Você está na página 1de 14

8

The TL Plane Bar


Element:
Formulation

8–1
Chapter 8: THE TL PLANE BAR ELEMENT: FORMULATION

TABLE OF CONTENTS

Page
§8.1. Introduction 8–3
§8.2. The TL Plane Bar Element 8–3
§8.2.1. Element Kinematics . . . . . . . . . . . . . . . 8–3
§8.2.2. GL Axial Strain . . . . . . . . . . . . . . . . 8–5
§8.2.3. Strain Derivatives . . . . . . . . . . . . . . . . 8–7
§8.2.4. PK2 Axial Stress . . . . . . . . . . . . . . . . 8–7
§8.2.5. Total Potential Energy . . . . . . . . . . . . . . 8–8
§8.2.6. Internal Force Vector . . . . . . . . . . . . . . 8–8
§8.2.7. Tangent Stiffness Matrix . . . . . . . . . . . . . . 8–8
§8.2.8. *Quick Refresher on Matrix Calculus . . . . . . . . . 8–11
§8. Notes and Bibliography
. . . . . . . . . . . . . . . . . . . . . . 8–12
§8. Exercises . . . . . . . . . . . . . . . . . . . . . . 8–13

8–2
§8.2 THE TL PLANE BAR ELEMENT

§8.1. Introduction
The basic concepts of nonlinear continuum mechanics reviewed in Chapter 7 are applied to the
development of finite element equations of a two-dimensional (plane), two-node bar element based
on the Total Lagrangian (TL) kinematic description. This will be referred to as a TL bar element
for brevity. There are two ways to construct TL finite elements:
1. The Standard Formulation (SF)
2. The Core Congruential Formulation (CCF).
The first method is easier to describe and is that followed in this Chapter. The second one is more
flexible and powerful but it is far more difficult to teach because it proceeds in stages. The CCF
will not be covered in this course, but older Chapters explaining it are posted on the web site.

§8.2. The TL Plane Bar Element


The element developed in this Chapter is a prismatic bar element that can be used to model pin-
jointed plane truss structures of the type sketched in Figure 8.1. Such structures may undergo
large displacements and rotations but their strains are assumed to remain small so that the material
behavior stays in the linear elastic range. These assumptions allows us to consider only geometric
nonlinear effects.1
Figure 8.2 shows a two-node bar element appropriate to model members of such truss structures.
The element moves in the (X, Y ) plane. In the reference (base) configuration the element has cross
section area A0 (constant along the element) and length L 0 . In the current configuration the cross
section area and length become A and L, respectively. The material has an elastic modulus E that
links the axial-stress and axial-strain measures defined later.
Because this Chapter deals primarily with the formulation of an individual element, the identification
superscript e will be omitted to reduce clutter until assemblies are considered.
The element has four node displacements and associated node forces. These quantities are collected
in the vectors    
u X1 fX1
u   f 
u =  Y1  , f =  Y1  , (8.1)
u X2 f X2
uY 2 fY 2

§8.2.1. Element Kinematics


In accordance with theory, the bar element remains straight in any configuration. Therefore, to
describe the element motion it is sufficient to consider a generic point P0 of material coordinates
X located on the longitudinal axis of the reference configuration C0 . That point maps to point P
of spatial coordinates x in the current configuration C. See Figure 8.3. The C0 coordinates can be
parametrically interpolated from the end nodes as
 1 
X (ξ ) (1 − ξ )X 1 + 12 (1 + ξ )X 2 N1 (ξ )X 1 + N2 (ξ )X 2
X(ξ ) = = 1 2
= , (8.2)
Y (ξ ) (1 − ξ )Y1 + 1 (1 + ξ )Y2 N1 (ξ )Y1 + N2 (ξ )Y2
2 2

1 An important application in Aerospace is the deployment of space trusses. In Civil Engineering it would be the deployment
of geodesic domes. For those applications a three-dimensional version of this bar element would be required.

8–3
Chapter 8: THE TL PLANE BAR ELEMENT: FORMULATION

Reference configuration (same


as base in TL description)
Motion

Current configuration

Figure 8.1. A plane truss structure undergoing large displacements while its material stays in the
linear elastic range.

2 (X2 ,Y2)
Reference configuration C 0
PK2 stress s0 , GL strain e0 = 0

X-sec area A0 , length L 0 ,


elastic modulus E

Y, y 1 (X1 ,Y1)
uY2
X, x uY1 fY1
X-sec area A, length L
X,Y : material frame
uX1 fX1
x,y : spatial frame fY2
1 (x1 ,y1)

Current configuration C fX2


uX2
PK2 stress s, GL strain e 2 (x2 ,y2)
Figure 8.2. The geometrically nonlinear, two-node, plane bar element in the Total Lagrangian (TL)
kinematic description. Can be used to model members of a plane truss such as that shown in Figure 8.1.

with a similar formula for the C coordinates:


 1 
x(ξ ) (1 − ξ )x1 + 12 (1 + ξ )x2 N1 (ξ )x1 + N2 (ξ )x2
x(ξ ) = = 1 2
= . (8.3)
y(ξ ) (1 − ξ )y1 + 12 (1 + ξ )y2 N1 (ξ )y1 + N2 (ξ )y2
2

Here ξ is the usual isoparametric coordinate that varies from −1 at node 1 to +1 at node 2, whereas
N1 (ξ ) = 12 (1 − ξ ) and N2 (ξ ) = 12 (1 + ξ ) are the well known linear shape functions in terms of ξ .
The displacement field is obtained by subtraction:
 
u X (ξ ) N1 (ξ ) u X 1 + N2 (ξ ) u X 2
u(ξ ) = x(ξ ) − X(ξ ) = = , (8.4)
u Y (ξ ) N1 (ξ ) u Y 1 + N2 (ξ ) u Y 2

8–4
§8.2 THE TL PLANE BAR ELEMENT

Bar longitudinal axis


2 (X2 ,Y2)

C0

P0 (X) X(ξ) N1(ξ) X1 + N2(ξ) X2


X(ξ) = =
Y(ξ) N1(ξ) Y1 + N2(ξ) Y2
Y, y 1 (X1 ,Y1)

u X(ξ) N1(ξ) u X1 + N2 (ξ) uX2


u(ξ) = x(ξ) − X(ξ) = =
uY (ξ) N1(ξ) u Y1 + N2 (ξ) uY2

X, x
x(ξ) N1(ξ) x1 + N2(ξ) x2
x(ξ) = =
y(ξ) N1(ξ) y1 + N2(ξ) y2
1 (x1 ,y1)
P(x) Bar longitudinal axis
C
2 (x2 ,y2)
Figure 8.3. Displacement field for the plane TL bar element.

Equation (8.4) may be expressed in matrix form as

 
  u X1
u X (ξ ) N1 (ξ ) 0 N2 (ξ ) 0  uY 1 
u(ξ ) = =   = N(ξ ) u. (8.5)
u Y (ξ ) 0 N1 (ξ ) 0 N2 (ξ ) u X2
uY 2

Here N(ξ ) is the 2 × 4 shape function matrix. The element kinematics defined by these equations
is pictured in Figure 8.3

§8.2.2. GL Axial Strain

As discussed in Chapter 7, in the Total Lagrangian (TL) description the Green-Lagrange (GL)
strains and the second Piola-Kirchhoff (PK2) stresses are frequently used as conjugate measures in
the formulation of the internal energy. The only GL strain that appears in the energy expression is
the GL axial strain e1 ≡ e, which is most expediciously defined using the length change as

L 2 − L 20
e= , (8.6)
2L 20

rather than through displacement gradients. Because of the linear displacement assumptions (8.5)
the strain e is constant over the element.2

2 This is in fact the only use of the displacement interpolation (8.5) in the ensuing derivations.

8–5
Chapter 8: THE TL PLANE BAR ELEMENT: FORMULATION

_
X
2 (X2 ,Y2) ψ0
X 21

C0 X 21
Y21 c0X = = cos ψ0
L0
Y21
L0 c0Y = = sin ψ0
L0
Y, y 1 (X1 ,Y1)
x21 L
cx = = cos ψ
L0 L0
y21 L
X, x cy = = sin ψ
L0 L0
1 (x1 = X1+uX1, y1=Y1+ uY1)
x21 = X21 + uX21

ψ y21 = Y21+ u Y21


C
L
2 (x2 = X2 +uX2 , y2=Y2 + u Y2)

Figure 8.4. Geometric interpretation of quantities used in the study of element kinematics.

Expression (8.6) can be maneuvered into a matrix function of the node displacements. To expedite
the procedure it is convenient to introduce the following auxiliary variables:
X 21 = X 2 − X 1 , Y21 = Y2 − Y1 , u X 21 = u X 2 − u X 1 , u Y 21 = u Y 2 − u Y 1 ,
X 21 Y21
c0X = = cos ψ0 , c0Y = = sin ψ0 ,
L0 L0 (8.7)
u x21 L u y21 L
cx = c X + X 21 = = cos ψ, c y = cY + Y 21 = = sin ψ.
L0 L0 L0 L0 L0 L0
Some of these quantities can be geometrically interpreted as illustrated in Figure 8.4. In particular,
c0X and c0Y are the {X, Y } direction cosines of the bar longitudinal axis in the reference configuration.
On the other hand, cx and c y are scaled (by L/L 0 ) direction cosines of the current configuration.
The squared bar expressions in e in terms of nodal displacements are

L 2 = (X 21 + u X 21 )2 + (Y21 + u Y 21 )2 , L 2 − L 20 = 2X 21 u X 21 + 2Y21 u Y 21 + X 21
2
+ Y21
2
, (8.8)

whence the GL axial strain (8.6) becomes


L 2 − L 20 1 1 2
e= = (c u + c u ) + u + u Y 21 .
2
(8.9)
2L 20 L 0 0X X 21 0Y Y 21
2L 20 X 21
This may be compactly written as

e = B0 u + 12 uT H u, (8.10)

8–6
§8.2 THE TL PLANE BAR ELEMENT

in which
 
1 0 −1 0
1 1  0 1 0 −1 
B0 = [ −c0X −c0Y c0X c0Y ] , H= 2 . (8.11)
L0 L 0 −1 0 1 0
0 −1 0 1
For future use it is convenient to introduce the matrix
1
B= [ −cx −c y cx c y ] = B0 + H u, (8.12)
L0
which is similar to B0 but with direction cosines evaluated in the current configuration. (The fact
that B − B0 = H u may be easily verified by direct computation.) The expression (8.10) shows
that the GL strain splits naturally into two parts: e = e L + e N , in which
e L = B0 u, where B0 is a constant-over-element, 1 × 4 rectangular matrix given by (8.11), depends
linearly on the node displacements u. This is the linear part of the GL strain.
e N = 12 uT H u, where H is a constant-over-element, 4 × 4 symmetric square matrix given by (8.11),
depends quadratically on the node displacements. This is the nonlinear part of the GL strain.

§8.2.3. Strain Derivatives


For further use in the computation of internal forces and stiffness matrix, the first and second
derivatives of e with respect to the nodal displacements will be needed.3 The derivative of e with
respect to u is

∂e ∂(B0 u + 12 uT H u)
= = B0T + uT HT = B0T + uT H = BT , (8.13)
∂u ∂u

in which use is made of (8.12). The second derivative of e with respect to u is

∂ 2e ∂ ∂e ∂(B0T + uT H) ∂BT
= = = = H. (8.14)
∂u ∂u ∂u ∂u ∂u ∂u

§8.2.4. PK2 Axial Stress


The stress measure conjugate to GL strains is the second Piola-Kirchhoff (PK2) stress tensor. The
only component that appears in the internal energy is the axial stress s, which is related to e through
the constitutive equation
s = s0 + E e, (8.15)
where s0 is the axial stress in the reference configuration (assumed constant over the element), and
E is the elastic modulus. The axial force based on this stress is

F = A0 s. (8.16)

3 If you are rusty in matrix calculus as regards taking derivatives with respect to a vector, please skim §8.2.8.

8–7
Chapter 8: THE TL PLANE BAR ELEMENT: FORMULATION

Note that this is not the true axial force in the current configuration C , which would be

Ftr ue = A σ, (8.17)

in which σ denotes the true or Cauchy stress in C and A is the actual cross-section area there.
Since s 0 and E are fixed, the stress derivatives with respect to the state vector are

∂s ∂e ∂ 2s ∂ 2e
=E = E BT , =E = EH (8.18)
∂u ∂u ∂u ∂u ∂u ∂u

§8.2.5. Total Potential Energy


In what follows it is assumed that the element is subjected only to node forces f that are conservative
and proportional. Consequently
f = λq. (8.19)
Here q denotes the incremental load vector. The Total Potential Energy (TPE) of the element in the
current configuration, expressed in terms of GL strains and PK2 stresses, is

=U −W = (s0 e + 2 E e ) d V0 − f u =
1 2 T
A0 (s0 e + 12 E e2 ) d X̄ − λqT u. (8.20)
V0 L0

where X̄ is directed along the bar longitudinal axis in C0 , as shown in Figure 8.4. All integrals are
carried out over the reference (=base) configuration C 0 . Since the integrands are constant, we get

= U − W, U = V0 (s0 e + 12 Ee2 ), W = λ qT u. (8.21)

Here V0 = A0 L 0 is the bar volume in C0 . This energy expression is separable because the internal
energy U depends only on u through e, and not on λ.

§8.2.6. Internal Force Vector


The finite element residual equations are obtained by taking the gradient of (8.21) with respect to
u. Since is separable, r = p − f, in which f = ∂ W/∂u = λq is the external force. The internal
force can be expanded as follows
 
∂U ∂e ∂e ∂e
p= = V0 s0 + Ee = V0 s . (8.22)
∂u ∂u ∂u ∂u
Using (8.13) we arrive at the detailed expression

p = V0 s BT = F L 0 BT = F [ −cx −c y cx c y ]T . (8.23)

Here F = A0 s is the axial force in the current configuration measured per unit area of the reference
configuration.4 The geometric interpretation of cx and c y is depicted in Figure 8.5. The relation
between F = A0 s and the true axial force Ftr ue = A σ is worked out in an Exercise.
4 This is the PK2 axial force; cf. (8.16) .

8–8
§8.2 THE TL PLANE BAR ELEMENT

fY1 = −F0 a y

fX1= −F0 ax 1
F0
2 fX2 = F0 ax

fY2 = F0 a y

Figure 8.5. Geometrical interpretation of the internal force vector (8.22). The axial force
F0 = A0 s would be positive as shown.

§8.2.7. Tangent Stiffness Matrix


Because the residual equations are separable, the tangent stiffness matrix is obtained simply by
differentiating the internal force p given by (8.22) with respect to the node displacements u:
 
∂ ∂e ∂e ∂e ∂ 2e
K= V0 s = V0 E ⊗ + V0 s = K M + KG , (8.24)
∂u ∂u ∂u ∂u ∂u ∂u

in which ⊗ denotes the dyadic (outer) product of two vectors. This expression indicates that K splits
naturally into two parts: K M and KG , which are called the material stiffness matrix and geometric
stiffness matrix, respectively, in the FEM literature. For the material stiffness we get
 c2 cx c y −cx2 −cx c y 
x
∂e ∂e E A0  cx c y c2y −cx c y −c2y 
KM = V0 E ⊗ = E A0 L 0 BT B =  .
∂u ∂u L0 −cx2 −cx c y cx2 cx c y
−cx c y −c2y cx c y c2y
(8.25)
Here we have replaced ∂e/∂u = BT from the first of (8.18), and applied the dyadic product formula
recalled in (8.34). The matrix expression in (8.25) looks formally similar to the stiffness matrix of
a linear bar element,5 except that B is now a function of u. The dependence of K M on material
properties (here just the elastic modulus E) explains the name “material stiffness,” which has
become standard in FEM publications. For the geometric stiffness, use (8.14) to get
 
1 0 −1 0
F  0 1 0 −1 
KG = V0 s H = F L 0 H =  . (8.26)
L 0 −1 0 1 0
0 −1 0 1

This component of K depends only on the stress state in the current configuration, since F = A0 s.
No material properties appear. Thus the name “geometric stiffness” applied to KG .6

5 To which it reduces if u = 0. In the linear case cx and c y become the cosine and sine, respective, of the orientation angle
ψ0 in the reference configuration. That angle is shown in in Figure 8.4.
6 As described under Notes and Bibliography, the confusing name “initial stress stiffness” was often used for KG in
pre-1970 FEM publications. It is also noted there that K M was presented as a two-matrix combination by some authors.

8–9
Chapter 8: THE TL PLANE BAR ELEMENT: FORMULATION

Remark 8.1. Assuming that E, A0 and L 0 are nonzero, the rank of K M is obviously one since B is a 1 × 4
matrix. On the other hand, the rank of the numerical matrix in H is 2 (because its eigenvalues are 2, 2, 0, 0).
Consequently KG has rank 2 if s is nonzero and 0 otherwise. Combining these results it can be shown that the
rank of K = K M + KG is 1 if the current configuration is unstressed and 2 otherwise. In other words, the rank
deficiency is 3 and 2, respectively. The implications of this property in the analysis of stability are considered
in later Chapters.

Remark 8.2. The addition of KG increases the bar stiffness if the current configuration is in tension (s > 0),
but it reduces it if the current configuration is in compression (s < 0). This is in accord with physical intuition.
The main effect of this stiffness is on the rotational rigid-body motions of the bar about the Z axis.

Reference configuration C 0 X-sec area A0 = 6, length L 0 = 4,


PK2 stress s0 = 25, GL strain e0 = 0 elastic modulus E = 160
1 (1,2) 2 (5,2)

Y, y uY1 =−1
1 (3,1)
uX1=2
uY2 =−5
X, x

X-sec area A, length L

Current configuration C
PK2 stress s, GL strain e
2 (6,−3)
uX2 =1
Figure 8.6. TL plane bar element example, with numerical data as shown.

Example 8.1. This numerical example pertains to the bar element shown in Figure 8.6. The reference
configuration geometry is defined by the position coordinates
 (X 1 = 1, Y1 = 2) and (X 2 = 5, Y2 = 2) of
the end nodes. The reference bar length is L 0 = (5 − 1) + (2 − 2)2 = 4. Other input data includes:
2

cross-section area A0 = 6, elastic modulus E = 160, reference (initial) PK2 stress s0 = 25, and reference
GL strain e0 = 0. The node displacements are (u X 1 , u Y 1 ) = (2, −1) and (u X 2 , u Y 2 ) = (1, −5). As a
result, the end nodes in the current configuration are located at (x1 = 1 + 2 = 3, y1 = 2 − 1 = 1) and
(x2 = 5 + 1 = 6, y2 = 2 − 5 = −3). Compute the internal force and tangent stiffness matrix in the current
configuration.

The current length is L = (6 − 3)2 + (−3 − 1)2 = 5. The current GL strain is e = (L 2 − L 20 )/(2L 20 ) =
(25 − 16)/32 = 9/32 = 28.125%, whence the PK2 stress is s = s0 + E e = 70. The scaled direction cosines
in C are cx = x21 /L 0 = 3/4 and c y = y21 /L 0 = −4/4 = −1. The matrices required for the internal force and
tangent stiffness computations are
 
1 0 −1 0
1 1  0 1 0 −1 
B= [ −3/4 −1 ] , H=  . (8.27)
0 
1 3/4
4 16 −1 0 1
0 −1 0 1

8–10
§8.2 THE TL PLANE BAR ELEMENT

Replacing these in equations (8.22) through (8.26) yields


   
−315 135 −180 −135 180
 420   −180 240 180 −240 
p= , KM =  ,
315 −135 180 135 −180
420 180 −240 −180 240
    (8.28)
105 0 −105 0 240 −180 −240 180
 0 105 0 −105   −180 345 180 −345 
KG =  , K =  .
−105 0 105 0 −240 180 240 −180
0 −105 0 105 180 −345 −180 345

§8.2.8. *Quick Refresher on Matrix Calculus


The material here is intended to help readers unfamiliar (or rusty) with matrix calculus as regards some
derivations carried out in §8.2.3, §8.2.6 and §8.2.7. These involve taking derivatives of a scalar or vector with
respect to a vector. The exposition is done using full-form examples. In what follows a, b, and c denote
scalars, x and y are 2 × 1 column vectors, A is an arbitrary 2 × 2 matrix, and S is a symmetric 2 × 2 matrix:
       
x1 y1 A11 A12 S11 S12
x= , y= , A= , S= . (8.29)
x2 y2 A21 A22 S12 S22
Matrix, full and indicial notations are used as appropriate; for the latter indices i, j run over 1,2.
Derivative of a scalar with respect to a vector:
 ∂c 
∂c   ∂ x1
= ∂∂c = (8.30)
∂x x i ∂c
∂ x2
Derivative of a vector with respect to a vector:
 ∂ x1 ∂ x1 
∂x  ∂ xi  ∂ y1 ∂ y2
= ∂y = . (8.31)
∂y j ∂ x2 ∂ x2
∂ y1 ∂ y2
Second derivative of a scalar with respect to two vectors:
 
  ∂ 2c ∂ 2c
∂ c
2
∂ c2 ∂x ∂y ∂ x1 ∂ y2 
= ∂ xi ∂ y j =  12 1 (8.32)
∂x ∂y ∂ c ∂ 2c
∂ x2 ∂ y1 ∂ x2 ∂ y2
This yields a symmetric matrix if x = y, as in the second derivative of the GL strain given in (8.14)
Dyadic product (a.k.a. outer product) of two scalar-with-respect-to-vector derivatives:
 ∂a ∂b ∂a ∂b   ∂a   T
∂a ∂b  ∂a ∂b  ∂ x1 ∂ y1 ∂ x1 ∂ y2 ∂ x1  ∂b ∂b  ∂a ∂b
⊗ = ∂x ∂y = ∂a ∂b ∂a ∂b = ∂a ∂ y1 ∂ y2 = ∂x . (8.33)
∂x ∂y i j ∂y
∂ x2 ∂ y1 ∂ x2 ∂ y2 ∂ x2
In particular, if y = x and a = b = c, the resulting matrix is symmetric:
 ∂c ∂c ∂c ∂c
  ∂c   T
∂c ∂c  ∂c ∂c  ∂ x1 ∂ x1 ∂ x1 ∂ x2 ∂ x1  ∂c ∂c  ∂c ∂c
⊗ = ∂x ∂x = ∂c ∂c ∂c ∂c = ∂ x1 ∂ x2 = ∂x . (8.34)
∂x ∂x i j ∂c ∂x
∂ x2 ∂ x1 ∂ x2 ∂ x2 ∂ x2

8–11
Chapter 8: THE TL PLANE BAR ELEMENT: FORMULATION

Some consequences of above formulas for dot product forms:



∂(xT y) ∂(xT y) ∂(xT x) ∂ xT x x
= y, = x, = 2x, = √ . (8.35)
∂x ∂y ∂x ∂x xT x
And for quadratic forms:

∂(xT A y) ∂(xT A y) ∂(xT A x) ∂(xT S x)


= A y, = AT x, = (A + AT ) x, = 2 S x. (8.36)
∂x ∂y ∂x ∂x

Notes and Bibliography


Work in DSM-based finite elements for geometrically nonlinear analysis started at Boeing in the late 1950s.
Much of it appeared in internal reports and memoranda difficult to access. The first systematic account of
this early period was provided by Martin7 at the 1965 Wright-Patterson conference [450]. His paper starts
“Development and application of the direct stiffness method to geometrically nonlinear problems has been
underway since 1958. Only a portion of that work has appeared in the general technical literature. Furthermore,
the derivations which have been given for stiffness matrices required for the large deflection problem have
often been confusing and, in some cases, have led to incorrect conclusions.”
Most of the early work, as well as the confusion alluded to by Martin, pertains to what is now called the
geometric stiffness matrix. The early name for it was initial stress matrix. A bar geometric stiffness was first
presented in [729]. If this is compared to (8.26) one may observe that half of the nonzero entries are missing.
As a consequence KG is not invariant with respect to the choice of coordinate frames, and has the wrong
rank. For beams and plates, a veritable “geometric stiffness matrix zoo” appeared in print during the 1960s.8
Such discrepancies were clarified in review articles by Martin [452] and Carey [122], who showed it to be a
consequence of using various order theories in the strain energy to build author-dependent approximations.
Another contributor to zoo crowding was use of different shape functions for different energy terms.
Further clarification on matrix derivations was provided by Rajasekaran and Murray [581]. This was the
departing point for the CCF formulation [217,221]. From current perpective those past controversies are
useless. The present development is consistent with continuum mechanics, and leaves no room for ambiguities.
What is here called the material stiffness matrix K M , such as (8.25) for the TL plane bar, is not identified as
such by some authors. Instead that component is presented as a combination of two matrices:
The linear stiffness matrix evaluated at the reference configuration, often assuming zero initial stresses.
The initial displacement matrix, which accounts for the motion from reference to current configuration.
The end result is that the tangent stiffness matrix appears as a combination of three matrices: linear, initial dis-
placement, and geometric. This three-way decomposition was introduced by Mallet and Marcal in 1968 [445].
It offers advantages for certain linearization arguments in stability calculations. It is, however, unnecessary
for our developments.

7 Harold C. Martin was Professor of Aeronautics and Astronautics at the University of Washington, Seattle, WA. He was an
active consultant to Boeing and JPL. His work with Jon Turner and Ray Clough over 1954–1955 resulted in the seminal
paper [727], widely considered as the start of the FEM as used today.
8 Some of the published results were plainly wrong, for example those in [23]. Incorrect conclusions were unintended
consequences of use of purely intuitive geometric arguments.

8–12
Exercises

Homework Exercises for Chapter 8


The TL Bar Element

fY = λ

uY
2 uX

E, A0 (1) (2) E, A0

H Y, y

;; ;;
α
1 X, x 3

S
S
(3)
E, A0

;; 4

Figure E8.1. A 3-bar FEM model for Exercise 8.1.

EXERCISE 8.1 [A/C:25] You go to work as a nonlinear-FEM engineer for a car company. Your supervisor
assigns you the job of designing a component of a wheel suspension system that can be modeled by the 3-bar
structure depicted in Figure E8.1. The model has the dimensions and properties shown and is only subjected
to vertical loads at node 2. The length S, bar section areas A0 and elastic modulus E are known, but the rise
angle α > 0 is a design variable. Find the largest α for which bifurcation, which is bad for the wheel,
√ cannot
occur. (For the 2-bar arch example structure that maximum α was shown to be defined by tan α ≤ 2/2.)
To study bifurcation, it is enough to set up the tangent stiffness matrix assuming that the X (horizontal)
displacement u X is zero. The following tangent stiffness matrix is obtained for S = 2, E = 1, A0 = 1 and
u X = 0:

 2
+ 2H u Y + u 2Y 
uY + uY + 2 
  0
KXX K XY  4 16 (1 + H 2 )3 
K= = 
K XY KY Y 2H + 6H u Y +
6u 2Y 2
3u 2Y
0 1
2
+ 3u4Y + 16 + 
(1 + H 2 )3
(E8.1)
Bifurcation points occur if K X X , which is quadratic in u Y : a + b u Y + c u 2Y , vanishes. Real bifurcation points
occur if the discriminant b2 − 4ac ≥ 0. Study when this happens as a function of H , and deduce the largest
α for which bifurcation cannot occur.

8–13
Chapter 8: THE TL PLANE BAR ELEMENT: FORMULATION

;;
Y, y

Bar element models


weightless string in
reference configuration
E, A0
L

(1)
Gravity field g

uY

1 uX X, x

Point mass m
Figure E8.2. Model of a classic pendulum for Exercise 8.2.

EXERCISE 8.2 [A:C:20] Although this course focuses on statics, this exercise deals with the effect of the
geometric stiffness on vibrations. Consider the pendulum configuration idealized in Figure E8.2. A lumped
mass m is suspended by a weightless elastic string. The string is modeled as a 2-node bar element. This
element is under a tensile prestress s0 = mg/A0 , where g is the acceleration of gravity. The tangent stiffness
matrix for the cable element in the reference configuration is K = K M + KG , which is 2 × 2 upon removing
the degrees of freedom at the fixed node 2. Because of the prestress the geometric stiffness does not vanish.
The order-2 vibration eigenproblem is

Kzi = ωi2 M zi , i = 1, 2 (E8.2)

where i is the mode index, ωi is the i th circular frequency in radians per second, zi the associated eigenvector
that include the horizontal and vertical displacements of node 1, and the mass matrix is
 
m 0
M= (E8.3)
0 m
Compute the two frequencies ω1 and ω2 . One of them, say ω1 , describes pendulum motions while the other
one pertains to a “bar mode” associated with axial motions. Discuss what happens to ω1 and ω2 if E → ∞,
which characterizes
√ the “inextensional string” limit, and whether the classical pendulum small-oscillations
frequency ω P = g/L is correct.

EXERCISE 8.3 (Requires knowledge of continuum mechanics.) [A:15] Suppose that the bar-element material
is linear isotropic, with elastic modulus E and ν is Poisson’s ratio ν. Find the relation between the true (Cauchy)
axial stress σ = σx x in the bar and the PK2 axial stress s = s X X . Hint: study the change in cross section area
as function of ν.

8–14

Você também pode gostar