21
Aeroelastic instability analysis of composite rotating blades based on loewy’s and theodorsen’s unsteady aerodynamics *Touraj Farsadi 1) and Altan Kayran 2) 1), 2) METUWind centre, Department of Aerospace Engineering, Middle East Technical University, Ankara, Turkey 1) [email protected] ABSTRACT Classical aeroelastic stability analysis approach is presented for the simplified composite blade model. For the purpose of the study, the composite wind turbine blade is modeled as an elastic cantilevered rotating thin-walled composite box beam with the developed Circumferentially Asymmetric Stiffness (CAS) structural model. Circumferentially Asymmetric Stiffness structural model takes into account non- classical effects such as transverse shear, material anisotropy and warping inhibition. For the aeroelastic stability analysis, aerodynamic analysis approaches used in the present study are based on classical Theodorsen’s theory and Loewy’s returning wake method in the frequency domain in conjunction with the CAS structural model. Hamilton’s principle and extended Galerkin method are used to obtain the coupled system of equations which are then posed as an eigenvalue problem. Results show that Theoderson and Loewy theories give close results in low rotor speeds but when the rotor speed is increased, the difference between these two methods becomes apparent. It is also shown that the fiber angle of the CAS structural model affects the aeroelastic instability speed. 1. INTRODUCTION Wind energy conversion is a fast growing source among renewable energies in the world. Wind turbines are getting larger for past ten years to capture more energy from the wind. Increasing the size of the turbines has resulted in a price reduction for the electricity per kWh, but it has also created problems associated with the length of the blades. One probable problem area is the aeroelastic stability problems of long wind turbine blades due to the increased bending and torsional flexibility. Increased bending and torsional flexibility of long wind turbine blades may cause torsional divergence and flapwise bending-torsion flutter at high speeds. Stiffness tailoring is one way to control the elastic deformation of blades made of composite materials in a 1) PhD Candidate 2) Professor

Aeroelastic instability analysis of composite rotating ...€¦ · Aeroelastic instability analysis of composite rotating blades based on loewy’s and theodorsen’s unsteady aerodynamics

Embed Size (px)

Citation preview

  • Aeroelastic instability analysis of composite rotating blades based on

    loewys and theodorsens unsteady aerodynamics

    *Touraj Farsadi1) and Altan Kayran2)

    1), 2) METUWind centre, Department of Aerospace Engineering,

    Middle East Technical University, Ankara, Turkey 1)

    [email protected]

    ABSTRACT

    Classical aeroelastic stability analysis approach is presented for the simplified composite blade model. For the purpose of the study, the composite wind turbine blade is modeled as an elastic cantilevered rotating thin-walled composite box beam with the developed Circumferentially Asymmetric Stiffness (CAS) structural model. Circumferentially Asymmetric Stiffness structural model takes into account non-classical effects such as transverse shear, material anisotropy and warping inhibition. For the aeroelastic stability analysis, aerodynamic analysis approaches used in the present study are based on classical Theodorsens theory and Loewys returning wake method in the frequency domain in conjunction with the CAS structural model. Hamiltons principle and extended Galerkin method are used to obtain the coupled system of equations which are then posed as an eigenvalue problem. Results show that Theoderson and Loewy theories give close results in low rotor speeds but when the rotor speed is increased, the difference between these two methods becomes apparent. It is also shown that the fiber angle of the CAS structural model affects the aeroelastic instability speed. 1. INTRODUCTION Wind energy conversion is a fast growing source among renewable energies in the world. Wind turbines are getting larger for past ten years to capture more energy from the wind. Increasing the size of the turbines has resulted in a price reduction for the electricity per kWh, but it has also created problems associated with the length of the blades. One probable problem area is the aeroelastic stability problems of long wind turbine blades due to the increased bending and torsional flexibility. Increased bending and torsional flexibility of long wind turbine blades may cause torsional divergence and flapwise bending-torsion flutter at high speeds. Stiffness tailoring is one way to control the elastic deformation of blades made of composite materials in a

    1)

    PhD Candidate 2)

    Professor

    mailto:[email protected]

  • structural optimization framework. Tangler (2000) pointed out that composite material could be used in the wind turbine blade to efficiently sustain the complex loading for long blades. A Thin-Walled Beam (TWB) is a slender structural element whose distinctive geometric dimensions are all of different orders of magnitude such that its thickness is small compared to the cross-sectional dimensions, while its length greatly exceeds the dimensions of its cross-section (Librescu 2006). Composite TWBs are used in jet engines, tilt rotor aircrafts, helicopters, turbo machinery, and in wind turbines as rotor blades, and their dynamic behavior is the topic of many studies. In the case of the rotating beam, Song (2001) performed the free vibration analysis of rotating pre-twisted TWBs, incorporated flaplag elastic coupling and adaptive capabilities using the Extended Galerkins method (EGM). Sina (2011) investigated the rotation effects in eigenvalue analysis of single cell-laminated composite TWB with closed cross-section. The effects of rotation, ply angle, taper ratio, slenderness, and hub ratios on natural frequencies and mode shapes of rotating TWB with flaptwist elastic coupling are studied. In wind turbine blades, for the aerodynamic solution, most codes use the Blade Element Momentum (BEM) method, as described by Glauert (1963). Blade element moment method is very fast and yields accurate results, provided that reliable airfoil data exist. Since wind turbines operate in unsteady flow environment for most of their time, it is important for the analyst to recognize that many of the tools to model unsteady aerodynamic effects on airfoils have already been laid down. Results for incompressible, unsteady airfoil problems are formulated in the frequency domain, primarily by Theodorsen (1935) and Loewy (1957). An authoritative source documenting these classical theories is given by Leishman (2000). These solutions have the same root in unsteady thin-airfoil theory, and give exact analytic solutions for airloads for different forcing conditions. Classical flutter is defined as the violent instability when the torsional mode of the blade couples with the flap-wise bending mode, resulting in rapid growth of the amplitude of the flap-wise and torsional motions (Politakis 2008 and Vatne 2011). Pitch-flap flutter and divergence are very similar to the classical flutter and divergence of fixed wing airplanes. Of course, the centrifugal action on the wind turbine blade is not part of the classical flutter of fixed wing aircraft. It should be noted that due to the centrifugal forces, the flapwise stiffness effectively increases in rotor blades. Janetzke (1983) at the NASA Lewis Research Center published the first paper directly related to the aeroelasticity of the wind turbine. Their study explored the possibility of whirl flutter and searched the effect of pitch-flap-coupling on the teetering motion of a 2 blade wind turbine. In 1998, Sandia Laboratories & National Renewable Energy Laboratory published an article on the aeoelastic tailoring of the wind turbine blade (Veers 1998). In this study, aeroelastic property of the blades is utilized to shape the power curve and reduce loads. Early and recent investigations of stability in wind turbine blades address the issue of classical flutter for smaller rotors (10m blades) (Lobitz 1998). Lobitz (2004) investigated the flutter limit of a MW sized wind turbine blade based on isolated blade stability analysis using quasi-steady and unsteady Theodorsen aerodynamics. In this study, it is shown that the predicted flutter speed of the blade using quasi-steady aerodynamics is lower than the flutter speed obtained using unsteady aerodynamics. Hansen (2004) indicated that the flutter rotational speed

  • is in the realm of two times the operational speed of the rotor. Recently, Farsadi (2016) demonstrated the aeroelastic stability analysis of wind turbine blades using the thin walled beam structural model and Wagners function in a coupled analysis framework both in frequency and time domain. In the present study, a unified model of aeroelastic systems in incompressible flow regime using aerodynamic states based on the theories of Theodorsen and Loewy is presented. The main goal of the study is to compare the effect of using Theodorsens unsteady aerodynamics and Loewys method on the aeroelastic stability characteristics of wind turbine blades. It should be noted that Theodorsens aerodynamics neglects the effect of the wake formed by the preceding blade, whereas Loewys method includes the effect of returning wake of the preceding blade. Using the two different aerodynamic models, numerical results are obtained for the aeroelastic stability of wind turbine blades modeled as rotating composite TWBs with the previously developed Circumferentially Asymmetric Stiffness (CAS) structural model. Circumferentially asymmetric stiffness structural model takes into account a group of non-classical effects such as the transverse shear, the material anisotropy and warping inhibition. Hamiltons principle and the extended Galerkin method are used to obtain the coupled linear governing system of dynamic equations. Governing equations are posed as eigenvalue problems involving Theodorsen and Loewys unsteady aerodynamics in the

    frequency domain. U-g method is used to determine the instability limit of the wind turbine blade. 2. ROTATING THIN WALLED BEAM The analysis of rotating blade structures is more complex than that of their nonrotating counterparts. In the rotating case, in addition to the accelerations resulting from elastic structural deformations, the centrifugal and Coriolis accelerations have to be included in the modeling. In this present study, the structural model is similar to those developed by Librescu (2006) and Sina (2011). Inertial reference coordinate is attached to the hub center, while the local coordinate system is fixed to the blade and used to define the complex cross-sectional properties, as shown in Figs. 1, 2.

    Fig. 1 Schematic description of the rotating blade structure simulated by

    "Rotating Thin-Walled Composite Beam"

    Fig. 2 Cross section of the TWB and the

    displacement field

  • Notice that for the thin walled beam theory considered herein, the six kinematic

    variables, u0(z,t), v0(z,t), w0(z,t), x(z,t), y(z,t), (z,t) defined with respect to the rotating local coordinate system in Fig. 2, represent 1-D displacement measures. a(s) and rn(s) shown in Fig. 2 are the perpendicular distances from the shear center to the normal and to the tangent of the mid-line beam contour and they are defined in Eqs. (4) and (5). The present structural model of the rotating thin-walled composite box beam, shown in Fig. 3, includes some non-classical effects such as material anisotropy, transverse shear, primary and secondary warping inhibition, non-uniform torsional model and rotary moment of inertia. Also, Circumferentially Asymmetric Stiffness (CAS) method is chosen among various lay-up methods in order to induce bending-twisting coupling in the composite rotor blade, as shown in Fig. 3. CAS ply-angle configuration, achievable via the usual filament winding technology, results in an exact decoupling between the extension - twist on one hand and the flap shear elastic coupling on the other hand. It should be noted that in the present study, thin walled beam is taken as non-tapered beam with no pre-twist in order to not to complicate the equations further. However, solution method based on extended Galerkin method is general and also applicable to tapered and pre-twisted beams.

    Fig. 3 Thin walled- composite beam and

    the CAS configuration

    Fig. 4 Schematic description of the wing structure and aerodynamic and structural

    coordinates

    Figs. 3 and 4 show that two coordinate systems exist, (x,y,z) as the local

    coordinate associated with the blade and (n,s,z) used to define complex cross-section

    profiles. The angular velocity of the rotor is assumed to be constant in the global y

    direction. The position vector " R " of a point in the deformed rotating beam, measured from the centre of the hub, can be expressed as:

    = + +D

    = = + +

    D = + +

    0

    0 0 ,u

    u

    R R R

    R R k R xi yj zk

    ui vj wk

    (1)

  • Considering angular velocity () and Eq. (1), "R&" and "R&&" can be written as:

    0

    2 20

    ( ) ( )

    2 ( ) 2 ( )

    R u R z w i vj w x u k

    R u w x u i vj w u R z w k

    = +W + + + + - W + = + W- + W + + - W- + + W

    & & &&

    && & && &&&& &

    where " R0" and " Ru " are the hub radius and the undeformed position vector of a beam point, respectively, and "" represents the displacement vector, whose components are defined in Eq. (1).

    The 3-D displacement quantities ( ), ,u v w are assumed to be: 0

    0

    '0

    ( , , , ) ( , ) ( , )

    ( , , , ) ( , ) ( , )

    ( , , , ) ( , ) ( , ) ( ) ( , ) ( ) ( , ) ( ) ( )x y w

    u x y z t u z t y z t

    v x y z t v z t x z t

    dx dyw x y z t w z t z t y s n z t x s n z t F s na s

    ds ds

    = -

    = +

    = + - + + - +

    f

    f

    q q f

    where

    '0

    '0

    ( , ) ( , ) ( , )

    ( , ) ( , ) ( , )

    ( ) ( ) ( )

    x yz

    y xz

    z t z t v z t

    z t z t u z t

    dy dxa s y s x s

    ds ds

    q g

    q g

    = -

    = -

    = - -

    The primary Warping Function ( wF ) is expressed as (Librescu 2006 and Sina 2011):

    0

    ( )

    ( )[ ( ) ] , , ( ) ( ) ( )

    ( )

    ns

    Cw n n

    C

    r sds

    h s dy dxF r s ds r s x s y s

    ds ds ds

    h s

    = - Y Y = = -

    The axial strains associated with the displacement field are given by Librescu (2006) and Sina (2011):

    0

    0 ' ' ' ''0

    ' ' ''

    ( , , , ) ( , , ) ( , , )

    ( , , ) ( , ) ( , ) ( ) ( , ) ( ) ( , ) ( )

    ( , , ) ( , ) ( , ) ( , ) ( )

    nzz zz zz

    zz x y w

    nzz y x

    n s z t s z t n s z t

    s z t w z t z t y s z t x s z t F s

    dy dxs z t z t z t z t a s

    ds ds

    e e e

    e q q f

    e q q f

    = +

    = + + -

    = - -

    where prime denotes derivative with respect to the z coordinate. The tangential shear strain components can be defined as (Librescu 2006 and Sina 2011):

    0 '

    0 ' '0 0

    ' '0 0

    ( , , ) ( , , ) 2 ( , )

    ( , , ) [ ( , ) ( , )] [ ( , ) ( , )]

    ( , , ) [ ( , ) ( , )] [ ( , ) ( , )]

    Csz sz

    sz y x

    nz y x

    As z t s z t z t

    dx dys z t u z t z t v z t z t

    ds dsdy dx

    s z t u z t z t v z t z tds ds

    g g fb

    g q q

    g q q

    = +

    = + + +

    = + - +

    (2)

    (3)

    (4)

    (5)

    (6)

    (7)

  • where 0 0,zz sze g are the normal strain and the in-plane shear strain components at

    the mid-surface of the thin-walled box beam, respectively. The kinetic energy, the strain energy and the work of external forces are calculated using the strain-displacement relationships, sectional effective stiffness matrix and external forces. By substituting these expressions into the Hamiltons Principle, given by Eq. (8), the governing system of equations can be obtained. The rotating thin-walled composite box beam theory produces a linear relationship between the sectional structural loads and the strain measures.

    Hamilton`s principle and variational formulation can be written as shown in Eq. (8) 2

    1

    1 2 0 0 0( ) 0 , 0

    t

    x y

    t

    T V W dt at t t t u v w- + = = = = = = = = d d d d d d dq dq df

    The kinetic energy of the system is given by Eq. (9).

    T R R dndsdzd r d= ? & & The potential energy of the system is given by Eq. (10).

    ( )10 ( )

    1[ ]

    2

    L N

    ij ij zz zz sz sz nz nx kkC h k

    V dndsdz dndsdz=

    = = + +

    d s de d s e s e s e

    The work of the external forces is given by Eq. (11).

    [ ]00

    ( , ) ( , ) ( , ) ,

    L

    ae aeW L z t v z t T z t dzd d df= +

    Where Lae and Tae are the unsteady aerodynamic lift and pitching moment about the reference axis. Substituting Eqs. (2) to (7) into Eqs. (9) and (10) and using Hamiltons principle, one gets Eqs. (12) and (13) for the variation of the potential and kinetic energy.

    0

    0 0 0 0 0

    0

    0

    Lz y x y x y x w z r

    x z y z

    z y y x x w w z r

    x z

    T w M Q M Q B M TV dz

    Q T u u Q T v v

    T w M M B B M T

    Q T u

    d dq dq f dfd

    d d

    d dq dq df f df

    ?+ - + - + + + + = - + + + + ?+ + - + + + +

    +

    0 0 0 0

    L

    y zu Q T v vd d

    + +

    2 2

    1 0 0 0 0 1 0 0 1 0 0 0 0 0

    2 25 15 4 5 4

    2 24 5 4 4 5 10 18

    2 2

    2

    2

    y y y x x x

    x

    b u w u u bv v b w u R z w w

    T b b b b b

    b b b b b b b

    d d d

    d q q dq q q f dq

    f q f f

    + W- W + + - W- + + W + = - + - W + + - W + W +

    + - W - - W - + - W

    & && &&&& &

    && && &

    && & &&0

    210 18 0

    L

    L

    dz

    b b

    f df

    f f df

    -

    + - W

    &&

    (9)

    (10)

    (11)

    (8)

    (13)

    (10)

    (11)

    (12)

    (13)

  • The Euler Lagrange equations can be obtained from Eqs. (8), (12) and (13) and they are given by Eq. (14).

    20 0 1 0 0 0: 2 0x zu Q Tu b u w ud

    + - + W- W = &&&

    0 0 1: 0y zv Q Tv bvd+ - =

    20 1 0 0 1 0 0: T 2 0yw b w u b R z wd - - W + W + + = && &

    25 15: M 0y y x y yQ b bdq q q

    - - + - W = &&

    24 14 4: 2 0x x y x xM Q b b bdq q q f

    - - + - W - W = && &

    2 24 5 4 5 4 10 18: B 2 0w z r xM T b b b b b b bdf f f f q f f

    + + - + + - W + W+ + - W = && & &&

    In Eq. (14), one-dimensional stress measures, , , , , , ,z r x y y x zT T Q Q M M M and wB

    can be defined in terms of stress resultants and stress couples as:

    ( , )z zzT z t N ds= ( , ) ( , )r z pT z t T z t I=

    ( , )x sz zndx dy

    Q z t N N dsds ds

    = +

    ( , )y sz zndy dx

    Q z t N N dsds ds

    = -

    ( , )x zz zzdx

    M z t yN L dsds

    = -

    ( , )y zz zzdy

    M z t xN L dsds

    = +

    ( , ) 2z sz szM z t N L dsy= +

    ( , ) ( ) ( )w w zz zzB z t F s N a s L ds= +

    where , ,zz sz znN N N are the stress resultants and szL and zzL are the stress couples defined by Eq. (16).

    2 2 2

    2 2 2

    , 1, , , 1, , 1,

    t t t

    zz zz zz sz sz sz nz nz

    t t t

    N L n dn N L n dn N n dns s s

    - - -

    = = = ?

    The reduced mass terms, ,i pb I in Eqs. (14) and (15) are defined in the work Sina

    (2011) Boundary conditions at two edges of the blade can also be obtained from Hamiltons principle as:

    (14-a)

    (15)

    (14-b)

    (14-c)

    (14-d)

    (14-e)

    (14-f)

    (16)

  • 0 0

    0 0

    0

    210 18

    0 0

    0 0

    0 0

    0 0

    0 0

    0 0

    0 0

    x z

    y z

    z

    x x

    y y

    w z r

    w

    u or Q T u

    v or Q T v

    w or T

    or M

    or M

    or B M T b b

    or B

    d

    d

    d

    dq

    dq

    df f f f

    df

    = + =

    = + =

    = =

    = =

    = =

    = + + + + - W = = =

    &&

    In Eq. (17), the left-hand side boundary conditions are called as essential boundary conditions and the right-hand side ones are called as natural boundary conditions. Constitutive relations for a general orthotropic material can be written as:

    11 12 13 16

    12 22 23 26

    13 23 33 36

    44 45

    45 55

    16 26 36 66

    0 0

    0 0

    0 0

    0 0 0 0

    0 0 0 0

    0 0

    ss ss

    zz zz

    nn nn

    zn zn

    sn sn

    sz sz

    C C C C

    C C C C

    C C C C

    C C

    C C

    C C C C

    s e

    s e

    s e

    s g

    s g

    s g

    ? ? ? ? ? ? ? ? ? ? ?= ? ? ? ?

    where ijC represent stiffness coefficients.

    The stress resultants " `N s " and stress couples " `L s " can be reduced to the expressions given by Eqs. (19) and (20).

    0

    011 12 13 14

    '21 22 23 24

    244 45 55

    zz

    zz sz

    sz

    nzz

    nz nz

    N K K K K

    N K K K K

    N A A A

    e

    g

    f

    e

    g

    = = -

    0

    041 42 43 44

    '51 52 53 54

    zz

    zz sz

    sz

    nzz

    L K K K K

    L K K K K

    e

    g

    f

    e

    =

    Where reduced stiffness coefficients ( )ijK are defined in Farsadi (2016).

    Taking into account the present CAS structural configuration, the entire system of

    equations splits exactly into flap / torsion / vertical transverse shear 0, , xv f q , and

    extension / lateral bending / lateral transverse shear0 0, , yu w q

    , respectively as shown in

    Eq. (14).

    (19)

    (20)

    (17)

    (18)

  • 3. UNSTEADY AERODYNAMIC MODELS

    3.1 Theodorsens unsteady aerodynamics The aerodynamics for a single blade is similar to that of a fixed wing with a free stream velocity that varies linearly from the root to the tip, assuming that the shed wake of the preceding blade dies out sufficiently fast so that the oncoming blade will encounter essentially still air (Lobitz 2004). Based on Theodorsen aerodynamics, for the general motion of thin airfoil of chord

    length 2b undergoing a combination of pitching and plunging motion in a flow of steady

    velocity U, unsteady lift and pitching moment about the reference axis are given by Eqs. (21) and (22), respectively.

    2 2

    3 2

    1( ) 2 ( )( ( ) ]

    2

    1 1( ( ) ( ) )

    2 8

    ae rel rel rel

    noncirculatorycirculatory

    ae rel

    noncirculatory

    L b v ba U bU C k U b a v

    T b av U a b a

    pr f f pr f f

    pr f f

    = - + + + + - +

    = - - - + +

    && & &&& &14444444444244444444443 14444444444444444244444444444444443

    & &&&&1 2

    2 1 12 ( ) ( )( ( ) )2 2rel rel

    circulatory

    U b a C k v U b apr f f+ + + + - &&44444444444444444 444444444444444443 14444444444444444444244444444444444444443

    The first term in Eq. (21) is the non-circulatory or apparent mass part, which

    results from the flow acceleration effect. The second group of terms is the circulatory components arising from the generation of circulation about the airfoil. Theodorsens

    function, also named as lift deficiency function, C(k) is a complex-valued function which

    depends on the reduced frequency k, ( b/ U). Theodorsens function is defined in terms of Hankel functions by Eq. (22).

    = +

    (2)1

    (2) (2)1 0

    ( )H

    C kH iH

    3.2 Loewys unsteady aerodynamics Loewy (1957) postulated a two dimensional model representing the aerodynamics of an oscillating rotary wing operating at low inflow. Loewys rotary wing unsteady aerodynamics theory is an incompressible theory, somewhat similar to the Theodorsens theory, but main difference is that in Loewy approach, the effect of the spiral returning wake behind the rotor is taken into account approximately. Fig. 5 is a schematic of Loewys two-dimensional model that is used to determine the effects of previously shed wakes on the lift deficiency function. Loewy assumed that there are infinite number of wakes behind the reference blade and applied the Biot-Savart law to each layer of shed vorticity to add together the effects on the differential downwash equation. Two indices that are used to account for the vorticity shed by a given wake are n, which indicates the revolution number of the reference blade, and q which indicates the blade number that wake belongs to.

    (22)

    (21)

  • Fig. 5 Loewys unsteady aerodynamic model for multi-blade rotor system (Couch 2003)

    Loewys theory considers only steady simple harmonic motion which has taken place over an infinite period of time. Therefore, Loewys theory assumes infinite number of wake layers behind the rotor. Loewy has shown that the unsteady aerodynamic lift and moment can be written in a form identical to the classical Theodorsen theory, except that the lift deficiency function must be replaced by the more complicated lift deficiency function denoted by , ,C k h m . In the new definition of the lift deficiency

    function, k is the reduced frequency and h is the wake spacing, which is a function of the period of the rotor revolution and the inflow velocity, and m is frequency ratio. Derivation of Loewys function is similar to the Theodorsens function except the terms which account for the vorticity generated by the reference blade and subsequent blades in previous revolutions. Loewys function in terms of Hankel functions is given by Eq. (23) (Loewy 1957).

    2

    1 1

    2 2

    1 0 1 0

    2 , ,, ,

    2 , ,

    H k J k W k h mC k h m

    H k iH k J k iJ k W k h m

    where

    1

    2 2, , 1 , ,

    kh im kr UW k h m e m h

    Qb bQ

    In Eq. (24), r is radial position of the airfoil, R is the radius of the blade, Q

    is the number of blades, b is the semi-chord, is the rotational velocity of the rotor, U is the inflow velocity. Comparing Loewys function with the Theodorsens function shows that additional terms in the numerator and the denominator are scaled

    by the weighting factor W. For large wake spacing, weighting factor aproaches zero

    (24)

    (23)

    (24)

  • and the Loewys function reduces to the Theodorsen function, as expected. Fig. 6

    compares the real parts and imaginary parts of the Theodorsens and the Loewys

    functions for U=15 m/s, =4 RPM and Q=3, respectively. The limitations of the Loewy lift deficiency function are that it is restricted to a rotary blade such as wind turbine blade with not high inflow velocity. For the case investigated in Fig. 6 Loewy function can be used in a reasonable reduced frequency range between 0 and 1. It can be seen

    that when k 1 which corresponds to high inflow, the effects of any shed layer of vorticity is negligible, and Theoderson and Loewy functions give the same results.

    Fig. 6 Comparison of real and imaginary parts of Theodorsens and Loewys

    functions versus reduced frequency for r/R=0.1, U=15m/s,Q=3,=4RPM

    Loewy's function changes the problem from that of a thin airfoil in harmonic motion to that of a rotary wing, undergoing harmonic motion. For high inflow, Loewys function approaches the Theodorsens function. For the high inflow case, it would be expected that all shed vorticity beyond a small fraction of a revolution would be too far below the reference blade to have a significant effect. The most important result from Loewy is that the wake geometry and phasing is the primary cause of unsteady rotor loading. 4. GOVERNING SYSTEM EQUATIONS OF THE AEROELASTIC SYSTEM

    In the present study, an anisotropic, rotating thin walled composite box beam is used in order to study the effect of employing the Theodorsens function and Loewys returning wake on the aeroelastic stability of the wind turbine blade in incompressible flow. For the derivation of the governing system of equations, the kinetic energy, the strain energy and the work by external forces of the wind turbine blade structure are calculated using the strain displacement relations, the sectional effective stiffness matrix and the external loads. Furthermore, the assumption of small deformations and small strain theory results in the linear relationship between the cross sectional external loads and the strain measures. The present rotating thin walled composite box beam

  • model, by means of the Hamilton`s principle and the variational calculus, yields two sets of governing system of equations. The first set is elastically coupled by the "flap / torsion / vertical transverse shear" deformation modes and the second set is elastically coupled by the "extension / lateral bending / lateral transverse shear" deformation modes. In this study, the second set is not taken into account. The equations of motion corresponding to the first set, are given by Eq. (25).

    '' ' '''0 55 0 56 0 1 0

    '' '' ' '' 233 37 55 0 56 4 14

    ''' '' '''' '' '' 265 0 66 73 77 4 5 4 5

    10 18

    0: ( ) ( )

    0: ( ) ( )( )

    0: ( ) ( ) ( ) ( )

    ( )(

    x z ae

    x x x x x

    x x r ae

    v a v a T v L bv

    a a a v a b b

    a v a a a T b b T b b

    b b

    d q f

    dq q f q f q q

    df q f q f f f f

    f

    = + + + + =

    = + - + - = + - W

    = + - + + + + - W + = + -

    +

    &&

    &&

    &&

    && 2 )f- W

    where the unsteady aerodynamic lift Lae and the unsteady aerodynamic pitching moment Tae are defined in Eq. (21). Centrifugal force that appears in the flap and the

    torsion equations as Tz and Tr represent centrifugal stiffening expressions. Tr plays the role of torsional stiffness induced by the centrifugal force field.

    The boundary conditions at the root of the blade (z=0) are given by:

    '0 0 0x

    dv alsodzfq f f= = = = =

    At the tip of the blade (z=L), from the Hamiltonss principle, boundary conditions

    come out as:

    ' ' ''' 20 55 0 56 1

    ' '33 37

    '' ' ''' ' ' ' 265 0 66 73 77 10 18 1

    ' ' ''56 0 66

    0: ( ) ( ) 0

    0: 0

    0: ( ) ( ) ( ) 0

    0: ( ) 0

    x

    x x

    x x p

    x

    v a v a b R z

    a a

    a v a a a b b b R z I

    a v a

    d q f

    dq q f

    df q f q f f f

    df q f

    = + - + W =

    = + =

    = + - + + + + + W =

    = + - =

    &&

    where Tz , Tr and R(z) are given by 2 2

    0

    21

    4 50 2

    1( ) ( ) ( ),2

    ( , ) ( ),

    2, ,

    ( , )

    z

    p pmp pm ph

    r z ph

    R z R L z L z

    T z t b R z

    I II b b I I

    m mT z t T I

    b

    = - + -

    = W

    = + = =+

    =

    and aij and bij and m0 , m2 in Eqs. (25-28) are defined in Librescu (2006).

    In order to solve the set of equations given by Eq. (25), Extended Galerkin Method (EGM) is employed. For this purpose, the aerodynamic and the structural

    (26)

    (27)

    (28)

    (25)

    (27)

    (28)

  • variables are separated in time and spatial coordinate z and the assumed mode shapes are defined by Eq. (29).

    11

    22

    33

    10

    2

    3

    ( , ) ( ) ( )( , ) ( ) ( )

    ( , ) ( ) ( ) ( , ) ( ) ( )

    ( , ) ( ) ( ) ( , ) ( ) ( )

    TTBBv v

    T Tx x x BB

    T TBB

    B z t z q tv z t z q t

    z t z q t B z t z q t

    z t z q t B z t z q tf f

    q

    f

    ==

    = =

    = =

    YY

    Y Y

    Y Y

    where

    ,1 ,2 , ,1 ,2 ,

    , 1 , 2 ,2 , 1 , 2

    ,2 1 ,2 2 ,3

    1 1,3 1 1,3 2 1,4

    2 2,4 1 1,4 2 1,5

    ... ...

    ...

    ... ,

    ...

    ...

    T T

    v v v v N v v v v NT

    N v N N N v NT

    x x N x N x NT

    B B N B N B NT

    B B N B N B N

    q q q q

    q q qf f f f f+ + + +

    + +

    + +

    + +

    = = = == = =

    Y Y Y Y

    Y Y Y Y

    Y Y Y Y

    Y Y Y Y

    Y Y Y Y

    ,2

    ,2 1 ,2 2 ,3

    1 1,3 1 1,3 2 1,4

    2 2,4 1 1,4 2 1,5

    ...

    ...

    ...

    ...

    T

    NT

    x x N x N x NT

    B B N B N B NT

    B B N B N B N

    q

    q q q q

    q q q q

    q q q q

    f

    + +

    + +

    + +

    = = =

    In Eq. (30), N is the degree of the polynomial assumed for admissible function

    and N is taken as 5 in the current study because of the convergence achieved. " iY "

    are the so-called admissible functions which satisfy the geometric boundary conditions of the problem. Admissible functions are assumed in form of five degree polynomial and the coefficients of the polynomial are considered as one as shown in Eq. (31).

    = + + + +2 3 4 5i z z z z zY

    It should be noted that based on the convergence study related to the order of the polynomial in Eq. (31), convergence in the results are obtained for a polynomial of degree five. The state vector or the column matrix of time dependent variables is defined by Eq. (32).

    1 2 3

    TT T T T T Tv x B B B

    q q q q q q qf =

    Applying Galerkin method to the governing equations given by Eq. (25), after some manipulations, the general equations of the aeroelastic system can be obtained in terms of time dependent variables as shown in Eq. (33).

    ( ) ( ) ( ) 0s ae ae s aeM M q C q K K q+ + + + =&& &

    where " Ms " and "Mae " are the structural and aerodynamic mass matrices and "Cae" is the aerodynamic damping matrix and " Ks" and "Kae" are the structural and aerodynamic stiffness matrices, respectively. In order to the solve the governing system of equations given by Eq. (33) based

    on Theodorsens and Loewys unsteady aerodynamics in frequency domain, the U-g

    (33)

    (30)

    (29)

    (32)

    (31)

    (29)

  • method is used. In the U-g method, an eigenvalue problem is built and solved by the addition of an artificial damping term. In this method, to determine the aeroelastic instability boundary, damping versus speed curves are plotted, although damping values determined are physically meaningless except around the flutter boundary where the damping value is equal to zero.

    In the U-g method, adding the artificial structural damping is a tricky way to eliminate some complex terms that come from the existence of the reduced frequency

    in aerodynamic matrices. By adding the artificial damping g to the structure, the governing differential equation in time domain can be rewritten as:

    {( ) ( ) ( (1 ) ) } 0s ae s ae s aeM M q C C q K ig K q+ + + + + + =&& &

    Damping of the aeroelastic system is zero at the instability boundary. Thus, assuming undamped harmonic oscillation of the blade, one can express the time dependent variables as in Eq. (35).

    w= i tq qe

    By substituting Eq. (35) into Eq. (34), one gets Eq. (36).

    2{ ( ) ( ) ( (1 ) )} 0s ae s ae s aeM M C C i K ig K qw w- + + + + + + =

    Dividing two sides of Eq. (36) by 2U and defining the reduced frequency as in Eq.

    (37), Eq. (36) can be re-written as Eq. (38).

    w

    =b

    kU

    2

    2 2

    (1 )( )

    s s ae ae ae

    ig k kK M M C i K

    bU b

    += + - -

    Where, 2

    ae aeK U K= and

    ae aeC UC=

    It should be noted that U appearing in Eqs. (37-39) is defined as the relative velocity at the blade tip due to the free stream velocity and rotational speed of the blade. Eq. (38) can further be written as Eq. (40).

    [ ( ) ] 0s

    A k K ql+ =

    Where,

    (34)

    (35)

    (36)

    (37)

    )

    (31)

    (38)

    (39)

    (40)

  • 2

    2

    2

    ( ) ( )

    (1 )

    s ae ae ae

    k kA k M M C i K

    bbig

    Ul

    = + - -

    +=

    The real and imaginary part of l can be used to calculate the velocity and the artificial damping g based on Eq. (42) as:

    1 Im( ),

    Re( ) Re( )U g

    l

    l l= =

    In the U-g method, by iterating over the reduced frequency k, aerodynamic matrices can be calculated and eigenvalue problem can be solved for each reduced frequency. Since artificial damping does not exist in real life, solution of the eigenvalue problem is accurate only at the in instability boundary. Instability velocity is calculated

    by iterating over the reduced frequency k until the artificial damping introduced becomes zero.

    5. NUMERICAL RESULTS In the present study, fiberglass/epoxy is taken as the material of the blade. The geometric and material specifications of the rotating thin-walled composite box beam with the CAS lay-up are given in Tables 1 and 2. Thin-walled beam has rectangular cross-section and constant cell wall thickness in order to demonstrate the effect of using Theodorsens and Loewys lift deficiency functions on the aeroelastic instability boundary.

    Table 1 Geometric and material properties of the rotating composite TWB

    Geometric Properties Material Properties

    Length, L m 40 1E GPa 17

    Width, 2w m 1.4 2 3E E GPa 3.5

    Depth, 2d m 0.5 12 13G G GPa 0.8

    Wall thickness, h m 0.02 23G GPa 0.65

    Density, 3kg / m 1950 Poissons ratio, 0.28

    (41)

    (42)

    ) (31)

    (43)

  • Figs. 7-18 investigate effects of the fiber angle, rotation speed and number of blades (Q = 1,2 and 3) on flutter speeds. Iterative eigenvalue solution is performed for two different fiber angles (0o and -150) and for two different fixed rotational speeds, 5 rpm and 10 rpm, for determining the flutter speeds. Unsteady aerodynamic models of Theoderson and Loewy are used in the unified aero elastic system introduced in the previous section. Loewy used the Biot-Savart law instead of potential flow theory to account for the layers of shed vorticity beneath the reference rotor blade caused by the reference blade and other blades in previous revolutions. It can be seen that the flutter speed obtained via Theodorsens function does not depend on number of blades. Theodorsens function gives same flutter speed for different wind turbine type with two or three blades. The difference in prediction of the flutter speed between Theodorsens and Loewys theories increases by increasing the number of blades. As the modern wind turbines have three blades, utilizing Loewy theory is more reasonable than Theodersons theory. Aeroelasticity of the isolated Wind turbine blade with three blades as popular wind turbine type are comprehensively investigated in continue. For the 0o fiber angle case, Figs. 11 and 12 show the variation of the artificial damping with the relative velocity at the blade tip, given by Eq. (44), for the flapwise bending and torsional deformation modes. In these figures a is a distance between the aerodynamic center and the elastic center. In Figs. 11 and 12, it is noticed that when the rotational speed of the blade is increased, flutter speed determined by employing Loewys lift deficiency function in the aeroelastic system of equations separates from the flutter speed determined by the use of Theodorsens lift deficiency function. With the Theodorsens lift deficiency function, the flutter speed for the rotational speed of 10 RPM is obtained as 29.5 m/s and the use of Loewys lift deficiency function gives a flutter speed of 36 m/s. When the rotational speed is decreased, as in the case of 5 RPM, flutter speeds obtained are almost close by the use of either lift deficiency functions in the unsteady aerodynamic model.

    2 2

    inf( )

    low at r LU U r

    == + W

    Table 2 Gross properties chosen for the MW size wind Turbine blade

    Rating 1.5 MW Blade number 3 Rotor diameter 60 m

    Blade chord 3.5 m Maximum Tip speed 50 m/s

    (44)

  • Fig. 7 U-g diagram For = 0

    0, = 5 rpm, a=0.3,

    Q=1

    Fig. 8 U-g diagram For = 0

    0, = 10 rpm, a=0.3,

    Q=1

    Fig. 9 U-g diagram For = 0

    0, = 5 rpm, a=0.3,

    Q=2

    Fig. 10 U-g diagram For = 0

    0, = 10 rpm, a=0.3,

    Q=2

    Fig. 11 U-g diagram For = 00, = 5 rpm, a=0.3,

    Q=3

    Fig. 12 U-g diagram For = 0

    0, = 10 rpm, a=0.3,

    Q=3

    Torsional

    mode

    Torsional

    mode

    Flapwise bending

    mode Flapwise bending

    mode

    Flapwise bending

    mode

    Torsional

    mode

    Flapwise bending

    mode

    Torsional

    mode

    Flapwise bending

    mode

    Torsional

    mode

    Flapwise bending

    mode

    Torsional

    mode

  • Figs. 17 and 18 present the U-g plots for fiber angle of -15o. Similar to the 0o case, when the rotational speed is increased, flutter speed determined by employing Loewys lift deficiency function in the aeroelastic system of equations separates from the flutter speed determined by the use of Theodorsens lift deficiency function. For the rotational speed of 10 rpm, the use of Theodorsens unsteady aerodynamic theory in the aeroelastic system gives a flutter speed of 32 m/s and Loewys theory gives a flutter speed of 39 m/s. For the rotational speed of 5 rpm, the difference in flutter speeds calculated by the two theories are decreased and they are 31.5 for Theodersons theory and 35 m/s for Loewys theory.

    For both fiber angle and rotational speed cases, Table 3 summarizes the flutter

    characteristics of the wind turbine blade. In Table 3, m is the ratio of oscillatory flutter frequency to the rotational speed and h is the wake spacing, shown in Fig. 5. Results presented in Table 3 show that high wake spacing, such as 2.22 or 2.4, corresponds to high inflow, and the effect of any shed layer of vorticity on the flutter speed is negligible. It is concluded that when the spacing between the wake spirals is small, as would the case for high rotational speed or low inflow case, flutter speed increases. Theodorsens unsteady aerodynamics theory cannot predict the increase of the flutter speed for small wake spacing.

    Fig. 13 U-g diagram For = -150, = 5 rpm, a=0.3,

    Q=1

    Fig. 14 U-g diagram For = -150, = 10 rpm, a=0.3,

    Q=1

    Flapwise bending

    mode

    Torsional

    mode Torsional

    mode

    Flapwise bending

    mode

  • Fig. 9 U-g diagram For = -150, = 5 rpm, a=0.3,

    Q=2

    Fig. 16 U-g diagram For = -15

    0, = 10 rpm, a=0.3,

    Q=2

    Fig. 17 U-g diagram For = -15

    0, = 5 rpm, a=0.3,

    Q=3

    Fig. 18 U-g diagram For = -15

    0, = 10 rpm, a=0.3,

    Q=3

    Furthermore, it is seen that the frequency ratio m, which measures the phase relationship between shed layers of vorticity, in the Loewys lift deficiency function also affects the flutter speeds. For the particular wind turbine blade studied, for both fiber angle cases, it is seen that at the flutter boundary low frequency ratio corresponds to higher flutter speeds. For both fiber angle cases, flutter mode is identified as flapwise bending mode. Loewy has shown that damping coefficient associated with the flapwise bending mode has sharp drops at the integer values of the frequency ratio (Loewy 1957). Viswanathan (1977) reported that at low integer values of the frequency ratio, this drop is confined to a small neighborhood near the integral values, but at higher values of the frequency ratio, the width of low damping increases. Lower flutter speed

    Flapwise bending

    mode Flapwise bending

    mode

    Flapwise bending

    mode

    Torsional

    mode

    Torsional

    mode

    Torsional

    mode

    Flapwise bending

    mode

    Torsional

    mode

  • for higher value of the frequency ratio, given in Table 3 for both fiber angle cases, could be due to the widening of the width of the low flapwise damping.

    Table 3. Effect of Theodorsens and Loewys unsteady aerodynamic models and fiber angle on the flutter characteristics of the wind turbine blade

    Number of

    blades

    Fiber angle (Deg.)

    Rotation speed (RPM)

    Unsteady Aerodynamic

    model

    Flutter tip speed (relative velocity)

    (m/s)

    Inflow speed (m/s)

    Flutter frequency

    (Rad/s)

    =

    h

    Q = 3

    0

    5 Theoderson 29 24.97 1.35 - -

    Loewy 32 27.9 1.35 2.6 2.22

    10 Theoderson 29.5 4.8 1.5 - -

    Loewy 36 18.9 1.6 1.53 1.21

    -15

    5 Theoderson 31.5 27.36 1.6 - -

    Loewy 35 31.3 1.6 3.07 2.4

    10 Theoderson 32 9.36 1.7 - -

    Loewy 39 24.17 1.8 1.73 1.34

    Table 3 also shows that flutter speeds of the blade with negative off-axis fiber angle are higher than the flutter speeds of the on-axis fiber angle case. Both, Theodorsens and Loewys unsteady aerodynamics, predict the same behaviour. The increase of the flutter speed of the blade with the negative off-axis fiber angle case is attributed to the positive effect of bending-twisting coupling effect that the negative off-axis ply angles induces in the blade. Figs. 3 and 4 show that when the ply angles are negative, fibers are aligned towards the leading edge and when bending occurs trailing edge deflects more than the stiffer leading edge resulting in reduction in the effective angle of attack of the blade section. Reduction in the effective angle of attack ultimately causes increase of the flutter speed. 6. CONCLUSIONS In the present study, Circumferentially Asymmetric Stiffness (CAS) structural model of a rotating thin-walled composite box beam is used as the simplified model of the wind turbine blade in conjunction with Theodorsens and Loewys functions in unsteady incompressible flow to study the aeroelastic instability of the composite blade. Hamiltons principle and the extended Galerkin method are used to obtain the coupled linear governing system of dynamic aeroelastic equations. Frequency domain solutions are performed for fixed rotational speeds to investigate the effect of using Theodorsens and Loewys unsteady aerodynamics on the flutter speeds of composite wind turbine blades. As a by-product of the study, the effect of fiber angle of the CAS model on the flutter speed is also show. Flutter speed results show that Theodorsens lift deficiency function, while applicable to the fixed-wing case, may not be as valid for rotary-wing aircraft for cases when the effect of previously shed layers of vorticity should be considered. In the present study, Theodorsens lift deficiency function is used primarily to set a baseline for the flutter calculations so that comparisons with the other lift deficiency functions can be made. It is concluded that when the spacing between the wake spirals is small, as would the case for high rotational speed or low inflow case, flutter speed increases. Theodorsens unsteady aerodynamics theory cannot predict

  • the increase of the flutter speed for small wake spacing. It is also shown that flutter speeds of the blade with negative off-axis fiber angle are higher than the flutter speeds of the on-axis fiber angle case mainly due to the decrease of the effective angle of attack due to the bending-twisting coupling induced by the negative off-axis fiber angle. ACKNOWLEDGEMENTS This work was supported by the METU Centre for Wind Energy and Scientific and Technological Research Council of Turkey (TBTAK), Project No: 213M611. REFERENCES Couch, M. A., Three-dimensional flutter theory for rotor blades with trailing-edge flaps,

    PhD thesis, Naval Postgraduate School, Monterey, California, June 2003. Farsadi, T., and Kayran, A., Classical Aeroelastic Stability Analysis of Large

    Composite Wind Turbine Blades, AIAA 2016-1959, AIAA SciTech, 57th AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 4-8 January, 2016, San Diego, USA,

    Glauert H. Airplane propellers In: Durand WF, editor. Aerodynamic theory. New York: Dover Publications; 1963.

    Hansen, M. H., Stability Analysis of Three-Bladed Turbines Using an Eigenvalue Approach, Proc. of the 2004 ASME/AIAA Wind Energy Symposium, Reno, pp. 192202, 2004.

    Janetzke, C. D. and Kaza, K., Whirl flutter analysis of a horizontal-axis wind turbine with a two-bladed teetering rotor, Solar Energy, vol. 31, no. 2, pp. 173182, 1983.

    Librescu, L. and Song. O., Thin-walled composite beams: theory and application, Springer, The Netherlands, 2006

    Loewy, R.G., A two-dimensional approximation to the unsteady aerodynamics of rotary wings, Aeronautical science, Vol. 24, No.2, 1957.

    Lobitz, D. L. and Veers, P. S., Aeroelastic Behavior of Twist-Coupled HAWT Blades, AIAA-98- 0029, Proc. 1998 ASME Wind Energy Symposium held at 36th AIAA Aerospace Sciences Meeting and Exhibition, Reno, NV, Jan. 12-15, 1998.

    Lobitz, D. W., Aeroelastic Stability Predictions for a MW-Sized Blade, Wind Energy 7 (13), pp. 211224, 2004.

    Leishman, J. G., Principles of Helicopter Aerodynamics, Cambridge University Press, New York, 2000.

    Politakis, G., Haans, W., Van Bussel, G. J. W., Suppression of Classical Flutter Using a Smart Blade. 46th AIAA Aerospace Sciences Meeting and Exhibit. Reno, Nevada. 7-10 January 2008.

    Sina S. A., Ashrafi M.J., Haddadpour H., and Shadmehri F., Flexuraltorsional vibrations of rotating tapered thin-walled composite beam Proceedings of the Institution of Mechanical Engineers, Part G: Journal of Aerospace Engineering 2011 225: 387

    Song, O., Librescu, L., and Oh, S.-Y., Vibration of pretwisted adaptive rotating blades modeled as anisotropic thin-walled beams, AIAA J., 2001, 39(2), 285295

    Tangler, J.L., The evolution of rotor and blade design, National Renewable Energy Laboratory, 2000

    Theodorsen, T. General Theory of Aerodynamic Instability and the Mechanism of Flutter, NACA Report 496, 1935.

    Vatne, S. R., Aeroelastic Instability and Flutter for a 10MW Wind Turbine, Norwegian University of Science and Technology, 2011.

    Veers, P., Lobitz, D., and Bir, G., Aeroelastic tailoring in wind-turbine blade applications, Windpower98, Bakersfield, CA (United States), 27 Apr-1 May 1998.

    Viswanathan, S.P., An Analysis of the flutter and damping characteristics of helicopter rotors, PhD thesis, Georgia Institute of Technology, Monterey, California, January 1977.