞ ¼ fðLs Þ þ ðx  Ls Þf0 ðLs Þ (32)

Z Ls þLm 3

    hðxÞ dx
Ls
€  us GCL0
qðtÞ 2 _
qðtÞ þ u2s qðtÞ þ 4us Gq2 ðtÞqðtÞ
_ Z Ls þLm 2.2.4. BWT4: A conic rigid cylinder mounted on a flexible beam
hðxÞ dx The configuration of BWT4 is similar to BWT3 except for the
Ls shape of the mast which is changed to a conic shape (Fig. 1(d)). The
 
_
yðtÞ conical shape of the mast enhances the lock-in phenomenon over a
¼ us F (28)
D wider range of wind speeds. To derive the model, by considering a
linear variation in diameter in (27), the reduced order equations of
where M, K, G and hðxÞ are defined as, motion for the first mode of the vibrations are obtained as,
A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643 637

  the CFD model in the commercial ANSYS Fluent package through a


2a User-Defined Function (UDF).
€ þ
M yðtÞ _ þ KyðtÞ ¼ LqðtÞ
LyðtÞ (33)
Dus The CFD model considers a non-stationary laminar incom-
pressible flow. The BWT4 is placed in a rectangular cuboid
  LZs þLm computational domain with downstream and upstream boundaries
€  us GCL0
qðtÞ 2 _
qðtÞ þ u2s qðtÞ hðxÞ dx extended to Ld
Dtop ¼ 160 and Lu
Dtop ¼ 80, respectively. The lateral
Ls boundaries are considered at the distance W ¼ 120Dtop (Fig. 3 (a)).
In order to avoid excessive simulation error due to the sharp
  LZs þLm
þ 4us Gq2 ðtÞqðtÞ
_ hðxÞ3 dx pressure gradients in cells near cylinder boundary, a moving mesh
technique is used that produces high-quality structured cells
Ls
around the BWT (Fig. 3 (a)). As BWT oscillates, the moving/
LZ
s þLm deforming mesh function regenerates the grid cells at each itera-
hðxÞ
¼ us F yðtÞ
_ dx (34) tion. The computational domain is partitioned into three distinct
DðxÞ
Ls blocks with map type quad-cells in the center cylindrical block
enclosing cylinder, unstructured T-grid-type tetrahedral/hybrid
where L is defined as, cells in the surrounding rectangular block and map type quad-cells
for the farther sides. The aspect ratio of the quad-cells, the grid size,
LZ
s þLm and the time step size are chosen fine enough to achieve mesh
1 2 h 
L¼ rV DðxÞ f2 ðLs Þ þ ðx  Ls Þ fðLs Þf0 ðLs Þ independence and ensure acceptable errors in unsteady flow sim-
2 ulations [31]. The resultant mesh has about 4 million cells for the
Ls
i BWT4 dimensions given in Fig. 3.
2
þ ðx  Ls Þf0 ðLs Þ dx (35) In the CFD model, the first-order pressure-based implicit solver
is used for transient formulation and momentum equations that
lead to the aerodynamic forces. Furthermore, to improve the ac-
curacy of the second order time integration, the non-iterative
3. Results and discussions fractional step method is used for the velocity-pressure coupling.
The structural vibration of BWT4 is determined by a finite
In the previous section, reduced order models of the vortex- element model in ANSYS which is linked with the CFD model
induced vibrations of the four proposed BWTs were developed. In through a UDF. The transient-structural model considers the lift
this section, first to evaluate the accuracy of the reduced order force as the only external force causing the vibrations of BWT4. The
models, their results are compared with the results of a 3D CFD- entire length of BWT4 (both the stand and the mast) is considered
FEM numerical simulation. Then, we aim to determine the power to be flexible. The computational domain consists of 50,000 un-
level that can be transmitted to BWTs and to study variations of the structured tetrahedral cells. The size of the cells and the time-step
level with the air velocity through numerical simulations of the are chosen fine enough to guarantee an adequate accuracy in
reduced order models. transient response.
The CFD-FEM modeling procedure for time instance t is outlined
3.1. Model validation as follows:

A 3D CFD-FEM simulation has been performed to validate the 1. Pressure and velocity fields are computed by the CFD model.
reduced order models. The configuration of BWT4 is chosen for 2. Aerodynamic load CL ðx; tÞ is determined by the CFD model.
validation purpose because it combines the main features of the 3. Aerodynamic load is applied on BWT4 in the FEM model.
fluid-solid interaction of the other configurations. The 3D CFD-FEM 4. Transverse motion yðx; tÞ is computed by the FEM model.
is performed by linking the finite element model of the structure to

Fig. 4. Results of the CFD-FEM simulation and the reduced-order model of BWT4.
638 A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643

Table 1
Comparison of the reduced order model and the CFD-FEM model.

Mean amplitude of Mean frequency of Mean amplitude of lift Mean frequency of lift RMSE of RMSE of lift
displacement (m) displacement (Hz) force (N) force (Hz) displacement force

Reduced order 0.297 1.98 14.60 1.99 1.13% 3.27%


model
3D CFD-FEM 0.300 2.00 14.55 2.00
model

Table 2 (MaxfCL ðx; tÞg) for various wind speeds close to the lock-in region
Dimensions and properties of the BWTs. x;t
versus the non-dimensional frequency U for BWT1 and BWT2. In
Physical Properties BWT1 BWT2 BWT3 BWT4
both turbines, the lift force increases dramatically as the non-
Lengths (Lm, Ls) (m) (4,0) (4,0) (4, 0.15) (4, 0.15) dimensional frequency approaches 1. Also in both turbines, at
Mean diameter of the mast D (m) 0.18 0.225 0.375 0.375
U < 1, the lift coefficient (CL ) diagram has an upward trend with
Mast thickness t (m) 0.001 0.001 0.001 0.001
Stand diameter Ds (mm) e e 0.018 0.018 increasing wind speed (i.e. increasing U) due to the intensification
Mass density (rm,rs) (103 kg/m3) (1.04,0) (1.04,0) (1.04, 8) (1.04, 8) of the vortices. As the non-dimensional frequency exceeds U ¼ 1
Mass (mm,ms) (kg) (2.6,0) (4.4,0) (5.21,0.39) (5.27,0.39) (due to high wind speed), although the vortices are intensified
Young's modulus Es (GPa) 2.2 2.2 2.2 2.2 further, the slope of the lift coefficient diagram reduces signifi-
Taper ratio c e 2 e 2
cantly and becomes negative due to a mismatch between the vortex
shedding frequency and the natural frequency.
At U < 1, BWT1 produces a larger lift than BWT2, however at
5. The displacement field of the BWT is passed to the moving mesh
U > 1, BWT2 has a larger lift than BWT1 (Fig. 8). The reason is that
the vortex shedding frequency (us) is not constant along the height
function for the next instance.
of BWT2 due to its conic shape. As air speed increases, the vortex
shedding frequency at sections closer to the upper end of the tur-
Since the main purpose of the current study is to investigate the
bine matches the natural frequency of the structure. The upper end
performance of the power generation, we consider the parameters
has a large frontal area and a long moment arm; hence the flow
of the simulation close to the lock-in region (Table 2) with the free
  induces a dominant resultant moment at the corresponding vortex
air speed of 5 ms . Fig. 4 compares the results of the 3D CFD-FEM shedding frequency and subsequently larger deflections. Large
simulation with the reduced-order model of BWT4 in the same deflections in turn enhance the vortex shedding and increase the
working conditions. Fig. 4 shows that the computed transverse lift force. However, at low air speeds, the lock-in occurs at the lower
displacement of the tip (YðLs þ Lm ; tÞÞ and resultant lift force end of the turbine which cannot cause a dominant moment at the
R lock-in frequency due to relatively small frontal area and moment
( 12 rV 2 DðxÞCL ðx; tÞdx) of the proposed reduced order model are in a
good match with those of the 3D CFD-FEM simulation. arm. In contrast, in BWT1 the vortex shedding frequencies of all
In terms of the steady state characteristics of the response, the sections along the height are the same, hence the lift force of all
amplitude and the frequency of the oscillations in both models sections are in the same frequency. Consequently, the lift force has
match with a satisfactory accuracy (Table 1). The transient char- only one powerful frequency component that enhances the forced
acteristics of the response have also been compared by means of vibrations in low air speeds.
computing the root mean square error (RMSE) of the reduced order In conclusion, in the conic shape BWT2, adaptation of the
model and the 3D CFD-FEM model (Table 1). Comparison of both shedding frequency improves the performance at high wind
steady state and transient characteristics shows an acceptable speeds. However, at low wind speeds, the lift force at a single
agreement between the models. Hence, the derived reduced order shedding frequency along the right circular cylinder BWT1 results
models of all BWTs are used for further investigations in this study. in a better performance than that of the conic BWT2.
Figs. 9 and 10 show the tip deflection and the lift force acting on
the turbine structure at free stream velocities of 3.6 m/s and 4.9 m/s
3.2. Simulation results that corresponds to U ¼ 0.85 and U ¼ 1.15, respectively. The struc-
ture in BWT3 and BWT4 are more flexible comparing with BWT1
Now, let us determine the power level that is transmitted to and BWT2, therefore the maximum deflections in Figs. 9 and 10 are
BWTs and investigate variations of the level with the free-stream larger than those of Figs. 6 and 7. Therefore, the power transmitted
velocity. In order to have a fair comparison between proposed rates of BWT3 and BWT4 are considerably higher.
configurations, the properties of BWTs (Table 2) are chosen to Due to the conical shape of BWT4, it has a larger moment of
achieve similar dynamic characteristics in all BWTs. The most inertia comparing with a BWT3 with the same mean diameter.
important dynamic characteristic of a BWT is lock-in, hence all Consequently, BWT4 has a lower natural frequency which pro-
BWTs are designed to experience lock-in around the air speed of motes the lock-in phenomenon in lower air speed comparing to
V ¼ 4:3 ms. The responses of the proposed BWTs at lock-in are BWT3.
depicted in Fig. 5. Fig. 11 shows the maximum lift coefficient at steady state for
Let us study the effect of the variations of the free-stream ve- various wind speeds close to the lock-in region versus the non-
locity on the dynamic response of BWTs. The non-dimensional dimensional frequency U for BWT3 and BWT4. For both turbines,
frequency U is introduced as U ¼ uuns to indicate the closeness to the lift force increases dramatically as the non-dimensional fre-
lock-in (lock-in occurs at U ¼ 1). Figs. 6 and 7 show the response of quency approaches 1.
BWT1 and BWT2 at free stream velocities of 3.6 m/s and V ¼ 4.9 m/s Similar to BWT1 and BWT2, in the stand-mast configurations,
that correspond to U ¼ 0.85 and U ¼ 1.15, respectively. the conic mast BWT4 has a higher performance at high wind
Fig. 8 shows the maximum lift coefficient at steady state
A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643 639

Fig. 5. BWT tip deflection (blue dashed line) and lift force (red line) for (a) BWT1, (b) BWT2, (c) BWT3, (d) BWT4. (For interpretation of the references to colour in this figure legend,
the reader is referred to the Web version of this article.)

speeds, while, at low wind speeds, the right circular cylinder BWT3 dimensions and material properties of the BWT in [32].
exhibits a better performance. This can be verified from Figs. 9 and
10, by comparing the deflections of BWT3 and BWT4 at low and 4. Conclusion
high wind speeds.
Fig. 12 illustrates the maximum steady tip deflection of the Four configurations of bladeless wind turbines (BWTs) have
BWTs in the lock-in region. The flexibly supported turbines (i.e. been investigated and compared. The BWTs consist of a relatively
BWT3 and BWT4) have larger tip deflections due to their flexibility long (right or conic) cylinder that is either flexible or mounted on a
which in turn enhance the vortex shedding and increase the lift flexible structure exposed to a uniform air flow. The nonlinear wake
force. oscillator semiempirical model was used to obtain an expression
Fig. 13 shows the mean power transmitted to the BWTs. We can for the crosswise flow-induced fluctuating lift force due to peri-
conclude that in low wind speeds the right circular flexibly odically shedding vortices. To derive nonlinear distributed-
mounted turbine (BWT3) can deliver a higher power while in high parameter models of the BWTs under the fluctuating lift force,
wind speeds the conical flexibly mounted turbine (BWT4) has a the Euler-Bernoulli beam theory and the Galerkin procedure were
better power output. Also note that a similar technology [32] is utilized. The predictions of the lift force and the deflection of the
reported to have a power output of 100 Watts that is close to the BWT based on the derived model were in a good agreement with
power transmitted to BWT3 and BWT4. An exact comparison with the results of the 3D CFD-FEM numerical simulations. The power
[32] is not possible due to the lack of information about the outputs of BWTs were studied in various wind speeds through
640 A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643

Fig. 6. BWT1 tip deflection (blue dashed line) and lift force (red line) (a) V ¼ 3.6 m/s (b) V ¼ 4.9 m/s. (For interpretation of the references to colour in this figure legend, the reader is
referred to the Web version of this article.)

Fig. 7. BWT2 tip deflection (blue dashed line) and lift force (red line); (a) V ¼ 3.7 m/s, (b) V ¼ 5 m/s. (For interpretation of the references to colour in this figure legend, the reader is
referred to the Web version of this article.)

numerical simulations of the derived models. It has been confirmed


that the amplitudes of the vibrations of the BWTs surged when the
vortex shedding frequency matches the natural frequency of the
structure, which is known as lock-in phenomenon. Furthermore,
the results of the simulations demonstrated that, the conic BWTs
have a higher performance at post-synchronization region (i.e. high
wind speeds); whereas the right circular cylinder BWTs exhibits a
better performance at pre-synchronization region (i.e. low wind
speeds). It was also demonstrated that mounting the bluff body on
a flexible structure, in comparison with using a flexible bluff body,
results in a significant increase in the power transmitted to the
BWT.

Fig. 8. Lift coefficient curves of BWT1 and BWT2.


A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643 641

Fig. 9. BWT3 tip deflection (blue dashed line) and lift force (red line) (a) V ¼ 3.6 m/s (b) V ¼ 4.9 m/s. (For interpretation of the references to colour in this figure legend, the reader is
referred to the Web version of this article.)

Fig. 10. BWT4 tip deflection (blue dashed line) and lift force (red line) (a) V ¼ 3.6 m/s (b) V ¼ 4.9 m/s. (For interpretation of the references to colour in this figure legend, the reader is
referred to the Web version of this article.)

Fig. 11. Lift coefficient curves of BWT3 and BWT4. Fig. 12. Maximum steady tip deflection response of the BWTs in the lock-in bound.
642 A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643

mode shape equations using Bessel functions is presented in [33].


Considering a fixed-free boundary condition for the truncated
conical cantilever, the first natural frequency and the correspond-
ing mode shape of BWT2 can be calculated using the method of
[33].
Using (36)e(39), the following coupled ODEs of motion are
derived for BWT2,

ZL
00 00 0
EyðtÞ fðxÞððI ðxÞf ðxÞÞ þ ð2I ðxÞfð3Þ ðxÞÞ þ ðIðxÞfð4Þ ðxÞÞÞ dx
0
2 L
Fig. 13. Power level transmitted to BWTs in the lock-in bound. ZL Z
1 24
þrs yðtÞ
€ AðxÞf ðxÞdx ¼ rV q DðxÞf2 ðxÞdx
2
2
Appendix 0 0
3
ZL
2a
Since the BWT2 is conical, the moment of inertia and the cross _
 yðtÞ f ðxÞdx5
2
(40)
section area of the turbine are functions of the length x (Fig. 14),
us
0
given by:

Fig. 14. The tapered conical BWT.

 Z L
 x4
qðtÞ  us GCL0 qðtÞ þ us qðtÞ
€ 2 _ 2
f2 ðxÞdx
IðxÞ ¼ I0 1 þ c (36)
L 0

 x2 n oZ L ZL 2
f ðxÞ
AðxÞ ¼ A0 1 þ c (37) þ us G4q2 ðtÞqðtÞ
_ f4 ðxÞdx ¼ us F yðtÞ
_ dx (41)
L DðxÞ
0 0
where I0 and A0 are the moment of inertia and the cross section
area at x ¼ 0, and c is the taper ratio.
Therefore, combining (2) and (14), the equations of motion are
written as: References

 00 00 0
 v2 y
[1] C.H. Williamson, R. Govardhan, A brief review of recent results in vortex-
E I ðxÞY ðx; tÞ þ 2I ðxÞY ð3Þ ðx; tÞ þ IðxÞY ð4Þ ðx; tÞ þ rs AðxÞ 2 induced vibrations, J. Wind Eng. Ind. Aerodyn. 96 (6) (2008 Jul 31) 713e735.
vt [2] A.B. Rostami, M. Armandei, Renewable energy harvesting by vortex-induced
  motions: review and benchmarking of technologies, Renew. Sustain. Energy
1 2 2a _
¼ rV DðxÞQ ðx; tÞ  Yðx; tÞ Rev. 70 (2017 Apr 30) 193e214.
2 us [3] A. Abdelkefi, Aeroelastic energy harvesting: a review, Int. J. Eng. Sci. 100 (2016
Mar 31) 112e135.
(38) [4] A.A. Khan, A. Shahzad, I. Hayat, M.S. Miah, Recovery of flow conditions for
optimum electricity generation through micro hydro turbines, Renew. Energy
96 (2016 Oct 31) 940e948.
  _ tÞ
Yðx; [5] A.A. Khan, A.M. Khan, M. Zahid, R. Rizwan, Flow acceleration by converging
Q€ ðx; tÞ  us G CL0
2
 4Q 2 ðx; tÞ Q_ ðx; tÞ þ u2s Q ðx; tÞ ¼ us F nozzles for power generation in existing canal system, Renew. Energy 60
DðxÞ
(2013 Dec 31) 548e552.
(39) [6] H.D. Akaydin, N. Elvin, Y. Andreopoulos, The performance of a self-excited
fluidic energy harvester, Smart Mater. Struct. 21 (2) (2012 Jan 24), 025007.
Reduced-order model for the BWT2 configuration can be ob- [7] R.D. Gabbai, H. Benaroya, An overview of modeling and experiments of
tained using the same approach as of BWT1. However, the natural vortex-induced vibration of circular cylinders, J. Sound Vib. 282 (3) (2005 Apr
22) 575e616.
frequencies and the mode shapes are different due to the non- [8] A. Abdelkefi, A.H. Nayfeh, M.R. Hajj, Design of piezoaeroelastic energy har-
uniform diameter of the structure. Analytical solution of the vesters, Nonlinear Dynam. 68 (4) (2012 Jun 1) 519e530.
A. Chizfahm et al. / Renewable Energy 121 (2018) 632e643 643

[9] A. Erturk, W.G. Vieira, C. De Marqui Jr., D.J. Inman, On the energy harvesting [21] R.A. Skop, O.M. Griffin, A model for the vortex-excited resonant response of
potential of piezoaeroelastic systems, Appl. Phys. Lett. 96 (18) (2010 May 3) bluff cylinders, J. Sound Vib. 27 (2) (1973 Mar 22) 225e233.
184103. [22] O.M. Griffin, Vortex-excited cross-flow vibrations of a single cylindrical tube,
[10] A. Abdelkefi, M.R. Hajj, A.H. Nayfeh, Piezoelectric energy harvesting from J. Pressure Vessel Technol. 102 (2) (1980 May 1) 158e166.
transverse galloping of bluff bodies, Smart Mater. Struct. 22 (1) (2012 Dec 10), [23] M.D. Dahleh, W.T. Thomson, Theory of Vibration with Applications, Prentice-
015014. Hall Inc, 1998.
[11] Y. Yang, L. Zhao, L. Tang, Comparative study of tip cross-sections for efficient [24] H.L. Dai, A. Abdelkefi, Y. Yang, L. Wang, Orientation of bluff body for designing
galloping energy harvesting, Appl. Phys. Lett. 102 (6) (2013 Feb 11), 064105. efficient energy harvesters from vortex-induced vibrations, Appl. Phys. Lett.
[12] M.M. Bernitsas, K. Raghavan, Y. Ben-Simon, E.M. Garcia, VIVACE (Vortex 108 (5) (2016 Feb 1), 053902.
Induced Vibration Aquatic Clean Energy): a new concept in generation of [25] S.S. Rao, Vibration of Continuous Systems, John Wiley & Sons, 2007 Feb 9.
clean and renewable energy from fluid flow, J. Offshore Mech. Arctic Eng. 130 [26] F.M. Besem, J.P. Thomas, R.E. Kielb, E.H. Dowell, An aeroelastic model for
(4) (2008 Nov 1), 041101. vortex-induced vibrating cylinders subject to frequency lock-in, J. Fluid Struct.
[13] Cajas JC, Houzeaux G, Ya n~ ez DJ, Mier-Torrecilla M. SHAPE Project Vortex 61 (2016 Feb 29) 42e59.
Bladeless: Parallel Multi-code Coupling for Fluid-structure Interaction in [27] R. Sanchez, R. Palacios, T.D. Economon, H.L. Kline, J.J. Alonso, F. Palacios, To-
Wind Energy Generation. wards a Fluid-structure Interaction Solver for Problems with Large De-
[14] R.A. Skop, S. Balasubramanian, A new twist on an old model for vortex-excited formations within the Open-source SU2 Suite, AIAA SciTech, 2016.
vibrations, J. Fluid Struct. 11 (4) (1997 May 1) 395e412. [28] L. Song, S. Fu, J. Cao, L. Ma, J. Wu, An investigation into the hydrodynamics of a
[15] S.M. Hasheminejad, A.H. Rabiee, M. Jarrahi, Semi-active vortex induced vi- flexible riser undergoing vortex-induced vibration, J. Fluid Struct. 63 (2016
bration control of an elastic elliptical cylinder with energy regeneration May 31) 325e350.
capability, Int. J. Struct. Stabil. Dyn. (2017 Feb 7), 1750107. [29] M.S. Triantafyllou, M.A. Grosenbaugh, R. Gopalkrishnan, Vortex-induced Vi-
[16] T. Wu, A. Kareem, Vortex-induced vibration of bridge decks: volterra series- brations in a Sheared Flow: a New Predictive Method, Woods Hole Oceano-
based model, J. Eng. Mech. 139 (12) (2013 Mar 16) 1831e1843. graphic Institution Ma, 1994.
[17] D.J. Olinger, A low-order model for vortex shedding patterns behind vibrating [30] A. Protos, V.W. Goldschmidt, G.H. Toebes, Hydroelastic forces on bluff cylin-
flexible cables, Phys. Fluids 10 (8) (1998 Aug) 1953e1961. ders, J. Basic Eng. 90 (3) (1968 Sep 1) 378e386.
[18] P.A. Monkewitz, C.H. Williamson, G.D. Miller, Phase dynamics of K arman [31] S.M. Hasheminejad, M. Jarrahi, Numerical simulation of two dimensional
vortices in cylinder wakes, Phys. Fluids 8 (1) (1996 Jan) 91e96. vortex-induced vibrations of an elliptic cylinder at low Reynolds numbers,
[19] R.E. Bishop, A.Y. Hassan, The lift and drag forces on a circular cylinder oscil- Comput. Fluids 107 (2015 Jan 31) 25e42.
lating in a flowing fluid, Proc. Roy. Soc. Lond. A Math Phys Eng Sci 277 (1368) [32] R. Whitlock, The power of the vortex: an interview, Renew. Energy Mag. 15
(1964 Jan 7) 51e75. The Royal Society. (3) (2015 April 07).
[20] M.L. Facchinetti, E. De Langre, F. Biolley, Coupling of structure and wake os- [33] S. Naguleswaran, A direct solution for the transverse vibration of Euler-
cillators in vortex-induced vibrations, J. Fluid Struct. 19 (2) (2004 Feb 29) Bernoulli wedge and cone beams, J. Sound Vib. 172 (3) (1994 May 5)
123e140. 289e304.

You might also like