Sei sulla pagina 1di 25

Introduction to Timoshenko Beam Theory

Aamer Haque

Abstract
Timoshenko beam theory includes the effect of shear deformation which is ignored in Euler-Bernoulli beam theory. An
elementary derivation is provided for Timoshenko beam theory. Energy principles, the stiffness matrix, and Green’s functions
are formulated. Solutions are provided for some common beam problems.

1. Formulation

1.1. Introduction
The goal of beam theory is to simplify the equations of solid mechanics to beams. All loads on a beam act parallel to
an axis that is transverse to its longitudinal axis. The length of a beam is much larger than its width and depth. Beam
theory provides equations for the deflection and internal forces of the beam. Figure 1 shows a typical deflection of a beam
that is loaded from above. We assume that the initial unloaded beam’s longitudinal axis coincides with the x axis. The sign
convention for the deflection u is that positive deflections have positive y values. The deflection u and angle of deflection θ
are assumed to be small. Applied loads are considered positive if they act in the direction of the positive y axis.
Euler-Bernoulli beam theory is extensively applied to structural analysis and design. Most design guides and manuals will
exclusively use Euler-Bernoulli beams. The derivation of Euler-Bernoulli beam theory is given in: Hibbeler [3, 4], Hjelmstad
[5], Kassimali [7], Krenk [9].
The problem with Euler-Bernoulli beam theory is that it is inaccurate for deep beams. Deep beams are ones in which the
depth is not negligible compared to the length. Either complete solid mechanics or a more accurate beam theory is required
for deep beams. Timoshenko beam theory is a simple extension to Euler-Bernoulli beam theory. Shear deformations, which
are absent in Euler-Bernoulli beam theory, are included in Timoshenko beam theory.
The basic assumptions for Timoshenko beam theory are:
• The longitudinal axis of the unloaded beam is straight.
• All applied loads act transverse to the longitudinal axis.
• All deformations and strains are small.
• Hooke’s law can be used to relate stresses and strains.

• Plane cross sections, which are initially normal to the longitudinal axis, will remain plane after deformation.
The last requirement differs from Euler-Bernoulli theory which requires that the plane cross sections also remain normal to
the beam axis after deformation. For Timoshenko beams, plane cross sections will rotate due to shear forces.

Figure 1: Deflection of a beam

Preprint submitted to Elsevier October 26, 2016


1.2. Differential Equations

Figure 2: Forces on a beam segment

This section briefly presents an elementary derivation of the equations of Timoshenko beam theory. A complete derivation
from the equations of solid mechanics is provided in Hjelmstad [5] and Wang et al. [16]. Figure 2 shows a small segment of
the beam. The shear force V and bending moment M are displayed as positive in the figure according to the usual beam
convention. The load w is positive when applied upwards. We assume that the applied load is approximately constant for
the small beam segment shown in figure 2. Vertical equilibrium requires:
V + dV = V + w dx
dV
= w (1.1)
dx
For purposes of static equilibrium, the constant load w can be assumed to act as a point load in the middle of the segment.
Equilibrium of moments results in:
w
M + dM = M + V dx + (dx)2
2
dM
= V (1.2)
dx

Figure 3: Pure bending deformation

The next step in deriving Timoshenko beam theory is to relate the internal forces V and M to deformation. Consider
the pure bending of the beam segment shown in figure 3. Pure bending means that the segment is bent only by constant
end moments M . The y coordinate is measured from the neutral axis of the beam and is positive upwards. The neutral axis
does not change length during the deformation of the beam. The original length of the beam segment is ds. Let dθ be the
angle between the planar cross sections at the ends of the beam segment. The curvature κ is defined through the equation:
1 dθ
κ= =
R ds
The elongation of a longitudinal fiber in the beam is given by:
ds′ = (R − y) dθ = (R − y)κ ds
Since strains are assumed to be small, we shall use the engineering strain:
ds′ − ds
ε= = −κy
ds

2
Notice that the strain is zero at the neutral axis. Hooke’s law gives the stress-strain relation:

σ = Eε = −κEy

where σ is the normal stress at the end planes of the segment and E is Young’s modulus. The bending moment M is related
to normal stress σ through: ˆ ˆ
M = − σy dA = κE y 2 dA = κEI

For beam theory, we need to relate the bending moment M to the deflection u. This is accomplished by substituting the
calculus definition of curvature:
d2 u
M dx2 d2 u
=κ= h ≈ 2 (1.3)
EI du 2
 i3/2 dx
1 + dx
du
The approximation is valid because dx is assumed to be small and hence its square is smaller.

Figure 4: Pure shear deformation

Now consider the case of pure shear shown in figure 4. The constitutive relation between shear V and angle of shear γ is:

V = GAs γ (1.4)

where G is the shear modulus and As is the shear area. As is usually not equal to the area of the cross section because the
distribution of shear stress is not constant along the cross section.
Notice that shear deformation does not result in longitudinal strain. Also remember that bending deformation did not
cause any shearing strain. Thus bending and shearing deformations are separable and the total angle of deformation is:
du
=θ−γ (1.5)
dx
Using equations (1.1)-(1.5), we arrive at the equations for Timoshenko beam theory:

V ′ (x) = w(x) (1.6)



M (x) = V (x) (1.7)
M (x)
θ′ (x) = (1.8)
EI(x)
V (x)
u′ (x) = θ(x) − (1.9)
GAs (x)

These equations can be reduced to the pair of differential equations for 0 < x < L:
   
d dθ du
EI = GAs θ − (1.10)
dx dx dx
  
d du
GAs θ − = w (1.11)
dx dx

These equations imply that θ and u are two unknown functions whose solutions are to be determined. This is in contrast to
Euler-Bernoulli beam theory where a single 4th order differential equation can be written for the displacement u. Equations
(1.10) and (1.11) are used to formulate variational and energy principles in the next section.

3
Figure 5: Beam deflection using dimensionless distance ξ

Figure 6: Internal beam forces using dimensionless distance ξ

The equations (1.6)-(1.9) can be integrated to yield the following solutions:


ˆ x
V (x) = w(s) ds + V0 (1.12)
ˆ0 x
M (x) = V (s) ds + M0 (1.13)
ˆ0 x
M (s)
θ(x) = ds + θ0 (1.14)
EI(s)
ˆ0 x ˆ x
V (s)
u(x) = θ(s) ds − ds + u0 (1.15)
0 0 GA s (s)

The constants of integration V0 , M0 , θ0 , and u0 are determined using the boundary conditions for the particular problem
being solved.
The solution to the beam equations is more conveniently written using the dimensionless spatial variable ξ = x/L. We
then reformulate functions f (x) as f (ξ). Since this involves reformulation of the function instead of mere replacement of the
dependent variable, we notate this process as: f (x) → f (ξ). Integrals of f (x) are transformed in the following manner:
s
τ= , L dτ = ds
L
ˆ x ˆ ξ
f (s) ds → L f (τ ) dτ (1.16)
0 0

Using (1.16), the solutions (1.12)-(1.15) are now written as:


ˆ ξ
V (ξ) = L w(τ ) dτ + V0 (1.17)
0
ˆ ξ
M (ξ) = L V (τ ) dτ + M0 (1.18)
0
ξ
M (τ )
ˆ
θ(ξ) = L dτ + θ0 (1.19)
0 EI(τ )
ˆ ξ ˆ ξ
V (τ )
u(ξ) = L θ(τ ) dτ − L dτ + u0 (1.20)
0 0 GA s (τ )

Henceforth we shall assume prismatic beams with constant EI and GAs .

4
2. Energy Principles

2.1. Principle of Virtual Work


2.1.1. Formulation
The pair of differential equations (1.10) and (1.11) are the starting point for developing energy principles for Timoshenko
beams. We first multiply (1.10) by a function φ and integrate by parts:
ˆ L ˆ L
′ ′
(EIθ ) φ dx = GAs (θ − u′ ) φ dx
0 0
ˆ L ˆ L
L
[EIθ′ φ]|0 − EIθ′ φ′ dx = GAs (θ − u′ ) φ dx
0 0

Remember that M = EIθ′ . Moving all terms to the left hand side of the equation we have:
ˆ L ˆ L
′ ′ L
EIθ φ dx + GAs (θ − u′ ) φ dx − [M φ]|0 = 0 (2.1)
0 0

φ is interpreted as a virtual angle of rotation of the plane cross section of the beam.
We now multiply (1.11) by a function v and integrate by parts:
ˆ L ˆ L
′ ′
[GAs (θ − u )] v dx = wv dx
0 0
ˆ L ˆ L
L
[GAs (θ − u′ ) v]|0 − GAs (θ − u′ ) v ′ dx = wv dx
0 0

Notice that V = GAs (θ − u′ ). Again moving all terms to the left hand side:
ˆ L ˆ L
L
GAs (θ − u′ ) v ′ dx + wv dx − [V v]|0 = 0 (2.2)
0 0

For this equation, v represents a virtual vertical displacement of the beam.


Setting equations (2.1) and (2.2) equal to each-other and rearranging:
ˆ L ˆ L ˆ L
L L
EIθ′ φ′ dx + GAs (θ − u′ ) (φ − v ′ ) dx = wv dx + [M φ]|0 − [V v]|0 (2.3)
0 0 0
The left hand side of this equation is internal virtual work and the right hand side is external virtual work. Now let
 T  T
u = θ u  and v = φ v be the actual and virtual generalized displacements. Also define the generalized force
T
vector f = M −V . Then we can write the Principle of Virtual Work:

δWint (u, v) = δWext (v) (2.4)

Internal virtual work δWint and external virtual work δWext are defined as:
ˆ L  
T EI 0
δWint (u, v) = ũ ṽ dx (2.5)
0 0 GAs
ˆ L
0 w v dx + f T (L) v(L) − f T (0) v(0)
 
δWext (v) = (2.6)
0
T
where ũT =
  
θ′ θ − u′ and ṽ = φ′ φ − v′ .

2.1.2. Total Potential Energy


The total potential energy for Timoshenko beams is defined as:
1
Π(u) =
δWint (u, u) − δWext (u) (2.7)
2
The mathematical theory of transforming from virtual work equations to total potential energy is given in Hjelmstad [5],
Johnson [6].

5
2.1.3. Boundary Conditions

Essential Natural

u = value, V =? or u =?, V = value

θ = value, M =? or θ =?, M = value

Table 1: Essential vs natural boundary conditions

The boundary terms in equation (2.3) provide the essential and natural boundary conditions. Either M or θ can be
specified at the end points but both cannot be specified simultaneously. The same argument applies to V and u. In the
case that θ or u is specified, the corresponding virtual displacement is not permitted to vary. The possibilities for boundary
conditions are summarized in table 2.

Location Type Enforced Value Virtual Displacement


x=0 Essential u(0) = u0 v(0) = 0
x=0 Natural V (0) = V0 v(0) is free
x=0 Essential θ(0) = θ0 φ(0) = 0
x=0 Natural M (0) = M0 φ(0) is free
x=L Essential u(L) = uL v(L) = 0
x=L Natural V (L) = VL v(L) is free
x=L Essential θ(L) = θL φ(L) = 0
x=L Natural M (L) = ML φ(L) is free

Table 2: Essential and natural boundary conditions

It should be noted that some combinations of boundary conditions are not permissible since they would violate the
existence and uniqueness principles of the differential equations. These issues will not be discussed here but are found in
mathematical texts on boundary value problems.

2.1.4. Function Spaces


Solution and variational spaces must be specified so that the integrals in equation (2.3) exist and the boundary conditions
are satisfied. If the actual displacement u coincides with the virtual displacement v, then we see that both the displacement
function and its derivative are required to be square-integrable. The set of square-integrable vector-valued functions v(x) =
 T
φ(x) v(x) on [0, L] is denoted as:
( ˆ )
L ˆ L
2 2 2 2
L = v : [0, L] → R [φ(x)] dx < ∞ and [v(x)] dx < ∞
0 0

 T
The set of functions v(x) in L2 with first derivatives v′ (x) = φ′ (x) v ′ (x) also in L2 is called:

H1 = v : [0, L] → R2 v ∈ L2 and v′ ∈ L2


The subset of H1 which satisfies the essential boundary conditions is notated as Hbc 1
. The actual displacements u =
 T 1
θ u come from Hbc . The virtual displacements v must be zero at essential boundary conditions. The function space
1 1 1
for virtual displacements is denoted as Hbc0 . Note that Hbc and Hbc0 will be depend on the boundary conditions of the
1 1
particular beam problem. We always have the following set relations: Hbc0 ⊆ Hbc ⊆ H1 .
For the example of a beam clamped at x = 0 and prescribed displacement u(L) = uL at x = L, we have:
1

= u : [0, L] → R u ∈ H1 , θ(0) = 0, u(0) = 0, u(L) = uL

Hbc
1

= v : [0, L] → R v ∈ H1 , φ(0) = 0, v(0) = 0, v(L) = 0

Hbc0

6
2.1.5. Variational Problem
The Principle of Virtual Work states that the solution u to the Timoshenko beam problem has the property that internal
virtual work equals external virtual work for all admissible virtual displacements.
1 1
Find u ∈ Hbc such that δWint (u, v) = δWext (v) ∀v ∈ Hbc0 (2.8)

2.1.6. Minimization Problem


The Principle of Minimization of Total Potential Energy states that the solution u to the Timoshenko beam problem is
the one that minimizes the total potential energy as defined by equation (2.7).
1 1
Find u ∈ Hbc such that Π(u) ≤ Π(v) ∀v ∈ Hbc (2.9)

2.1.7. Concentrated Applied Moments and Forces


Suppose that N concentrated moments Mi and point loads pi are applied at xi ∈ (0, L). We allow for the possibility that
the applied moment or point load could be zero at xi . The interval [0, L] is partitioned into subintervals:
N[
+1
[0, L] = [xi−1 , xi ]
i=1

where x0 = 0 and xN +1 = L. We included the end points of the beam in order to simplify the book-keeping. The Principle
of Virtual Work (2.3) for each segment of the beam is:
ˆ xi ˆ xi ˆ xi
′ ′
EIθ φ dx + ′ ′
GAs (θ − u ) (φ − v ) dx = wv dx + [M φ]|xxii−1 − [V v]|xxii−1
xi−1 xi−1 xi−1

Summing over all segments, we have:


ˆ L ˆ L ˆ L
EIθ′ φ′ dx + GAs (θ − u′ ) (φ − v ′ ) dx = wv dx + M (L) φ(L) − M (0) φ(0) − V (L) v(L) + V (0) v(0)
0 0 0
N
X
M (x+ +
 − −

− i ) φ(xi ) − M (xi ) φ(xi )
i=1
N
X
V (x+ +
 − −

+ i ) v(xi ) − V (xi ) v(xi ) (2.10)
i=1

where the plus and minus superscripts indicate right


 and left hand limits respectively. We associate concentrated moments
with jumps in moment: Mi = − M (x+ . Similarly, we equate point loads with jumps in shear: pi = V (x+
 −
i ) − M (xi ) i )−

V (xi ). The sign convention for applied concentrated forces is as follows: Mi is positive counter-clockwise and pi is positive
upwards. Equation (2.10) allows for discontinuities in internal forces and virtual displacements.
If all virtual displacements are continuous, then φ(xi ) = φ(x+ − + −
i ) = φ(xi ) and v(xi ) = v(xi ) = v(xi ). The Principle of
Virtual Work simplifies to:
ˆ L ˆ L ˆ L
′ ′ ′ ′
EIθ φ dx + GAs (θ − u ) (φ − v ) dx = wv dx + M (L) φ(L) − M (0) φ(0) − V (L) v(L) + V (0) v(0)
0 0 0
N
X N
X
+ Mi φ(xi ) + pi v(xi ) (2.11)
i=1 i=1

7
2.2. Principle of Complementary Virtual Work
2.2.1. Formulation
Let δM (x) be a virtual moment distribution and δV (x) be a virtual shear distribution for the beam. We restrict these
distributions to be ones that are in static equilibrium with the virtual applied load δw(x). Multiply both sides of equation
(1.8) by δM and integrate by parts:
L L
M δM
ˆ ˆ
θ′ δM dx = dx
0 0 EI
L L
M δM
ˆ ˆ
L ′
[δM θ]|0 − θ δM dx = dx
0 0 EI
L L
M δM
ˆ ˆ
L
dx − [δM θ]|0 + θ δM ′ dx = 0 (2.12)
0 EI 0

Now multiply equation (1.9) by δV , integrate by parts, and use the fact that δV = δM ′ and δV ′ = δw to get:
L L
V δV
ˆ ˆ

(θ − u )δV dx = dx
0 0 GAs
L L L
V δV
ˆ ˆ ˆ
θ δV dx − u′ δV dx = dx
0 0 0 GAs
L L L
V δV
ˆ ˆ ˆ
θ δM ′ dx − u′ δV dx = dx
0 0 0 GAs
L L L
V δV
ˆ ˆ ˆ
L
θ δM ′ dx − [δV u]|0 + u δV ′ dx = dx
0 0 0 GAs
L L L
V δV
ˆ ˆ ˆ
L
θ δM ′ dx − [δV u]|0 + δw u dx = dx
0 0 0 GAs
L L L
V δV
ˆ ˆ ˆ
L
θ δM ′ dx − [δV u]|0 + δw u dx − dx = 0 (2.13)
0 0 0 GAs
Set equation (2.12) equal to equation (2.13) and rearrange terms:
L L L
M δM V δV
ˆ ˆ ˆ
dx + dx = δw u dx + [δM θ]|L L
0 − [δV u]|0 (2.14)
0 EI 0 GAs 0

The left hand side is complementary internal virtual work and the right hand side is complementary external virtual work.
 T  T
Equation (2.14) can be written as using matrix and vector notation. Let p = M V and q = δM δV . Then
the Principle of Complementary Virtual Work is:
∗ ∗
δWint (p, q) = δWext (q) (2.15)
∗ ∗
where complementary virtual internal work Wint (p, q) and complementary virtual external work Wext (q) are:
L  1

0
ˆ

Wint (p, q) = pT EI
1 q dx
0
0 GAs
ˆ L
u dx + qT (L) u(L) − qT (0) u(0)

 
Wext (q) = 0 δw
0
 T
where u = θ u is the actual displacement field. Remember that p is related to u through the differential equations
(1.6)-(1.9).

2.2.2. Complementary Total Potential Energy


The complementary total potential energy for the Timoshenko beam is defined as:
1
Π∗ (p) = ∗
δWint ∗
(p, p) − δWext (p) (2.16)
2

8
2.2.3. Function Spaces
Suppose the virtual forces coincide with the actual forces in equation (2.14). Then the integrals exist only if the forces
are square-integrable functions. However, the virtual forces must also be in equilibrium. We denote the set of virtual forces
as L2eq . The actual forces are derived from displacements which satisfy the essential boundary conditions. The set of actual
forces is notated as L2eqbc . Note the relationship between the sets is: L2eqbc ⊆ L2eq ⊆ L2 . The function spaces are determined
by the particular problem being solved. Unfortunately, the L2eq spaces are more cumbersome to specify than the Hbc 1
spaces of
virtual work methods. This fact helps to explain why complementary virtual work is seldom mentioned in the mathematical
theory of continuum mechanics.

2.2.4. Variational Problem


The Principle of Complementary Virtual Work states that the solution u to the Timoshenko beam problem has the
property that complementary internal virtual work equals complementary external virtual work for all admissible virtual
forces.

Find p ∈ L2eqbc such that δWint


∗ ∗
(p, q) = δWext (q) ∀q ∈ L2eq (2.17)

2.2.5. Minimization Problem


The Principle of Minimization of Total Complementary Potential Energy states that the solution p to the Timoshenko
beam problem is the one that minimizes the total complementary potential energy as defined by equation (2.16).

Find p ∈ L2eqbc such that Π∗ (p) ≤ Π∗ (q) ∀q ∈ L2eqbc (2.18)

2.2.6. Concentrated Virtual Moments and Forces


We now allow for the possibility of N concentrated virtual moments δMi and virtual point loads δpi applied at xi ∈ (0, L).
At any location, either of these virtual forces could be zero. The interval [0, L] is partitioned into subintervals:
N[
+1
[0, L] = [xi−1 , xi ]
i=1

where x0 = 0 and xN +1 = L. The Principle of Complementary Virtual Work (2.14) for each segment is:
ˆ xi ˆ xi ˆ xi
M δM V δV
dx + dx = δw u dx + [δM θ]|xxii−1 − [δV u]|xxii−1
xi−1 EI xi−1 GA s xi−1

Summing up all the segments:


L L L
M δM V δV
ˆ ˆ ˆ
dx + dx = δw u dx + δM (L) θ(L) − δM (0) θ(0) − δV (L) u(L) + δV (0) u(0)
0 EI 0 GAs 0
N
X
δM (x+ +
 − −

− i ) θ(xi ) − δM (xi ) θ(xi )
i=1
N
X
δV (x+ +
 − −

+ i ) u(xi ) − δV (xi ) u(xi )
i=1

Concentrated virtual moments are associated with jumps in virtual moment: δMi = − δM (x+
 −

i ) − δM (xi ) . We also equate
virtual point loads with jumps in virtual shear: δpi = δV (x+ −
i ) − δV (xi ). The actual displacements θ and u are continuous.
Thus the Principle of Complementary Virtual Work becomes:
L L L
M δM V δV
ˆ ˆ ˆ
dx + dx = δw u dx + δM (L) θ(L) − δM (0) θ(0) − δV (L) u(L) + δV (0) u(0)
0 EI 0 GAs 0
N
X N
X
+ δMi θ(xi ) − δpi u(xi ) (2.19)
i=1 i=1

9
3. Stiffness Matrix
3.1. Derivation
3.1.1. Beam element

Figure 7: Sign convention for the beam element

Matrix structural analysis requires stiffness matrices for every structural member. Only the derivation of the stiffness
matrix is described in this paper. Details of matrix structural analysis are found in: Harrison [2], Kassimali [8], McCormac
[10], McGuire et al. [11]. The Timoshenko beam element is shown in figure 7. The displacements and end forces are positive
in the direction of the positive y axis. End rotations and moments are positive if they are counter-clockwise. This sign
convention is demonstrated in figure 7.
The stiffness matrix K relates the beam end displacements u to the applied loads f :

Ku = f (3.1)
T
θ1 ]T and f = Fu0 FM0 Fu1 FM1

where u = [ u0 θ0 u1 . Equation (3.1)is explicitly written as:
    
K11 K12 K13 K14 u0 Fu0
 K21 K22 K23 K24   θ0 
 =  FM0 
 
 
 K31 K32 K33 K34   (3.2)
u1   Fu1 
K41 K42 K43 K44 θ1 FM1
The most direct way of computing the elements of K is to consider the meaning of the j-th column. Kij is the end force fi
that is required to maintain the unit displacement uj = 1 while keeping all other displacements at zero. Thus the procedure
to compute the j-th column of K is as follows:
1. Solve the beam equations (1.17)-(1.20) with boundary conditions:
(
1 if i = j
ui = δij =
0 if i 6= j

2. Compute the end shears V0 , V1 and end moments M0 , M1 from the solution of the beam equations.
3. Translate from the beam sign convention to the element sign convention using:

K1j = V0 , K2j = −M0 , K3j = −V1 , K4j = M1

This procedure will be applied in the next several pages of this paper.

3.1.2. Beam equations


The general solutions of the beam equations are required in order to compute the stiffness matrix. Since w(ξ) = 0, the
solution of the beam equations (1.17)-(1.20) are:

V (ξ) = V0 (3.3)
M (ξ) = LV0 ξ + M0 (3.4)
L2 V0 2 LM0
θ(ξ) = ξ + ξ + θ0 (3.5)
2EI EI
L3 V0 3 L2 M0 2
 
V0
u(ξ) = ξ + ξ + L θ0 − ξ + u0 (3.6)
6EI 2EI GAs
Boundary conditions are used to determine the values of the constants V0 , M0 , θ0 , and u0 .
We also define the parameter Φ which will be used in forming the stiffness matrix.
12EI
Φ= (3.7)
L2 GAs

10
3.1.3. Unit displacement at left end

Figure 8: Unit displacement u0 = 1

Left BC: u(0) = 1, θ(0) = 0; Right BC: u(1) = 0, θ(1) = 0


The left end boundary conditions give: u(0) = 1 ⇒ u0 = 1 and θ(0) = 0 ⇒ θ0 = 0.
Solving for M0 in (3.5) with θ(1) = 0, we have:
LV0
M0 = −
2
Substituting this expression into (3.6) with u(1) = 0:

L3 V0 L3 V0 LV0
0 = − − +1
6EI 4EI GAs
 3 
L L
0 = − + V0 + 1
12EI GAs
L3
 
12EI
1+ 2 V0 = 1
12EI L GAs
12EI
V0 = (1 + Φ)−1
L3
The end moments are computed to be:
6EI
M0 = − (1 + Φ)−1
L2
6EI
M1 = (1 + Φ)−1
L2
Translating from beam sign convention to matrix sign convention, the first column of the stiffness matrix is:
     
K11 V0 12
 K21   −M0  EI  6L 
 K31  = 
   =  
−V1  (1 + Φ)L3  −12 
K41 M1 6L

11
3.1.4. Unit rotation at left end

Figure 9: Unit displacement θ0 = 1

Left BC: u(0) = 0, θ(0) = 1; Right BC: u(1) = 0, θ(1) = 0


The left end boundary conditions give: u(0) = 0 ⇒ u0 = 0 and θ(0) = 1 ⇒ θ0 = 1.
Solving for M0 in (3.5) with θ(1) = 0, we have:

LV0 EI
M0 = − −
2 L
Substituting this expression into (3.6) with u(1) = 0:

L3 V0 L3 V0 L LV0
0 = − − +L−
6EI 4EI 2 GAs
 2 
L 1 1
0 = − + V0 +
12EI GAs 2
L2
 
12EI 1
1+ 2 V0 =
12EI L GAs 2
6EI −1
V0 = (1 + Φ)
L2
This value is used to compute M0 :
3EI EI
M0 = − (1 + Φ)−1 −
L L
3EI −1 EI
M0 = − (1 + Φ) − (1 + Φ)(1 + Φ)−1
L L
EI
M0 = − (4 + Φ)(1 + Φ)−1
L
M1 is then calculated to be:

M1 = LV0 + M0
6EI EI
M1 = (1 + Φ)−1 − (4 + Φ)(1 + Φ)−1
L L
EI
M1 = (2 − Φ)(1 + Φ)−1
L
Translating from beam sign convention to matrix sign convention, the second column of the stiffness matrix is:
     
K12 V0 6L
 K22   −M0  EI  (4 + Φ)L2 
 = =  
 K32   −V1  (1 + Φ)L3  −6L 
K42 M1 (2 − Φ)L2

12
3.1.5. Unit displacement at right end

Figure 10: Unit displacement u1 = 1

Left BC: u(0) = 0, θ(0) = 0; Right BC: u(1) = 1, θ(1) = 0


The left end boundary conditions give: u(0) = 0 ⇒ u0 = 0 and θ(0) = 0 ⇒ θ0 = 0.
Solving for M0 in (3.5) with θ(1) = 0, we have:
LV0
M0 = −
2
Substituting this expression into (3.6) with u(1) = 1:

L3 V0 L3 V0 LV0
1 = − −
6EI 4EI GAs
 3 
L L
1 = − + V0
12EI GAs
L3
 
12EI
1 = − 1+ 2 V0
12EI L GAs
12EI
V0 = − 3 (1 + Φ)−1
L
The end moments are computed to be:
6EI
M0 = (1 + Φ)−1
L2
6EI
M1 = − 3 (1 + Φ)−1
L
Translating from beam sign convention to matrix sign convention, the third column of the stiffness matrix is:
     
K13 V0 −12
 K23   −M0 
= EI  −6L 
  −V1  = (1 + Φ)L3  12 
   
 K33
K43 M1 −6L

13
3.1.6. Unit rotation at right end

Figure 11: Unit displacement θ1 = 1

Left BC: u(0) = 0, θ(0) = 0; Right BC: u(1) = 0, θ(1) = 1


The left end boundary conditions give: u(0) = 0 ⇒ u0 = 0 and θ(0) = 0 ⇒ θ0 = 0.
Solving for M0 in (3.5) with θ(1) = 1, we have:

LV0 EI
M0 = − +
2 L
Substituting this expression into (3.6) with u(1) = 0:

L3 V0 L3 V0 L LV0
0 = − + −
6EI 4EI 2 GAs
 2 
L 1 1
0 = − + V0 +
12EI GAs 2
L2
 
12EI 1
1+ 2 V0 =
12EI L GAs 2
6EI −1
V0 = (1 + Φ)
L2
This value is used to compute M0 :
3EI EI
M0 = − (1 + Φ)−1 +
L L
3EI −1 EI
M0 = − (1 + Φ) + (1 + Φ)(1 + Φ)−1
L L
EI
M0 = − (2 − Φ)(1 + Φ)−1
L
M1 is then calculated to be:

M1 = LV0 + M0
6EI EI
M1 = (1 + Φ)−1 − (2 − Φ)(1 + Φ)−1
L L
EI
M1 = (4 + Φ)(1 + Φ)−1
L
Translating from beam sign convention to matrix sign convention, the fourth column of the stiffness matrix is:
     
K14 V0 6L
 K24   −M0  EI  (2 − Φ)L2 
 = =  
 K34   −V1  (1 + Φ)L3  −6L 
K44 M1 (4 + Φ)L2

14
3.2. Matrix Form
Collecting the information from the previous pages, the stiffness matrix for the Timoshenko beam element is:
 
12 6L −12 6L
EI  6L (4 + Φ)L2 −6L (2 − Φ)L2 
K=   (3.8)
(1 + Φ)L3  −12 −6L 12 −6L 
6L (2 − Φ)L2 −6L (4 + Φ)L2

The equilibrium equation is thus:


    
12 6L −12 6L u0 Fu0
EI  6L (4 + Φ)L2 −6L (2 − Φ)L2   θ0   FM0 
  =  (3.9)
(1 + Φ)L3  −12 −6L 12 −6L   u1   Fu1 
6L (2 − Φ)L2 −6L (4 + Φ)L 2
θ1 FM1

Remember that K is only the element stiffness matrix. The global stiffness matrix is constructed by assembling the stiffness
matrices for each element. The right hand side contains the applied forces. Forces occurring along the beam are represented
as fixed end shears and moments at the ends. These statically equivalent forces are placed in the right hand side vector.
Details concerning this procedure are described in texts on matrix structural analysis such as: Kassimali [8], McCormac
[10], McGuire et al. [11], Timoshenko and Young [15].

3.3. Meaning of Φ
The parameter Φ, defined by equation (3.7), appears in every term of the element stiffness matrix. Φ measures the ratio
of the bending stiffness to shear stiffness. For a very large shear stiffness, Φ is close to zero. In this case, the solutions for
the Timoshenko beam match the solutions to Euler-Bernoulli beam theory. This is expected since shear deformations are
ignored for Euler-Bernoulli beams.
The relative importance of shear deformation can be explored further by examining the material and shape constants in
Φ. We will ignore all numerical constants contained in Φ and focus on dimensional factors instead. For rectangular sections,
the moment of inertia and shear area scale according to the following dimensions:

I ∼ wd3 , As ∼ wd

where w is the width and d is the depth of the section. Noting that E and G are usually the same order of magnitude, we
have:  2
d
Φ∼
L
Thus, Φ ∼ 0 when L ≫ d. For I-sections, the analysis is slightly different:

I ∼ wf tf d2 , As ∼ tw d

where wf is the flange width, tf is the flange thickness, and tw is the web thickness. For most sections, tf and tw are much
nearer in magnitude than they are to d. Also, wf is usually smaller than d. Thus we can write:
 2
wf d d
Φ∼ <
L2 L

We again observe the justification of Euler-Bernoulli beam theory when L ≫ d.

15
4. Green’s Functions

4.1. Formulation

Figure 12: Download unit point load at ξ = µ

The Green’s functions are the solutions to the beam equations (1.17)-(1.20) for a downward unit point load acting at
location x = a. The point load is represented as a delta function: w(x) = −δ(x − a). In terms of the dimensionless variable
ξ, we set µ = a/L. The delta function obeys the scaling property:
1
δ(x − a) = δ(L(ξ − µ)) =
δ(ξ − µ) (4.1)
L
Thus we write: w(ξ) = −δ(ξ − µ)/L. We also require the fact that the integral of the delta function is the Heaviside function:
ˆ ξ (
1 if ξ ≥ µ
δ(τ − µ) dµ = H(ξ − µ) = (4.2)
0 0 if ξ < µ

Detailed information on Green’s functions and the delta function are found in: Richards and Youn [12], Richtmyer [13],
Stakgold [14]. Successive integrals of the Heaviside function are given by the formula:
ξ
(ξ − µ)n+1
ˆ
(τ − µ)n H(τ − µ) dτ = H(ξ − µ), n≥0 (4.3)
0 n+1

These properties allow explicit integration of the solution (1.17)-(1.20).


The shear force is computed to be:
ˆ ξ
V (ξ) = L w(τ ) dτ + V0
0
ξ
1
ˆ
V (ξ) = L −
δ(τ − µ) dτ + V0
0 L
V (ξ) = −H(ξ − µ) + V0 (4.4)

The bending moment is given as:


ˆ ξ
M (ξ) = L V (τ ) dτ + M0
0
ˆ ξ
M (ξ) = L [−H(τ − µ) + V0 ] dτ + M0
0
M (ξ) = L [−(ξ − µ) H(ξ − µ) + V0 ξ] + M0 (4.5)

The angle of deflection can be simplified:


ξ
M (τ )
ˆ
θ(ξ) = L dτ + θ0
0 EI
L2 ξ
ˆ  
1
θ(ξ) = −(τ − µ) H(τ − µ) + V0 τ + M0 dτ + θ0
EI 0 L
2
 
L 1 2 1 2 1
θ(ξ) = − (ξ − µ) H(ξ − µ) + V0 ξ + M0 ξ + θ0
EI 2 2 L
2
 
L 2 2 2
θ(ξ) = −(ξ − µ) H(ξ − µ) + V0 ξ + M0 ξ + θ0 (4.6)
2EI L

16
Finally, the deflection is calculated in the form:
ξ ξ
L
ˆ ˆ
u(ξ) = L θ(τ ) dτ − V (τ ) dτ + u0
0 GAs 0
3 ξ
  ˆ ξ
L 2 2EI L
ˆ
u(ξ) = −(τ − µ)2 H(τ − µ) + V0 τ 2 + M0 τ + 2 θ0 dτ − [−H(ξ − µ) + V0 ] dτ + u0
2EI 0 L L GAs 0
L3
 
1 3 1 3 1 2 2EI L
u(ξ) = − (ξ − µ) H(ξ − µ) + V0 ξ + M0 ξ + 2 θ0 ξ + [(ξ − µ) H(ξ − µ) − V0 ξ] + u0
2EI 3 3 L L GAs
L3
 
3 L
u(ξ) = −(ξ − µ)3 H(ξ − µ) + V0 ξ 3 + M0 ξ 2 + [(ξ − µ) H(ξ − µ) − V0 ξ] + Lθ0 ξ + u0 (4.7)
6EI L GAs

Expressions (4.4)-(4.7) are the Green’s functions for the Euler-Bernoulli beam. Let Gq (ξ, µ) denote the Green’s function
for the value of quantity q at location ξ due to the unit downward load at location µ. We let q ∈ {V, M, θ, u} and list all the
corresponding Green’s functions:

GV (ξ, µ) = −H(ξ − µ) + V0 (µ) (4.8)


GM (ξ, µ) = L [−(ξ − µ) H(ξ − µ) + V0 (µ) ξ] + M0 (µ) (4.9)
L2
 
2
Gθ (ξ, µ) = −(ξ − µ)2 H(ξ − µ) + V0 (µ) ξ 2 + M0 (µ) ξ + θ0 (µ) (4.10)
2EI L
L3
 
3
Gu (ξ, µ) = −(ξ − µ)3 H(ξ − µ) + V0 (µ) ξ 3 + M0 (µ) ξ 2
6EI L
L
+ [(ξ − µ) H(ξ − µ) − V0 (µ) ξ] + L θ0 (µ) ξ + u0 (µ) (4.11)
GAs
Notice that the constants of integration V0 , M0 , θ0 , and u0 depend on the boundary conditions and are functions of the
location µ of the point load.

4.2. Arbitrary Loads


Consider the case of N point loads pi at location µi . By the principle of superposition, the total response for quantity q
is the sum of responses due to each point load pi . Thus the solution to the Euler-Bernoulli beam equations is:
N
X
q(ξ) = pi Gq (ξ, µi ) (4.12)
i=1

where q ∈ {V, M, θ, u} and Gq (ξ, µ) is the corresponding Green’s function.


For an arbitrary load distribution w(ξ) acting downwards, the total response is given by the integral:
ˆ 1
q(ξ) = L w(µ) Gq (ξ, µ) dµ (4.13)
0

17
4.3. Uniform Loads
For a uniform load w (weight per unit length) acting downwards, the solution (4.13) becomes:
ˆ 1
q(ξ) = wL Gq (ξ, µ) dµ (4.14)
0
In order to proceed further, we require the evaluation of certain types of definite integrals. The first of which involves
non-negative integer powers of ξ − µ multiplied by H(ξ − µ):
1
ξ n+1
ˆ
(ξ − µ)n H(ξ − µ) dµ = , n≥0 (4.15)
0 n+1

Another useful definite integral is:


1
1
ˆ
(1 − µ)n dµ = , n≥0 (4.16)
0 n+1
Formulas (4.15) and (4.16) are proved in Haque [1].
Using the Green’s functions (4.8)-(4.11) and applying the formula (4.15) to (4.14):
 ˆ 1 
V (ξ) = wL −ξ + V0 (µ) dµ (4.17)
0
 ˆ 1  ˆ 1
2 1 2
M (ξ) = wL − ξ + ξ V0 (µ) dµ + wL M0 (µ) dµ (4.18)
2 0 0
1 1 ˆ 1
wL3
 
1 3 2
ˆ ˆ
2
θ(ξ) = − ξ +ξ V0 (µ) dµ + ξ M0 (µ) dµ + wL θ0 (µ) dµ (4.19)
2EI 3 0 L 0 0
ˆ 1 ˆ 1
wL4
 
1 3
u(ξ) = − ξ4 + ξ3 V0 (µ) dµ + ξ 2 M0 (µ) dµ
6EI 4 0 L 0
2
 ˆ 1  ˆ 1 ˆ 1
wL 1 2
+ ξ −ξ V0 (µ) dµ + wL2 ξ θ0 (µ) dµ + wL u0 (µ) dµ (4.20)
GAs 2 0 0 0

18
5. Examples

5.1. Simply-supported Beam


5.1.1. Boundary Conditions

Figure 13: Simply-supported beam

5.1.2. Green’s Functions


The left end boundary conditions give: u(0) = 0 ⇒ u0 = 0 and M (0) = 0 ⇒ M0 = 0. The right end boundary conditions
are used to determine θ0 and V0 .
Using M (1) = 0 in equation (4.5) we compute the value of V0 :

0 = L [−(1 − µ) + V0 ]
V0 = 1−µ

The value of θ0 is computed using u(1) = 0 in equation (4.7):

L3
−(1 − µ)3 + (1 − µ) + Lθ0
 
0 =
6EI
L2
(1 − µ)3 − (1 − µ)
 
θ0 =
6EI
The Green’s functions are:

GV (ξ, µ) = −H(ξ − µ) + (1 − µ)
GM (ξ, µ) = L [−(ξ − µ) H(ξ − µ) + (1 − µ)ξ]
L2  L2 
−(ξ − µ)2 H(ξ − µ) + (1 − µ)ξ 2 + (1 − µ)3 − (1 − µ)
 
Gθ (ξ, µ) =
2EI 6EI
L3 
−(ξ − µ)3 H(ξ − µ) + (1 − µ)ξ 3 + (1 − µ)3 ξ − (1 − µ)ξ

Gu (ξ, µ) =
6EI
L
+ [(ξ − µ) H(ξ − µ) − (1 − µ)ξ]
GAs

5.1.3. Uniform Load

 
1
V (ξ) = wL −ξ
2
wL2
M (ξ) = ξ(1 − ξ)
2
wL3
θ(ξ) = − (1 − 6ξ 2 + 4ξ 3 )
24EI
wL4 wL2
u(ξ) = − (ξ − 2ξ 3 + ξ 4 ) − ξ(1 − ξ)
24EI 2GAs

19
5.2. Fixed Beam
5.2.1. Boundary Conditions

Figure 14: Fixed beam

5.2.2. Green’s Functions


The left end boundary conditions give: u(0) = 0 ⇒ u0 = 0 and θ(0) = 0 ⇒ θ0 = 0. The right end boundary conditions
are used to determine M0 and V0 .
Using θ(1) = 0 in equation (4.6) and u(1) = 0 in equation (4.7), we get the system of equations:

L2
 
2
−(1 − µ)2 + V0 + M0 = 0
2EI L
L3
 
3
−(1 − µ)3 + β(1 − µ) + (1 − β)V0 + M0 = 0
6EI L

where β = 6EI/L2 GAs . The solution to this system is:

3(1 − µ)2 − 2(1 − µ)3 + 2β(1 − µ)


V0 =
1 + 2β
 
L (1 − µ) + (β − 1)(1 − µ)2 − β(1 − µ)
3
M0 =
1 + 2β
For sake of brevity, the Green’s functions are:

GV (ξ, µ) = −H(ξ − µ) + V0
GM (ξ, µ) = L [−(ξ − µ) H(ξ − µ) + V0 ξ] + M0
L2
 
2
Gθ (ξ, µ) = −(ξ − µ)2 H(ξ − µ) + V0 ξ 2 + M0 ξ
2EI L
L3
 
3 L
Gu (ξ, µ) = −(ξ − µ)3 H(ξ − µ) + V0 ξ 3 + 2
M0 ξ + [(ξ − µ) H(ξ − µ) − V0 ξ]
6EI L GAs

5.2.3. Uniform Load

 
1
V (ξ) = wL −ξ
2
wL2 1
 
M (ξ) = − − ξ(1 − ξ)
2 6
wL3
θ(ξ) = − (ξ − 3ξ 2 + 2ξ 3 )
12EI
wL4 2 wL2
u(ξ) = − ξ (ξ − 1)2 − ξ(1 − ξ)
24EI 2GAs

20
5.3. Cantilever Beam
5.3.1. Boundary Conditions

Figure 15: Cantilever beam

5.3.2. Green’s Functions


The left end boundary conditions give: u(0) = 0 ⇒ u0 = 0 and θ(0) = 0 ⇒ θ0 = 0. The right end boundary conditions
are used to determine M0 and V0 .
Using V (1) = 0 in equation (4.4) we compute the value of V0 :

0 = −1 + V0
V0 = 1

Then we use M (1) = 0 in equation (4.5)to determine M0 :

0 = L [−(1 − µ) + 1] + M0
M0 = −Lµ

The Green’s functions are:

GV (ξ, µ) = −H(ξ − µ) + 1
GM (ξ, µ) = L [−(ξ − µ) H(ξ − µ) + ξ] − Lµ
L2 
−(ξ − µ)2 H(ξ − µ) + ξ 2 − 2µ ξ

Gθ (ξ, µ) =
2EI
L3 
−(ξ − µ)3 H(ξ − µ) + ξ 3 − 3µ ξ 2

Gu (ξ, µ) =
6EI
L
+ [(ξ − µ) H(ξ − µ) − ξ]
GAs

5.3.3. Uniform Load

V (ξ) = wL(1 − ξ)
wL2
M (ξ) = − (1 − ξ)2
2
wL3
θ(ξ) = − (3ξ − 3ξ 2 + ξ 3 )
6EI
wL4 wL2
6ξ 2 − 4ξ 3 + ξ 4 −

u(ξ) = − ξ(2 − ξ)
24EI 2GAs

21
5.4. Spring-supported Beam
5.4.1. Boundary Conditions

Figure 16: Spring-supported beam

5.4.2. Green’s Functions


The left end moment boundary condition gives M (0) = 0 ⇒ M0 = 0.
Using M (1) = 0 we compute the value of V0 :

0 = L [−(1 − µ) + V0 ]
V0 = 1−µ

The value of V (1) is thus:


V (1) = −µ
The spring equations provide the value of the end displacements:
1
u(0) = − (1 − µ)
k0
µ
u(1) = −
k1
Thus u0 = −(1 − µ)/k0 and u1 = −µ/k1 . Notice that ki → ∞ implies ui → 0 for i = 1, 2.
The value of θ0 is computed using u(1) = u1 :

L3
−(1 − µ)3 + (1 − µ) + Lθ0 + u0
 
u1 =
6EI
L2  1
(1 − µ)3 − (1 − µ) + (u1 − u0 )

θ0 =
6EI L
The Green’s functions are:

GV (ξ, µ) = −H(ξ − µ) + (1 − µ)
GM (ξ, µ) = L [−(ξ − µ) H(ξ − µ) + (1 − µ)ξ]
L2  L2   1
−(ξ − µ)2 H(ξ − µ) + (1 − µ)ξ 2 + (1 − µ)3 − (1 − µ) + (u1 − u0 )

Gθ (ξ, µ) =
2EI 6EI L
L3  3 3 3

Gu (ξ, µ) = −(ξ − µ) H(ξ − µ) + (1 − µ)ξ + (1 − µ) ξ − (1 − µ)ξ
6EI
L
+ [(ξ − µ) H(ξ − µ) − (1 − µ)ξ] + ξu1 + (1 − ξ)u0
GAs
It is seen that Green’s functions for θ and u are the sum of the corresponding Green’s functions for simple beams and a linear
displacement field connecting the ends.

5.4.3. Uniform Load

 
1
V (ξ) = wL −ξ
2
wL2
M (ξ) = ξ(1 − ξ)
2
wL3 1
θ(ξ) = − (1 − 6ξ 2 + 4ξ 3 ) + (u1 − u0 )
24EI L
wL4 wL2
u(ξ) = − (ξ − 2ξ 3 + ξ 4 ) − ξ(1 − ξ) + ξu1 + (1 − ξ)u0
24EI 2GAs

22
5.5. End-restrained Beam
5.5.1. Boundary Conditions

Figure 17: End-restrained beam

5.5.2. End Moments


The left end displacement boundary condition gives u(0) = 0 ⇒ u0 = 0.
We first use equation (4.5) to derive an expression for V0 in terms of the end moments: M0 = M (0) and M1 = M (1).

M1 = L [−(1 − µ) + V0 ] + M0
M1 − M0
V0 = + (1 − µ)
L
This expression and the end moment-angle relations are incorporated into equations (4.6) and (4.7):

L2
 
M1 2 M1 − M0 2 M0
− = −(1 − µ) + + (1 − µ) + M0 +
β1 2EI L L β0
L3
 
M 1 − M 0 3 1 M0
0 = −(1 − µ)3 + + (1 − µ) + M0 − [M1 − M0 ] + L
6EI L L GAs β0

The solution of this system of equations for M0 and M1 is:


 
Lκ0 1
M0 = − µ(1 − µ) (1 − µ)(2 + κ1 ) + 2 + Φκ1
γ 2
 
Lκ1 1
M1 = µ(1 − µ) (1 − µ)(2 + κ0 ) − 4 − κ0 − Φκ0
γ 2

where
Lβ0 Lβ1 12EI
κ0 = , κ1 = , Φ= , γ = κ0 κ1 + 4(κ0 + κ1 ) + 12 + Φ (κ0 κ1 + κ0 + κ1 )
EI EI L2 GAs

5.5.3. Green’s Functions


For sake of brevity, the Green’s functions (4.8)-(4.11) are:

GV (ξ, µ) = −H(ξ − µ) + V0
GM (ξ, µ) = L [−(ξ − µ) H(ξ − µ) + V0 ξ] + M0
L2
 
2 M0
Gθ (ξ, µ) = −(ξ − µ)2 H(ξ − µ) + V0 ξ 2 + M0 ξ +
2EI L β0
L3
 
3
Gu (ξ, µ) = −(ξ − µ)3 H(ξ − µ) + V0 ξ 3 + M0 ξ 2
6EI L
L M0
+ [(ξ − µ) H(ξ − µ) − V0 (µ) ξ] + L ξ
GAs β0
where the values of V0 and M0 were computed above for this problem.

23
5.5.4. Uniform Load
For a uniform load of w, the following integrals are first computed:
ˆ 1  
Lκ0 1 κ1
M0 (µ) dµ = − + (1 + Φ)
0 γ 2 12
ˆ 1  
Lκ1 1 κ0
M1 (µ) dµ = − + (1 + Φ)
0 γ 2 12
ˆ 1    
κ1 1 κ0 κ0 1 κ1 1
V0 (µ) dµ = − + (1 + Φ) + + (1 + Φ) +
0 γ 2 12 γ 2 12 2
ˆ 1 ˆ 1
1
θ0 (µ) dµ = M0 (µ) dµ
0 β 0 0

These values are substituted into the uniform load solution:


 ˆ 1 
V (ξ) = wL −ξ + V0 (µ) dµ
0
 ˆ 1  ˆ 1
1
M (ξ) = wL2 − ξ 2 + ξ V0 (µ) dµ + wL M0 (µ) dµ
2 0 0
ˆ 1 ˆ 1 ˆ 1
wL3
 
1 2
θ(ξ) = − ξ3 + ξ2 V0 (µ) dµ + ξ M0 (µ) dµ + wL θ0 (µ) dµ
2EI 3 0 L 0 0
ˆ 1
wL4 3 2 1
 
1 4
ˆ
3
u(ξ) = − ξ +ξ V0 (µ) dµ + ξ M0 (µ) dµ
6EI 4 0 L 0
1 1
wL2 1 2
 ˆ  ˆ
+ ξ −ξ V0 (µ) dµ + wL2 ξ θ0 (µ) dµ
GAs 2 0 0

24
References

[1] Aamer Haque. Green’s functions for euler-bernoulli beams. 2015.


[2] H.B. Harrison. Computer Methods in Structural Analysis. Civil Engineering and Engineering Mechanics Series. Prentice-
Hall, Enlgewood Cliffs, NJ, 1973.
[3] R.C. Hibbeler. Mechanics of Materials. Prentice Hall, Upper Saddle River, New Jersey, seventh edition, 2008.
[4] R.C. Hibbeler. Structural Analysis. Prentice Hall, Upper Saddle River, New Jersey, eighth edition, 2012.
[5] Keith D. Hjelmstad. Fundamentals of Structural Mechanics. Springer, New York, second edition, 2005.
[6] Claes Johnson. Numerical Solution of Partial Differential Equations by the Finite Element Method. Studentlitteratur,
Lund, Sweden, 1987.
[7] Aslam Kassimali. Structural Analysis. PWS-Kent, Boston, 1993.
[8] Aslam Kassimali. Matrix Analysis of Structures. Brooks/Cole, Pacific Grove, CA, 1999.
[9] Steen Krenk. Mechanics and Analysis of Beams, Columns, and Cables. Springer-Verlag, Berlin, second edition, 2001.
[10] J.C. McCormac. Structural Analysis using Classical and Matrix Methods. Wiley, New York, fourth edition, 2007.
[11] W. McGuire, R.H. Gallagher, and R.D. Ziemian. Matrix Structural Analysis. Wiley, New York, second edition, 2000.
[12] J Ian Richards and Heekyung K. Youn. Theory of Distributions, A Nontechnical Introduction. Cambridge University
Press, Cambridge, 1990.
[13] Robert D. Richtmyer. Principles of Advanced Mathematics Physics, volume 1 of Texts and Monographs in Physics.
Springer-Verlag, Berlin, 1978.
[14] Ivar Stakgold. Green’s Functions and Boudnary Value Problems. Wiley-Interscience, New York, 1979.
[15] S.P. Timoshenko and D.H. Young. Theory of Structures. McGraw-Hill, New York, second edition, 1965.

[16] C.M. Wang, J.N. Reddy, and K.H. Lee. Shear Deformable Beams and Plates, Relationships with Classical Solutions.
Elsevier, Oxford, 2000.

25

Potrebbero piacerti anche