Sei sulla pagina 1di 23

10

The TL Plane
Beam Element:
Formulation

10–1
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–2

TABLE OF CONTENTS

Page
§10.1. Introduction 10–3
§10.2. Beam Models 10–3
§10.2.1. Basic Concepts and Terminology . . . . . . . . . . . 10–3
§10.2.2. Beam Mathematical Models . . . . . . . . . . . . 10–4
§10.2.3. Finite Element Models . . . . . . . . . . . . . . 10–5
§10.2.4. Shear Locking . . . . . . . . . . . . . . . . 10–7
§10.3. X -Aligned Reference Configuration 10–8
§10.3.1. Element Description . . . . . . . . . . . . . . . 10–8
§10.3.2. Motion . . . . . . . . . . . . . . . . . . . 10–9
§10.3.3. Displacement Interpolation . . . . . . . . . . . . . 10–10
§10.3.4. Strain-Displacement Relations . . . . . . . . . . . 10–11
§10.3.5. *Consistent Linearization . . . . . . . . . . . . . 10–12
§10.3.6. *Rigid Body Motion Check . . . . . . . . . . . . 10–12
§10.4. Arbitrary Reference Configuration 10–13
§10.4.1. Strain-Displacement Matrix . . . . . . . . . . . . 10–14
§10.4.2. Constitutive Equations . . . . . . . . . . . . . . 10–15
§10.5. Strain Energy 10–15
§10.6. Internal Force Vector 10–16
§10.7. Tangent Stiffness Matrix 10–16
§10.7.1. Material Stiffness Matrix . . . . . . . . . . . . . 10–16
§10.7.2. Eliminating Shear Locking by RBF . . . . . . . . . 10–17
§10.7.3. Geometric Stiffness Matrix . . . . . . . . . . . . . 10–19
§10.8. Commentary on Element Performance 10–22
§10.9. Derivation Summary 10–22
§10. Notes and Bibliography
. . . . . . . . . . . . . . . . . . . . . . 10–22
§10. Exercises . . . . . . . . . . . . . . . . . . . . . . 10–23

10–2
10–3 §10.2 BEAM MODELS

§10.1. Introduction
In the present Chapter the standard formulation of Total Lagrangian (TL) kinematics is used to
derive the finite element equations of a two-node Timoshenko plane beam element. This derivation
is more typical of the general case. It is still short, however, of the enormous complexity involved,
for instance, in the FEM analysis of nonlinear three-dimensional beams or shells. In fact the latter
are still doctoral thesis topics.
In the formulation of the bar element in Chapter 8, advantage was taken of the direct expression of
the axial strain in terms of reference and current element lengths. That shortcut bypasses the use
of displacement gradients, and allows the reference configuration to be arbitrarily oriented. The
simplification works equally well for straight bars in three-dimensional space.
A more systematic but lengthier procedure is unavoidable with more complicated elements. The
procedure requires going through the displacement gradients to construct a strain measure. Some-
times this measure is too complex and must be simplified while retaining physical correctness. Then
work-conjugate stresses are introduced and paired with strains to form the strain energy function
of the element. Repeated differentiations with respect to node displacements yield the expressions
of the internal force vector and tangent stiffness matrix. Finally, a transformation to the global
coordinate system may be required in some cases.
In addition to giving a better picture of the general procedure just outlined, the beam element
illustrates the treatment of rotational degrees of freedom.

§10.2. Beam Models

§10.2.1. Basic Concepts and Terminology


Beams represent the most common structural component found in civil and mechanical structures.
Because of their ubiquity they are extensively studied, from an analytical viewpoint, in Mechanics
of Materials courses. Such a basic knowledge is assumed here. The following material recapitulates
definitions and concepts that are needed in the finite element formulation.
A beam is a rod-like structural member that can resist transverse loading applied between its
supports. By “rod-like” it is meant that one of the dimensions is considerably larger than the other
two. This one is called the longitudinal dimension and defines the longitudinal direction or axial
direction. Directions normal to the longitudinal direction are called transverse. The intersection
of planes normal to the longitudinal direction with the beam are called cross sections, just as for
bar elements. The beam longitudinal axis is directed along the longitudinal direction and passes
through the centroid of the cross sections.1 .
Beams may be used as isolated structures. But they can also be combined to form framework
structures. This is actually the most common form of high-rise building construction. Individual
beam components of a framework are called members, which are connected at joints. Frameworks
can be distinguished from trusses by the fact that their joints are sufficiently rigid to transmit bending
moments between members.
1 If the beam is built of several materials, as in the case of reinforced concrete, the longitudinal axis passes through the
centroid of a modified cross section. The modified-area technique is explained in elementary courses of Mechanics of
Materials

10–3
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–4

reference configuration

motion

current configuration

Figure 10.1. A geometrically nonlinear plane framework structure.

In practical structures beam members can take up a great variety of loads, including biaxial bending,
transverse shears, axial forces and even torsion. Such complicated actions are typical of spatial
beams, which are used in three-dimensional frameworks and are subject to forces applied along
arbitrary directions.
A plane beam resists primarily loading applied in one plane and has a cross section that is symmetric
with respect to that plane. Plane frameworks, such as the one illustrated in Figure 10.1, are
assemblies of plane beams that share that symmetry. Those structures can be analyzed with two-
dimensional idealizations.
A beam is straight if the longitudinal direction is a straight line. A beam is prismatic if its cross
section is uniform. Only straight, prismatic, plane beams will be considered in this Chapter.

§10.2.2. Beam Mathematical Models


Beams are actually three-dimensional solids. One-dimensional mathematical models of plane
beams are constructed on the basis of beam theories. All such theories involve some form of
approximation that describes the behavior of the cross sections in terms of quantities evaluated on
the longitudinal axis. More precisely, the element kinematics of a plane beam is completely defined
by the two models described below if the following functions are given:
• Axial displacement u X (X )
• Transverse displacement u Y (X ) (also called lateral displacement)
• Cross section rotation θ Z (X ) ≡ θ(X ): angle by which cross section rotates.2
Here X denotes the longitudinal coordinate in the reference configuration. See Figure 10.2. In the
sequel we will drop the subscript Z from θ for brevity.
Two beam mathematical models are commonly used in structural mechanics:
Bernoulli-Euler (BE) Model. This model is also called classical beam theory or engineering beam
theory, and is the one covered in elementary treatments of Mechanics of Materials. It accounts for
bending moment effects on stresses and deformations. Transverse shear forces are recovered from
equilibrium but their effect on beam deformations is ignored (more precisely: the strain energy
due to shear stresses is neglected). The fundamental kinematic assumption is that cross sections

2 This angle suffices if the cross section remains plane, an assumption verified by both theories described here.

10–4
10–5 §10.2 BEAM MODELS

θZ (X) ≡ θ(X)

Current
current configuration cross section

Y, y
motion uY (X)
X, x

uX (X)

reference configuration Reference


X cross section
Figure 10.2. Definition of beam kinematics in terms of the three displacement functions: u X (X ),
u Y (X ), and θ (X ). The figure actually depicts the BE model kinematics. In the Timoshenko model, θ (X )
is not constrained by normality, as illustrated in Figure 10.3.

current configuration finite element idealization


of current configuration

motion

reference configuration finite element idealization


of reference configuration
Figure 10.3. Idealization of a geometrically nonlinear beam member (as taken, for example, from a
plane framework structure like the one in Figure 10.1) as an assembly of finite elements.

remain plane and normal to the deformed longitudinal axis. This rotation occurs about a neutral
axis parallel to Z that passes through the centroid of the cross section.
Timoshenko Model. This model corrects the BE theory with first-order shear deformation effects.
The key assumption is that cross sections remain plane and rotate about the same neutral axis, but
do not remain normal to the deformed longitudinal axis. The deviation from normality is produced
by a transverse shear stress that is assumed to be constant over the cross section.
Both the BE and Timoshenko models rest on the assumptions of small deformations and linear-
elastic isotropic material behavior. In addition both models neglect any change in dimensions of the
cross sections as the beam deforms. Either theory can account for geometrically nonlinear behavior
due to large displacements and rotations, as long as the other assumptions hold.

10–5
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–6

(a) C 1 (BE) model θ2 (b) C 0 (Timoshenko) model θ2

θ1 θ1

Y, y uY2 uY2
uY1 uY1
X, x
uX1 uX2 uX1 uX2
1 2 1 2
Figure 10.4. Two-node beam elements have six DOF, regardless of which model is used.

§10.2.3. Finite Element Models


To carry out the geometrically nonlinear finite element analysis of a framework structure, beam
members are idealized as an assembly of finite elements, as illustrated in Figure 10.3. Beam
elements used in practice have usually two end nodes. The i th node has three DOF: two node
displacements u Xi and u Y i , and one nodal rotation θi , positive counterclockwise in radians, about
the Z axis. See Figure 10.4.
The cross section rotation from the reference to the current configuration is called θ in both models.
In the BE model this is the same as the rotation ψ of the longitudinal axis. In the Timoshenko
model, the shear distortion angle is γ = ψ − θ − psi, as shown in Figure 10.5. The mean shear
strain has the opposite sign: γ̄ = −γ = ψ − θ, so as to make the shear strain e X Y positive, as per
the usual conventions of structural mechanics.
Either the BE or the Timoshenko model may be used as the basis for the TL beam element for-
mulation. Superficially it appears that one should select the latter only when shear effects are to
be considered, as in “deep beams” whereas the BE model is used for ordinary beams. But here a
“twist” appear because of finite element considerations. This twist is one that has caused significant
confusion among FEM users over the past 25 years. See Appendix S for the tragicomical story of
the topic.
Although the Timoshenko beam model appears to be more complex because of the inclusion of
shear deformation, finite elements based on this model are in fact simpler to construct! Here are
the two key reasons:
(i) Separate kinematic assumptions on the variation of cross-section rotations are possible, as
made evident by Figure 10.5. Mathematically: θ(X ) may be assumed independently of u X (X )
and u Y (X ). As a consequence, two-node Timoshenko elements may use linear variations in
both displacement and rotations. On the other hand, a two-node BE model requires a cubic
polynomial for u Y (X ) because the rotation θ(X ) is not independent.
(ii) The linear transverse displacement variation matches that commonly assumed for the axial
deformation (bar-like behavior). The transverse and axial displacement assumptions are then
said to be consistent.

10–6
10–7 §10.2 BEAM MODELS

Angles are positive normal to reference


θ beam axis X
as shown. Note
that _γ = θ − ψ _
γ ψ
whereas γ = −γ = ψ − θ normal to deformed
γ beam axis
direction of
deformed
ψ
cross section _
90
// X (X = X)

ds

Figure 10.5. Illustrates total and BE section rotations θ and ψ, respectively, in the plane beam Timoshenko
model. The mean shear distortion angle is γ = θ − ψ, but we take γ̄ = −γ = θ − ψ as shear strain measure, to
match usual sign conventions of structural mechanics. For elastic deformations of engineering materials |γ̄ | << 1.
Typical values for |γ̄ | would be O(10−3 ) radians whereas rotations ψ and θ may be much larger, say 1-2 radians.
The magnitude of γ̄ is grossly exaggerated in the Figure for visualization convenience.

This simplicity is even more important in geometrically nonlinear analysis, as strikingly illustrated
by the elements contrasted in Figure 10.6. Although as pictured there both elements have six
degrees of freedom, the internal kinematics of the Timoshenko model is far simpler.

§10.2.4. Shear Locking


In the FEM literature, a BE-based model such as the one shown in Figure 10.4(a) is called a C 1
beam because this is the kind of mathematical continuity achieved in the longitudinal direction
when a beam member is divided into several elements (cf. Figure 10.3). On the other hand, the
Timoshenko-based element pictured in Figure 10.4(b) is called a C 0 beam because both transverse
displacements, as well as the rotation angle θ, preserve only C 0 continuity.
What would be the first reaction of an experienced but old-fashioned (i.e, “never heard about
FEM”) structural engineer on looking at Figure 10.6? The engineer would pronounce the C 0
element unsuitable for practical use. And indeed the kinematics looks way wrong. The shear
distortion grossly violates the basic assumptions of beam behavior. And indeed, a huge amount of
shear energy would be require to keep the element straight as depicted.
The engineer would be both right and wrong. If the two-node element of Figure 10.6(b) were
constructed with actual shear properties and exact integration, an overstiff model results. This
phenomenom is well known in the FEM literature and receives the name of shear locking. To avoid
locking while retaining the element simplicity it is necessary to use certain computational devices
that have nothing to do with physics. The most common are:
1. Selective integration for the shear energy.
2. Residual energy balancing.
The second device will be used with only a cursory explanation at the end of this Chapter. For
explanatory literature references see Notes and Bibliography and Appendix S. In this Chapter the
C 0 model will be used to illustrated the TL formulation of a two-node, geometrically nonlinear
beam element.

10–7
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–8

2-node C 1 (cubic) element 2-node C 0 linear displacement & rotations


for Euler-Bernoulli beam model: element for Timoshenko beam model:
plane sections remain plane and plane sections remain plane but not
normal to deformed longitudinal axis normal to deformed longitudinal axis

(a) (b)
C 1 element
with same DOFs

1 2 1 2

Figure 10.6. Contrasting kinematics of 2-node beam FEM models based on (a) BE beam theory, and (b)
Timoshenko beam theory. These are called C 1 and C 0 beam elements, respectively, in the FEM literature.

Remark 10.1. As a result of the application of the aforementioned devices the beam element behaves like a BE
beam although the underlying model is Timoshenko’s. This represents a curious paradox: shear deformation
is used to simplify the kinematics, but then most of the shear will be removed to restore the correct physical
behavior.3 As a result, the name “C 0 element” is more appropriate than “Timoshenko element” because
capturing the actual shear deformation is not the main objective.

Remark 10.2. The two-node C 1 beam element is used primarily in linear structural mechanics. It is in fact
the beam model used in the IFEM course [220, Chapter 12].) This is because some of the easier-construction
advantages cited for the C 0 element are less noticeable, while no artificial devices to eliminate locking are
needed. The C 1 element is also called the Hermitian beam element because the shape functions are cubic
polynomials specified by Hermite interpolation formulas.

§10.3. X -Aligned Reference Configuration

§10.3.1. Element Description


We consider a two-node, straight, prismatic C 0 plane beam element moving in the (X, Y ) plane, as
depicted in Figure 10.7(a). For simplicity in the following derivation the X axis system is initially
aligned with the longitudinal direction in the reference configuration, with origin at node 1. This
assumption is relaxed in the next section, once invariant strain measured are obtained.
The reference element length is L 0 . The cross section area A0 and second moment of inertia I0
with respect to the neutral axis4 are defined by the area integrals
  
A0 = d A, Y d A = 0, I0 = Y 2 d A, (10.1)
A0 A0 A0

3 The FEM analysis of plates and shells is also rife with such “two wrongs make a right” paradoxes.
4 For a plane prismatic beam, the neutral axis at a particular section is the intersection of the cross section plane X =
constant with the plane Y = 0.

10–8
10–9 §10.3 X -ALIGNED REFERENCE CONFIGURATION

(a) (b)
L _
Y, y θ(X)
θ=ψ−γ 2
ψ 2
x P(x,y)
xC C(xC ,yC ) C C(X+uX ,uY )
y
1 uYC 1
yC uYP
uY (X) = uYC
X uXC uX (X)=uXC
X
P0 (X,Y) uXP
Y X, x C0(X)
1 C0(X,0) 2 1 2
C0
L0

Figure 10.7. Lagrangian kinematics of C 0 beam element with X -aligned reference configuration: (a)
plane beam moving as a 2D body; (b) reduction of motion description to 1D as measured by coordinate X .

In the current configuration those quantities become A, I and L, respectively, but only L is frequently
used in the TL formulation. The material remains linearly elastic with elastic modulus E relating
the stress and strain measures defined below.
As in the previous Chapter the element identification superscript e will be omitted to reduce clutter
until it is necessary to distinguish elements within structural assemblies.
The element has the six degrees of freedom depicted in Figure 10.4. These degrees of freedom and
the associated node forces are collected in the node displacement and node force vectors
   
u X1 f X1
 uY 1   fY 1 
   
 θ1   fθ 1 
u= , f= . (10.2)
 u X2   f X2 
   
uY 2 fY 2
θ2 fθ 2
The external loads applied to the nodes will be assumed to be conservative.

§10.3.2. Motion
The kinematic assumptions of the Timoshenko model element have been outlined in §10.2.2.
Basically they state that cross sections remain plane upon deformation, but not necessarily normal
to the deformed longitudinal axis. In addition, changes in cross section geometry are neglected.
To analyze the Lagrangian kinematics of the element shown in Figure 10.6(a), we study the motion
of a particle originally located at P0 (X, Y ) in the reference configuration. The particle moves to
P(x, y) in the current configuration. The projections of P0 and P on the neutral axis, along the
cross sections at C 0 and C are called C0 (X, 0) and C(xC , yC ), respectively. The exact kinematics is


x xC − Y (sin ψ+ sin γ cos ψ) xC − Y [sin θ + (1− cos γ ) sin ψ]


= = . (10.3)
y yC + Y (cos ψ− sin γ sin ψ) yC + Y [cos θ + (1− cos γ ) cos ψ]

10–9
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–10

in which ψ + γ has been replaced by θ . But xC = X + u XC and yC = u Y C . From now we denote


u XC and u Y C simply as u X and u Y , respectively. Assumed for simplicity that γ is constant over the
element. Expand (10.3) in Taylor series in γ up to O(γ 4 ):

x u X + X + 12 Y γ 3 cos θ − Y (1 + 12 γ 2 + 24 7
γ 4 ) sin θ
= (10.4)
y u Y + Y (1 + 12 γ 2 − 24
7
γ 4 ) cos θ + 12 Y γ 3 sin θ
If we assume the shear distortion is very small, setting γ → 0 in (10.4) gives the simplified
Lagrangian description of the motion

x X + u X − Y sin θ
= , (10.5)
y u Y + Y cos θ

in which u X , u Y and θ are functions of X only. This form can be also directly obtained from the last
of (10.3) by replacing 1− cos γ by 0. This concludes the reduction to a one-dimensional model, as
sketched in Figure 10.7(b).
For future use it is convenient to define an “extended” internal displacement vector w, and its X
derivative:

u X (X ) dw du X /d X uX

w = u Y (X ) , w = = du Y /d X = u
Y , (10.6)
θ(X ) dX dθ/d X θ

in which primes denote derivatives with respect to X . The derivative θ


will be also denoted as κ,
which has the meaning of beam curvature in the current configuration. Also useful are the following
differential relations
ds




1 + u X = s cos ψ, u Y = s sin ψ, s = = (1 + u
X )2 + (u
Y )2 , (10.7)
dX
in which ds is the differential arclength in the current configuration; see Figure 10.5. If u
X and u
Y
are constant over the element,
s
= L/L 0 , 1 + u
X = L cos ψ/L 0 , u
Y = L sin ψ/L 0 . (10.8)

Remark 10.3. Replacing 1 − cos γ by zero in (10.3) is equivalent to saying that Y (1 − cos γ ) can be neglected
in comparison to other cross section dimensions. This is consistent with uncertainties in cross section changes.
Such changes would depend on the normal stress as well as Poisson’s ratio effects. The motion (10.5) has the
virtue of being purely kinematic (i.e., no material properties appear) and of leading to (exactly) eY Y = 0.

§10.3.3. Displacement Interpolation


For a 2-node C 0 element it is natural to express the displacements and rotation functions as linear
in the node displacements:
 
u X1
 uY 1 
u X (X ) 1−ξ 0 0 1+ξ 0 0  
 θ1 
w = u Y (X ) = 12 0 1−ξ 0 0 1+ξ 0   = N u, (10.9)
 u X2 
θ(X ) 0 0 1−ξ 0 0 1+ξ  
uY 2
θ2

10–10
10–11 §10.3 X -ALIGNED REFERENCE CONFIGURATION

in which ξ = (2X/L 0 ) − 1 is the isoparametric coordinate that varies from ξ = −1 at node 1 to


ξ = 1 at node 2, and N as usual denotes the shape function matrix. Differentiating this expression
with respect to X yields the displacement gradient interpolation:
 
u X1
u 
u
X −1 0 0 1 0 0  Y1 
1  θ 

w = u
Y = 0 −1 0 0 1 0  1  = N
u. (10.10)
L0 u 
θ
0 0 −1 0 0 1  X2 
uY 2
θ2

§10.3.4. Strain-Displacement Relations


The deformation gradient matrix of the motion (10.5) is
∂x ∂x

1 + u
X − Y κ cos θ − sin θ
F = ∂ X ∂Y = , (10.11)
∂y ∂y u
Y − Y κ sin θ cos θ
∂ X ∂Y
in which primes denote derivatives with respect to X , and κ = θ
is the curvature. The displacement
gradient matrix is

u X − Y κ cos θ − sin θ
G=F−I= , (10.12)
u
Y − Y κ sin θ cos θ − 1
From the definition (7.16), the plane portion of the Green-Lagrange (GL) strain tensor follows as

eX X eX Y
e= = 12 (FT F − I) = 12 (G + GT ) + 12 GT G. (10.13)
eY X eY Y

This expands to the component form




1 2(u X −Y κ cos θ) + (u X −Y κ cos θ) + (u Y −Y κ sin θ)


2 2
−(1 + u
X ) sin θ + u
Y cos θ
e= 2

−(1 + u X ) sin θ + u Y cos θ 0


(10.14)
It is seen that the only nonzero strains are the axial strain e X X and the shear strain e X Y +eY X = 2e X Y ,
whereas eY Y vanishes. Through the consistent-linearization techniques described in §10.3.5 below,
it can be shown that under the small-strain assumptions made precise therein, the axial strain e X X
can be replaced by the simpler form

e X X = (1 + u
X ) cos θ + u
Y sin θ − Y κ − 1, (10.15)

in which all quantities appear linearly except θ . The nonzero axial and shear strains will be arranged
in the strain vector



e1 eX X (1 + u
X ) cos θ + u
Y sin θ − Y θ
− 1 e − Yκ
e= = = = .
e2 2e X Y −(1 + u
X ) sin θ + u
Y cos θ γ
(10.16)

10–11
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–12

The three strain quantities introduced in (10.16)

e = (1 + u
X ) cos θ + u
Y sin θ − 1, γ = −(1 + u
X ) sin θ + u
Y cos θ, κ = θ
,
(10.17)
characterize axial strains, shear strains and curvatures, respectively. These are collected in the
following generalized strain vector:
e
h= γ . (10.18)
κ
Because of the assumed linear variation in X of u X (X ), u Y (X ) and θ(X ), e and γ only depend on
θ whereas κ is constant over the element. Making use of (10.8) one can express e and γ in the
geometrically invariant form

L cos γ̄ L sin γ̄
1 + e = s
cos(θ − ψ) = , γ = −s
sin(θ − ψ) = (10.19)
L0 L0
In theory one could further reduce e to L/L 0 and γ to L γ̄ /L 0 , but those “simplifications” would
actually complicate the strain variations taken in the following Section.
§10.3.5. *Consistent Linearization
The derivation of the consistent linearization (10.15) s based on the following study, known in continuum
mechanics as a polar decomposition analysis of the deformation gradient. Introduce the matrix
 
cos α − sin α
Ω(α) = , (10.20)
sin α cos α
which represents a two-dimensional rotation (about Z ) through an angle α. Since Ω is an orthogonal matrix,
ΩT = Ω−1 . The deformation gradient (10.11) can be written
   
s
0 −Y θ
0
F = Ω(ψ) + Ω(θ) . (10.21)
0 0 0 1
where s
is defined in (10.7) Premultiplying both sides of (10.21) by Ω(−θ ) gives the modified deformation
gradient 

s cos(θ − ψ) − Y θ
0
F̄ = Ω(−θ )F = (10.22)
−s
sin(θ − ψ) 1
Now the GL strain tensor 2e = FT F − I does not change if F is premultiplied by the orthogonal matrix Ω
T
because FT ΩT ΩF = FT F. Consequently 2e = F̄ F̄ − I. But if the strains remain small, as it is assumed in
the Timoshenko model, the following are small quantities:
(i) s
− 1 = (L/L 0 ) − 1 because the axial strains are small;
(ii) Y θ
= Y κ because the curvature κ is of order 1/R, R being the radius of curvature, and |R| >> |Y |
according to beam theory because Y can vary only up to the cross section in-plane dimension;
(iii) γ̄ = ψ − θ, which is the mean angular shear deformation.
Then F̄ = I + L + higher order terms, where L is a first-order linearization in the small quantities s
− 1, Y θ

and γ̄ = ψ − θ . It follows that


e = 12 (L + LT ) + higher order terms (10.23)
Carrying out this linearization one finds that e X Y and eY Y do not change, but that e X X simplifies to (10.15). It
.
can also be shown that 2e X Y = γ = γ̄ within the order of approximation of (10.23).

10–12
10–13 §10.4 ARBITRARY REFERENCE CONFIGURATION

§10.3.6. *Rigid Body Motion Check


It is instructive to check whether the foregoing displacements and strain fields can accuractely represent rogid
body motions (RBM). An arbitrary RBD of the element can be represented by a rigid translation {u X R , u Y R },
along X and Y , resectively, and a rotation of angle θ R about the translated position. Denoting the new
coordinates by {x R , y R } the motion is given by
    
xR cos θ R sin θ R X + uXR
= . (10.24)
yR − sin θ R cos θ R Y + uY R
The displacements are u X = x R − X R and u Y = y R − Y . The X derivatives are

u
X = cos θ R − 1, u
Y = − sin θ, θ R
= κ R = 0. (10.25)

This motion is inserted into the followiung strain measures

e(1)

X X = (u X −Y κ cos θ ) + 2 (u X −Y κ cos θ ) + 2 (u Y −Y κ sin θ ) ,


1 2 1 2

e(2)

X X = (1 + u X ) cos θ + u Y sin θ − Y κ − 1,
(10.26)
e(1)

X Y = −(1 + u X ) sin θ + u Y cos θ

e(2) (1)

X Y = e X Y − 2 cos θ (u Y − βY κ sin θ ).

in which β is arbitrary. In (10.27) e(1) (1) (2)


X X and e X Y are those in the GL strain tensor (10.13). Normal strain e X X is
(2)
the result of the linearization (10.15), while the shear strain e X X includes a correction not derivable from the
GL formulas. Inserting the RBM described by (10.24) and (10.25) into the shear and normal strain measures
(10.27) one obtains

e(1)
X X = 0, e(2)
X X = 2 sin θ R ,
2
e(1)
X Y = − 2 sin(2θ R ),
1
e(2)
X Y = 0. (10.27)

Conclusion: the strain measures e = e(2) (1)


X X and γ = e X Y used in the strain vector (10.18) do not capture exactly
RBM as the rotation angle θ R gets large. For that to be achieved, it would be necessary to pick e = e(1) X X and
(2)
γ = e X Y . There is no displacement field, however, that produces both of these as components of the GL strain
tensor. See also Homework 10.1.

§10.4. Arbitrary Reference Configuration


In the general case the reference configuration C 0 of the element is not aligned with X . The
longitudinal axis X̄ forms an angle ϕ with X , as illustrated in Figure 10.8. The six degrees of
freedom of the element are indicated in that Figure. Note that the section rotation angle θ is
measured from the direction Ȳ , normal to X̄ , and no longer from Y as in Figure 10.6.
Given the node coordinates (X 1 , Y1 ) and (X 2 , Y2 ), the reference angle ϕ is determined by

cos ϕ = X 21 /L 0 , sin ϕ = Y21 /L 0 , X 21 = X 2 − X 1 , Y21 = Y2 − Y1 , L 20 = X 21


2
+ Y21
2
. (10.28)

The angle φ = ψ +ϕ formed by the current longitudinal axis with X (see Figure 10.5) is determined
by
x21 y21
cos φ = cos(ψ + ϕ) = sin φ = sin(ψ + ϕ) = , (10.29)
L L
in which x21 = x2 − x1 , y21 = y2 − y1 and L may be obtained from the motion as

x21 = X 21 + u X 2 − u X 1 , y21 = Y21 + u Y 2 − u Y 1 , L 2 = x21


2
+ y21
2
. (10.30)

10–13
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–14

_
// γY φ = ψ+ϕ_
θ2 ψ
// γX
ϕ
2(x 2 ,y2) // γX

_
// γY C
uY2
θ1
_
1(x1 ,y1) Xγ

uX2 ϕ
Y, y u Y1 // γX
_ 2(X2,Y2)

X, x C0
uX1
1(X1,Y1)

Figure 10.8. Plane beam element with arbitrarily oriented reference configuration.

Solving the foregoing trigonometric relations for ψ gives


X 21 x21 + Y21 y21 X 21 (X 21 + u X 2 − u X 1 ) + Y21 (Y21 + u Y 2 − u Y 1 )
cos ψ = = ,
L L0 L L0
(10.31)
X 21 y21 − Y21 x21 X 21 (Y21 + u Y 2 − u Y 1 ) − Y21 (X 21 + u X 2 − u X 1 )
sin ψ = = .
L L0 L L0
It follows that L sin ψ and L cos ψ are exactly linear in the translational node displacements. This
property simplifies considerably the calculations that follow.

§10.4.1. Strain-Displacement Matrix


For the generalized strains it is convenient to use the invariant form (10.24), which does not depend
on ϕ. The variations δe, δγ and δκ with respect to nodal displacement variations are required in the
formation of the strain displacement relation δh = B δu. To form B we take partial derivatives of e,
γ and κ with respect to node displacements. Here is a sample of the kind of calculations involved:
∂e ∂[L cos(θ − ψ)/L 0 − 1] ∂[L(cos θ cos ψ + sin θ sin ψ)/L 0 − 1]
= =
∂u X 1 ∂u X 1 ∂u X 1
(10.32)
−X 21 cos θ + Y21 sin θ − cos ϕ cos θ + sin ϕ sin θ cos ω
= 2
= =− ,
L0 L0 L0
in which ω = θ + ϕ. Use was made of (10.31) in a key step. These derivatives were checked with
Mathematica. Collecting all of them into matrix B:

1 − cos ω − sin ω L 0 N1 γ cos ω sin ω L 0 N2 γ
B= sin ω − cos ω −L 0 N1 (1 + e) − sin ω cos ω −L 0 N2 (1 + e) . (10.33)
L0 0 0 −1 0 0 1

10–14
10–15 §10.5 STRAIN ENERGY

N
M
V

M0
N0
V0
C0

Figure 10.9. Beam stress resultants (internal forces) in the reference and current configurations.

Here N1 = (1 − ξ )/2 and N2 = (1 + ξ )/2 are abbreviations for the element shape functions
(caligraphic symbols are used to lessen the chance of clash against axial force symbols).

§10.4.2. Constitutive Equations


Because the beam material is assumed to be homogeneous and isotropic, the only nonzero PK2
stresses are the axial stress s X X and the shear stress s X Y . These are collected in a stress vector s
related to the GL strains by the linear elastic relations


0
0

sX X s1 s1 + Ee1 s1 E 0 e1
s= = = 0 = 0 + = s0 + Ee, (10.34)
sX Y s2 s2 + Ge2 s2 0 G e2

in which E is the modulus of elasticity and G is the shear modulus. We introduce the prestress
resultants   
N =
0 0
s1 d A, V =
0 0
s2 d A, M =
0
−Y s10 d A. (10.35)
A0 A0 A0

These define the axial forces, transverse shear forces and bending moments, respectively, in the
reference configuration. We also define the stress resultants

N = N 0 + E A0 e, V = V 0 + G A0 γ , M = M 0 + E I0 κ. (10.36)

These represent axial forces, tranverse shear forces and bending moments in the current configura-
tion, respectively, defined in terms of PK2 stresses. See Figure 10.9 for signs. These are collected
in the stress-resultant vector
z = [ N V M ]T . (10.37)

10–15
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–16

§10.5. Strain Energy

As in the case of the TL bar element, the total potential energy  = U − P is separable because
P = λqT u. The strain (internal) energy is given by
  
 0 T   0 
U= (s ) e + 2 e Ee d V =
1 T
(s1 e1 + s20 e2 ) + 12 (Ee12 + Ge22 ) d A d X̄ . (10.38)
V0 A0 L0

Carrying out the area integrals while making use of (10.34) through (10.37), U can be written as
the sum of three length integrals:
  
U= (N e +
0 1
2
E A0 e2 ) d X̄ + (V γ +
0 1
2
G A0 γ 2 ) d X̄ + (M 0 κ + 12 E I0 κ 2 ) d X̄ ,
L0 L0 L0
(10.39)
The three terms in (10.39) define the energy stored through bar-like axial deformations, shear
distortion and pure bending, respectively.

§10.6. Internal Force Vector

The internal force vector can be obtained by taking the first variation of the internal energy with
respect to the node displacements. This can be compactly expressed as
  
 
δU = N δe + V δγ + M δκ d X̄ = z δh d X̄ =
T
zT B d X̄ δu. (10.40)
L0 L0 L0

Here h and B are defined in equations (10.23) and (10.33) whereas z collects the stress resultants
in C as defined in (10.35) through (10.37). Because δU = p T δu, we get

p= BT z d X̄ . (10.41)
L0

This expression may be evaluated by a one point Gauss integration rule with the sample point at
ξ = 0 (beam midpoint). Let θm = (θ1 + θ2 )/2, ωm = θm + ϕ, cm = cos ωm , sm = sin ωm ,
em = L cos(θm − ψ)/L 0 − 1, γm = L sin(ψ − θm )/L 0 , and
 
−c −sm − 12 L 0 γm cm sm − 12 L 0 γm
1  m
Bm = B|ξ =0 = sm −cm 21 L 0 (1 + em ) sm −cm 12 L 0 (1 + em )  (10.42)
L0
0 0 −1 0 0 1

where subscript m stands for “at beam midpoint.” Then


 T
−cm −sm 1
L γ
2 0 m
cm sm 1
L γ
2 0 m N
p = L 0 Bm z =
T  sm −cm − 12 L 0 (1 + em ) −sm cm − 12 L 0 (1 + em )  V (10.43)
0 0 −1 0 0 1 M

10–16
10–17 §10.7 TANGENT STIFFNESS MATRIX

§10.7. Tangent Stiffness Matrix


The first variation of the internal force vector (10.41) defines the tangent stiffness matrix

 T 
δp = B δz + δBT z d X̄ = (K M + KG ) δu = K δu. (10.44)
L0
This is again the sum of the material stifness K M and the geometric stiffness KG .
§10.7.1. Material Stiffness Matrix
The material stiffness comes from the variation δz of the stress resultants while keeping B fixed.
This is easily obtained by noting that

δN E A0 0 0 δe
δz = δV = 0 G A0 0 δγ = S δh, (10.45)
δM 0 0 E I0 δκ
where S is the diagonal constitutive matrix with entries E A0 , G A0 and E I0 . Because δh = B δu,
the term BT δz becomes BT SB δu = K M δu, whence the material matrix is

KM = BT SB d X̄ . (10.46)
L0

This integral is evaluated by the one-point Gauss rule at ξ = 0. Denoting again by Bm the matrix
(10.42), we find

KM = BmT SBm d X̄ = KaM + KbM + KsM (10.47)
L0

where KaM , KbM and KsM


are due to axial (bar), bending, and shear stiffness, respectively:
 
cm2 cm sm −cm γm L 0 /2 −cm2 −cm sm −cm γm L 0 /2
 cm sm sm2 −γm L 0 sm /2 −cm sm −sm2 −γm L 0 sm /2 
 
E A 0  −c γ L /2 −γ L s /2 γ 2
L 2
/4 c γ L /2 γ L s /2 γm2 L 20 /4 
KaM =  m m2 0 m 0 m m 0 m m 0 m 0 m

L0  −cm −cm sm cm γm L 0 /2 cm2 cm sm cm γm L 0 /2 
 
−cm sm −sm2 γm L 0 sm /2 cm sm sm2 γm L 0 sm /2
−cm γm L 0 /2 −γm L 0 sm /2 γm2 L 20 /4 cm γm L 0 /2 γm L 0 sm /2 γm2 L 20 /4
  (10.48)
0 0 0 0 0 0
0 0 0 0 0 0 
E I0  0 0 1 0 0 −1 

KM =
b
  (10.49)
L0  0 0 0 0 0 0 
 
0 0 0 0 0 0
0 0 −1 0 0 1
 
sm2 −cm sm −a1 L 0 sm /2 −sm2 c m sm −a1 L 0 sm /2
 −cm sm cm2 cm a1 L 0 /2 cm sm −cm2 cm a1 L 0 /2 
 
G A 0  −a L s /2 c a L /2 a 2
L 2
/4 a L s /2 −c a L /2 a12 L 20 /4 
KsM =  1 02 m m 1 0 1 0 1 0 m m 1 0

L0  −sm cm sm a1 L 0 sm /2 sm2 −cm sm a1 L 0 sm /2 
 
cm sm −cm2 −cm a1 L 0 /2 −cm sm cm2 −cm a1 L 0 /2
−a1 L 0 sm /2 cm a1 L 0 /2 a12 L 20 /4 a1 L 0 sm /2 −cm a1 L 0 /2 a12 L 20 /4
(10.50)
in which a1 = 1 + em .

10–17
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–18

§10.7.2. Eliminating Shear Locking by RBF

How good is the nonlinear material stiffness (10.49)-(10.50)? If evaluated at the reference config-
uration aligned with the X axis, cm = 1, sm = em = γm = 0, and we get

 E A0 
L0 0 0 − ELA0 0 0
0
 G A0 
 0
 L0
1
2
G A0 0 − GLA0 

1
2
G A0
 0

 0 1 E I0 + 1 G A L
G A0 0 − E I 0 + − 12 G A0
1
G A L 
 2 L0 4 0 0 L0 4 0 0 
KM =  EA 
− 0 0 0 E A 0 0 0 
 L L0 
 0 
 0 G
− L 0 A − 2 G A0
1
0 G A 0 − 1
G A 
 0 L0 2 0 
0 1
2
G A0 − ELI0 + 14 G A0 L 0 0 − 12 G A0 E I0 + 1 G A L
L0 4 0 0
0
(10.51)
This is the well known linear stiffness of the C 0 beam. As noted in the discussion of §10.2.4, this
element does not perform as well as the C 1 beam when the beam is thin because too much strain
energy is taken by shear. The following substitution device, introduced by MacNeal,5 removes
that deficiency in a simple way. The shear rigidity G A0 is formally replaced by 12E I0 /L 20 , and
magically (10.51) becomes

 EA 
L
0 0 0 − ELA0 0 0
 0 0 
 0 12E I0 6E I0 − 12E3 I0 6E I0
 L 30 L 20
0
L 20
 L0 
 6E I0 4E I0 
 0
 L0 0 − 2I0
6E 2E I0
L0


L 20 L0
K̂ M =
 E A0 E A0
.
 (10.52)
− L 0 0 0 0 
 0 L0 
 
 0 − 12E3 I0 − 6E2I0 0 12E I0 6E
− 2 I 0
 L0 L0 L 30 L0 
 
0 6E I0 2E I0 0 − 6E2I0 4E I0
L 20 L0 L0 L0

This is the well known linear stiffness matrix of the C 1 (Hermitian) beam based on the Bernoulli-
Euler model. That substitution device is called the residual bending flexibility (RBF) correction.6
Its effect is to get rid of the spurious shear energy due to the linear kinematic assumptions. If the
RBF is formally applied to the nonlinear material stiffness one gets K̂ M = KaM + K̂ Mb , where KaM
is the same as in (10.50) (because the axial stiffness if not affected by the substitution), whereas

5 See reference in Notes and Biblioghraphy.

6 RBF can be rigurously justified through the use of a mixed variational principle, or through a flexibility calculation.

10–18
10–19 §10.7 TANGENT STIFFNESS MATRIX

KbM and KsM merge into

 
12sm2 −12cm sm 6a1 L 0 sm −12sm2 12cm sm 6a1 L 0 sm
 −12cm sm 12cm2 −6cm a1 L 0 12cm sm −12cm2 −6cm a1 L 0 
b EI  6a L s −6cm a1 L 0 a2 L 20 −6a1 L 0 sm 6cm a1 L 0 a3 L 20 

K̂ M = 3  1 0 2m 
L 0  −12sm 12cm sm −6a1 L 0 sm 12sm2 −12cm sm −6a1 L 0 sm 
 
12cm sm −12cm2 6cm a1 L 0 −12cm sm 12cm2 6cm a1 L 0
6a1 L 0 sm −6cm a1 L 0 a3 L 20 −6a1 L 0 sm 6cm a1 L 0 a2 L 20
(10.53)
in which a1 = 1 + em , a2 = 4 + 6em + 3em2 and a3 = 2 + 6em + 3em2 .

Remark 10.4. MacNeal actually proposed the more refined substitution

1 1 L 20
replace by + (10.54)
G A0 G As 12E I0

where G As is the actual shear rigidity; that is, As is the shear-reduced cross section studied in Mechanics
of Materials. The result of (10.54) is the C 1 Hermitian beam corrected by shear deformations computed
from equilibrium considerations. (This beam is derived in Chapter 13 of IFEM Notes [220].) If the shear
deformation is negligible, the right hand side of (10.54) is approximately L 20 /(12E I0 ), which leads to the
substitution used above.

§10.7.3. Geometric Stiffness Matrix

The geometric stiffness KG comes from the variation of B while the stress resultants in z are kept
fixed. To get a closed form expression it is convenient to pass to indicial notation, reverting to
matrix notation later upon “index contraction.” Let the entries of KG , B, u and z be denoted as
K Gi j , Bki , u j and z k , where indices i, j and k range over 1–6, 1–6, and 1–3, respectively. Call
A j = ∂B/∂u j , j = 1, . . . 6. Then using the summation convention,

  
∂ Bki j
K Gi j δu j = δB z d X =
T
δu j z k d X = Aki z k d X δu j , (10.55)
L0 L0 ∂u j L0

whence

j
K Gi j = z k Aki d X̄ , (10.56)
L0

Note that in carrying out the derivatives in (10.56) by hand one must use the chain rule because B
is a function of e, γ and θ , which in turn are functions of the node displacements u j . To implement
this scheme we differentiate B with respect to each node displacement in turn, to obtain:

10–19
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–20


∂B 1 0 0 N1 sin ω 0 0 N2 sin ω
A1 = = 0 0 N1 cos ω 0 0 N2 cos ω ,
∂u X 1 L0 0 0 0 0 0 0

∂B 1 0 0 −N1 cos ω 0 0 −N2 cos ω
A2 = = 0 0 N1 sin ω 0 0 N2 sin ω ,
∂u Y 1 L0 0 0 0 0 0 0

∂B N1 sin ω − cos ω −N1 L 0 (1 + e) − sin ω cos ω −N2 L 0 (1 + e)
A3 = = cos ω sin ω −N1 L 0 γ − cos ω − sin ω −N2 L 0 γ ,
∂θ1 L0 0 0 0 0 0 0
(10.57)
∂B 1 0 0 −N1 sin ω 0 0 −N2 sin ω
A4 = = 0 0 −N1 cos ω 0 0 −N2 cos ω ,
∂u X 2 L0 0 0 0 0 0 0

∂B 1 0 0 N1 cos ω 0 0 N2 cos ω
A5 = = 0 0 −N1 sin ω 0 0 −N2 sin ω ,
∂u Y 2 L0 0 0 0 0 0 0

∂B N2 sin ω − cos ω −N1 L 0 (1 + e) − sin ω cos ω −N2 L 0 (1 + e)
A6 = = cos ω sin ω −N1 L 0 γ − cos ω − sin ω −N2 L 0 γ .
∂θ2 L0 0 0 0 0 0 0

To restore matrix notation it is convenient to define


j j j
W N i j = A1i , WV i j = A2i , W Mi j = A3i , (10.58)

as the entries of three 6 × 6 “weighting matrices” W N , WV and W M that isolate the effect of the
stress resultants z 1 = N , z 2 = V and z 3 = M. The first, second and third row of each A j becomes
the j th column of W N , WV and W M , respectively. The end result is
 
0 0 N1 sin ω 0 0 N2 sin ω
 0 0 −N1 cos ω 0 0 −N2 cos ω 
1  N1 sin ω −N1 cos ω −N12 L 0 (1 + e) −N1 sin ω N1 cos ω −N1 N2 L 0 (1 + e) 

WN =  
L0  0 0 −N1 sin ω 0 0 −N2 sin ω 
 
0 0 N1 cos ω 0 0 N2 cos ω
N2 sin ω −N2 cos ω −N1 N2 L 0 (1 + e) −N2 sin ω N2 cos ω −N22 L 0 (1 + e)
 (10.59)

0 0 N1 cos ω 0 0 N2 cos ω
 0 0 N1 sin ω 0 0 N2 sin ω 

1  N1 cos ω N1 sin ω 
−N1 L 0 γ
2
−N1 cos ω −N1 sin ω −N1 N2 L 0 γ 
WV =  
L0  0 0 −N1 cos ω 0 0 −N2 cos ω 
 
0 0 −N1 sin ω 0 0 −N2 sin ω
N2 cos ω N2 sin ω −N1 N2 L 0 γ −N2 cos ω −N2 sin ω −N22 L 0 γ
(10.60)
and W M = 0. Notice that the matrices must be symmetric, since KG derives from a potential. Then

KG = (W N N + WV V ) d X̄ = KG N + KGV . (10.61)
L0

10–20
10–21 §10.7 TANGENT STIFFNESS MATRIX

Node Displacements u

Eqs (10.5), (10.8)

Element displacement field w = [uX , uY , θ]T

Eq. (10.9)

Displacement gradients w' = [u'X , u'Y , θ' ]


T

Eq. (10.15)

Generalized strains h = [ e , γ, κ ]
T

Eq. (10.29) Strain energy U

Stress resultants z = [ N ,V, M ]T


∫ _
vary U: δU = L zT BT dX δu = pT δu
0
Eq. (10.34)

Internal forces p

vary p: δp = ∫ L0
_
(B T δz + δBT z) dX δu = (KM + KG )

Eq. (10.40)

Tangent stiffness matrix K = KM + KG

Figure 10.10. Roadmap for the derivation of the TL plane beam element.

Again the length integral should be done with the one-point Gauss rule at ξ = 0. Denoting again
quantities evaluated at ξ = 0 by an m subscript, one obtains the closed form
 
0 0 sm 0 0 sm
 0 0 −cm 0 0 −cm 
Nm  sm −cm − 2 L 0 (1 + em ) −sm cm − 2 L 0 (1 + em ) 
1 1 
KG =  
2  0 0 −sm 0 0 −sm 
 
0 0 cm 0 0 cm
sm −cm − 12 L 0 (1 + em ) −sm cm − 12 L 0 (1 + em )
  (10.62)
0 0 cm 0 0 cm
 0 0 sm 0 0 sm 
Vm  cm sm − 2 L 0 γm −cm −sm − 2 L 0 γm 
1 1 
+  .
2  0 0 −cm 0 0 −cm 
 
0 0 −sm 0 0 −sm
cm sm − 12 L 0 γm −cm −sm − 12 L 0 γm

10–21
Chapter 10: THE TL PLANE BEAM ELEMENT: FORMULATION 10–22

in which Nm and Vm are N and V evaluated at the midpoint.


Result of running the fragment of

§10.8. Commentary on Element Performance


The material stiffness of the present element works fairly well once MacNeal’s RBF device is done.
On the other hand, simple buckling test problems, as in Exercise 11.3, show that the geometric stiff-
ness is not so good as that of the C 1 Hermitian beam element.7 Unfortunately a simple substitution
device such as RBF cannot be used to improve KG , and the problem should be viewed as open.
An intrinsic limitation of the present element is the restriction to small axial strains. This was
done to facilitate close form derivation. The restriction is adequate for many structural problems,
particularly in Aerospace (example: deployment). However, it means that the element cannot model
correctly problems like the snap-through and bifurcation of the arch example used in Chapter 8, in
which large axial strains prior to collapse necessarily occur.

§10.9. Derivation Summary


Figure 10.10 is a roadmap that summarizes the key steps in the derivation of the internal force and
tangent stiffness matrix for the C 0 plane beam element.

Notes and Bibliography


For detailed justification the curious reader may consult advanced FEM books such as Hughes’ [331]. The
RBF correction was introduced by R. H. MacNeal in [382]. The derivation of the Timoshenko plane beam
element for linear FEM is presented in Chapter 13 of the IFEM Notes [220].

7 In the sense that one must use more elements to get equivalent accuracy.

10–22
10–23 Exercises

Homework Exercises for Chapter 10


The TL Plane Beam Element: Formulation

EXERCISE 10.1
[A+C:20] Show that the displacement field that generates the measures e(1) (2)
X Y and e X Y given in (10.27) simulta-
neously as GL strains is    
x X + u X − Y sin θ
= . (E10.1)
y u Y + Y cos θ + 2X sin θ
Show that if this displacement field is selected, all GL strains exactly vanish for an arbitrary RBM.

EXERCISE 10.2
[A:20] Obtain the linearized strain e X X associated with the field (E10.1) using the polar decomposition of its
deformation gradient tensor.

EXERCISE 10.3
[A+C:30] (Research level) Rederive p and K matrix for the displacement field (E10.1), using the exact GL
strain e X X .

10–23

Potrebbero piacerti anche