You are on page 1of 16

Article

Journal of Vibration and Control


2019, Vol. 25(23–24) 2894–2909

On the controllers’ design to stabilize ! The Author(s) 2019


Article reuse guidelines:
sagepub.com/journals-permissions
ground resonance helicopter DOI: 10.1177/1077546319873797
journals.sagepub.com/home/jvc

José A. Ignácio da Silva1 , Douglas D. Bueno2 and


Gustavo L. C. M. de Abreu1

Abstract
Ground resonance (GR) in helicopters is a potentially catastrophic instability commonly caused by coalescence of the
regressive cyclic blade lag mode with the fuselage motion in certain rotor speed ranges. It can limit the helicopter
operational envelope and the design of this type of vehicle can become a difficult task. Although a broad class of actuators
allows the use of active and semi-active techniques to design feedback-based control systems, a limited number of works
in the literature introduce formulations to compute the controller gain to suppress this phenomenon. Also, commonly,
a control approach defines a feedback, particularly to a specific rotor speed. In this context, this work introduces an
alternative methodology to design an active control system to stabilize GR of a helicopter. The proposed approach can
suppress this instability in all rotor speed ranges by using only one control gain. Two strategies are proposed based on
linear matrix inequalities (LMIs). The Lyapunov stability criteria are used and the unstable rotor speed is considered as an
uncertain parameter to define an associated convex space. Using convex optimization, a robust control gain is computed
until all the unstable rotor speed range is stabilized. Numerical simulations are carried out to demonstrate the effect-
iveness of this methodology. The results confirm the viability of the proposed approach to design active and semi-active
controllers.

Keywords
Ground resonance helicopter, controller design, linear matrix inequalities, rotor speed range stabilization, robust
control gain

1. Introduction
Hammond (1974) used Floquet’s theory to predict
Ground resonance (GR) in helicopters is a potentially the GR boundaries of instabilities (i.e., the unstable
catastrophic instability commonly caused by the coales- rotor speed range). That approach can be used to
cence of the regressive cyclic blade lag mode with the study anisotropic rotors (blades different among
fuselage motion in certain rotor speed ranges (Chopra, them). In addition, different studies involving paramet-
1990). This phenomenon typically occurs in hinged ric vibrations and nonlinear analysis about GR phe-
(articulated) rotors. However, helicopters using nomenology for isotropic and anisotropic rotors can
the soft-inplane hingeless (Lytwyn et al., 1971) and be found in Janowski and Tongue (1988), Gandhi
bearingless rotors (Jang and Chopra, 1998) are also
susceptible to this problem. A classical analysis of this 1
Department of Mechanical Engineering, São Paulo State University
phenomenon was developed by Coleman and Feingold
(UNESP), Faculty of Engineering of Ilha Solteira, São Paulo State, Brazil
(1956) using a simple linear mechanical model. 2
Department of Mathematics, São Paulo State University (UNESP),
Despite its relative simplicity, the method accur- Faculty of Engineering of Ilha Solteira, São Paulo State, Brazil
ately predicts the GR boundaries of instabilities for Received: 18 September 2018; accepted: 24 January 2019
isotropic rotors (i.e., all blades similar among
them). The accuracy of that proposed method was Corresponding author:
José A. Ignácio da Silva, Department of Mechanical Engineering, São Paulo
verified theoretically and experimentally in the works State University (UNESP), Faculty of Engineering of Ilha Solteira, Avenida
of Gabel (1962) and Friedmann and Venkatesan Brasil 56, São Paulo State, Brazil.
(1985). Email: jose.i.silva@unesp.br
da Silva et al. 2895

and Malovrh (1999), Kowaleczko and Dzygadlo designing several controllers, is to switch the gain
(2002), Kunz (2002), Sanches et al. (2011, 2012, 2014), appropriately, which can involve unsafe procedures
Bergeot et al. (2016), Gourc et al. (2016), and references and require high computational capacity. In this con-
therein. text, this article proposes the use of linear matrix
The GR occurrence implies excessive vibrations inequalities (LMIs) to compute only one control gain
which can lead to a catastrophic destruction of a heli- to stabilize a rotor speed range in which GR of a heli-
copter if they are not reduced on time (Chopra, 1990; copter occurs.
Afolabi, 1993). In this sense, conventional approaches The LMIs have been used to solve mechanical and
to mitigate this instability involve the installation of lag aeromechanical problems—some of these are presented
dampers on the blade root. In general, they are passive in Johnson and Erkus (2005, 2006), Bueno et al. (2014),
hydraulic or elastomeric lag devices. Elastomeric dam- Mura and Lovera (2015), Koroishi et al. (2016),
pers were extensively studied by Kunz (1999) and Soliman et al. (2016), Wang et al. (2017); Yazici and
Brackbill et al. (2000). Bauchau and Liu (2006) pro- Sever (2018), and these formulations are commonly
posed a model of hydraulics lag dampers and validated based on the well-known Lyapunov’s theory and poli-
their results experimentally. Bergeot et al. (2017) stu- topic convex hulls (Boyd et al., 1994). The methods can
died the application of nonlinear energy sinks absorbers involve pole placement restrictions (Chilali and
mounted on the blades for GR suppression. Kim and Gahinet, 1996), saturation constraints (Benzaouia
Yun (2001) proposed the use of a new lag damper based et al., 2006; Soliman et al., 2016), energy-based specifi-
on piezoelectric actuation and a shunt circuit, and com- cations (Boyd et al., 1994; Gao et al., 2009), system
pare the results with a classical elastomeric lag damper. norms as H1 , H2, and L2 (Boyd et al., 1993; Chilali
According to these studies, passive lag dampers are and Gahinet, 1996; Li et al., 2013; Yazici and Sever,
designed to work under particular conditions and 2018). The typical solutions of these methodologies
their performance can be substantially reduced when involve convex optimization which can be efficiently
they are applied in different situations. This is an solved by using a powerful class of computational algo-
important limitation for practical engineering applica- rithms (Gahinet et al., 1995).
tions. In another context, the use of semi-active and In this context, the present article proposes a new
active control strategies is an attractive option to sup- approach to compute only one control gain to stabilize
pression of GR since they are capable of adjusting their GR instabilities in a rotor speed range. The rotor speed
dissipation levels for each rotor operational condition is considered as an uncertain parameter to formulate
(Panda et al., 2004). the method. The set of equations involving this uncer-
Recent researches have employed semi-active tech- tain parameter is easily described by a convex polytopic
niques to suppress the GR vibrations, also using mag- hull in a GR boundary. An LMI version of the linear–
netorheological (MR) lag dampers. Gandhi et al. (2001) quadratic regulator (LQR)- controller is used. Results
investigated the use of a MR damper with the robust and discussions based on the analytical developments
feedback linearization control law to alleviate the limit and numerical simulations are introduced to show the
cycle instabilities including nonlinear effects. Zhao et al. proposed strategy to stabilize the GR.
(2004) studied the application of the on–off, sliding
mode, and feedback linearization control methodolo-
gies using MR dampers for isotropic and anisotropic
2. GR dynamic model
rotor conditions. Other applications of this kind of The mechanical model used in this work captures the
device to suppress the GR phenomenon can be found essential features of the GR phenomenon. It is assumed
in Marathe et al. (1998), Kamath et al. (1999), and Hu that the helicopter on its landing gear can be repre-
and Wereley (2008). Similarly, a semi-active selective sented by effective parameters applied at the rotor
Coulomb friction-based lag damper is reported by hub (Coleman and Feingold, 1956). The fuselage on
Bauchau et al. (2010). the landing gear is considered anisotropic, which can
Although different methodologies have been pro- be represented by a rigid block with equivalent masses
posed in the literature to suppress GR vibrations in mfx along the x-direction and mfy along the y-direction;
helicopters, commonly they comprise control design respectively linked to linear springs with stiffness kfx
approaches which define controllers’ gain for each par- and kfy along the x, y directions. Only inplane motions
ticular steady rotor speed. However, this kind of of the rotor hub are further considered, on the longitu-
instability typically occurs in a continuous range of dinal x and lateral y directions, described by the xf and
rotation (Coleman and Feingold, 1956). In this sense, yf degrees of freedom (DOFs), respectively.
to consider a different control gain for each discrete The rotor-blades system is considered isotropic
rotor speed in a range can limit the use of active and (all blades are similar among them), operating through
semi-active techniques. The main difficulty, besides constant angular rotor speed () and have radius R. It
2896 Journal of Vibration and Control 25(23–24)

is composed by four blades (Nb ¼ 4) symmetrically According to the literature, for this problem
mounted on the rotor hub. Each blade presents the aerodynamic effects can be neglected (Coleman and
lead-lag motion nk (the blade in-plane motion) and its Feingold, 1956; Hammond,. 1974; Janowski and
location is determined by the azimuthal angle Tongue, 1988; Gandhi and Malovrh, 1999; Kim and
wk ¼ t þ 2 ðk  1Þ with x-axis (Coleman and Yun, 2001; Kunz, 2002; Zhao et al., 2004; Bielawa,
Feingold, 1956). The blades are considered uniform 2006; Bauchau et al., 2010; Sanches et al., 2011, 2012,
and rigid, also including the same mass mb, mass 2014; Bergeot et al., 2016, 2017); however, this is
moment Sb, and moment of inertia Ib around the dependent on the levels of flexibility and rotor speed.
z-axis in relation to the lag rotating center (point A in Also, the fuselage and the blade structural damping are
Figure 1 (b)). They have a root linear torsional spring neglected because the dissipative nature of damping
kb on the lag hinge. An illustrative scheme of this contributes to stabilize the system. In practice, neglect-
system is shown in Figures 1 (a) and 1 (b), where e is ing this energy dissipation the designed controller is
the lag hinge offset, b is the location of the blade’s more conservative. Neglecting damping, the estimated
center of mass (CM)—the point B in Figure 1 (b) level of control force can be higher than necessary to
measured from the lag hinge (point A in Figures 1 stabilize the damped system. This strategy is typically
(b)). This illustration is an important scheme to simplify employed for the problems of self-excited vibration, as
the understanding of the physical meaning behind the for an instance, the aeroelastic flutter (Broadbent and
equations presented herein. Williams, 1960; Bueno et al., 2015). In this context, if
the structural damping is considered, the formulation
presented herein remains valid. The mechanical model
used to illustrate the proposed approach is described by
(a)
six DOFs.

2.1. Mathematical formulation


The position of each kth blade CM (point B in Figure 1
(b)) is projected in the nonrotating coordinate system
(inertial coordinate system) and it is described by

xbk ¼ xf þ e cosð kÞ þ b cosð k  k Þ ð1aÞ

ybk ¼ yf þ e sinð kÞ þ b sinð k  k Þ ð1bÞ

(b) In this case, the total kinetic energy can be


expressed by

X
Nb
T ¼Tfþ T bk ð2aÞ
k¼1

where

1 
Tf¼ mfx x_ 2f þ mfy y_ 2f ð2bÞ
2
1 1  
T bk ¼ Ib ð  _k Þ2 þ mb x_ 2bk þ y_ 2bk ð2cÞ
2 2

and T f is the kinetic energy of the fuselage, and T bk


represents the kinetic energy of each kth blade.
Similarly, the total potential energy can be written by
Figure 1. Mechanical model of a four-bladed helicopter rotor
for ground resonance analysis: (a) top view from the model; and X
Nb

(b) detailed representation of the rotor-blade and hub for a k-th V ¼ Vf þ V bk ð3aÞ
blade. k¼1
da Silva et al. 2897

where moment, Sfx ¼ Sb =RM ~ fx and Sfy ¼ Sb =RM ~ fy are the


nondimensional
 longitudinal
 and lateral fuselage iner-
1 
tia, !2b ¼ Ibkb 2 þ Re Sb is the nonrotating lag blade fre-
Vf ¼ kfx x2f þ kfy y2f ð3bÞ ~ fx and !2 ¼ kfy =M ~ fy are the
2 quency, and !2fx ¼ kfx =M fy
longitudinal and lateral fuselage frequencies, with
1 M ~ fx ¼ mfx þ Nb mb and M ~ fy ¼ mfy þ Nb mb . This proced-
V bk ¼ kb k2 ð3cÞ
2 ure is a classical approach used to study the GR
problem, as shown in Coleman and Feingold (1956),
and V f is the potential energy of the fuselage, and V bk Hammond (1974), Janowski and Tongue (1988),
represents the potential energy of each kth blade. Gandhi and Malovrh (1999), Kim and Yun (2001),
The ground resonance model (GRM) is a system of Kunz (2002), Zhao et al. (2004), Bielawa (2006);
equations obtained through the direct application of Sanches et al. (2011, 2012, 2014), Bergeot et al.
the Lagrange’s equations. Considering the first-time (2016), Gourc et al. (2016), Bergeot et al. (2017), and
derivative of equation (1) and introducing the result the references therein.
into the equations of energy, the Lagrange’s equations It is important to highlight that in the literature there
can be used for each DOF. This is considered to be the are different models to study the GR, such as including
absence of external forces acting on the system. In add- the stiffness of the blades, additional DOFs of fuselage
ition, a first-order Taylor series expansion for the trig- and dynamics of shock absorbers mounted on the land-
onometric blade lag angle terms is applied and the ing gears (Lytwyn et al., 1971; Janowski and Tongue,
nonlinear terms are neglected, as often these are 1988; Jang and Chopra, 1998; Bielawa, 2006; Lee and
employed to study the GR stability. The result is a Kim, 2010). However, the model used herein is also
linear time periodic system of differential equations. widely used for studying the GR (e.g., Coleman and
For the cases where the controller is designed to sup- Feingold, 1956; Gandhi and Malovrh, 1999; Bielawa
press the system response due to external forces, as for 2006; Sanches et al., 2014; Bergeot et al., 2017), and it
an instance a turbulence load, the inclusion of non- is employed in this work to introduce the idea of stabi-
linear terms can be more relevant. However, these lizing a range of rotor speed using a single controller.
terms can be neglected if the designed controller is For a four-bladed helicopter rotor, equations
used to stabilize the system (Gandhi et al., 2001; (4a)–(4c) can be rewritten in a matrix form by
Zhao et al., 2004; Bauchau et al., 2010). In particular,
to simplify the design of the controller, it is considered MðÞq00 ðÞ þ GðÞq0 ðÞ þ KðÞqðÞ ¼ 0 ð5Þ
a nondimensional form by dividing these equations by
(2 R), as discussed in subsection 3.3. The six resultant where qðÞ ¼ fx f y f 1 2 3 4 gT is the vector of
nondimensional equations can be written by displacements, and MðÞ, GðÞ and KðÞ are, respect-
ively, the mass, gyroscopic, and stiffness matrices.
k00 þ Sb x 00f sinð kÞ  Sb y00f cosð kÞ þ !2b k ¼ 0 ð4aÞ This system of equations depends on the nondimensio-
nalized time s which physically corresponds to the time
X
Nb X
Nb change of scale.
x 00f þ Sfx k00 sinð kÞ þ2 Sfx 0k cosð kÞ
k¼1 k¼1
ð4bÞ 2.2. Coleman transformation
X
Nb
!2fx
 Sfx k sinð kÞ þ x f ¼ 0 The GRM contains periodic terms that depend on the
2
k¼1
azimuthal angle wk of each kth blade and it is commonly
named linear periodic system (Bielawa, 2006). In this
X
Nb X
Nb
y 00f  Sfy k00 cosð Sfy 0k sinð case, classical stability analysis and linear control tech-
kÞ þ 2 kÞ
k¼1 k¼1 niques are not suitable to deal with the periodic terms
2
ð4cÞ presented in the GRM. Therefore, it is convenient to
X
Nb ! fy
þ Sfy k cosð kÞ þ y ¼0 introduce the multiblade coordinate transformation
k¼1
2 f (MBC), also referred to as the Coleman transformation
(Coleman and Feingold, 1956). The MBC is a mathem-
where  ¼ t is nondimensional time, atical tool that transforms the system of equations from
ðÞ0 ¼ d ðÞ=d ¼ d ðÞ=dt and ðÞ00 ¼ d2 ðÞ=d 2 ¼ d2 ðÞ=2 dt2 a linear periodic rotating coordinate system (physical
indicate the first and second nondimensional time domain) to a nonrotating coordinate one (Coleman
derivatives. x f ¼ xf =R and y f ¼ yf =R are the normalized domain). This second domain is a linear time-invariant
longitudinal and lateral displacements, respectively. (LTI) mathematical system. For a helicopter with four
Sb ¼ Sb R=Ib is the nondimensional blade mass rotor-blades, the blade lag angle, nk, can be written in
2898 Journal of Vibration and Control 25(23–24)

terms of MBC coordinates by (Bielawa, 2006) Based on the above matrices the following state-
space realization can be used
k ¼ 0 þ 1c cosð kÞ þ 1s sinð kÞ þ ð1Þk 0 ð6Þ
x0 ¼ Ax ð10Þ
where the multi-blade coordinates are defined by

1 X
Nb
1 X
Nb where xðÞ ¼ fqðÞ q0 ðÞgT 2 R2n1 is the state vector,
0 ¼ k , 0 ¼ k ð1Þk ð7aÞ A 2 R2n2n is the dynamic matrix defined below and its
Nb k¼1 Nb k¼1
eigenvalues allow for understanding the system stability
(Chen, 1998). n ¼ 4 is the number of DOFs in this new
2 X Nb
2 X Nb
coordinate system.
1s ¼ k sinð k Þ, 1c ¼ k cosð kÞ ð7bÞ
Nb k¼1 Nb k¼1
 
044 I44
The multi-blade coordinates n0 and 0 represent the A¼ ð11Þ
collective and differential modes of the coupled mech- M1
c Kc M1
c Gc
anical system (rotor-blades/fuselage), respectively. 1c
and 1s are the cosine and sine components of the cyclic Using the above matrix, the GR can be identified by
modes, respectively (Bielawa, 2006). Substituting equa- studying the real part of its eigenvalues for each steady
tion (6) into equations. (4a)–(4c) and considering the  in the operational helicopter envelope (Bielawa,
MBC coordinates defined in equations (7a) and (7b) a 2006). Figure 2 shows an illustrative example of this
new system of equations can be defined. In particular, idea in which min and max define an unstable rotor
the collective and differential rotor modes are speed range. In this sense, according to the theories
decoupled from the other ones (fuselage and cyclic introduced in Section 1, the question herein is to com-
rotor modes) and the used transformation involves pute a unique control gain to stabilize this rotation
exact algebraic manipulation (no approximations). range, which makes the practical use of active and
The resulting system contains four DOFs, that is, the semi-active technologies easier.
longitudinal and lateral hub DOFs and the sine and
cosine ones. In this case, the new matrix form is given
by (see a detailed demonstration in Appendix A):

Mc q00c þ Gc q0c þ Kc qc ¼ 0 ð8Þ

where qc ¼ fx f yf 1s 1c gT is the new nondimen-


sional vector of displacements (described in the nonro-
tating coordinate system), Mc , Gc and Kc are,
respectively, the mass, gyroscopic, and stiffness new
matrices given by
2 3
1 0 2Sfx 0
6 7
6 0 1 0 2Sfy 7
Mc ¼ 6
6 
7
7 ð9aÞ
4 Sb 0 1 0 5
0 Sb 0 1
2 3
0 0 0 0
60 0 0 0 7
6 7
Gc ¼ 6 7 ð9bÞ
40 0 0 2 5
0 0 2 0
2 !2 3
fx
2 0 0 0
6 7
6 !2fy 7
6 7
Kc ¼ 6 0 2 0 0 7 ð9cÞ
6 7
4 0 0 !2b  1 0 5 Figure 2. Schematic representation of a typical ground
resonance boundary of instability, based on the Coleman
0 0 0 !2b 1 diagram.
da Silva et al. 2899

following MBC sine and cosine components of the con-


3. Controller design to suppress GR trol force
In this section, a new approach to compute a control
gain to stabilize the GR in a rotation range is intro- 2 X Nb
2 X Nb

duced. It is considered an actuator placed on each blade Ta1s ¼ Ta sinð k Þ, Ta1c ¼ Ta cosð kÞ
Nb k¼1 k Nb k¼1 k
as illustrated in Figure 3. To compute the control gain,
an LMI formulation of the LQR and a polytopic ð13Þ
description of the continuous rotation range are used.
In this case, the system of equations (equation (8))
can be rewritten, also including the actuators force
3.1. Actuator placement on each blade vector, as
To provide the control action, a hypothetical actuator
mounted on each blade is considered. The interaction 1
Mc q00c þ Gc q0c þ Kc qc ¼  Ta ð14Þ
of the control force (Fak ) with the actuator arm (see Ib  2 c
Figure 3) results in a control moment (Tak ) applied
just in a lag direction that is considered opposite to where Tac ¼ f0 0 Ta1s Ta1c gT is the vector of
the blade lag angle. A schematic illustration for control forces described in the nonrotating frame; the
the actuation attachment mechanism is shown in control input vector defined by uc ¼ fTa1s Ta1c gT can also
Figure 3. A similar mechanism can be found in conven- be used to consider the input matrix e
B such that
tional helicopters such as the Sikorsky CH-53E and
UH-60 (Warrier and Ali, 2017). Therefore, equation 2 3
0 0
(4a) can be rewritten including the control action 60
Tak in its left-hand side (instead of using zero). 6 07
7
Tac ¼ e
Buc , where e
B¼6 7 ð15Þ
Similarly, for a four-bladed helicopter rotor equation 41 05
(5) can be rewritten also including the left-hand side 0 1
term I1
b
2 Ta , where Ta ¼ f0 0 Ta1 Ta2 Ta3 Ta4 gT
is the control forces vector, resulting in equation 12
Introducing equation 15 into equation (14), the
00 0 1 state-space realization is given by
MðÞq ðÞ þ GðÞq ðÞ þ KðÞqðÞ ¼  Ta ðÞ ð12Þ
Ib  2
x0 ¼ Ax þ Bu uc ð16Þ
To write the control force vector in the nonrotating
coordinate system, it is necessary to consider the where Bu 2 R2nm is the control input matrix and m is
the amount of control inputs.
 
1 042
Bu ¼ e ð17Þ
Ib 2 M1
c B

3.2. LMI formulation to LQR problem


Given an LTI system, it is well-known that Lyapunov’s
theory introduces the existence of a quadratic energy
function VðxÞ ¼ xT Px. If a positive defined matrix
P 4 0 exists that satisfies the Lyapunov’s stability cri-
teria (Boyd et al. 1994)
:
_
VðxÞ ¼ x_ T Px þ xT P x 5 0, 8 x 6¼ 0 ð18Þ

the system is asymptotically stable (all trajectories con-


verge to zero as t ! 1) (Boyd et al., 1994). In particu-
lar for the nondimensional time (s) it is also possible to
Figure 3. Attachment of the actuation’s mechanism for ground assure the asymptotic stability considering the follow-
resonance mitigation. ing proposition.
2900 Journal of Vibration and Control 25(23–24)

Proposition 1. The LTI system, described in terms of W ¼ WT 2 R2n2n , Z 2 Rm2n and X 2 R2n2n exist,
nondimensionalized time s (equation (10)) is asymptot- such that they satisfy the following convex minimization
ically stable based on a state quadratic Lyapunov energy problem
function VðxÞ ¼ xT Px, if a positive defined matrix,
P 4 0, exists that satisfies the condition min TrðQWÞ þ TrðXÞ
subject to
dVðxÞ
5 0, 8x 6¼ 0 ð19Þ AW þ Bu Z þ WAT þ ZT BTu þ 2W 5 0 ð23Þ
d " #
X R1=2 Z
4 0, W 4 0
Proof. Differentiating the quadratic Lyapunov energy ZT R1=2 W
function VðxÞ ¼ xT Px on the nondimensionalized time
(s) and substituting in equation (19), results in where the Lyapunov matrix and the state feedback gain
are, respectively, given by P ¼ W1 and F ¼ W1 Z.

dVðxÞ dxT dx Proof. See Appendix A. The matrices W, X, and Z in


¼ Px þ xT P 5 0, P40 ð20Þ
inequality (23) are obtained by solving the convex mini-
d d d
mization problem and TrðÞ denotes the matrix trace.
Considering that  ¼ t, inequality (20) can be
rewritten as 3.3. Polytopic approach and GR boundaries
  of instability
1 dxT T dx
Px þ x P 50 ð21Þ
The main point in this article is to write each GR
 dt dt
boundary of instability (see Figure 2) in terms of poly-
with P 4 0. Once the rotor speed is always a positive topic parametric uncertainty. The strategy is to con-
number ð 4 0Þ, it is possible to multiply both sides of sider the rotor speed  as an uncertain parameter and
inequality (21) by  without changing its signal. In this define the associated polytopic convex hull. However,
case, the final result is equal to that in inequality (18). once the system of equations contains a quadratic term
To design the controller, an LMI based LQR prob- (1=2 , equation (16)) involving the rotor speed, an alge-
lem is implemented. Details of this formulation can be braic manipulation must be previously used to write an
found in Johnson and Erkus (2006) and it is used as a LMI. In this sense, a new uncertain parameter
decay rate  to improve the LQR transient response.  ¼ 1=2 is introduced and the associated system
The LQR–LMI controller has been used in the litera- ½A ¼ AðÞ, Bu ¼ Bu ðÞ is considered which depends on
ture as shown in, for example, Olalla et al. (2009), this new linear uncertain parameter limited through the
Soliman and Bajabaa (2013), and Soliman et al. (2016). minimum (min) and maximum (max) values, that is,
Consider an LTI system described by equation (16),  2 ½min , max  ¼ ½2 2
max , min .
the LQR-–LMI controller problem consists of obtain- In practice, the admissible values of h are con-
ing a control gain F 2 Rm2n such that uc ¼ Fx, that strained in a polytope in the parameter space Rnp
minimizes the quadratic performance index with n ¼ 2np vertices (in this case np ¼ 1). This poly-
Z tope is built by the convex combination of each vertex
1
(Boyd et al., 1994). For this problem, the polytope of
J¼ e2 ðxT Qx þ uTc Ruc Þd ð22Þ
0 an uncertain parameter is a line with two vertices,
n ¼ 2, in parameter space R that can be expressed by
subject to V0 ðxÞ 5 0 (Boyd et al., 1993), where
Q ¼ QT 4 0 2 R2n2n and R ¼ RT 4 0 2 Rmm are
P ¼ fg 2 Co f1 , 2 g
symmetric positive matrices that weigh the states and ( )
control energies, respectively (Anderson and Moore, Xn X
n
 li  i , li ¼ 1, li  0 , i ¼ 1, . . . , n
2007), and the pair ½A, Bu  must be controllable. The
i¼1 i¼1
solution of the LQR-–LMI controller problem is
ð24Þ
introduced by Theorem 1.

Theorem 1. (Olalla et al., 2009) The system described where 1 ¼ min and 2 ¼ max . Based on these values it
by equation (16) is asymptotically stable, in the sense of is possible to define the polytopic system by
Lyapunov’s theory, by a state feedback control law, 
u ¼ Fx, using a decay rate greater than or equal to ½A, Bu ðÞ 2 Co ½A, Bu i , . . . , ½A, Bu n 
 and guaranteed cost J, if the matrices
da Silva et al. 2901

( ) Table 1. Helicopter parameters used in the numerical


X
n X
n
 li ½A, Bu i , li ¼ 1, li  0 , i ¼ 1, . . . , n application (Hammond, 1974).
i¼1 i¼1 Parameter Value
ð25Þ
Number of blades, Nb 4
Rotor radius, R 5.64 m
The introduced model nondimensionalization allows Operational rotor speed, 0 31.42 rad/s
for the reduction of the polytope dimension. If the Blade mass, mb 94.9 kg
dimensional system is used, there are two terms to be Blade mass moment, Sb 289.1 kg.m
defined as uncertain parameters ( and 2). In this Blade mass moment of inertia, Ib 1084.7 kg.m2
case, the polytope must be a rectangle (four vertices)
Lag hinge offset, e 0.3048 m
which can be more complex to be solved by the convex
Lag spring, kb 0 N.m/rad
optimization algorithm in comparison with the solving
approach herein. Longitudinal hub mass, mfx 8026.6 kg
The main advantage of describing this problem by Lateral hub mass, mfy 3283.6 kg
convex polytopic hulls is the guarantee of achieving the Longitudinal hub spring, kfx 1240481.08 N/m
stability in all rotor speed ranges where the GR occurs. Lateral hub spring, kfy 1240481.08 N/m
Also, in practice, the LMI solutions must be done only
at the polytopic vertices instead of all points inside it
(Boyd et al., 1994). In other words, if it is possible to
find a pair of matrices Z and W such that all LMIs are modal damping diagrams (Figures 4 (a) and 4 (b)) are
simultaneously satisfied. used to characterize the stability of the rotor/fuselage
To achieve an asymptotically stable system behavior system in the context of GR.
in all rotor speed ranges ½min , max  using a unique Figure 4 (a) shows the Coleman diagram, which pre-
control gain it is necessary to write the LMI in equation sents the frequencies’ evolution through the rotor
(23) for each vertex of the polytope defined in equation speed. It is used as a normalized rotor speed to simplify
(25), as given by the following LMI the understanding. In addition, Figure 4 (b) shows the
damping behavior from which two system instabilities
min½TrðQWÞ þ TrðXÞ subject to are detected (the range of negative relative modal
Ai W þ Bui Z þ WATi þ Z BTui þ 2W 5 0
T damping). By comparing these results, it is possible to
identify the system instability in the coalescence of
i ¼ 1, 2 ð26Þ regressive blade mode (1s ) with the longitudinal fusel-
" #
X R 1=2
Z age mode (xf) and, subsequently, the lateral fuselage
T 1=2
4 0, W40 mode (yf) as the rotor speed increases, characterizing
Z R W
the GR occurrence. In this case two boundaries of
instabilities are observed because the fuselage is con-
where the Lyapunov matrix and the state feedback gain sidered anisotropic.
are given by P ¼ W1 and F ¼ W1 Z, respectively. Based on these results the unstable ranges of rotor
speed are min1 ¼ 0:4480    max1 ¼ 0:61260
and min2 ¼ 0:6690    max2 ¼ 1:020 . It is
4. Numerical application noted that the second range requires more damping
To illustrate the proposed approach numerical simula- to stabilize the GR.
tions are carried out considering the helicopter’s par-
ameters defined in Hammond (1974), for which the
4.2. Controller design
system’s physical and geometric properties are shown
in Table 1. The GR stability analysis is performed to To design the controller two strategies were considered.
find the GR boundaries and the control design is done The fuselage is assumed anisotropic which implies that
by using the introduced formulation. two unstable rotor speed ranges exist. For the first
strategy (Approach A), a unique polytope is used to
comprise the two boundaries of instabilities. The
4.1. GR identification system is represented by ½A, Bu ðÞ ¼ CofðAmin1 , Bmin1 Þ,
An eigenvalue analysis is initially performed using the ðAmax2 , Bmax2 Þg and the computed gain is FT .
matrix A (equation (11)) to determine the frequencies The second strategy (Approach B) consists of con-
(!j , 8j ¼ 1, . . . , 4) and modal damping (j , 8j ¼ sidering one polytope for each unstable rotor range. In
1, . . . , 4) for each system mode. The Coleman and the this case, a specific controller gain is computed to
2902 Journal of Vibration and Control 25(23–24)

(a)

(b)

Figure 5. Feedback scheme for active suppression of the


ground resonance phenomenon for both control Approaches
(A and B).

transformation is applied on the control inputs as fol-


lows (see Figure 5)
2 3
sinð 1Þ cosð 1Þ
6 sinð cosð 7
6 2Þ 2Þ 7
uðÞ ¼ Pc ðÞuc ðÞ, Pc ðÞ ¼ 6 7
4 sinð 3 Þ cosð 3Þ 5
sinð 4Þ cosð 4Þ
ð27Þ
Figure 4. (a) Coleman diagram for an undamped rotor system;
and (b) modal damping diagram for an undamped rotor system. where u ¼ fTa1 Ta2 Ta3 Ta4 gT is the physical con-
trol input vector and Pc is the inverse MBC coordinate
transform matrix; the resulting feedback gain is peri-
odic once it depends on the azimuthal angle of blades.
stabilize each boundary of instability. It is defined F1 For this numerical application the control gains are
for the first GR boundary, for which the system is computed by using Q ¼ 260  I88 and R ¼ 0:001
described by the polytope ½A, Bu ðÞ ¼ CofðAmin1 , I22 , where I is the identity matrix, a decay rate
Bumin1 Þ, ðAmax1 , Bumax1 Þg, and min1    max1 . T ¼ 0.01 is employed in Approach A, and 1 ¼ 0.16
Similarly, F2 is the control gain designed to stabilize and 2 ¼ 0.08 are used for Approach B. These values
the second GR boundary. The system polytopic were defined arbitrarily.
description is ½A, Bu ðÞ ¼ CofðAmin2 , Bumin2 Þ, ðAmax2 , Figures 6 (a) and 6 (b) show the Coleman diagram
Bumax2 Þg, where min2    max2 is the rotor speed and the modal damping diagram for the controlled
range. In practice, Approach B consists of switching system considering Approach A. It is possible to note
the control gain appropriately for each rotor oper- an important changing of relative eigenfrequencies
ational speed. (Figure 6 (a)) in comparison with the uncontrolled
To compute the control gain, the new coordinate system (Figure 4 (a)) once the coalescence effect does
system obtained from the Coleman transformation is not exist. Also, the stability can be verified through
considered. Based on this gain, the control forces uc Figure 6 (b) for the rotor speed range using one con-
are defined in the nonrotating coordinate frame and troller gain.
they need to be converted to the physical coordinates Based on the control Approach B, the Coleman dia-
for practical applications. In this case, an inverse MBC gram and the modal damping diagram are shown in
da Silva et al. 2903

(a) (a)

(b)
(b)

Figure 7. (a) Coleman diagram for the controlled system


(Approach B); and (b) modal damping diagram for the controlled
Figure 6. (a) Coleman diagram for the controlled system system (Approach B).
(Approach A); and (b) modal damping diagram for the controlled
system (Approach A).

Figures 7 (a) and 7 (b), respectively. A similar effect in


comparison with Approach A was noted. However, the
controlled system exhibits low values of relative eigen-
frequencies and modal damping once a specific control-
ler is designed for each unstable rotor speed range. In
this case, the control gain must stabilize a small convex
space, and because of this the second strategy can be
more convenient for practical applications.

4.3. Time response analysis


To demonstrate the controllers’ performances over time
the system responses are evaluated for  ¼ 0:560 and
 ¼ 0:80 . These values correspond to the rotor speeds
with smaller relative modal damping for the first and
second GR boundaries, respectively (see Figure 4 (b)).
The time responses are obtained by direct time integra-
tion of equations (5) and (12), respectively, for the Figure 8. Time response of normalized longitudinal fuselage
uncontrolled and controlled conditions. A fourth (xf =R) motion ( ¼ 0:560 ).
2904 Journal of Vibration and Control 25(23–24)

order Runge–Kutta method is used with initial condi- maximum control efforts are 768 N and 5.9 kN for
tion q0 ¼ f0:01 0:01 0 0 0 0gT and a time step Approach A, and 610 N and 1.9 kN for Approach B.
equal to 104 seconds. Figure 12 shows the lateral fuselage time response at
Figure 8 shows the longitudinal fuselage time  ¼ 0:80 for both uncontrolled and controlled condi-
response at  ¼ 0:560 . The results demonstrate the tions. The stability is also achieved using the proposed
assured stability for both Approaches A and B (the strategies. Figure 13 shows the phase map of lateral
equilibrium position is reached in approximately 1.0 fuselage motion for this rotor speed and some differ-
second). In addition, Figure 9 shows the phase map ences between the approaches. Figures 14 and 15 show
to be asymptotically stable and the differences between the control efforts acting on blades 1 and 2. Based on
these two control strategies. The control forces for Approach A, the maximum control force is 1.46 kN on
the actuators placed on blades 1 and 2 are shown in blade 1 and 5.9 kN on blade 2. For Approach B these
Figures 10 and 11, respectively. For these blades the maximum values are 1.26 kN and 2.6 kN, respectively.

Figure 11. Control effort on blade 2 (Fa2 ) – ( ¼ 0:560 ).


Figure 9. Phase map of longitudinal fuselage motion
( ¼ 0:560 ).

Figure 12. Time response of normalized lateral fuselage (yf =R)


Figure 10. Control effort on blade 1 (Fa1 ) – ( ¼ 0:560 ). motion for  ¼ 0:80.
da Silva et al. 2905

Figure 13. Phase map of lateral fuselage motion ( ¼ 0:80 ). Figure 15. Control effort on blade 2 (Fa2 ) – ( ¼ 0:80 ).

helicopter blades (Bauchau and Liu, 2006), modern


semi-active MR lag dampers (Zhao et al., 2004), and
semi-active lag friction dampers (Bauchau et al., 2010).

5. Conclusions
In this work an alternative approach to stabilize the
GR phenomenon in helicopters using active control
techniques and LMIs is presented. The classical mech-
anical model for GR analysis is employed.
The main contribution of this work consists of con-
sidering the GR boundaries of instabilities (i.e., an
unstable rotor speed range) as an uncertain parameter
which can be easily described by a convex polytopic.
An LMI version of the LQR controller with a pre-
scribed degree of the stability problem is employed.
The quadratic robust criteria ensure that all previously
Figure 14. Control effort on blade 1 (Fa1 ) – ( ¼ 0:80 ).
unstable rotor speed ranges become asymptotically
stable by using only one control gain. In particular,
Based on these results it is possible to conclude that the proposed method is applied to solve the GR prob-
both control Approaches A and B can be used to sta- lem considering the helicopter’s parameters defined in
bilize the GR phenomenon. In particular, Approach A Hammond (1974). Based on the present study, the fol-
can be more difficult to establish a numerical solution lowing conclusions are obtained:
once it considers just one convex space to comprise two
unstable rotor speed ranges. It can require large control 1. The nondimensionalization of the used GRM is
forces, but the use of the small decay rate  can avoid an important strategy to solve the problem once
high levels. On the other hand, the use of two control it reduces the convex space dimension. This
gains computed through Approach B requires a select- procedure allows the use of Lyapunov’s energy
ive feedback produce to switch the gain appropriately function to design the controller via LMIs
according to the rotor speed range. (Proposition 1).
The control efforts are calculated for each torque 2. The LQR-–LMI controller demonstrates high
Tai, i ¼ 1, . . . , 4, and they are converted to control performance through fast temporal system stabil-
forces by using the geometric actuator arm. ization. Time responses are shown for the more
The maximum force level is equal to 5.9 kN, which is critical rotor speeds (i.e., rotor speeds that require
representative of a typical hydraulic actuator for more damping for the GR suppression).
2906 Journal of Vibration and Control 25(23–24)

3. The two control Approaches (A and B) can be Boyd S, Balakrishnan V, Feron E, et al. (1993)Control system
used to suppress the GR phenomenon. Approach analysis and synthesis via linear matrix inequalities. In:
A is simpler to implement, but it can require Proceedings of 1993 American Control Conference, San
higher levels of control forces. On the other Francisco, California, USA, 2–4 June 1993. Piscataway,
hand, although Approach B requires a switching NJ: Institute of Electrical and Electronics Engineers,
pp.2147–2154.
procedure to set the control gain for each unstable
Boyd S, El Ghaoui E, Feron E, et al. (1994) Linear Matrix
rotor speed range, the convex optimization solu-
Inequalities in Systems and Control Theory, Volume 15,
tion may be simpler once it involves stabilizing a Studies in Applied Mathematics. Philadelphia, PA:
more limited convex space in comparison with Society for Industrial and Applied Mathematics.
Approach A. Brackbill CR, Lesieutre GA, Smith EC, et al. (2000)
Characterization of the low strain amplitude and fre-
quency dependent behavior of elastomeric damper mater-
Declaration of Conflicting Interests ials. Journal of the American Helicopter Society 45(1):
The author(s) declared no potential conflicts of interest with 34–42.
respect to the research, authorship, and/or publication of this Broadbent EG and Williams M (1960) The Effect of
article. Structural Damping on Binary Flutter, R & M 3169,
London: Ministry of Aviation, Aeronautical Research
Funding Council. Available at: http://naca.central.cranfield.ac.uk/
The author(s) would like to thank the São Paulo Research reports/arc/rm/3169.pdf (accessed September 2018).
Foundation (FAPESP), under Grant No. 2016/05988-2, for Bueno DD, Goes LCS and Goncalves PJP (2014) Control of
the financial support to this research. limit cycle oscillation in a three degrees of freedom airfoil
section using fuzzy Takagi–Sugeno modeling. Shock and
Vibration. Article ID 597827. pp.12. Available at: https://
ORCID iD www.hindawi.com/journals/sv/2014/597827/ (accessed
José A. Ignácio da Silva https://orcid.org/0000-0001-6960- September 2018).
250X Bueno DD, Goes LCS and Goncalves PJP (2015) Flutter
analysis including structural uncertainties. Meccanica
References 50(8): 2093–2101.
Chen C-T (1998) Linear System Theory and Design, 3rd ed.
Afolabi D (1993) The cusp catastrophe and the stability prob-
lem of helicopter ground resonance. Proceedings of the New York: Oxford University Press, Inc.
Royal Society of London A: Mathematical, Physical and Chilali M and Gahinet P (1996) H1 : design with pole place-
Engineering Sciences 441(1912): 399–406. ment constraints: An LMI approach. IEEE Transactions
Anderson BD and Moore JB (2007) Optimal Control: Linear on Automatic Control 41(3): 358–367.
Quadratic Methods. North Chelmsford, MA: Courier Chopra I (1990) Perspectives in aeromechanical stability of
Corporation. helicopter rotors. Vertica 14: 457–508.
Bauchau OA and Liu H (2006) On the modeling of hydraulic Coleman RP and Feingold AM (1956) Theory of self-excited
components in rotorcraft systems. Journal of the American mechanical oscillations of helicopter rotors with hinged
Helicopter Society 51(2): 175–184. blades. Techreport 1351, NACA. Available at: https://
Bauchau OA, Van Weddingen Y and Agarwal S (2010) ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/
Semiactive coulomb friction lead-lag dampers. Journal of 19930092339.pdf (accessed September 2018) .
the American Helicopter Society 55(1): 12005-1–12005-12. Feron E, Balakrishnan V, Boyd S, et al. (1992) Numerical
Benzaouia A, Tadeo F and Mesquine F (2006) The regulator methods for H2 related problems. In: Proceedings
problem for linear systems with saturations on the control of 1992 American Control Conference, Chicago,
and its increments or rate: An LMI approach. IEEE Illinois, USA, 24–26 June 1992. Piscataway, NJ:
Transactions on Circuits and Systems I: Regular Papers Institute of Electrical and Electronics Engineers,
53(12): 2681–2691. pp.2921–2922.
Bergeot B, Bellizzi S and Cochelin B (2016) Analysis of Friedmann P and Venkatesan C (1985) Coupled helicopter
steady-state response regimes of a helicopter ground res- rotor/ body aeromechanical stability comparison of theor-
onance model including a non-linear energy sink attach- etical and experimental results. Journal of Aircraft 22(2):
ment. International Journal of Non-Linear Mechanics 78: 148–155.
72–89. Gabel R (1962) Exact mechanical instability boundaries as
Bergeot B, Bellizzi S and Cochelin B (2017) Passive suppres- determined from the Coleman equation. Journal of the
sion of helicopter ground resonance using nonlinear American Helicopter Society 7(1): 17–23.
energy sinks attached on the helicopter blades. Journal of Gahinet P, Nemirovski A, Laub AJ, et al. (1995). LMI
Sound and Vibration 392: 41–55. Control Toolbox for Use With MATLAB. The
Bielawa R (2006) Rotary Wing Structural Dynamics and MathWorks, MA: Inc Natick. Available at: https://www.
Aeroelasticity. 2nd edition. Reston, VA: American researchgate.net/publication/3611303_LMI_control_tool-
Institute of Aeronautics and Astronautics, Inc. box_user’s_guide (accessed September 2018).
da Silva et al. 2907

Gandhi F and Malovrh B (1999) Influence of balanced rotor Li H, Liu H, Hilton C, et al. (2013) Non-fragile H1 control
anisotropy on helicopter aeromechanical stability. AIAA for half-vehicle active suspension systems with actuator
Journal 37(10): 1152–1160. uncertainties. Journal of Vibration and Control 19(4):
Gandhi F, Wang KW and Xia L (2001) Magnetorheological 560–575.
fluid damper feedback linearization control for helicopter Lytwyn RT, Miao W and Woitsch W (1971) Airborne and
rotor application. Smart Materials and Structures 10(1): ground resonance of hingeless rotors. Journal of the
96–103. American Helicopter Society 16(2): 2–9.
Gao C, Duan G and Jiang C (2009) Robust controller design Marathe S, Gandhi F and Wang KW (1998) Helicopter blade
for active flutter suppression of a two-dimensional airfoil. response and aeromechanical stability with a magnetor-
Nonlinear Dynamics and Systems Theory 9(3): 287–299. heological fluid based lag damper. Journal of Intelligent
Gourc E, Sanches L, Michon G, et al. (2016) Post-critical Materials Systems and Structures 9(4): 272–282.
analysis of ground resonance phenomenon: Effect of Mura R and Lovera M (2015) Rotorcraft vibration control:
stator asymmetry. Nonlinear Dynamics 83(1): 201–215. Adaptive vs LPV methods. IFAC-PapersOnLine 48(26):
Hammond CE (1974) An application of Floquet theory to 109–114.
prediction of mechanical instability. Journal of the Olalla C, Leyva R, Aroudi AE, et al. (2009) Robust
American Helicopter Society 19(4): 14–23. LQR control for PWM converters: An LMI approach.
Hu W and Wereley NM (2008) Hybrid magnetorheological IEEE Transactions on Industrial Electronics 56(7):
fluid-elastomeric lag dampers for helicopter stability aug- 2548–2558.
mentation. Smart Materials and Structures 17(4): 045021. Panda B, Mychalowycz E, Kothmann B, et al. (2004) Active
Jang J and Chopra I (1998) Ground and air resonance of an controller for Comanche air resonance stability augmen-
advanced bearingless rotor in hover. Journal of the tation. In: Proceedings of the 60th Annual American
American Helicopter Society 33(3): 20–29. Helicopter Society Conference, Volume 3, Baltimore,
Janowski M and Tongue B (1988) Construction and analysis Maryland, USA, 7–10 June 2004. New York: Vertical
of a simplified non-linear ground resonance model. Flight Society, pp.2048–2060.
Journal of Sound and Vibration 122(2): 233–241. Sanches L, Michon G, Berlioz A, et al. (2011) Instability
Johnson EA and Erkus B (2005) Investigation of dissipativity zones for isotropic and anisotropic multibladed rotor con-
for control of smart dampers via LMI synthesis. In: figurations. Mechanism and Machine Theory 46(8):
Proceedings of the 2005, American Control Conference, 1054–1065.
Volume 5, Portland, Oregon, USA, 8–10 June 2005. Sanches L, Michon G, Berlioz A, et al. (2012) Parametrically
Piscataway, NJ: Institute of Electrical and Electronics excited helicopter ground resonance dynamics with high
Engineers, pp.3084–3089. blade asymmetries. Journal of Sound and Vibration
Johnson EA and Erkus B (2006) Dissipativity and perform- 331(16): 3897–3913.
ance analysis of smart dampers via LMI synthesis. Sanches L, Michon G, Berlioz A, et al. (2014) Response and
Structural Control and Health Monitoring 14(3): 471–496. instability prediction of helicopter dynamics on the
Kamath GM, Wereley NM and Jolly MR (1999) Analysis ground. International Journal of Non-Linear Mechanics
and testing of a model scale magnetorheological fluid heli- 65: 213–225.
copter lag mode damper. Journal of American Helicopter Soliman HM and Bajabaa NS (2013) Robust guaranteed-cost
Society 44: 234–248. control with regional pole placement of active suspensions.
Kim SJ and Yun CY (2001) Performance comparison Journal of Vibration and Control 19(8): 1170–1186.
between piezoelectric and elastomeric lag dampers on Soliman HM, Benzaouia A and Yousef H (2016) Saturated
ground resonance stability of helicopter. Journal of robust control with regional pole placement and applica-
Intelligent Material Systems and Structures 12(4): 215–222. tion to car active suspension. Journal of Vibration and
Koroishi E, Lara-Molina F, Borges A, et al. (2016) Robust Control 22(1): 258–269.
control in rotating machinery using linear matrix inequal- Wang M, Cole MOT and Keogh PS (2017) New LMI based
ities. Journal of Vibration and Control 22(17): 3767–3778. gain-scheduling control for recovering contact-free oper-
Kowaleczko G and Dzygadlo Z (2002) Effect of free play in ation of a magnetically levitated rotor. Mechanical
rotor blades hinges on the non-linear ground resonance of Systems and Signal Processing 96: 104–124.
a helicopter. Aircraft Engineering and Aerospace Warrier J and Ali SF (2017) Control of ground resonance in
Technology 74(6): 508–516. helicopters using semi active damping, In: Proceedings of
Kunz D (1999) Elastomer modelling for use in predicting 2017 Indian Control Conference (ICC), Guwahati, India,
helicopter lag damper behavior. Journal of Sound and 4–6 January 2017. Piscataway, NJ: Institute of Electrical
Vibration 226(3): 585–594. and Electronics Engineers, pp.111–116.
Kunz DL (2002) Nonlinear analysis of helicopter ground res- Yazici H and Sever M (2018) L2 gain state derivative feed-
onance. Nonlinear Analysis: Real World Applications 3(3): back control of uncertain vehicle suspension systems.
383–395. Journal of Vibration and Control 24(16): 3779–3794.
Lee Y-K and Kim K-J (2010) Ground resonance analysis for Zhao Y, Choi YT and Werely N (2004) Semi-active damping
an eight-degrees-of-freedom rotorcraft with double-stage of ground resonance in helicopters using magnetorheolo-
oleo-pneumatic shock absorbers. Journal of Aircraft gical dampers. Journal of American Helicopter Society 15:
47(5): 1647–1655. 468–492.
2908 Journal of Vibration and Control 25(23–24)

In addition, Feron et al. (1992) introduced that the


Appendix A: Proof of Theorem 1 nonlinear term, TrðR1=2 ZW1 ZT R1=2 Þ can be replaced
x ¼ e x and e
Let e uc ¼ e . These new variables are used by a second auxiliary variable X, which satisfies the
to define the linear–quadratic regulator (LQR)- per- following convex minimization problem
formance index J from the classical LQR technique
such that min TrðXÞ subject to X 4 R1=2 ZW1 ZT R1=2 ð35Þ
Z 1
T which applying the Schur’s complement (Boyd et al.,
J~ ¼ x Qe
ðe uTc Re
x þe uc Þd ð28Þ
0 1994) becomes
" #
subject to V0 ð~
xÞ 5 0. In addition, the system dynamics X R1=2 Z
considering the decay rate  is given by Anderson and min TrðXÞ subject to 40
ZT R1=2 W
Moore, (2007)
ð36Þ
0
~ ¼ ðA þ IÞ~
x x þ Bu ~
uc ð29Þ
Finally, the complete formulation for the LQR-–
and the state feedback control law in terms of the new linear matrix inequality (LMI) problem is
variables e
x and euTc is ~ ~ x, where F
uc ¼ F~ ~ ¼ F.
Consider a typical Lyapunov’s function min ½TrðQWÞ þ TrðXÞ subject to
Vð~
xÞ ¼ x~T P~
x, substitute equation (29) into the AW þ Bu Z þ WAT þ ZT BT þ 2W 5 0
Lyapunov’s inequality (equation (28)) and introduce " # ð37Þ
the state feedback control law ~ x, the perfor-
uc ¼ F~ X R1=2 Z
4 0, W 4 0
mance index J~ (equation (28)) is rewritten by ZT R1=2 W
Z 1
T where W ¼ P1 and the gain of state feedback is
J~ ¼ x ðQ þ FT RFÞe
e x d ð30Þ
0 obtained by F ¼ ZW1 . Based on the above
results the LQR-–LMI is introduced, quod erat
subject to ðAT þ FT BTu ÞP þ PðA þ Bu FÞ þ 2P 5 0. Pre demonstrandum.
and post-multiply this inequality by P1 , and let
W ¼ P1 , to obtain Appendix B: Ground resonance
(GR) model
WðAT þ FT BTu Þ þ ðA þ Bu FÞW þ 2W 5 0 ð31Þ
The Lagrange’s equations applied for the GR problem
Applying the matrix trace operator TrðÞ on the per- can be defined as
formance index J~ (equation (30)) and after some alge-

d @T @T @V
braic rearrangements involving matrix trace operator  þ ¼ Qx ð38aÞ
(Chen 1998), is obtained dt @x_ f @xf @xf
!
d @T @T @V
 þ ¼ Qy ð38bÞ
J~ ¼ TrðQWÞ þ ðFT RFWÞ ð32Þ dt @y_f @yf @yf

R1

where W ¼ 0 xxT d satisfies the Lyapunov’s inequal- d @T @T @V
 þ ¼ Qk ð38cÞ
ity in equation (31) (Boyd et al., 1993; Johnson and dt @_k @k @k
Erkus, 2006). A new variable Z ¼ FW is introduced
once equation (31) is not linear. In this case where, Qx, Qy, and Qk are the external forces acting on
the system. Considering the first-time derivative of
WAT þ ZT BTu þ WA þ Bu Z þ 2W 5 0 ð33Þ equation (1) and introducing the result into equation
(2c), the blade’s kinetic energy is accounted for. Then,
The matrix R can be decomposed as R ¼ R1=2 R1=2 substituting the energy expressions (equations (2a)–(3c))
because it is symmetric and positive defined. Based on in the Lagrange’s equations (equations (38a)–(38c)) con-
this result, and using equation (32), the LQR- perfor- sidering the absence of external forces acting on the
mance index, J~ can be rewritten by system, the equations of motion can be derived as

TrðQWÞ þ TrðR1=2 ZW1 ZT R1=2 Þ ð34Þ Ib €k þ Sb x€ f sinð k  k Þ  Sb y€f cosð k  k Þ


da Silva et al. 2909

The system defined in equations (40a)–(40c) can be


þ eSb 2 sinðk Þ þ kb k ¼ 0, 8k ¼ 1, . . . , Nb ¼ 4
re-written in terms of multiblade coordinate transform-
ð39aÞ ation (MBC) coordinates, by introducing equation (6)
into equation (40a) and writing equations (40b) and
X
Nb
(40c) in terms of MBC coordinates, resulting in
~ fx x€ f þ Sb
M €k sinð k  k Þ
k¼1
ð39bÞ 000 þ ð!2b  1Þ0 ¼ 0 ð41aÞ
X
Nb
2
 Sb ð  _k Þ cosð k  k Þ þ kfx xf ¼ 0
k¼1 000 þ ð!2b  1Þ0 ¼ 0 ð41bÞ

X
Nb 00
1s þ Sb x 00f  21c
0
þ ð!2b  1Þ1s ¼ 0 ð41cÞ
~ fy y€ f  Sb
M €k cosð k  k Þ
k¼1
ð39cÞ 00
1c  Sb y 00f þ 21s
0
þ ð!2b  1Þ1c ¼ 0 ð41dÞ
X
Nb
2
 Sb ð  _k Þ sinð k  k Þ þ kfy yf ¼ 0 !fx2
00
k¼1 x f þ 2Sfx 1s
00
þ x f ¼ 0 ð41eÞ
2
Following Coleman and Feingold (1956), Hammond 00 !fy2
(1974), Janowski and Tongue (1988), Gandhi and y f  2Sfy 1c
00
þ y f ¼ 0 ð41fÞ
2
Malovrh (1999); Kim and Yun (2001), Kunz (2002),
Zhao et al. (2004), Bielawa (2006), Bauchau et al. The collective and differential rotor modes are
(2010), Sanches et al. (2011, 2012, 2014), Bergeot decoupled from the other ones (fuselage and cyclic
et al. (2016, 2017), Warrier and Ali (2017) and the ref- rotor modes), thence they can be eliminated from the
erences therein, a small lagging angle motion (k  1) analysis. The resulting system contains four degrees of
around the trivial equilibrium position of the system freedom (DOFs), that is, the longitudinal and lateral
was assumed, thus equations (39a)–(39c) are expanded hub DOFs and the sine and cosine ones, that can be
in a first-order Taylor series resulting in a linear system compacted in the following matrix form
with periodic coefficients. A nondimensional form of
this is considered by dividing these equations by Mc q00c þ Gc q0c þ Kc qc ¼ 0 ð42Þ
(2 R), resulting in
where qc ¼ fx f y f 1s 1c gT is the nondimensional
k00 þ Sb x 00f sinð   00 cosð
k Þ  Sb y f kÞ þ !2b k ¼0 ð40aÞ vector of displacements (described in the non-rotating
coordinate system), Mc , Gc and Kc are respectively the
X
Nb X
Nb mass, gyroscopic and stiffness matrices given in equa-
x 00f þ Sfx k00 sinð kÞ þ2 Sfx 0k cosð kÞ tions (9a)–(9c).
k¼1 k¼1
ð40bÞ
X
Nb
!2fx
 Sfx k sinð kÞ þ x f ¼ 0
k¼1
2

X
Nb X
Nb
y 00f  Sfy k00 cosð kÞ þ2 Sfy 0k sinð kÞ
k¼1 k¼1
2
ð40bÞ
X
Nb ! fy
þ Sfy k cosð kÞ þ y ¼0
k¼1
2 f

You might also like