Bridge Dyanmic
Bridge Dyanmic
DANIEL EKSTRÖM
LARS-LEVI KIERI
ABSTRACT
Today, the design of railway bridges to withstand dynamic loads caused by high
speed trains is both complicated and time consuming. Therefore simplifications are
desirable in order to facilitate the calculations. Based on literature studies some
simplified methods used to analyse beams subjected to dynamic loads are compiled
and also compared with finite element analysis in order to verify the results.
The method of transforming and reducing deformable structures into a single degree
of freedom system, giving calculations that are easy to handle, is discussed and
investigated in this thesis. When a beam is simplified into a single degree of freedom
system the beam is assumed to have a specific shape of deformation and therefore
tabled beam equations can be used in order to estimate the capacity of the beam.
These simplifications of the beam equations can also be applicable for dynamic loads,
and used to describe the system response.
In this thesis a single degree of freedom system is examined for various simple types
of loading, where the main focus has been to compare the response of displacement
and acceleration between different load types. This comparison gives a basic
understanding of the dynamic phenomena that causes oscillations and also their
effects.
Further the loading in the single degree of freedom system is modified to resemble a
train load model, HSLM-A. The results between the system and a finite element
model of a simply supported beam are compared, where the results show that the
system can be used to give an approximation of the response on the beam.
The simplified methods discussed above are only investigated for mutual parameters.
Since railway bridges often are made of reinforced concrete, which have a complex
structure and behaviour, the responses from the simplified methods due to dynamic
loads needs to be studied more in detail.
SAMMANFATTNING
Balkar, och andra deformerbara kroppar, kan omvandlas till ett enfrihetsgradsystem
som tillskrivs ekvivalenta egenskaper för att ge samma deformation som den
deformerbara kroppen. När balken har omvandlats till ett enfrihetsgradsystem kan
tabellerade, så kallade, balkekvationer användas för att direkt uppskatta balkens
respons.
SAMMANFATTNING II
CONTENTS III
PREFACE VIII
NOTATIONS IX
1 INTRODUCTION 1
1.1 Background 1
1.2 Aim 1
1.3 Method 1
1.4 Limitations 1
1.5 General layout 2
2 MATERIAL 3
2.1 Linear elastic material 3
2.2 Viscous elastic material 4
2.2.1 Newton material 4
2.2.2 Kelvin material 6
3 BASIC DYNAMICS 8
3.1 Kinematics 8
3.1.1 Velocity 8
3.1.2 Acceleration 9
3.2 Kinetics 9
3.2.1 Newton’s second law 9
3.2.2 Single degree of freedom (SDOF) system 10
3.2.3 Free vibration – Undamped 10
3.2.4 Free vibration - Damped 12
3.2.4.1 Critical damping ξ =1 13
3.2.4.2 Strong damping ξ >1 14
3.2.4.3 Weak damping ξ <1 14
3.2.5 Forced vibration – Undamped with a harmonic load 15
3.2.6 Forced vibration – Damped with a harmonic load 17
3.3 Resonance and dynamic amplification factor 18
3.3.1 Undamped system 19
3.3.2 Damped system 19
4 BEAM DYNAMICS 21
7 SDOF ANALYSIS 41
7.1 General 41
7.2 Analytical solution 41
7.2.1 Undamped system 41
7.2.2 Damped system 42
7.2.3 Assumptions and analysed parameters 43
7.2.4 Excitation 44
7.3 Numerical solution 45
7.3.1 Numerical formulation 45
7.3.2 Assumptions and analysed parameters 46
7.3.3 Excitation 46
7.3.4 Train load 48
11 CONCLUSIONS 97
11.1 General 97
11.2 Further investigations 99
12 REFERENCES 100
Morgan Johansson, PhD, from Reinertsen Sverige AB, has supervised us with
engagement, and we specially thank him for his support and good ideas. Thanks also
to the rest of Reinertsen’s staff for answering various questions that appeared during
the work.
Our opponents, Simon Skoglund and Johan Nesset, have continually assisted with the
development of this thesis and we thank them for grateful help and good ideas of
improvements.
Dynamic analyses of railway bridges are today very complicated and time consuming,
because of the large, complex 2D- and 3D computer models and the different load
cases from high speed train models. Therefore the company Reinertsen has interest in
examining the dynamic phenomena’s and find out if there is a more simple way to
analyse bridge structures than the 2D modelling.
1.2 Aim
The aim of this thesis is analysing and increasing the knowledge of the dynamic
behaviour of railway bridges. This thesis should give a basic understanding of
dynamic phenomena that causes bridges to oscillate and give a guidance of
simplifications that can be made during modelling of bridges. The simplifications are
mainly aimed to be used in the early stages in the design process and lead to simpler
models than those used today. Further, the study aims to make it possible to decide if
any of the high speed train models can be neglected.
1.3 Method
Literature studies have been done in order to find, understand and compile different
simple methods used when analysing the behaviour of structures exposed for dynamic
loads. Literature studies have also been made in order to get a deeper understanding of
dynamics, their appearance and effects. The agreement between such simple methods
is investigated by comparing results from simple models with the real behaviour,
assumed to be found by using a multi degree finite element model. The finite element
analyses are made by means of the commercial finite element software ADINA
(2004).
1.4 Limitations
The methods described in this thesis, used in order to simplify analyses of structures
subjected to dynamic loads, can be used on different types of deformable structures.
However, only the application on single degree of freedom systems (SDOF systems)
and simply supported beams are treated in this thesis.
When investigating the train load, only the model HSLM-A1 is used and the
investigated responses are displacement, velocity and acceleration. The reason for
using only one train model is due to the limited time for this thesis.
Dynamic loads and their effects on railway bridges is a huge subject which requires
long time to fully understand. Due to the limited time and in order to keep this scope
within reasonable limits only SDOF systems and simply supported beams that are
subjected to a dynamic load are discussed in this thesis.
Chapter 6 describes some of the rules that the Swedish railway administration has for
designing bridges for train speeds above 200 km/h.
Chapters 7 and 8 describe the SDOF system and how it is examined for various types
of loading. The responses of the SDOF system are calculated and the behaviour of the
system is examined for both undamped and damped systems with various loading.
Chapters 9 and 10 describe the finite element model of a simply supported beam that
is investigated for various loading. The response of the beam is calculated by use of
this finite element model and its behaviour is examined for both undamped and
damped beams with various loading. The results from the finite element model are
compared with the results from the SDOF system.
σ = Eε (2.1)
The proportional constant E is called the modulus of elasticity. The principle relation
between stress and strain of a linear elastic material is shown in Figure 2.1(a). The
linear elastic material can be described by a spring, see Figure 2.1(b). A loading of the
structure with the stress σ=σ0 gives the response of a strain ε0. If the load is removed
at time t1 the strain will also disappear, see Figure 2.1(c-d).
σ0
E
σ ε0
ε ε t t
t1 t1
EA
FE = σA = EAε = u = ku (2.2)
L
where k is the stiffness of the 1D spring. A principle relation between the elastic force
and the displacement for a 1D linear elastic material is shown in Figure 2.2.
Figure 2.2: The principal relation between the elastic force and the displacement
for a 1D linear elastic material.
σ σ
ε
ε0 σ0 σ0
σ0 ε0
ε
ε0 ε0 ε
Figure 2.3: The creep and relaxation phenomena for a viscous elastic material.
dε σ
= ε& = ε& N = (2.3)
dt η
The stress is linearly proportional to the time dependent strain and η is the constant of
viscosity, see Figure 2.4(a). An instantaneous loading of the structure with the stress
σ=σ0 gives the response of a time dependent strain. If the load is removed at time t1
σ σ t
t1
ε r = ∫ 0 dt = 0 1 (2.4)
0
η η
The Newton material can be described by a model of a damper, see Figure 2.4(d).
σ
σ ε
σ0
η εr σ
t ε
ε& t
t1 t1
(a) (b) (c) (d)
The damping force FD in a 1D structure with area A subjected to a load will thus be
linearly proportional to the velocity u& , i.e.:
where c is the damping of the 1D structure. A schematic relation between the damping
force and the velocity for a viscous elastic material is shown in Figure 2.5.
FD
u&
Figure 2.5: The principle relation between damping force and the velocity for a
viscous elastic material.
ε1 = ε H ε2 = εN
σ1 = σ H σ2 =σN
(1) (2)
σ ε
The constitutive relation and the differential equation for the Kelvin material can be
derived as:
E σ
σ = σ H + σ N = Eε H + ηε N = Eε + ηε& where ε& + ε= (2.6)
η η
To be able to describe instantaneous loading of the structure with a stress σ=σ0, the
derivative of the strain has to be rewritten by use of the chain rule as:
dε dε dσ dε
ε& = = = σ& (2.7)
dt dσ dt dσ
dε 1 E 1σ
+ ε= (2.8)
dσ σ& η σ& η
For instantaneous loading at time t=0 it holds that if σ& → ∞ Equation (2.8) gives:
dε
= 0 ⇒ ε (σ ) = Constant
dσ
(2.9)
dε
but since ε = 0 , when σ = 0 this means that =0
dσ
The initial condition of ε(0)=0 gives C=-σ0/E and the strain is then:
The instantaneous loading at t=0 affects at first the viscous part only which carries the
whole stress σ=σ0. When t→∞ the strain limit is obtained:
σ0
ε∞ = (2.12)
E
σ 0 t1
−E −E
σ0 σ0 t1
ε1 = − e η
= 1− e η (2.13)
E E E
If the loading is released at time t=t1, such that σ(t<t1)≠0 and σ(t1)=0, see Figure
2.4(b), this will give no jump in the strain ε according to Equation (2.9). The
differential equation in Equation (2.6) can now be solved with the condition σ(t1)=0.
E
t ≥ t1: ε& + ε =0 (2.14)
η
The initial condition is ε(t1)=ε1 according to Equation (2.14) which gives the solution
of Equation (2.16):
−E
(t −t1 )
ε = ε 1e η for t > t1 (2.16)
When t → ∞ means that ε → 0 and there is no remaining deformation, see Figure 2.7.
ε
Hooke material
Newton material
ε
∞
σ /E Kelvin material
0
3.1 Kinematics
The linear motion of a particle is the simplest way to describe a movement of the
particle. A particle P, see Figure 3.1, is restricted to move along the s-axis and the
position is described by a function f (t), where t is the time. At time t the particle has
the position s and with a provided time step of ∆t the particle moves a distance ∆s.
P P′
s
s s + ∆s
t t + ∆t
3.1.1 Velocity
The velocity for the same particle as described in Figure 3.1 is derived by studying
how fast the position of the particle is changing. When the time changes from t to t+∆t
the particle moves a distance ∆s and by that the mean velocity during the movement
can be stated as:
∆s
v= (3.1)
∆t
The velocity of the particle is stated by letting the time step ∆t go towards zero. That
will lead to P′ moves closer to P and the mean velocity will approach a boundary
value. Therefore the velocity of the particle at time t is defined by the boundary value
as:
∆s ds
v(t ) = v = lim = ≡ s& (3.2)
∆t →0 ∆t dt
3.1.2 Acceleration
It is often interesting to know how fast the velocity varies as the particle is moving,
therefore the velocity of the particle is studied in the points P and P′, see Figure 3.1.
The particle in these points has a velocity of v and v+∆v respectively. The mean value
of the acceleration is defined as the mean velocity change per time unit with a particle
movement from point P to P′. The mean value of the acceleration can be stated as:
∆v
a= (3.3)
∆t
In the same way as when deriving the velocity of the particle the acceleration can be
written as:
∆v dv d 2s
a(t ) = a = lim = ≡ v& = 2 ≡ &s& (3.4)
∆t →0 ∆t dt dt
An important aspect of describing the particle movement is that there can occur
several combinations of sign changes of the values v and a. In the case when the
particle moves in a positive direction along the s-axis, the velocity has a positive value
and an increasing acceleration will lead to an increasing velocity. If the acceleration
would decrease, this instead corresponds to a decreasing velocity. The same
phenomena occur if the particle moves in the negative direction along the s-axis, but
instead with a negative velocity.
3.2 Kinetics
The response of bodies subjected to dynamic forces can be described by means of
differential equations abbreviated as DE. This chapter will only describe linear
vibrations with a single degree of freedom abbreviated as SDOF. In a SDOF system
the position for the body is defined by one coordinate. Before deriving these equations
of motion for dynamic loads the Newton’s second law is defined.
d
FI = k (mv ) (3.5)
dt
d
FI = (mv ) = ma = m&s& (3.6)
dt
In a stable equilibrium position the force in the spring is equal to the gravity force of
the mass. The force changes with the deformation of the spring, while the gravity
force is independent of the position.
k c
Dynamic vibrations occur when the system is disturbed from its stable equilibrium
position. The disturbance creates internal forces that try to bring back the system to its
equilibrium position and this phenomena causes oscillations. The system will oscillate
around its equilibrium position until the damping has reduced the oscillation to zero
and finally a new stable equilibrium has occurred.
FE = ku (3.7)
u m
mg
When the body is moved a distance u from the unloaded equilibrium position and then
released the system will undergo an undamped free vibration about the unloaded
equilibrium position. The forces acting on the isolated body is shown in Figure 3.3
where mg is the dead weight of the system.
Due to dynamic equilibrium conditions the sum of the forces shall be zero.
mu&& + ku = 0 (3.9)
k
u&& + ω 2 u = 0 where ω = (Circular frequency) (3.10)
m
where the A and θ respectively C1 and C2 are constants of integration and they are
determined from the boundary conditions.
When the system has started to oscillate, it will oscillate endlessly with the same
amplitude A since the system is not affected by any kind of damping, see Figure 3.4.
1 2π
T = = - Period
t f ω
θ /ω ω t + θ - Phase angular
(If A>0)
T
θ - Phase constant
Figure 3.4: Oscillation of an undamped system.
The system in Figure 3.5, reminds a lot about the Kelvin material described in Section
2.2.2. So, here the properties for the spring and damper are combined together. The
damping of the system is noted as c and the damping force FD for the system,
described in Section 2.2.1, can be expressed as:
FD = cu& (3.12)
k c
ue k c
cu&
mg + ku
m
mu&&
u m
mg
Due to dynamic equilibrium conditions the sum of the forces shall be zero.
c c k
u&& + 2ξωu& + ω 2 u = 0 where ξ = = and ω = (3.15)
ccr 2 km m
As can be seen in Equation (3.15), ξ is a percentage of the critical damping ccr, see
Section 3.2.4.1. Setting u=eλt gives the characteristic equation as:
(
λ2 + 2ξωuλ + ω 2 = 0 with roots λ1, 2 = − ξ ± ξ 2 − 1 ω ) (3.16)
The two roots of Equation (3.16) have the same value and that leads to a solution that
contains a polynomial. In this case of a first order equation as:
u (t ) = ( At + B )e −ωt (3.18)
where A and B is constants of integration and they are determined from the boundary
conditions. This function is also a non-periodic and has the same principal shape as in
Figure 3.6.
If the roots for Equation (3.16) are both real this result in a general solution as:
where C1 and C2 are constants of integration that are real and determined from the
boundary conditions.
ξ=1
ξ >1
Figure 3.6: Example of typical oscillation of a damped system with critical and
strong damping.
The boundary conditions for the system decide the curvature of the oscillation and a
period can not be found. The amplitude approaches exponentially towards zero with
time due to the roots of Equation (3.16) are negative, see Figure 3.6.
If the roots for Equation (3.16) are imaginary, then the general solutions of the
differential equation (3.14) are:
B1 and B2 are constants of integration that are real and determined from the boundary
conditions.
u(t)=Ae ξω
- t
A1 A2
0 t
The difference between the undamped and the weak damped case is that the amplitude
is decreasing exponentially with time and the oscillation has a lower circular
frequency ωd, which leads to a longer period Td, see Figure 3.7. According to Bergan
and Larsen (1986), the case of critical and strong damping rarely or never occurs in
real structures and therefore only weak damping will be treated further in this thesis.
Whenever damping is discussed or mentioned it is the case of weak damping.
k
ue k
m
mg + ku
p (t ) u m
mu&&
mg
p(t ) p(t )
Assume that the load in this case is harmonic and therefore periodic and has a shape
of sinus function. The load is defined as:
u p (t ) = C 3 sin ω p t (3.26)
(− mω p
2
)
+ k C 3 sin ω p t = p0 sin ω p t where
p0 p0 1 p0 1 ωp (3.27)
C3 = = = and β =
k −ω p m
2
k ω p
2
k 1− β 2
ω
1 −
ω
p0 1
u (t ) = (C1 sin ωt + C 2 cos ωt ) + sin ω p t (3.28)
k 1− β 2
where C1 and C2 are constants of integration that are determined from the boundary
conditions.
k c
ue k c
m
cu&
mg + ku
u m
p (t )
mu&&
mg
p(t)
p(t )
Due to dynamic equilibrium conditions the sum of the forces shall be zero.
Assume that the load in this case is harmonic and therefore periodic and has a shape
of a sinus function. The load is defined as:
The constant C1 and C2 are solved by combing Equation (3.32) and Equation (3.34).
p0 1− β 2 ωp
C1 = where β =
(
k 1 − β 2 + (2ξβ )2
2
) ω
(3.35)
p 2ξβ
C2 = − 0
(
k 1 − β 2 + (2ξβ )2
2
)
The particular solution for the DE is then:
u p (t ) =
p0 1
[( )
1 − β 2 sin ω p t − 2ξβ cos ω p t ] (3.36)
(
k 1 − β 2 + (2ξβ )2
2
)
The general solution for the DE is then:
where B1 and B2 are constants of integration that are determined from the boundary
conditions.
u (t ) = u p = R sin (ω p t − θ ) where
( ) p0
1
1
R = C12 + C 22 2 = and
k
(1 − β ) + (2ξβ )
2 2 2
(3.38)
C2 2ξβ
θ = arctan − = arctan
2
C1 1− β
Then the constant of integration C1 and C2 are solved from Equation (3.28) and the
total solution is derived as:
u& p β p 1
u (t ) = 0 − 0 sin ωt + u 0 cos ωt + 0
2
sin ω p t (3.40)
ω k 1− β k 1− β 2
Since the boundary conditions in Equation (3.28) describe that the system initially
was in a stable equilibrium position, i.e. u (0) = u 0 = 0 and u& (0) = u& 0 = 0 , the response
can be stated as:
u (t ) = u static
1
(sin ω p t − β sin ωt ) (3.41)
1− β 2
where ustatic=p0/k describes the static displacement of the system. The dynamic
amplification factor is derived by:
u p (t )max 1 ωp
DAF = = where β = (3.42)
u static 1− β 2 ω
The dynamic amplification factor increases rapidly when the load frequency closes
the natural frequency of the system and decreases when it has larger frequencies.
u max 1
DAF = = (3.43)
u static (1 − β ) + (2ξβ )
2 2 2
The dynamic amplification factor has the same effects as in the undamped case, but it
will be smaller when the system is damped, see Figure 3.10.
5
ξ = 0.1
4
ξ = 0.125
ξ = 0.15
DAF
ξ = 0.2
2
ξ = 0.25
ξ = 0.4
1
ξ=1
ξ=2
0
0 0.5 1 1.5 2 2.5 3
β
Figure 3.10 Illustrate the effects of DAF, damping coefficient and relationship
between the load frequency and the natural frequency of the system.
To determine the beam response versus time the eigenmodes for a simply supported
beam can be computed, by verifying the eigenvalue equation with the homogeneous
boundary conditions.
d 4u
EI − ω 2 mu = 0 for 0 < x <l
dx 4
(4.1)
2
d u
u= = 0 at x = 0, l
dx 2
2 nπx
u ( n ) ( x) = sin where n = 1, 2, 3,….. (4.2)
ml l
EI
ω (n ) 2 = (nπ )4 (4.3)
ml 4
The above stated expressions are based and derived from energy expressions for a
continuous beam. An attempt to give a background and derive the expressions above,
in a simplified way, can be seen in Appendix A.
In Figure 4.1, the three first eigenmodes for a simply supported beam are shown. The
first eigenmode corresponds to the lowest eigenfrequency.
Figure 4.1: The three first eigenmodes for a simply supported beam.
Normally, when a beam is subjected to a dynamic load, the load frequency will not
coincide with the eigenfrequencies and therefore the resulting shape of deformation
q ( x, t )
ke ce
M , L, E , I , C , σ y
me
pe (t )
The simplification to a SDOF system implies that the properties of the continuous
system has to be assigned with equivalent quantities for the mass m, the internal force
FI, the damping force FD and the load p(t) applied to a certain system point. The
deflection in the system point is assumed to be described by the same function as for
the SDOF system. The system point is chosen to coincide with the point that normally
will achieve the largest displacement, i.e. the midspan in the case of a simply
supported beam, see Figure 4.3. One condition, for the transformation of the
properties to be possible, is that a uniform change of the deformation is assumed. This
means that if the displacement increases in one point the displacements in all other
points will increase proportional to this displacement.
Shape of deformation
System point at time t = t1
Shape of deformation at time
t = t2
As discussed earlier in Section 3.2.6, the differential equation for the SDOF system in
Figure 4.2 can be stated as Equation (3.30). Using the notations from Figure 4.2 the
differential equation can be stated as:
The equivalent quantities for the mass, internal force and the load can be expressed by
means of transformation factors.
Combining Equation (4.4) and (4.5), we obtain the definition of the transformation
factors.
me
κM = (4.6)
m
keus
κK = (4.7)
ku
p e (t )
κP = (4.8)
p(t )
If the expression in Equation (4.5) is divided with Equation (4.8), we can state three
new transformation factors.
κM
κ MP = (4.9)
κP
κK
κ KP = (4.10)
κP
The kinetic energy generated by the equivalent mass in the SDOF system is:
2
SDOF me v s
Wk = (4.12)
2
∆u s
where v s = is the velocity of the system point in the vertical direction.
∆t
ρ density [kg/m3]
∆u
v = v( x) = velocity of arbitrary point in the vertical direction [m/s]
∆t
The change of the displacement in an arbitrary point in the beam can be expressed as:
∆u = u ( x, t2 ) − u ( x, t1 ) = αu ( x, t1 ) − u ( x, t1 ) = (α − 1)u ( x, t1 ) (4.15)
u ( x , t1 )
u ( x, t2 ) = α ⋅ u ( x, t1 )
Shape of deformation
at time t = t1
System point
Shape of deformation at
time t = t2
u( x, t2 ) − u( x, t1 ) = (α − 1)u( x, t1 )
Where u(x,t1) is the displacement at time t=t1 at the distance x from the end of the
beam and u(x,t2) is the displacement at the same point in the longitudinal direction of
the beam at time t=t2. Due to that the change in displacement is uniform, it can
according to Figure 4.4 be said that:
u ( x , t 2 ) = α ( x , t1 ) (4.16)
The change of displacement in the system point, when time goes from t=t1 to t=t2, can
be expressed as, see Figure 4.4:
and since the assumption of uniform displacement in valid for all times t, the general
form of Equation (4.15) and (4.17) can be written as:
∆u = (α − 1)u ( x, t ) (4.18)
∆u s = (α − 1)u s (t ) (4.19)
Using Equation (4.18) and (4.19) together with that the velocity in any arbitrary point
x in vertical direction can be expressed as v=∆u/∆t, Equation (4.14) can be rewritten
as:
x=L
((α − 1)u (x, t ))2 x=L
u ( x, t )
2
me = ∫
x = 0 ((α − 1)u s (t ))
2
ρAdx = ∫
x =0 u s (t )
2
ρAdx (4.20)
Now using Equation (4.6) together with the assumption of uniformly distributed mass
along the beam length, the transformation factor for the mass can be written as:
1 u ( x, t ) 1 u ( x, t )
x= L 2 x=L 2
m x∫=0 u s (t ) L x∫=0 u s (t )
κM = ρAdx = dx (4.21)
i.e. the transformation factor for the mass is depending on the assumed shape of the
deformation.
The internal force and the work it performs are depending on the behaviour of the
material. For the SDOF system this is shown in Figure 4.5, where the shaded areas
represent the total internal work for the material. FI,e,max is the maximum value of the
equivalent internal force. In the case of linear elastic material the maximum internal
force is corresponding to FI,e,max=Ke·us,max.
FI ,e
FI ,e ,max
Ke
1
u s ,max us
Figure 4.5: Work for a SDOF system for a linear elastic material.
The internal force for the SDOF system can be expressed by the the spring relation
shown in Figure 4.5.
FI ,e = k e u s (4.22)
Following Samuelsson and Wiberg (1999) the work of deformation for the beam
made of linear elastic material can be derived by studying a lamella of length ∆x and
the sectional forces, N , and deformations, ∆n , belonging to it, see Figure 4.6.
V
M
M z
N N
V
∆x
EA 0 0 N ∆n
1
N = 0 GA 0 ∆n , N = V ∆n = ∆t
(4.23)
∆x β
0 0 EI M ∆m
E
G= shear modulus [Pa]
2(1 + υ )
The meanings of the deformations ∆n, ∆t and ∆m are shown in Figure 4.7 .
∆n ∆m
∆t
The constant β can be derived from the statement that the work of deformation due to
shear force shall be equal to the work of deformation due to shear stress.
z =h
Vβ
Vγ = V
GA
= ∫ τ ( z )γ ( z )b( z )dz
z =0
(4.24)
Vβ
where γ = average value of shear angle [-]
GA
τ
γ = shear angle [-]
G
EA GA EI
dΠ is = ∆n d∆n + ∆td∆t + ∆m d∆m = N t d∆n (4.26)
∆x β∆x ∆x
In order to get the total work of deformation of the segment, Equation (4.26) will be
integrated over the deformation ∆n .
∆n ∆t ∆m
EA GA EI
Π is = ∫
∆n = 0
∆x
∆n d∆n + ∫
∆t = 0
β∆x
∆t d∆t + ∫
∆m = 0
∆x
∆m d∆m =
(4.27)
= EA(∆n ) +
2 GA
(∆t )2 + EI (∆m )2 1
β 2∆x
Once again using Hooke’s law and integrating the work of deformation for the
segment over the length, L, of the beam will give the total work of deformation for the
beam.
x=L x=L
Π is N 2 βV 2 1
Π beam
i = ∫ dx = ∫x = 0 EA GA
+ + M ( x )u ′′( x ) dx (4.28)
x = 0 ∆x 2
If the influences from the normal- and shear forces are neglected the total work of
deformation for the beam can be written as:
x=L
1
Π = ∫ M ( x)u′′( x)dx
beam
i (4.29)
2 x =0
ξ
Figure 4.8: Mass in a) equilibrium and b) moved ξ from equilibrium position.
Study the undamped SDOF system in Figure 4.8. The displacement ξ causes an
internal work for the SDOF system which by use of Equation (4.22) can be written as:
ξ =u s 2 2
ku ku
Π SDOF
i = ∫ k eξdξ = e s = κ K s (4.30)
ξ =0 2 2
As stated earlier the total internal work of the SDOF system shall be equal to the total
work of deformation of the beam, meaning that Equation (4.29) shall be equal to
Equation (4.30).
2 x=L
ku s 1
κK = ∫ M ( x )u ′′( x )dx (4.31)
2 2 x =0
The stiffness k of the beam is depending on shape of the load and is determined by:
x=L
∫ q(x, t )dx = ku
x =0
s (4.32)
The definition of stiffness k of the beam according to Equation (4.32) together with
Equation (4.31) gives the final expression of the transformation factor for the internal
force when having a linear elastic material.
x=L
1
x= L
1
∫ M (x )u ′′(x )dx
κK = ∫x=0M (x )u ′′(x )dx = u s
x =0
2 x=L (4.33)
ku s
∫ q(x, t )dx
x =0
For high beams it might be necessary to include the influences from the shear forces
to get adequate results, see Section 4.3.5 for further discussion.
k
c = ξccr = ξ 2mω = ξ 2m = ξ 2 km (4.34)
m
As can be seen in Equation (4.34), the damping is determined by the mass m and the
stiffness k. By inserting the equivalent values derived earlier, and thereby adjusted,
the expression for the equivalent critical damping can be stated as:
ke
ce = ξccr ,e = ξ 2meω = ξ 2me = ξ 2 k e me (4.35)
me
Later in the thesis this assumption is verified by comparing the results between
different solution methods, see Section 10.2.2. Here the results between a SDOF
solution using the regular c, and a solution based on the transformed equivalent ce, are
compared and shows coinciding results.
The work generated by the equivalent load in the SDOF system during a time
increment ∆t is:
Π SDOF = p e (t )u s (t ) (4.36)
where: x is the coordinate with origin at one end of the beam [m]
x=L
u ( x, t )
x=L x=L
p e (t )u s (t ) = ∫ q(x, t )u (x, t )dx ⇔ p e (t ) = ∫ q(x, t ) u (t ) dx (4.38)
x =0 x =0 s
The transformation factor for the load, see Equation (4.8), can now be written as:
x=L
u ( x, t )
∫ u s (t )
q ( x, t )dx
κP = x =0
x=L (4.39)
∫ q( x, t )dx
x =0
Also the transformation factor for the external load is depending on the assumed
shape of deformation. It is further depending on the shape of the load.
Figure 4.9: Simply supported beam loaded with a concentrated load in midspan.
Material κP κM κK κ MP κ KP
FI (t ) + FD (t ) + FE (t ) = p(t ) (5.1)
where FI(t) are the inertia forces, FD(t) are the damping forces and FE(t) are the
internal forces and p(t) are the external forces. The forces in Equation (5.1) are all
time-dependent. From the equation of equilibrium governing the linear dynamic
response this can be stated as:
where M, C and K are the mass, damping and stiffness matrices respectively and P is
the vector of externally applied loads. U , U& and U&& are the displacement, velocity and
acceleration vectors. The Equation (5.2) is a linear DE of the second order with
constant coefficients.
In the direct integration method in Equation (5.2) it is assumed that at time t=0 the
displacement, velocity and acceleration vectors are known and they are donated as
0
U , 0U& and 0U&& . The solution is made for the time span 0≤ t ≤T, where the time is
divided into n time steps ∆t=T/n. The step-by-step procedure calculates the solution to
the next required time from the solutions at the previous times. Assume that the
solutions at times tt-∆t and t∆t are known, then the next solution is calculated at time
tt+∆t. Finally the integration procedure will give an approximate solution to the DE
within the time interval at times ∆t, 2∆t,…, t, t+∆t,..T. In this report only linear
analysis will be used with a constant time step ∆t.
2
The parameters α and δ determines the accuracy and stability of the integration. When
δ=1/2 and α=1/6, the relations between Equation (5.3) corresponds to the linear
acceleration method, while δ=1/2 and α=1/4, the relations corresponds to the
constant-average-acceleration method (also called trapezoidal rule). The trapezoidal
rule is an unconditionally stable scheme, meaning that there is no demand on the
incremental time step to reach a stable solution, see Figure 5.1.
t t + ∆t
The solution of the displacements, velocities and accelerations at time t+∆t are
achieved by combining Equation (5.3) and Equation (5.2). The implicit time
integration is one of the available methods to solve dynamic problems in the finite
element program ADINA. The complete solution procedure according to
Bathe (1996) is shown in Table B.1 in Appendix B.
The central difference method is derived from the Newmark method by using the
values of δ=1/2 and α=0 in Equation (5.3). When combining the two equations with
each other this leads to:
1
t + ∆t
U = t U + t + ∆t U&∆t − t + ∆tU&&∆t 2 (5.4)
2
U
∆U
2∆t t
t t + 2∆t
∆U dU
v(t ) = v = lim = ≡ U&
∆t →0 ∆t dt
(5.5)
t + ∆t &
U=
1
2∆t
( t + 2 ∆t
U −tU )
The acceleration and the velocity are approximated in terms of displacements, and by
combining Equation (5.5) into Equation (5.4) and changing the variable t+2∆t to
t+∆t. The acceleration is derived as:
t 1
U&& = 2 ( t − ∆t
U − 2 t U + t + ∆t U ) (5.6)
∆t
t
U& =
1
2∆t
(
− t − ∆t U + t + ∆t U ) (5.7)
By inserting Equation (5.6) and Equation (5.7) into Equation (5.8) gives:
1 1 t + ∆t 2 1 1 t − ∆t
2M+ C U = t P − K − 2 M tU − 2 M − C U (5.9)
∆t 2∆t ∆t ∆t 2∆t
∆t 0 && 2
− ∆t
U = 0 U − ∆t ⋅0 U& + U (5.10)
2
To have a more effective procedure of solving the DE, the mass matrix and damping
matrix are chosen to be diagonal. This results in that the mass matrix and the load
vector can be lumped and introduced into Equation (5.9). The lumped mass matrix is
calculated as:
1 1
Mˆ = 2 M + C
∆t 2∆t
M 1 0 . 0
(5.11)
0 M2 . 0
Mˆ =
. . . .
0 0 . Mn
2 1 1 t − ∆t
t
Pˆ = t P − K − 2 M tU − 2 M − C U (5.12)
∆t ∆t 2∆t
Mˆ t + ∆tU = t Pˆ (5.13)
To have accuracy of the solution based on the central difference method, the time step
needs to be small enough. The time step should be smaller than the value of a critical
time step, which can be calculated as:
Tn 2
∆t ≤ ∆t cr = = (5.15)
π ω max
A time step that is too large is easily noticed from the solution, by that the responses
of the displacements, velocities and accelerations grows very fast and finally becomes
unrealistic. The complete solution procedure according to Bathe (1996) is shown in
Table B.2 in Appendix B.
C Rayleigh = αM + βK (5.16)
where M is the mass matrix in Equation (5.2), which can be lumped or consistent, and
K is the stiffness matrix that corresponds to zero initial displacements. The Rayleigh
damping constants α and β can be determined from a minimum of one given damping
ratios ξi that correspond to two unequal circular frequencies of vibration ωi and ωj, see
Figure 5.3.
ξi
ω
ωi ωj
By using the two unequal circular frequencies the damping ratio can be stated in two
different expressions as can be seen in Equation (5.17). The Rayleigh constants α and
β are then determined by combining the two expressions with each other:
α βω i α βω j
ξi = + and ξ i = + (5.17)
2ω i 2 2ω ´ j 2
A disadvantage with Rayleigh damping is that the higher modes are considerably
more damped than the lower modes, for which the Rayleigh constants have been
selected. Also the damping is incorrect for all other modes except for the two mode
shapes related to the given circular frequencies.
The mass proportional damping in the explicit time integration is defined as:
Where M̂ is the lumped mass matrix and the damping matrix CRayleigh is replacing the
damping matrix C in Equation (5.2).
where φi, i=r,…, s are the mode shapes calculated in a frequency analysis, and the ui
are the corresponding unknown generalized displacements. The displacements ui are
calculated by solving the decoupled modal equations:
where ξi is the critical damping ratio corresponding to the circular frequency ωi, and
pi=φiP. This equation can be recognised from Chapter 3. The mode superposition
method can be solved with many different time integration methods, but the Newmark
method, following trapezoidal rule, is used in ADINA, see Equation (5.3). Mode
superposition is effective when the time integration has to be carried out over many
time steps. The cost of calculating the required frequencies and mode shapes is
reasonable.
In the mode superposition method the damping is specified for each mode and the
values of the modal damping ξi, i=r,...., s can all be different. The modal damping for
each mode can be determined by using Rayleigh damping or it is also possible to
define it directly for each mode in ADINA.
D NxD D
d d d d
3 11 3 D
3,525
The model HSLM-B shall be used for simple bridges, for bridges with one span and a
span length less than 7.0 m. By simple bridges means simply supported beam bridges
and simply supported slab bridges. The model HSLM-B consists of N number of point
loads with a value of 170 kN and a spacing of d, see Figure 6.2 - Figure 6.3.
N x 170kN
d d d d d d d d d d d d d d d
Figure 6.3: The spacing d between loads and number of loads N dependent on the
span width.
Table 6.2: Damping factor for different spans and type of bridges.
When using the numerical solution different types of loading are used but the total
impulse of each load pulse is kept constant in order to determine if some type of load
is more dangerous than others. Finally a train load based on the train model HSLM-A
is used as excitation of the SDOF system. The results from the analysis of a SDOF
system will later be compared with the results from a finite element model. This in
order to examine to what degree a simplified analysis, such as the SDOF system, can
be used.
Based on the assumptions, material properties, load magnitude etc. used in the
calculations and that will be presented later, none of the calculated values in the thesis
are intended as real values, but are only intended as mutual comparison with each
other. The inputs used have been chosen in order to simplify the comparison between
different load cases.
All the solutions described in this chapter is programmed and solved with the
commercial computer software MATLAB 7.0.1.
p (t )
If assuming that the system initially is at rest, the displacement can be stated as:
u (t ) =
p0 1
(sin ω p t − β sin ωt )
(
k 1− β 2 ) (7.1)
v(t ) =
p0 1
(ω p cosω p t − βω cosωt )
(
k 1− β 2 ) (7.2)
a (t ) =
p0 1
k (1 − β )
2
(
− ω p sin ω p t + βω 2 sin ωt
2
) (7.3)
k c
p (t )
[ ]
u (t ) = G e −ξ ⋅ω ⋅t [ A ⋅ sin ω d t + sin θ cos ω d t ] + sin (ω P t − θ ) (7.4)
ξω ω
A = sin θ − P cosθ
ωd ωd
(7.5)
p 1
G= 0
k (1 − β ) + (2ξβ )
2 2 2
[
u& (t ) = G e −ξωt [(ω d A − ξω sin θ ) cos ω d t − (ω d sin θ + ξωA)sin ω d t ]
+ω p cos(ω p t − θ ) ] (7.6)
[ [( ) (
u&&(t ) = G e −ξωt − ω d A + ξω sin θ sin ω d t − ω d sin θ + ω d ξωA cos ω d t −
2 2
)
( ) (
− ξωω d A − ξ 2ω 2 sin θ cos ω d t + ξωω d sin θ + ξ 2ω 2 A sin ω d t − ) ] (7.7)
−ω p sin (ω p t − θ ) ]
The complete derivation and solution, for both the undamped and damped case, can
be seen in Appendix C.
Obviously, the influence of the ratio between the frequency of the applied load and
the natural frequency of the system is of interest. The ratio is described by the circular
frequency and stated as β=ωp/ω. The influence on the system is examined for both the
undamped and damped system, and the value of β is varied between 0.1-10. The
main focus in this thesis is consistently on values of β <1, since it in reality is not
Mainly during the analyses the damping factor is set to 2.5 %, but to determine the
influence of the damping, the system is provided with different values of the damping
factor ξ during excitation by the harmonic load. The reason for the damping factor to
be chosen to be 2.5 % is based on damping factors normally used for bridges and the
formulations in Table 6.2.
7.2.4 Excitation
Two different load cases were examined by using the analytical solution.
p(t)
p
0
Load case 1 is the normal sinus function that oscillates around 0, see Figure 7.3.
p(t)
p
0
p(t)
2p0
Figure 7.5: Load case 2, the uplifted harmonic loading with an angular shift.
In order to prevent the instantaneous loading p0 that occur in Figure 7.4, an angular
shift is introduced to retrieve the load in load case 2, see Figure 7.5 and
Equation (7.10). Now the load do not starts so suddenly and therefore corresponds
more to a real train load.
−1
t + ∆t 2 1 1 t − ∆t 1 1
U = t P − K − 2 M tU − 2 M − C U 2 M + C (7.11)
∆t ∆t 2∆t ∆t 2∆t
C = 2 Mω ξ (7.12)
t
U& =
1
2∆t
(
− t − ∆t U + t + ∆t U ) (7.13)
The calculations have been made mainly for the damped case, but for calculations for
the undamped case the damping matrix C in Equation (7.12) is set to zero.
As for the analytical solution described in Section 7.2, the same parameters are of
interest and studied here. When damping is introduced to the system the damping
factor is set to 2.5 % in all cases.
Further it is studied what influence a change in the natural frequency fn has on the
system, when the value of the load frequencies fp1 and fp2 are kept constant.
7.3.3 Excitation
Load case 1 and 2, i.e. the two harmonic load cases that also are used in the analytical
analysis, are mainly used as a verification of the numerical solution. By comparing the
results from the numerical analysis with the analytical result, the numerical analysis
can be verified.
f t
p1
The third load consists of a continuous triangular load, see Figure 7.6. The triangle is
basically a simplification of the sinusoidal load in load case 2, see Figure 7.5, and has
the same total area. The triangular load is dependent of the two frequencies fp1 and fp2,
where the width of the load is described by the frequency fp2 and the distance between
the peaks by frequency fp1, see Figure 7.6. In this case where it is continuity in the
load, fp1 and fp2 will be equal, and here the continuity is based on a system that is not
unloaded at any time. The result from this load case is compared with load case 2,
because they have similar shape, and to elucidate if there are any differences even
though they have similar load shape and area.
p(t) f
p2
2p
0
f t
p1
The fourth load case consists of a rectangular load, and this load is an attempt to
describe the bogie axle load generated by the train, described in Section 6.1.2, with
only one impulse. The width of the load is dependent of the frequency fp2 and the
distance between the peaks by frequency fp1, see Figure 7.7. The area under the
continuous triangular load and the rectangular load is chosen to be equal.
f t
p1
1
0.8
0.6
0.4
0.2
0
0 10 20 30 40
Time [s]
Figure 7.9: Train load HSLM-A1 with β=0.1, i.e. loading generated by a train
velocity v=1.8 m/s. The figure displays the power car, end coach and
the two first intermediate coaches.
As can be seen in Table 6.1 there are different values on the load for the different
models. Earlier, to keep the system as simple as possible, the unit load 1 N is used. To
keep the simplicity and to be able to make comparison between different trainloads,
the loading needs to keep its mutual relationship. Therefore the tabled load values are
divided by 105, i.e. the load for A1 is PA1=1.7 N.
1.5
1
Load [N]
0.5
-0.5
-1
0 5 10 15 20 25
Time [s]
Figure 8.1: Different harmonic loads acting on the system in the analytical solution.
As can be seen in Figure 8.2 - Figure 8.3, the system response for an undamped
system that is loaded by a harmonic force, is increasing when β→1. When β=1
resonance effects appear in the system and the highest responses are achieved. If the
value of β >1 and increases, i.e. β→∞, the system response have the opposite effect
-5
-10
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.2: Displacements for load case 1 with β=0.1, β=0.5 and β=0.9.
200
100
0
-100
-200
-300
-400
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.3: Accelerations for load case 1 with β=0.1, β=0.5 and β=0.9.
Umax 1.0998 1.0825 1.4182 1.6249 1.7320 2.5000 3.3191 4.9237 9.9652
|Umin| 1.0998 1.0825 1.4182 1.6249 1.7320 2.5000 3.3191 4.9237 9.9652
Amax 4.3816 9.5574 16.871 25.970 36.001 59.218 91.767 155.63 354.09
|Amin| 4.3816 9.5574 16.871 25.970 36.001 59.218 91.767 155.63 354.09
As can be seen in Table 8.1 the maximum and minimum values are the same, this
because of that the load oscillates around its equilibrium at 0.
In Figure 8.4 - Figure 8.5 the results in form of displacements and accelerations from
an undamped system loaded by load case 2 are displayed. The difference in loading
between load case 1 and 2 is that load case 2 has no negative values and contains an
angular shift, see Figure 8.1. As mentioned earlier, the system response for an
undamped system that is loaded by a harmonic force, increases when β→1, and
resonance effects appear when β=1. This is therefore also obtained in this case. The
maximum and minimum values of displacement and acceleration for β <1, are listed
in Table 8.2.
10
Displacement [ m ]
-5
-10
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.4: Displacements for load case 2 with β=0.1, β=0.5 and β=0.9.
200
100
0
-100
-200
-300
-400
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.5: Acceleration for load case 2 with β=0.1, β=0.5 and β=0.9.
Table 8.2: Maximum and minimum values of displacement and acceleration for the
undamped case, load case 2.
Umax 2.0202 2.0000 2.1978 2.2671 2.6667 2.9732 3.9216 5.2840 10.526
Amax 0.7782 2.9866 7.6331 15.039 14.804 41.236 74.605 140.37 332.03
|Amin| 0.7975 2.9866 7.8089 13.787 26.319 41.236 75.860 132.00 336.61
9
Displacement [m]
0
0 0.2 0.4 0.6 0.8 1
β
Figure 8.6: Variation in maximum and minimum displacements with changing β for
the undamped load cases 1 and 2.
300
200
100
0
0 0.2 0.4 0.6 0.8 1
β
Figure 8.7: Variation in maximum and minimum accelerations with changing β for
the undamped load cases 1 and 2.
In the results, when comparing Table 8.1 - Table 8.2, it can be seen that generally by
assuming that the load do not contain any negative values and that the loading is
applied with a phase shift, the maximum and minimum values decrease, except for the
maximum displacements where it has the opposite effect, see Figure 8.6 - Figure 8.7.
However, the load in load case 2 is only applied in one direction so this is an expected
effect.
5
Displacement [m]
-5
-10
0 5 10 15
Time [s]
In practice this means that the oscillation frequency and amplitude of the system will
go towards and finally coincide with the frequency of the applied load. The system
response for the damped system follows the same pattern as for the undamped system,
i.e. that the response increases when β→1, and that resonance effects, with large
system responses, appear when β=1. As can be seen in Figure 3.10 it is when β ≈1
that the influence of the damping has largest effect. The main difference between the
damped and undamped systems can be seen by comparing Figure 8.9 with the
undamped system in Figure 8.2. For the first 5 seconds the response looks very
similar and it is first after this time the influence of the damping starts to get
noticeable. During the time while the homogeneous solution damps out, the system
response goes toward a stabile oscillation with the same frequency as the applied load.
The system response in terms of displacements and accelerations for a damped system
applied with load case 1 is displayed in Figure 8.9 - Figure 8.10.
4
2
0
-2
-4
-6
-8
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.9: Displacement for load case 1 with β=0.1, β=0.5 and β=0.9.
100
50
0
-50
-100
-150
-200
-250
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.10: Acceleration for load case 1 with β=0.1, β=0.5 and β=0.9.
The maximum and minimum values of displacement and acceleration for β<1, are
listed in Table 8.3.
Table 8.3: Maximum and minimum values of displacement, and acceleration for
the damped case, load case 1.
Umax 1.0651 1.0606 1.3805 1.5750 1.6741 2.0799 2.9804 4.1261 7.2849
|Umin| 1.0328 1.0411 1.2436 1.4000 1.6242 2.3296 2.7720 4.2535 7.2891
Amax 3.7666 7.4085 12.5670 17.943 30.965 52.373 74.839 128.92 249.25
|Amin| 3.7152 8.6098 15.3740 23.880 33.300 42.614 78.037 123.08 247.84
Applying the same procedure as for the undamped system, the load is changed and the
damped system is loaded by load case 2. As for the undamped system when the
applied load was changed from load case 1 to load case 2, the basic appearance of the
system response for the damped system is very similar to the previous case, see
Figure 8.11 - Figure 8.12.
4
2
0
-2
-4
-6
-8
-10
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.11: Displacement for load case 2 with β=0.1, β=0.5 and β=0.9.
100
50
0
-50
-100
-150
-200
-250
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.12: Acceleration for load case 2 with β =0.1, β=0.5 and β =0.9.
The maximum and minimum values of displacement, and acceleration for β <1, are
listed in Table 8.4.
Table 8.4: Maximum and minimum values of displacement and acceleration for the
damped case, load case 2.
Umax 2.0147 2.0415 2.1438 2.2497 2.6168 2.8994 3.4965 4.8165 8.0004
|Umin| 0.0101 0.0415 0.1360 0.3185 0.3317 0.8721 1.6684 2.9496 6.0658
Amax 0.7487 2.8720 6.0643 12.590 14.225 36.464 66.778 116.72 241.71
|Amin| 0.5805 2.5382 6.1131 12.643 24.384 38.352 60.581 114.12 239.42
0
0 0.2 0.4 0.6 0.8 1
β
Figure 8.13: Variation in maximum and minimum displacements with changing β for
the damped load cases 1 and 2.
200
Acceleration [m/s2]
150
100
50
0
0 0.2 0.4 0.6 0.8 1
β
Figure 8.14: Variation in maximum and minimum displacements with changing β for
the damped load cases 1 and 2.
The response of the system has also been studied for different values for the natural
frequency of the system. When the eigenfrequency increases from fn=1 Hz to fn=5 Hz
and fn=10 Hz, respectively, the displacement is not affected in any case and that is
because the relation between the eigenfrequency and load frequency remains the
same. This relationship between the eigenfrequency and displacement is what makes
the simplifications regarding mass, stiffness and applied loads justified. The results
40
ξ = 0.01
30 ξ = 0.02
ξ = 0.03
ξ = 0.04
DAF
20 ξ = 0.05
ξ = 0.06
ξ = 0.07
10
ξ = 0.08
ξ = 0.09
0
0 0.5 1 1.5 2 2.5 3
β
By applying that the focus of interest are on results for β <1 and that there seems to be
no drastic change of DAF until β >0.5, the axis in Figure 8.15 can be scaled down, see
Figure 8.16.
20
ξ = 0.06
15 ξ = 0.07
10 ξ = 0.08
ξ=0.09
ξ = 0.09
5
0
0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
β
Figure 8.16: Theoretical DAF for the analytical solution, scaled from Figure 8.15.
Both Figure 8.15 and Figure 8.16 are plotted based on the equation stated in
Section 3.3.2, i.e. Equation (3.43). According to Figure 8.15, considering a value of
β=0.9 and ξ=2.5 %, the total displacement would be ~5 times the static displacement
of the system. According to the results stated in Table 8.3, the largest value of
displacement is for |Umin|=7.2891ustatic. The difference between these two results can
be explained by the fact that according to the derivation of the DAF, the homogeneous
solution is assumed to die out and is therefore not included in the expression. So the
expression for DAF is in other words only valid when the damping has totally
excluded the influence of the homogeneous solution, see Figure 8.17
-5
Homogeneous solution damped out
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.17: Only up is considered in calculation of DAF.
When calculating DAF as stated above, the real maximum displacement is not
considered, which leads to an underestimation of the displacement. If the
displacement is plotted in the same way as DAF it can be displayed as in Figure 8.18.
25 ξ = 0.04
ξ = 0.05
20
ξ = 0.06
15 ξ = 0.07
|Umin|=7.2891ustatic ξ=0.09
10 ξ = 0.08
ξ = 0.09
5
0
0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
β
Remembering the |Umin| value stated above and comparing it in the same way as
earlier it can be seen in Figure 8.18 that now the result seems more reasonable, and
therefore this is the value that should be set as a reference value and not the
theoretical DAF.
To be able to make the comparison between the two solutions some modifications of
the analytical solution are made. In the analytical solution, the static displacement
ustatic, can be neglected or set to 1. In the numerical solution this is not possible, and
variables such as stiffness k, mass m and applied load p0 are needed in the expression,
see Equation (7.11). To make the expression as simple as possible, this is solved by
giving both the mass and applied load a value equal to 1, this is also described earlier
in Section 7.3.2. With these variables set to 1, the stiffness is adjusted so that the
system has an eigenfrequency of 1 Hz. To modify the results from the analytical
solution it is then needed to divide the results with the applied stiffness.
0.8
0.6
Load [N]
0.4
0.2
0
0 Time [s] 4.4444
Figure 8.19: Continuous triangular load compared with the sinusoidal load.
As can be seen in Figure 8.19 the sinusoidal and continuous triangular load reminds a
lot of each other, and the total impulse of each load pulse is the same for both cases.
When the system response for the two loads is compared it appears that the value of β
determines how well the result of the sinusoidal load and the continuous triangular
load coincides with each other.
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.20: Displacement for the continuous triangular load compared with the
sinusoidal load with β=0.9.
2
2
1
0
-1
-2
-3
-4
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.21: Acceleration for the continuous triangular load compared with the
sinusoidal load with β=0.9.
In Figure 8.20 - Figure 8.21, β=0.9 is displayed and it can be seen that the shape of
the principle response for the system excited by the continuous triangular load is
identical to the response with the sinusoidal load, apart from the magnitude of the
amplitude which is smaller for the continuous triangular load, for both displacements
and accelerations. The difference in amplitude can probably be explained by that the
sinusoidal load has higher impulse area at the peak. By examining every 1/10 of β,
this behaviour can be seen down to β≈0.4. At this ratio the acceleration starts to
become larger for the continuous triangular load than for the sinusoidal load, and
finally if the β value decreases even more, the total system response is larger for the
continuous triangular load.
0.02
0.015
0.01
0.005
0
-0.005
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.22: Displacement for the continuous triangular load compared with the
sinusoidal load with β=0.1.
As can be seen in Figure 8.22 the difference in displacements are very small even for
β=0.1, but when comparing the accelerations there can be seen a large difference, see
Figure 8.23.
0.02
0.01
0
-0.01
-0.02
-0.03
-0.04
-0.05
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.23: Acceleration for the continuous triangular load compared with the
sinusoidal load with β=0.1.
The analysis were performed for every 1/10 in the interval 0.1≤ β ≤0.9, and, if
considering only the continuous triangular load, the expected behaviour was that the
magnitude of the system response would have a constant decrease when β→0, but a
noticeable increase in magnitude started to appear at β=0.4, see Table 8.5 and
Figure 8.24 - Figure 8.25.
Umax 0.0256 0.0325 0.0313 0.0265 0.0255 0.0297 0.0364 0.0499 0.0832
|Umin| 0.0001 0.0071 0.0055 0.0000 0.0001 0.0043 0.0123 0.0258 0.0585
Amax 0.0437 0.3758 0.2887 0.2523 0.1651 0.2986 0.5882 1.0804 2.4258
|Amin| 0.0473 0.3762 0.2757 0.2621 0.1754 0.3230 0.5437 1.1079 2.4182
0.09
0.06
0.03
0
0 0.2 0.4 0.6 0.8 1
β
Figure 8.24: Variation in maximum and minimum displacements with changing β for
the damped load cases 2 and 3.
0
0 0.2 0.4 0.6 0.8 1
β
Figure 8.25: Variation in maximum and minimum accelerations with changing β for
the damped load cases 2 and 3.
0.02
0.01
0
-0.01
-0.02
-0.03
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.26: Displacement response for the continuous triangular and the sinusoidal
load with β=1/3.
0.5
-0.5
-1
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.27: Acceleration response for the continuous triangular and the sinusoidal
load with β=1/3.
The two figures above indicates that the response grows dramatically for β=1/3. If the
result from the response with β=1/3 is compared with, e.g. β=0.4, the difference is
almost a factor 2 for the displacement and at least a factor 3 for the acceleration, see
Figure 8.28 - Figure 8.29.
0.02
Displacement [m]
0.01
-0.01
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.28: Displacement for the continuous triangular load compared with the
sinusoidal load with β=0.4.
0.2
Acceleration [m/s2]
0.1
-0.1
-0.2
-0.3
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.29: Acceleration for the continuous triangular load compared with the
sinusoidal load with β=0.4.
0.1
Displacement [m]
-0.1
-0.2
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.30: Resonance behaviour for the continuous triangular load with β=1/3
and ξ=0.
The result is that when β=1/(2n+1), the response has a typical resonance behaviour,
i.e. the system response grows towards infinity, see Figure 8.30. This behaviour can
not be seen for any other β <1 except for β=1/(2n+1).
0.8
0.6
Load [N]
0.4
0.2
0
0 1 Time
2 [s] 3 4 4.4444
Figure 8.31: Rectangular load compared with the continuous triangular load.
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
-0.12
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.32: Displacement for the rectangular load compared with the continuous
triangular load with β=0.9.
2
1
0
-1
-2
-3
-4
-5
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.33: Acceleration for the rectangular load compared with the continuous
triangular load with β=0.9.
0.03
0.02
0.01
0
-0.01
-0.02
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.34: Displacement for the rectangular load compared with the continuous
triangular load with β=0.1.
0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
-1
0 5 10 15 20 25 30 35 40
Time [s]
Figure 8.35: Acceleration for the rectangular load compared with the continuous
triangular load with β=0.1.
For β=0.1 the rectangular load generates even more drastic oscillation than the
continuous triangular load. An explanation for this may be that the change in loading,
that becomes more dramatic when going from the sinusoidal load to the continuous
triangular and finally the rectangular, has bigger influence for lower velocities. Also
for the rectangular load the phenomena with β=1/(2n+1) occurs. The maximum and
minimum values of displacement and acceleration for β <1, are listed in Table 8.6.
Umax 0.0487 0.0972 0.0632 0.0487 0.0296 0.0487 0.0566 0.0776 0.1271
|Umin| 0.0160 0.0719 0.0359 0.0299 0.0043 0.0224 0.0350 0.0547 0.1036
Amax 0.1531 0.4700 0.2478 0.1812 0.07509 0.1531 0.2117 0.3422 0.6701
|Amin| 0.1415 0.4704 0.2353 0.1960 0.0797 0.1469 0.2290 0.3581 0.6775
0.12
0.08
0.04
0
0 0.2 0.4 0.6 0.8 1
β
Figure 8.36: Variation in maximum and minimum displacements with changing β for
the damped load cases 3 and 4.
4
Acceleration [m/s2]
0
0 0.2 0.4 0.6 0.8 1
β
Figure 8.37: Variation in maximum and minimum accelerations with changing β for
the damped load cases 3 and 4.
0.8
0.6
Load [N]
0.4
0.2
0
0 0.5 1 Time 1.5
[s] 2 2.5 2.73
In this part of the analysis it is of interest to find out what influence the frequency fp2,
i.e. the local frequency in the double triangular load, has on the total system response.
This load case represents the intermediate coaches in the train and to be able to make
any conclusions, the worst possible case is investigated, which is when β2= fp2/ fn=1.
The number of intermediate coaches used in the HSLM-A varies between 11-18. By
analysing different relationships between fp1 and fp2 it could be seen during how many
impulses the system response is still growing. The analysis shows that already at the
ratio fp1/fp2 ~6 the system response has stopped to grow within the minimum number
of intermediate coaches, i.e. 11 coaches. It should be remembered though that this
result is obtained with no concern taken to the power car or end coach.
0.05
Displacement [m]
-0.05
-0.1
0 5 10 15 20 25 30 35 40 45
Time [s]
As can be seen in Figure 8.39, the system response stabilizes and reaches maximum
value of the system response after only a few loading cycles when the ratio between
fp1 and fp2 is close to 6. Compare this to when the ratio between fp1/fp2 = 3, when the
maximum displacement is not reached until the number of 22 intermediate coaches
has passed, see Figure 8.40.
Double triangular
0.2
0.15
Displacement [m]
0.1
0.05
0
-0.05
-0.1
0 5 10 15 20 25 30 35 40 45
Time [s]
Figure 8.40: When fp1=3fp2, the number of intermediate coaches reaches 22 until the
maximum displacement is obtained, β2=1.0.
β2 β2 β2 β2
βC2 βC2
A1-β=0.2
0.1
0.08
0.06
Displacement [m]
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0 20 40 60 80 100 120
Time [s]
A1-β=0.9
0.04
0.03
Displacement [m]
0.02
0.01
0
-0.01
-0.02
-0.03
-0.04
0 5 10 15 20 25 30
Time [s]
1
0.5
0
-0.5
-1
-1.5
-2
0 5 10 15 20 25 30
Time [s]
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 0.9 1.8 2.7 2.97 3.6 4.5 5.4 6.3 7.2 8.1
βc1= 0.13 0.26 0.39 0.43 0.51 0.64 0.77 0.90 1.03 1.16
βc3= 0.51 1.02 1.53 1.70 2.04 2.55 3.06 3.57 4.09 4.60
In Appendix E.3 all the values of the five different β for the different train loads are
presented.
These variations in β makes it difficult to predict the structural responses. For
instance, it would be easy to consider that the most severe response would appear for
β≈1, but instead it appears already at β=0.2 for HSLM-A1. In Figure 8.42 it can be
seen that the maximum response takes place very early in the response history, and
according to Table 8.7 both βc2 and βc3 are close to 1. Even though no results are
presented, analysis in SDOF is made for all the train loads HSLM-A1-10, and the
analysis clearly shows that this variation of β has great influence. Generally, except
for HSLM-A1-2, the most dangerous ratio for each load case is β=0.5 and is
generated by the intermediate coaches. For HSLM-A1-2 the maximum response
occurred for β=0.2, where the response for HSLM-A2 was the total maximum of all
cases and generated by the power car and end coach. It can also be seen that
displacements and accelerations have the same decisive β-values, i.e. the maximum
occurs for the same value of β. It should still be remembered that in this case the
resonance velocity is based upon β=β1.
9.1.1 Geometry
The beam is chosen to have the length L=10 m and a cross-section with a width
b=1 m and a height h=1 m, see Figure 9.1.
z z
P P
x y
L
b
9.1.2 Material
The material is modelled as isotropic and linear elastic, where the strains and
displacements are assumed to be small. The mass and modulus of elasticity is
calculated so that the first natural frequency f1 of the beam is equal to one. First the
mass is calculated from Equation (4.6) to m=2.06 kg. Then inserting Equation (4.6)
into Equation (4.3), the modulus of elasticity can be calculated to E=10013 N/m2, see
Appendix F.1. Poisson’s ratio is set to υ=0.2, according to Swedish concrete codes.
The linear static equilibrium for the finite element model in ADINA is:
KU = P (9.1)
The static analysis is solved in ADINA by using a direct solution method, where the
stiffness matrix is assumed to be symmetric and positive definite. The equation solver
used in ADINA is a sparse matrix solver, which according to ADINA (2004) are one
or two orders of magnitude faster than other available solution methods based on
Gauss elimination. The solver is also very reliable and robust.
The natural frequencies and the mode shapes of vibration of the FE-model are
calculated in ADINA by solving the eigenvalue problem:
Kφ i = ω i2 Mφ i (9.2)
where K is the stiffness matrix corresponding to the time the solution start, M is the
mass matrix corresponding to the time of solution start, ωi and φi are the circular
frequency and mode shape, respectively, for mode i. Note that the frequencies are
extracted in the eigenvalue solution in numerically ascending sequence. The
eigenvectors are M-orthonormal, i.e. orthogonal and with a length equal to one:
φ iT Mφ i = 1 (9.3)
There are several solution methods to choose between when deciding how ADINA
should calculate the frequencies and mode shapes. The used method is the subspace
iteration method, where the starting subspace is generated by the Lanczos method, see
Bathe (1996). The subspace iteration method is default in ADINA.
The following procedures are available in ADINA for solution of the finite element
equations in a linear dynamic analysis, and they are described in Chapter 5. The
procedures are the central difference method, Newmark method and mode
superposition. The Newmark and mode superposition method are solved by the
trapezoidal rule.
9.1.6 Damping
The damping is chosen to be 2.5 %, i.e. the same as in the SDOF model. When using
the central difference method and Newmark method, the damping must be modelled
as Rayleigh damping, see Section 5.1.3. Using the mode superposition the modal
damping can be set to the value 2.5 % for each mode, see Section 5.2.
P P
L/2 L/2
b
P P
L/2 L/2
b
Figure 9.4: Beam loaded by a varying point load applied in the middle of the beam.
The FE-model is also loaded by the same train load that is modelled in SDOF, see
Section 7.3.4. The train load is applied in the same way as above, see Figure 9.4, but
here only the damped FE-models and SDOF systems are compared.
L
b
P v P P
2m
h
L
b
3m 11m 3m
h
L
b
Figure 9.7: Travelling power car load.
Model Displacement [mm] Error [‰] Moment [Nm] Reaction force [N]
Beam Theory 24.97 - 2.50 0.50
20 Beam Elements 24.97 0.001 2.50 0.50
800 Beam Elements 24.97 0.003 2.50 0.50
0.10
0.05
0.00
-0.05
-0.10
-0.15
-0.20
0 5 10 15 20
Time [s]
Figure 10.1: Displacement for an undamped beam loaded by a sinusoidal load with
β=0.9.
2
1
0
-1
-2
-3
-4
-5
0 5 10 15 20
Time [s]
Figure 10.2: Acceleration for an undamped beam loaded by a sinusoidal load with
β=0.9.
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
-0.12
0 5 10 15 20
Time [s]
Figure 10.3: Displacement for a damped beam loaded by a sinusoidal load with
β=0.9.
2
1
0
-1
-2
-3
-4
0 5 10 15 20
Time [s]
Figure 10.4: Acceleration for a damped beam loaded by a sinusoidal load with
β=0.9.
The results from the implicit integration method and the mode superposition method
are similar, since both methods use the Newmark trapezoidal rule. Further analyses
are only made with the mode superposition method, because it is easier to model the
damping with this method, than for the implicit time integration method. As
mentioned earlier, in the implicit method the damping must be modelled as Rayleigh
damping, while in the mode superposition method the damping can be introduced
directly to the eigenfrequency modes. In the problem modelled here, the results from
using Rayleigh damping are similar to those when using modal damping, but if the
analysed FE-model was more complex the results would start to differ. The reason is
that in the Rayleigh damping the damping ratio varies for different eigenmodes, but in
this case the amount of structural mass that is active in the first eigenmodes are high
and therefore the higher modes has less importance to the response, see Appendix F.3.
-0.005
-0.01
-0.015
-0.02
-0.025
-0.03
-0.035
0 5 10 15 20
Time [s]
Figure 10.5: Convergence analysis of the displacement for a damped beam loaded
by a sinusoidal load with β=0.5.
0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
0 5 10 15 20
Time [s]
Figure 10.6: Convergence analysis of the acceleration for a damped beam loaded by
a sinusoidal load with β=0.5.
For the undamped FE-model, the convergence analysis shows similar results
regarding the time step as for the damped FE-model, see Appendix F.6.
SDOF ADINA
0.1
0.08
0.06
Displacement [m]
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
-0.12
0 5 10 15 20
Time [s]
SDOF ADINA
5
4
3
Acceleration [m/s ]
2
2
1
0
-1
-2
-3
-4
-5
0 5 10 15 20
Time [s]
P (t )
t1 t2
Figure 10.9: The change of loading time for the SDOF system.
The change of the loading time from t1 to t2 in the SDOF system, results in a response
for the displacement and acceleration more similar to the response from the analysis
of the FE-model, see Figure 10.10 - Figure 10.11.
0.02
0
-0.02
-0.04
-0.06
-0.08
0 2 4 6 8 10
Time [s]
2
1
0
-1
-2
-3
0 2 4 6 8 10
Time [s]
The responses for β=0.9, shown in the figures above, are not only similar in
amplitude but also almost in phase, but in the lower β-values the load SDOF 2 has a
small phase angular difference compared to the results of the FE-model. The
responses of load SDOF 1 are out of phase by an angle of about 90˚-180˚ and, with
the exception of β=0.2, it generally has lower amplitudes on the response, see
Appendix F.8.
P (t )
2m
t, x
t1
The second modified load SDOF 2 has a loading history of a superposition of two
triangular loads. The loading time for the triangular loads is t2, which is the time for
the single travelling load to move the distance of 10 m along the beam. The second
triangular is delayed with the time t3 that corresponds to the spacing within the bogie
axles in the train model HSLM-A1, where the spacing is 2 m, see Figure 10.13.
P (t ) P (t )
t t
t2 t3 t2
t3
The change in loading from SDOF 1 to SDOF 2 for the system shows the same
phenomena as for the comparison of the single travelling load, see Section 10.3.1. The
results show that the response of the displacement and acceleration becomes more
similar to the response from the analysis of the FE-model, see
Figure 10.14 - Figure 10.15.
0.1
Displacement [m]
0.05
-0.05
-0.1
-0.15
0 2 4 6 8 10
Time [s]
Figure 10.14: Comparison of displacements between different load cases with β=0.9.
2.5
2
-2.5
-5
0 2 4 6 8 10
Time [s]
Figure 10.15: Comparison of accelerations between different load cases with β=0.9.
The responses for β=0.9, shown in the figures above, are not only similar in
amplitude but also almost in phase, but in the lower β-values the load SDOF 2 has a
small phase angular difference compared to the results of the FE-model. The
responses of load SDOF 1 are out of phase by an angle of about 90˚-180˚. Further, it
generally has lower amplitudes on the response, except for the velocity and the
acceleration, when β=0.2, and for the acceleration when β=0.5, see Appendix F.9.
P (t )
3m 11m 3m
t, x
t1
The second modified load, SDOF 2, has a load history of a superposition of four
triangular loads, where the triangular loads are combined in the same way as in
Figure 10.13. The loading time for the triangular loads is t2, which is the time for the
single travelling load to move the distance of 10 m along the beam. The second
triangular is delayed with the time t3 that corresponds to the spacing within the bogie
axle loads in HSLM-A1. The appearance time for the following combination of two
triangular loads is delayed by the time that corresponds to a spacing of 1 m, i.e. since
the length of the beam is 10 m, see Figure 10.17.
P (t )
t, x
1m
t3 t2
The change in loading from SDOF 1 to SDOF 2 for the system, has the same
phenomena as for the comparison of the single travelling load and the double
travelling load, see Section 10.3.1 - Section 10.3.2. The results show that the response
of the displacement and acceleration becomes more similar to the responses from the
analysis of the FE-model by using SDOF 2, see Figure 10.18 - Figure 10.19.
0.05
0
-0.05
-0.1
-0.15
-0.2
0 2 4 6 8 10
Time [s]
Figure 10.18: Comparison of displacements between different load cases with β=0.9.
4
2
2
0
-2
-4
-6
-8
0 2 4 6 8 10
Time [s]
Figure 10.19: Comparison of accelerations between different load cases with β=0.9.
The responses for β=0.9, shown in the figures above, are not only similar in
amplitude but also almost in phase, but in the lower β-values the load SDOF 2 has a
small phase angular difference compared to the results of the FE-model. The
responses of load SDOF 1 are out of phase by an angle of about 90˚-180˚. Further, it
generally has lower amplitudes on the response, except for the velocity and the
acceleration, when β=0.2, and for the acceleration when β=0.5, see Appendix F.10.
SDOF ADINA
0.08
0.04
Displacement [m]
-0.04
-0.08
-0.12
0 20 40 60 80 100 120
Time [s]
SDOF ADINA
4
3
Acceleration [m/s2]
2
1
0
-1
-2
-3
-4
0 20 40 60 80 100 120
Time [s]
SDOF ADINA
0.3
0.2
Displacement [m]
0.1
-0.1
-0.2
-0.3
0 10 20 30 40 50
Time [s]
SDOF ADINA
10
8
6
Acceleration [m/s ]
2
4
2
0
-2
-4
-6
-8
-10
0 10 20 30 40 50
Time [s]
In this thesis large focus is put on what influence the frequency ratio β=ωp/ωn has on
a SDOF system for different types of loading The calculations in this thesis are not
intended as real values, but are only intended to act as mutual comparison with each
other. The inputs have been chosen in order to simplify the comparison between
different load cases.
Different types of excitations have been studied in order to determine if some types
are more critical or dangerous than others. Three different types of load shapes were
used: sinusoidal, triangular and rectangular loads. By using different combinations a
total of 5 load cases are examined. When determining which type of load that is more
critical or dangerous the total impulse of the load pulse and loading frequency is kept
constant. The analysis showed that the rectangular load had the most critical response.
Why this is the most critical load is believed to depend on two reasons. First of all,
this load has the maximum magnitude during the entire impulse, compared with the
other examined loads that have maximum magnitude during only one ∆t. The second
reason is how the load is applied and unloaded. The rectangular load is applied very
suddenly and reaches maximum magnitude instantly and is also removed in the same
way, while the continuous triangular and sinusoidal load grows until maximum
magnitude during half the impulse time. The difference between the different load
types was velocity dependent, i.e. dependent of the ratio between the frequencies β.
During faster applied loads the response shapes were very similar, with only
difference in magnitude, while the slowly applied loads have large differences both in
shape and magnitude. This means that for fast applied loads, all load cases can be
reasonably well described by a harmonic excitation, considering the response shape.
As stated above the type of load has great influence on the response of the system, as
well as the ratio between circular frequencies. It can also be seen that a combination
by the two also seems to have great influence. For both the continuous triangular and
rectangular load it could be seen that when the frequency ratio β=1/(2n+1) the
response started to grow dramatically and when undamped even show resonance
behaviour. Exact reason for this phenomenon is not clearly investigated. This
phenomenon could not be seen during harmonic excitation so the type of loading
clearly has an influence, and especially how the load is applied and unloaded.
Double triangular loads were used to resemble the bogie axle loads generated by the
intermediate coaches. The sensitivity of the relation between the two frequencies fp1
and fp2, where fp1 and fp2 are the frequencies based upon the distances D and d in the
train load model HSLM-A, was investigated. It is shown that already when fp1~ 6fp2
the maximum displacements are obtained for less number of intermediate coaches
HSLM-A1 was considered as a varying point load, and applied and compared
between the SDOF system and FE-model. The response for displacements the results
coincide very well, while for the accelerations the response coincides in terms of
response shape but there is a large difference in magnitude. No clear explanation can
be given for why the large difference appears for the acceleration, but one reason may
be that the transformation factors not entirely capture the behaviour of the beam,
which causes this difference. When applying the travelling load to the FE-model and
comparing the results with the same point load as above, they can at first seem to have
no similarities at all. However, after normalizing the values from the SDOF system it
seems like the transformation from a deformable body into a SDOF system is a good
approximation when interested in the system response shape.
In an attempt to make the SDOF system able to resemble the travelling point loads in
a better way, the time that the load is applied were modified. It shows that by
experimenting with load time and shape, it is possible to describe the response from at
least up to four travelling point loads that is intended to resemble the power car in
HSLM-A.
Based on analysis of the analytical solution it has been shown that the, commonly
used Dynamic Amplification Factor DAF not are satisfying when looking at short
term dynamic loading, since it neglects the time dependence of the damping and
therefore will underestimate the maximum response.
Previously it could be seen that the results from modifying the SDOF system by
superposition of triangular loads gave a satisfactory result in terms to resemble the
moving power car. Therefore the entire HSLM-A needs to be modified to determine if
this captures the behaviour in a better way.
To verify this method real material properties and load magnitudes should be
introduced into the expressions and determine if the conclusion stated above still
withhold. The method should also be investigated for the behaviour if the geometrical
properties or boundary conditions change. In order to make the analysis method even
more complex the response of moments and section forces should be included.
ADINA (2004): Theory and Modelling Guide Version 8.2.2. ADINA R&D, Inc.,
Volume 1, Watertown, USA.
Banverket, (2006): BV BRO – Banverkets ändringar och tillägg till Vägverkets Bro
2004, (Revision and additions by the Swedish Railway Administration to the
Swedish Bridge code, Bro 2004. In Swedish), inklusive supplement nr 1, Utgåva 8,
BVS 583.10.
Bathe K.-J. (1996): Finite Element Procedures. Prentice Hall, Englewood Cliffs, New
Jersey, USA.
Geradin, M., Rixen, D. (1992): Mechanical vibrations. John Wiley & sons, Inc., No 2,
New York, USA.
Grahn, R., Jansson, P-Å. (2002): Mekanik – Statik och dynamik, (Mechanics – Static
and dynamic. In Swedish), Studentlitteratur, No 2, Lund, Sweden.
WEB SOURCES:
KTH (2006) Notes for lesson in Structural Dynamics at the Royal institute of
Technology, Stockholm. http://www2.mech.kth.se/~jean/dynamics/dc8.ppt#6.
The theory presented in Chapter 3 and systems represented by discrete models are
usually an idealized view. To be able to formulate the governing equations for a
continuous system, the theory of continuum mechanics have to be applied. Here the
equation of motion is expressed in terms of a displacement field:
The space variables x, y, z are continuous and therefore the system contains an infinity
of degree of freedom. Continuous systems may be considered as limiting cases of
discrete systems and therefore, the specific geometry of the continuous bodies allows
simplified formulations of the equation of motion, only expressed by one or two
displacement components, and themselves as functions of one or two space variables
and time.
for either longitudinal or transverse motion. The equation of motion for such type of
system can be obtained by Hamilton’s principle, or it is also possible to generalize the
formulation of the Lagrange equations to continuous one-dimensional systems.
δ ∫ Ldt = 0 (A.1)
t1
- The potential energy Wp is a function of the displacement v and its first and
second derivatives with respect to the space variable.
- The kinetic energy Wk is always a function of the velocity field v& , but may
also depend on the instantaneous configuration v, it is first derivative in space
and the associated velocity v& ′ .
The general dynamic behaviour in terms of Lagrangian density can be stated as:
Equation of motion:
∂L ∂ ∂L ∂ 2 ∂L ∂ ∂L ∂ ∂ ∂L
− + − + = 0 (A.7)
∂v ∂x ∂v ′ ∂x 2 ∂v ′′ ∂t ∂v& ∂t ∂x ∂v& ′
Boundary conditions:
∂L ∂ ∂L ∂ ∂L x = 0
−
∂v ′ ∂x ∂v ′′ − ∂v = 0 at (A.8)
∂t ∂v& ′ x = l
and
∂L x = 0
δv ′ = 0 at (A.9)
∂v ′′ x = l
z, w
p (x) V (l ) M (l )
or or
q( x) u (l ) φ (l )
x
M (0) V (0)
or or
φ (0) u (0)
Transverse vibrations of beams, see Figure A.1, are simplest described by using a
model with the kinematic assumptions that:
w = w( x ), v=0 (A.10)
3. The axial displacement components results from the rotation of the cross
section. The rotation is such that the cross section remains orthogonal to the
neutral axis.
∂w
u ( x, z ) = − z (A.11)
∂x
With the assumption of geometric linearity, the strain expressions can be written as:
∂u ∂2w
εx = = −z 2
∂x ∂x
∂w
εz = =0 (A.12)
∂z
1 ∂u ∂u
ε xz = + =0
2 ∂x ∂z
∂w
u = −z
z, w ∂x
x
∂w
∂x
where I ( x ) = ∫ z 2 dA
A
(A.13)
EI is the bending stiffness of the cross section.
∂2w
is the beam curvature.
∂x 2
l l
2 δw& 2
1
2 0 A( x )
& 2
& 2
(1
)
Wk = ∫ ∫ ρ u + w dAdx = ∫ ∫ ρ z + w& 2 dAdx =
2 0 A( x ) δx
2
(A.14)
l l
1 ∂w& 1
= ∫ ρI dx + ∫ ρAw 2 dx
2 0 ∂x 20
By introducing
Here the first term describes the kinetic energy for vertical translation, while the
second term describes the rotational kinetic energy of the cross section.
To be able to compute the potential energy of the external forces, imagine, as can be
seen in Figure A.1, that the beam is subjected to a distributed vertical load p(x,t) and
distributed moments q(x,t) per unit length. At the beam ends, either the shear loads
V or the bending moments M are applied or the displacements w or rotations Φ are
imposed.
The potential associated with a bending moment M, shown in Figure A.3, can be
computed as:
∂w ∂w
W p ,ext ,M = − ∫ uσ x dA = − ∫ − z σ x dA = − M (A.16)
A A
∂x ∂x
z, w
σx
x Mx
Now applying either Hamilton’s principle or Lagrange equation with w as the only
independent function, the equation of motion can be stated as:
∂ 2 ∂w && ∂ 2 ∂2w ∂q
&& −
mw mr + EI 2 = p − (A.18)
∂x ∂x ∂x 2 ∂x ∂x
∂w ∂ 2 w
w=w or V = mr 2 EI 2 − q = V (A.19)
∂x ∂x
- on the rotation
Where M and T are, respectively, the bending moments and shear loads.
The equation for free vibration of the beam is obtained from Equation (A.18) by the
assumption of a harmonic motion w(x,t) = w(x)sinωt
d2 d 2w d 2 dw
EI − ω 2 mw + ω 2
2 mr =0 (A.21)
dx 2 dx dx dx
The kinetic assumption of no shear deformation remains valid provided that the ratio
I/A=r2 remains small. It is thus consistent in this case to neglect the rotary inertia of
the cross sections, so that the free vibration equation of the beam becomes:
d2 d 2w
EI − ω 2 mw = 0
2 (A.22)
dx 2 dx
- on the displacement
d d 2w
w=0 or V = EI 2 (A.23)
dx dx
- on the rotation
dw d 2w
=0 or M = EI =0 (A.24)
dx dx 2
When the bending stiffness EI and the mass per unit length m remain constant over
the beam length, the eigenvalue problem can be stated as:
d 4w m
4
−ω 2 w=0 (A.25)
dx EI
To determine the beam response versus time the eigenmodes for a simply supported
beam can be computed, by verifying the eigenvalue equation with the homogeneous
boundary conditions.
d 4u
EI 4
− ω 2 mu = 0 for 0 < x <l
dx
(A.26)
2
d u
u= = 0 at x = 0, l
dx 2
2 nπx
u ( n ) ( x) = sin where n = 1, 2, 3,….. (A.27)
ml l
EI
ω (n ) 2 = (nπ )4 (A.28)
ml 4
In Figure 4.1, the three first eigenmodes for a simply supported beam are shown. The
first eigenmode corresponds to the lowest eigenfrequency.
Figure A.4: The three first eigenmodes for a simply supported beam.
Normally, when a beam is subjected to a dynamic load, the load frequency will not
coincide with the eigenfrequencies and therefore the resulting shape of deformation
will not be the same as any of the eigenmodes. However, the dominating shape of
deformation is usually the first eigenmode but it is influenced by higher modes. SDOF
systems have only one eigenmode and hence there are no influences from higher
modes.
A: Initial calculations:
δ ≥ 0.50 α ≥ 0.25(0.5 + δ )2
1 δ 1 1
a0 = a1 = a2 = a3 = −1
α∆t 2 α∆t α∆t 2α
δ ∆t δ
a4 = −1 a5 = − 2 a 6 = ∆t (1 − δ ) a 7 = δ∆t
α 2 α
Kˆ = K + a 0 M + a1C
Kˆ t + ∆tU = t + ∆t Pˆ
t + ∆t
U& = t U& + a 6 U&& + a 7 U&&
t + ∆t t
A: Initial calculations:
∆t 0 && 2
Calculate − ∆t
U = U − ∆t U& +
0
U 0
1 1
Mˆ = 2 M + C
∆t 2∆t
2 1 1
t
Pˆ = t P − K − 2 M tU − 2 M − C t − ∆tU
∆t ∆t 2∆t 2
Mˆ t + ∆tU = t Pˆ
t 1
U&& = 2 ( t − ∆t
U − 2 t U + t + ∆t U )
∆t
t
U& =
1
2∆t
(
− t − ∆t U + t + ∆t U )
mu&& + cu& + ku = 0 1)
Assume:
u = Ce λt
u&& = λCe λt
u&& = λ2Ce λt ⇒
Insert into 1)
c k k
λ2 + λ + = 0, ω=
m m m
c
λ2 + λ + ω2 = 0
m
2
−c c
λ= ± −ω
2
⇒
2m 2 m
−c c
2 c
λ =ω ± − 1 , ξ=
2mω 2mω 2mω
( )
λ = ω − ξ ± ξ 2 − 1 = −ξω ± iω 1 − ξ 2
ωd = ω 1 − ξ 2
( )
u (t ) = e −ξωt C1eiω d t + C2e −iω d t = e −ξωt ( A sin ωd t + B cos ωd t ) = uh
u p = C1 sin ω pt + C2 cos ω pt
u& p = ω pC1 cos ω pt − ω pC2 sin ω pt
u&&p = −ω p C1 sin ω pt − ω p C2 cos ω pt
2 2
Insert into 1)
− mω p C1 sin ω pt − mω p C2 cos ω pt +
2 2
p0 1− β 2
C1 =
k 1 − β 2 + (2ξβ )2
(
2
)
⇒
p0 2ξβ
C2 = −
( )
k 1 − β 2 2 + (2ξβ )2
up =
p0 1
[( )
1 − β 2 sin ω pt − 2ξβ cos ω pt ]
(2 2
)
k 1 − β + (2ξβ )2
u p = R sin (ω pt − θ )
where
R= (C 2
)
+ C 22 =
p0 1
(1 − β ) + (2ξβ )
1
k 2 2 2
C2 2ξβ
θ = arctan − = arctan for β < 1
1 − β
2
C1
C2 2ξβ
θ = π + arctan − = π + arctan for β > 1
1 − β
2
C1
u0 = B + R sin (− θ ) ⇒
B = u0 + R sin(−θ )
Set u0 = u&0 = 0
ξω ω
A= R sin (θ ) − R p cos(θ ), B = R sin (θ ) ⇒
ωd ωd
ξω ωp
u (t ) = R e −ξωt sin θ cos ω d t + sin θ sin ω d t − cosθ sin ω d t + sin (ω p t − θ )
ωd ωd
ωp ξω
u& (t ) = R e −ξωt ξω cosθ sin ω d t − sin θ sin ω d t − sin θ cos ω d t −
ω d ωd
− ω d sin θ sin ω d t − ξω sin θ cos ω d t + ω p cosθ cos ω d t + ω p cos(ω pt − θ )
ξω ωp
u&&(t ) = R e −ξωt ξ 2ω 2 sin θ cos ω d t + sin θ sin ω d t − cosθ sin ω d t +
ωd ωd
+ 2ξω ω d sin θ sin ω d t − ξω sin θ cos ω d t + ω p cosθ cos ω d t −
− ω d sin θ cos ω d t − ξωω d sin θ sin ω d t − ω pω d cosθ sin ω d t − ω p sin (ω pt − θ )
2 2
u p = C1 sin ω pt + C2 cos ω pt
Insert into 2)
− mω p C1 sin ω pt − mω p C2 cos ω pt +
2 2
p0 2ξβ
C1 = −
( )
k 1 − β + (2ξβ )
2 2 2
⇒
p0 1− β 2
C2 = −
( )
k 1 − β 2 2 + (2ξβ )2
up = −
p0 1
[ ( )
2ξβ sin ω pt + 1 − β 2 cos ω pt ]
( 2
)
k 1 − β + (2ξβ )
2 2
u p = R cos(ω pt − θ )
R= (C 2
)
+ C22 =
p0 1
=
p0
G
( )
1
k 1 − β + (2ξβ )
2 2 k
2
14442444 3
G
C1 2ξβ
θ = arctan = arctan
2
for β > 1
C2 1− β
C1 2ξβ
θ = π + arctan = π + arctan
2
for β < 1
C2 1− β
Set u (t ) = uh + u p ⇒
mu&& + cu& + ku = p0
c k
u&& + u& + u = p0 ⇒
m m
p
up = 0
k
G cos(ω pt − θ ) + 0
p0 p
u (t ) = e −ξωt ( A sin ω d t + B cos ω d t ) +
k k
Gω p sin (ω pt − θ )
p0
−
k
p0 p
u0 = B + G cos(− θ ) + 0 ⇒
k k
p p
B = u0 − 0 G cosθ − 0
k k
Set u0 = u&0 = 0
p0 ξω ω p0
A=− (G cosθ + 1) + G p sin θ , B=− (G cosθ + 1)
k ω p ωd k
Then set
ξω ω
Acos = (G cosθ + 1) + G p sin θ , Bcos = (G cosθ + 1) ⇒
ω p ωd
u (t ) = [
e (− Acos sin ω d t − Bcos cos ω d t ) + G cos(ω pt − θ ) + 1
p0 −ξωt
k
]
u& (t ) =
k
[
p0 −ξωt
e (− ω d Acos cos ω d t + ω d Bcos sin ω d t +ξω Acos sin ω d t + ξωBcos cos ω d t ) −
−Gω p si n(ω pt − θ ) ]
u&&(t ) =
k
[ [(
p0 ξωt
)
e ω d Acos − 2ξωω d Bcos − ξ 2ω 2 Acos sin ω d t +
2
( ) ]
+ ω d Bcos + 2ξωω d Acos − ξ 2ω 2 cos cos ω d t − Gω p cos(ω pt − θ )
2
]
p(t)
Load Case 1
p
0
Harmonic load
t p(t ) = p 0 sin ω P t
p(t)
2p0
Load Case 2
Harmonic load
p(t ) = p 0 − p 0 cos ω P t
p(t) f
2p
p2 Load Case 3
0
Continuous triangular
f t
p1
p(t) f
2p
p2 Load Case 4
0
Rectangular
f t
p1
p(t) f
p2 Load Case 5
2p
0
Double triangular
f t
p1
Mean Error
Mean Error
β2 β2 β2 β2
βC1 βC3 β1
βC2 βC2
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 0.9 1.8 2.7 2.97 3.6 4.5 5.4 6.3 7.2 8.1
βc1= 0.13 0.26 0.39 0.43 0.51 0.64 0.77 0.90 1.03 1.16
βc3= 0.51 1.02 1.53 1.70 2.04 2.55 3.06 3.57 4.09 4.60
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 0.54 1.09 1.63 1.81 2.17 2.71 3.26 3.80 4.34 4.89
βc1= 0.14 0.27 0.41 0.45 0.54 0.68 0.81 0.95 1.09 1.22
βc2= 0.63 1.27 1.90 2.11 2.53 3.17 3.80 4.43 5.07 5.70
βc3= 0.54 1.08 1.62 1.80 2.16 2.70 3.23 3.77 4.31 4.85
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 1.00 2.00 3.00 3.33 4.00 5.00 6.00 7.00 8.00 9.00
βc1= 0.14 0.29 0.43 0.48 0.57 0.71 0.86 1.00 1.14 1.29
βc2= 0.67 1.33 2.00 2.22 2.67 3.33 4.00 4.67 5.33 6.00
βc3= 0.57 1.13 1.70 1.89 2.27 2.84 3.40 3.97 4.54 5.11
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 0.70 1.40 2.10 2.33 2.80 3.50 4.20 4.90 5.60 6.30
βc1= 0.15 0.30 0.45 0.50 0.60 0.75 0.90 1.05 1.20 1.35
βc2= 0.70 1.40 2.10 2.33 2.80 3.50 4.20 4.90 5.60 6.30
βc3= 0.60 1.19 1.79 1.99 2.38 2.98 3.57 4.17 4.77 5.36
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 1.10 2.20 3.30 3.67 4.40 5.50 6.60 7.70 8.80 9.90
βc1= 0.16 0.31 0.47 0.52 0.63 0.79 0.94 1.10 1.26 1.41
βc2= 0.73 1.47 2.20 2.44 2.93 3.67 4.40 5.13 5.87 6.60
βc3= 0.62 1.25 1.87 2.08 2.50 3.12 3.74 4.37 4.99 5.62
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 1.15 2.30 3.45 3.83 4.60 5.75 6.90 8.05 9.20 10.35
βc1= 0.16 0.33 0.49 0.55 0.66 0.82 0.99 1.15 1.31 1.48
βc2= 0.77 1.53 2.30 2.56 3.07 3.83 4.60 5.37 6.13 6.90
βc3= 0.65 1.30 1.96 2.17 2.61 3.26 3.91 4.57 5.22 5.87
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 1.20 2.40 3.60 4.00 4.80 6.00 7.20 8.40 9.60 10.80
βc1= 0.17 0.34 0.51 0.57 0.69 0.86 1.03 1.20 1.37 1.54
βc2= 0.80 1.60 2.40 2.67 3.20 4.00 4.80 5.60 6.40 7.20
βc3= 0.68 1.36 2.04 2.27 2.72 3.40 4.09 4.77 5.45 6.13
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 1.00 2.00 3.00 3.33 4.00 5.00 6.00 7.00 8.00 9.00
βc1= 0.18 0.36 0.54 0.60 0.71 0.89 1.07 1.25 1.43 1.61
βc2= 0.83 1.67 2.50 2.78 3.33 4.17 5.00 5.83 6.67 7.50
βc3= 0.71 1.42 2.13 2.36 2.84 3.55 4.26 4.96 5.67 6.38
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 1.30 2.60 3.90 4.33 5.20 6.50 7.80 9.10 10.40 11.70
βc1= 0.19 0.37 0.56 0.62 0.74 0.93 1.11 1.30 1.49 1.67
βc2= 0.87 1.73 2.60 2.89 3.47 4.33 5.20 6.07 6.93 7.80
βc3= 0.74 1.48 2.21 2.46 2.95 3.69 4.43 5.16 5.90 6.64
β1= 0.1 0.2 0.3 0.33 0.4 0.5 0.6 0.7 0.8 0.9
β2= 1.35 2.70 4.05 4.50 5.40 6.75 8.10 9.45 10.80 12.15
βc1= 0.19 0.39 0.58 0.64 0.77 0.96 1.16 1.35 1.54 1.74
βc2= 0.90 1.80 2.70 3.00 3.60 4.50 5.40 6.30 7.20 8.10
βc3= 0.77 1.53 2.30 2.55 3.06 3.83 4.60 5.36 6.13 6.89
SDOF
0.1
0.08 β=0.2
0.06
Displacement [m]
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
0 20 40 60 80 100 120
Time [s]
SDOF
0.1
0.08 β=0.5
0.06
Displacement [m]
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
0 10 20 30 40 50
Time [s]
SDOF
0.04
0.03 β=0.9
0.02
Displacement [m]
0.01
0
-0.01
-0.02
-0.03
-0.04
0 5 10 15 20 25 30
Time [s]
SDOF
0.6
0.4
β=0.2
Velocity [m/s]
0.2
-0.2
-0.4
-0.6
0 20 40 60 80 100 120
Time [s]
SDOF
0.6
0.4 β=0.5
Velocity [m/s]
0.2
-0.2
-0.4
-0.6
0 10 20 30 40 50
Time [s]
SDOF
0.25
0.2 β=0.9
0.15
0.1
Velocity [m/s]
0.05
0
-0.05
-0.1
-0.15
-0.2
-0.25
0 5 10 15 20 25 30
Time [s]
SDOF
4
3 β=0.2
Acceleration [m/s2]
2
1
0
-1
-2
-3
-4
0 20 40 60 80 100 120
Time [s]
SDOF
4
3 β=0.5
Acceleration [m/s2]
2
1
0
-1
-2
-3
-4
0 10 20 30 40 50
Time [s]
SDOF
1.5
1 β=0.9
Acceleration [m/s2]
0.5
0
-0.5
-1
-1.5
-2
0 5 10 15 20 25 30
Time [s]
P P
L/2 L/2
b
The mass of the SDOF system is transformed into a mass of the beam by using
Equation (4.31):
me 1
m= = = 2.06 kg (F.1)
κM 0.4857
bh 3
I= = 0.083m 4 (F.2)
12
The modulus of elasticity E is calculated from the relation between the circular
eigenfrequencies ωn and the chosen natural frequency fn.
EI
ω n = (nπ )2
mL4
ω n2 mL4
E= = 10013.21N/m 2
(nπ ) I
4
P
R1 = R2 = = 0.50 N (F.4)
2
PL
M L/2 = = 2.50 Nm (F.5)
4
PL3
u= = 24.97 mm (F.6)
48EI
f1,800=1.00 Hz f2,800=4.00 Hz
f3,800=9.00 Hz f4,800=16.00 Hz
f5,800=25.00 Hz f6,800=36.00 Hz
f7,800=49.00 Hz f8,800=64.00 Hz
f9,800=81.00 Hz f10,800=100.00 Hz
f11,800=121.00 Hz f12,800=144.00 Hz
f13,800=169.00 Hz f14,800=196.00 Hz
f15,800=225.00 Hz f16,800=256.00 Hz
f17,800=289.00 Hz f18,800=324.00 Hz
f19, 20 =277.36 Hz
f19,800=361.00 Hz
20 Elements 800Elements
Mode Freq. fn [Hz] Mass [%] Acc. mass [%] Freq. fn [Hz] Mass [%] Acc. mass [%]
-0.005
-0.010
-0.015
-0.020
-0.025
-0.030
0 5 10 15 20
Time [s]
:
-0.005
-0.010
-0.015
-0.020
-0.025
-0.030
-0.035
0 5 10 15 20
Time [s]
0.10
0.05
0.00
-0.05
-0.10
-0.15
-0.20
0 5 10 15 20
Time [s]
0.002
0.000
-0.002
-0.004
-0.006
-0.008
-0.010
0 5 10 15 20
Time [s]
0.02
0.00
-0.02
-0.04
-0.06
-0.08
0 5 10 15 20
Time [s]
0.2
0.0
-0.2
-0.4
-0.6
-0.8
0 5 10 15 20
Time [s]
0.004
0.002
0.000
-0.002
-0.004
-0.006
-0.008
-0.010
0 5 10 15 20
Time [s]
0.15
0.05
-0.05
-0.15
-0.25
-0.35
0 5 10 15 20
Time [s]
2
1
0
-1
-2
-3
-4
-5
0 5 10 15 20
Time [s]
-0.005
-0.01
-0.015
-0.02
-0.025
-0.03
0 5 10 15 20
Time [s]
-0.01
-0.015
-0.02
-0.025
-0.03
-0.035
0 5 10 15 20
Time [s]
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
-0.12
0 5 10 15 20
Time [s]
0.002
0
-0.002
-0.004
-0.006
-0.008
-0.01
0 5 10 15 20
Time [s]
0.02
0
-0.02
-0.04
-0.06
-0.08
0 5 10 15 20
Time [s]
0.2
0.0
-0.2
-0.4
-0.6
0 5 10 15 20
Time [s]
0.004
0.002
0
-0.002
-0.004
-0.006
-0.008
-0.01
0 5 10 15 20
Time [s]
0.2
0.1
0
-0.1
-0.2
-0.3
0 5 10 15 20
Time [s]
2
1
0
-1
-2
-3
-4
0 5 10 15 20
Time [s]
0
-0.005
-0.01
-0.015
-0.02
-0.025
-0.03
-0.035
-0.04
0 5 10 15 20
Time [s]
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
0 5 10 15 20
Time [s]
0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
0 5 10 15 20
Time [s]
-0.005
-0.01
-0.015
-0.02
-0.025
-0.03
-0.035
0 5 10 15 20
Time [s]
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0 5 10 15 20
Time [s]
0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
0 5 10 15 20
Time [s]
SDOF ADINA
0,1
0,08 β=0.2
0,06
Displacement [m]
0,04
0,02
0
-0,02
-0,04
-0,06
-0,08
-0,1
-0,12
0 20 40 60 80 100 120
Time [s]
SDOF ADINA
0,6
0,4 β=0.2
0,2
Velocity [m/s]
0
-0,2
-0,4
-0,6
-0,8
0 20 40 60 80 100 120
Time [s]
SDOF ADINA
5
4 β=0.2
3
Acceleration [m/s2]
2
1
0
-1
-2
-3
-4
-5
0 20 40 60 80 100 120
Time [s]
SDOF ADINA
0.05
0.04 β=0.4
0.03
Displacement [m]
0.02
0.01
0
-0.01
-0.02
-0.03
-0.04
-0.05
0 10 20 30 40 50 60
Time [s]
SDOF ADINA
0.3
β=0.4
0.2
Velocity [m/s]
0.1
-0.1
-0.2
-0.3
0 10 20 30 40 50 60
Time [s]
SDOF ADINA
3
β=0.4
2
Acceleration [m/s2]
-1
-2
-3
0 10 20 30 40 50 60
Time [s]
SDOF ADINA
0,1
0,08 β=0.5
0,06
Displacement [m]
0,04
0,02
0
-0,02
-0,04
-0,06
-0,08
-0,1
0 10 20 30 40 50
Time [s]
SDOF ADINA
0,6
β=0.5
0,4
Velocity [m/s]
0,2
-0,2
-0,4
-0,6
0 10 20 30 40 50
Time [s]
SDOF ADINA
8
6 β=0.5
Acceleration [m/s ]
2
4
2
0
-2
-4
-6
0 10 20 30 40 50
Time [s]
SDOF ADINA
0.02
β=0.7
0.01
Displacement [m]
-0.01
-0.02
-0.03
0 10 20 30 40
Time [s]
SDOF ADINA
0.2
0.15 β=0.7
0.1
Velocity [m/s]
0.05
0
-0.05
-0.1
-0.15
-0.2
0 10 20 30 40
Time [s]
SDOF ADINA
5
4 β=0.7
3
Acceleration [m/s2]
2
1
0
-1
-2
-3
-4
0 10 20 30 40
Time [s]
SDOF ADINA
0,04
0,03 β=0.9
0,02
Displacement [m]
0,01
0
-0,01
-0,02
-0,03
-0,04
0 5 10 15 20 25 30
Time [s]
SDOF ADINA
0,3
β=0.9
0,2
Velocity [m/s]
0,1
-0,1
-0,2
-0,3
0 5 10 15 20 25 30
Time [s]
SDOF ADINA
6
β=0.9
4
Acceleration [m/s ]
2
-2
-4
-6
0 5 10 15 20 25 30
Time [s]
0.01
0
-0.01
-0.02
-0.03
-0.04
-0.05
0 2 4 6 8 10
Time [s]
0.08
0.04
0
-0.04
-0.08
-0.12
-0.16
-0.2
0 2 4 6 8 10
Time [s]
0.5
0
-0.5
-1
-1.5
-2
0 2 4 6 8 10
Time [s]
0.02
0
-0.02
-0.04
-0.06
-0.08
0 2 4 6 8 10
Time [s]
0.1
-0.1
-0.2
-0.3
0 2 4 6 8 10
Time [s]
1.5
1
0.5
0
-0.5
-1
-1.5
-2
0 2 4 6 8 10
Time [s]
0.02
0
-0.02
-0.04
-0.06
-0.08
0 2 4 6 8 10
Time [s]
0.2
-0.2
-0.4
0 2 4 6 8 10
Time [s]
2
1
0
-1
-2
-3
0 2 4 6 8 10
Time [s]
-0.02
-0.04
-0.06
-0.08
-0.1
0 2 4 6 8 10
Time [s]
0.05
0
-0.05
-0.1
-0.15
-0.2
0 2 4 6 8 10
Time [s]
0.5
0
-0.5
-1
-1.5
-2
-2.5
0 2 4 6 8 10
Time [s]
-0.05
-0.1
-0.15
0 2 4 6 8 10
Time [s]
0.2
-0.2
-0.4
-0.6
0 2 4 6 8 10
Time [s]
2
1
0
-1
-2
-3
0 2 4 6 8 10
Time [s]
0.05
-0.05
-0.1
-0.15
0 2 4 6 8 10
Time [s]
0.2
0
-0.2
-0.4
-0.6
-0.8
0 2 4 6 8 10
Time [s]
0
0 2 4 6 8 10
-2.5
-5
Time [s]
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
0 2 4 6 8 10
Time [s]
0.2
-0.2
-0.4
-0.6
0 2 4 6 8 10
Time [s]
2
2
1
0
-1
-2
-3
-4
0 2 4 6 8 10
Time [s]
0
-0.02
-0.04
-0.06
-0.08
-0.1
-0.12
0 2 4 6 8 10
Time [s]
0.1
0
-0.1
-0.2
-0.3
-0.4
0 2 4 6 8 10
Time [s]
1
0
-1
-2
-3
-4
0 2 4 6 8 10
Time [s]
0.05
0
-0.05
-0.1
-0.15
-0.2
0 2 4 6 8 10
Time [s]
0.5
-0.5
-1
-1.5
0 2 4 6 8 10
Time [s]
4
2
2
0
-2
-4
-6
-8
0 2 4 6 8 10
Time [s]
SDOF ADINA
0.08
β=0.2
0.04
Displacement [m]
-0.04
-0.08
-0.12
0 20 40 60 80 100 120
Time [s]
SDOF ADINA
0.6
0.4 β=0.2
0.2
Velocity [m/s]
0
-0.2
-0.4
-0.6
-0.8
0 20 40 60 80 100 120
Time [s]
SDOF ADINA
4
3 β=0.2
Acceleration [m/s2]
2
1
0
-1
-2
-3
-4
0 20 40 60 80 100 120
Time [s]
SDOF ADINA
0,3
β=0.5
0,2
Displacement [m]
0,1
-0,1
-0,2
-0,3
0 10 20 30 40 50
Time [s]
SDOF ADINA
1,6
1,2 β=0.5
0,8
Velocity [m/s]
0,4
0
-0,4
-0,8
-1,2
-1,6
0 10 20 30 40 50
Time [s]
SDOF ADINA
10
8 β=0.5
6
Acceleration [m/s2]
4
2
0
-2
-4
-6
-8
-10
0 10 20 30 40 50
Time [s]
SDOF ADINA
0,3
0,2 β=0.9
Displacement [m]
0,1
0
-0,1
-0,2
-0,3
-0,4
0 5 10 15 20 25 30
Time [s]
SDOF ADINA
2
1,5 β=0.9
1
Velocity [m/s]
0,5
0
-0,5
-1
-1,5
-2
0 5 10 15 20 25 30
Time[s]
SDOF ADINA
12
β=0.9
8
Acceleration [m/s2]
-4
-8
-12
0 5 10 15 20 25 30
Time [s]