You are on page 1of 45

ROSE School, Pavia, Italy Instructor: Enrico Spacone

NONLINEAR STRUCTURAL ANALYSIS Teaching Assistant: K. Wijesundara


Spring Term 2009

The class notes© for this course were developed jointly by


Prof. E. Spacone at the University “G D’Annunzio” of Chieti-Pescara, Italy
and Prof. J. P. Conte at the University of California, San Diego, USA

6. BEAM THEORY – WEAK FORM


FINITE ELEMENT FORMULATIONS

6.1 INTRODUCTION

Comparison of different approaches to solve structural analysis/mechanics problems:

STRONG FORM
The strong form of the problem consists of the differential equations of the problem.

The exact solution of the differential equations is only available for simple cases.

When the geometry, the boundary conditions and the loads become complex, it is

basically impossible to find a closed-form solution. For a simple beam with constant

cross section, the governing equation EIv0′′′′ + wy = 0 can be easily solved,

provided wy is a simple function. On the other hand, if the beam has a tapered cross

section and the material is nonlinear, i.e., E = E ( ε , x ) , the beam differential

equation becomes ( E ( ε , x) I ( x) v′′)′′ + w


0 y = 0 and is not solvable.

u ( x) b ( x)

Compatibility Equilibrium
ε = Lε u LTσ σ + b = 0

Constitutive Law
ε ( x) σ = σ (ε) σ ( x)

Figure 1 Tonti’s diagram of Strong Form

6-1
6-2

WEAK FORM (Rayleigh-Ritz Method)

The Rayleigh-Ritz method proposes to solve the problem using an approximate,

kinematically admissible solution, i.e., a solution that satisfies the essential (or

kinematic) boundary conditions. Rather than solving the differential equation exactly,

the R-R method suggests to use an approximate solution that applies to the entire

domain of the problem. The solution is given in the form

n
u ( x ) ≅ φ0 + ∑ ci φi ( x )
i =1

and the unknown coefficients ci are found by imposing the Principle of Stationary

Potential Energy ( ⇒ ∂Π ∂ci = 0 ). The solution improves as more terms are used,

i.e., as n →∞.

WEAK FORM (FE Method):

Finite Element Methods are based on the realization that finding a “good” solution on

the entire problem domain (say a three story building) may be quite cumbersome.

The structure is thus divided into several subdomains (the “elements”) and an

approximate solution to the displacement fields is found inside each element. At the

limit, as the number of elements becomes infinite (and their size infinitely small), the

solution should converge to the exact solution.

The remainder of this chapter will deal with application of the FE method to the beam problem.

6.2 DISPLACEMENT-BASED FORMULATIONS

A displacement-based finite element may be formulated in a number of ways. Two equivalent alternatives

are presented hereafter, one based on the Principle of Virtual Displacements, the other on the Principle of

Stationary Potential Energy. After presenting these two approaches, their applications to the formulation of

beam finite elements will be presented.


6-3

6.2.1 Principle of Virtual Displacements

(a) Compatibility (enforced in the Strong Form)

Section Deformations ε ( x ) = Lε u ( x ) (6.1)

where ε ( x) = generalized deformation fields along element

Lε = strain differential operator

u ( x) = displacement fields

(b) Section Constitutive Law

σ ( x) = k ( x) ε ( x) (6.2)

where σ ( x) = generalized force fields along element

k ( x ) = section stiffness matrix

(c) Equilibrium (enforced in the Weak Form through the Principle of Virtual Displacements)

L L
δ U P = ∫ δε
T T
( x ) σ ( x ) dx = ∫ δ uT LTε ( x ) σ ( x ) dx (6.3)
0 0

6.2.1.1 Finite Element Approximation

u ( x ) = NU ( x ) U (6.4)

Thus,

ε ( x ) = Lε u ( x ) = Lε N U ( x ) U (6.5)

The following expression is typically used in FE derivations:

B ( x ) = Lε N U ( x ) (6.6)

Thus,
6-4

ε ( x) = BU (6.7)

and

L
δ U P = ∫ δ UT BT ( x ) σ ( x ) dx
T

 L

δ UT  P − ∫ BT ( x ) σ ( x ) dx  = 0
 0 

The above equation must hold for an arbitrary δ U . Thus,

L
P = ∫ BT ( x ) σ ( x ) dx (6.8)
0

The above equation can be further developed to yield the element stiffness matrix K.

L
P = ∫ BT ( x ) σ ( x ) dx
0
L
= ∫ BT ( x ) k ( x ) ε ( x ) dx
0
L
= ∫ BT ( x ) k ( x ) B ( x ) dx U

0
 
K

or

P = KD (6.9)

where

L
K = ∫ BT ( x ) k ( x ) B ( x ) dx (6.10)
0

P and K are weighted integrals (sums) of the section forces and stiffness. This will become clear in a
later chapter, where issues related to the numerical implementation of the elements are discussed.
6-5

6.2.2 Principle of Stationary Potential Energy

The procedure is identical to that followed in the previous approach, that uses the Principle of Virtual

Displacements. In step (c), Equilibrium is enforced through the Principle of Stationary Potential Energy.

For convenience, we assume that we only have nodal loads P.

(c) Equilibrium (enforced in the Weak Form through the Principle of Stationary Potential Energy)

L
1 T
ε ( x ) σ ( x ) dx − UT P
2 ∫0
Π=

L
1 T
= ∫ ε ( x ) k ( x ) ε ( x ) dx − UT P
20

L
δΠ = ∫ δεT ( x ) k ( x ) ε ( x ) dx − δ UT P
0

According to the Principle of Stationary Potential Energy, the equilibrium configuration is found by setting

δΠ = 0 , which leads to

∫ δε ( x ) k ( x ) ε ( x ) dx − δ UT P = 0
T

or

∫ δε ( x ) σ ( x ) dx − δ UT P = 0
T

which is equivalent to the Principle of Virtual Displacements. Thus, the same formulation is obtained as in

Eq. (6.3).

6.2.3 Interpolation (or Shape) Functions

In the FE Method, the displacement field inside an element is typically approximated through a polynomial

interpolation, which can be written in the Rayleigh-Ritz form as

(A) u ( x) = a0 + a1 x + a2 x 2 + ⋯

In this case, the physical meaning of the coefficients a0 , a1 , a2 , ... is lost.


6-6

Alternatively, the approximate displacement field can be written as a function of the nodal displacements

U as

n
(B) u ( x) = N1 ( x) U1 + N 2 ( x) U 2 + N 3 ( x) U 3 + ⋯ = ∑ N i ( x) U i = N U ( x) U
i =1

where U i denotes the nodal displacement along degree of freedom i and N i ( x ) is the corresponding

interpolation (or shape) function. The displacement interpolation functions must satisfy the following

conditions:

N i ( x ) = 1 at U i

N i ( x ) = 0 at U j , j ≠ i

We will distinguish between two classes of problems:

(a) those involving displacements only

(b) those involving displacements and their derivatives

The first class (a) requires continuity only of the displacements, and will be referred to as C 0 problems
(e.g., truss, torsion), whereas the second class (b) requires continuity of slopes and will be referred to as

C1 problems.

Strong form
u ( x) b ( x)
Weak form

Compatibility Equilibrium
L L
ε = Lε u
∫ δε σ dx - ∫ δ u b dx = 0
Τ T

0 0

Constitutive Law
ε ( x) σ = σ (ε ) σ ( x)

Figure 1 Tonti’s diagram of Weak Form (Displacement-Based Formulation)


6-7

6.2.4 Example 1: Axial/Torsional Problems - Two-Node Element

Expression (A) u ( x ) = a0 + a1 x

Find a0 and a1 by imposing u ( 0 ) = U1 and u ( L ) = U 2

1
a0 = U1 , a1 = (U 2 − U1 )
L

1
u ( x ) = U1 + (U 2 − U1 ) x
L

U 
Expression (B) u ( x ) = N U1 ( x ) U1 + N U 2 ( x ) U 2 = { N U1 ( x ) N U 2 ( x )}  1  = N U ( x ) U
U 2 

x x
where N U1 ( x ) = 1 − and NU2 ( x ) = (6.11)
L L

1 1
N U1 ( x ) NU2 ( x )

x x
0 L 0 L

Figure 2 Shape functions of two-node bar element

Why can’t we use higher-order terms?


6-8

6.2.5 Example 2: Bending Problem (Euler-Bernoulli Beam) – Two Node Element

Expression (A) v ( x ) = a0 + a1 x + a2 x 2 + a3 x3

Find a0 , a1 , a2 , a3 by imposing v ( 0 ) = U1 , v ′ ( 0) = U 2 , v ( L ) = U 3

and v ′ ( L) = U 4 .

a0 = U1 , a1 = U 2 , a2 =
1
L3
( −3LU1 − 2 L2U 2 + 3LU 3 − L2U 4 ) ,
1
a3 = ( 2U1 + LU 2 − 2U 3 + LU 4 )
L3

v ( x ) = U1 + U 2 x

3 (
−3LU1 − 2 L2U 2 + 3LU 3 − L2U 4 ) x 2
1
+
L
1
+ 3 ( 2U1 + LU 2 − 2U 3 + LU 4 ) x 3
L

Expression (B) v ( x ) = N U1 ( x ) U1 + N U 2 ( x ) U 2 + N U 3 ( x ) U 3 + N U 4 ( x ) U 4

U1 
U 
 
= { N U1 ( x ) N U 2 ( x ) N U 3 ( x ) N U 4 ( x )}  2  = N U ( x ) U
U 3 
U 4 

where

2 3 2
x x  x
N U1 ( x ) = 1 − 3   + 2   N U 2 ( x ) = x 1 − 
L L  L
(6.12)
2 3 2
x x x x
NU3 ( x ) = 3   − 2   NU4 ( x ) = − x   + x  
L  L L L

The polynomials in (6.12) are called the cubic hermitian polynomials.


6-9

1 1
N U1 ( x ) N U3 ( x )

x x
0 L 0 L

x
NU2 ( x ) 0 L
1
1 x NU4 ( x )
0 L

Figure 3 Shape functions of two-node Euler-Bernoulli beam element

6.2.6 Two-node Displacement-Based Bar Element

U1, N1 U2, N2

Figure 4 Two-node bar element

U  N 
Nodal Displacements: U =  1 ; Nodal Forces: P =  1
U 2  N2 

Section Displacements: u ( x ) = u0 ( x ) = N U ( x ) U

Shape Functions: N U ( x ) = { N U1 ( x ) N U 2 ( x )}

x
N U1 ( x ) = 1 −
L
x
NU2 ( x ) =
L

Section Deformations: ε ( x ) = { ε 0 ( x )} ; Section Forces: σ ( x ) = { N ( x )}


6-10

Section Constitutive Law (Eq. (6.2)): N ( x ) = EA( x) ε 0 ( x )

k ( x ) = EA( x)

From the compatibility condition:

d
ε ( x) = ε 0 ( x) = u0 ( x) = L ε u ( x)
dx

d
Lε =
dx

It follows from eq. (6.6) that:

B ( x ) = Lε N U ( x )

=
d
dx
{ N U1 ( x ) N U 2 ( x )}

d  x x
= 1 − 
dx  L L

B ( x) = { N U′ 1 ( x) N U′ 2 ( x)}

1
N U′ 1 ( x) = −
L
1
N U′ 2 ( x) =
L

Thus,

 1 1
B ( x ) = − 
 L L

The element forces P and the element stiffness matrix K are


6-11

P = ∫ B ( x) σ ( x) dx
T

1  −1
L

= ∫ L  1  N dx
0
L
 N U′ 1 ( x) 
= ∫  N ′ ( x)  N dx
0 U2

−1
Pbar = N   (6.13)
1
L

K bar = ∫ B ( x) k ( x) B ( x) dx
T

0
L
 N U′ 1 ( x ) 
= ∫  N ′ ( x)  EA ( x) { N ′ ( x)
0 U2
U1 N U′ 2 ( x )} dx

1 −1
L
1
= ∫   EA ( x ) { −1 1} dx
0
L1 L

For EA ( x ) constant over the length of the bar,

EA  1 −1
K bar =
L  −1 1 
(6.14)

The equilibrium expressed by Eq. (6.13) is illustrated in Figure 5.

P1 N N P2

Figure 5 Equilibrium between section and nodal forces

Observations:

(a) If the cross sectional area is constant and the material is linear elastic, the above

formulation is "exact". The linear shape functions for u0 ( x ) are the exact solutions of

the governing differential equation of the problem, namely EAu0′′ = 0 .


6-12

(b) If the cross sectional area varies along the bar and/or the material is not linear elastic,

the above formulation is not "exact". The linear shape functions for u0 ( x ) are

approximation of the exact solution of the differential equation of the problem. The

stiffness matrix is approximate.


6-13

6.2.7 Two-node Displacement-Based Euler-Bernoulli Beam Element (Flexure Only)

v1, V1 v2 , V 2
M1, θ1 M2, θ2

Figure 6 Two-node Euler-Bernoulli beam element

 v1  U1 
θ  U 
   
Nodal Displacements: U =  1 =  2
 v2  U 3 
θ 2  U 4 

 V1   P1 
 M  P 
   
Nodal Forces: P =  1 =  2
 V2   P3 
 M 2   P4 

Section Displacements: u = v0 ( x ) = N U ( x ) U

Shape Functions: N U ( x ) = { N U1 ( x ) N U 2 ( x ) N U 3 ( x ) N U 4 ( x )}

2 3
x x
N U1 ( x ) = 1 − 3   + 2  
L L
2
 x
N U 2 ( x ) = x 1 − 
 L
2 3
x x
N U3 ( x ) = 3   − 2  
L L
2
x x
NU4 ( x ) = − x   + x  
 L  L

Section Deformations: ε ( x) = {κ ( x )} ; Section Forces: σ ( x ) = { M ( x )}

Section Constitutive Law (Eq. (6.2)): M ( x ) = EI ( x) κ ( x )

k ( x ) = EI ( x)
6-14

From the compatibility condition:

d2
ε ( x) = κ ( x ) = v0 ( x) = L ε u ( x)
dx 2

d2
Lε =
dx 2
It follows from Eq. (6.6) that:

B ( x ) = Lε N U ( x )
d2
= 2 { N U1 ( x ) NU2 ( x) N U 3 ( x ) N U 4 ( x )}
dx

B ( x) = { N U′′1 ( x) N U′′2 ( x ) N U′′3 ( x) N U′′4 ( x)}

where

6 12
N U′′3 ( x) =
6 12
N U′′1 ( x ) = + x − x
L2 L3 L2 L3
4 6 2 6
N U′′2 ( x ) = − + 2 x N U′′4 ( x) = − + 2 x
L L L L

Thus, the element end (or nodal) forces P and the element stiffness matrix K are

P flexure = ∫ B ( x) σ ( x) dx
T

 N U′′1 ( x) 
L 
 N U′′2 ( x) 
= ∫  M ( x) dx
0  U3 ( ) 
N ′′ x
 N U′′4 ( x) 
 6 12 
 L2 + L3 x 
 
 4 6 
L − + x
 L L2 
= ∫  M ( x) dx
0
6 12 
− x
 L2 L3  (6.15)
 2 6 
− + 2 x 
 L L 
6-15

L
K flexure = ∫ BT ( x ) k ( x ) B ( x ) dx
0

 N U′′1 ( x ) 
L 
 N U′′ 2 ( x ) 
= ∫  EI ( x ) { N U′′1 ( x ) N U′′ 2 ( x ) N U′′ 3 ( x ) N U′′ 4 ( x )} dx
0  U3 ( )
N ′′ x
 N U′′ 4 ( x ) 
 
 6 12 
 L2 + L3 x 
 
 4 6 
− + x
 6 12   4 6  
L
 2
  6 12   2 6
= ∫  L L  EI ( x )  2 + 3 x   − + 2 x   2 − 3 x − + 2 x   dx
0
6 12 
− 3x  L L   L L  L L   L L 
L L 
2

 2 6 
− + 2 x 
 L L 

For EI ( x ) constant over the length of the beam,

 12 EI 6 EI 12 EI 6 EI 
 L3 −
L2 L3 L2 
 
 6 EI 4 EI 6 EI
− 2
2 EI 
 L2 L L L 
K flexure =
 − 12 EI 6 EI 12 EI 6 EI 
− 2 − 2 
 L3 L L3 L 
 6 EI 2 EI 6 EI 4 EI 
 − 2 
 L2 L L L 

Observations:

(a) if the cross section is constant (i.e., prismatic beam) and the material is linear elastic,

the above formulation is "exact". The cubic shape functions for v0 ( x ) are the exact

solutions of the governing differential equation of the problem, namely EIv0′′′′ = 0 .

(b) if the cross section varies over of the length of the beam and/or the material is not
linear elastic, the above formulation is not "exact". The cubic shape functions for

v0 ( x ) represent an approximation of the exact solution of the governing differential

equation of the problem, ( EIv0′′)′′ = 0 . The stiffness matrix is approximate.


6-16

6.2.7.1 Example of Tapered Beam (Flexure only)

P
2I0 I0

2 4
1 3

Figure 7 Euler-Bernoulli cantilever beam-column with variable moment of inertia

Consider the cantilever beam with variable moment of inertia shown in Figure 7. A single element is used

to model the beam response. The beam stiffness is computed as

K flexure = ∫ B ( x) k ( x) B ( x) dx
T

 6 12 
 L2 + L3 x 
 
 4 6 
L − + x
 L L2   6 12   4 6   6 12   2 6 
= ∫  EI ( x )  2 + 3 x  − + 2 x  2 − 3 x  − + 2 x  dx
0
6 12 
− 3x  L L   L L  L L L L 
L L 
2

 2 6 
− + 2 x 
 L L 
 L/2 L

= I 0  2 ∫ {⋮} E {⋯} dx + ∫ {⋮} E {⋯} dx 
 0 L/2 

Consider the cantilever beam with variable moment of inertia shown in the figure.
6-17

6.2.8 Two-Node Displacement-Based Euler-Bernoulli Beam Element (Axial + Flexure)

v1 v2
u1 u2

θ1 θ2

Figure 8 Two-node Euler-Bernoulli beam-column element

 u1  U1 
 v  U 
 1  2
θ  U 
Nodal Displacements: U =  1 =  3
u2  U 4 
 v2  U 5 
   
θ 2  U 6 

 N1   P1 
 V  P 
 1   2
 M   P 
Nodal Forces: P =  1 =  3
 N 2   P4 
 V2   P5 
   
 M 2   P6 

u ( x ) 
Section Displacements: u ( x) =  0  = NU ( x ) U
 v0 ( x ) 

Shape Functions:
N ( x) 0 0 NU2 ( x ) 0 0 
N U ( x ) =  U1 
 0 NU3 ( x ) NU4 ( x ) 0 NU5 ( x ) NU6 ( x )

where
6-18

2
 x
N U1 ( x ) = 1 −
x N U 4 ( x ) = x 1 − 
L  L
2 3
x x x
NU2 ( x ) = NU5 ( x ) = 3  − 2  
L  L L
2 3
x x x x
2
NU3 ( x ) = 1 − 3   + 2   NU6 ( x ) = − x   + x  
L L L L

ε ( x)  u ′ ( x) 
Section Deformations: ε ( x) =  0  =  0 
 κ ( x)   v0′′( x) 

 N ( x) 
Section Forces: σ ( x) =  
 M ( x ) 

 N ( x )   EA( x ) 0   ε ( x ) 
 =   
EI ( x )  κ ( x ) 
Section Constitutive Law (Eq. (6.2)):
 M ( x )   0

 EA( x) 0 
k ( x) = 
EI ( x) 
Section stiffness:
 0

From the compatibility condition:

d 
ε ( x )   dx
0 
u ( x) 
ε ( x) =  0  =    0  = L ε u ( x)
 κ ( x )   0 d   v0 ( x) 
2

 dx 2 

d 
 dx 0 
Lε =  
0 d2 
 dx 2 

It follows from Eq. (6.6) that:


6-19

B ( x ) = Lε N U ( x )
d 
0 
 dx N ( x) 0 0 NU2 ( x ) 0 0 
=   U1 
0
2
d  0 NU3 ( x ) NU 4 ( x ) 0 N U5 ( x ) N U 6 ( x )
 dx 2 

 N ′ ( x) 0 0 N U′ 2 ( x ) 0 0 
B ( x ) =  U1 
 0 N U′′3 ( x) N U′′4 ( x) 0 N U′′5 ( x ) N U′′6 ( x) 

where

1 4 6
N U′ 1 ( x) = − N U′′4 ( x) = − + x
L L L2
1 6 12
N U′ 2 ( x) = N U′′5 ( x) = 2 − 3 x
L L L
6 12 2 6
N U′′3 ( x) = 2 + 3 x N U′′6 ( x) = − + 2 x
L L L L

Thus, the element end (or nodal) forces P and the element stiffness matrix K are

 N U′ 1 ( x) 0 
 
 0 N U′′3 ( x) 
L L
 0 N U′′4 ( x)   N ( x) 
PEB − beam = ∫ BT ( x) σ ( x) dx = ∫    dx (6.16)
0  U2 ( )
0
N′ x 0   M ( x) 
 0 N U′′5 ( x) 
 
 0 N U′′6 ( x) 

K EB −beam = ∫ B ( x) k ( x) B ( x) dx
T

 N U′ 1 0 
 0 N U′′3 
  (6.17)
L
 0 N U′′4   EA( x ) 0   N U′ 1 0 0 N U′ 2 0 0 
= ∫  dx
N′
0  U2
0  0 EI ( x)   0 N U′′3 N U′′4 0 N U′′5 N U′′6 
 0 N U′′5 
 
 0 N U′′6 
6-20

For EI ( x ) constant over the length of the beam:

 EA EA 
 L 0 0 − 0 0 
L
 
 0 12 EI 6 EI 12 EI 6 EI 
0 −
 L3 L2 L3 L2 
 6 EI 4 EI 6 EI 2 EI 
 0 0 − 2
= 
L2 L L L 
K EB −beam 
EA EA
− 0 0 0 0 
 L L 
 12 EI 6 EI 12 EI 6 EI 
 0 − − 0 − 2 
 L3 L2 L3 L 
 6 EI 2 EI 6 EI 4 EI 
 0 0 − 2
L2 L L L 
6-21

6.2.9 Three-Node Displacement-Based Timoshenko Beam-Column Element

α1 α2 α3
v1 v2 v3
u1 u2 u3
1 2 3

Figure 9 Three-node Timoshenko beam-column element: degrees of freedom

 u1  U1 
 v  U 
 1   2
α1  U 3 
   
 u2  U 4 
   
Nodal Displacements: U =  v2  = U 5 
α  U 
 2  6
 u3  U 7 
 v  U 
 3   8
α 3  U 9 

 N1   P1 
 V  P 
 1   2
 M 1   P3 
   
 N 2   P4 
   
Nodal Forces: P =  V2  =  P5 
M   P 
 2  6
 N 3   P7 
 V  P 
 3   8
 M 3   P9 

The stiffness approach is based on the selection of approximate displacement fields along the element. In

the case of a Timoshenko beam, there are three independent fields: the axial displacement u0 ( x ) ,

transversal displacement v0 ( x ) , and cross-section rotation α 0 ( x ) . The three independent fields are
6-22

necessary, because the rotation α0 ( x) is not the first derivative of v0 ( x ) , as in the case of the Euler-

Bernoulli beam.

A two-node element would yield a linear approximation of u0 ( x ) , v0 ( x ) and α 0 ( x ) along the beam. This

approximation would be too crude. The next choice is a three-node element. This implies parabolic

displacement fields for u0 ( x ) , v0 ( x ) and α 0 ( x ) .

 u0 ( x ) 
 
Section Displacements: u ( x ) =  v0 ( x )  = N U ( x ) U
α ( x ) 
 0 

1
Shape Functions :

 N U1 ( x ) 0 0 NU 2 ( x ) 0 0 NU3 ( x ) 0 0 
 
NU ( x ) =  0 N U1 ( x ) 0 0 NU 2 ( x ) 0 0 NU3 ( x ) 0 
 0 0 N U1 ( x ) 0 0 NU2 ( x ) 0 0 N U 3 ( x ) 

where

x2 x
N U1 ( x ) = 2 2
− 3 +1
L L
2
x x
NU2 ( x ) = − 4 2 + 4
L L
2
x x
NU3 ( x ) = 2 2 −
L L

1
N1 (ξ ) = ξ (ξ − 1) N 2 (ξ ) = (1 − ξ 2 ) 1
N3 (ξ ) = ξ (ξ + 1)
2 2

1 1 1

-1 0 +1 ξ
1
6-23

ε 0 ( x )   u0′ ( x) 
   
Section Deformations: ε ( x ) =  κ ( x )  =  α 0′ ( x) 
 γ ( x )  v ′ ( x ) − α ( x ) 
   0 0 

 N ( x) 
 
Section Forces: σ ( x) =  M ( x ) 
 V ( x) 
 

Section Constitutive Law (Eq. (6.2)):

 N ( x )   EA( x ) 0 0  ε 0 ( x ) 
    
M ( x ) =  0 EI ( x ) 0  κ ( x) 
V ( x)   0 0 GAs ( x )   γ ( x ) 
  

 EA( x) 0 0 
 
Section stiffness: k ( x) = =  0 EI ( x) 0 
 0 0 GAs ( x ) 

From the compatibility condition:

d 
 0 0
ε 0 ( x )   dx   u0 ( x ) 
  d  
ε ( x) =  κ ( x)  =  0 0  v0 ( x)  = L ε u ( x)
 dx  
 γ ( x)    α 0 ( x) 
  d
0 −1 
 dx 

d 
 dx 0 0
 
d 
Lε =  0 0
 dx 
 d 
0 −1 
 dx 

It follows from Eq. (6.6) that:

B ( x ) = Lε N U ( x )
6-24

 N U′ 1 ( x) 0 0 N U′ 2 ( x) 0 0 N U′ 3 ( x) 0 0 
 
B ( x) =  0 0 N U′ 1 ( x) 0 0 N U′ 2 ( x) 0 0 N U′ 3 ( x) 
 0 N U′ 1 ( x) − N U1 ( x) 0 N U′ 2 ( x) − N U 2 ( x) 0 N U′ 3 ( x) − N U 3 ( x) 

where

4x 3
N U′ 1 ( x) = −
L2 L
8x 4
N U′ 2 ( x) = − 2 +
L L
4x 1
N U′ 3 ( x) = 2 −
L L

Thus, the element end (or nodal) forces P and the element stiffness matrix K are

L
PT −beam = ∫ BT ( x ) σ ( x ) dx (6.18)
0

 N U′ 1 ( x) 0 0 
 
 0 0 N U′ 1 ( x) 
 0 N U′ 1 ( x) − N U1 ( x ) 
 
L  U2 ( )   N ( x) 
N′ x 0 0
  
PT −beam = ∫ 0 0 N U′ 2 ( x)   M ( x)  dx
0
0 N U′ 2 ( x) − N U 2 ( x)   V ( x) 
 
 N U′ 3 ( x ) 0 0 
 0 0 N U′ 3 ( x) 
 
 0 N U′ 3 ( x) − N U 3 ( x) 

L
K T −beam = ∫ B ( x ) k ( x ) B ( x ) dx
T

0
6-25

Observations:

(a) The above formulation is approximate, even for constant cross section properties, since

parabolic distribution of the displacement fields do not satisfy the beam governing

differential equations.

(b) The element may suffer from "locking" problems. This phenomenon is studied in more

advanced classes (such as Finite Elements).


6-26

6.2.10 Summary of Displacement-based Approach

PVD

Strong Form
Compatibility

Lε u ( x ) = ε ( x )

Weak Form
Equilibrium
δ We + δ Wi = 0

or

L
δ U T P = ∫ δεT ( x ) σ ( x ) dx
0

Constitutive Law σ ( x) = k ( x) ε ( x)

Hence L
δ U P = ∫ δ uT ( x ) LTε k ( x ) L ε u ( x ) dx
T

FE approximation u ( x ) = NU ( x ) U

∫ N ( x ) L σ ( x ) dx
Element Equations
P = T
U
T
ε
0

 L T 
P = ∫ N U ( x ) LTε k ( x ) N U ( x ) dx  ⋅ U
 0 
6-27

6.2.11 General displacement-based formulation (Including initial effects and element loads)

Principle of Virtual Displacements

(a) Compatibility (enforced in the Strong Form)

Displacements: u ( x ) = NU ( x ) U

where u ( x ) = displacement fields along the element

N U ( x ) = shape functions

U = nodal displacements

Deformations: ε ( x) = B ( x) U (6.19)

where ε ( x ) = deformation fields inside element

B ( x ) = L ε N U ( x ) = derivatives of shape functions

(b) Section Constitutive Law

σ ( x ) = k ( x ) ( ε ( x ) − ε0 ( x )) (6.20)

where σ ( x ) = section generalized stresses

k ( x ) = section stiffness matrix

ε 0 ( x ) = section initial deformations (stress free)

(c) Equilibrium (enforced in the Weak Form through the Principle of Virtual Displacements)

L L

∫ δε ( x ) σ ( x ) dx = ∫ δ u ( x ) b ( x ) dx + δ U P
T T T
(6.21)
0 0

where b ( x ) = vector of distributed loads


6-28

L L

∫ δε ( x ) k ( x ) ( ε ( x ) − ε 0 ( x ) ) dx = ∫ δ u ( x ) b ( x ) dx + δ U P
T T T

0 0

L L L

∫ δε ( x ) k ( x ) ε ( x ) dx − ∫ δε ( x ) k ( x ) ε 0 ( x ) dx = ∫ δ u ( x ) b ( x ) dx + δ U P
T T T T

0 0 0

or

L L L

∫ δ U B ( x ) k ( x ) ε ( x ) dx − ∫ δ U B ( x ) k ( x ) ε 0 ( x ) dx = ∫ δ U N U ( x ) b ( x ) dx + δ U P
T T T T T T T

0 0 0

or

L L L

∫ B ( x ) k ( x ) ε ( x ) dx − ∫ B ( x ) k ( x ) ε ( x ) dx = ∫ N ( x ) b ( x ) dx + P
T T T
0 U
0 0 0

or

L L L

∫0 B ( x) k ( x) B ( x) dx ⋅ U − ∫0 B ( x) k ( x) ε 0 ( x) dx = ∫0 N U ( x) b ( x) dx + P
T T T

   


K Pi P0

Thus, K U − Pi = P + P0

∫ B ( x ) k ( x ) B ( x ) dx
T
Element Stiffness Matrix: K =
0

L
Element Initial Force Vector: Pi = ∫ BT ( x ) k ( x ) ε 0 ( x ) dx
0

L
P0 = ∫ N ( x ) b ( x ) dx
T
Element Equivalent Load Vector: U
0
6-29

6.2.12 Intertia term: consistent and lumped mass matrices

If the intertia terms are included in the beam element formulation, then the equilibrium equation becomes:

LTσ σ + b - ρ u
ɺɺ = 0

ɺɺ = {uɺɺ( x, y, z ) vɺɺ( x, y, z ) w
ɺɺ ( x, y, z )} are the accelerations in the
T
where ρ is the mass density and u

x, y, z directions and the double dots indicate double derivative with respect to time.

6.2.13 Displacement-Based Formulations

The FE formulations remain identical, except for the last step, equilibrium where an additional term must

be added

L L L

∫ δε ( x ) σ ( x ) dx = ∫ δ u ( x ) b ( x ) dx − ∫ δ u ( x ) ρ uɺɺ ( x ) dx + δ U
T T T T
P (6.22)
0 0

0
inertia term

The inertia term is expandes as

L L

∫ δ u ( x ) ρ uɺɺ ( x ) dx = δ U ∫ N U ( x ) ρ N U ( x ) dx Uɺɺ = δ U M Uɺɺ


T T T T

0 0

L
M = ∫ N ( x ) ρ N ( x ) dx
T
where U U is the element mass matrix
0

ɺɺ = second time derivatives of the nodal displacements.


U

With this new term, the element matrix equilibrium becomes;

ɺɺ − P = P + P
KU + MU (6.23)
i 0

If there are no initial strains or element loads, the element equation is written:

ɺɺ = P
KU + MU (6.24)
6-30

For a two-node, 2D Euler-Bernoulli beam the consistent mass matrix is:

140 0 0 70 0 0 
 0 156 22 L 0 54 −13L 

ρ AL  0 22 L 4 L2 0 13L −3L 
M=  
420  70 0 0 140 0 0 
 0 54 13L 0 156 −22 L 
 
 0 −13L −3L 0 −22 L 4 L2 
2

Sometimes the mass is lumped at the two end nodes to obtain the lumped mass matrix:

1 0 0 0 0 0
0 1 0 0 0 0 

ρ AL 0 0 0 0 0 0
M=  
2 0 0 0 1 0 0
0 0 0 0 1 0
 
0 0 0 0 0 0 

6.2.14 Mode frequencies and mode shapes

The unforced vibrations of a linear system:

ɺɺ = 0
KU + MU

The above equation has a solution that may be written as follows:

Ui = Φi sin (ωi t + α i ) ( i = 1, 2,..., n )

where n is the number of degrees of freedom. In this harmonic expression, Φi is a vector of nodal

amplitudes (the mode shape) the ith mode of vibration. The symbol ωi represents the angular frequency

of mode i, and αi denotes the phase angle.

ɺɺ = −ω 2 Φ sin (ω t + α )
U i i i i i

Thus:
6-31

( K − ω M ) Φ sin (ω t + α ) = 0
i
2
i i i

The above manipulation has the effect of separating the variable time from those of space, and we are left

with a set of n homogeneous algebraic equations. The term sin (ωit + α i ) can be canceled out, since it

cannot be equal to zero, thus

(K − ω M ) Φ = 0
i
2
i

The above equation has the form of the algebraic eigenvalue problem. From the theory of homogeneous

equations, nontrivial solutions exist only if the determinant of the matrix is equal to zero, thus

det ( K − ω i2M ) = 0

Expansion of this determinant yields a polynomial of order n called the characteristic equation. The n

roots ωi2 of this polynomial are the characteristic values, or eigenvalues. Substitution of these roots (one

at a time) into the homogeneous equations produces the characteristic vectors, or eigenvectors Φi (or

mode shapes) within arbitrary constants.

6.2.15 Example: Shear Building

m v1

h k1 = k2 /2

2m
v2

h k2=12EI/h^3

m 0   k -k 
M=  K= 
 0 2m  -k 3k 
6-32

KΦ = ω 2 M Φ

 1 -1 1 0 
k  Φ = ω 2
m 0 2 Φ
-1 3   

 1 -1 1 0  ω 2m
-1 3  Φ = λ 0 2  Φ with λ=
    k

1- λ -1
=0
-1 3 - 2λ

2λ 2 − 5λ + 2 = 0 characteristic equation

1
λ1,2 = , 2
2

ω12 m 1 k ω 22 m 2k
= ⇒ ω1 = =2 ⇒ ω2 =
k 2 2m k m

1 2 -1 -1 -1


 -1 2  Φ1 = 0 -1 -1 Φ2 = 0
   

φ11   1   φ21   1 
Φ1 =   =   Φ2 =   =  
φ12  1 2  φ22  −1

Φ1
1

1/2
6-33

Φ2
1

-1

Sometimes the mode shapes are normalized so that ΦiΤ M Φi = 1

In the present example:

m 0   1  3 m 0   1 
Φ1Τ M Φ1 = {1 1 2}    = m Φ2Τ M Φ2 = {1 −1}     = 3m
 0 2m  1 2  2  0 2m  −1

 2 3   1 3 
ˆ = Φ1 = m 
Φ  ˆ 2 = Φ2 = m 
Φ 
3 3m
1

m  1 6  − 1 3 
2
6-34

6.3 FORCE-BASED FORMULATIONS

Force-based formulations are not “naturally” fit for the FE method, because they yield the element

flexibility matrix rather than the element stiffness matrix. There are however some cases where it is worth

to compute the element flexibility matrix F (without rigid body modes) and to invert it to get the

corresponding element stiffness matrix K . This is particularly interesting in those instances where

displacement-based beam formulations are approximate and force-based formulations are exact (for

example tapered elements, material nonlinear elements).

6.3.1 Principle of Virtual Forces (Principle of Complementary Virtual Work)

a) Equilibrium (enforced in the Strong Form)

2
Section Forces : LTσ σ ( x ) = 0 (6.25)

where σ ( x) = forces fields inside element

Lσ = force differential operator

b) Section Constitutive Law

ε ( x) = f ( x) σ ( x) (6.26)

where f ( x) = section flexibility matrix

c) Compatibility (enforced in the Weak Form through the Principle of Virtual Forces)

L
δ P T U = ∫ δσT ( x ) ε ( x ) dx (6.27)
0

2
It is assumed that there are no element forces, that is b ( x) = 0
6-35

6.3.2 Finite Element Approximation

The section forces are written as a function of the element end forces (nodal forces):

σ ( x ) = NP ( x ) P (6.28)

Eq. (6.27) is further expanded:

δ P T U = ∫ δσT ( x ) ε ( x ) dx
0
L
= ∫ δ P T NTP ( x ) ε ( x ) dx
0
L
= δP T
∫ NTP ( x ) ε ( x ) dx
0

Thus,

L
U = ∫ NTP ( x ) ε ( x ) dx (6.29)
0

Eq. (6.29) is further expanded to yield the element flexibility matrix:

The following expression is typically used in FE derivations:

U = ∫ NTP ( x ) ε ( x ) dx
0
L
= ∫ NTP ( x ) f ( x ) σ ( x ) dx (6.30)
0
L
= ∫ NTP ( x ) f ( x ) N P ( x ) P dx
0

Thus,

L
U = ∫ N ( x ) f ( x ) N ( x ) dx
T
P P P
0

or

U = F P (6.31)
6-36

where F is the element flexibility matrix:

L
F = ∫ N ( x ) f ( x ) N ( x ) dx
T
P P (6.32)
0

The same formulation is obtained by applying the principle of Stationary Complementary Potential Energy.

Strong form
u ( x) b ( x)
Weak form

Compatibility Equilibrium
L L
LTσ σ + b = 0
∫ δσ ε dx - ∫ δ b u dx = 0
Τ T

0 0

Constitutive Law
ε ( x) ε = ε (σ) σ ( x)

Figure 10 Tonti’s diagram of Weak Form (Force-Based Formulation)


6-37

6.3.3 Two-Node Force-Based Euler-Bernoulli Beam-Column Finite Element

The element is formulated without rigid body modes, for reasons that will soon be clear. It is important to

recall that the selection of a system without rigid body modes (basic system) is not unique.

θ1 θ2
u

Element Deformations

Μ1 Μ2
N

Element Forces

Figure 11 Two-node beam-column element: deformations and forces without rigid body modes

θ1   M1 
   
Nodal Deformations: U = θ 2  ; Nodal Forces: P = M 2 
u  N 
   

ε ( x )   N ( x ) 
Section Deformations: ε ( x) =  0 ; Section Forces: σ ( x) =  
 κ ( x )   M ( x ) 

Section Constitutive Law: ε ( x) = f ( x)σ( x)

 1 
 EA( x) 0 
The simplest form of f ( x ) is: f ( x) =  .
 1 
 0 EI ( x) 

Shape Functions. The selection of the force shape functions stems from the requirement that

equilibrium, as expressed in eq. (6.25), be satisfied in the strong form. For the Euler-Bernoulli beam:

LΤσ σ ( x ) = 0
6-38

d 
0 
 dx  N ( x ) 
   = 0 (6.33)
0 d   M ( x ) 
2

 dx 2 

The satisfaction of eq. (6.33) yields:

N ( x ) = c1
M ( x ) = c2 x + c3

The three constants are found by imposing that (Figure 11):

N ( L) = N
M ( 0) = − M1
M ( L) = M 2

Thus,

N ( x) = N
x  x
M ( x ) =  − 1 M 1 +   M 2
L  L

or

 0 1  M1 
 N ( x )  
0
 = x   M 
 M ( x )   L − 1 L
x
0  
2

  N

With the notation of (6.28), we can write the matrix of force shape functions as

 0 0 1
N P ( x ) =  x  x
 (6.34)
 − 1 0
 L  L 

The element flexibility matrix is:


6-39

  x 
0  L − 1   1 
L
   EA( x) 0   0 0 1
 x    dx
F = ∫ 0  
L   x − 1 x (6.35)
0   1  0
  0  L  L 
1 0   EI ( x) 
 
 

Observations:

a) There is no approximation involved in any of the steps shown above, thus the element flexibility

matrix F is "exact" for a section tapered along the element and/or for a nonlinear material!

In the case of a constant cross section and a linear elastic material:

  x 
0  L − 1 
   1 
L
 EA 0   0 0 1
 x    dx
F = ∫ 0    x 
L 
x
0   0 1   − 1 0
  L  
EI  
 L
1 0 
 
 

 L L 
 3EI − 0 
6 EI
 
L L
F = − 0  (6.36)
 6 EI 3EI 
 
 0 L 
0
 EA 

b) The element flexibility matrix F can then be inverted to get the element stiffness matrix:

K = F −1 (6.37)

Inversion requires non-singularity of F , hence the formulation without rigid body modes.

c) The element forces are more difficult to derive. In the linear elastic case (of interest here), they

can be computed using the linear relation between forces and deformations, thus

P = KU (6.38)
6-40

d) Rigid body modes must then be added to get the 6x6 (in the 2D case) element stiffness matrix

K . The same procedure is applied to get the 6x1 element force vector P .

K = Γ TRBM K Γ RBM (6.39)

P = Γ ΤRBM P (6.40)

f) Special procedures must be followed to get the element forces in the nonlinear case.

6.3.4 2-Node Force-Based Timoshenko Beam-Column Finite Element

The element degrees of freedom are identical to that used for the Euler-Bernoulli beam of Figure 11.

θ1   M1 
   
Nodal Deformations: U = θ 2  ; Nodal Forces: P = M 2 
u  N 
   

ε 0 ( x )   N ( x) 
   
Section Deformations: ε ( x) =  κ ( x)  ; Section Forces: σ ( x ) = M ( x )
 γ ( x)   V ( x) 
   

Section Constitutive Law: ε ( x) = f ( x) σ ( x)

 1 
 0 0 
 EA( x) 
 1 
f ( x) =  0 0 
 EI ( x) 
 1 
 0 0 
 GAs ( x) 

Shape Functions. The selection of the force shape functions stems from the requirement that

equilibrium, as expressed by eq. (6.25), be satisfied in the strong form. For the Timoshenko beam,

LTσ σ ( x ) = 0
6-41

d 
 dx 0 0
   N ( x) 
0 d2  
 0   M ( x)  = 0 (6.41)
dx 2
   V ( x) 
0 d
−1
 dx 

The satisfaction of (6.33) yields:

N ( x ) = c1
M ( x ) = c2 x + c3

The three constants are found by imposing that (Figure 11):

N ( L) = N
M ( 0) = − M1
M ( L) = M 2

Thus,

N ( x) = N
x   x
M ( x) =  − 1 M 1 +   M 2
L   L
 1  1
V ( x) =   M 1 +   M 2
 L  L

or

 
 0 0 1
 N ( x)     M1 
   x  x   
 M ( x )  =  − 1 0 M 2 

 V ( x )   L  L
  N 
 
 1 1
0
 L L 

With the notation of (6.28), we can write the matrix of force shape functions as:
6-42

 
 0 0 1
 
x 
N P ( x) =  − 1 0
x
(6.42)
 L  L 
 
 1 1
0
 L L 

The element flexibility matrix is:

L
F = ∫ NTP ( x ) f ( x ) N P ( x ) dx
0

 x 1  1   
0 −1  0 0 
L L  EA( x)   0 0 1
L   
x 1  1   x −1 x
F = ∫ 0  0 0  0  dx (6.43)
 L L EI ( x) L L 
1  
0 
0
  1 
 0 1  
1
0
   0 0   L L 
 GAs ( x) 

Observations:

a) There is no approximation involved in any of the steps shown above, thus, also in the case of

the Timoshenko beam, the element flexibility matrix F is "exact" for a section varying along the

element and/or for a nonlinear material!

b) In the case of a constant cross section and linear elastic material:

 L 1 L 1 
 3EI + L GA − +
6 EI L GAs
0 
 s

 L 1 L 1 
F = − + + 0  (6.44)
6 EI L GAs 3EI L GAs
 
 L 
 0 0 
 EA 
6-43

c) The element stiffness matrix K and the element force vector P without rigid body modes are

found using (6.37)and (6.38). The element stiffness matrix K and the element force vector P

are found using (6.39) and (6.40).

d) There is no approximation involved in any of the steps shown above, thus, also in the case of

the Timoshenko beam, the element flexibility matrix F is "exact" for a nonlinear section and/or

for a nonlinear material!


6-44

6.3.5 Summary of Force-Based Approach

PVF

Weak Form
Compatibility

δ We* + δ Wi * = 0

or

L
δ P T U = ∫ δσT ( x ) ε ( x ) dx
0

Strong Form
Equilibrium

Lσ σ ( x ) + b ( x ) = 0

Const. Law ε ( x) = f ( x) σ ( x)

Hence L
δ P T U = ∫ δσ T ( x ) f ( x ) σ ( x ) dx
0

FE approx.s σ ( x ) = NP ( x ) P

∫ N ( x ) ε ( x ) dx
Element EQs.
U = T
P
0

L
U = ∫ N ( x ) f ( x ) N ( x ) dx
T
P P P
0
6-45

6.4 SUMMARY OF DISPLACEMENT-BASED AND FORCE-BASED APPROACHES

PVD PVF

Strong Form Weak Form


Compatibility

Lε u ( x ) = ε ( x ) δ We* + δ Wi* = 0

or

L
δ P U = ∫ δσT ( x ) ε ( x ) dx
T

Weak Form Strong Form


Equilibrium
δ We + δ Wi = 0 Lσ σ ( x ) + b ( x ) = 0

or

δ U T P = ∫ δεT ( x ) σ ( x ) dx
0

Const. Law σ ( x) = k ( x) ε ( x) ε ( x) = f ( x) σ ( x)

Hence L L
δU P = ∫δ u
T T
( x) L T
ε k ( x ) L ε u ( x ) dx δ P U = ∫ δσT ( x ) f ( x ) σ ( x ) dx
T

0 0

FE approx.s u ( x ) = NU ( x ) U σ ( x ) = NP ( x ) P

L L

∫ N ( x ) L σ ( x ) dx ∫ N ( x ) ε ( x ) dx
Element EQs.
P = T
U
T
ε U = T
P
0 0

L L
P = ∫ N U ( x ) L ε k ( x ) N U ( x ) dx U U = ∫ N ( x ) f ( x ) N ( x ) dx
T T T
P P P
0 0

You might also like