Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

AIAA 2016-1242

AIAA SciTech
4-8 January 2016, San Diego, California, USA
57th AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference

Nonlinear Modelling of Axially Deformable Elastica


based on Hyperelasticity

Fang Jiang∗, Su Tian† and Wenbin Yu‡


Purdue University, West Lafayette, IN 47907-2045, USA
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

Considering the cross-sectional dimension of the elastica much smaller than the wave-
length of its global deformation, the three-dimensional hyperelasticity is rigorously split
to a one-dimensional macroscopic global analysis and a two-dimensional cross-sectional
analysis. These two analyses are featuring both geometric and material nonlinearities.
The two-dimensional nonlinear cross-sectional analysis utilizes the three-dimensional hy-
perelastic material models to homogenize the nonlinear beam cross-sectional constitutive
relations, which are input as material properties for the global one-dimensional geomet-
ric exact beam analysis. Restricted to small strains, analytical solutions can be obtained
and proves that the sectional analysis of the Euler-Bernoulli type results in an effect that
the bending stiffness will vary with the axial deformation. Without ad hoc assumptions,
exact solution of such extension effect for the strain energy up to the third order is given
analytically for isotropic beams with neo-Hookean model. In addition, the present the-
ory without the small strain restriction is implemented using the finite element method
in VABS, a general-purpose cross-sectional analysis tool, of which the solutions are com-
pared with the analytical results. The geometrical exact beam theory is extended to handle
one-dimensional nonlinear material model so that both geometry and material nonlinear-
ities can be handled. Finally, the present theory is validated using the three-dimensional
analysis in commercial finite element software.

I. Introduction
One-dimensional (1D) beam models with finite strain deformation and hyperelastic and/or elastoplastic
constitutive relations are widely studied.1–11 However, all these models are based on ad hoc assumptions
such as rigid cross-sections.12, 13 Focusing on predicting the nonlinear global behavior of the beam, the
cross-sectional properties (EI, EA, etc. E is the Young’s modulus. I is the second moment of the cross-
sectional area. A is the cross-sectional area.) are evaluated by approaches which are usually based on ad
hoc assumptions and are only valid for small deformation and linear elastic behavior. However, accurate
evaluation of the cross-sectional properties is a prerequisite for predictive global behavior analysis.
To account for the effects of nonlinear beam cross-sectional properties, the cantilever beam was previously
studied with an assumed nonlinear moment-curvature relation.14 Another matter of expediency for nonlinearly
elastic materials is to take the nonlinear strain-stress data curves from the experiments and plug into EI, EA,
etc. with rigid cross section. However, this approach heavily depends on the size of the data base from the
experimental activities and not applicable for finite deformation with complicated loading conditions, which
is the very case of the elastica. The concept of elastica is originally proposed for problems on large deflections
of slender structures without axial deformation and has been implemented into finite element approach.15
The bending stiffness (or the flexural rigidity) is estimated as EI. When the axial load is considered the
axial deformation rigidity is assumed to be EA. In addition, in some cases the models are also created under
the assumption of inextensibility and linear elasticity. Such a theory is widely applied in solving the buckling
problems,16–19 designing the compliant mechanisms,15, 20–24 building the soft robots,25–28 and developing the
microelectromechanical systems (MEMS).29, 30
∗ Graduate Research Assistant, Ph.D. Candidate, School of Aeronautics and Astronautics.
† Graduate Research Assistant, Master Student, School of Aeronautics and Astronautics.
‡ Associate Professor, School of Aeronautics and Astronautics, Associate Fellow, AIAA; Fellow, ASME; Member AHS.

1 of 17

American Institute of Aeronautics and Astronautics


Copyright © 2015 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
On the other hand, beam theories on the arbitrary cross-sections and linear elastic composite materials
have been well developed with the restriction of small strains since 1980s.31–34 The systematic methodology
contributed by these works split the original three-dimensional (3D) geometrically nonlinear, materially linear
problem of the slender structure into a 1D geometrically exact beam theory (GEBT)35 and two-dimensional
(2D) variational asymptotic beam sectional (VABS) analysis. GEBT was first proposed by Ressiner1, 36 and
further developed by Simo,6 Simo and Vu-Quoc,7 Jelenić and Crisfield,37 Betsch and Steinmann,38 Ibrahim-
begović,39 Ibrahimbegović and Mikdad.40 VABS takes care of the cross-sectional analysis to provide the
structural and inertial properties for the 1D beam analysis.41–47 VABS also provides a set of algebraic rela-
tions to recover the complete 3D fields. The only restriction is that the strains are small which is needed for
geometrically nonlinear analysis and the structure to be slender which is the very motivation and justifica-
tion for using beam models. All the deformations of the cross-section are rigorously captured in VABS. By
means of these theoretical foundations, Jiang et al.48, 49 extended the sectional analysis to the finite strain
field featuring both geometric and material nonlinearities. Using such an advanced theory, various nonlinear
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

effect can be explained both analytically and numerically.


In this paper, a further extension is proposed to analyze geometrically and materially nonlinear elastica.
A shown in Figure 1, succeeding the spirits of nonlinear composite beam analysis,50 the original 3D nonlin-
ear anisotropic continuum mechanics is rigorously split into a 2D cross-sectional analysis and a 1D beam
analysis. By asymptotic analysis and utilizing hyperelastic material models, the 2D cross-sectional analysis,
that is, VABS, provides the structural properties featuring on both material and geometrical nonlinearities.
Following the methodology in Jiang, Yu, and Hodges48, 49 analytical sectional analysis is given for the elastica
undergoing combined stretching and bending. Exact solution is obtained for isotropic case with neo-Hookean
material model based on the strain energy asymptotically correct up to the third order of the strain. The
finite element method is used to implement this theory, of which the solutions are compared with the ana-
lytical results for validations. The nonlinear cross-sectional properties provided by VABS is used as material
properties in the 1D beam analysis tool, GEBT, which is further developed here mainly based on the work
done by Hodges,50 Yu and Blair35 and Wang et al.51 The mixed formulation is used, which is the combination
of exact intrinsic equations and kinematical equations by Lagrange multipliers. More details can be found
in Hodges.50 Analytically, the sectional forces and moments in the nonlinear algebraic equations of GEBT
are substituted by using the nonlinear beam constitutive relations. These relations express the the sectional
forces and moments in terms of the axial strain, curvature, and cross-sectional geometry, nonlinearly. Finally,
3D finite element model of a slender cantilevered elastica is created in the commercial finite element software
ANSYS. The results of this 3D analysis prove that the axial deformation is triggered by the pure bending
moment applied at the free end. This fact is found to match the results of present method very well.

3-D Anisotropic Continuum Mechanics

Dimensional Reduction Using VAM

VABS GEBT
Cross-Sectional Analysis Nonlinear Global Beam Analysis
(Material and Geometrical Sectional (Material and Geometrical
Nonlinearities) Properties Nonlinearities)

Nonlinear Recovery Relations Global Behavior

3-D Displacement/Strain/Stress Fields

Figure 1. Present Methodology for Elastica with both Material and Geometrical Nonlinearities.

2 of 17

American Institute of Aeronautics and Astronautics


II. Nonlinear Variational Asymptotic Beam Sectional Analysis
A. Nonlinear Asymptotic Formulation
In Figure 2, ei for i=1, 2, 3 are fixed dextral, mutually perpendicular unit vectors in the absolute reference
frame, and r0 and R0 denote the position vector of the material point on the reference line of the undeformed
and deformed configurations, respectively. (Here and throughout all the paper, Greek indices assume values
2 and 3 while Latin indices assume 1, 2, and 3. Repeated indices are summed over their range except
where explicitly indicated.) bi and Bi are the rigid orthogonal triads attached to the cross-section in the
undeformed and deformed configurations, respectively.
b3 b2
g2
g3 b
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

g1 1

r0 u B3
G1 G2
e3 B2
R0 G3
e1 e2
B1

Figure 2. Schematic of underformed and deformed beam.

Then the material position vectors in the undeformed and deformed beam body can be expressed as

r = r0 + xα bα (1)

R = R0 + xα Bα + wi (x1 , x2 , x3 )Bi (2)


with wi representing the 3D unknown warping functions to describe the difference between the position of
deformation body and those can be described by deformation of the reference curve x1 in terms of R0 +xα Bα .
R0 can also be expressed as
R0 = r0 + u (3)
where u denotes the beam displacement. Note u is not the displacement of some material point in the
original structure. Rather it is the displacement field of the beam model we are constructing. Note in
Eq. (2), we actually express R in terms of R0 , B i , and wi , which is six times redundant. Six constraints
are needed to ensure a unique mapping. We are free to choose the triad B i and we can choose B 1 to be
tangent to x1 which introduces two constraints. As discussed in Ref. [44], we can additionally introduce the
following four constraints for the warping functions:

hwi i = 0 hw2,3 − w3,2 i = 0 (4)

From here and throughout the paper we assume a prismatic beam with uniform cross-sectional geometry.
To derive a theory of the classical type, we define the following generalized 1D strains:

R00 = (1 + γ)B1 (5)

Bi0 = κj Bj × Bi (6)
in which the upper prime denotes derivative to x1 , γ the axial strain, κ1 the twist, and κα the curvature of
the deformed beam reference line. In Figure 2, gi denote the covariant base vectors of the undeformed body.
And let the controvariant base vectors of the undeformed body denoted by g i . Then we have

gi = g i = bi (7)

3 of 17

American Institute of Aeronautics and Astronautics


The covariant base vectors of the deformed configuration can be evaluated as
∂R
Gk = (8)
∂xk
Together with Eq. (6), we have

G1 =[1 + γ + w10 − (x2 + w2 )κ3 + (x3 + w3 )κ2 ]B1


+ [w20 − (x3 + w3 )κ1 + w1 κ3 ]B2 (9)
+ [w30 + (x2 + w2 )κ1 − w1 κ2 ]B3
G2 =w1,2 B1 + (1 + w2,2 )B2 + w3,2 B3 (10)
G3 =w1,3 B1 + w2,3 B2 + (1 + w3,3 )B3 (11)
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

Then the deformation gradient tensor can be formulated as

F = Gi bi (12)

Denoting the order of the beam sectional dimension as h and that of the axial dimension as L, and using
the book-keeping parameter ε to express the maximum strain magnitude, we have
h
1 (13)
L
ε = max(γ, hκ1 , hκ2 , hκ3 ) (14)
Static hyperelasticity is governed by a strain energy density function W which can be ultimately written in
terms of the deformation gradient tensor F , that is,

W = W (F ) (15)

The static behavior of the structure is governed by the principle of virtual work such that
Z L
δ hW i dx1 − δΨ = 0 (16)
0

where δΨ is the virtual work due to applied loads in the original 3D structure. Note for simplicity, we used
the principle of virtual work governing the static behavior but what is derived here is directly applicable to
dynamic behavior.44 Carrying out an asymptotical analysis similar as that in Ref. [44], we can conclude that
virtual work done by applied loads to the warping functions are negligible. Thus, to solve for the warping
function, we require the variation of the strain energy Π vanishes, that is

δΠ = δ hW i = 0 (17)

Based on the previous work on VABS,31–34, 41–46 it has been shown that

wi = O(hε) (18)

Then it is clear that the following is still true for the situation under large warping displacements and finite
beam strains
wi h 
wi0 ∼ =O ε ε (19)
L L
In light of Eq. (19), the wi0 terms in Eq. (9) can be neglected because their contributions to the strain energy
are much smaller than the other terms.

4 of 17

American Institute of Aeronautics and Astronautics


B. Finite Element Implementation
This section is slightly adapted from Ref. [48] on developing a numerical implementation to model beams
made of general hyperelastic material and subject to large strains. For prismatic bars, the variation of the
functional in Eq. (17) can be carried out as
D ∂W E
δFij = hPij δFij i = 0 (20)
∂Fij

where Pij is the so-called first Piola-Kirchhoff stress tensor (PK1 ) defined by

∂W
Pij ≡ (21)
∂Fij
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

To solve the minimization problem in Eq. (20) using the finite element method, we need to introduce the
following matrix notations for the components of deformation gradients in Eq. (12):
h iT
F = F11 F12 F13 F21 F22 F23 F31 F32 F33 (22)
h iT
∆= 1 0 0 0 1 0 0 0 1 (23)
h iT
w = w1 w2 w3 (24)
 
1 0 x3 −x2
 0 0 0 0 
 
 
 0 0 0 0 
 
 0 −x 0 0 
 3 
F =  0 0 0 0  (25)
 
 
 0 0 0 0 
 
 0 x 0 0 
 2 
 0 0 0 0 
 

0 0 0 0
 
0 −κ3 κ2
 ∂ 
 ∂x
2
0 0 
 
 ∂ 0 0 
 ∂x3 
−κ1 
 
 κ3 0
 ∂

Fh =  0 ∂x2 0  (26)

 0 0
 
 ∂x3 

 −κ2 κ1 0 
 
 0 ∂
 0 
∂x2 

0 0 ∂x3

So that the components in Eq. (12) can be rewritten in the following matrix form:

F = ∆ + F  + Fh w (27)

where the beam strains are arranged in


h iT h iT
 = (1) (2) (3) (4) = γ κ1 κ2 κ3 (28)

Beam strains in Eq. (28) are user inputs as strain loadings. In addition, let w be discretized using finite
elements as
w = SV (29)

5 of 17

American Institute of Aeronautics and Astronautics


where S denotes the shape functions, and V denotes a column matrix of the nodal values of the warping
functions. Substituting Eq. (29) into Eq. (27) then into Eq. (20), we arrive at the following nonlinear
algebraic equations
Ω(V ) = h(Fh S)T P i = 0 (30)
where PK1 stress components are carried by
h iT
P = P11 P12 P13 P21 P22 P23 P31 P32 P33 (31)

And the integration, h i, is conducted by using the Gaussian quadrature.


Standard approaches such as the Newton-Raphson method can be used to solve the nonlinear equations
in Eq. (30) iteratively for V . Let Vo denote a given numerical column matrix of the nodal warping values
which usually starts with a null vector, and let dV denote a correction to the Vo . Then in the neighborhood
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

of Vo , Ω in Eq. (30) can be expanded in a Taylor series with the high order terms neglected as
 
∂Ω
Ω(Vo + dV ) ≈ Ω(Vo ) + dV = 0 (32)
∂V V =Vo

Solving Eq. (32) will give the correction dV as


 −1
∂Ω
dV = − Ω(Vo ) (33)
∂V V =Vo

where
∂Ω
= h(Fh S)T [A](Fh S)i (34)
∂V
with [A] as the 9 × 9 matrix condensed from the components of the fourth-order first elasticity tensor Aijkl ,
which, for numerical implementation, is defined as

∂Pij ∂2W
Aijkl ≡ ≡ (35)
∂Fkl ∂Fij ∂Fkl

Meanwhile, the four constraints in Eq. (4) can be written in matrix form as

h(Fc S)iV = 0 (36)

with  
1 0 0
 0 1 0 
Fc =  (37)
 

 0 0 1 
∂ ∂
0 ∂x3 − ∂x 2

Giving  as a strain loading in forms of stepwise updating of o by incremental d as

 = o + d (38)

we can solve for dV iteratively for each loading step using


" # " #
dV h(Fh S)T P i
Ξ =− (39)
Λ [0]4×1

where Ξ is the tangent coefficient matrix


" #
h(Fh S)T [A](Fh S)i h(Fc S)T i
Ξ= (40)
h(Fc S)i [0]4×4

6 of 17

American Institute of Aeronautics and Astronautics


and Λ is a column vector consisting of Lagrange multipliers for applying the four warping constraints to the
problem
h iT
Λ = Λ1 Λ2 Λ3 Λ4 (41)

Then update the warping Vn by the converged correction dV during iterations of Eq. (39) as

Vn = Vo + dV (42)

which is followed by obtaining the deformation gradient F and the PK1 stress P in terms of  and Vn .
To calculate the sectional equivalent force and moments, define a column vector in order to carry out the
derivative of  in Eq. (28) with respect to beam strains
∂ h iT
δ̄(i) = = δi1 δi2 δi3 δi4 i = 1, 2, 3, 4 (43)
∂(i)
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

where δ(ij) = 1 if i = j and δ(ij) = 0 if i 6= j, in which i, j = 1, 2, 3, 4. In addition, denote R as a column


matrix arranging the resultant equivalent axial extensional force F1 , twisting moment M1 , and bending
moments M2 and M3 , that is
h iT h iT
R = R(1) R(2) R(3) R(4) = F1 M1 M2 M3 (44)

Finally in light of Eq. (30), we can evaluate the components in R corresponding to the loading strains  as
* T +
∂hW i ∂(Fh S)
R(i) = = F δ̄(i) + Vn P i = 1, 2, 3, 4 (45)
∂(i) ∂(i)

In summary, as long as a strain energy function W and the strain loading  in Eq. (38) are given, iteratively
solving Eq. (39) will provide a finite warping field for Eq. (45) to obtain an set of accurate sectional force
and moments as a hyperelastic beam constitutive relation.

III. Global Beam Analysis


The advantage of GEBT is that it can capture all geometrical nonlinearalities in a systematical way,
which is very important for structures or mechanisms containing large deformations. Another feature of this
approach is that the formulation remains the same no matter what kind of material is used. The intrinsic
equations of motion are derived from Hamilton’s extanded principle:
Z t2 Z l
 
δ(K − U) + δW dx1 dt = δA (46)
t1 0

where t1 and t2 are arbitrary fixed time, δ is the usual Lagrangean variation for a fixed time, K and U are
the kinetic and strain energy densities, respectively, δW is the virtual work density and δA is the virtual
action at the ends of the beam and at the ends of time interval. The bars over variations are used to indicate
that the virtual quantities need not be the variations of functions. We can expand Eq. (46) as:
Z t2 Z l h i
T T
δV T P + δΩT H − δγ T F − δκT M + δq f + δψ m dx1 dt
t1 0
Z l Z t (47)
T T t2 2
T T l
= (δq P̂ + δψ Ĥ) dx1 − (δq F̂ + δψ M̂ ) dt
0 t1 t1 0

where V and Ω are the linear and angular velocities of the beam reference line, respectively; P and H are the
sectional linear and angular momentum, respectively; γ and κ are the force and moment strain measures,
respectively; F and M are the sectional force resultant and moment resultant, respectively; δq and δψ are
the virtual displacement and rotation, respectively; f and m are the distributed forces and moments per unit
length, respectively; P̂ and Ĥ are the linear momentum and angular momentum, respectively, evaluated at
the ends of time interval; F̂ and M̂ are the forces and moments, respectively, evaluated at the ends of space
interval.

7 of 17

American Institute of Aeronautics and Astronautics


To complete the mixed formulation, we need to introduce a group of kinematical relations through La-
grange multipliers, so that all necessary variables for solving a problem are included in the single formulation
and can vary independently. These relations can be written as:

u0 = C bB (e1 + γ) − e1 − k̃u u̇ = C bB V − v − ω̃u (48)


0 −1 Bb −1 Bb
c =Q (κ + k − C k) ċ = Q (Ω − C )ω (49)

where u is the 1D displacement; v and ω are the linear and angular velocities, respectively, of the undeformed
triad b in a global inertial frame; k is the initial curvatures; e1 is defined as a column matrix as e1 = [1 0 0]T ;
c is the Wiener-Milenković rotation parameter; C bB and C Bb are the rotation matrices; Q is also a quantity
related with rotation. Those rotation related quantities are defined as:
φ
c = 4ntan (50)
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

4
cT c
 

c0 = 2 1 − tan =2− (51)
4 8
1
C Bb = (c20 − cT c)∆ − 2c0 c̃ + 2ccT
 
2
(52)
(4 − c0 )
C bB = (C Bb )T (53)
  
1 1 T 1 T
Q= 4 − c c ∆ − 2c̃ + cc (54)
(4 − c0 )2 4 2

where n is the unit vector of rotation axis and φ is the rotation magnitude according to Euler’s rotation
theorem; ∆ is a 3 × 3 identity matrix.
After identifying the Lagrange multipliers, the complete mixed formulation can be written as
Z l
0T T
h i
T
δu0a Fa + δψ a Ma + δψ a Ḣa + ω̃a Ha + Ṽa Pa − C aB (ẽ1 + γ̃)FB
0
T  0T 0T
+ δuTa (Ṗa + ω̃a Pa ) − δF a C aB (e1 + γ) − C ab e1 − δF a ua − δM a ca


T T
− δM a Q−1 ab (55)
a C κ + δP a (Va − va − ω̃a ua − u̇a )
T T
o
+ δH a (ΩB − ωB − C ba Qa ċa ) − δuTa fa − δψ a ma dx1
 T T T
l
= δuTa F̂a + δψ a M̂a − δF a ûa − δM a ĉa
0

where subscript indicates the frame where quantities are defined and a represents the global reference frame;
C Ba = C ba C with C ba the relation between frame a and undeformed frame b, which is determined and time
independent, and C the same as C Bb in Eq. (52) where c is replaced by ca ; and Qa = C ab QC ba .
To carry out the finite element implementation of the variational statement in Eq. (55), we use linear
shape functions for the test functions δua , δψ a , δF a and δM a and use constants for other functions. Dividing
the beam into N elements with the starting node of ith element numbered as i and the ending node numbered
as i + 1, and using shape functions provided above, the integration in Eq. (55) can be evaluated analytically,
resulting in
N n
X T T T T T − T
δuTi fu−i + δuTi+1 fu+i + δψ i fψ−i + δψ i+1 fψ+i + δF i fF−i + δF i+1 fF+i + δM i fM i
+
+ δM i+1 fM i
i=1
T T
o
+ δP i fPi + δH i fHi (56)
T T T
= δuTN +1 F̂N +1 + δψ N +1 M̂N +1 − δF N +1 ûN +1 − δM N +1 ĉN +1
T T T
− δuT1 F̂1 − δψ 1 M̂1 + δF 1 û1 + δM 1 ĉ1

8 of 17

American Institute of Aeronautics and Astronautics


with
∆Li h ˙
i
fu∓i = ∓C T C ab Fi − f¯i∓ + ω̃a C T C ab Pi + C T C ab Pi (57)
2
∓ T ab ∓ ∆Li h ˙
i
fψi = ∓C C Mi − m̄i + ω̃a C T C ab Hi + C T C ab Hi + C T C ab (Ṽi Pi − (ẽ1 + γ̃i )Fi ) (58)
2
∓ ∆Li  T ab
C C (e1 + γi ) − C ab e1

fFi = ±ui − (59)
2
∓ ∆Li −1 ab
fMi = ±ci − Qa C κi (60)
2
fPi = C T C ab Vi − vi − ω̃a ui − u̇i (61)
ba ba
fHi = Ωi − C Cωa − C Qa ċi (62)
and
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

Z 1 Z 1
f¯i− = (1 − η)fa ∆Li dη f¯i+ = ηfa ∆Li dη (63)
0 0
Z 1 Z 1
m̄−
i = (1 − η)ma ∆Li dη m̄+
i = ηma ∆Li dη (64)
0 0

where ∆Li is the length of the ith element; Li is the x1 coordinate of the starting node; and η is a general
coordinate defined as
x1 − Li
η= (65)
∆Li
For a simple system containing a single beam with loads applied at boundaries, we can acuqire following
equations from Eq. (56):
fu−1 − F1∗ = 0 fψ−1 − M1∗ = 0 fF−1 − û1 = 0 −
fM 1
− ĉ1 = 0 (66)
at the starting point,
fu+N − FN∗ +1 = 0 ∗
fψ+N − MN +1 = 0 fF+N + ûN +1 = 0 +
fM N
+ ĉN +1 = 0 (67)
at the ending point. Note F1∗ , M1∗ , FN∗ +1 and MN

+1 are external forces/moments balancing the internal
resultants. At each intermediate point, we have
fu+i + fu−i+1 = 0 fψ+i + fψ−i+1 = 0 fF+i + fF−i+1 = 0 +
fM i

+ fM i+1
=0 (68)
for i = 1, . . . , N − 1. Also for each element, we have
fPi = 0 fHi = 0 (69)
for i = 1, . . . , N .

IV. Examples and Validations


A. Analytical Solution by Neo-Hookean Model
If the strain is still relatively small so that ε3 is negligible, it is possible for us to find analytical solution for
the nonlinear cross-sectional analysis with the 1D strain energy Π corrected to the third order of ε. Focusing
on modeling the combined bending-stretching behavior for isotropic beams, we assume
w1 = κ1 = 0 (70)
Then the components of deformation gradient tensor defined by Eq. (12) can be written as
 
1 + γ + (x3 + w3 )κ2 − (x2 + w2 )κ3 0 0
[Fij ] =  0 1 + w2,2 w3,2  (71)
 

0 w2,3 1 + w3,3
In this paper we follow the methodology of Ref. [48, 49] of which the steps are:

9 of 17

American Institute of Aeronautics and Astronautics


Step (1): Define hyperelastic strain energy density W = W (F );
Step (2): Asymptotic expand Π to the third order with respect to ε: Π = hW i ≈ ΠO(ε2 ) + ΠO(ε3 ) ;
Step (3): Utilize variational statement, δΠO(ε2 ) = 0, to obtain governing equations for wi ;
Step (4): Solve for wi and plug them into Π ≈ ΠO(ε2 ) + ΠO(ε3 ) ;
Step (5): Take derivatives of Π with respect to γ and κi to obtain the constitutive relation.

In the next section, we will use neo-Hookean material to illustrate the methodology introduced here. However
it is not limited to the neo-Hookean material. As shown in Ref. [48, 49] various type of hyperelastic energy
function is compatible with this theory. In addition, the solution reflecting the bending-stretching effect by
using St-Venant/Kirchhoff material has been obtained in Ref. [48].
The neo-Hookean material model is a natural extension of that adopted in the study of small elastic
deformations, is described as
1
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

W N H = c10 (I¯1 − 3) + (J − 1)2 (72)


d1
The superscript N H stands for neo-Hookean type of hyperelastic material. Here J is the Jacobian of
the deformation gradient tensor F , I¯1 = J −2/3 I1 , and I¯2 = J −4/3 I2 , with I1 and I2 as the first two
invariants of right Cauchy-Green deformation tensor C = F T · F or the left Cauchy-Green deformation
tensor B = F · F T since their invariants are the same. The coefficients c01 and d1 can be related to the
equivalent Lamé parameters by
 
µ 1 K 1 2
c10 = = = λ+ µ (73)
2 d1 2 2 3

where K is the bulk modulus, λ and µ the Lamé parameters which can be expressed in terms of Young’s
modulus E and Poisson’s ratio ν as
Eν E
λ= µ= (74)
(1 + ν)(1 − 2ν) 2(1 + ν)

Integrating over the cross-section, we obtain that ΠO(ε2 ) is just the zeroth-order energy with the following
expression,43

2ΠO(ε2 ) = (λ + 2µ) Γ211 + w2,2


2 2
+ w3,3 + 2λ hΓ11 w2,2 + Γ11 w3,3 + w2,2 w3,3 i + µ Γ223 (75)

in which
Γ11 = γ + x3 κ2 − x2 κ3 Γ23 = w3,2 + w2,3 (76)
In addition, ΠO(ε3 ) takes the form of

ΠO(ε3 ) = λ hΥλ i + µ hΥµ i (77)

in which

Υλ = (Γ11 + w2,2 + w3,3 ) [Γ11 (w2,2 + w3,3 ) + w2,2 w3,3 − w2,3 w3,2 + (w3 κ2 − w2 κ3 )] (78)
 
7 2 3
Γ11 + (w2,2 + w3,3 )3

Υµ = (w2,2 + w3,3 ) [Γ11 (Γ11 + w2,2 + w3,3 ) + 3w2,2 w3,3 ] −
9 3
1
− (w2,3 w3,2 )(w2,2 + w3,3 ) − (Γ11 + w2,2 + w3,3 )(w2,3 + w3,2 )2 (79)
 3 
4
+ Γ11 w2,2 w3,3 + 2(w3 κ2 − w3 κ3 )
3

The warping functions that minimize 2ΠO(ε2 ) in Eq. (75) are governed by the Euler-Lagrange equations
of this functional

2 (1 − ν) w2,22 + (1 − 2ν) w2,33 + w3,23 + 2νΓ11,2 = 0 (80)


2 (1 − ν) w3,33 + (1 − 2ν) w3,22 + w2,23 + 2νΓ11,3 = 0 (81)

10 of 17

American Institute of Aeronautics and Astronautics


and the associated boundary conditions
2n2
n3 (w2,3 + w3,2 ) + [ν (Γ11 + w3,3 ) + (1 − ν) w2,2 ] = 0 (82)
1 − 2ν
2n3
n2 (w2,3 + w3,2 ) + [ν (Γ11 + w2,2 ) + (1 − ν) w3,3 ] = 0 (83)
1 − 2ν
where nα is the direction cosine of outward normal with respect to xα .
The solutions of the warping functions are given by Yu and Hodges43 as
   
I23 I2 − I3 ν
w2 = wb2 = −x2 νγ + − x2 x3 νκ2 + x22 − x23 + κ3 (84)
A A 2
   
I23 I2 − I3 ν
w3 = wb3 = −x3 νγ + x2 x3 − νκ3 + x22 − x23 + κ2 (85)
A A 2
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

with
A = h1i I2 = x23 I3 = x22 I23 = hx2 x3 i (86)
According to the variational asymptotic method (VAM),48, 49, 52 we can show that the warping function
we have solved is sufficient to obtain the strain energy asymptotically correct up to the order of ε3 . For
simplicity, we further choose a centroidal coordinate system. The generalized axial force and the bending
moment can be obtained by
∂Π h γ i E
−7 − 14ν + 20ν 2 γ + I2 κ22 + I3 κ23 −7 − 23ν + 20ν 2
 
F1 = = EA 1 + (87)
∂γ 9 9
Eκ22
 
∂Π 2γ
−7 − 23ν + 20ν 2 κ2 + hx3 27x22 ν + x23 −14 − 55ν + 40ν 2 i
  
M2 = = EI2 1 +
∂κ2 9 18
(88)
Eκ2 κ3 Eκ23
hx2 −9x22 ν + x23 14 + 73ν − 40ν 2 i + hx3 9x23 ν + x22 −14 − 73ν + 40ν 2 i
   
+
9 18
2
 
∂Π 2γ Eκ 3
−7 − 23ν + 20ν 2 κ3 + hx2 −27x23 ν + x22 14 + 55ν − 40ν 2 i
  
M3 = = EI3 1 +
∂κ3 9 18
(89)
Eκ2 κ3 Eκ22
hx3 9x23 ν + x22 −14 − 73ν + 40ν 2 i + hx2 −9x22 ν + x23 14 + 73ν − 40ν 2 i
   
+
9 18

B. Analytical Solution versus Numerical Solution


A rubber elastica with a rectangular cross section shown in Figure 3 is under consideration. With such
x3
h = 1cm

x2

h = 1cm

Figure 3. Cross-sectional geometry of rubber elastica. (Length of FE model 50h)

11 of 17

American Institute of Aeronautics and Astronautics


section, Eqs. (87), (88), and (89) reduce to the following concise form
h γ i Eh4 2
F1 = Eh2 1 + −7 − 14ν + 20ν 2 γ + κ2 + κ23 −7 − 23ν + 20ν 2
 
(90)
9  108
Eh4


−7 − 23ν + 20ν 2 κ2

M2 = 1+ (91)
12 9
Eh4
 
2γ 2

M3 = 1+ −7 − 23ν + 20ν κ3 (92)
12 9

The bulk modulus of the rubber is K = 1.9123725 MPa and the shear modulus µ = 0.3825 MPa.
The beam-structural constitutive relations obtained by using numerical VABS and analytical formulas in
Eq. (90)-(92) are plotted in Figure 4. It is noted that VABS does not restrict to the small strain yet the
analytical solution rely on the assumption that strains are small. Figure 4a shows that F1 does not depend
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

on the curvature very much. The analytical solution can provide the accurate solution when the axial strain
is under 0.15. However as shown in Figure 4b, the bending stiffness can be heavily influenced by the axial
strain. And the analytical solution can hardly provide the accurate solution when the axial strain is relatively
large. A safe range of axial strain for the accuracy of the analytical solution is 0.0 to 0.05.

(a) F1 (γ, κ2 ) (b) M2 (γ, κ2 )

Figure 4. Nonlinear beam-structural constitutive relations.

C. Validation with Three-Dimensional Finite Element Analysis


As a control group, three-dimensional finite element (FE) model according to Figure 3 is built in ANSYS
with length L = 50 cm. The bulk modulus K = 1.9123725 MPa and the shear modulus µ = 0.3825 MPa. In
ANSYS the 3D body of the cantilevered elastica is meshed by 20-node quadrilateral elements with elemental
volume size of 0.1 × 0.1 × 0.1cm3 . The schematic of the FEM in ANSYS is shown in Figure 5. One tip of
the elastica is clamped and the other is made rigid by using multiple point constraint element (MPC184).
Bending moment M2 is applied to master node on the free end. An example of large nonlinear bending
deformation of the 3D elastica FEM results from the applied M2 is shown in Figure 6. The deformed axial
length L1 is obtained by summing the distances between neighboring nodes located in the center of the 3D
body. Tip rotation angle θ2 are measured from the master node. Consequently, the axial strain γ and the
curvature can be calculated from these measurements as
L1 θ2
γ= −1 κ2 = (93)
L L1

Nonlinear beam element BEAM189 in ANSYS is also used to solve this problem with the same bulk and
shear modulus to see the difference between the present theory and the traditional elastica model. Note
ANSYS BEAM189 computes the beam stiffness based on the rigid cross-section assumption.

12 of 17

American Institute of Aeronautics and Astronautics


Master Node

Slave Node

Clamped Node
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

Multiple Point Constraint Element


(MPC 184)

Figure 5. Schematic of finite element modeling of the cantilevered rubber elastica.

Figure 6. Contour plot of the displacement UY (U3 ) of ANSYS nonlinear solution for M2 = 0.01 Nm.

13 of 17

American Institute of Aeronautics and Astronautics


On the other hand, substituting the following into Eqs. (87), (88), and (89)

F1 = M3 = κ3 = 0 M2 = M2ANSYS (94)

one can solve for the γ and κ2 by nonlinear system of equations


h γ i Eh4 2
0 = Eh2 1 + −7 − 14ν + 20ν 2 γ + κ2 −7 − 23ν + 20ν 2

9  108 (95)
Eh4

ANSYS 2γ 2

M2 = 1+ −7 − 23ν + 20ν κ2
12 9

to obtained the axial strain and the curvature of the deformed elastica.
This example is also calculated by using Eqs. (90)-(92) and GEBT. First a series of simplifications can
be done to those governing equations based on the problem. All terms related with velocities, momenta and
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

time derivatives can be cancelled. Frame a and b coincide at the undeformed state, so that C ab is an identity
matrix. Only concentrated loads are applied at the boundary of the elastica, then distributed forces and
moments f and m can be dropped. After substituting the constitutive relations Eqs. (90)- (92) into those
equations, we will have a total of 12N + 12 equations to solve for a total of 12N + 12 unknowns: ui , ci , γi
and κi for each element plus 6 unknowns from each boundary. For this particular example, we only apply a
concentrated moment M2 at the free end and to see the response of axial strain γ. Hence, we can further
reduce the number of equations and unknows to 5N + 5: ui,1 , ui,2 , ci,2 , γi,1 and κi,2 for each element plus
F1,1 and M1,2 at the fixed end and uN +1,1 , uN +1,2 and cN +1,2 at the free end.
The γ - M2 curves obtained by ANSYS and Eq. (95) are both plotted in Figure 7. Predictions by present
theory match the result of 3D FEA. However, the traditional nonlinear beam element fails on predicting
the extension-bending coupling behavior of the elastica. This error may not be very critical for the analysis
of a single beam structure.53 Nevertheless, for analyses of cellular structures consisting of huge amount of
micro-beams,53–55 the cumulation of such errors will result in considerable difference in the prediction of the
behavior of the global cellular structures.
0 .0 0 2 0
A N S Y S (3 D , N e o -H o o k e a n M o d e l)
A N S Y S (1 D , R ig id S e c tio n )
P re se n t (A n a ly tic a l V A B S )
0 .0 0 1 5 P re se n t (A n a ly tic a l V A B S + G E B T )

0 .0 0 1 0

0 .0 0 0 5

0 .0 0 0 0
0 .0 0 0 0 .0 0 2 0 .0 0 4 0 .0 0 6 0 .0 0 8 0 .0 1 0 0 .0 1 2
M 2 (N m )

Figure 7. Axial strain γ induced by bending moment M2 .

V. Conclusions
Present theory of the sectional analysis of elastica is found to be more rigorous. One more nonlinear con-
tribution in the formulae is obtained which makes the updated model more accurately predict the extension

14 of 17

American Institute of Aeronautics and Astronautics


results from the applied bending moment. The numerical and the analytical solution of VABS approach
are compared and the analytical model is validated to be applicable when the axial strain is under 0.05.
Numerical example of a 3D rectangular elastica is analyzed by using ANSYS as a control group. The results
of analyzing this example by using different approaches show that with the present theory fully intrinsic
nonlinear behavior of elastica can be captured by a 1D beam model with accuracy comparable to 3D finite
element analysis. Such a theory could be a nice design tool for honeycomb structures, compliant mechanisms,
soft robots, and MEMS, of which the validating applications will be the future works.

Acknowledgements
This research is supported, in part, by the Army Vertical Lift Research Center of Excellence at Georgia
Institute of Technology and its affiliate program through subcontract at Purdue University. The technical
monitor is Dr. Mahendra J Bhagwat. The views and conclusions contained herein are those of the authors
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

and should not be interpreted as necessarily representing the official policies or endorsement, either expressed
or implied, of the funding agency.
The authors are also indebted to Professor Haijun Su in the Mechanical and Aerospace Engineering
Department at the Ohio State University for providing the background knowledge of potential applications
of our theory in design of compliant mechanisms.

References
1 Reissner, E., “On One-Dimensional Finite-Strain Beam Theory: the Plane Problem,” Zeitschrift für angewandte Math-

ematik und Physik ZAMP , Vol. 23, No. 5, 1972, pp. 795–804.
2 Lewis, G. and Monasa, F., “Large Deflections of Cantilever Beams of Non-Linear Materials of the Ludwick Type Subjected

to an End Moment,” International Journal of Non-Linear Mechanics, Vol. 17, No. 1, 1982, pp. 1–6.
3 Simo, J. and Ortiz, M., “A Unified Approach to Finite Deformation Elastoplastic Analysis based on the Use of Hyperelastic

Constitutive Equations,” Computer Methods in Applied Mechanics and Engineering, Vol. 49, No. 2, 1985, pp. 221–245.
4 Simo, J., Taylor, R., and Pister, K., “Variational and Projection Methods for the Volume Constraint in Finite Deformation

Elasto-Plasticity,” Computer Methods in Applied Mechanics and Engineering, Vol. 51, No. 1, 1985, pp. 177–208.
5 Kounadis, A. N. and Mallis, J. G., “Elastica Type Buckling Analysis of Bars from Non-Linearly Elastic Material,”

International Journal of Non-Linear Mechanics, Vol. 22, No. 2, 1987, pp. 99–107.
6 Simo, J., “A Finite Strain Beam Formulation. The Three-Dimensional Dynamic Problem. Part I,” Computer Methods in

Applied Mechanics and Engineering, Vol. 49, No. 1, 1985, pp. 55–70.
7 Simo, J. and Vu-Quoc, L., “A Three-Dimensional Finite-Strain Rod Model. Part II: Computational Aspects,” Computer

Methods in Applied Mechanics and Engineering, Vol. 58, No. 1, 1986, pp. 79–116.
8 Simo, J. C. and Vu-Quoc, L., “On the Dynamics in Space of Rods Undergoing Large Motions—A Geometrically Exact

Approach,” Computer Methods in Applied Mechanics and Engineering, Vol. 66, No. 2, 1988, pp. 125–161.
9 Simo, J. C. and Taylor, R. L., “Quasi-Incompressible Finite Elasticity in Principal Stretches. Continuum Basis and

Numerical Algorithms,” Computer Methods in Applied Mechanics and Engineering, Vol. 85, No. 3, 1991, pp. 273–310.
10 Simo, J. C. and Vu-Quoc, L., “A Geometrically-Exact Rod Model Incorporating Shear and Torsion-Warping Deforma-

tion,” International Journal of Solids and Structures, Vol. 27, No. 3, 1991, pp. 371–393.
11 Lee, K., “Large Deflections of Cantilever Beams of Non-Linear Elastic Material under a Combined Loading,” International

Journal of Non-Linear Mechanics, Vol. 37, No. 3, 2002, pp. 439–443.


12 Attard, M. M., “Finite Strain—Beam Theory,” International Journal of Solids and Structures, Vol. 40, No. 17, 2003,

pp. 4563–4584.
13 Attard, M. M., “Finite Strain—Isotropic Hyperelasticity,” International Journal of Solids and Structures, Vol. 40, No. 17,

2003, pp. 4353–4378.


14 Oden, J. T. and Childs, S. B., “Finite Deflections of a Nonlinearly Elastic Bar,” Journal of Applied Mechanics, Vol. 37,

No. 1, 1970, pp. 48–52.


15 Tada, Y. and Lee, G. C., “Finite Element Solution to An Elastica Problem of Beams,” International Journal for Numerical

Methods in Engineering, Vol. 2, No. 2, 1970, pp. 229–241.


16 Wang, C. Y., “A Critical Review of the Heavy Elastica,” International Journal of Mechanical Sciences, Vol. 28, No. 8,

1986, pp. 549–559.


17 Wang, C. Y., “Post-Buckling of a Slamped-Simply Supported Elastica,” International Journal of Non-Linear Mechanics,

Vol. 32, No. 6, 1997, pp. 1115–1122.


18 Coffin, D. W. and Bloom, F., “Elastica Solution for the Hygrothermal Buckling of a Beam,” International Journal of

Non-Linear Mechanics, Vol. 34, No. 5, 1999, pp. 935–947.


19 Ibrahim, R. A., “Recent Advances in Nonlinear Passive Vibration Isolators,” Journal of Sound and Vibration, Vol. 314,

No. 3, 2008, pp. 371–452.


20 Saggere, L. and Kota, S., “Static Shape Control of Smart Structures Using Compliant Mechanisms,” AIAA journal,

Vol. 37, No. 5, 1999, pp. 572–578.

15 of 17

American Institute of Aeronautics and Astronautics


21 Howell, L. L. and Midha, A., “A Method for the Design of Compliant Mechanisms with Small-Length Flexural Pivots,”

Journal of Mechanical Design, Vol. 116, No. 1, 1994, pp. 280–290.


22 Howell, L. L. and Midha, A., “Parametric Deflection Approximations for End-Loaded, Large-Deflection Beams in Com-

pliant Mechanisms,” Journal of Mechanical Design, Vol. 117, No. 1, 1995, pp. 156–165.
23 Howell, L. L., Midha, A., and Norton, T. W., “Evaluation of Equivalent Spring Stiffness for Use in a Pseudo-Rigid-Body

Model of Large-Deflection Compliant Mechanisms,” Journal of Mechanical Design, Vol. 118, No. 1, 1996, pp. 126–131.
24 Su, H.-J., “A Pseudorigid-Body 3R Model for Determining Large Deflection of Cantilever Beams Subject to Tip Loads,”

Journal of Mechanisms and Robotics, Vol. 1, No. 2, 2009, pp. 021008.


25 Majidi, C., “Remarks on Formulating an Adhesion Problem Using Euler’s Elastica (draft),” Mechanics Research Com-

munications, Vol. 34, No. 1, 2007, pp. 85–90.


26 Majidi, C., “Shear Adhesion between an Elastica and a Rigid Flat Surface,” Mechanics Research Communications,

Vol. 36, No. 3, 2009, pp. 369–372.


27 Majidi, C., O’Reilly, O. M., and Williams, J. A., “On the Stability of a Rod Adhering to a Rigid Surface: Shear-

Induced Stable Adhesion and the Instability of Peeling,” Journal of the Mechanics and Physics of Solids, Vol. 60, No. 5, 2012,
pp. 827–843.
28 Zhou, X., Majidi, C., and O’Reilly, O. M., “Flexing into Motion: A Locomotion Mechanism for Soft Robots,” Interna-
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

tional Journal of Non-Linear Mechanics, Vol. 74, 2015, pp. 7–17.


29 Sadeghian, H., Rezazadeh, G., and Osterberg, P. M., “Application of the Generalized Differential Quadrature Method

to the Study of Pull-in Phenomena of MEMS Switches,” Journal of Microelectromechanical Systems, Vol. 16, No. 6, 2007,
pp. 1334–1340.
30 Xia, W., Wang, L., and Yin, L., “Nonlinear Non-Classical Microscale Beams: Static Bending, Postbuckling and Free

Vibration,” International Journal of Engineering Science, Vol. 48, No. 12, 2010, pp. 2044–2053.
31 Danielson, D. A. and Hodges, D. H., “Nonlinear Beam Kinematics by Decomposition of the Rotation Tensor,” Journal

of Applied Mechanics, Vol. 54, No. 2, 1987, pp. 258–262.


32 Hodges, D. H., “Non-linear Inplane Deformation and Buckling of Rings and High Arches,” International Journal of

Non-linear Mechanics, Vol. 34, No. 4, 1999, pp. 723–737.


33 Hodges, D. H., “A Mixed Variational Formulation based on Exact Intrinsic Equations for Dynamics of Moving Beams,”

International Journal of Solids and Structures, Vol. 26, No. 11, 1990, pp. 1253–1273.
34 Hodges, D. H., Atilgan, A. R., Cesnik, C. E., and Fulton, M. V., “On a Simplified Strain Energy Function for Geometrically

Nonlinear Behaviour of Anisotropic Beams,” Composites Engineering, Vol. 2, No. 5, 1992, pp. 513–526.
35 Yu, W. and Blair, M., “GEBT: A General-Purpose Nonlinear Analysis Tool for Composite Beams,” Composite Structures,

Vol. 94, No. 9, 2012, pp. 2677–2689.


36 Reissner, E., “On One-Dimensional Large-Displacement Finite-Strain Beam Theory,” Studies in Applied Mathematics,

Vol. 52, 1973, pp. 87–95.


37 Jelenić, G. and Crisfield, M. A., “Geometrically Exact 3D Beam Theory: Implementation of A Strain-Invariant Finite

Element for Statics and Dynamics,” Computer Methods in Applied Mechanics and Engineering, Vol. 171, No. 1, 1999, pp. 141–
171.
38 Betsch, P. and Steinmann, P., “Frame-Indifferent Beam Finite Elements Based Upon the Geometrically Exact Beam

Theory,” International Journal for Numerical Methods in Engineering, Vol. 54, No. 12, 2002, pp. 1775–1788.
39 Ibrahimbegović, A., “On Finite Element Implementation of Geometrically Nonlinear Reissner’s Beam Theory: Three-

Dimensional Curved Beam Elements,” Computer Methods in Applied Mechanics and Engineering, Vol. 122, No. 1, 1995,
pp. 11–26.
40 Ibrahimbegović, A. and Mikdad, M. A., “Finite Rotations in Dynamics of Beams and Implicit Time-Stepping Schemes,”

International Journal for Numerical Methods in Engineering, Vol. 41, No. 5, 1998, pp. 781–814.
41 Popescu, B. and Hodges, D. H., “On Asymptotically Correct Timoshenko-like Anisotropic Beam Theory,” International

Journal of Solids and Structures, Vol. 37, No. 3, 2000, pp. 535–558.
42 Popescu, B. and Hodges, D. H., “Asymptotic Treatment of the Trapeze Effect in Finite Element Cross-Sectional Analysis

of Composite Beams,” International Journal of Non-linear Mechanics, Vol. 34, No. 4, 1999, pp. 709–721.
43 Yu, W. and Hodges, D. H., “Elasticity Solutions Versus Asymptotic Sectional Analysis of Homogeneous, Isotropic,

Prismatic Beams,” Journal of Applied Mechanics, Vol. 71, No. 1, 2004, pp. 15–23.
44 Yu, W., Hodges, D. H., and Ho, J. C., “Variational Asymptotic Beam Sectional Analysis—An Updated Version,” Inter-

national Journal of Engineering Science, Vol. 59, 2012, pp. 40–64.


45 Yu, W., Hodges, D. H., Volovoi, V., and Cesnik, C. E., “On Timoshenko-like Modeling of Initially Curved and Twisted

Composite Beams,” International Journal of Solids and Structures, Vol. 39, No. 19, 2002, pp. 5101–5121.
46 Yu, W., Volovoi, V. V., Hodges, D. H., and Hong, X., “Validation of the Variational Asymptotic Beam Sectional Analysis,”

AIAA Journal, Vol. 40, No. 10, 2002, pp. 2105–2112.


47 Hodges, D. H., Harursampath, D., Volovoi, V. V., and Cesnik, C. E., “Non-classical Effects in Non-linear Analysis of

Pretwisted Anisotropic Strips,” International Journal of Non-linear Mechanics, Vol. 34, No. 2, 1999, pp. 259–277.
48 Jiang, F. and Yu, W., “Non-linear Sectional Analysis of Composite Beams with Finite Deformation and Hyperelastic

Materials,” 56th AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, AIAA SciTech, January
2015.
49 Jiang, F., Yu, W., and Hodges, D. H., “Analytical Modeling of Trapeze and Poynting Effects of Initially Twisted Beams,”

Journal of Applied Mechanics, Vol. 82, No. 6, 2015, pp. 061003.


50 Hodges, D. H., “Nonlinear Composite Beam Theory,” Progress in Astronautics and Aeronautics, Vol. 213, 2006, pp. 304.
51 Wang, Q., Yu, W., and Sprague, M. A., “Geometric Nonlinear Analysis of Composite Beams Using Wiener-Milenković

Parameters,” 54th AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, AIAA SciTech, April
2013.

16 of 17

American Institute of Aeronautics and Astronautics


52 Berdichevsky, V. L., Variational Principles of Continuum Mechanics, Springer-Verlag, LagVergne, TN, USA, 2009.
53 Haghpanah, B., Papadopoulos, J., Mousanezhad, D., Nayeb-Hashemi, H., and Vaziri, A., “Buckling of Regular, Chiral
and Hierarchical Honeycombs under a General Macroscopic Stress State,” Proceedings of the Royal Society of London A:
Mathematical, Physical and Engineering Sciences, Vol. 470, No. 2167, 2014, pp. 20130856.
54 Mousanezhad, D., Ghosh, R., Ajdari, A., Hamouda, A., Nayeb-Hashemi, H., and Vaziri, A., “Impact Resistance and

Energy Absorption of Regular and Functionally Graded Hexagonal Honeycombs with Cell Wall Material Strain Hardening,”
International Journal of Mechanical Sciences, Vol. 89, 2014, pp. 413–422.
55 Mousanezhad, D., Babaee, S., Ghosh, R., Mahdi, E., Bertoldi, K., and Vaziri, A., “Honeycomb Phononic Crystals with

Self-Similar Hierarchy,” Physical Review B , Vol. 92, No. 10, 2015, pp. 104304.
Downloaded by KUNGLIGA TEKNISKA HOGSKOLEN KTH on January 20, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-1242

17 of 17

American Institute of Aeronautics and Astronautics

You might also like