Você está na página 1de 10

Cross-sectional Analysis of

Anisotropic, Thin-Walled, Closed-Section Beams


with Embedded Strain Actuation
Mayuresh J. Patil∗ and Eric R. Johnson†
Virginia Polytechnic Institute and State University, Blacksburg, VA 24061-0203

The paper presents a theory for the analysis of active anisotropic beams. The focus
of the paper is on the calculation of 2-D, beam, cross-sectional properties that can then
be incorporated into any 1-D beam analysis. Closed-form expressions for the cross-section
flexibilities as well as actuation strains for an active, anisotropic, thin-walled, closed-section
beam are derived and presented in a form that can be easily applied. Test cases presented
in the paper indicate that the theory can be used to efficiently calculate accurate cross-
sectional properties for preliminary design and optimization.

I. Introduction
The advent of composite materials transformed the field of structural design, especially as applied to
aerospace design. Anisotropic composite materials are being used at an increasing rate in the aerospace
industry. Beam-like composite structures are ideal for helicopter blades as well as high-aspect ratio wings
used in uninhabited aerial vehicles. There has been considerable research effort in the aerospace structures
community to develop efficient and effective modeling techniques to help in the design and development of
composite wings.
The development of smart materials is further revolutionizing the industry. The combination of smart
materials and composites has the potential to drastically improve future aerospace vehicles, making them
safer as well as adaptable. The analysis of anisotropic structures with embedded actuation is essential to
reach the full potential of these new materials.
The present paper develops a theory for the analysis of anisotropic, thin-walled, closed-section beams
with embedded strain actuation. The focus of the paper is on the 2-D cross-sectional analysis of active
anisotropic beams. The analysis leads to closed-form analytical expressions for the cross-sectional flexibility
coefficients and the actuation strain measures. The theory is based on thin-walled assumption and thus shell
bending strains are neglected. The present analysis extends the work of Johnson et al.1 to include the effect
of active anisotropic strain actuation.
There are presently very few theoretical developments for the comprehensive cross-sectional analysis of
active beams. duPlessis and Hagood2 extended the thin-walled beams theory of Rehfield3 to present one
of the first cross-sectional analysis for active beams. Cesnik and Shin4 based their work on the asymptotic
analysis of thin-walled beams as developed by Berdichevsky et al.5 and thus leads to better estimates of the
cross-sectional stiffnesses and actuation forces. Both the theories give closed-form analytical expressions for
the stiffnesses and actuation forces. In addition, both the theories are displacement-based and give only the
classical (extension, bending and torsion) stiffnesses and actuation forces. Timoshenko-like shear deformation
is not included. The present work extends the flexibility results of Ref. 1 to include embedded anisotropic
actuation. Ref. 1 presents a thin-walled analysis which includes transverse shear deformation so the cross-
sectional flexibility matrix is 6×6. It should be noted that all the above analyses neglect shell bending strains.
Volovoi and Hodges6 have shown that for certain special cases of beams with circumferentially non-uniform,
strongly anti-symmetric lay-ups, the hoop bending strain/moment can play a significant role, leading to
∗ Assistant Professor, Department of Aerospace and Ocean Engineering. Senior Member AIAA.
† Professor Emeritus, Department of Aerospace and Ocean Engineering. Senior Member Retired AIAA.

1 of 10

American Institute of Aeronautics and Astronautics


in-plane warping and reduction in torsional rigidity. Analysis that include shell bending effects increase the
complexity considerably, and for closely spaced wing ribs, cross-sectional distortion is minimized.
There is theoretical development done by Cesnik and Ortega-Morales7 and Cesnik and Palacios8 for
numerical (finite-element) solution to a general anisotropic cross-section with embedded actuation. The
former work presents a classical (4 × 4) formulation while the later includes shear deformation. The work is
an extension of passive theories by Cesnik and Hodges9 and Popescu and Hodges.10 The numerical work has
led to the development of various version of Variational-Asymptotic Beam Sectional (VABS) analysis that
include anisotropic strain actuation.
Other work, though for passive beams, relevant to the present work is the thin/thick-walled, cross-
sectional analysis of Jung et al.11 This work is described as able to provide closed-form expression but these
expressions are not presented in the paper. Giavotto et al.12 have developed a numerical, finite-element
based cross-sectional analysis of beams.

II. Theory
The beam cross-sectional analysis presented here follows closely the development in Ref. 1. The strain
energy derivation will be presented briefly for the sake of completeness, but the focus will be on the active
part of the derivation which is new in this paper.

A. Active laminate analysis


For an active composite lamina the in-plane stress and in-plane strain are related in its principal (fiber) axis
system (l, t) as,13       
 l 
 
1
El − νEltl 0   σl   dl 

t = − νEtlt 1
0 σ + d E (1)
 
Et  t t

  1

    
γlt 0 0 τlt 0
 
Glt

where, l , t , γlt are the normal strain (longitudinal) along the fiber direction, normal strain (transverse)
perpendicular to the fibers, and shear strain respectively. σl , σt , τlt are the corresponding stresses. El , Et ,
Glt are the modulus in the longitudinal, transverse and shear directions respectively. The Poisson’s ratio
are related as νtl = νltEEl t . dl and dt are the actuation strain coefficients and E is the applied electric field.
Normally the electric field is generated by the voltage applied across the laminate and is given by the ratio
V
of the applied voltage to the thickness between the electrodes E = telec . The above equation assumes a
unidirectional composite. For a general laminated lay-up the flexibility matrix and the actuation-vector may
be fully populated. In addition, if the electric field is applied in multiple directions then one would need to
consider a d-matrix (rather than a vector).
The in-plane stresses can be written as:
      
 σl 
  Qll Qlt 0  l 
    Qll dl + Qlt dt 

σt = Qlt Qtt 0  t − Qlt dl + Qtt dt E (2)
 

  
    
τlt 0 0 Qlt,lt γlt 0
  

where, the Q’s are the in-plane stiffness coefficients.


Let Tσ− transform the stresses from the lamina principal axis (l, t) system to the shell axis system (z, s,
see Figure 1(b)) and let T+ transforms the strains from the shell axis system to the lamina principal axis
system.14 Thus:
       
 σz 
  Qll Qlt 0  z 
   Qll dl + Qlt dt 
 
−  + −
σs = Tσ Qlt Qtt 0  T s − Tσ Qlt dl + Qtt dt E (3)

  
   
τzs 0 0 Qlt,lt γzs 0
   

For an active composite shell composed of multiple laminae, the in-plane shell stress can be obtained by
integrating Eq. 3 through the shell thickness. Thus, the active composite shell constitutive equations for

2 of 10

American Institute of Aeronautics and Astronautics


in-plane stress resultants is given by:
      
e
n
  B11 B12 B13   n 
  
ns = B12 B22 B23  s − nes (4)
 

  
   e
 
q B13 B23 B33 γ q
 

where, n, ns , q are the in-plane stress resultants along the beam axis, along the cross-sectional contour, and
the shear flow tangent to the contour, respectively (Figure 1(b)), and,
   
B11 B12 B13 X Qll Qlt 0
Tσ− Qlt Qtt
 +
B12 B22 B23  = 0  T (5)
  

B13 B23 B33 lamina 0 0 Qlt,lt


   
e
n 
  X

 Q ll d l + Qlt d t 

e −
ns = T σ Qlt dl + Qtt dt E (6)

 e  lamina  
q 0
 

Qy s ns (s)
Qx
Mx q(s)
z
My P T

z n(s)
(a) Beam forces and moments (b) In-plane stress resultants

Figure 1. Coordinate Axis System

B. Cross-sectional analysis
Following the derivation of Ref. 1, the no hoop stress (ns = 0) assumption leads to:

(n + ne ) = B + b(q + q e ) (7)
(q + q e ) = Bs γ + Bz  (8)

and
1
= [(n + ne ) − b(q + q e )] (9)
B
1
γ = [a(q + q e ) − b(n + ne )] (10)
B

3 of 10

American Institute of Aeronautics and Astronautics


where,
2
B12 B2
B = B11 − − bBz Bs = B33 − 23 (11)
B22 B22
2
 
B12 B23 1 B12 Bz
Bz = B13 − a= B11 − b= (12)
B22 Bs B22 Bs
B 12 B23 e
n e = ne − ne qe = qe − n (13)
B22 s B22 s

1. Normal stress resultant


Now assuming that the axial strain is linear function of the cross-sectional coordinates (free warping) we get:

 = W 0 + xθy0 + yθx0 (14)

where, W (z) is the axial deformation of the cross-section, and θx (z) and θy (z) are the rotations of the cross-
section about the x and y axes respectively. The sign convention for the rotations follow the sign conventions
of the moments as defined in Figure 1(a). ( )0 denotes a partial derivative with respect to z.
Now, the stress resultant can be written in terms of the deformation measures as:

(n + ne ) − b(q + q e ) = B W 0 + xθy0 + yθx0



(15)

By integrating the above equation over the contour we can reduce the cross-sectional stress distribution
into the cross-sectional axial force P , bending moments Mx and My , and the corresponding active forces
and moments P e , Mxe and Mye . The sign convention is given in Figure 1(a). Thus we can write the beam
axial force and bending moments as a function of the axial displacement and the rotations. We can then
solve for the axial displacements and rotations as:

P kx0   ky0 
W0 =

− 0 M y − x0 P − kx (M x − y0 P ) − 0 M x − y0 P − ky (M y − x0 P ) (16)
S Dy Dx
k
θx0 = 0 M x − y0 P − ky (M y − x0 P )
 
(17)
Dx
k 
θy0 = 0 M y − x0 P − kx (M x − y0 P )

(18)
Dy

where,
I I I
P = P + P e − b(q + q e ) ds P = n ds P e = ne ds (19)
I I I
Mx = Mx + Mx − b(q + q e )y ds
e
Mx = ny ds e
Mx = ne y ds (20)
I I I
My = My + Mye − b(q + q e )x ds My = nx ds Mye = ne x ds (21)

(22)

The rest of the parameters in Eqs. 16-18 are cross-sectional characteristics and are given by:
Sy Sx
x0 = y0 = (23)
I SI S I
S= B ds Sx = By ds Sy = Bx ds (24)
0 0
1 Dxy Dxy
k= kx = ky = (25)
1 − kx ky Dx0 Dy0
Dx0 = Dx − y02 S Dy0 = Dy − x20 S 0
Dxy = Dxy − x0 y0 S (26)
I I I
Dx = By 2 ds Dy = Bx2 ds Dxy = Bxy ds (27)

4 of 10

American Institute of Aeronautics and Astronautics


Eqs. 16-18 give the expressions for the displacements. Substituting in the Eq. 15 we can now derive an
expression for the normal stress resultant distribution. The normal stress resultant expression does have the
shear stress resultant (q) in it, but once the shear stress resultant is calculated we can be substituted it into
the above expression.

2. Shear stress resultant (shear flow)


The equation of equilibrium for the shell stress resultant is given by:
n0 + q̇ = 0 (28)
where, (˙) represents the partial derivative with respect to s.
Now, using q 0 = 0, P 0 = 0, Mx0 = Qy and My0 = Qx , we can write the shear flow in terms of the beam
shear forces Qx and Qy as:  
y x
q̇ = −Bk Qy + Qx (29)
Dx0 Dy0
where,
x = x − x0 − kx (y − y0 ) y = y − y0 − ky (x − x0 ) (30)
Thus, we get the shear flow as:
 Z s Z s 
Qy Qx
q = −k Byds + Bxds + q0 (31)
Dx0 0 Dy0 0

where, q0 is the shear flow at the origin of the contour coordinate. q0 can be calculated by satisfying torque
equivalence: I
T = rq ds (32)

where, r(s) is the perpendicular distance from the origin to the contour at any point on the contour. Using
the above equation one can solve for q0 in term of T to get:
1
q = Fx Qx + Fy Qy + T (33)
2A
where, Z s I Z s  
k 1
Fx = − Bx ds − Bx ds r ds (34)
Dy0 0 2A 0
Z s I Z s  
k 1
Fy = − By ds − By ds r ds (35)
Dx0
0 2A 0
I
1
A= r ds (36)
2
Now,substituting the shear flow into the normal stress resultant equation, we get:
 
B Bky0 y Bkx0 x Bky Bkx
(n + ne ) = − − (P + P e ) + 0 (Mx + Mxe ) + (My + Mye )
S Dx0 Dy0 Dx Dy0
(37)
Φ
+ Φx Qx + Φy Qy + T + Φe
2A
where, I  I I 
B y x
Φx = bFx − bFx ds − Bk bF x (y − y 0 ) ds + bF x (x − x 0 ) ds (38)
S Dx0 Dy0
I  I I 
B y x
Φy = bFy − bFy ds − Bk bFy (y − y0 ) ds + 0 bFy (x − x0 ) ds (39)
S Dx0 Dy
I  I I 
B y x
Φ=b− b ds − Bk b(y − y 0 ) ds + b(x − x 0 ) ds (40)
S Dx0 Dy0
 I I I
e e B Bky0 y Bkx0 x e Bky e Bkx
Φ = bq − − − bq ds − 0 byq ds − bxq e ds (41)
S Dx0 Dy0 Dx Dy0

5 of 10

American Institute of Aeronautics and Astronautics


3. Strain energy
The complementary strain energy of the beam per unit length can be written as,
I
∗ 1
U = [(n + ne ) + (q + q e )γ] ds (42)
2
where, the strain-resultant expressions are given in Eqs. 9 and 10.
The above equation is a quadratic function of the beam generalized forces and moments (P , Mx , My ,
Qx , Qy , T ). Using Castigliano’s theorem we have:
∂U ∗ ∂U ∗ ∂U ∗
W0 = θx0 = θy0 = (43)
∂P ∂Mx ∂My
∂U ∗ ∂U ∗ ∂U ∗
ψx = ψy = θ0 = (44)
∂Qx ∂Qy ∂T
where, θ is the torsional rotation. ψx and ψy are the shear deformations given by:
ψx = U 0 + θy ψy = V 0 + θx (45)
where, U and V are the displacements in the x and y directions.
After application of Castigliano’s theorem, we get the cross-sectional constitutive law as,
      
0  e


 θ x  C 11 C 12 C 13 C 14 C 15 C 16 
M x + M x  
E1 
 0   e
  
θ C C C C C C M + M E
  
 
 
 
 


 y 
  12 22 23 24 25 26  y
  y 
 
 2
 0    e
  

W C13 C23 C33 C34 C35 C36  
P + P
 
E 3

=C
 + (46)
Ψx   14 C24 C34 C44 C45 C46    Qx E4 
   

 
   
   
Ψy  C15 C25 C35 C45 C55 C56   Qy E5 

 
  
 
 
 

   
 
 


 0    
  
θ C16 C26 C36 C46 C56 C66 T E6
 

where,
I
k
E1 = − bq e y ds (47)
Dx0
I
k
E2 = − 0 bq e x ds (48)
Dy
I I I
1 e ky0 e kx0
E3 = − bq ds + 0 bq y ds + 0 bq e x ds (49)
S Dx Dy
I
1
E4 = (Φx Φe − bFx Φe − bΦx q e + aFx q e ) ds (50)
B
I
1
E5 = (Φy Φe − bFy Φe − bΦy q e + aFy q e ) ds (51)
B
I
1 1
E6 = (ΦΦe − bΦe − bΦq e + aq e ) ds (52)
2A B
The cross-sectional flexibility coefficients are as presented in Ref. 1 and are given in the Appendix.
Thus,       


 θx0 
 C11 C12
 C13 C14 C15 C16  Mx  
E1 
0     
θ  C12 C22 C23 C24 C25 C26   M E

 
 
 
 


 y 
   y
 
 2

 0    
W C13 C23 C33 C34 C35 C36  P E
     
3
=  + (53)
  14 C24
C C34 C44 C45 C46 
 Ψ
 x
   Qx 
  E4 
  
 
 Ψy  C15 C25 C35 C45 C55 C56  
 Qy  E5 

   
 
 
     


 0  
    
θ C16 C26 C36 C46 C56 C66 T E6
 

where,
E1 = C11 Mxe + C12 Mye + C13 P e + E 1 (54)
E2 = C12 Mxe + C22 Mye e
+ C23 P + E 2 (55)
E3 = C13 Mxe + C23 Mye + C33 P e + E 3 (56)

6 of 10

American Institute of Aeronautics and Astronautics


III. Results

CEC1 0° CEC1 +45°


AFC +45° (+2800 V) CEC1 +45°
AFC -45° (-1200 V) CEC1 +45°
CEC1 0° CEC1 +45°
24.49 mm

CEC1 0° CEC1 0° AFC -45° (±2000 V) AFC -45° (±2000 V)

25 mm
CEC1 0° CEC1 0° AFC -45° (±2000 V) AFC -45° (±2000 V)
CEC1 0° CEC1 0° AFC -45° (±2000 V) AFC -45° (±2000 V)
CEC1 0° CEC1 0° AFC -45° (±2000 V) AFC -45° (±2000 V)

0.5 in
CEC2 [15°]6
CEC1 0° CEC1 +45°
AFC +45° (+2800 V) CEC1 +45°
AFC -45° (-1200 V) CEC1 +45°
CEC1 0° CEC1 +45°

24.49 mm 25 mm 0.923 in

(a) Box-Beam BB17 (b) Box-Beam BB38 (c) Box-Beam B110

Figure 2. Geometry and ply lay-up for the test cases

CEC AFC
El (GPa) 142 42.2 psi
Et (GPa) 9.80 17.5
Glt (Pa) 6.00 5.50
νlt 0.300 0.354
tply (mm) 0.127 0.127
dl (nm/V) - 0.381
dt (nm/V) - -0.160
telec (mm) - 1.143

Table 1. Material properties for the test cases

To validate the theory, various box beam cases are studied. The geometry and ply lay-up for the three
test cases is shown in Figure 2. The properties of the material used in the test cases is presented in Table 1.
First test case is box beam BB1 from Cesnik and Ortega-Morales.7 It is 4-ply, 0◦ , graphite/epoxy box
with 2 plies at the top and bottom replaced by +45◦ /−45◦ active fibre composites. The cross-sectional
flexibilities and strain actuation coefficients are presented in Table 2.
As can be seen from the Table 2, the difference between the present results and the results from Cesnik
and Shin (a thin-walled, classical theory) are negligibly small (< 0.1%) for all the classical flexibilities and
actuation strains. The difference in the present results and the results of Cesnik and Ortega-Morales (a
classical FEM theory) are also quite small. The differences can be attributed to a combination of wall
thickness effects and numerical errors associated with FEM results. The convergence of FEM results for
very thin-walled beams is especially slow due to errors associated with high-aspect-ratio solid elements.
The results using the present theory also compare well with the results from Cesnik and Palacios (a FEM
Timoshenko-like theory). The values shear flexibilities differ only by around 1%.
Box beam BB1 does not exhibit any couplings. The second test case is that of box beam BB3 from
Cesnik and Palacios.8 The box beam is composed of 4-ply, +45◦ , graphite/epoxy on the top and bottom
side and 4-ply, −45◦ , active fiber composite on the sides. This lay-up leads to extension-torsion coupling
and bending-shear coupling. As seen from Table 3, the present theory accurately calculates the classical
flexibility coefficients and the corresponding strain actuation coefficients.
The shear and shear coupling coefficient estimates from the two theories are quite different. It is not
clear which of the answers is correct. Thus another standard for comparison is required. Such a theory is
not available for active cross-sectional analysis, but there are published results for passive cross-sectional
flexibilities.
The third test case considered here is the box beam configuration B1 from Popescu and Hodges.10 This
box beam configuration is a circumferentially-uniform, 6-ply, −15◦ , rectangular box. The paper compares

7 of 10

American Institute of Aeronautics and Astronautics


Flexibilities Present Cesnik Cesnik Cesnik
and Shin4 Ortega-Morales7 Palacios8
Actuation Thin-Wall (6×6) Thin-Wall (4×4) FEM (4×4) FEM (6×6)
C11 (bend vert) ×103 2.109 2.109 2.107 2.107
C22 (bend horiz) ×103 1.590 1.589 1.590 1.589
C33 (ext) ×107 1.812 1.812 1.812 1.812
C44 (shear horiz) ×106 5.956 - - 5.892
C55 (shear vert) ×106 7.455 - - 7.411
C66 (tors) ×102 1.842 1.843 1.829 1.829
E3 (ext) ×106 6.350 6.345 6.265 6.419
E6 (tors) ×102 -3.167 -3.169 -3.069 -3.090

Table 2. Results for box beam BB1 from Ref. 7

Flexibilities Present Cesnik % Diff.


and Palacios8
Actuation Thin-Wall (6×6) FEM (6×6)
C11 (bend vert) ×102 1.250 1.249 -0.1
C22 (bend horiz) ×102 1.190 1.189 -0.1
C33 (ext) ×106 1.270 1.270 0.0
C44 (shear horiz) ×106 4.683 4.695 +0.3
C55 (shear vert) ×106 4.298 5.083 +18.3
C66 (tors) ×102 1.128 1.119 -0.8
C14 ×105 -9.847 -9.799 -0.5
C25 ×105 -1.510 -1.542 +2.1
C36 ×105 -2.173 -2.157 -0.7
Twist Actuation
E3 (ext) ×104 1.015 1.009 -0.6
E6 (tors) ×102 -4.171 -4.140 -0.7
Bending Actuation
E2 (bend horiz) ×102 1.188 1.181 -0.6
E5 (shear vert) ×104 9.841 9.820 -0.2

Table 3. Results for box beam BB3 from Ref. 8

Flexibilities Present Jung et al.11 NABSA12 VABS10


Thin-Wall Thick-Wall (%) FEM (%) FEM (%)
C11 (bend vert) ×105 2.825 2.721 (3.7) 2.763 (2.2) 2.678 (5.2)
C22 (bend horiz) ×105 1.118 1.104 (1.3) 1.089 (2.6) 1.078 (3.6)
C33 (ext) ×106 1.352 1.334 (1.4) 1.335 (1.3) 1.327 (1.9)
C44 (shear horiz) ×105 2.050 2.057 (-0.4) 2.029 (1.0) 2.902 (-41.6)
C55 (shear vert) ×105 5.075 4.990 (1.7) 4.782 (5.8) 6.943 (-36.8)
C66 (tors) ×104 1.178 1.148 (2.5) 1.143 (2.9) 1.115 (5.3)
C14 ×105 1.656 1.614 (2.6) 1.595 (3.7) 1.567 (5.4)
C25 ×105 -1.656 -1.619 (2.3) -1.562 (5.7) -1.532 (7.5)
C36 ×106 8.794 8.538 (2.9) 8.550 (2.8) 8.411 (4.3)

Table 4. Results for box beam B1 from Ref. 10

8 of 10

American Institute of Aeronautics and Astronautics


the results with those obtained using NABSA,12 another FEM cross-sectional analysis code which accounts
for shear deformation. This case was also studied by Jung et al.11 The results from the present closed-form
solution is presented in Table 4, and compared with the three other results. All the four results are quite
close for the classical flexibilities. The shear and shear coupling flexibilities calculated using the present
approach are close to the results of Ref. 11 and Ref. 12. The shear flexibility results of Ref. 10 differ by
around 40% from the other three results. Overall the results from the present work are around 1 - 4 % off
from the results of Ref. 11 and 1 - 6 % off from the results of Ref. 12.

IV. Conclusions
The paper presents the cross-sectional analysis of anisotropic, thin-walled, closed section beams with
embedded strain actuation. Closed-form expressions are derived for the cross-sectional flexibilities and
actuation strains. Preliminary results are obtained for three cases, including cases with active anisotropic
strain actuators and structural coupling. These results compare very well with published results. Available
thin-walled theories only give the classical flexibilities and the corresponding actuation strain. Timoshenko-
like active results can be generated using an available FEM-based theory, but there is some discrepancy in
predicting the shear flexibilities between the present theory and theories based on variational asymptotic
method (displacement formulation). The present theory compares well with other theories based on stress
formulation. The present theory can be used to generate Timoshenko-like (6 × 6) active results efficiently
using closed-form expressions and thus is ideal for preliminary design and optimization.

Appendix

k
C11 = (57)
Dx0
kky
C12 = − (58)
Dx0
k kkx
C13 = − y0 + 0 x0 (59)
Dx0 Dy
I
k
C14 = − 0 bFx y ds (60)
Dx
I
k
C15 = − 0 bFy y ds (61)
Dx
I
k
C16 = − by ds (62)
2ADx0
k
C22 = (63)
Dy0
k kky
C23 = − x0 + 0 y0 (64)
Dy0 Dx
I
k
C24 = − 0 bFx x ds (65)
Dy
I
k
C25 = − 0 bFy x ds (66)
Dy
I
k
C26 = − bx ds (67)
2ADy0
1 k k
+ 0 y02 − ky x0 y0 + 0 x20 − kx x0 y0
 
C33 = (68)
S Dx Dy

9 of 10

American Institute of Aeronautics and Astronautics


I I I
1 ky0 kx0
C34 = − bFx ds + bFx y ds + bFx x ds (69)
S Dx0 Dy0
I I I
1 ky0 kx0
C35 = − bFy ds + bFy y ds + bFy x ds (70)
S Dx0 Dy0
I I I
1 ky0 kx0
C36 = − b ds + by ds + bx ds (71)
2AS 2ADx0 Dy0
I
1  2
Φx − 2bFx Φx + aFx2 ds

C44 = (72)
B
I
1
C45 = [Φx Φy − bFx Φy − bFy Φx + aFx Fy ] ds (73)
B
I
1 1
C46 = [Φx Φ − bFx Φ − bΦx + aFx ] ds (74)
2A B
I
1  2
Φ − 2bFy Φy + aFy2 ds

C55 = (75)
B y
I
1 1
C56 = [Φy Φ − bFy Φ − bΦy + aFy ] ds (76)
2A B
I
1 1  2 
C66 = 2
Φ − 2bΦ + a ds (77)
4A B

Acknowledgement
The authors would like to acknowledge some thoughtful discussions with, as well as box beam data
provided by, Prof. Carlos Cesnik and Dr. Rafael Palacios from University of Michigan, Ann Arbor.

References
1 Johnson, E. R., Vasiliev, V. V., and Vasiliev, D. V., “Anisotropic Thin-Walled Beams with Closed Cross-Sectional

Contours,” AIAA Journal, Vol. 39, No. 12, 2001, pp. 2389–2393.
2 duPlessis, A. J. and Hagood, N. W., “Modeling and experimental testing of twist actuated single cell composite beams

for helicopter rotor blade control,” Tech. Rep. AMSL 96-1, Active Material Systems Laboratory, Massachusetts Institute of
Technology, 1996.
3 Rehfield, L. W., “Design Analysis Methodology for Composite Rotor Blades,” Proceedings of the Seventh DoD/NASA

Conference on Fibrous Composites in Structural Design, Denver, Colorado, June, 1985.


4 Cesnik, C. E. S. and Shin, S.-J., “On the modeling of active helicopter blades,” International Journal of Solids and

Structures, Vol. 38, No. 10-13, 2001, pp. 1765 – 1789.


5 Berdichevsky, V. L., Armanios, E. A., and Badir, A. M., “Theory of Anisotropic Thin-Walled Closed-Section Beams,”

Composites Engineering, Vol. 2, No. 5 – 7, 1992, pp. 411 – 432.


6 Volovoi, V. V. and Hodges, D. H., “Theory of Anisotropic Thin-Walled Beams,” Journal of Applied Mechanics, Vol. 67,

No. 3, 2000, pp. 453 – 459.


7 Cesnik, C. E. S. and Ortega-Morales, M., “Active Beam Cross-Sectional Modeling,” Journal of Intelligent Material

Systems and Structures, Vol. 12, No. 7, July 2001, pp. 483 – 496.
8 Cesnik, C. E. S. and Palacios, R., “Modeling Piezocomposite Actuators Embedded in Slender Structures,” Proceedings

of the 44th Structures, Structural Dynamics, and Materials Conference, Norfolk, Virginia, April 2003, AIAA Paper 2003-1803.
9 Cesnik, C. E. S. and Hodges, D. H., “VABS: A New Concept for Composite Rotor Blade Cross-Sectional Modeling,”

Journal of the American Helicopter Society, Vol. 42, No. 1, January 1997, pp. 27 – 38.
10 Popescu, B. and Hodges, D. H., “On asymptotically correct Timoshenko-like anisotropic beam theory,” International

Journal of Solids and Structures, Vol. 37, No. 3, Oct. 2000, pp. 535 – 558.
11 Jung, S. N., Nagaraj, V. T., and Chopra, I., “Refined structural model for thin- and thick-walled composite rotor blades,”

AIAA Journal, Vol. 40, No. 1, Jan. 2002, pp. 105 – 116.
12 Giavotto, V., Borri, M., Mantegazza, P., Ghiringhelli, G., Carmaschi, V., Maffioli, G. C., and Mussi, F., “Anisotropic

Beam Theory and Applications,” Computers and Structures, Vol. 16, No. 1-4, 1983, pp. 403 – 413.
13 Bent, A. A., Hagood, N. W., and Rodgers, J. P., “Anisotropic Actuation With Piezoelectric Fiber Composites,” Journal

Of Intelligent Material Systems And Structures, Vol. 6, No. 3, May 1995, pp. 338 – 349.
14 Tsai, S. W. and Hahn, H. T., Introduction to Composite Materials, Technomic Publishing Company, Inc., Lancaster,

Pennsylvania, 1980.

10 of 10

American Institute of Aeronautics and Astronautics

Você também pode gostar