Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Accepted Manuscript

Dynamic modeling of vortex induced vibration wind turbines

A. Chizfahm, E. Azadi Yazdi, M. Eghtesad

PII: S0960-1481(18)30038-7
DOI: 10.1016/j.renene.2018.01.038
Reference: RENE 9643

To appear in: Renewable Energy

Received Date: 26 October 2017


Revised Date: 18 December 2017
Accepted Date: 13 January 2018

Please cite this article as: Chizfahm A, Yazdi EA, Eghtesad M, Dynamic modeling of vortex induced
vibration wind turbines, Renewable Energy (2018), doi: 10.1016/j.renene.2018.01.038.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
Lift coefficient curves of BWT1 and BWT2 versus non- Lift coefficient curves of BWT3 and BWT4 versus non-
dimensional frequencies Ω dimensional frequencies Ω

RI
Geometrical model of the bladeless wind
turbine

SC
(a) BWT1, (b) BWT2, (c) BWT3, (d) BWT4

Power level transmitted to BWTs in the lock-in bound

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

1 Dynamic Modeling of Vortex Induced Vibration Wind Turbines

2 A. Chizfahm,
3 M.Sc. Student, School of Mechanical Engineering, Shiraz University, Shiraz, Iran
4 E. Azadi Yazdi*,
5 Assistant Professor, School of Mechanical Engineering, Shiraz University, Shiraz, Iran,
6 Email: ehsanazadi@shirazu.ac.ir, Telephone: +98-938-415-3797

PT
7 M. Eghtesad,
8 Professor, School of Mechanical Engineering, Shiraz University, Shiraz, Iran
9
10

RI
11 Abstract:

SC
12 This paper studies the dynamic modeling of four configurations of vortex-induced vibrations of a
13 bladeless wind turbine (BWT). The BWTs consist of a bluff body mounted on a flexible
14 structure in the flow field. The shape of the bluff body and its mounting structure are different
15
U
among the proposed BWTs. The Euler-Bernoulli beam theory and the Galerkin procedure are
AN
16 used to derive a nonlinear distributed-parameter model for the BWTs under a fluctuating lift
17 force due to periodically shedding vortices. The derived dynamic model is validated through
M

18 comparison with a 3D CFD-FEM numerical simulation. The effects of the wind speed on the
19 induced lift force, turbine deflection, and generated power of four BWTs are investigated. It is
verified that the amplitude of the vibrations of the BWT increases significantly when the vortex
D

20
21 shedding is synchronized with the structural oscillations. The results show that, while conic
TE

22 BWTs have a higher performance at post-synchronization region (i.e. high wind speeds), the
23 right circular cylinder BWTs exhibits a better performance at pre-synchronization region (i.e.
EP

24 low wind speeds).

25 Keywords: Bladeless Wind Turbine; Vortex Induced Vibration; Dynamic Modeling; lock-in
C

26 phenomenon
AC

27 1 Introduction
28 Submerged bodies may experience periodically shedding vortices in the flow fields close to
29 their surface. The periodic shedding vortices in the wind or water currents result in fluctuations
30 in the pressures around the submerged body which yields to vortex-induced vibrations (VIV) in
31 flexible structures. The structural VIV is often a detrimental phenomenon due to the fatigue
32 damage consequences (for example in power transmission lines, towing cables, and mooring

1
ACCEPTED MANUSCRIPT

33 lines); however, it can be beneficial in wind power [1-3] and hydropower utilities [4, 5]. The
34 practical significance of VIV energy in renewable energy and energy harvesting has led to a
35 large number of fundamental studies, for instance VIV of circular cylinders [6, 7], flutter of
36 airfoils [8, 9], and galloping of sliding structures [10, 11]. Most of the previous research studies
37 have considered a VIV harvester similar to the so-called VIVACE (Vortex Induced Vibration

PT
38 Aquatic Clean Energy) device presented in [12]. The VIVACE is a spring mounted cylinder that
39 exhibits transverse oscillations in the flow field. VIVACE and similar devices cannot be up-

RI
40 scaled easily due to the relatively high cost of the required support structure. In contrast, recently
41 a BWT is studied in [13] that eliminates the need of the linear support by mounting the cylinder

SC
42 on the tip of a flexible beam.

43 The proposed VIV bladeless wind turbine (BWT) consists of a bluff body (often a relatively

U
44 long cylinder) mounted on a flexible structure in the flow field. The absence of rigidly moving
AN
45 parts is the main difference between VIV bladeless and conventional wind turbines. The flow
46 power transmitted to the VIV bladeless wind turbine is considerable when the lock-in
47 phenomenon occurs, i.e. the frequency of vortex shedding is close to the structural natural
M

48 frequency. Hence, to study and improve the energy production rate of VIV bladeless wind
49 turbines, the effects of the turbine design parameters on the lock-in phenomenon should be
D

50 carefully analyzed through dynamic modeling.


TE

51 An accurate model of VIV bladeless wind turbines requires a model for the forced-vibrations
52 of the structure subjected to fluid flow forces attained by the Navier–Stokes equations in the
EP

53 presence of moving boundaries [14]. However, due to computational complexity, the numerical
54 model is not useful for structural design and controller synthesis purposes. Moreover,
C

55 computational difficulties arise for simulating three-dimensional (3D) flexible VIV wind turbines
56 that often have large aspect ratios [15].
AC

57 Alternatively, the computational complexity associated with the fluid flow modeling can be
58 reduced by considering a two-dimensional (2D) flow model that exerts a crosswise flow-induced
59 force on each slice of the structure. In this approach the sophisticated CFD calculations are
60 replaced by reduced order models [16], semiempirical models [17], or analytical models [18]. In
61 general, the traditional semiempirical models are the most favorable, because they have neither
62 the conservative simplifications of the analytical models nor the computational complexity

2
ACCEPTED MANUSCRIPT

63 associated with the reduced order models. The most proper semiempirical model of the crosswise
64 flow-induced force follows the idea of a wake oscillator with a single flow variable that models
65 the harmonic nature of the vortex shedding [19]. The flow variable satisfies van der Pol or
66 Rayleigh equation which results in a stable and nearly harmonic oscillation [20]. The wake
67 oscillator semiempirical model has been shown to have the potential to demonstrate basic

PT
68 features of VIV, such as limit cycle oscillations and lock-in phenomenon [21, 22].

RI
69 Moreover, the modeling complexity of the VIV bladeless wind turbine caused by the
70 structural forced-vibration model, which leads to partial differential equations (PDEs) of motion,

SC
71 can be reduced by various methods. The PDE can be reduced to an ordinary differential equation
72 (ODE) via Galerkin approach [14]. Several other methods have also been proposed to simplify
73 the structural vibrations modeling such as the method of lumped parameters [23, 24], Rayleigh-

U
74 Ritz method [25], assumed modes method [26], collocation method [27] and least squares method
AN
75 [28].

76 The purpose of this paper is to compare and contrast various aspects of four different concepts
M

77 for VIV bladeless wind turbines. The proposed BWTs are different in the shape of their bluff
78 body and their mounting structure. Specifically, we aim to determine the power levels that can be
D

79 transmitted to BWTs and variations of these levels with the free-stream air speed. To this end,
80 simplified analytical dynamic models have been derived for the transverse displacement degree
TE

81 of freedom (DOF) of BWTs. The dynamic models have been used to analyze the lock-in
82 phenomenon and investigate its dependence on the free-stream velocity for the proposed BWTs.
EP

83 The paper is organized as follows. Section 2 presents the dynamic models of the proposed
84 BWTs. The wake oscillator semiempirical model is used to derive a simplified expression for the
C

85 crosswise flow-induced force on the BWTs. The wake oscillator model has been combined with
AC

86 the transverse vibrations model of the structure for each BWT. The proposed dynamic models
87 have been validated through comparison with 3D CFD and FEM numerical simulations. Then,
88 we determine the power level that can be transmitted to BWTs through numerical simulations of
89 the dynamic models. We also study variations of the power level with the air speed. Finally,
90 conclusions are presented in Section 4.

3
ACCEPTED MANUSCRIPT

91 2 Dynamic modeling of VIV Bladeless Wind Turbines


92 The VIV bladeless wind turbines consist of a relatively long cylinder that is either flexible or
93 mounted on a flexible structure exposed to air flow. We consider four BWTs that differ in the
94 shape of the cylinder and its mounting structure (Figure 1):

PT
95 • BWT1: A right circular flexible cylinder
96 • BWT2: A conic flexible cylinder

RI
97 • BWT3: A right circular rigid cylinder mounted on a flexible beam
98 • BWT4: A conic rigid cylinder mounted on a flexible beam

SC
99 In BWTs the periodic shedding vortices in the air flow along the z-direction induce vibrations
100 in the y-direction. The amplitude of the vibrations (and consequently the flow power transmitted

U
101 to the BWT) is considerable when the frequency of vortex shedding is close to the structural
AN
102 natural frequency that is known as lock-in phenomenon. Hence, to study the lock-in
103 phenomenon, the dynamic models of the proposed BWTs are determined in this section. Note
104 that the computational complexity associated with the CFD and FEM models are avoided by
M

105 deriving simplified dynamic models. Also note that a power take-off unit should be installed near
106 the bottom end of the BWT which is not considered in this paper for the sake of modeling
D

107 simplicity.
TE
C EP
AC

4
ACCEPTED MANUSCRIPT

Movements
direction

PT
RI
U SC
X X
X X
Y Y Y Y
AN
(a) (b) (c) (d)
108

109 Figure 1-The bladeless wind turbines; (a) BWT1 (b) BWT2 (c) BWT3 (d) BWT4
M

110
D

111 2.1 Aerodynamic force model


112 The wake oscillator semiempirical model is used to obtain a simplified expression for the
TE

113 crosswise flow-induced force on the BWTs. The nonlinear wake oscillator model was proposed
114 for the first time by Skop and Griffin [21]. The model has been further developed by Skop and
EP

115 Balasubramanian [14] to be able to accurately capture the asymptotic, self-limiting structural
116 response for flexible cylinders. In the model a stall term is incorporated to describe the
C

117 dependence of the cross-flow force on the transverse velocity of the structure. The velocity
118 coupling enables the model to capture the lock-in phenomenon. Moreover, the displacement
AC

119 coupling and the acceleration coupling were also introduced in [20] to achieve a more accurate
120 model of the lock-in phenomenon. Based on the wake oscillator semiempirical model, the
121 fluctuating lift coefficient CL(x,t) at the time t on a slice of the BWT located at a distance x from
122 the origin is defined as:

5
ACCEPTED MANUSCRIPT

2α ɺ
C L ( x , t ) = Q (x , t ) − Y (x , t ) (1)
D ωs

123 where Q ( x , t ) is the excitation component of the fluctuating lift coefficient and Yɺ ( x, t ) is the

124 transverse velocity of the slice of the BWT. The second term on the right side is called the stall

PT
125 term, where α is an empirical constant, ωs is the vortex shedding frequency that is defined as

126  = 2 , S is the Strouhal number [14], V is the wind velocity and D is the mean diameter of

RI


127 the slice of the BWT. In this model, the stall parameter, , is defined to limit the fluctuating


SC
128 lift coefficient for large deflections based on the observations of Triantafyllou [29]. The van der
129 Pol equation is satisfied by using the excitation component of the fluctuating lift coefficient
Q(x,t) as:

U
130 AN
Yɺ ( x , t )
Qɺɺ( x , t ) − ωsG (C L2 0 − 4Q 2 ( x , t ))Qɺ ( x , t ) + ωs2Q ( x , t ) = ωs F (2)
D
M

131 where, CL0, G and F are empirical parameters [14]. For stationary cylinders, the right-hand side
132 of the equation (2) is equal to zero, so that the above equation has a self-excited, self-limited
D

133 solution given by,


TE

Q = C L 0 sin(ωs t ) (3)
EP

134 This means that the fluctuating lift coefficient on stationary cylinder has the amplitude equal to
135 CL0, where based on Protos et al. [30]  ≪ 1 for circular sections.
C

136 2.2 Structural VIV models


137 Since the structure of the proposed BWTs differs, the equations of motion have been derived
AC

138 separately for each BWT. In all BWT configurations, the wind flow is assumed to be uniformly
139 steady with a constant velocity and the boundary layer effect on the clamped end is neglected.

140 2.2.1 BWT1: A right circular flexible cylinder


141 In the BWT1 configuration presented in Figure 1(a), the turbine is a uniform right cylinder that
142 can be modeled as a clamped-free Euler-Bernoulli beam exposed to air flow. Considering the

6
ACCEPTED MANUSCRIPT


143 distributed aerodynamic force expression on the beam element as 
 , , the equation

144 of motion is expressed by Euler-Bernoulli equation as:

∂4 ∂ 2Y (x, t) 1
( EIY ( x , t ) ) + ρ = ρ V 2 DCL ( x, t ) (4)
∂x 4 ∂t 2
s
2

PT
145 where E is the Young modulus of elasticity of the cylinder, I is the second moment of area of the
146 cross section, ρs is mass per unit length of the cylinder, and ρ is the air density. Let us consider

RI
147 the boundary conditions of the clamped-free beam as,

∂Y

SC
Y (0, t ) = 0; (0, t ) = 0
∂x (5)
∂Y2
∂ 3Y
(L , t ) = 0; (L , t ) = 0
∂x 2 ∂x 3

U
Substituting (1) into (4), the two coupled dynamic equations of motion are obtained as,
AN
148

 ∂4 ∂ 2Y (x , t ) 1  2α  (6)
 4 ( EIY (x , t ) ) + ρs = ρV 2  DQ (x , t ) − Yɺ (x , t ) 
 ∂x ∂t 2
ωs
M

2  

 ɺɺ ɺ
Y (x , t ) (7)
Q (x , t ) − ωsG (C L 0 − 4Q (x , t ))Q (x , t ) + ωs Q (x , t ) = ωs F D
2 2 ɺ 2
D

149 A simplified form of the partial differential equation (PDE) of motion can be developed using
TE

150 the Galerkin method, where the solution to the PDE is assumed to be of the form:

Y (x , t ) = ∑ φi (x ) y i (t ) (8)
EP

151 where   and   are called the mode shape and the modal response factor, respectively.
C

152 Furthermore, based on the results of Skop and Balasubramanian [14], mode shapes of ,  can
AC

153 be approximated to be the same as the mode shapes of !, . There are two possible
154 explanations for the above statement. First, the response of the BWT is dominated by the lock-in
155 phenomenon. Hence, only mode shapes of the fluctuating lift coefficient which are close to
156 structural mode shapes have a considerable contribution to the response of the system near lock-
157 in region. Second, if van der Pol equation (1) is linearized, it can be easily shown that the mode
158 shapes of ,  exactly match the mode shapes of !,  [14]. Therefore, for ωs ≈ ωn ,i , , 

159 may be assumed as:

7
ACCEPTED MANUSCRIPT

Q (x , t ) = ∑ φi (x ) q i (t ) (9)

160 where "  is the modal response factor for the fluctuating lift excitation component. To comply
161 with the boundary conditions (5), the ith mode shape of the clamped-free cantilever is chosen as,

PT
(cos βi l + cosh βi l )
φi (x ) = (cos βi x − cosh β i x ) + (sin βi x − sinh β i x ) (10)
(sin β i l + sinh βi l )

RI
162 The corresponding natural frequency is given by:

SC
EI
ωn = ( βi l )
2
(11)
i
ms l 4

163
U
The largest amount of wind energy transmitted to the BWT is in the first mode, where we have
AN
164 ( β1l ) 2 = 3.52 [23]. Hence, a dynamic model that considers only the first mode has an acceptable
165 error in the prediction of the energy production rate of the BWT. By considering only the first
M

166 mode shape, substituting (8) and (9) into (6) and (7) and multiplying the mode shape of the
167 system by both sides of the equation and integrating on  , the reduced-order equations (ODEs) of
D

168 the motion are written as:


TE

1  2α  L
EIy ( t ) ∫ φ ( x)φ ( 4) ( x)dx + ρ s ɺɺ
y ( t ) ∫ φ 2 ( x)dx = ρ V 2  Dq ( t ) − yɺ ( t )  ∫ φ 2 ( x)dx
L L
(12)
0 0 2  ωs  0
EP

( qɺɺ( t ) − ω GC qɺ ( t ) + ω q ( t ) ) ∫ φ 2 ( x)dx + ( 4ωs Gq 2 ( t ) qɺ ( t ) ) ∫ φ 4 ( x)dx


L L
2 2
s L0 s 0 0

 yɺ ( t )  L 2 (13)
=  ωs F  φ ( x)dx
D  ∫0
C


AC

where, φ ′ and φ ( n ) represent d φ dx and d n φ dx respectively, and for simplicity of notation


n
169

170   ≔  x,  ≔  t, and " ≔ " t. The coupled equations of motion are linear
171 with respect to  and its derivatives, but nonlinear with respect to " due to the appearance
172 of the term "
" in (13).

8
ACCEPTED MANUSCRIPT

173 2.2.2 BWT2: A conic flexible cylinder


174 Considering BWT2 (Figure 1(b)), the turbine is a tapered conical flexible cylinder attached to the
175 ground. Similar to the previous section, the equation of motion is derived based on the Euler-
176 Bernoulli beam theory as:

∂2 ∂2 ∂ 2Y (x , t ) 1

PT
[ EI ( x ) Y ( x , t )] + ρ A ( x ) = ρV 2 D (x )C L (x , t ) (14)
∂x ∂x ∂t
2 2 s 2
2
177 The following coupled ODEs of motion are derived for BWT2 (Please see Appendix for the

RI
178 proof),

( )
Ey ( t ) ∫ φ ( x) ( I ′′( x)φ ′′( x)) + (2 I ′( x)φ ( 3) ( x)) + ( I ( x)φ (4) ( x)) dx + ρ s ɺɺ
y ( t ) ∫ A( x)φ 2 ( x)dx
L L

SC
0 0

1  L 2α  (15)
yɺ ( t ) ∫ φ 2 ( x)dx 
L
= ρ V 2  q ∫ D( x)φ 2 ( x)dx −
2 
0 ωs 0

( qɺɺ( t ) − ω GC qɺ ( t ) + ω q ( t ) ) ∫ U
φ 2 ( x)dx + {ωs G 4q 2 ( t ) qɺ ( t )} ∫ φ 4 ( x)dx = ωs Fyɺ ( t ) ∫
φ 2 ( x)
AN
L L L
2 2
s L0 s dx (16)
0 0 0 D( x)
M

179 2.2.3 BWT3: A right circular rigid cylinder mounted on a flexible beam

180 The third configuration, BWT3 (Figure 1(c)), consists of a flexible stand and a rigid mast. The
D

181 flexible stand enables the BWT to vibrate, while the mast causes the vortex shedding and
182 generates an oscillatory aerodynamic force. In this configuration, the aerodynamic force on the
TE

183 stand can be neglected because it is relatively narrow. Hence, to study the lock-in phenomenon,
184 the stand can be modeled as a clamped-free Euler-Bernoulli beam. The effect of mass and inertia
EP

185 of the mast and the distributed fluctuating lift force on the mast can be replaced by a resultant
186 point force and a resultant point moment on the tip of the stand as shown in Figure 2. Hence,
C

187 using (8), the kinetic energy and the potential energy of the system are given by,
AC

1 Ls ɺ 2 1 1
( ) ∑∑ ( ) ( ) φi ( x ) φ j ( x ) mdx = ∑ M i yɺi2 ( t )
Ls
T=
2 ∫0
Y x , t mdx =
2 i j
y
ɺ i t y
ɺ j t ∫0 2
(17)

1 Ls 1 1
U = ∫ EI Y ′′2 ( x, t ) dx = ∑∑ yi ( t ) y j ( t ) ∫ EIφi′′( x ) φ ′′j ( x ) dx = ∑ Ki yi2 ( t )
Ls
(18)
2 0 2 i j 0 2

 
188 where & = ' 
 ( ) is the generalized mass and * = ' +, - "/
) is the
189 generalized stiffness. The work done by the distributed fluctuating lift force and the mass and
190 inertia of the mast is given by,

9
ACCEPTED MANUSCRIPT

δ W = FM δ Y ( Ls , t ) + M M δθ ( Ls , t ) = FM ( ∑ φi ( Ls ) δ yi ( t ) ) + M M ( ∑φi′ ( Ls ) δ yi ( t ) ) (19)

191 01 and &1 are respectively the force and moment exerted from the mast on the tip of the stand,
192 defined by,

PT
Ls + Lm 1
FM = ∫ ρV 2 DCL ( x, t ) dx − mm ɺɺym (20)
Ls 2
Ls + Lm 1
MM = ∫ ρV 2 DCL ( x, t )( x − Ls ) dx − mm Lm ɺɺym − I mθɺɺm (21)

RI
Ls 2

193 where for the center of mass of the mast;, 234 is the distance to the tip of the stand, and ,4
̅ is the

SC
194 moment of inertia about z-axis. Also 34 and 74 are written as

θ m = Y ′ ( x, t ) x = L = ∑ φi′ ( Ls ) yi ( t ) (22)

U
s

Lm L  L 
ym ≈ Y ( x, t ) x = L + θm = ∑φi ( Ls ) yi ( t ) + m ∑φi′ ( Ls ) yi ( t ) = ∑ φi ( Ls ) + m φi′( Ls ) yi ( t ) (23)
AN
s 2 2  2 
8 :; :; :B
195 Now, substituting (17) - (19) into Lagrange’s equation ( − + =  ), the ODEs of
89 :<=> :<@ :<@
M

196 motion are given as,

yi ( t ) + Ki yi ( t ) = FM φi ( Ls ) + M M φi′ ( Ls ) (24)
D

M i ɺɺ
TE

197
C EP
AC

10
ACCEPTED MANUSCRIPT

C L (x, t) C L (x, t)

mm ɺɺ
ym

I mθɺɺm
m m , Lm

PT
Lm

RI
SC
m s , Ls

U
X X

Y Y
AN
198

199 Figure 2- BWT3 and BWT4 (a) The schematic model (b) Free-body diagram.
M

200 Since the mast is assumed to be a rigid body, the velocity of the mast (that is required for the lift
201 coefficient equation (1)) is written as,
D

Yɺ ( x, t ) ≈ Yɺ ( x, t ) + ( x − Ls ) θɺm
TE

x = Ls (25)
= ∑ φi ( Ls ) yɺi ( t ) + ( x − Ls ) ∑ φi′ ( Ls ) yɺi ( t ) Ls ≤ x ≤ Ls + Lm

202 As it has been explained earlier, the most significant contributors among the mode shapes of the
EP

203 excitation component of the fluctuating lift coefficient are the ones that match the structural
204 mode shapes, therefore,
C

Q ( x , t ) = q i (t ) φi ( Ls ) + ( x − Ls ) φi′ ( Ls )  (26)


AC

205 Considering only the first mode of the vibrations of the stand, from (1) and (20)-(26) the
206 reduced-order model is obtained as:

11
ACCEPTED MANUSCRIPT

 2α 
Myɺɺ(t ) +   Γyɺ (t ) + Ky (t ) = Γq (t ) (27)
 D ωs 

Ls + Lm 3

∫ h( x) dx  yɺ ( t ) 
( qɺɺ ( t ) − ω GC qɺ ( t ) + ω q ( t ) ) + ( 4ω Gq ( t ) qɺ ( t ) )
s
2
L0
2
s s
2 Ls
Ls + Lm
=  ωs F  (28)
∫ h( x) dx  D 

PT
Ls

207 where M, K, Γ and h( x) are defined as,

RI
  L   Lm Lm I m  2 
M = ∫ φ 2 (x)m s dx + m m φ 2 ( L s ) +  L m + m  φφ ′ ( Ls ) +   φ ′ ( Ls )
Ls
+

SC
(29)
0
  2   2 mm  

Ls
K = ∫ EI [φ ′′(x)]2 dx

U
0
AN (30)

φ 2 ( Ls ) + ( x − Ls ) ( 2φ ( Ls ) φ ′ ( Ls ) + 2 ( x − Ls ) φ ′2 ( Ls ) )  dx
1 Ls + Lm
Γ= ρV 2 D ∫   (31)
2 Ls
M

h( x) = φ ( Ls ) + ( x − Ls ) φ ′ ( Ls ) (32)
D

208 2.2.4 BWT4: A conic rigid cylinder mounted on a flexible beam


TE

209 The configuration of BWT4 is similar to BWT3 except for the shape of the mast which is
210 changed to a conic shape (Figure 1(d)). The conical shape of the mast enhances the lock-in
211 phenomenon over a wider range of wind speeds. To derive the model, by considering a linear
EP

212 variation in diameter in (27), the reduced order equations of motion for the first mode of the
213 vibrations are obtained as,
C

 2α 
AC

Myɺɺ (t ) +   Λyɺ (t ) + Ky (t ) = Λq (t ) (33)


 D ωs 

( qɺɺ( t ) − ω GC qɺ ( t ) + ω q ( t ) ) ∫ ( )∫
Ls + Lm Ls + Lm 3

s
2
L0
2
s h( x) dx + 4ωs Gq 2 ( t ) qɺ ( t ) h( x) dx
Ls Ls

Ls + Lm h( x ) (34)
= ωs Fyɺ ( t ) ∫ dx
Ls D( x)

214 where Λ is defined as,

12
ACCEPTED MANUSCRIPT

D( x) φ 2 ( Ls ) + ( x − Ls ) (φ ( Ls ) φ ′ ( Ls ) + ( x − Ls ) φ ′2 ( Ls ) )  dx
1 Ls + Lm
Λ= ρV 2 ∫ (35)
2 L s

215 3 Results and Discussions

PT
216 In the previous section, reduced order models of the vortex-induced vibrations of the four
217 proposed BWTs were developed. In this section, first to evaluate the accuracy of the reduced
218 order models, their results are compared with the results of a 3D CFD-FEM numerical

RI
219 simulation. Then, we aim to determine the power level that can be transmitted to BWTs and to
220 study variations of the level with the air velocity through numerical simulations of the reduced

SC
221 order models.

222 3.1 Model validation

U
223 A 3D CFD-FEM simulation has been performed to validate the reduced order models. The
AN
224 configuration of BWT4 is chosen for validation purpose because it combines the main features of
225 the fluid-solid interaction of the other configurations. The 3D CFD-FEM is performed by linking
226 the finite element model of the structure to the CFD model in the commercial ANSYS Fluent
M

227 package through a User-Defined Function (UDF).


D

228 The CFD model considers a non-stationary laminar incompressible flow. The BWT4 is placed in
229 a rectangular cuboid computational domain with downstream and upstream boundaries extended
TE

C I
230 to = 160 and = 80, respectively. The lateral boundaries are considered at the distance
DEF DEF

231 K = 120DMNO (Figure 3 (a)). In order to avoid excessive simulation error due to the sharp
EP

232 pressure gradients in cells near cylinder boundary, a moving mesh technique is used that
233 produces high-quality structured cells around the BWT (Figure 3 (a)). As BWT oscillates, the
C

234 moving/deforming mesh function regenerates the grid cells at each iteration. The computational
AC

235 domain is partitioned into three distinct blocks with map type quad-cells in the center cylindrical
236 block enclosing cylinder, unstructured T-grid-type tetrahedral/hybrid cells in the surrounding
237 rectangular block and map type quad-cells for the farther sides. The aspect ratio of the quad-
238 cells, the grid size, and the time step size are chosen fine enough to achieve mesh independence
239 and ensure acceptable errors in unsteady flow simulations [31]. The resultant mesh has about 4
240 million cells for the BWT4 dimensions given in Figure 3.

13
ACCEPTED MANUSCRIPT

241 In the CFD model, the first-order pressure-based implicit solver is used for transient formulation
242 and momentum equations that lead to the aerodynamic forces. Furthermore, to improve the
243 accuracy of the second order time integration, the non-iterative fractional step method is used for
244 the velocity-pressure coupling.

PT
v y = 0, ∂v x ∂y = 0

RI
v x = 0,v y = yɺ

SC
Y
∂v x ∂x = 0
X ∂v y ∂x = 0

U
AN
v y = 0, ∂v x ∂y = 0
245
M

246 (a)
D
TE
C EP

247
AC

248 (b)

249 Figure 3- (a) Fluid flow problem configuration and boundary conditions. (b) Details of the
250 constructed mesh.

251 The structural vibration of BWT4 is determined by a finite element model in ANSYS which is
252 linked with the CFD model through a UDF. The transient-structural model considers the lift
253 force as the only external force causing the vibrations of BWT4. The entire length of BWT4
254 (both the stand and the mast) is considered to be flexible. The computational domain consists of

14
ACCEPTED MANUSCRIPT

255 50,000 unstructured tetrahedral cells. The size of the cells and the time-step are chosen fine
256 enough to guarantee an adequate accuracy in transient response.

257 The CFD-FEM modeling procedure for time instance  is outlined as follows:

258 1. Pressure and velocity fields are computed by the CFD model.

PT
259 2. Aerodynamic load  ,  is determined by the CFD model.
260 3. Aerodynamic load is applied on BWT4 in the FEM model.

RI
261 4. Transverse motion ,  is computed by the FEM model.
262 5. The displacement field of the BWT is passed to the moving mesh function for the next

SC
263 instance.

264 Since the main purpose of the current study is to investigate the performance of the power

U
265 generation, we consider the parameters of the simulation close to the lock-in region (Table 2)
Q
AN
266 with the free air speed of 5 . Figure 4 compares the results of the 3D CFD-FEM simulation
R

267 with the reduced-order model of BWT4 in the same working conditions. Figure 4 shows that the
computed transverse displacement of the tip (!2 + 24 ,  and resultant lift force
M

268

269 ('

 , )) of the proposed reduced order model are in a good match with those of
D

270 the 3D CFD-FEM simulation.


TE

271

272 Figure 4- Results of the CFD-FEM simulation and the reduced-order model of BWT4
EP

273 In terms of the steady state characteristics of the response, the amplitude and the frequency of the
274 oscillations in both models match with a satisfactory accuracy (Table 1). The transient
C

275 characteristics of the response have also been compared by means of computing the root mean
AC

276 square error (RMSE) of the reduced order model and the 3D CFD-FEM model (Table 1).
277 Comparison of both steady state and transient characteristics shows an acceptable agreement
278 between the models. Hence, the derived reduced order models of all BWTs are used for further
279 investigations in this study.

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

Mean Mean Mean amplitude Mean frequency RMSE of RMSE of


amplitude of frequency of of lift force of lift force displacement lift force

15
ACCEPTED MANUSCRIPT

displacement displacement (N) (Hz)


(m) (Hz)

Reduced
0.297 1.98 14.60 1.99
order model
1.13 % 3.27 %
3D CFD-
0.300 2.00 14.55 2.00
FEM model

281 3.2 Simulation results

PT
282 Now, let us determine the power level that is transmitted to BWTs and investigate variations of
283 the level with the free-stream velocity. In order to have a fair comparison between proposed

RI
284 configurations, the properties of BWTs (Table 2) are chosen to achieve similar dynamic
285 characteristics in all BWTs. The most important dynamic characteristic of a BWT is lock-in,

SC
4
286 hence all BWTs are designed to experience lock-in around the air speed of  = 4.3 . The


287 responses of the proposed BWTs at lock-in are depicted in Figure 5.

U
288 Table 2- Dimensions and properties of the BWTs
AN
Physical Properties BWT1 BWT2 BWT3 BWT4
Lengths (Lm, Ls) (m) (4,0) (4,0) (4, 0.15) (4, 0.15)
M

Mean outer diameter of the mast D (m) 0.18 0.225 0.375 0.375
Mast thickness t (m) 0.001 0.001 0.001 0.001
D

Stand diameter Ds (mm) - - 0.018 0.018


3
Mass density (ρm ,ρs) (kg/m ) (1040,0) (1040,0) (1040, 8000) (1040, 8000)
TE

Mass (mm ,ms ) (kg) (2.6,0) (4.4,0) (5.21,0.39) (5.27,0.39)


Young’s modulus Es (GPa) 2.2 2.2 2.2 2.2
Taper ratio c - 2 - 2
EP

289
C

0.02 0.02
Tip deflection (m), Lift force ( × 10 N)

Tip deflection (m), Lift force ( × 10 N)

0.015 0.015
3

3
AC

0.01 0.01

0.005 0.005

0 0

-0.005 -0.005

-0.01 -0.01

-0.015 -0.015

-0.02 -0.02
0 1 2 3 4 5 0 1 2 3 4 5
Time (s) Time (s)
290
291 (a) (b)

16
ACCEPTED MANUSCRIPT

0.15 0.15
Tip deflection (m), Lift force ( × 10 N)

Tip deflection (m), Lift force ( × 10 N)


2

2
0.1 0.1

0.05 0.05

0 0

PT
-0.05 -0.05

-0.1 -0.1

RI
-0.15 -0.15
0 2 4 6 8 10 0 2 4 6 8 10
Time (s) Time (s)
292

SC
293 (c) (d)

294 Figure 5- BWT tip deflection (blue dashed line) and lift force (red line) for (a) BWT1, (b) BWT2, (c)
295 BWT3, (d) BWT4.

296 Let us study the effect of the variations of the free-stream velocity on the dynamic response of
U
AN
ωs
297 BWTs. The non-dimensional frequency Ω is introduced as Ω = to indicate the closeness to
ωn
lock-in (lock-in occurs at Ω = 1). Figure 6 and Figure 7 show the response of BWT1 and BWT2
M

298
299 at free stream velocities of 3.6 m/s and V=4.9 m/s that correspond to Ω=0.85 and Ω=1.15,
300 respectively.
D
TE

-3
x 10
8 0.06
Tip deflection (m), Lift force ( × 10 N)

Tip deflection (m), Lift force ( × 10 N)

6
0.04
3

4
EP

0.02
2

0 0

-2
C

-0.02

-4

-0.04
AC

-6

-8 -0.06
0 1 2 3 4 5 0 2 4 6 8 10
Time (s) Time (s)
301
302 (a) (b)

303 Figure 6- BWT1 tip deflection (blue dashed line) and lift force (red line) (a) V=3.6 m/s (b) V=4.9 m/s

17
ACCEPTED MANUSCRIPT

-3
x 10
6 0.08
Tip deflection (m), Lift force ( × 10 N)

Tip deflection (m), Lift force ( × 10 N)


0.06
4
3

3
0.04

2
0.02

0 0

PT
-0.02
-2

-0.04

-4
-0.06

RI
-6 -0.08
0 1 2 3 4 5 0 2 4 6 8 10
Time (s) Time (s)
304
305

SC
(a) (b)

306 Figure 7- BWT2 tip deflection (blue dashed line) and lift force (red line); (a) V=3.7 m/s, (b) V=5 m/s

307 Figure 8 shows the maximum lift coefficient at steady state ( Max {CL ( x, t )} ) for various wind

U
x ,t

308 speeds close to the lock-in region versus the non-dimensional frequency Ω for BWT1 and
AN
309 BWT2. In both turbines, the lift force increases dramatically as the non-dimensional frequency
310 approaches 1. Also in both turbines, at Ω < 1, the lift coefficient ( ) diagram has an upward
M

311 trend with increasing wind speed (i.e. increasing Ω) due to the intensification of the vortices. As
312 the non-dimensional frequency exceeds Ω = 1 (due to high wind speed), although the vortices
D

313 are intensified further, the slope of the lift coefficient diagram reduces significantly and becomes
TE

314 negative due to a mismatch between the vortex shedding frequency and the natural frequency.

315 At Ω < 1, BWT1 produces a larger lift than BWT2, however at Ω > 1 , BWT2 has a larger lift
EP

316 than BWT1 (Figure 8). The reason is that the vortex shedding frequency (ωs) is not constant
317 along the height of BWT2 due to its conic shape. As air speed increases, the vortex shedding
318 frequency at sections closer to the upper end of the turbine matches the natural frequency of the
C

319 structure. The upper end has a large frontal area and a long moment arm; hence the flow induces
AC

320 a dominant resultant moment at the corresponding vortex shedding frequency and subsequently
321 larger deflections. Large deflections in turn enhance the vortex shedding and increase the lift
322 force. However, at low air speeds, the lock-in occurs at the lower end of the turbine which cannot
323 cause a dominant moment at the lock-in frequency due to relatively small frontal area and
324 moment arm. In contrast, in BWT1 the vortex shedding frequencies of all sections along the
325 height are the same, hence the lift force of all sections are in the same frequency. Consequently,

18
ACCEPTED MANUSCRIPT

326 the lift force has only one powerful frequency component that enhances the forced vibrations in
327 low air speeds.

328 In conclusion, in the conic shape BWT2, adaptation of the shedding frequency improves the
329 performance at high wind speeds. However, at low wind speeds, the lift force at a single

PT
330 shedding frequency along the right circular cylinder BWT1 results in a better performance than
331 that of the conic BWT2.

RI
BWT1 BWT2

0.75

SC
Maximum lift coefficient

0.7
0.65
0.6

U
0.55
AN
0.5
0.45
0.85 0.9 0.95 1 1.05 1.1 1.15 1.2 1.25 1.3 1.35 1.4 1.45 1.5
Ω=ωs/ωn
M

332

333 Figure 8- Lift coefficient curves of BWT1 and BWT2


D

334 Figure 9 and Figure 10 show the tip deflection and the lift force acting on the turbine structure at
TE

335 free stream velocities of 3.6 m/s and 4.9 m/s that corresponds to Ω=0.85 and Ω=1.15,
336 respectively. The structure in BWT3 and BWT4 are more flexible comparing with BWT1 and
EP

337 BWT2, therefore the maximum deflections in Figure 9 and Figure 10 are larger than those of
338 Figure 6 and Figure 7. Therefore, the power transmitted rates of BWT3 and BWT4 are
339 considerably higher.
C
AC

340 Due to the conical shape of BWT4, it has a larger moment of inertia comparing with a BWT3
341 with the same mean diameter. Consequently, BWT4 has a lower natural frequency which
342 promotes the lock-in phenomenon in lower air speed comparing to BWT3.

19
ACCEPTED MANUSCRIPT

0.1
0.25
0.08
Tip deflection (m), Lift force ( × 10 N)

Tip deflection (m), Lift force ( × 10 N)


0.2
2

2
0.06
0.15
0.04 0.1

0.02 0.05

0 0

PT
-0.02 -0.05

-0.04 -0.1

-0.15
-0.06
-0.2

RI
-0.08
-0.25
-0.1
0 2 4 6 8 10 0 2 4 6 8 10
Time (s) Time (s)
343
344

SC
(a) (b)

345 Figure 9- BWT3 tip deflection (blue dashed line) and lift force (red line) (a) V=3.6 m/s (b) V=4.9 m/s

U
0.25
0.06
Tip deflection (m), Lift force ( × 10 N)

Tip deflection (m), Lift force ( × 10 N)

0.2
2

AN
0.04 0.15

0.1
0.02
0.05

0 0
M

-0.05
-0.02
-0.1

-0.04 -0.15
D

-0.2
-0.06
-0.25
0 2 4 6 8 10 0 2 4 6 8 10
TE

Time (s) Time (s)


346
347 (a) (b)

348 Figure 10- BWT4 tip deflection (blue dashed line) and lift force (red line) (a) V=3.6 m/s (b) V=4.9
EP

349 m/s

350 Figure 11 shows the maximum lift coefficient at steady state for various wind speeds close to the
lock-in region versus the non-dimensional frequency Ω for BWT3 and BWT4. For both turbines,
C

351
352 the lift force increases dramatically as the non-dimensional frequency approaches 1.
AC

353 Similar to BWT1 and BWT2, in the stand-mast configurations, the conic mast BWT4 has a
354 higher performance at high wind speeds, while, at low wind speeds, the right circular cylinder
355 BWT3 exhibits a better performance. This can be verified from Figure 9 and Figure 10, by
356 comparing the deflections of BWT3 and BWT4 at low and high wind speeds.

20
ACCEPTED MANUSCRIPT

BWT3 BWT4

0.65
Maximum lift coefficient 0.6

0.55

PT
0.5

0.45

RI
0.4
0.85 0.9 0.95 1 1.05 1.1 1.15 1.2 1.25 1.3 1.35 1.4 1.45 1.5
Ω=ωs/ωn

SC
357

358 Figure 11- Lift coefficient curves of BWT3 and BWT4

U
359 Figure 12 illustrates the maximum steady tip deflection of the BWTs in the lock-in region. The
AN
360 flexibly supported turbines (i.e. BWT3 and BWT4) have larger tip deflections due to their
361 flexibility which in turn enhance the vortex shedding and increase the lift force.
M

362 Figure 13 shows the mean power transmitted to the BWTs. We can conclude that in low wind
363 speeds the right circular flexibly mounted turbine (BWT3) can deliver a higher power while in
D

364 high wind speeds the conical flexibly mounted turbine (BWT4) has a better power output. Also
365 note that a similar technology [32] is reported to have a power output of 100 Watts that is close to
TE

366 the power transmitted to BWT3 and BWT4. An exact comparison with [32] is not possible due to
367 the lack of information about the dimensions and material properties of the BWT in [32].
EP

BWT1 BWT2 BWT3 BWT4

0.5
C
BWTs maximum steady state tip

0.4
AC deflection (m)

0.3

0.2

0.1

0
0.85 0.9 0.95 1 1.05 1.1 1.15 1.2 1.25 1.3 1.35 1.4
Ω=ωs/ωn
368

21
ACCEPTED MANUSCRIPT

369 Figure 12- Maximum steady tip deflection response of the BWTs in the lock-in bound.

BWT1 BWT2 BWT3 BWT4

120
BWTs generated power (w)

100

PT
80
60

RI
40
20

SC
0
0.85 0.9 0.95 1 1.05 1.1 1.15 1.2 1.25 1.3 1.35 1.4
Ω=ωs/ωn
370

371
U
Figure 13- Power level transmitted to BWTs in the lock-in bound
AN
372 4 Conclusion
M

373 Four configurations of bladeless wind turbines (BWTs) have been investigated and compared.
374 The BWTs consist of a relatively long (right or conic) cylinder that is either flexible or mounted
D

375 on a flexible structure exposed to a uniform air flow. The nonlinear wake oscillator
semiempirical model was used to obtain an expression for the crosswise flow-induced fluctuating
TE

376
377 lift force due to periodically shedding vortices. To derive nonlinear distributed-parameter models
378 of the BWTs under the fluctuating lift force, the Euler-Bernoulli beam theory and the Galerkin
EP

379 procedure were utilized. The predictions of the lift force and the deflection of the BWT based on
380 the derived model were in a good agreement with the results of the 3D CFD-FEM numerical
C

381 simulations. The power outputs of BWTs were studied in various wind speeds through numerical
382 simulations of the derived models. It has been confirmed that the amplitudes of the vibrations of
AC

383 the BWTs surged when the vortex shedding frequency matches the natural frequency of the
384 structure, which is known as lock-in phenomenon. Furthermore, the results of the simulations
385 demonstrated that, the conic BWTs have a higher performance at post-synchronization region
386 (i.e. high wind speeds); whereas the right circular cylinder BWTs exhibits a better performance
387 at pre-synchronization region (i.e. low wind speeds). It was also demonstrated that mounting the

22
ACCEPTED MANUSCRIPT

388 bluff body on a flexible structure, in comparison with using a flexible bluff body, results in a
389 significant increase in the power transmitted to the BWT.

390 5 References

PT
391 [1] Williamson CH, Govardhan R. A brief review of recent results in vortex-induced vibrations. Journal of Wind
392 Engineering and Industrial Aerodynamics. 2008 Jul 31;96(6):713-35.

393 [2] Rostami AB, Armandei M. Renewable energy harvesting by vortex-induced motions: review and benchmarking

RI
394 of technologies. Renewable and Sustainable Energy Reviews. 2017 Apr 30;70:193-214.

395 [3] Abdelkefi A. Aeroelastic energy harvesting: A review. International Journal of Engineering Science. 2016 Mar
396 31;100:112-35.

SC
397 [4] Khan AA, Shahzad A, Hayat I, Miah MS. Recovery of flow conditions for optimum electricity generation
398 through micro hydro turbines. Renewable Energy. 2016 Oct 31;96:940-8.

399 [5] Khan AA, Khan AM, Zahid M, Rizwan R. Flow acceleration by converging nozzles for power generation in

U
400 existing canal system. Renewable energy. 2013 Dec 31;60:548-52.
AN
401 [6] Akaydin HD, Elvin N, Andreopoulos Y. The performance of a self-excited fluidic energy harvester. Smart
402 materials and Structures. 2012 Jan 24;21(2):025007.

403 [7] Gabbai RD, Benaroya H. An overview of modeling and experiments of vortex-induced vibration of circular
404 cylinders. Journal of Sound and Vibration. 2005 Apr 22;282(3):575-616.
M

405 [8] Abdelkefi A, Nayfeh AH, Hajj MR. Design of piezoaeroelastic energy harvesters. Nonlinear Dynamics. 2012
406 Jun 1;68(4):519-30.
D

407 [9] Erturk A, Vieira WG, De Marqui Jr C, Inman DJ. On the energy harvesting potential of piezoaeroelastic systems.
408 Applied Physics Letters. 2010 May 3;96(18):184103.
TE

409 [10] Abdelkefi A, Hajj MR, Nayfeh AH. Piezoelectric energy harvesting from transverse galloping of bluff bodies.
410 Smart Materials and Structures. 2012 Dec 10;22(1):015014.

411 [11] Yang Y, Zhao L, Tang L. Comparative study of tip cross-sections for efficient galloping energy harvesting.
EP

412 Applied Physics Letters. 2013 Feb 11;102(6):064105.

413 [12] Bernitsas MM, Raghavan K, Ben-Simon Y, Garcia EM. VIVACE (Vortex Induced Vibration Aquatic Clean
414 Energy): A new concept in generation of clean and renewable energy from fluid flow. Journal of offshore mechanics
C

415 and Arctic engineering. 2008 Nov 1;130(4):041101.

416 [13] Cajas JC, Houzeaux G, Yáñez DJ, Mier-Torrecilla M. SHAPE Project Vortex Bladeless: Parallel multi-code
AC

417 coupling for Fluid-Structure Interaction in Wind Energy Generation.

418 [14] Skop RA, Balasubramanian S. A new twist on an old model for vortex-excited vibrations. Journal of Fluids and
419 Structures. 1997 May 1;11(4):395-412.

420 [15] Hasheminejad SM, Rabiee AH, Jarrahi M. Semi-Active Vortex Induced Vibration Control of an Elastic
421 Elliptical Cylinder with Energy Regeneration Capability. International Journal of Structural Stability and Dynamics.
422 2017 Feb 7:1750107.

423 [16] Wu T, Kareem A. Vortex-induced vibration of bridge decks: Volterra series-based model. Journal of
424 Engineering Mechanics. 2013 Mar 16;139(12):1831-43.

23
ACCEPTED MANUSCRIPT

425 [17] Olinger DJ. A low-order model for vortex shedding patterns behind vibrating flexible cables. Physics of Fluids.
426 1998 Aug;10(8):1953-61

427 [18] Monkewitz PA, Williamson CH, Miller GD. Phase dynamics of Kármán vortices in cylinder wakes. Physics of
428 Fluids. 1996 Jan;8(1):91-6.

429 [19] Bishop RE, Hassan AY. The lift and drag forces on a circular cylinder oscillating in a flowing fluid.
430 InProceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 1964 Jan 7

PT
431 (Vol. 277, No. 1368, pp. 51-75). The Royal Society.

432 [20] Facchinetti ML, De Langre E, Biolley F. Coupling of structure and wake oscillators in vortex-induced
433 vibrations. Journal of Fluids and structures. 2004 Feb 29;19(2):123-40.

RI
434 [21] Skop RA, Griffin OM. A model for the vortex-excited resonant response of bluff cylinders. Journal of Sound
435 and Vibration. 1973 Mar 22;27(2):225-33.

SC
436 [22] Griffin OM. Vortex-excited cross-flow vibrations of a single cylindrical tube. Journal of Pressure Vessel
437 Technology. 1980 May 1;102(2):158-66.

438 [23] Dahleh MD, Thomson WT. Theory of vibration with applications. Prentice-Hall Inc. 1998.

U
439 [24] Dai HL, Abdelkefi A, Yang Y, Wang L. Orientation of bluff body for designing efficient energy harvesters
440 from vortex-induced vibrations. Applied Physics Letters. 2016 Feb 1;108(5):053902.
AN
441 [25] Rao SS. Vibration of continuous systems. John Wiley & Sons; 2007 Feb 9.

442 [26] Besem FM, Thomas JP, Kielb RE, Dowell EH. An aeroelastic model for vortex-induced vibrating cylinders
443 subject to frequency lock-in. Journal of Fluids and Structures. 2016 Feb 29;61:42-59.
M

444 [27] Sanchez R, Palacios R, Economon TD, Kline HL, Alonso JJ, Palacios F. Towards a Fluid-Structure Interaction
445 solver for problems with large deformations within the open-source SU2 suite. AIAA SciTech. 2016.
D

446 [28] Song L, Fu S, Cao J, Ma L, Wu J. An investigation into the hydrodynamics of a flexible riser undergoing
447 vortex-induced vibration. Journal of Fluids and Structures. 2016 May 31;63:325-50.
TE

448 [29] Triantafyllou MS, Grosenbaugh MA, Gopalkrishnan R. Vortex-Induced Vibrations in a Sheared Flow: a New
449 Predictive Method. WOODS HOLE OCEANOGRAPHIC INSTITUTION MA; 1994.

450 [30] Protos A, Goldschmidt VW, Toebes GH. Hydroelastic forces on bluff cylinders. Journal of Basic Engineering.
EP

451 1968 Sep 1;90(3):378-86.

452 [31] Hasheminejad SM, Jarrahi M. Numerical simulation of two dimensional vortex-induced vibrations of an elliptic
453 cylinder at low Reynolds numbers. Computers & Fluids. 2015 Jan 31;107:25-42.
C

454 [32] Whitlock R. The Power of the Vortex: An Interview. Renewable Energy Magazine. 2015 April 07.
AC

455 [33] Naguleswaran S. A direct solution for the transverse vibration of Euler-Bernoulli wedge and cone beams.
456 Journal of Sound and Vibration. 1994 May 5;172(3):289-304.

457 Appendix

458 Since the BWT2 is conical, the moment of inertia and the cross section area of the turbine are
459 functions of the length  (Figure 14), given by:

24
ACCEPTED MANUSCRIPT

4
 x  (36)
I (x ) = I 0 1 + c ( ) 
 L 
2
 x  (37)
A (x ) = A 0 1 + c ( ) 
 L 

where , and Y are the moment of inertia and the cross section area at  = 0, and Z is the taper

PT
460
461 ratio.

RI
U SC
462 AN
463 Figure 14- The tapered conical BWT.

464 Therefore, combining (2) and (14), the equations of motion are written as:
M

( )
E I ′′ ( x)Y ′′( x, t ) + 2 I ′( x) Y ( 3) ( x, t ) + I ( x)Y (4) ( x, t ) + ρ s A( x)
∂2 y
∂t 2
(38)
1  2α ɺ 
D

= ρ V 2  D( x)Q( x, t ) − Y ( x, t ) 
2  ωs 
TE

ɺ
ɺɺ( x, t ) − ω G (C 2 − 4Q 2 ( x, t ))Qɺ ( x, t ) + ω 2Q( x, t ) = ω F Y ( x, t )
Q (39)
s L0 s s
D( x)
EP

465 Reduced-order model for the BWT2 configuration can be obtained using the same approach as of
466 BWT1. However, the natural frequencies and the mode shapes are different due to the non-
467 uniform diameter of the structure. Analytical solution of the mode shape equations using Bessel
C

468 functions is presented in [33]. Considering a fixed-free boundary condition for the truncated
AC

469 conical cantilever, the first natural frequency and the corresponding mode shape of BWT2 can be
470 calculated using the method of [33].

471 Using (36)-(39), the following coupled ODEs of motion are derived for BWT2,

25
ACCEPTED MANUSCRIPT

( )
Ey ( t ) ∫ φ ( x) ( I ′′( x)φ ′′( x)) + (2 I ′( x)φ ( 3) ( x)) + ( I ( x)φ (4) ( x)) dx + ρ s ɺɺ
y ( t ) ∫ A( x)φ 2 ( x)dx
L L

0 0

1  L 2α  (40)
yɺ ( t ) ∫ φ 2 ( x)dx 
L
= ρ V 2  q ∫ D( x)φ 2 ( x)dx −
2 
0 ωs 0

φ 2 ( x)
( qɺɺ( t ) − ω GC qɺ ( t ) + ω q ( t ) ) ∫ φ 2 ( x)dx + {ωs G 4q 2 ( t ) qɺ ( t )} ∫ φ 4 ( x)dx = ωs Fyɺ ( t ) ∫
L L L
2 2

PT
s L0 s dx (41)
0 0 0 D( x)

RI
472

U SC
AN
M
D
TE
C EP
AC

26
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

Highlights
• The governing equation of a bladeless wind turbine (BWT) is derived.
• The model has been validated through a 3D CFD-FEM numerical simulation.
• The effects of the wind speed on the performance of BWT are investigated.
• The conic BWTs have a higher performance at high wind speeds.
• The right circular cylinder BWTs exhibits a better performance at low wind speeds.

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like