Vous êtes sur la page 1sur 195

DYNAMICS IN THE

PRACTICE OF
STRUCTURAL DESIGN

0. Sircovich Saar

WITPRESS Southampton, Boston


DYNAMICS IN
THE PRACTICE
OF STRUCTURAL DESIGN

0. Sircovich Saar -·''

Published by

WIT Press -,._~:-.,~-, _


Ashurst Lodge, Ashurst, ·southampton, S040 7AA, UK
Tel: 44 (0) 238 029 3223; Fax: 44 (0) 238 029 2853
E-Mail: witpress@witpress.com
http://www.witpress.com

For USA, Canada and Mexico

WIT Press
25 Bridge Street, Billerica, MA 01821, USA
Tel: 978 667 5841; Fax: 978 667 7582
E-Mail: infousa@witpress.com
http://www.witpress.com

British Library Cataloguing-in-Publication Data

A Catalogue record for this book is available


from the British Library

ISBN: 1-84564-161-2

Library of Congress Catalog Card Number: 2005928180

No responsibility is assumed by the Publisher, the Editors and Authors for any injury
and/or damage to persons or property as a matter of products liability, negligence or
otherwise, or from any use or operation of any methods, products, instructions or ideas
contained in the material herein.

© WIT Press 2006

Printed in Great Britain by Athenaeum Press Ltd., Gateshead

All rights reserved. No part of this publication may be reproduced, stored in a retrieval
system, or transmitted in any fonn or by any means, electronic, mechanical, photocopying,
recording, or otherwise, without the prior written pennission of the Publisher.
Contents

Preface ix

Acknowledgements xi

List of symbols xv

Chapter 1 Structural design in dynamic situations ..•................. 1

PART! Basics ...........•...........•............•.............. 3

Chapter 2 General overview ...........•........•.................... 5


I The dynamic response of structures ......................... 5
2 Dynamic equilibrium of forces ............................. 14
3 Energy in structural vibrations ............................. 20
4 Linear and nonlinear response of structures .................. 24
5 Vibrations in existing structures ............................ 28
6 Working with the computer ............................... 29
7 Superposing static and dynamic effects ..................... 30

PART2 Dynamics in design .......••......•....•................. 33

Chapter 3 Dynamic events and effects ..•............................ 35


I Introduction ............................................. 3 5
2 Dynamic events and effects ............................... 35
3 Response of structures .................................... 3 7

Chapter4 Dynamic loads on structures ...............•............. 39


I General description ....................................... 39
2 Working machines ....................................... 41
3 Vehicle loads ............................................ 44
4 Human activities ......................................... 46
5 Construction loads ....................................... 49
Chapter 5 Vibrations .............................................. 51
I Introduction ............................................. 51
2 Vibrations of structures ................................... 51
3 Sources of vibrations ..................................... 52
4 Vibrations in the design process ............................ 53

PART3 Structural materials in dynamic situations ...........•.... 57

Chapter 6 The rate of strain ....................................•.. 59


I Introduction ............................................. 59
2 The rate ofloadiug ....................................... 60
3 The rate of strain ......................................... 60
4 Structural materials ....................................... 62
5 Structural elements ....................................... 64

Chapter 7 Fatigue ....................................•.....•...... 67


I The phenomenon ......................................... 67
2 Fatigue in structural materials ............................. 70
3 Fatigue in structural elements .............................. 72

Chapter 8 Damping ........................•...................... 77


I Introduction ............................................. 77
2 Viscous dampiug ......................................... 78
3 Hysteretic damping ....................................... 78
4 Damping in structures .................................... 80

PART4 Structural dynamics design •.................•.....•..... 81

Chapter 9 Mathematics .....•........•.......................•..... 83


1 Introduction ............................................. 83

Chapter 10 Single degree of freedom system ............•...........•. 87


I Introduction to SDOF ..................................... 87
2 Free vibration ............................................ 88
3 Forced vibrations ........................................ 93
4 Dynamic modification factor .............................. 96
5 Resonance .............................................. 98
6 Dynamic loads ........................................... 99
7 Mathematical approaches ................................ 100
8 The time domaiu ........................................ I 00
9 The frequency domain ................................... l 0 I
l 0 Elastoplastic systems .................................... 106
II Nonlinear systems ....................................... 109
12 Torsion dynamic forces .................................. 110
Chapter 11 Multidegree of freedom: lumped mass system ......•..... 111
1 Introduction to MDOF ................................... 111
2 Vibration modes ........................................ 114
3 Forced vibrations ....................................... 118
4 Pulsating load .......................................... 119
5 Modal analysis .......................................... 123
6 Damping in MDOF ...................................... 124
7 The lumped masses ...................................... 124

Chapter 12 Distributed mass system ..................•............. 127


1 Introduction ............................................ 127
2 Mathematical approach .................................. 129
3 Design of a beam ....................................... 132

PARTS Natural dynamic loads ................................. 133

Chapter 13 Earthquakes ........................................... 135


1 Introduction ............................................ 135
2 Earthquakes ............................................ 136
3 Earthquake loads ........................................ 138
4 Earthquake response analysis ............................. 139
5 Static force procedure ................................... 140
6 Linear elastic response spectrum .......................... 142
7 Analytic procedures for linear elastic response .............. 145
8 Nonlinear inelastic response .............................. 147
9 Ductility ............................................... 149
10 Pushover analysis ....................................... 151
11 Soil-foundation interaction ............................... 154

Chapter 14 Wind loads ............................................ 161


1 Introduction ............................................ 161
2 Quasi-static wind loads .................................. 163
3 The dynamic response ................................... 166
4 Aeroe!astic phenomena .................................. 169
5 Vortex ................................................. 169
6 Buffeting ............................................... 171
7 Galloping .............................................. 172
8 Flutter instabilities ...................................... 173

Selected bibliography 177

Index 179
Preface
Structural dynamics is a theme of wide-ranging knowledge covering a variety of
topics, some of them of direct application in structural design. Among the latter
a clear distinction can be made between those necessary for the engineer in the
daily practice of structural design and those related to academic activities,
research, and the development of commercial products.
This book was conceived and is written as an overview of some aspects of
structural dynamics. It is intended for engineers who normally tackle design
situations involving dynamic loads with the appropriate computer software in the
daily practice of design. Presumably, the book will be complemented by (a) the
technical information required for any particular dynamic problem; (b)
consultation with experts about unusual dynamic design situations.
The usual practice in structural design offices is to continuously collect and
update valuable technical information in this field of practice. This collection
includes textbooks, a variety of Codes of Practice, Committee's Guidelines,
National and International Reports on natural events and disasters, expert and
especially professionally oriented reports, scientific publications, professional
journals, professional catalogues of commercial products, and more.
Each chapter of the book deals independently with a subject in structural
dynamics without a necessary link to the foregoing chapters, as is the case with
textbooks. This approach allows the engineer to go directly to the topic of his
interest at any given moment, with one exception: this is the sequence of
chapters 9-12, where a certain continuity and correlation was considered for
reasons of clarity of presentation. To minimize difficulties for the reader, in
these chapters only a schematic elaboration of the mathematical treatment of
structural dynamics is included.
In the other chapters mathematical formulations are given just by way of
complementary information, so as to refresh the background of the practicing
engineer. This particular means is necessary for the acquisition of a thorough
insight into structural dynamics.
In some chapters portions of text are set in italic, which is designed to bring
the reader's attention to topics or other text of particular importance.
This book has benefited from countless publications, mainly textbooks,
reports, papers, symposium and conference proceedings, journals, and many
more, dealing with structural dynamics, published in the course of decades.
Together they have created a vast, rich, and deep panorama on the theme.
Structural engineers, working in design and consulting engineering, like the
author of this book, are indebted to them for furnishing the professional
knowledge required in their daily practice.
Last but not least, this book was conceived and written with the most
profound regard for the author's colleagues, structural engineers in the daily
practice of design, who carry on their shoulders enormous responsibility for the
stability of structures built everywhere.

Oscar Sircovich Saar


2006
Acknowledgements
J wish to thank to Professor Carlos A. Brebbia who suggested the idea of this
book and to whom I remain in debt for encouraging me to do it.
Thanks to Mr. Murray Rosovsky who generously gave his time to ensure that
the text of this book was written in the correct language.
Thanks to two academic students, Ayala Even Zohar for her dedicated work
on my drawings and Eyal Cohen for his intelligent typing of the mathematical
formulations.
~'

I
f
§

I
I
I

I
I
List of symbols

f
I
A
A,,
A,,(t)
amplitude
amplitude of the nth normal mode
time varying amplitude
11 c damping coefficient, viscous damping coefficient
CAD air damping
1 c" critical damping coefficient
equivalent viscous damping coefficient

I
Ceq
ch,eq equivalent viscous damping coefficient for a hysteretic mechanism
C,C1,C2 constants
:w dt, dr time increment
i
I E Young's modulus
I I natural frequency, natural frequency of the nth normal mode/,= l!T,,

It fc
is
F
stress in concrete
stress in steel
external force
_fil'
F, external earthquake force
''
-M Fk restoring force

I F1 inertial force
Fo damping force
I FL wind lift force
II
g
h
acceleration due to gravity
height, thickness

I
-I
Hz

I
Hertz (cycles per second)
impulse
moment of inertia
i k spring constant
I~ L, I length
~- m uniform distributed mass
M mass, lumped mass
!- M moment of external load, moment at internal section
'
n integer, normal mode number
N axial load

If;
,i-
p external load
p(t) dynamic load
p load
P(t) dynamic load
Pc(t) dynamic radial load
Ph(t) dynamic horizontal component ofload
P,(t) dynamic vertical component ofload
P"(t) dynamic load at the nth mode
P-11 effect in compression members
R structural resistance, material resistance
Sa spectral acceleration
Sd spectral displacement
Sv spectral velocity
St Strohual number
t time
lct short period of time
T natural period of vibration
T,, natural period of vibration of normal modes T1, T2, ... , Tn
v wind velocity
v deflection, displacement
v velocity
v acceleration
w weight
x,y,z geometric coordinates
x, earthquake structures motion
x, earthquake ground motion

a angle
y weight
11 increment
o;, static deflection
s, dynamic deflection
41 plastic deflection
Sct(t) time-varying deflection
& strain
concrete strain
steel strain
phase angle
friction coefficient, ductility ratio
damping ratio
air density
stress
time
steel bar transversal section
modal shape
OJ natural circular frequency
Wo damped natural circular frequency
OJn natural circular frequency of the nth normal mode
Q forcing force frequency

Abbreviations

CG center of gravity
DE damping energy
DMF Dynamic Magnification Factor
DMS distributed mass system
EI flexural stiffness
KE kinetic energy
MDOF multidegrec of freedom
PSv pseudo-velocity
SDOF single degree of freedom
SE strain energy
(
CHAPTERl

Structural design in dynamic situations

A comprehensive structural design should simultaneously include the static and the
dynamic aspects of the problems to be solved. The basic relations of the two aspects,
mainly the deformations~stresses relation, are the same for structural analysis.
The main difference between a static and a dynamic design situation can be
described through a comparison of a simple, statically determinate, linear structure
under static or dynamic load. In the static design situation, calculations can be
performed straightforwardly; in the dynamic design situation, the external load,
which varies with time, imparts to the structure dynamic forces that are also variable
with time, creating an interdependence between cause and effect.
Structural calculations in dynamic design situations are far more complicated
than those in static design situations. They can evolve to highly sophisticated lev-
els, requiring mathematical training for anyone who attempts to perform them, an
uncommon circumstance in the practice of structural design.
Mathematical treatment of a dynamic case is indeed the right way to determine
the response of a structure in a particular dynamic design situation. This makes
the problem closer to the activity of an expert than to that of an engineer facing a
practical problem to be solved. In the present computer age, it is quite normal for
the engineer to rely on the correct use of computers to find a solution that satisfies
the requirements of safety and performance. This is so even though the engineer's
knowledge, which includes mathematics and physics, applied mechanics, struc-
tural materials, structural dynamics, seisrnicity; and generally the basic training for
complementary self-instruction across the entire range of his or her daily practice,
enables him or her, if required, to deal directly with the dynamic issue.
The practicing engineer should have a clear and thorough understanding of the
dynamic issue in order to deal with problems of design in all its aspects. He or
she should not have to deal with the mathematical formulation and solution of the
differential equations of equilibrium.
2 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Actually, the work of the practicing engineer in the computer age is slightly
different from that of an expert in structural dynamics, as becomes apparent from
the following comparison.

Expert's approach Structural designer

I. Physical and mathematical modeling; 1. Geometrical-mechanical definition


formulation of the dynamic equation of the structure
of motion
2. Mathematical solution of equation 2. Input of model and load definition
of motion into computer
3. Formulation of a structure's response 3. Interpretation of graphic and
functions; elaboration of response numerical output for the structure
curves

The response of a structure under the effects of a dynamic event should be taken
into account at the design stage. This is for the safety and reliability of the structure,
user's convenience, fulfillment of the expected performance, and prevention or
minimization of structural damage.
The dynamic design situation deals with the relation between the structural
response and the external dynamic load. Mass, stiffness, and damping capacity
of the structure are the main factors to be taken into consideration; the engineer
can change or modify them while learning about the dynamic response of a partic-
ular case.
Designing for static and dynamic loads alike is always a "back and forth" process.
Part 1

Basics
CHAPTER2

General overview

1 The dynamic response of structures

Engineers know fully well that a load which changes with time creates a design
situation that may require due consideration of structural dynamics; it depends on
the type of structure and the kind of loads superimposed on them.
A dynamic load is by definition a time-varying load. Its effect in a given structure
depends on the interrelation between its particular characteristics and those of the
structure, as can be seen in Fig. 2.1.

(a) _i;:=======:::,.
Yi
_Ls:

Figure 2.1: Simple supported beam, 1200, A= 3340mm2 , I= 2.140 x 107 mm4 .
(a) Concentrated load lying on top of it. (b) Same load, but applied on
the beam in a very short increment of time !ct.

Figure 2.1 a shows the value of the static deflection 8" of the beam in the middle
of the span. The load is applied slowly enough to create a static design situation,
and it can be calculated using formulas found in any textbook. Figure 2.1 b shows
values of the dynamic deflection 8d of the beam, at the same point and for the
same load, but applied suddenly in a very short increment of time. For different
6 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

values of ta, the values of 8d can be calculated on a computer using any software
for dynamic calculations.
The formula can be written as

8d = 8,, · (DMF) (2.1)

Hence, the vertical deflection on the beam 8d under a dynamic load P(t) can
be obtained from the static deflection 8,,by multiplying the latter by a Dynamic
Magnification Factor (DMF), whose value fluctuates with time and varies, loosely
speaking, between 1 and 2 for this particular example.
A superimposed load on a structure, even if it is intended to be a static load, can
initially be a dynamic load, depending on how it will be applied to the structure.
Applying the load on the beam in a short period of time td, as in Fig. 2. lb, creates
a dynamic design situation. If, for example, load P is built up gradually, in its set
place, it will obviously create a design situation for a static load, as in Fig. 2.la.
Figure 2.2 depicts the vertical deflection of the beam 8d under a dynamic load.
It is an oscillating deformation with Ost as a medium value.

(0,0_)+-------+------,1'---<>

ll st-k
- p

_rDMF
!r-=T,----z---r-,tf-~T=-_,,.-7"' JDMF
ll
Figure 2.2: The same beam as in Fig. 2.1. Vertical deflection under dynamic load
P(t).

The fluctuation curve of 8d(t) is a sinusoidal function of time and it' amplitude
is a function of DMF.
The DMF depends on two factors:
I. M the combination of a percentage of the mass of the beam with the mass of
the dynamic load and the spring constant k of the beam;
2. The ratio between

Id (the period of time for which the load is applied on the beam) and
T (natural period of the beam).

The presentation of 8d as function of 8,, as in eqn. (2.1) is very illustrative. Its


purpose is to remind us that both the static and the dynamic response are related to
the same formula:
F =k · 8.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 7

This is the basic relation that links

the force F, the spring constant k of the structnre, and the deflection 8,

and remains valid whether F is static or dynamic.


Other structural design considerations, such as internal forces, stresses, and
strains, are usually obtained from the time-varying displacement 8(t).

l,l Definitions

To become acquainted with the concept of period T, let us take a look at Fig. 2.3.

II II II II

P----<>

(a) (b) ( c) (d)

Figure 2.3: (a) Vertical flexible bar between articulate supports. (b) Deformation of
barunder external load P. (c) Starting position for horizontal oscillating
movement of bar on sudden removal of P. (d) Extreme position of
movement of bar at left.

In Fig. 2.3c the vertical bar is at an initial position of free motion from right
to left at the very moment of removal of load P. In Fig. 2.3d the bar has reached
the end position of its initial movement, which becomes the starting position of a
returning free motion to the right. The bar's motion will continue back and forth
unless some physical characteristic causes it to stop.
One complete motion of the bar in both directions is called

one cycle.

The period of time required for the motion of the bar to go through one cycle is

T (the natural period of the bar)

measured in seconds.
The number of cycles the moving structure performs in l s is
1
f = T (the natnral frequency)
measured in Hertz.
8 DYNAMICS JN THE PRACTICE OF STRUCTURAL DESIGN

In the mathematical formulation of motion one cycle is considered a 360°, or


2n, motion; hence

2rr
w= T (the natural circular frequency)

measured in cycles per second (c.p.s).


The motion of the structure, as represented in Fig. 2.3c and d, after removal of
the external load is called
free vibration
(see additional comment at the end of next paragraph).

1.2 Free modes of vibration

The free vibration movement ofa bar ofuniform mass m and unifonn spring constant
k all along, is in reality a superposition of different vibrating modes, graphically
represented in Fig. 2.4.

static rosition l0"1 inode znd mode 3"1 inode 4 111 1node
II II II II
\
node
I
node
I
t+l node t+J <+l node t+l····
node
I
node
I I

(a) (b) (c) (d) (e)

Figure 2.4: Characteristic vibration modes of a vertical bar in free vibration.

The number of modes of vibration.for a structure with uniform mass all along is
theoretically infinite. Each niode has its own natural period, T1, Tz, ... , T11 , usually
ordered decreasingly, and its own shape, the first mode of vibration is known as the
fundamental mode.

w = n2rr2 {Ei
fn = 1/Tn, Wn = 2rr/Tn, n 12 y-;;;
where n is the mode number.
The mode shapes are sine curves, with an amplitude An(t), which vary with time.
The nodes of each mode are at fixed locations; they do not change during vibration.
A 11 can be calculated, using a computer, for regular dynamic design situations.
In section I.I, Fig. 2.3, only the fundamental mode was considered, where w
represents the natural circular frequency of that mode.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 9

1.3 Damping

The fluctuation of the beam in Fig. 2.2 cannot continue indefinitely because it
is energy consuming. Each structure has its own internal mechanism of energy
absorption, the damping mechanism of the structural system, which is the factor
responsible for this fading out of the movement.
An indefinite free vibration of a beam is not real. It disappears with time, as can
be seen in Fig. 2.5.

Figure 2.5: lld stabilizes, with time, at a static value llst.

The damping mechanism can originate in different sources of dissipation of


energy.
In the beam in Fig. 2.1, with friction-free articulations and friction-free support
movements, the external energy submitted by the dynamic load to the beam is
absorbed by an internal mechanism of friction in the material of the beam. This is
hysteretic loss.
Figure 2.6 shows a stress-strain curve characteristic of a cyclic deformation.
The area enclosed in the loop is a measure of the energy absorbed in one cycle.

Figure 2.6: A hysteretic loop.


1Q DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

The dimension of "a" depends on many factors inherent to the material and on the
stressing limits which promote either elastic or nonelastic hysteresis.

1.3.1 Sources of damping


Damping is a factor of the utmost importance in the dynamic response ofstructures.
It attenuates the response and reduces the amplitude ofvibration; strain and stresses
are reduced accordingly.
Damping contributes to the dissipation of the energy imparted to the vibrating
structure by the external dynamic load. In most structures, a complex mechanism of
damping affects the response of the structure. Only in rare structures, like a simple
friction-free supported steel beam with a punctual dynamic load, as in Fig. 2.1,
vibrating slowly in the elastic range of stresses, may a single source of damping
effect be considered in the design process.
Sources of damping can be (a) inherent in the structure, (b) typical of the con-
struction system, (c) external to the structure, or (d) from the soil foundation.
Damping of type (a) arises from the structural material. Damping of type (b) arises
from composite structural materials and construction systems. Damping of type
(c) is the contribution of nonstructural coniponents, such as pavements, flooring,
partition elements, railings. Damping of type (d) originates from the interaction
between the soil and the structural foundation, especially for soft soils.
An additional source for energy dissipation originates when the amplitude of
deformation of the structure exceeds the elastic limits, towards the plastic range.
Figure 2. 7 shows a cyclic inelastic deformation curve.
As already mentioned, the area of the hysteretic loop is a measure of the energy
dissipated in a cyclic response. This kind of damping mechanism, by inelastic

strain e I

Figure 2.7: Force deformation in an inelastic hysteretic loop.

I
I
}
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 11

deformation, is more effective than others when a significant amount of energy has
to be dissipated; for example, if the excitation is an earthquake motion.
Figures 2.6 and 2. 7 show a graphic comparison of the energy dissipated in a
cyclic response of a structure with elastic versus plastic behavior, respectively.
Damping in real structures is a complex phenomenon, usually combining dif-
ferent types of energy dissipation mechanisms. Therefore, the engineer should be
careful when assessing the value of the damping coefficient. Actual values to be
included in design should be taken from dynamic experimental tests already per-
fonned on existing structures. These values can be found in appropriate reports
and books.
For existing structures, it is possible to improve their capacity to dissipate some
of the energy imparted by an external dynamic event. This can be achieved by
attaching an external damping mechanism to the structure (see section 5 below).

1.4 Forced vibration

Figure 2.Sa shows a beam with a rotating machine on top of it. The unbalanced
rotating weight of the machine imparts a dynamic vertical load P(t).
In this particular case P(t) is a fluctuating load with a "circular frequency ff'.
Figure 2.8b shows the diagram of the load.
The pulsating load P(t) forces the beam into an undulating deflection. After some
initial perturbation the beam moves mainly with the periodicity of the load.

(a) l P(t)

(b) P(t)

(c)

Figure 2.8: (a) The same beam as in Fig. 2.1, under a pulsating load P(t). (b) Dia-
gram of the pulsating load. (c) Undulating deflection of beam.
J2 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

This forced movement is called the steady-state response, and it lasts as long as
the pulsating load acts on the beam.
Figure 2.Sc presents a diagram of the deflection of the beam subjected to the
pulsating load. This deflection combines two superposing movements: the natural
free motion of the beam and a forced motion.
The free motion of this combined undulation is called the

transient motion

which disappears with time because of damping, as shown above in Fig. 2.5.
This transient motion can be considered an initial perturbation. Thereafter, the
beam vibrates regularly in what is called the

steady state.

1.4.1 Resonance
The deflection of a beam subjected to a pulsating load can be written, again,
8a = 8,1 • (DMF), as in the example in Fig. 2.1.
The DMF depends on the ratio r:l./w of the circular frequency of the pulsating
load and the natural circular frequency of the system.

DMF

5.0

~~O.l

1.0

JL I
l
w

Figure 2.9: The DMF for sinusoidal pulsating load.


!
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 13

When the frequency Q of the pulsating load approaches the value w of the
natural frequency of the structure, the amplitude of the beam's vibration increases,
theoretically, towards infinity; a state of resonance is created.
Actually, this mathematical conclusion should be understood as a simplification
of a very complex issue. Real structures do not reach infinite resonance ampli-
tude because they become nonlinear, or because of plasticity or other effects; but,
deformations can indeed become too large to be tolerated.
In fact, the amplitude of vibration close to resonance is mitigated, primarily, by
damping, as seen in Fig. 2.9. Factor~ which is a measure of damping is given by
~ = c/c"
where
c is the actual damping capacity of the real structure;
Ccris the critical damping: a situation where the vibration is totally eliminated
by damping.
In any case, a structure does not reach the resonance amplitude at once. It builds
up with time and requires many cycles (Fig. 2.10).

/j

Figure 2.10: The initial response of a structure at resonance.

1.4.2 Response phases in a forced vibration


Figure 2.11 depicts a damped motion of a beam subjected to a pulsating load. The
load acts on the beam for a limited time, but long enough to produce a forced
vibration with its maximal amplitude.

( 1) initial transient vibration


(2) increasing a111p!itude because
of resonance
(3) steady state
(4) free vibration

/j

Figure 2.11: The response of a beam with damping, subjected to a pulsating load.
The load is removed after a limited time.
14 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

The diagram in Fig. 2. l l shows four consecutive phases of vibrations:


1. an initial transient 1novemcnt, corresponding to the natural vibration of the
structure, superimposed on lhe guiding curve of vibration; this phase lasts for
a short number of cycles;
2. the amplitude of vibration increases with each cycle towards the maximal
amplitude of the forced vibration;
3. steady-state motion at resonance, with an amplitude of vibration according to
the damping characteristics of the structure (Fig. 2.9);
4. fading out of the motion of the structure once the external load has been
removed.

2 Dynamic equilibrium of forces

2.1 Free motion

Two particular components should be taken into consideration in any dynamic


design situation, in addition to the spring component (flexibility or rigidity), which
is characteristic of the static design situation. These components are the inertial
and damping forces that originate owing to the motion of the structural system
(Fig. 2.12).

II II II

I \
I I
M

\ I I

(a) (b) (c)

Figure 2.12: A linear elastic system with damping. (a) A flexible bar as in Fig. 2.3a;
M = mass of the bar. (b) Bar in free motion, as in Fig. 2.3c and d.
(c) Forces in the bar in motion.

The bar in Fig. 2.12b is in a state of free vibration with damping. Three forces,
changing continuously with time, originate owing to the motion of the bar, namely

Fk (the restoring force),


F1 (the inertial force), and
Fo (the damping force).

The restoring force originates owing to the deflection of the beam.


DYNAMICS IN THE PRACTICE OP STRUCTURAL DESIGN 15

A mass in motion develops a fictitious force, F 1 = M · V, called the inertial force,


proportional to its acceleration (v). It results from applying D 'Alembert's principle,
which evolves from Newton's second law.
The damping force, Fo, depends on the type of structure under consideration. It
can be of diverse physical nature, as already noted in section 2.4.
At any given moment of motion these forces are in a state of dynamic equilibrium,
so the equation of equilibrium of forces is

Fk(t) + F1(t) + Fo(t) = 0 (2.2)

Because of damping, both the motion and the forces fade out gradually until the
bar comes to a complete stop. Fb F1, and Fo are functions of time F = f(t), as
already mentioned.

2.2 Forced vibration

In Fig. 2.13a the pulsating load P(t) originates an oscillating motion on the beam;
the bar is at the stage of forced vibration. In Fig. 2. l 3b the deflected position
of the beam at a given moment is shown. In Fig. 2. l 3c the forces Fk, F1, and Fo
are developed by the motion of the beam in a direction opposite to that of the
external load.

P(t)=P1sin.Qt

(a) -------'~-----
extreme positions of vibrating motion

( c) tFk, F1, F0 forces originated by 1notion of bcan1s


2~----L----;_;;~-l v - direction of oscillation

Figure 2.13: A horizontal beam subjected to a pulsating load P(t).

For this dynamic design situation, the equation of equilibrium is

F1(t) + Fo(t) + Fk(t) = P(t) (2.3)

The oscillating motion of the beam continues as long as the dynamic load pulsates
on the beam; it is a
forced vibration.
This motion does not disappear because of damping, as it takes place in free motion.
The effect of damping on the forced vibration is to mitigate the amplitude of
motion, up to a certain limit.
16 DYNAMICS IN THE PRACTICE OF S1RUCTURAL DESIGN

Since forces F = f(t) are functions of time, eqns. (2.2) and (2.3) are differential
equations. The solution of these equations, in the form of deflection v, is also a
function of time,
v =f(t)
and so are velocity (V) and acceleration (V) of motion.

2.3 Mathematical treatment of a dynamic design sitnation

2.3.1 Single degree of freedom (SDOF) system

~v(t)
P"(t)
M
2/~~2-""='~tirrace level P" (t)
I
M,
I
I k
I
Pv(t) p (t)
~· //' ;

(a) (b) (c) ~P"(t) (d)

Figure 2.14: Schematic observatory tower.

Figure 2.14 depicts:


(a) A cantilever column of height h, with uniform mass m and rigidity EI, all
along; the static equation of equilibrium for the horizontal load is Fo = k · vo,
where k = 3EI/h 3 .
(b) A schematic tower with an observatory terrace on top of the cantilever colun1n.
A rotating machine, as in Fig. 2.8, is attached to the upper slab.
(c) The dynamic force o,f the unbalanced weight of the rotating machine is repre-
sented by the vector Pr(t), 1-vith the vertical component Pv(t) along the column.
(d) A spring ofconstant k and a concentrated mass M = M 1+M2. It represents the
vibrating tower of (b), provided the column's mass and Pv(t) can be neglected
and also if Ph(t) can be considered as operating in the center of gravity
(CG) ofM.

This simplification o,f the dyna1nic design situation in Fig. 2. l 4b is known as an


SDOF system. It is valid for small amplitudes of motion and allows a simple mathe-
matical treatn1ent ofthe dynamic problem. Usually the results are very close to those
of the exact mathe1natical calculation for slende1; uniform, and light 1nass columns.
The simplified dynamic pmblem in Fig. 2. I 4d represents a single concentrated
mass M on a linear elastic spring of constant k, excited by an external dynamic
load. The internal forces Fr(t), Fo(t), and Fk(t) are in constant equilibrium with
the external horizantal load P(t).
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 17

Hence, the equation of equilibrium of dynamic forces for an SDOF system is

(2.4)

where
Ph(t), is as already explained in connection with the load P(t) in Fig. 2.8;
Fr(t) =M · ii = inertial force = mass x acceleration;
Fo(t) = c · v = damping force = damping coefficient x velocity;
- Fk(t) = k · v =spring force= elastic spring constant x deflection.
An SDOF system means one single mass attached to a linear elastic spring which
has no mass. This equation is valid for any type of external excitation.
P(t) represents, for example, a suddenly applied load, a step load, a harmonic
or periodic load, or even a random irregular load.

2.3.2 Damping
The mathematical expression chosen for the damping force in eqn. (2.4 ), Fo(t) =
c · V, requires an explanation: coefficient c stands for a viscous type of damping.
It represents the damping force that will affect a structural element moving in a
fluid, not characteristic of the dissipating mechanism that operates in a structure.
Viscous damping is chosen because it is proportional to the velocity of movement of
the vibrating structure. Introducing this type of damping force leads to the simplest
mathematical formulation of the dynamic equation of equilibrium, which for a free
vibration is
M ·ii+c·v+k·v=O (2.5)
a linear differential equation of the second order with constant parameters. This
equation has a very well known solution.
An alternative selection of hysteresis as the type of damping, for example, pro-
duces a complicated dynamic equation of equilibrium, as follows:

M ·v+(c/w)·v+k·v=O (2.6)
The friction type, known as Coulomb's damping, is dependent on the amplitude
of movement and the force acting on the suifaces in friction. Therefore, introducing
Coulomb S damping into the equation of equilibrium of dynamic forces gives rise
to a complicated differential equation.
In fact, the comprehensive effect of damping in a structure is the result of super-
position ofdifferent types of damping, which is difficult to identify. For this reason,
the viscous coefficient introduced in the equation is evaluated so as to represent an
equivalent of the overall effect of energy dissipation during vibration.

2.3.3 Solution of eqnation of motion


2.3.3.1 Harmonic load Let us assume an external dynamic load, Ph(t) = P1 ·
sin(Q · t), that is, a load of amplitude P1, varies with time according to a sinusoidal
18 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

function with frequency Q;for such a load the solution of eqn. (2.4) is in the form
of a dynamic displacenzent v(t) that varies with time:

Equation (2.7) can be rewritten as

v(t) =free vibration+ forced vibration

(a) Remarks on the "free vibration" component ofv(t).

I. e-~wt tends to zero with increasing time, which means that the free vibration
component ofthe motion will disappear after some cycles. Therefore, it is known
as a transient component, or initial perturbation, of the forced vibration.

2. w= /k7M is the natural circular frequency of the vibrating system.


S= c /2M w is a measure of the actual damping in the vibrating system, as
a fraction of the critical damping.

Usually Wo r-J w, for real structures, since S < 0.1.

3. C1, C2 are constants to be calculated according to the initial conditions of


motion of the vibrating structure.
(b) Remarks on the ''forced vibration" component ofv(t), the steady state. It con-
tains three parts:
I. v,, = P1/k is the static deflection of the structure.

2.

The DMF for a harmonic external load is represented in Fig. 2.9. It can be
seen that without damping

DMF = 1/(1- (n/w) 2 ]


tends to oo, when n tends to w. This is known as a case of resonance.
The effect of damping is to avoid the danger of resonance and to mitigate
the amplitude ofv(t).

3. The fluctuating component:

sin(Q · t + 6)
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 19

represents the function of time of the forced vibration. From this expression it
becomes clear that the motion of the response is out of phase with the motion
of the external forcing function; the phase angle is e.

2.3.3.2 Constant load Let us assume now a suddenly applied load, as an alter-
native to the periodic load that we analyzed before: P1 is a constant load, applied
in a very short time, on a structure initially at rest. The applied load remains there.
The fact that the load is applied suddenly creates a dynamic design situation.
The equation of equilibrium is similar to eqn. (2.4 ), but with an external constant
load that does not change with time:

M · v+ c · v+ k · v = P1 (2.8)

The solution of this equation, for a structure initially at rest, is

(2.9)

which can be written as


v(t) = v,, · DMF
where v,, = P1/k, and DMF = [ 1 - e-~""(cos wot+~· sin wot)].
The fluctuating component disappears with time, since e-.;wt -+ O; hence the
DMF tends to I; that is,
v(t)at rest = Vst

2.3.4 Vibration of a real strncture

I n:2 /Ef
wn= (n-2) 72 ...jfil m;EI
e
-~- 1~~ 1 mode

re' ffiI -~----


w2= 2.2512 ...j m ?i fixed points ~

W3 · 2
2
~(3-..L)~
ei
'/II
-
m
- "~ixedpoints
------ ·
-----.
fixedpomts
--- 3'amode

Figure 2.15: A cantilever beam with uniform mass and rigidity all along. Normal
modes of free vibration.

The natural modes of vibration of the cantilever beam in Fig. 2.15, and their
natural circular period Wn, can be found in any textbook, or can be calculated
using a computer. The deflection of the beam, at any given instant of time t, is the
sum of all the mode curves, each with its own period.
20 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Real structures, like the cantilever beam, have distributed mass and continuous
rigidity, not always unifonn all along. Still, vibration of a real structure implies the
superposition of an infinite number of vibration modes.
Attempts to simplify the problem, as in Fig. 2.14d, are oriented to simplify the
mathematical treatment, and hence to clarify the physical response.
Note the similarity between the cantilever deformation curve in Fig. 2.14a and
the first mode of vibration in Fig. 2.15.

3 Energy in structural vibrations


The vibration of a structure is an energy-consuming process, which starts with an
external dynamic action on the structure, such as a dynamic load or an imposed
motion. Removal of the external dynamic action means stoppage of the energy sup-
ply, and hence stoppage of the vibration owing to damping, actually in a gradually
decreasing process (see phase 4 in Fig. 2.11 ).

3.1 Undamped free vibration

A discussion of energy in structural vibration should better begin with an undamped


free vibration, in which two types of energy continuously balance each other, in an
oscillating motion, like that shown in Fig. 2.3c and d:

SE + KE = constant (2.10)

The strain energy, SE= f(t), is the energy accumulated in the deflected beam at
any given moment of the motion, and varies with time. Kinetic energy, KE= f(t),
is the energy the mass of the structure accumulates at the same given moment,
during motion, according to the velocity of movement; it is also a function of time.
For comparison, see the pendulum in Fig. 2.16.

/ '
I '
/
'
/
'
/
'
/
'
/
/
''
/
/
'

-v M +v

Figure 2.16: Mass M in a pendulum motion.


DYNAMICS JN THE PRACTICE OF STRUCTURAL DESIGN 21

The pendulum movement of mass M encompasses two types of energy in a stage


of continuous balance if any damping is neglected: potential energy and kinetic
energy. The first is comparable to the strain energy of a deflected beam, while the
second is the energy generated by the motion of the mass of the structure.
The potential energy reaches its maximal value at the extreme positions when
motion stops. Since the mass is at rest in this position, kinetic energy is zero.
With movement of the mass M towards the opposite extreme position, potential
energy decreases progressively, while the value of kinetic energy increases. When
the mass M traverses the position x = 0, the kinetic energy is maximum and the
potential energy is zero, and so on.

3.2 Damped forced vibration

In a practical case of a structure under a dynamic load, two other types of energy
should be taken in consideration:

WE (the energy supplied by the work of the external load) and


DE (the energy dissipated by damping during motion).

A loose equation of equilibrium of energy will be of the form

WE=DE+SE+KE (2.11)

which means

work of the external load = change of internal energy in the structure

Remark: The differential equation of equilibrium of dynamic forces, eqn. (2.3), can
also be obtained by an appropriate mathematical formulation of the equations of
energy.

3.3 Work and Energy

In Fig. 2.17 the beam is the same as in Fig. 2.1. Load P1 is dropped onto the beam,
from a certain height h; the dynamic elastic deflection of the beam is lid = f (t).

l 10 [ kN
k-=7.02 n1m
J

Figure 2.17: A simple supported beam under an accidental load.

The work performed by the load P1, all the way down h +lid, is

(2.12)
22 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

The beam is at rest when it reaches its maximal deflection amplitude 8d; hence
KE= 0, as is the case in the extreme position of the pendulum in Fig. 2.16.
Neglecting any initial stored energy, the strain energy in the beam at its m.aximal
amplitude of de,fiection is
(2.13)

Assuming there is no damping, DE = 0, eqn. (2. I I) becomes (WE = SE), see also
Fig. 2.18:

Pi· (h + 8d) = ~k · 8~ (2.14)

8d = P/k ± J(P/k) 2 + 2h(P/k), or

8d = 8,,. (1 ± J1 +2h/8,,)
where JI+ 2h/8,. represents the maximum amplitude of DMF.

h LI dl

Figure 2.18: A graphical representation of the equilibrium of energy, eqn. (2.14).

3.4 Energy dissipation

3.4.1 Damping versus strain energy


Hysteretic damping was already mentioned in section 2.4. This typical damping is
a source of dissipation of energy of the material, when a structure ;..~ subjected to
an oscillatory external load.
In Fig. 2.19 a graphical comparison is sho1:vn of the elastic strain energy (SE)
and the da1nping energy (DE). The ratio o,f these t1:vo areas is a nieasure o,f the
hysteretic damping capacity of the material to dissipate energy. Area DE, inside
the hysteretic loop, represents the mechanical energy dissipated in one cycle. Note
points a and b. At a stre.vses are zero and strains are not; the opposite holds at b.
In the hysteretic loop stresses and strains are not in phase.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 23

1«>'---SE

' e (LI)

Figure 2.19: Hysteretic loop, u/E, or P/8.

3.4.2 Damping in real structnres


As mentioned in section 2.3.2, the effect of damping in a real structure is the result
of a superposition of different types of damping mechanisms, each with its own
contribution to the overall dissipation of energy.
To simplify the mathematical treatment of the dynamic problem, those types
of damping are represented sometimes by Ceq. which is an "equivalent viscous
coefficient, " based on a comparison of the energy dissipated by those mechanisms
with that of a viscous damping.
The work done by a viscous force, opposing an external dynamic load, can be
better evaluated by calculating the energy dissipated in one cycle of a hannonic
motion, as follows:

DE= 4.
rn/D.
lo Fo. v. dt = 4.
r1
lo
Q
(cv). v. dt (2.15)

For a forcing load, P ·sin ( Q · t), the steady state component ofthe response function,
from eqn. (2. 7), can be expressed as

v(t) =A ·sin (Q · t - 9)

where
9 =tan-I ( 2~. (Q/w) )
1 - (Q/w)2
is the phase angle between the forcing function and the steady-state response
function.
Introducing the derivative of the steady-state function v(t) into the integral of
eqn. (2.15 ), and peiforming the relevant mathematical operations gives

DE= c · (nQ ·A 2 ) (2.16)


24 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

which is the energy dissipated by the viscous mechanism, in one cycle ofvibration of
amplitude A and frequency n, proportional to c = coefficient of viscous damping.
From eqn. (2.16) it is possible to establish

Ceq = DE/rrfJ.A2
for any damping mechanism. DE is a nieasure of its energy dissipation
characteristic.
For example, in Fig, 2.19 the area of the ellipse is DE = Ch · k · rr · A2 (where
A = 8,,); hence
Ch,eq =Ch' k/Q (2.17)
where Ch,eq' is the viscous equivalent damping coefficient for a hysteretic
mechanism.
A second example: the energy dissipated in one vibration cycle, by a Coulomb
damping mechanism, is , DE = 4 · µP ·A, where P is the force acting perpen-
dicularly to the moving areas in friction, µ is the friction coefficient, and A is the
amplitude of movement of these areas; hence

cc,oq = 4 · µP /(nQ ·A) (2.18)

where cc,cq is the viscous equivalent damping coefficientfor a Coulomb mechanism.

4 Linear and nonlinear response of structures


Structural materials are usually ductile. Their load-resistance curve (P/R) is ini-
tially a straight line as indicated in Fig. 2.20, between 0 and a, which thereafter
bends progressively into a plastic zone of deformation.

I
I

R
0,0

Figure 2.20: An elastoplastic material.

If the deflection of the structure remains in the elastic range during the dynamic
vibration, which means a linear spring constant, the period of the structure remains
unchanged and the structural response, as described previously, is correct. The
motion is known as the linear elastic response of the structure.
DYNAMICS IN TIIE PRACTICE OF STRUCI1JRAL DESIGN 25

The stiffness and consequently the natural frequency of the structure undergo
changes when plastic articulations are created; axial forces may affect the geometric
stiffness too.
Plastic deformation of the material affects the damping coefficients since a certain
similarity exists between damping and plasticity. These design factors and others
disrupt the initial linear elastic motion, turning it into a plastic response.
Sometimes keeping the structure at the stage of elastic response is impractical.
A certain amount of plasticity allows energy absorption, and hence a much more
elegant architectonic design and a more economical solution.
Creating a design situation for a nonlinear response can be achieved by properly
manipulating the physical characteristics of the structure for input into the computer
(Fig. 2.21). Figure 2.21 offers an example of a structure that can be intentionally
designed to vibrate with a certain amount of plasticity.

P(t}

~
(a)
~ P(t)
~
(b) [j. ~ .§
P(t)

(c)
~
Figure 2.21: A fully restrained beam under a dynamic load.

In Fig. 2.21 a we see a beam fully restrained at the supports, vibrating under a
dynamic load P(t); stresses and motion are kept within the linear elastic range.
In Fig. 2.2lb a certain amount of plasticity is allowed at the supports. In this
case a new system can be considered for the response of the structure beyond some
elastic limit value of deformation or stresses. This type of partial plastic hinge can
be modeled using a computer, for example, with an appropriate plastic spring link.
In Fig. 2.2lc the procedure is the same as in Fig. 2.2lb, but with an additional
plastic hinge. ·
In practice, such a transformation of an elastic beam into another with plastic
hinges can be achieved with appropriate design details.
A plastic dynamic design situation has to be created with care. The plastic
response of a structure under dynamic loads can yield major unwanted conse-
quences, such as undesirable permanent deformations and/or malfunction of the
structure.
Nevertheless, for certain types of dynamic loads, such as explosions and earth-
quakes, the plastic response of the structure is usually anticipated. These loads are
in reality drastic events that require absorption of an enormous amount of energy.
26 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

4.1 Elastoplastic design situations

The mathematical treatment ofa nonlinear design situation is a step-by-step proce-


dure, with differential increments, instead of analytical treatment as in the case of
linear elastic systems. In each of those steps the dynamic system can be considered
linearly elastic. Each step starts with the structural and load characteristics at the
end of the preceding incremental step.
Unfortunately, this procedure does not allow thorough insight into the basic
aspects of an elastoplastic dynamic design situation. Each individual case leads
to its own conclusion. Two particular cases described here may be of some help
to clarify the issue. These cases are in respect of a structural material with a
deformation diagram, as shown in Fig. 2.22.

a a

Up]

e e
real diagram bilinear simplification

Figure 2.22: Stress-strain diagrams.

Case a: A beam like that in Fig. 2.1, but overstressed by an external load applied '
suddenly. The load, P1, remains on the structure for an interval of time greater than -:1
ten times the period of the structure. The work performed by the external load is
WE=P·8am
where 8dm• which is the 1naximum deflection ofthe loaded beam, reaches a value 8pi·
This work is equilibrated by the internal energy ofthe deflected beam, represented
by the hatched area in Fig. 2.23:

Rp1 -----~

Figure 2.23: Resistance-deformation diagram.


DYNAMICS IN THE PRACTlCE OF STRUCTURAL DESIGN 27

From the equation of equilibrium WE ; : :; SE, the resistance of the beam is

P = Rpt · (1 - 2~) (2.19)

whereµ= 8p1/8e1 is the ratio of ductility of the structure.


From eqn. (2.19) two conclusions are evident: first, P increases with increasing
degree of plasticity; second, the required resistance of the beam is independent of
the natural period of the structure, provided the load is applied suddenly and left
for a long term on the beam.
Case b: The beam is the same as in case a, subjected to a load P great enough to
produce plastic defonnation in the structure. The load is applied for an extremely
short period of time and removed immediately. Figure 2.24 shows the diagram- of
loading.

P(t)

Figure 2.24: Short-term load.

The concepts of momentum and impulse are useful for the analysis of this case:
- momentum is the product of mass Mand velocity V;
- impulse is the product of load P and dt, a differential increment of time.
The impulse of P for a very short interval of time is

i = p. ~dt

The impulse i of the external load imparts to the beam a motion with initial
velocity V,
i=M ·v
The kinetic energy developed in mass M by the motion is

The movement of the beam stops at its maximal deflection 8p1. and KE, the kinetic
energy, transforms into strain energy (SE) according to eqn. (2.10),

SE=Rpt. (sp1- ~s.1) = ~i2/M


2
Rpt · 8e1(2µ - I)= i /M
28 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Introducing w 2 = k/M and oo1 = Rp1/k,

Rp1 = i · wj.}2µ. - 1 (2.20)

From eqn. (2.20) two conclusions are again evident. First, as in case a the required
resistance R of the structure can be reduced by allowing a greater degree ofplastic
deformation. Second, contrary to case a, the required resistance of the beam is a
function of the natural period of the structure.

5 Vibrations in existing structures


The basic factors of mass and inertial forces, rigidity of the structure, damping
and energy dissipation, so well known in the dynamic design of new structures,
are precisely the factors to be kept in mind with regard to vibration problems
on existing structures, even though the approaches to the solution can
differ.
Passive or active sttuctural control is implemented to modify and adjust the
existing structure to the dynamic perturbation, in an effort to obviate frequencies
close to resonance, reduce amplitude of response and risk of fatigue, improve
structural and functional performance, and minimize user's discomfort.
Passive control techniques are so called because once they are implemented in
the structure they cannot modify their contribution to the response to an external
perturbation. Active control techniques are powered systems that operate by oppos-
ing the external perturbation, both in amplitude and in frequency, each time in such
a manner as to minimize their effect on the structure.
The vibration of an existing structure starts always with an external dynamic
effect and is the consequence of the dynamic characteristics of the structure itself.
Accordingly, the study of the problem should start with a thorough check of the
external source of the vibration and then proceed to the structure.
Perturbation of machinery and equipment on a structure can be changed, for
example, by providing an extra isolated inertial block or frame. Perturbation of
machinery operating close to the structure can be probably isolated.
If the vibration of the structure cannot be reduced to an acceptable level, by
intervention in the sources of the dynamic perturbation, the damping capacity of
the structure should be increased. This can be achieved by the addition of external
mechanical systems.
The range of engineering possibilities among active or passive devices is very
rich. It depends on the characteristics of the perturbing loads; it also depends on
practical or architectural limitations, budget, and many more considerations.
The engineer relies most in the industry when looking after devices for control of
dynamic performance of structures. These include moving masses, electromagnetic
mechanisms, viscous or dry friction devices, etc. He or she does not have to start
from zero.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 29

6 Working with the computer


6.1 Study models

With time, practicing structural engineers develop a certain "feel" for structures in
their daily work, which is actually the result of extensive practice together with
deep and correct professional knowledge.
To develop such a sense in a dynamic design situation, special attention should be
given to other factors, in addition to the stiffness k of a static design, the damping
ratio ~. and mass M. The engineer can consider alternatives to these variables,
creating more than one acceptable study model for input into the computer.
Figure 2.25 shows two diagrams of DMF for two acceptable study models of
the structure with the same degree of damping.

DMF

(1 =(,

,.
I I /'' ,,
/I I
I
I I
I
DMF1 for ~~ <1
/
/ I
~ I
\
,I
' ,_,.a /b
' '

Do Do Q_
-=l -=l w
w, w,
Figure 2.25: (a) The resonance curve for a structure with frequency w1 under a
dynamic load with frequency no. (b) The same as in (a), but for a
structure with frequency wz. DMF2 for no = Wj at structure a. DMF1
for f2o = WJ at structure b.

A comparison of the outputs of these models will help the engineer develop his
or her feel for the dynamic response of the structure and will produce options for
dynamic design considerations.
Let us assume, for example. a predetermined dynamic load of frequency no for
a structure to be designed. It might be a printing press in a building. Introducing
some modifications in mass M and stiffness k in the structure at the design stage
changes the natural circular frequency w of the structure under design from w1 to w2.
30 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

It is advisable to start with valuable differences of one of those factors in the


diverse study models. Then, after comparison of the results, some refinements can
be introduced in them. Differences can be introduced by working with notional
thickness or a priori detailing of the proposed articulations.
A justifiable higher coefficient of damping~ is also an acceptable way to improve
the response of the structure under design, provided it is chosen with care. When
trying to evaluate the damping ratio one should keep in mind that a structure is usu-
ally a combination of structural elements, sometimes of different materials, with a
variety of possible joints. This holds even for homogeneous and monolithic struc-
tures, because of components such as foundations, attached nonstructural elements,
and more.

6.2 Equivalent substructures

If the study model of the structure contains an enormous amount of information,


which makes the program too heavy to run, a simplification can be considered by
zoning the structure and proceeding step by step with substructures. The dynamic
analysis is pe1formed for a global model of equivalent substructures.

7 Superposing static and dynamic effects

In design situations for simultaneous static and dynamic loads, static and dynamic
deflections must be summed for stress and strain control, and for other design
purposes. Superposition of effects may be considered only under certain restrictions.

7.1 Linear and nonlinear design situation

Figure 2.26 depicts an ordinary structure. a beam, subjected to static load, such
as dead load, live load, temperature and more. Proper calculation should yield the
regular, very well known static deflection 8,,m.

F(t)

!
Figure 2.26: Dynamic load on a beam.

The structure under consideration is also subjected to a dynamic IoadF(t). Proper


calculation should yield a deflection o(t), variable with time, with maximal ampli-
tude Odin.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 31

Two different situations should be checked for Ii = listm + liam:


I. 8 remains under the linear elastic limit of deformation; in this case the super-
position of effects is valid;
2. 8 exceeds the linear elastic limit of deformation; in such a situation the superpo-
sition of effects is not valid. The dynamic analysis should start from the results
of the static analysis; it leads to a nonlinear dynamic analysis.

7.2 Buckling effect

Figure 2.27 portrays the same beam as in Fig. 2.26, but with an axial load N.

F(t)

~
Figure 2.27: Dynamic load on a beam subjected to an axial load.

If the beam has a tendency to buckle because of the external static load N, this
effect should be checked in advance. P-1'>. effects should be considered as part of
the stage of static analysis, as explained in section 7. l.
The effect of the axial load is also considered in the dynamic analysis by adding
the effect of the geometric stiffness to the equation of equilibrium.
Part 2

Dynamics in design
I
:
I
II
CHAPTER3

Dynamic events and effects

1 Introduction
Any dynamic structural design situation is created by an external dynamic event.
Therefore, it is of the utmost importance to become familiar with those events, and
their consequences, for due consideration in the process of structural design.
Engineers in the daily practice of structural design are well aware of the different
types of loads for static design. They are also aware of the dynamic effects of
earthquake and wind loads, since they are compelled by the Codes of Practice to
consider them in the process of structural design.
"Live loads" as defined in the Codes are time-varying loads, but their variation
with time is relatively slow and they do not excite the structure, hence their inclusion
in the static design. Sometimes prescribed live loads include a certain dynamic
factor. These "quasi-static" loads are also dealt with in a static design instead of a
dynamic design, as, for example, in the case of short to moderate span heavy bridges.
Engineers are usually less aware of the variety of other dynamic events that
can affect the structure under design after or during construction. However. proper
consideration for dynamic effects on structures is becoming more important with
the modem tendency to design slender and elegant structures. Such structures are
the result of conscious professional efforts to stress the engineer's capabilities in the
fields of design knowledge, the use of stronger materials, and the implementation
of sophisticated building technologies.

2 Dynamic events and effects


2.1 Dynamic events

In general, a dynamic event can be in the form of a vibrating load, like that created
by a working machine, or an external time-varying force, like those created by
construction work in the close vicinity, or by a major event such as an air blast.
36 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

With reference to loads and forces, the basic relation between weight, mass, and
force should be kept in mind:

A weight of 1 kg mass is 1 kg force, or


1 kg force= 1 kg m. g (acceleration due to gravity).

External dynamic loads can be produced by human activity (dancing, running,


skipping, etc.), operating machinery (turbine, printing press, forging, etc.), road/rail
traffic, etc.
External dynamic forces can be produced by construction activities (piling, blast
excavation, heavy compaction, etc.), collisions of cars or of airplanes with struc-
tures, etc.
External dynamic events can arise from ground motion such as an earthquake or
collapse of ground foundations. They can also result, for example, from the rupture
of a localized structural element.
Clearly, a wide range of dynamic events may require consideration in structural
dynamic design situations.

2.2 Dynamic effects

The main dynamic effects on strnctural design are


overs tressing,
- vibrations, and
- fatigue of structural materials.
Dynamic effects do not usually occur concomitantly. A situation of overstressing
does not always arise together with intolerable vibrations. Problematic vibrations
can appear in a structure that is not overstressed and has no notable deflections.
Material fatigue can cause intolerable damage to strnctures that until then have
performed quite well. Accordingly, the engineer should consider the effects of a
dynamic design situation separately.
Unfortunately, for some of these effects there are neither satisfactory answers
in the Codes nor sound and strong applicable rules or reliable guidelines. This
sometimes creates an intolerable design situation for the practicing engineer.
Overstressing should be considered as an ultimate limit state, while intolerable
vibrations are a matter of serviceability limit state.
The problem of fatigue should underlie the design process since it is related to
restrictions in the acceptable stress limits.

2.3 Consequences from dynamic situations

Among the consequences that the engineer should be aware of are building dam-
ages, physiological and psychological effects on people, degradation of machine
performances, and more.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 37

These effects cannot always be quantified from outputs of computer calculations,


but the engineer should take adequate preventive measures whenever possible.
This would be necessary, for example, when construction work is performed with
heavy equipment in dense building areas.

2.4 Dynamic problems

Certain types of dynamic problems like those caused, for example, by heavy
machinery close to highly sensitive laboratory instruments, or music recording
studios. are clearly problems to be dealt with through the intervention of an expert.

3 Response of structures

The response of a structure to an external dynamic event depends on three things:


the dynamic event, the dynamic characteristics of the structure, and the way the
dynamic forces reach the structure.
Any structure with predefined characteristics, such as mass distribution, stiffness
components, and inherent damping values, will respond differently for different
dynamic loads. Alternatively, changes in the structural characteristics will produce
a different response to the same external dynamic event.

3.1 Resonance

Any structure is constructed of structural elements with continuous mass, and has,
by nature, an infinite number of modes of vibrations. Any mode of vibration is
characterized by its shape and its natural circular period of vibration wn.
The first mode, with the lowest value of w, is named the fundamental mode.
Subsequent modes, in increasing order, are known as the second mode, the third
mode, and so on.
If the external dynamic load has a natural frequency Q close to any of the nat-
ural frequencies of the structure, a situation of resonance is created. Resonance
with the fundamental mode can become the origin of overstressing and intolerable
deformations.
Resonance with subsequent modes may be less significant in respect of the ampli-
tude of deformation and stresses, mainly because of damping, but it can be impor-
tant, for example, in respect of the performance of delicate instruments.

3.2 Damping

The response of a structure subjected to a dynamic load can be mitigated by increas-


ing the inherent damping of the strncture or by the addition of localized shock
absorbers, energy dissipaters, or even sophisticated external damping devices,
which sometimes extend to involve many storeys of the building.
38 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Since damping of a structure also depends on the structural material and/or


the construction details, structural response can be improved by an appropriate
structural design.
I
For dynamic loads of machine origin the structural response can be mitigated by
the choice of a suitable location for the disturbing machine.
I
I
I
CHAPTER4

Dynamic loads on structures

1 General description

Dynamic loads on structures originate from a variety of sources such as:


- human activities (e.g., walking, running, dancing, or skipping);
- working machines, inside a building or in the surroundings;
- construction work (e.g., piling, mechanical excavations or drilling, and blasting
activities);
moving loads on bridges;
car, train, or airplane accidents;
- impact loads (e.g., falling debris);
- collapse of a structural element;
wind loads, wind gusts;
air blast pressure;
- loss of support because of ground failure;
earthquake.
Some of these dynamic loads may be described as a particular function of time,
while others are random dynamic loads. The former are known as deterministic
loads and the latter as nondeterministic loads.
Loads that present a regular time variation in a large number of cycles are called
periodic loads. If this variation has a sinusoidal pattern then it is known as a
harmonic load (Fig. 4.la and b).
Nonperiodic loads can be of diverse forms, usually of short duration (Fig. 4. lc).
If the load is applied in a truly short lapse of time, dt, it is known as an impulsive
load (Fig. 4. ld). A blast load can be either an impulsive load, if the structure is close
to the focus of the explosion, or a pulse load, of short duration, if the structure is
distant enough. Nonperiodic loads can have a nonsystematic form, for example, an
earthquake (Fig. 4. le).
40 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

F[kN] periodic load

(a)

t[ sec]

F[kN] ham1onic load

(b)
t[sec]

short
F[kN] F[kN] F[kN] duration
pulse loads

(c) t[scc]
t[scc] t[sec]

F[kN] F[kN]
in1pulse non periodic load

(d) t[sec] (e)


-I---'----~

Figure 4.1: Illustrations of typical dynamic loads.

Whe11 the input for static calculations is that of a time-varying load described
by a function, the response of the structure can be obtained as a time history of
displacements, strains, and stresses.
When the input is that of a random dynamic load it cannot be described by an
analytical function. As a result, it is no longer possible, any longer, to obtain a time
history of the internal forces. stresses, or displacements. This load is dealt with by
statistical techniques. The input provides information valid only for that statistically
defined load and the structural response is evaluated by means of a nondeterministic
methods of analysis.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 41

Curiously, even an earthquake load can become a deterministic one if the structure
is designed on the basis of the accelerogram of a measured earthquake. Dynamic
loads can be considered as belonging to one of the fo11owing two categories:
1. those that can produce serious structural damage or cause collapse of the
structure;
2. those that produce problematic vibrations and minor structural damage.
Loads of the first category should be considered in the ultimate limit state design,
while loads of the second category should be included in the serviceability limit
state design.

2 Working machines
The dynamic design situation for structures intended to carry a working machine
requires full information from the machine factory, such as total weight, weight
of moving parts, frequency of movement, expected imbalance of moving parts,
and more.
As early as possible, a thorough check should be made for the possibility of
obtaining a vibration-isolated machine, which will minimize or avoid transmission
of dynamic loads to the structure. If not, the factory may provide suggestions about
the recommended way of setting up the machine in place, whether directly on the
structure or on a special base with an intermediate damping element.
Even working machines "expected to be static" can become the origin of a
dynamic load. It can arise from poor manufacture, faulty design, imperfect mount-
ing of an unusually large machine supplied by parts, defects of maintenance, etc.
In the following subsection a simplified description is given of some typical
machine dynamic loads.

2.1 Harmonic loads

The circular motion of the rotating mass M in Fig. 4.2 produces a centrifugal force
F, acting in a plane perpendicular to the axis of movement. This force increases
with increasing velocity of the rotating mass.
The machine's frequency is given in revolutions per minute (r.p.m.). For dynamic
calculations this frequency, Q, is measured in cycles per second (c.p.s.), which
yields a correlation c.p.s. = r.p.m./60. Alternatively, this frequency f is measured
in hertz and the equivalence is

Q = 2rrf

Figure 4.2c shows the vertical and horizontal component of the centrifugal force.
Because of their periodicity these fluctuating vector forces can create design situa-
tions of resonance in cases where Q is close to the natural circular frequency w of
the structure, or of the supporting structural element itself.
42 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

(a) (c)
M

(b) (d) F

Figure 4.2: (a) A machine with rotating mass M. (b) The centrifugal force of the
rotating mass. (c) Vectorial components. (d) Diagram of the harmonic
dynamic load.

The engineer's concern is usually directed to the vertical fluctuating component


of the centrifugal force acting on the beam, but due consideration should be given to
the horizontal components too. At times they can be of importance, if they introduce
significant axial forces.
High speed rotating machines can be mounted directly on a structural element
since they will not create a situation of resonance, Q >> w. Machines with a vibrat-
ing frequency Q close to the natural circular frequency of the structure w should
be mounted on dampers (elastomeric pads, springs, etc.).
Rotating machines with very large dynamic loads, especially with low mass, can
be mounted on an artificial base which supports an additional mass. This additional
mass, in turn, will be mounted on dampers.

2.1.1 ~eriodic loads


Periodic loads, arising from reciprocating machines such as pistons, compressors,
or screening primarily create an oscillating force in the structures, usually together
with a rotating force.
Even though the main force induced by the crankshaft is in a horizontal direction,
the reciprocating machines' forces are still periodic, and hence capable of creating
a situation of resonance. The general comments for rotating machines in section 2.1
may be valid for machines inducing periodic loads.
Figure 4.3 shows some illustrations of periodic load functions.
DYNAMICS lN THE PRACTICE OF SlRUCTURAL DESIGN 43

t,
CJ CJ CJ

t
'
V vV
Figure 4.3: Examples of functions for periodic loads.

2.1.2 Impulsive loads


Loads arising from hammers, punching, cutting, molding presses, and the like,
are loads characterized by very short duration. They are induced at random or at
prescribed intervals.
Loads at nonregular intervals are known as impulsive loads. Dynamic calcula-
tions can be performed with correct consideration of the momentum created by the
moving mass.
For the type of impulsive loads induced at prescribed intervals t; (Fig. 4.4), the
ratio of the length of this interval to the natural period of the structure T can be of
importance in creating a situation of resonance. This danger does not exist if the
interval ti between impact loads is long enough, and the vibration that originates in
the structure decays because of damping, since no cumulative process of amplitudes
of vibration is created.

,, l;

Figure 4.4: Regular impulsive loads.


I
l
44 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN I
Some machines of the type reviewed here also have significant oscillating masses,
!
which also give rise to oscillating loads.

3 Vehicle loads

3.1 Collision

Usually a design situation where a moving object of mass M1 strikes a structural


element of mass Mz, for example, when a car collides with a column, a crane with
tilting load overturns, or a structural element collapses, is named impact.
The impact is a dynamic event that happens in a very short measure of time,
to. During this event the moving mass M1 strikes mass M2 (usually at rest). I
The kinetic energy of the moving mass gives rise to the force of collision Po. I
Fracture of material, together with plastic and elastic deformation, originates in
the areas of contact, with transient velocities v1 and v2 in the moving masses, up
to the moment when the entire event comes to rest. A mathematical elaboration
I
,[

~1
of the event includes equations of equilibrium of kinetic energy and momentum.
A differentiation is made between hard impact and soft impact, as a function of
the elastic or plastic characteristics of the event, and as a function of the structural
materials involved (concrete, steel, wood, etc.). J
Because of the randomness of collision events, Codes of Practice, Norms and
Guidelines prescribe quasi-static loads which represent the normative forces for
particular design situations. These forces should be taken as a minimum. The prac-
I
ticing engineer can try to foresee the possible effects of a particular collision event
by creating a design situation with impulsive loads, as described in Fig. 4.5.
I
Figure 4.5a represents the time history for a soft, highly plastic event, such as I
a crane with tilting load overturning. Figure 4.5b depicts the time history for a
relatively elastic dynamic event, like a brief accidental punch of a truck into an
elastic steel beam. Adequate values of td can be chosen, as an initial trial, such
that Id ::::; ~ T, where Tis the natural period of the fundamental mode of the struc-
ture; Pd is the weight of the moving load W multiplied by a factor::::; 125%; this
factor accounts for the unknown velocity of the moving load and the elastoplastic
deformations in the area of collision.

F F

(a) -j-..,.-1.---t> (b) -1-,.---'----t>t


+-14 ~
Figure 4.5: Time histories of impulsive loads. t
I
lj

I
;&:
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 45

3.2 Translational moving loads

The main characteristic of this type of loads is that they are concentrated loads of
a constant value, but in a position that changes with time (Fig. 4.6).

F
------?>direction of
1novement
Fconstant

Figure 4.6: A moving load along a supporting structure.

This category ofloads differs from working machines' loads. The latter concerns
loads varying with time, a situation that arises from a moving mass in a prefixed
location.
Typical translational loads are cars, trams, and pedestrians traversing a bridge
with a certain velocity (Fig. 4.7).

(a) ------?>direction of
~~carload move1nent

~
(b)

Figure 4.7: Translational load.

The dynamic effect originates in the velocity of movement. Aracecar traveling at


very high velocity can induce dynamic effects similar to resonance, but, of course,
no steady-state effect can be induced. An extreme case is that of a mass moving
very slowly, for example, crawling; this clearly will not produce a dynamic effect.
Figure 4.7a shows a simple supporting beam for a car bridge. In Fig. 4.7b the
envelope of deflection is at the center of the beam for any position of the moving
load. The unbroken line depicts the deflection under a static load, and the dotted
lines depict the fluctuation of this deflection owing to the dynamic effect.
46 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Prescribed live loads in Codes or Guidelines usually include this dynamic mag-
nification factor (DMF) for design of regular bridges. A thorough check of these
loads should be performed for unusual and lightweight structures.
The dynamic effect of translational motion thus increases with bouncing and
spring of the car masses or with haphazard factors such as uneven road surface,
bumping, wrong details of bridge joints, and more.

4 Human activities
The anecdote of the bridge that collapsed because of the resonance effect of the tread
of a column of soldiers marching across it has regaled generations of engineering
students as an introduction to structural dynamics.
Still, the Codes consider human activity load simply as a live load, albeit with
some special consideration of the marching pace. Increasing attention to vibrations
can be found in the Codes, but the structural engineer, responsible for the stability
of the structure as a whole and in its parts, should also be mindful of the dangers
of overstressing that can arise from dynamic loads caused by human actions.
Examples of sources of dynamic loads are the following very common human
activities:
fast and energetic walking or running across a pedestrian bridge;
rhythmic jumping, as in the gym;
- body bouncing and swaying, as in the stands of a stadium.

4.1 Walking and running

The range of frequencies of these activities is relatively broad. Most lie between 1.5
and 5 Hz, which is also the range for a variety of structural components. Accord-
ingly, some dynamic magnification of the so-called "static live load" can appear.
Attention should be paid especially to very slender steel structural elements since
their damping value is very low, of the order of2-5%o, whereas for regular structures
in concrete it is usually more than 1o/o.
It may be illuminating to trace changes in length, time load function, and fre-
quency of steps of the translating motion of a single person along a pedestrian
bridge (Fig. 4.8).
The forcing function for normal walk has three stages: landing with the foot-
heel, pushing off for the next step, and, in between, bouncing of the pedestrian
weight. The forcing function for running has only two stages: landing and pushing
off. The length of the step increases with increasing velocity of movement, while
decreasing the time of contact with the walking surface.
In Fig. 4.9a the walking movement is characterized by a continuous contact with
the walking surface, and hence a superposition of the steps forcing function. In
Fig. 4.9b the running movement is characterized by a noncontinuous contact with
the ground.
DYNAMlCS IN THE PRACTICE OF STRUCTURAL DESIGN 47
(a) F
l.O -Go
easy walk
;:::; 50 cm' length
t of step
.,,__,. _,__...,_
heel toe
ta

(b) F

1.0 fast \Valk


~90 to 100 cm'
length of step
t

(c) F

LO - - -Go running
~1.so to 2.00 m'
length of step
t
tc << tb
_,___.,_
Figure 4.8: Forcing function of one footstep.

Walking and running are time-load functions that can have frequencies close to
that of the structure. Nevertheless, a steady-state forcing situation is not likely to
be created since pedestrian movement does not load a fixed position, and it lasts
only till the moment the pedestrian leaves the walking area on the structure.

F F

/'
\ r-

(a)-+--~~---~~~-!>

T T

Figure 4.9: Periods of pedestrian translating movement.

If, for example, many pedestrians are crossing a bridge at the same time, the
deformation originated by one person should increase. It is very simple to conclude
48 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

that if two people are walking side by side in step, the deformation should be mul-
tiplied by a factor k = 2. If many persons are walking along the bridge at the same
time, they try, subconsciously and as a natural reaction, to accommodate their pace
to the vibrations they feel under their feet. As a result, individual-originated defor-
mations will be superimposed, but not linearly, because of perturbations induced
by walkers of different sex and weight arriving at random. The square root of the
number of pedestrians walking at the same time on the bridge may well represent
an adequate value for the factor k mentioned above, but it should be confirmed by
new researches.

4.2 Gymnastic loads

Skipping or jumping on the spot can originate dangerous overstressing because of


resonance if the gymnast, for example, jumps on a slender rib supporting the floor.
An approximate forcing function may be in the form shown in Fig. 4.10.

1.0

,y Id ,y ,y t d ,y
T

Figure 4. I 0: Forcing function for skipping or jumping at a fixed location.

According to the aforementioned upper limits of frequency, the period T can


be close to S Hz, and the forcing load between three and five times the gymnast's
weight. If many gymnasts are jumping together, the deformation should be multi-
plied by their number.

4.2.1 Crowded areas


Live loads laid down by the Codes for crowded areas usually cover actually static
situations.
Even so, dynamic problems can arise in particular cases, such as dancing areas,
grandstands, spectator areas at pop concerts, and the like. People moving, jump-
ing, clapping hands, and swaying rhythmically at frequencies perhaps close to the
natural frequency of the structure do induce dynamic magnification factors, which
can reach 100%.
DYNAMICS IN THE PRACTICE OF STRUCTIJRAL DESIGN 49

For these design situations three more types of loading should be considered:
live load increased by a certain DMF;
swaying horizontal loads, in both in x and y directions;
resonance problems, when the structure has a natural frequency ranging from
I to5Hz.

5 Construction loads
Construction work can become a source of dynamic loads and the origin of direct
or indirect structural damage such as cracks, foundation settlement, towers out of
vertical, etc. Damage can also be caused to nonstructural building elements, such
as partition walls, ceiling, windowpanes, and more.
Structural damage caused by construction loads or transient construction con-
ditions does not usually include total collapse of a structure except by accident.
The reason is that the vibrations that originate owing to these operations are kept
under certain limits in order to avoid both human annoyance and cosmetic build-
ing cracks. The main problem is that since these damages are not always detected,
safety factors are diminished.
Low intensity vibrations of long duration are characteristic of the operation of
site construction machinery. Vibration-induced structural fissures cause a reduction
in stiffness, which eventually leads to the risk of resonance, resulting in further
widening of these fissures.
Construction loads can be divided in two different groups: those taking place
outside the boundaries of the structure and those inside the building.

5.1 Site construction

Construction work performed outside the boundaries of the structure include, for
example, excavation with heavy equipment, buried blasting operations, heavy com-
pactors, heavy hammer piling, performed at building sites close to existing struc-
tures. Such operations, when performed nearby, say within 50 m, should be checked
in advance. If disturbances are foreseen, restrictive rules are imperative.
Ground motions that originate as a result of construction operations produce
three types of waves: longitudinal or compressive waves L, shear waves S, and
surface waves R. These waves, characteristic of ground motions, differ from waves
radiating from an open air event, such as an air blast, where one single type of wave
radiates spherically, provided there are no atmospheric perturbations.
In ground motions, the motion of the waves should be distinguished from the
motion of any single ground particle. Movement of ground particles does not always
accompany the wave movement, and in certain cases they can even move in opposite
directions.
To visualize the difference between wave and particle motion, let us imagine a
stone dropped into a pool of water with a dirty surface: more or less circular waves
50 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

will move away radially from the focus, but at the same time, dirty spots, which
represent single particles, will move vertically.
The S waves reaching the building at low frequency can create a state of resonance
with the vibrating subsoil. This dynamic situation can cause damage to the structure.
Dynamic forces propagate in the ground as elastic waves; particles' motion and
ground strains accumulate energy, which is transmitted to the building elements.
This energy can be calculated using ground wave characteristics such as length,
amplitude, and velocity. Unfortunately, construction workloads are of random char-
acter, and despite extensive and experimental measurements on site, suggested
parameters are useless for structural calculations, and the engineer should rely on
expert advice, guidelines, recommendations, and Code regulations.

5,2 Building construction

Operating equipment, lifting devices, transport of materials or machinery com-


ponents, heavy percussion equipment, careless or accidental dropping of building
components, friction forces, and many more of the like can create an unforeseen
dynamic building situation. Incorporation of prefabricated elements into an ordi-
nary construction system may we11 become the source of dynamic forces.
Plastic deformations and transitory opening of cracks, not always detected,
degrade the structure's designed strength.
CHAPTERS

Vibrations

1 Introduction
The vibration of structures or the vibration of structural elements is a frequent
phenomenon.
The appropriate engineering knowledge is still only evolving, and its practical
application in structural design is limited to a few cases. Uncertainty in basic param-
eters of design sometimes renders computer calculations insufficiently reliable.
Vibration of structures is mainly an issue of serviceability. Only in particular cases
can they be the 01igin of structural damage, such as structural cracks or fatigue of
materials. Usually the amplitude and frequency of vibration, which people find
annoying, malfunctioning of machines, out of balance laboratory equipment, or an
irritating noise level are smaller than those required to produce structural damage.
The engineer is increasing!y called upon to pay proper attention to various aspects
of the problem of vibrations since the consequences may be serious complaints
about the finished structure or an extreme human response, such as anxiety and fear,
up to the point of people refusing to use the building, the bridge, the observatory
tower, or the like.
The issue of vibrations will be considered from three different angles for the
practice of structural design:
1. vibrations,
2. the source of vibrations, and
3. vibrations in the process of design.

2 Vibrations of structures
Modern structures are usually stronger, of lighter mass, and more slender than in the
past. This is the result of a combination of three factors: stronger structural material,

I
modern building technologies, and sound knowledge of engineering design.

1
.I
52 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Such structures tend to vibrate under dynamic effects, at times to the point of
impairing the expected performance. These vibrations are not recorded on massive
structures.
Complaints by end users may start in respect of vibration levels lower than
those that cause fear or damage. Their origin may be because of rattling objects,
sound levels, affected sensitive equipment, cosmetic cracks, annoyance, and many
more.
Human levels of sensitivity to vibration are lower than the level required to
produce structural damage. Unfortunately, it is not always possible to take human
sensitivity as a limiting criterion. This is attested by comparing the sensitivity of a
worker in a factory with heavy vibrating and noisy machinery with that of a clerk
sitting at an office desk, on a very light floor. The latter may be disturbed merely
by vibrations arising from normal human activity around him/her.
Footbridges are typical examples of structures where annoyance, disquiet, or
even fear can be produced by vibrations that are not of the type that can produce
physical damage.
Annoying vibration levels are different, for example, in hospitals as against apart-
ment blocks or offices.
Acceptable vibration levels are different for aparttnent blocks in isolated resi-
dential areas compared with those on busy city thoroughfares, and so on.

3 Sources of vibration
Three main categories of sources can be established according to the type of vibra-
tion under consideration.

3.1 Dynamic vibrations that can be foreseen

Human activities, working machines, musical events, etc. This type of vibration
can be defined up to a limited degree of reliability, in terms of the frequency
and amplitude, as required in the process of design. Thereafter, adequate counter-
measures can be considered, whenever necessary, such as tuning of frequency or
damping, or changing some characteristic of the sources, such as machinery location
or affixation.

3.2 Vibrations originating at construction sites

These vibrations are due to construction operations, heavy machinery excavation,


hammer piling, compaction, deep ground blasting, percussion machinery, etc.
Here the source can be clearly defined, and physical parameters such as amplitude
and frequency of vibration, energy of impact, and so on can be established. How-
ever, the ground waves that are created propagate through an uncertain medium.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 53

Therefore, it is impractical to try to calculate countermeasures mathematically.


Instead, well-defined rules for vibration levels are given in the literature.

3.3 Dynamic vibrations at random

Overpressure from air blast, heavy lorries on bumpy roads, collisions, and so forth,
generate vibrations at random. Normative regulations are usually applicable.

4 Vibrations in the design process


Vibrations in the stage of structural design may be classified into three categories:
1. vibrations that should be considered in the design process. according to the
requirements of the Codes of Practice;
2. vibrations that can be foreseen in the design process as their sources are known
to the engineer;
3. vibrations that cannot be included in the calculations but should be kept in
mind towards the construction phases.
Dynamic loads can induce vibrations that are not dangerous to the structure, but
are nevertheless annoying or even intolerable to human sensitivity, and that can
be mathematically described, such as those caused by human activity or machine
operations. They may be introduced into computer calculations in the design phase.
The magnitude of the induced vibrations depends mainly on the ratio between the
natural frequency of the structure, or the structural element, and the dominant fre-
quency of the excitation. If the excitation frequency ranges approximately between
2 and 5 Hz, it is advisable to try to maintain the fundamental mode of the structure
above this range.
Yet certain structures, such as light floors or pedestrian bridges, may have fre-
quencies close to that of the excitation frequency. Any attempt to change them
drastically, whether through their mass or structural rigidity, may well prove imprac-
ticable for esthetic, functional, or economic reasons.
In such cases the engineer should consider the possibility of increasing the damp-
ing characteristic of the structure. A naked steel beam, for example, with friction
free supports, can have a low damping coefficient, around 2%o for steel and 5%o for
concrete, while for a real structure this coefficient can reach 10 times this value. The
real characteristics of the structure sometimes involve a combination of more than
one st.Iuctural material, attached secondary construction elements such as flooring,
heavy partition walls, railings, and so on. Each of these components contributes to
increasing the amount of damping of a real structure.
Vibrations of the kind considered here belong to the limit state of control of
serviceability. Hence calculations performed for elastic response should exclude
damping values arising from a plastic or even an elastoplastic deformation of the
material. Values of damping of 10% or even more may be included only forultimate
capacity control.
54 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

4.1 Normative magnitudes of vibrations

Normative values of displacement, velocity, or acceleration of induced vibrations


are very few. Therefore, the key question of whether vibration effects are admissible
or not, remains open.
Some national Codes prescribe certain limits for vertical acceleration in pedes-
trian bridges or displacement and velocity of measured induced vibrations in build-
ings. These values are empirically based, so some relatively wide discrepancies can
be found among them.
Nevertheless, the engineer is obliged to work according to these limits when they
are established by the norms or the directives contained in his or her job assignment
conditions.

4.2 Tolerable magnitudes of vibrations

An extensive and large number of vibrations on structures are covered only partially,
or not at all, in the Codes. Still, the engineer is required to pay attention to them
in the design process because of the intolerable consequences that may arise from
those perturbations.
Some of the dynamic actions that produce these perturbations can be mathemat-
ically defined, at least as required for input in computer calculations. The problem,
however, lies in the difficulties besetting a correct appreciation of the displacement,
velocity, or acceleration response values.
Any attempt to define acceptable levels for the response characteristics should
consider closely human response, nonstructural damage to buildings, snags in
machine operations, and so on; none of these can be quantified in absolute values.
In general, the effect of vibration can be estimated through its amplitude and
frequency, whose eftect proves to be in inverse relation: a similar eftect will be
produced by increasing amplitude with decreasing frequency.
The practicing engineer should consult national and international Codes, and
institutional regulations and guidelines on acceptable levels of vibrations.
For harmonic vibrations, a formulation may be given for pseudo-velocity or
pseudo-acceleration
2
v = 2rr · f · d or a = (2rr · f) . d

These relations can be of some help in an effort to estimate the response


characteristics.

4.2.l Human tolerance


Value limits are usually given in terms of displacement or accelerations, as a function
of the frequency of vibration.
Tolerance levels are physiologically or psychologically based:
physiological: these refer to a sense of well being, induced fatigue, or health
preservation.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 55

psychological: these refer to human perception and sensitivity, which vary with
expectations and diverse surrounding conditions, for example:
busy commercial districts;
industrial areas;
schools, hospitals, and libraries;
residential areas.

4.2.2 Acceptance levels in buildings


Acceptance levels for buildings are adjusted to minimize damages, usually secon-
dary damages in structural and nonstructural elements.
The issue here concerns damages that may be considered mainly as irritants, not
those that endanger the skeleton of the building. A notable exception is vibration-
induced fatigue, which should be dealt with separately.
The problem is to establish acceptable vibration levels to avoid damages
such as
cracks,
widening of existing cracks,
spalling of wall plaster,
detaching of affixed nonstructural elements, and
falling of cladding components.
Vibration levels are formulated as a function of the type of skeleton, the quality
of the building, and the construction system. Vibration levels are usually based on
observations and measurements of experimentally induced vibrations. Hence the
limiting values cannot be considered absolute, but are 1nainly intended to give the
order of the expected effects.
Vibration velocities, with a certain dependence on their frequency ranges, are
commonly used to establish acceptance limits for building damages. It has been
found that the effect of the magnitude of displacement is a function of its velocity,
and differs for different sources of vibration.

4.2.3 Acceptable vibrations for particular cases


Tolerance levels of vibrations in machinery and equipment cannot be defined as
pertaining to a common family.
No universal acceptance levels for vibration characteristics can be defined for:
sensitive laboratory apparatus,
machines for high precision products,
precision line production machinery,
material or product control and calibration,
laboratory balances, etc.
The practicing engineer is entitled to obtain, well in advance in the design stage,
a clear specification of the expected limitations on the vibration characteristics,
displacement, velocity, acceleration, and frequency that are expected from the
building.
!
I
Part3

Structural materials in
dynamic situations
CHAPTER6

The rate of strain

1 Introduction
This chapter deals with the material properties of structural elements subjected to
dynamic loads, in respect of reinforced concrete, prestressed or not, and structural
steel.
The properties of structural materials change in dynamic situations, as compared
with these characteristics under slow and quasi-static loads. Some of these changes
should be taken into consideration in the design process.
The main topics involved in such considerations are:
I. The rate of change with time of the superimposed load, q = dq/dt. This affects,
for example, the strength of steel and the strength of concrete and its modulus
of elasticity.
2. The number and amplitude of fluctuations of the superimposed load, during
the design lifetime of the structure. This can reduce the ultimate capacity of
the structure through fatigue.
3. Damping: the capacity of structures and structural materials to absorb energy,
and hence mitigate the effect of external dynamic events.
Not all dynamic situations so affect the material's properties as to be included
in the design of structures. Normal buildings, bridges, and other structures, whose
design is governed by rules set down in the Codes, are in fact usually designed for
quasi-static loads. For the most part, these Codes also include particular consider-
ations for dynamic situations.
The design of steel structures in dynamic design situations is supported by sound
and extensive knowledge. For concrete this knowledge is still limited, largely
because steel is a more homogeneous material. The fatigue mechanism, for example,
is very well known for steel, while it is more complex for reinforced concrete since
it depends on the interaction of many parameters such as adherence, reinforcement
protuberances, crack distribution, and more.
60 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Since the issue of structural materials in dynamic design situations is still


evolving, the practicing engineer should carefully peruse guidelines and special
publications dealing with the material properties and resistance capacity of struc-
tural elements.
Codes for concrete structures are very well worked out for quasi-static loads, but -
still lack a comprehensive approach to dynamic design situations.

2 The rate of loading


Practicing engineers are familiar with structural material properties such as the
modulus of elasticity, stresses at ultimate limit, the yield limit, and elastic and
plastic strain, since they appear in day-to-day static calculations. It is of interest
to learn how these properties change with the rate of loading of a structure. This
knowledge also enriches the understanding of the response of structural materials.
In design practice, live loads are usually considered static loads, because their
rate of loading qis usually measured in minutes, while dynamic loads are measured
in seconds or fractions of seconds; impact events occur in even shorter lengths of
time.
The choice of the yield limit as the criterion for acceptable stresses in structural
design is based on the knowledge of yield stress, derived from static experiments.
This limit differentiates between the elastic or plastic response of a structure.
Plasticity of structural materials is a phenomenon by which the material flows
under load. This flow is sensitive to the strain rate, 6 = dc/dt, whose measure-
ment is considered in laboratory experiments as an adequate reference for checking
changes of material properties under dynamic loads. The strain rate affects the plas-
tic behavior of a structural material by increasing the yield limit and the stresses
above this new yield limit. This holds especially for mild steel, which is a type of
structural steel. The strain rate has a similar effect in concrete. In general, load-
ing at high rates causes a strengthening effect, but at the same time the material
becomes more brittle. The fracture strain decreases with increase in the rate of
strain. Accordingly, the plastic response should be confined whenever possible to
low percentages of the material's known plastic elongation limit.
The strengthening effect starts at a relatively low strain rate, of the order of i; = 1
[per second], which is typical for some cases of structures subjected to dynamic
loads.

3 The rate of strain


The engineer himself or herself can get an initial sense of the "rate of strain" if
he or she makes an approximate calculation of the velocity of a cyclic elongation,
as illustrated in Fig. 6.la: a steel hanging rod carries a load Q applied suddenly
in a length of time t at its bottom. Any transverse section is subjected to an axial,
uniform distributed tensile stress.
I
I
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 61

(a)

(b) (c) (d)

!::. est
Q

Figure 6.1: Cyclic elongation.

The well-known formulas are:

Q A·E
k=-=-
1:>.l l

1 2rr
T = 2rr (M = 2rrJ g_ · ; w=-
VT g A·E/l T

t>.l,, is the total elongation of the rod, measured at the point of application of the
external load Q, for a static situation, while s is the elongation, originated by the
same load, in any unit of length of the rod; hence both can be represented by a
similar vector. Figure 6.1 b shows the static strain s,,.
If load Q is dropped into
its position, the strain will grow dynamically, as shown in Fig. 6. lc, and it will
fluctuate between two extremes, as shown in Fig. 6.ld.
If we neglect damping, which can be accepted as a legitimate approximation for
the beginning of the movement, one cycle of the strain fluctuation is represented
in Fig. 6.2a, which is a "sinusoid." The velocity of the motion is given by the
derivative, which is again a "sinusoid," as can be seen in Fig. 6.2b. Figure 6.2c
shows a part of the strain variation taken from ·Fig. 6.2a, and Fig. 6.2d presents a
linear approximation of the velocity diagram from Fig. 6.2b.
Let us assume that the part of the strain indicated in Fig. 6.2c is, with good
approximation, proportional to Bst. and l:J,.t, also with good approximation, measures
the length of time of occurrence of this motion. In this case we can calculate the
rate of strain of this event.
For l = 250 cm, A = 5 cm2 , Q = 7.5 t, Sst = 0.0007, T = 0.085 s

~ Sst 0.0007
s= - = = 0.05 [per second]
t>.t 0.085/6
62 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

''

--: f-, \ ~:_ __ _


,'
(a) (c)

(b)

Figure 6.2: Strain fluctuation cycle.

The mathematical solution, for the sake of comparison, is

£d =&st· DMF

DMF = I - cos (w · t) = 1 - cos ( 2n ~)

and
. d 2Jt . (
£d =&st· dtDMF =Est· T · sm 2ny:t )

When the sine function in the DMF becomes unity, the rate of strain becomes a
maximum:
2Jt
fd,m'"' = &st • T = 0.05298 [per second]

The order of the rate of change in structural dynamic situations is usually similar
to that in the calculated example. In extreme dynamic situations, such as an impact
or a blast, the order of the rate of change can be hundreds of times higher.

4 Structural materials
4,1 Concrete subjected to high rate of strain

The compression strength of concrete increases remarkably with increasing rate


of strain. The well-known limits of ultimate compression strength of concrete for
static design situations are controlled in laboratory tests for axial loads applied at
a very slow rate of strain, £, of the order of 1 x 10- 4 per second. These values
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 63

test at high
strain rate

standard test

strain<;[%]

Figure 6.3: Effect of the rate of strain in the stress-strain relation.

of ultimate compression strength are almost doubled (around 175%) for dynamic
design situations with rate of strains of 1-10 per second (Fig. 6.3).
The tensile strength of concrete is a different issue. The tensile stress-strain
diagram is usually established in flexural tests instead of axial tensile tests. In both
alternatives the results should be considered with caution. The specimen cracks at
the weakest point in the material. No internal bridging mechanism is known, which
is most probably the reason for the considerable diversity in data collected from
experiments. In bending, stresses and strains are linearly distributed between the
bottom fiber and the neutral axis, which makes interpretation of the results still
more complex.
No strengthening effects due to the rate of strain can be taken into account for
the ultimate tensile capacity of concrete.
The modulus of elasticity of concrete E =Young's modulus is likewise affected
by the rate of strain. In dynamic design situations an increase of 10% in the value
of E can be considered.

4.2 Structural steel subjected to high rate of strain

For reinforcing steel bars and regular grades of structural steel, the yield stress
with increasing rate of strain increases in dynamic design situations up to some
25% more than the value for a slow quasi-static rate of loading (Fig. 6.4). The
modulus of elasticity and the ultimate capacity do not change with the rate of
strain.
For high-strength prestressing steel no noticeable changes are registered in the
stress-strain diagram or in the modulus of elasticity with change from a slow to a
high rate of strain.
64 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

1 standard test
2 test ay high
rate of strain

2 I

strain, E

Figure 6.4: Effect of the rate of strain in the stress-strain relation.

S Structural elements
Reinforced structural elements are the combination of concrete and steel bars. The
bonding of the two is essential to assure adequate performance. In a beam subjected
to a variable moment along its axis, changes in the tension force in the bar depend
on the bonding mechanism. In columns subjected to centric compression loads, the
bonding mechanism assures its proper distribution between the two materials.
To benefit from the increasing values of concrete compression strength and the
yield limits of reinforcement for increasing rate of strain, it is essential to ensure
overall good bonding conditions in the structural element.

5.1 Reinforced concrete columns

Columns that are centrically loaded, and subjected to a high rate of strain. present
the designer with a dilemma. The loading capacity of a column loaded axially is
given essentially by the compression capacity of the concrete plus the yield capacity
of the reinforcing bars:
Pu,column = Pcu + Psy
This means that Pu increases with increasing rate of strain. But the designer
should be aware that this rise in the column's ultimate capacity can impair the
equilibrium conditions of buckling.

5.2 Bending at high rate of strain

Structural elements subjected to bending, such as beams or slabs, benefit from the
increasing resistance capacity of materials under increasing rate of strain.
A common practice is to design structural elements for bending as under-
reinforced structures. This ensures a ductile mode of failure, which means that
overloading to ultimate capacity will produce a noticeable increase in deflection

l
l
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 65
(a) A

(<--?'--, -'11 ! i f
A
I
(b)
rct
(c)
es
(d)
rs

e~ strain in concrete
on modulus of rupture

eE strain in the
plastic stage
(e)
f! f
E, (f) f,y (g) E,y

Figure 6.5: Diagrams of progressive stresses and strains in a flexural beam.

before collapse. This mechanism, characteristic of the ductile mode of failure, sig-
nals a warning before disaster.
The process of increasing stresses in a section of an underreinforced element
subjected to bending is well known to the practicing engineer. The beam undergoes
successive stages with progressive increase of the external load (Fig. 6.5).
Figure 6.5a shows part of a beam from section A-A to the left, subjected to
bending. Figure 6.Sb is a diagram of elastic stresses at stage I: there are no tensile
cracks. Figure 6.5c is the strain diagram for stage I. Figure 6.5d shows the elastic
stresses at stage II, characterized by tensile cracks. Figure 6.5e is the relevant linear
strain diagram. Figure 6.Sf is the diagram of stresses in the plastic stage.
The internal moment of forces is given by the integral of stresses in the concrete
plastic compressed area versus the integral of tensile stresses in the reinforcement.
Figure 6.5g is the strain diagram at ultimate capacity, consistent with the principle
that plane sections, like section A-A, remain plane throughout the process up to the
flexural failure of the beam.
Underreinforced flexural elements subjected to an increasing overload approach
failure with an increasing deflection and an almost constant tensile force in the
reinforcement, stressed at yield. The required increase of moment of the internal
forces is attained by an increase of the internal lever arm. This is caused by raising
the neutral axis and increasing the compression stresses towards full plasticity of
the concrete.
Clearly, underreinforced structural elements subjected to bending at increasing
rate of strain benefit from the increasing compression stresses of concrete and from
the raising of the yield limit of reinforcement. The result is an increasing moment
capacity under high rate of strain.
Rigidity at bending: For certain types of dynamic design situations, the engineer
wishes to keep stresses in the elastic range, as close as possible to the modulus of
rupture.
66 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Since the limit for elastic strain at tension is a random issue, as already noted it
is advisable to test two alternatives:
I. A section in stage I, with a modulus of elasticity (E =Young's modulus) for a
high rate of strain. E can be taken as some 10% more than its value for a static
design situation.
2. A section with cracks, namely stage II, and the adequate moment of inertia and
the regular value of the modulus of elasticity.
CHAPTER 7

Fatigue

1 The phenomenon

A structural element in a dynamic situation can fail at a stress level lower than that
of the known stresses in a static design situation.
This event is known as a fatigue failure. It originates in the deterioration of the
structural material caused by the live loads because of their variable nature. It is a
long-term, progressive type of failure.
Fatigue failure can take place in steel and reinforced concrete structures alike.
The latter naturally includes prestressed elements.
1\vo characteristics of the changing loads are owing to the reduced resistance
capacity of the structural material: ( 1) the amplitude of the dynamic load at each
cycle and (2) the total number of those cycles, regular or not, during the lifetime
of the structure. If the amplitude, or the number of cycles, is relatively small the
fatigue phenomenon will not affect the structure's resistance capacity, as explained
later.
The amplitude of changing stresses in the fatigue phenomenon is not related to
the rate of strain.
The lifetime of a structure is its designed working life. It depends principally
on the type or nature of the structure. It may be a temporary structure with an
expected life of no more than 5 years at one extreme, or one of 120 years for a
civil engineering structure like a bridge, at the other. A structure's lifetime can be
extended if a structural element susceptible to fatigue is replaced by a new one at
the right time. For the lifetime of particular structures the engineer should refer to
the Codes of Practice.
The allowable amplitude of variable stresses and the correlated number of dynamic
cycles for structural materials are determined for an unlimited life criterion.
In a structural element under static loads, the stress regime, loosely speaking,
is permanent. The stresses change temporarily whenever an external live load
68 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

is superimposed. These changes may show different patterns; some examples are
shown in Fig. 7.1.

time time
(a) oscillant (b) impact

amplitude with
changing signs
I
time time
(c) irregular (d) alternate

Figure 7.1: Dynamic live load patterns.

The state of fatigue in a structural material should be seen as resulting from the
cumulative effect of the stress changes that occur during the structure's lifetime.
This cumulative process refers only to amplitude :fluctuations with stresses above
a certain value, known as the endurance limit.
The structural material "remembers," so to speak, the different relevant dynamic
events that have taken place throughout the years, and "adds up" all of them (Fig. 7 .2).

········--Pl +e> --oj

time
(years, 1nost probable)

Figure 7.2: Along-time addition process of stress changes.

Fluctuation amplitudes with stresses below this limit do not cause fatigue. The
endurance limit can confidently be established for structural materials (steel, con-
crete, reinforcing bars, or prestressing steel), as well as for reinforced concrete or
steel structural components. The endurance level of stresses can be established by
tests for constant amplitude.
Figure 7 .3 represents the curve a as a function of n, with reference to the
endurance limit. If aM, the maximum stress of the amplitude of fluctuation is below
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 69

endurance
lin1it

Figure 7.3: Endurance limit of stresses.

the endurance limit, and the number of stress fluctuations n tends to be unlimited,
no fatigue damage will be registered.
If aM is greater than the endurance limit, there is a particular number of stress
fluctuations that causes fatigue failure of the material. This number na is a function
of the amplitude a of the fluctuation stresses:

where "M is the lower limit of fluctuation. The latter can be of the same sign of
"M' or not luml :':: luMI·
Three types of fluctuation can be identified. Figure 7.4a shows the oscillating
cycles between two stress limits, O'M and am, both of the same sign, whether tension

±o

OM

Om I· n
(a) oscillaiting cycles

±o

OM
(b) repetitive cycles
a
n

o
(c) alternating cycles
a n

Om

Figure 7.4: Fluctuation loads.


70 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

or compression; Fig. 7 .4b shows the repetitive cycles, when am = O; Fig. 7 .4c shows
the alternating cycles, when the limits are of different sign.
For each type and for different ranges of amplitudes of vibrations, the number
of cycles differs in the failure of any structural material. The maximum allowable
stress is a function of the initial stresses as given for the static permanent loads, on
top of which the amplitude and number of fluctuations are superimposed.
As stated, during its designed working life a structure can be subjected to a
random succession of fluctuations, each of different amplitude, for example, a 1 ,
in a number of fluctuations n~ smaller than n i, which is the failure limit for this
particular amplitude.
The cumulative effect of these random fluctuations may be expressed as

where k must be less than unity to avoid fatigue failure.


This adding-up process is done without discrimination of type, amplitude, or
sequence of these fluctuating stresses. Each adds its own relative value to the total.
In practice, not all fluctuations that a structure undergoes during its design life
are above the endurance limit. For example, a bridge with a high live/dead load
ratio will probably not always be in a situation of full live load. Hence not all
amplitude fluctuations will have a maximum stress above the endurance level, so
only a certain percentage of them will be considered in the fatigue design.
The fatigue phenomenon relates, basically, to deterioration of the structural mate-
1ia1, even though the variable stresses and strains change in the range of the elastic
response. Deterioration of the structural material is accelerated by plasticity. There-
fore, distinctions should be made when one is faced with the possibility of plastic
fatigue. Plasticity should be avoided, for example, when one designs a beam to
support the oscillating load produced by a machine.

2 Fatigue in structural materials


2.1 Concrete

Endurance levels of stresses in concrete are established in experiments done on


laboratory specimens subjected mainly to compression, namely experiments for
repetitive or oscillating loads.
Since neither the strain rate nor frequency is a factor, all cycles of loading are
conducted to seek the number of loadings n until failure at any given amplitude, a,
of fluctuating stresses.
One loading is conducted from zero to failure to establish the ultimate capacity
of a specimen,fcu. If the same specimen is subjected to loading from 0 to 90% of
fcu only, the specimen will fail after some 100 repeated cycles. If the amplitude of
the loading cycles is from 0 to less than 50% offcu, more than 1,000,000 cycles
will be required until failure.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 71

An interesting situation arises when the same specimen is submitted to an oscil-


lating load of amplitude

The number of loading cycles until failure n increases with decrease of maximum
stress aM and decrease of amplitude a.
The reference made above to "the same" specimen is merely a figurative form of
expression. In practice, an adequate number of concrete laboratory specimens are
prepared in advance, all of the same age, curing, and ultimate capacity. Experiments
on concrete specimens should consider quality of concrete, age, curing, ambient
humidity, frequency of loading cycles, and more.
Last but not least, an interesting finding from these experiments is that if the
amplitude a of the loading cycle is kept very small, the limit value of fatigue does
not tend to 100% of/,u, as might be expected, but only to 85%, which is the known
value of ultimate compression stresses in concrete under Jong-term, permanent
static load; the concrete becomes "tired" with time.

2.2 Reinforcing bars

As with concrete, the amplitude of stress fluctuation and the maximum stress of
this fluctuation define the number of cycles the bars can undergo before fatigue
failure. They differ slightly with smooth bars of different grades. For ribbed bars
this difference becomes immaterial.
Irregularities in the round surface of the bars, with the exclusion of superficial
oxidation film, scales, or flaxes, become the natural location for the initiation of
cracks, so that the area around those ribs are prone to fatigue cracks. An unusual
situation is created by those ribs, which is that the endurance limit is almost the
same for the different grades of reinforcing bars. An amplitude

of some 150 MPa can be considered the endurance limit for usual grades of rein-
forcing ribbed bars, which is the amplitude of stresses the bar can stand, without
failure, for an unlimited number of cycles, including sometimes alternate stresses.

2.3 Pretensioning steel

Since this type of steel, whether on wires, strands, or alloy bars has a smooth
round surface, the number of cycles to fatigue failure changes with the amplitude
of stresses as is the case a smooth bar of reinforced ordinary concrete.
Several mathematical formulations based on experiments, for n, the number
of cycles to fatigue failure, as a function of a, the stresses amplitude, have been
suggested by investigators, but the limit of endurance should be given in the Codes.
Assuming that the design of pretensioned elements are based on the ultimate
resistance stress of the active steel, the number of cycles to fatigue failure increases
with increasing of the minimum fluctuating stress.
72 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

2.4 Structural steel

Structural steel may be stressed indistinctly, through tension or compression. But


with fluctuating loads a distinction should be made for oscillating stresses (fluc-
tuation of one sign, either compression or tension), or alternating stresses, from
tension to compression and vice versa.
The fatigue phenomenon of structural steel depends mainly on the same two
parameters as those for reinforced concrete material, namely the amplitude of the
fluctuating stresses and the number of cycles. The decay of steel resistance is very
low, up to 10,000 cycles. In fact, it can be considered as being among the usual
material factors of safety even for amplitudes with a maximum stress close to the
yield limit.
The maximum stress of oscillation amplitudes diminishes progressively for higher
numbers of cycles. up to 100 million cycles, for which the endurance limit is estab-
lished. With alternate fluctuation the endurance limit diminishes further.
With these considerations, an acceptable endurance limit for direct stress on
nmmal grade steel should be of the order of 100 MPa. This endurance limit increases
for increasing grades of steel.
As mentioned in respect of reinforcing bars, concentration of stresses reduces
the fatigue resistance of structural steel. Geometrical defects, holes, and the like
give rise to such concentration of stresses.

3 Fatigue in structural elements


The fatigue phenomenon in structures has characteristics somewhat different from
those described for their basic structural material. The latter were subjected to direct
stresses, compression, or tension, which were nearly constant along the entire length
of the specimen tested. Structural elements are usually subjected to a regime of
stresses that are more complex, for example, bending, torsion, or even the so-called
direct normal solicitations. The stress distribution is not constant either through a
transverse section or along the structural element.
If we consider the distribution of elastic stress in a transverse section of an
isostatic beam, subjected to bending, as in Fig. 7.5. it becomes apparent that the
fatigue deterioration of the material will happen close to the extreme fibers. at the
top and at the bottom, while a significant percentage of the area of the section will
be subject to a fluctuating amplitude of stresses that are below the endurance limit.
The same linear distribution of stresses exists in any transverse section all along
the beam but with absolute values decreasing gradually towards the supports. There-
fore, the bending fatigue phenomenon will start only in the central region of the
length of the beam.
A similar picture is to be found in an ordinary reinforced concrete beam, as
depicted in Fig. 7.6, with a random distribution of bending cracks. The moment
distribution creates a regime of stresses along the beam, which causes the fatigue
phenomenon to appear in the central region of the length of the beam.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 73

~---===-:sp(t)

(a)
_K 11 ::L
rt
(b) ----=::::::::-=____~ (c)1=-4
Figure 7.5: (a) Steel beam; (b) moments diagram; (c) section I.

i P(t)

(b)
-[]-Jc
1-1 fs

Figure 7.6: (a) Isostatic concrete reinforced beam; (b) cracked section;
(c) uncracked section.

Attention should be paid to some allowances included in the Codes: redistribu-


tion of moments in calculations for continuous beams or frames leads to overstress-
ing and plasticity effects. For fluctuating loads such redistribution can adversely
affect the real level of stresses, which becomes higher than the endurance limit.
Accordingly, it seems more appropriate in dynamic design situations to perform
calculations based on elastic analysis with no redistribution.

3.1 Fatigue in steel structures

For a thorough study of the eventual problem of fatigue, the engineer should con-
sider the steel structure as a collection of possible defects, some of them of such
magnitude that they may become a focus of stress concentration fatigue. These
defects may be cuts, holes, mounds, accidental results of improper use of tools,
or simply intents for the correction of geometric deformations with their residual
stresses.
A steel structure is in reality an ensemble of rolled structural elements held
together by connectors, in pin, bolt, or welded joints. These connectors should
be deemed a potential focus for fatigue failure, each at a different level and with
different characteristics. Examples are fatigue in concentration of stresses around
74 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN I
1
perforations, shear fatigue in pins, or the complex fatigue mechanism in welded
connections, both in the weld and in its surrounding area. All these "defects" in the
rolled steel component can become a focus for fatigue stress concentration, usually
forgotten in the design and calculations for fluctuating loads.
The process of fatigue starts with the appearance of very fine cracks, which
enlarge with the dynamic motion until the moment when the structural element
cracks suddenly and the fatigue fracture occurs.

3.2 Fatigue iu reinforced concrete structures

Reinforced concrete frames are cast by parts, but proper detailing and good con-
struction practice ensure continuity, to the extent that a frame's structural elements
can be considered both homogeneous and continuous. Since connections cast in
place between different parts do not impair continuity, those areas are not regarded
as focus for fatigue stresses, as is the case with steel structures.
The problem of fatigue in reinforced concrete mainly concerns cracks and slip-
page. Cracks in flexural members subjected to oscillating loads should be consid-
ered as a process and not as a static created situation. Figure 7 .7a shows some cracks
in the tensioned region of the beam subjected to flexure in Fig. 7 .6. Figure 7 .7b is
a diagram of the tensile force which is not uniform along a reinforcing steel bar:

Tsl =Ac· act +As· asl


Tsu = As · asll
where as1 = n · act < crsrr and n = Es/ Ee.
Because of these changes the diameter of the bar diminishes progressively
(Poisson effect) in the transition area towards the crack, as depicted in Fig. 7.7c.
The transition area is a region of no adherence and of slippage of the steel along
the concrete.

(a)
Crack opening

(b)
(
--··-P ill'
Figure 7 .7: Cracks in reinforced concrete flexural members.
DYNAMJCS IN THE PRACTICE OF STRUCTURAL DESIGN 75

Regions of no adherence and slippage are also to be found close to supports of


a beam. Such regions may also be found in prestressed elements, inside the post-
tensioning sheaths, in cases of insufficient grouting.
Tensile force fluctuates in a flexural member subjected to oscillating loads. This
fluctuation gives rise to slippage of steel bars with no adherence to the surrounding
concrete. This movement creates focus of fatigue.

3.2.1 Ordinary reinforced concrete


Fluctuation of internal forces increases the widening of cracks, thereby progres-
sively reducing the compression area and consequently increasing the internal arm
of those forces.
Since compression stresses within the serviceability limit are usually below
the endurance limit (~50% ), fatigue failure of underreinforced ordinary flexural
members is expected to be similar in nature to failure in static situations. Failure
of the beam will occur after large deflections and through the crushing of concrete.
In general, it is advisable to consider that a flexural reinforced concrete member
will have an endurance limit of some two-thirds of the ultimate static resistance.

3.2.2 Prestressed beams


A differentiation should be made as to whether tensile stresses are or are not allowed
in the concrete.
Figure 7 .Sa shows a prestressed isostatic beam under an arbitrary load that fluc-
tuates with time. The pretension tendon is embedded in the beam.
If flexural cracks are not allowed, the internal stresses will fluctuate between
extreme limits, as in Fig. 7.8b. Such a beam, properly designed and cast under
good and sound practice conditions, can fluctuate more than 100 million cycles,
which means practically an unlimited number of cycles during the design lifetime
of the structure.
If tensile flexural cracks are allowed, the design lifetime of the structure is
limited.
The fatigue failure mechanism is different from that of ordinary reinforced con-
crete flexural elements. Under oscillating loads cracks open and close again in any
single cycle, and adherence at either side of the cracks is damaged. This damage

· p(x,t)
JIID !! ! j 11 l ! ! I !flC

(a) l l
(b)
(c) ~ :;;::'.:"f neutral axis

Figure 7.8: Pretensioned beam.


76 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN rI
increases with additional fluctuations. For tendons inside a duct damage also occurs
between the steel and the grout.
The stresses in the active steel in the region of a crack fluctuate with a certain
amplitude
a= O"M - O"m

If this amplitude and the number of cycles together are above the endurance limit,
the beam is expected to fail because of fatigue failure of the active prestressing steel.
This is a different pattern of failure from that expected in an ordinary reinforced
concrete beam or in a fully prestressed beam.
CHAPTERS

Damping

1 Introduction
Damping is a characteristic of structures that becomes apparent only in dynamic
situations. It opposes the movement, dissipates external energy imparted to the
dynamic system by the external action, and consequently reduces the amplitude of
motion.
Absorption of the external energy submitted by the dynamic load is performed by
a complex combination of different factors. We differentiate between two families
of damping. These are:
1. Structures damping: A variety of energy dissipation mechanisms activated
inside the structure itself by the dynamic load. Even tangible quasi-static loads,
such as persons, contribute to the damping capacity of a structure.
2. Radiating dan1ping: The energy the st.Iucture dissipates during motion in its
interaction with the foundation and with the external media in its surrounding.
Structures damping is the result of the inherent energy absorption capacity of the
basic structural material being randomly combined with friction in the supports and
an undetectable continuous internal reaccommodation of the building itself, which
takes place during motion. The latter originates in internal friction and irreversible
microdeformations in the superimposed nonstructural materials and building parts.
The inherent damping of structural materials itself originates in the internal molec-
ular friction, which dissipates energy, as heat, during motion.
Radiating energy arises chiefly under dramatic external dynamic loads such
as strong wind motions, earthquakes, and blast events, and should be considered
separately.
Structures damping is generally intended to describe energy-dissipating capacity
in the elastic response, even though some neglect of permanent deformation must
come with such an effect.
Plasticity in the structural element, almost always present in the response of
structures, deserves special consideration. The motion of a structure in a dynamic
78 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

situation is by nature an energy-consuming process. This energy comes with the


external action. Since plasticity exerts an energy-consuming effect in the structural
material, it can be considered, by extrapolation, one more factor included in the
damping characteristic of the structure. A small amount of plasticity is introduced
into the damping of a structure by the choice of higher values for the damping
coefficient, but it will cause some permanent deformation. If an elastic response is
required, plasticity effects should be avoided.

2 Viscous damping
Damping is a factor of the utmost importance in the dynamic response of structures.
As noted, it is the result of superposition of different types of damping mechanisms,
each with its own contribution to the overall dissipation of energy.
The formulation of these damping factors produces complex expressions which
encumber clear and easy understanding of the subject of damping and its further
mathematical treatment. Furthermore, a separate evaluation of their contribution to
structures damping is extremely difficult or impossible, except in very simple cases.
There the tendency is to introduce the effect of damping in the form of "viscous
damping," namely the force opposing the movement of a solid in a viscous fluid,
which is directly proportional to the velocity of that movement. Its inclusion in the
differential equation of motion sometimes facilitates formulation of the response
function. This in turn promotes clear understanding of the dynamic situation.
Although viscous damping is not typical for structures, in most practical cases,
besides its advantage as a mathematical instrument, its approximate value has
proved representative of the effect of all the damping factors together in the response
of a structure.
Since it is difficult to establish the absolute value of the damping factor for any
particular dynamic situation, it is useful to consider the damping coefficient as a
percentage of the critical value:

The critical damping represents the minimum value that eliminates any free oscil-
lation of the structural element. It is a function of mass M and spring constant k
characteristic of any given structure in a dynamic design situation.

3 Hysteretic damping

Structural engineers are familiar with hysteresis, so the phenomenon deserves some
treatment even though it is only one of the factors contributing to the value of the
viscous damping coefficient.
Hysteretic damping is defined as the energy dissipated in one cycle of a structure
subjected to a fluctuating load. It can be represented by the area enclosed in one
load-deflection hysteretic loop (Fig. 8.1).
DYNAMICS IN TIIE PRACTICE OF STRUCTURAL DESIGN 79

P(t)

Figure 8.1: A load-deflection hysteretic loop.

When the structure deflects from 0 to A, the work pe1formed by the structure,
and hence the energy absorbed, is proportional to the area OAB. When it returns
from A to C, it restores an amount of energy proportional to the work represented by
the area ABC. This energy absorbing and restoring process is repeated four times
in one hysteretic loop, hence the area A CDEA represents the total energy absorbed
by the structural material in one cycle of deformation.
Three possible locations for point A in Fig. 8.1 may be distinguished:
I. Point A is between the elastic and the proportionality limits, represented in
Fig. 8.1. The loop is closed and it is repeated in any cycle. It does not pro-
duce deformation of the structure worth consideration in the design. Energy
dissipation is mainly performed through heat.
2. Point A is below the elastic limit. The loop is very narrow and the deformation
of the structure is elastic (Fig. 8.2a). Energy dissipation is by heat.
3. Point A represents a deflection of the structure that exceeds the proportionality
limit. The hysteretic loop is open and its area swells with each additional
cycle (Fig. 8.2b). Energy dissipation is mainly by plastic deformation and
increasing deflection of the structure.

P(t) P(t)

(aj CW
Figure 8.2: Load-deflection hysteretic loops.
80 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

4 Damping in structures
Three levels of damping can be established for practical use in dynamic design
situations:
1. For ordinary linear situations
0.005 0: c 0: 0.05
or O.So/o · Ccr :S ; _:::; 5% · Ccr

Even though damping is basically an inelastic characteristic of structures,


in dynamic situations with small displacements it is considered an elastic
mechanism:
for a naked isostatic beam with friction-free supports, for example, the
damping coefficient may hardly be considered greater than 0.7% for con-
crete and 0.5o/o for steel; this coefficient is mainly composed of the inherent
damping characteristic of the structural material;
for typical buildings a coefficient of 1-2% is usual considered adequate;
higher values, up to 5%, include some irreversible deformations and some
amount of energy radiation into the surrounding ground.
2. For situations where the expected response may include plastic deformations

0.05 0: c 0: 0.20
or 5% · Ccr :S ; :S 20% · Ccr

The higher values lead to deformations of an order double that calculated for
the yield limit of the material. In these cases the choice of working with the
viscous damping coefficient is only for simplicity's sake. It should be kept in
mind that this calculation yields roughly approximated results.
3. Miscellaneous
When higher amounts of energy dissipation are required to avoid a catas-
trophe, the damping calculation technique should be replaced by plastic-
ity studies. Real extreme dynamic situations, such as design of structures
close to earthquake collapse, are handled directly by calculations in the
plastic domain.
Air damping becomes important in relatively light and elastic structures
such as cable bridges. For the engineer this is a problematic design situ-
ation: under certain conditions it changes its sign and the damping force
becomes negative. Hence, this force increases the deformation of the oscil-
lating structure instead of mitigating it. A wind velocity, Ver. can be cal-
culated for a given structure which establishes a minimum limit. Any
wind velocity greater than this minimum will cause a negative air damp-
ing coefficient, which in tum will bring about inevitable collapse of the
structure.
Part4

Structural dynamics design


J
CHAPTER9

Mathematics

1 Introduction
The main characteristics of a structural system in a dynamic design situation are the
mass and stiffness of the structure and the global capacity of the system to dissipate
some percentage of the energy provided by the external dynamic effect in order to
move the structure.
The motion of a structure can be conceived as a design situation of equilibrium
of forces. They are time varying, so they change continuously, but at any given
moment a mathematical equation of equilibrium can be written:

Finertial + Faamping + Fstiffness + Fex:tcmal load = 0

where

Fmertial = F1 = M · V;
Faamping = Fo = c · V;
Fstiffness = Fs = k · v;
Fextemallood = F(t) = F1 ·/(t).

Ifthe displacement in the structural system is represented by v, then the velocity


v v,
and the acceleration will be and respectively.
Amass in motion originates, by D' Alembert's principle, an inertial force, which
is the product of mass and acceleration, M · V.
The damping capacity of the structural dynamic system is usually represented as
a force which is the product of a viscous damping coefficient and the velocity of
motion, c · V.
The spring stiffness force Fs, is the product of displacement and a spring coeffi-
cient k, which represents, at any moment, the elastic deformation of the structural
system k · v.
84 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

The external dynamic effect is represented, in the equation of forces, by F(t),


which is the product of a particular constant value and a particular function of time:
F1 ·f(t).
If the movement of the mass is a rotation, measured by the angle <p, instead of a
displacement then velocity and acceleration become ip and !p, respectively, and the
inertial force will be given as the product of the mass moment of inertia, Im, and ;p.
A systematic mathematical study of dynamic systems is usually performed in
order of increasing complexity by classifying structures into three main categories.

Single degree of freedom system (SDOF): This is a single concentrated mass,


linked to the support by means of a single spring, as in Fig. 9.1. The spring mass
can be neglected, in contrast to the supported mass.

water tank with


massM

column with
spring constant k

Figure 9.1: Schematic representation of an elevated water tankou a slender column.

Multidegree offreedom system (MDOF): There are more than one concentrated
mass and more than one interconnecting spring, as in Fig. 9.2. Spring _masses can
be neglected, in contrast to the concentrated masses.

hi kl
mass 3

,. ,~

<~ < ~h2 k2 mass2


,< '<
,. ' <

hl kl mass 1

. .. / , .- / ' . . ,-/ ' '//

Figure 9.2: Three-level frame with a concentrated mass at each level and the
column's spring constant between levels.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 85

mass distribution
~--'--

j_' 2

Figure 9.3: A continuous beam with distributed mass, not always uniform, all along.

Distributed mass system (DMS): Here a structural element has a continuously


distributed mass all along, as in Fig. 9.3.
Consideration of a given structure as pertaining to one of the categories SDOF,
MDOF, or DMS requires a simplification, sometimes significant, of the real struc-
ture. This simplification opens the way to a mathematical study of the dynamic
problem.
This mathematical treatment of the dynamic issue includes the application of
highly elaborate and sophisticated techniques, analytical at times, numerical when
this suffices. This mathematical study is the field of researchers and scientists.
In practice, dynamic problems and design of structures in dynamic situations
are solved principally by means of appropriate computer programs. Nevertheless,
mathematics is the only way to obtain thorough knowledge and deep understanding
of a structure's response in a dynamic situation, and it is the basis of all the computer
programs.
The following chapters present a simplified presentation of the mathematics of
structural dynamics, which is necessary to elucidate the effect of dynamic properties
and the dynamic characteristics of the response of structures to dynamic excitations.
CHAPTERlO

Single degree of freedom system

1 Introduction to SDOF

F(t)
(b)~
(a)
(c) ~)
/ FR r1

Figure 10.1: An SDOF, light steel frame with a rotating machine.

The horizontal dynamic component F(t), which is a horizontal force changing


with time, introduced into the frame by a working machine, originates a horizontal
vibrating movement, henceforth, whose amplitude x(t) is time varying too.
Three internal forces oppose this horizontal movement, namely

Fs = k · x = spring force, which represents the elastic horizontal


deformability of the frame;
Fi = M · X= inertial force that originates in· the moving mass;
Fo = c ·X = a damping force, inherent in the moving system.

Figure 10.lb is a classic representation of the real structural system, especially


for the mathematical study of a dynamic system. The moving mass of the system
is represented as a lumped, concentrated mass whose motion, originated by the
external dynamic force F(t), is regulated by a physical spring of constant k and a
damping mechanism of constant c.
Figure 10. lc shows all the forces, which are constantly in equilibrium, even
though they change continuously during the motion.
88 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

The equation of equilibrium of forces is:

F1 + FD + Fs = F(t) (10.1)

or
M ·x+c·x+k·x=F1 ·f(t) (10.2)

For a variety of functions f(t), an analytical expression for the solution of this
equation of motion may be formulated

X = fx(t)

which facilitates a clear and deep understanding of the response of structures in


dynamic situations, as will become apparent in the following sections.

2 Free vibration
Assume that an exte111al transitory action works on a structure, initiates its motion,
and then disappears. The structure will continue its vibrations for a while, until
it stops in a situation of static equilibrium. The external action creates a transient
dynamic situation.
This vibration of the structure, after removal of the external dynamic effect, is
called
"damped free vibration,"

and its foremost characteristic is that eventually it will come to a full stop.
The equation of equilibrium during this motion is

M·x+c·x+k·x=O (10.3)

which is the same equation as before, eqn. (I 0.2), but without the external force.
Assume next that no damping mechanism exists in the structural system. Then
the equation of equilibrium is further reduced to

M·x+k·x=O (10.4)

whic~ represents an

"undamped free vibration";

it will continue indefinitely.


The assumption that the damping force does not exist, FD = 0, is purely theo-
retical since no real structure is without damping.
In practice, what really does exist is a certain type of structure, for example, a
slender, high-grade steel cantilever beam, whose free motion will continue for a
remarkably high number of cycles.
DYNAMICS IN THE PRAcrrcE OF STRUCTURAL DESIGN 89

2.1 Undamped free motion

An analytical expression for the undamped free motion of the SDOF system is

x= fi·sin(fft·t)+x;·cos(fft·t) (10.5)

which is a function of the moving mass M and the spring constant k. The basic
assumption underlying this analytical expression for x is that the internal elastic
force in the moving system is linearly proportional to the deformation and is rep-
resented by the spring constant k.
The wavy, undamped free motion is basically a sinusoid, best known as "harmonic
motion"; Xi and Xi are, respectively, the displacement and the velocity of the vibrat-
ing system the moment the external force is removed.
Figure 10.2b shows a steel cantilever beam with concentrated mass M, after
removal of F(t), the dynamic external load.

(+)
F 1 · f(t)

(a)
~ *
8M
t(-)

t
----
max(+)li
(b) --
111111 (-)Ii

Figure 10.2: A cantilever beam.

At the position of maximum displacement Xmax, the velocity of motion is zero.


Alternatively, the position of maximum velocity imax occurs at zero displacement.
Figure 10.3 presents diagrams of undamped free motion for three different cases,
each representing the instant when the vibrating structure is given to free vibration
by removal of the external dynamic load.
Figure 10.3a shows free motion for x; = O: Fig. 10.3b shows free motion for
xi= O; Fig. 10.3c shows free motion for initial conditions Xi and Xi, different from
zero.
The length T of one complete cycle, measured in Fig. 10.3 along the axis of time,
is the same for a, b, or c:

T = 2rr/if [s]
90 DYNAMICS JN THE PRACTICE OF STRUCTURAL DESIGN

(a)
p~~ cl. x:4: ,;,(~ l)

(b) r,v D \:/ l


I> x~x;w{~ l)

x
T

Figure 10.3: Undamped free motion.

T =period of the structure, is the length of time necessary for the SDOF undamped
system to perform one cycle of free motion; it is a constant value, characteristic of
each system, and it depends only on the mass M and the spring constant k.
The inverse of the period is the frequency f of the system:

f = -1 = - 1
T 2n
g -
M
[Hertz or Hz]

The natural circular frequency, w, is the preferred parameter for the mathematical
expression of the response deformation x of the dynamic system:

w = 2n -f = g [c.p.s.]

Hence eqn. (10.5) becomes:

X; .
x = - · sm (w · t) + x; ·cos (w · t) (IO.Sa)
w
and the acceleration is obtained by derivation, twice, of eqn. (JO.Sa):

x= -i; · w ·sin (w · t) - x; · w2 ·cos (w · t) (10.6)

The introduction of values given by eqns. (10.Sa) and (10.6) into the equation
of forces (10.4) satisfies the latter.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 91

The square of the natural circular frequency w is known as the eigenvalue:

2 k .
w = M = eigenvalue
The eigenvalue is extremely useful in the mathematical treatment of MDOF and
DMS systems.
The four parameters mentioned above, f, T, w, and w2, are interrelated such that
any one of them determines all the others. Essentially they are, the function of the
spring constant k and the mass M in motion.

2.2 Damped free motion

Real structures are damped systems. The mechanism of damping differs from one
structural system to another, but its effect is always the same, namely to reduce
progressively the amplitude of vibration, and hence the number of cycles in the
free motion, to the point when the moving system comes to a full stop and reaches
a static situation of equilibrium. The amount of damping likewise differs from one
structural system to another. The amount of damping of a particular structure is
measured as a certain percentage of critical damping and is represented by the
damping coefficient ratio, or damping factor

Critical damping is a function of the mass in movement and the spring constant
of the dynamic structural system

l'critical = 2~

The critical amount of damping is defined as that amount that causes the structure to
return from the initial displacement Xi to the position of static equilibrium, without
any vibration (Fig. 10.4).

Figure 10.4: Displacement fades out because of damping.

Critical damping should be considered as a reference only. Usually, real structures


are underdamped. The effective amount of damping is only a small percentage of
the critical.
92 DYNAMICS TN THE PRACTICE OF STRUCTURAL DESIGN

For a dynamic structure whose damping mechanism can be represented by vis-


cous damping, eqn. (10.3), which contains the damping force, can be written in an
equivalent expression

M ·x+ 2M~w ·x+k ·x = o (10.3a)

The solution of eqn. (10.3) gives the analytical expression for the SDOF system's
response, in damped free motion:

x = e-<wt (''; +;~· ~w · sin(wo · t) +x; · cos(wo · t)) (10.7)

where wo is the natural circular frequency of the damped system.


For real structures the difference between wo and w can be neglected. Hence
expression (10.7) can be substituted, with good approximation, by

x = e-<wt ( (~ +x; · ~) · sin(w · t)+x; · cos(w ·I)) (10.8)

Equation (10.8) becomes eqn. (IO.Sa) for an undamped system where~= 0. This
equation is represented in Fig. 10.5c.

(a) x

-1

(b) x

x
(c)

Figure 10.5: Damped free motion.


DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 93

Figure 1O.Sa and b shows the response of an underdamped SDOF system accord-
ing to the particular initial conditions of motion, which are, respectively, the initial
velocity Xi = 0 and the initial displacement xi= O.
Comparing the diagrams in Fig. 10.5 with those in Fig. 10.3 highlights the effect
of damping, which is to diminish, cycle after cycle, the amplitude of motion. The
harmonic motion is the same, and the period T "' To does not change either. The
ratio between two consecutive amplitudes an and an+l is constant, and is given by
the logarithmic decrement 8 which can be written, approximately, as

8 =In ( -a,,- ) "'2n · ~


Gn+1

The free damped motion is performed between limits given by the exponential
functions portrayed in Fig. 10.5. k in Fig. JO.Sc represents a constant that can
be calculated by an equation, albeit rather complex, which contains the initial
conditions x; and .X;, the period of the structure T, and the damping coefficient~-

3 Forced vibrations
If an external dynamic action acts upon a structural system for a certain length of
time, the vibration of the structure, during this length of time, is called a

"forced vibration."

If the external dynamic action is removed, the structural system will continue to
vibrate for a while, in what is called

"damped free vibration."

Note that the forced vibration becomes a "clean" vibration only after some time.
First, it starts as a superposition of the damped free motion on forced motion.
Accordingly, the dynamic response of a structure subjected to a transitory dynamic
action can be considered subdivided into three successive stages of vibration:

x(t) = Xstagel --+ Xstage2--+ Xstage3

where

Xstage 1 is the superposition of damped free and forced vibration;


Xstage 2 is a "clean" forced vibration;
Xstage 3 is a damped free vibration.

In practice, stage 2 is the one to consider for the eventual ultimate limit design
or control of the structure.
94 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

3.1 Pulsating force

The effect of a pulsating force on an isostatic beam can be considered the classic
choice for a study of the effects of external dynamic action on a structural system
(Fig. 10.6).

(a)
El

(b)

f(t) =Q·sini.2t

Figure 10. 6: A pulsating force on a beam.

In this particular case the external action can be, for example, the force originated
by a machine working with an unbalanced rotating mass.
F(t) represents a rotating force which at any given moment has two com-
ponents:
a horizontal component FH(I) = Q ·cos (Q · t) and
a vertical component Fv(t) = Q ·sin (Q ·I)
FH(t) introduces a longitudinal harmonic force into the beam. It is not
relevant for our present study.
- Fv(t) is the pulsating transversal force that creates the dynamic design
situation.
- Q is the weight of the rotating mass.
Qin [cycles per second] is the circular frequency of the rotating mass, given
by the factory in the form of (r.p.m.), the machine's revolutions per minute.
The amount of damping of the dynamic structural system is usually represented
by an equivalent viscous coefficient c [kN · s/m]; the spring constant k represents
the elasticity of the beam, as a vertical spring,

k = Q [kN/m]
!c.
where~ =vertical deflection under static load Q. This equation considers a linearly
elastic deformation.
The equation of equilibrium of the dynamic forces is

M ·x+c·x+k·x=Q·sin(rl·t) (10.9)
which is similar to eqn. (10.2).
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 95

Introducing the already explained expressions

c=2M·w·~ and w=f-:;

into eqn. ( 10.9), the latter becomes

M ·x+2Mw · ~ ·x+Mw2 ·x = Q · sin(Q. t) (10.10)

A general solution for this equation is

x(t) = e-t'"1 (C1 · sin(wot) + C2 · cos(wot))


Qo I
+ 2
k (1- rnl2) + (2~m2

2~ £cos (Q · t)]
2
x [ ( 1 - ( £) ) sin (Q · t) - (10.11)

This general expression of x(t) is very useful forunderstanding the way tbe structural
system responds to an external harmonic forcing function.
First, it responds in 'stage l ', which means the superposition of damped free
motion:

(JO.Ila)

and damped forced motion:

Qo
k (1 - (g}2)
1
2
+ (2~£)2
[(1 -(~) ) w
2
sin(Q · t)-2~[1 cos(Q · t)]
w
(10.llb)

The damped free motion, eqn. (10.1 la), fades out with time, while the forced motion,
eqn. (10.llb), is continuous as long as the external action continues to act on the
structure (stage 2), a circumstance known as the steady-state motion, Xss(t).
C1 and C2 are constants reflecting the initial conditions of the motion, displace-
ment, and velocity; see eqn. (10.7).
Second, the forced component of the motion, Xss(t), can be rewritten, after so1ne
mathematical manipulations, as

(10.12)

where e is only a phase angle.


96 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

This expression includes three factors:

x,,. = Qo
k
= the static deflection under load Qo

1
DMF = ---------~

[(1 - rnl2)2 + (2.;m2r/2


sin (Q. t - 8) = the harmonic function which defines the characteristics of
the motion itself; a sinusoidal movement.

The phase angle


2.;!l. )
(~) 2
1
8 = tan- ( _
1
indicates that the response motion of the structure lags behind the motion of the
external forcing function.
Third, the response is a function of both the natural circular frequency of the
structure, w, and the circular frequency of the external action, Q. The response
greatly depends on the ratio Q/w.

4 Dynamic modification factor


The termx,,(t) in eqn. (10.12) becomes a maximum when sin (Q · t-8) becomes I.
Hence
Xss,max = Xst · DMF

DMF represents the effect of the dynamic action when compared with that of the
same load but statically applied.

In practice, cases where DMF > 1 are important because of the direct correlation
between the displacement, Xss,max• and the internal forces created by the dynamic
design situation in a linearly elastic structural system.
Figure 10.7 is a diagrammatic representation of the dynamic effect of a pulsating
force represented by the value of the DMF.
Three outcomes of this graphic representation merit attention:
I. ihe rapid rise in the value of DMF, again for ratios Q / w, close to 1; for this
reason it is important to try to keep Q and w as far apart as possible;
2. the mitigating effect of the amount of damping, .;, in the value of DMF for a
given ratio Q/w close to l;
3. for high values of the ratio Q/w the effect of DMF is to fall below the value
of Xst; it is rather like saying that the rotating mass has no time to sit fully on
the structure; for such situations the maximum deflection should be x," which
will become real each tin1e the machine stops working.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 97

DMF
5.0

~=10°!0

Figure 10.7: DMF for a pulsating load.

The DMF explained and analyzed here at length is valid for a pulsating load.
The correct formulation for each type of dynamic loading should be worked out
separate! y.
Just for comparison, consider the case of an isostatic beam (Fig. IO.Sa) with a
suddenly applied permanent load (Fig. JO.Sb).

(a) (b) F

.LS:
El
rm, f,m ~
Oo

td

(c)

21t 2n

Figure 10.S: An isostatic beam under a permanent load, rapidly applied.


98 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

The external permanent load is applied in a really short period of time td, say
one-tenth of the period T of the structure.
In such a case the dynamic vertical displacement of the structure is given by the
expression

x(t) = ~o { 1 - e-"'"[ cos(w · t) + ~ · sin(w · tll} (10.13)

where the difference between cvo and cv is neglected, which is adequate for real
practical cases (for ~ 5 14%, w0 :o> 99% · w ). Figure 10.8c shows the undulating
motion of the beam which fades out with time.
The maximum deflection of the beam will be approximately at w · t = n, and
will be given by

x(t)= ~o {I-e-<"[cos(n)+~·sin(n)J}
x(t) = x, 1(1 + e-<n)
For different values of~ = 20%, 10%, and 5%,

x(t) = 1.53 · x", 1.73 · x", 1.85 . x,,


The upper limit will be for a theoretical case of a structure with zero damping, in
this case ~ = 0, and
DMF=l+e- 0 =2.

5 Resonance
The effect of the pulsating load greatly depends on the ratio between the circular
frequency of the load Q and the natural circular frequency cv of the structural system,
as is apparent from the curves in Fig. 10.7. The particular case where Q equals (J)
is known as a case of resonance.
In practice, ratios of Q / w close to I should be considered cases of resonance
too, since DMF reaches values that endanger the structure.
Damage to the structure, in cases of resonance, should be interpreted very
broadly; it is not always a situation of collapse. With increasing deformations the
structural response is not linear; plastic deformations, open cracks, and unendurable
permanent deformations may make the structure unusable.
In resonance, the motion of the structure synchronizes with the pulsating load.
At the beginning of each cycle of movement the structure receives a fresh push
from the pulsating load and the amplitude of each successive cycle increases.
This physical description illuminates two characteristics of the process of
resonance:
I. Since the amplitude of motion grows from one cycle to the next, the maximum
amplitude at resonance may require a high number of cycles. Unfortunately, a
high number of cycles may need only a short time interval.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 99

2. Since the amplitude at resonance is reached through a cumulative process, the


absolute value of Qo is of no significant importance.

6 Dynamic loads
Structures in dynamic situations may be subject to a variety of dynamic loads, as
shown in Fig. 10.9.

tQ (d) F

~
(a)
~ ;t
I (.____o,"-/ •
td v
(b) V>
td
,L'> (e)~

td

(c)~ (f)~
td

Figure 10.9: (a) Harmonic load; (b) periodic load; (c) irregular load; (d) arbitrary
load; (e) blast load; (f) pulse load.

The Structure's response always presents the same general characteristics already
set forth for pulsating loads, namely:
1. the existence of a DMF; practicing engineers are naturally interested in cases
whereDMF> I;
2. the resonance phenomenon and its magnifying effect on induced deformations
in the structure;
the mitigating effect of damping;
the delay between the exact moment of the structure's maximum response
versus the exact moment of maximum amplitude of the external dynamic
load.
Sadly, no rules of thumb can help foster an i_ntelligent guess at the expected
value of DMF or of the phase angle, as it can be seen through the results of the
appropriate mathematical calculations pursued for the examples in Fig. 10.10.
In all these cases, as almost always in dynamic situations, it is advantageous to
relate the time Id to T, the natural period of the system.
For an undamped linear elastic SDOF system, the maximum values of DMF are
case a: DMF = 2, for ld/T = 1.00,
case b: DMF = 2, for td/T = 0.50,
case c: DMF = 1.50, for td/T = 0.93,
cased: DMF = 1.52, for ld/T = 0.75.
}00 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

(a) F (c) F

t
td td

(b) F (d) F

_t
td

Figure IO.IO: (a) Gradually applied load; (b) rectangular pulse; (c) blast; (d) grad-
ually increasing and decreasing pulse.

and the maximum values of the delay time, or the time of maximum response, for
the same ratios td/T as before are:
- case a: no delay;
- case b: no delay;
case c: maximum deflection after the blast starts and before it reaches a half
of the structure's period;
- case d: maximum deflection occurs after the peak of the load and before it
vanishes.

7 Mathematical approaches

Two main approaches to the mathematical treatment of a dynamic strnctnral prob-


lem are highly illustrative:
1. formulation in the time domain;
2. formulation in the frequency domain.

8 The time domain

Let us assume that the external dynamic load is represented by an irregular diagram,
as in Fig. I0.11, which can be considered a case of a general dynamic load F1 ·f (t).
We shall start our mathematical consideration with a vertical strip of amplitude
F 1, width dr, at a time r. The load represented by the area of this hatched strip
is known as an impulse. It looks similar to the pulse load given as an example in
Fig. 10.10, but the measure of time ld of the pulse load in Fig. IO. IO can be variable,
shorter or longer; instead, dr is a true short length of time, not measured as a ratio of
the period T of the SDOF strnctnre, but measured in absolute fractions of a second.
For mathematical formulations dr is a differential of time.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 101

--.,>~----''--++ dt

Figure 10.11: Diagram of an arbitrary dynamic load.

Since the effect of an impulse on a structure can be formulated analytically, the


summation of the effect of all impulses composing the general dynamic load is
conducive to a formulation of x(t), the steady-state response of a SDOF structure
(after the free vibration has faded out)

x(t) = x" · w fo' f(r) · e-«o(t-r) ·sin (w(t - r)) · dt (10.14)

The solution of the integral in eqn. (10.14), known as the "Duhamel integral,"
will include a DMF factor of the static deflection value and an expression for
the harmonic motion of the structure. Equation (10.14) is known as the response
through the time domain. Unfortunately, it is not always possible to solve this
integral analytically. Where this is the case the integration will have to be performed
numerically.

9 The frequency domain

When one works through the frequency domain, any external general dynamic load
is transformed into a summation of its hannonic components. This is instead of a
summation of the impulse components, as is characteristic of working through the
time domain. In the frequency domain the mathematical technique for a periodic
load differs from that adequate for an arbitrary load.

9.1 Periodic loads

Let us start with an overview of periodic loads (Fig. 10.12). A periodic function is
characterized by its period Tp and the conditionf(t) = f(t + n · Tp). The graphical
representation of this condition is given in Fig. 10.12 by the values off,.fb, and/,;
each one repeats itself, period after period.
For an external dynamic load represented by

F(t) = F1 · f(t)

f (t) can be any arbitrary function of the period Tp, which repeats itself along the
positive axis of time or along the axis of time from -oo to +oo. It is just a matter
of mathematical convenience.
102 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

(a) ~ 'C7
Tp
/["\
-v
f,

Tp
f,
/["\
'C7
f,
cc i;1

(b)

(c)

Tp Tp

Figure 10.12: Periodic functions.

Basically, the mathematical technique is to represent any periodic load by its


Fourier series, which is a trigonometric series:

F(t) =Ko+ K1 ·sin (Q · t + <p1) + Kz ·sin (2Q · t + <p2)


+ K3 · sin (3Q · t + <p3) + .. · (10.15)

where Ko, K1, ... , Kn. and cpo, cp1, ... , cp11 are constants to be calculated.
Representation of an arbitrary periodic external load F(t) in the Fourier series
is the result of a summation of an infinite number of hannonic functions. It also
reveals that the natural circular frequency of all these harmonics are multiples of
one basic circular frequency, called the fundamental frequency Q.
Expanding the trigonometric sine function in eqn. (I 0.15) sometimes leads to a
more useful expression of the Fourier series:
00 00

F(t) = F1 · f(t) = ao +La,, ·cos (n · Qt)+ L bn ·sin (n · flt) (10.16)


11=1 11=1

where

ao
Q
= -F1
12rr/Q f(t) · dt (10.17)
2rr 0

a,, = -F1
Q 12rr/Q f(t) ·cos (n · Qt)dt (10.18)
lt 0

bn =
Q
-F1
12rr/Q f(t) ·sin (n · Qt)dt (10.19)
lt 0
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 103

This transformation of the expression of the dynamic load is possible so far as a


solution of
r2rr/Q
lo f(t) · dt

is known. In eqn. (10.15), Ko = ao, K,, = .,/a~ + b~, 'Pn = tan- 1(a,,/b,,) where
ao, an, and bn are constant values, calculated with finite integrals.

9.1.1 Graphical representation

F (t) F (t)

t 11'°4
(a) (d)
TT n

F (t) F (t)

+ sin·t
-Y;,sin·2t
+Y;sin·3t
-!,,\sin-4t n~l5

(b) --<'--_,_,,-<-++-~--<> t (e)


t
TT

F (t) F (t)

n=c:o
(c) (f) _,~----~TT~-1>
TT

Figure 10.13: Graphical representation off(t)=t on the Fourier series (0 < t < rr).

Figure 10.13 is a graphic representation of the Fourier series for a particular function
f(t) = t. The function is valid for -rr < x < rr.
104 DYNAMICS IN THE PRACTlCE OF STRUCTURAL DESIGN

The limits of the function, in this example, were chosen only for clarity in the
mathematical representation. Performing calculations of ao, a11 , and bn for this
particular function yields

2
ao = 0; a11 = O;
bn = - ·cos(n· rr)
n
sin (t) sin (21) sin (3t) )
f (t)=t=2 ( ~- - - - + - - - · · · (10.20)
1 2 3

Figure 10.13a depicts the load function/(t) = t; Fig. 10.13b shows the first four
terms of the Fourier series in eqn. (I 0.20); Fig. 10. l 3c, d and e shows the summation
of n terms of the series; Fig. 10.13f shows the summation of n = oo terms of the
series, which is exactly the original function/(t) = t, since the function does not
exist at limit TI.

9.1.2 Steady state in Fourier series


The steady-state response of the SDOF structure, subjected to the effect of a periodic
load, is given, after the free vibration has faded out, by the expression

00

x(t) = x,, · DMFo + L DMF,,[A ·sin (n · flt)+ B ·cos (n · flt)] (10.21)


n=l

where

DMFo = -ao (constant value)


Ft

Again, as when working through the time domain, the response is a harmonic
function; see eqn. (10.21), with the already known three components: a DMFx,..
factor, and a sinusoidal function, which in this case too is a summation.

9.1.3 Amplitudes of harmonic fnnctions


The amplitudes of the harmonic functions in the Fourier series is sometimes a given
datum, as in Fig. 10.14.
DYNAMICS IN TIIE PRACTICE OF STRUCTURAL DESIGN 105

6
/
2
~' I
I , \
4
I 1 /
v~
' 5t '
o 20 30 40 so 60 no
hannonic frequencies

Figure 10.14: Discrete Fourier amplitude spectrum.

9.2 Arbitrary load

F (t)

Figure 10.15: An arbitrary load.

An arbitrary dynamic load, with a general representation as in Fig. 10.15, is not


periodic. To apply the frequency domain approach to this problem a different tech-
nique should be used. The arbitrary load will be considered part of a periodic load
with period Tp --> oo. The graphical representation of this notion could be taken
as that of Fig. 10.12, but with a significant modification (Fig. 10.16).
If the period Tp is assumed to grow to oo, the other components of the series will
move away and disappear. This technique is known as the Fourier Transform, and
results from an expansion of the Fourier Series technique.

h \....} I
(
\
'
t
t>

Figure 10.16: Periodic simulation of a one-time function.


106 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

If
i:f(t)dt < 00

this technique yields

F(n · Q) = 1"'
-00
f(t) · e-;·n·D·t ·di,
(10.22)
f(t) = - 1
2n
1"'
-oo
f(n. Q). e;·n·D·t. d(n. Q)

The trigonometric functions of sines and cosines have been written in their expo-
nential form
et+ e-t er - e-t
sin(t) = ; cos (t) = - - 2 -
2
andi=R.
The expressions in eqn. (I 0.22) are known as the Fourier Pair Transform, for
arbitrary loads, which is an alternative technique to the Fourier Series Transform,
for periodic loads. The frequency equation "F(n · Q)" is the transform of the time
functionf(t); the time function is the inverse transform of the frequency function.
By extending the period Tp to infinity, the summation of trigonometric functions
becomes an integral of an infinite number of sines and cosines. The dashed line curve
in Fig. 10.14 becomes a continuous line, called a continuous Fourier amplitude
spectrum, with amplitudes infinitesimally close to each other.

10 Elastoplastic systems
The mathematical treatment of dynamic design situations through the time domain
or the frequency domain is valid as long as the structure's response takes place
within the range of elastic deformations. Linear elastic deformations is a sine qua
non for validity of the summation technique of superposing individual responses
in the expressions for x(t), as implied in equations such as (10.14) and (10.21).
The basic equation of equilibrium of dynamic forces in motion, expressed in
eqns. (10.1) and (10.2), assumed a linear elastic ratio of the spring force

F, = k · x
with displacement x, without any upper limit.
Figure 10.17a shows the representative resistance-deformation curve for
medium to high quality structural steel.
Figure 10.17b is a bilinear diagram, admitted by the Codes of Practice for design
purposes, instead of the real curve. It is known as a bilinear resistance curve. Line
a-b represents the linear elastic spring part of R,

R=k·x
DYNAMICS IN THE PRACTICE OF S1RUCTURAL DESIGN 107

R R R
b c

x x
-,,+-~----"'x
a Xe!

(a) (b) (c)

Figure 10.17: Resistance-deformation curves.

which increases linearly up to the limit of elastic deformation, XcJ, while b-c rep-
resents the plastic constant resistance

RpI = constant

for further deformation x; it will continue as long as the ductility of the material
permits.
A dynamic design situation that involves a material with a bilinear resistance-
deformation curve will still respond to the same equation [eqn. ( 10.1 )] of equilib-
rium of dynamic forces. The problem can be solved by systematic application of
some of the formulas already given in the preceding pages of this chapter, and as
in the following example.

(a) (b)
LS...

td

Figure 10.18: Plastic deformation of an isostatic steel beam in a dynamic situation.

Figure 10. l Sa shows an isostatic steel beam subjected to a concentrated dynamic


load. Figure 10.1 Sb is a diagram of a load F 1, like that in Fig. 10. lOa, dropped by
accident instead of being gradually laid down, as it should have been,

Id --> 0

As already known, such an accident implies a DMF on the absolute value of the
load, which can easily increase by more than 50%.
If through effect of this dynamic load deformation x of the beam is greater than
the elastic limit,
X > XeI

the structure's response can be divided into three stages, as follows.


108 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

In stage one the equation of equilibrium of forces, similar to eqn. (10.2), is

M ·x+ c ·x+k ·X = F1

This stage ends when x = Xel; when this takes place the structure reaches a certain
velocity Xe1.
The second stage starts with the structure in the aforementioned initial conditions,

X = Xel• X = Xel
The dynamic forces are naturally in a condition of equilibrium their equation also
evolves from eqn. (10.1):
M ·x+c ·x+Rv1 = F1
The spring force is no longer proportional to deformation, but constant with x.
The ratio c · i/Rp1 can be neglected since absorption of energy is mainly regulated
by plasticity. Hence, the equation of equilibrium reduces to
M ·x+Rp1 =F1
This stage ends when the motion of the structure stops; displacement reaches its
maximum
Xmax = Xel + ~Xpl
and the spring force in the deformed structure equals the external load F1.
The third stage is the point of maximum deformation from where the structure
returns toward the position of equilibrium. The return path is parallel to the linear
elastic deformation line, as shown in Fig. 10.17c. The equation of equilibrium of
dynamic forces should include, again, a coefficient for damping.

t
"
I
'
Xsi
x,,
-1 -
T

Figure 10.19: Vibration curve of Fig. 10.18.

Figure 10.19 shows the damped vibration motion of the beam, which follows
the return stage; motion will definitely stop because of damping. In this example
the final position of equilibrium after the accident will include a permanent plastic
deformation ~Xpl· Fortunately, from here on, the beam will remain in the elastic
range, for further loading.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 109

11 Nonlinear systems
Figure 10.20 shows the resistance-deformation curve for two different structural
materials.

R R
I

Re

x I ' x
(a) (b)

Figure 10.20: Load-deformation curves.

Figure I 0.20a is the load-deformation diagram for normal grade structural steel.
Up to a certain limit the curve and the tangent at x = 0 are the same. Only after
that limit, R,, does the curve bend, slowly at the beginning.As long as the response
of the structure x 0 is below that limit, eqn. (10.1), the equation of equilibrium of
dynamic forces, is valid.
Figure 10.20b depicts the load-deformation curve of a hardening system that
starts with a zero horizontal tangent and rises with deformation, up to a certain
point of inflection, i; thereafter the curve bends towards a plastic region. Between
zero and i the curve represents a structure of nonlinear elastic resistance; after point
i the curve represents a structure of a nonlinear plastic resistance. The case of a
hardening structure is given only as an example.
Nonlinearity may lie in the stress-strain characteristics of a structural material, or
in the load-deformation ratio of a real structure, or even in the damping coefficient.
An analytical mathematical solution can hardly be expected when a dynamic
design situation involves some nonlinear components. Instead, numerical techni-
ques are implemented. It is the step-by-step process and a widely applied technique.

R R

/
n+l /
/
n
n-1
(a) - + - - - - - - - - - { > x (b) - t - - - - - - - l > x

Figure 10.21: Linear segmentation of continuous curves.

Figure 10.2la is an enlargement of a certain part of the representative curve of


an elastic system, chosen at random. The curve is divided into segments, such as
n - I, n, and n + 1, and each segment turns into a straight line, as in Fig. 10.2lb.
11 Q DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

For each linear segment an equation like (10.2) and a solution like eqn. (10.11)
may be written. The motion in any linear segment starts with the displacement and
velocity of the end of the motion in the preceding segment:

This step-by-step process enjoys the advantage of adaptability to computer calcu-


lations. Like any other simplification, this process has built-in inaccuracies; sophis-
ticated mathematical techniques can reduce them to an acceptable minimum.
This representation of a curve by a succession of finite elements with an analytical
mathematical solution in a chain, one after the other, offers a mathematical way to
establish the response of a structural dynamic system with nonlinear components.

12 Torsion dynamic forces

m
(c) ;;c.===========t'll transverse
beam

Figure 10.22: Steel frame.

Figure 10.22a shows a one-storey steel frame with a lumped mass at the beam-
column joint.
The frame is one of several parallel structures carrying a transverse double heavy
beam for a crane (Fig. 10.22c). The mass of the frame's column and beam can be
neglected, and the axis of the transverse beam is located at the center of the joint.
Horizontal displacement of the system will include a rotation of the mass
(Fig. !0.22b). Hence the equation of equilibrium of dynamic forces should include
a dynamic rotational inertial force
M · x +Im ·ii+ k · x + k* ·a= F(t)
where
a =f(x),
k* = f(k), and
Im = rotational mass moment of inertia.
CHAPTERll

Multidegree of freedom: lumped mass system

1 Introduction to MDOF
1.1 A multistorey building

Figure II.la shows a simplified example of a MDOF system. It is a multistorey


building with a steel frame and h » fo.
In the following mathematical study, the total distributed mass of each floor is
considered concentrated in one single mass M. Hence M 11 represents the total mass
M of floor n.
If the vertical flexural stiffness of the floors is markedly greater than the flexural
stiffness of all the columns attached to them, the horizontal displacement of the

Xn+l
1--1>
Mn11

l, l, x,. k n+l
lb f-1>
M,.
l, l, k,.

f l:f lb f 0
M,

lo J,
x, k,
lb M,
k,
I, I,
(a) (b) (c) /,
./ /

Figure 11.1: A multistorey building.


112 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

building can be considered as originating in the deformation of the columns only.


The horizontal displacement can be represented as in Fig. 11.1 b. Descents offloor
levels can be neglected if the displacements are small. Hence the position of each
mass Mn is described by one single coordinate x11 •
Figure 11. lc is a schematic representation of this particular structural dynamic
system: lumped masses M, located in the joints of the springs k. The building is
represented by a vertical chain of springs, each one with the characteristics of the
columns' stiffness of the respective floor. The springs are interconnected at the
joints so as to ensure that the tangent of the deformed structure will remain vertical.
Lumped masses, each representing the total mass of the respective floor, are located
at the joints. The columns' masses have also been added to the lumped masses.
The horizontal motion of the building, when subjected to an external dynamic
event, can be represented by the horizontal motion of the lumped masses.
The position of each mass is described, at any given moment of the motion, by
one single independent coordinate x. The number of these coordinates defines the
number of degrees of freedom of the system.
For each of these masses in motion a differential equation of equilibrium of
dynamic forces can be written, similar to eqn. (10.2) in chapter 10:

M1 ·x1 +c1 ·x1 +kr ·xr =F1(1J


M2 · x2 + ci · .i:2 + ki ·xi = F2(1) (11.1)

Mn . Xn + Cn . Xn + k,~ . x~ = Fn(t)

The set of equations in (11.1) describe in full the motion of the structure at any
given moment. In matrix formulation this set of equations can be represented

M · x + C · x + K · x = F(t) (11.2)

The similarity of eqns. ( 10.1) and (10.2) in chapter 10 is apparent. An equation


of equilibrium of dynamic forces is written for each of the lumped masses, as for
a single degree of freedom (SDOF) system.
The superscript on k; in eqn. (11.l) is intended to indicate that the elastic force
in each equation is given by the contribution of all spring constants related to the
particular mass Mn. The superscript on x~ represents the relative displacements of
mass M~1 , not the absolute displacement x11 •

1.2 Displacement and rotation

In the example in Fig. I l .2a, a mass M is supported on shock absorbers of spring


constant k; the mass M is assumed to be the foundation of a machine.
The working machine introduces a vertical dynamic loadF(t) into its foundation.
This case can be analyzed as an SDOF system since the position of the mass in
motion is defined at any given moment by one single coordinate, which is the
DYNAMICS IN TIIB PRACTICE OF STRUCTURAL DESIGN 113

M
0
C.G.

Figure 11.2: Concentrated mass on spring supports.

vertical displacement x(t). The equilibrium of forces is described by eqn. (I 0.2) in


chapter 10.
If an external factor introduces an additional dynamic moment M, as indicated
in Fig. l 1.2b, the position of the mass M will be described by two independent
coordinates: vertical displacements x(t) and rotational angle a(t). Even though it
is a single mass system the structure should be analyzed for such a motion as an
MDOF system since two differential equations are necessary to completely describe
the equilibrium of dynamic forces:

M · x + k, · x =Fi · f(t) (11.3)


IM ·a+ ka ·a =Mi · f(t) (11.4)

where IM is the rotational moment of inertia of mass M and ka its rotational spring
constant; k 8 represents the vertical spring constant k of the shock absorbers. The
rotational movement of a mass originates on it an inertial rotational moment M1 =
a, x,
JM · similar to the inertial force Fr = M · originated by a linear translation x
of mass M. The effect of ka · a is to oppose the rotational motion.

1.3 Analysis of an MDOF system

The analysis of an MDOF system should start wiih determination of the number of
lumped masses and the number of independent coordinates necessary to define the
position of each mass at any given moment, since it may require more than one, as
already seen in the examples given in sections I.I and 1.2.
The number of independent coordinates defines the number of degrees of freedom
of the structure, not the number of lumped masses.
Once the number oflumped masses is established and calculated, and the number
of degrees of freedom is determined, calculation starts by finding the natural circular
frequencies of vibrations and the modal characteristic shapes of deformations for
each degree of freedom.
114 DYNAMICS IN Tiffi PRACI1CE OF STRUCTURAL DESIGN

2 Vibration modes

The natural circular frequencies Wn and modal shape ¢ 11 are both characteristics of
the structure, no matter what type of external dynamic load acts upon it. There-
fore, the prescribed mathematical way to determine Wn and </Jn is by working with
differential equations of the free motion, which in matrix form and disregarding
damping is:
M·x+K·x=O (l l.2a)

2.1 Mathematical process

Figure 11.3a, depicts a three-storey building with the basic stiffness and mass
characteristics of the building in Fig. 11.2.
Expansion of the matrix formulation (ll.2a) resolves into the following set of
simultaneous equations:

M1 · x1 + k1 (x1 - x2) = 0
M2 · x2 + k1(x2 - X3) - k1 (x1 - x2) = 0 (11.5)
M3 · x3 + k3 · x3 - k1(x2 - x3) = 0

Each of these equations, as already pointed out, is similar to that of an SDOF


system. Hence the solution will be in the form of a harmonic function; their general
expressions will be

x1 =di · sin(w · t + 9) (11.6)


x2 = d1 · sin(w · t + 9) (11.7)
x3 = d3 · sin(w · t + 9) (11.8)

where d1, d2, and d3 are constant values representing the maximum amplitude of
each function; 9 is a phase angle.

level I
lb
!, I, k
m,
• level 2 X2'4==:rl
lb
I, I, k2
ID3
level 3
lb
I, I, k3

(a) .. . . . , ... . (b)

Figure 11.3: Lateral deformation of a three-storey building.


DYNAMICS IN TIIE PRACTICE OF STRUCTIJRAL DESIGN 115

The second derivative results in functions for accelerations:


2
x1 = -w · d1 · sin(w · t + 9) (l l.6a)
2
x2 = -w · d2 · sin(w · t + 9) (ll.7a)
x3 = -w2 · d3 · sin(w · t + 9) (II.Sa)
Introducing these expressions for x and xinto eqn. (11.5) and rearranging, results
in a new set of equations:
(-M1 · w 2 +k1) · d1 -k1·d2=0
2
-k1 · d1 + (-M2 · w + ki + k2) · d2 - k2 · d3 = 0 (11.9)
-k2 · d2 + (-M3 · w + k1 + k3) · d3 =
2
0
with unknowns d1, d2, and d3.
A nontrivial solution of eqn. (11.9) exists insofar as the determinant of the
unknown' s coefficients is zero
(-M1w 2 +k1) -k1 0
-k1 (-M2w2 + ki + k2) -k2 =0 (I I.JO)
0 -k2 (-M3w 2 + k1 + k3)
Since masses M and spring constants k are known, solving this determinant
yields a frequency equation in w 2 . This equation has three real roots, each for one
degree of motion. Written in column vectors there are:

[:!] = eigenvlaues for


w1] natural circular
2
w2
eachDOF, [ = frequencies
wi
W3 eachDOF
for

3
A frequency equation like (11.10) can be written for any MDOF structure by
following the same simple mathematical path, and the result will be n natural
circular frequencies for n degrees of freedom.
This mathematical path becomes cumbersome for more than four degrees of
freedom, but it has the advantage of a clear introduction to MDOF structures. In
practice, matrix expressions and sophisticated mathematical techniques for eigen-
value problems are applied.

2.2 Modal shape

With eqn. (11.9), sets ofreal values for d1, d2, and d3 can be calculated, by substi-
tuting w by one of those determined for w1, wz, or w3.
To differentiate results, a special notation is used, as follows, again written in
column vectors:
116 DYNAMICS IN TIIE PRACTICE OF STRUCTURAL DESIGN

Each of these sets ford represents the geometric coordinates of a curve of dynamic
motion of the structure. Figure 11.4 is a general schematic representation of these
curves, which also helps to clarify the meaning of the subscripts.
Since eqn. (11.9) contains no information on the initial conditions of motion, nor
of the external dynamic force, values of d are relative and not absolute. Usually,
values of d in each set are normalized to one of them, which is given the value 1;
for example,

a1 = d11 · [d12~d11]
d13/d11
= [:::]
a13

which is a column vector of dimensionless coordinates describing the shape of the


curve of vibration of mode 1.
The square matrix of all modal shapes is

In reality the motion curve of a structure in a dynamic situation is only a single


one, with geometry changing continuously with time, and it represents at any given
moment the sum ofcoordinates of all modal curves at the same moment. In Fig. 11.4
the schematic representation shows the curves at a rare moment, when d11, dt2, and
d13, are all in the same direction, for maximum summation of displacements at the
top of the building. In practice, coordinates are summed by means of one of several
possible rules based on statistical considerations.

du
4-< ..r
=I

dz!
i- I
i
d31
I
-I
r I

fundamental second third


mode mode mode

Figure 11.4: Vibration modes.


DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 117

2.3 Example

As an example, let us give real values to masses and spring constants

[M1]-
M2
M3
- [0.013]
0.019
0.027
[t.s2].
-
cm
, [k1]- [
kz
kJ
- 7.78] [ - t ]
11.33
16.11 cm

Substituting these values in eqns. (11.9) and (11.10) results in:

WI]
w2 =
[12.53]
28.64 [c.p.s.]; a11] [1.000]
0.737 ;
[W3 40.51 [ = a21
a31 0.361

a12] [ l.000] ; a13] = [-1.744


1.000]
[ = -0.372
a22
a32 -0.799 [ 1.168
a23
a33

These values were calculated for a three-storey reinforced concrete building with
square slabs on four columns (Fig. 11.5).

8001800
40 ' ' rib slab roof
24\24

40 rib slab for heavy loads


0
0
27\2 7
"'
40 massi ve slab

0
0 30130
"'
/

Figure 11.5: Three-storey building with lslabs » Icolumns·

The characteristic shapes of the three modes are represented in Fig. 11.6.
Each curve represents a characteristic shape of a planar vibration which moves
harmonically with time, as expressed in eqns. (11.6), (11.7), and (11.8). The first or
fundamental mode has one fixed point Ai in its motion; the second mode has two
fixed points, A1 and Bz. The third mode has three fixed points, AJ, B3, and C3. By
extrapolation, motion of the nth mode will have n fixed points.
118 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

1.000 1.000 1.000

C3

'.,__-0.372 -1.744
I
I
I B3
I I
0.361 r- -0.799 --..., 1.168

A3
/ /

fundamental second third


mode mode mode

Figure 11.6: Modal characteristic shapes.

2.4 Natural periods of vibration

The natural periods of vibration T, instead of the natural circular frequencies w,


are of interest; for example, in Fig. 11.3,

Ti] [0.50] [s]


[T,T1 = 0.22
2ir
T=-·
w'
0.15
The largest period of movement is given for the first mode. The structure moves
relatively slowly in this mode and it has time to develop relatively large amplitudes.
Successive higher modes move with increasing velocity and decreasing amplitudes
of motion.
If the engineer is interested in deflections of a structure, for example, it may be
enough to consider the summation of the first three to five modes.
If the engineer is considering problems of vibration, such as those that can irritate
people, or if his or her attention is drawn to problems of acceleration that may affect
delicate laboratory equipment, higher modes become significant.

3 Forced vibrations
Even though each one of the equations in (11.1) represents the equilibrium of
dynamic forces, one for each mass of the system, it is the set of simultaneous
equations that should be used to obtain the circular frequencies w11 and the relative
coordinates of characteristic shapes of the modes of vibration </>M11 •
Once the physical components of the structure are known, and calculations of w11
and <PM,, have been performed, a new concept defining the mass that participates in
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 119

,,,,
F1· f(t)
~
F1·f1(t)
""'~ F2

F2·f2(t)
~ - F

(a) . % (b) , % (c) .. .%


Figure 11.7: Horizontal dynamic loads in a multistorey building.

any single modal vibration can be elaborated, whicb is equivalent to the contribution
of acceleration of masses M 11 in any mode n. It is the representative mass M of
mode n. Likewise, an equivalent spring constant, Kn, is defined, which represents
the influence of the stiffness of the bars of the system in any mode n.
The equation of equilibrium of dynamic forces for any selected mode is now

(11.11)
where Fn is an equivalent load, which considers the effect on the selected mode of
the external loads acting in the masses of the system. Elaboration of F11 requires
that all the loads acting on the system have the same functionf(t) as in Fig. 11.?a
and c; not that in Fig. 11.?b.
Equation (11.11) is now self-standing, which means it can be solved separately
for each mode and not as a part of a set of equations. This procedure is known as
uncoupling of the set of equations in (II.I).
Equation (11.11) is very interesting and useful. It enables the engineer to consider
each mode independently of all others. In reality the interdependence between
modes is now built in into the definition of equivalent mass, equivalent spring
constant, and equivalent external load for each mode.
Equation (11.11) is highly illustrative. It shows that under certain conditions,
which are fulfilled by the majority of linear elastic MDOF systems, a dynamic
situation of an MDOF system, subjected to a forced vibration, can be calculated
as the sum of SDOF systems. Hence consideration and conclusions known for the
latter, such as the effect of damping, the existence of a Dynamic Magnification
Factor (DMF), the resonance phenomenon, and more, are valid for each mode of
the MDOF systems.

4 Pulsating load
Let us consider the case of a building like that in Fig. 11.5, a machine with a heavy
rotating mass located at the center of gravity (CG) of the top slab.
120 DYNAMICS IN TIIE PRACTICE OF STRUCTURAL DESIGN

• .

,
Figure 11.8: Horizontal dynamic load on a three-storey building.

Any eccentricity of the rotating mass will produce two pulsating component
loads, one vertical and one horizontal. If the building is symmetrical, and we are
interested in the horizontal displacement caused by this dynamic load, the vertical
component can be disregarded.
Figure 11.8 is a schematic representation of the dynamic situation.

4.1 Equivalent valnes of mass, stiffness, and external force for


uncoupled modes

The weight of the machine should be included in the equation of equilibrium of


dynamic forces.
The masses in eqn. (II.I) include all masses in motion at the moment of the
dynamic event, even those defined as transitory or live loads in other static
considerations.
Let us assume here that the quantity of the forcing rotating mass is included in
the value for Mi, already assigned in the example in Fig. 11.5.
The equivalent mass, stiffness, and load for the fundamental mode are

+ m2 · a~i + m3 · a~i = 0.027 t · s2/cm


Mi = mi · afi (11.12)
- 2 2 2
Ki= ki(au - a2i) + k1(a2i - a3i) + k3(a3i) = 4.24 t/cm (11.13)

. Fi =Pi ·au + (P2 = 0) · a2i + (P3 = 0) · a3i =Put (11.14)

An alternative way to calculate the equivalent spring constant is


- 2 -
Ki =Wi ·Mi

Since any characteristic mode can be treated as an SDOF system, the equation of
equilibrium of dynamic forces can be written as eqn. (IO. IO) in chapter IO.

(11.15)
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 121

where ~i is the percentage of damping for mode I. The response of the MDOF
system, in the fundamental mode only, can be written as eqn. (10.12) in chapter 10

(11.16)

or
(11.17)

where
- -
Fi Fi
Xtst =-=- = 2
Ki "'i ·Mi
(11.18)
XJst = 0.236. P1

Performing the same calculations for modes 2 and 3 results in the following
values:
M2=0.100t·s 2 ·cm-i M3 =0.108t·s 2 ·cm- 1
K2 = 84.97 t · cm- 1 K3 = 100.75 t. cm-i
F2 =Pit F3 =Pit
X2nd = 0.0012 · P1; X3rd = 0.010 ·Pi
The horizontal displacement of the top level, for example, is given by the super-
position of the contribution of the three modes at that level:

Xtop level =Xi st. au . (DMF)i . sin (Q. t +Bi)


+ X2nd · a12 · (DMF)z · sin (Q · t + 62)
+ XJrd · a13 · (DMF)3 ·sin (Q · t + 63) (11.19)

Assuming that the values of all sine functions become unity at a given moment, the
maximum displacement of the top level is given by

Xtop level = Xist ·au · (DMF)1 + X2nd ·au · (DMF)z + X3rd · a13 · (DMF)3
(11.20)

Since this assumption is hardly likely, the value of displacement so calculated


can be considered a maximum.

4.2 Contribntion of the uncoupled modes

The real contribution of each of the modes to the horizontal displacement, see
eqn. (11.19), depends on the values of DMF, which in turn depend on the circular
frequency of the dynamic perturbation Q (Fig. 11.9).
122 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

DMF

5
~~!0%
n~14 c.p.s.~ 840 r.p.m.

DMF2+---1f
DMF3
I

2 3

Figure 11.9: Example of a DMF curve.

4.3 Modal effects

The calculation of moments induced in the columns depends on the relative dis-
placement between the bottom and top levels of each. For this particular case, where
the assumption I,1,b » Icolumn was made, see Fig. 11.10:

P·h3
L\= 12 EI
h
M=\~ 1 A

Figure 11.10: Horizontal deformation of a restrained colunm.


DYNAMlCS IN THE PRACTICE OF STRUCTURAL DESIGN 123

where

1'xki = ( 1'xki" 1'xk12 , 1'xk 13 ] = ( lau - a21 I , la12 - a22I , la13 - az3 I ]
6EI
Mkim~ = J;2 · (x1" · la11 - az1 I · (DMF)1 + X2nd · la12 - a22I · (DMF)2

+ X3rd · la13 - a23I · (DMF)3]

The maximum value of moment M also assumes that all sine functions of the
modes become unity at a given moment. Hence Mkimax can be considered larger
than the real expected value.
Even though the fundamental mode definitely appears dominant over higher
modes, a conclusion based on this observation should be made with some reserva-
tions; first, the picture that emerges from Fig. 11.9 changes from one structure to
another, and differences between amplitudes of modal displacements may not be
so large; second, if Q is not in the surroundings of w1, the participation of higher
modes, in the total horizontal displacement, becomes more valuable; third, higher
modes can be dominant if the engineer is looking for the acceleration response of
the structure at a specific point on it; and so on.

5 Modal analysis
The response of an MDOF structure subjected to an external dynamic effect,
whether dynamic load or foundation motion, can be analyzed in the way already
described for a lumped mass system under a pulsating load. It is known as the
characteristic modal approach.
Dynamic design situations exist where a direct analytical approach is more suit-
able, but the modal approach has the advantage of a clear understanding of the
structure's motion and response. Uncoupled modal equations allow extrapolation
of the knowledge of SDOF systems to the analysis of MDOF systems.
A variety of mathematical techniques facilitate the analysis of MDOF systems
with a large number of lumped masses. Some of them are apt for finding only a
few of the lowest consecutive modes, when this is deemed suitable for the purposes
of the particular case.

5.1 Summation of modal responses

The real motion of the structure is given not by each modal displacement but by the
summation of all modal coordinates, at any given moment, for each mass separately;
see, for example, eqn. (11.19).
If, for example, x11 defines the maximum amplitude of mass I, at mode I, then

x11 = X(n~l\, ·au · (DMF)1

X1n = Xnst • a111 • (DMF)n

I
124 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

and in general, for coordinate of mass M at mode n,

where this expression states: the amplitude of motion of a mass M at mode n is


obtained by multiplying the so-called "modal static displacement" by the modal
participation factor and by the DMF for mode n. The DMF is to be taken from a
case of an SDOF system with Wn, subjected to the same dynamic load.
The general expression describing the motion of mass M is

which represents the summation of harmonic functions, with particular amplitudes


XMri of all modal motions.
Determination of XM from eqn. (11.21) is a task usually performed by statistical
tools such as:
- AVS, the summation of absolute values of all modes at any joint coordinate
(This method is considered as giving exaggerated values of deformation);
SRSS, the square root of the sum of the squares;
- GMC, general mode combination;
- CRC, complete quadratic combination.
These rules of summation are valid for a structure subjected to an external
dynamic force, with or without initial conditions, Xi, Xi, of the motion.

6 Damping in MDOF

The estimation of adequate values of damping to be introduced into calculations


of a strncture is a task that requires some degree of judgment for MDOF as well as
for SDOF systems.
There is a certain interaction between the values of the amount of damping to
be introduced into the different modes, especially for nonlinear or plastic response.
But since there is no way to determine the structure's critical damping a priori, at the
design stage, the practice is based on some agreed values for the probable percentage
of the critical, 1% :S $ :S 10%. It depends on whether it is a "naked" strncture or has
additional nonstructural elements, whether the skeleton is monolithic or assembled
from prefabricated components, whether it is steel or reinforced concrete, whether
prestressed or not, and so on.

7 The lumped masses


MDOF systems are based on the assumption that masses of the structural elements
are concentrated in the joints, which is, of course, a simplification of reality.
The mass of a beam should be distributed towards its two ends, for example,
with consideration of both the distribution of the mass and rigidity along the beam,
and conditions of support at both ends.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 125

(a) ~,.,-------,;=1,-------:z::,_,,_B

A
(b) l~t-1-------0~'0-l-----::6_7'B
Figure 11. ll: Schematic beams.

Figure 11.11 a shows a simple, supported, isostatic beam, with uniform mass and
rigidity all along. The lumped masses will be MA = Mn = m · 1/2. Figure 11.11 b
presents the same beam but with an elastic restrain atA; MA > MB. The calculation
of lumped masses is based on the acceleration of motion of the beam under the effect
of an external dynamic rotational moment at one of the supports, such as MaB.

S1
clL cl
'
s,
c'L c'R

S3

c1 c'
'
,
Figure 11.12: Vertical frame.

In practice, for a frame like that in Fig. 11.12, it is acceptable to first consider
lumped masses at joints, resulting from halving beams', columns', and slabs' masses
for a preliminary design, since at this stage not all the physical characteristics of
the structural elements are fully defined. Later more sophisticated definitions of
the structure are possible, with a higher number of degrees of freedom and more
detailed physical characteristics.
The lumped mass system is a sound tool for solving structural dynamic problems.
However, it is a tool based on approximations, and the accuracy of the results can
always be further refined if necessary or requested.
Powerful computer programs enable the engineer to solve highly complex struc-
tures with relative ease.
CHAPTER12

Distributed mass system

1 Introduction

Real structures are continuous distributed mass systems (DMSs), not necessarily
of uniform mass distribution. Accordingly, the response of a structure subjected
to a dynamic load will be better evaluated by a mathematical approach that con-
siders continuous mass structural elements instead of the approximation present in
the mathematical treatment, which transforms distributed mass structural elements
into discrete coordinates systems with lumped masses, whether SDOF or MDOF
systems.
Formulation of the response of a space structure with distributed mass elements
and distributed dynamic loads will include time-varying position and rotation space
coordinates. The formulation for such dynamic situations can lead to mathematical
expressions with unknown analytical solutions. Some of the latter can be managed
with numerical mathematical techniques. But then the question arises whether in
such cases it may be more convenient to work with an MDOF structural system.
Because of the mathematical difficulties, the DMS approach is confined mainly
to one- and two-dimensional systems, such as beams and plates.
Basically, the vibrating response of structural elements with continuous mass
distribution can be described as a superposition of modal displacements, each with
its own characteristic shape and natural circular frequency, as was done for MDOF
systems.
Figure 12.1 shows an ordinary beam with uniform distribution of mass m and
stiffness EI all along; in Fig. 12.lb the beam length is subdivided into segments a
and b; in Fig. 12. le the uniform distributed mass has been transformed into lumped
masses M, and Mb, interconnected with supports by massless springs; Fig. 12.ld,
e and f shows the three characteristic shapes of the modal displacements of the
MDOF system with length b =length a; Fig. 12. lg, hand i shows the characteristic
shapes of the first three modal displacements of the DMS system.
128 DYNAMICS IN THE PRACTICE OF STRUCTIJRAL DESIGN

Note the similarity of the modal shapes of the MDOF and of the DMS. Dis-
crepancies in values of w result from the simplification that represents the lumped
mass system when compared with the real DMS. If, for example, the length of bin
Fig. 12.1 is changed to 50% of a, the values of w1, w2, and w3, become, respectively,
28.07, 142.5, and 220.4.
If the number of segments, and hence the number of lumped masses, is increased,
the number of modal shapes of the MDOF system will increase too, and conse-
quently the precision of the values calculated both for w aud for the modal shape
characteristics will also increase. Obviously, if the number of segments is infinite,
the results will be equal to those calculated for a DMS.

IP200
(a)L
m.EI ::L A""' 33.65 cm1
I'°" 2200 cm 4
--fJ c:1-
(b) L ::L
y, a b a y,
(c) _LS;___ 0 0 0 ::6._

(d) MDOF syste1n


' - - , c. M" Mb M.-,, - - - - 1st 1node
-o-----Et-----\:/- ro1 =2s.20

(e)
----0---- M"DOF system
2nd mode
---o--- 002,.,116.8!

---0, _-0--_ MDOFsystem


(f)
3rd n1ode
003"" 260.72

DMS system
(g) 1st mode
OOJ = 28.48

(h) DMS system


211d n1ode
002 = !!3.92

(i) ···--··-·-----

-----
--·-- DMSsystem
3rd motle
003 = 256.32

Figure 12.1: Isostatic beam; modal shapes.


DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 129

The dynamic response of a beam calculated as a system with continuous mass


distribution results from the superposition of an infinite number of modal shapes'
vibrations. In practice, it is usually enough to consider only a few modal shapes.

2 Mathematical approach
Figure 12.2 illustrates a beam subjected to an external arbitrary load q(x). The beam
is of uniform rigidity EI and constant transversal section all along.

1IrI[q(x)
II
m; El; I

Figure 12.2: An isostatic beam subjected to arbitrary load.

For a vertical displacement y of the beam, the basic relations between moment,
shear, and external load can be written, from statics,
d2y
M=-EI- (12.la)
dx2
d3y
V=-EI- (12.lb)
dx3
d4y
q(x) =-EI- (12.lc)
dx4
Equation (12. l) is valid for a slender beam and small displacements. The effect
of shear is disregarded in these equations.
The equation of equilibrium of motion of the beam subjected to an arbitrary and
time-varying load p(x, t) can be elaborated by writing the equiliblium of dynamic
forces in any differential element of length ds as follows (Fig. 12.3c). For small
displacements ds ~ dx. The derivative of vertical displacement is written as a partial
derivative, since y is function of x and t. In eqn. (12. lc), load q(x) should be replaced
by the effective load in the beam in motion, which is
a2y
q(x, t) = p(x, t) - m at2. (12.2)

Replacing expression (12.2) in eqn. (12. lc) resolves as


a4y a2y
EI . ax4 +m . a12 = p(x, t) (12.3)

Equation (12.3) is the elementary differential equation of equiliblium of dynamic


forces for the motion of a slender beam. The effect of the inertial rotational mass
moments is neglected, which is a valid assumption for small displacements.
PJip(x,t)
130 DYNAMICS IN: THE PRACTICE OF STRUCTURAL DESIGN

m; El;~

(b)
elastic axis

(c)

Figure 12.3: An isostatic beam subjected to an arbitrary dynamic load.

This equation is very interesting because it contains two variables, which means
that the expression for y should contain the same two variables

y(x, t) = fn(t) · </Jn(X) (12.4)

where f(t) and ¢(x) are functions of one single variable each: one for the time-
varying function and the other for the location coordinate function.
The elementary differential equation of motion (12.3) can be extended to include
damping:
a•y a2 y ay
EI · -
ax4
+ m·-
a12
+ c · -ar = p(x t)

(12.3a)

Introducing expression (12.4) into eqn. (12.3) yields the function ofresponse of the
structure, y(x, t). Hence the following conclusions:
1. The free motion of the beam is characterized by modal vibrations.
2. The modal characteristic shapes are sine curves

¢,,(x) =sin (n·TI·X)


1
DYNAMICS IN THE PRACTICE OF S1RUCTIJRAL DESIGN 131

---::c,._-.- <..>n= n2.1t2 /Ef


(a) -"-"-"'----,-11;-:E"'I-;-1""· pi ,.,;m

(b) %11--------=---::c,._-.-
~ m; EI; f

(c) ~1----,-11,'°"·E"'J-;-1""·----l~
Figure 12.4: Natural circular frequency for beams with different support conditions.

3. The time function of the modal vibrations is harmonic with the natural circular
frequency w 11 for each mode

fn(t) = c' sin (wnt +en)

4. The number of modes is theoretically infinite, but in practice the study of


particular cases is usually based on a limited number n of the lowest modes only

l<n<oo

5. Each mode of vibration has its own natural frequency w11 , which is a function of
mass m and rigidity EI. The expression for Wn varies according to the support
conditions, as shown in the examples in Fig. 12.4.
6. The response of the beam subjected to an external loadp(x, t) can be elaborated
separately for each normal mode by means of an equivalent SDOF system as
already explained under section 3 of chapter 11.

Yn(X, t) = LAn(t) · <f>n(x)

Ynm,, = LAn,t · (DMF)n · </>n(X) (12.5)

where (DMF)n is the same Dynamic Magnification Factor as that for a one-
degree lumped mass system with the same circular natural frequency w, and
subjected to the same external load.
7. The vertical displacement of a beam in a forced motion is obtained by the
superposition of the modal shapes:

y(x, t) = LAn(t) · </>n(x)

where An(t) is the function of the modal amplitude.


132 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

3 Design of a beam

For static situations, the design of a beam is a simple and straightforward process:
calculation of moments and shear forces, and control of stresses and deflections.
By contrast, in dynamic situations the process is the reverse since finding the
response of the structure usually means establishing the maximal vertical displace-
ment of the vibrating beam [see eqn. (12.5)] and then proceeding with the design
of the beam by a process that is the reverse of that just described.
A question arises as to how many normal modes should be taken into consider-
ation in addition to the fundamental mode. It is an issue that the engineer should
tackle separately for each particular case; usually, a few of the lowest modes will
be enough.
In certain cases such as a beam with uniform load, in which the natural frequencies
are not too close in the lowest modes, the fundamental mode can possibly give
satisfactory results. But it is better to consider a few of the lowest modes.
Part 5

Natural dynamic loads


CHAPTER13

Earthquakes

1 Introduction
Earthquakes are dramatic natural events that release enormous amounts of energy
with the potential to cause serious damage of disastrous consequences.
To try to meet the human need for structures of diverse kinds, with an acceptable
level of building strength, antiseismic design requires from structural engineer-
ing high theoretical knowledge, expert interpretation of observations collected on
disaster sites, and a thorough understanding of the catastrophic process.
A sound knowledge of the different factors that affect the earthquake resistance
capacity of a structure should underlie the engineer's design to Codes of Practice,
since the antiseismic weakness of a structure unfortunately only becomes evident
when it is just too late for appropriate correction measures. Poor conceptual design,
incorrect static calculations, insufficient or wrong detailing of the structure to be
built, or bad implementation of design requirements at the construction site will
result in a structure incapable of withstanding the seismic event expected for the
region where it is erected.
Design for earthquake resistance should start with the choice of an adequate
structural concept and anticipation of the probable structural dimensions for the
expected dynamic loads. Since those loads, in turn, depend on the characteristics of
the structure, the final dimensions will result from a somewhat iterative process: a
preliminary calculation of forces will lead to a check of stresses and deformations,
which in turn will lead to a redesign of the structural elements; the latter will affect
the value of the dynamic forces for a new cycle of check and design. This iterative
process should hopefully yield levels of stresses and deformations, in the frame of
an economic solution, which meets predefined functional and esthetic requirements.
When considering the effects of an earthquake in the design of a structure, the
engineer should always think first of the possible tragic consequences of these
natural events, and second of the cost of his solution.
136 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Obviously, when designing a structure the engineer's first concern is for the
forces arising from a seismic event, but attention should also be paid to the possible
effects of the seismic activity on foundations and soils, which are not automatically
included in the dynamic analysis of forces and deformations. Examples are earth
movements and cracks that can dislocate supports, liquefaction of soils that can
produce dangerous tilting, unbalance of loads and even overturning of a structure,
landslides, and more.
The generally constant need to design within economic constraints results in
elastoplastic static design and calculations, with due consideration for adequate
levels of ductility. Some of the expected consequences of such an approach may
be damage to nonstructural and infill construction parts and finishes as well. So
far, no standard economic technological solutions have been elaborated for these
"side" consequences. If severe damage to these secondary building components is
to be avoided, static calculations should be based on the linear elastic response of
structures, whose outcome will most probably be an expensive solution. Serious
damage to nonstructural elements can also be caused by self-inertial forces, which
may challenge their strength and overall stability.
Special emphasis should be placed on good and proper detailing of structural
members and their connections, with particular attention to the reversal effect of
earthquake forces that may cause alternate tension and compression stresses and
strains. Proper detailing, sometimes even with some redundancy, may prove crucial
for the structure to successfully withstand unexpected earthquake events.

2 Earthquakes

Earthquakes are natural sporadic events that may happen randomly in time, place,
or intensity anywhere on the world map.
The most accepted explanation for these natural events is that they are the result
of fractures, unpredictable in time, in the crust of the earth, caused by tectonic
forces and volcanic activities. Tectonic activity can be of such a magnitude that it
creates valleys and mountains. But these events happen once in thousands of years.
Its more frequent effects can for the most part be resisted by structures designed
and constructed adequately.
The earth is covered by a mantle with a crust on its surface. Inside this mantle
there are continuous tendencies to local motions, which initiate elastic strains in the
rock. When the rock can no longer withstand these strains it cracks, and a certain
amount of energy is suddenly released. The fracture of the rock can take place in a
bounded region at any depth. This is known as the hypocenter of the earthquake. Or
the fracture can happen along a plane that is kilometers long, known as a fault, or
a fault trace when it ruptures up to the surface of the crust. The vertical projection
of the hypocenter to the surface of the crust is known as the epicenter.
Internal motions in the earth take place continuously, and most of them are
recorded by instruments called seismographs. The study of these motions belongs
to the field of seismology. These motions are of different levels of strength. The
engineer is concerned with "moderate-to-strong" motions only. Lower levels are
DYNAMJCS IN THE PRACTICE OF S1RUCTURAL DESIGN 137

mainly of the interest to seismologists, while higher levels belong to a category of


catastrophic events that structures will not withstand.
The shear rupture of the rock, when it cannot sustain the continuously increasing
elastic strain, originates a release of energy together with slippage movements of
antagonistic parts of the rock. It generates ground waves which propagate in the
crust, with a very complex pattern including reflections and refractions, until they
reach the surface of the earth.
The damage that the earthquake will cause in human construction work is the
principal objective of the engineer's study of this natural phenomena. It depends
on many characteristics, which can be detected by appropriate instruments that are
conveniently distributed. Seismographs record ground acceleration, velocity, and
amplitude of displacements. Of the utmost importance is the complementary infor-
mation provided by observers' descriptions and factual details from disaster areas.
Antiseismic engineering design practice combines continuously evolving profes-
sional theoretical knowledge, interpretation of extensive collected field data, and
on-site witness descriptions. The engineering design criteria also include consider-
ations of economic risk.
Two expressions are commonly used to describe the strength of an earthquake
are intensity and magnitude.
Intensity is a quantitative measure, besides being a qualitative description, of the
size and gravity of the seismic event. A known scale is tbe twelve-grade "Modified
Mercalli" (MM) scale, given in Roman numerals (I-XII). Earthquake intensity
diminishes with distance outwards from the epicenter. Although scales of intensity
provide rich and valuable information on the earthquake event, they do not furnish
measures to be used in practical design since they are not based on recording
by instruments. The description of an event is usually depicted on an isoseismic
map (Fig. 13.1). The somewhat concentric lines are defined by local geological
anomalies.

VI

VII

natural margin

Figure 13. l: Diagrammatic presentation of an isoseismic map.


13 8 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Magnitude is related to the energy released by an earthquake and is measured on


a decimal logarithm scale
M=logA-C
where

- A is the maximum amplitude recorded in a seismograph and


- C is a constant, selected from a particular earthquake.
Magnitude is independent of the location of the seismograph, owing to ade-
quate correction of distances between its location and the source. Even though the
magnitude of a registered earthquake provides a quantitative measure of the event,
recorded diagrams of accelerations, velocities, and displacements provide the basis
for practical design, especially the amplitude, frequency, and duration of the event
in the acceleration diagram.

3 Earthquake loads

Design for earthquakes presents the engineer with the most problematic situation
he or she can find himself or herself in as a designer considering dynamic loads.
It means describing tbe load, qualitative and quantitative, to a certain acceptable
level of reliability. Moreover, once he or sbe has made the choice about the load,
another difficult question arises, namely if the chosen load will appear during the
life of the structure. These two uncertainties, the type and magnitude of the load
and the probability of the load occurring, exert a strong and permanent influence
on the entire process, which includes design, static calculations, and detailing and
construction of any structure to be erected in a seismic area.
Loads of earthquake origin are highly irregular, variable in amplitude, and involve
a superposition of frequencies, of long or short duration, that can act on the structure
in any possible direction, horizontal or vertical. Depending on the length of the
structure, a strange situation can arise: the extremes of one single structure, for
example, a long bridge, will be submitted to different earthquake motious. Not
only external loads can be created by ground motion but external moments too.
Because of this uncertainty in the earthquake loads the common practice is to
consider antiseismic design as divided into two stages:
1. Design for moderate seismic loads: in this case it is expected that the structure
will resist the event without damage at all, or atmostwith no significant damage.
This stage of design can in some ways be likened to a serviceability limit stage.
Of course, cracks and deformations cannot be predefined, but slight damage
that will require only moderate repair work can be expected.
2. Design for strong and severe earthquakes: the engineer's chief concern is to
prevent the collapse of the structure, in part or whole, to ensure no serious
human injury or loss of life. This stage can in some way be likened to an
ultimate limit design. Repair of the structure may prove too expensive or even
impossible. This approach is justified when the economic consequences are
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 139

weighed against the investment cost required for further upgrading of designs.
It is actually a design-to-risk approach.

4 Earthquake response analysis

At the beginning of the analysis it should be clearly established whether the response
of the structure is expected to be linearly elastic, for a relatively moderate ground
motion; note that this task is extremely difficult to accomplish. Structural response
usually occurs in the plastic range for relatively severe but moderate earthquakes.
The structural material responds elastoplastically, and the infilling nonstructural
components will evidence some cracks and dislocation. All these responses are cl-
assified into categories ranging from negligible to moderate or to serious damage.
Considering, first, the horizontal displacement of a structure, it is simple to imag-
ine that the inertial force originated by the motion is represented by a certain per-
centage of the weight:
F = ry · W = (% · g)W.
Still to be established, of course, are the amount of this percentage and the
distribution of force F along the structure.
This static approach is widely implemented in the local codes because of its
simplicity. The value of ry is established with due consideration to the type of
structure and its structural material, its natural period T, soil characteristics, and
acceptable values of ductility. The distribution ofload along the structure is provided
by the appropriate formula. This is basically a static, lateral, seismic shear force
approach. It can be used in design calculations, provided the structure meets very
strict requirements and limitations. If these are not fulfilled another approach to
design calculations should be taken.
Since structures are in reality MDOF systems, the modal approach is, for linear
elastic response cases, second best to the static shear coefficient approach just men-
tioned. Two alternatives can be considered for MDOF systems: the linear response
spectrum or the step-by-step analytical calculation. The linear response spectrum
approach is relative simple and has the advantage that the engineer can "feel" the
contribution of each mode to the structure's response. The step-by-step alternative
has the advantage that it is adequate for computer calculations.
Nonlinear inelastic response is treated by solving the equation of equilibrium of
forces for a previously selected earthquake record. Nonlinear analysis is by math-
ematical methods for numerical integration since analytical solutions are usually
not possible. Inelastic dynamic analysis is highly complex and not justifiable in the
light of the uncertainties involved in the selected description of the earthquake that
is assumed will affect the structure in the future.
A relatively new approach to nonlinear response analysis is becoming used
more frequently, namely the pushover analysis for multistorey buildings. The
behavior of a structural system is learned by applying a lateral load, which is
increased consistently up to the point where the top level of the building reaches a
140 DYNAMICS TN THE PRACTICE OF STRUCTURAL DESIGN

preset measure of displacement; usually the lateral load distribution fits the
fundamental mode.
The pushover approach has two advantages over the lateral shear approach: first,
it represents more adequately the inertial characteristics of the structure and it can
also be adapted to the history of a ground excitation; second, it can be elaborated
for a multimode response. This latter possibility can provide a practical way to
mitigate the existing conflict between architectural design and the stiff limitations
demanded by experts for good performance in earthquakes.
The approaches mentioned so far assume that the anticipated ground motion can
be formulated between certain limits of probability. Hence these approaches can
be called deterministic.
Information on severe ground motions is still inadequate since both earthquakes
and structural response to seismic events are highly random, so a stochastic approach
is also deemed appropriate for seismic structural response analysis. This approach
is definitely the field for experts.

5 Static force procedure

This procedure enables the engineer to perform the static calculations very simply.
It is a two-step procedure.
First, a preliminary design should include considerations related to the "archi-
tecture of the structure," which means fulfillment of the basic requirements, clearly
established in the Codes, aimed at obtaining a more or less regular structure in
the three space directions. The intention is to obtain as uniform a distribution of
masses as possible, with no discontinuities in either geometry or strength. The main
objective of this step in design is to ensure a noninterrupted, regular, and smooth
path for the flow of the seismic loads down to the foundations. A sound structural
configuration will provide for an adequate response to moderate earthquakes while
ensuring the required strength to overcome local weaknesses, unexpected member
forces that originate in some small torsion asymmetry, or local concentrations of
stresses resulting from constructional defects.
Second, the seismic input is given as a quasi-static load, and assuming a certain
amount of damping ~,

F = K(µ.) · K(s) · Z · W (13.1)

where
K (µ.) considers the structural material;
K (s) considers the type of soil under the foundations, type of structural framing,
and a certain amount of ductility in the deformation of the structure;
Z is the seismic acceleration input as a function of T, the natural period of the
structure;
W is the total weight of structure.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 141

Though the force obtained from eqn. (13.1) represents an external static load, due
consideration has been given to the dynamic characteristics and the soil foundation
interaction of the structure subjected to a seismic event.
The seismic acceleration input is given in the form of a specific spectrum; accel-
eration a= % . g versus period T. It represents the characteristic value of a chosen
ground motion with a particular magnitude, at a certain distance from the site of
erection of the structure.
The force obtained from eqn. (13.1) represents a horizontal seismic static load.
The vertical static load is usually only a certain percentage of that force.
If the particular structure under design is a high building, the force obtained from
eqn. (13.1) is intended to represent the maximum shear force at the bottom of the
structure. It is usually assumed to be distributed over the height of the building, as
described by the following expression:

W·h
Fi=-'-'F (13.2)
w
where W; is the weight of floor i and h; is the height of floor i from the bottom of
the structure.
The shear force procedure for the building assumes that there are no weak and
no soft floors.
The vertical distribution of forces F; given by eqn. (13.2) was adopted after
observations that the vertical lateral seismic deformation of buildings is closer to a
straight line than to that of a vertical cantilever (Fig. 13.2).
The static lateral force procedure is a legitimate approach to antiseismic design
for structural configurations as close to the symmetry as possible, and with a clear
chain of sequence yielding progress throughout, from the supersttucture to the
foundations.

I
I

(a) (b) (c)


. ., . . . . / 1 cantilever line
2 building defonnation

Figure 13.2: Cantilever high-rise building.


142 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

6 Linear elastic response spectrum


Figure 13.3a shows a very simple structure: a one-storey building frame with a
horizontal slab, on light columns. It can be considered, for earthquake response
study, a single mass Mon a single spring k.

,v Xg

OM I <l--O

l MX:g
I I
I I

(a)
Xg 7r i 7v X
/

(b)

Figure 13.3: Base displacement of a symmetric frame.

For a horizontal earthquake motion, the equation of equilibrium of forces is

M·Xg+c·x+k·x=O
M (x8 + x) + c . x + k . x = o
or
M ·x+ c ·x+k ·X = -M ·Xg
(13.3)
M · x+ c · x+ k · x = Fg(t)

where Fg(I) =IM · Xg I; the sign of Fg(t) can be disregarded since the earthquake
motion can have any direction.
Equation ( 13.3) is the very well known SDOF system's equation of equilibrium of
forces. Here it becomes evident that the mathematical treatment of the problem does
not differ whether the external load that shakes the structure is from an earthquake
motion or from any other dynamic load. Note that the seismic load originates in the
ground displacement Xg and not in the relative motion xg between the oscillating
mass M and its supports.
If the structure is under the effect of any other dynamic load, Fp(t), at the moment
the eart~quake occurs the equation of motion becomes

M · x+ c · x+ k · x = Fg(t) + Fp(t). (13.3a)

Solving eqn. (13.3) is merely a deterministic problem once the ground motion
is defined. For a linear system the integration of the differential equation can be
performed with the Duhamel integral:

x(t) = - 11'
w 0
Xg(r) · e-<w(t-<) ·sin w(t - r) · dr. (13.4)
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 143

Equation (13.4) is solved using a computer, applying a numerical step-by-step


method.
As already mentioned, x(t) is the motion of mass M relative to its foundation.
This is actually the objective of this study, not the value of Xg, which can be easily
calculated by addition of Xg . Internal forces and deformations, stresses and strain,
velocities, and accelerations in the structure can be calculated from x(t).
Usually the integral of eqn. (13.4) is defined, with quite adequate accuracy, as
the pseudo-velocity spectrum

PSv(T, ?) = [lo' Xg(r) · e-«o(I-•) ·sin w(t - r) · dr] lmox

and following this definition

Sd = 1/ w · PSv, the relative spectral displacement,


Sa = w · PSv, the absolute spectral acceleration.

The difference between the spectral velocity Sv that can be obtained by differ-
entiation of eqn. (13.4) and PSv is negligible if Xg is long enough, as happens in
reality.

6.1 SDOF spectral response

Assuming that a series of SDOF oscillators with the same damping factor$, but
different periods T, are subjected to the same earthquake motion, the result will be
a family of the three spectral responses for each oscillator:

Sd(T, ?); Sv(T, ?); Sa(T, ?l

Since the loads are originated by seismic ground motion, it is only natural to have
recourse to past earthquakes records and attempt to extrapolate from them to pos-
sible corning seismic events. The number of available earthquakes for a particular
region is not always enough to rely on for practical applications; this information
is complemented with statistical techniques. Artificial records of earthquakes are
mathematically elaborated, together with proposals for some "representative" hypo-
thetical ground motions, and many more ways that the experts in such a special field
constantly elaborate on the basis of the rising numbef of recorded natural real events.
Elaborated ground motions always reflect the basic knowledge that peak ampli-
tudes of accelerations are found at higher frequencies, while those of velocities on
the moderate and displacements appear at lower frequencies.
A simple way to represent the expected peak structural response amplitudes
of acceleration, velocities, and displacements for an anticipated earthquake is by
means of a design spectra diagram, drawn on logarithmic scale, on a tripartite grid.
One single graph for these three spectral response values is usually drawn, either as
a function of the natural frequencies/ or as a function of the natural period T of these
oscillators. The frequency is the reciprocal,/ = l/T, of the period T. The diagrams
144 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Sv [crn\scc]

I
J
elastic design spectrum
(~ = cy;, Ccr)

T[see]

Figure 13.4: The response spectrum for a linear SDOF system.

are elaborated for rigid supports and a predefined value of the damping factor s
(Fig. 13.4).
The representation in one single curve enjoys advantages from the approximate 1
relations between the three response spectra:
Sd = 1/ w · PSv = spectral displacement,
Sa = w · PSv = spectral acceleration
where PSv = pseudo-spectral velocity.
The response curve approaches ground acceleration for rigid system'\ and ground
displacements for flexible systems. A rigid or stiff system has short periods; a
flexible one is characterized by long periods (small frequencies). Response curves
are given indiscriminately as a function of frequencies or period of the structure.

6.2 MDOF response spectrum procednre

Real structures are continuous mass systems, although most can be conceptually
reduced to multidiscrete systems. For earthquake analysis, their dynamic response
can be obtained by superposition of modal shapes, easily performed using a com-
puter. The real structure is thus transformed in an MDOF system.
The modal shapes of vibration and their natural periods are inherent in any struc-
ture and can be determined independently of the characteristics of the external load.
The equation of equilibrium of forces for the MDOF system is similar to that of
an SDOF system, eqn. (13.3), but in a matrix form

M · x + C · x + K · x = F g(t) (13.5)

where M, C, and K, are matrices of mass, damping, and stiffness, respectively.


DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 145

The eigenvalues and the eigenvectors resulting from the mathematical process
involved in the solution of eqn. (13.5) are w, the natural circular frequencies, and
<j;,, the modal shapes of the natural modes of vibration.
For each mode of vibration an equation of equilibrium of dynamic forces can be
written
(13.6)
where

Mn* = <P~ · M · <Pn = generalized mass,


c; = .p!, . c . .p,, = generalized damping,
K,; = <P~ · K · </J11 = generalized stiffness,
F;(t) = .p;, ·Fa · <j;,, =generalized ground motion load.
Instead of continuing hereafter with the mathematical process, a simpler way is
possible by relying on the response spectrum, like that of Fig 13.4, for the appro-
priate ground motion in consideration: for each mode the maximum amplitude of
response, Sdn, can be extracted from the response spectrum as a function of the
period of the mode, which is

In practice, one single value for the damping factor~ can be selected for all modes.
The uncertainties about introducing different values for damping, whatever they
may be, in each mode, do not seem to be justified.
The next step is to proceed to the summation of the maximum amplitudes of
response. Usually it is enough to consider a short number of modes. With computer
programs a selection is generally conducted for the first 10 modes.
As for the summation of responses, a statistical rule should be applied since not
all the amplitudes of response reach a maximum at the same time. Some of these
rules are:
AVS, the summation of absolute values of all modes at any joint coordinate (this
method is thought to give exaggerated values of deformation);
SRSS, the square root of the sum of the squares;
GMC, general mode combination;
CRC, complete quadratic combination.
These rules of summations are valid for linear elastic structures.

7 Analytic procedures for linear elastic response

The response spectrum procedure described in the preceding section is a simple and
acceptable way to analyze a linear structure, for which the principle of superposition
holds. An essential condition is to have at hand, of course, the response spectrum
146 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

for the selected ground motion to happen during the expected design life of the
structure.
In general, the dynamic analysis of a structure subjected to an earthquake starts
with the equation of equilibrium of dynamic forces, namely eqn. (13.5).
By neglecting damping and considering the structure itself, before the earthquake
takes place, the equation of equilibrium is reduced to
M·x+K·x=O
from which the eigenvalue w;,
and the modal vectors f/>n, of the natural modes are
obtained. From here displacement can be written for each mode separately
x,,(t) = t/Jn • q,,(t)

where q,,(t) is the amplitude of the generalized coordinates of mode n. For each
mode it is possible to define new characteristics, known as the generalized charac-
teristics already mentioned in the preceding section.
Equations of motion can be written separately for each mode of vibration:

q.. . + (lV11*) 2 ' q11 =


+ 2"5 • Wn* • qn F*/M*
n 11 • (13.7)

7.1 The time domain

Integration of this differential equation can be performed in the time domain with
the Duhamel integral, which in tum is solved by a step-by-step process. During
this process the stiffness of the structure is unchanged as it is given in the form of
a linear elastic function of displacement. The irregular function of the seismic load
is decomposed in a succession of linear steps (Fig. 13.5).
For each step a linear equation [eqn. (13.7)] is written, so that the initial con-
ditions of displacement, velocity, and acceleration are taken from the end of the

F(t+ll)t

F (t)

Figure 13.5: A chain of successive linear segments in substitution of a curve.


DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 147

previous step. There are several numerical methods to work out this step-by-step
integration in computers.

7.2 The frequency domain

Equation (13.7) can also be solved in the frequency domain, with the Fast Fourier
Transform, applicable for linear systems.
The irregular input of the ground motion is expanded into a summation of sine
and cosine functions:
00

F;(t) = ao +L [an· cos(Qn · t) + bn · sin(Qn · t)] (13.8)


n=l

The advantage of a Fourier series representation of the earthquake load is that it


provides information on the frequencies and amplitudes of the harmonic compo-
nents of the irregular function representing the earthquake input, thus pointing to
eventual situations of resonance with one of the natural modal frequencies. Espe-
cially, situations can arise of resonance with the fundamental or some of the lower
natural modes of frequencies.

8 Nonlinear inelastic response


From time to time ordinary structures are subjected to earthquake forces sev-
eral times greater than those that maintain their response within the linear elastic
range. They undergo inelastic deformations and local damage.Actually the inelastic
response is the mechanism that helps structures to resist severe ground motion. The
yielding of structural materials indeed contributes to the capacity of the structure
to resist earthquake forces. But bear in mind that this comes with a certain loss
of strength. The yield limitfy and the ductility ratioµ, are two main factors in the
analysis of inelastic response.

8.1 Step-by-step integration

These dynamic situations can be analyzed using computers, applying a step-by-step


program. This mathematical technique is simifar to that implemented for linear
response. The main difference is that a continuous degradation of the structural
capacity has to be introduced in the calculations. The mathematical procedure is by
integration of a sequence of differential equations of dynamic forces, at the begin-
ning of each step, for linear structures with new degraded properties (Fig. 13.6).

8.2 Linear approximation

Some simplification can be introduced in the analysis of inelastic response by


linearization of the stiffness characteristic of the structure (Fig. 13.7).
148 DYNAMICS IN TIIE PRACTICE OF STRUCTURAL DESIGN

Pk (t+At) F k (t+iit)
F k (t) h(t)

~< At ~v
x(t) x(t+At)

(a) restoring force (b)extemal fOrcc

Figure 13.6: Linear step in a curve.

Figure 13.7 is the diagram of an elastoplastic material, OAB, and an extrapolation


of the linear stiffness, OD.

Fk

Fn D

) D'

I
I
A
Fy B
A

E c x
0

Figure 13.7: Alternative linear stiffness for an elastoplastic material.

Are~s OABC and ODE represent, respectively, the energy absorbed by the elasto-
plastic and the elastic system. Equating these two areas gives

(13.9)

The conclusion is that it is possible to determine a maximum displacement Xd for a


particular ductility ratio, and up to that limit the analysis can be performed for an
equivalent linear system.
This linear simplification should be restricted by some considerations regarding
the ductility ratio and the loss of strength of the nonlinear system. The selected value
DYNAMICS IN THE PRACTICE OF S1RUCTURAL DESIGN 149

ofµ, should be established with due consideration for the probable consequences,
in physical damage and life injuries, of a relative large deformation. The straight
line representing the linear stiffness of the structure has to be modified to take into
account some loss of strength; most probably the line should be like OD' inFig 13.7.
For multilevel building structures the linear simplification should be comple-
mented with some restricting considerations about the concentration of local ductile
deformations:
In frame structures local ductile deformation should be induced in the beams
where local plasticity effects can be redistributed, and not in the columns.
- Weak storeys should be avoided since there is a tendency for plastic concentra-
tion of deformation on those storeys. The weak storey is characterized by having
a stiffness lower than that of the upper or lower storeys; nonlinear deformation
and ductility are several times greater for the weak storeys than for the adjacent
storeys.
- concentration of deformations should also be expected in the upper and the
lower storeys of buildings, charncterized by uniformity all along their height.
With the linear approximation procedure, these zones of plastic deformation
concentrations require special attention to detailing.

9 Ductility
Ductility can be defined as the property of the structure to dissipate, through inelas-
tic deformations, most of the energy introduced by the ground motion. The ratio of
ductility should provide for an adequate capacity of the structure to resist the earth-
quake forces and also for it not to reach collapse or even a high level of damage.
Ductility likewise contributes to achieving an economic design, since structural
elements critical for the structure's stability are designed for forces smaller than
those that should be required for a linear elastic response.
In spite of the positive contribution of ductility to the overall antiseisrnic design,
certain structural elements cannot be designed for ductile response, for example, a
stay cable bridge with one single column. Excessive plastic asymmetric deformation
of the column can endanger the entire stability of the structure.
It is of the utmost importance that structural elements have an adequate level of
ductility so that when undergoing yielding the necessary strength for the minimum
required structural performance wi11 sti11 remain.

9.1 Ductility ratios

Ductility is basically connected with the plastic deformation that the structural
material undergoes beyond the yield limit; for example, the ductility ratio is
Xm
µ,= -
Xy

for the elastoplastic material represented in Fig. 13.8.


150 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

Figure 13.8: Simplified force-deformation diagram of an elastoplastic material.

However, regarding a hinge on a structural element the question arises of whether


rotation of the section is possible, and to what extent. Not always will the ductility
µ, of a section and µ,g, the rotation ductility of that section, represent concomitant
values of the same level of ductility. The question also arises as to the uniformity
of distribution of the rotation ductility along the structural element, defined as the
curvature ductility. These three aspects of ductility govern the antiseismic design.
The effect of ductility in shear resistance can be better understood in a hinge
mechanism. A hinge created in a flexural element by a seismic cyclic motion can be
described as a section of material flowing alternately from tension to compression
and vice versa (Fig. 13.9a). The shear resistance mechanism changes (Fig. 13.9b);
accordingly, allowable shear stresses should be reduced for seismic design.

(a)
flextural hinge

(b)

flextural hinge
with shear

Figure 13.9: Plastic hinge in a flexural element.


DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 151

Ductility is a phenomenon arising from structural material characteristics.


It happens in the post-yield region, while engineers' calculations are based on
the yield limit, and not over that limit. The effects of ductility should be studied
through analytical experiments. The findings of these experiments are finally elab-
orated as "do-and-do-not" rules for practical applications.
The amount of ductility chosen for a specific antiseismic structural design should
be linked to the expected performance for a given ground motion. The performance
criterion represents an acceptable level of physical damage to the structure. This
level of damage is in turn connected to life-safety considerations, after the inclusion
of the structure's seismic functional status and feasibility evaluations of economic
cost before and after the earthquake in the decision on the amount of the ductility
required.

10 Pushover analysis
Ordinary structures under seismic loads have mostly a nonlinear inelastic response.
These situations can be handled by dynamic analysis programs, but their implemen-
tation requires mathematical techniques, elaborated mainly for academic
applications.
For regular structures with a more or less uniform symmetric distribution of
masses, simplified procedures are given in the Codes or Technical Guides.
For complex structures sophisticated computer programs can be of use. The main
problem lies in the necessity for certain input simplifications, which may produce
incorrect results. The difficulty in making use of these sophisticated programs is
that it is like working with a black box. To avoid errors, computer calculations
should be augmented with preliminary approximate calculations. The latter will
enable the engineer to make a critical interpretation of the results.
A relatively simple procedure for buildings of many storeys, which has gained
ground in recent years, is the pushover static analysis, which can be considered as
an upgrading of the lateral base shear analysis. It allows the study of nonlinear,
post-yielding response of the structure, by control of the displacement of the top of
the building.
The pushover procedure is conceptually simple. The structure is analyzed in the
fundamental mode and the lateral inertial seismic forces are vertically distributed
accordingly. These forces are increased monotonically and controlled versus the
displacement of the top of the building. Basically the procedure is a displacement-
based approach. The building is assumed to reach an inelastic deformation.
Figure 13 .10 represents the process schematically. Figure 13. !0a and b are easily
elaborated with any dynamic computer program for plastic deformation. The values
ofrelative amplitudes of the fundamental mode shape q>(¢1, ¢2, ... , <f>n) are directly
provided by the computer. The vertical distribution of forces follows the vertical
distribution of masses.
Figure 13. !0b represents the displacement of the top of the building while the
horizontal forces are incremented at a constant pace.
152 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

-+-4'- x !<>il v
(a) (b)
L\e,r
Ap
--...(> m,,._1 ' I'

11
-<>
.....
.....
Vy r;
pushover curve Xu
-<> µu= Xy
-e> m2
..,.. nl1 x
Xy x.,

Sa
(c) (d)

Sa ~-~
1'~ 2n( VIX
_Fi ) '

(e) Sa (f) V-Sa·M

-1-
-~
V:v r

sa
I

Sd Sd
Sd Xy Sd

Sa fsismic
(g) (h)

~
sa µ,
. ---·-µn

Sd x
sct Xy (µ,-xy) Xu

(i) t cot exnectcf1 nan1a9e


cate!!ones U1>\U.u structural eiem'~ struc e1en1
• dimrnuc
2' 0%
20%
40%

colb1"se
" 100%

Figure 13.10: Successive steps in a basic static pushover analysis.


DYNAMICS IN THE PRACTICE OF S1RUCTURAL DESIGN 15 3

In Fig. 13.lOc the building is replaced by an SDOF system with generalized


mass, displacement, and shear force:
n
M=L_m;</J;; i 1 =xt/r; v=v/r
i=l

where
r = L:7~1 m;</J;
L7~1 m;(</>;)2
is a participation factor.
Figure 13. lOd is the chosen acceleration spectrum. It can be taken from a diagram
artificially drawn for the design purpose. The spectral acceleration Sa is a function
of the period of the SDOF system.
Figure 13. l Oe is a correlation curve that provides the expected displacement it
of the top of the building for the earthquake acceleration.
Figure 13.lOf is the capacity diagram of the structure. Actually similar to that of
Fig. 13. IOb on a different scale.
Figure 13. l Og is the Sa/Sd diagram for different ductility ratios µ.. It provides a
measure of the inelastic deformation the structure will undergo to resist the earth-
quake event.
Figure 13.lOh it is the elastoplastic capacity diagram of the structure.µ. should
be compared here with the ultimate ductility capacity up to collapse of the structure.
The static pushover analysis is a "lateral displacement/base shear force" design
procedure. It has two advantages over the base shear lateral force procedure: first, it
is somehow related to the expected deformation of the structure; second, it provides
a tool to measure approximately the plastic deformation that the structure will
undergo when subjected to the strong earthquake motion for which it is designed.
Because of the latter, the pushover analysis could be named a "seismic structure
performance" procedure for analysis.
This performace is measured by several criteria, among them danger to life safety
and exit way obstructions, level of damage to structural elements as well as to non-
structural building components, the extent to which mechanical/electrical/plumbing/
utility services will be impaired.
Figure 13.lOd, e, g and i, performance tables, are to be found in the Codes and
appropriate Design Guidelines.

10.1 Limitations of the pushover procedure

Two main imperfections of the static pushover analysis should be mentioned:


1. Even though the procedure is a nonlinear static analysis that provides an insight
to the performance in the post-yield range, it is actually based on a linear system
for which the acceleration spectrum is elaborated.
2. The procedure is suitable for a building whose seismic response is mainly in
the fundamental mode.
154 DYNAMICS IN TIIE PRACTICE OF STRUCTURAL DESIGN

Introduction of the effect of higher modes in the pushover analysis can help to
overcome this imperfection. Some techniques for multimode pushover analysis are
under development at present. The problem is not only to find the right way to do
it, but also not to transform into a cumbersome approach the existing, relatively
simple, pushover technique for practical use.

11 Soil-foundation interaction

Foundations of ordinary buildings in static design situations will usually be designed


for an allowable bearing stress for spread foundations or friction stress between the
piles and the ground. This will be complemented with control of settlement in
compressible soils.
A seismic foundation design is a more complex issue. However, for ordinary
buildings, again, it is most probable that a seismic design will emerge with an
increase of the vertical load, some lift forces, and a horizontal load transmitted to
the soil by friction in shallow foundations or by lateral bearing pressure between
the piles and the surrounding ground. This simplistic design approach does not
provide an acceptable answer to the potential damage that an earthquake can cause
on a structure, besides those resulting from the lateral induced loads. Moreover, it
is not enough to understand why similar structures in the same region are damaged
differently.
Foundation design should ultimately ensure overall stability-not just allowable
stresses and adequate settlements levels.
A construction site should be examined for potential liquefaction, landslide and
slope instability or the possibility of fault motion, and more.

11.1 Geological soil conditions

When designing in a seismic zone it is worth the engineer's while to spend some time
becoming acquainted with the particular geological conditions that can affect the
soil-foundation interaction, particularly the site response. For the same underlying
earthquake waves arriving from the hypocenter, differences in geology and soil will
produce different ground motions under foundations in two adjacent sites.
The soil below the structural foundations should itself be considered a structure
with dynamic parameters: shear modulus, mass density, damping, and fundamental
natural period. This soil structure is subjected to the earthquake motion, and its
response is in turn transmitted through the foundations into the building structure.
Geological factors may affect the soil response and cause degradation and reduction
of the initial strength of the soil. Some of these, like liquefaction or underwater
discharge, can have catastrophic consequences.
Some of the soil conditions that merit attention are:
I. Superposition of horizontal homogeneous layers extending horizontally, long
enough, and to a depth greater than the dimensions of the structure may be
considered good enough for a seismic foundation.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 15 5

2. Different types of soil under a building's set of foundations can impair its safety
if they move away from each other.
3. The depth of a soil deposit affects its frequency; the natural period increases
with increasing thickness of the deposit.
4. Sloping bedding layers can be problematic if they can move differently.
5. A soil deposit on a sloping rock can have unacceptable earthquake displacements.
6. The water content of soils can exert an effect:
a saturated sand layer can lose its shear strength under earthquake motion;
a layer of loose sand beneath a reliable layer of foundation soil can undergo
liquefaction under the combined effect of water and ground motion; the
result can be a catastrophic settlement;
an earthquake can cause underground erosion or local streams.

11.2 Dynamic particularities of soil

11.2.1 Liquefaction
This is a phenomenon that can happen in a layer of cohesionless saturated soil,
subjected to the shaking motion of an earthquake. If the layer is located at a depth
of a few meters, this effect can cause ejections of sand blows and water through the
upper layer of soil, originated in the overburden pressure of the upper deposit of
soil. Liquefaction can cause the cohesionless layer to disappear simply by flowing
away, with the consequent collapse of the above deposit or layerof soil (Fig. 13.11 ).

foundation soil layer (b) collapse of upper layer

relative stiff soil


~--~

reduced sandy layer

Figure 13.11: Before (a) and after (b) liquefaction of an inner sandy layer.

In saturated sandy soils, the cyclic earthquake motion will cause a progressive
reduction to zero of the shear strength. If the layer is relatively compacted and the
earthquake not too strong, the saturated sand may not reach the liquefied state, or
it will take many ground shocks, or liquefaction will occur only on an earthquake
of long duration.
The danger ofliquefaction can be reduced by increasing the layer density through
compaction or appropriate drainage.

11.2.2 Landslide
A soil layer beneath a building may well move aside in a strong earthquake, if the
soil lies on a sloping bedrock; this motion is known as landslide. The consequences
can be horizontal motion, titling, and even overturning of the building.
156 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

11.2.3 Excessive settlement of soil foundation


Soil may become vertically deformed during an earthquake, especially with dry,
loose sands. The settlement increases inversely to compaction of soil and directly
with increasing severity of ground motion. Earthquake vibration produces com-
paction of soil and consequently cracks on the building (Fig. 13.12).

Figure 13.12: Irregular foundation settlement of a building during an earthquake.

11.3 Dynamic properties of soil

The mass of soil beneath a structure extends actually for long distances in any
direction and for profound depths; sometimes it is homogeneous, sometimes it
changes abruptly at short distances. The bedrock may be at a depth of 1000 m or
may rise above the building site. This variety affects the response under earthquake
motions, but for design considerations some dynamic properties of the given mass
of soil can be established from a combination of field and laboratory tests:
- mass density,
- shear modulus,
- damping,
- natural frequency.
Together with these properties a study is made of the type of the foundation soil,
including size and shape of the soil particles, and physical and chemical properties;
the soil is classified as cohesionless or cohesive.
Cohesionless soils, sand, and gravel resist shear forces by internal friction and
interlocking of particles, while cohesive clays do so by cohesion and friction
together. This difference is related to the further response to ground motion since
the water in the inter particle voids behaves differently. The outflow of water in
clay pores lasts long, while the opposite is the case in sandy cohesionless soils.
The field tests for dynamic studies include the Standard Penetration Test (SPT)
to establish the liquefaction potential of sandy soil; shear wave velocity, to estimate
the soil shear modulus G; vibration test, for the fundamental period of soil; and
water table level. These field tests are complemented by laboratory tests, which
include particle size composition for potential soil liquefaction, the cyclic tri-axial
test for an estimate of the damping ratio, and weight measurements for mass density.

11.4 Soil-structure response analysis

Earthquake loads are induced in the structure through the ground motion, as this
occurs at the surface of the soil medium on which it is founded.
DYNAMICS IN TIIB PRACTICE OF STRUCTURAL DESIGN 157

A (structure)

soil medium

bed rock

Figure 13.13: Ground soil under a structure; vertical section.

Figure 13.13 shows a structure built on the surface of a soil medium which in
tum covers the bedrock beneath. Many questions concern the effect of the soil both
in the bedrock ground motion and in the structure.
Mathematical approaches to learning the response of a structure under a seismic
event are usually based on two assumptions: the foundation of the structure moves
rigidly, and the effects of the ground motion are one way, namely from the soil to the
structure. In fact, both assumptions can be challenged by reality, since foundations
are mostly separate units and the vibration of the building can affect the ground
motion in the surroundings of the structure, depending on the ratio of masses and
relative stiffness.
The soil medium has two properties which strongly differentiate it from
structures:
the soil structure is definitely nonlinear and
the boundaries are not clearly defined.
Besides, mass density, shear modulus, or shear strength and damping can be
established by local soil investigations, or at least the initial elastic values of shear
modulus and damping for the mathematical formulation of the problem. The Poisson
ratio is measured; the modulus of elasticity E is estimated from the experimental
stress-strain curve.
The soil-structure issue can be subdivided into three different interaction stages:
bedrock-soil medium,
soil medium-structure, and
a complete bedrock-soil medium-structure interaction.

11.4.1 The free-field motion


If no structure is built on an area, the ground motiOn registered on its sutiace results
from the earthquake vibration waves in the bedrock traveling upward. The vibration
recorded at the sutiace of the soil medium is known as the "free-field motion."
If the soil medium can be recognized as a superposition of horizontal layers
resting on the bedrock and extending horizontally for a long enough distance, an
equation can be written which describes the vertical propagation of the waves
through any of those layers for a shear beam of mass density p, shear modulus G,
and damping coefficient Cg:

(13.10)
15 8 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

This equation is idealized for a one-dimensional model and facilitates analysis


of ground motion waves traveling vertically from the bedrock to the surface, and
vice versa.
This analysis procedure, suitable for a soil with horizontal layers resting on
bedrock, allows, for example, the description of the free-field motion in a building
site with no earthquake records on the basis of existing records for a similar soil
but located far away.

11.4.2 The bedrock-soil medium-structure iuteraction


The complete interaction of a structure in a soil foundation, while the latter lies on
top of the bedrock through which the earthquake motion arrives, can be managed
by the finite element method (Fig. 13.14).
The complete model to analyze is subdivided into finite elements. This sub-
divided model makes for due consideration of nonhorizontal layers, irregular soil
conditions, inside lenses of sand or soft soils, and more. The model can tackle
translations and rotations in any space direction.
The usual equation of equilibrium of dynamic forces can be expanded to encom-
pass the different components:

M · V + C · V + K · V = - (mr1 ·Yr+ Cr· Vr + kr · Vr) (13.11)

where

M =Mm+ M,; C = Cm+ C,; K =Km+ K,

The subscript "m" stands for soil medium, "s" for structure and "r" for bedrock.
The right hand side of eqn. (13.11) represents the earthquake input motion in the
model, arriving through the bedrock.

Figure 13.14: Characteristic finite element representation of a structure-soil situa-


tion, for mathematical study.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 159

If the bedrock is too deep the bottom layer is chosen somewhere in the vertical
direction with a sudden increase in the soil properties. As for the vertical boundaries,
these are usually selected at a distance many times greater than the dimensions of
the structure, so as to fulfill the requirements of nonenergy reflection.
Equation (13.11) may be solved in many ways, for example, with the usual step-
by-step procedure, which is appropriate for the nonlinearity of the soil medium;
another approach is with an equivalent linearization in the frequency domain.

11,4.3 Soil medium-structure interaction


The solution of eqn. (13.11), as already explained for the full model, may become
too expensive and time-consuming. If the records for the surface ground motion can
be considered as already containing the effects of the soil medium, then no need
exists for the bedrock earthquake motion input. In such a case equation (13.11)
can be simplified: the bedrock components disappear and the soil medium input is
transposed from the left-hand side to the right-hand side:

(13.12)

where the subscript "m" now represents the earthquake input motion.
If a further simplification is introduced by considering the structure on a rigid
foundation, and the earthquake input is represented only by the surface acceleration
record, then eqn. (13.12) reduces to the very well known equation of equilibrium
of dynamic forces
m·X+c ·X+k ·X = -m ·Xg
usually used for a thorough study of the effects of an earthquake on any structure.
CHAPTER14

Wind loads

1 Introduction
The effects of winds on structures depend on the characteristics of the flow, changing
continuously with time and location, and on the characteristics of the structures and
their surrounding.
The structural response analysis is not always dictated by wind loads only but
also by aeroelastic peculiarities of the structure.
Slender structures are sensible to the action of wind. They oscillate and this
movement can induce changes in the wind flow nearby, creating an active wind-
structure interaction. This situation may give rise, in critical circumstances, to
a growing deformation in the structure up to its total collapse. In other cases,
the amplitude of structural motions is controlled by damping, but remains still
large enough to exceed any acceptable serviceability limit, based either in human
sensibility criteria or in relation to a specific functional requisite.
Long structures such as bridges may very well be subjected to loads not uniform
all along, and tall structures are subjected to a changing wind profile along its height.
This nonuniformity is caused by wind turbulences created close to the surface of
the earth.
Wind loads on a structure are not always working only in the direction of the
flow; across wind loads may be originated by a special aerodynamic across wind
loads mechanism.
For design purposes wind loads are established from interpretations of the physics
of fluid flows and the effects of a variety of factors representing both the surround-
ings and the geometry and stiffness of the structure. The uncertainty in the effect
of these factors conduces, sometimes, to wind tunnel experiments as a prerequisite
for structural design, especially of major structures.
Structural analysis for wind dynamic situations includes structural dynamics,
aerodynamics, and aeroelasticity, complemented with an analysis of probabilities.
162 DYNAMIC.I) IN THE PRACTICE OF STRUCTURAL DESIGN

Aerodynamics deals with the forces originated by the airflow when disturbed
by the presence of a body in its way. This oncoming wind itself can be affected by
local turbulences in the airflow before arriving to the bluff body in which the study
of those forces is of interest (Fig. 14.1).
Aeroelasticity deals with the forces that originate in the bluff body by the inter-
action between the wind flow and the motion of the body itself. If the structure is
flexible enough, its motion under the effect of the wind affects the characteristics of
the flow and, consequently, modifies the wind forces themselves. If the bluff body
is a relatively stiff stmcture the wind forces are originated by the fluid flow only.
The analysis of probabilities considers the irregularity of the flow. It is based on
statistical extrapolations of wind measurements already collected for a particular
region. The wind time-varying force F(t) is represented by its mean value plus the
fluctuating components

F(t) = Fmcan + F(u, V, W, t) (14.1)

where
Fmean represents the probable value of the force during a certain period of
time, trncan;
F(u, v, w, t) represents the variation of the fluctuating components with time and
space directions; its peak value Fpk and its "spectrum" S(f) usually complement
this description.
By definition
fmelli

f F(u, v, w, t)dt = 0

Two topics are of the engineers main concern for structural design: first, cases
where quasi-static wind loads can be applied for structural calculations; second,
cases where aeroelastic phenomena and self-exited forces can be dominant in the

(a)

(b)

disturbed wind flow

Figure 14.1: Types of wind flow approaching a bluff body.


DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 163

wind design situation; this phenomena, relevant for aerodynamic sensitive struc-
tures, includes vortex shedding, galloping, buffeting, torsion divergence, and flutter.

2 Quasi-static wind loads


An undisturbed motion of wind in the atmosphere is said to fl.ow along streamlines;
the latter are lines of unidirectional force (Fig. 14.2).

I streamlines

) obstacle
wind
direction

IA B

Figure 14.2: Wind flow divergence on an obstacle.

In any given section "Pl' upstream, the streamlines are parallel, moving undis-
turbed at a certain velocity V, arriving at an obstacle "B". Some of those streamlines
are stopped and others deviate. The kinetic energy of the flow

KE= ~p·v
2
(14.2)

is partially transformed into velocity pressure on the obstacle. Those are actually
aerodynamic forces acting on the structure. They are directly dependent on the
velocity of the flow and on p, the air density. p at sea level = 1.225 kg mass/m3 .
If the airflow stops totally on arriving at a structure, such as the obstacle in
Fig. 14.2, the kinetic energy transforms in a static pressure
2
= ~ p [kg· s = ~ p.
2
q _l ] . V 2 [m ] yl [kg] (14.3)
2 m m3 s2 2 m2

For a rigid obstacle, such as a stiff low-rise building, for example, it is usual to
transform wind forces acting on the external surface into static loads. This simplifi-
cation, for design purposes, is not possible in the case of flexible structures such as
high-rise buildings, light cable bridges, high-rise stacks, light membrane structures,
and many more that can reveal some deformations under wind loads.

2.1 Wind flow

2.1.1 Wind velocity


Undisturbed flow can be found at hundreds of meters above the ground level or at
shorter heights above a quite water level, like that of a lake surface.
164 DYNAMICS IN Tiffi PRACTICE OF STRUCTURAL DESIGN

Wind flows close to ground level are characterized by turbulence that originates
in the ground surface irregularities; the velocity of wind is affected by those turbu-
lences and by thermal effects.
In order to establish the maximum wind load for design purposes, a mean value
of wind velocity, Vmean, should be estimated. This is done by collecting records of
wind velocities in open areas. The mean wind value is estimated by averaging the
records collected during a certain period of time, and using statistical techniques.
This mean value is usually calculated for a time of 1 h or 10 min.
The velocity of wind for design purposes, VR, is greater than the mean value

VR = Vmean + D. V(u, v, w, t) (14.4)

where !:J. V is a function of the turbulences in space directions, changing continu-


ously with time. !:J. V represents the incidence of short duration wind gusts in the
total value of the wind velocity that should be considered in wind load calculations.
A gust factor can be defined as follows:

This gust factor increases inversely to gust duration; for that it is important to
choose adequately the gust duration to be taken for design purpose. A gust time
measure of 1 min, for example, may result in an increase of some 20% for the mean
value averaged for 1 h. The adequate selection of gust duration depends on two
considerations:
- the time taken for the aerodynamic forces to build up and the
- duration of gust that is most effective in deflecting the structure.
Impulsive forces lasting for about half the period of the natural oscillation of the
structure are very effective in deflecting it.
Since the flow velocity of the wind selected for design purposes is an issue of
statistics, due consideration should be given to the probability of annual exceedence
of the chosen value; for example, a return period r of 50 years represents a 2 % annual
probability of exceedence, l/r. Larger values selected for the return period will
result in larger values of the velocity pressure for design.

2.1.2 Flow pattern


A wind flow approaching an obstacle such as a stiff building will be partially stopped
and partially diverted so as to flow around the building and over its roof. Figure 14.3
shows some patterns of the airflow around a stiff body.
Sharp comers. such as comer numbers 1 and 3 in Fig. l 4.3a, may cause detach-
ment of the oncoming flow. Turbulences tend to produce reattachment of the flow,
as in Fig. l 4.3b and f, or to move the points of detachment backwards in round
surfaces like in Fig. 14.3d compared with Fig. 14.3c. The detachment effect of a
sharp corner can be neutralized by an inclination of the side surface (Fig. 14.3e),
or a pitch roof, (Fig. 14.3h) if the slope angle" is greater than a certain limit.
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 165

(a) sn1oth flow_........ -... (bl


turbulent flow

(c) (d)
smooth flow turbulent flow

(e)
:moth flown: ~a

(g)

Figure 14.3: Airflow pattern around bluff bodies.

Surface of contact with the wind flow are surfaces of wind pressure, like surface
1~3, in Fig. 14.3a; surface of detachment are surfaces under wind suction, like
surfaces 1-2, 3-4 or 2-4 in Fig. l 4.3a.
Wind forces are always perpendicular to the body surface, whatever the wind
flow direction.
In zones of contact the airflow adheres to the surface like 3-4 in Fig. l 4.3e and the
wind motion stops; immediately away from the sutface the wind velocity increases
gradually from zero to the wind flow velocity; this situation creates a very thin
layer of along shear forces between the surface and the flow; this is called a skin
friction drag.
The flow that follows the detachment points are characterized by the local rota-
tional motions of the air particles that originate high local suction forces; they are
created by shear forces in the flow (Fig. 14.4).

region of
-~rt-? shear forces

buildling corner

Figure 14.4: Localized rotational motion.


166 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

2.2 Static wind loads

The wind flow is dynamic by nature, but rigid structures are not sensitive to wind
and so a simplification can be done for the purpose of structural analysis and
calculations, in such a manner as to consider wind effects through their equivalent
static forces. The energy of the wind fluctuations, as represented in the power
spectrum of the longitudinal component u of the turbulences , is quite small for the
range of frequencies typical of rigid structures with natural periods T < 2 s.
The wind velocity pressure should be taken directly from gusts of a few seconds
only, with no relation to any !mean average wind velocity, since this pressure builds
up immediately for this type of building. This fact has been confirmed through
measurements performed in existing buildings. A gradual transition from gusts to
10 min mean average of wind velocity could be done by considering the spatial
correlation of turbulence effects in large structures with greater natural periods T,
by introducing the turbulence intensity; it is given by:
fu(z) = O"u(z)/V(z)
whereuu(Z) is the turbulence standard deviation and V(z) is themean wind velocity,
both as a function of height above ground level.
Wind loads for rigid structures are given in the National Codes as a function of
the basic velocity pressure q, multiplied by a number of coefficients each repre-
senting the influence of a particular aerodynamic characteristic, such as: the terrain
roughness, height above the ground level, site topographic conditions, geometry,
and more; they are clearly presented in the Codes.

3 The dynamic response

The simple approach of static loads, elaborated for rigid structures to be applied
on relative low-rise buildings, cannot be extended for structures sensible to wind
flow.
Sensible structures vibrate under the fluctuating load generated by wind turbu-
lences. In such cases detailed analysis and investigations should be performed.
These structures are usually grouped in families of similar aerodynamic character-
istics such as: modern cable bridges, stacks, high-rise buildings, light membranes,
cables and thin structural components, and more.
The motion of the structure does not reflect directly the pattern of the aerodynamic
load; the latter includes a multiplicity of frequencies (Fig. 14.Sa), while structures
motion is somehow harmonic, see the response of a flexible building in Fig. 14.Sb.
The motion of a structure interacting with wind is very complex. Several aero-
dynamic phenomena can be detected in this motion. Besides the characteristics of
motion and induced forces, attention should be focused on the aeroelastic instability
that originates under a wind flowing at velocities higher than a certain limit known
as the "divergence speed" or the "critical speed".
The wind contributes an additional damping effect to the structures' mechan-
ical damping. Even though there is a problem with this air damping coefficient,
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 167

(a) F (t)

(b) x (t)

Figure 14.5: Response of a flexible building. (a) Aerodynamic load; (b) motion of
the structure.

it changes its sign at high wind velocities, and the amplitude of the motion increases
progressively. At velocities higher than the above-mentioned critical limits insta-
bility or collapse of the structure happens. These limits can be approximately cal-
culated with adequate mathematical formulas. Design considerations of mass and
geometric proportions allow the engineer to raise their values away from the wind
regime levels at the building site.
The actual tendency is to establish a growing number of rules, mathematical
ready formulas, suggested geometric proportions, recommended structural fre-
quency ranges, drift and acceleration limits for families of structures. All those
oriented to reduce expensive time-consuming experiments, and to facilitate struc-
tural design.
The wind forces and their effect on a structure _cannot be formulated on definite
mathematical functions of time. Only statistical techniques can be implemented.
Calculations are usually performed by specialists, supported at times by local mea-
surements and, when necessary, with wind tunnel experiments.
The design of a particular structure includes the superposition of three different
considerations:
- static wind loads;
- an aeroelastic investigation of the expected motion and its effects;
- dynamic loads created in the air-structure interaction; these should be super-
posed to the static loads.
168 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

3.1 Winds on flexible structnres

The mathematical treatment of the effects of wind on flexible structures is a highly


complex and still evolving issue. In the following, just by way of example, a scheme
of this treatment for a simplified SDOF system.
The wind force on a concentrated mass can be expressed as

F(t) = ~p · V(t) 2 ·Co ·A (14.5)

where
V(t) = Vmcan + u(t)
u(t) is the along wind turbulence (V, w, are neglected); Co is the along wind aero-
dynamic force coefficient; A is the facade of the concentrated mass to the wind flow.
The wind load is considered as the summation of a constant plus a fluctuating
load,
F(t) = F mean + p(t) (14.6)

The equation of equilibrium of forces for the turbulent wind load is:

(14.7)

for wind sensitive structures eqn. (14.7) should include two particular additions to
the ordinary equation of dynamic forces: first, the air viscosity damping coefficient
CAn; second, the self-excited force resulting from the air-structure interaction.
Equation (14.7) becomes

M ·x+c·x+k·x= ~p·Co ·A(Vmean +u-x)2 (14.8)

The known equation of equilibrium of forces for an SDOF system subjected to


a constant external load F mean is now complemented with the fluctuating force, Fu
and the air damping force, FAD:

Fu= p· Cn ·A· Vmean ·U

FAn=p·Cn·A·Vmean·X

Neglecting small values (u2 , ±2 , ux),

where FAD is the air viscosity damping force; CAD is the air damping coefficient.
The total damping coefficient for wind sensitive structure design situations becomes

Since wind loads vary randomly with time they can be represented by a superpo-
sition of harmonic functions with different frequencies. The spectrum of the force
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 169

so represented provides information on the frequency content and, consequently,


the margins of resonance.
The analysis of structures in this situation is usually performed in the frequency
domain instead of the time domain, which means working with the frequency
components instead of the wind flow fluctuations with time; S(n) the power spectral
density function, is related to p(t) through the Fourier Pair Transform.
Introducing
roo ]'12
p(t) = [lo Sp(n) · dn
the response is given also as a spectral function

roo ]1/2
x(t) = [lo Sx(n) · dn
where
Sx(n) =f (:~; 2 2
IH(n)l ; IX(n)l ; Su(n))

- H(n) is the structural admittance; a transfer function for p(t) and x(t) of the
structure into a power spectral density function;
X(n) is the aerodynamic admittance; it transforms the properties of the airflow
into wind loads;
- Su(n) is a turbulence spectrum.

Since mathematical approaches to analyze air-structure interaction involve


several assumptions, they are complemented with wind tunnel tests. The reliability
of their results depends on an adequate structural modeling, a correct introduc-
tion of the wind shear profile and a good simulation of the surrounding turbulence
conditions.

4 Aeroelastic phenomena
It is the result of an interaction between aerodynamic forces and structural motions.
The phenomena is characterized by its direct relation to wind flow velocity and,
in particular, to a critical value which can be formulated for design calculations, as
seen in the following section.

5 Vortex
Wind flowing around a bluff body sheds vortices from both sides of the disturbing
body, into the wake created behind it (Fig. 14.6).
These swirling eddies result from shear forces that originate owing to the internal
friction between the flow lines close to the bluff body. It increases continuously
towards the body's surface, as much as wind speed decreases from streamline
velocity to zero.
J 70 DYNAMICS IN THE PRACTICE OF STRUC11JRAL DESIGN

;;;:> ;;;:>
(a)
'.::>wake
~
,r a A'

(b)
;;;:>
wake

Figure 14.6: Vortices behind a bluff body.

Vortices are shed into the wake alternately from one side against the other. They
are created together with changes in wind velocity. These changes produce alternate
side loads on the body harmonically.
For rigid bluff bodies the distance a between vortices, and hence the time interval
between their shedding, is constant for a constant wind flow.
The frequency of vortex shedding, and hence the frequency of the across wind
alternate forces, is a function of the wind characteristics, roughness of the surface,
and the shape of the cross section
v
f =St· -
D
where St is the Strouhal number, a coefficient which for a round section is also a
function of the Reynolds number, R.
Both Rand St are aerodynamic coefficients tabulated as function of the cross sec-
tion of the body. Sensible vibrations of flexible structures introduce some changes
in the values of St, but for small initial motions the basic concept remains the same.
Subjected to the lateral harmonic forces, the structure initiates a transversal
motion, relatively small for frequencies f which are far from one of the natural
frequencies of the structure, n, especially of the fundamental mode. With progres-
sive increasing or decreasing velocity of the wind flow, the frequency f approaches
n and a ~tate of resonance is created. This is the vortex critical wind velocity:
n·D
Ver= - -
St
when f = n. An interesting experimental observation to remark is that at this
state of resonance an interchange of effects takes place: the motion of the struc-
ture dominates the wind flow and for small changes of wind velocity the effect
of resonance continues. This aeroelastic phenomenon is called" lock-in". Turbu-
lences in the wind and structural vibrations affect the lock-in range. Wind velocities
DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN ] 71

lock-in
region
n
,,__ _ __,.lock-in range
v

Figure 14.7: The lock-in phenomenon.

that can induce vortex vibrations are, for certain structures, relatively low, in the
expected range of normal wind flows; somewhere around the 40 km/h range; vibra-
tions amplitude are relatively small, usually much less than half the across wind
dimension of the structure.
Figure 14.7 shows a graphical representation of the lock-in phenomenon.
Since moderate wind flows causes the "lock-in" effect in structures, the latter are
frequently subjected to stress vibrations. Hence, due consideration should be given,
in the design stage, to effects of fatigue. The effect of vortex-induced vibrations in
people, such as those that can appear in light pedestrian bridges, are also of short
duration and they will produce, mostly, a light transient annoyance.

6 Buffeting

A structure A situated at a certain distance I in the street of vortices created in


the lee of an upstream bluff body B, as shown in Fig. 14.8, may be subjected to
harmonic fluctuating wind loads arriving in the steady stream.

r:·

B A

Figure 14.8: Street of vortices between two bluff bodies.

Oscillations of large amplitude can be induced in the structure, especially if the


frequency f of the vortices loads is close to the natural frequency of the structure.
This buffeting effect on the structure does not depend on the aerodynamic
loads in the bluff body. The distance l and frequency f are among the factors
of more importance. Oscillations of structure A can be in the across wind or in the
longitudinal wind direction.
172 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

7 Galloping

This aerodynamic effect is characterized by oscillations of large amplitude in the


across wind direction. The amplitude of these oscillations can reach values much
larger than the dimension of the structures' section, measured normal to the wind
flow direction.
The description of the phenomenon may start with the angle of incidence a(t) of
the wind flow on the structures' "front facade" (Fig. 14.9); it usually changes with
time
>
a(t) <0
V, in Fig. 14.9, indicates the direction of the average wind.
If the structure initiates an across wind motion, its velocity can be represented
by y; VR in Fig. 14.9b is the vector resultant of V. This wind flow generates two
wind forces

Fo = ~P · v~ · D ·Co
h = ~p · v~ · D · CL
where
- Fo is the along wind drag force;
~ FL is the across wind force, named the lift force;
- Co and CL are the drag and lift aerodynamic coefficients, which are the func-
tions of the geometry of the structure and the wind flow characteristics and are
determined, mostly, in wind tunnel experiments.
The effect of the lift force is to increase the initial across wind motion and it can
conduce to galloping instability if the following mathematical condition is fulfilled

- d· I
dCr + Cola~o < 0
a a=O

This is known as the Den Hartog criterion which establishes that the incipient
galloping phenomenon is the result of a negative aerodynamic damping.

y
·tI
v . ID;(
- · . . . .. v

(a)
t
a__. ---------~---~·-
-X
!D
-------<
(b)
t
y

Figure 14.9: Incidence of wind force on a bluff body.


DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN 173

Circular cylindrical structures cannot evidence galloping effects since, because


of symmetry,
dCL
-=0
da
and Co(O) > O; but if a cylindrical chimney is located in the wake of another
structure, the galloping effect can occur because of the vortex-generated transversal
forces. This effect is known as Wake Galloping.
CL is an aerodynamic coefficient affected by wind velocity, and hence a minimum
value of a critical velocity, Vee. for galloping instability can be calculated.

8 Flutter instabilities

Flutter for structural design, such as classical flutter or torsion instability are aero-
elastic phenomena that can have catastrophic consequences for structures subjected
to wind velocities larger than certain limits known as flutter Ver· Because of this
engineers should assure that the critical velocity for the particular structure under
design will be much higher than the maximum expected for the region of the
building site. This aeroelastic instability is typical of structures with an airfoil
section such as cable bridge slabs.
Two modes of motion characterize those structures: bending and torsion move-
ments. A combined motion of these two modes is called classical flutter, while only
torsion motion characterizes an SDOF flutter; translational motion alone is usually
not considered a dangerous phenomenon.
The flutter motion extracts energy from the wind flow. If the damping mech-
anism dissipates this energy the motion will decay. If not, the flutter motion will
increase continuously up to final collapse or to a stage where the structure undergoes
a substantial change in its conceptual design. This situation is known as the critical
flutter motion.

8.1 Classical flutter

It is a motion that includes rotation and translation simultaneously(Fig.14.IO), with


a certain phase angle between these two displacements. The phase angle creates a
certain balance in the process of extracting energy from the wind flow to feed the
two degrees of freedom motion.
The equations of equilibrium of the dynamic forces can be written as

M ·y+2Mw, ·~, ·y+Mw, ·y = h (14.9)

Im . a + 2Im . Wa . ~a . a + w~ . Im . Ci = Ma (14. IO)

where
- FL and M, are disturbing winds effects;
- w 8 and Wa are natural circular frequencies of the two degrees of motion;
174 DYNAMICS IN THE PRACTICE OF STRUCTURAL DESIGN

·······i-
-~;;::L......

Figure 14.10: Airfoil section on a wind flow.

~' and ~a are structure and air damping factors;


M and Im are mass and mass moment of inertia;
- CL and CM are aerodynamic lift and moment coefficients.
Equations (14.9) and (14.10) represent the equilibrium of dynamic forces on an
uncoupled motion, while for a two degrees coupled motion equivalent mass and
mass moment of inertia Me and fem should be introduced. For the two degrees of
freedom, translation y and rotation a, the natural circular frequencies are defined
as follows:

Damping effects in the motion are the result of a combination of structures


damping and an aerodynamic damping supplied by air. If both damping coefficients
are positive, flutter oscillation will be mitigated. But~" diminishes with increasing
air velocity, up to an instant where it becomes negative and with an increasing
absolute value. A situation will arrive where the overall damping is zero. If airflow
velocity increases over this limit, known as the flutter critical velocity, Yer. the
amplitudes of motion will continue to grow, and the unsustainable critical flutter
motion wiH be created.
The.critical wind speed limit, Vero can be raised high above the expected wind
velocities by increasing the ratio wa/Ws, and increasing the mass and the mass
moment of inertia of the structure.

8.2 Torsional flutter

For bridges where the two degrees of motion are uncoupled, an SDOF motion
instability can be created, known as torsion flutter. It may happen in structures
undergoing separated wind flows (Fig. 14.11).
DYNAMICS JN THE PRACTICE OF STRUCTURAL DESIGN 175

centre of torsion
elasticity

v
-------!>


]3

Figure 14.11: Transversal section on a wind flow.

The structure is subjected to a strong aerodynamic rotational moment under


an increasing value of a. For low velocities this situation is stable, but for wind
velocities exceeding a certain critical value known as the divergence velocity,
a flutter torsion instability is created,

2 ]1/2
2wa ·fem
Vmv = [p·B2~
where Lem is the equivalent mass moment of inertia. This flutter instability may be
compared with the static phenomenon of buckling in that it can happen even for
very small angles of incidence.
Selected bibliography

Bachmann, H. & Ammann, W., Vibrations in Structures, IABSE-AIPC-IBVH,


1987.
Bachmann, H. et al., Vibration Problems in Structures, Birkhauser Verlag: Basel,
1995.
Beards, C.F., Structural Vibration, Analysis and Damping, Arnold: London, 1996.
Biggs, J.M., Introduction to Structural Dynamics, McGraw Hill, Inc.: New York,
1964.
Chopra, A.K., Dynamics of Structures, Prentice Hall, 2001.
Clough, R.W. & Penzien, J., Dynamics ofStructures, McGraw Hill, Inc.: New York,
1975.
Dyrbye, C. & Hansen, S.O., Wind Loads on Structures, John Wiley & Sons:
Chichester, 1996.
Goschy, B., Design of Buildings to Withstand Abnormal Loading, Butterworths,
1990.
Jeary, A., Designer's Guide to the Dynamic Response of Structures, E & FN Spon:
London, 1997.
Komodromos, P., Seismic Isolation for Earthquake Resistant Structures,
WIT Press: Southampton, 2000.
Lawson, T., Building Aerodynamics, Imperial College Press, 2001.
Naeim, F., The Seismic Design Handbook, Kluwer Academic Publishers: New York,
2002.
Rosenblueth, E., Design of Earthquake Resistant Structures, Pentech Press:
London, 1980.
Scanlan, S., Wind Effects on Structures, John Wiley & Sons, Inc.: New York, 1996.
Smith, J.W., Vibration of Structures, Chapman & Hall, 1988.
Wakabayashi, M., Design of Earthquake Resistant Buildings, McGraw Hill, Inc.:
New York, 1986.
Yu-Xian, H., Liu, S.-C. & Dong, W., Earthquake Engineering, E & FN Spon, 1996.
Index

active structural control 28 elastoplastic


aerodynamics 162 design 26
admittance 169 material 24
forces 163 systems 106
lift coefficient 174 elongation 60
aeroelasticity 162 cyclic 60, 61
instability 166 endurance limit 68, 69
phenomena 169 energy 20
airfoil section (wind) 173 damping 21
dissipation 22
brittle material 60 kinetic 20
buckling effect 31 strain 20
buffeting (wind) 171 work 21
epicenter 136
collision 44 equivalent mass 119, 120
cycle of motion 23 equivalent stiffness 120

D' Alembert 15 fatigue 36, 67


damping 9, 17, 18, 29, 37, 46, 77, 80, 124 concrete 70
critical 13, 91 failure 67
energy dissipation 77 lifetime 67
force 14 pretensioning steel 71
hysteretic 78 reinforcing bars 71
mechanism I0, 11 structural elements 72
sources 10 prestressed beams 75
types 10, 17 reinforced concrete 74
viscous 17 steel structures 73
wind 166, 168 structural steel 72
Den Hartog 172 variable stresses 70
OMS 85 with different load patterns 68
Duhamel 101, 142, 146 fault.(earth) 136
Dynamic Magnification Factor (DMF) 6, 12, flexibility 14
18, 22, 96, 97 flutter 173
dynamic loads 99 classical 173
dynamic problems 37 critical wind speed 174
instability 173-175
earthquake torsional 173, 174
artificial records 143 torsion instability 173
intensity 137 divergence 163
load 138 force 7, 83
magnitude 138 damping 14, 17, 87
records 143 dynamic 7, 16, 17
eigenvalue 91, 115 inertial 14, 17, 83
]80 INDEX

force (contd.) MDOF 84, lll, 113, 144


restoring 14 earthquake analysis 144
static 7 Merca!li 137
time-varying 35, 162 modal response summation 123, 145
Fourier Pair Transform 106, 169 AVS 145
Fourier series I02-104 SRSS 145
free-field 157, 158 GMC 145
free motion 12, 14 CRC 145
free vibration (motion) 8, 17, 18, 88, 92, 93 shapes 8, 115, 128
frequency 8, 37, 90 modal analysis 123
circular 8, 18, 90, 91, 115, 131 modes of vibration
frequency domain l 0 I fundamental 8
momentum 27
galloping 172
instability 172
wake 173 nonlinear systems 109
ground acceleration
displacement 137, 142 passive structural control 28
spectra 144 performance (earthquakes) 153
velocity 137 period 7, 90, 118
plastic hinge 25, 150
human loads 46 plasticity 25, 60
construction 49, 50 pseudo-velocity (earthquakes) 143
crowded areas 48 pushover 139, 151
gymnastic 48
jumping 46 rate of loading 60
running 46 rate of strain 59, 60
swaying 46 in bending 64
walking 46 in concrete 62
human tolerance 54
in reinforced concrete 64
hypocenter 136
in steel 63
hysteresis 9
resonance 13, 29, 37, 98
inelastic 10
response of structures 24
loop 9, 78, 79
linear 24
loss 9
nonlinear 24
impulse 27 Reynolds number 170
inertial force 14 rigidity 14

kinetic energy (wind) 163 seismic acceleration 141


ground motion 142
landslide 155 SDOF 16, 84, 87
liquefaction 155 soil foundation settlement 156
load 46 dynamic properties 156
arbitrary dynamic 130
soil-structure response 156
dynamic 5, 39
spectral
frequency 37 acceleration 143
harmonic 39, 41
displacement 143
impulsive 39, 43, 44
response (SDOF) 143, 144
nonperiodic 39
velocity I43
periodfo 39, 42
spring constant 24
short pulse 40
steady state 12
static 35
stiffness 25
translational 45
strain fluctuation 61
vehicle 44
Strouhal number 170
working machines 41
study models 30
logarithmic decrement 92
lumped masses 124
time domain 100
mass 16 torsion dynamic forces I 10
concentrated 16 transient motion 12, 18
INDEX 181

vibrations 51 force, drag 172


acceptance levels 55 lift 172
in design 53 spectrum 168
normative values 54 gust factor 164
serviceability limit 51 loads 161
source 51, 52 dynamic 173
tolerable values 54 self-excited 168
vibration free modes 8 static 166
vmiex 169 power spectral density function 169
critical velocity 170 pressure 165
lock-in 171 profile 161
shed 169, 170 streamlines 163
street 171 suction 165
tunnel 172
wind 161 turbulence 161, 168
critical speed 166 spectrum 169
flow 163 velocity, mean value 164
pattern 164 design value 164

Vous aimerez peut-être aussi