You are on page 1of 18

Journal of Sound and Vibration 493 (2021) 115810

Contents lists available at ScienceDirect

Journal of Sound and Vibration


journal homepage: www.elsevier.com/locate/jsv

A novel methodology for analyzing modal dynamics of


multi-rotor wind turbines
Oliver Tierdad Filsoof a,b,∗, Morten Hartvig Hansen c, Anders Yde b,
Peter Bøttcher b, Xuping Zhang a,∗
a
Aarhus University, Department of Engineering, Inge Lehmanns Gade 10, Aarhus, Denmark
b
Vestas Wind System A/S, Innovation & Concepts, Hedeager 42, Aarhus, Denmark
c
SDU Mechatronics, University of Southern Denmark, Alsion 2, Sonderborg, Denmark

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

Article history: A multi-rotor wind turbine (MRWT) is accepted as a novel solution in reducing the ro-
Received 7 April 2020 tor size without decreasing the energy harvesting capability compared to a single-rotor
Revised 2 October 2020
wind turbine (SRWT). However, there is no systematic methodology available for obtaining
Accepted 22 October 2020
the modal properties of MRWTs. This paper presents a novel methodology and systematic
Available online 23 October 2020
modeling method of modal dynamics with comprehensive numerical simulations and anal-
Keywords: yses of MRWTs. In the methodology, Coleman transformation is employed to convert the
Multi-rotor wind turbine nonlinear time-variant dynamics formulation of a MRWT into a set of linear time-invariant
Modal analysis dynamic equations. Using HAWC2 and HAWCStab2 a method to establishing high-fidelity
Linear time-invariant model linear time-invariant dynamic models is presented. The structural modal dynamics of a tri-
High-fidelity model rotor wind turbine is systematically investigated by using modal analysis theory and inter-
preting the rotor speed-dependent dynamics visualized in a Campbell diagram. The inves-
tigation shows that rotor modes on a MRWT have similar behavior as for a SRWT and that
low torsional stiffness of the rotor attachment point affects the first flapwise backward
whirling modes by decreasing the corresponding natural frequencies. Moreover, the low
torsional stiffness results in a delayed splitting of the asymmetric flapwise rotor modes.
The simulations and analyses result verified the proposed dynamic modeling method. The
major contributions in this paper provide the essential insights and guidance in the design
of MRWTs but also for SRWT with slender towers.
© 2020 Elsevier Ltd. All rights reserved.

1. Introduction

Today most wind turbines consist of a single-rotor with three blades attached to a tower as shown in Fig. 1(a). For
onshore wind turbines, the rotor diameter can reach up to 170 m and 222 m for offshore wind turbines. The future trend
is to keep increasing the rotor diameter to increase the wind turbine capacity factor and decrease its cost. However, large
rotors lead to some challenging aspects. Firstly, the rotor mass is proportional to the cubic of the rotor radius while the
power output to the square making it difficult to keep the cost down when up scaling wind turbines. The second challenge is
the decreasing natural frequencies of the rotor blades. The consequence of this is an increased risk of instabilities including


Corresponding authors.
E-mail addresses: otf@eng.au.dk (O.T. Filsoof), mortenhhansen@mci.sdu.dk (M.H. Hansen), anyde@vestas.com (A. Yde), pebot@vestas.com (P. Bøttcher),
xuzh@eng.au.dk (X. Zhang).

https://doi.org/10.1016/j.jsv.2020.115810
0022-460X/© 2020 Elsevier Ltd. All rights reserved.
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 1. Splitting the rotor area of a (a) single-rotor wind turbine into a three smaller rotor areas in a (b) multi-rotor configuration makes it possible to keep
the same power output and, at the same time, reduces the rotor size by 57 %. This challenges the normal scaling law for wind turbines.

flutter [1–3]. Thirdly, it is challenging to manufacture, transport, and install large rotor blades. A novel way to overcome
these challenges is to make a multi-rotor wind turbine (MRWT) shown in Fig. 1(b). The idea of a MRWT is to divide the
power output of a single-rotor wind turbine (SRWT) into n rotors resulting in a reduced length of the rotor blades.
The modal dynamics of SRWT have been a research topic for decades and are often analyzed by linearizing the system
around an equilibrium point while assuming constant rotor speed. Doing so result in a linear time-periodic system with the
period T = 2π where  is the rotor speed. For isotropic rotors with three or more blades, e.g. having equally spaced and
identical blades, it is possible to use Coleman transformation [4] to make the system matrix time-invariant. This transfor-
mation was used by Malcolm [5] to linearize the equation of motions of an operating three-bladed wind turbine. Hansen
[6] later used this transformation to show that the forward and backward whirling edge modes have different damping
properties. Due to its rather simple implementation, different software tools use Coleman transformation. These include
FAST [7] and HAWCStab developed by Hansen [8] and later upgraded to HAWCStab2 [9] by deriving the kinematics from a
nonlinear co-rotational formulation. For anisotropic rotors, it is not possible to utilize Coleman transformation due to mul-
tiple harmonics components in the solution space. Instead, Floquet theory [10] or Hill’s method [11] can be used to find
the principle harmonics in the response. Skjoldan and Hansen [12] used Floquet theory to analyze a SRWT. In this work the
anisotropy was introduced by varying the blades stiffness properties. Hansen [13] used Hill’s method to analyze the modal
dynamics of a two-bladed SRWT and found that two-bladed wind turbines experience larger and higher harmonic terms
in the system matrix. This shows that two-bladed SRWTs experience modal couplings that do not exist for three-bladed
turbines. The finding by Hansen was confirmed by Kim et al. [14]. He analyzed the modal frequencies of the Envision 3.6
MW two-bladed SRWT.
A limited number of studies have been performed on analyzing the modal response of MRWT, including a study by
Filsoof et al. [15]. Analytically they showed that applying Coleman transformation for each rotor in a MRWT setup yields a
time-invariant system as for a SRWT. They further found that the period of each rotor does not have to be the same. For
the experimental dynamic of MRWTs, acceleration data from the Vestas four rotor wind turbine has been analyzed [16].
Another study on a scaled-down model of this wind turbine show identical behavior in the rotor modes between a SRWT
and a MRWT [17].
This paper is an extension of the preliminary work [18]. In this work, rotor modes from the DTU 30 MW tri-rotor wind
turbine (TRWT) was related to a SRWT with the same rotor. But no detailed description of the method, methodology, and
verification was given. Therefore, this work aims to include this by presenting the general novel methodology to perform
modal analysis of MRWTs with no restrictions on the number of rotors and the individual rotor speed. Using this, the
method to evaluate the stability of a high-fidelity linear time-invariant MRWT is developed from which modal frequen-
cies, damping ratios, and mode-shapes are extracted by direct eigenvalue analysis. To demonstrate the applicability of the
method, a high-fidelity linear model of the DTU 30 MW tri-rotor wind turbine (TRWT) (Fig. 1(b)) is made. This model is
used to understand the underlying fundamental structural modal dynamics of the TRWT and to verify the method. The ver-
ification involves comparing natural frequencies from the TRWT to observed resonance frequencies from a fully nonlinear
model in HAWC2.

2. Problem formulation

If the power output of a SRWT splits into n rotors in a MRWT setup, the rotor radius equals √R where R is the radius
n
R3
of the SRWT [19]. This makes the power output scaling with R2 but the mass now becomes proportional to √
n
compared
to R3 for a SRWT. As an example, dividing the power output of an SRWT with 200 m long rotor blades into three smaller

2
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 2. Structure of the paper. One part is on establishing the novel methodology and method for analyzing the modal properties of multi-rotor wind
turbines, and the second on verification.

and identical rotors, the corresponding MRWT blade radius is reduced to 115.5 m. This reduction in blade length lowers the
risk of blade instabilities and ease manufacturing, transportation, and installation of the blades. Furthermore, standard rotor
nacelles might extend their product lifetime by being used in a MRWT concept.
A drawback with a MRWT design is that it becomes much more complicated to access the modal dynamics compared to
a SRWT. This is an essential part of the design process to avoid instabilities and natural frequencies. If these are not avoided
or handled, they can lead to an unstable design, high fatigue loads and in the worst-case drive the system to complete
failure. The natural frequencies and damping ratios are not constant for a wind turbine but are highly affected by the rotor
speed due to gyroscopic coupling effects, stress stiffening of the blades, and the structure/wind interactions. Therefore, the
natural frequencies and damping ratios are often shown as a function of either rotor speed (Campbell diagram) or wind
speed.
This overview of the fundamental dynamics is the key in the design phase due to three reasons. The first is that the pitch
or generator torque controllers can interfere with the natural frequencies. Secondly, the rotors input periodic excitations
into the system at given multiples of the rotor speed [20]. Thirdly, a prerequisite is to know the natural frequencies so it is
possible to make control strategies to go fast pass them when speeding up the rotor.
This paper focuses on the methodology of modal dynamic analysis of MRWTs and the technical framework is structured
as shown in Fig. 2. In the Theoretical Framework section, the methodology of extending Coleman transformation to MRWTs
is presented. Subsequently, a method to establish high-fidelity linear time-invariant MRWT models is presented by using
HAWC2 and HAWCStab2. In the Results and Discussion section, the method is used to obtain a Campbell diagram of the
DTU 30 MW TRWT. This diagram is analyzed in detail to give an insight into the structural dynamical behavior of MRWTs.
To verify the linear model, Campbell diagrams of a fully nonlinear model of the TRWT are made in HAWC2 by applying
excitation forces and analyzing outputs in the frequency domain. The important findings in this paper are listed in the
Conclusion section.

3. Theoretical framework

A rotating mechanical system like a wind turbine is time-variant by nature which makes it difficult to analyze its modal
properties. A direct eigenvalue analysis of the system matrix is not an option because it yields time-varying eigenvalues.
However, with proper assumptions, the system can be changed to be a time-variant periodic system and even time-invariant
by utilizing appropriate transformations. This section presents Lyapunov-Floquet and Coleman transformation that can be
used to adapt the modal properties for SRWTs, and how Coleman transformation can be extended to MRWTs. This extension
makes it possible to build high-fidelity linear time-invariant models of MRWTs, which is demonstrated by showing how to
combine the aeroelastic tools HAWC2 [21] and HAWCStab2 [22]. To verify the linear model, a method to obtain Campbell
diagrams from a nonlinear model in HAWC2 is also presented.

3.1. Modal analysis of wind turbines

The modal dynamics of a non-rotating wind turbine is assessed by analyzing the following time-invariant first-order
ordinary differential equation
x˙ (t ) = A x(t ) (1)

3
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

where x(t ) = [x1 (t ), x2 (t ), x3 (t ), . . . , xN (t )]T contains the N states in the system, (•˙ ) denotes the time derivative and A is
the system matrix containing information about the mass, damping, and stiffness properties of the wind turbine. Assuming
that the solution of Eq. (1) is on the form x(t ) = v eλt , where λ is the complex eigenvalue and v is the eigenvector, the
following eigenvalue problem is formulated

Av = λv (2)

From the eigenvalues λk = σk + iωk (i = −1) the natural frequencies in Hz, fn,k , and damping ratios, ξk , can be determined

ωk −σk
fn,k = ξk =  (3)
2π σk2 + ωk2
Where k determines the mode number. The mode-shapes are given by the eigenvectors vk revealing how the wind turbine
vibrates when excited at fn,k . As the wind turbine starts rotating at a constant speed, , the system matrix becomes periodic

x˙ (t ) = A(t ) x(t ) = A(t + T ) x(t ) (4)


with period T = 2π . Due to this periodicity in the system matrix, it is infeasible to perform a direct eigenvalue analysis
since the eigenvalues will become time-depended. Instead, the system states can be transformed into a new set of states
using the Lyapunov-Floquet transformation matrix, L(t ) [23]

x(t ) = L(t ) z(t ) = L(t + T ) z(t ) (5)

L(t ) = φ(t ) eRt φ(0 )−1 L(0 ) (6)


where φ(t ) is a fundamental solution matrix to Eq. (4) so that φ
˙ (t ) = A(t )φ(t ) and R is a constant non-singular matrix
defined in terms of the monodromy matrix C = φ(0 )−1 φ(T ) = eRT . Inserting Eq. (5)into Eq. (4) makes the system time-
invariant
 
z˙ (t ) = L(t )−1 A(t ) L(t ) − L˙ (t ) z(t ) = AL z(t ) (7)
where AL is the new system matrix. The solution to Eq. (7) takes the form

z(t ) = eAL t z(0 ) (8)

z(0 ) = L(0 )−1 x(0 ) (9)


from which the modal dynamics can be analyzed. However, it requires solving N linearly independent solutions of Eq. (4) to
establish the fundamental solution matrix used in Eq. (6). This is computationally expensive and only practical for low-order
systems. However, there exist other methodologies [24] to solve the eigenvalue problem of time-variant periodic systems like
Hill’s method [25]. In this method, the eigenvalues are approximated by expanding the system matrix and modes-shapes by
Fourier series, but this is also computationally expensive and not applicable for high-fidelity models with a high number of
degree-of-freedoms (DOFs).
However, for isotropic rotors (identical and equidistantly spaced blades), with three or more blades, there exist a more
practical approach in making the system matrix time-invariant. That is to use Coleman transformation [4]. The principle in
this transformation is to transform the rotor DOFs from a rotating frame into the ground-fixed frame. This is achieved by
using the Coleman transformation matrix, B(t ), which for a three-bladed isotropic rotor is given

xb (t ) = B(t ) zb (t ) (10)

 
INb INb cos(ψ1 (t )) INb sin(ψ1 (t ))
B(t ) = INb INb cos(ψ2 (t )) INb sin(ψ2 (t )) (11)
INb INb cos(ψ3 (t )) INb sin(ψ3 (t ))

where I is the identity matrix and ψi (t ) =  t + 23 (i − 1 )π is the position of blade number i. The subscript (• )Nb refers
to the number of rotor DOFs and defines the size of I. The new states introduced by Coleman transformation zb (t ) =
[a0 (t ) , a1 (t ) , b1 (t )]T is called multi-blade coordinates. It has been shown that Coleman transformation is a special case
of the Lyapunov-Floquet transformation by letting L(t ) = B(t ) in Eq. (7) [26].
Assuming a harmonic solution of the form zb (t ) = v eλ t so that xb (t ) = B(t ) v eλ t it is possible to arrive at the following
expression for the motion of for blade number i
  

  2 2
xbk.i (t ) = eσk t A0 cos ωkt + ϕ +A
0 BW
cos (ωk + )t + (i − 1 ) + ϕ BW +A FW
cos (ωk − )t − (i − 1 ) + ϕ F W
3 3
(12)

4
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

where A(• ) and φ (• ) are amplitudes and phase angles. It is seen that the rotor modes split into three different modes. One
is a symmetric mode, (• )0 which does not depend on the rotor speed whereas the asymmetric parts namely the backward
whirling (BW) (• )BW and forward whirling (FW) (• )F W have a rotor speed dependens by ±. The amplitudes are given in
terms of the multi-blade coordinates

A0 = |a0 |
|a1 − ib1 |
ABW = (13)
2
|a1 + ib1 |
AF W =
2
and phase angles by

ϕ 0 = Arg(a0 )
ϕ BW = Arg(a1 − ib1 ) (14)
ϕ F W = Arg(a1 + ib1 )
Filsoof et al. [15] showed that applying Coleman transformation to each rotor of the MRWT makes the system matrix time-
invariant. Therefore, Coleman transformation matrix is extended to a block diagonal matrix for a MRWT system
⎡ ⎤
B1 (t ) 0 ··· 0
⎢ 0 B2 (t ) ··· 0 ⎥
xb (t ) = ⎢ . .. .. ⎥zb (t ) (15)
⎣ .. .
..
. .

0 0 ··· B p̄ (t )
where the subscript (• ) p̄ refers to total number of rotors on the MRWT. This means that the motion of the individual rotor
blades on a MRWT is given
  

  2 2
xbk.i.p (t ) = eσk t A0p cos ωkt + φ p0 + ABW
p cos ( ω k +  p )t + ( i − 1 ) + φ BW
p + A FW
p cos ( ω k −  p )t − ( i − 1 ) + φ FW
p
3 3
(16)
which is almost identical with Eq. (12). The only difference is the rotor index (• ) p on the amplitudes, phase angle, and rotor
speed. Note, that the asymmetric modes do not depend on the rotor speed from the other rotors. This shows that a linear
MRWT system, which is a complicated time-varying system, can elegantly be turnd time-invariant by applying Coleman
transformation on each rotor even though the rotors have different periods.

3.2. High-fidelity linear time-invariant multi-rotor wind turbine model

The previous section showed that a linear MRWT system is turned time-invariant by applying Coleman transformation
on each rotor. This important finding makes it feasible to take linear time-invariant rotor models and attach them to a
linear support-structure model. In this work, the linear support-structure model is made with the aeroelastic multibody tool
HAWC2 while the rotor-nacelle-assembly (RNA) model is established with the aeroelastic SRWT stability tool HAWCStab2.
The advantage of using HAWCStab2 for the RNA model is that it outputs the mass, damping, and stiffness matrices in multi-
blade coordinates. HAWCStab2 have further been validated and verified in several publications [6,8,27–30].
The assembling procedure to achieve the system matrix of a MRWT and performing a series of eigenvalue analysis for
various rotor speeds are shown in Fig. 3. In the first step, HAWC2 is used to make a model of the support-structure. The
rotors are then exported from HAWCStab2 and connected to the support-structure using a generic matrix formulation. After
assembling the global mass, stiffness, and damping matrix, the system is rewritten into a first-order form, and modal pa-
rameters are extracted by direct eigenvalue analysis. The procedure is repeated for several rotor speeds to get a Campbell
diagram. The step is presented in more detail in the upcoming subsections.

3.2.1. Rotor-nacelle-assembly model


The RNA refers to the nacelle, hub, shaft, generator, and blades of a wind turbine, and the linear model is made in the
aeroelastic code HAWCStab2. The structural model in HAWCStab2 is derived from a co-rotational formulation of articulated
Timoshenko beam elements with six DOFs per node (three translations and three Euler angle rotations), consists of flexible
beam elements representing the important dynamics of the RNA including nacelle motion, drive-train rotations, and three
rotor blades including the hub. The discretization of these components into beam elements is shown in Fig. 4. Notice that
only the low-speed shaft is included, but the inertia and stiffness from the generator are compensated for by altering the
shaft properties as a low speed equivalent system. Gravity is not included in the model and the structural tilt of the nacelle
is not taken into account for the system to have a steady-state solution.

5
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 3. This schematic diagram shows how to make a modal analysis of a multi-rotor wind turbine. For multi-rotor wind turbines, i is a vector defining
the rotational speed of each rotor, and for a single-rotor, it is a scalar quantity.

Fig. 4. Discretization of the rotor-nacelle-assembly into Timoshenko beam elements in HAWCStab2. The yaw interface and drive-train is described in the
ground-fixed frame while the hub and blades are given in multi-blade coordinates (Eqs. (13) and (14)).

The model of the RNA can be exported from HAWCStab2 in a second order form
M p z̈ p + C p z˙ p + (S p + K p ) z p = 0 (17)
where the constant matrices are the mass matrix, M p , the combined gyroscopic and structural damping matrix C p , the
centrifugal stiffness matrix S p , and the elastic stiffness K p . The subscript (• ) p denotes the index of the RNA on the MRWT
(the support structure will be denoted by p = 0).
The DOF vector z p can be written on the form
⎧ ⎫

⎨dg,p ⎪

a0,p
zp = (18)

⎩a1,p ⎪

b1,p

6
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

where the vector dg,p contains the yaw interface, generator bearing and shaft DOFs which all are described in the ground-
fixed frame in HAWCStab2. The multi-blade coordinates, a0,p ,a1,p and b1,p are the transformed hub and blade DOFs
(Eq. (10)).
The co-rotational beam formulation in HAWCStab2 is special in the way that the six DOFs of each node are the local
translations and rotations of the endpoint of the particular element. This kinematic formulation means that the very first six
DOFs of the articulated beam element model (being those of the second node because the first node is always clamped and
does not lead to any DOFs) describe the three translations and three rotations of the entire RNA in the ground-fixed frame
of reference. The subsequent DOFs describe translations and rotations relative to the motion described by DOFs below them.
This results in a very populated mass matrix which is inconvenient for time integration with an inversion in each time step,
but it does not matter for eigenvalue and frequency-domain analyses which the high-fidelity linear time-invariant model is
developed for.
The mass, stiffness, and damping matrices of the RNA from HAWCStab2 is written in a generic form


00,p 01,p
p = (19)
10,p 11,p
where  can be replaced with M, C or K, and subscript (• )αβ ,p denotes which DOFs the matrix entries are coupling to. The
block matrices K10,p and K01,p of the elastic stiffness matrix are zero in the co-rotational formulation of HAWCStab2.

3.2.2. Support-structure model


The linear model of the support-structure (for the TRWT the tower, wires, and arms) is made in HAWC2. The structural
formulation in HAWC2 is based on the floating frame of reference formulation where each body has an attached coordinate
system, and bodies are connected using nonlinear algebraic constraint equations [31]. Elastic deformations are taken into
account by discretizing each body into a number of linear Timoshenko beam elements. Wires are modeled using a general-
ized bar element formulation [32]. The equations of motions for the support structure can, as for the RNA, be exported in
HAWC2 in second-order form
M0 z̈0 + C0 z˙ 0 + K0 z0 = 0 (20)
where the DOF vector z0 is given
⎧ ⎫

⎪ d0,0 ⎪
⎪ d0,1 ⎪
⎪ ⎪
⎨ ⎪ ⎬
z0 = d0,2 (21)

⎪ ⎪
.. ⎪

⎪ ⎪
⎩ . ⎪⎭
d0, p̄
where d0,1 ,d0,2 ,. . . ,d0, p̄ are the DOFs of the connection points on the support-structure which the RNA are connected to,
and d0,0 are the remaining DOFs of the model.
The mass matrix M0 , the structural damping matrix C0 , and the elastic stiffness K0 is, as for the RNA matrices, written
in the following generic form
⎡ ⎤
00,0 01,0 02,0 ... 0 p̄,0
⎢10,0 11,0 0 ... 0 ⎥
⎢ 22,0 ...

0 = ⎢
⎢ .
20,0 0 0 ⎥
⎥ (22)
.. .. .. ..
⎣ .. . . . . ⎦
 p̄0,0 0 0 ...  p̄ p̄,0
where the damping and stiffness matrices follow the same notation as the RNA denoting which DOFs the matrix entries are
coupling to.

3.2.3. Coupled model for eigenvalue analysis


The RNA models are coupled directly to the support-structure model through their common DOFs leading to the follow-
ing generic matrix formulation
⎡ ⎤
00,0 01,0 0 ... 0 p̄,0 0
⎢10,0 11,0 + 00.1 01,1 ... 0 0 ⎥
⎢ 0 10,1 11,1 ... 0 0 ⎥
⎢ ⎥
f = ⎢ . .. .. .. .. .. ⎥ (23)
⎢ .. . . . . . ⎥
⎣ ⎦
 p̄0,0 0 0 ...  p̄ p̄,0 + 00. p̄ 0 p̄, p̄
0 0 0 ...  p̄0, p̄  p̄ p̄, p̄
where the subscript (• ) f refers to the full system matrices. Due to the inertia forces of the rotating parts, the gyroscopic
component cannot be separated from the structural damping components of the C p matrix from HAWCStab2. Therefore, the

7
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 5. This figure shows the procedure for generating a Campbell diagram from a nonlinear wind turbine model. The first step is to generate a low-pass
filtered white-noise signal. Using this a random binary sequence is made by looking at the sign of the filtered white-noise signal. This signal is used to
excite the nonlinear model. The output signals from the wind turbine model are transferred into the frequency domain to find the natural frequencies of
the system.

structural damping matrix for the complete system is established by assigning a modal damping coefficient to each mode
according to Chopra [33]
 

N/2
2 ζk.s ωk
C f.s = M f vk vTk M f (24)
Mk
k=1

Where N/2 is the total number of DOFs in the system, M f is found by replacing  with M in Eq. (23), and Mk and ξk.s are
the generalized modal mass and assigned structural damping ratio to mode number k respectively. This way of assigning
structural damping to the system means that C0 is not used.
Now using M f from Eq. (24) and replacing  with C and K in Eq. (23), the complete system in first order form is given


0 I  x
x˙ = (25)
−M−1
f
Kf −M−1
f
C f + C f.s
  
Af

from which it is possible to make a direct eigenvalue analysis to get natural frequencies and damping ratios using Eq. (3).
The eigenvectors for the hub and blades are transformed back to its original coordinates by multiplying them with
Eq. (15) (needs slight modification to take account for DOFs in ground-fixed frame).

3.3. Campbell diagram from nonlinear model

To verify that the high-fidelity time-invariant linear model represents the nonlinear dynamics around the point of lin-
earization, its natural frequencies are compared against a nonlinear model. The nonlinear model is made in HAWC2, and to
get an overview of how its natural frequencies develops as a function of rotor speed, several simulations are performed. The
simulation procedure is illustrated in Fig. 5 including how sensor data is post processed. In each simulation, the rotor speeds
are fixed, and a sequence of impulse forces are applied at 25 different locations. These locations are based on experience
from Kim et al. [14], and includes two impulse-forces on each blade in the flap and edge direction to excite blade modes,
three on the tower in fore-aft and side-side direction to get first and second tower modes, and, lastly, two impulses are
applied on each arm to get various combinations of arm-bending modes.
Exciting the complete frequency spectrum of the system requires an impulse sequence with a flat frequency response.
This property is achieved by using a random binary signal (RBS) [34]. A white zero-mean Gaussian noise signal is generated
and filtered through a low-pass filter. The RBS is made by evaluating the sign of the filtered signal. A negative sign results
in the value − 1 and +1 for a positive sign. Each RBS signal has its own seed to make them uncorrelated. Sensor outputs

8
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 6. Campbell diagram for the tri-rotor wind turbine where all rotors have the same speed.

(accelerations, forces, etc.) from simulations are analyzed by examining its power spectral density (PSD). The PSD is calcu-
lated by taking the squared magnitude of the output signal in the frequency domain, Y(ω ), and divide it with the number
of samples, m, multiplied with the sampling time, t.
|Y(ω )|2
PSD = (26)
m t
For a wind turbine peaks appearing in the PSD are natural frequencies and periodic excitations from the rotors. The ones
from periodic excitations are characterized by having no damping. Therefore, they show up in the PSD as sharp peaks and
can be distinguished from the natural frequencies of the system. After covering the rotor speed range of interest, the data
is visualized in a waterfall plot where the x -axis is rotor speed, y -axis frequency, and z -axis power. This plot can then be
compared with the results of the high-fidelity linear model.

4. Results and discussion

In this section a model of the DTU 30 MW TRWT is established using the developed method. The wind turbine is il-
lustrated in Fig. 1(b), and consists of a tower supported by three wires at the bottom, a rotor attached to the top of the
tower, and two lower rotors connected to the tower using two arms and two wires. From the high-fidelity linear model, the
low-order modal dynamics is accessed by analyzing its natural frequencies and mode-shapes as a function of rotor speed. At
the end of this section, the linear model is verified by comparing its natural frequencies with those from a nonlinear model
in HAWC2.

4.1. Campbell diagram of tri-rotor wind turbine

A Campbell diagram gives important insight into the dynamic characteristics of the system and is very beneficial in
the design phase of wind turbines. It helps to understand the nonlinear behavior of the system and shows if there is any
coincidence between natural frequencies and periodic excitations from the rotors. This knowledge can be used to develop
smart control strategies and fatigue load reduction. Following the procedure described in Fig. 3, a Campbell diagram for the
TRWT is established to analyze how the natural frequencies develop as function of rotor speed. The Campbell diagram is
shown in Fig. 6 where the three rotors have the same rotational speed. The Campbell diagram ranges from standstill to 15
rpm enabling proper visualization of gyroscopic and blade stiffening effects on natural frequencies. The frequency ranges
from 0–1.25 Hz because low-order modes dominate the system response. As seen in the Campbell diagram, the natural
frequencies cross as the rotor speed increases. Therefore, these works as an inadequate choice of reference for sorting.
Instead, if the eigenvalue analysis is performed at sufficiently small rotor speed increments, mode-shapes only experience
small changes. These can be compared using the modal assurance criterion [35,36] which is used to sort the Campbell
diagram.
The increased number of modes in the Campbell diagram, compared to a SRWT, is due to a more complex support-
structure but it is mostly due to the additional rotors. Adding another rotor introduces an additional set of symmetric and
asymmetric rotor modes as seen in Eq. (16).
To simplify the overview of the Campbell diagram, modes are divided into three categories: Wire modes, support-
structure modes, and rotor modes. Wire modes are defined as those with almost pure wire vibrations. Rotor modes are

9
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 7. The two rotors configurations on the tri-rotor wind turbine. (a) shows one of the lower rotors supported by the arm-structure and a wire, while (b)
the top rotor attached at the tower top.

Table 1
This figure shows rotor modes for the tri-rotor wind turbine. RS and AS refer to rotor symmetric and asymmetric
respectively for the rotor as illustrated in Fig. 7. Top rotor modes have the index T, and the lower rotors have the
index L. The first horizontal/vertical edge modes do not have an index because the first vertical edge modes for the
lower rotors are closely spaced with vertical and horizontal edge modes for the top rotor which makes them couple.
The superscript ∗ indicates same coupling to the support-structure, but that it couples to the second flapwise blade
bending frequency.

Rotor modes

Description Mode Frequency, Hz Mode-shape

First yaw (L, RS) 7 0.447 Fig. 8(c)


First yaw (L, RA) 8 0.448 Fig. 8(c)
First tilt (L, RS) 10 0.560 Fig. 8(b)
First tilt (L, RA) 11 0.561 Fig. 8(b)
First yaw (T) 12 0.581 Fig. 8(c)
First tilt (T) 13 0.592 Fig. 8(b)
First horizontal flap (T) 14 0.620 Fig. 8(a)
First horizontal flap (L, RA) 15 0.628 Fig. 8(a)
First horizontal flap (L, RS) 16 0.628 Fig. 8(a)
First horizontal edge (L, RA) 21 0.900 Fig. 8(a)
First horizontal edge (L, RS) 22 0.904 Fig. 8(f)
First horizontal/vertical edge 23, 24, 25, 26 0.930 Fig. 8(d-e)
Second yaw (L, RS) 27, 28 1.135 Fig. 8(c)∗
Second yaw (L, AS) 29, 30 1.136 Fig. 8(c)∗

dominated by the multi-blade coordinates etc. either a0 ,a1 or b1 (Eq. (13)) or combinations, wheres in support-structure
modes the multi-blade coordinates do not dominate the response.

4.1.1. Rotor modes


Examining Eq. (12), which describes the response of the individual blades in multi-blade coordinate the whirling modes
are only affected by its rotor speed. This can be interpreted as each rotor is going to behave like a SRWT. The only thing that
can make the response differ on a MRWT is the attachment point of the rotors being different. Fig. 7 illustrate the rotors
configuration on the TRWT. The lower rotors are each supported by a circular profile and a wire. As the wires transfer the
gravitational load from the two rotors to the tower it is possible to make a slender circular profile. But as a consequence,
the overall torsional stiffness is reduced since the wires cannot transfer torsion. Moreover, the wire introduces a direction-
dependent bending stiffness. The attachment point of the top rotor has a larger torsional stiffness due to the increased
tower diameter and symmetric bending properties.
Fig. 8 shows the flapwise and edgewise rotor modes at standstill for a SRWT and which modes they couples to as the
rotor speeds up. For a detailed explanation refers to work done by Hansen [37]. Each rotor on the TRWT is seen as a SRWT.
The reason for this is to adopt the same terminology used in Fig. 8. It can directly be used for the top rotor. However, for
the lower rotors motions related to yaw and tilt and vertical and horizontal motions are interchanged.
The various stiffness properties of the rotor attachment points result in three distinguished motions of the rotor modes:
1) Motion of the top rotor, 2) symmetric rotor motion of the lower rotors, and 3) asymmetric motion of the lower rotors.
An example of this behavior is visualized in Fig. 9 showing different variations of the symmetric flapwise modes (A0 ) at
standstill. Other combinations of simultaneous motion of all rotors or movement between the top rotor and one of the
lower rotors due occur as the natural frequencies of the top rotor and two lower rotors approaches.
The rotor modes present in Fig. 6 is given in Table 1. It is seen that the low torsional stiffness of the arm-structures has
a significant effect on the yaw modes for the lower rotors. The natural frequencies of the rotor symmetric and asymmetric
yaw modes are 0.45 Hz (mode 7 and 8). In comparison, the same mode for the top rotor is 0.58 Hz (mode 12). This is closer

10
O.T. Filsoof, M.H. Hansen, A. Yde et al.
11

Journal of Sound and Vibration 493 (2021) 115810


Fig. 8. The different variations of first flapwise and edgewise rotor modes at standstill. The text states which symmetric or asymmetric modes they couple to when rotating.
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 9. Combinations of symmetric flap modes for the tri-rotor wind turbine at standstill. Wires are colored red. In the symmetric case (a) only one rotor
is moving while in the rotor symmetric (b) and asymmetric (c) the two lower rotors couples. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

Fig. 10. Difference between the rotor symmetric (a) and asymmetric (b) second tilt mode. In the asymmetric mode, the arms bend out of phase which
couples to tower torsion resulting in a higher natural frequency compared to the symmetric mode.

to the first flapwise blade bending natural frequency at 0.61 Hz as also seen for SRWT. When rotating, the modes couples
to the first flapwise BW modes. This makes the natural frequencies decrease with −. The delayed splitting of the lower
rotors is due to the high asymmetry of the tilt/yaw stiffness. The two asymmetric flap modes cannot easily couple to create
the whirling because they are too separated in natural frequency at standstill. It needs a certain rotor speed before this
asymmetry is balanced out to make the whirling modes appear. Analyzing the modal amplitudes shows they are dominated
at by either the cosine (a1 ) or sine (b1 ) components at low rotors speeds (related to tilt or yaw). At higher rotor speeds,
the modes have both cosine and sine components representing whirling modes. The slope is not linear due to centrifugal
stiffening effects.
The tilt natural frequency for the top rotor at 0.59 Hz (mode 13) is slightly higher than for the lower rotors at 0.56 Hz
(mode 10 and 11) due to the lower tilt stiffness of the arm-structures compared to the tower top. The tilt modes couples
to the flapwise FW modes as the rotors starting rotating making them increases with +. The last modes related to the
first flapwise blade bending are the symmetric modes (Fig. 9). The symmetric mode for the top rotor equals 0.62 Hz (mode
14) and for the lower rotors 0.63 Hz (mode 15 and 16). The frequencies are marginally higher than the first flapwise blade
bending natural frequency because as the blades deflect forward the tower top or arm-tip makes a deflection in the opposite
direction as seen in Fig. 8. This reduces the relative movement of the blades, and, thus, reducing the modal mass.
As for the first yaw modes the second yaw modes related to the lower rotors at 1.14 Hz (mode 27 and 28) are affected
by the low torsional stiffness of the arm-structures. These modes represent a coupled motion between yawing at the ro-
tor attachment point and the second flapwise blade bending mode. The natural frequencies of the second tilt modes (not
present in Fig. 6) are 1.35 Hz for the rotor symmetric case, and 1.39 Hz for the rotor asymmetric case. The small difference
is found in how the symmetric and asymmetric modes couples to the tower as shown in Fig. 10. In the symmetric case, the
bending makes the tower twist in torsion whereas in the asymmetric case, the arms bend in-phase resulting in no coupling
in tower torsion. The second flapwise BW and FW modes are not affected significantly by the rotor speed. This is due to a
part of the blade deflection is out of phase with the yaw/tilt rotations. The resulting smaller reaction moments cause the
asymmetric stiffness of the tilt/yaw rotations of the support to have a larger impact on these modes, e.g. the gyroscopic
effect is minimized and, thereby, contributes less to the frequency split.
The first edgewise blade bending is related to the roll, vertical, and horizontal edge deflections as shown in Fig. 8. The
horizontal (mode 21, 22 and 23) and vertical (mode 24, 25, and 26) modes are all centered close to the first edgewise blade
bending natural frequency at 0.93 Hz in the Campbell diagram. There are no symmetric edgewise modes due to the fixed-
free boundary condition of the drive-train (the generator DOF is free in the RNA models). Modes for the top rotor are more
centered at standstill compared to the lower rotors and have higher natural frequencies. The mode-shapes revels that the
differences are because of the lower rotor couples in tilt. This coupling is stronger for the vertical modes explaining the
larger difference from the edgewise blade bending natural frequency. The coupling in tilt does not only lower the natural
frequencies but also introduces an amount of flapwise blade motion into the modes. At standstill, the horizontal edge mode
of the top rotor coincides with the vertical edge modes of the lower rotors, resulting in modes-shapes where all rotors
deflect simultaneously. As the rotor speed increases, the horizontal mode couples to the first FW edgewise mode which
increases with + while the vertical mode decrease with −. At a certain rotor speed, the modes decouple.

12
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Table 2
Support-structure modes for the tri-rotor wind turbine. For the first tower
fore-aft modes, symmetric (sym) refers to that the two lower rotors moves
in-phase with the tower and asymmetric (asym) out of phase. For the lateral
arm-bending modes, symmetric and asymmetric refers to in-phase and out
of phase arm-bending respectively.

Support-structure modes

Description Mode Frequency, Hz Mode-shape

First tower torsion 1 0.135 Fig. 11(a)


First tower fore-aft (sym) 2 0.137 Fig. 11(b)
First tower side-side 3 0.140 Fig. 11(c)
First tower fore-aft (asym) 4 0.168 Fig. 11(d)
Lateral arm-bending (asym) 5 0.411 Fig. 12(a)
Lateral arm-bending (sym) 6 0.413 Fig. 12(b)
Second tower side-side 17 0.635 Fig. 13(a)
Second tower fore-aft 20 0.798 Fig. 13(b)

Fig. 11. The four lowest modes for the tri-rotor wind turbine. Wires are colored red. (a) Tower torsion (mode 1), (b) first tower fore-aft, symmetric arm-
bending (mode 2), (c) first tower side-side (mode 3), (d) first tower fore-aft, asymmetric arm-bending (mode 4). (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of this article.)

The last observation is that the directional dependent bending stiffness of the arm-structure does not show any signifi-
cant effect on the natural frequencies. All frequency splits are centered close to the first edgewise blade bending frequency
and do split with ±, giving them almost the same behavior as seen for a SRWT.

4.1.2. Support-structure modes


The support-structure modes in the Campbell diagram are listed in Table 2. The tower torsion (mode 1), first tower fore-
aft with symmetric (mode 2) and asymmetric (mode 4) arm-bending, and first tower side-side (mode 3), have the lowest
natural frequencies in the system. Their modes-shapes are shown in Fig. 11. The first three modes have similar natural
frequencies and for mode 2 and 3 it is because they are dominated by similar bending stiffness and modal mass. For mode

13
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 12. Arm-bending modes for the tri-rotor wind turbine. Wires are colored red. (a) Lateral arm-bending, symmetric (mode 5), (b) lateral arm-bending,
asymmetric (mode 6). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 13. The two second tower modes. Wires are colored red. (a) Second tower side-side (mode 17), (b) Second tower fore-aft (mode 20). (For interpretation
of the references to colour in this figure legend, the reader is referred to the web version of this article.)

1 it is a result of the off-centered rotor mass from the tower combined with the tower torsional stiffness of the tower. Mode
4 where the arms move out of phase with the tower has a higher frequency due to the reduced movement of the rotors
compared to mode 2 when the tower and arm-structures are in phase. Generally, the natural frequencies of the mode 1 − 4
are unaffected by the rotor speed although mode 1 has a slight decrease in frequency at high rotor speeds. This decrease is
due to that the gyroscopic effects stabilize the torsional movement which lowers the natural frequency.
The symmetric and asymmetric lateral arm-bending modes (mode 5 and 6) have a natural frequency close to 0.41 Hz,
and their mode shapes are seen in Fig. 12. The geometrical symmetry condition of the TRWT makes the modes show up
in a symmetric and asymmetric pair. At standstill, the lateral arm-bending motion couples to the torsional motion of the
arm. As the rotors speed increases the torsional motion reduces significantly due to the increased angular momentum of
the rotors give a stabilizing effect. Both modes have a constant behavior in the diagram and are only affected locally at 5
rpm by coupling to the first flapwise BW modes of the lower rotors.
The second tower side-side, mode 17, and fore-aft, mode 20, shown in Fig. 13. Mode 17 is closely spaced with the flapwise
rotor and wire modes, which is seen in the mode-shape. Mode 20 shapes the tower as the second bending mode of a
cantilever beam which tilts the top rotor and bends the arm-structures. The bending of the arm-structure does not result in
a translational movement of the lower rotors, but a tilt motion instead. As the rotor speed increases, mode 20 couples with
the first edgewise BW and first flapwise FW modes at approximately 7 rpm.

4.1.3. Wire modes


The natural frequencies of the wire modes for the TRWT can simply be approximated by the natural frequency of free
vibrating string [38]

n T
fwire,n = (27)
2L μ
14
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Table 3
n T
First wire frequencies for the tri-rotor wind turbine using f wire,n = 2L μ for n = 1
and the tri-rotor wind turbine. Mode numbers for the wires are not given because
they are not present in Fig. 6. The mass per unit length for each wire is 12 kg m−1 .
The length of the wires supporting the tower is 96 m and the tension force for the
front wires is 132.210 kN, and the thus behind the rotor 174.816 kN. The wires sup-
porting the arm-structures is 119.3 m in length, and the tension is 11855.236 kN. The
initial length is used in the calculations.

Wire modes

Frequency, Hz
Description Mode
Wire equation Linear model

First tower wire (front of rotors) 9 0.55 0.56


First tower wire (behind of rotors) 18, 19 0.63 0.64
Second tower wire (front of rotors) 29, 30 1.10 1.14
First arm wire - 4.17 4.19

Fig. 14. Forward and backward whirling contributions to the first edgewise FW modes showing that at 9 rpm the modes changes from being the first FW
edge mode to a second BW flap mode. Modal amplitudes is normalized.

Where n = 1, 2, 3, . . . , ∞ is an integer giving the nth natural frequency, L is the wire length, T is the tension force in the
wire, and μ is the mass per unit length. Table 3 show the wire frequencies using Eq. (27) and the once identified in the
Campbell diagram. The error is less than 0.5 %, and the trend indicates that Eq. (27) underestimates the frequencies. This
underestimation is because Eq. (27) assumes that the wire is fixed in both ends, but in the coupled model the attachment
points are flexible. The lower natural frequency of the front wires compared to the wires behind is due to the small rotor
overhang tilts the wind turbine forward. Wire frequencies for multiples of n also match, and, as expected, the wire modes
are unaffected by the rotors speed.

4.1.4. Coupling effects


Several mode coupling effects occur in the Campbell diagram when two or more modes with similar deflections are
getting close. An example of this coupling effect is the sudden increase of one of the second flapwise BW modes at ap-
proximately 9 rpm. At this point, the mode starts following the first edgewise FW mode related to the top rotor. At the
same time, the first edgewise FW modes of the lower rotors decreases. This effect is seen in Fig. 14 showing how much the
edge FW and flap BW amplitudes contribute to the first edgewise FW mode for the lower rotors. From 2 to 9 rpm the BW
contribution increases until it reaches 1. At this rotor speed, the mode is influenced both by BW and FW. However, as the
rotor speed further increases the contribution from FW drops, and the local coupling effects vanish.

4.2. Comparison with nonlinear model

The high-fidelity linear model of TRWT is a linearization of a nonlinear model in HAWC2. Consequently, the linear model
shall represent the dynamic response of the nonlinear model around the point of linearization. To verify this, two Camp-
bell diagrams are made from the nonlinear model. The procedure for this is shown in Fig. 5. Fig. 15 show diagrams for
synchronous rotor speed while Fig. 16 shows when the speed of the top rotor and lower rotors differs.
Figs. 15 (a) and 16(a) shows the frequency content in the tower root torsional moment from the nonlinear model,
whereas (b)-(c) show how rotor modes develop for the top rotor and the lower rotors by showing the PSD for a sensor
at the arm-structure and one at the tower top.
Analyzing Fig. 15 the linear model captures the dynamical behavior of the nonlinear model well. The nonlinear model
shows several modes diverging from the first flapwise blade bending natural frequency at 0.6 Hz, This coincides with the
linear model. The nonlinear model confirms that the BW modes of the lower rotors are 0.45 Hz. Furthermore, the nonlinear

15
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

Fig. 15. Campbell diagrams from the nonlinear HAWC2 model of the tri-rotor wind turbine for (a) tower root torsional moment and the superimposed
solid lines are from the linear model, (b) tower top torsional moment and (c) arm-structure shear-force. All rotors have the same speed.

Fig. 16. Campbell diagrams from the nonlinear HAWC2 model of the tri-rotor wind turbine for (a) tower root torsional moment and the superimposed
solid lines are from the linear model, (b) tower top torsional moment and (c) arm-structure shear-force. The top rotor has twice the speed as the bottom
rotors. The x-axis in the diagrams refers to the speed of the bottom rotors.

model shows the tendency of the first flapwise whirling modes having a delay before attending the ± slope. The first
edgewise whirling modes are also present in the diagram for the nonlinear model at 0.9 Hz. The second flapwise BW modes
of the lower rotors are seen at 1.14 Hz. These are not seen clearly for rotors speeds less than 5 rpm. At 9 rpm the modes
start to attend the slope of the FW edge mode of the top rotors while the FW edge mode of lower rotors flattens out like
predicted by the linear model. The Campbell diagram in Fig. 16 shows that even if the speed differs between the rotors, the
linear model can predict the behavior of the nonlinear model.
Figs. 15 (b) and 16(b) only show frequency splitting of the asymmetric modes for the top rotor, while Figs. 15(c) and
16(c) shows almost identical splittings although the top rotor has twice the speed in Fig. 16. This means that the lower
rotors and top rotor behaves decoupled, and revisiting Eq. (12) it states that whirling modes only depend on the specific
rotor speed for a MRWT, which is confirmed by this observation.

5. Conclusions

This paper presented a novel methodology to convert the dynamic formulation of a MRWT to a linear time-invariant
system by applying Coleman transformation on each rotor. The methodology also works if the rotational speeds on a MRWT

16
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

are different. From the linear time-invariant system, modal properties can be extracted by direct eigenvalue analysis. The
developed methodology is a major contribution to the modeling of MRWTs. It shows that the stability of MRWTs can be
assessed using existing multi-body tools or analytical methods through proper linearization of the system matrices and
transformations.
To demonstrate this, a method to make high-fidelity linear time-invariant MRWT models has been proposed that uses
two aeroelastic tools: HAWC2 and HAWCStab2. A model of the DTU 30 MW TRWT is established focusing on analyzing the
TRWTs fundamental dynamics and verify the methodology and method. The verification involved comparing modal frequen-
cies from the linear model with ones obtained from a nonlinear model. The first verification case was for synchronous rotor
speeds. In the second case, the top rotor had twice the rotor speed as for the lower rotors. In both cases, the linear model
was able to capture the fundamental dynamics of the nonlinear model.
From the comprehensive analysis of the natural frequencies of the DTU 30 MW TRWT, the following is concluded:

• Rotor modes on a MRWT behave similarly to a SRWT.


• Geometrical symmetry in MRWTs results in mode-coupling effects. This is seen for the lower rotors on TRWT resulting
in a rotor symmetric and rotor asymmetric mode pairs.
• Low torsional stiffness influences rotor yaw modes by 1) lowering the natural frequency significantly and 2) results in a
delayed splitting of the first flapwise asymmetric modes due to the high asymmetry of the tilt/yaw stiffness.
• The first edgewise rotor modes are not significantly affected by the direction depended bending stiffness introduced by
the wires.

These important findings give a general fundamental understanding of the dynamic behavior of MRWTs and can be used
for control and design optimization. The results are not only limited to MRWTs but also provide important insight into the
modal dynamics of SRWTs with slender towers, e.g. cable-stayed towers.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

CRediT authorship contribution statement

Oliver Tierdad Filsoof: Conceptualization, Methodology, Software, Validation, Formal analysis, Data curation, Visualiza-
tion, Writing - original draft. Morten Hartvig Hansen: Conceptualization, Methodology, Writing - review & editing, Supervi-
sion. Anders Yde: Investigation, Writing - review & editing, Supervision. Peter Bøttcher: Conceptualization, Writing - review
& editing, Supervision. Xuping Zhang: Conceptualization, Methodology, Validation, Formal analysis, Writing - review & edit-
ing, Visualization, Supervision.

Acknowledgment

This work was support by Innovation Fund Denmark [grant 7038-00113A]; and Vestas Wind System A/S. Thanks to the
HAWCStab2 group at the DTU, Risø, Denmark for their assistance with the aeroelastic stability tool HAWCStab2.

References

[1] D.C. Janetzke, K.R. Kaza, Whirl flutter analysis of a horizontal-axis wind turbine with a two-bladed teetering rotor, Sol. Energy 31 (2) (1983) 173–182,
doi:10.1016/0 038-092X(83)90 079-8.
[2] M. Hansen, Vibrations of a three-bladed wind turbine rotor due to classical flutter, in: ASME 2002 Wind Energy Symposium, American Society of
Mechanical Engineers Digital Collection, 2002, pp. 256–266.
[3] B. Chen, Z. Zhang, X. Hua, S.R. Nielsen, B. Basu, Enhancement of flutter stability in wind turbines with a new type of passive damper of torsional
rotation of blades, J. Wind Eng. Ind. Aerodyn. 173 (December 2017) (2018) 171–179, doi:10.1016/j.jweia.2017.12.011.
[4] R.P. Coleman, A.M. Feingold, Theory of Self-Excited Mechanical Oscillations of Helicopter Rotors with Hinged Blades, Technical Report, 1957.
[5] D.J. Malcolm, Modal response of 3-bladed wind turbines, J. Sol. Energy Eng. (2002), doi:10.1115/1.1509479.
[6] M.H. Hansen, Improved modal dynamics of wind turbines to avoid stall-induced vibrations, Wind Energy 6 (2) (2003) 179–195, doi:10.1002/we.79.
[7] J.M. Jonkman, M.L.J. Buhl, FAST User’s Guide, Technical Report, National Renewable Energy Laboratory (NREL), Golden, CO, 2005, doi:10.2172/15020796.
[8] M.H. Hansen, Aeroelastic stability analysis of wind turbines using an eigenvalue approach, Wind Energy 7 (2) (2004) 133–143, doi:10.1002/we.116.
[9] M.H. Hansen, Aeroelastic properties of backward swept blades, in: 49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and
Aerospace Exposition, 2011, pp. 1–19, doi:10.2514/6.2011-260.
[10] E.A. Coddington, N. Levinson, Theory of Ordinary Differential Equations, Tata McGraw-Hill Education, 1955.
[11] G.W. Hill, On the part of the motion of the lunar perigee which is a function of the mean motions of the sun and moon, Acta Math. 8 (1) (1886) 1–36,
doi:10.1007/BF02417081.
[12] P.F. Skjoldan, Modal dynamics of wind turbines with anisotropic rotors, 47th AIAA Aerospace Sciences Meeting including the New Horizons Forum and
Aerospace Exposition, 2009, doi:10.2514/6.2009-1036.
[13] M.H. Hansen, Modal dynamics of structures with bladed isotropic rotors and its complexity for 2-bladed rotors, Wind Energy Sci. Discuss. 1 (2) (2016)
1–37, doi:10.5194/wes- 2016- 27.
[14] T. Kim, T.J. Larsen, A. Yde, Investigation of potential extreme load reduction for a two-bladed upwind turbine with partial pitch, Wind Energy 18 (8)
(2015) 1403–1419, doi:10.1002/we.1766.

17
O.T. Filsoof, M.H. Hansen, A. Yde et al. Journal of Sound and Vibration 493 (2021) 115810

[15] Dynamic modeling and stability analysis of a dual-rotor wind turbine, Volume 6: 14th International Conference on Multibody Systems, Nonlinear
Dynamics, and Control (2018), doi:10.1115/DETC2018-86142. International Design Engineering Technical Conferences and Computers and Information
in Engineering Conference, V006T09A022
[16] O.T. Filsoof, A. Yde, X. Zhang, Operational modal analysis of a multi-rotor wind turbine, in: The 29th International Ocean and Polar Engineering
Conference, International Society of Offshore and Polar Engineers, 2019.
[17] O.T. Filsoof, A. Yde, X. Zhang, Operational modal analysis of a multi-rotor wind turbine, The Science of Making Torque from Wind (TORQUE 2020),
2020.
[18] O.T. Filsoof, M. Hansen, A. Yde, P. Bøttcher, X. Zhang, Modal dynamics of a three-bladed tri-rotor wind turbine, Adv. Eng. Mater. Struct.Syst. (2019)
681–686.
[19] P. Jamieson, Innovation in Wind Turbine Design, Wiley Online Library, 2011, doi:10.1002/9781119975441.
[20] D. Tcherniak, S. Chauhan, M.H. Hansen, Applicability limits of operational modal analysis to operational wind turbines (2011) 317–327. doi:10.1007/
978- 1- 4419- 9716- 6_29.
[21] HAWC2 website, (http://hawc2.dk/a) Accessed: 2020-02-09.
[22] HAWCStab2 website, (http://www.hawcstab2.vindenergi.dtu.dk/b) Accessed: 2020-02-09.
[23] S. Sinha, R. Pandiyan, Analysis of quasilinear dynamical systems with periodic coefficients via Liapunov Floquet transformation, Int. J. Non Linear Mech.
29 (5) (1994) 687–702, doi:10.1016/0 020-7462(94)90 065-5.
[24] P.F. Skjoldan, Aeroelastic modal dynamics of wind turbines including anisotropic effects, vol. 66, Danmarks Tekniske Universitet, Risø Nationallabora-
toriet for Bæredygtig Energi, 2011.
[25] L. Meirovitch, Methods of Analytical Dynamics, Courier Corporation, 2010.
[26] P.F. Skjoldan, M.H. Hansen, On the similarity of the Coleman and Lyapunov-Floquet transformations for modal analysis of bladed rotor structures, J.
Sound. Vib. 327 (3–5) (2009) 424–439, doi:10.1016/j.jsv.2009.07.007.
[27] B. Akay, D. Ragni, C.S. Ferreira, G.J.W.V. Bussel, Wind turbine fatigue damage evaluation based on a linear model and a spectral method, Wind Energy
(September 2015) (2013) 1–20, doi:10.1002/we.
[28] I. Sønderby, M.H. Hansen, Open-loop frequency response analysis of a wind turbine using a high-order linear aeroelastic model, Wind Energy (2014),
doi:10.1002/we.1624.
[29] T. Larsen, T. Kim, Experimental and numerical study of rotor dynamics of a two- and three-bladed wind turbine, Int. J. Offshore Polar Eng. 26 (4)
(2016) 355–361, doi:10.17736/ijope.2016.mmr12.
[30] L. Bergami, M.H. Hansen, High-fidelity linear time-invariant model of a smart rotor with adaptive trailing edge flaps, Wind Energy 20 (3) (2017)
431–447.
[31] A.A. Shabana, Dynamics of Multibody Systems, 2013. doi:10.1017/CBO9781107337213.
[32] A. Yde, T.J. Larsen, A.M. Hansen, M. Fernandez, S. Bellew, Comparison of simulations and offshore measurement data of a combined floating wind and
wave energy demonstration platform, J. Ocean Wind Energy 2 (3) (2015) 129–137.
[33] A.K. Chopra, Dynamics of Structures - Theory and Applications to Earthquake Engineering, Prentice Hall, 2012.
[34] L. Ljung, System identification theory for user, 1987, doi:10.1016/0 0 05-1098(89)90 019-8.
[35] R.J. Allemang, D.L. Brown, A correlation coefficient for modal vector analysis, in: Proceedings of the 1st International Modal Analysis Conference, 1,
1982, pp. 110–116.
[36] R.J. Allemang, The modal assurance criterion–twenty years of use and abuse, Philos. Trans. R. Soc. A 37 (8) (2003) 14–23.
[37] M.H. Hansen, Aeroelastic instability problems for wind turbines, Wind Energy 10 (6) (2007) 551–577, doi:10.1002/we.242.
[38] S.S. Rao, Mechanical Vibrations Fifth Edition, 2010.

18

You might also like