Lecture Notes CT 4145
Lecture Notes CT 4145
A.V. Metrikine
ii
iii
iv
CONTENTS
CHAPTER 1. SIGNIFICANCE OF DYNAMICS FOR CIVIL ENGINEERING.1
1.1. Earthquake-Induced Vibrations. 1
1.2. Wind-Induced Vibrations...... 2
1.3. Traffic-Induced Vibrations 4
1.4. Construction- and Production-Induced Vibrations... 5
1.5. Wave- and Flow-Induced Vibrations 6
CHAPTER 2. DYNAMICS OF MASS-SPRING-DASHPOT SYSTEMS.. 8
2.1. Fundamental Assumptions Leading to Models with Finite Number
of Degrees of Freedom (Mass-Spring-Dashpot Models... 8
2.2. One Degree of Freedom Systems without Damping. 9
2.2.1. Free Vibrations..10
2.2.2. Forced Vibrations under Harmonic Force 14
2.2.3. Forced Vibrations under General Disturbing Force......19
2.3. One Degree of Freedom Systems with Viscous Damping...29
2.3.1. Free Vibrations..30
2.3.2. Forced Vibrations under Harmonic Force.... 35
2.3.3. Forced Vibrations under General Disturbing Force..43
2.4. Two Degrees of Freedom Systems without Damping
2.4.1. Free Vibrations..47
2.4.2. Forced Vibrations under Harmonic Force.53
CHAPTER 3. ONE-DIMENSIONAL (SLENDER) STRUCTURES... 57
3.1. Fundamental Assumptions Leading to One-Dimensional
Continuous Models....... 57
3.2. Formulation of Equations of Motion for One-dimensional Models 59
3.2.1. Transverse Motion of a String...59
3.2.2. Longitudinal Motion of a Rod...60
3.2.3. Torsion of a Rod62
3.2.4. Pure Shear of a Beam64
3.2.5. Bending of a Beam66
3.2.6. Viscous-Elastic Kelvin Foundation...68
3.2.7. Combination of Models.....69
3.3. Boundary and Interface Conditions for One-dimensional Models.................. 70
3.3.1. Transverse Motion of a String ..... 71
3.3.2. Longitudinal Motion of a Rod...................72
3.3.3. Torsion of a Rod73
3.3.4. Pure Shear of a Beam74
3.3.5. Bending of a Beam75
3.4. Introduction to Dynamics of One-Dimensional Structures......................78
3.4.1. Longitudinal Vibrations of a Rod..79
3.4.2. Bending Vibrations of a Beam..85
3.5. Elastostatics of One-Dimensional Structures...................88
3.5.1. Static Response of a Rod to Axial Loading.. 89
3.5.2. Static Response of a Rod on Elastic Foundation to Axial Loading.. 99
3.5.3. Static Response of a Beam..104
3.5.4. Static Response of a Beam on Elastic Foundation..118
vi
vii
NOTATIONS
f0
f
G
g
G (t )
H (...)
H
I
i
J
Jt
K
k
K
Kd
kd
L
M
m
Mt
mt
n
P
P
Q
q
q1
s
T
Heaviside step-function
horizontal force
moment of inertia of a cross-section
1
polar moment of inertia
torque constant
kinetic energy
spring constant
spring constant, dimensionless stiffness
dimensionless stiffness
stiffness per unit length
Lagrange function
mass, moment of force
mass
torque
linear density of torque
amount of viscous damping, integer
force
potential energy
generalized force
body force
force per unit length
characteristic exponent, Laplace variable
tension
A
A0
a
B
b
A, B, C
c
cd
E
E
F
F0
g ( s ) image function
viii
Tn
t, t
t
U ( x)
u
V
v0
v
w
X
Xc, Xs
x0
x
x
natural period
time
time interval
function of coordinate
axial displacement
transverse force, shear force
initial velocity
velocity increment
transverse displacement
complex amplitude
amplitudes
initial displacement
coordinate, displacement from the equilibrium position
velocity
angle
wavenumber
dimensionless amount of viscous damping, shear strain
axial strain
initial phase
phase lag
( t ) function of time
a , b
n , n
angle of twist
mass ratio
slope
mass density
normal stress
shear stress
radial frequency, radial frequency of excitation
partial frequencies
natural frequencies
x , F vector quantities
K , M matrix quantities
CHAPTER 1
SIGNIFICANCE OF DYNAMICS FOR CIVIL ENGINEERING
The analysis and design of structures subject to time dependent forces constitute the
field of structural dynamics. For example, the dynamic forces that are transmitted to a
structure that supports oscillating machinery must be considered in its analysis and
design. This is also true for a bridge facilitating car or train traffic, for a structure that
is subjected to blast, wind or water waves, or for a building whose foundation is
disturbed by an earthquake. In this Section, these and other examples of civil
engineering systems and structures are discussed to underline the significance of
dynamics for modern civil engineering. All examples are subdivided into 5 categories
in accordance with different types of causes of vibrations. These causes are:
earthquake, wind, traffic, construction and production, and water waves.
1.1.
Earthquake-Induced Vibrations
An earthquake is the shaking of the earth caused by pieces of the crust of the Earth
that suddenly shift. The crust, the thin outer layer of the Earth, is mostly cold and
brittle rock compared to the hot rock deeper inside. This crust is full of large and
small cracks called faults. Although these faults can be hundreds of miles long,
usually you cannot see the cracks because they are buried deep underground and
because the pieces of crust are compressed together very tightly. The powerful forces
that compress these crust pieces also cause them to move very slowly. When two
pieces that are next to each other get pushed in different directions, they will stick
together for a long time (many years), but eventually the forces pushing on them will
force them to break apart and move. This sudden shift in the rock shakes all of the
rock around it. The resulting vibrations propagate in the form of seismic waves and
can cause significant damage of buildings. Figure 1.1 shows a high-rise building
damaged by the Taiwan Earthquake in 1999.
F(t)
Figure 1.2. Earthquake-induced excitation and simple structural model for a building
such an analysis utilize beams, which describe both the bearing walls and the floors,
see Figure 1.2. The force exerted by an earthquake is often modelled as a shear force
applied to the building footing.
1.2.
Wind-Induced Vibrations
As modern structures move towards taller and more flexible designs, the problems of
wind effect have become increasingly apparent. The most well known examples of
structures vulnerable to wind effects are suspension bridges (see Figure 1.3 of the
famous Tacoma Narrows Bridge), skyscrapers (Figure 1.4), chimneys (Figure 1.5)
and wind-energy generators (Figure 1.6).
lead to catastrophic failure as in the case of the Tacoma Narrows Bridge see Figure
1.3. Flutter is associated with so-called self-excited vibrations, during which the
structural velocity is in phase with the aerodynamic force. The term self-excited
originates from the fact that the structure itself chooses to move in phase with the
aerodynamic excitation. Such a motion is associated with negative damping
(because positive damping implies that the force is in anti-phase with structural
velocity), which leads to structural instability. In an unstable scenario, the structure
extracts energy from the wind amplifying its own motion, which can lead to
catastrophic failure. Flutter is of significant danger for structures and therefore must
be avoided for the wind velocity range of interest by means of structural optimization
and aerodynamic tailoring.
There exist a wide variety of prediction models for wind-induced vibrations. They
range from simplistic one- and two-degree of freedom systems to sophisticated threedimensional elasticity models. In the majority of these models, however, the
aeroelastic effects are described simplistically, especially at high Reynolds numbers
associated with strong wind and large structural dimensions. This is because it is
extremely difficult to model the shear layers around the structure. The modeling is
normally done by introducing added mass, negative damping and additional dynamic
stiffness of the structure. All these added properties are associated with the
aerodynamic force, induced by structural motion.
1.3.
Traffic-Induced Vibrations
Bridges, Railway Lines
When a vehicle such as a car or a train passes through an urban area (see Figure 1.7 of
a bullet train in Kyoto City), building occupants in the neighbourhood may experience
significant discomfort associated with vibrations induced by a vehicle. There can be
various reasons for these vibrations to occur: uneven or bumpy roads, curves of the
traffic lines, imperfection of vehicle suspension system, unevenness of train wheels,
corrugation of rails, etc. To predict the level of building vibrations induced by traffic,
dynamic model for a vehicle and for a building have to be integrated into a module
that predicts propagation of elastic waves through the soil from the vehicle to the
building.
Traffic-induced vibrations are important for civil engineering not only because of
the discomfort they can cause to people but also because of their significant influence
on durability and reliability of structures such as bridges, see Figure 1.8. Bridges must
be designed such that no perceptible traffic-induced vibrations occur during
exploitation. This is especially important for bridges facilitating high-speed train
traffic, since such a train can cause considerable dynamic amplification of bridge
vibrations. Sometimes, train-induced vibrations of rails can lead to catastrophic
consequences. This can happen with continuously welded railway tracks during a hot
summer. Due to the temperature extension, a considerable axial compression can
occur in the rails. This compression makes the rails vulnerable both to static and
dynamic buckling see Figure 1.9 and 1.10. The static buckling occurs in poorly
ballasted tracks (Figure 1.9), while the dynamic buckling can be initiated even in a
well-ballasted track under the action of a passing train (Figure 1.10).
1.4.
Construction sites are well known as a source of noise and ground vibrations. The
latter are mainly associated with pile driving, see Figure 1.11. Every nock of the
hammer on the pile is accompanied by powerful vibrations of the surrounding ground.
In some cases these vibrations can be damaging to surrounding buildings let alone
their annoyance to people living around. Thus, before opening a new construction site,
the level of vibrations must be assessed that can be expected in the surrounding
buildings.
1.5.
Waves and currents in the ocean cause vibrations of floating vessels and submerged
structural elements. The wave-induced vibrations play an increasingly important role
in dynamics of offshore platforms, especially in that of the semi-submersible
platforms, see Figure 1.13. These platforms facilitate the deep-water oil and gas
exploitation, which is becoming more and more important once easily accessible
oil/gas reservoirs are getting exhausted. These platforms float in the Sea being moored
to the seabed by flexible cables, which keep the platform at place but almost do not
resist to relatively small vibrational motions. These motions, however, can be
damaging to the pipes, through which oil/gas is transported up to the platform.
Therefore, the floating platforms should be designed such that their response to waves
is minimized. Normally, this is achieved by building platforms with Frequency
Response Functions, which do not intersect with the Wave Spectrum.
CHAPTER 2
DYNAMICS OF MASS-SPRING-DASHPOT SYSTEMS
Many of the vibratory phenomena associated with complicated systems may be
understood by studying the behaviour of simple systems. The most basic vibrating
system consists of a single lumped mass and a spring. This is said to have one degree
of freedom, implying that there is only one possible direction of movement for the
lumped mass. In reality, structures may move in a few different directions, rotate, etc.
Therefore, real structures are systems with many degrees of freedom. In many
practical situations, however, just a few degrees of freedom are activated by the load.
In these situations, models accounting for one or two degrees of freedom can provide
a good estimate for dynamic structural processes. In this Chapter, attention is focused
on one- and two-degrees of freedom systems, which are most facilitating for
understanding of the fundamentals of structural dynamics.
x(t)
m
F(t)
Unloaded
static
position
x(t)
F(t)
Figure 2.2. Idealization of vertical vibration of a lorry on a bridge
In Figure 2.1 the relatively heavy deck of the offshore platform is replaced by a
lumped mass while the braced steel jacket structure is replaced by an elastic spring,
which models the lateral stiffness of the structure. The wave load is assumed to be
applied to the deck. In Figure 2.2the bridge girder is idealized as a leaf spring with its
mass concentrated at midspan being dynamically loaded by a heavy lorry. Both these
idealizations disregard all but one possible motion and implicitly assume that the
loading of the structure is such that only one mode of structural vibrations is activated.
Although, the idealizations may be thought to be apparently crude, a skilled engineer
can make surprisingly accurate predictions of the behavior of real structures by
intelligent choice of the parameters of simple systems.
2.2.
F(t)
x
(2.1)
Both force and acceleration are vector quantities, mass being scalar, therefore a force
exerted on mass causes acceleration in the direction of the force.
10
In the case under consideration, there are two forces acting on the mass: the
external force F ( t ) and the force exerted by the spring FSPRING ( t ) . Therefore, by
Newtons second law
F ( t ) + FSPRING ( t ) = m
x
(2.2)
where x ( t ) is the displacement of the mass from its equilibrium position, in which
the spring is undisturbed. Let us underline that in Eq.(2.2) all terms are vectors. In
accordance with Hookes law,
FSPRING ( t ) = kx
(2.3)
Substituting this expression into Eq.(2.2) and projecting the result onto the x axis,
which is chosen to be directed rightward, the following scalar equation is obtained
mx+ kx = F ( t )
(2.4)
This is the equation of motion, which describes vibrations of the mass-spring system
around its equilibrium position under the action of the external force F ( t ) .
2.2.1. Free Vibrations. If the external force F ( t ) is absent, the equation of motion
(it is a second order ordinary linear differential equation) for the mass-spring system
reduces to
mx+ kx = 0
(2.5)
In this case, the system can move only if given an initial displacement or/and an initial
velocity:
x ( 0 ) = x0
x ( 0 ) = v0
(2.6)
To find x ( t ) that satisfies both the equation of motion Eq.(2.5) and the initial
conditions Eq.(2.6), two steps are to be made. First, the general solution of Eq.(2.5)
should be found, which will contain two unknown constants. The number of these
unknown constants is equal to the order of the differential equation. Secondly, these
two constants should be determined by employing the initial conditions.
To find the general solution of Eq.(2.5), a theorem can be used that states that the
general solution of an ordinary linear differential equation of N th order with
constant coefficients can be written as ( x ( t ) is the unknown function in this equation)
N
x ( t ) = X n exp ( sn t )
n =1
(2.7)
11
where X n and sn are complex constants and are referred to as the complex amplitude
and the characteristic exponent, respectively.
For the second order differential equation Eq.(2.5), the general form of solution
Eq.(2.7) reduces to
2
x ( t ) = X n exp ( sn t )
(2.8)
n =1
n =1
n =1
X ( ms
n
2
n
n =1
(2.9)
+ k ) exp ( sn t ) = 0
If a non-trivial solution X n 0 is sought for, then Eq.(2.9) can be satisfied if and only
if the following characteristic equation is satisfied:
msn2 + k = 0
(2.10)
The characteristic exponents sn are the roots of this equation. In the case under
consideration s1,2 read
s1 = i k m ,
s2 = i k m
(2.11)
x ( t ) = X 1 exp it k m + X 2 exp it k m
(2.12)
where the exponents with the imaginary arguments are understood through the Euler
formula, which reads
exp ( i ) = cos ( ) + i sin ( )
(2.13)
( (
)) + X ( cos (t k m ) i sin (t
) cos ( t k m ) + i ( X X ) sin ( t k m )
)
x ( t ) = X 1 cos t k m + i sin t k m
= ( X1 + X 2
k m
))
(2.14)
12
n = k m
(2.15)
is the natural frequency of the mass-spring system. As can be seen from Eq.(2.14),
this is the frequency, with which the mass-spring system would vibrate given arbitrary
initial conditions.
To determine the unknown constants A and B , Eq.(2.14) is to be substituted into
the initial conditions Eq.(2.6). This yields
A = x0
(2.16)
Bn = v0
Thus, the deflection of the mass-spring system from its equilibrium position reads
x ( t ) = x0 cos (n t ) +
v0
sin (n t )
(2.17)
Introducing the amplitude A0 and initial phase 0 in accordance with the following
relationships:
x0 = A0 cos (0 )
(2.18)
v0 n = A0 sin (0 )
and combining the right-hand side of Eq.(2.17) trigonometrically, we can rewrite this
equation in the following, widely used form
x ( t ) = A0 cos (n t 0 )
(2.19)
where
A0 = x02 + ( v0 n )
v
0 = arctan 0
x0n
(2.20)
The displacement x ( t ) of the mass-spring system can be plotted versus time as shown
in Figure 2.4
13
2/n
v0
A0
x0
Figure 2.4 shows that vibrations of the mass-spring system are perfectly sinusoidal
and last forever. This is because of an assumption of zero damping, which is not
accurate. In reality, there almost always exists some damping (energy loss) in
vibratory systems. This damping causes the oscillatory motion induced by the initial
disturbance be reduced to zero in the course of time.
For a better understanding of the dynamic behaviour of the mass-spring system it is
useful to consider variation of the mechanical energy in the system with time. The
energy E of the system is composed of the kinetic energy K, which is associated with
the mass motion, and the potential energy P, which is associated with contractions
and expansions of the spring. This energy is given as
E =K+P =
1
1
1
2
2
m ( x ) + kx 2 = m ( x ) + n2 x 2
2
2
2
(2.21)
Substituting into this equation the expression for the mass-spring motion Eq.(2.17),
one obtains
2
2
v0
1
2
E = m ( x0n sin ( n t ) + v0 cos (n t ) ) + n x0 cos (n t ) + sin (n t )
2
n
1
1
1
m ( x02n2 + v02 ) = mv02 + kx02
2
2
2
(2.22)
Eq.(2.22) shows that the mechanical energy of the mass-spring system is time
invariant and equals the sum of the kinetic and potential energies, which the system is
supplied with at the initial time moment.
14
2.2.2. Forced Vibrations under Harmonic Force. Consider the dynamic response of
the mass-spring system to a harmonic (sinusoidal) force. The importance of
considering such a force descends from the fact that any time-dependent force can be
represented as a superposition of harmonic forces with different amplitudes and
phases (see Eq.(2.19) for definition of the amplitude and the phase of the harmonic
motion). Thus, the dynamic behaviour of the mass-spring system under a harmonic
force may be considered as a fundamental knowledge, which facilitates understanding
of the behaviour of the system under a force of arbitrary time signature.
Assume that the external force acting on the mass-spring system is given as
F ( t ) = F0 cos (t ) . In this case the equation of motion Eq.(2.4) takes the form
kx = F0 cos (t )
mx+
(2.23)
(2.24)
Substituting Eq.(2.24) into the equation of motion Eq.(2.23), the following expression
can be found for the constant X in the particular solution:
X=
F0
F
1
= 0
2
k m
k 1 2 n2
(2.25)
F0
1
cos (t )
k 1 2 n2
(2.26)
The unknown constants A and B depend on the initial conditions. Employing the
initial conditions Eq.(2.6), the following system of two algebraic equations can be
formulated for determining A and B:
15
A+
F0
1
= x0
k 1 2 n2
(2.27)
Bn = v0
F0
1
k 1 2 n2
(2.28)
v0
Substituting Eq.(2.28) into Eq.(2.26), the following final expression is obtained that
describes the dynamic behavior of the mass-spring system:
x ( t ) = x0 cos (n t ) +
v0
sin (n t ) +
F0
1
( cos (t ) cos (nt ) )
k 1 2 n2
(2.29)
This expression clearly shows that the mass-spring system vibrates at two frequencies,
one being the natural frequency n and the other being the frequency of the
external force. This is also true if the initial displacement x0 and the initial velocity v0
are zero.
If the natural frequency and the frequency of the force are close, an interesting
phenomenon occurs, which is called beating. This is shown in Figure 2.5, where the
ratio x ( t ) ( F0 k ) (note that F0 k is the static displacement of the mass spring
system under the force F0 ) is plotted versus time in accordance with Eq.(2.29). To
plot this figure, the initial conditions were assumed as zero,
= 2 rad s , n = 2.1rad s.
30
20
x(t)k/F 0
10
0
-10
-20
-30
0
50
100
150
time t [s]
200
250
Figure 2.5. Vibration of the mass-spring system in the case that the natural frequency is close to the
forcing frequency
16
Figure 2.5 shows that the vibrations beat with a long period, which is inverse
proportional to the difference between n and . The beating can emerge only in the
systems without (which is unrealistic) of with a small damping, which suppresses it in
the course of time. As will be discussed in Section 2.3, after a sufficiently long time,
the dynamic behavior of the mass spring system subject to a small damping is
approximately described by the particular solution Eq.(2.24). In this case this solution
is called the steady-state solution. It is important to underline that the steady-state
solution is independent of the initial conditions.
Let us study the steady-state solution, which, in accordance with Eqs.(2.24) and
(2.25) is given as
xsteady = X cos (t ) =
F0
1
cos (t )
k 1 2 n2
(2.30)
The dependence X xstatic , xstatic = F0 k being the static deflection, on the ratio n
of the forcing and the natural frequencies is shown in Figure 2.6.
5
4
3
X/xstatic
2
1
0
-1
-2
-3
-4
-5
0
/n
Figure 2.6, as well as Eq.(2.30), show that for n < 1 the ratio X xstatic is positive ,
whereas for n > 1 it is negative. In order to understand the meaning of this sign
change, we return to Eq.(2.23) and the assumption that the steady-state solution has
the form given by Eq.(2.24). It appears that in the region n > 1 the value of X is
negative. But we can write
X cos (t ) = + X cos (t + )
(2.31)
17
rightward, for n > 1 the mass is located to the left from the equilibrium position
when the force pushes rightward.
Thus, one may distinguish the amplitude-frequency characteristic and the phasefrequency characteristic in vibrations of the mass spring system (the former is often
referred to as the frequency-response function). To find expressions for these
dependencies, we rewrite Eq.(2.30) as
xsteady
F0
1
cos (t ) ,
k 1 2 n2
n < 1
F0
1
cos (t + ) ,
k 1 2 n2
n > 1
(2.32)
This expression shows that the amplitude of vibrations of the mass-spring system is
given by
X =
F0
1
k 1 2 n2
(2.33)
for all frequencies of the force, whereas the phase lag between the force and the mass
changes abruptly from 0 to as the forcing frequency becomes larger than the
natural frequency:
0,
=
,
n < 1
(2.34)
n > 1
The amplitude- and phase-frequency dependencies are shown in Figures 2.7a and
2.7b.
5
X/xstatic
0
0
/n
18
180
90
0
0
/n
F0
1
( cos (t ) cos (nt ) )
k 1 2 n2
(2.35)
Obviously, at n = 1 both the numerator and the denominator of this expression are
zero. To get rid of this uncertainty, one may differentiate the numerator and the
denominator with respect to (according to the LHospitals rule). This yields
x (t ) =
F0 t sin (t )
F
= 0 t sin (t )
2
k 2 n = 2k
n
(2.36)
19
Eq.(2.36) shows that the amplitude of resonant vibrations grows in time linearly
and the displacement of the mass-spring system increases as shown in Figure 2.8 for
= 2 rad s .
20
x(t)k/F0
10
-10
-20
0
10
15
20
25
t [s]
As it was shown in the previous section, the energy of free vibrations of the massspring system does not change in time. The presence of an external force, naturally,
changes the situation leading to variation of the energy of the mass-spring system in
time. This energy remains limited unless resonance occurs.
2.2.3. Forced Vibrations under General Disturbing Force. In most practical
applications the dynamic loading F ( t ) is irregular and non-periodic. In this section
two ways of analysis are presented of the response of the mass-spring system to such
loading. The first way is based on the representation of a time-dependent force by a
sequence of short impulses. Superposition of the system responses to these impulses
is found in the form of a convolution integral, which is often called Duhamels
integral after the French mathematician J.M.C. Duhamel. The second way is based on
application of the Laplace transform (named after the French scientist P.S. Laplace).
In many cases, application of the Laplace transform implies physically that the
loading F ( t ) is first represented as a continuous superposition (integral) of its
harmonic components so that the response to each of these harmonics can be found
separately. These responses are then integrated by means of the inverse Laplace
transform to give the dynamic response to the original non-harmonic load.
Consider Eq.(2.4) with the initial conditions Eq.(2.6):
mx+ kx = F ( t )
x ( 0 ) = x0
(2.37)
x ( 0 ) = v0
20
t
t'
dt'
1
F ( t ) sin (n ( t t ) ) dt
mn 0
(2.39)
This expression still does not include the effect of any initial displacement x0 or any
initial velocity v0 when t = 0 . These effects, however, are exactly those represented
by Eq.(2.17). Hence, for a complete solution of Eq.(2.37), we may write
v0
1
x ( t ) = x0 cos (n t ) + sin (n t ) +
F ( t ) sin (n ( t t ) ) dt
mn 0
n
(2.40)
21
account both free and forced vibrations and all this together is accounted for by the
integral Eq.(2.39).
The solution Eq.(2.40) was obtained using some physical considerations. It is
necessary, therefore, to check whether this solution indeed satisfies the original initialvalue problem Eq.(2.37). To perform this check, one has to substitute Eq.(2.40) into
both the equation of motion and the initial conditions. Such a substitution requires
taking the time derivative of the integral in Eq.(2.40), which has the upper limit
dependent on the time t . The following general formula should be used in this case,
which can be found in any book on advanced mathematical analysis:
b( t )
b(t )
db
da
f ( , t ) d =
f ( , t ) d + f ( b ( t ) , t ) f ( a ( t ) , t )
dt a ( t )
dt
dt
t
a(t )
(2.41)
Using this formula the first and the second time-derivatives of Eq.(2.40) can be
written as
x ( t ) = n x0 sin (n t ) + v0 cos (nt ) +
1
F ( t ) cos (n ( t t ) ) dt
m 0
x ( t ) = n2 x0 cos (nt ) n v0 sin (n t )
n
m
F ( t) sin ( ( t t)) dt +
n
F (t )
(2.42)
Substitution of these expressions into the initial-value problem Eq.(2.37) proves that
Eq.(2.40) is indeed the solution to this problem.
For some applications, it is convenient to rewrite the integral in Eq.(2.40) using the
following trigonometric relation:
sin (nt nt ) = sin (n t ) cos (nt ) cos (n t ) sin (n t )
Substituting this relation in Eq.(2.40), we obtain
x ( t ) = x0 cos (n t ) +
v0
(2.43)
where
A (t ) =
1
F ( t ) sin (n t )dt ,
mn 0
B (t ) =
1
F ( t ) cos (n t )dt
mn 0
(2.44)
22
t
F
1
F ( t ) sin (n ( t t ) )dt = 0 cos (t ) sin (n ( t t ) )dt
mn 0
mn 0
(2.45)
F0
F
1
cos (t ) cos (nt ) ) = 0
=
( cos (t ) cos (nt ) )
2
2 (
k 1 2 n2
m ( n )
Substituting this into Eq.(2.40), we obtain the following expression, which coincides
with Eq.(2.29)
x ( t ) = x0 cos (n t ) +
v0
sin (n t ) +
F0
1
( cos (t ) cos (nt ) )
k 1 2 n2
(2.46)
(b)
(a)
2
t2
t1
t0
/n
-1
1
-2
/n
If the pulses follow each other with a constant time delay equal to 2 n as show in
Figure 2.10(a), this delay being equal to the period of natural oscillations of the massspring system, then Eq.(2.46) takes the form (which is obtained by substitution
tn = t0 + 2 n n )
23
x=
=
=
=
sin (n ( t t0 ) ) vk
k =0
This expression clearly shows that the amplitude of the response will be built up
without limit in the course of time. The same result will follow from a succession of
positive and negative impulses, which are shown in Figure 2.10(b). In this case the
time delay equals n so that tn = t0 + n n should be substituted in Eq.(2.46) to
give
x=
( v sin ( ( t t ) ) + v sin ( ( t t
0
n ) )
( v sin ( ( t t ) ) + v sin ( ( t t ) )
0
( v sin (
0
( t t0 ) ) v1 sin (n ( t t0 ) )
sin (n ( t t0 ) ) vk + vk
n
k = 0,2,4
k =1,3,5
These observations lead to the important practical conclusion that any periodic
external force, the period of which equals the period of natural vibrations causes
resonance, i.e. a large-amplitude response.
As a third example consider the external force to be suddenly applied to the mass
and then remain constant, i.e.
0,
F (t ) =
F0 ,
t<0
t0
(2.47)
24
x ( t ) = x0 cos (n t ) +
v0
sin (n t ) +
F0
(1 cos (nt ) )
mn2
(2.48)
Noting that
F0
F
= 0 = xstatic
2
mn
k
v0
sin (n t )
(2.49)
From this expression we see that a suddenly applied constant force produces free
vibrations of amplitude
A=
( x0 xstatic )
v
+ 0
n
(2.50)
superimposed upon the static displacement. In the particular case where the initial
conditions are zero, Eq.(2.50) for the amplitude, reduces to A = xstatic and we have
xmax = 2 xstatic and xmin = 0 . Thus, a suddenly applied constant force produces a
maximum deflection twice as great as the static effect of the same force.
Eq.(2.47) corresponds to the constant force, once applied, to act indefinitely. If it
acts only for an interval of time t and then is suddenly removed, i.e.
t < 0 and t > t
0,
F (t ) =
F0 ,
(2.51)
0 t t
(2.53)
25
This shows that after removal of the force, the mass-spring system oscillates at its
natural frequency. The amplitude of these vibrations in the case of zero initial
conditions reads
A = xstatic
( cos ( t ) 1)
n
t
2
t
+ sin (n t ) = 2 xstatic sin n = 2 xstatic sin
(2.54)
2
Tn
g ( s ) = G ( t ) exp ( st ) dt
(2.55)
0,
H ( t t0 ) =
1,
t < t0
t > t0
(2.56)
26
Original Function G ( t )
G (t )
Definition
Image Function g ( s )
G ( t ) exp ( st ) dt
0
g ( s)
+ i
Inversion
Formula
1
g ( s ) exp ( st ) dt
2 i i
Linearity
Property
Differentiation
AG1 ( t ) + BG2 ( t )
G ( t )
sg ( s ) G ( 0 )
G ( t )
G(
n)
Ag1 ( s ) + Bg 2 ( s )
s 2 g ( s ) sG ( 0 ) G ( 0 )
(t )
s n g ( s ) s n 1G ( 0 ) ...G (
Integration
Convolution
Theorem
G ( t ) G ( ) d
1
(0)
1
g (s)
s
G ( ) d
n 1)
g1 ( s ) g 2 ( s )
Translation
G (t b) H (t b) , b > 0
exp ( bs ) g ( s )
(2.57)
f ( s ) + msx0 + mv0
(2.58)
ms 2 + k
The remaining task is to apply the inversion formula to determine the original
function x ( t ) . This can be done by applying the contour integration method
supplemented by the residue theorem. In this development, however, there is no room
for studying this method. Therefore, a table of Laplace transforms will be used, a
short version of which is presented below. For finding other pairs of Laplace
transforms, one may use MAPLE.
g ( s)
G (t )
1s
1 s2
1
t
1 sn
t n 1 ( n 1)!
27
g ( s)
G (t )
1 t
( n +1 2 )
, n = 1, 2,3,...
2n t n 1 2
1 (s + a)
exp ( at )
1 (s + a)
t exp ( at )
1 (s + a)
t n 1 exp ( at ) ( n 1) !
(a b)
( s + a )( s + b )
s
(a b)
( s + a )( s + b )
exp ( at ) exp ( bt )
ba
a exp ( at ) b exp ( bt )
a b
1
s + a2
s ( s2 + a2 )
1
sin ( at )
a
cos ( at )
1 ( s2 a2 )
sinh ( at ) a
cosh ( at )
s
s a2
1
2
s ( s + a2 )
2
1
(1 cos ( at ) )
a2
1
s ( s + a2 )
1
( at sin ( at ) )
a3
1
( sin ( at ) at cos ( at ) )
2a 3
(s
+a
2 2
(s
+a
(s
t
sin ( at )
2a
2 2
1
( sin ( at ) + at cos ( at ) )
2a
+ a2 )
t cos ( at )
s2 a2
(s
(s
+a
+a
)( s
+b
2 2
(( 2n 1)! )
(a
1
( s + a 2 ) + b2
2
s+a
( s + a 2 ) + b2
2
b2 )
cos ( at ) cos ( bt )
b2 a2
1
exp ( at ) sin ( bt )
b
exp ( at ) cos ( bt )
28
Let us now transform Eq.(2.58) to the time domain. To this end we rewrite this
equation as follows:
x (s) =
1
1
s
1
+ x0 2
+ v0 2
f (s) 2
2
2
m
s + n
s + n
s + n2
(2.59)
In accordance with the linearity property (see Table 2.1), every of the three terms in
Eq.(2.59) may be inverted separately. To invert the first term, we use the convolution
theorem (Table 2.1), which states that a multiplication of images is inverted into
convolution of the originals. Since, by definition, the original to f ( s ) is F ( t ) ,
whereas the original to 1 ( s 2 + n2 ) is 1 n sin (n t ) (Table 2.2), the first term is
inverted into the Duhamels integral:
t
1
1
1
f (s) 2
F ( t ) sin (n ( t t ) )dt
2
m
s + n mn 0
(2.60)
In Eq.(2.60) and in what follows the sign indicates a pair of Laplace-original and
Laplace-image.
The second and third terms in Eq.(2.60), according to Table 2.2, have the following
originals:
x0
s
x0 cos (nt )
s + n2
2
(2.61)
v
1
0 sin (n t )
v0 2
2
s + n n
1
x ( t ) = x0 cos (n t ) + sin (n t ) +
F ( t ) sin (n ( t t ) ) dt
n
mn 0
(2.62)
Let us consider the case of a suddenly applied constant load given by Eq.(2.47). The
Laplace image f ( s ) of this load, according to the first row of Table 2.2 reads
f (s) =
F0
s
Thus, the first term in the general image-expression Eq.(2.59) takes the form
F0
1
2
m s ( s + n2 )
29
F0
F
1
0 2 (1 cos (n t ) )
2
2
m s ( s + n ) mn
(2.63)
2.3.
F(t)
x
The dashpot introduces an additional force FDASHPOT ( t ) on the mass relative to the
mass-spring system depicted in Figure 2.3. Therefore, the balance of forces as dictated
by Newtons second law, has one additional term as compared to Eq.(2.2). This
balance reads
(2.64)
FDASHPOT ( t ) = cx
(2.65)
Substituting this expression into Eq.(2.64) and projecting the result onto the x axis,
the following scalar equation is obtained
kx = F ( t )
mx+ cx+
(2.66)
30
(2.67)
c m = 2n
(2.68)
(2.69)
x ( t ) = X n exp ( sn t )
(2.70)
n =1
(2.71)
Note that the characteristic equation Eq.(2.71) could be obtained by applying the
Laplace transform with zero initial conditions to Eq.(2.69).
The characteristic exponents sn (the roots of Eq.(2.71)) can be easily found for the
second-order polynomial Eq.(2.71) to give
s1 = n + n 2 n2 ,
s2 = n n 2 n2
(2.72)
(2.73)
31
n > n or n < n . Going back to notations introduced by Eq.(2.68), we see that this
rests on the relative magnitude of the damping coefficient c and the spring constant
k . Generally speaking, a large coefficient of damping and a small spring constant will
result in real values of s1 and s2 while, for the reverse of these conditions, s1 and s2
will be complex numbers.
Case 1, n > n . In this case, s1 and s2 have real values. To evaluate the constants
X 1 and X 2 in the general solution Eq.(2.73), we must impose the initial conditions of
the motion. Assuming as a particular case, that
x ( 0 ) = x0 ,
x ( 0 ) = 0
(2.74)
s2 x0
,
s1 s2
X2 =
s1 x0
s1 s2
(2.75)
x0
( s1 exp ( s2t ) s2 exp ( s1t ) )
s1 s2
(2.76)
In connection with this solution, it should be noted that both s1 and s2 are negative,
s2 having a bigger absolute value. Thus, the displacement x of the mass has the same
sign as x0 and approaches zero as a limit when the time tends to infinity. The
displacement-time diagram plotted in accordance with Eq.(2.76) is shown in Figure
2.12.
x0
We see that the motion is not a vibration at all but simply one in which the suspended
mass, after its initial displacement, gradually creeps back toward the equilibrium
position but takes theoretically infinite time to get there. This is called aperiodic
motion. It results from the fact that the damping coefficient is too large relative to the
spring constant.
In the special case n = n , we also obtain aperiodic motion, and the corresponding
value of the damping coefficient
32
c = 2nm = 2 km
(2.77)
1 = n2 n2
(2.78)
(2.79)
Using these expressions for the characteristic exponents, and applying the Euler
formula in the same way, as it was done in Eq.(2.14), Eq.(2.73) may be rewritten as
x ( t ) = exp ( nt ) ( A cos (1t ) + B sin (1t ) )
(2.80)
where A and B are new arbitrary constants. To evaluate these constants, we assume
the following general initial conditions:
x ( 0 ) = x0
(2.81)
x ( 0 ) = v0
Substituting Eq.(2.80) into the initial conditions Eq.(2.81) and resolving the soobtained system of two algebraic equations, we find
A = x0 ,
B=
v0
nx0
(2.82)
v nx
x ( t ) = exp ( nt ) x0 cos (1t ) + 0 + 0 sin (1t )
1 1
(2.83)
(2.84)
33
v nx
A0 = x + 0 + 0
1 1
2
0
(2.85)
v + nx0
0 = arctan 0
x01
A0exp(-nt)
A0
T 1=
2
1
x0
0
1
t
-A0 exp(-nt)
Figure 2.13 shows that each time that cos (1t 0 ) becomes equal to 1 , the
34
coincide exactly with the points of the curve representing extreme positions of the
vibrating mass and that, owing to damping, the time interval needed for the mass to
move from an extreme position to a subsequent middle position is greater than that
needed to move from a middle position to the next extreme position. The rate at which
the amplitude diminishes depends upon the damping factor n and can be calculated in
the following manner. Let x1 be the displacement at the first point of tangency with
the exponential curve and xs the displacement after ( s 1) complete cycles. Then it is
evident from Eq.(2.84) that
x1
= exp ( ( s 1) nT1 ) or
xs
x
ln 1 = ( s 1) nT1
xs
(2.86)
These expressions show that the amplitudes at the ends of successive cycles diminish
as a geometric progression. The quantity
nT1 =
2 n
(2.87)
n2 n 2
on which the rate of decay depends, is called the logarithmic decrement of the
amplitude. It can be determined experimentally by observing in what proportion the
amplitude is diminished after an arbitrary number of cycles and then using Eq.(2.86).
Using notation (2.78), we can write the expression for the period of damped free
vibrations as
T1 =
(2.88)
n2 n 2
As the value of n varies from 0 (no damping) to n (critical damping), the period T1
varies from the natural period of undamped vibrations 2 n to infinity, while the
frequency of vibration
n2 n 2
1
f1 = =
2
T1
(2.89)
varies from the frequency of the undamped vibrations n 2 to 0. For small values
of n n , which we usually have in practice, Eq.(2.89), taking the first two terms in
the Taylors expansion of the radical, can be simplified to
n2
f1 = f n 1 2 ,
2n
where
fn =
n
2
(2.90)
This expression shows that the frequency of the damped vibrations differs from that of
the undamped vibrations only by a small quantity of second order. Consequently, in
35
(2.92)
(2.93)
(2.94)
like it was done for the undamped mass-spring system. The reason is that substituting
Eq.(2.94) into Eq.(2.93) we would get the following equation
(2.95)
(2.96)
where X c and X s are constants, the subscripts of which show that they are the
multipliers of the cosinus and sinus, respectively. To determine these constants we
substitute Eq.(2.96) into Eq.(2.93) and obtain
36
( X
2
This equation can be satisfied for all values of t only if the expressions in parentheses
vanish. Thus, for calculating X c and X s , we have the following system of linear
algebraic equations
2 X c + 2n X s + n2 X c = f0
(2.98)
2 X s 2n X c + n2 X s = 0
2
n
2
n
2
n
2 )
2 ) + 4n2 2
2
(2.99)
2n
2 ) + 4n2 2
2
x part = Re X exp ( it )
Substituting
Eq.(2.100)
into
Eq.(2.93)
cos (t ) = Re ( exp ( it ) ) , we obtain
(2.100)
and
using
the
fact
that
(2.101)
This equation can be satisfied for all values of t only if the following equation with
complex coefficients is satisfied:
+ 2in + n2 X = f 0
(2.102)
f0
+ 2in + n2
2
(2.103)
37
f0
exp ( it )
x part = Re
2
2
+ 2in + n
f0
cos (t ) + i sin (t ) )
= Re
2
2 (
+ 2in + n
( 2 2 ) 2in
= f 0 Re
( cos (t ) + i sin (t ) )
( 2 2 )2 + 4n 2 2
n
f0
=
(n2 2 ) cos (t ) + 2n sin (t )
2
2
2
2 2
( n ) + 4 n
(2.104)
which coincides with Eq.(2.96) supplemented by Eq. (2.99). The latter way of
obtaining the particular solution might seem to be more laborious but it is the way that
is mainly used in advanced research on dynamics of systems. The reason for that is
that the complex form of the particular solution
x part = X exp ( it )
(2.105)
is equally suitable for finding this solution in the case of the cosinusoidal external
force F ( t ) = F0 cos (t ) and the sinusoidal one F ( t ) = F0 sin (t ) . In the later case
the particular solution is simply given by
x part = Im X exp ( it )
(2.106)
(2.107)
as discussed in the preceding section, and the forced vibrations have the period
T=
38
identical with the period of the external force that produces them.
Because of the factor exp ( nt ) , which is due to the viscous damping, the free
vibrations gradually subside, leaving only the steady forced vibrations represented by
the last two terms, i.e., by the particular solution Eq. (2.96). These forced vibrations
are maintained indefinitely by the action of the disturbing force and therefore are of
great practical importance.
Let us consider the forced vibrations in detail. To facilitate this, we rewrite the
steady-state solution Eq.(2.96) in a new form, introducing the amplitude and phase lag
of vibrations, similar to Eq.(2.31). This new form reads
x part = X cos (t )
(2.108)
where
X = X c2 + X s2 =
Xc
Xs
= arctan
f0
(2.109)
(n2 2 ) + 4n2 2
2
2n
arctan
=
( n 2 )
(2.110)
Note that the amplitude X and the phase lag can be obtained directly from the
complex amplitude X of the steady-state vibrations, which is defined by Eq.(2.103).
Indeed, as any complex value, this complex amplitude can be rewritten as
( ( ))
X = X exp i arg X
(2.111)
where
( )
X = Re X
( )
+ Im X
( )
( )
f0
2
n
2 ) + 4n 2 2
2
Im X
= arctan 2n =
arg X = arctan
2
2
Re X
( n )
( )
(2.112)
It is easy to see that the amplitude and phase lag of the steady-state vibrations are
equal to the modulus and the argument (with minus) of the complex amplitude X .
From Eq.(2.108) we see that steady-state forced vibration with viscous damping is
a simple harmonic motion having constant amplitude X and phase lag with the
force. As mentioned before, the period of the forced vibration T = 2 is the same
as that of the external force regardless of the natural period of the system and
regardless of the amount of damping.
39
2
2
n2
n n n
n
(2.113)
2n
(2.114)
The ratio of the amplitude of the steady-state vibration X and the corresponding
static deflection xstatic = F0 k is referred to as magnification factor or dynamic
amplification factor. In accordance with Eq.(2.113), it reads
X
xstatic
(2.115)
2
1 2 + 2
n
n
2
One can see that this magnification factor depends on the amount of damping, as
represented by , and on the ratio n of the frequency of the external force and
the natural frequency of the undamped vibrations.
In Figure 2.14, the values of the magnification factor X xstatic are plotted for
various values of the amount of damping against the ratio n .
5.0
=0
4.5
4.0
= 0.2
X /xstatic
3.5
3.0
| 2.5
| 2.0
1.5
1.0
0.5
= 0.3
= 0.5
=1
=2
0.0
0.0
0.5
1.0
1.5
2.0
2.5
/n
From this figure we see that when the frequency of the force is small compared
with the natural frequency n , the value of the magnification factor is not greatly
40
different from unity, regardless of the amount of damping. This means that the
amplitude of the steady-state vibration at the frequency band << n is
approximately equal to the static displacement xstatic . Owing to this property, the
vibrations at this frequency band are often called quasi-static vibrations.
When >> n , i.e., when the frequency of the force is much greater that the
natural frequency, the value of the magnification factor tends to zero, regardless of the
amount of damping. This means that a high-frequency external force produces
practically no forced vibrations of a system that has a low natural frequency.
Since in both extreme cases << n and >> n , the damping has only a
marginal effect on the magnification factor, it is justifiable in discussing forced
vibrations to neglect the effect of damping entirely, in which case Eq.(2.115) takes the
much simpler form given by Eq.(2.33):
X =
F0
1
k 1 2 n2
As the value of approaches that of n , i.e. as the frequency of the force gets
close to the natural frequency, the magnification factor grows rapidly and, as we see
from Figure 2.14, its value is very sensitive to the amount of damping as represented
by . It can also be noted that the maximum value of the magnification factor occurs
for a value of n slightly smaller than unity, i.e., slightly below resonance. Setting
the derivative of the magnification factor with respect to n to zero, we find that
the maximum occurs when
2
= 1
n
2
(2.116)
Substituting this relationship into Eq.(2.115), it can be found that the maximum
magnification depends on the amount of damping according to the following formula:
X
2
=
2
xstatic max 4
(2.117)
From this formula it follows that if 2 , the magnification factor as a function of the
frequency of external force has no maximum. This means that in this case, the largest
value of the magnification factor is equal to unity and is achieved at the zero
frequency of the external force. In yet other words, this implies that no dynamic
amplification is possible if 2 . Note that the damping that corresponds to = 2 is
precisely the critical damping ccrit = 2 km as defined by Eq.(2.77).
In most practical situations, << 1 and it is justifiable to consider that the
maximum value of the magnification factor occurs when n and the maximum of
the magnification factor can be approximated by
X
1
xstatic max
(2.118)
41
Both Eq.(2.117) and Eq.(2.118) show that for small damping, the amplitude of forced
vibration can become extremely large under the condition of resonance. Since large
amplitudes mean large stresses in the spring, the condition of resonance is usually to
be regarded as a dangerous one and is to be avoided whenever possible.
Let us turn now to the phase relationship between the forced vibrations and the
external force producing them. This is represented by the phase angle in Eq.(2.108)
, the expression for which can be rewritten making use of , Eq.(2.114), as
2
2
n 1 n
= arctan
(2.119)
We recall that since the external force varies according to cos (t ) and the forced
vibrations according to cos (t ) , the phase angle represents the phase lag of
the vibrations behind the external force. That is, when the external force F ( t ) in
Figure 2.11 is directed rightward, the mass on which it acts is not yet in its utmost
right position but arrives there seconds later.
Eq.(2.119) shows that the value of the phase lag, like that of the magnification
factor, depends both upon the relative amount of damping and upon the ratio
n . The curves in Figure 2.15 show the variation of the phase lag with the ratio
n for several values of the damping factor . Note that arctan ( x ) is a multivalued function, therefore, to obtain the continuous graphs in Figure 2.15, one has to
plot the principal value Arctan ( x ) if the argument x is positive and the principal
value plus (if the graph is in radians) or plus 180o (if the graph is in degrees) if the
argument is negative.
180.0
=0
= 0.25
[deg]
= 0.5
90.0
=2
=1
0.0
0.0
0.5
1.0
1.5
2.0
2.5
/n
We see that when = 0 , the phase lag is discontinuous (this was also shown in
Figure 2.8). We recall that in this case the forced vibrations are exactly in phase with
the disturbing force for all values of n smaller than unity and out of phase for all
values of n greater than unity. When damping is present, we can see a continual
change in as the ratio n increases. Also, regardless of damping, = 90 0 = 2
42
at resonance. That is, at resonance, the forced vibrations lag behind the external force
by one quarter cycle.
For values of n either well below resonance or well above it, we note that a
moderate amount of damping ( << 1) has only a secondary effect on the phase lag.
That is, well below resonance, it is practically zero, while well above resonance, it is
practically 180 0 = . This means, again, that in discussing forced vibrations well
away from the condition of resonance, the effect of damping can be ignored.
The frequency dependence of the phase lag presented in Figure 2.14 allows us to
identify the frequency bands, at which particular elements of the mass-spring-dashpot
system play a major role. To carry out such identification, we can use Eq.(2.101),
which, after multiplying through by m , reads
Re ( m 2 + ic + k ) X exp ( it ) F0 exp ( it ) = 0
(2.120)
Employing the representation for the complex amplitude X given by Eq. (2.111) and
taking into account that X = X and arg X = , we rewrite Eq.(2.120) as
( )
(2.121)
To satisfy this equation, the expression in the figure brackets must vanish, hence
dividing by exp ( it i ) we obtain
( m
+ ic + k X = F0 exp ( i )
This equation, using the following equalities substantiated by the Euler formula
Eq.(2.13),
1 = exp ( i ) , i = exp ( i 2 )
(2.122)
can be rewritten as
m 2 X exp ( i ) + c X exp ( i 2 ) + k X = F0 exp ( i )
(2.123)
The three terms on the left-hand side represent the forces produced by the mass,
dashpot and spring respectively, whereas the term on the right-hand side is the
external force. We can clearly see now that while the external force leads the
displacement by the phase angle , the phases of the other forces are as follows:
the spring-force is in phase with the displacement,
the dashpot-force leads the displacement by 2 ,
the mass-force leads the displacement by .
Using this information, we can take a fresh look at the frequency dependence of
the phase lag of vibrations presented by Fig. 2.15. This figure shows that when
<< n , the phase lag between the force and the displacement is almost zero, which
coincides with the phase of the spring-force. This implies that the major role in low-
43
(2.124)
x ( 0 ) = v0
assuming that the external force per unit mass f ( t ) is an arbitrary function of time.
f ( t ) dt
sin (1 ( t t ) )
(2.125)
where 1 = n2 n2 . Note that Eq.(2.125) is valid both for n2 > n 2 and n2 < n 2 . In
the latter case, 1 has an imaginary value.
Since each impulse f ( t ) dt between t = 0 and t = t has a like effect, we obtain,
as a result of the continuous action of the external force, the following displacement
of the mass
x (t ) =
1 0
f ( t ) exp ( n ( t t ) ) sin (1 ( t t ) ) dt
(2.126)
44
v nx
x ( t ) = exp ( nt ) x0 cos (1t ) + 0 + 0 sin (1t )
1 1
1 0
(2.127)
f ( t ) exp ( n ( t t ) ) sin (1 ( t t ) ) dt
0,
F (t ) =
F0 ,
t<0
(2.128)
t0
Substituting this expression in the general formula Eq.(2.40), and assuming that the
initial displacement x0 and initial velocity v0 are zero, we obtain
x=
(2.129)
= n t ,
(2.130)
we rewrite Eq.(2.129) as
x
xstatic
4 2
4 2
= 1 exp cos
+
sin
2
2
2
4 2
(2.131)
The curves in Figure 2.16 show the variation of the relative displacement x xstatic with
the dimensionless time for three values of the damping as represented by .
2.0
=0
= 0.5
x/xstatic
1.5
1.0
0.5
= 1.9
0.0
0
10
15
20
25
45
This figure shows that when = 0 the mass oscillates indefinitely long with the
amplitude equal to the static deflection around the static equilibrium x xstatic = 1 . This
result coincides with that of Section 2.2.3. When 0 , i.e., in the presence of
damping, the mass sooner or later comes to its static equilibrium. The larger the
damping, the faster this happens. Two scenarios are possible here. If damping is
moderate as exemplified by = 0.5 some oscillations around the static equilibrium
take place before the mass gets in rest in its static position. If damping is large, as
represented by = 1.9 , the mass just creeps to the static position demonstrating no
oscillations whatsoever. Thus, under the action of a suddenly applied constant force,
any system with damping (read: any engineering system) will sooner or later get to its
static equilibrium, regardless of the initial conditions. If the constant force is removed
at a time instant the system, owing to damping, will get to its undisturbed equilibrium
x =0.
2.4.
In the preceding sections we discussed the theory of the vibrations of a system with a
single degree of freedom. Though the exact idealized system with which the theory
dealt occurs rarely, a number of actual cases are sufficiently close to the ideal to
permit conclusions of practical importance. The theory of the single-degree-offreedom system enables us to explain the resonance phenomenon in many machines,
to calculate natural frequencies of a number of structures, and to discuss the main
principles of vibration isolation.
This exhausts the possibilities of application pretty thoroughly, and in order to
explain additional phenomena it is necessary to develop the theory of more
complicated systems. As a first step, in this section we consider a system with two
degrees of freedom, which will yield the explanation of most vibration dampers
applied in practice.
The most general undamped two-degree-of-freedom system can be reduced to that
shown in Figure 2.17 and consists of two masses, which rest on a frictionless
horizontal plane, and three springs. The springs k1 and k2 connect the respective
masses to the walls, while the spring k3 ties the masses together, thereby coupling
their motion. The external loading is represented by the forces F1 ( t ) and F2 ( t ) ,
which are applied to the masses m1 and m2 , respectively.
m1
k1
F1(t)
F2(t)
x1(t)
k3
x2(t)
m2
k2
x
To write down the equations of motion for this system, we use the second Newtons
law. Each mass is subject to three forces, namely the external force and two springforces. Therefore, the vector-form of the equations of motion reads
46
(2.132)
where x1 ( t ) and x2 ( t ) are the displacements of the masses from their equilibrium
positions, in which the springs are undisturbed. The spring-forces imply the
following: FSPRING 1-1 is the vector-force imposed on m1 by spring k1 ; FSPRING 2 2 is
the vector-force imposed on m2 by spring k2 ; FSPRING 3-1 is the vector-force imposed
on m1 by spring k3 ; FSPRING 3-2 is the vector-force imposed on m2 by spring k3 .
According to the Hookes law, the vector-spring-forces can be expressed as
FSPRING 1-1 = k1 x1 ,
FSPRING 2-2 = k2 x2 ,
(2.133)
Substituting these expressions into Eqs.(2.132) and projecting the result onto the
x axis, the following scalar equations are obtained
m1
x1 + k1 x1 + k3 ( x1 x2 ) = F1 ( t )
m2
x2 + k2 x2 + k3 ( x2 x1 ) = F2 ( t )
(2.134)
These are the equations of motion, which describe small vibrations of the undamped
two-degrees-of-freedom system near its equilibrium position under the action of the
external forces F1 ( t ) and F2 ( t ) .
Retrieving the equations of motion of a dynamic system, one has to project the
balances of forces written in the vector-form on the chosen coordinate axes. If the
system can move in one direction only, the operation of projecting the forces is easy.
If, however, a planar or spatial motion is considered, this operation can prove to be
quite laborious and vulnerable to men-introduced errors. To avoid such errors, the socalled Lagrangian formalism can be used, which is shortly described below.
The Lagrangian formalism is based on the Lagrangian equations, which read
d L L
= Qs
dt qs qs
(2.135)
where
L =KP
(2.136)
is the Lagrange function, which is defined as the difference between the kinetic
energy and the potential energy of the system under consideration; qs and qs ,
s = 1, 2,3,... are the generalized coordinates and velocities of the system; Qs are the
generalized forces acting on the system. The term generalized reflects the fact that
qs can represent not only a displacement but an angle as well, or, for example, the
voltage in an electric machine. The system presented in Figure 2.16 has two degrees
47
of freedom, which are specified by the two displacements. This implies that for this
system s = 1, 2 , q1 = x1 , q2 = x2 , Q1 = F1 , Q2 = F2 .
The Lagrangian equations are based on a very general principle that any system is
smart enough and if it had to move from one point to another one, being
unconstraint, it would choose the way, which requires the minimum amount of energy
to be spent. The mathematical developments, which lead to the Lagrangian equations,
can be found in any text book on theoretical mechanics.
As an example, let us apply the Lagrangian equations to obtain the equations of
motion for the system shown in Figure 2.17. To this end, we have to write the kinetic
and the potential energies of this system. The kinetic energy is the sum of the kinetic
energies of the two masses, whereas the potential energy is the sum of the potential
energies of the three springs. Therefore, the following expressions hold
(
(
1
2
2
K
= m1 ( x1 ) + m2 ( x2 )
2
1
2
2
2
P = k1 ( x1 ) + k2 ( x2 ) + k3 ( x1 x2 )
2
(2.137)
L L K
=
=
= m2 x2 ,
q2 x2 x2
L L
P
=
=
= ( k1 x1 + k3 ( x1 x2 ) ) ,
q1 x1
x1
(2.138)
L L
P
=
=
= ( k2 x2 k3 ( x1 x2 ) )
q2 x2
x2
Substitution of these expressions into the Lagrangian equations Eq.(2.135) gives the
equations of motion, which coincide with Eq.(2.134).
The equations of notion, Eq.(2.134) can be written in the matrix form, which is
widely used for numerical analyses. The matrix form of Eq. (2.134) reads
M
x+K x=F
(2.139)
where x = { x1 , x2 }
vector, and
m 0
M = 1
,
0 m2
and F = { F1 , F2 }
k + k
K = 1 3
k3
k3
k2 + k3
(2.140)
48
m1
x1 + k1 x1 + k3 ( x1 x2 ) = 0
(2.141)
m2
x2 + k2 x2 + k3 ( x2 x1 ) = 0
x1 ( t ) = X n( ) exp ( snt ),
x2 ( t ) = X n( ) exp ( sn t )
n =1
(2.142)
n =1
1
1
2
m1 X n( ) sn2 exp ( snt ) + k1 X n( ) exp ( snt ) + k3 X n( ) exp ( snt ) X n( ) exp ( snt ) = 0
n =1
n =1
n =1
n =1
4
4
4
4
2
2
2
1
m2 X n( ) sn2 exp ( snt ) + k2 X n( ) exp ( snt ) + k3 X n( ) exp ( snt ) X n( ) exp ( snt ) = 0
n =1
n =1
n =1
n =1
{ X ( ) ( m s
+ k1 + k3 ) + X n(
{ X ( m s
+ k2 + k3 ) + X n ( k3 ) exp ( snt ) = 0
1
n
2
1 n
n =1
4
( 2)
n =1
2
2 n
2)
( k3 )} exp ( snt ) = 0
(1)
(2.143)
In order these equations be satisfied at any instant of time, the expressions in the
figure brackets must vanish, that is the following system of algebraic equations must
be satisfied:
X n( ) ( m1sn2 + k1 + k3 ) + X n(
1
X n(
2)
(m s
2
2 n
2)
( k3 ) = 0
+ k2 + k3 ) + X n( ) ( k3 ) = 0
1
(2.144)
This system of equations (2.144) has a non-trivial solution if the determinant of its
coefficient matrix vanishes. The coefficient matrix of Eq.(2.144) is given as
m1sn2 + k1 + k3
k3
2
= Msn + K
2
k
m
s
+
k
+
k
3
2 n
2
3
(2.145)
(m s
1
+ k1 + k3 )( m2 s 2 + k2 + k3 ) k32 = 0
49
k +k k +k k k +k k +k k
s4 + s2 1 3 + 2 3 + 1 2 1 3 2 3 = 0
m2
m1m2
m1
(2.146)
a2 =
k1 + k3
k +k
k3
, b2 = 2 3 , ab2 =
m1
m2
m1m2
(2.147)
The quantities a and b are the frequencies of the system in which one of the
masses is held clamped (the so-called partial frequencies), whereas ab represents the
strength of the coupling between the masses. With these notations and s = i ,
Eq.(2.146) can be written as
(2.148)
The roots of this equation are the natural frequencies of the two-degrees-of-freedom
system under consideration.
It is interesting to note that Eq.(2.148) can be analysed geometrically with the help
of Mohrs circle diagram, which may be familiar to the reader from two-dimensional
elasticity. The Mohrs circle diagram for Eq.(2.147) is shown in Fig.2.18.
C
O
In this diagram
OA = a2
OB = b2
BC = ab2
The circle is drawn through C about the mid-point between A and B as centre. The
new points D and E thus found, determine the natural frequencies of the system:
n21 = OD
n22 = OE
50
which can be verified from the equation. Figure 2.17 shows, in particular, that the
interval between the natural frequencies of n1 and n 2 is always greater than or
equal to that between the partial frequencies a and b .
The analytical expressions for the natural frequencies can be found from
Eq.(2.148) to give
1
a2 + b2
2
1
n 2 =
a2 + b2 +
2
n1 =
2
a
b2 ) + 4ab4
2
(2.149)
2
a
2 2
b
+ 4
4
ab
It is obvious that the expression under the inner square root in Eq.(2.149) is positive.
It can also be easily checked that a2 + b2 >
2
a
both natural frequencies are real. Thus, the general solution Eq.(2.142) of Eq.(2.141),
following similar rearrangements as presented in Eq.(2.14), can be rewritten as
x1 ( t ) = A1 cos (n1t ) + B1 sin (n1t ) + C1 cos (n 2t ) + D1 sin (n 2t )
(2.150)
Eq.(2.150) shows that the free vibrations of the system take place at two distinct
frequencies. The amplitudes of vibrations at these frequencies depend on specific
initial conditions. Correspondingly, the complete vibration pattern can be quite
different, depending on the natural frequencies and initial conditions. We may expect,
however, that if the natural frequencies are close to each other, the beats should occur
leading to a time-displacement diagram similar to that depicted in Figure 2.5.
The free vibrations of the one-degree-of freedom system considered in the
preceding sections are fully characterised by its natural frequencies and the initial
conditions. This is not the case, however, with the two-degree-of freedom system. The
latter requires knowledge of the relationship between the amplitudes of vibrations of
the two degrees-of-freedom, that is A2 A1 , B2 B1 , C2 C1 , D2 D1 . It should be
underlined that these ratios are independent of the initial conditions. To find these
ratios, Eq.(2.150) can be substituted into the equations of motion Eq.(2.141), which,
employing the notations Eq.(2.147), can be rearranged to
x1 + a2 x1 ab2 x2 = 0
x2 + b2 x2 ab2 x1 = 0
(2.151)
51
cos (n1t ) A1 (a2 n21 ) A2ab2 + sin (n1t ) B1 (a2 n21 ) B2ab2
cos (n1t ) A2 (b2 n21 ) A1ab2 + sin (n1t ) B2 (b2 n21 ) B1ab2
To have these equations satisfied at any time instant, all expression in the figure
brackets must vanish. Thus, the following equations must be satisfied:
A2
1
=
(a2 n21 ) ,
A1 ab2
A2
ab2
=
A1 (b2 n21 )
B2
1
=
(a2 n21 ) ,
B1 ab2
ab2
B2
=
B1 (b2 n21 )
C2
1
=
2 n22 ) ,
2 ( a
C1 ab
C2
=
C1 (b2 n22 )
D2
1
=
2 n22 ) ,
2 ( a
D1 ab
ab2
D2
=
D1 (b2 n22 )
(2.152)
2
ab
We see that there are two expressions for every ratio of the amplitudes, which should
be satisfied simultaneously. It is easy to check that they are satisfied simultaneously
indeed, since n21 and n22 are the roots of Eq.(2.148). This, of course, is to be
expected since Eq.(2.148) is nothing else but the condition of linear dependence of
the equations in the system Eq.(2.141), which implies that if one of the two equation is
satisfied then the other one is satisfied automatically.
Using the ratios of amplitudes given by Eq.(2.152), the general solution Eq.(2.150)
can be given the following more specific form:
x1 ( t ) = A1 cos (n1t ) + B1 sin (n1t ) + C1 cos (n 2t ) + D1 sin (n 2t )
x2 ( t ) = ( A1 cos (n1t ) + B1 sin (n1t ) )
+ ( C1 cos (n 2t ) + D1 sin (n 2t ) )
2
ab
2
ab
2
a
n21 )
2
a
(2.153)
n22 )
This form of the general solution contains four unknown constants, which can be
uniquely determined employing the initial conditions (the initial displacement and
initial velocity for x1 and those for x2 ).
The general solution Eq.(2.153) shows that the amplitudes of the free vibrations of
the masses m1 and m2 are intrinsically related to each other. Thus, the free vibrations
of the two-degrees-of-freedom system are characterised not only by the natural
frequencies but also by the shapes of vibrations at these frequencies. These shapes are
52
normally referred to as normal modes and imply a specific motion of the masses with
respect to each other.
To gain insight in what the normal modes mean physically, let us simplify the
system somewhat by making it symmetrical. To this end we assume that k1 = k2 = k
and m1 = m2 = m . In this case, a2 = b2 , = 1 and the expressions Eq.(2.149) for the
natural frequencies reduce to
n1 = a2 ab2 = k m
n 2 = a2 + ab2 =
( k + 2k3 )
(2.154)
(2.155)
To find the natural frequencies of the system, the following determinant should be
set to zero:
53
det 2 M + K = 0
m1
x1 + k1 x1 + k3 ( x1 x2 ) = F0( ) cos (t )
1
(2.156)
m2
x2 + k2 x2 + k3 ( x2 x1 ) = F0( ) cos (t + 0 )
2
We will study only the forced steady-state vibrations of the system, since these
vibrations are commonly of most practical importance.
We already know that the steady-state vibrations take place at the frequency of the
force. Therefore, these vibrations may be sought in the form, which is a supermosition
of cosinusoidal vibrations dictated by the forces on the right-hand side of Eq. (2.156),
x1 = A1 cos (t ) + B1 cos (t + 0 )
(2.157)
x2 = A2 cos (t ) + B2 cos (t + 0 )
{ m A + k A + k A k A F ( )} cos (t )
1
1 1
1 1
+ { 2 m1 B1 + k1 B1 + k3 B1 k3 B2 } (t + 0 ) = 0
{ m A + k A
2
+ k3 A2 k3 A1} cos (t )
(2.158)
+ 2 m2 B2 + k2 B2 + k3 B2 k3 B1 F0( 2) (t + 0 ) = 0
To satisfy these equations at any time instant the expressions in the figure brackets
must vanish. This leads to the following two systems of algebraic equations, from
which the amplitudes of the steady-state vibrations can be found:
2 m1 A1 + k1 A1 + k3 A1 k3 A2 = F0(1)
2
m2 A2 + k2 A2 + k3 A2 k3 A1 = 0
(2.159)
2 m1 B1 + k1 B1 + k3 B1 k3 B2 = 0
2
(2)
m2 B2 + k2 B2 + k3 B2 k3 B1 = F0
(2.160)
54
Let us consider from hereon a simplified but a very practically important example,
which is depicted in Figure 2.19. In this example, we consider a two-degrees-offreedom system, which consists of a main system K , M subject to a harmonic force
F0 cos (t ) and an attached auxiliary system k , m . This model can schematise, for
example a bridge with an auxiliary mass elastically attached in the middle, Fig.
2.20(a), or a vibrating machine with an auxiliary mass-spring system, Fig. 2.20(b).
This auxiliary mass-spring system is normally referred to as an undamped dynamic
vibration absorber. It will be shown that the parameters of this absorber can be
chosen such that the main system (the bridge or the machine) does not vibrate
although it is subject to the external harmonic force.
F0 cos (t )
x1(t)
k
x2(t)
m
K
x
Figure 2.19. Vibrations absorber k-m attached to the main system K-M (bridge, machine, etc.)
(a)
Bridge beam
(b)
k
k
m
K/2
K/2
Auxiliary mass
Figure 2.20. Examples of application of the vibration absorber: (a) bridge with an elastically
suspended mass; (b) machine with an elastically mounted mass
(2.161)
k2 = 0, m2 = m, F0( ) = 0
2
As follows from Eq.(2.160), since F0( ) = 0 , the amplitudes B1 and B2 are trivially
zero. The system of equations for A1 and A2 , Eq.(2.159), is simplified to
2
(2.162)
55
A1 (a2 2 ) ab2 A2 = F0 M
2
2
2
A2 (b ) ab A2 = 0
(2.164)
2n (b2 2 )
2nab2
(2.165)
56
A1
xstatic
A2
xstatic
=
=
n2 (b2 2 )
2
n21 )( 2 n22 )
(
2
2
n
2
n1
2
ab
)(
(2.166)
n22 )
A/x static
In Figure 2.21 the displacement-frequency curves A1 xstatic and A2 xstatic are plotted
qualitatively versus the frequency of the external force.
n1
n2
A1
A2
Figure 2.21 shows clearly two resonances at the natural frequencies of the system.
Furthermore, as noticed earlier, we see that the main system does not move at the
partial frequency b of the auxiliary system. Concerning the phases of vibrations,
Figure 2.21 allows to conclude that (i) for < n1 the masses vibrate in phase with
the load and the phase lag between vibrations of the masses is zero, although the
amplitudes of the vibrations are different; (ii) for n1 < < b the masses are out of
phase with the load but the phase lag between vibrations of the masses is still zero;
(iii) for b < < n 2 the masses start to vibrate out of phase with each other, the main
mass keeping in phase with the load; (iii) for n 2 < the masses are out of phase
with each other, but the main mass vibrates out of phase with the load.
Concluding this section, it is important to note that the displacement-frequency
curves in Figure 2.21 are plotted for the undamped two-degrees-of-freedom system. If
the viscous damping were included, these curves would be different, especially in the
vicinity of resonances and around b .
57
CHAPTER 3
ONE-DIMENSIONAL STRUCTURES
Most civil engineering systems and structures are continuous and have distributed
material properties. A bridge, a chimney, and a railway track, for example, can be
considered as beams with uniform mass per unit length and uniform bending stiffness.
Landing strips of an airport have properties distributed over their surface. The soil
properties are distributed all over the volume. The fact that the material properties are
distributed continuously suggests that to describe the mechanical behaviour of these
systems and structures, we need to quantify their mechanical behaviour with
characteristics (displacements, forces, etc.) which depend both on time and position.
Analyses of models, which take into account both these dependencies, form the
subject of continuum mechanics.
The mechanical behaviour of many structures such as long bridges, chimneys,
railway tracks, skyscrapers, etc., which are much longer in one direction than in the
two other ones, depends mainly on the time and position along the structure.
Consequently, modelling these structures, one may assume a priory that the
displacements, forces and moments in these structures depend on time and the alongthe-structure co-ordinate only. Such models are referred to as one-dimensional models
and the structures, which these models represent, are often called one-dimensional
structures.
Though the mechanical characteristics of the one-dimensional structures depend on
one co-ordinate only, this dependence is continuous. This makes a principal
theoretical difference between the lumped models considered in the previous chapter
and the continuous models, which will be treated in this chapter. To describe the
mechanical behaviour of a continuous system, theoretically speaking, we have to track
the displacement time diagrams, for example, of infinitely many points. Therefore,
continuous models are often referred to as models with infinite number of degrees of
freedom. In practice, however, mechanical properties of structures vary in space
relatively slowly. Therefore, it is always possible to characterise the mechanical
behaviour of such structures using a finite-degree-of-freedom system, increasing the
accuracy of characterization by increasing the number of the degrees of freedom to be
considered.
3.1.
58
F(t)
u (t, z )
w (t, x )
x
F(t)
Figure 3.2. Idealization of vertical vibration of a lorry on a bridge by a simply supported beam
with a lumped mass in the middle
In Figure 3.1. the relatively heavy deck of the offshore platform is replaced by a
lumped mass while the braced steel jacket structure is replaced by a beam, which may
perform the bending motion in the plane of the figure. The model in Figure 3.1 is
more advanced than that in Figure 2.1 because the distributed mass and distributed
stiffness of the supporting structure is now taken into account. Basically, accounting
for the distributed properties of the supporting structure implies that we take account
of the dependence of the support reaction on both the amplitude and frequency of the
platform vibrations. This was not the case with the model in Figure 2.1, where the
reaction was considered as proportional to the displacement of the platform only.
In Figure 3.2. the bridge girder is idealized as a simply supported beam with its
mass and stiffness uniformly distributed over the bridge length. The advantage of the
model in Figure 3.2 relative to that in Figure 2.2 is that now the reaction of the bridge
against the lorry depends not only on the amplitude but also on the frequency of the
lorry vibrations.
Finalizing this section, it is worth noting that the larger the mass of the platform
and the lorry in Figures 3.1 and 3.2, the more accurate predictions can be obtained of
vibrations of the platform and lorry using the one-degrees-of-freedom systems
depicted in Figures 2.1 and 2.2. Thus, the one-degrees-of-freedom models are often
capable of predicting vibrations of heavy bodies attached to elastic structures. What
they are incapable of, however, is the prediction of the shape of elastic structures such
as the braced steel jacket structure and the bridge in Figures 3.1 and 3.2. Thus, to
study this shape and, accordingly, the deformations and stresses in these structures,
the continuous models should be applied.
In the sections to follow five one-dimensional continuous models will be first
formulated and then analyzed. These models represent the following basic
deformations of one-dimensional structural elements: (1) transverse deflection of a
cable; (2) longitudinal compression and extension of a rod; (3) torsion of a rod; (4)
pure shear of a beam; (5) bending of a beam.
59
3.2.
q1 ( x, t )
T
x
V
T
T
q1 ( x, t )
w
s
w
(a)
V + V
H
T + T
x
(b)
Figure 3.1. Taut string: (a) taut string subject to a distributed force; (b) differential element of a
displaced from the rest position (horizontal line) string
Consider a differential element x of the string. The forces acting on this element
once it is displaced from its rest position are shown in Figure 3.1(b). Assuming that
the horizontal motion of the string is negligible so that every element of the string
moves just vertically, we may write the following equation, which is the projection of
the vector form of second Newtons law on the vertical (directed downward) axis:
Ax
2w
= V + V V + q1x = V + q1x
t 2
(3.1)
where w ( x, t ) is the vertical deflection of the string, is the mass density of the
string material and A is the cross-sectional area of the string. Note that on the lefthand side of Eq.(3.1) we introduced the partial derivative with respect to time, which
is not necessarily equal to the ordinary derivative as it is applied to a function of both
time and space.
Dividing Eq.(3.1) by x and taking into account that the vertical component V of
the tensile force T is related to its horizontal component H as
60
V = H tan ( )
(3.2)
2 w ( H tan ( ) )
=
+ q1
x
t 2
(3.3)
Assume that the vibrations of the string are small so that the following relation
holds
FG w IJ
H x K
<< 1.
(3.4)
which implies that the slope of the string should be small. This condition is usually
referred to as the condition of small vibrations. Under this condition, the following
approximation holds
tan ( )
w
x
(3.5)
Substituting Eq.(3.5) into Eq.(3.3), and taking the limit x 0 we obtain the
following equation
2 w ( x, t )
t 2
w ( x, t )
H
+ q1 ( x, t )
x
x
(3.6)
2 w ( x, t )
t 2
w ( x, t )
T
+ q1 ( x, t )
x
x
(3.7)
Equations (3.6) and (3.7) are partial differential equations that describe small
vibrations of the string about its equilibrium. These equations can predict the string
response both to initial conditions and the external loading q1 ( x, t ) provided that the
predicted vibrations do not violate the assumption Eq.(3.4). Note that tension T can
be a function of the co-ordinate x in Eq.(3.7), which is the case, for example, of long
catenary risers is considered.
3.2.2. Longitudinal Motion of a Rod. Modelling the longitudinal motion of a thin
rod it is assumed that its particles move along the main axis only. Such motion is
important for piles driven into soil. In statics, the rod model can be applied for
studying stresses in such structures as chimneys and skyscrapers to roughly analyze
61
stresses induced by the dead weight. Another important application of the model is
found when stresses are analysed in constraint structural elements undergoing thermal
expansion.
Consider a straight, prismatic rod as shown in Figure 3.2(a). The co-ordinate x
refers to a cross-section of the rod, while the longitudinal displacement of this section
is given by u x , t . We presume the rod to be under a dynamically varying stress field
x , t , so that adjacent sections are subjected to varying stresses. A body force
q x , t is also assumed present.
b g
b g
b g
(a)
bg
x + x
(b)
Figure 3.2. A thin rod: (a) with co-ordinate x and displacement
on element x of the rod
With these assumptions the equation of motion of the rod in the x direction reads
Ax
2u
= ( x ) A + ( x + x ) A + qAx
t 2
(3.8)
where is the mass density of the rod material and A is the cross-sectional area of
the rod. The latter is a constant in this development, since we are considering a
prismatic rod. We note that tensile stress is assumed positive.
By taking into account the following Taylor expansion of x + x
g bg
x + x x +
x
x
(3.9)
Eq.(3.8) reduces to
2u
=
+q
t 2 x
(3.10)
Material effects have not been introduced yet, so equation (3.10) is applicable to nonelastic as well as elastic problems. We now presume that the material behaves
elastically and follows the simple Hookes law
62
= E
(3.11)
where E is the Youngs modulus and is the axial strain that for the case at hand is
defined by
= u x
(3.12)
2u u
=
E
+q
t 2 x x
(3.13)
One can easily recognise this equation to be identical in form to Eq.(3.6) derived for
the taut string.
If the rod is homogeneous so that E does not vary with x , equation (3.13) may be
written as
2u
2u
2 = E 2 +q
t
x
(3.14)
On the other hand, if the rod is not prismatic, so that its cross-sectional area A is a
function of x , Eq.(3.13) can be generalized to
2u
u
A 2 = EA + qA
t
x
x
(3.15)
These equations can predict the longitudinal motion of the thin rod both due to initial
conditions and the external loading q ( x, t ) provided that the strain-state remains
elastic and approximately uniaxial.
3.2.3. Torsion of a Rod. The torque may occur in many structures. For example, in
bridges loaded by heavy vehicles, in skyscrapers under wind loading, in rails loaded
by fright trains, etc. In reality, any rotational motion of a structure is accompanying by
warping, that is by deviation of every cross section from its original plane shape. The
less similar the structural cross-section to the circle, the greater the warping.
Performing a rough structural analysis, however, the warping can be neglected, which
will be followed in this development.
Let us obtain the equation of motion governing small torsion of a generalized rod.
To this end, we consider a straight, cylindrical rod, a differential element of which is
shown in Figure 3.3 (the circular cross-section is shown as an example). The coordinate x refers to a cross-section of the rod, while the angle of twist is given by
( x, t ) .
63
( x + x )
M t ( x + x )
Mt ( x)
x
m1
x + x
Figure 3.3. Differential element of a rod subject to end torques and distributed body torque m1 (x,t)
J x
2
= M t ( x ) + M t ( x + x ) + m1 x
t 2
(3.16)
where is the mass density of the rod material and J is the polar moment of inertia.
By employing the following Taylor expansion of M t ( x + x )
M t ( x + x ) M t ( x ) +
M t
x
x
(3.17)
2 M t
=
+ m1
x
t 2
(3.18)
(3.19)
where G is the shear modulus of the rod material and J t is the torque constant. For
one-dimensional structural elements with circular cross-section this constant is
numerically equal to the polar moment of inertia J . For all other shapes of the crosssection, J and J t differ. As established from elasticity analysis, the torque constants
for the following four cross-sections are:
64
circle of radius a
Jt = J =
a 3b 3
a2 + b2
a4
a4
15 3
a 3b
3
= GJ t
2
x
x
t
+ m1
(3.20)
One can easily recognise this equation to be identical in form to Eq.(3.6) derived for
the taut string and Eq.(3.13) derived for the rod in longitudinal motion.
Equation can predict a small rotational motion of one-dimensional structures both
due to initial conditions and the external torque m1 ( x, t ) in the case that the structural
shape does not cause considerable wrapping.
3.2.4. Pure Shear of a Beam. Some structures such as multi-storey buildings with
weak main walls may experience deformations which are very close to pure shear
deformations. For example, each storey of a high-rise building shaken by an
earthquake would move nearly horizontally. This deflection can be roughly predicted
with a model, which can be referred to as a shear beam, which is introduced in this
section.
Consider a straight, prismatic beam as shown in Figure 3.4. The co-ordinate x
refers to a cross-section of the beam, while the vertical displacement of this section is
given by w ( x, t ) . We presume the beam to be under a dynamically varying shear
b g
65
x
x
x
w
( x + x )
( x)
q
x
With these assumptions the equation of motion of the beam in the vertical direction
reads
Ax
2w
= ( x ) A + ( x + x ) A + qAx
t 2
(3.21)
where is the mass density of the beam material and A is the cross-sectional area of
the beam, the latter being a constant in this development.
Expanding ( x + x ) into the Taylor series and dividing by Ax we obtain from
Eq.(3.21)
2 w
=
+q
t 2 x
(3.22)
We now assume that the beam material behaves elastically and follows the simple
Hookes law while experiencing shear deformation:
= G
(3.23)
where G is the shear modulus and is the shear strain that for the case at hand is
defined by
= w x
(3.24)
Substituting the constitutive relation Eq.(3.23) into the equation of motion Eq.(3.22),
and using the kinematic relation Eq.(3.24), we obtain
2 w w
=
G
+q
t 2 x x
(3.25)
66
This equation is once again identical in form to the equations of motion obtained in
the preceding sections.
If the beam is homogeneous so that G does not vary with x , equation (3.13) may
be written as
2w
2w
=
+q
G
t 2
x 2
(3.26)
2w
w
= GA + qA
2
t
x
x
(3.27)
These equations can predict the shear motion of the thin beam both due to initial
conditions and the external loading q ( x, t ) provided that the shear strain plays the
major role in the deformation process.
3.2.5. Bending of a Beam. Bending of beams is the most common phenomenon in
structural engineering. Nearly every structure contains beams as structural elements
and a beam-in-bending model can be extremely instrumental both for dynamic and
static analyses of many structures.
Consider a straight beam undergoing transverse motion, as shown in Figure 3.5(a)
and consider a differential element of the beam as isolated in Figure 3.5(b).
q1 ( x, t )
x
M (x )
q1 ( x, t )
M ( x + x )
V (x )
V ( x + x )
(a)
(b)
Figure 3.5. (a) Beam undergoing transverse motion and (b) differential element of the beam subject
to the shear force, bending moment and external load
Bending moment M , shear force V and a distributed external force with density
per unit length q1 ( x, t ) act on the beam element. We invoke the basic hypothesis of
the Euler-Bernoulli theory of beams: namely that the plane cross-sections initially
perpendicular to the axis of the beam remain plane and perpendicular to the neutral
axis during bending. This assumption implies that the longitudinal strains vary
linearly across the beam and that, for elastic behaviour, the neutral axis of the beam
passes through the centroid of the cross-section. According to this theory the bending
moment and curvature are related as
67
M = EI
2w
x 2
(3.28)
where w( x, t ) is the transverse deflection of the neutral axis of the beam, E is the
Youngs modulus, and I is the moment of inertia of the cross-section. The result
Eq.(3.28) carries the assumption that the slope of the beam is small, see Eq.(3.4).
In accordance with Figure 3.5, the equation of motion of the beam element in the
vertical direction (if the slopes are small, the shear forces V ( x ) and V ( x + x ) can be
considered as if they were directed vertically) can be written as
Ax
2 w
= V ( x ) + V ( x + x ) + q1 x
t 2
(3.29)
Applying the Taylor expansion and dividing through by x , this equation can be
reduced to
2 w V
=
+ q1
x
t 2
(3.30)
M
x
(3.31)
where the higher order contributions of the external load q to the moment are
neglected.
Substitution of Eq.(3.31) into the equation of motion Eq.(3.30), yields
2 w 2 M
A 2 = 2 + q1
t
x
(3.32)
Employing the constitutive relation Eq.(3.28), this equation can be written in terms of
the transverse deflection w of the beam neutral line:
2 w 2
+
t 2 x 2
2 w
EI 2 = q1
x
(3.33)
This equation governs the transverse motion of the beam as it responds to the external
loading q ( x, t ) and initial conditions. The application of this equation is constraint by
the requirements that the slopes of the beam are small, the rotational inertia effects are
neglected and the Euler-Bernoulli assumptions are not violated significantly.
It is important to note that Eq.(3.33) is qualitatively different from all equations of
motion obtained in the preceding sections. In contrast to all previously obtained
equations of motion, Eq.(3.33) is of the order four with respect to the spatial coordinate, whereas the other equations are of the second order. The implications of this
difference will become clear later in this chapter.
68
x
Figure 3.6. Beam on Kelvin foundation as a model for a railway track
The vertical motion of the beam is described by Eq.(3.33), which, assuming that
the railway track is homogeneous along its length, reduces to
2 w
4 w
+
= q1
EI
t 2
x 4
(3.34)
There are two external vertical forces, which act on the railway track (the beam) in the
absence of a train. One force is the gravity force, whereas the other one is the soil
reaction. The density per unit length of the gravity force is Ag , where g is the
gravity acceleration. The soil reaction we model by the Kelvin foundation, which
implies that the density per unit length of this reaction is given by
q1,SOIL = kd w cd
w
t
(3.35)
where k d is the soil stiffness per unit length and cd is the soil damping coefficient per
unit length. Substituting this expression into Eq.(3.34) and accounting for the gravity
force, we obtain
2w
4w
w
+
+ kd w + cd
= Ag
EI
2
4
t
x
t
(3.36)
This equation allows to analyse vibrations of the railway track in the gravity field due
to imposed initial conditions.
69
b g
u x, t
x
kd
motion of the rod on viscous-elastic foundation. The equation of motion for this
model can be deduced from the equation of motion for the rod. Assuming for
simplicity that the rod is prismatic and homogeneous, we employ Eq.(3.14) and
replace in this equation the body force q by the reaction of the foundation, that is by
q1 =
1
u
kd u cd
A
t
(3.37)
2u
2u 1
u
E
+ kd u + cd
=0
2
2
A
t
x
t
(3.38)
70
To obtain the equation of motion for this beam, we could repeat the regular procedure
of equilibrating the forces and moments, like in Section 3.2.4, accounting for the
additional tensile force. There is, however, a more economical approach of just
combining the equation of motion for the beam, Eq.(3.33), and that for the string,
Eq.(3.7). Indeed, these two equations can be considered as two particular cases of the
equation of motion for the tensioned beam. In this perspective, Eq.(3.7) for the string
can be considered as that for the tensioned beam with the restoring force associated
with the bending stiffness neglected. On the other hand, Eq.(3.33) for the beam can be
considered as that for the tensioned beam with the restoring force associated with the
tension neglected. Thus, to account for these restoring forces simultaneously, we may
add the term responsible for the bending force from Eq.(3.33) into Eq.(3.7), or,
alternatively, we may add the term responsible for the tensile force from Eq.(3.7) into
Eq.(3.33). In both cases, the result reads
2 w 2
+
t 2 x 2
2 w w ( x, t )
+ q1
EI 2 = T
x
x x
(3.39)
This is the equation for the transverse motion of the tensioned beam.
If the beam were not tensioned but axially compressed, the form of Eq.(3.39)
would stay intact, but the tensile force T would have to be replaced by a compressive
force P .
3.3.
Every structure has a finite length and is fixed or attached to another structure at its
ends. Thus, to predict the mechanical behaviour of a structure, we have to describe
mathematically the conditions at the ends of the structure. This is done with the help
of the boundary conditions.
There are two types of the boundary conditions, namely: kinematic and dynamic.
Often, these are also referred to as essential and natural. A boundary condition is
called kinematic or essential if the kinematic characteristics of the structure are
prescribed at its ends such as displacements, angles or slopes. A boundary condition is
called dynamic or natural if forces or moments are prescribed at the ends. The number
of the boundary conditions, which have to be formulated at an end of the structure,
depends on the equation of motion, which is adopted for prediction of structural
mechanics. This number is always twice as small as the order of this equation with
respect to the spatial co-ordinate.
Sometimes, the mechanical characteristics of a structure, such as the crosssectional area or an elasticity modulus, change abruptly along the structure. The crosssection at which this happens is called an interface. At the interface, the equations of
motion are not valid and the interface conditions have to be formulated. These
conditions can also be subdivided into the kinematic and dynamic. However, at any
interface, both kinematic and dynamic conditions must be applied simultaneously,
which is not a necessity in the case of the boundary conditions. The number of the
interface conditions is always equal to the order of the equation of motion with
respect to the spatial co-ordinate.
In what follows, the boundary and interface conditions are formulated for the onedimensional models, which were introduced in the preceding section.
71
w
x
(3.40)
In correspondence with Fig.3.1(b) this expression has to be taken with the minus sign
if applied to the right-end of the string.
Besides the force imposed by the string, the dynamic boundary conditions depend
on the external element, to which the boundary is attached. This can be a spring, a
dashpot or a mass. The forces imposed by these elements are known to us from the
courses on basic mechanics and Chapter 2 of this development.
Table 3.1 shows all basic boundary conditions (BC), which the string can be
subject to. In this table, A ( t ) and B ( t ) are the vertical displacements of the ends, k ,
c, and m are the stiffness, dashpot coefficient and mass of the boundary element,
respectively.
Type of BC
Kinematic
(displacement)
Left-end
w = A (t )
Dynamic
(external force)
w
= P
x
Dynamic
(elastic)
w
= kw
x
Dynamic
(viscous)
w
w
=c
x
t
Dynamic
(inertial)
w
2w
=m 2
x
t
Right-end
w = B (t )
w
=P
x
w
= kw
x
w
w
= c
x
t
w
2w
= m 2
x
t
Two particular, practically important, boundary conditions can be deduced from this
table, namely: the fixed end and the free end. For the fixed and not displaced end, the
displacements A ( t ) or B ( t ) are zero and the boundary condition reads w = 0 . For
the free end (subject to no external reaction), k = c = m = 0, so that the boundary
condition reads w x = 0 . Let us note that if we assumed that the string were
supported by the viscous-elastic Kelvin foundation, the boundary conditions would
72
not have changed. The reason for this is that the Kelvin foundation is not able to
provide a reaction at a cross-section (at an end) but only at an element of onedimensional system of a finite extension.
The interface conditions at a cross-section of the string, where the tension changes
abruptly should ensure that the string does not break and the balance of vertical forces
is not violated. This means that at interface x = c the following conditions must hold
simultaneously:
w ( x, t )
= w ( x, t ) x = c 0
x =c + 0
w
w
=T
T
x x =c 0
x x =c + 0
(3.41)
Naturally, these conditions must also hold at any other cross-section of the string.
3.3.2. Longitudinal Motion of a Rod. The equation of motion for the longitudinal
motion of a rod, Eq.(3.15) is of the second order with respect to the spatial coordinate, as well as the string-equation. Therefore, the boundary conditions for the rod
are mathematically similar to those for the string.
The kinematic boundary condition for the rod prescribes the longitudinal
displacement, whereas the dynamic boundary condition represents the balance of
longitudinal forces. The expression for the force imposed by the rod on its left-end can
be deduced from Eqs.(3.11) and (3.12) to give
f L = A = EA u x
(3.42)
In correspondence with Fig.3.2(b) this expression has to be taken with the minus sign
if applied to the right-end of the rod.
The basic boundary conditions for the rod in longitudinal motion are summarized
in Table 3.2. As in the previous sub-section, A ( t ) and B ( t ) are the displacements of
the ends, now longitudinal.
Type of BC
Kinematic
(displacement)
Left-end
A(t )
u = A(t )
Dynamic
(external force)
EA
u
= P
x
Dynamic
(elastic)
EA
u
= ku
x
Dynamic
(viscous)
EA
u
u
=c
x
t
Dynamic
(inertial)
EA
u
2u
=m 2
x
t
u = B (t )
EA
u
=P
x
EA
u
= ku
x
EA
u
u
= c
x
t
EA
u
2u
= m 2
x
t
Right-end
B (t )
73
u
u
= EA
EA
x x =c 0
x x =c + 0
(3.43)
3.3.3. Torsion of a Rod. The equation of motion for the rotational motion of a rod,
Eq.(3.20), is of exactly the same form as that for the string and the rod in longitudinal
motion. Therefore, the boundary and interface conditions must have the same form as
those in the preceding two sub-sections.
The kinematic boundary condition for the rod in torsion prescribes the angle of
rotation of a cross-section, whereas the dynamic boundary condition represents the
balance of torques. The expression for the torque imposed by the rod on its left-end is
given by Eq.(3.19) and reads
M t = GJ t
(3.44)
In correspondence with Fig.3.3 this expression has to be taken with the minus sign if
applied to the right-end of the rod.
The basic boundary conditions for the rod in torsion are summarized in Table 3.3.
In this table, A ( t ) and B ( t ) are the angles of rotation of the end-cross-sections,
kt [ N m ] , ct [ N s m] , and I t kg m 2 are the stiffness, dashpot coefficient, and
moment of inertia of the boundary element opposing the torque.
Type of BC
Kinematic
(displacement)
Dynamic
(external torque)
Dynamic
(elastic)
Left-end
= A (t )
GJt
Right-end
= B (t )
= M text
x
GJt
= M text
x
= kt
x
GJt
= kt
x
GJt
Dynamic
(viscous)
GJ t
= ct
x
t
GJ t
= ct
x
t
Dynamic
(inertial)
GJ t
2
= It 2
x
t
GJ t
2
= It
x
t2
74
conditions must ensure continuity of the rotation angle and torque. Therefore, at
interface x = c the following conditions must hold:
( x, t )
= ( x, t ) x = c 0
x=c+ 0
= GJ t
GJ t
x x =c + 0
x x = c 0
(3.45)
3.3.4. Pure Shear of a Beam. The shear motion of a beam is governed by Eq.(3.27),
which is identical in form to the equations of motion for the string and rod in both
longitudinal and rotational motion.
The kinematic boundary condition for the beam in pure shear prescribes the
transverse displacement. The dynamic boundary condition represents the balance of
shear forces. The expression for the shear force imposed by the beam on its left-end
can be deduced from Eqs.(3.23) and (3.24) to give
f S = A = GA
w
x
(3.46)
In correspondence with Fig.3.4 this expression has to be taken with the minus sign if
applied to the right-end of the beam.
The basic boundary conditions for the beam in pure shear are summarized in Table
3.4, using the same notations as for the string (sub-section 3.3.1).
Type of BC
Kinematic
(displacement)
Left-end
w = A (t )
Dynamic
(external force)
GA
Dynamic
(elastic)
GA
w
= kw
x
Dynamic
(viscous)
GA
w
w
=c
x
t
Dynamic
(inertial)
GA
w
2w
=m 2
x
t
Right-end
w = B (t )
w
= P
x
GA
P
w
=P
x
GA
w
=kw
x
w
w
GA =c
x
t
w
2 w
GA =m 2
x
t
75
w ( x, t )
= w ( x, t ) x = c 0
x =c + 0
w
w
= GA
GA
x x = c + 0
x x =c 0
(3.47)
3.3.5. Bending of a Beam. The equation of motion for the beam in bending, Eq.(3.33)
, is of the order four with respect to the spatial co-ordinate. Therefore, at each end of
the beam, two boundary conditions have to be formulated, in contrast to all previously
considered cases.
The kinematic condition may prescribe the transverse deflection and the slope of
the beam. The dynamic boundary conditions represent the balance of forces and the
balance of moments acting on the end of the beam. The beam participates in these
boundary conditions with the shear force and moment, which it imposes on the endcross-sections. The expressions for the moment and shear force as they act on the leftend of the beam are given by Eqs.(3.28) and (3.31), and read
2w
x 2
3w
V = EI 3
x
M = EI
(3.48)
(3.49)
In correspondence with Fig.3.5(b) these expressions are to be taken with the minus
signs if applied to the right-end of the beam.
In all preceding cases, at an end-cross-section, it was allowed to pose one boundary
condition (because the equations were of the second order), implying that we could
pose either kinematic or dynamic boundary condition and never both of them. In the
case of the beam in bending, two boundary conditions should be posed at each endcross-section. These can be two kinematic, two dynamic or kinematic and dynamic
boundary conditions.
In Table 3.5 all basic boundary conditions are presented, which the beam can be
subject to. In this table,
A ( t ) and B ( t ) are the vertical displacements of the ends
A ( t ) and B ( t ) are the slopes at the ends
k , c, and m are the stiffness, dashpot coefficient and mass of the boundary
element, acting in the vertical direction.
kr [ N m] , cr [ N s m ] , and I r kg m 2 are the stiffness, dashpot coefficient, and
76
Type of BC
Kinematic
(displacement)+
Kinematic
(slope)
Kinematic
(displacement)+
Dynamic Moment
(elastic)
Kinematic
(displacement)+
Dynamic Moment
(viscous)
Kinematic
(displacement)+
Dynamic Moment
(inertial)
Kinematic (slope)
+
Dynamic Force
(elastic)
Left-end
w = A(t )
A
w
= A ( t )
x
w
= B (t )
x
w = A(t )
2 w kr w
=
x2 EI x
kr
w = A( t )
2 w cr 2 w
=
x2 EI t x
cr
w = A( t )
2w Ir 3w
=
x2 EI t 2 x
Ir
w
= A ( t )
x
k
k
3w
w
=
EI
x3
Kinematic (slope)
+
Dynamic Force
(inertial)
w
= A ( t )
x
m
m 2w
3w
=
EI t 2
x3
EI
2w
=M
x2
EI
w
=P
x3
EI
w
=
x2
w = B (t )
k w
2w
= r
2
EI x
x
w = B(t )
cr 2w
2w
=
EI t x
x2
kr
B
cr
w = B( t )
w
= A ( t )
x
c w c
3w
=
EI t
x3
Dynamic Moment
(elastic + viscous
+inertial)
+
Dynamic Force
(elastic + viscous
+inertial)
w = B (t )
Kinematic (slope)
+
Dynamic Force
(viscous)
Dynamic Moment
(external)
+
Dynamic Force
(external)
Right-end
I r 3w
2w
=
EI t 2 x
x2
w
= B (t )
x
3w k
w
=
x3 EI
w
= B (t )
x
3 w c w
=
x3 EI t
w
= B (t )
x
3w m 2 w
=
x3 EI t 2
EI
2w
= M
x2
EI
w
= P
x3
EI
w
=
x2
Ir
w
2w
k
w
c
I
+
+
r
r
r
x
t
t 2
3w
w
2w
EI 3 = kw + c + m 2
x
t
t
w
2w
+
+
k
w
c
I
r
r
r
x
t
t 2
3w
w
2w
EI 3 = kw + c
+m 2
x
t
t
77
The last row in this table presents the two dynamic boundary conditions in the most
general form, not separating the elastic, viscous and inertial cases.
There are a few types of the beam fixation, which are widely used in practice,
namely:
pinned-pinned (at both ends the displacement and moment are zero);
clamped-clamped (at both ends the displacement and slope are zero);
clamped-free or cantilever (at one end the displacement and slope are zero,
whereas at the other end the moment and force are zero);
clamped-pinned (at one end the displacement and slope are zero, whereas at the
other end the displacement and moment are zero);
pinned-free (at one end the displacement and moment are zero, whereas at the other
end the moment and force are zero).
Though all these types of fixation are particular cases of the general ones presented in
Table 3.5, it is customary to collect these in a separate table. This is done in Table 3.6
presented below.
Type of fixation
Pinned-Pinned
Schematization
x=a
x=b
Boundary Conditions
x = a : w = 2 w x 2 = 0
x = b : w = 2 w x 2 = 0
Clamped-Clamped
Clamped-Free
x=a
x=b
x=a
x=b
x = a : w = w x = 0
x = b : w = w x = 0
x = a : w = w x = 0
2
x = b : 2 w x = 3 w x = 0
Clamped-Pinned
x=a
x=b
x = a : w = w x = 0
x = b : w = w2 x 2 = 0
Pinned-Free
x=a
x=b
x = a : w = 2 w x = 0
2
x = b : 2 w x = 3 w x = 0
Table 3.6 Typical beam fixations
There must be four interface conditions (in accordance with the spatial order of the
equation of motion) imposed at each cross-section of the beam, where the bending
stiffness changes abruptly. These conditions should ensure that the displacement,
slope, moment and shear force at the interface are continuous. Thus, at interface x = c
the following conditions must be satisfied simultaneously:
78
w ( x, t )
= w ( x, t ) x = c 0
x =c + 0
w
w
=
x
x x =c 0
x =c + 0
2w
2w
=
EI
EI
x 2
x 2 x =c 0
x=c+0
3w
3w
EI 3
= EI 3
x x = c 0
x x =c + 0
(3.50)
3.4.
In the preceding sections we formulated the equations of motion for a series of onedimensional structural models and discussed the boundary conditions, which
correspond to these models. This gives us the mathematical basis for studying the
mechanical behaviour of one-dimensional structures.
In this section, we will introduce the method of separation of variables, which
generally allows to study the dynamic behaviour of a large class of one-dimensional
models, as long as they are governed by linear differential equations.
Two models will be considered in this section, namely: the model for longitudinal
motion of a rod and that for the bending motion of a beam. It will be assumed that
both the rod and the beam are cylindrical and homogeneous. The results, which will
be obtained for the rod in longitudinal motion, can be easily generalized to describe
the motion of the string, rod in torsion and beam in pure shear. The possibility of such
generalization is based on the equations of motion for these four models, which have
exactly the same format. Indeed, each of these equations can be written in the
following generalized form (note that we assume all parameters to be independent of
the co-ordinate)
2
2w
2 w
=c
+Q
t 2
x 2
(3.51)
where c has the physical meaning of the wave speed, whereas Q represents a
generalized external force. To obtained from Eq.(3.51) the particular models, which
were introduced in Section 2 of this Chapter, the following substitutions are to be
made:
c 2 = T ( A) , Q = q1 ( A ) for the string, Eq.(3.7);
c 2 = E , Q = q for the rod in longitudinal motion, Eq.(3.14);
c 2 = GJ t ( J ) , Q = m1 ( J ) for the rod in torsion, Eq.(3.20);
c 2 = G , Q = q for the beam in pure shear, Eq.(3.26);
79
The bending motion of a beam will be considered separately since the equation of
motion in this case is different from the other ones.
3.4.1. Longitudinal Vibrations of a Rod. The complete mathematical statement of a
problem describing the dynamic behaviour of a continuous system includes three
components, namely: the equation(s) of motion, the boundary conditions and the
initial conditions. Depending on the loading type and the type of the boundary
conditions, different mathematical methods may be applied to achieve the most
efficient solution procedure. In this section, the method of separation of variables will
be employed, which can handle almost any loading type but, in its classical form, is
only applicable to time-independent boundary conditions.
In this sub-section we consider vibrations of a rod, which is cylindrical,
homogeneous and has the length L . The equation of motion governing the
longitudinal motion of such a rod and the most general form of the initial conditions
read
2u
2u
=
+ q ( x, t )
E
t 2
x 2
(3.52)
u ( x, 0 ) = u 0 ( x )
u
= v0 ( x )
t t =0
(3.53)
To demonstrate the method of separation of variables we will assume that the rod is
fixed at both ends as shown in Figure 3.9. Note, that as mentioned in the beginning of
this section, the method can be equally easily applied to any other time-independent
boundary conditions, which account for a constant displacement or a constant force
applied at the ends of the rod.
u ( x, t )
x=0
q ( x, t )
x=L
Figure 3.9. Fixed-fixed rod
80
The boundary conditions for the rod fixed at both ends read
u ( 0, t ) = u ( L, t ) = 0
(3.54)
The method of separation of variables is based on the assumption that the time- and
space- dependent generalized displacement of a continuous system can be represented
as a multiplication of two functions, one dependent only on the time, and the other
dependent only on the spatial coordinate. In accordance with this method, we assume
that the longitudinal displacement of the rod can be represented in the following form
u ( x, t ) = U ( x ) ( t )
(3.55)
where U ( x ) and ( t ) are unknown functions of the coordinate along the rod and
time, respectively.
Let us first assume that the external loading q is absent and the equation of
motion, Eq.(3.52), reads
2
2u
2 u
=c
t 2
x 2
(3.56)
d 2
2U
2
c
=
dt 2
x 2
(3.57)
(3.58)
dt 2
2
c 2 1 U = 2
U x 2
(3.59)
81
d 2
+ 2 = 0
2
dt
(3.60)
The general solution of this equation is known to us, see Eq.(2.12). It reads
= A sin (t ) + B cos (t )
(3.61)
This form of the solution represents simple harmonic motion at the frequency ,
provided that the separation constant in Eq.(3.59) is positive, i.e., 2 > 0 . If we had
chosen a separation constant less than zero or complex in Eq.(3.60), we would than
obtain a non-harmonic solution, which is definitely incapable of representing the
periodic motion, which, as the common sense suggests, must be possible in the rod
without damping. Thus, the separation constant 2 in Eq.(3.59) must be positive.
Consequently, can be given the physical sense of a frequency of vibrations.
The second equation in Eq.(3.59) can be written as
d 2U 2
+ U =0
dx 2 c 2
(3.62)
The general solution of this equation under the condition that > 0 reads
U = C sin ( x ) + D cos ( x )
(3.63)
where
(3.64)
Now we will use the boundary conditions. Eq.(3.54). Substituting Eq.(3.55) into
these boundary conditions we obtain the following equations
U ( 0 ) ( t ) = 0
U ( L ) ( t ) = 0
which have to be satisfied at every time instant. This can be achieved if and only if the
following equations hold
U ( 0) = U ( L ) = 0
(3.65)
Substituting the general expression for U into these boundary conditions, and
imposing the condition of non-triviality that the coefficients C and D may not vanish
simultaneously, we obtain
D=0
(3.66)
sin ( L ) = 0
(3.67)
82
Equation (3.67) is called the frequency equation, since it determines the natural
frequencies of the system. For the case at hand, we have
n L = n
n = c n = n c L,
n = 1, 2,3,...
(3.68)
The frequencies n in Eq.(3.68) represent the discrete frequencies at which the rod is
capable of undergoing harmonic motion. For a given value of n , we thus have the
vibration pattern of the rod described by
U n ( x ) = Cn sin ( n x )
(3.69)
n=1
n=2
u
x
n=3
x
n=4
Figure 3.10. Symmetric (n = 1,3) and antisymmetric (n = 2,4) modes of a fixed-fixed rod
From this figure we can note that the modes with n = 1,3,... give symmetric motion
while antisymmetric modes result from n = 2,4,. The points of zero displacement
are called the nodes of vibration, while the points of maximum vibration are called the
antinodes.
Combining the time dependence, Eq.(3.61), and the spatial dependence, Eq.(3.63)
for a given value of n, we obtain
un ( x, t ) = ( An sin (n t ) + Bn cos (n t ) ) sin ( n x )
(3.70)
where the constant Cn of Eq.(3.69) has been absorbed in the constants An and Bn .
The general solution of Eq.(3.56) is obtained by the superposition of all particular
solutions of the form (3.70). This solution reads
83
(3.71)
n =1
n =1
sin ( n x ) = u0 ( x )
(3.72)
A
n
n =1
sin ( n x ) = v0 ( x )
To find the constants An and Bn from Eqs.(3.72), we multiply each of these equations
by sin ( m x ) and integrate over the interval 0 x L . Noting that
sin ( x ) sin ( x ) dx = 0,
L
L 2,
m=n
mn
(3.73)
we obtain
L
2
An =
v0 ( x ) sin ( n x ) dx
Ln 0
L
(3.74)
2
Bn = u0 ( x ) sin ( n x ) dx
L0
u ( x, t ) = n ( t ) sin ( n x )
(3.75)
n =1
dependent character of q ( x, t ) .
Substitution of Eq.(3.75) into Eq.(3.52) gives
84
d 2 n
1
2
2 + n n sin ( n x ) = q ( x, t )
n =1 dt
(3.76)
sin
sin
( n ) ( m )
2
n n
0
n =1 dt
0
(3.77)
This equation has exactly the same form as Eq.(2.36), therefore it must have the
solution of the form given by Eq.(2.39), i.e.,
n ( t ) = n ( 0 ) cos (nt ) +
t
1 d n ( 0 )
1
sin (nt ) + Qn ( t ) sin (n ( t t ) ) dt (3.78)
n dt
n 0
( 0 ) sin ( x ) = u ( x )
n
n =1
d n ( 0 )
n =1
dt
(3.79)
sin ( n x ) = v0 ( x )
This system of equations can be resolved in exactly the same manner as Eqs.(3.72) to
give
d n ( 0 )
dt
2
= v0 ( x ) sin ( n x ) dx
L0
L
(3.80)
2
n ( 0 ) = u0 ( x ) sin ( n x ) dx
L0
85
2u
=0
x 2
(3.81)
(3.82)
From the first glance, it seems that there can not be any relationship between the static
solution, Eq.(3.82), and the dynamic solution, Eq.(3.70). Note, however, that the
dynamic solution, Eq.(3.70), was obtained under the requirement that the frequency
is not greater than zero. But in statics we have = 0 . Therefore, the general
solution of Eq.(3.62), has to be rewritten for the case = 0 to correspond to statics.
This gives the following expression for U ( x ) (instead of Eq.(3.63))
U = C1 + C 2 x
(3.83)
which, taking into account that there is only one natural frequency, which is equal to
zero, represents the complete solution to the problem.
Thus, the dynamic solution can be reduced to the static one by taking the limit of
all natural frequencies going to zero. This confirms mathematically the obvious fact
that statics is a particular case of dynamics.
3.4.2. Bending Vibrations of a Beam. The method of separation of variables can be
applied to the problem of bending motion of a beam in the same manner as described
in the previous sub-section.
Consider the bending motion of a beam, which is cylindrical, homogeneous and
has the length L . In this sub-section we will consider only the free vibrations of the
beam, i.e., the external force q1 will be neglected. The equation of motion governing
the beam motion in this case and the general form of the initial conditions read
4
2w
2 w
a
+
= 0,
t 2
x 4
a2 =
EI
A
(3.84)
u ( x, 0 ) = u 0 ( x )
u
= v0 ( x )
t t =0
(3.85)
Let us assume that the beam is pinned at both ends so that the boundary conditions at
x = 0 (the left-end) and x = L (the right-end) are given as
2w
w ( 0, t ) = 2 ( 0, t ) = 0
x
(3.86)
2w
( L, t ) = 0
x 2
(3.87)
w ( L, t ) =
86
(3.88)
where W ( x ) and ( t ) are unknown functions of the coordinate along the rod and
time, respectively.
Substituting Eq.(3.88) into the equation of motion, Eq.(3.84) and dividing the
obtained result by W , we obtain
4
1 d 2
2 1 W
a
+
=0
W x 4
dt 2
(3.89)
Since the first term in Eq.(3.89) depends only on time, whereas the second term
depends only on the coordinate, this equation can be satisfied if and only if these
terms are equal to a separation constant. This constant has to be introduced in the
following manner, to ensure that the beam is capable of performing harmonic
vibrations:
1 d 2
= 2
2
dt
4
a 2 1 W = 2
W x 4
(3.90)
(3.91)
(3.92)
The general solution of this equation, in accordance with the general representation,
Eq.(2.7), can be written as
4
W ( x ) = C n exp ( sn x )
(3.93)
n =1
2
a2
=0
(3.94)
87
(3.95)
where
=
4
(3.96)
a2
where
(3.97)
The boundary conditions, Eqs.(3.86) and (3.87) can be rewritten in the terms of
W ( x ) as (since these should be valid at any time instant)
W (0) =
2W
(0) = 0
x 2
(3.98)
W ( L) =
2W
( L) = 0
x 2
(3.99)
Substituting the general solution, Eq.(3.97), into these boundary conditions, and
imposing the non-triviality condition that the coefficients C1 C4 may not vanish
simultaneously, we obtain
C2 = C3 = C4 = 0
sin ( L ) = 0
(3.100)
n L = n ,
n = 1, 2,3,...
(3.101)
The frequency equation, Eq.(3.101), determines the natural frequencies of the beam
vibrations, which, taking into account Eq.(3.96), can be found as
n
n = a
L
(3.102)
88
Correspondingly, the normal mode of the beam vibrations, for a given value of n ,
reads
Wn ( x ) = Cn sin ( n x )
(3.103)
which is precisely the same in the form as Eq.(3.69) derived for the rod. Therefore,
the shape of the normal modes repeats that shown in Figure 3.10.
Combining the time dependence, Eq.(3.91), and the spatial dependence, Eq.(3.103)
we obtain the following general expression describing the beam vibrations caused by
the initial conditions:
(3.104)
n =1
The constants An and Bn can be determined from the initial conditions to give the
expressions presented by Eq.(3.74), in which the natural frequencies are defined by
Eq.(3.102).
Note, that if the boundary conditions at the beam ends were chosen differently, the
frequency equation would be more complicated (transcendental) as well as the normal
modes and, consequently, the overall solution.
3.5.
In a great deal of practical situations, the major contribution to the stress-state of civil
engineering structures is associated with static loads. These loads may be associated
with the direct gravity loading, the dead weight of various loads, the mean component
of the drag force in a flow, the temperature extension of materials, etc. Therefore, it is
of immense importance for civil engineering to study the statics response of structures
to time-invariant loads.
The static response of an elastic structure is governed by the balance of forces and
moments acting on this structure. This balance can be straightforwardly deduced from
the corresponding equation of motion by setting the time derivatives to zero and
neglecting the time dependence of the loading term. For example, consider the
equation of longitudinal motion of the rod, Eq.(3.15):
2u
u
= EA + q ( x, t ) A
2
t
x
x
(3.105)
Assume that the external force q ( x ) acting on the rod is time-independent. In this
case, after a sufficiently long time after the application of the load, the structure will
assume the static equilibrium, in which its displacement u is time independent. In this
case the time derivatives of u vanish and Eq.(3.105) takes the form
u
EA = q ( x ) A
x
x
89
which represents the balance of axial forces acting on a differential element of the rod.
The same idea holds for the boundary conditions. To study the static response of
structures, we may use the boundary conditions formulated in Section 3.3 setting to
zero those terms in these conditions, which include the time derivatives.
In this section we will consider the static response of the rod to longitudinal forces
and that of the bending beam to transverse forces and moments. Various forces and
boundary conditions will be treated. Additionally, both the rod and beam will be
considered subject to the elastic foundation and the effect will be studied of this
foundation on the static response and the method of analysis. The static response of
the other three above-introduced models (string, rod in torsion and beam in shear) will
not be studied, because of the mathematical similarity of these models to that of the
rod.
3.5.1. Static Response of a Rod to Axial Loading. Let us shortly repeat the
equations, which govern the axial deflection of the rod. These are:
the equilibrium of axial forces, as follows from Eq.(3.8):
d
( A ) = qA
dx
(3.106)
Note that this form of the equilibrium is valid not only for the cylindrical rods but
also for the rods with variable cross-sections.
the constitutive equation, Eq.(3.11):
= E
(3.107)
= u x
(3.108)
(3.109)
du
dx
(3.110)
Making use of the normal force, the equilibrium equation can be written as
dN
= qA
dx
(3.111)
90
(3.112)
N ( x ) = q ( x ) A ( x ) dx + N ( 0 )
(3.113)
u ( x) =
0
N ( x )
dx + u ( 0 )
E ( x ) A ( x )
(3.114)
If the cross-sectional area A , the Youngs modulus E , and the body force q are
invariant along the rod, the general solutions, Eqs.(3.113) and (3.114) reduce to
N ( x ) = q1 x + C1
u ( x) =
(3.115)
1 1
2
q1 x + C1 x + C2
EA 2
(3.116)
Left-end
u=A
Dynamic
(external force)
N = P
Dynamic
(elastic)
N = ku
Right-end
u=B
N=P
N = ku
Note that the dynamic boundary condition in Table 3.7 is given in terms of the normal
force, which is more suitable for static analyses.
The interface conditions at a cross-section x = c of the rod, where an abrupt
change of the Youngs modulus or the cross-sectional area takes place, are given by
Eq.(3.43). These conditions can also be written making use of the normal force to give
u ( x )
= u ( x ) x =c 0
x =c + 0
N ( x ) x =c + 0 = N ( x ) x = c 0
(3.117)
91
Equations (3.110) and (3.111) supplemented by Table 3.7 and the interface
conditions, Eq.(3.117), allow us to analyse the axial deflections and normal forces in
the rod induced by any axial force. In what follows we consider a number of possible
loads and boundary conditions.
Example A. Consider a fixed-free cylindrical and homogeneous rod of the length L as
shown in Figure 3.11(a). Assume that this rod is subject to a distributed axial force
with the constant density per unit length q1 .
The boundary conditions in the case at hand are given as
u ( 0) = 0
(3.118)
N ( L) = 0
(3.119)
Since the rod is cylindrical and homogeneous, and the external force is constant, the
simplified expressions can be applied for the normal force and the axial displacement,
Eqs.(3.115) and (3.116).
Substituting the simplified expression for the normal force, Eq. (3.115), into the
boundary condition at the right end, Eq.(3.119), gives C1 = q1 L , which implies that
the normal force in the rod is given as:
N = q1 ( L x )
(3.120)
Thus, the axial force varies linearly along the rod as shown in Figure 3.11(b). Because
the rod is cylindrical, the variation of the axial stress along the rod is also linear, as
well as that of the axial strain, in correspondence with the constitutive equation,
Eq.(3.107).
(a)
q1
u=0
N =0
L
(b)
N
(c)
u
Figure 3.11. (a) Fixed-free rod under uniform axial loading; (b) normal force; (c) axial deflection
To find the axial displacement of the rod, we substitute the simplified expression for
this displacement, Eq.(3.116), into the boundary condition at the right end of the rod.
This yields
92
0=
1
( 0 + 0 + C2 )
EA
C2 = 0
(3.121)
Thus, the variation of the axial deflection along the rod is parabolic and given as
u=
1
q1 x ( 2 L x )
2 EA
(3.122)
(3.123)
u ( L) = 0
(3.124)
To solve this problem we apply the simplified expression for axial displacement,
Eq.(3.116). Substituting this expression into the boundary conditions, we obtain:
(a)
q1
u=0
u=0
L
(b)
(c)
Figure 3.12. (a) Fixed-fixed rod under uniform axial loading; (b) normal force; (c) axial deflection
93
0=
1
( 0 + 0 + C2 )
EA
C2 = 0
0=
1 1 2
q1 L + C1 L + C2
EA 2
C1 =
(3.125)
1
q1 L
2
Thus, the normal force and axial displacement, in accordance with Eqs.(3.115) and
(3.116) become
1
N = q1 L x ,
2
u=
1
q1 x ( L x )
2 EA
(3.126)
q1
u=0
u=0
L
(b)
q1
(c)
N
(d)
u
Figure3.13. (a) Fixed-fixed rod under cosinusoidal axial loading; (b) external axial force; (c)
normal force in the rod; (d) axial deflection
94
N ( x) =
x
sin
+ N ( 0)
q0 L
(3.127)
where N ( 0 ) is unknown. Substituting the obtained expression for the normal force,
Eq.(3.127), into the general expression for the axial displacement, Eq.(3.114), we
obtain
u ( x) =
q0 L2
x N (0)
cos
1 +
x + u (0)
2
EA L EA
(3.128)
u (0) = 0
0 = 2
q0 L2 N ( 0 )
+
L + u ( 0)
2 EA EA
N (0) = 2
q0 L
Thus, the normal force and axial displacement, in accordance with Eqs.(3.127) and
(3.128) become
N=
q0 L
x
2 sin
2
L
(3.129)
qL
x
u = 20 L cos
L + 2x
EA
L
x, u
q1 ( x )
Figure 3.14. A vertical column with a variable cross-section loaded by its deadweight
95
The body force, which the gravity imposes in the column is q = g , where is the
mass density of the column material and g is the gravity acceleration. Consequently,
the equilibrium of vertical forces in the column (expressed through the normal stress),
Eq.(3.106), takes the form
0
dA ( x )
dx
= gA ( x )
(3.130)
This is an ordinary differential equation of the first order, the solution to which can be
sought in the form
A ( x ) = A0 exp ( sx )
Substituting this into Eq.(3.130), we obtain
0s = g
s = g 0
(3.131)
Thus the cross-section should have the following exponentially depending of the
height form:
g
A ( x ) = A0 exp
0
(3.132)
The unknown constant A0 should be found from the equilibrium condition at the top
of the column, which reads
N ( 0) = P
(3.133)
By definition, the normal force equals the multiplication of the normal stress and the
cross-sectional area, see Eq.(3.109). Thus, Eq.(3.133) can be written as
0 A ( 0 ) = P ,
which, after substitution of Eq.(3.132), yields the following expression for A0 and the
final expression for the cross-sectional area:
A0 =
A( x) =
g
exp
0
0
P
(3.134)
Example E. Consider a cylindrical and homogeneous rod of the length L , the left end
of which is fixed, whereas the right end is attached to a spring with the stiffness k as
shown in Figure 3.15(a). Assume that this rod is subject to a distributed axial force
with the constant density per unit length q1 .
96
(a)
q1
u=0
3L 5
(b)
N = ku
2
q1 L
5
N
1
q1 L2
10
(c)
u
Figure 3.15. (a) Fixed-elastically supported rod under uniform axial loading; (b) normal force; (c)
axial deflection. The results are presented for K = 4
The boundary conditions in this case, in correspondence with Table 3.7, read
u ( 0) = 0
(3.135)
N ( L ) = ku ( L )
(3.136)
Since the rod is cylindrical and homogeneous, and the external force is constant, the
simplified expressions can be applied for the normal force and the axial displacement,
Eqs.(3.115) and (3.116).
Substituting the simplified expression for the displacement, Eq.(3.116), into the
boundary condition at the left end of the rod, Eq.(3.135), we obtain
0=
1
( 0 + 0 + C2 )
EA
C2 = 0
With this constant defined, substitution of the simplified expression for the normal
force, Eq. (3.115), into the boundary condition at the right end, Eq.(3.136), gives
q1 L + C1 =
k 1 2
q1 L + C1 L
EA 2
(3.137)
1+ K 2
1+ K
(3.138)
97
where K = kL ( EA ) represents the ratio of the spring stiffness k to the overall axial
stiffness of the rod EA L .
Thus, the normal force in and the axial displacement of the rod are given as
1+ K 2
N = q1 L
x
1+ K
q 1+ K 2 1 2
u = 1 xL
x
1+ K
2
EA
(3.139)
u ( L) = 0
(3.140)
(a)
q1
u=0
L 2
7 L 12
(b)
L 2
u=0
5
q1 L
12
N
(c)
Figure 3.16. (a) Fixed-fixed combined rod under uniform axial loading; (b) normal force; (c) axial
deflection
98
N ( L 2 + 0 ) = N ( L 2 0 )
(3.141)
Since both parts of the rod are cylindrical and homogeneous, the simplified
expressions for the normal force and axial displacement, Eqs.(3.115) and (3.116),
hold for each part. Thus, introducing the subscripts L and R for the left- and righthand-part related quantities, respectively, we may write
N L ( x ) = q1 x + C1
(3.142)
2
q1 x + C1 x + C2
2
N R ( x ) = q1 x + C3
uL ( x ) =
uR ( x ) =
1
EAL
1
EAR
(3.143)
(3.144)
2
q1 x + C3 x + C4
2
(3.145)
from
u ( 0 ) = uL ( 0 ) = 0 :
from
u ( R ) = uR ( L ) = 0 :
from
from
C2 = 0
1
q1 L2 + C3 L + C4 = 0
2
1 1 2 1
1 1 2 1
u L ( L 2 ) = uR ( L 2 ) :
q1 L + C1 L + C2 =
q1 L + C3 L + C4
AL 8
2
2
AR 8
1
1
N L ( L 2) = N R ( L 2 ) : q1 L + C1 = q1 L + C3
2
2
1
1 + 3 A
q1 L
,
4
1 + A
C2 = 0,
C3 = C1 ,
C4 =
1 2 1 A
q1 L
4
1 + A
(3.146)
where A = AL AR .
Thus, the problem has been solved. For graphical representation of the results, let
us assume that the left-hand part of the rod has twice greater cross-section than the
right-hand part, i.e., A = 2 . In this case the expressions for the constants, Eq.(3.146),
reduce to
C1 =
7
q1 L,
12
C2 = 0,
C3 = C1 ,
C4 =
1
q1 L2
12
99
and the expressions for the normal forces and axial displacements, Eqs.(3.142)(3.145) become
7
N L ( x ) = q1 L x ,
12
q x 7
uL ( x ) = 1 L x ,
2 EAL 6
N R ( x ) = q1 L x
12
q 7
1
u R ( x ) = 1 Lx L2 x 2
2 EAR 6
6
(3.147)
The distribution of the normal force and that of the axial displacement along the rod
are plotted in Figure 3.16(b,c).
3.5.2. Static Response of a Rod on Elastic Foundation to Axial Loading. In this
sub-section we consider the rod as it is subject to a distributed reaction of an elastic
foundation as shown in Figure 3.17.
kd
x, u
Figure 3.17. Rod on elastic foundation
The equilibrium equation, governing the balance of axial forces in this rod reads
d
du
EA = qA + kd u
dx
dx
(3.148)
d 2u
kd u = qA
dx 2
(3.149)
If, additionally, the body force q and the stiffness per unit length kd of the elastic
foundation are also coordinate-invariant, then the general solution of this equation can
be easily found to give
kd
kd qA
u = C1 cosh
x + C2 sinh
x +
EA
EA k
d
(3.150)
100
kd
kd
du
= EAkd C1 sinh
x + C2 cosh
x
dx
EA
EA
(3.151)
k qA
u = C1 exp
x + C2 exp d x +
EA
EA kd
kd
k
N = EAkd C1 exp
x C2 exp d x
EA
EA
(3.152)
(3.153)
Equations (3.150) and (3.151) show that both the axial displacement of and the
normal force in the rod on elastic foundation vary along the rod exponentially, even in
the case of a constant external force. To grasp other peculiarities of the static response
of the rod on elastic foundation to an axial force, we consider two examples, which
are discussed below.
Example A. Consider a fixed-free cylindrical and homogeneous rod of the length L
on elastic foundation as shown in Figure 3.18(a). Assume that this rod is subject to a
distributed axial force with the constant density per unit length q1 .
x
(a)
u=0
q1
N =0
(b)
N
(c)
Figure 3.18. (a) Fixed-free rod on elastic foundation under uniform axial loading; (b) normal force; (c)
axial deflection
101
(3.154)
N ( L) = 0
(3.155)
Since the rod is cylindrical and homogeneous, the elastic foundation is uniform and
the external force is constant, the simplified expressions can be applied for the normal
force and axial displacement, Eqs.(3.150) and (3.151).
Substituting these expressions into the boundary conditions, we obtain the
following system of algebraic equations:
0 = C1 +
k
kd
0 = C1 sinh d L + C2 cosh
L
EA
EA
q1
,
kd
q1
,
kd
C2 =
kd
q1
tanh
L
EA
kd
Consequently, the normal force and the axial displacement are given as
kd
k
kd
x + tanh d L sinh
x
1 cosh
EA
EA
EA
k
kd
kd
EA
N = q1
L cosh
x
sinh d x tanh
EA
EA
EA
kd
u=
q1
kd
(3.156)
(3.157)
102
N = Mg
L 2
gA
L 2
kd
N =0
Figure 3.16. Vertical column (water tower) whose lower part is in the ground
(3.158)
NL ( L) = 0
(3.159)
where the subscripts U and L mark the upper- and lower-part related quantities.
Since at x = L 2 the foundation starts acting on the rod abruptly, we have to
formulate the interface conditions at this point. In accordance with Eq.(3.117), these
conditions read
uU ( L 2 ) = uL ( L 2 )
(3.160)
NU ( L 2 ) = N L ( L 2 )
(3.161)
The normal force in and the displacement of the upper-part of the rod, since it has
no foundation, are given by Eqs.(3.115) and (3.116), whereas those of the supported
lower-part are given by Eqs.(3.152) and (3.153). Thus, these quantities read
NU = q1 x + C1
uU =
(3.162)
1 1
2
q1 x + C1 x + C2
EA 2
kd
N L = EAkd C3 exp
x C4 exp
EA
kd
k
u L = C3 exp
x + C4 exp d x +
EA
EA
(3.163)
kd
x
EA
(3.164)
q1
kd
(3.165)
103
C1 = P
and
NU = q1 x P
(3.166)
Now we will make use of the dynamic interface condition, Eq.(3.161), and the
boundary condition at the lower end, Eq.(3.159). Substitution of Eqs.(3.162) and
(3.164) into these two conditions yields the following system of two linear algebraic
equations with respect to C3 and C4 :
k L
k L
L
P = EAkd C3 exp d
C4 exp d
EA 2
2
EA 2
kd
k
C3 exp
L C4 exp d L = 0
EA
EA
q1
L Q + 2P
3
exp
K d 1 exp K d
EA 2 K d
2
L Q + 2P
C4 =
exp
EA 2 K d
Kd
) (1 exp (
Kd
))
))
(3.167)
kd L q1
1 1 2
L
+ C4 exp
+
q1 L + C1 + C2 = C3 exp
2
EA 8
EA 2 kd
EA 2
This equation can be solved by employing Eq.(3.167) to give the following expression
for C2
1 + exp
1
C2 =
4 K d P + 8Q + K d Q 4 K d ( Q + 2 P )
8K d
1 exp
(
(
)
)
Kd
Kd
(3.168)
104
column is equal to the load at the top and that the elastic foundation is four times
softer than the rod.
0.0
17.0
16.8
N/Q
u=(EA/q1/L2)
-0.4
16.6
16.4
-0.8
16.2
-1.2
16.0
-1.6
15.8
0.00
0.25
0.50
0.75
1.00
0.00
0.25
u/L
0.50
0.75
1.00
u/L
Figure 3.17 shows that for the chosen set of parameters the displacement varies along
the rod only slightly (less than 6%) relative to the average displacement of the rod
( 16.3 q1L ( EA) ) . The latter displacement is almost equal to the displacement, which
(3.169)
M
x
(3.170)
The relation between the bending moment and the transverse displacement,
Eq.(3.28):
M = EI
2w
x 2
(3.171)
105
w
=
x
(3.172)
V ( x ) = q1 ( x ) dx + V ( 0 )
(3.173)
M ( x ) = V ( x ) dx + M ( 0 )
(3.174)
0
x
()
() ()
M x
( 0 ) x + w ( 0 )
w ( x ) =
dxdx
E
x
I
x
0 0
x x
(3.175)
If the loading q1 is invariant along the beam, the general expressions for the shear
force V and the bending moment M , Eqs.(3.173) and (3.174), reduce to
V = q1 x + C1
1
M = q1 x 2 + C1 x + C2
2
(3.176)
(3.177)
If the bending stiffness is also (as well as q1 ) invariant along the beam, the general
expression for the transverse displacement, Eq.(3.175), becomes
w=
1 1
1
1
4
3
2
q1 x + C1 x + C2 x + C3 x + C4
6
2
EI 24
(3.178)
The boundary conditions, which are applicable to the bending of a beam, can be
deduced from Table 3.5 to form the following table:
Type of BC
Kinematic
(displacement)+
Kinematic
(slope)
Kinematic
(displacement)+
Dynamic Moment
(elastic)
Kinematic (slope)
+
Dynamic Force
(elastic)
Dynamic Moment
(external)
+
Dynamic Force
(external)
Left-end
w= A
w
= A
x
Right-end
w=B
w
= B
x
w= A
M = kr
w=B
A
w
x
M = kr
kr
w
= A
x
V = kw
A
k
M = M 0
V = P
w
x
w
= B
x
V = kw
M = M0
V =P
M0
kr
B
P
106
Type of BC
Dynamic Moment
(elastic)
+
Dynamic Force
(elastic)
Left-end
w
M = kr
x
V = kw
Right-end
M = kr
w
x
V = kw
Table 3.8 covers all possible types of the boundary conditions, which may be imposed
on the beam in statics. If no elastic element is attached to an end of the beam, then
only four boundary conditions are applicable at this end. These conditions are
collected in the table below.
Type of BC
Pinned
Schematization
Boundary Conditions
w=M =0
Clamped
w = = 0
Free
M =V = 0
Sliding
=V = 0
Table 3.9. The simplest boundary conditions for a beam
w
w
,
=
x x =c + 0 x x =c 0
2 w
2 w
,
EI 2
= EI 2
x x = c + 0
x x =c 0
3 w
3 w
EI 3
= EI 3
x x = c + 0
x x =c 0
(3.179)
These conditions can also be written in the terms of the slope , the bending moment
M and the shear force V :
w ( c + 0) = w ( c 0) ,
M ( c + 0) = M ( c 0) ,
(c + 0) = ( c 0) ,
V (c + 0) = V ( c 0)
(3.180)
Equations (3.180) underline that the displacement, slope, bending moment and shear
force must be continuous at every interface of the beam provided that neither external
moment nor external transverse force are applied at this interface.
Using the above collected equations, any elastostatic problem of the bending of a
beam can be solved. In what follows we consider a number of examples to
demonstrate the methods of solution.
107
(a)
w = 0, w x = 0
M = 0, V = 0
(b)
q1 L
V
1 2
q1 L
2
(c)
M
(d)
1 q1 L4
8 EI
Figure 3.18. (a) Clamped-free beam under uniform transverse loading; (b) shear force; (c) bending
moment; (d) transverse displacement
V ( L) = 0
(3.181)
(3.182)
(3.183)
(3.184)
Since the beam is cylindrical and homogeneous and the external force is constant, the
simplified expressions can be applied for the shear force, bending moment, and
transverse displacement, Eqs.(3.176)-(3.178).
Substitution of the simplified expressions for the shear force and bending moment,
Eqs. (3.176) and (3.177), into the boundary conditions at the right end of the beam,
Eqs.(3.183) and (3.184), gives the following system of equations:
0 = q1 L + C1
1
0 = q1 L2 + C1 L + C2
2
108
C1 = q1 L
(3.185)
1
C2 = q1 L2
2
This results in the following expressions for the shear force and bending moment:
V = q1 ( L x )
(3.186)
1
M = q1 ( x 2 + L ( L 2 x ) )
2
To find the displacement of the beam, we substitute the simplified expression for this
displacement, Eq.(3.178), into the boundary conditions at the left end, Eqs.(3.181) and
(3.182). This results in C3 = C4 = 0 and, consequently, in the following expression
for the beam displacement:
w=
q1 x 2 2
1 1
1
1
4
3
2
q
x
+
C
x
+
C
x
=
1
1
2
24 EI ( x + 2 L ( 3L 2 x ) )
6
2
EI 24
(3.187)
This displacement is shown in Figure 3.18 (d) together with the shear force and
bending moment, which are depicted in Figures 3.18 (b) and 3.18 (c), respectively.
Example B. Consider a cylindrical and homogeneous beam of the length L as shown
in Figure 3.19. Assume that this beam is clamped at the left end, and is subject to a
concentrated transverse force P and a concentrated moment M 0 at the right end.
x
P
M0
w = 0, w x = 0
M = M0, V = P
L
Figure 3.19. Beam clamped at the left end and subject to concentrated force and moment
at the right end
V ( L) = P
(3.188)
(3.189)
(3.190)
(3.191)
109
As in the previous example, to solve this problem, we may apply the simplified
expressions for the shear force, bending moment, and transverse displacement,
Eqs.(3.176)-(3.178).
Since no distributed force is applied to the beam, in accordance with Eqs. (3.176)
and (3.177), the shear force remains constant along the beam, whereas the bending
moment varies linearly. The constants C1 and C2 , which correspond to this variation,
can be easily found from the boundary conditions at the right end to give
C1 = P
C2 = M 0 PL
(3.192)
(3.193)
To find the displacement of the beam, we substitute the simplified expression for this
displacement, Eq.(3.178), into the boundary conditions at the left end, Eqs.(3.181) and
(3.182). This results in C3 = C4 = 0 and, consequently, in the following expression
for the beam displacement:
1 1
1
x2 1
3
2
w = C1 x + C2 x =
P ( x 3L ) + M 0
EI 6
2
2 EI 3
(3.194)
This displacement is visualised by the cubic parabola, which would reduce to the
ordinary parabola if the force P were absent.
Example C. Consider a pinned-pinned (simply supported) cylindrical and homogeneous beam of the length L as shown in Figure 3.20 (a). Assume that this beam is
subject to a distributed transverse force of the constant density per unit length q1 .
The boundary conditions in this case are given as
w (0) = 0
M (0) = 0
w( L) = 0
M ( L) = 0
(3.195)
(3.196)
(3.197)
(3.198)
To solve this problem we may apply the simplified expressions for the bending
moment and displacement, Eqs.(3.177) and (3.178). Substitution of these equations
into the boundary conditions results in the following system of four algebraic
equations with respect to the unknown constants C1 C4 :
0 = C4
0 = C2
110
1
1
1
q1 L4 + C1 L3 + C2 L2 + C3 L + C4
24
6
2
1 2
0 = q1 L + C1 L + C2
2
0=
1
q1 L,
2
C2 = 0,
C3 =
1
q1 L3 ,
24
C4 = 0
q1
(a)
(3.199)
x
w = 0, M = 0
w = 0, M = 0
L
q1 L 2
(b)
q1 L 2
V
L 2
(c)
1 2
q1 L
8
M
(d)
5 q1 L4
384 EI
Figure 3.20. (a) Pinned-pinned beam under uniform transverse loading; (b) shear force; (c) bending
moment; (d) transverse displacement
Thus, the shear force, bending moment and transverse displacement in the case at
hand are given as
1
q1 ( L 2 x )
2
1
M = q1 x ( L x )
2
qx
w= 1
x3 + L ( L2 2 x 2 )
24 EI
V=
(3.200)
111
(a)
w = 0, M = 0
w = 0, M = 0
L
q0
(b)
q0
q1
(c)
2
V
q0 L
(d)
M
(e)
w
Figure 3.21. (a) Pinned-pinned beam under uniform transverse loading; (b) external force; (c) shear
force; (d) bending moment; (e) transverse displacement
Since the external force is not constant along the beam, we should apply the
general expressions for the shear force, bending moment and displacement of the
beam, Eqs.(3.173)-(3.175). Substituting q1 = q0 cos ( x L ) into Eq. (3.173) and
evaluating the integral, we obtain
x
V ( x ) = q1 ( x ) dx + V ( 0 ) =
0
x
sin
+ V ( 0)
q0 L
(3.201)
112
M ( x ) = V ( x ) dx + M ( 0 ) =
0
q0 L x x
sin
+ V ( 0 ) dx + M ( 0 )
0 L
q0 L2
x
= 2 cos
1 + V (0) x + M ( 0)
L
(3.202)
1
EI
x x
( 0) x + w ( 0) =
M ( x ) dxdx
0 0
q0 L2 x
+ M ( 0) dxdx
( 0) x + w ( 0)
V
x
cos
1
0
(
)
0 0
1 q L2 L x 1
= 0 2 sin x + V ( 0) x 2 + M ( 0) x dx ( 0) x + w ( 0)
EI 0
L 2
1
=
EI
x x
2
1 q0 L2 L2 L2
x x V ( 0) x M ( 0) x
+
2 2 2 cos +
EI
6
2
L 2
3
(3.203)
( 0) x + w ( 0)
0 = M ( 0)
1 q0 L4 2 1 V ( 0 ) L M ( 0 ) L
0 = 2 2 +
+
EI
2
6
2
3
0 = 2
q0 L2
( 0 ) L + w ( 0 )
+ V (0) L + M ( 0)
qL
1 q0 L3 2 1
, M ( 0 ) = 0, V ( 0 ) = 2 0 2
2
2
6
EI
Thus, the expressions for the shear force, bending moment and transverse
displacement can be written as
113
q0 L 2
x
sin
2
L
q L
x
M ( x ) = 0 2 L cos
L + 2x
L
V ( x) =
w( x) =
(3.204)
x
1 q0 L L2
x
2
2
L L cos
2 x + ( 2 x 3Lx + L )
2
2
EI
L
6
(3.205)
w
(0) = 0
x
(3.206)
M ( L ) = kr
w
x
V ( L ) = kw
(3.207)
(3.208)
To solve this problem, the simplified expressions for the bending moment and
displacement, Eqs.(3.177) and (3.178), may be applied. Substitution of these
equations into the boundary conditions results in the following system of four
algebraic equations with respect to the unknown constants C1 C4 :
0 = C4
0 = C3
k 1
1
1
q1 L2 + C1 L + C2 = r q1 L3 + C1 L2 + C2 L + C3
2
2
EI 6
q1 L + C1 =
1
1
k 1
4
3
2
q1 L + C1 L + C2 L + C3 L + C4
EI 24
6
2
K ( 5 + K r ) + 24 (1 + K r )
1
q1 L
,
2
K ( 4 + K r ) + 12 (1 + K r )
C2 =
K ( 6 + K r ) + 24 ( 3 + K r )
1
q1 L2
,
12
K ( 4 + K r ) + 12 (1 + K r )
C3 = 0, C4 = 0
(3.209)
114
(a)
w = 0, w x = 0
M = kr w x , V = kw
L
(b)
11
q1 L
14
V
2 2
q1 L
7
(c)
M
(d)
1 q1 L4
7 EI
Figure 3.22. (a) Clamped-elastically supported beam under uniform transverse loading; (b) shear
force; (c) bending moment; (d) transverse displacement. The results are presented for K = 4, K r = 0
V=
(3.210)
115
Figure 3.23, presents another particular case, in which the translational spring is
absent, whereas the stiffness of the rotational spring is four times greater than the
overall rotational stiffness of the beam, i.e., K r = 4 and K = 0 . In this case the
expressions for the shear force, bending stiffness and transverse displacement read
V = q1 ( L x )
q1
30 Lx 15 x 2 11L2 )
(
30
q x2
w= 1
20 Lx 5 x 2 44 L2 )
(
120 EI
M=
(3.211)
q1
(a)
w = 0, w x = 0
M = kr w x , V = kw
L
(b)
11
q1 L
14
V
11 2
q1 L
30
(c)
M
(d)
11 q1 L4
30 EI
Figure 3.23. (a) Clamped-elastically supported beam under uniform transverse loading; (b) shear
force; (c) bending moment; (d) transverse displacement
If both the translational and the rotational springs are present, the beam pattern
depends significantly of the stiffness of these springs relative to the beam stiffness.
Example F. Consider a homogeneous beam of the length L , which consists of two
cylindrical parts, each of the length L 2 as shown in Figure 3.24(a). Both parts are
cylindrical but the cross-sectional moment of inertia of the left-hand part is I L ,
whereas the cross-sectional moment of inertia of the right-hand part is I R . Assume
116
that this beam is subject to a distributed axial force with the constant density per unit
length q1 , clamped at x = 0 and free at x = L .
The boundary conditions in the case at hand read
w (0) = 0
(3.212)
w
(0) = 0
x
M ( L) = 0
(3.213)
(3.214)
V ( L) = 0
(3.215)
w ( L 2 + 0) = w ( L 2 0) ,
M ( L 2 + 0) = M ( L 2 0) ,
q1
(3.216)
(a)
M = 0, V = 0
w = 0, w x = 0
L 2
L 2
(b)
q1 L
V
1 2
q1 L
2
(c)
M
(d)
1 q1 L4
8 EI
Figure 3.24. (a) Clamped-free combined beam under uniform transverse loading; (b) shear force; (c)
bending moment; (d) transverse displacement
Since both parts of the beam are cylindrical and homogeneous, the simplified
expressions for the shear force, bending moment and transverse displacement,
Eqs.(3.176)-(3.178), can be applied to each part of the beam. Thus, introducing the
117
4
3
2
wL =
q1 x + C1 x + C2 x + C3 x + C4
6
2
EI L 24
VR = q1 x + C5
1
M R = q1 x 2 + C5 x + C6
2
1 1
1
1
4
3
2
wR =
q1 x + C5 x + C6 x + C7 x + C8
6
2
EI R 24
(3.217)
(3.218)
(3.219)
(3.220)
(3.221)
(3.222)
(3.223)
Further, subsisting the expressions for the bending moment and shear force in the
right-hand-part of the beam into the boundary conditions at the right end, Eqs.(3.214)
and (3.215), we readily obtain the following expressions for C5 and C6 :
C5 = q1 L,
1
C6 = q1 L2
2
(3.224)
q1 L4 + C7 L + C8 =
q1 L4 + C1 L3 + C2 L2
I R 384
2
48
8
I L 384
1 7
1
1
1 1
3
3
2
q1 L + C7 = q1 L + C1 L + C2 L
I R 48
8
2
I L 48
1
1
1
q1 L2 = q1 L2 + C1 L + C2
8
8
2
1
1
q1 L = q1 L + C1
2
2
(3.225)
118
and the shear force, bending moment, and transverse displacement take the form
shown in Figures 3.24 (b-d). Note that the slope of the beam is continuous at the
interface.
3.5.4. Static Response of a Beam on Elastic Foundation. In this sub-section we
consider the beam on elastic foundation as shown in Figure 3.25.
kd
x
Figure 3.25. Beam on elastic foundation
(3.226)
d 4w
+ kd w = q1
dx 4
(3.227)
If, additionally, the external force q1 and the stiffness per unit length kd of the elastic
foundation are also coordinate-invariant, then the general solution of this equation can
be readily found. This solution is the superposition of a particular solution of
Eq.(3.227), which may be chosen as
wpart =
q1
kd
(3.228)
119
and the general solution of the homogeneous part of this equation. The latter solution
is to be sought in the form
4
wgen,hom = C n exp ( sn x )
(3.229)
n =1
Substituting Eq.(3.229) into Eq.(3.227) with the zero right-hand side, we obtain the
following characteristic equation
k
(3.230)
sn4 + d = 0
EI
Solving this characteristic equation, we find
s1 = (1 + i ) ,
s2 = (1 i ) ,
(3.231)
s3 = ( 1 + i ) , s4 = (1 + i )
where
4 4 =
kd
EI
(3.232)
= exp ( x ) C1 exp ( i x ) + C2 exp ( i x ) + exp ( x ) C3 exp ( i x ) + C4 exp ( i x )
= exp ( x ) ( C1 cos ( x ) + C2 sin ( x ) ) + exp ( x ) ( C3 cos ( x ) + C4 sin ( x ) )
where
C1 = C1 + C 2 , C2 = i C1 C 2 , C3 = C 3 + C 4 , C4 = i C 3 C 4
Thus, the general solution of Eq.(3.227) under the condition that EI , kd , and q1 are
invariant along the beam, reads
w = exp ( x ) ( C1 cos ( x ) + C2 sin ( x ) ) + exp ( x ) ( C3 cos ( x ) + C4 sin ( x ) ) +
q1
kd
(3.233)
Consequently, employing Eqs.(3.170) and (3.171), the bending moment and shear
force can be found as
(3.234)
120
)}
(3.235)
These expressions show that variation of the transverse displacement of, and the
bending moment and shear force in the beam on elastic foundation are represented by
the superposition of multiplications of the sinusoidal and exponential functions.
Normally, the solution of elastostatic problems concerning the beam on elastic
foundation, requires application of a program like MAPLE, which is capable of
performing symbolic calculations. This is beyond the scope of this course. In what
follows, only one example is considered, which has a practical significance and can be
analyzed employing simple evaluations.
Example A. Consider an infinitely long beam on elastic foundation, which is loaded
by a concentrated force P as shown in Figure 3.26. This model proves useful for
studying the static deformation of rails under a train wheel.
x=0
x
wL
wR
Figure 3.26. Infinitely long beam on elastic foundation subject to concentrated load
The boundary conditions in this case should be formulated not at the ends of the
beam but at the infinite distance from the load. Obviously, the larger the distance from
the load, the smaller must be the static response of the beam. Therefore, the boundary
conditions at infinity may be written as
lim w ( x ) = 0
(3.236)
At the cross section x = 0 , where the load is applied, the interface conditions
should be formulated. These are also particular in this case, since the external load P
must be accounted for. It is not difficult, however, to understand, using Table 3.8 and
the interface conditions at a load-free cross-section, Eqs.(3.179), that the following
conditions must be satisfied at x = 0 :
w ( +0 ) = w ( 0 )
w ( +0 ) w ( 0 )
=
x
x
2
2
w ( +0 ) w ( 0 )
=
x 2
x 2
3 w ( +0 ) 3 w ( 0 )
EI
=P
3
x3
x
(3.237)
121
The first three conditions represent continuity of the displacement, slope and bending
moment. The last condition is the balance of vertical forces, which act on the crosssection of the beam loaded by the external force.
The solution procedure of the problem at hand may be simplified by making use of
the symmetry of the problem. Indeed, the problem is symmetric with respect to the
loading point. Therefore, the response of the beam must be symmetric with respect to
this point as well. To take advantage of this symmetry, we introduce
wR ( x ) = w ( x ) x > 0
wL ( x ) = w ( x ) x <0
(3.238)
where the subscripts L and R stand for the parts of the beam, which lay at the left
and at the right from the load, respectively. Employing these notations, the interface
conditions at x = 0 can be rewritten as
wR ( 0 ) = wL ( 0 )
wR ( 0 ) wL ( 0 )
=
x
x
2
2
wR ( 0 ) wL ( 0 )
=
x 2
x 2
3 wR ( 0 ) 3 wL ( 0 )
EI
=P
3
x 3
x
(3.239)
(3.240)
(3.241)
(3.242)
The requirement of the symmetry of the beam displacement with respect to the
loading point can be formulated mathematically as
wL ( x ) = wR ( x )
(3.243)
With this requirement, the first and the third boundary conditions, Eqs.(3.239)
and(3.241), are satisfied automatically, whereas the second and the fourth,
Eqs.(3.240) and (3.242), can be written in the terms of wR only to give
wR ( 0 )
=0
x
3 wR ( 0 ) 1
= P
EI
x3
2
(3.244)
(3.245)
Let us now derive wR . We start with the general expression, Eq.(3.233), which in
the case of q1 = 0 , reads
wR = exp ( x ) ( C1 cos ( x ) + C2 sin ( x ) ) + exp ( x ) ( C3 cos ( x ) + C4 sin ( x ) ) (3.246)
122
(3.247)
C3 = C4 =
P
P
=
3
4 EI
kd
Thus, the deflection of the beam at the right of the load is given as
P
exp ( x ) ( cos ( x ) + sin ( x ) )
kd
wR =
(3.248)
Correspondingly, the bending moment and shear force at the right of the load read
P
exp ( x ) ( cos ( x ) sin ( x ) )
2
VR = P exp ( x ) cos ( x )
MR =
Using the symmetry requirement, the solution of the problem can be written such that
it becomes applicable to the entire beam. This may be accomplished by replacing x
by its absolute value in all expressions. This yields
w=
P
exp ( x ) cos ( x ) + sin ( x )
kd
M=
P
exp ( x ) cos ( x ) sin ( x )
2
(3.249)
V = P exp ( x ) cos ( x )
Note that the argument of cosinus is kept intact, since cosinus is an even function.
Equations (3.249) clearly show that the transverse displacement, bending moment
and shear force decay exponentially with the distance from the load. This decay,
however, is not monotonic but oscillatory along the beam.
The deflection of the beam along with the bending moment and shear force are
visualized in Figure 3.27 considering the vertical deflection of a railway track (two
rails) under an axle load of a train. To plot this figure, parameters of the UIC 60 rails
were employed, namely EUIC 60 =2.1 1011 N m-2, I UIC 60 = 3055 10-8 m4, and an average
effective stiffness of the soil against the railway track, k e ff = 108 N m-2 . From these
parameters we may retrieve the physical parameters of our model as
123
E = EUIC 60 = 2.1 1011 N m-2, I = 2 I UIC 60 = 6110 10-8 m4, k d = k eff = 108 N
(3.250)
1.2
1.0
40.0
0.8
30.0
M kN m
w mm
Bending Moment
50.0
0.6
0.4
20.0
10.0
0.2
0.0
0.0
-10.0
-0.2
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00
wm
wm
Shear Force
100.0
80.0
V kN
60.0
40.0
20.0
0.0
-20.0
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00
wm
Figure 3.27. Vertical displacement, bending moment and shear force for the UIC 60 railway track
under an axle loading
Figure 3.27 clearly shows that the rails are perceptibly deformed only within seven
meters from the load. Further away from the load, the rails are practically not
disturbed. This is due to the soil reaction, which we modelled with the help of elastic
foundation. Note that the softer the soil, the larger the response of the railway track
and the slower the decay of the deflection with the distance from the load.
124