Velocity feedback damping of piezo-actuated wings

Velocity feedback damping of piezo-actuated wings

Accepted Manuscript Velocity feedback damping of piezo-actuated wings Xiao Wang, Marco Morandini, Pierangelo Masarati PII: DOI: Reference: S0263-8223...

1MB Sizes 1 Downloads 40 Views

Accepted Manuscript Velocity feedback damping of piezo-actuated wings Xiao Wang, Marco Morandini, Pierangelo Masarati PII: DOI: Reference:

S0263-8223(17)30423-3 http://dx.doi.org/10.1016/j.compstruct.2017.04.016 COST 8447

To appear in:

Composite Structures

Please cite this article as: Wang, X., Morandini, M., Masarati, P., Velocity feedback damping of piezo-actuated wings, Composite Structures (2017), doi: http://dx.doi.org/10.1016/j.compstruct.2017.04.016

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

Velocity feedback damping of piezo-actuated wings Xiao Wang∗, Marco Morandini, Pierangelo Masarati Dipartimento di Scienze e Tecnologie Aerospaziali, Politecnico di Milano, Via La Masa 34, 20156 Milano, Italy

Abstract A geometrical nonlinear model of thin-walled beams with fiber-reinforced and piezo-composite is developed for smart aircraft wing structures. Some nonclassical effects such as warping inhibition and three-dimensional (3-D) strain are accounted for in the beam model. The governing equations and the corresponding boundary conditions are derived using the Hamilton’s principle. The Extended Galerkin’s Method is used for the numerical study. A negative velocity feedback control algorithm is adopted to control the aircraft wing response. The effective damping performance is optimized by studying anisotropic characteristics of piezo-composite and elastic tailoring of the fiber-reinforced host structure. The relations between active vibration control effect and design factors, such as the size and position of piezo-actuator are investigated in detailed. Keywords: thin-walled beam, fiber-reinforced, piezo-composite, active control

Nomenclature aij

1-D global stiffness coefficients

AF i

Piezo-actuator coefficients, see Appendix A

bij

Inertial coefficients

2b, 2d

Width and depth of the beam cross-section, see Fig. 1

˜w , B ¯w Pure mechanical bimoment, piezo-bimoment actuation, and Bw , B

∗ Corresponding

author Email addresses: [email protected] (Xiao Wang ), [email protected] (Marco Morandini ), [email protected] (Pierangelo Masarati )

Preprint submitted to Composite Structure

April 15, 2017

external bioment on the beam tip Ei

Electric field intensity

Fw , a(s)

Primary and secondary warping functions

ki

Feed back control gain in Eq. (26)

ˆ h

Electrode spacing of the interdigitated electrode

L

Length of the beam, see Fig. 1

Mx , My , Mz 1-D stress couples of pure mechanical contributions ˜ x, M ˜y, M ˜ z Piezo-moment actuations about the x, y and z axes M ¯ x, M ¯y, M ¯ z External stress couples on the beam tip M N

Total number of constituent layers in the wall

Np , Nh

Numbers of piezocomposite and host layers in the wall

Qx , Qz

Shear forces in the x− and z−directions

˜x, Q ˜z Q

Piezo-transverse-shear actuations in the x− and z−directions

¯x, Q ¯z Q

Applied transverse shear forces at the beam tip

R(y)

Distribution function along span for the actuator

(s, y, n)

Local coordinate system on the cross-section, see Fig. 1

Ty , T˜y , T¯y

Generalized axial force per unit span, piezo-extension actuation, and applied axial force at the beam tip, respectively

u0 , v0 , w0

Displacement components of the cross-section along x, y, z axes, see Fig. 1

αik , βik

Weighting coefficients for control gains in Eq. (26)

θ

Ply-angle of layer, see Fig. 1

θh , θp

Ply-angles of host structure and piezo-actuator

θx , θz , φ

Rotations of the cross-section about the x, z and y axes, see Fig. 1

ψ(s)

Torsional function

˜t Γt , Γ

Pure mechanical nonlinear terms in twist motion and piezo-nonlinear actuation, respectively

δ ˙ (),() ¨ 0 (),

Variation operator

XT

Transpose of the matrix or vector X

∂()/∂t, ∂ 2 ()/∂t2 , ∂()/∂y

2

1. Introduction Composite thin-walled beams are widely used because of their potential advantages. Among them, light weight, specific high stiffness, and elastic couplings 5

are worth citing. Refined thin-walled beam models have been proposed for open or closed cross-section (see e.g. [1–6]). Among these efforts, Cortinez et al. [3] and Vo and Lee [4] introduced warping shear to improve the model’s accuracy. More generally, any geometrically exact, intrinsic theory of anisotropic beams can be used provided the beam constitutive equation are computed correctly.

10

Many similar approaches, based on a semi-analytical discretization of the beam section displacement field, have been proposed to compute the stiffness matrix of arbitrarily complex beam sections, see e.g. [1, 7, 8]. Basically the same approach can be used for the characterization of composite beams with piezoelectric patches [9, 10]. That said, simplified models such as the one used in

15

this paper are still interesting, as they allow to get a better inside into the dependence of the elastic solution of the beam section parameters. Piezoelectric material are advantageous because of their fast response. As such, they are well suited for the active control of deformable beams [11–15]. Due to the brittle nature of ceramics, they are however vulnerable to damage and can

20

hardly conform to a curved surface. These drawbacks are overcame by piezocomposite materials such as the Active Fiber Composite (AFC) [16] and the Macro-Fiber Composite (MFC) [17]. Piezo-composite materials can be shaped and bonded to surfaces or embedded into structures. Thanks to their anisotropic characteristics, piezo-composite actuators can produce a twisting actuation that

25

can help controlling the vibration of helicopter blades, see e.g. [18, 19]. In this paper, a geometrical nonlinear thin-walled beam theory incorporating fiber-reinforced and piezo-composite is developed. Transverse shear strain, primary and secondary warping inhibition and three-dimensional strain are accounted for. A typical circumferential asymmetric stiffness (CAS) lay-up con-

3

30

figuration is adopted to model the aircraft wing structure. The twist-bending elastic coupling induced by the CAS lay-up is beneficial for the aeroelastic response behavior [14, 20, 21], especially for the suppression of the flutter instability [22–24]. Numerical studies based on the Extended Galerkin’s Method are performed, and the relation between actuation performance and voltages is

35

investigated. Based on a negative velocity feedback control algorithm, the dynamic response of an aircraft wing excited by an impulsive load is investigated. Furthermore, the control efficiency is optimized via the study of anisotropic characteristics of piezo-actuator and the elastic tailoring of fiber-reinforced host structure. Finally, considering the high cost and high density of piezo-composite

40

materials, the effect of size and position of the piezo-actuator is investigated.

2. Structural elements of a geometrically nonlinear theory of anisotropic thin-walled beams A geometrical nonlinear, single-cell, closed cross-section, fiber-reinforced composite thin-walled beam model incorporating piezoelectric materials is consid45

ered. The geometric configuration and the chosen corresponding coordinate system are indicated in Figs. 1. In addition to the global coordinate (x, y, z), a local coordinate (s, y, n) is defined on the mid-line contour of the cross-section, see Fig. 1. 2.1. Constitutive relations We assume that both fiber-reinforced composite material and the piezocomposite material can be modeled with linear piezoelectric constitutive relationships. The constitutive equations of a 3-D piezoelectric continuum can be expressed as [25, 26] σij = cE ijkl εkl − ekij Ek ,

(1a)

Di = eikl εkl + κεik Ek ,

(1b)

4

y (v ) 0

z (w0)

ϕ

θz θ

2d

s

θx

L

x

n

(u0)

2b Figure 1: Geometry of the beam with a rectangular cross-section (CAS lay-up).

50

ε where, cE ijkl , ekij , and κik denote the elastic stiffness coefficients, the piezoelec-

tric stress tensor and the dielectric constant tensor, respectively. The superscripts E and ε denote constant electric field and constant strain, respectively. Symbols σij and εkl denote the stress and strain components, while Ek and Di denote the electric field intensity and electric displacement vector, respectively. 55

Eq. (1a) describes the converse piezoelectric effect which is used for distributed sensing while Eq. (1b) describes the direct piezoelectric effect that is used for the active distributed control. The constitutive equations for the kth layer, reduced to the plane stress condition σnn = 0 and referred to the local coordinate system (s, y, n) are            ¯ 11 Q ¯ 12 ¯ 16       σss  Q 0 0 Q εss  ess                        ¯            ¯ ¯        σ Q Q 0 0 Q ε e yy  12 22 26 yy  yy                ¯ 44 Q ¯ 45 = 0 − E1(k) . τyn 0 Q 0  γyn 0                         ¯ 45 Q ¯ 55    0 γsn     τsn  0 Q 0   0                            τ     ¯ ¯ ¯ Q16 Q26 0 0 Q66 γsz esy  sy (k)

(k)

(k)

(k)

(2) We assume constant electric filed through the actuator thickness, i.e., E1 = ˆ where V and h ˆ are the applied voltage and electrode spacing of the −(V /h), 5

interdigitated electrode for the actuator layer, respectively. And the reduced piezoelectric stress coefficients can be computed as C¯13 ess = m2 e11 + n2 e12 − ¯ e13 , C33

C¯23 eyy = n2 e11 + m2 e12 − ¯ e13 , C33

C¯36 esy = mn (e11 − e12 ) − ¯ e13 , C33

(3a)

(3b)

where m ≡ cos θ and n ≡ sin θ, θ ∈ [0, 2π]. 2.2. 2-D piezoelectric constitutive equations The host structure is assumed composed of Nh layers; the piezo-actuator is composed of Np piezocomposite layers. The distribution function R(·) of actuators can be given by (see Fig. 2):

60

Rk (n) = H(n − n(k−1) ) − H(n − n(k) ),

(4a)

Rk (s) = H(s − sk1 ) − H(s − sk2 ),

(4b)

Rk (y) = H(y − yk1 ) − H(y − yk2 ),

(4c)

where H(·) denotes Heaviside’s distribution, while (n(k−1) , n(k) ), (sk1 , sk2 ) and (yk1 , yk2 ) denote, in sequence, the top and bottom heights of kth piezoelectric layer measured across the beam thickness, its location along the beam circumference and its span location. Considering the definition of the stretching quantity Aij , of the bendingstretching coupling stiffness quantity Bij and of the thermal and hygric moments Dij , (Aij , Bij , Dij ) =

N Z X k=1

n(k)

¯ (k) (1, n, n2 ) d n, Q ij

(N = Nh + Np ),

(5)

n(k−1)

as well as the assumption Nss = Nsn = 0 [27], (Nss , Nyy , Nsy , Nyn , Nsn ) =

N Z X k=1

n(k)

(σss , σyy , σsy , σyn , σsn )(k) d n,

n(k−1)

6

(6a)

yk2 yk1 y x n s

z

N

VT

VL

k

VR x

n

n n(k-1) (k) s

sk1

VB

sk2

2 1

Figure 2: Piezo-actuator location.

(Lyy , Lsy ) =

N Z X k=1

n(k)

(σyy , σsy )(k) n d n,

(6b)

n(k−1)

the stress resultants and stress couples reduce to the following expressions        0       ˜      N K K K K  N yy  12 13 14 yy     yy     11                    0  N   K    ˜ γys Nsy  ys 21 K22 K23 K24    = − , (7)   L     ˜ yy    K41 K42 K43 K44       φ0  L yy                        ˜   L    K51 K52 K53 K54  1yy   L sy sy and  Nyn =

A44 −

A45 2 A55

 γyn .

(8)

0 In these equations, (1) Kij , (2) 0yy , (3) 1yy , (4) γys and (5) γyn denote (1)

the modified local stiffness coefficients, (2) the axial strain associated with the primary warping, (3) a measure of curvature associated with the secondary warping, (4) the tangential shear strain and (5) the transverse shear strain, respectively. Their expressions can be found in Ref. [27]. The piezo-actuator

7

˜yy , N ˜sy and stress couples L ˜ yy , L ˜ sy are defined as induced stress resultants N     ˜yy = PNp eyy − A12 ess E1(k) [n(k) − n(k−1) ]R(k) (s)R(k) (y)  N  k=1  A11       P A  N 16 p ˜sy =  ess E1(k) [n(k) − n(k−1) ]R(k) (s)R(k) (y) N k=1 esy − A 11     ˜ yy = PNp 1 eyy [n(k) + n(k−1) ] − B12 ess E1(k) [n(k) − n(k−1) ]R(k) (s)R(k) (y)  L  k=1  A11   2   P 1 B  N 16 p  ˜ sy = L esy [n(k) + n(k−1) ] − ess E1(k) [n(k) − n(k−1) ]R(k) (s)R(k) (y) k=1 2 A11 (9)

3. Formulation of the governing system Hamilton’s principle is used to derive the governing equations and the corresponding boundary conditions. The path of motion renders the following variational form stationary: Z t2

 δT + δV − δWe d t = 0,

(10a)

t1

with ( at t = t1 , t2 ) δu0 = δv0 = δw0 = δθx = δθz = δφ = 0, 65

(10b)

where T , V and δWe denote the kinetic energy, strain energy and the virtual work due to the external forces, respectively. 3.1. Governing equations After lengthy variation manipulations, the governing equations are ˜ z )φ0 sin φ + (Mx + M ˜ x )φ0 cos φ δu0 : [(Ty + T˜y )u00 − (Mz + M

(11a)

˜ x ) cos φ + (Qz + Q ˜ z ) sin φ]0 + px − b1 u + (Qx + Q ¨0 = 0, δv0 : (Ty + T˜y )0 + py − b1 v¨0 = 0,

˜ z )φ0 cos φ − (Mx + M ˜ x )φ0 sin φ δw0 : [(Ty + T˜y )w00 − (Mz + M ˜ x ) sin φ + (Qz + Q ˜ z ) cos φ]0 + pz − b1 w − (Qx + Q ¨0 = 0, 8

(11b)

(11c)

˜ y )0 − (Bw + B ˜w )00 + [(Mx + M ˜ x )(u00 cos φ − w00 sin φ) δφ : (My + M ˜ z )(w00 cos φ + u00 sin φ) + (Γt + Γ ˜ t )φ0 ]0 − (Mz + M ˜ x )(u00 φ0 sin φ + w00 φ0 cos φ) + (Mx + M ˜ z )(w00 φ0 sin φ − u00 φ0 cos φ) − (Mz + M

(11d)

˜ x )(u0 sin φ + w0 cos φ) + (Qx + Q 0 0 ˜ z )(u00 cos φ − w00 sin φ) − (Qz + Q + my + b0w − (b4 + b5 )φ¨ + (b10 + b18 )φ¨00 = 0, ˜ x )0 − (Qz + Q ˜ z ) + mx − b4 θ¨x = 0, δθx : (Mx + M

(11e)

˜ z )0 − (Qx + Q ˜ x ) + mz − b5 θ¨z = 0, δθz : (Mz + M

(11f)

where px , py and pz are the external forces per unit span length and mx , my and mz are the moments about x−, y− and z − axes, respectively; bw stands for the bimoment of the surface traction. The essential and natural boundary conditions are: δu0 : u0 = u ¯0

or (12a)

¯x, Ty u00 − Mz φ0 sin φ + Mx φ0 cos φ + Qx cos φ + Qz sin φ = Q or Ty = T¯y ,

δv0 : v0 = v¯0

δw0 : w0 = w ¯0

(12b)

or (12c)

¯z, Ty w00 − Mz φ0 cos φ − Mx φ0 sin φ − Qx sin φ + Qz cos φ = Q

δφ : φ = φ¯ or 0 − Bw + My + Mx (u00 cos φ − w00 sin φ)

¯ y − (b10 + b18 )φ¨0 , − Mz (w00 cos φ + u00 sin φ) + Γt φ0 + Γt φ0 = M 9

(12d)

δφ0 : φ0 = φ¯0

¯w , or Bw = B

(12e)

δθx : θx = θ¯x

¯ x, or Mx = M

(12f)

δθz : θz = θ¯z

¯z. or Mz = M

(12g)

In Eqs. (11) and (12), the terms without and with over-tilde (˜·) identify the pure mechanical and piezo-actuator contributions, respectively. The expressions 70

for stress and stress couple resultants (e.g. Ty , T˜y , Mx , etc.) will be discussed in the following section. Explicit expressions for inertial terms bij can be found in the appendix of Ref. [27].

4. Governing equation system for circumferential asymmetric stiffness lay-up configuration 75

4.1. Force-displacement relationship The linearized beam constitutive law reads Fi = aij Dj

(13)

where Fi , aij and Dj denotes the stress and couple stress resultants (i.e. the beam internal actions), the stiffness coefficients and the beam deformation measure work-conjugated to the internal actions, respectively; see e.g. Ref. [27] for an explicit expression of aij . The stiffness matrix [aij ] is fully populated in the general case, so that bending, twisting, extension, transverse shearing and warping are all coupled together. For a circumferential asymmetric stiffness (CAS) lay-up configuration the stiffness matrix [aij ] can be decoupled into two types

10

of elastic coupling, viz, extension-transverse shear       Ty        Q  

 a  11  a14 x =     a15 Qz           Γ  a t

18

a14

a15

a44

0

0

a55

a48

0

and bending-twist coupling,       M  a 0 0   z   22      M   0 a 0 x 33 =      0 Bw  0 a66          M  0 a37 0 y

 1 0 2 1 0 2 0 v + (u ) + (w0 )    a18  0    2 0 2      0 0   a48 θz + u0 cos φ − w0 sin φ  ,  θx + u00 sin φ + w00 cos φ  0          1 0 2  a88  (φ ) 2

(14a)

  0 0 0 0 0     θ − w φ cos φ − u φ sin φ  z 0 0        0 0 0 0 0  a37  θx + u0 φ cos φ − w0 φ sin φ  ;     0  φ00          a77  φ0

(14b)

0

the stiffness coefficient a15 is equal to zero for balanced lay-ups on the left and right beam spars. 4.2. Force-voltage relationship We introduce two actuator pairs, viz., flange-actuator-pair (top and bottom plates) and web-actuator-pair (left and right spars). Thus, four voltage parameters VT , VB , VL and VR can be defined, i.e. the voltages applied on the actuators located at the top, bottom, left and right plates of the beam, see Fig. 2. The relation between the piezo-actuator induced stress and stress couple resultants ˜ and the voltages (V) can be expressed as (F) ˜ = [AF F i ]R(y)V, 80

(15)

˜ = where R(y) of Eq. (4c) denotes the location along span of the actuator, F ˜y B ˜w Γ ˜t M ˜z M ˜x Q ˜x Q ˜ z }T and V = {V1 V2 V3 V4 }T = 1 {(VT − {T˜y M 2 VB ) (VT + VB ) (VL − VR ) (VL + VR )}T . The piezo-actuator coefficients AF i (i = 1, 2, 3, 4) are defined in Appendix A. If the piezo-composite actuators are distributed in CAS configuration, Eq. (15)

11

can be split in two, viz., extension-transverse coupling       T˜y (y, t)   AT2 y AT4 y            Q    ˜ x (y, t) AQx  0 V (t) 2 2  = R(y),    ˜ z (y, t)    0  V4 (t) Q AQz   4        Γ  ˜ (y, t)  AΓt AΓt t

and bending-twist coupling     ˜   M (y, t) z        M  ˜ x (y, t)



2

4

0

z AM 3

  Mx A1 =   ˜w (y, t)     0 B          y M  ˜ y (y, t) AM 1

(16a)



   0  V1 (t)  R(y).  0  V3 (t) 

(16b)

y AM 3

4.3. Linearized governing equations 85

Approximating the trigonometric functions of Eqs. (11)-(12) with their Taylor series expansion, i.e. sin φ ≈ φ and cos φ ≈ 1, the CAS lay-up configuration equations can be split into two subsystem, the Lateral Bending-Extension coupling subsystem (u0 −v0 −θz ) and the Twist-Vertical Bending coupling subsystem (w0 − φ − θx ). BE-subsystem (Lateral Bending-Extension coupling subsystem). 0 δu0 : a14 v000 + a44 (u000 + θz0 ) − b1 u ¨0 + px + δp AQx 2 V2 R (y) = 0,

(17a)

δv0 : a11 v000 + a14 (u000 + θz0 ) − b1 v¨0 + py + δp AT2 y V2 R0 (y) + δp AT4 y V4 R0 (y) = 0, (17b) z 0 δθz : a22 θz00 − a14 v00 + a44 (u00 + θz ) − (b5 + b15 )θ¨z + mz + δp AM 3 V3 R (y)

− (δp + δs )AQx 2 V2 R(y) = 0. (17c) The boundary conditions for cantilevered beams are at y = 0: u0 = v0 = θz = 0, 12

(18)

and at y = L: ¯ δu0 : a14 v00 + a44 (u00 + θz ) + δs AQx 2 V2 = Qx ,

(19a)

δv0 : a11 v00 + a14 (u00 + θz ) + δs AT2 y V2 + δs AT4 y V4 = T¯y ,

(19b)

z ¯ δθz : a22 θz0 + δs AM 3 V3 = M z .

(19c)

TB-subsystem (Twist-Vertical Bending coupling subsystem). 0 δw0 : a55 (w000 + θx0 ) − b1 w ¨0 + pz + δp AQz 4 V4 R (y) = 0,

δφ : a37 θx00 + a77 φ00 − a66 φ(iv) − (b4 + b5 )φ¨ + (b10 + b18 )φ¨00 + my + b0w y My 0 0 + δp AM 1 V1 R (y) + δp A3 V3 R (y) = 0,

(20a)

(20b)

x 0 δθx : a33 θx00 + a37 φ00 − a55 (w00 + θx ) − (b4 + b14 )θ¨x + mx + δp AM 1 V1 R (y)

− (δp + δs )AQz 4 V4 R(y) = 0, (20c) The boundary conditions, for cantilevered beams are at y = 0: w0 = φ = φ0 = θx = 0,

(21)

¯ δw0 : a55 (w00 + θx ) + δs AQz 4 V4 = 0, = Qz ,

(22a)

and at y = L:

My My ¨0 ¯ δφ : a37 θx0 + a77 φ0 − a000 66 φ + (b10 + b18 )φ + δs A1 V1 + δs A3 V3 = My ,

(22b) 13

90

¯w , δφ0 : a66 φ00 = B

(22c)

x ¯ δθx : a33 θx0 + a37 φ0 + δs AM 1 V1 = Mx ,

(22d)

If the actuator is spread over the entire beam span, the traces are δp = 0 and δs = 1; otherwise, if the actuator is a single patch, the traces assume the values δp = 1 and δs = 0. Note that the BE-subsystem and the TB-subsystem are not entirely independent, since they are coupled by the voltage parameters V3 and V4 .

95

5. Solution methodology The Extend Galerkin’s Method (EGM) [28, 29] is used to discretize the system. EGM adopts weighting (or shape) functions that need to fulfill only the essential boundary conditions. Natural boundary conditions do not appear explicitly in the functional, and are satisfied in a weak sense. Symmetric mass and stiffness matrices will be obtained for thin-walled beams with CAS lay-ups. From now on, bold fonts will denote vectors or matrices, with regular fonts denoting scalar variables. Thus, u0 (y, t) = ΨTu (y)qu (t), φ( y, t) = ΨTφ (y)qφ (t),

v0 (y, t) = ΨTv (y)qv (t), θx (y, t) = ΨTx (y)qx (t),

w0 (y, t) = ΨTw (y)qw (t), θz(y, t) = ΨTz (y)qz (t), (23)

where the shape functions ΨTu (y), ΨTv (y), · · · ΨTz (y) are required to fulfill only the essential boundary conditions. Equations (17)-(19) and (20)-(22) lead to the following discretized equations of motion, [M](B,T ) {¨ q}(B,T ) + [K](B,T ) {q}(B,T ) + [A](B,T ) {V}(B,T ) = [Q](B,T ) ,

(24)

where the subscript B, T denote the BE-subsystem and TB-subsystem, respectively, and qB = {qTu qTv qTz }T , qT = {qTw qTφ qTx }T , VB = {V2 V3 V4 }T , 14

100

VT = {V1 V3 V4 }T . The expressions for mass matrix M(B,T ) , stiffness matrix K(B,T ) , actuation matrix A(B,T ) and external excitation vector Q(B,T ) are given in Appendix B.

6. Model validations The beam simplified model is firstly validated by comparing the vibration 105

frequency predicted for a composite thin-walled beam with the analytical results of Ref. [30] and the experimental data of Ref. [31]. The geometry and material properties of the cantilever thin-walled box beam of Fig. 1 are specified in Table 1. The results, reported in Table 2, show a better agreement of the present model with experimental data than that of Ref. [30]. Table 1: Details of thin-walled composite box beam for validation [31] 2

E11

1.42 × 1011 N/m

E22 = E33

9.8 × 109 N/m

G12 = G13

6.0 × 109 N/m

G23

4.83 × 109 N/m

µ12 = µ13

0.42

µ23

0.50 a

Density (ρ )

1.442 × 103 Kg/m

2

Width (2ba )

2.268 × 10−2 m

2

Depth (2da )

1.212 × 10−2 m

Number of layers (Nh )

6

Layer thickness

1.270 × 10−4 m

Length (L)

0.8446 m

2

3

Inner dimensions of the cross section.

Table 2: Natural frequency (Hz) for [45]6 CAS lay-up configuration

Mode

Exp. [31]

Analytical [30]

Error (%)

Present

Error

1TB

16.67

14.69

-11.9

15.20

-8.8

2TB

96.15

92.02

-4.3

95.09

-1.1

1BE

29.48

25.13

-14.8

26.64

-9.6

a

110

a

(%)

Relative error, (Present − Exp.)/Exp. × 100%.

The piezo-composite actuator model validation is based on the results obtained for a 1/16th scale blade with NACA 0012 airfoil single-cell cross-section [32].

15

NACA 0012 cross section E-glass [90o/90o]

0.02 Unit: [m]

AFC [±45o]

0.01 0 0.006 -0.01 0.02

0.0275

-0.02 -0.02

0.05

0

0.02

0.04

Figure 3: NACA0012 airfoil cross-section.

The cross-section geometry is shown in Fig. 3. The AFC fibers are aligned at +45o and −45o at the blade top and bottom, respectively. The E-glass and AFC material properties are reported in Table 3. Fig. 4 shows the induced tip 115

twist angle as a function of the applied voltage, with a good agreement with Ref. [32].

7. Static study The material property (Graphite-Epoxy) and geometric specification for the thin-walled box beam of Fig. 1 are shown in Table 4. In order to obtain better 120

actuating performance, the piezo-actuators are manufactured by single crystal MFC [33, 35], whose material property is specified in Table 3. Moreover, the piezo-composite laminate is distributed over the entire cross-section. The layup configurations (both for host structure and piezo-actuator) can be found in Table 5.

16

Table 3: Material properties of E-glass, AFC and single crystal MFC (S-MFC)

E-Glass [32]

AFC [32]

S-MFC [33]

E1 (Gpa)

14.8

30.54

6.23

E2 (Gpa)

13.6

16.11

11.08

G12 (Gpa)

1.9

5.5

2.01

µ12

0.19

0.36

0.229

d11 (×10−12 m/V)

N/A

381

1896.5

d12 (×10−12 m/V)

N/A

-160

-838.2

−3

1700

4810

5338.3

2.032

1.689

17

N/A

1.143

1.7

ρ (Kg m

)

Thickness (×10−4 m) Electrode spacing (×10

−3

m)

Table 4: Material property and geometric specification of the host structure [34, p. 131] 2

206.8 × 109 N/m

Width (2ba )

0.254 m

2

Depth (2da )

0.0681 m

2.55 × 10 N/m

2

Wall thickness (h)

0.0102 m

G23

3.10 × 109 N/m

2

Number of layers (Nh )

6

µ12 = µ13 = µ23

0.25

Layer thickness

0.0017 m

Length (L)

2.032 m

E11

5.17 × 109 N/m

E22 = E33

9

G12 = G13

1.528 × 103 Kg/m

ρ a

125

3

The length is measured on the mid-line contour.

7.1. Piezo-actuator coefficients study The non-zero CAS configuration piezo-actuator coefficients of Eq.(16) are depicted in Figs. 5 and 6 as a function of the piezo-actuator ply-angle θp . Two distinct trends emerge from the results of Figs. 5 and 6. x Mz The first trend characterizes the bending coefficients (AM 1 , A3 ) and the

130

extension coefficients (AT2 y , AT4 y ). The coefficients increase from θp = 0o to θp = 90o , then decrease until θp = 180o . Their values equal zero when θp ≈ 40o and ≈ 140o . Note that, because of the reverse definition of θx (see Fig. 1), x z coefficients AM and AM present the exactly opposite trends. 1 3

17

1.5

Blade tip twist

(deg)

Present duPlessis & Hagood Experimental 1

0.5

0 0

500

1000

1500

2000

Applied voltage (V) Figure 4: Tip deflection

Qz The second trend characterizes the transverse shear coefficients (AQx 2 , A4 ) 135

y My and the twist coefficients (AM 1 , A3 ). The previous groups of coefficients have

a symmetric dependence centered around θp = 90o . These coefficients, instead, show an anti-symmetric trend around θp = 90o . They are equal to zero when θp = 0o , 90o , 180o , with the maximum absolute values reached for θp ≈ 42o and θp ≈ 138o .

Table 5: Lay-up configurations for beam with CAS lay-up [unit:deg].

Flanges Material

Layer

Piezo-actuator

CAS (7)

Host structure

CAS (1-6)

a

a

Webs

Top

Bottom

Left

Right

[θp ]

[θp ]

[θp ]

[θp ]

[θh ]6

[θh ]6

[θh / − θh ]3

[θh / − θh ]3

The piezo-actuator is positioned of the outer side of the laminate.

18

Moment-coefficients (N m-1 V-1)

0.4

A My 1

0.3

A Mx 1

0.2 0.1 0 -0.1

A My 3

-0.2

A Mz 3

-0.3 -0.4 0

30

60

90

Ply angle

120 p

150

180

(deg)

Figure 5: Actuating moment coefficients with single crystal MFC as a function of the ply-angle θp in CAS lay-up configuration.

Force-coefficients (N m-2 V-1)

5

A Ty 2 A Ty 4

2.5

0

A Qz 4

-2.5

A Qx 2 -5 0

30

60

90

Ply angle

120 p

150

180

(deg)

Figure 6: Actuating force coefficients with single crystal MFC as a function of the ply-angle θp in CAS lay-up configuration.

19

140

7.2. Actuation performance study In this subsection, the relationship between actuation performance and voltage parameters (V1 , V2 , V3 , V4 ) is specifically discussed. 7.2.1. TB-subsystem The equations governing the TB-subsystem in Eqs. (20), have the voltage

145

vector VT = [V1 , V3 , V4 ]T as forcing parameters. Specifically, the voltage parameter V1 influences both twist and bending, while V3 and V4 work for twist and transverse shear, respectively. Note that the elastic coupling of the structure has a significant effect on the actuation performance. The twist-bending elastic coupling in Eq. (14b) is related to the stiffness coefficient a37 . Two typ-

150

ical host structure ply-angles, viz., θh = 15o and θh = 75o , are considered here. The corresponding values for a37 were computed in Ref. [27], i.e., a37 equals −4.05 × 103 N.m2 and 3.92 × 105 N.m2 , respectively. Thus, the elastic coupling can be ignored for θh = 15o , while it is significant for θh = 75o . Figures 7 and 8 depict the tip deflections obtained from V1 = 1000 V (black curve), V3 = 1000 V (red curve) and V4 = 1000 V (green curve) for weak and strong elastic coupling cases, respectively. The non-dimensional quantities are defined as u ˆ0 =

155

u0 , 2b

vˆ0 =

v0 , L

w ˆ0 =

w0 , 2b

φˆ = φ,

θˆx = θx ,

θˆz = θz .

(25)

It can be seen that the voltage parameter V4 , which is related to transverseshear-actuation can be ignored in both cases. In addition, the voltage parameter V1 dominates the TB-subsystem actuation performance. Note that, for strong elastic coupling case in Fig. 8, twist-actuation shows better performance on bending deflection than direct bending-actuation.

160

7.2.2. BE-subsystem According to the governing equations (17) of the BE-subsystem, voltage V2 is related to both extension and transverse shear actuation, while V3 and V4 are

20

V 1=1000V

0.025

Tip deflection

0.02 0.015 0.01 V 3=1000V

0.005 0 -0.005 -0.01

V 4=1000V

-0.015 0

30

60

90

Ply-angle

120 p

150

180

(deg)

Figure 7: Beam tip deflections of the TB-subsystem ( θh = 15o ) as a function of the piezoactuator ply-angle θp .

10-3

8 6

V 1=1000V

Tip deflection

4 V 3=1000V 2 0 -2 V 4=1000V

-4 -6 0

30

60

90

Ply-angle

120 p

150

180

(deg)

Figure 8: Beam tip deflections of the TB-subsystem ( θh = 75o ) as a function of the piezoactuator ply-angle θp .

21

related to bending and extension actuation, respectively. The results obtained for two typical host structure cases, θh = 15o and θh = 75o , are reported in Figs. 9 and 10. 10-3

4 3

Tip deflection

V 3=-1000V

V 2=1000V

2 V 4=1000V

1 0 -1 -2 0

30

60

90

Ply-angle

120 p

150

180

(deg)

Figure 9: Beam tip deflections of BE-subsystem ( θh = 15o ) as a function of the piezo-actuator ply-angle θp . 165

The beam extension stiffness coefficient a11 is much higher than the other terms. Thus, the extension actuation induced by voltage parameter V4 can be ignored, see both Figs. 9 and 10. The voltage V2 has a significant effect on actuation performance for both cases: the transverse-shear-actuation, weak in 170

the TB-subsystem, is much stronger in the BE-subsystem. One reason for this is that the size of the flange-actuator-pair is almost four times than that of the web-actuator-pair. In addition, the direct bending actuation induced by voltage V3 (red line) in the θh = 75o case in Fig. 10 is much weaker than that in the θh = 15o case in Fig. 9.

175

7.2.3. Conclusion In a nutshell, the TB-subsystem is dominated by V1 for the weak elastic coupling case (see Fig. 7), while the BE-subsystem is dominated by V2 when 22

10-3 1 V 3=-1000V

Tip deflection

0.5

0 V 4=1000V

-0.5

V 2=1000V -1 0

30

60

90

Ply-angle

120 p

150

180

(deg)

Figure 10: Beam tip deflections of BE-subsystem ( θh = 75o ) as a function of the piezoactuator ply-angle θp .

the elastic coupling effect is strong (see Fig. 10). As a result, although the TBsubsystem and the BE-system are actually coupled by V3 and V4 , it is nonethe180

less reasonable to treat these two subsystems as independent.

8. Dynamic control: negative velocity feedback control 8.1. Governing equations including negative velocity feedback control A negative velocity feedback control is considered here. Since we are interested in the aeroelastic control of an aircraft wing we choose to focus this investigation into the torsion/bending (TB) subsystem. If we assume the sensor can offer the velocity information at y = Ys , then actuating voltage vector VT for the negative velocity feedback control can be computed       ˆ˙ s , t) + β k θˆ˙x (Ys , t)]      V1  −k1 [α1k φ(Y     1     ˙ ˆ s , t) VT = V 3 = = P(Ys )q˙ T (t), −k3 φ(Y             ˙ kˆ V  −k [αk w  ˙ 4 4 4 ˆ0 (Ys , t) + β4 θx (Ys , t)]

23

(26)

where, ki (i = 1, 3, 4) are the feedback control gains, αik and βik are weighting coefficients of the control gains, and matrix P is defined in Appendix B. As a result, the closed-loop discretized system Eq. (24) becomes ¨ T (t) + AT Pq˙ T (t) + KT qT (t) = QT (t). MT q

(27)

8.2. Control gain weighting coefficients discussion The first step in designing the controller is to choose suitable control weighting coefficients. Since the flapping and twisting motions usually have a significant phase difference, we can simplify the control system by considering only two cases for the voltage parameter V1 , i.e., the bending control (α1k = 0, β1k = −1) and the twist control (α1k = 1, β1k = 0). Furthermore, as evidenced by the static study, the shear force induced by V4 is immaterial in the TB-subsystem. As a result, the velocity feedback Eq. (26) can be simplified to a combination of these two cases: the Bending Control Methodology,     V   k θ˙  1 1 x VT = = V  −k φ˙ 

(28)

3

3

and the Twist Control Methodology     V  −k φ˙  1 1 VT = = V  −k φ˙  3

(29)

3

The static study in Section 7.2.1 evidenced a significant effect of the twist185

bending elastic coupling on the actuation performance. Thus, the weak and strong elastic coupling cases are separately investigated here. Bending control is considered first because twist control has a weak bending authority for the weak elastic coupling case. Fig. 11 depicts the damping ratios obtained for the first four modes of the weak elastic coupling case θh = 15o . The figure allows to

190

easily identify the positive damping area 90o 6 θp 6 138o . The first four modes are denoted as Flap1, Flap2, Twist1 and Flap3, respectively. Since the twistbending coupling can be ignored in the first four modes, the damping ratios of y x the flapping and the twist modes follow the trend of coefficients AM and AM 1 3

variation in Fig. 5, respectively. 24

0.02 2TB (Flap2) 4TB (Flap3)

Damping ratio

0.01

0

1TB (Flap1)

-0.01

Positive damping area 3TB (Twist1)

-0.02 0

30

60

90

120

Ply-angle in piezo-actuator

p

150

180

(deg)

Figure 11: Damping ratios as a function of piezo-composite ply-angle θp ; bending control methodology, ki = 100, θh = 15o .

195

Considering the strong twist-bending elastic coupling case of θh = 75o , the results obtained with the bending control methodology and the twist control methodology are shown in Figs. 12 and 13, respectively. Note that, due to the strong elastic coupling, there will be no pure bending mode or twist mode for the TB-subsystem. The twist control methodology is more efficient, especially for

200

the third mode (3TB), whose modal shape is dominated by the twist component. This is because the negative twist damping will be induced by direct bending actuation via twist-bending elastic coupling, see Fig. 12. In a nutshell, the optimal control strategy for the θh = 15o case is the bending control methodology with θp ≈ 120o , while for the θh = 75o case is the

205

twist control methodology with θp ≈ 135o . 8.3. Optimized by tailoring It is clear that the twist-bending elastic coupling has a significant effect on control efficiency. Thus, in order to optimize the control performance by host structure tailoring, the effect of the ply-angle θh at the vibration control

25

0.04 4TB 3TB just induced by V1

Damping ratio

0.02

0 1TB

3TB

-0.02 2TB

-0.04 0

30

60

90

120

Ply-angle in piezo-actuator

p

150

180

(deg)

Figure 12: Damping ratios as a function of the piezo-composite ply-angle θp ; bending control methodology, ki = 100, θh = 75o .

0.08 2TB

0.06

Damping ratio

0.04

4TB 1TB

0.02 0 -0.02

3TB

-0.04

Positive damping area

-0.06 -0.08 0

30

60

90

120

Ply-angle in piezo-actuator

p

150

180

(deg)

Figure 13: Damping ratios as a function of piezo-composite ply-angle θp ; twist control methodology, ki = 100, θh = 75o .

26

350 4TB

300

3TB(Twist1)

Frequency (Hz)

250 200 150

2TB(Flap2)

100 1TB(Flap1) 50 0 0

15

30

45

60

Ply-angle in host structure

h

75

90

(deg)

Figure 14: First four modes frequencies as a function of host structure ply-angle θh ; θp = 120o and ki = 100.

210

authority is investigated. Figure 14 depicts the frequencies of the first four modes as a function of angle θh . A typical mode cross phenomenon between the third and fourth modes can be found in Fig. 14. In order to avoid misunderstanding in the following study, we denote 3TB and 4TB the 4th and 3rd modes in the mode crossed range

215

33o < θh < 60o , respectively. The frequency of mode 1TB increases from θh = 0o to θh = 90o ; modes 2TB and 4TB reach their maximum frequency for θh ≈ 80o . In order to compare the bending and twist control methodologies, θp = 120o in piezo-actuator is adopted here, since this angle allows to achieve compara-

220

ble bending moment and twist authorities. The damping ratios obtained with the bending and the twist control methodology are shown in Figs. 15 and 16, respectively.

A sudden change around the mode cross points θh ≈ 33o and

θh ≈ 58o can be found in Figs. 15 and 16. Especially near the point θh ≈ 58o in Fig. 15, a jump phenomenon is observed. This is because near this point, not 225

only the eigen-frequencies of the 3TB and 4TB modes are almost the same, but 27

Mode cross region (3TB and 4TB modes)

0.04

2TB

3TB

4TB

Damping ratio

0.03

0.02

0.01

1TB

0 0

15

30

45

60

Ply-angle in host structure

75

h

90

(deg)

Figure 15: Damping ratio as a function of the host structure ply-angle θh , bending control methodology; θp = 120o and ki = 100.

0.2

2TB Mode cross region (3TB and 4TB modes)

Damping ratio

0.15

4TB 3TB

0.1

0.05

0 1TB 0

15

30

45

60

Ply-angle in host structure

h

75

90

(deg)

Figure 16: Damping ratio as a function of host structure ply-angle θh ; twist control methodology, θp = 120o and ki = 100.

28

their mode shapes are similar as well. Outside of the mode cross regions, the damping ratios for both control methodologies increase for the 1TB, 2TB and 4TB modes, and decreases for the 3TB mode. Furthermore, for bending control methodology of Fig. 15, mode 3TB has a negative damping above θh ≈ 80o , 230

and the damping ratios of modes 1TB, 2TB and 3TB significantly decrease near θh = 90o . The absolute value of the first mode (1TB) eigenvalue real part is plotted as a function of θh in Fig. 17. The corresponding curves for the 2TB, 3TB and 4TB modes are reported in Fig. 18. The plots of Figs. 17 and 18 can be splitted

235

into two different regions, i.e. Area 1 for 0 < θh < 45o and Area 2 for 690 < θh < 87o ; these regions are characterized by weak and strong twist-bending elastic coupling, respectively. In Area 1, the twist control methodology (red lines) has almost no flapping damping; thus, the bending control methodology (black lines) should be chosen. Within Area 2, instead, the flapping damping

240

induced by the elastic coupling allows to achieve a damping that is higher than that of the direct flapping control; thus, the twist control methodology would be a better choice here. 8.4. Vibration control under an impulsive load An impulsive load with coefficients pz = my = 1, and mx = 0 in Eqs. (20) is

245

applied to the θh = 75o structure. Figs. 19 and 20 depict the time responses of ˆ t), respectively. the tip flapping displacement w ˆ0 (L, t) and twist rotation φ(L, Piezo-actuator θp = 90o with bending control methodology, that was chosen in Refs. [14, 15, 24], is compared with the piezo-actuator θp = 135o with twist control methodology. From the results of Figs. 19 and 20, it can be seen that the

250

twist control shows a significant advantage both for the flapping and twisting motions. 8.5. The influence of position of piezo-actuator Investigating the effect of size and position of piezo-actuators allows to strike a balance between their cost and efficiency. As reported by Librescu

29

Reverse real part of eigenvalue

3 Twisting Control 2.5 Area 1 2 1TB 1.5 1

Bending Control Area 2

0.5 0 0

20

40

60

Ply-angle in host structure

80 h

(deg)

Figure 17: Absolute value of the real part of 1TB eigenvalue as a function of the host structure ply-angle θh ; θp = 120o , ki = 100.

Reverse real part of eigenvalue

160 Area 2

Twisting Control

140 120 100

Area 1

80

2TB 3TB 4TB

60 Bending Control

40 20 0 -20 0

20

40

60

Ply-angle in host structure

80 h

(deg)

Figure 18: Absolute value of the real part of 2TB, 3TB and 4TB eigenvalues as a function of the host structure ply-angle θh ; θp = 120o , ki = 100.

30

0.01

0.005

0

-0.005

-0.01 0

0.2

0.4

0.6

0.8

1

Figure 19: Tip flapping displacement w ˆ0 in TB-subsystem (θh = 75o ) subject to an impulse load; ki = 100.

0.01

0.005

0

-0.005

-0.01 0

0.1

0.2

0.3

0.4

0.5

Figure 20: Tip twist rotation φˆ in TB-subsystem (θh = 75o ) subject to an impulse load; ki = 100.

31

Host structure

Piezo-actuator

Po s

itio n Si ze

Figure 21: Position and size of the piezo-actuator.

255

et al [20, 21], twist-bending elastic coupling is beneficial for the aeroelastic response behavior. Thus the piezo-actuator θp = 135o with twist control methodology and strong elastic coupling host structure θh = 75o is considered here. The size of the piezo-actuator is first fixed at 10% of the beam span, and the sensors are applied at the outer end of the piezo-actuator, see Fig. 21. The

260

resulting first four modes damping ratios are depicted in Fig. 22 as a function of the piezo-actuator position. When the piezo is positioned between 48% and 67% of the wing span, the 4TB mode damping is negative. A good compromise is achieved at 40% of the span. However, for a large size piezo-actuator, e.g., 70% length of span see Fig. 23, the damping ratios of the first four modes are

265

all positive. In a nutshell, the optimized position for piezo-actuator is around central point of the span.

9. Summary and conclusions A geometrical nonlinear fiber-reinforced composite thin-walled beam theory incorporating piezo-composite actuators is developed and used to model a smart 270

aircraft wing. A circumferential asymmetric stiffness (CAS) lay-up configuration is adopted in order to split the linear system into two subsystems, viz., 32

15

10-3

3TB 2TB

Damping ratio

10

4TB 5

1TB

0 Negative damping -5 10

20

30

40

50

60

70

80

90

Position of piezo-actuator (size: 0.1L) Figure 22: Damping ratios as a function of the piezo-actuator position; θh = 75o , θp = 135o , ki = 100.

0.09

Damping ratio

0.08

2TB

0.06

4TB

3TB

0.04

1TB 0.02

0 35

40

45

50

55

60

65

Position of piezo-actuator (size: 0.7L) Figure 23: Damping ratios as a function of the piezo-actuator position; θh = 75o , θp = 135o , ki = 100.

33

the lateral bending-extension coupled subsystem and the twist-vertical bending coupled subsystem. A simple negative velocity feedback control algorithm is proposed, and used to asses how the piezo-actuator and the fiber-reinforced 275

host structure elastic tailoring influence the vibration suppression control. The main conclusions are that 1. the twist bending elastic coupling of the host structure is beneficial for active damping performance; 2. the transverse shear force actuation is immaterial in TB-subsystem, and

280

the inducing voltage V4 can be ignored in dynamical control; 3. for the strong twist elastic coupling case, the twist control presents a better flapping control efficiency than the direct bending control; 4. optimizing the size and position of the piezo-actuator can improve the damping control authority and reduce the overall cost of the structure.

34

285

Appendix A. The piezo-actuator coefficients AF i The subscript i = 1, 2, 3, 4 of piezo-actuator coefficients AF i denote the operation Z

AF 1

AF T

=

Z ds −

T

AF 3 =

Z

AF B

d s,

AF 2

Z =

B

AF L ds −

Z

L

AF T

Z ds +

T

AF R d s,

AF 4 =

R

Z L

AF B d s,

(A.1a)

AF R d s,

(A.1b)

B

AF L ds +

Z R

where T, B, L, R denotes top, bottom, left and right wall, respectively. The definitions of AF are given as Ty

A

=

Np  X k=1

A12 eyy − ess A11



[n(k) − n(k−1) ] R(k) (s), ˆ h

(A.2a)

 Np n  X [n(k) − n(k−1) ] A12 A = x eyy − ess R(k) (s) ˆ A11 h k=1   o [n(k) − n(k−1) ] dz 1 B12 − eyy [n(k) + n(k−1) ] − ess R(k) (s) , ˆ ds 2 A11 h

(A.2b)

 Np n  X [n(k) − n(k−1) ] A12 A = z eyy − ess R(k) (s) ˆ A11 h k=1   o [n(k) − n(k−1) ] dx 1 B12 + eyy [n(k) + n(k−1) ] − ess R(k) (s) , ˆ ds 2 A11 h

(A.2c)

Mz

Mx

Qx

A

 Np X dx

[n(k) − n(k−1) ] R(k) (s), ˆ h

(A.2d)

  Np X [n(k) − n(k−1) ] dz A16 = esy − ess R(k) (s), ˆ ds A11 h

(A.2e)

=

k=1

Qz

A

ds

A16 esy − ess A11



k=1

  Np n X [n(k) − n(k−1) ] A12 A =− Fw eyy − ess R(k) (s) ˆ A11 h k=1   o [n(k) − n(k−1) ] 1 B12 + a(s) eyy [n(k) + n(k−1) ] − ess R(k) (s) , ˆ 2 A11 h Bw

35

(A.2f)

  Np n X [n(k) − n(k−1) ] A16 A = ψ(s) esy − ess R(k) (s) ˆ A11 h k=1   o [n(k) − n(k−1) ] 1 B16 + 2 esy [n(k) + n(k−1) ] − ess R(k) (s) , ˆ 2 A11 h

(A.2g)

  Np n X [n(k) − n(k−1) ] A12 (x2 + z 2 ) eyy − ess R(k) (s) ˆ A11 h k=1   o [n(k) − n(k−1) ] 1 B12 + 2rn eyy [n(k) + n(k−1) ] − ess R(k) (s) . ˆ 2 A11 h

(A.2h)

My

AΓt =

Appendix B. Matrix via the Extended Galerkin’s Method Mass matrix.  T Z L b1 Ψu Ψu  MB =  0  Symm

 b1 Ψw ΨTw RL  M= 0   Symm

0

0

b1 Ψv ΨTv

0



(b5 + b15 )Ψz ΨTz

0

   d y, 

0 T

(b4 + b5 )Ψφ ΨTφ + (b10 + b18 )Ψ0 φ Ψ0 φ

0 (b4 + b14 )Ψx ΨTx

(B.1)

    d y. 

(B.2) Stiffness matrix.  0 0 T Z L a44 Ψu Ψu  KB =  0  Symm

 T a55 Ψ0w Ψ0w RL  KT = 0   Symm

a14 Ψ0u Ψ0v

T

a44 Ψ0u Ψz T



a11 Ψ0v Ψ0v

T

a14 Ψ0v Ψz T

   d y, 

(B.3)

T



T

   d y. 

T

a22 Ψ0z Ψ0z + a44 Ψz Ψz T

a55 Ψ0w Ψ0x

0 T

a77 Ψ0φ Ψ0φ + a66 Ψ00φ Ψ00φ

T

a37 Ψ0φ Ψ0x T

a33 Ψ0x Ψ0x + a55 Ψx Ψx T

(B.4)

36

Piezo-actuator Matrix.   Qx 0 A Ψ R (y) 0 0 u Z L 2   Ty  Ty 0 AB =  A2 Ψv R0 (y) 0 A4 Ψv R (y) d y,   0 Qz 0 −AQx Ψ R(y) A Ψ R (y) 0 z z 2 3  Z AT = 0

L

(B.5)



0 AQz 4 Ψw R (y) 

0 0   My  My 0 A1 Ψφ R (y) A3 Ψφ R0 (y)  d y, 0   Qz x 0 AM Ψ R (y) 0 −A Ψ R(y) x x 1 4

(B.6)

External forces vector. R  L  ¯ x Ψu (L)    p Ψ d y + Q x u    R0  L ¯ QB = , p Ψ d y + T Ψ (L) y v y v 0     R   L  m Ψ dy + M ¯ z Ψz (L) z z 0

(B.7)

  RL   ¯   p Ψ d y + Q Ψ (L) z w z w   0 R  L 0 0 ¯ ¯ QT = (my + bw )Ψφ d y + [My Ψφ (L) + Bw Ψφ (L)] .    0  RL     ¯ x Ψx (L) m Ψ d y + M x x 0

(B.8)

Control matrix P in Eq. (26).   T T k k 0 −k1 α1 Ψφ (Ys ) −k1 β1 Ψx (Ys )      0 −k3 Ψφ T (Ys ) 0 P=   T   Ψ (Y ) w s T −k4 α4k 0 −k4 β4k Ψx (Ys ) 2b

37

(B.9)

References [1] W. Yu, D. Hodges, J. Ho, Variational asymptotic beam sectional analysisan updated version, Int. J. Eng. Sci. 59 (2012) 40–64. 290

[2] S. Ibrahim, E. Carrera, M. Petrolo, E. Zappino, Buckling of composite thin walled beams by refined theory, Compos. Struct. 94 (2012) 563–570. [3] V. Cortinez, M. Piovan, Vibration and buckling of composite thin-walled beams with shear deformability, J. Sound Vib. 258 (2002) 701–723. [4] T. P. Vo, J. Lee, Flexural-torsional behavior of thin-walled composite box

295

beams using shear-deformable beam theory, Eng. Struct. 30 (2008) 1958– 1968. [5] A. Varello, E. Carrera, Free vibration response of thin and thick nonhomogeneous shells by refined one-dimensional analysis, J. Vib. Acoust. 136 (6) (2014) 061001.

300

[6] E. Carrera, M. Filippi, P. K. Mahato, A. Pagani, Advanced models for free vibration analysis of laminated beams with compact and thin-walled open/closed sections, J. Compos. Mater. (2014) 0021998314541570. [7] V. Giavotto, M. Borri, P. Mantegazza, G. Ghiringhelli, V. Carmaschi, G. Maffioli, F. Mussi, Anisotropic beam theory and ap-

305

plications,

Computers

&

Structures

16

(1-4)

(1983)

403–413.

doi:10.1016/0045-7949(83)90179-7. URL

http://www.sciencedirect.com/science/article/

B6V28-47S0Y51-1F/2/3dcd227ef680245164977e0448c9f3a6 [8] M. Morandini, M. Chierichetti, P. Mantegazza, Characteristic behavior 310

of prismatic anisotropic beam via generalized eigenvectors, International Journal of Solids and Structures 47 (10) (2010) 1327–1337. doi:10.1016/j.ijsolstr.2010.01.017. URL

http://www.sciencedirect.com/science/article/pii/

S0020768310000284 38

315

[9] P. Masarati,

G. L. Ghiringhelli,

non-homogeneous

plates

with

Characterization of anisotropic,

piezoelectric

inclusions,

ers & Structures 83 (15-16) (2005) 1171–1190.

Comput-

doi:doi:DOI:

10.1016/j.compstruc.2004.10.017. URL 320

http://www.sciencedirect.com/science/article/

B6V28-4FKY8WG-1/2/c4da0f3cbc7e9cbd655e84b284c805c8 [10] C. Brillante, M. Morandini, P. Mantegazza, Characterization of beam stiffness matrix with embedded piezoelectric devices via generalized eigenvectors, International Journal of Solids and Structures 59 (2015) 37–45. doi:10.1016/j.ijsolstr.2015.01.003.

325

URL

//www.sciencedirect.com/science/article/pii/

S0020768315000050 [11] L. Librescu, S. S. Na, Dynamic response of cantilevered thin-walled beams to blast and sonic-boom loadings, Shock and Vibration 5 (1) (1998) 23–33. [12] S. Na, L. Librescu, Oscillation control of cantilevers via smart materials 330

technology and optimal feedback control: actuator location and power consumption issues, Smart Materials and Structures 7 (6) (1998) 833. [13] S. Na, L. Librescu, M.-H. Kim, I.-J. Jeong, P. Marzocca, Robust aeroelastic control of flapped wing systems using a sliding mode observer, Aerosp. Sci. Technol. 10 (2) (2006) 120–126.

335

[14] L. Librescu, S. Na, Z. Qin, B. Lee, Active aeroelastic control of aircraft composite wings impacted by explosive blasts, J. Sound Vib. 318 (1) (2008) 74–92. [15] S.-J. Cha, J.-S. Song, H.-H. Lee, S. Na, J.-H. Shim, P. Marzocca, Dynamic response control of rotating thin-walled composite blade exposed to exter-

340

nal excitations, J. Aerosp. Eng. 27 (5) (2014) 04014025. [16] A. A. Bent, Active fiber composites for structural actuation, Ph.D. thesis, Massachusetts Institute of Technology (1997). 39

[17] W. K. Wilkie, R. G. Bryant, J. W. High, R. L. Fox, R. F. Hellbaum, A. Jalink Jr, B. D. Little, P. H. Mirick, Low-cost piezocomposite actuator 345

for structural control applications, in: SPIE’s 7th Annual International Symposium on Smart Structures and Materials, International Society for Optics and Photonics, 2000, pp. 323–334. [18] G. L. Ghiringhelli, P. Masarati, M. Morandini, D. Muffo, Integrated Aeroservoelastic Analysis of Induced Strain Rotor Blades, Mechanics of

350

Advanced Materials and Structures 15 (3) (2008) 291.

doi:10.1080/

15376490801907822. URL http://www.informaworld.com/10.1080/15376490801907822 [19] C. Brillante, M. Morandini, P. Mantegazza, Periodic controllers for vibration reduction using actively twisted blades, The Aeronautical Journal 355

120 (1233) (2016) 1763–1784. doi:10.1017/aer.2016.80. URL

https://www.cambridge.org/core/

journals/aeronautical-journal/article/

div-classtitleperiodic-controllers-for-vibration-reduction-using-actively-twisted-blade 89330CE640F7C750DE4FF561D31DF3D4 360

[20] L. Librescu, O. Song, Behavior of thin-walled beams made of advanced composite materials and incorporating non-classical effects, Appl. Mech. Rev. 44 (11) (1991) S174–80. [21] L. Librescu, A. Khdeir, Aeroelastic divergence of swept-forward composite wings including warping restraint effect, AIAA journal 26 (11) (1988) 1373–

365

1377. [22] Z. Qin, L. Librescu, Dynamic aeroelastic response of aircraft wings modeled as anisotropic thin-walled beams, Journal of Aircraft 40 (3) (2003) 532–543. [23] Z. Qin, L. Librescu, P. Marzocca, Aeroelasticity of composite aerovehicle wings in supersonic flows, Journal of spacecraft and rockets 40 (2) (2003)

370

162–173.

40

[24] J.-S. Song, J. Choo, S.-J. Cha, S. Na, Z. Qin, Robust aeroelastic instability suppression of an advanced wing with model uncertainty in subsonic compressible flow field, Aerosp. Sci. Technol. 25 (1) (2013) 242–252. [25] J. Yang, An introduction to the theory of piezoelectricity, Vol. 9, Springer 375

Science & Business Media, 2004. [26] R. M. Jones, Mechanics of composite materials, CRC press, 1998. [27] X. Wang, Z. Qin, Nonlinear modal interactions in composite thin-walled beam structures with simultaneous 1: 2 internal and 1: 1 external resonances, Nonlinear Dynamics 86 (2) (2016) 1381–1405.

380

[28] L. Librescu, S. Na, Dyanmic response of cantilevered thin-wlled beams to blast and sonic-boom loadings, Shock. Vib. 5 (1) (1998) 23–33. [29] A. Palazotto, P. Linnemann, Vibration and buckling characteristics of composite cylindrical panels incorporating the effects of a higher order shear theory, Int. J. Solids Struct. 28 (3) (1991) 341–361.

385

[30] E. A. Armanios, A. M. Badir, Free vibration analysis of anisotropic thinwalled closed-section beams, AIAA journal 33 (10) (1995) 1905–1910. [31] R. Chandra, I. Chopra, Experimental-theoretical investigation of the vibration characteristics of rotating composite box beams, Journal of Aircraft 29 (4) (1992) 657–664.

390

[32] A. J. Du Plessis, Modeling and experimental testing of twist actuated single cell composite beams for helicopter blade control, Ph.D. thesis, Massachusetts Institute of Technology (1996). [33] J.-S. Park, J.-H. Kim, Analytical development of single crystal macro fiber composite actuators for active twist rotor blades, Smart Mater. Struct.

395

14 (4) (2005) 745. [34] L. Librescu, O. Song, Thin-Walled Composite Beams: Theory and Application, Springer, New York, 2006, chap.8, pp. 213-232. 41

[35] S.-C. Choi, J.-S. Park, J.-H. Kim, Vibration control of pre-twisted rotating composite thin-walled beams with piezoelectric fiber composites, J. Sound 400

Vib. 300 (1) (2007) 176–196.

42

List of Figures 1

405

Geometry of the beam with a rectangular cross-section (CAS layup). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5

2

Piezo-actuator location. . . . . . . . . . . . . . . . . . . . . . . .

7

3

NACA0012 airfoil cross-section. . . . . . . . . . . . . . . . . . . .

16

4

Tip deflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

18

5

Actuating moment coefficients with single crystal MFC as a function of the ply-angle θp in CAS lay-up configuration. . . . . . . .

6

Actuating force coefficients with single crystal MFC as a function of the ply-angle θp in CAS lay-up configuration. . . . . . . . . . .

410

7

Beam tip deflections of the TB-subsystem ( θh = 15 ) as a func-

9

Beam tip deflections of the TB-subsystem ( θh = 75 ) as a func-

Beam tip deflections of BE-subsystem ( θh = 15 ) as a function

Beam tip deflections of BE-subsystem ( θh = 75 ) as a function

12

14

. . . . . . . . . .

430

27

Damping ratio as a function of the host structure ply-angle θh , bending control methodology; θp = 120o and ki = 100. . . . . . .

16

26

First four modes frequencies as a function of host structure plyangle θh ; θp = 120o and ki = 100. . . . . . . . . . . . . . . . . . .

15

26

Damping ratios as a function of piezo-composite ply-angle θp ; twist control methodology, ki = 100, θh = 75o .

425

25

Damping ratios as a function of the piezo-composite ply-angle θp ; bending control methodology, ki = 100, θh = 75o . . . . . . . . . .

13

23

Damping ratios as a function of piezo-composite ply-angle θp ; bending control methodology, ki = 100, θh = 15o . . . . . . . . . .

420

22

o

of the piezo-actuator ply-angle θp . . . . . . . . . . . . . . . . . . 11

21

o

of the piezo-actuator ply-angle θp . . . . . . . . . . . . . . . . . . 10

21

o

tion of the piezo-actuator ply-angle θp . . . . . . . . . . . . . . . . 415

19

o

tion of the piezo-actuator ply-angle θp . . . . . . . . . . . . . . . . 8

19

28

Damping ratio as a function of host structure ply-angle θh ; twist control methodology, θp = 120o and ki = 100. . . . . . . . . . . .

43

28

17

Absolute value of the real part of 1TB eigenvalue as a function of the host structure ply-angle θh ; θp = 120o , ki = 100. . . . . . .

18

30

Absolute value of the real part of 2TB, 3TB and 4TB eigenvalues as a function of the host structure ply-angle θh ; θp = 120o , ki = 100. 30

435

19

Tip flapping displacement w ˆ0 in TB-subsystem (θh = 75o ) subject to an impulse load; ki = 100. . . . . . . . . . . . . . . . . . .

20

impulse load; ki = 100.

440

31

Tip twist rotation φˆ in TB-subsystem (θh = 75o ) subject to an . . . . . . . . . . . . . . . . . . . . . . .

31

21

Position and size of the piezo-actuator. . . . . . . . . . . . . . . .

32

22

Damping ratios as a function of the piezo-actuator position; θh = 75o , θp = 135o , ki = 100. . . . . . . . . . . . . . . . . . . . . . . .

23

33

Damping ratios as a function of the piezo-actuator position; θh = 75o , θp = 135o , ki = 100. . . . . . . . . . . . . . . . . . . . . . . .

44

33

List of Tables 445

1

Details of thin-walled composite box beam for validation [31] . .

15

2

Natural frequency (Hz) for [45]6 CAS lay-up configuration . . . .

15

3

Material properties of E-glass, AFC and single crystal MFC (SMFC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4 450

5

17

Material property and geometric specification of the host structure [34, p. 131] . . . . . . . . . . . . . . . . . . . . . . . . . . . .

17

Lay-up configurations for beam with CAS lay-up [unit:deg]. . . .

18

45