You are on page 1of 188

Structural Engineering Documents

Hugo BACHMANN
Walter AMMANN

Induced by Man
and Machines

International Association for Bridge and Structural Engineering IABSE


Association Internationale des Ponts et Charpentes AIPC
Internationale Vereinigung fur Bruckenbau und Hochbau IVBH
About the Authors:

Hugo BACHMANN

Born 1935 in Olten, Swit-


zerland. Diploma in Civil Engi-
neering at the Swiss Federal
Institute of Technology in
Zurich (ETH). Professional ex-
perience in consulting engine-
ering offices and construction
industry in the domain of steel
structures, reinforced and pre-
stressed concrete structures.
Degree of Dr. sc. techno from
ETH. Ordinary professor for
Structural Engineering at this
school. Research work in con-
crete plasticity, shear capacity
design, lightweight concrete
structures, partial prestres-
sing, structural dynamics and
earthquake engineering.
Active as a consultant and
expert. Member of various
national and international
committees and associations.

Walter AMMANN

Born 1949 in Ennetblihl, Swit-


zerland. Diploma in Civil
Engineering at the Swiss Fed-
eral Institute of Technology in
Zurich (ETH). Professional ex-
perience in a geotechnical en-
gineering office. Degree of Dr.
sc. techno from ETH. Research
work in impact loading on
reinforced and prestressed
concrete structures, material
behaviour at elevated strain
rates, structural dynamics and
earthquake engineering.
Member of national and inter-
national committees and -
since 1984 - again part-time in
the planning and design prac-
tice.
Vibrations in Structures
Induced by Man and Machines

Vibrations in Structures)) concentrates on vibrations in structures as excited by human


motion or machine operation. Man-induced vibrations may arise from walking, running,
skipping, dancing, etc. They occur mostly in pedestrian structures, office buildings, gym-
nasia and sports halls, dancing and concert halls, stadia, etc. Existing publications treat by
and large some isolated aspects of the problem; the present one attempts, for the first time,
a systematic survey of man-induced vibrations. Machine-induced vibrations occur during
the operation of all sorts of machinery and tools with rotating, oscillating orthrusting parts.
The study concentrates rather on small and medium size machinery placed on floors of
industrial buildings and creating a potential source of undesirable vibrations. The associ-
ated questions have rarely been tackled to date; they entail problems similar to those of
man-induced vibrations.

The book is consciously intended to serve the practising structural engineer and not primar-
ily the dynamic specialist. It should be noted that its aim is not to provide directions on how
to perform comprehensive dynamic computations. Instead, it attempts the following:

1. to show where dynamic problems could occur and where a word of caution is good
advice;

2. to further the understanding of the phenomena encountered as well as of the underlying


principles;

3. to impart the baSic knowledge for assessing the dynamic behaviour of the structures or
structural elements;

4. to describe suitable measures, both preventive to be applied in the design stage and
remedial in the case of rehabilitation.
Structural Engineering Documents
.... .. ,

Sec~ao de Estruturas
Biblioteca

Requisitado em: Devolvido em:

f ..tt!!.I~? .... .1.... .1.....


.....1.....1...... .... .1.... .1.....

.....1.....1..... .... .1.... .1.....

.....1.... .1..... .... .1.... .1..... Hugo BACHMANN


.....1.... .1..... .....1.....1..... Walter AMMANN
.... .1.... .1..... .....1.....1.....

.....1.....1..... .... .1.....1..... VIBRATIONS


.....1.... .1.....

.... .1.... .1.....


.....1.....1.....

.....1.... .1.....
IN STRUCTURES
.....1.....1..... .....1.... .1.....
Induced by Man
.....1.....1..... .... .1.... .1.....
and Machines
.....1.....1..... .....1.... .1.....

.....1.... .1..... .... .1.... .1.....

.....1.....1..... .....1.... .1.....

.... .1.... .1..... .....1.....1.....

.....1.....1..... .... .1.... .1.....

.....1.....1..... .... .1.... .1.....


:opyright 1987 by
Iternational Association for Bridge and Structural Engineering

.II rights reserved. No part ofthis book may be reproduced in any form or by any means, electronic or
lechanical, including photocopying, recording, or by any information storage and retrieval system, without
ermission in writing from the publisher.

>EN 3-85748-052-X
rinted in Switzerland

ublisher:
\BSE-AIPC-IVBH
TH - Honggerberg
H-8093 Zurich, Switzerland

el.: lnt + 411377 26 47


elex: 822186 lABS CH
elegr.: IABSE, CH-8093 Zurich
Increasingly, the practising engineer sees himself confronted with the effects of loads on
a structure which vary with time, that is to say, which are dynamic by nature. The following
are some of the main reasons:
- Structures planned or constructed today are often more prone to vibrations than those of
former years. Because of the use of high-quality material and the exploitation of their
strength, larger and larger spans are chosen while the thicknesses of beams and slabs
have kept decreasing. Hence, of the total load - mostly assumed to be non-dynamic - the
dead or permanent load constitutes a smaller part. Slender structures with a small
stiffness and a low mass are more prone to vibrations than more squat and stiffer
structures with a larger mass. If such a modern-type structure is exposed to dynamic
loads, considerable vibration problems can arise.
- Several dynamic actions have increased in intensity latterly. For example, rationali-
zation has led to the installation of more efficient machinery in manufacturing plants.
Efficiency is enhanced by, for instance, a higher production frequency which may raise
the demands on the structure to sustain higher dynamic loads.
- Also, the demands on a structure for serviceability to the user have been increased. The
environmental stress level being generally higher, people have become more sensitive to
vibrations. Therefore, normative values for admissible vibrations are becoming,stricter,
and vibration amplitudes must be reduced. There is also a higher demand from the point
of the precision of certain manufacturing processes (e. g. milling, weaving) requiring a
lower vibration level of their machines.

Apart from this, new engineering tasks have called for a more exact consideration of
loading types and of the dynamic behaviour of certain structures. This has been the case,
for example, with offshore structures subject to high dynamic wave loads and with nuclear
power plants required to remain safe under exceptional loading such as earthquakes.
These kinds of problems, however, have evolved into comprehensive subjects of their own
and are not treated here.

This book is more or less limited to vibrations of structures as excited by human motion
or machine operation. Man-induced vibrations may arise from walking, running, skipping,
dancing, etc. They occur mostly in pedestrian structures and particular buildings, such as
office buildings, gymnasia and sports halls, dancing and concert halls, spectator galleries,
etc. Existing publications treat by and large some isolated aspects of the problem; the
present one attempts a systematic survey. Machine-induced vibrations occur during the
operation of all sorts of machinery and tools with rotating, oscillating or thrusting parts. A
substantial literature can be found on machine foundations, concerned with the support of
mostly heavy machines on purpose-built substructures. The present treatise concentrates
rather on small and medium size machinery placed on ordinary floors of industrial
buildings and creating a source of undesirable vibrations. The associated questions have
rarely been tackled to date. They entail problems similar to those of man-induced vibra-
tions. Questions beyond this range, such as, for instance, machine-induced ground vibra-
tions interacting with adjacent buildings or affecting other parts of the structure, are
excluded.
The book is consciously intended to serve the practising structural engineer and not
primarily the dynamics specialist. It should be noted that its aim is not to provide directions
on how to perform comprehensive dynamic computations.
lnstead, it attempts the following:
1. To show where dynamic problems could occur and where caution is necessary.
2. To further the understanding of the encountered phenomena as well as of the under-
lying principles.
,. To impart the basic knowledge for assessing the dynamic behaviour of the structures or
structural members.
t To describe suitable measures, either preventative to be applied in the planning stage or
improving inadequate performance.

This limitation often compels simplifications and approximations to be made, which


llways need to be verified in terms of their applicability. More complex situations may well
~equire more detailed investigations with the aid of the references given or the help of a
;pecialist.

fhe book is divided into the following sections:


- Chapter 1 gives a concise survey of the diverse dynamic problems encountered in civil
engineering. They are characterized in broad terms, whereby some themes are men-
tioned which are not covered in this book.
- Chapter 2 is concerned with man-induced vibrations. The dynamic loads resulting from
human motion, their effects and possible countermeasures are treated in depth, discus-
sing vibrations in pedestrian structures, office buildings, gymnasia, dancing and concert
halls, and high-diving platforms in swimming pools.
Chapter 3 deals with machine-induced vibrations of the category defined above. The
dynamic loads, their effects and some countermeasures are treated within these limits.
Chapter 4 derives acceptance criteria for the effects of dynamic loading. Depending on
the circumstances, they could be used as admissible values for an appraisal of a given
vibration problem.
- Appendix A reports on cases where man- or machine-induced vibrations in structures
were a problem. Improvements are discussed for some of them. Practical cases like
these, particularly when damage occurred, can provide valuable insight.
- Appendix B furnishes the fundamentals of structural dynamics. Included are a collection
of formulas to determine the eigenfrequencies of beams and slabs, basics of tuned
vibration absorber design, and comments on the dynamic properties of important
construction materials including the question of damping.

For some of the questions and problems addressed, research results and practical
experience are still rather limited; this is particularly the case for man-induced vibrations.
Therefore, any advice, criticism or additional practical case reports will be very welcome.

The authors wish to express their gratitude to all colleagues who contributed case reports
or supported the work by their open discussion and competent advice. They particularly
acknowledge the substantial contributions of Dr. D. Somaini to several sections. They
further feel indebted to other former and present members of the Institute of Structural
Engineering at the Swiss Federal Institute of Technology (ETH) , who typed and proofread
the manuscript and produced the figures: Ms. S. Burki, Ms. R. Feusi, Mrs. S. Schenkel
(Dipl. lng. ETH) , Ms. H. Ungricht (Dipl. lng. ETH) , Mr. K. Baumann (Dipl. Ing. ETH) ,
Mr. R. Vogt (Dipl. lng. ETH) and Mr. L. Sieger. The translation from German by
Mr. J.-M. Hohberg (MSc.) and Dr. E. G. Prater is also gratefully acknowledged.

Zurich, July 1987 H. Bachmann, W. Ammann


Page

Preface

Table of contents

Z Survey ofproblems in structural dynamics 1

1.1 Characterization of dynamic loads 1


1. 1.1 General aspects 1
1. 1.2 Harmonic loads 3
1. 1.3 Periodic loads 3
1.1.4 Transient loads 4
1.1.5 Impulsive loads 4
1.1.6 Additional loading criteria 5
1.2 Source-dependent effects of dynamic loads and countermeasures 5
1. 2.1 Classification of effects 5
1. 2.2 Man-induced vibrations 6
1. 2.3 Machine-induced vibrations 7
1.2.4 Wind 8
1.2.5 Waterwaves 8
1. 2.6 Earthquakes 9
1. 2.7 Rail and road traffic 9
1. 2.8 Construction works 10
1. 2.9 Impact 10
1. 2.10 Blast waves and explosions 11
1. 2.11 Loss of support 11

~ Man-induced vibrations 13

~.1 General aspects 13


~.2 Dynamic loading from human motions 13
2.2.1 Walking and running 14
(a) General characterization 14
(b) Time function of the vertical load 17
(c) Time function of the horizontal load 22
(d) Influence of the number of people 24
(e) Deliberate excitation 26
2.2.2 Skipping 26
(a) General characterization 26
(b) Load-time function 27
(c) Influence of the number of people 29
(d) Deliberate excitation 30
2.2.3 Dancing 30
2.2.4 Joltingmotion 31
2.3 Effects of man-induced vibrations 31
2.4 Measures against man-induced vibrations 33
2.4.1 Pedestrian structures 33
(a) Frequency tuning 33
(b) Calculation of a forced vibration 35
(c) Special measures 35
2.4.2 Office buildings 36
(a) Frequency tuning 36
(b) Calculation of a forced vibration 37
(c) Special measures 38
2.4.3 Gymnasia and sports halls 38
(a) Frequency tuning 38
(b) Calculation of a forced vibration 39
(c) Special measures 40
2.4.4 Dancing and concert halls 40
(a) Frequency tuning 40
(b) Calculation of a forced vibration 41
(c) Special measures 41
2.4.5 High-diving platforms 42

3 Machine-induced vibrations 45
3.1 General aspects 45
3.2 Dynamic loads of various types of machinery 45
3.2.1 Machines with rotating parts 45
3.2.2 Machines with oscillating parts 50
3.2.3 Machines with impacting parts 53
3.3 Effects of machine-induced vibrations 56
3. 3.1 Inertial vibrations 56
3.3.2 Structure- and air-borne acoustic waves 57
3.4 Measures against machine-induced vibrations 58
3.4.1 Machines with rotating and oscillating parts 58
(a) Frequency tuning 59
(b) Calculation of a forced vibration 65
(c) Special measures 65
3.4.2 Machines with impacting parts 65
(a) Frequency tuning 65
(b) Calculation of a forced vibration 66

cceptance criteria 67

General aspects 67
Structural criteria 68
4.2.1 Standard DIN 4150, Part 3 (1983) 69
4.2.2 Standard SN 640312 (1978) 70
4.2.3 Directive KDT 046/72 (1972) 70
4.2.4 Draft ISO/DIS 4866 (1984) 70
Physiological criteria 73
4.3.1 Standard DIN 4150, Part 2 (1975) 75
4.3.2 Standard ISO 2631 (1980) 78
4.3.3 Directive VDI 2057 (1983/1981/1979) 80
4.3.4 BRE Digest 278 (1983) 80
4.3.5 British Standard BS 5400, Part 2 (1978) 80
4.3.6 British Standard BS 6472 (1984) 81
4.3.7 NBC Canada, Commentary A (1985) 82
4.3.8 Regulation SBA 123 (1982) 82
4.3.9 Criteria from the literature 82
Production-quality criteria 83
Suggested overall acceptance levels 85

)ENDIXA: Casereports 87
1: Footbridge of 1. 92 Hz 87
2: Footbridge of ~ 2.3 Hz 89
3: Footbridge of ~ 4.0 Hz 89
~: Footbridge of ~ 1.1 Hz lateral 89
5: Office building 91
5: Exhibition pavilion 93
7: Gymnasium building 93
~: Sports hall adjacent to grandstand 96
~: Gymnastics room 97
LO: Concert hall with pop music 98
11: Open-air theatre 100
L2: 10 m diving platform 100
L3: 5 m diving platform 101
L4: 10 m indoor-diving platform 101
L5: Weaving factory building 101
L6: Factory building with plastics-forming machines 103
No. 17: Factory building with metal-forming press 105
No. 18: Factory building with automatic boring machine 106
No. 19: Dairy with butter churns 107
No. 20: Factory building with stamping machines 107
No. 21: Production and assembly building with milling and grinding machines 108
No. 22: School building with heat pump 108

APPENDIX B: Fundamentals a/vibration theory 111


B 1 SDOF systems 112
B 1.1 Free vibration 112
B 1.2 Forced vibration 116
B 2 MDOFsystems 125
B 2.1 Free vibration 125
B 2.2 Forced vibration 126
B 3 Distributed-parameter systems (eigenfrequencies) 132
B 3.1 Single-span beams 132
B 3.2 Continuous beams 135
B3.3 Frames 137
B3.4 Arches 137
B 3.5 Plates and slabs 138
B 3.6 Substitute SDOFsystems 140
B 4 Harmonic analysis 144
B 5 Frequency tuning 149
B 5.1 Low tuning 149
B 5.2 High tuning 153
B 6 Structural behaviour under impact 154
B 6.1 Single impact 154
B 6.2 Periodic impact 155
B 7 Vibration absorbers 157
B 7.1 Definition 157
B 7.2 Categories 157
B 7.3 Linear vibration absorbers 158
B 7.4 Shock absorbers 160
B 8 Dynamic material properties 161
B 8.1 Strength and ductility 161
B 8.2 Flexural rigidity of cracked reinforced-concrete members 162
B 8.3 Damping 164

References 167
Notation 173
1

In this chapter the different types of dynamic loads are characterized, followed by a
description of their possible effects and existing measures for their limitation.

nrrHUlrnr Loads

1.1.1 General rJA.""'u ........."

In broad terms, a major part of the loads encountered in civil engineering can be
designated as dynamic because they vary with time. In practice, however, slowly varying
loads can be treated as quasi-static since the inertia and damping forces are negligible. The
presence of inertia and damping forces is in fact the important distinction between dynamic
and static loading. These forces arise from the accelerations and velocities induced in the
structural member, and they have to be included in the calculation of stress resultants and
support reactions. The magnitude of these forces and their function against time depend
both on the kind of excitation from outside and on the intrinsic dynamic behaviour of the
structural member, i. e. its dynamic properties.

According to their time function, dynamic loads can be categorized (Figs. 1.1 and 1.2) as
- harmonic
periodic
- transient
impulsive.

Additional aspects and criteria, respectively, are the number of load cycles, the ensuing
loading or strain rate in the affected member, the probability of occurrence of the load or
the peak values it may attain (Tab. 1.1, Section 1.1.6).
2 Survey of Problems in Structural Dynamics

dead load

static f<E::'-------------I permanent load

slowly varying
life load

machine operation

loads
human motion

harmonic

wind

water waves

periodic

earthquakes

dynamic
rail and road
traffic
transient
construction
works

impact
impulsive

blast wave l
explosion

loss of
support

Fig. 1.1 Types of loading in civil engineering


Characterization of Dynamic Loads 3

(a) harmonic load ( b) periodic load

~ ~ ~
""-.7 ""-.7
.. t
~ ~.~ period Tp
/ .. t

(c) transient load (d) impulsive load

~qnAv~~ .. t
~, ~
V v .. t

Fig. 1.2 Typical time functions of dynamic loads

1.1.2 Harmonic Loads


Harmonic loads vary according to a sine function with or without a certain phase shift.
They affect the structure for a sufficiently long time to permit a steady-state vibration
response (Fig. 1.2a, see also Appendix B 1). They may be caused by
machines with synchronously rotating masses which are slightly out of balance
(e. g. generators)
machines with intentional out-of-balance forces (e. g. vibrators).

1.1.3 Periodic Loads


Periodic loads exhibit a time variation which is repeated at regular intervals called
periods. Although the function within one period is arbitrary (Fig. 1.2b), its repetition
allows it to be decomposed into a series of harmonic loads through a Fourier transfor-
mation (<<harmonic analysis, Appendix B 4). Again, the duration of the load is long
enough for steady-state response to develop.
4 Survey of Problems in Structural Dynamics

They may arise from


- human motion such as walking, skipping, dancing, deliberate excitation, etc.,
- machines having
more than one unbalanced mass (e. g. rotary presses, centrifugal pumps, blowers,
centrifugal separators, vibrators, etc.)
oscillating parts (e. g. combustion engines, reciprocating compressors, textile ma-
chines, frame saws, bells, etc.)
periodically impacting parts (e. g. punching machines, presses, etc.).

Periodic loads can also be caused by wind, such as by vortex shedding, even though its
general nature is transient.

1.1.4 Transient Loads


Transient loads exhibit an arbitrary time variation without any periodicity. Also the
duration of the load is arbitrary (Fig. 1.2c). Such loads can be due to
- wind, namely in the wind direction as opposed to vortex shedding in the lateral direction
- water waves
- earthquakes
- rail or road traffic, either acting directly on a structure or interacting through the ground
(ground vibrations)
- construction works (e. g. sheet or pile driving by ramming or drilling, use of vibrating
rollers, blasting operations, etc.).

1.1.5 Iml~ul~;ive Loads


Impulsive loads are also of a transient nature. Their duration, however, is so short that
the structural member affected reacts in quite a different manner (Fig. 1.2d). This type of
loading can result from
- machines operating in single impacts (e. g. forge hammers)
- construction works (e. g. in the immediate vicinity of blasting operations)
- impact (e. g. crash of vehicles, airplanes, ships, or projectiles hitting the structure, rock
fall on a protective gallery, etc.)
- blast waves
- sudden collapse of load-bearing members (e. g. columns in buildings or piers of bridges).
Characterization of Dynamic Loads 5

1.1.6 Additional LuaUlHI~ Criteria

Apart from the load-time function, other features can be of importance (Tab. 1.1).
These include
- the number of load cycles over a given period or over the entire life of the structure
(N in Tab. 1.1)
- the strain rate (E in Tab. 1.1) or rate of loading
- the peak values of the dynamic load and the rise time
- the probability of occurrence of exceptional loads, such as earthquakes, blast waves,
etc., during the life time of the structure.

static and quasi - static dynamic

short term long term fatigue low-cycle fatigue impulsive

10 3 < N < 10 7+10 8 10 < N < 10 3 1 < N < 20

t < 10- 5 t < 10- 5 10- 5 < t < 10- 3 10- 5 < t < 10- 2
10- 3 < t < 10 + 10
1 2

. creep and . traffic earthquake impact


relaxation . machinery blast wave
temperature collapse of
support

Tab. 1.1 Characteristics of static and dynamic loads:


N = number of load cycles, E = strain rate [s -1]

1.2 IlvrHII1i11C Loads and lLounter'm(~asun~s

1.2.1 Classification of Effects


Dynamic load effects can be divided into three categories, i. e.
- effects on structures
- effects on people
- effects on machinery and installations.

Effects on Structures
A structure can be affected in both its load-bearing capacity and its serviceability.
6 Survey of Problems in Structural Dynamics

Impairment of the load-bearing capacity can take the following forms:


- Fatigue:
The stress fluctuation associated with the dynamic load can critically impair the strength
of the material after a certain number of cycles, all the more the higher the stress
excursions are. Design against fatigue, as it is called, does not just depend on the type of
loading, but also on the planned or expected life time of the structure. Distress due to
large stress excursions, of which a few are sufficient to cause collapse, is termed low-
cycle fatigue (Tab. 1.1).
Local plastification:
Under extraordinary high loads with a low probability of occurrence, such as vehicle
crash against columns and earthquakes, localized or limited plastic deformation may be
accepted in order to enable an economic design of the respective structural member.
- Alteration to the material properties under high-speed loading:
A very short rise time of the load causes a high strain rate in the respective structural
member (Tab. 1.1). The strength and stiffness parameters of most materials are sensitive
to the strain rate (see e. g. [1.1] and [1.2]). For steel and concrete, the usual construction
materials in civil engineering, the strain-rate effect is mostly beneficial to the load-
bearing capacity of the structure.

Impairment of the serviceability entails damage mostly to nonstructural elements of a


building, such as cracks in partitions, loss of cladding, etc.

Effects on People
The serviceability is impaired when the vibration of the building under dynamic load
causes disturbance and discomfort to the occupants.

Effects on Machinery and Installations


The serviceability to manufacturing processes can be impaired by vibrations transmitted
through the structure. Secondary vibration of the installations may hinder the production
process, affect the proper functioning of machines, and cause associated material-tech-
nological problems.

1.2.2 Man-Induced Vibrations


Human motion is sufficient to cause various dynamic loads. They can be periodic in
nature (e. g. due to walking, running, skipping, dancing, etc.) or transient (e. g. by jumping
off a high-diving platform). Man-induced vibrations are thus to be generally expected in
Source-Dependent Effects of Dynamic Loads and Countermeasures 7

- pedestrian structures
- gymnasia and sports halls
- high-diving platforms in swimming pools.

The resulting vibrations can lead to the following forms of distress:


- overstressing of the structure, in extreme cases to the loss of structural integrity
- damage to nonstructural elements (e. g. cladding on sports halls)
- intolerable vibration velocities and accelerations disturbing and discomforting the
respective users
- excessive noise (e. g. due to reverberating equipment).

Of all possible countermeasures, the foremost idea is to avoid resonance phenomena,


for instance by
increasing the stiffness of the structure or the particular structural member
- installation of (tuned) vibration absorbers
- regulating the use to avoid critical loading.

1.2.3 Machine-Induced Vibrations


Machinery can affect buildings or structural members in several ways with quite differ-
ent types of loading, depending on whether the critical forces arise from a rotating,
oscillating or intermittent motion of the machine parts. Thus, machines can produce
harmonic, periodic or transient loads (Fig. 1.1). The vibrations induced in the structure
may cause
- failure of structural members in fatigue
- damage to nonstructural elements (e. g. partitions)
- intolerable vibration velocities and accelerations impairing the well-being of operating
personnel and, possibly, persons in the neighbourhood
- excessive deformation interfering with the ongoing production (e. g. problems of
tolerances in the manufactured items).

Again, avoidance of resonance is the primary precaution, with provision of additional


damping to suppress remaining vibrations. Thus, the available countermeasures are
- stiffening of the structure or specific structural members
- disconnecting parts of the structure to cut off vibration transmission
installation of springe-damper) elements for isolation against vibration
- adjusting the speed of the machines (i. e. the period of loading)
- provision of tuned vibration absorbers.
8 Survey of Problems in Structural Dynamics

1.2.4 Wind
Slender structures exposed to wind can essentially be subject to two different types of
vibration:
In the wind direction, vibrations are excited by gusts. The resulting structural vibrations
interfere with the wind flow leading to motion-induced wind loads, which may enhance
the vibrations.
- In the lateral direction, vibrations are excited by periodic vortex shedding on alternating
sides of the structure.

These kinds of vibration are to be taken into consideration when designing


high-rising and slender structures (e.g. towers, chimneys, skyscrapers)
- structures of large spans or slim profile (e. g. suspension bridges).

The wind-induced vibrations may affect these structures through


- deterioration caused by fatigue
- complete loss of load-bearing capacity
- failure of the fasteners of nonstructural elements (e. g. cladding)
- discomfort to occupants (e. g. excessive sway of a high-rise building).

To mitigate these effects, the following are the primary countermeasures taken:
- stiffening of the structure to avoid resonance phenomena
roughening of its surface to moderate vortex shedding
- aerodynamic optimization of the profile
- installation of tuned vibration absorbers.

1.2.5 Water Waves


Water waves, caused by the interaction of the bottom air layer with the water surface,
are characterized by their height, which determines the size of the loads on a structure, and
their wave frequency or period. Vibrations due to water waves have to be considered in the
case of the following structures:
offshore structures (oil-drilling rigs)
- harbour structures.

Detrimental effects may include


loss of load-bearing capacity
- excessive deformation, intolerable to the proper functioning of the app'liances or the
well-being of people.
Source-Dependent Effects of Dynamic Loads and Countermeasures 9

The basic countermeasure is once again the stiffening of the structure to avoid reson-
ance.

1.2.6
Geotectonic activity in the earth's crust and the upper mantle trigger earthquakes,
primarily of the dislocation type. Over a longer period of time elastic strain energy
accumulates in a fault zone, often of historic origin, until the rock strength in the critical
direction is reached. During the rupture a substantial part of the energy released is radiated
as kinetic energy in the form of seismic waves. They induce in buildings both horizontal and
vertical motions.

The possible damage due to seismic vibrations encompasses


- loss of the load-bearing capacity
- local irreversible deformation (plastic zones) of the primary structural members
destruction of secondary elements, installations and equipment.

To control such damaging effects, the precautions have to include


- the choice of a suitable structural system with emphasis on symmetry
- appropriate seismic design loads for the structure
careful detailing of structural members and nonstructural elements.

1.2.7 Rail and Road Traffic


Traffic-induced vibrations have various sources and emanate, for instance, from
unevenness in rail track and road surfaces and the corresponding motions induced to
vehicles. They occur in a directly loaded structure (e. g. abridge) or are observed in
buildings adjacent to traffic routes and railway lines. The vibrations can produce
cracks in structural members or nonstructural elements
- disturbance to people on the structure or within the building.

Provisons may include


stiffening of the structure or installation of tuned vibration absorbers in the case of direct
loading
- repaving the roads and bridges
- for railway tracks, elastic support of the ties, of the concrete slab-track, etc.
- disconnecting parts of the structure to prevent transmission.
10 Survey of Problems in Structural Dynamics

1.2.8 Construction Works


Construction works may be a source of vibrations when they include ramming or vibro-
driving of piles and sheet piles, soil compaction with vibratory rollers, blasting, etc.

The effects of such vibrations may be


cracks in nonstructural elements of buildings in the neighbourhood
- disturbance to people living in these buildings.

The vibrations can be kept within bounds by


- using a smaller pile hammer
- adjusting the frequency of the vibrator
choice of an alternative method of construction (e. g. drilling instead of ramming)
- limiting the explosive charge in blasting
restricting the time and duration of execution.

1.2.9
With regard to impact loads, local and global effects need to be clearly distinguished.
While the local behaviour refers to influences in the zone of impact, the global behaviour
describes the loading and vibrations spreading through individual structural members or
the structure as a whole. Depending on the deformability of both the impacting object and
the part of the structure directly affected, the impact may be characterized as hard or soft
[1.2].
Impact loads may result from
- vehicles crashing into piers, parapets, etc.
- airplane crashes on buildings
- rockfall on protective galleries.

The possible effects comprise


spalling of concrete in the impact area or on the reverse side
- punching failure
failure of the impacted structural member in bending
- inadmissible vibrations in adjacent members.

Besides provisions to eliminate the chance of impact, measures to reduce the effects on a
structural member include
- design for an appropriate impact load
- detailing for a large absorption capacity prior to failure (ductile behaviour)
Source-Dependent Effects of Dynamic Loads and Countermeasures 11

- protection of the load-bearing structure with cushioning material (e. g. covering a


rockfall-protection gallery with gravel).

1.2.10 Blast Waves and ~xpl{~sJ(ms

Explosions generally produce a blast wave affecting the entire structure as a pressure
front (gas explosion, nuclear warfare), but they may result alternatively in local impact if a
charge is applied directly to a structural member.

Accordingly, the damage may involve


- local destruction of particular members due to direct hit (which may involve total loss of
the structure by progressive collapse)
- development of plastic zones in the primary structural members (possibly to the point of
failure) under large-scale surface pressure loading.

To avoid these kinds of damage, the following countermeasures may be introduced:


- design of the structure specifically against progressive collapse
design of the structural members for the pressure wave load
covering the structural members at risk with cushioning material.

1.2.11 Loss of :-Sulmort


If a load-bearing member fails (e. g. a column, a wall, etc.), a step load is exerted on
adjacent structural members previously supported (e. g. a floor slab). The effect can also
be one of a sudden load removal.
Under these circumstances - often a direct consequence of an external impact -, the
subsequent collapse of neighbouring structural members or the progressive collapse of the
structure as a whole can only be prevented if these members have the ability to redistribute
the forces quickly and to sustain the resulting stresses.
Apart from protective measures against impact (e. g. crash barriers around bridge
piers), the structure could specifically be designed to have the capacity to survive the loss of
a particular structural member and to redistribute the forces.
13

In an increasing number of cases, structures are found to be unduly responsive (<<lively)


to human motion, resulting in disturbing or even harmful vibrations. Structures intended
for moderate live loads of an assumed static nature are often designed with rather slender
dimensions, the possibility that dynamic loading might govern the design being overlooked
or underestimated. Another factor is the advent of new kinds of user activities such as
fitness classes to the accompaniment of strongly rhythmic music in gymnasia, which may
bring about a considerable dynamic loading. Although overloading of the structure itself is
not the primary concern, secondary building elements (e. g. cladding, windows) can be
damaged and the comfort of people impaired. They may feel alarmed enough to leave their
places precipitately (see [2.1] and Case No.7). One is usually more sensitive to vibrations
induced by someone else than those which are partly due to one's own activity. Thus, man-
induced vibrations are basically a problem of serviceability.

This chapter first describes the dynamic loading functions (force vs. time) resulting from
various human motions. It then shows their effects. The countermeasures are classified
according to categories of structures in which man-induced vibrations are to be expected.

2.2 Dynamic Loading from Human Motions

Man can cause various types of dynamic loads by his physical activity. The loads may be
of periodic or transient nature (Fig. 1.2).
Periodic loads are mainly the result of the following forms of human motion:
- walking
running
skipping
- dancing.
Of course, this is just a crude categorization. Other forms of motion such as rhythmic
skipping during fitness classes, jazz dance sessions, foot stamping, hand clapping and body
rocking at a concert, etc., are included or may be a combination of those forms.
Transient loads result primarily from a jolting motion imparting a single impulse to a
structural member (e. g. take-off from a diving platform, landing on a floor after jumping
from an elevated position, bumping against a wall with the shoulder, etc.).
14 Man-Induced Vibrations

Periodic loads depend in their time function as well as in their frequency on the pace and
the kind of motion. A closer look at the time function of the periodic loading reveals that
considerable forces are not just transmitted in the actual frequency of the walking,
skipping or dancing rhythm, but also in the frequencies of upper and lower harmonics.
A systematic categorization of man-induced loading is difficult. Apart from the
aforementioned factors, the number, for example, of the people involved in the excitation
plays a role as well. The following section deals first with periodic loads due to walking and
running, skipping and dancing. Then the transient loads caused by a jolting motion are
briefly discussed (Section 2.2.4).

2.2.1 and KUlllnill~

(a) General Characterization


The motion forms of walking and running may give rise to considerable dynamic loading
on pedestrian structures, for example footbridges, overpasses, connecting passageways,
etc. These forms can be characterized by the pacing rate, the forward speed and
specifically - by the time function of the loading.

Pacing Rate
The pacing rate (fs) dominates the resulting dynamic load. It is sometimes given as
footfalls per second [FF/s], but its nature as loading frequency is more adequately expres-
sed in Hz. For normal walk on horizontal surface, both Matsumoto [2.2] and Schulze [2.3]
found it to range between 1.5 and 2.5 Hz (Fig. 2.1). Assuming a Gaussian normal
distribution around the mean of 2.0 Hz yields a standard deviation of 0.13 Hz [2.3] to 0.18
Hz [2.2]. Kramer [2.4] gives a slightly different average of 2.2 Hz with 0.3 Hz standard
deviation.
For normal jog, the mean pace rate varies from 2.4 to 2.7 Hz [2.5], [2.10]. For sprinting it
may be as high as 5.0 Hz [2.6]. On public pedestrian structures, however, pacing rates
above 3.5 Hz are rare [2.5].

Forward Speed
The speed or velocity of pedestrian propagation (vs) is coupled with the pacing rate (fs)
through the stride length (Is)' Naturally, different people may possess quite different stride
lengths and paces for the same forward speed. Figure 2.2 gives average values for this
interrelation from numerous tests [2.6], which are compiled in the following Tab. 2.1.
Dynamic Loading from Human Motions 15

100

III
c:
.2
~
~ 50 measured
o distribution
o
c: Gaussian
distribution

Fig. 2.1
a 'ff-L.-"--+-.--t-----.----f""="l'--+__---
1.6 2.0 2.4 2.8 Distribution of pacing rates
for normal walk (from [2.2])

forward speed Vs [m/s 1


..--.2.0 2 3 4 5 6 7 8
...s
...J'
.c:
0.
c:
1.5
~
Q)
"0
:EIII 1.0

0.5

Fig. 2.2
a Forward speed and stride
a 1.0 2.0 3.0 4.0 5.0
pacing rate f s [HZ] length during walking and
running (after [2.6])

fs [Hz] Vs [m/s] Is [m]


slow walk ~ 1.7 1.1 0.60
normal walk ~ 2.0 1.5 0.75
fast walk ~ 2.3 2.2 1.00
slow running (jog) ~ 2.5 3.3 1.30
fast running (sprint) > 3.2 5.5 1.75

Tab. 2.1 Correlation of pacing rate, forward speed and stride length for walking and
running (average values after [2.6])
16 Man-Induced Vibrations

Load-Time Function
When walking or running, a person exerts a vertical and horizontal dynamic load (in
forward and lateral direction), which, in fact, has the physical dimension of a force. The
main parameters affecting the load-time function are the following (see also [2.10]):

"0

-
'-
If)

0
)(
0
200f
100: --
n --n--rrlt
--- ,fotic we;g"f

10.921 ~ 12.451 13.381


12

z==s:--o--rvr-f\-
0
..c.
If)
Cl
C
:.;::
200f
u
0
..-
If)
'00:
11.051 ill 12.671 13.481

;=\
-0
c
If)
"0

-'-
0
)(
0
20001::
'00: --r\---fj
[JJ]] ~
--fvr- - 12.321 @.TIJ
0

200f
-Z=S---C\ --f\--A-
(/)
If)
Cl
C
:.;::
u
0
..-
If)
100:
~ ~ 12.671 13.381

'-
(I)
..0
..0
:::J
\-
-
If)
"0
'-
0
)(
0
200f
100: z::s--n---(\ -fr-
10.851 ~ ~. lMJ
\-
(I)
>
200f
0
-0
0
C
(/)
If)
Cl
.S:
oX

E
u
100: Z3 -M---fv(-fr-
If) [O:83J !T.65l[[@ 13.3 01

I I I I I I I r-----i 1-----1
0 050 1.00 1.50 S 0 0.50 1.00 0 0.50 0 0.50
"
Fig. 2.3 Load-time functions for various pacing rates, footwear, and surface conditions
(from [2.7])
Dynamic Loading from Human Motions 17

- pacing rate
- stepping particularities (heel/ball contribution)
- person's weight
- person's sex
- type of footwear (or lack of it)
- floor surface condition (softness of cover).

(b) Time Function of the Vertical Load


Given the multitude of parameters, the results of different investigations vary greatly,
influenced also by the test procedure and measuring technique adopted. The influence of
pacing rate, footwear and floor surface on the development of the dynamic vertical load
was thoroughly examined in [2.7]. As Fig. 2.3 reveals, the latter parameters are of minor
importance compared to the pacing rate. The weight of the test person was 1100 N. The
shape of the load-time function for walking with a medium pacing rate is more or less that
of a saddle; the two observable load maxima result from stepping with the heel and pushing

1.4 1.4
:E :E
CJl
'(j) 1.0 ! - next footfall CJl
'(j) 1.0
3: I 3:
,~
slow walk ,~
fast walk
:(J)
0.6
1.67 Hz :g(J)
0.6
2,94 Hz
....... .......
"0 0.2 "0 0.2
c c
.9 .9
0 0.5 1.0 5 0 0.5

:E 3.0
CJl :E
'(j)
1.0 ,~
3: Q)

walk 3: 2.0
,~ jog
'0 0.6 2,38 Hz ~
iii .E
(J) 1.0
....... .......
"0 "0
c 0.2 c
.9 .9
0.5' 1.0 5 0 0.1 0.2 0.3
1.4

:E :E 3.0
CJl CJl
'cv 1.0 'cv
3: brisk walk 3:
() ()
spri nt
0.6 200 Hz
(J) (J)
I.
....... .......
"0 0.2 "0
c c
.9 .9
0.5 1.0 5 0 0.1 0.2 0.3

Fig. 2.4 Modification of load-time function with pacing rate [2.6]


18 Man-Induced Vibrations

off with the ball of the foot. This feature disappears with increasing pacing rate and
degenerates to a single maximum of sharp rise and descent when the person is running.
From low to high pacing rates, the width of the signal decreases, and the load maximum
increases. While for strolling with a frequency below 1 Hz the maximum load hardly
exceeds the weight of the person, it increases by a quarter or a third for 2 Hz and by a half
around 2.5 at about 3.5 Hz the maximum reaches about double the weight of the test
person.
For relatively high pacing rates (above ca. 2.5 Hz) somewhat larger maxima are indi-
cated in Fig. 2.4, which is taken from Wheeler [2.6] and partly based on Harper's work
[2.8]. For fast running the maximum load can increase to three times the weight (Fig. 2.5).
The same figure shows a clear relation between the pacing rate and the duration of foot
contact with the ground.

:E ~
01
'(i)
\ peak load/static weight
0.

~ 3.0
\ 0.6 6
u \ "-6
\
E Vl \
0.55
"0
:;; 2.0 \ 0.4t>
o \
.s .8
'\ 0.36
-'" u
o "-
~ 1.0 "- "- 0.2
.......
...............
contact duration - - -
-- 0.1
Fig. 2.5
Contact duration and peak
O-'-----i'----+---t---j------t----'-0
o 1.0 2.0 3.0 40 5.0 load (reI. to static weight)
pacing rate fs [ Hz] as depending on the pacing
rate (from [2.6])

1: 1. 5 - - - r - - - - - - r - - - - , - - - - . , - - - - - - . - - - - - - , - - - , - - - - - - ,
Cl
'(j)
~
u
~
in 1.0 - ; r . - - f - - - t - - - \ t - - - - + - - - + - - - - , , - ' i ( - - - - - J ' - - j - - - - - - \ j
.......
"0
o
2

0.5 -+-JL.--+---j---+---+---j------t--t--+---if-----j

0.1 0.2 0.3 OA 0.5 0.6 0.7 s

Fig. 2.6 Load-time function resulting from footfall overlap during walking
(from [2.10])
Dynamic Loading from Human Motions 19

During walking, one of the feet is always in contact with the ground at a given time (see
in Fig. 2.4 the time function of the second foot in broken line), such that the contact
duration of the two feet and thus the loads overlap. Figure 2.6 [2.10] shows the resulting
time variation of the total dynamic load during walking. The load function clearly has
components in the 2nd and 3rd harmonic, i. e. a pulsating load is also exerted at double and
triple the pacing rate. The time function given in [2.3], although slightly different, exhibits
basically the same phenomenon.
In contrast to the behaviour during walking, ground contact during running is interrup-
ted, the ratio of contact duration to pace period becoming smaller with increasing pacing
rate.
An idealized mathematical formulation of the dynamic load variation must take this
difference between walking and running into account by choosing different approaches for
continuous ground contact and discontinuous ground contact.

Continuous Ground Contact


The load-time function for walking which exhibits an overlap of the individual contact
times of either foot (Fig. 2.6) can be idealized by the following expression:

F p (t) = G + ~Glsin(2Jtfst)
+ ~Gzsin(4Jtfst-CfJz) + ~G3sin(6Jtfst-CP3) (2.1)

where:
G weight of the person (generally assumed to G 800 N)
~Gl = load component (amplitude) of 1st harmonic
~Gz = load component (amplitude) of 2nd harmonic
~G3 = load component (amplitude) of 3rd harmonic
fs = pacing rate
CfJz = phase angle of the 2nd harmonic relative to the 1st harmonic
CP3 = phase angle of the 3rd harmonic relative to the 1st harmonic.

The force component (i. e. the Fourier amplitude coefficient) of the 1st harmonic may be
taken from the literature (e. g. [2.3], [2.6]) and from own test results [2.10] to be
~Gl O.4G for fs = 2.0Hz (2.2)
~Gl = 0.5G for fs 2.4 Hz (2.3)
with linear interpolation in between.
The load components of the 2nd and 3rd harmonic in the range of fs ::::= 2 Hz can be taken
from [2.3], [2.10] as ~Gz ::::= ~G3 ::::= O.lG.
20 Man-Induced Vibrations

The phase angles crz and cr3 exhibit in reality a large scatter because of the many
parameters they depend on [2.10]. In a computation they may be approximated to crz = n/2
and cr3 = n/2 (partly different values are given in [2.30]). If the most unfavourable
combination of the different harmonics is to be captured, the phase angles need to be
varied. In most cases, however, a forced vibration induced by walking is governed by just
one harmonic, so that phase angles become immaterial.

Discontinuous Ground Contact


The load-time function for running, which is generally characterized by a single load
maximum, can be expressed by a sequence of semi-sinusoidal pulses (<<half-sine model,
Fig.2.7a).

The function within one period is given by

(2.4)

with:
= Fp,max/G dynamic impact factor
= peak dynamic load
weight of the jogger (generally assumed to G = 800 N)
contact duration
= Vfs = pace period.

The impact factor k p results from the condition of constant potential energy, i. e. that the
integral of the load-time function over one pace period must equalize the load at rest (static
weight). Figure 2.7 shows how k p varies with the ratio tpiTp.
It is interesting to compare the theoretical values of k p from the half-sine model (noting
that tpfs = tplTp in Fig. 2.7b) with the average experimental values given in Fig. 2.5 for a
certain contact duration t p and pacing rate fs : In the range of about 3 Hz they agree rather
well, whereas for 2 Hz and 4 Hz the half-sine model produces roughly a 30% larger
dynamic impact factor. Thus, the half-sine model can, in this case, be considered conserva-
tive with respect to the magnitude oftheload. Experiments described in [2.10] yielded (for
running) a k p of more than factor three, which is slightly higher than theoretical values.
Altogether, the half-sine model seems generally to be appropriate.
The time function according to the half-sine model can be brought into the same format
as Eq. (2.1), i. e. as the sum of static weight G and harmonic load components:
Dynamic Loading from Human Motions 21

(0)

7 ,
u
.E 5
1\
U
CJ
0.4
.
\
3
\
2
~
----- Fig. 2.7
Idealized load-time function
for jogging and skipping:
0.00 0.25 0.50 0.75 1.00
(a) half-sine model
t p /T p (b) impact factor depending
on reI. contact duration

00 t
F (t) = G
p
+ n=O
L: L\.Gncos[2nonofs(t - _P )]
2'n
(2.5)

with:
G = weight of the person (generally assumed to G = 800 N)
L\.Gn load component (amplitude) of n-th harmonic
n = number of n-th harmonic
fs = pacing rate
tp contact duration.
22 Man-Induced Vibrations

1.5

e..:>
'c: w- ,,
e..:>
<l
\ ,
\ "
\
0.5
\ \ ""
\ ,
'. " '\ Fig. 2.8
0.0+- ' . /~' ><:-.::-:::.=:: . <---
+--''-L---'-~--'-'--'''''--~:.>..jL~~~=j
:.; Load components (amplitudes)
0.00 0.25 0.50 0.75 1.00
t p /T p of several harmonics according
to the half-sine model

Figure 2.8 shows the load component amplitudes L\G n of the first four harmonics versus
the contact duration ratio tplTp as calculated by harmonic analysis (see Appendix B 4).
Note that, depending on this ratio, higher harmonics may still contribute substantially to
the loading. However, similar to walking, computation of a forced vibration is dominated
by just one harmonic, so that the phase shift (2nfstpl2 for all harmonics) does not matter.

Loading due to walking or running can usually be assumed as stationary (i. e. spatially
fixed) excitation. The relatively slow forward speed has hardly any effect on the vertical
excitation of the structure as investigations have confirmed (e. g. [2.7] and [2.11]). If one
were to consider the effect that the forward speed of a single walking or running person has
on a forced vibration, the response would be transient with smaller amplitudes, since the
steady state is not reached before the person leaves the structure.

Time Function of the Horizontal Load


The loading from human walking or running is much smaller in the horizontal (longitudi-
nal and lateral) directions than in the vertical direction. Even then it may become a
problem for very flexible structures (see [2.12] and Case No.4).

Investigations for a pacing rate of 2.0 Hz are presented in [2.3]. Looking at the time
function of the lateral vibration displacement in Fig. 2.9, it becomes obvious that the
lateral sway of the person's centre of gravity occurs with half the pacing rate (i. e. 1.0 Hz),
whereas the longitudinal displacement is dominated by the full pacing rate as in the vertical
direction.
A harmonic analysis of the pertinent loading according to Fig. 2.9, exerted by a
relatively light person of 587 N, reveals more details, Fig. 2.10. The major lateral load
Dynamic Loading from Human Motions 23

test I test .II


(0) IR IL _
IR L +200
/ - ......J\ /1
vertical
I ( I r \ \ ( \ { \ o
\ I \ I \ "- \ { \
\ J \ \ J '-...1 '- - 200
" \( J \I \ v 'l
- 400

test I test TI
R L R L
lateral +2.0 [em]
~

........... l7 -- -........ 1/
-.;;

displacement
\.
--- ./ '- ./
- 2.0

0 0.5 s 0 0.5 5
+ 200 [em/5 2 ]
/'\.....7 / '\......-/" o
""v'
'-- 1\ ..--...7 .......- ...........11 .\.......-
- 200
acceleration

(c) test I t es t]I


R L R L
longitudina I +2.0 [em]
I '\. ~ '\ / '\. 7'" o
""'-
displacement
~
f "-
- / '\ '\.. "-
2.0

0 0.5 s 6 0.5 5
~
+200 [em/5 2 ]
/ /1 / /'
/\ / \ , / 1\ / 1\ O
/ \ / \ / / \ / \ /
\ / \ / \ / \ I \/ -200
v v
V
V 'V
acceleration

Fig. 2.9 Time function of the vertical dynamic loading with resulting horizontal vibra-
tion displacements and accelerations for a pedestrian walking with 2 Hz (from
[2.3])

components are associated with frequencies of f/2 or 3f/2; in the longitudinal direction
the major components have fs and 2fs ' accompanied by a component in f/2 due to a more
pronounced footfall on one side.
24 Man-Induced Vibrations

216 N ::::: 37% of static weight

120 N
62N 69N
49N
23N 25N

LJ.G 1 LJ.G z LJ.G 3 6G 4 6G s LJ.G 1/ Z 6G i 6G 3/ Z LJ.G z 6G s/z


2Hz 4Hz 6Hz 8Hz 10Hz 1Hz 2Hz 3Hz 4Hz 5Hz 1Hz 2Hz 3Hz 4Hz 5Hz

(a) vertical load (b) lateral load (c) longitudinal load

Fig. 2.10 Harmonic load components (Fourier amplitudes) of the directional load-time
functions of Fig. 2.9 (from [2.3])

Influence of the Number of


The results above are for a single person. More interesting in practice is the combined
effect of more than one person at a time, particularly in walking. For that purpose the
following aspects have to be considered:
- The density of pedestrians is limited by the feasibility of uninhibited walking. According
to [2.3] an upper limit is reached with 1.6 to 1.8 persons/m2 , which corresponds to a static
load of about 1100 to 1400 N/m2 ; however, a value of 1 person/m 2 (as assumed in [2.2])
seems to be more realistic.
- Based on observations it is proposed in [2.3] that pedestrians walking initially with
individual pace on a footbridge will try to adjust their step subconsciously to any
vibration of the pavement. This phenomenon of feedback and synchronization becomes
more pronounced with larger vibration of the structure. One observation in the case of a
laterally swaying footbridge is reported in [2.12] (see Section 2.3). During tests by the
authors [2.10] it appeared to be rather difficult to keep to one's usual step when
disturbed by vertical vibration displacements of more than 10 or 20 mm; the person tends
to fall out of step and adjusts more or less to the motion of the pavement. However,
displacements have to be that large for the phenomenon to be noticeable.

The phenomena of more than one person walking C,an be described as follows: As long as
just one person with a pacing rate between 1.5 and 2.5 Hz walks near the centre of a larger
Dynamic Loading from Human Motions 25

span, he will produce more or less a forced steady-state response (after an initial
tramW::;lll phase). A vibration produced by a differently walking second person will super-
,~,"'''<'O the first response so that at certain times the two vibration amplitudes will be
~rirI1TH1{-> or subtractive - depending on frequency and phase. Thus the amplitude of the
displacement will not be constant but exhibit a typical surge from interference of
(see Case No.1, Fig. ALIa). With the number of pedestrians increasing and their
enhancing one another at times, larger vibration amplitudes may occur. As men-
tioned before, the capacity of the bridge deck limits the number of people who can walk at
certain speed (about 1.5 m/s at fs = 2 and with the necessary clearance (about 1 min
all directions at fs = 2

A mathematical description of excitation by more than one pedestrian is difficult. After


Matsumoto [2.2] one can assume a Poisson distribution for the arrival probability of
pedestrians and derive an enhancement factor (m) to be applied to the vibration amplitude
caused by a single pedestrian in the centre of the span:

m = VATo (2.6)

where:
A mean flow rate (persons/s over width of the deck) for a certain period of
time (maximum Amax ~ 1.5 person/s'm)
= time necessary to cross the bridge of length L at speed vs , i. e. To L/v s
ATo number of persons simultaneously on the bridge at the given mean flow
rate.

This formula has not been verified in the field. Extensive computer studies [2.6], in
which arrival times, individual weights and pace rates were varied at random, tended to
confirm it, especially for bridges with a natural frequency near to the most probable pacing
rate at 2 Hz. For example, the simulation with 100 persons/min. on a footbridge of 2 m
width and 26 m span yielded m 5.5, comparing well with m 5.4 from Eq. (2.6) for V s =
1.5 m/s. At the current state of knowledge the formula seems to hold true. In practice, the
flow rates often will be relatively small and thus the factor m too. Possible infrequent
excess of vibration acceptance criteria could be tolerated anyway.
Equation (2.6) may be directly applied to pedestrian structures with a natural frequency
f1 in the range of normal predominant pacing rate between 1.8 and 2.2 Hz (Fig. 2.1).
Higher and lower pacing rates are quite rare; hence it is suggested that for structural
frequencies f 1 between 2.2 Hz and 2.4 Hz and between 1.8 Hz and 1.6 Hz the enhancement
factor is reduced linearly to mmin 2. O. This figure corresponds to two people marching in
step.
26 Man-Induced Vibrations

Deliberate Excitation
Besides the response to one or more pedestrians walking at random, to joggers etc.,
there is the possibility that footbridges and similar structures are deliberately excited to
resonance by a group of people (<<vandal loading). This may happen in the vertical
direction through marching in step or skipping to time, or in a horizontal direction through
rhythmical shift of the centre of gravity. This aspect is given closer qualitative consider-
ation in [2.9], for example.
The regulations [2.13] introduce a load case extraordinary live loads to cover groups
marching in step, spectators rocking to music, and deliberate excitation. Equivalent static
loads are prescribed, under which the behaviour of the structure has to be checked.

2.2.2 SkilPpil12
General Characterization
This kind of motion is likely to produce considerable dynamic loads particularly in
gymnasia and sports halls, but it may also become critical in other types of buildings where
jazz dance sessions are practised. The deliberate excitation of pedestrian structures may
also be effected by skipping. Skipping can be characterized by the skipping frequency (fs)
and the time function of the dynamic load.

Skipping Frequency
More recently gymnastics to rhythmic music have become quite common. The items of
music are chosen such that their rhythm suits various skipping, jumping and running
exercises. During fitness classes, frequencies for different forms of motion ranging from
2.0 to 3.2 Hz were found [2.1]. Jazz dance sessions with skipping showed the same
frequency range [2.14]. Tests with skipping for a longer time (1 to 2 min.) on the same spot
revealed frequencies from 1.0 to 2.8 Hz [2.4]. It seems unlikely, that even for a short time
(say 20 s) the human physiology would permit a frequency of more than 3.5 Hz. This leaves
a range between 1.8 and 3.4 Hz for calculation purposes.

Load-Time Function
A person skipping exerts primarily a vertical load on the floor. The main parameters
affecting the time function of this dynamic load are the following:
- skipping frequency
- intensity (moderate or maximum height)
person's weight
- type of footwear
- floor surface condition (softness of cover).
Dynamic Loading from Human Motions 27

(0) 3000~-'----

2000

15 1000
.s

1.0 1.5 2.0 25


time [s]

800~
600
't:J
400
Fig. 2.11
.so Loading caused by a person's
200
skipping (from [2.15]):
(a) Load-time function
2 3 4 5 6 7 8
frequency [Hz] (low-pass filtered at 9 Hz)
(b) Fourier amplitude spectrum

Load-Time Function
Although the range of possible frequencies is bounded as shown above, the variability of
rhythmic exercises during gymnastics or other sporting events leaves much heterogeneity
in terms of the load-time function. The following concentrates on skipping on the spot
(both feet simultaneously) as a form of motion, which is easy enough to describe but still
representative with view to the frequency range as well as to the maximum possible load
amplitude.
Skipping of a single person with a frequency of fs = 2.1 Hz is shown in Fig. 2.11 [2.15],
both in terms of the resulting time function of the load and the associated Fourier
amplitude spectrum. The latter exhibits a clear load component in the 2nd harmonic. The
ratio of maximum load to the person's weight, i. e. the impact factor k p = Fp,maJG (see
Section 2. 2.1), reaches kp 3.4in this example. The maximum peak load of a single person
skipping is measured in [2.10] to be six times his weight (i. e. kp == 6 with possible extension
to kp == 7 for extremely high skipping).

For an idealized mathematical description and calculation of a forced vibration, the


half-sine model - a sequence of semi-sinusoidal pulses (Fig. 2.7a) - can be used again.
Equations (2.4) and (2.5), which were derived for walking, are valid with Tp and fs as
skipping period and skipping frequency. The resulting impact factor increases with
decreasing ratio tpiTp According to experiments reported in [2.10], this contact ratio may
vary between 0.25 and 0.6.
28 Man-Induced Vibrations

Because of individual differences in the motion, allowing for softer or harder bounce on
the floor irrespective of the frequency, for skipping the relation between contact duration
and frequency is not a priori defined as it was for walking and running (see Fig. 2.5).
Hence, various contact durations t p for the same frequency fs (i. e. various tiTp) yield a
spectrum of impact factors k p (see Fig. 2.7b). It was observed in [2.10] that physiology
hardly permits a contact duration below t p =:= 0.15 s. From that the minimum contact
duration ratio is:

(2.7)

Inserting the practical minimum skipping frequency of fs =:= 1. 8 one arrives at tiT p =
0.26 and can take from 2.7b a factor of k p =:= 6; this is the maximum impact factor
reported in [2.10].
Going back to the results given in Fig. 2.11a, the half-sine approximation of the time
function yields tiTp =:= 0.50, i. e. for the measured fs =:= 2.1 Hz an idealized contact duration
of t p =:= 0.23 s. Figure 2.7b gives a dynamic impact factor of k p =:= 3.1, which compares well
with the measured ratio of maximum dynamic load to weight of k p =:= 3.4.
To further assess the validity of the half-sine model, a look at the Fourier amplitude
spectrum of Fig. 2.11b is informative. Generally, skipping on the spot produces a major
loading in the frequencies of the 1st, 2nd and 3rd harmonics, i. e. in fs , 2fs and 3fs ' But the
spectrum shows that loads can also be exerted in f/2 and 3f/2, or other intermediate
harmonics, as the time variation of the load is apparently not exactly periodic in the
skipping frequency. This phenomenon can become much more pronounced, if the loads
exerted from either foot differ, e. g. when leaping from one foot to the other. Complex
physical exercises as occur during jazz dance sessions give rise to load-time functions
considerably in variance with the half-sine model. Figure 2.12 shows the Fourier amplitude
spectrum of measured accelerations of a gymnasium floor. Substantial response was
observed at 3f/2 (besides fs 2.3 Hz and 2fs = 4.6 which happened to coincide with
the 3.5 Hz fundamental frequency of the floor structure. This kind of excitation can result,
for instance, when one foot touches the floor with a slight time lag during skipping.
Important as these subtleties may appear in certain cases, for general calculation purposes
forced vibrations in particular the half-sine model is sufficiently accurate.

c:: 0.008 3.5 4.6 Hz


o
'';:: 0.006 t-- 2.3 Fig. 2.12
e
"*
u
0,004 t-- Fourier amplitude spectrum
g 0.002 r- of the accelerations of a
o ""'\.A..;./'I. ,1'0. J">.

0 1 2 "" 3 4 5 6 7 gymnasium floor with


frequency [Hz] f1 = 3.5 Hz at fs = 2.3 Hz
(from [2.15] after [2.14])
Dynamic Loading from Human Motions 29

Influence of the Number of


When comparing Fig. 2.13 to Fig. 2.11, one finds little difference in the load-time
for skipping of one person or a group of eight persons. This is due to the
.c.,.~"i-HYr~C
remarkable synchronization by virtue of the accompanying music. A difference exists,
however, with respect to the minimum possible duration of contact.

(0) 20000...r::----.r-----

10000
-0
0
.3

0
0 0.5 1.0 1.5 2.0 2.5
time [5]
(b) 5000

4 000

3000
-0 Fig. 2.13
0 2000
.3
i 000

0
0 2
\.
3
j
4 5
-.A
6 7
Loading due to 8 persons
skipping in phase (from [2.15]):
(a) Load-time function
8
frequency [H z] (low-pass filtered at 9. Hz)
(b) Fourier amplitude spectrum

Due to slight variations of the individual motions in a larger group of participants, which
are never completely synchronized, the minimum contact duration increases to t p 2: 0.20 s,
resulting in a minimum contact duration ratio achievable by groups of

(2.8)

Inserting fs 2: 1.8 Hz as a realistic minimum frequency for skipping brings the contact
duration ratio to tplTp = 0.35 and leads in Fig. 2.7b to an upper bound of the impact factor
of kp 2: 4.5 for a larger group.
In contrast to the case of randomly walking pedestrians (see Section 2.2.1), where the
dynamic load increases with the square root of the number of people involved, for
skipping-type motion the dynamic load increases as a linear function. The upper limit of
30 Man-Induced Vibrations

occupancy in gymnasia, as observed by the authors in fitness classes with skipping and
jogging exercises, is about 1 person per 4 m 2 , or 0.25 person/m2 In a gymnasium of 20 x 30
m2 floor area, this would amount to ca. 150 participants. A denser occupancy was assumed
in [2.16] with 1 person per 2 m 2 , or 0.5 person/m2

Deliberate Excitation
Skipping is the easiest method to deliberately induce resonance in a structure vibrating
vertically. In tests [2.6] on 22 footbridges, a group of two or three persons tried to excite the
structure to the largest possible displacements in its fundamental frequency. In some cases,
the attained displacement amplitudes hardly exceeded those due to normal pedestrian
traffic, whereas in all the others they did not even approach them. Hence, vandal loading
was not considered critical. The same conclusion was reached in [2.5]. However, our Case
No.1 deals with a footbridge of 40 m span which three persons, skipping continuously in
the centre of the span, could excite to a displacement amplitude of 9.4 mm; this was more
than double the amplitude of vibrations under a flow rate of 29 pedestrians per minute.

2.2.3 UaJIlCIIU!

During normal dance events, but also during concerts with a clapping, stamping and
rocking audience (see Cases No. 10 and 11), serious dynamic loads may arise. Not much is
known about the load-time functions of the resulting dynamic loading since it is rather
difficult to measure these loads directly.

Dancing Frequency
Clapping of hands, foot stamping or body rocking occurs with the beat of the music,
which may vary between 1.6 Hz for slow and 3 Hz for fast pieces. As the exerted forces
usually vary from one beat to the other (unequal load on either foot, etc.), the load
spectrum may also contain frequencies below the dancing frequency, for instance semi-
harmonic frequencies as described in Section 2.2.2.

Load-Time Function
To measure directly the load-time function induced by dancing is difficult. Since floor
contact usually is maintained, the maximum dynamic force is unlikely - according to Figs.
2.3 and 2.4 - to greatly exceed the static load. However, the total load can be substantial
due to the dense occupancy of the floor, particularly at concerts. In [2.17] the density is
estimated to be 2 persons/m2 , but can possibly increase locally to 6 persons/m2
For the purpose of calculating a forced vibration (see Section 2.4.4), an idealized
mathematical load-time function can be taken from the motion type of walking with
continuous ground contact (Section 2.2.1). The pertinent equation is (2.1). Following
Effects of Man-Induced Vibrations 31

the amplitude coefficient of the 1st harmonic may be taken as = 0.5G, that of
2nd harmonic is estimated by the authors to be 0.15 G. the very nature of the
of dancing motion, this recommendation produces results which are approximate
best.

til UJIIL.lUUI~ Motion

A person may exert a single impulse by taking off from a high-diving platform, by
on a gymnasium floor after jumping from an elevated position, or by bumping
lClUUHJ'F,

a wall with his shoulder. These loads are transient. The importance of these kinds
of impact loading is minor and has been little investigated. The time function of the load is
seldom available. A most extensive discussion of the subject is given in [2.18]. Further data
is found in [2.19] or [2.20] and [2.21], wherein an equivalent spring stiffness is derived for a
person jumping onto the floor.

2.3

Man-induced vibrations affect the structure and occupants in rather different ways.
For the structural integrity, they are rarely a problem. More seriously affected is the
serviceability of the structure. Besides damage to nonstructural elements the major
concern is for the wellbeing of the people affected. They tend to feel those vibrations to be
particularly disturbing which are not caused by themselves. Section 4.3 gives acceptance
criteria to evaluate the severity of possible discomfort.

Effects of man-induced vibrations on structures or structural members include


loss of the load-bearing capacity; although rare, collapse may occur due to overloading
of the structure (Hyatt Regency walkway in Kansas City [2.22]) or failure in fatigue;
cracking of structural members, cladding and partitions, falling away of plaster, etc.

People may feel affected by man-induced vibrations in several ways:


Mechanically as, for instance, by the vibrations of a footbridge or an intermediate floor
in a gymnasium. High-diving platforms in swimming-pools are a special case, their
vibration can irritate the athletes to the point that the platform becomes useless.
- Acoustically by the noise from reverberation and rattling of equipment; this effect can
be quite dramatic, evoking pronounced anxiety and flight from the place (see Case
No.7).
32 Man-Induced Vibrations

. concrete
0 steel or composite
kZ3 sensitive
frequency range

..
... ...
0

0 0
o
o. 0
0

2 ;::; :%
0

Fig. 2.14
o Fundamental natural frequencies
o 10 20 30 40 50 60
span [m] of footbridges of different span
(from [2.5])

Optically by impressions like irritatingly large deflections of the ceiling under a gym-
nasium hall or frightening sway of the diving platform from the viewpoint of the athlete.

The magnitude of man-induced vibrations in structures or of structural members


depends primarily on the ratio between the dominant excitation frequency (see Section
3.3.1) and the fundamental natural frequency of the structure or structural member. The
natural frequencies depend on the span and the statical system of the structure, its flexural
rigidity and mass. Figure 2.14 gives a survey of spans and fundamental frequencies of 44
footbridges examined in Britain [2.5]. Generally, the fundamental frequency decreases
with increasing span, but it can differ widely in structures of similar span, so that definite
conclusions cannot be drawn.
The fundamental frequency, but also higher frequencies of a structure can be excited. A
case is reported in [2.6] in which a steel footbridge with a natural frequency f1 8 Hz
exhibited resonance at a pacing rate offs 3.7 Hz. The reason was that the 3rd harmonic of
the dynamic load (f = 33.7 = 11.1 Hz) excited the 2nd structural frequency f2 = 11.1 Hz.
In some cases a footbridge can also get excited in longitudinal or lateral direction. This
can be caused by dynamic loads with half the pace rate or a multiple of it (see Figs. 2.9 and
2.10). Note that the synchronization phenomenon described in Section 2.2.1 is more
pronounced for lateral vibrations. It was held partly responsible in [2.12] for the observed
amplification of substantial lateral vibrations of a steel bridge. Presumably, the pedestrian,
having noticed the lateral sway, attempts to reestablish his balance by moving his body in
the opposite direction; the load he thereby exerts on the pavement, however, is directed so
as to enhance the structural vibration.
Measures Against Man-Induced Vibrations 33

As for machine-induced vibrations (see Chapter 3), the first and foremost counter-
measure is to change the structural frequency (called frequency tuning, see Appendix B 5).
Normally, this provides a sure solution. In some cases the calculation of a forced vibration
is to be recommended. Special measures, e. g. installation of tuned vibration absorbers,
are mainly suitable for improvement of existing structures (see Appendix B 7).
In the following, frequency tuning, calculation of a forced vibration, and special
measures are discussed individually for each type of structure. High-diving platforms are
an exception in that, besides the frequency, the stiffness of the structure enters as a second
criterion. The goal of each measure is to suppress the structural vibrations to admissible
upper bounds. Acceptance criteria for these bounds are given in Chapter 4 (for high-diving
platforms, see Section 2.4.5). Appendix B 3 contains approximate formulas for determin-
ing the fundamental frequency of the respective structure, Appendix B 8 data on material
properties.

2.4.1 Pedestrian Structures


The following countermeasures against man-induced vibrations are applicable to foot-
bridges and similarly loaded structures such as overpasses, connecting passageways (<<sky-
walks), long-span staircases, boat landing stages, etc.

(a) ~relquenc:v

Considering the statistical distribution of pacing rate (see Section 2. 2.1), the largest risk
of excitation in the vertical direction is for pedestrian structures having a natural frequency
f1 between ca. 1.6 and 2.4 Hz. Fundamental and higher frequencies should be kept out of
this range by all means:

f. -----< 1.6 Hz (2.9)


1 > 2.4 Hz

Moreover, since dynamic force components from walking are also exerted at double the
pacing rate (2nd harmonic, see Section 2.2.1 and Case No.3), bridge structures with
relatively low stiffness and damping, particularly steel and composite steel bridges, should
be designed not to have fundamental and higher frequencies in the range between ca. 3.5
and 4.5 Hz either:

f. < 3.5 Hz
(2.10)
1 ----> 4.5 Hz
34 Man-Induced Vibrations

The frequency clearance need not be as large as for the 1st harmonic (Eq. 2.9), as the
forces excerted in the 2nd harmonic are smaller.
It is conservative in practically all cases to have a high tuning of the fundamental
frequency to ca. 5 i. e.

(2.11)

Compliance with the recommendations (2.9) and (2.10) does not eliminate occasional
resonance in frequencies between the two clearance ranges, but this will be a rare loading
for most structures. It could occur at brisk walk with fs ~ 2.5 or if ajogger passed with fs
3 Hz. Other pedestrians, however, walking at the same time with fs ~ 2 Hz would disturb
the vibration thus tending to hinder resonance effects.
There rarely are structures with low stiffness and simultaneously low damping in the
lateral direction, i. e. transverse to the general traffic axis (see Case No.3, for example).
Frequencies in the range 0.8 to 1.2 Hz should be avoided and, depending on circumstances,
also in the range 2.6 to 3.4 Hz (see Fig. 2.10b).
Also in rare cases certain structures possess very low stiffness in the longitudinal
direction, i. e. in the direction of traffic (e. g. a frame structure with rather flexible struts),
or are supported on very soft bearings. Frequencies in the range 0.8 to 1.2 Hz should be
avoided here as also in the second range from ca. 1.6 to 2.4 Hz (see Fig. 2.10c).

In the following, some regulations in codes of practice are drawn attention to:
The British load specifications for bridges [2.25] waive vibration serviceability checks for
pedestrian and cyclist overpasses if the fundamental frequency f 1 without live load lies
above 5 Hz. See also Section 4.3.5.
- The Canadian code [2.16] adopts a frequency of f1 = 6 Hz as threshold of checking
requirements.
The East-German code for bridges, footbridges and stairs [2.13] uses as a criterion the
fundamental frequency (vertical, longitudinal and lateral) of the structure when loaded
with 300 kg/m 2 additional mass, f 1,g+p' If this frequency is above 6 Hz, a check is not
required; if it lies between 6 and 3 then the dynamic performance has to be evaluated
for stairs but not for the other types of structure. For these checks in the vertical and
longitudinal directions, the relevant pacing rates are taken as fs and 2fs , while in the
lateral direction in addition f/2 and 3f/2 are investigated (as follows from the harmonic
analysis of the load-time function of walking, see Fig. 2.10). The code specifies an
equivalent static load procedure for computational purposes. For a range 1 Hz < f1,g+p <
3 pedestrian bridges have to be checked in the same way. Genuine dynamic
investigation is stipulated for f1,g+p < 1 Hz.
Measures Against Man-Induced Vibrations 35

The natural frequencies of pedestrian structures should always be computed with


conservative upper and lower bounds to account for secondary elements such as pave-
ments or railings, the magnitude of the dynamic modulus of elasticity, the transition to the
cracked state and the tension stiffening of the concrete between cracks in reinforced
concrete structures, etc. (see also Appendix B 8).

In the cases of
- a very flexible structure
very low damping properties (e. g. in welded steel structures)
- impossibility of frequency tuning as recommended
high demands for the comfort of the user,
one should either calculate a forced vibration or - if that proves to be inadequate adopt
special measures.

Calculation of a Forced Vibration


Calculating a forced vibration in general at matched frequency as steady state response
Appendix B 1) will give an idea of the dynamic performance of the structure. The
load-time function can be chosen for walking according to (2.1). Several uncertainties
may arise: The often inaccurate determination of the fundamental structural frequency is
one; others are the assumed damping properties (see Appendix B 8) and the influence of
more than one pedestrian (see Section 2.2.1d); the latter is to be estimated from an
average density of people and the mean flow rate of pedestrians respectively. The results of
such a computation must be interpreted with reservation. are to be compared to the
acceptance criteria in Chapter 4.

~plecl:[U Measures
If the structural frequency cannot be properly shifted or the calculated forced vibration
yields inadmissible amplitudes, special measures have to be considered. They are particu-
larly appropriate to improve an existing structure.
Special measures may, for example, be
- installing a stiffer hand rail
- adding tie-down cables
increasing the damping
- attaching a tuned vibration absorber.
The first two measures aim at stiffening the structure. For instance, a hand rail may be
substituted or complemented by a truss construction. A pedestrian overpass across a
motorway could be tied down in midspan. If symmetrically attached up to the middle third
of the span, cables can effectively suppress horizontal vibrations in the 1st and 2nd mode
shape. Arranged in an inclined plane, they simultaneously reduce vertical vibrations.
36 Man-Induced Vibrations

Substantially increasing the damping properties proves more difficult, unless the struc-
ture's intrinsic damping is very low. In structures with ~ == 0.5% (of critical damping),
friction devices may be installed, for instance, in hand rails or at bearings. However, as
soon as the intrinsic damping is close to ~ == 1%, damping devices become too bulky for
realization or too expensive.
Sometimes, the installation of one or more vibration absorbers is more promising. Such
a vibration absorber consists of a mass-spring-damper system which is attached to the
structure or its vibrating part at a selected spot and carefully tuned with respect to the
primary structure's vibration characteristic. It is a typical improvement device. Some
actual applications are reported in [2.2], [2.5], [2.11], [2.23] and [2.24]. How such an
absorber works and how its behaviour is calculated, is shown in Appendix B 7. The
effectiveness of the absorber is primarily determined by the ratio of the absorber mass to
the (modal) mass of the structure and increases with smaller damping of the structure. This
favours steel structures for possible application.
The extent to which encountered vibrations are to be reduced is given by the acceptance
criteria in Section 4.3, especially those in Sections 4.3.5 (BS 5400) and 4.3.9 (values taken
from the literature on pedestrian structures and not laid down in codes).

2.4.2 Office HUlildiJD2S


The countermeasures discussed in the following sections concern vibrations caused by
the random walk of people in office and administation buildings and disturbing personnel
at their working places.

People spending some eight hours a day sitting or standing at their working places are
rather sensitive to building vibrations. The tolerance of acceptable values is thus neces-
sarily stricter (see e. g. [4.25]), and - in contrast to pedestrian structures, gymnasia and
sports halls, etc. - the 3rd harmonic of the dynamic load-time function must be taken into
consideration as well.
For floor slabs in office buildings (see Case No.5), our present state of knowledge would
suggest high tuning above the frequency ofthe 3rd harmonic of the load-time function (see
Fig. 2.10a). Moreover, the stiffness, mass and damping of the type of structure in question
must be taken into account.
As shown in Section 2.2.1, the walking pacing rate centres around 2 Hz and rarely
reaches more than 2.4 Hz. Consequently, the structure should comply with the following
fundamental frequencies:
- reinforced concrete f 1 > 7.5 Hz
prestressed concrete f 1 > 8.0 Hz
Measures Against Man-Induced Vibrations 37

composite (in-situ concrete slab on steel girders) f1 > 8.5 Hz


steel (e. g. corrugated sheets with concrete infill on girders) f1 > 9.0 Hz

The order of ascending frequencies is due to the accompanying decrease in stiffness,


mass and damping.
Observing these lower frequency bounds should generally lead to an acceptable vibra-
behaviour (see Chapter 4), although exceptionally they may prove insufficient for
structures with low stiffness and damping. On the other hand, massive concrete
structures (e. g. large-span joist floors) may perform well with fundamental frequencies
these bounds. Massive reinforced concrete slabs have frequencies well above 7.5 Hz
anyway. Hence the bounds given above for office buildings are mainly relevant for light-
composite and, in particular, steel floor structures.
Calculating the fundamental frequency of slabs in office buildings calls for conservative
estimates as to load sharing by the floor finish, the dynamic modulus of elasticity, the
transition to the cracked state of reinforced concrete members and the tension stiffening of
the concrete, etc. (see Appendix B 8). Stiffening due to nonstructural partitions should be
HV!;;'~V''-'~V'-< unless structural details such as load-carrying connections to the structure secure
the participation.

Special cases, e. g.
still higher demands of comfort to the user
vibration-sensitive equipment or installations
may require raising the lower frequency bound above the given values.
Otherwise, one can attempt to calculate the expected amplitudes of a forced vibration,
or one can resort to special measures.

(b) Calculation of a Forced Vibration


Random walk of people considered as a forcing function (see Appendices Bland B 2 for
the general procedure) can be tackled in analogy to pedestrian structures. For the excita-
tion from a single person, in most cases the dynamic load component of the 3rd harmonic
(Eq. (2.1)) must be considered. The assumption of a matched-frequency loading, e. g.
resonance steady state between this component and the fundamental frequency, will be
safe in connection with a conservative estimate of the damping properties (see Appendix B
8). The damping effect even due to soft floor covering and partitions or other installations
can be surprisingly small [2.26]. The influence of the number of people on the displacement
amplitude can be approximately assessed using Matsumoto's formula (Eq. (2.6)).
38 Man-Induced Vibrations

~DleCI:al Measures

The special measures applicable to office floor slabs are


- stiffening the structure
increasing the damping
- attaching tuned vibration absorbers (see Case No.6).

Increasing the damping, however, is rather difficult to achieve.

2.4.3 :iYlll1n~lsja and

The following measures to reduce man-induced vibrations in gymnasia and sports halls
apply by and large also for places of similar use, such as jazz dance studios, although
sometimes to a slightly relaxed standard (see Section 2.2.2).

Special attention has to be devoted to the design of the floors that are exposed to
rhythmic skipping, jumping and running exercises (see Cases No.7 and 8). The present
state of knowledge suggests for this kind of building high tuning with respect to the 2nd
harmonic of the load-time function, with skipping on the spot as relevant type of motion.
Moreover, this has to be done with due consideration of stiffness, mass and damping,
which vary with different floor designs.
Since critical frequencies for skipping lie between 2.8 and 3.4 Hz (see Section 2. 2.2), the
following values are recommended as lower bound for the fundamental structural fre-
quency (which happen to coincide with those for office buildings before):
reinforced concrete f1 >7.5Hz
- prestressed concrete f1 >8.0Hz
- composite (in-situ concrete slab on steel girders) f1 >8.5Hz
steel (e. g. corrugated sheets with concrete infill on girders) f1 > 9.0Hz

The order of ascending frequencies is due to the accompanying decrease in stiffness,


mass and damping.
Under normal circumstances these bounds ensure that the vibrations remain within the
acceptable range (e. g. with maximum accelerations below 5% g, see Chapter 4). In case of
doubt, a forced vibration analysis should be carried out and the results compared with the
acceptance criteria.
Following the relevant standards, gymnasium floors need only be designed for static,
i. e. non-dynamic loading. One exception is the Dutch code [2.27], which requires a
fundamental structural frequency of at least 5 Hz. The commentary on the Canadian code
of 1980 [2.28] contains the same provision, but is being revised with more detailed
Measures Against Man-Induced Vibrations 39

rec:olllmen<JatlOlls in the new draft of the commentary [2.16]. They distinguish between
of use of the structure and possible excitations, the mass and type of design of the
structure (e. g. also wooden floors), and the required level of comfort for the people
LJ.J-.1-'v~~--' Similar recommendations are given in [2.15].

Again, the calculation of the fundamental floor frequency should be based on conserva-
estimates of participation of the floor finish, dynamic modulus of elasticity, transition
the cracked state and the tension stiffening in reinforced concrete (see Appendix B 8).

cases in which
the above frequency bounds cannot be assured with certainty or
the demand for comfort is higher,
computation of a forced vibration or special measures are advisable.

Calculation of a Forced Vibration


Computing a forced vibration - generally for steady-state conditions, see Appendix B 1-
the dynamic behaviour of a gymnasium floor to be approximately assessed. As an
forcing function one can normally take the semi-sinusoidal loading function (<<half-
model) presented in Fig. 2.7a for the motion type skipping on the spot, starting
Eq. (2.8). To capture resonance conditions, the skipping frequency has to be chosen
respect to the fundamental structural frequency (an integer fraction in most cases).
The impact factor follows from Fig. 2.7b, while the Fourier amplitudes of the several
harmonics of the loading function can be taken from Fig. 2.8. Note, however, t~at small
Fourier amplitudes are subject to considerable uncertainties due to their sensitivity to
deviations of the actual loading function from the half-sine model. One should not
use the values of DG2 , DG3 and DG4 to the right hand side of the first zero points. In tests
[2.10] it proved difficult to excite a structure by the 3rd harmonic of the loading function
from a single person's skipping. Unless the skipping frequency is kept precisely, the
structural response drops substantially; this sensitivity is the more pronounced the higher
the fundamental frequency of the structure. From our present knowledge it seems unlikely
that structural frequencies above 7.5 Hz (to be matched by the 3rd harmonic of 2.5 Hz
skipping) could be excited to strong vibrations even by a larger group of people.
The dynamic load should be based on the assumption of 1 person per 2 m 2 maximum,
corresponding to 0.5 person/m 2 (see Section 2.2.2).
The damping properties to be used in the calculation may be taken from Appendix B 8.
They are assumed to be somewhat higher than those for pedestrian structures because of
the contributions of floor finish, cladding elements, etc. It has occasionally been men-
tioned that the damping may sometimes rise markedly (up to double its basic value) as soon
as people are present on the floor (e. g. [2.16]); this is only true during decay measurements
40 Man-Induced Vibrations

where the human body absorbs part of the vibration energy, but not for steady-state
response to excitation by the person himself.
Results of these computations may be judged against the acceptance criteria of Chapter 4.

(c) Measures
In analogy to the sections on pedestrian structures and office buildings, possible special
measures are
stiffening the structure
increasing the damping
attaching tuned vibration absorbers.

Of course, critical activities like rhythmic skipping or running may simply be prohibited,
but this would restrain the use of the building.

2.4.4 D3lnCJlni! and Concert Halls


In the following, measures against the type of man-induced vibrations described in
Section 2.2.3 are discussed. In concert halls, pop concerts that animate the audience to
activities such as stamping, clapping, rocking, etc., are critical. Depending on circumstan-
ces, the same basic considerations could be applied to other types of buildings hosting
similar events, for instance to grandstands in stadia.

The considerations are more or less those derived for gymnasia in the previous section,
namely a request for high tuning with respect to the 2nd harmonic of the dynamic load.
When comparing the extent to which the 2nd harmonic participates in skipping (~G2 ==
l.4G fortplT p == 0.30 from Fig. 2.8) and in dancing (~G2 == 0.15G, see Section 2. 2.3), one
arrives still at the same magnitude if the very different possible density of people is
accounted for as well (compare the respective sections on calculation of a forced vibra-
tion). As stated in Section 2. 2.3, the dominant dancing frequencies lie between 2 and 3 Hz.
Generally, the critical range of 3 to 3.4 Hz for gymnasia is avoided so that the 2nd harmonic
as criterion yields lower bounds on the fundamental frequency as before. For different
floor designs, the recommended bounds are now:

- reinforced concrete f1 > 6.5 Hz


prestressed concrete f1 > 7.0 Hz
- composite (in-situ concrete slab on steel girders) f1 > 7.5 Hz
- steel (e. g. corrugated sheets with concrete infill on girders) f1 > 8.0 Hz
Measures Against Man-Induced Vibrations 41

The increase in bounding frequency is again associated with the decrease in stiffness,
mass and damping.
For calculation of the fundamental frequency of floors in dancing or concert halls, the
body masses of the people on the floor have to be included in the correct way. Note, that
particularly during pop concerts in halls without (fixed) seats, the audience may gather
around the stage, leading to a loading density of up to 6 persons/m 2 This uneven distribu-
tion of people, together with the variety of their active or passive behaviour, may have a
strong influence on the size of the moving mass and hence the floor frequency (see Case
No. 10).
In addition to general conservativeness, the assumptions for the frequency computation
should reflect the concentration of the added human mass in active and passive areas
such that the given structural system is exposed to the most unfavourable mass distribu-
tion. The comments on frequency calculation given in Section 2.4.3 for gymnasia and
sports halls should be remembered.
Although it may be appropriate in some cases to calculate a forced vibration, the
uncertainty in the assumptions is much worse than for pedestrian structures or gymnasia.
Special measures are therefore often preferred.

(b) Calculation of a Forced Vibration


For the above reasons, a steady-state computation (see Appendix B 1) for dancing and
concert halls can yield just a crude approximation to the true dynamic behaviour. The load-
time function would be assumed as in Section 2.2.3.
As regards the density of people, one can assume a maximum of 1 person per ,0.17 m\
corresponding to 6 persons/m2 , unseated and 2 persons/m2 with fixed seats. Assumptions
on density and distribution of the dynamic load are similar to the considerations previously
made for the frequency computation.
Possible damping factors are to be found in Appendix B 8 and are generally identical
with those for gymnasium floors.
The results of the computation can be judged against the acceptance criteria in Chapter 4.

~D~ecl:lU Measures
As in the case of gymnasia and sports halls the following measures might be considered,
particularly for improvement of existing structures:
- stiffening the structure
- increasing the damping
- attaching tuned vibration absorbers
- restricting the use of the building.
42 Man-Induced Vibrations

Platforms
Vibrations of high-diving platforms need to be suppressed not to hamper the athletes
during their performance. They would be disturbed in their concentration by
- swaying of the shaft in a direction not necessarily coinciding with the direction of take-off
- rigid-body motion of the platform with similar effects
- vibration within the platform, which is particularly unpleasant as the athlete may be
given an unwanted spin.

Following [2.29], two types of criteria are to be met:


criteria of frequency equivalent to high tuning, necessitating a frequency computation
- criteria of stiffness to be checked by relatively simple static calculations.

Frequency Criteria
The bounds to be observed are listed in Tab. 2.2. They concern shaft sway, rigid-body
motion and platform vibration. A major distinction is to be made whether or not a spring-
board for figure diving is mounted on the platform. The rhythmic skipping on the spring-
board contributes much to the excitation of the platform, so that stricter frequency bounds
apply.

frequency bounds without with


springboard springboard
shaft vibrations f1 ? 3.5 Hz f1 ? 5Hz
(all fundamental modes in nodding,
lateral sway, twist)
rigid-body vibration f1 ? 7 Hz f1 ? 10 Hz
(flexibility of foundation)
slab vibration f1 ? 10 Hz f1 ? 10 Hz
Tab. 2.2 Frequency criteria for high-diving platforms (from [2.29])

Stiffness Criteria

The spatial vector displacement of the front edge of the platform (Fig. 2.15) must remain
under static loading (with or without springboard)
Measures Against Man-Induced Vibrations 43

d (Fx,Fy ) < 1 mm (2.12)


and the lateral front displacement alone
dx = 0.5d < 0.5 mm (2.13)
for a set of forces of = 1kN.

slab

r - - - - - - - , L - - - - - - . Fx = 0.5 kN

Fy = 1.0 kN

Fz = 1.0kN
:::::: shaft
Fig. 2.15
Multiaxial static loading
to check stiffness of diving
platforms (after [2.29])

The stiffness criteria are relatively severe. According to practical experience, however,
reinforced concrete platforms can be assumed to maintain their uncracked stiffness in
bending as well as torsion. The listed bounds were derived for high-performance platforms
suited to high and figure diving alike. For less professional demands in recreational indoor
or outdoor swimming facilities, these bounds could well be lowered (see Cases No. 12, 13
and 14).
45

3.1

Machine-induced vibrations of buildings and their structural members have an increas-


ing practical importance due to several reasons. A contribution from the structural side is
the continuing trend to higher exploitation of material strength and larger spans; another
part of the problem is the switch to larger machinery or one operated at higher speed. The
observed vibrations may possibly impede or even prevent the opening of new production
facilities. In most cases not the structural integrity and the overall safety are at risk, but
either the personnel operating these machines is continuously or temporarily irritated, or
production problems arise such as substandard goods due to unforeseen motion of the
machinery. Machine-induced vibrations are thus - similar to man-induced vibrations -
mainly a problem of serviceability.
The following chapter characterizes in its first part the dynamic loading exerted from
various types of machinery. The second part deals with possible effects on buildings or
structural members and with suitable countermeasures. The emphasis is thereby on small
and medium size machinery which is typically installed on floors of buildings and liable to
cause vibration problems there.

3.2 JVIHU11111t" Loads of Various Types

Depending on its manufacturing' purpose, state of maintenance and design details, a


machine causes distinct loads on the structural member it rests on. The loads depend
primarily on the type of motion the machine parts describe: whether it is rotating,
oscillating, or of an impacting type. The following descripton of the dynamic loads is hence
grouped into these three categories.
The time function of the exerted loading is periodic in general, sometimes even har-
monic (see Section 1.1, Figs. 1.1 and 1.2). In any case can a periodic load be decomposed
by means of a Fourier analysis into a number of harmonic components. The relevant
theory is explained in Appendix B 4.

3.2.1 Machines with Rotating Parts


Dynamic loads may arise from rotating parts of machinery if they are insufficiently
balanced or if electrodynamic fields are active.
46 Machine-Induced Vibrations

Loads due to rotating parts may exhibit either of two time functions:
- quadratic excitation
- constant-load excitation.

Quadratic excita~ion is always related to out-of-balance forces and hence frequency


dependent. It arises whenever the centre of mass of a rotating part does not coincide with
the axis of rotation. The product of mass and eccentricity is called unbalance.

Out-of-balance forces may originate from:


- excessive tolerance in manufacturing the rotating parts
insufficient flexural rigidity of the axle
- accidental influence on the operation (dust collecting on blower blades, etc.)
intentionally eccentric masses (as in vibrators)
- bearing clearance (due to wear)
- passage through a critical speed
accidental unbalance (e. g. failure of a turbine blade).

Machines with predominantly rotating parts are for instance:


- blowers and ventilators
centrifugal separators
vibrators
- washing machines
lathes
- centrifugal pumps
- rotary presses
plastics-moulding presses
- turbines
- generators.

The excitation due to out-of-balance forces is quadratic since the amplitude of the forces
increases with the square of the excitation frequency. If a single rotating part is out of
balance, it exerts a harmonic load in any direction in the plane perpendicular to the axis of
rotation. The amplitude of this load is:
2
4n 2 m 'e
=m , e--nb= 4n
2 f2
b=m , ew2 (3.1)
3600
where:
F(j) centrifugal force
m' mass of the rotating part (unbalanced fraction)
e eccentricity of the unbalanced mass
nb speed of revolution of the part [r. p.m.]
Dynamic Loads of Various Types of Machinery 47

fb excitation frequency (fb nb/60)


(D angular velocity of the rotating part.

In an arbitrary direction the harmonic load may thus be defined as

F(t) = Fwsin((Dt) = m' e(D2sin((Dt) (3.2)


which acts on the total mass m of the machine (inc!. the mass m' of the rotating part).

Taking a damped oscillator with a single degree of freedom (SDOF) as a rather


simplified model, this dynamic load produces the following displacement of the total mass
with time (see also Appendix B 1):

m' e(D2
x(t) = . C sin(Dt + cp) (3.3)
k

where:
1
C

2'~'(D'(D
= arctan (Dl2 - (D 21 = phase angle

with:
x displacement of the total mass m
(Dl = circular eigenfrequency of the SDOF oscillator
k = stiffness of the SDOF oscillator
~ = damping ratio (with respect to critical damping).

The displacement amplitude amounts to

1
xmax = (3.4)
k

or rewritten to a dimensionless dynamic magnification factor on the displacements:

v = Xmax'm
(3.5)
q m"e

with:
m k/ (DI = total mass.
48 Machine-Induced Vibrations

Figure 3.1a shows the influence of the damping ratio on the magnification factor V q and
the accompanying lag in the phase angle between loading and displacement function.

(0)

3.01---f---Jil-Ht-t--j

o 1.0 2.0 3.0 4.0 5.0


frequency ratio w/w i

o 1.0 2.0 ~o 4D ~o

frequency ratio ::;1

oO\)
3 1---l---HJ4-lli-"'-'0~.\ _0---1
o.\\)

1 2 3 4 5
frequency ratio w/Wj Fig. 3.1
Magnification factor and
shift in phase angle for
various damping ratios:
(a) V q for quadratic
o 2 3 4 5 excitation
W
frequency ratio W1 (b) V k for constant-load
excitation

Engineering practice has the notion of an admissible unbalance. This figure depends on
the type of machinery and its speed of revolution and is either provided by the manufac-
turer or can possibly be taken from guidelines (e. g. [3.1]). Data on admissible excen-
tricities can also be found in [3.2].
If more rotating parts with individual unbalances are mounted on the same shaft, they
rotate with identical speed but different phase angle, and hence they also produce a
resultant harmonic load.
If several unbalanced parts rotate with different speeds, the superposition of their
individual harmonic loads will, in general, not result in a strictly periodic load.
Dynamic Loads of Various Types of Machinery 49

Not only unbalances but also alternating electromagnetic fields can give rise to periodic
loads with considerable participation of higher and sometimes lower harmonics with
respect to the operating frequency.

The forcing function of constant-load excitation is of the form

F(t) Fosin(wt) (3.6)

with:
Fa amplitude of the force (constant).

The amplitude remaining constant, F(t) varies with sinewt). If the circular frequency
w of the rotating part is constant, the quadratic excitation (3.2) degenerates to a constant-
load excitation (3.6).

Taking again a damped SDOF oscillator as a rather simplified model of the structure or a
member thereof, a constant-load excitation produces the following displacement of the
mass with time (see Appendix B 1):

x(t) = k . C sin(wt + cp) (3.7)

with C and cp as in Eq. (3.3).

The displacement amplitude amounts to

1
(3.8)

In relation to the static displacement, one obtains the dimensionless magnification factor

V k-- X max xmax = 1 (3.9)


X stat Folk V (1- (WIW1)2)2 + (2~wlw1)2

Figure 3.1b gives the magnification factor and the phase angle for various damping ratios
under constant-load excitation. Note that the phase angle is the same for constant-load as
for quadratic excitation.
50 Machine-Induced Vibrations

3.2.2 Machines with Usciljlatjln~ Parts


Oscillating parts of machines always excite dynamic loads. The causative motion can be
translational, rotational with small angle, or a combination of both (pendular motion).

Machines with predominantly oscillating parts are for instance


- weaving machines
- piston engines
reciprocating compressors
- reciprocating pumps
- emergency power generators (diesel engines)
- flat-bed printing presses
- frame saws
- crushing machines
screening machines
tool machines
- bells.

Although all reciprocating machines or engines exert primarily an oscillating force in the
direction of the piston motion, they also give rise to a rotational load component due to the
excentric hinging of the connecting rod to the crankshaft. Either component is of the
quadratic excitation type. It is a question, however, of the number of pistons and of their
suitable arrangement with respect to each other to minimize the resulting load (see e. g.
[3.3]). Higher harmonics of the operating frequency sometimes become important.

Typical load-time functions from machines with oscillating parts are shown in Figs. 3.2a
and 3.3. As for weaving, Fig. 3.2a gives for various types of machines the time function of
the vertical load transmitted at the footings and Fig. 3.2b the pertinent spectra of Fourier
amplitudes. It becomes apparent that only for air-jet machines and shuttle looms do the
maxima of the transmitted loads coincide with the operating frequency (weft insertion
rate). With all other types the maximum occurs in a higher harmonic frequency, with
projectile machines, for instance, in the 4th harmonic (see Appendix B 4). This effect
mainly reflects the mode of operation of the various types. Without sley stop the weft
insertion rate governs the load-time function (Fig. 3.2a) and the Fourier amplitude
spectrum, whereas with sley stop this influence is clearly superseded by higher harmonics
(see [3.4]).
The static and dynamic footing loads of the four types of weaving machines in Fig. 3.2 are
compiled in Tab. 3.1. Apart from the vertical, considerable dynamic load components
result in warp thread direction (x-component), the ones in weft thread direction (y-
component) being much smaller. As shown by Natke [3.4], the spectrum of a machine
group is characterized by the individual spectrum of the respective type, i. e. no frequency
shift results, for instance, from varying the number of machines on a common base slab.
Dynamic Loads of Various Types of Machinery 51

(0) (b)

_~:::~Jw~
2000.,--------------,

6
LL
w

'"
0

o time [s] 1 ";;


air- jet weaving machine ROTI L5000

;ooo~

E
.~
3:

-woooL--------,---------J,
o time [s] i

r""""~
~v~+JY
0000

n
o

-:000
o timers] 1
ropier weaving machine DORNIER

-~:::~
~E

o time [s] 1
projectile weaving machine SULZER projectile weaving machine SULZER

Fig. 3.2 Machines with predominantly oscillating parts, here the vertical loading from
various types of weaving machines (from [3.4]):
(a) load-time function, (b) derived Fourier spectrum

Another example of a load due to an oscillating mass is the one exerted in horizontal
direction by a bell. Figure 3.3 gives a load-time function and its Fourier amplitude
spectrum.
In contrast to the load-time functions of most weaving machines, bells always exert the
largest load component in the fundamental harmonic, i. e. in their swinging frequency of
half the stroke rate. As the 2nd and 4th harmonics are vertical, the next structurally
important harmonic of only slightly smaller amplitude is the 3rd. Load components of the
5th or higher harmonics are mostly negligible for bells with a swinging angle of 70 to 90.
Resonance phenomena may thus be encountered for bells with 60 to 65 strokes per minute,
equivalent to a swinging frequency of about 0.5 Hz. In that case, the 3rd harmonic of the
bell load (f == 30.5 Hz = 1.5 Hz) may coincide with the fundamental frequency of the bell
tower, which is most often one of rotational bending (due to elastic encastrement in the
ground).
52 Machine-Induced Vibrations

warp beam
weaving machine weft insertion weft width range of

:L
Ruti shuttle 249/min 160cm spectrum +2 3+
[Hz]
foot static loads [kN] dynamic loads [kN]
no. Fz Fz Fx Fy
y
1 4.13 7.58 3.33 1.52 + 1 4+
2 6.14 6.77 3.26 1.10 3-16
cloth beam
3 8.26 6.26 3.46 L07
4 5.27 7.91 3.27 1.78

total 23.80

weaving machine weft insertion weft width range of


Dornier rapier 184/ min 190cm spectrum
[Hz]
foot static loads [kN] dynamic loads [kN]
no. Fz Fz Fx Fy

1 6.27 6.74 2.25 0.83


2 10.81 5.78 1.22 0.83 3-60
3 6.07 6.18 2.12 0.70
4 9.03 8.65 6.20 0.72

total 3218

weaving machine weft insertion weft width range of


Su Izer proj ectile 258/min 85 spectrum
[Hz]
foot static loads [kN] dynamic loads [kN]
no. Fz Fz Fx Fy
1 6.81 4.45 0.75
2 ( n. o. ) 9.58 5.52 0.56 3-40
3 9.35 4.45 0.60
4 6.23 5.17 1.01

weaving machine weft insertion weft width range of


Sulzer projectile 209/min 85 spectrum
[Hz]
foot static loads [kN] dynamic loads [kN]
no. Fz Fz Fx Fy
1 4.55 2.94 0.4 I
2 (n. 0.) 6.26 3.57 0.39 3-40
3 6.14 2.97 0.44
4 4.94 3.96 0.74

weaving machine weft insertion weft width range of


Ruti air - jet 465/min 150 cm spectrum
[Hz]
foot static loads [kNJ dynamic loads [kN]
no. Fz Fz Fx Fy Tab. 3.1
1 3.4 5 3.31 2.22 0.33 Compilation of static
2 5.75 3.18 1.00 0.22 5-25 and dynamic footing loads
3 10.35 2.66 2.04 0.47
exerted by the weaving
4 1.73 2.62 1.08 0.52
machine types of Fig. 3.2
total 21.28
(from [3.4])
Dynamic Loads of Various Types of Machinery 53

3.3
Load-time function caused
-4
-6 by bell ringing:
-8
harmonics 1, 3 and 5,
-10 T
1------------------1 and their superposition
(from [3.5])

0
Bell assemblies with larger swinging angles (up to 360 for the English bell type) may well
exhibit load components up to the 20th harmonic.

The two examples of weaving machines and bells sufficiently demonstrate that both the
time function of the load and the pertinent Fourier spectrum depend strongly on the
source. Additional and more detailed data on the load-time functions due to oscillating
parts can be found, for instance, in [3.2], [3.5], [3.6], [3.7.]' [3.8], [3.9], [3.10] and [4.24].

3.2.3 Machines with Impa4~tnlle Parts


Machines with impacting parts often develop large intermittent dynamic force~. Skilful
design, however, will attempt to balance (e. g. with a counterblow hammer) the major part
of the force within the machine frame. This reduces the residual forces on the structure
very much.

Machines with impacting parts are for instance


- moulding presses
punching machines
- power hammers
- forging hammers.

While power and forging hammers exert a truly intermittent impact load, systems like
moulding presses or punching machines have at least some oscillating parts. In contrast to
true impact loading, these mixed dynamic loads of the second type of machine exhibit a
periodic peak followed by a more or less free vibration decay of the impacted body. As the
operating frequencies of these machines can be rather high, complete vibration decay
between impacts may not always be possible.
54 Machine-Induced Vibrations

For power and forging hammers, the load-time function is in most cases transient, i. e.
rarely periodic. The exact shape of the function depends on the operating mode of the
machine, but also on the moulding properties of the processed material. One can distin-
guish the loading phase (duration of impact) from the following, sometimes much longer,
phase without load. In general, the load-free phase of vibration decay is not of constant
duration, hence the nature of the loads is transient (single impulses). There are some
machines on the market, though, which do have a constant load-free phase, resulting then
in a periodic type of loading.
The loading can be characterized by the following parameters of the impact phase
3.4):
- duration of impact (tp )
- momentum (I)
- rise time of the load (t a)
- peak force (Fp,max) .

c 2.0+----j---f::,..---!!---+--L.-i----'---+---i--+----.<
.2
"8
;;::
1.6-1--j-l-~A~~=t:~~d:::==:;===l==J
'2
g 1. 2+---+~-+-7""'-I--1-_-'-_--L--'----I---+--1
E
.S? 0.8+--+-+-7--+----1--1 Fig. 3.4
E
o Dynamic magnification for
~ 0.4 +-I77'-+--+--I----+---'\---,-"---,--I---t---1
-0
various loading functions
0.2 0.4 0.6 0.8 i.O 1.2 1.4 1.6 1.8 2.0 as depending on the ratio
t p /T 1 between load duration and
structural frequency

The combination of these parameters determines the shape of the load-time function,
for instance semi-sinusoidal, triangular or even quite rectangular. In most cases just the
maximum of a dynamic quantity resulting from the impact is of primary interest, e. g. the
peak displacement X max of the centre of mass of the affected structural member. Rep-
resented by an SDOF oscillator, Fig. 3.4 shows for various shapes of load-time functions
the dynamic magnification factor versus the ratio tpiTl (duration of impact to the natural
period of the oscillator). The dynamic magnification factor on the displacement is defined
as in (3.9):

(3.10)
Dynamic Loads of Various Types of Machinery 55

where:
Fp,max peak force
k stiffness of the impacted structural member
Fp,ma/k = static displacement under the peak force.

Figure 3.4 leads to the following conclusions:


(1) In the range tlT l > 1, the magnification factor is dominated by the rise time.
Instantaneously applied loads induce the highest possible value 2.0, whereas
gradually applied loads lead to a lower bound on the magnification of 1. O.

(2) In the range tlT l < 0.25, the magnification Vs depends strongly on the shape of the
load-time function, but the peak displacement is largely determined through the
applied momentum. The relationship can be approximated to

(3.11)

where:
mass of the SDOF oscillator (structural member)
circular eigenfrequency of the oscillator
tp
I f F(t)dt= applied momentum
o
-
t

and for the peak displacement

(3.12)

Because of the typically very short impact duration t p, considerable load components
contribute over a wide frequency band. Structural damping does not have much effect
under these loading conditions.
More details on the load-time functions can be found e. g. in [3.7]'
56 Machine-Induced Vibrations

3.3.1 Inertial Vibrations


The effects of the vibrations do not depend qualitatively on whether they are due to
rotational, oscillatory, or impacting type of motion; in this respect, no distinction is made
in the following.
As mentioned in Section 1.2 the effects of structural vibrations can show a wide variety,
the basic distinction being drawn between effects on structures or structural members on
the one hand and those on people, installations, machinery and their products on the
other. The severity of the effects depends primarily on the frequency and the amplitude of
the vibration. Acceptance criteria for these have been compiled in Chapter 4; they mark
thresholds of the vibrations being possibly detrimental to the structure, disturbing or even
harmful to people, etc.

Effects of machine-induced vibrations on structures may include:


appearance of cracking, crumbling plaster, loosening of screws, etc.
- problems of fatigue of steel girders or steel reinforcement with consequent damage,
ultimately to the extent of collapse
loss of the load-bearing capacity (in rare cases of overstressing).

People working temporarily or permanently near to the vibration emitting machine or


the co-vibrating structural member are concerned to various degrees. The intensity may
range from barely noticeable over slightly and severely disturbing to harmful. They can
occur in three different ways:
- as mechanical effects (e. g. vibrations of floor or ceiling, see Cases No. 20 and 21)
as acoustic effects (e. g. noise from reverberating installations and pieces of equipment,
also structure-borne or air-borne sound, see Case No. 22)
optical effects (e. g. visible motion of building elements, installations or objects).

Effects on machinery and installations can induce:


- problems of material behaviour in the machine itself (deformations, strength, fatigue)
problems of material technology in the manufactured goods (e. g. excessive tolerance
due to unplanned motion of tools and installations, see Cases No. 15, 17, 18 and 20).

The amplitude of structural vibrations due to dynamic loads from machinery depends
strongly on the ratio of the natural frequencies of the structure to the dominant harmonics
of the machinery loads. It may well happen that the largest vibrations do neither occur at
the fundamental structural frequency nor at the operating frequency of the machine, but at
a higher harmonic common to both the structural and the loading frequency spectra. For
Effects of Machine-Induced Vibrations 57

instance, the 3rd harmonic of a structural member may resonate with the 5th harmonic of
the load from a weaving machine (depending on the type of machine).
It should be pointed out that, despite the distinct vertical direction in which the major
part of the dynamic load is exerted on the structure, other spatial components of the load
cannot be neglected without closer inspection. In particular, rather flexible structures may
also undergo severe horizontal vibrations under horizontal dynamic load components.

3.3.2 Structure- and Air-Borne Acoustic Waves


Vibrations of machinery and its neighbourhood are also apt to induce acoustic waves in
the structure or in air [3.11]. Air-borne acoustic waves are emitted from the machine
(mostly from those running at higher speed) in the form of direct sound, which usually is
sensed by the human ear as more or less disturbing noise (see Case No. 22).

r
r~~~m~~~~~/ / vibrations perceptible
to unpleasant
~
radiated sound well
audible to unpleasant Fig. 3.5
audibility-\
limit " Critical frequency domains
for the structure's inertial
2 5 10 20 50 100 200
vibration and the radiated,
frequency [Hz] structure-borne sound
(from [3.12])

On the other hand, vibrations may travel long distances through various transmitting
media, such as columns, walls, foundations, piping, etc., in the form of structure-borne
acoustic waves. They in turn can be radiated from the wall and ceiling surfaces of a room
with an enclosed volume of air as indirect sound (called also secondary or radiated sound)
and sensed by the ear as noise. This kind of indirect sound is often felt as particularly
unpleasant because its source is mostly not locatable. Another and rather critical factor is
the frequency content of the noise. The range of audibility is limited to about 20 Hz as
lowest frequency. As Fig. 3.5 illustrates, the audibility increases with higher frequency,
that is a lower sound level suffices for the same degree of perception. It also shows that in
the range beyond 50 Hz the sound radiated from the structure becomes the dominating
criterion for the disturbance of people, superseding the direct effect of vibration of the
machine and its neighbourhood.
58 Machine-Induced Vibrations

Moreover, acoustic vibrations may also impair the structural members and installations
through which they are transmitted. Even though the resulting alternating stresses are in
most cases of small amplitude, their very high number of cycles may still present a certain
danger in terms of acoustic fatigue.

Measures LJII,..;:;;U....lll.:ll1i- Machine-Induced '/IlIh.... <l'lhi"'.nlQ

Vibrations due to rotational and oscillatory motions are more or less similar in cause,
effects, and hence also in countermeasures; thus they are treated together. Fundamentally
different are alone vibrations due to intermittent motion of machinery with impacting
parts, which require special measures.
As for man-induced vibrations (see Chapter 2), separating the frequencies by tuning is
the most effective measure (see Appendix B 5). Some cases may need the calculation of a
forced vibration or warrant special measures. However, priority should always be given to
suppressing vibrations at their source, for instance by keeping out-of-balance forces to a
minimum, reducing impact, etc.
Frequency tuning between machine and structure or structural members is commonly
referred to as vibration isolation. In the present context of preventing load vibrations from
being transmitted to neighbouring zones, one speaks of active isolation. The opposite
action of preventing external vibrations from reaching an object is termed accordingly
passive isolation.

3.4.1 Machines with Kotatme and UsciIJlatin2 Parts


Adjusting the fundamental frequency so as not to coincide with the operating frequency
nor with a higher harmonic of the load-time function is effective in most cases of rotary or
oscillatory motion (see Cases No. 15, 21 and 22). Different measures should be chosen
according to the respective type of machine, characterized in the following by its operating
frequency. Of the three categories given below, a certain overlap between the first two
cannot be avoided.
Group 1:
Low to medium operating frequency (nb 1 7 600 r. p.m., fb = 0.02 7 10 examples
are reciprocating pumps and compressors, weaving machines, rotary presses, etc. Ma-
chines with low operating frequency (fb < 1 can be expected to produce only small
dynamic loads, the exception being bells (see Section 3.2.2).
Group 2:
Medium to high operating frequency (nb = 300 7 900 r.p.m., fb = 5 7 15 Hz); e.g. large
diesel engines, blowers, certain weaving machines, etc.
Measures Against Machine-Induced Vibrations 59

3:
operating frequency (nb > 1000 r. p.m., fb > 15 -7- 20 Hz); e. g. turbines, small diesel
engines, centrifugal separators, vibrators, etc.

Machines with predominantly rotational or oscillatory motion allow, in general, low


tuning as well as high tuning. Since the objective is to completely avoid any state of
resonance, the following parameters become important:
- the dominant natural frequencies of the structure as a whole and of the member
immediately under the machine, taking into account existing springe-damper) elements
or a base slab (called in the following the machine base, see Fig. 3.6)
- the frequencies of the dominant dynamic load components (operating frequency of the
machine and relevant higher harmonics).

(0) (b) (c)

/base

~~~;S: } mounting

" ' - - structural membir (floor slab) --..-/

Fig. 3.6 Low-tuning of machinery with rotational or oscillatory motion:


(a) direct mounting on a structural member
(b) support on springe-damper) elements
(c) mounting on a vibration-isolated base

If the fundamental frequency f 1 of the isolated system falls clearly below the operating
frequency fb of the machine, low tuning materializes. For effective high tuning, the
fundamental frequency f1 of the isolated system must be established well above the highest
frequency component with an appreciable contribution to the dynamic loading (higher
harmonics of the operating frequency). A short excursion into the theory of low and high
tuning will elucidate the difference (see also Appendix B 5).

Theoretical Background
The load a machine with quadratic or constant-load excitation exerts on its base can be
visualized as composed of a spring force and a damping force. The maximum reaction force
at the base becomes
60 Machine-Induced Vibrations

for quadratic excitation:

(3.13)

for constant-load excitation:

(3.14)

whereby Vk is in both equations the dynamic magnification factor of Eq. (3.9). The
frequency ratio W/Wl (or flf1) in this context is called the tuning ratio (or sometimes the
tuning factor).

The ratio of the maximum reaction force to the maximum load coming from the machine
gives the so-called transmissibility V R with

VR- -R-max
- (3.15)
Fo

This transmissibility V R is plotted in Fig. 3.7 for various damping ratios. Note that the
influence of damping is advantageous only up to the tuning ratio W/Wl ::::; V2, whereas
damping reduces the isolation effect for tuning ratios higher than that. Besides, the
transmissibility is also the same for passive isolation, so that Fig. 3.7 can be used for both
kinds of isolation alike.

3.0

2.5
a:
>
~ 2.0
:0
"iii
III
"EIII 1.5
C
e
<!>
1.0
~ Fig. 3.7
.e
0.5 Transmissibility for various
damping ratios as depending
0.4 0.6 0.8 1 ./2 2 3 4 5 6 7 8 on the tuning ratio between
tuning ratio flf; operating and structural
frequency
Measures Against Machine-Induced Vibrations 61

Instead of the transmissibility V R one may prefer to work with the complementary part,
which is the isolation effectiveness:

(3.16)

Under the assumption of negligibly small damping one obtains for wlwl > \/2:

(WIWI)2 2
(3.17)
(wiWI)2 - 1

The eigenfrequency WI Vk/m of the isolated system can be related through m = Gig to
the combined weight of machine and base and thus to the static deflection produced in the
springe-damper) elements [3.13]:

(3.18)

where:
f frequency of the dynamic loading
g acceleration of gravity (9.81 m/s2)
Xstat static displacement under the weight of machine and base.

The evaluation of (3.18) in Fig. 3.8 allows to determine from the known excitation
frequency (f) the support displacement (xstat ) necessary to achieve any desired level of
isolation effectiveness (l-V R), assuming that the isolators have little damping.

45
r;:;
E 40
35
>. 30
u
c
Q)
::;)
25
0"
20
~ Fig. 3.8
c 15
0
Effectiveness of vibration
~
u
10
5 isolation depending on the
)(
Q)
0
0 5 iO 15 static displacement due to
static deflection Xstat [mm] machine weight and operating
frequency
62 Machine-Induced Vibrations

Low Tuning
On the one hand, low tuning reduces the reaction forces substantially (see the trend of
the transmissibility function in Fig. 3.7), while on the other the resulting soft support yields
large displacements already under static load. Certain conditions must be fulfilled for low
tuning to be effective:
(1) The machine should be operated at a speed of more than 4 to 6 Hz.
(2) Higher structural eigenfrequencies must not coincide with the dominating harmonic of
the dynamic load.
(3) The soft base of the machine must not cause manufacturing problems.
(4) On passing the fundamental and higher natural frequencies of the base system during
start-up and shut-down of the machine, no intolerably large vibrations may occur.

Condition (1) reflects the constructional limitation on the degree of softening of the base
(Fig. 3.6) that can be achieved. The limit is often reached which a stiffness resulting in an
isolation-system frequency of ca. 1.5 to 3 this necessitates a minimum operating
frequency of ca. 4 to 6 Hz. The idea is to achieve a tuning ratio of about 2 (see Appendix
B 5). However, a tuning ratio between 2.5 and 4 is to be preferred since a larger ratio yields
smaller reaction forces.
Condition (2) prevents resonance phenomena in the higher frequencies of the base.
Condition (3) results from production requirements, as a softer support yields larger
displacements of the machine.
Condition (4) can often be ensured by precise instructions on start-up and shut-down of
the machinery, possibly in combination with braking devices in the slowing-down phase or
enhanced damping of the base.
In some cases of employing a reduced tuning ratio (owing to extremely soft spring
(-damper) elements, a large base mass, etc.), the recommended operating frequency need
not be adhered to if condition (3) is observed with particular care.
Of the three machine frequency groups mentioned at the beginning of the chapter, the
third, i. e. machines with high operating frequencies, is with respect to low tuning the most
suited. Accepting a tuning ratio near the lower limit (2.0 -;- 2.5) opens the possibility of low
tuning also to the second group of medium to high-frequency machines.

It applicable, low tuning can be put into practice in three possible ways (Fig. 3.6):

Direct mounting of the machine on the structural member (Fig. 3.6a)


As a structure or its structural members cannot be softened at will, this type of mounting
is appropriate only for high-speed machines (group 3). If the structure or structural
members were to be constructed with a softer response (2 -;- 3 Hz), the ensuing
displacement amplitudes might then become so large as to critically affect operation by
the personnel and the functioning of the machine itself.
Measures Against Machine-Induced Vibrations 63

Mounting of the machine on spring(-damper) elements (Fig. 3.6b)


For this variant the frame of the machine needs to be designed sufficiently stiff in order
to equalize the differences in motion of the individual support points. In addition, the
basic vibration behaviour of the machine must not be substantially altered (e. g. intro-
ducing new torsional vibrations). Often the machine manufacturer himself will design
for this solution, because it has the advantage that the springe-damper) elements could
easily be modified or exchanged should the isolation prove ineffective.
The springe-damper) elements differ with the machine characteristics (its dimensions,
mass, genuine stiffness, operating frequency, etc. ). The different types on the market
comprise steel springs, elastomeric or rubber pads, and air suspension.
Steel springs can be used in the frequency range between 1 and 8 Hz. One distinguishes
between helical, cup, and laminated springs. The latter two types are in principle one-
dimensional elements, whereas helical springs may be used in spatial isolation and are
hence often preferred for relatively low-frequency excitation due to unbalanced rotating
masses. All springs are attached either on one side to the machine frame alone (with
guard against self-movements) or on both sides to machine and building structure.
Elastomeric and rubber pads have a range of applicability from about 5 to 20 Hz.
Mounting can be effected in three ways:
floating support:
The machine is continuously bedded on elastomeric or rubber mats. (This type of
support also uses cork or felt above frequencies of at minimum 10 Hz with cork and 20
Hz with felt.)
- one-sided fixation at the machine:
The springe-damper) elements are attached locally to the machine allowing for
construction tolerances and unevenness of the base.
- two-sided fixation at both machine and floor:
Attaching the springe-damper) elements on both sides still allows adjustment in
height, but is additionally suitable for resisting some limited horizontal load compo-
nents.
Air suspension is employed for frequencies from 0.5 to 5 Hz. This is the range of very
low frequencies in which steel springs would yield too large displacements .

Mounting on a vibration-isolated base (Fig. 3.6c)


This measure is taken into consideration for machines with
- very large dynamic loads
- insufficient stiffness for isolation according to Fig. 3.6b
- very small mass which would demand extremely soft springs for low tuning
very eccentric centre of mass.
64 Machine-Induced Vibrations

Due to the additional mass of the base one can employ stiffer springe-damper)
elements than one could in direct support, thereby reducing the vibration amplitudes.
The stabilizing mass should have at least the same weight as the mass of the machine
alone.
Mounting on a machine base is very advantageous for isolation but is relatively costly.
It may also interfere with the production process (impeded access, problematic level
difference for supply of raw materials and removal of end products, etc., even when
several manufacturing machines are mounted on the same base). Sometimes, however,
an over-thick floor slab or foundation enables the stabilizing mass to be accommodated
in a deep cavity thus ensuring free access. Consideration must be given to a sufficient
stiffness of the base itself.

High Tuning
In general, high tuning is only feasible with a rigid connection to the structure and is
chosen especially in those cases in which low tuning is impractical. The requirements for
high tuning are:
(1) The frequency of the highest relevant load component does not exceed about 20 Hz.
(2) Production demands that machine vibrations be kept small.

The first requirement results from the fact that a structure or structural member cannot
be designed for high stiffness at will within a given expenditure. Since the computational
prediction of the structural frequency contains some uncertainty, the tuning ratio is
augmented by an additional safety factor. The lower bound to the fundamental frequency
of the structure or the respective structural member thus becomes

(3.19)

with:
fl required fundamental structural frequency
fb operating frequency of the machine (r. p.m.l60)
n max order of the highest relevant harmonic of the loading
WI/W = reciprocal of the tuning ratio (see Appendix B 5)
Sf safety factor (usually taken as 1.1 to 1.2).

For calculating the expected frequency of stiffly designed structures one usually cannot
neglect the flexibility of the ground, which may under some circumstances substantially
reduce the combined frequency of the soil-structure system [3.4], [3.5].
Figure 3.7 demonstrates that the forces under high tuning always have a factor V R > 1
with respect to rigid base (equivalent to the static case). High tuning is the preferred
measure for all machines of the first (lowest) frequency group (see Case No. 15).
Measures Against Machine-Induced Vibrations 65

Calculation of a Forced Vibration


If the success of frequency tuning is somewhat doubtful (e. g. because of a wide
spectrum of the dynamic loading, etc.), the calculation of the structural
under forced vibration is to be recommended. Such a calculation yields informa-
on the maximum expected vibration amplitudes, which then could be compared with
acceptance criteria given in Chapter 4. However, such a calculation is itself subject to
considerable uncertainty, especially as far as the actual dynamic loads from the machine
and realistic damping properties are concerned. The procedure for such a computation is
outlined in the Appendices B 1, B 4 and B 5. Appendix B 8 suggests appropriate damping
coefficients.

(c) Measures
A further option for the reduction of vibration is just as for man-induced vibrations-
the provision of tuned vibration absorbers. They are chosen mainly for the reduction of
vibrations in existing structures with little damping or relatively small mass, when the
dynamic loads cannot be reduced and when alternative measures (such as stiffening the
structure) would lead to unreasonable costs. One should realize, however, that any tuned
vibration absorber is a mechanical system vibrating at a considerable amplitude and that it
needs proper maintenance. The theoretical basis and the various construction types of
absorbers are described in Appendix B 7.

3.4.2 Machines with Impa~~tlIl~ Parts

The following countermeasures apply to machinery with an intermittent kind o(motion,


giving rise to either periodic or single impulses (transient load, see Appendix B 6). The
solution of a practical problem is largely determined by the load-time function (see Cases
No. 16, 17, 19 and 20). As discussed previously in Section 3.2.3, the important parameters
are the duration of the impulse, the momentum and the peak force from the loading.

(a) Frequency
Vibrations of machines with intermittent motion are best counteracted by low tuning,
i. e. by providing low stiffness of supports. High tuning, in contrast, is most likely not an
alternative as the Fourier spectrum of the loading is typically rather wide, i. e. relevant load
components are exerted over a wide frequency band.
Low tuning is applicable under the following conditions:
(1) The resulting displacement amplitudes of relatively large magnitude must not interfere
with production requirements.
(2) The loading on the structure must be accommodated without difficulty.
66 Machine-Induced Vibrations

Condition (1) is frequently not met for machines in the metal-working or plastics-
producing industry because of the usual production requirement of a very stiff mounting.
This problem can be mitigated considerably by using a stabilizing mass and thereby
rendering low tuning feasible. Low tuning finds a common application in machine hammer
design, often in combination with a stabilizing mass.
Condition (2) can normally be achieved by an appropriate design of the structure or the
respective structural member.

The above conditions being met, low tuning of machines with predominantly intermit-
tent motion can proceed along the same lines as discussed previously in Section 3.4.1 for
rotating or oscillating motion.
Of the feasible solutions depicted in Fig. 3.6 solution (c), suggesting mounting on an
isolated base, is the best for low tuning. The resulting displacements can thereby be kept
substantially smaller than in solution (b), using only springe-damper) elements as supports.
Direct mounting of the machine on a relatively soft structure or member is only feasible
for small impacting loads. Attention has then to be given to the possibility of substantial
excitation of higher structural frequencies.

Calculation of a Forced Vibration


If the success of frequency tuning cannot be guaranteed, resort may again be made to the
approximate computation of a forced vibration. For impacting loads and direct mounting
(Fig. 3.6a), an appropriate method is outlined in Appendix B 6.
A periodic loading can be decomposed by Fourier analysis into its constitutive har-
monics, of which subsequently a small number are selected as forcing functions.
If the loading is of a transient nature and exhibits strongly the features of Section 3.2.3,
then the computation of a forced vibration must proceed in two phases (see Appendix B 6):
Phase 1: Computation for the effective duration of the loading (impact).
Phase 2: Computation of the subsequent free vibration (after the loading has ceased), for
which the current values of the state variables of motion at the end of phase 1
serve as initial conditions for evaluating the integration constants of phase 2.
67

In civil engineering practice, measured or calculated vibration magnitudes (e. g. acceler-


velocities, displacements) need to be evaluated as to whether or not the vibration
can be tolerated. To this end, the following acceptance criteria are introduced.
""".(rnI1TllfH-''-' exceeding these values or falling short of them do not necessarily indicate an

in3ldnl1S~Slble state of vibration. In only a few cases the acceptance criteria represent more
less agreed bounds; much more often they indicate a practicable order of magnitude
a certain scatter. Nevertheless, these criteria may be used as bounds of admissibility,
for instance if made a part of the contract or some other form of obligation.
Criteria for vibrations of buildings and pedestrian structures may be stipulated with
respect to the following effects:
- overstressing of structural members (deformation, fatigue, strength)
- physiological effects on people (mechanical, acoustic, optical)
- impediment of production processes (problems of product tolerances, etc.) as well as
overstressing of machinery (deformation, fatigue, strength).

How to derive the acceptance criteria, is quite a complex problem. Tolerable values for
vibration effects on machinery and installations and - still more so - for the physiological
effects on people are most difficult to agree upon, implying a considerable range of
discretion. Somewhat more reliable are the values of tolerable stressing of structures,
because their vibrations are relatively easy to measure and to assess.
Vibration bounds may be given as physical quantities such as
- displacement amplitude
- velocity amplitude
- acceleration amplitude
or as
- empirically derived quantities (e. g. KB intensity [4.1]).

The above categorization of the various vibration effects leads to the following division
into
- structural criteria
- physiological criteria
- production-quality criteria.
68 Acceptance Criteria

A concise compilation of criteria as this does not really permit their direct application.
Rather more, the reader should routinely consult the complete edition of the respective
regulation or reference literature, in the choice of which he will- hopefully - be assisted by
this survey.
The last paragraph of this chapter recommends global criteria derived for the most
critical vibration effect, depending on types of buildings and their use. They may suffice for
feasibility studies, in cases with large uncertainties regarding the dynamic loads or the
structural properties, or whenever no specific regulation is to be adhered to.

Vibrations induced by man, machinery, traffic, construction works, etc., may cause
deformations and smaller or larger forms of distress in buildings, their structural members
and particularly - nonstructural elements. Forms of distress may be
- cracking of walls and slabs
- aggravation of existing cracking in structural members and nonstructural elements
falling down of equipment or cladding thereby endangering occupants.
Continuous vibration, however, can also lead to problems of fatigue and overstress in
principal load-bearing members.

Acceptance criteria must take account of the following parameters (among others):
type and quality of the building material (especially its ductility)
- type of construction
properties of the building foundation
- main dimensions of the principal load-bearing members
- age of the building
- duration of the vibration effects
- characterization of vibration (frequency, etc.).

Figure 4.1 shows the amount of expected structural damage to depend on various
parameters. Although the acceptance criteria are taken to be independent of frequency,
the quantity most suitable as indicator differs over the frequency range: While the bound
should be on vibration velocities for low frequencies, it should be on accelerations for high
frequencies. This notwithstanding, most regulations define their bounds in terms of
velocity.
The following gives a concise survey of various regulations, codes of practice and some
reference literature. Another, more extensive survey on acceptance criteria used in
different countries is given in [4.3] with emphasis on vibrations of buildings due to nearby
blasting operations.
Structural Criteria 69

z:-
'0
o
Q)
>
~
oQ) c:
o Fig. 4.1
0.Qi~--~~+~-~~""""---~~
'';::
0. e
Q)
Response spectrum for vibra-
Q)
u tion effects on people and
u
o structure (from [4.21]);
1 in. = 25.4 mm
0.1 o Origin of data:
0, y
!? U.S. Bureau of Mines
frequency [Hz] Rausch [4.25]
Reiher/Meister [4.23]

4.2.1 Standard DIN Part 3 [4.4]


The third part of Standard DIN 4150 treats effects on buildings and structural members
due to an internal or external source of vibration. For their assessment, the vibration
velocities or - if necessary - the stresses due to dynamic loads are to be compared with the
given criteria. With predominant use of (mostly measured) vibration velocities, different
criteria are used for
- short-term structural vibrations (transient)
- steady-state structural vibrations
steady-state vibrations, particularly of floor slabs.

Transient vibrations, as excited e. g. by blasting operations, pile driving, etc., are to be


limited in terms of maximum foundation velocities, the values of which depend on the
accepted degree of damage, Fig. 4.2: 20 -;- 50 mm/s (from f::::; 10 Hz to f 100 to avoid
severe damage, 5 -;- 20 mm/s to avoid slight damage, and 3 -;- 5 mm/s for particularly
sensitive environments. For steady-state structural vibrations, floor slabs in particular,
10 mm/s is considered admissible, even though it does not entirely preclude slight damage
such as cracking.
70 Acceptance Criteria

.....--.
VI

"E 50
~
45
z:.
'u 40
~
(I.)
> 35
~
0
(I.) 30
Q.

25

20

15

10

5 Fig. 4.2
3
0 Bounds on the foundation
0 10 20 30 40 50 60 70 80 90 100
vibration velocities for
frequency [Hz] building categories of
DIN 4150 Part 3 (from [4.4])

4.2.2 Standard SN 640312 [4.6]


The Association of Swiss Highway Engineers distinguishes in their Standard SN 640312
four different categories of buildings, mainly according to the type of construction, Tab.
4.1. The acceptance criteria are again peak vibration velocities, which are quoted in Tab.
4.2. Two different groups of vibration sources are considered: The first group comprises
machinery, traffic and construction equipment, the second group blasting operations
assumed to occur less frequently and thus permitted for higher bounds. The criteria are
defined for two frequency ranges and are based on the spatial vector of the vibration
velocity.

4.2.3 Directive KDT 046/72 [4.7]


The Directive issued by the Chamber of Technology of the German Democratic Repub-
lic also distinguishes four categories of buildings and defines admissible vibration velocities
in function of frequency, Tab. 4.3. Bounds on vibrations due to blasting for the same
categories are presented in Fig. 4.3. The Directive is discussed e. g. in [4.8].

4.2.4 Draft ISO/DIS 4866 [4.9]


The draft of Standard ISO/DIS 4866 deals primarily with types and methods of vibration
measurement on structures and does not contain any acceptance criteria. Interesting in the
present context, however, is the distinguishing of 14 different building categories accord-
Structural Criteria 71

structural definition
category
I reinforced-concrete and steel structures (without plaster) such as indus-
trial buildings, bridges, masts, retaining walls, unburied pipelines;
underground structures such as caverns, tunnels, galleries, lined and
unlined
II buildings with concrete floors and basement walls, above-grade walls of
concrete, brick or ashlar masonry; ashlar retaining walls, buried pipe-
lines;
underground structures such as caverns, tunnels, galleries, with masonry
lining
III buildings with concrete basement floors and walls, above-grade masonry
walls, timber joist floors
IV buildings which are particularly vulnerable or worth protecting

Tab. 4.1 Structural categories according to SN 640312 (from [4.6])

structural source M source S


category f [Hz] V rnax [mm/s] f [Hz] V rnax [mm/s]
I 10 -7- 30 12 10 -7- 60 30
30 -7- 60 12 -7- 18* 60 -7- 90 30 -7- 40**
II 10 -7- 30 8 10 -7- 60 18
30 -7- 60 8 -7- 12* 60 -7- 90 18 -7- 25**
III 10 -7- 30 5 10 -7- 60 12
30 -7- 60 5 -7- 8* 60 -7- 90 12 -7- 18**
IV 10 -7- 30 3 10 -7- 60 8
30 -7- 60 3 -7- 5* 60 -7- 90 8 -7- 12**
source M: machinery, traffic, construction works - (*) the lower value applies to
30 Hz, the upper to 60 Hz, with interpolation in between.
source S: blasting operations - (**) the lower value applies to 60 the upper to
90 Hz, with interpolation in between.
Tab. 4.2 Acceptance criteria of SN 640312 for the structural categories
of Tab. 4.1 (from [4.6])
72 Acceptance Criteria

ing to type of construction, foundation, ground condition, and importance of the building.
The usefulness of such a detailed categorization in view of the many uncertainties remains
questionable.

building category vz,adm [mm/s]


I historical monuments 2

II half-timbered houses 5
III wall construction (e.g. buildings of 10
slab walls, blocks, masonry)
IV framed construction (e.g. buildings of steel, 30
reinforced concrete, timber)
Tab. 4.3 Building categories and admissible velocities in KDT 046172 (from [4.8])

I
J:'
'0 100
o
Q)
>
Q)
:0
'iii
b' iI iog COfe'O,m:::
.~ 30
TIl. I)
'"C
o V I
I I

II
J1I
10

11
5

2 I V
Fig. 4.3
Bounds on vibration velocities
2 10 30 ~oo
due to blasting in function of
frequency [Hz]
building category (Tab. 4.3)
in KDT 046172 (from [4.8])
Physiological Criteria 73

In principle, the human sensitivity to mechanical vibrations is very subtle. As an


the human body notices vibration displacement amplitudes of only 0.001 mm,
finger-tips can even detect amplitudes 20 times smaller. However, the human reaction
a given vibration depends very much on circumstances. Personal discomfort is realized
a different degree if, for instance, one sits at an office desk, operates a machine, or drives
a car. Obviously the person's attitude to the vibration source matters - whether it is his own
IJ~L'''''~'~.L activity or external- as well as the degree of accustoming.

U"1'",n,pTprc influencing the human sensitivity are the following:


- position of the affected person (standing, sitting, lying)
- direction of incidence with respect to the human spine
- activity of the affected person (resting. walking, running)
- community (existence of fellow-sufferers)
- age
- sex
- frequency of occurrence and time of day
- duration of decay (damping).

The intensity of perception depends, of course, also on the objective physical vibration
parameters
- displacement amplitude
velocity amplitude
acceleration amplitude
duration of effect (exposure time)
vibration frequency.
This is illustrated in the following figures.
Figure 4.4 shows the human sensitivity spectra of displacement amplitudes of different
frequency (from [4.10]). According to Fig. 4.5, also the acceptance criteria for velocity
amplitudes depend on frequency (from [4.11]). Table 4.4 gives acceptance criteria for
effects on people in function of velocity and acceleration amplitudes, which apply to
harmonic vibrations with displacement amplitudes of less than 1 mm ([3.2] after [4.12]).

Different codes of practice and numerous publications have attempted to find the most
realistic criteria for physiological vibration effects. Concepts and some figures of several
codes are cited and discussed in the following.
74 Acceptance Criteria

0.0500 +----k--I--I-'\l-H+H---II--+---1r-t-t-H+1

........... 0.0100
~
-c
Q)

E
0.0050

Q)
(.)
c
Q.
III
:c
0.0010

0.0005

Fig. 4.4
0.0001 +----L-....-l---'-j---.l--l.....l---l..1r--'--.l---Jr-t--'-'-'...J...j Frequency dependence of the
1 5 10 50 100
human perception of vibration
frequency [Hz]
displacements (from [4.10]);
1 in. = 25.4 mm

vibration effects on people frequencies 1 -;- 10 Hz frequencies 10 -;- 100 Hz


amax [mm/s2] V max [mm/s]

imperceptible 10 0.16
just perceptible 40 0.64
clearly perceptible 125 2.0
annoying 400 6.4
unpleasant, 1000 16.0
painful if lasting
harmful > 1000 > 16.0
Tab.4.4 Acceptance criteria for physiological vibration effects
(from [3.2] after [4.12])
Physiological Criteria 75

m intolerable

1I unpleasant

I threshold of
perception

10-3 -+--.----,-_,---,-_--,----,- ___ Fig. 4.5


1 2 5 10 20 50 100
Frequency dependence of the
frequency [Hz]
human perception of vibration
accelerations (from [4.11])

4.3.1 Standard DIN Part 2 (1975) [4.1]


The second part of Standard DIN 4150 deals with the effects of vibrations from mostly
external sources on people in residental buildings. The frequencies considered range from
1 to 80 Hz. Measured vibration quantities (displacement, velocity, acceleration) serve as
input for an empirically derived intensity of perception, called KB value:

KB d. 0.82 (4.1)
VI + 0.0322
where:
d displacement amplitude [mm]
f vibration frequency [Hz].

The KB value can alternatively be formulated in terms of velocity (v) or acceleration (a)
of the vibration, which if the vibration is harmonic - are interrelated by
v a
d (4.2)
2Jtf 4Jt2.2
76 Acceptance Criteria

The thus calculated KB value of the examined vibration is to be compared with the
reference value in the code according to
- use of the building
frequency of occurrence
- duration of effects
time of day of occurrence.

Figure 4.6 depicts one of the double-logarithmic diagrams in [4.1]. Table 4.5 gives
acceptance criteria for assessing vibrations in apartments and similar environments.

building zone acceptable KB intensity


(actual utilization and development of time
the estate within radius of vibration continuous or infrequent
emission) repeatedly
purely residential, housing estate, day 0.2 (0.15*) 4
holiday resort
night 0.15 (0.1 * ) 0.15
village and small business, day 0.3 (0.2* ) 8
town-centres
night 0.2 0.2
business and trade (incl. offices) day 0.4 12
night 0.3 0.3
industrial day 0.6 12
night 0.4 0.4
exceptional areas day 0.1 -:- 0.6 4 -:- 12
(acc. to residential content)
night 0.1 -:- 0.4 0.15 -:- 0.4
(*) Values in brackets should be complied with if buildings are exited horizontally with
a frequency below ~ 5 Hz.

Tab. 4.5 Acceptable KB intensities for residential buildings (from [4.1])


Physiological Criteria 77

" "'" " "-


5
"!'.. "-
""",-
"""~
5
" ""'ro-.
3

2 3
""l\.. " '" "'- """'" ~ ~ ro-."",
"""1-- .... KB = 25.6-1-

~ ~ I II
2 KB = 12.~-1-
,,~ "~

,.........
8
10 i~ I
"~.... " ................... KB = 6.4
II
~
l/)
.......
E 8 "'" "-
.....
~ 5
" ""I'\.. . . . r-....
........ i-- ... I

" " "",,- ""-


>-
5 KB = 3.2-~

~
u
0 3
Q)
>
CI> 2 3 ""- l ' I

"'' -~
.::: I-..r-.
t> """"--- KB = 1.6

-.e
I
2
CI>
"'- ...... I I
~ 1\. ""'r-. """r--..,..
8
, k ~ ,..,.
KB = 0.8

5
-
....
"-
"'-
......L
""-.
-;...

,..., ... -~ I-- KB=0.56


I I
b.4-~
I-

,
"""'I-... I-- KB' =
5 - ........
"'~ r---r--, I
3 i !'----- KB = 0.28
.....1'-.
3 !'.. . ......1'-- I I I
KB = 0.2-~
2
.... ~ ......1'--. ..... 1 1
1'-- ...
2
~ F-- KB = 10 .14

~ i'- .... ~ ........ I I I


KB = 0.1 -I-
"'-c... I-
8 -f----
1 I I I
2 3 5 B 10 20 30 50 80 100

frequency [HZ]

Fig. 4.6 Iso-perception curves in function of frequency (from [4.1])


78 Acceptance Criteria

4.3.2 Standard ISO 2631 [4.13]


The Standard ISO 2631 covers all effects on people from periodic or transient vibrations
in the frequency range of 1 to 80 Hz. Three different levels of human discomfort are
distinguished:
- The reduced comfort boundary applies to tolerable disturbance during activities like
eating, reading or writing.
- The fatigue-decreased proficiency boundary describes the level at which recurrent
vibrations cause fatigue to personnel with consequent reduction of efficiency; this occurs
at about three times the first discomfort level.
The exposure limit defines the maximum vibration tolerable with respect to human
health and safety and is about six times higher than the first level.

...---. 20

x~
N
til
"-
15 r--
E 12.5 r--
'---'
10
'+-
r
~
Q)
8.0 /
N .;
C 6.3
I / -'
c: r--... /'. ~
.~
0+-

ecv
5.0
4.0
~
~~

r--... ......... ~I-"" , /


.-/"
v/
.-'/'
/
/
-,~
-,v
..V
/

",
3.15 1 min.
cv
(.) ~ ....... ~ / -,V V I
./
(.)
2.5
r""- ........ ..... ....... l6-tnin.
/ V V ./
'" , ..... ffik=F(
1........
/
C

,
2.0
........ 2 ~
~ V 1/ /'
1.6

1.25
~
-I'-..... V
~

1//
/ Z r
V -'
V -'
,
~

--
1 h. ./
1.0
......... i"'-...

~V
0.8
i'..
.......
V i/ V )V
V V
0.63

0.50
........ ........ "
......... .........
.......
........ 2.5 h.
4 h.
/
V V -' V
V
~

0.40
~ /' ~
........ ""-
~X
0.315 .....
......... "- 1/
~ .......... 8 h.
0.25 ........ ~

~
0.20
V
.........
0.16 II'
/
"
0.125
0.1
10.016 0.4 0.5 0.63 0.8 1.0 1.25 1.6 2.0 2.5 3.15 4.0 5.0 6.3 8.0 10 12.5 1.6 20 25 31.5 40 50 63 80

frequency [Hz]

Fig. 4.8 Bounds on z-acceleration for fatigue-decreased proficiency


depending on exposure duration (from [4.13]);
the exposure limit is obtained by multiplying by 2,
the reduced-comfort boundary by dividing by 3.15
Physiological Criteria 79

z
t z

t
x x

Fig. 4.7 Reference coordinate system for vibration effects on man (from [4.13])

20
N
Vl 15
"-
.... 12.5
10
8.0

6.3

OJ 5.0
>< 4.0
c
3.15
c:
.~ 2.5

~
Q)
2.0
Q) 1.6
u
u 1.25
c
1.0

0.8
0.63

0.50
0.40
0.315

0.25
0.20

0.16
0.125
0.10
0.016 0.4 0.5 0.63 0.8 1.0 1.25 1.6 2.0 2.5 3.15 4.0 5.0 6.3 8.0 10 12.5 1.6 20 25 31.5 40 50 63 80

frequency [HZ]

Fig. 4.9 Bounds on x- and y-accelerations for fatigue-decreased proficiency


depending on exposure duration (from [4.13]);
other bounds are obtained by the same factors as in Fig. 4.8
80 Acceptance Criteria

The bounds given in tables and diagrams depend on the direction of incidence to the
human body, using a reference coordinate system with the z-axis in the direction of the
human spine, Fig. 4.7.
Criteria are formulated in terms of an effective acceleration which is defined as the root-
mean-square value over the exposure time T:
j

1 T
, - J a (t)dt
2
(4.3)
To

Figure 4.8 gives the bounds on the effective acceleration in the spine direction, Fig. 4.9
those on the effective accelerations perpendicular to the spine. The values represent the
fatigue-decreased proficiency boundary, from which the values for the two other levels
can be derived by division or multiplication with the aforementioned factors.

4.3.3 Directive VDI 2057 [4.14]


The Directive VDI 2057 comprises four Parts (Folios), of which only Part 2 has been
passed whereas the others exist in draft form. Particularly Parts 1 and 2 are similar to ISO
2631 [4.13], so that the reader is referred to the outline of the latter above. One main
difference lies in the definition of the vibration intensity: The effective accelerations of ISO
2631 are already weighted with frequency and direction of incidence (see Fig. 4.7) to derive
direction-dependent KX, KYand KZ values. Given for various durations of vibration
effects, these bounds can now be applied irrespective of frequency and represent the same
discomfort levels as in ISO 2631.

4.3.4 BRE 278 [4.15]


The British Building Research Establishment defined K intensities in borrowing the KB
levels of perception from DIN 4150 part 2 [4.1]. Table 4.6 attributes the different K values
to the respective perception intensities, Tab. 4.7 gives the acceptable K intensities for
different exposure times and categories of buildings.

4.3.5 British Standard BS 5400, Part 2 (1978) [2.25]


Part 2 of the British Standard BS 5400 limits for pedestrian bridges the acceleration
amplitude for matched-frequency loading by a notional pedestrian. The upper bound is
specified as 0.5Vf;" [m/s 2], amounting to 0.7 m/s 2 at 2 Hz and to 1.1 m/s2 at 5 Hz.
Physiological Criteria 81

human perception K value commentary


not felt . < 0.1 Values are applicable to both horizontal and
threshold of perception 0.1 vertical vibrations.
barely noticeable 0.25 K values 25 and 63 of DIN 4150 were here not
noticeable 0.63 adopted as their effects on people can hardly
easily noticeable 1.6 be distinguished.
strongly detectable 4
very strongly detectable 10
Tab. 4.6 Relation between K values and human perception of motion (from [4.15])

building time acceptable K intensity


category
continuous repeatedly occasionally
hospitals and day 2.5
nursing homes 0.1 0.1
night 0.1
residential day 0.2 (0.1*) 4
0.1
night 0.1 0.1
city residential day 0.3 (0.15*) 0.63 (0.3*) 8
and business
night 0.1 0.1 0.1
industrial day 0.63 (0.3*) 0.8 (0.4*) 12
night 0.63 (0.3*) 0.8 (0.4*) 12
(*) Values in brackets apply to cases where the frequency of vibration is below 15 Hz.
Tab. 4.7 K intensity in function of building category and exposure time (from [4.15])

4.3.6 British Standard BS 6472 [4.25]


The British Standard BS 6472 serves to assess vibration effects on people in workshops,
offices, residential buildings and particularly sensitive environments such as operating
theatres. Similar to DIN 4150 [4.1] it categorizes according to vibration frequency and
exposure time and defines criteria either in velocity or acceleration, depending on the
frequency range. Accelerations are given as root-mean-square effective values (confer
Eq. (4.3)), whereas the velocities are specified as peak values.
82 Acceptance Criteria

4.3.7 NBC ,--,a.ll.llau,a, l:ommeJl1talry A [2.16]


The National Building Code of Canada gives in the pertinent Commentary A frequency-
independent acceptance criteria for accelerations in the following environments:
- dancing and dining (close to a dance floor) 2%g
- concert and sport events 5%g
- fitness classes 5%g

4.3.8 j{e~:ulalt10n SBA 123 [2.13]


The Regulation SBA 123 [2.13] of the State Construction Supervision Board of the
German Democratic Republic requires an explicit check on liveliness for pedestrian
bridges, stair cases, and similar structures. Essentially, the Regulation defines a physiolog-
ical stress value K, both for vertical and horizontal (longitudinal or lateral) vibration:

K = x stat f2 (4.4)

where:
xstat = displacement of a characteristic point under equivalent static load [mm]
f excitation frequency [Hz].

The upper bound for acceptance is K adm = 10.

4.3.9 Criteria from the Literature


The available literature contains several suggested acceptance criteria for various types
of structure, especially pedestrian structures. As the vibration effects on the structure (e. g.
fatigue) are virtually negligible, the objective is here too a limit of discomfort to people.
Criteria for pedestrian structures are always to be used with care. It makes a large
difference whether the magnitude of vibration to be assessed
- occurs frequently, rarely or exceptionally
- has been actually measured (on an existing structure)
- has been computed from more or less vague data, and then whether it is a mean expected
value or a conservative upper bound.

Acceptance criteria should pay regard to circumstances, such as the user-accepted or


design level of comfort, etc.
Production-Quality Criteria 83

generally used indicator is the maximum acceleration:


In the opinion of [2.2], 10% g is a threshold value above which pedestrians feel insecure
and the structural serviceability is impaired.
Report [2.5] advocates for footbridges that the acceleration bound stipulated by BS 5400
[2.25] be relaxed to 1.0v7;" [m/s2], provided f 1 is outside the critical range of 1.7 -:- 2.2
Hz. This criterion amounts to a bound of 13% gat 1.7 Hz and 15% gat 2.2 Hz.
Reference [2.6] suggests the peak vibration velocity as criterion and points to the
different intensity of perception during walking and standing: Vibration frequencies
above ca. 3 Hz are less bearable when standing than when walking, while it is the
opposite for slower vibrations. A walking pedestrian is said to sense as unpleasant a
velocity of 24 mm/s. On the basis of a harmonic vibration, this criterion can be trans-
formed to a bound on the displacement or the acceleration amplitude; it then corre-
sponds to 3% gat 2 Hz and 7.5% gat 5 i. e. much smaller values than in the previous
references.

For industrial or scientific work, bounds to vibration effects need to be formulated with
respect to the following problems:
- problem of production technology apparent in the manufactured goods (e. g. problems
of tolerance on lathes, milling machines, weaving machines, extrusion presses, etc.)
- diminished performance of apparatus (e. g. electron microscope, computer, etc.)
- problems of material technology of the machine itself (e. g. wear on shaft bearings,
excessive deformation, fatigue and strength limits of machinery parts or at supports,
etc. ).

Universally applicable criteria cannot be given but need to be specified individually for
types of machinery. Therefore, codes of practice and literature contain either very general
statements or values of rather limited applicability.
General statements try mostly to group various types of machinery in different sensitiv-
ity classes, such as the ones in Tab. 4.8. Depending on the frequency range, either the peak
vibration velocity or the peak acceleration is used as critical quantity, Tab. 4.9.
84 Acceptance Criteria

apparatus machinery and equipment


category
I optical instruments, such as microscopes, interferometer, opti-
meter, etc.; mechanical measuring instruments in the micron-
range, apparatus for precision scale calibration; finishing of opti-
cal lenses; precision cutters; rotor-balancing machines and other
heavy precision machinery; machine control stations
II machinery for grinding of ball bearings, cogwheels, razor blades,
etc.; coordinate grinding machines, milling and turning machin-
ery to a precision of some hundreths mm
III metal-working machinery for turning, cutting, drilling, milling,
etc., to usual precision; spinning, weaving and sewing machin-
ery; printing presses, etc.
IV rotary machines such as blowers, centrifugal separators, electric
engines, etc.; stamping machines and presses in light metal-
working industry; precision drilling machines; vibratory ma-
chines such as vibrators, jarring plates, riddlers, strewing ma-
chines, etc.
Tab. 4.8 Categories of apparatus sensitivity (from [3.2])

apparatus sensitivity to harmonic frequ.l..;- 10Hz frequ.lO..;- 100Hz


category vibrations amax [mm/s 2] vmax[mm/s]
I highly sensitive 6.3 0.1
II normally sensitive 63 1
III little sensitive 250 4
IV insensitive >250 >4
Tab. 4.9 Acceptance criteria for the categories defined in Tab. 4.8 (from [3.2])

More detailed data are to be found in the following standards and recommendations:
- ISO 2372, 1974 [4.16]
- ISO 2373, 1974 [4.17]
- VDI 2056, 1964 [4.18]
- VDI 2063, 1982 [4.19]
- ISA-Transactions, 1964 [4.20].
References [4.21] and more so - [4.22] quote further national codes and recommenda-
tions where acceptance criteria with regard to machinery can be found. Specific data on
individual types of machinery are given e. g. in [4.24] and [4.2].
Suggested Overall Acceptance Levels 85

lc(~eDtaIlce Levels

As is apparent from what has been said thus far, the assessment or the design limitation
vibrations has to observe bounds to the (simultaneous) effects on the structure and its
members, on people, and sometimes also on machinery and equipment.
lJeoenOlmg on the type of structure and its use, the three objectives may yield rather
.rl;+,rt:>r,:>-nt bounds, let alone the discrepancies in recommendations of different codes and

publications for the same type.

Therefore, it may be helpful in some cases - and particularly for provisional studies - to
operate with crude but simple global criteria. They can be quite appropriate in the absence
of exact data on the dynamic loads or in combination with relatively rough estimates on the
structural properties, necessary to calculate a forced vibration. In the light of such uncer-
tainties, frequency-dependent values and more sophisticated definitions (e. g. effective
values) are hardly practicable. Instead, of the three objectives of vibration limitation, as
mentioned above, only the most critical one is selected and observed. The resulting over-
all bounds may also be used sensibly in cases of not designing for compliance with a specific
regulation.
With this in mind, the authors undertook to derive from their own experience and
knowledge of the literature criteria for various environments, Tab. 4.10. The comments
indicate the limitations of their applicability.

structure acceptance level comments cases no.


pedestrian a:s::5 10% g the lower value does normally 1,2,3,4
structures not produce discomfort
office buildings a:S:: 2% g DIN 4150 and BS 6472 may 4,6
yield quite different values
gymnasia a:S:: 5 -;- 10% g the higher value only if 7,8,9
(sports halls) acoustic effect small
only participants on or near
the vibrating floor
dancing and a:S:: 5 -;- 10% g same as for gymnasia 10
concert halls
factory floors v:s:: ~ 10 mm/s e.g. for conventional loom; 15
high-quality production
needs much stricter bounds
Tab.4.10 Overall acceptance levels for various types of environments
87

:DD~~na]IX A

In the following, several practical cases are described where vibrations induced by man
machines were observed in structures. Most cases involve problems with interesting
teatuI'es, cases'in which there had been complaints and for which in part improvement
have been carried out. For the sake of the practical lessons they teach, these
are described in a practice-oriented manner, and more abstract aspects are deliber-
left out.
Some cases stem from the authors' own expertise, others are cited from the literature or
made available by colleagues of the engineering community with their permission for
IJU'JU.~'UUVH. Their courtesy is specially acknowledged. The authors would also be grateful
their readers' reporting further interesting cases, because they feel that in learning from
designs which were found inadequate in hindsight are often the most valuable.

1: f104)UJlrUI!!e Hz

Designed as a simply supported beam with a 40 m span, the bridge vibrated vertically
with clearly perceptible amplitude when used under normal conditions (pedestrian walk-
[ALI]. In order to clarify the nature and intensity of the vibrations, tests and
measurements were made.
Figure ALIa gives an excerpt of a typical plot of the vertical vibration velocity.
Characteristic is the surging effect which was observed under normal walking conditions as
well as under deliberate excitation by three persons skipping continuously in midspan.
Figure A1.lb shows a similar excerpt of a decay measurement following a single impulse
due to the synchronous jumping of three persons.
As results of two different readings, the peak displacement amplitude for a rate of ca. 29
people crossing per minute was 4.6 mm, corresponding to an acceleration of 6.7% g. The
damping ratio was evaluated from three decay measurements to 2.3%, 1.8% and 2.4%,
i. e. an average of 2.2% of critical damping. The three persons skipping produced a
displacement amplitude of 9.4 mm, more than double the one at crossing of 29 persons/
min.
88 Case Reports

(0)

z:. IIl-1ifl,IHI-II'IIII'iI'
'(3 IHtllltltflrWH
o 1I1HllHHf.flilff

~rlJlt.

( b)

time

(c) 20

10

:::::: -10
co
~
- 20
;;
- 30
h A

40
I ~rl ~l
o 2 4 6 8 10

frequency [Hz]

Fig. ALI Footbridge with 1.92 Hz fundamental frequency (from [ALI]):


(a) typical surge phenomenon in midspan vibration velocity
(b) displacement decay in an impulse loading test
(c) amplitude spectrum of quarter-span vibration velocity
Footbridges 89

The harmonic analysis of the response records substantiated that the fundamental
equency of the structure was 1.92 Hz (Fig. A1.1c). The closeness to the predominant
rate of about 2 Hz explains the sensitivity. As an improvement measure it was
~u."'.<:.~u~~~ to install a tuned vibration absorber (see Appendix B 7).

Hz

After finishing construction of a footbridge with a 2.40 m wide prestressed concrete box
. . .""J..L ... ~'L'
simply supported with a 34 m span, the structure was found to be very lively [A2.1].
However, an immediately adjacent railway bridge made it possible to join the two
with a rubber profile band as commonly used for simple bridge deck expansion
This measure reduced the vibrations to a relatively small magnitude which to date
not given rise to complaints.
Subsequent examination yielded a fundamental frequency of the footbridge of ~2.3
same as the theoretically computed value.
In order to assess its vibration behaviour, one may calculate a forced vibration. One
produces with a matched pacing rate of 2.3 Hz the following vertical amplitudes for
assumed damping ratio of 1%: 2.3 mm displacement, 33 mm/s velocity, and
m/s2 acceleration. Based on Matsumoto's formula (Eq. (2.6)), giving for a mean flow
of 30 persons/min. and 1.5 m/s forward speed a (reduced) enhancement factor m =
this results in an acceleration of 1.3 m/s2 This value exceeds by far the suggested
t'''r>'~a level of Tab. 4.10.
" f ' f ' a .....

It was fortunate coincidence that the flexible joint connection to the adjacent bridge was
possible as this measure increased the damping considerably.

3: Hz

When a steel footbridge showed noticeable vertical vibration, its frequency and
amplitude were measured during normal evening peak traffic [ALl]. The number of
people crossing varied between 30 and 55 per minute. It became obvious that the bridge
vibrated predominantly in its fundamental frequency of about 4 Hz. This was excited by
the second harmonic of the load-time function with a pacing rate frequency of about 2 Hz
(see also Fig. 2.10).

During the opening ceremony, about 300 to 400 persons moved across a new steel
footbridge which was excited to strong lateral vibration [2.12]. Figure A4.1a shows the
90 Case Reports

elevation
( a)

I
cross- sect ion I

( b) (c)
I I

W

I
I

~~
-I-- \
\
I
If
lateral mode shapes I II
.1 1.0 8.0 Hz

N N (J)
<X:l r<>
r<> r--:

Fig. A4.1 Footbridge with lateral vibrations at 1.1 Hz (from [2.12]):


(a) elevation, plan and cross-section
(b) first three mode shapes of lateral vibration
(c) transfer function for the lateral displacement in midspan
Office BUilding 91

system of an S-shaped continuous girder over several spans, of which the main of
m is suspended from an angular arch. The intermediate supports of the girder are
pivot joints.
Figure A4.1b gives the first mode shapes for lateral vibration, Fig. A4.1c is a double-
plot of the transfer function for the lateral displacement amplitude in the
l"rrr:>1'11"h1'nlf'

of the main span, calculated for stationary excitation over the length of the main
and a damping ratio of 0.5%.
The frequency of the lowest lateral mode is in the range of ~ 1.1 Hz and thus very close to
the mean pacing rate at ~2 Hz, producing an almost resonating vibration (see
2.10). Moreover, it could be supposed that in this case the pedestrians synchronized
step with the bridge vibration (as described in Section 2.2.1), thereby enhancing the
considerably.
h1'<:o,1"l'.n

Improvement was achieved by tuned vibration absorbers to be effective in the lateral

5:

A three-floor industrial office building was to be raised by two storeys, which - for load
reduction were designed as a lightweight steel structure [2.26]. When the new interme-
diate floor could be walked on, it was observed to respond disturbingly to human motion.
The structure of the extension consists of sixteen frames with a single span of 12.8 m
(Fig. A5.1), spaced 1.55 m apart. Each storey has a height of 2.9 m. The intermediate floor
is a composite construction with steel girders IPE 330 supporting corrugated steel sheets,
which are filled with cement mortar to 25 mm above their upper edge. The frame columns
are not centred on the inner structural walls of the old building but are supported indirectly
(i. e. elastically) by a steel girder grid. This grid made from quite heavy sections extends in
both directions about 3 m and 4.5 m, respectively, beyond the extension floor area to reach
the external walls of the old building. Additional supports are provided under the main
girders HEA 400 at about 6 m spacing.

Preliminary tests with one person skipping indicated a fundamental floor frequency of
~6 Hz, which was more precisely determined later on by means of impulse loading (using a
dropped ball) to 5.85 Hz at a damping ratio of ~1 %.
Further skipping tests revealed basically the same dependence as mentioned in [2.1] (see
Case No.7). The fundamental floor frequency was excited by a higher harmonic compo-
nent of the dynamic load. Graphically speaking, the higher harmonic excitation can be
visualized as skipping not with every, but every other or every third downward swing of the
floor (at a certain phase shift), i. e. at a skipping frequency of ~2.9 Hz with every other
swing (2nd harmonic of the loading matching the structural frequency), or at one of ~2Hz
with every third swing (3rd harmonic of the loading). The second harmonic (~2. 9
produced the larger amplitudes, because the by one third higher loading rate overcompen-
92 Case Reports

tl 12510 ;;o(

IPE 300 (200 400dl


I"-

0 0
0 0

~ '"0-w
t .'2." 'V~~IO/IV
W

'" C B P R A o Q E F \
0-

2000 3000 3600.4000 6255 6255 4000 3600 3000 2000


1 ! I 1 ! l j j I ~""")., ...
~
~ IPE 330
U II 11
6680 !! lJ 5830

If) I) 1900 :1
~ :Ii
ii i :
IU HE8300
HEA400
"""
~II ! Jig HEA400
=
HE8 300
~ I
IPE 270

I
.

3000 3460 I 6150 I. 3200 1 3400

Fig. A5.1 Office building (from [2.26]): Section through a typical frame, supported
elastically by the steel girder grid

sates the smaller impact factor at higher frequency (see Section 2.2.2) and imparts more
energy to the system.
The vibration due to three persons walking on the floor was also largely dominated by
the fundamental structural frequency of 5.85 Hz and had a peak acceleration of 10% g. The
serviceability criteria of both ISO 2631 [4.13] and VDI 2057 [4.14] indicated that normal
office work could not be performed on the intermediate floor without improvement
measures. The much stricter bounds of DIN 4150 [4.1] were exceeded by far, but also
according to BS 6472 [4.25] complaints were certainly to be expected.
Improvement was attempted primarily by fitting a number of additional columns MSH
120/80/5 into the lower floor of the extension (see Fig. A5.1). This raised the fundamental
frequency by a factor of almost two to 11.25 Hz and reduced the amplitudes noticeably, in
particular the maximum acceleration (due to three persons walking) to ~6.6% g. Judging
the new state by the aforementioned codes leads to diverging opinions: ISO 2631 is not
really applicable to buildings, VDI 2057 seems to lack the necessary distinctions for this
problem, according to DIN 4150 remaining vibrations are still far from being admissible,
and BS 6472 indicates that user complaints might be possible but not very probable.
Since then, a second row of additional columns RHS 120/60/5 has been fitted
(Fig. A5.1), and among other extra measures a specially soft floor cover has been
provided. People walking in offices and corridors still cause vibrations of clearly percep-
Exhibition Pavilion 93

magnitude, which, however, when questioned, the personnel in the extension storeys
not regard as disturbing. Hence, in cases like this, the British Standard BS 6472 [4.25]
to be best applicable.

In a two-storey exhibition pavilion strong vibrations occurred whenever people walked


the intermediate floor, restricting considerably the serviceability of the building [2.23].
structure vibrated both vertically and horizontally.
Designed as a lightweight steel frame, the loading from the intermediate floor is carried
corrugated steel sheets and single-span cross-beams to two main longitudinal beams of
section IPB 200, which run continuously over several spans.
Measuring the vibrations due to walking gave the frequency spectrum of Fig. A6.1a.
Peak values can be seen to occur around 6 Hz and 27 Hz. Under impulse loading the lowest
frequency of the corrugated sheet floor was determined to be 27 that of the cross-
beams to be 18.5 Hz and that of the longitudinal beams to be 6.2 Hz. In a similar manner to
that of the office building (Case No.5), mainly the fundamental frequency of the system
was excited by the 3rd harmonic of the dynamic load-time function (about 2 Hz pacing
rate).
Architectural reasons did not permit any visible alteration of the structure, so that the
stiffness could not be raised by fitting additional columns or strengthening the longitudinal
beams. The only practicable way of improvement was to install tuned vibration absorbers,
eight in total, carefully tuned and damped with a mass of 65 kg each. Figure A6.1b gives the
constructional details. They reduced the displacement amplitude of the main beams by a
factor of six.

7:

A two-storey gymnasium building showed severe vibrations soon after its opening [2.1].
They occurred regularly whenever the upper hall was used by modern-style fitness classes,
performing skipping, jumping and running exercises to strongly rhythmic music. They
became manifest mainly in the lower hall, where the vertical swinging of the intermediate
floor under the direct loading due to the fitness class was clearly noticed and where the
glazed exterior walls vibrated horizontally.
The secondary effects were also rather dramatic: rattling of the entrance doors and the
shutters to the apparatus store-room, clattering of equipment attached to floor and walls,
and a strong, alternating draught due to compression and decompression of the air volume
in the lower hall, which one could feel when standing near the open entrance door. Several
times these combined phenomena caused people to leave the lower hall hastily during
fitness classes going on above.
94 Case Reports

( a)

6 12 18 24 30
frequency [Hz]

tuned
vibration absorber

longitudinal
girder

Fig. A6.1 Exhibition pavilion (from [2.23]):


(a) acceleration frequency spectrum
(b) details of the vibration absorbers on the main girders
Gymnasia and Sports Halls 95

(0 ) (b) d [mm]
6 II
HEA900 I
I

,,
I
4
J \
intermediate floor
J
\ ,
2
I "
__ ",,4 " 0.......

'/l '// '/// 18.52

(c)
I

entrance
l 3.761 550 I 550 I 3.76"1

i 19,12 'I.

(d )
"V'

~ I' + of + "t"~
~~#I -t- + + + + -I- +
~~ ~
-t.&; ::o!J +1

I 17.70m
376m S,SOm

Fig. A7.1 Two-storey gymnasium hall (from [2.1]):


(a) cross-section of building and plan of the intermediate floor
(b) resonance curve of displacements
(c) detail of the displacement plot with the supposed incidents of loading with
period T p and structural period = T p /2
(d) elevation and cross-section of the improvement carried out
96 Case Reports

Figure A7.1a shows the building in plan and section. The intermediate floor is con-
structed as reinforced concrete beams of 1 m height, 18.5 m span, and at 3 m spacing. They
rest on edge beams supported by steel columns at 6 m intervals.
The dynamic properties of the floor and the magnitude of loading was measured during
controlled tests in which fitness classes of up to 130 participants were instructed to perform
various exercises at specified frequencies, synchronized by the loudspeaker-amplified beat
of a metronome.
Resonance curves were evaluated for approximately 20 seconds during skipping excita-
tions with frequencies varied from 2.0 to 3.2 Hz (Fig. A7.1b). The strongest vibration
occurred at ~2.48 Hz. Since the fundamental frequency of the floor was ~4.9 Hz,
resonance was excited by the 2nd harmonic of the load-time function as though the
gymnasts were loading the floor on every other downward swing (with a certain phase
shift) (Fig. A7.1c). The peak amplitudes were measured to 5.5 mm vertical displace-
ment, 167 mmls corresponding velocity, and 5.15 m/s 2 acceleration, i.e. 52% g (!).
The damping ratio was determined as 2.4% of critical damping by measuring the
vibration decay after a commanded sudden standstill of the group. (The damping ratio
would have been smaller without people on the floor.)
Improvement of the structure was imperative because the high stress level entailed the
risk of fatigue deterioration of the reinforcing steel and of other damage; moreover, the
serviceability was much reduced as in the upper hall fitness classes with the usual number of
participants would have to be discontinued.
In order to avoid resonance with the 2nd harmonic of all possible load-time functions,
the fundamental floor frequency had to be brought to some 7.5 Hz. The beams were
stiffened by adding a bottom flange, made of a steel plate 800 mm x 60 mm and two side
plates and filled with concrete. For a good bond the surface of the concrete beam was
chipped and the flange with the concrete infill pressed against the beam by means of
prestressed cross-bolting (Fig. A7.1d).
New excitation tests with fitness classes as before yielded a much reduced vertical
displacement amplitude of 0.3 mm, i. e. only 5% of the value prior to improvement, and
did not produce resonance. The new fundamental frequency was measured to 7.3 0.1 Hz
and the damping ratio to 2%. As this was done in a drop-test with a 100 kg sandbag from a
height of 1.8 m, this damping value cannot directly be compared to the previous one, in
which many more people were on the floor. Further details, discussion of theory and
improvement variants can be found in [2.1].

No. Adllacent to

The design project of a large two-storey sports hall involved 3 m high and 42 m spanning
prestressed concrete beams (Fig. A8.1a). On one side the columns had also to support a
grandstand.
Gymnasia and Sports Halls 97

o
o

o
o
N

Fig. A8.1
Sports hall connected
to grandstand:
(a) cross-section
(b) simple dynamic model

For a first assessment of the vibrational performance under rhythmic excitation as


during fitness classes, dynamic calculations were applied to a simplified but conservative
model structure (Fig. A8.1b) and yielded a fundamental frequency of ~2.4 Hz. The low
value was mainly due to a very adverse contribution of the horizontally swinging
grandstand mass. For the resonance case of the excitation frequency being equal to the
structural frequency, a forced vibration computation with 1% damping and a user density
of one person skipping per 4 m 2 indicated amplitudes of 34 mm displacement, 560 mrnls
velocity, and 8.6 rnls 2 (!) acceleration.
The high stress level was expected to lead to damage or even structural collapse.
Therefore, one had to consider substantial design modifications and strengthening with the
goal of reaching a fundamental frequency of the floor of about 8 Hz.

In a high-rise building with a reinforced-concrete framed structure, gymnastic exercises


held in the uppermost storey regularly produced considerable vibrations [A9.1]. They
were felt in the uppermost six storeys, i. e. over about a quarter of the building height, and
led to complaints of the office employees. At a skipping frequency of 2.2 Hz, maximum
floor accelerations were measured to be some 1% g, a magnitude which can be quite
disturbing.
98 Case Reports

According to a preliminary investigation, the office floors vibrate mostly with 4.4
even though their fundamental frequency lies around 10 Hz. This strange phenomenon
found some explanation in the observation that the 4.4 Hz vibrations originate from the
concrete columns acting as extensional springs, an effect which also could be confirmed
through calculation. The vibration is excited by the 2nd harmonic of the loading function
and migrates through the columns into the lower floors. This rather interesting finding
needs further investigation, but any economical improvement would certainly be rather
difficult to conceive.

No. 'O'1I1l ........... Hall Music

A medium-size hall of 32.8 m x 55 m area exhibited strong floor vibration during pop
concerts [A10.1]. These events were attended by up to 4000 people, a part of the audience
was always enthused to activities such as clapping of hands, body rocking or light bouncing
up and down to the music. Several hall areas without fixed seats brought about differences
in density of the audience with a particular grouping (estimated 6 persons/m2 maximum)
towards the stage whenever the musicians were on show.
The plan and the sections of the floor structure in Fig. A10.1 reveal a hollow-block slab
of 0.45 m thickness and about 450 kg/m 2 mass. The floor is supported by two rows of
columns spaced 6.62 m apart, across which runs in longitudinal direction an internal joist (a
cast-in strengthened steel section DIE 38). The structural action can thus be described asa
predominantly crosswise-spanning continuous plate over three openings of 8.6 m / 14 m /
8.6 m. The floor is bisected by a dilatational joint. Designated originally as an exhibition
hall, the building had merely been designed for static loads, so that the new problems are to
be attributed to the unanticipated utilization for concerts.
Measurements were made at several concerts, showing that the vibration intensity
depended more or less on the kind of music influencing the activities of the audience,
and on its beat frequency fm. For example, soft pop music (Band Simple Minds) with
fm =2.0 -;- 2.5 Hz and 3500 relatively quiet spectators yielded at fm = 2 -;- 3 Hz a peak
vibration amplitude of 4.6 mm vertical displacement and 1.66 m/s2 acceleration. During
hard rock music (Band Status Quo), however, with a beat of fm = 2.1 -;- 3 Hz and an
audience of ca. 2000 (of which about one third was quite electrified), the displacement atfm
= 2.1 Hz reached 9.5 mm, the peak velocity 177 mm/s and the maximum acceleration
2.7 m/s2
When excited by a local impulse, the unloaded floor showed a fundamental frequency of
~6.2 Hz. Dynamic loadings on a larger area would probably bring out the inherent
structural action in the crosswise direction of the building and with the additional mass of
the audience - a fundamental frequency in the range of double the beat frequency of the
music (ca. 4.0 -;- 4.5 Hz). This indicates a resonance-like excitation of the fundamental
frequency of the structure, as found for the gymnasium floor of Case No.7, by the 2nd
harmonic of the audience's bouncing-time function.
Concert Hall with Pop Music 99

The measured vibration magnitudes and concluded stressing of the floor exceeded the
aOJml~~Sl{)le values
by far so that the hall had to be closed for pop concerts. For improve-
the floor would have to be supported by additional beams in the cross-direction, or
props would have to be fitted and removed before and after every concert, if the
of the lower storey was not to be greatly restricted. Finally, the discussion became
as another location for such concerts was found with a much higher floor
1recluency of about 12 Hz (unloaded).

plan:

+- 8
.J

:Ldilatational joint

cross-section A - A :

long. section 8 - 8 :

Fig. A10.1 Concert hall with hollow-cast floor structure (from [A10.1])
100 Case Reports

nn:~u-,~JII Theatre

An open-air amphitheatre housed a religious gathering of the New Life People. While
music was played through the loudspeakers, now and then the delighted audience on the
main stand rose from their seats clapping their hands. Considerable vibrations occurred in
the stand structure which, according to unsophisticated measurements, had a fundamental
frequency of 2.5 Hz and a damping ratio of 1.5%. Installing additional columns improved
the situation.

No. Platform

The high-diving platform of A12.1 showed severe vibrations when used by visitors
to the pool, so it had to be closed and modified [ALl].
Before and after the structural modifications, test persons excited controlled vibrations
for velocity measurements at the two points and P 2 both during build-up and decay ofthe
motion. From a Fourier power spectrum, the following frequencies of the fundamental
modes were established:

before modification after modification


shaft nodding (PI) 2.12 Hz 3.27 Hz
lateral shaft sway (PI) 2.45 Hz 3.67 Hz
shaft twist (P2) 2.37 Hz 3.71 Hz

After modification, much smaller vibration amplitudes were obtained allowing the
facility to be reopened to the public.

Fig. A12.1
10 m diving platform with
measurement points
(from [ALl])
Diving Platforms 101

interest is the comparison between the post-modification frequencies and the lower
rt~(liUel[ll,;V
bound of 3.5 Hz recommended in Section 4.5. The shaft nodding mode is just
low this bound. Neither rigid-body vibration nor slab vibration seem to have been
probably due to favourable ground conditions and the existing platform joist,

The shaft of a 3 m and a 5 m platform showed intolerably large twisting vibration soon
completion. The two reinforced-concrete platforms are each supported by two steel
~VJ"""HJHU with cantelever beams, resulting in an eccentric layout of the facility with the four
forming the corners of an open rectangular shaft. The columns were braced
f'nolllrY,nc

only near the foundation.


The design engineer, although he had calculated the fundamental bending frequency
had found it uncritical, must have overlooked the low torsional rigidity of his design.
two concrete wall members were erected between the foundation and each
parallel to the flights of stairs, combining an architecturally satisfying solution
a much improved dynamic behaviour.

m InnoorJ

An indoor swimming pool with a 10 m platform of high architectural beauty in very


slender construction turned out to be useless because of its pronounced responsiveness.
Although many variants of design modifications were studied, no satisfactory modification
could be found, and thus the platform had to be demolished in favour of a much stiffer
structure.

More cases of inadequate platforms are described in [2.29].

No. 15: WIll'~Vllna li!fOl.lf>lhr'''''T JiUlla]lD~

Strong vibrations occurred in the factory floor (with a basement underneath) of a


twenty-year old weaving mill after modernization to faster running weaving machines
[A15.1].
Thorough measurements found vibration velocities of the floor structure under the
machines of up to 24 mm/s, exceeding by far the commonly used acceptance criteria of
10 mmls of DIN 4150. The machines themselves vibrated horizontally with a double
amplitude of 1.2 mm where only about 1 mm was tolerable. This limitation was not
demanded by the machine construction but by the risk of weaving flaws in the production
of delicate fabric due to weft irregularity.
102 Case Reports

( a) plan, with position of weaving machines


IX VIII VII VI

CD

0
I'-
<i It)

(,) a:
N
CD
CD
to
Cl

CD
OJ
to

section A- A section B- B

30

(b) plan with steel joist arrangement


X IX VIII VII VI V IV III

I I
I I

I
1- HEA 240

end plate detail cross-section of concrete joist


Factory Buildings 103

The floor structure consists of an 18 cm thick reinforced-concrete slab spanning continu-


4.70 -7- 5.88 m over joists, which in turn are supported at 5.88 m spacing, in a smaller
as prestressed joists at twice this spacing (Fig. A15.1a). The fundamental frequency of
floor was found to be ~21 Hz. Prior to modernization, the weaving machines had a
of about 200 r.p.m., equivalent to a frequency of ~3.3 Hz. The new machines,
run - depending on the produced fabric - at 240 to 290 r. p.m., equivalent to
Hz at maximum. Apparently, the floor vibrations were excited mainly by higher
and 5th) harmonics of the load-time function, details of which are unknown.
Low tuning by mounting the machines on soft springe-damper) elements was impossible
despite reduction in floor vibrations - because this would have aggravated the weaving
IHOlvHLHv vibrations beyond specified quality standards. Therefore, the floor was strength-
by fitting wide-flange steel sections every 2 m between the existing concrete joists
A15.1b). The upper flanges were glued with resin mortar to the underside of the
rOllcnete slab, and the section end plates were glued and dowelled~ to the concrete joists.
The improvement was successfull in that subsequent measurements showed the vibra-
velocity to exceed virtually nowhere the bound of 10 mmJs (with the exception of a
old concrete construction joint which demanded special attention). The weaving
vLUvL>,.,",U

HH-J'vHULvvibrations were also much smaller than before, and no further production-quality
VV.lvHhJoccurred.

No.

A newly erected factory building of about 17 m x 100 m in plan had two floors below
street level, the gound floor, and one upper floor, with a distinct design change at street
level (Fig. A16.1): The floor structure of the basement is a joistless flat slab of 22 cm
structural thickness and 3 cm cement topping, supported at 5.20 m column spacing in the
longitudinal direction, and at 5.60 m / 5.20 m /5.70 m spacing in the lateral direction ofthe
building. The floor structure above the ground floor, however, is a 10 cm corrugated-sheet
steel composite construction, supported by steel frames of sections HEA 700 as primary
and by 140 mm steel sections as secondary girders. The frames (without intermediate props
shown in Fig. A16.1) span about 16.5 m, while the secondary girders lie 2.30 m apart.
After completion, various plastics-production machines were installed in the ground
floor, among others extrusion presses, cutters and hot-moulding presses, which all caused
strongly felt floor vibrations. A measuring campaign was initiated, since the factory owner
intended to install similar machinery on the first floor as well, despite its planned use as
material store room and for the location of lighter machinery only.

Fig. A15.1 Weaving factory building (from [A15.1]):


(a) floor plan with location of machines and cross-sections
(b) lay-out and details of installed steel girders
104 Case Reports

L
1689
(a) 5.80 5.20 'l
5.89
1 Q
"-
'I
::!:

gJ coo 0 0 0 0 0 0 0 0t HEA 600


r I

- - HEB 500
r1')
I I HEB 500 - -

"-
~
"-
::!:' I I
l2Jr ooooooooood HEA 700

T
!i-steel pCOpS
HEB 180
I
r1')
"-
C\I
C\I 1 ~
r I

...-
./ RiC celums
50/30 cm
f./ RIC celums
30/30 cm
I /R/C celums
30/30 cm

I I
, r1')
"-
0
I
I I
C\I

I III 11\
D c B A

( b)

f-- effect of other machine:,

Fig. A16.1 Factory building with plastics-forming machines:


(a) cross-section
(b) induced vibration of the floor slab above the basement
Factory Buildings 105

It soon became obvious that the main sources of vibration was the stamping action of the
presses producing egg-boxes, cups, containers and similar products: In a
n-.I.JL.l.V\-J..I.'-I-_lUj::.,

motion, a table of e. g. 80 cm x 80 cm area moves upwards and is met by the stamp


n"....",l1lrrdownwards, resulting in a strong impact at closing as well as opening of the moulds.
lifting rate was counted to 13 --;-- 21 per minute, i. e. 26 --;-- 42 impacts per minute, that isa
of 1.4 --;-- 2.3 s.
The concrete floor, including the weight of the machinery, had a fundamental frequency
~ 19 i. e. a period of ~0.05 s. With this large difference in period, the floor slab is not
1.. to a forced periodic vibration but loaded more or less transiently by single thrusts
,.,,",',.1.\'-'-'-1-

a free vibration phase (in a conglomerate of frequencies) to almost complete decay


the next impact. Hence, the physical effect is that of single impulses (Fig. A16.1b),
though the rhythmic repetition suggests a periodic loading.
The velocity of the floor vibration amounted to 5 mmls, which was irrelevant to the floor
integrity. The K intensity, however, which describes the human perception
to VDI Directive 2057 [4.14], can be calculated to 3.5 (see Section 4.3). This
when one is exposed to it for hours, is categorized as strongly noticeable and

The investigation of the floor structure above the ground floor revealed frequencies of
6.3 Hz for the frame and ~ 12 Hz for the secondary composite structure, much less stiffthan
the basement floor. Installing there plastics hot-moulding presses was impossible without
extensive strengthening. After careful consideration, the planned improvement was to
install internal columns in the ground floor (at intervals of about one third of the span) with
the consequence of restricted production area, and to provide a 20 cm concrete slab
between the footings of the critical machines and the composite floor structure
(Fig. A16.1a).

No. l"'U~T~I=lforJnlln2: Press

The reinforced-concrete floor slab of a factory building had been designed for static
loads according to yield-line theory [AlO.1]. Thus, the 5.90 m (resp. 5.31 m) x 7.40 m large
floor bays between joists had a thickness of just 18 cm (Fig. Al7.1).
The floor vibrated considerably under the impacts of a metal-forming press it supported,
the peak vibration displacements being measured to 1.44 mm at a floor frequency of
~ 13 Hz. Fears that the structure might be overstressed turned out to be unfounded, but the
vibration of the press led to violation of specified product tolerances.
The required strengthening was realized in a first step by gluing two steel girders HEA
300 to the underside of the floor slab, reducing the displacement amplitude to about 60% .
As the remaining vibrations were still too large, the steel girders were replaced by concrete
beams with dowel connection to the floor, and the spans were reduced by haunching the
existing columns (Fig. A17.1). This brought the floor displacement amplitude down to
about 40% of its initial value.
106 Case Reports

press~ ~\~;rengthenin~ haunches

o()) columns I'

ui I 400x400
il
-----r-t-
I ..
,_JOists
I 580x500

6.77 7.40 7.40 m

Fig. A17.1 Plan and section of a factory floor with metal-forming press (from [AI0.l])

No.

Boring machine stations in the metal-working industry process the clamped workpiece
with a moveable boring head which is programme-controlled in spatial motion, and its
tools are changed automatically by a robot arm.
In this case, a prototype was put in operation and exhibited vibration of the boring head
which caused the working tolerances to be exceeded. To reduce the head vibration, the
machine shaft was filled with concrete, with the effect that the machine's intrinsic vibration
was reduced but that now the enlarged mass excited severely the reinforced-concrete flat
floor slab on which the machine stood; this in turn again led to machine vibrations with
tolerance violation.
The vibrations were reduced by doubling the time scale of the programme which
controlled the working motions. In the long run, however, a 50% reduction in productivity
was not acceptable, therefore a strengthening of the floor structure was envisaged. Owing
to refined programming in which the motion was only slowed down during the phase of
abrupt thrusting motion of the boring head, strengthening of the structure was electroni-
cally circumvented and postponed to the time when the installation of more automatic
stations would call for a comprehensive review of the vibration problem.
Factory Buildings 107

with R ...'1t1t..:.........'"" ..""","'"

On the first floor of a dairy building four butter churns were installed. The floor slab
supporting them is a 52 cm thick prestressed hollow-block slab spanning about 13 m. The
blocks are formed by top and bottom plates of 10 cm thickness and 25 cm wide ribs (in the
principal load-carrying direction) at 1.0 m spacing. These ribs also contain the prestressing
steel.
Each of the four butter churns rotates during production around a horizontal axis and
exerts thereby dynamic loads via their trestle to the floor. One cycle of 1.5 to 1.75 hours
produces between 1.3 and 2.5 tons of butter. The largest load results from 10 to 20 minutes
of kneading as the last process of the cycle: At every revolution of the churn of 3000 litre
capacity of cream (one of the four machines even had 6000 litres), the solidifying mass of
the nearly finished butter drops from the cover onto the bottom; this exerts an impulse on
the trestle and the floor.
For an investigation of the floor response, four acceleration pickups were installed to
transmit the vibration during production, from which the floor velocity was obtained by
integration. Its peak value was 6 mm/s near the smaller butter churns, while near the big
one several cycles exceeded 10 mm/s, the largest peak reaching 21 mm/s. A frequency
analyzer extracted two predominant floor frequencies from the records, the first at 12.5 Hz
and the second at 23.5 Hz.
As regards potential overstressing of the floor structure, Part 3 of DIN 4150 [4.4] points
out that vibration velocities of more than 10 mm/s may cause deterioration, so that a
calculational check of the stress level in the structural member is needed.

In a factory building, in which stamping machines with automatic retraction were


operated, the floor vibrated to a degree that the owner was afraid of damage to the
reinforced-concrete structure [4.20]. The vibrations occurred although the machines were
mounted on springe-damper) elements (according to Fig. 3. 6b), which, however, could not
be tuned low enough as then the machines would malfunction.
The vibrations were measured to clarify whether the structural integrity was at risk and
what kind of vibration isolation might be feasible. The fundamental floor frequency was
17 Hz, while the presses worked at 43 and 27 lifts/min, respectively. It was found that every
lift exerted an impulse to the floor slab which subsequently vibrated freely in its fundamen-
tal mode.
The peak velocities of the (predominantly vertical) floor vibrations were measured as
3.21 mm/s at 43 lifts/min and as 1.60 mm/s at 27 lifts/min. To these values, the Standard
DIN 4150 (Part 2) [4.1] assigns perception intensities of KB 2.2 and KB 1.1,
respectively (see Section 4.3). The conclusions according to the Standard were the
following:
108 Case Reports

(1) The magnitude of floor vibrations did not jeopardize the integrity of the structure.
(2) An intensity of KB = 2.2 is qualified as easily noticeable, but it would not impair
efficiency or well-being of the operating personnel.
(3) Reduction of the floor vibration without enhancing the machine vibration would
necessitate low tuning with a stabilizing mass (see Fig. 3.6c).
(4) Some reduction could also have been achieved by overhauling the machines (elimina-
tion of bearing clearances, optimizing the shape of cams, etc.).

No. IVlI111I1lg and ljrJllldlng Machines

The operation of a milling machine in an upper floor of a production and assembly


building caused considerable vibration over a large area of the reinforced-concrete floor
slab having a fundamental frequency of ~10 Hz [A20.1].
Close to the milling machine the vertical vibration velocity of the floor was measured
as 1.3 -7- 1.6 mrnJs, depending on the cutting depth varying between 3 and 6 mm. At the
spot where a new grinding machine was intended to be installed, the velocities were still
0.2 -7- 0.3 mrnJs, which was considered too large for an undisturbed independent operation.
The vibration transmission could be efficiently curbed by supporting the milling machine
on springe-damper) elements (see Fig. 3.6b): Close to the machine the peak vertical floor
velocities were reduced to 0.45 and 0.60 mrnJs for cutting depths of 3 and 6 mm, respec-
tively, and at position of the grinding machine to 0.5 mrnJs. When in operation, no
interference of the two machines was observed.

School tiulldJllltz Heat

The vibrations and noise emanating from the heating system in the basement of a school
building disturbed the classes [A20.1].
Measurements of vibration and noise showed that the vibration, mainly of the floor
above the basement, was largely caused by insufficient isolation of a heat pump and that
part of the noise had been also due to transmission of evaporator pump vibrations throqgh
its supports and the pipework. No isolation to structure-borne noise was provided between
the heat reservoirs and the building. The vibration isolation of the heat pump consisted of a
rubber mat between the structure and the concrete base on which the pump assembly was
mounted (see Fig. 3.6c).
The peak vibration velocities were measured as 10 mrnJs on this base, 0.31 mrnJs on the
floor adjacent to the base, and 0.21 mrnJs on the floor above the basement. For the
dominant frequency of 16.5 Hz which turned out to be the operating frequency of the
School Building with Heat Pump 109

compressor -, the perception intensities according to DIN 4150 (Part 2) [4.1] were
KB = 7.1, 0.21, and 0.14 for the respective point velocities; thus, the intensities ranged
from strongly to slightly noticeable.
Analyzing the isolated system of heat pump, base and rubber mat uncovered a natural
frequency of 10.6 Hz. For a ratio of 1.5 between compressor frequency and system
frequency, the effectiveness of the isolation is only 23 % and thus clearly insufficient. The
high natural frequency of the isolated system was due to an excessive contact pressure on
the rubber mat.
The vibrations could be considerably reduced by installing additional springe-damper)
elements, which - in order to cut off transmission of the high frequencies were put on
elastomeric pads. Likewise, the heat reservoirs were lined underneath with insulating
boards and all pipeworks fitted with rubber compensators. These measures remedied the
noise and vibration problems.
111

AllPenldlX B

Appendix B provides a summary of fundamental structural dynamics sufficient to


understand the material presented in Chapters 2 and 3. Examples are given for common
practical problems. Problems beyond this scope will necessitate the consultation of stan-
dard textbooks on the subject.

The following is divided into the sections:


B1 Systems with a single degree of freedom (SDOF)
B2 Systems with many degrees of freedom (MDOF)
B3 Systems with distributed parameters
B4 Harmonic analysis
B5 Frequency tuning
B6 Structural behaviour under impulsive loading
B7 Tuned vibration absorbers
B8 Material properties under dynamic loading.

The differential equation of motion is usually written in complex state variables which
include implicitly the phase angle. In some applications, however, the classical notation
with sine and cosine functions is preferable. The agreed nomenclature for displacement is
accordingly
u(t) = complex displacement
x(t) = real displacement.
112

Bl

As an introduction to more complicated systems, we consider a concentrated mass


acting on a springe-damper) element, Fig. Bl.l. It helps to illustrate some veryfundamen-
tal relations of general applicability.

Free 'Il./"lh""... 1h,n1l'll

One speaks of a free vibration if a dynamic system undergoes a transient motion without
external excitation. This state, assuming that damping is proportional to the vibration
velocity, is described by the differential equation:

m'x + c'x + kx = 0 (BI.I)


where:
x = displacement
m = mass
c = damping constant
k = spring stiffness.

The terms in Eq. (Bl.I) represent forces standing in equilibrium:


m'x = inertial force
C'x viscous damping force
kx = elastic restoring (spring) force.

The differential equation (Bl.I) is tried to solve with an exponential type of solution,
which will turn out later to be complex when damping is present:

---10=- F (t)

Fig. BLI
Dynamic system with a
single degree of freedom
(SDOF oscillator)
SDOF Systems 113

st
x(t) u(t) = D'e (B1.2)

Inserting the trial solution into Eq. (B 1.1), yields the characteristic polynomial

m 'S2 + cs + k = 0 (B1.3)

The following parameters are substituted:

? k (B1.4)
wI= m

(B1.5)

Their physical meaning is that WI is the undamped circular eigenfrequency and 1; the
damping ratio. (Note that here and in the following, the subscript 1 refers to the SDOF
system and does not necessarily imply an equivalence to the fundamental frequency of an
actual structure.)

Then the polynomial (B1.3) becomes

(B1.6)

with the two roots:

(B1.7)

The damping ratio 1; relates the existing damping to the critical damping

at which a vibration decays from the maximum displacement straight to the rest position
without actually performing a vibration at all (<<aperiodic vibration). Typical damping
ratios in civil engineering are

Thus, the roots of (B1.7) are complex:

(B1.8)

with:
i =
114 Fundamentals of Vibration Theory

The frequency COD = COl- VI - ~2 is the eigenfrequency of the damped SDOF system and
is always smaller than the undamped eigenfrequency COl- As the difference is small for
common damping ratios up to ~ 10 -;- 15 %, one usually sets (although not used in the
following development):

COD=COl-~=COl (B1.9)

If (B1.8) is substituted into (B1.2), the transient displacement solution is obtained:

(B1.10)

with two integration constants V A and VB to be determined from the initial conditions.

With aid of the Eulerian equations, the exponential function can be expressed in two
harmonic terms:

+i-CODt
e- = cos (CODt) isin (CODt) (B1.U)

so that (B1.10) can alternatively be written as

By substituting the complex vibration constants V by real constants X

the displacement may be rewritten in real format x(t) as

(B1.12a)

The physical meaning of this equation may be elucidated by introducing:

Xl amplitude constant
XB
CPl = arctan - phase angle
XA
SDOF Systems 115

x(t) (B1.12b)

the initial conditions


x(t=O) = 1 , x(t=O) = 0
'-'''U.~~H'f-, in the constants
= 0, 1,
(B1.12a) reduces to

This function has been plotted in Fig. B1.2 for a damping ratio of ~ = 10%.

(J)
x--..
x
---
o -/--+-------,f--------\;----+---~-__l

--- ---
B1.2
Free vibration of an SnOF
-1--t"'----...---------,-------,..----.,.------l oscillator with 10% damping
o 2 3 4 5

The exponential decay of the amplitude is a direct consequence of the presence of


damping. The ratio of two successive vibration amplitudes Xn and Xn +l defines the so-called
logarithmic decrement 6, which is related to the damping ratio ~ by

6 = In X
n
~ 2Jt~ (B1.13)
Xn +1

A structure with many degrees of freedom is usually characterized by its fundamental


natural frequency f 1 [Hz], which is computed from the first circular eigenfrequency
WI [rads/s] by

f1 = ~ = _1_.-'\ ~
2Jt 2Jt V ill
(B1.14)
116 Fundamentals of Vibration Theory

or alternatively by the inverse:

T i=2 ny1
1
o
(B1.15)

which is the fundamental period of vibration of the structure.

A forced vibration results when an external transient load F(t) acts on the mass,
extending the differential Eq. (B1.1) by a known forcing term on the right-hand side:

m'x + C'x + kx F(t) (B1.16)

The free vibration (B1.10) now constitutes only one part, the homogeneous solution, on
which the force-dependent, particular solution is superimposed.
The particular solution becomes especially simple if the time function of the force is
harmonic.

Harmonic Loading
The harmonic loading function may be caused by machinery with rotating parts. For the
so-called constant-load excitation (Section 3.2.1), the harmonic force takes the form

in which:
= amplitude of the loading
(00 circular frequency of the loading.

The differential equation (B1.16) of forced motion of a damped SDOF system then
becomes:

(B1.17)

The homogeneous solution corresponds to Egs. (B1.10) and (B1.12), respectively, the
particular solution is of the form:

(B 1.18)

The second term with X D is necessary as x(t) and F(t) usually are not in phase.
SDOF Systems 117

Inserting Eq. (B1.8) into (B1.17) leads after some arithmetic manipulation - to the
constants

(B1.19)

As with (B1.12), one can simplify (B1.18) to:

x(t) (Bl.20)

in which:

CPo = arctan

The sum of the homogeneous part (B 1.12) and the particular part (B 1.20) yields the total
solution

x(t) . (B1.21)

A steady-state vibration still continues after the free vibration caused by the initial
conditions has been damped out, so that finally only the particular solution - the second
term in Eq. (B1.21) will remain.

Generalization of the inhomogeneous differential equation (B 1.17) is best done by


reverting to complex notation. A general periodic forcing function can then be expressed as
a superposition of j harmonic contributions (see Appendix B 4 on harmonic analysis):
i'w.. t
+ cu + ku = j=O
00

mii I Ae
]
] (B1.22)

in which:
A j = Fourier amplitude of the j-th harmonic load component (may be complex)
Wj circular frequency of the j-th load component.
118 Fundamentals of Vibration Theory

In contrast to (B1.I7), the loading function of (B1.22) now contains an imaginary


contribution as well (remember the Eulerian equation (B1.11)). The homogeneous so-
lution is again Eq. (B1.10), while the particular solution becomes

00 i'w.. t
u(t) = I. Ue
j=O ]
] (B1.23)

For linear-elastic vibrational systems the total vibration response can be obtained by
superposition of the responses of individual harmonic components. (This is implied in the
following without writing the sum over all j.)

Inserting this with use of WI (B1.4) and ~ (BI.5) for the SDOF system properties, the
differential equation (B1.22) becomes for the j-th forcing component:

iwt iwt iwt A iwt


- w?Ue
] ]
] + 2'i::
'::> WI
iwUe
] ]
] + W2IUe
]
] = ::..:Ie
m ] (B1.24)

The common exponential term dropping out, the complex displacement amplitude U j is
derived to

(B1.25)

Dividing both numerator and denominator by wi yields

(B1.26)

the amplitude of which becomes for positive real A j

(B1.27)

The phase angle to the j-th component of the loading is

w w
cpo = arctan 2'i::
'::> 1 1 (B1.28)
] WI - wj
One obtains the total solution, considering all forcing functions, by superposing the
particular part (BI.23) on the homogeneous part (B1.10):

(B1.29)
SDOF Systems 119

As already mentioned for Eq. (Bl.21), the homogeneous solution accounts for the
initial conditions. As this part decays with the factor e-~Wlt exponentially due to the
damping~, the transient vibrational behaviour becomes irrelevant in many practical cases,
leaving just the particular part

iwt
u(t) = Ufe J (B1.30)

as the steady-state vibration under the j-th component of a general periodic forcing
function.

With the aid of the response amplitude (B1.27), one can define a dynamic magnification
factor to the static displacement under the load component A j :

1
(Bl.31)

This dynamic-to-static ratio of peak displacement depends on the damping ratio ~ and on
the ratio between the loading and the oscillator frequency, W/Wl. Plotting the dynamic
magnification factor versus W/Wl for various damping ratios ~ yields the so-called fre-
quency response curves, Fig. Bl. 3. They allow the rapid calculation of the expected
displacement amplitude for a loading at a given frequency or, respectively, of the critical
structural frequency range.

If the damping is small, the dynamic magnification has its maximum at resemance,
i. e. when Wj WI:

(B1.32)

25

20

1.5

. ';=0.02
>' 1.0 Fig. B1.3
Dynamic magnification with
5 reI. excitation frequency
for various damping ratios ~
0 (frequency response curves)
0.0 0.6 2.4 3.0
120 Fundamentals of Vibration Theory

~=O.05
C=O.02

90

Fig. B1.4
Phase shift of the displace-
ment of an SDOF oscillator with
o+-~~~::::::::::'-----r---,----.---,....---~ respect to the harmonic load
0.0 0.5 2.0 2.5 3.0

Figure B1.4 shows the phase angle as defined in Eq. (B1.2S). Note that at resonance
(Wj= WI) the loading and the displacement response exhibit a phase shift of CPj = 90.

General Loading
If the loading follows a general periodic (but non-harmonic) or a transient time function,
then a closed-form particular solution cannot be found.

One way to calculate the SDOF response to such a loading is to work in the frequency
domain. Firstly, the loading is decomposed into its harmonic components by Fourier
analysis (see Appendix B 4), transforming the loading function into the frequency domain
with complex Fourier amplitudes A j . Then, the appropriate values of the frequency
response curve (from Eq. (B1.26)) are multiplied with the Fourier amplitudes of the
loading and superimposed with the individual phase angles of the components. Having
thus obtained the displacement response in the frequency domain, the inverse Fourier
transformation yields the response in the time domain.

Example B1.l

The SDOF oscillator of Fig. B1.1 is excited by a half-sine impulse loading as used in the later
Example B4.3 (Fig. B4.1) with T p = 0.4 s, resulting e. g. from a gymnast skipping with fs 2.5 Hz.
The steady-state displacement response in the time domain is to be found.

The system has the properties:


10,000 kg
10 kN/m
2%
5 Hz i.e. WI = 2Jtf1 31.3 rad/s.
SDOF Systems 121

25

- - real part

15 imaginary part

5LJ
f----,
~
':.... -5

\
I
I
I!( Fig. B1.5
"- Example B1.l:
-:;j- I
-15 II Complex frequency response

-25
l curve of an SDOF oscillator
with f 1 5 Hz, ~ = 2%
0 50 100 150 200
(0)'
J
0.20

- - real part

0.12 - - - - imaginary part

0.04
"-
3
'-/

::J-.04

Fig. B1.6
-.12 Example B1.l:
Displacement amplitudes
-.20 in the frequency domain
0 200
(0)'
J
0.25

0.15

,,0.05
E
u
'--'

::J-.05

Fig. B1.7
-.15 Example B1.1:
Resulting displacement
-.25 in the time domain
0.0 0.2 0.4 0.6 0.8
t [s]

Figure B1.5 shows U/(A/k) as a function of Wj according to Eq. (B1.26). Taking from Fig. B4.l
the Fourier amplitudes of the loading A j and multiplying them with the magnification ordinates of
Fig. B1.5, one finds the response u(w) in the frequency domain, Fig. B1.6. The inverse Fourier
transformation yields the displacement u(t) in the time domain, Fig. B1. 7. As can be seen, the
displacement amplitude is 0.00124 m.
122 Fundamentals of Vibration Theory

For simple hand calculation purposes, one may assume that the response is dominated by the
loading component corresponding to resonance with the structural system. Then from Fig. B4.1c the
respective Fourier amplitude at WI is A 2 0.42 kN, which is superimposed to the static displacement
(at W o 0) under A o = 0.64 kN and yields

(B1.33)

Note that about 94(/,,0 of the total displacement is due to the dynamic effect, giving
== 1.05 mm. Because of the rather wide spread of frequencies in this case, the total response
Udyn
amplitude (Fig. Bl.7) is reasonably well approximated by Eq. (B1.33).

An equivalent approach for determining the response to a general loading function is


based on Duhamel's convolution integral. It describes the response of the SDOF oscillator
to a series of infinitely short impulses (Dirac functions). The response velocity due to such
an impulse of duration t p can be calculated from the equation of momentum to be

1 tp
-m . f F(t)'dt
0
(B1.34)

As the duration of the impulse is infinitely short, i.e.

(B1.35)

and the displacement x(t=tp ) is assumed to be zero (i.e. X B 0 at t = 0), the subsequent
free response x(t) can be calculated from Eq. (B1.12a) with Eq. (B1.34) to

x(t) == x(t)

P
(t
x(t) == e -~'WI.t. -1- ' fF(t)'dt )sm(wlt)
.
(B1.36)
m'wI 0

while neglecting the damping influence on the SDOF eigenfrequency according to


Eq. (B1.9).

The integral contains the momentum of the impuls, which for the Dirac function has the
value one with the appropriate units.
SDOF Systems 123

Example B1.2:

As an example of the Dirac impulse, Fig. B1.8 shows the resulting displacement response u(t) of a
similar SDOF oscillator as in Example B1.1 but with damping of ~ 5%.

0.20

0.12

r-1 O. 04
E
E
L-J

:J-.04 Fig. B1.8


Example Bl.2:
-.12 Free vibration of the SDOF
oscillator of example B1.1,
-.20 subsequent to a Dirac impulse
0 2
t [s]

The response to any general load-time function can be represented as the integral over
the responses to a series of Dirac impulses. At resonance (w o = WI) one obtains:

(B1.37)

This is Duhammel's integral.

Example Bl.3:

Figure Bl.9 shows a load-time function with the typical shape assumed for the crash of an airplane
on a reinforced concrete shell structure. This load is imposed on the SDOF system (Fig. B1.1) of
Example B1.1. With the aid of Eq. (B1.37), the time function of the displacement response u(t) has
been computed and plotted in Fig. B1.lO. In order to show how different the peak response can be,
the behaviour of SDOF oscillators with 2 Hz and 10 Hz, respectively, has additionally been plotted in
the same diagram.
Performing the same SDOF response calculation for a given load-time function and for the whole
range of frequencies, one can compose the peak responses to the so-called response spectrum,
Fig. B1.11.

Besides these two methods, the response in the time domain can be computed by
numerical methods, as, for instance, the direct integration of the differential equation of
motion in discrete time intervals (schemes of Newmark, Wilson, etc.). For more informa-
tion about these methods, the reader is referred to [3.13] and [B1.1].
124 Fundamentals of Vibration Theory

1.00 - . - - - . = 0 - - - - - - - - - - - - - - - - - - ,

80

z
..:::::, 60

LL

40

Fig. Blo9
20 Example Blo3:
Load-time function of
an airplane crash
0.25 0.50 0.75 1..00
t [sJ

1..0-.----------------------,

0.6

" 0.2
E Fig. BloW
u
L.J Example Bl.3:
:J -.2
Displacement history for
an SDOF oscillator with
-.6 s
m = 1000 kg, = 5%
and varying frequency of
-1..0+-----,------,-----------,-----1 f 1 = 2 Hz / 5 Hz / 10 Hz
0.00 0.25 0.50 0.75 1..00
t [sJ

0.5 - r - - - r - - - - - - - - - - - - - - - - - - - - ,

0.4

0.3

Fig. Bloll
x 0.2
(0
Example Blo3:
E
:J Response spectrum to crash
0.1. loading (Fig. Bl.9) for
various SDOF frequencies
0.0 -J-------,,.------~---===::::;:=======::j and S 5% damping
o 5 1.0 1.5 20
f [HzJ
125

:82

For the calculation of vibrational response, a structure has to be abstracted to an


appropriate simpler model with a reduced number of degrees of freedom. A beam, for
example, can be modelled as a series of lumped masses along its central axis, or a framed
structure with a lumped mass for each storey on a vertical bar. This is denoted as a multi-
degree of freedom (MDOF) system.

B Free Vibration

Without external excitation, an MDOF system vibrates in its natural modes each of
which is analogous to an SDOF oscillator. The differential equation of motion for free
vibration is written in real matrix-form as follows:

[M]'{x} + [C]'{x} + [K]'{x} {O} (B2.1)

where:
[M] mass matrix (diagonal for lumped masses)
[K] = stiffness matrix (symmetric)
[C] = viscous damping matrix (symmetric and orthogonal, see below)
{x } vector of displacement x(t).

The differential equation is of the same form as the one for the SDOF oscillator
Eq. (B1.1), except that now the single variables are substituted by vectors and matrices.

In the absence of damping, Eq. (B2.1) is simplified to:

[M]{x} + [K]'{x} {O} (B2.2)

Assuming the solution to be of the form (analogous to Eq. (B1.2)):

n iw t
{x(t)} = {u(t)} = I
k=l
{'lJd.e k (B2.3)
with:

{'lJd = vector of the k-th eigenmode at frequency Wk


126 Fundamentals of Vibration Theory

one obtains a system of homogeneous equations:

([K] wt[M]) {'tiJd {O} . (B2.4)

which represents a dynamic eigenvalue problem. It has non-trivial solutions (in the sense of
{'tiJk} =I=- {O} for all k) only if its determinant vanishes:

(B2.5)

This requirement leads to the characteristic equation of the problem with a polynomial
of degree n, which generally yields n different eigenvalues Wb the circular eigenfrequen-
cies. From (B2.5) then follow n eigenvectors {'tiJk} , the mode shapes, which form together
the matrix [tV] of the eigenvectors, the modal matrix. (For solution methods of the
eigenvalue problem, see e.g. [B1.l].)

Example B2.1

Figure B2.1 shows the shapes of the eigenmodes for a simplified three-storey building with three
degrees of freedom. The parameters are:

kl = 600 kN/m ml = 10 t WI 4.58 rads/s


k2 = 1200 kN/m m2 15 t W2 = 9.83 rads/s
k3 1800 kN/m m3 20 t W3 = 14.50 rads/s

giving the stiffness and the mass matrix of the system to

[ 1O~ 0 00]
600 -600 0 ]
[K] = -6 00 1800 -1200 [kN/m], [M] 15 [t]
[
0 -1200 3000 o 20

For a forced vibration the differential equation of free vibrations is augmented by the
forcing function vector {F(t)}:

[M]{x} + [C]{x} + [K]{x} {F(t)} (B2.6)

The solution can be obtained:


- in the frequency domain
- by modal analysis
- by direct integration in the time domain.
MDOF Systems 127

F(t)
---lIP- X1

~ k1
2 2
F( t) m2
---lIP- x2
k2 k2
"2 "2
F( t) m3
-..... -"",x 3
k3 k3
2 "2
/;

Fig. B2.1 Example B2.1:


Idealization of a three-storey building as 3-DOF system

Solution in the Frequency domain


In generalization of Eqs. (B1.22), (B1.23), the forcing function F(t) and the solution x(t)
can be transformed into the frequency domain. Periodic loading results in discrete fre-
quencies Wj, whereas aperiodic loading has to be described by a continuous frequency
band.
For the simplest case of a harmonic excitation, the Fourier transforms (see also Appen-
dix B 4) read

{u(t)} {U(w)}.ei . w .t (B2.7)

{F(t)} {A(w)}.ei . w.t

and the differential equation (B2.6) becomes after substitution:


iwt iwt
([K] + iw[C] w2 [M]),{U(w)}e = {A(w)}'e (B2.8)

Collecting the matrices in the bracket above into the so-called dynamic stiffness matrix

[S] = [K] + i'w,[C] w2 [M] (B2.9)

the equation systems in the frequency domain becomes:


128 Fundamentals of Vibration Theory

[S] {U(w)} = {A(w)} (B2.10)

This permits direct solution for the displacement amplitudes:

{U(w)} = [S]-l{A(w)} (B2.11)

By applying the inverse Fourier transformation one obtains the displacement amplitude
in the time domain. Because of the high computational effort, this method is feasible only
for systems with a moderate number of degrees of freedom.

Example B2.2:

For the 3-DOF system of the previous example, Fig. B2.2 gives the frequency response curve of
the displacements for the case that the same loading is synchronously applied at all lumped masses.
Then Eq. (B2.11) becomes

{V(w)} = [wW A(w) (B2.12)

The displacements in Fig. B2.2 are normalized with respect to A(w). Obviously, only the first
eigenmode matters under such a loading. The higher modes are of greater importance, if a load
excites only the mass closest to the base (Fig. B2.3). In this calculation the damping matrix [C] was
assumed proportional to the mass matrix (uo 0.0625) and the stiffness matrix (Ul = 0.00139) as so-
called Rayleigh damping (see Eq. (B2.21) in the following section).

Modal Analysis
This method allows to decouple the system of differential equations of motion into a set
of independent differential equations of SDOF oscillators. To achieve this, the displace-
ment variables are transformed with the modal matrix ['If] to modal coordinates:

{x(t)} = I {1JJdYk = ['If].{y} (B2.13)


k=l

where:
n = number of degrees of freedom considered
{y} vector of the modal coordinates Yk
{1JJk} k-th eigenmode vector
['If] = matrix of the n eigenmode vectors.

The eigenmodes of the damped system are identical with those of the undamped system.
MDOF Systems 129

0.5

0.4

r""1
Z
~ 0.3
E
Fig. B2.2
E Example B2.1:
L.-J

<i 0.2 Frequency response curve of


.........
::::> displacements under synchronous
u1
0.1. loading at all lumped masses
l0
~i~~~ with Rayleigh damping of
0.0 U o = 0.0625, Ul = 0.00139
0 4 8 1.2 1.6 20
(Vj [ 1/sJ

0.30

0.24

r""1
Z
~0.1.8
E
E
L.-J Fig. B2.3
-0.1.2
Example B2.1:
.........
::::> Frequency response curve of
0.06 displacements as in Fig. B2.2,
~Ui
U2
but due to only one time-varying
0.00
0
'. U3

8
A
1.2
- 1.6 20
load acting on the lowest mass

(Vj [1/sJ

Substituting the physical coordinates in Eq. (B2.6) by the modal coordinates (B2.13):

[M][W]{Y} + [C][W]{y} + [K][W]{y} = {F(t)} (B2.14)

and premultiplying all terms with [W]T yields

[W]T[M][W]{y} + [W]T[C][W]{Y} + [W]T[K][W]{y} = [W]T{F(t)} (B2.15)

The matrix products are termed generalized matrices, i. e.

[W]T[M][W] = [M*] (B2.16)

[W]T[C][W] [M*]2~[Q] [C*] (B2.17)

[W]T[K][W] = [M*][Q2] = [K*] (B2.18)


130 Fundamentals of Vibration Theory

[M*] is the generalized mass matrix, [C*] the generalized damping matrix, and [K*] the
generalized stiffness matrix. [Q] is a diagonal matrix with the circular eigenfrequencies.
The foregoing operation renders all three generalized matrices diagonal with the following
relation for the k-th eigenmode (as k-th entries in the matrices):

(B2.I9)

(B2.20)

To decouple the damping matrix [C], it is mostly assumed to be a linear combination of


mass and stiffness matrix, the so-called Rayleigh damping [3.13]:

(B2.2I)

or in generalized form

(B2.22)

The k-th generalized damping term becomes, due to (B2.I7),

(B2.23)

and a modal damping ratio ~k can be evaluated:

(B2.24)

Damping defined in the (mutually orthogonal) eigenmodes of the undamped structure is


more generally termed orthogonal or classical damping. In principle, a different damping
ratio could be specified for every mode. In many cases, however, one damping ratio
common to all modes suffices.

Now, inserting the above generalized matrices into (B2.I5) yields the differential
equation of motion in modal coordinates {y}:

[M*]{Y} + [C*]'{y} + [K*]'{Y} [WY'{F(t)} (B2.25)

As all matrices are diagonal, the individual differential equations are decoupled and can
be considered separately, i. e. the k-th mode becomes:

(B2.26)
MDOF Systems 131

Note, that this equation has the form of Eq. (B1.16) of an SDOF oscillator and can be
solved in the same way (see Appendix B I). The results of all equations are the modal
coordinates Yk(t) which serve as multipliers to the eigenmode shape vectors {'4Jk} to yield
the total displacement vector {x(t)} by superposition according to Eq. (B2.13).
Modal analysis is motivated by the advantage of being able to consider only the
dominant eigenmodes of vibration and thus to reduce the degrees of freedom to the number
of modal coordinates essential to capture the system's behaviour.

Direct integration in the Time Domain


The direct integration discretizes the system of differential equations (B2.6) in time with
either implicit or explicit operators. For the solution procedure, the reader is referred to
[3.13] and [B1.I].
132

B3

True continuous mass distribution can be used for evaluating the eigenfrequencies of
beams, frames, arches and plates. They are calculated from the partial differential equa-
tion of motion for free vibration.
The following is a compilation of eigenfrequencies of several structural systems. Finally
it will be shown how to calculated approximately a forced vibration on continuous systems
by using the analogy of an equivalent SDOF oscillator.

B ~1I1f!I~e;:"\ID3In Beams

Consider the partial differential equation of the simply supported and freely vibrating
beam with continuous distribution of mass:

(B3.1)

in which (slightly different from the notation in previous chapters):


u = dynamic deflection of the beam
x = longitudinal coordinate along beam axis
t time
!l = mass per unit length
E I = flexural rigidity (assumed constant).

The problem is solved by separating spatial and temporal variables in product form:

u(x,t) w(x)T(t) (B3.2)

The time-dependent term follows to

iwt
T(t) = Tee (B3.3)

and the spatial shape of the vibration becomes


Distributed-Parameter Systems 133

w(x) = WA'COS(~'X) + WBsin(~x) + WC'sinh(~'x) + WDcosh(~x) (B3.4)

with:

(B3.5)

Introducing the appropriate boundary conditions yields - from the frequency equation -
the eigenfrequencies (Dn, the constants in the shape equation Wi, ~n, and finally W(x,(Dn),
i. e. the vibration shape dependent on the eigenfrequency in question.

Eigenfrequencies have been compiled in various tables and textbooks. Figure B3.1
compares the free vibrations of single-span beams with different boundary conditions,
expressed in the variable CPn' Then the eigenfrequencies result from the common formula

(B3.6)

Fig. B3.1
support conditions factor tpn
Eigenfrequency coefficients CPn
tpl = 4.73
for flexural vibration modes of
tpz = 7.85
single-span beams with various
tp3 = 11.00
support conditions, Eq. (B3.6)
tp4 = 14.14

tpn = n '7T
A A n = 1,2, ...

tp, = 4.73 Fig. B3.2


tp z = 7.85 Eigenfrequency coefficients CPn
~ ~ tp3 = 11.00 for other vibration modes of
tp4 = 14.14 single-span beams with various
support conditions, Eq. (B3.7)
tpi = 1.88

tpz = 4.69
~ tp3 = 7.85
support condition factor tpn

tp4 = 11.00
tp n = n '7T

tpl = 3.93 ~ ~ n = 1,2, ...

tp z = 7.07
~ 7A tp 3 = 10.21 ~
7T
tpn=-2'+n7T

:;j
tp4 = 13.35 n = j, 2, " .
134 Fundamentals of Vibration Theory

with:
E Young's modulus of elasticity [N/m2]
I = flexural moment of inertia [m 4]
!l mass per unit length [kg/m]
I length of beam [m].

For torsional, longitudinal and shear vibrations of a beam, and for a vibrating string, the
eigenfrequencies are computed from basically the same equation (ern from Fig. B3.2) with
the respective properties under the root:

torsional longitudinal shear string

Ul
n
= '1'"
I
.yG.IX
T
=i"''A =i"'~ = i"''v1 (B3.7)

in which additionally:
G = shear modulus [N/m2]
IT = torsional moment of inertia [m4]
Q mass per unit volume [kg/m3]
X = rotational mass moment per unit length [kg'm2/m]
S tensionforceinthestring[N].

~ ~ ~ ~~ ~
-t--l--f- tt7~t ~~ ti-ti=t ~~
w=~ w=4~3
ml
w=.1.. .j3EIl
ob m w=s-J"Y
ml 3 w=o~~

~ ~
mb

~m, +-l----+
~
tl+J
2 l 2 j+Jj-
l
2 2

w=.j ~ I w= w=4J 3EI i w=14./ EI . I


l3(m+ O.5m b) l3( m+O.375mb)
m+'3 m s

Fig. B3.3 Eigenfrequency of single-span beams with a concentrated mass: upper row
without beam mass, lower row with beam mass (mb = WI)
Distributed-Parameter Systems 135

If an additional mass is attached to the beam, the fundamental flexural eigenfrequencies


can be taken from Fig. B3.3. The beam itself may be treated as massless or as having a
uniform mass distribution (with mb = WI).

B 3.2

,u.. [kg/m] II = 20.0 [m]


2
E [N/m ] - - - - l1= 15.0
I [m 4 ] ---------- II = 12.0
An =wn'~ - . - . - lj= 10.0
[m- 2 ] - .. - .. - II = 8.0
11. 3 11. 3 11. 3 11. 3 11. 2
0.5

~J~\.
\ 'I~.

0.4
11. 2 1 - ' - ,
\
\

i"- . . . . . ..\
\
\
\

.\
~.
.\,-.,
\
\
1-.-..

.
- ..........
..............
\

\ \1
\
\~
":~
\
\,..

\ --"
11.
2 ---- \.~~--
'--- --- \
"' , ---- ~...
0.3 -I-
.....
' ... ....
\
.'\.
.......
. i'..
....
..... .......... ,

....

---..-
...........
' ' .... "'-.,
.. -,.... .
11. .............
1

0.2-I-
11. 2 ~
\

..... - .. .. -
........
~:--
- ...---~- r:-~
"

e-_I_1~
~~
-'-
.....
,
I- .:-
.....
............ ' ............
11.

11.
1

2
--.--.
'-. ~ ..........
.......
-':- ......... -'-. 1--.-
.... .....
.... ................
:-.::.~-~
0.1 - 11. 1 ----r---- ----- ~
.........
...... -
-------- _-
----- 1"'::.:::::-:....,."
~ "---
---~
-
11. 1 1 - - 0 - -
1---
11. 1 i .... ~

0.5 1.0

Fig. B3.4a Chart for eigenfrequencies of two-span beams (after [B3.1])


136 Fundamentals of Vibration Theory

The eigenmodes and frequencies of continuous beams are subject to too many parame-
ters to permit a comprehensive survey in this booklet.
As an example, however, Figs. B3.4a/b show diagrams for finding the fundamental
eigenfrequencies of beams over two or three spans. They are based on [B3.1], where many
more tables of other cases may also be found.

fL [kg/m] l1 =20.0 [m]


E [N/m 2 ] - - - - l1 = 15.0
I [m 4 ] - . - - - l1 = 10.0
- .. -.0- l1 = 7.0
An=Wn~ -------- II = 5.0
[m- 2]
/"3
0.5

--~---t>J.
/"11'- __

I
0.4 I '\ I
I
I
/"3 f'. '. " ,
" 1\ .., " , , ,
1

"- i', I
\ .., " I

0.3 0 . . .'
I". ...~ t- .. _

-. -. -. "- "- "-


/"2 1-- 1-- \

-. '--- ~.'- 1- -
1--', 1--
'--.
/"1 '--- . ,~

.~""" 1-.... -
1- - 1- -

'~'l ...., I-- ..... r...- .... t-...


A3 ~.
0.2 ......... , "- .-..- ..
-..- .1"'---.
........... , 1'--'_
~." ~ .....
"
'- ....
A2 f-'-r'_ r-..
A1f-'-:j-._ .~ .~
-'-.--.- -'- ' -
. ~.
'- ~'-
:::::..::-.....: ... .. _.j.-- -I.~-
0.1 A3
A2 . . . . . . .r---
A11:=.=-:::~ ~~ --
A2
"'--;;;100 __ -'. r--.
-'- 1--.-
-- ---- --
- - ,..::-....: t::.:,;"
0
-" r-.
..-..- - -
::..~ :;:..::::r..-:-::.. - _~.

A1 '- '- .-+-


0.5 1.0 1.5

Fig. B3.4b Chart for eigenfrequencies of three-span beams (after [B3.1])


Distributed-Parameter Systems 137

8 Frames

For the vibrational properties of frames, the reader is referred to [B3.2], where numer-
ous cases are tabulated.

83.4

The fundamental eigenfrequencies for radial and flexural vibrations, i. e. for symmetric
and antimetric vibrations, of a circular arch are (from [B3.3]):

radial flexural

~
w =CPR."
1 r2 V ~ 01 (B3.8)

with:
a = sector angle [radians]
r radius [m].

The values CPR and CPB can be taken from Fig. B3. 5, where:

A = cross-sectional area [m2].

support cond it ions factor 'f>

~
4 i
Q)
'" ... ... 'f>R=
~1 7T I
+(X4. r 2'
'"0 ~/
0 A
E
0
:0
~

~
_ ~1 + 500.5' I
'f>R - (X4. r 2 . A

47T 2 -(X2

, 'f>B=~
t?1
,'.
Q)
'"0 ../ \ 1+~2
0
E
.. / 47T

~
::l
><
Fig. B3.5
Q)

Eigenfrequency coefficients of
~
;;::
... " \ = ~3803.2- 92.101 ()(2 + ()(4'
,..// \
'f>B 1 + 0.06054' 0(2
circular arches, Eq. (B3.8)
(from [B3.3])
138 Fundamentals of Vibration Theory

Following [B3.3], a similar formula applies to parabolic arches:

CPa." ~ (B3.9)
F V~

with the CPa values to be taken from Fig. B3. 6.

80
81.9 f~
'\ J l ~
70
'\ 70.8

60
'\
~4)0 : I

\ 55:8

50 1""1
~
44.0
40

30
I
I
"" ""
34.7
Fig. B3.6
Eigenfrequency coefficients of
0.1 0.2 0.3 0.4 0.5
foil parabolic arches, Eq. (B3.9)
(from [B3.3])

B 305 Plates and Slabs

Quadratic and Circular Plates


The eigenfrequencies W n for quadratic and circular plates with various boundary condi-
tions are determined from the formula given in the caption of Fig. B3.7 with the aid of the
tabulated coefficient Bn

Rectangular Plates
Similar for rectangular plates, the two lowest eigenfrequencies W n for various boundary
conditions follow from the formula in the caption of Fig. B3.8, for which the coefficients CPn
are to be evaluated in function of the side aspect ratio y.
Distributed-Parameter Systems 139

support
81 82 83 84 85 86 87
conditions

e 11.84 24.61 40.41 46.14 103.12

@ 6.09 10.53 14.19 23.80 40.88 44.68 61.38

@ 4.35 24.26 70.39 138.85

I~@\
\
a I 5.90
,_/ I

Fig. B3.7
I~ 1.01 2.47 6.20 7.94 9.01
Eigenfrequencies of circular
and quadratic plates for
I~I 10.40 21.21 31.29 38.04 38.22 47.73 various support conditions
(after [3.8]):

l~ 2.01 6.96 7.74 13.89 18.25


W
n Bn
'\ /
V Q'a '(1-v
4
Et2
2)

---
with
I~i 6.83 14.94 16.95 24.89 28.99 32.71
E = Young's modulus [N/m2 ]
--- v Poisson's ratio
I~I 8.37 15.82 20.03 27.34 29.54 37.31 = slab thickness [m]
a = diameter, side length [m]
--- Q = mass per volume [kg/m3]
II~I
0 a I 5.70 14.26 22.82 28.52 37.08 48.49
1_ _ _ 1
for the support conditions
- - - - free edge
~ 4.07 5.94 6.91 10.39 17.80 18.85 _____ hinged edge
====== fixed edge

Continuous Plates
For the eigenfrequencies of continuous plates, if not to be approximated as continuous
beams (Fig. B3.4), the reader is referred to [3.6].
140 Fundamentals of Vibration Theory

support conditions

'Ui <PI,1 = 1.57 (1 + y2)


------

I a I (j + 0. 25y 2)
I b I <P2,1 = 6.28
I I
------
<P 1 ,2 = 1.57
(1 + 4 y2)

'ui
<PI , 1 = 1.57 J1 + 2.5 y 2 + 5.14y 4'
I 0 I <P2,1 = 6.28Ji + 0.625 y 2 + 0.321 y4 i
I b I
I I
<Pl,2= 1.57 ~ 1 + 9.32y 2 + 39.06y 4'

------ <Pl,1 = 1.57 ~ 5.14 + 2.92y 2 + 2.44y 4'

lui <P2,1 = 9.82Ji +0.. 266 y 2 +0.0625y

> 1,2
41

= 1.57 ~ 5.14 + 1O.86y 2 + 25.63y 4'


Fig. B3.8
Eigenfrequencies of rectangular
------
<P 1, 1 1.57 ~ 1 + 2.33y 2 + 2.44y 4' plates for various support
I II
<P2,1 = 6.28 ~ 1 + 0.582y 2 + 0.152y 4'
conditions (after [B3.6]):
\I U
b 11 i
~ 1 + 8.69y 2 + 25.63y 41
<P 1,2 = 1.57 ill = 2.Jt. CPn2 .
n a
with
------ <P 1,1 = 1.57 ~2.44 + 2.72y 2 + 2.44y 4'
CPl = CPl,1
lui >2,1

<PL2
= 7.95 ~ 1 + 0.395y 2 + 0.095y 4'

= 1.57 J2.44 + 1O. 12y 2+ 25.63 y 4'


CfJ2
and
Min. {CPl,2;CfJ2,1}

E Young's modulus [N/m 2 ]


v Poisson's ratio
9'1,1 = 1.57.J5.14 + 3. 13y 2 + 5.14y 4'

lu, <P2,1 = 9.82 ~1 + 0.298y 2 +0.132y 4'

c.p 1,2 = 1.57 -}5.14 + 11.65y 2 + 39. 06y4'


y
slab thickness [m]
aspect ratio alb
a,b = side length [m]
~ mass per area [kg/m 2]

In certain cases, the system to be calculated, such as a slab or a beam, can be approxi-
mated by a substitute SDOF oscillator. For it to be equivalent, its properties have to be
chosen such as yield to the same eigenfrequency ot interest and the same maximum
displacement amplitude - under a substitute force F (t) as the original system. The
Distributed-Parameter Systems 141

parameter transformation is achieved through individual conversion factors <P, which are
derived by equilibrating the work done by external forces, the kinetic energy, and the
stresses of the substitute system with those of the original system. The conversion factors
for various boundary conditions have been compiled in Fig. B3.9 for single-span beams
and in Fig. B3.10 for slabs.

Equivalent Parameters
The load-factor <PL = F(t )/F(t) serves to derive the substitute loading amplitude F( t)
from the true loading F(t) of the original system, keeping the time function unaltered. It is
derived by equilibrating the external work done in the two systems.
The mass factor <PM = film determines the concentrated substitute mass fi from the
given total mass m of the original system and is derived by equilibrating the kinetic energy
of the two systems. The foregoing two factors can be combined in the load-mass factor
Cj)LM = Cj)MI<PL'

moss factor load -moss factor effective


loading load beam dynamic
'PLM
and support factor I----..,..---+----r----I collapse load stiffness support reaction
conditions 'P L lumped distributed lumped distributed
moss moss moss moss
R k v

8M p 384EI
0.64 0.5 0.781 0.39R + 0.11 P
[ 5[3

4M p 48EI
1.0 1.0 0.494 1.0 0.494 -l-3- 0.78R-0.28P
l

8M ps 185EI Vf =0.26R + 0.1 2P


0.578 0.448 0.775 -[3-
l V2 =0.43R + 0.19P

16M ps 107EI Vi =0.54R + 0.1 4P


1.0 1.0 0.43 1.0 0.43 -[3-
3l V2 =0.25R + 0.07P

12 Mps 384EI
0.533 0.406 0.765 -l- -l-3- 0.36R + 0.14 P

192EI
1.0 1.0 0.383 1.0 0.383 -L-3- 0.71 R - 0.21 P

Fig. B3.9 SDOF conversion factors for single-span beams with various support and
loading conditions (after [B3.3]);
Mp = moment of plastic resistance (s support, m = midspan)
142 Fundamentals of Vibration Theory

Multiplying the effective stiffness k (as ratio of the amplitude of load and displacement)
?! theoriginal system with the loading factor <PL gives the equivalent spring stiffness
k = <Pc k of the substitute system.

Eigenfrequencies
The eigenfrequency of the substitute SDOF system is then evaluated in analogy to
Eq. (B1.4) as

load mass load- effective dynamic support reactions


0
support conditions factor factor mass collaps lood plate
b
factor stiffness
lPL 'PM 'PLM R k Vo Vb

---- 271 ETo


1.0 0.45 0.31 0.68 1; (MfO + Mfb) 2-
-0- 0.1 8R +0.07P 0.18R + 0.07P

'l~'
I b I
I 248Elo
~ (12 Mfo + 11.0Mfb)
I 0
I I 0.9 0.47 0.33 0.70 - 0 2-- 0.1 6R + 0.06P 0.20R+0.08P
----
228Elo
E.<O<b
2 - - 0.8 0.49 0.35 0.71 ~(12MfO+ 10.3Mfb) - 0 2-- 0.1 4R +0.06P 0.22 R + 0.08P

216E 10
0.7 0.51 0.37 0.73 ~(12MfO+ 9.8Mfb) 2-
-0- 0.13R+0.05P 0.24R + 0.08P

212 El o
0.6 0.53 0.39 0.74 ~ (12 Mfa + 9.3 M fb ) -0-
2- 0.11 R +0.04P 0.26R +0.09P

216El o
0.5 0.55 0.41 0.75 ~(12MfO+ 9.0M fb ) 2-
-0- 0.09R + 0.04P 0.28R+0.09P

870El o
1.0 0.33 0.21 0.63 30.2Msb
---;;r- 0.1 5 R + O. 10 P 0.15R+0.10P

IEjJj 0.9 0.34 0.23 0.68 27.8 MSb


798EI o
2-
-0- 0.14R +0.09P 0.1 7R + 0.1 OP

b 757 El 0
2' $ a $ b 0.8 0.36 0.25 0.69 26.0Msb 2-
-0- 0.1 2R + 0.08 P 0.19R+0.11P

- 744El o
0.7 0.38 0.27 0.71 26.0 Msb - 0 -2- 0.11R+0.07P 0.21 R + O. 1 1P

778EI o
0.6 0.41 0.29 0.71 26.4 MSb -0-
2- 0.09R +0.06 P 0.23R+0.12P

0.5 0.43 0.31 0.72 25.0M sb


- 866El o
-0- 0.08R +0.05 P 0.25R+0.12P
2-

Fig. B3.10 Approximate SDOF conversion factors for rectangular plates with various
support conditions (after [B3 .3]); the factors have been derived for simultane-
ous excitation of more than one eigenmode (impact loading) and are not true
SDOF conversion factors;
Ia moment of inertia per unit width
Mfa total plastic field moment along yield line parallel to side a
Msa total plastic support moment along side a
Distributed-Parameter Systems 143

(B3.10)

Besides the factors CPL, CPM, CPLM and the equivalent spring stiffness k, the Figs. B3.9 and
B3.10 contain information on the load capacity of the respective system and on the
magnitude of its dynamic support reactions.

Example B3.1

The fundamental eigenfrequency of a single-span, simply supported beam is to be derived from an


equivalent substitute system. The result can be compared with the original beam frequency given in
Fig. B3.1.

Parameters of the original structure are:


length 1, mass !l per m length, flexural rigidity E I.

Conversion factors for the substitute SDOF system from Fig. B3.9:
384EI
<PL = 0.640, <PM 0.50, k=--
5P

giving

0.64384EI
0.50W l5P

In this simple case, the SDOF approximation compares very well with the fundamental eigenfre-
quency from Fig. B3.1, which yields the factor

<pi = ;Tt2 = 9.87 == 9.92

in front of the square root.


144 Fundamentals of Vibration Theory

B4

Any periodic loading F(t) can be decomposed into a constant part and several harmonic
load contributions which, when superimposed, result in the total load-time function given.
This harmonic decomposition is a discrete Fourier transformation of the loading and is
therefore also called a Fourier analysis.

The basic relation is

F(t) (B4.1)

with the Fourier coefficients


1 T
-T .f F(t)dt
0
(B4.2)

2 T 2n
an = T f F(t)ocos(noTot)odt (B4.3)
o

2 T 2n
0

-
TJ F(t)osin(no-ot)odt
oT
(B4.4)

in which w 2 niT is the lowest circular frequency of the loading.


0

The frequencies of the harmonic components are multiples of the loading frequency Wo
An alternative way of writing the n-th component of the load is

(B4.5)

with the amplitude An and the phase angle ern of the n-th component. (For the static
component n = 0 is F o = ao .) They are:

An Va~ + b~ (B406)

an
er n = arctan-
b (B4.7)
n
Distributed-Parameter Systems 145

The result of the harmonic analysis of the loading is commonly plotted as the Fourier
component amplitude An dependent on the loading frequency w. This kind of plot is
known as a Fourier amplitude spectrum.

Alternatively, the decomposition can be formulated in complex terms (with the compo-
nent indices n or j - as in Section B1.2 used interchangeably):

. 2Jt
ln-t
F(t) = A'e
n T (B4.8)
n=O

2Jt
1 T -inTt
f
-' F(t)e
To
(B4.9)

in which now is a complex Fourier Amplitude of the loading function F(t). Equation
(B4.8) is thus the inverse Fourier transform of a periodic loading, which is defined in
the frequency domain as (infinitely many) discretely spaced harmonic components at
Wn n2Jt/T.

Example B4.1:

Consider a periodic loading with the time function of Fig. B4.1a. It may describe the loading due to
a person's skipping on the spot and has a half-sine shape with a contact duration t p = 0.2 s and a return
period T p = 0.4 s, corresponding to a skipping frequency fs = 2.5 Hz.

The first five harmonic load components of a Fourier analysis are presented in Fig. B4.1b. The
constant part Fo is in this case equal to the weight force G of the person. Superposition of all five
components yields the dashed loading function in Fig. B4.1a. The original loading would be captured
more accurately if additional higher harmonics had been considered. The Fourier amplitude and
phase angles of the harmonic components are shown in Fig. B4.1c. They are fully equivalent to the
complex representation in Fig. B4.2, where the amplitude (Eq. (B4.9)) is split into the real parts
(solid lines) and the imaginary parts (broken lines).

Example B4.2:

A loading function actually measured for a skipping motion is shown in Fig. B4.3a. Note, that the
Fourier amplitude spectrum (Fig. B4.3b) is continuous in reality, in contrast to that of the loading
model in Fig. B4.1.
146 Fundamentals of Vibration Theory

(0) 2 0 0 . . , - - - - , . . , . . - - - - - - - - - - - z ; : : " ; - - - - - -

160

....,
'" 120
Q
Z
..l<:
I-J 80

.:=.
LL
40

'"'0.8
t[ S]

(b) 200

120

'" 40
<:>
Z
..l<:
L...J -40

-LLc _
120

-200
0.0 0.2 0.4 0.6 0.8
t [S]

(c) 200

160
...., Fig. B4.1
'"<:> 120
Example B4.1:
Harmonic analysis of
Z
..l<: a periodic loading:
80
c (a) loading function

(skipping on the spot)


40 (b) harmonic load components
AO A2 I (c) Fourier amplitude spectrum
0
...jL -,L- -,L- -,Ar-3_ _ ~ of the loading function
0.000 15.708 31. 416 47.124 62.832 78.540
Harmonic Analysis 147

80
real port of An

40 imaginary port of An

N 0 I I
g I
Z
..:>::.
L-J
-40
I
I
I
I
c
<r I Fig. B4.2
-80 I Example B4.1:
I
I Complex Fourier amplitudes
-120 of the loading of Fig. B4.1a
0.000 15.708 31.416 47.124 62.832 78.540
2'7T
n'T [rods/sJ

(0)

1.i..Q. 4.0
I f s =2A Hz I
3.0

2.0

1.0

O.i 0.2 0.5 0:6 0.7 0.8 0.9 i,2 1.3 1.4
l Tp
"II,. time [5]

(b) (\)
"0
:E
0.
E
o

o 10
frequency [Hz]

Fig. B4.3 Example B4.2:


Experimentally measured skipping load: (a) loading function,
(b) corresponding continuous Fourier amplitude spectrum
148 Fundamentals of Vibration Theory

If a loading is of transient nature, e. g. a single impact, it may be regarded as periodic with


an infinitely large period T p . For the purpose of a discrete Fourier transformation in
structural dynamics, the period may be approximated as finite, provided it is chosen large
enough to assure the complete decay of the structural vibration due to the transient load
within this period. Therefore, the choice of the period depends very much on the funda-
mental frequency of the structural member and on its damping properties.

Example B4.3:

Figure B4.4a shows a transient, half-sine impact lasting for t p = 0.2 s. The complex Fourier
amplitudes are plotted in Fig. B4.4b, derived by a discrete analysis under the assumption of aloading
period T p = 10 s. As this corresponds to 0.1 Hz loading frequency, the components are spaced at
discrete frequency intervals which are multiples of 0.1 Hz.

(0) 2 . 0 . . . , , - - - - - - - - - - - - - - - - - - - - - - - ,

1.6

..--. 1.2
z
~
L.-J

I..L 0.8

0.4

0.0 - P - - - - - - . - - - - - - - , - - - - - , - - - - - - - - 1
0.0 2.5 5.0 7.5 10.0
t [sJ

(b) 0 . 1 0 , - - - - - - - - - - - - - - - - - - - - - - ,
- - real port of An
- - - imaginary port of An
0.06

1,l l l l l l l \I, I I I\ I 'I 1~ I1 1 1 1 1 1 1 1 1 1 1 '1 '1 '1 '1 '1 1 1 1 1 1 I ,' ,'" ".,
Fig. B4.4
..--.0.02
z Example B4.3:
~
L.-J ,., Harmonic analysis of a single
c-.02 "111111111111111111111111111111111 1"
IT '11111111111111111111111 1" impact load, treated as
""1111111""
periodic with T p = 10 s:
-.06 (a) loading
(b) discrete Fourier amplitude
-.10 + - - - - - - . - - - - - - . - - - - - - , - - - - - - - - 1 spectrum
0.0 12.5 25.0 37.5 50.0
[Hz]
149

85

Adjusting the eigenfrequencies of a structure or a structural member is always aimed at


avoiding resonance with a given range of the loading frequency, i. e. the structure is
actually de-tuned with respect to given load component frequencies. Using the terminol-
ogy of high and low tuning - and given that appropriate constructional measures are
feasible -, the fundamental structural frequency can either be increased to exceed the
highest relevant loading frequency (high tuning), or be lowered to remain under the lowest
relevant loading frequency (low tuning). The possible regions for high and low tuning are
illustrated in Fig. B5 .1, which shows - for various damping ratios - the dynamic magnifica-
tion in the displacement of an SDOF oscillator. It depends on the ratio between system
eigenfrequency fl and the frequency f of the relevant harmonic loading, which is the
reciprocal value of the so-called tuning ratio. As can be seen, the success of tuning is closely
linked with the degree of damping and the nearness of the relevant excitation frequency to
the fundamental structural frequency (frequency separation). Although, mostly referring
to the lowest structural frequency, high or low tuning can sometimes become important
with respect to higher structural frequencies (see Section 3.4.1).

25

low tuning high tuning

20

15
~=O.O2
..>L
> Fig. B5.1
10
Frequency response curve of an
5 SDOF system under constant-load
excitation, showing the possible
0 range for low and high tuning
0.0 0.6 1.2 1.8 2.4 3.0
f 1/f

BLow . . . . . ,.. JUllJ;:,

The success of low tuning stands and falls with the frequency separation and the damping
(Fig. B5.1). To illustrate how much the residual dynamic displacement depends on the
damping ratio, Fig. B5.2 shows for an SDOF oscillators the damping necessary to keep the
fundamental vibration amplitude Xl at a desired fraction of the static displacement Xstat for a
150 Fundamentals of Vibration Theory

1.0...,...---------------------;;>\

0.8

0.6
)JJ\

0.4
Fig. B5.2
I Required damping ratio ~, depend-
0.2
I ing on the frequency ratio f l If and
'0.50 =x;/xstot
0.0 the tolerable displacement Xl IX stat
0.0 0.2 0.4 0.6 0.8 1.0
f 1/f

given reciprocal tuning ratio. The larger the damping, the smaller the frequency separation
can be. If the damping is small as usual in civil engineering -, then slight shifts on the
frequency axis result in much different maximum vibration amplitudes. On the other hand,
inexact determination of the fundamental structural frequency will, if close to resonance,
make a large difference in terms of resulting dynamic displacements and stresses. The
following example demonstrates that low tuning has an even stronger effect on stresses in a
structural member than on the displacements.

Example B5.1:

A machine idealized as a mass supported on springs and dampers (Fig. B5.3) be subjected to the
harmonic loading with constant amplitude F o , corresponding to just one component (n=l) in Eq.
(B4.8) without higher harmonics:

iwt
F.oe

m
Fig. B5.3
Example B5.1:
Dynamic model of a machine
(on rigid structure)
under harmonic excitation
Frequency Tuning 151

Initially, let the structure be in resonance with the loading frequency, i.e. fl = w/(2'n), and the
static displacement for the pertinent structural stiffness be xstat,reson.' Now the structural stiffness is
gradually reduced. Figure B5.4a gives the forces and Fig. B5.4b the displacements for various
damping ratios in function of the resulting fl/f. The effect oflow tuning is seen to be more pronounced
in the stress resultants than in the total displacements, which include the static displacement in this
formulation.

(0) 25

20

15
LL t=O.02~
"'-
0:::
10
t=O.05

'j
5 t=O.10

0
0.0 0.2 0.4 0.6 0.8 1.0
f 1/f

(b) 25

20

C
:;; 15
~
t=O.02
..;
(1l
+> 10
><
(f)
~=O.O5
Fig. B5.4
"'-x Example B5.1:
(1l
E
>< 5 Effectiveness of low tuning:
(a) reI. spring reaction
0 (b) reI. machine displacement
0.0 0.2 0.4 0.6 0.8 1.0
f 1/f

The different ways of achieving low tuning structurally, as presented in Section 3.4.1 and
sketched in Fig. 3.6, are now to be compared more closely. The example of Figs. B5.3 and
B5.4 corresponds to a type (a)>> design of direct mounting of the machine on a relatively
soft structural element. For this type, only the structural parameters k and c contribute. An
additional parameter mM as in Fig. B5.5 enters in the form of the mass of the machine itself
(<<type (b)>> design, Fig. 3.6) or represents the machine combined with the stabilizing mass
152 Fundamentals of Vibration Theory

situation model

JL, E I, l

Fig. B5.5 Dynamic model of a machine mounted on a structural member


under harmonic excitation

of a vibration-isolated base (<<type (c)>> design, Fig. 3.6). The stiffness and damping
coefficients k M and CM are those of the isolating springe-damper) elements. The mass ms
and the stiffness k s are the parameters of the equivalent SDOF systems substituting the
structural member (see Appendix B 3.6). The loading from the machine is assumed to be
harmonic.

The major question is whether the resulting 2-DOF system, Fig. B5.5, can be treated as
two independent SDOF oscillators, i. e. whether the two spring-mass systems can be
decoupled. To this end, Fig. B5.6 takes the ratio mM/mS as parameter for a set of curves,
which give the coupled eigenfrequencies (01 and (02 (relative to the structural eigenfre-
quency (Os) versus the ratio of the decoupled eigenfrequencies of machine and structure. It
can be concluded that decoupling (i. e. (Os not influenced by the existence of mM and vice
versa) is feasible only if the individual eigenfrequencies of the two partial systems are
spaced apart and if the mass mM (machine including stabilizing mass) is small compared to
the equivalent mass ms of the structural member. With mM approaching the size of ms, the
coupled frequencies divert quickly from the decoupled values.

The eigenfrequencies (01 and (02 of the coupled 2-DOF system are:

(B5.I)
Frequency Tuning 153

2.0

1..6

1..2
Ul
3
"-
3
0.8 y
...... ......
... , ..
---
~
Fig. B5.6
0.4 Eigenfrequencies of the 2-DOF
system with various ratios
0.0-1"'------,----,-----,------,----- of masses and stiffnesses
0.000 0.400 0.800 1..200 1..600 2.000
.jk M' ms Ik s ' m~= wMI Ws

If decoupled, the system can be analyzed as follows:


Step 1: Calculate the displacements and forces (in spring and damper) of the isolators due
to the external load F(t) = Foe i w t for the partial system of machine plus
stabilizing mass as though resting on firm ground.
Step 2: Calculate the displacement and stress resultants in the structure due to the spring
and damper forces of the isolators treating them as external forces.

As Fig. B5.1 demonstrates, high tuning unlike low tuning - is not capable of reducing
the displacements and stress resultants down to any desired level. The stiff limit of fixed
support corresponds to static loading and is almost reached for a reciprocal tuning ratio of
two. The effect is not particularly sensitive to the degree of damping.
154

B6

Systems with one or two degrees of freedom and linear-elastic properties are discussed
for single impact and - for the SDOF system also for periodic impact loading.

A single impact always constitutes a transient loading, for which a response calculation is
based on the convolution integral of Eq. (B1.37). The time function can be approximated
by idealized impulse shapes, for instance, by a rectangular, triangular, or half-sine impulse
(Fig. B6.1).
The ratio between impact duration t p and structural period T I determines the dynamic
magnification as given in Fig. 3.4 for an undamped SDOF oscillator, i. e. the maximum
displacement relative to the static under peak load. For a rectangular impulse (shape A)
the magnification is largely equal to two, for a half-sine impulse (shape D) it reaches a
maximum of 1.8 at t p == T 1/2.
The influence of damping on the magnification for the half-sine impulse is shown in
Fig. B6.2; it turns out to be negligible for the small damping ratios usually encountered, in
particular for durations t p < ~O.4TI'
If the impact is very short with respect to the fundamental structural period, one can
estimate the dynamic displacement by means of the differential equation of motion for the
Dirac impulse (B1.36) to

I
xmax =m,wI
-- (B6.1)

A
F
B
C
Fig. B6.1
o Types of temporal development
of impact loads:
A rectangular
- B - trapezoidal
C - triangular
D semi-sinusoidal
Structural Behaviour under Impact 155

2.0-,-------,--------------

1.6

...,~ 1.2
(f)
x
'-..
@ 0.8 Fig. B6.2
E
x Response spectra to a
0.4 semi-sinusoidal impulse
for various SDOF periods
0.0 + - - - - - - - r - - - - , - - - - . , - - - - - - , - . - - - - - - - j and damping ratios
0.0 0.4 1.6 2.0

in which I is the momentum

tp

1= f F(t)'dt (B6.2)
o

Note that for a very short relative duration of the impact, the maximum displacement
depends only on the momentum of the impulse irrespective of its shape. According to
Fig. B6.2, this approximation may be used for relative durations up to tJT 1 == 0.1 -:- 0.2.

For calculating a 2-DOF system under impact load, the primary question is that of
possible decoupling as discussed in Section B5.2 (Fig. B5.6).
If the system can be decoupled, the directly hit mass can be treated first according to the
procedure outlined above. The resulting reaction forces are then imposed as loading on the
second mass for calculating the latter's reaction forces and displacements. In any case,
however, a check on the decoupling assumption is recommended.
If the system cannot be decoupled, then the complete system must be analyzed with
two degrees of freedom. For a very short impact in relation to the system frequency
(tJT1 < ~0.1), the magnitude of the momentum can again be assumed to determine the
response. Otherwise, the impact loading has to be followed in the time domain.

In most vibration problems of human or mechanical ongm the impact loading is


periodic, that is a sequence of single impulses occurring with a return period T p'
In Fig. B6.3, the dynamic displacement magnification of an SDOF oscillator is plotted
against the ratio of impulse duration to structural frequency for various impulse shapes
(Fig. B6.1), assuming an impact period of T p = 2tp and ~ = 5% damping.
156 Fundamentals of Vibration Theory

For the particular case of a half-sine impulse, the Figs. B6.4 and B6.5 show the influence
of the relative impact period for two different damping ratios of S = 2% and S = 5%,
respectively.

20..,.---------------------,

16

Fig. B6.3
Response spectra to periodic
impact loads (tp ITp = 0.5)
of the impulse shapes in
O+===-=--,------,-------r-----r-----,,.---~ Fig. B6.1 for 5% damping
0.00 0.75 1.50 2.25 3.00 3.75 4.50
Tp /T1

20-,--------------------.

16

Fig. B6.4
Response spectra to periodic
semi-sinusoidal impact loads
of various durations tp ITp
for 2 % damping
0.75 1.50 2.25 3.00 3.75 4.50
Tp/T 1

20..,.----------------------,

16

~ 12
+>
([)
x
"""-
@
E
8
X

Fig. B6.5
Response spectra as in
O~~~~---===;:::==--__.--_,_--~--__1 Fig. B6.4, but for 5% damping
0.00 0.75 1.50 3.00 3.75 4.50
157

B7

B 7.1

A vibration absorber is a vibratory subsystem attached to and tuned with respect to a


primary vibrating system (a machine, a structure or structural member). Tuning accurately
the frequency of the absorber results in induced inertial forces, which counteract the
vibration of the primary system. While the primary vibration amplitudes can thus be
suppressed to a large degree, considerable displacement amplitudes must be accepted in
the absorber system.

B 7.2 Categories

Vibration absorbers fall into one of two categories (see for example [3.2]):
- linear vibration absorbers (tuned spring-mass-damper system)
shock absorbers.
In each category, sketched in Fig. B7.1, several further distinctions are possible.

I
(a) ~/~~~0{:2:.~::22 (b)

~;~~e%Y
~~~~~ber{ y

primary system

Fig. B7.1 Dynamic model of a linear absorber and a shock absorber

In the following only the working principles of the two absorber types are discussed. It is
recommended that designing a vibration absorber should be based on the relevant litera-
ture.
158 Fundamentals of Vibration Theory

B 7.3

For the primary system with an attached linear absorber (Fig. B7.la), the differential
equation of motion is formulated by analogy to the damped system with two degrees of
freedom (see Section B 2):

(B7.l)

(B7.2)

Parameter subscript S refers to the primary system (i. e. the equivalent SDOF oscillator
of a vibrating structure) and subscript T to the tuned vibration absorber. The degrees of
freedom Xs and XT are the total displacements of the mass of the primary system and that of
the absorber, respectively.
Solving this equation yields the following condition for a vibration absorber tuned to the
optimum:

(B7.3)

From this the optimum absorber frequency is determined as

(B7.4)

It turns out to be somewhat smaller than the frequency of the primary systems ( ~95% )
and to depend on the mass ratio mT/mS'
The graphs on the left hand side of Fig. B7.2 illustrate that the maximum displacement
amplitude of the primary systems can be much reduced near the resonance point, but that it
reacts very sensitively to even minor changes in the absorber frequency (Fig. B7.2 bottom
left). Theoretical calculation alone is insufficient to determine the final absorber parame-
ters; rather, the accurate structural frequency has to be measured in situ, and the absorber
design should allow for refined tuning after the installation. This is most easily done by
altering the absorber mass. Note also, that the absorber hardly affects the frequency
response curve away from the particular resonance point.
The graphs on the righthand side of Fig. B7.2 demonstrate that the absorber displace-
ments may reach a multitude of the structural displacements.
Vibration Absorbers 159

0 0

(0)
0
l!)
C\J , 0
0
C\J

primary system JI tuned absorber


0
0
II 0
0
0 II CD
C\J

/Irwithaut tuned absorber


0 0 0
0
II 0
l!)
II ~
II ~
0 I I (J)
c 0
~ 0
0 J I Q)
0

:s
0
CD
~
X
.~
Y-
o 0
0
0.. 0

E
l!)
"'"
0 xSslol
0
0
/ 0
0

0
0 . 00 0 0 . 00 wj

0 0
0 0
(b) l!) 0
C\J
C\J I
II
primary system tuned absorber
II
0 0
0 II 0
II ~
0
C\J
I~ without tuned absorber
II :::-'
0
0

~
I
I
g 0
0

~
~
I
I
r (J)
o x
,/
0
C 0 0
o
."6
0
~ cD

u
, J
..=
x
I
~ 0
/
0
0
Cl:ci "'"
~ xS stol

gw------ 0
0

0
0 . 00

Fig. B7.2 Frequency response curves of an SDOF oscillator (S = 2%)


with a damped dynamic absorber of mass mT = Ylooms:
(a) optimum tuning: fT/fs = 0.987, ST = 6.6%
(b) poor tuning: fT/fs = 0.900, ST 6.6%

The optimum damping ratio can be determined from

Sopt (B7.5)

or through stepwise variation of the numerical value of S. Figure B7.3 gives an example of
the effect of damping on the displacements of the primary system. In special cases,
damping need not be provided at all, in particular for a constant excitation frequency near
or exactly at the resonance point. See [3.2] for details.
160 Fundamentals of Vibration Theory

o
o
l \ i . , . . . - - - - - - - - , - - - - - - - - - , system mass ms = 21,830 kg WT1wS =0.974
o
absorber mass mT = 500 kg damping: 3% 9% 15%
o
~ without tuned
absorber

o
o
~

c g
o ~ Fig. B7.3
13<..> Example of the frequency
~
0. g response of a primary system
E to
o XS stat as depending on the degree of
I
gU-----
damping of the dynamic absorber
1.50 3.00 4.5 6.00 7.50 9.00

Vibration absorption is more efficient, the larger the mass ofthe absorber is compared to
that of the primary system and the smaller the primary damping. An appropriate choice is
about 1/20 of the primary mass [2.23], a larger absorber mass resulting in a negligible
further reduction of displacements. It follows that absorbers are particularly well suited to
reduce resonating vibrations of structures or structural members with not too large a mass
and small genuine damping; they will prove inadequate for structures having a large mass
or vibrating strongly despite substantial genuine damping.
As an absorber of the type shown above is only effective over a narrow frequency band
and tuned to a particular structural frequency, it is not satisfactory in structures with
several closely spaced eigenfrequencies which are all more or less excited by the dynamic
loading in question.
The application of vibration absorbers will be considered whenever critical dynamic
loading cannot be avoided, especially in existing structures which are difficult to stiffen
economically. It should also be borne in mind, that an absorber is a mechanical system with
large vibration amplitudes, which needs adequate maintenance.

B 7.4 Shock absorbers

Shock absorbers are nonlinear in behaviour and found on machines, chimneys, masts,
etc. As sketched schematically in Fig. B7.1b, they consist of a mass attached to the primary
vibrating system so as to strike on the latter one or two times in each vibration period. If the
absorber force is to act at the right moment to counteract the primary vibration, such a
system also requires careful tuning and maintenance.
161

B8

B ,tlrjplJrBath and lJIICtlll1:V

Under dynamic loading most material properties such as Young's modulus, strength
and strain limits - change to a greater or lesser extent, when compared to the values
characteristic for slow, quasi-static loading. The change accelerates with the rate ofloading
and is usually expressed in function of the strain rate
. dE
E=- (B8.1)
dt

Most often an average strain rate of loading can be considered. In inertial vibration
problems it hardly exceeds E :::= 0.1 S-l, so that the expected change in properties is
moderate [1.2]. A much larger influence is to be noted for strain rates E :::= 1 -7- 10 s-t, as
occur typically for high impact loads and will generally lead to plastic deformation in the
structure or the member affected.
Dynamic loads may also influence the fatigue resistance of the structural material, even
at a low number of cycles if the load is large enough (low-cycle fatigue).
The property changes for concrete and steel are briefly discussed in the fo1l9wing.
Usually they describe a straight line when plotted against the strain rate in semi-logarith-
mic scale and are given in form of a factor with respect to the quasi-static value. This
approach will be followed here to indicate the general trend of change; a more extensive
and quantitative discussion can be found in [1.2]. As the influence of the strain rates
encountered for typical dynamic loads (disregarding strong impact, blast loading, etc.) is
limited to a few percent, it may be neglected in most design cases.

Tensile strength of concrete


The effect of the strain rate is relatively large. For instance, a rate E 0.1 S-l will result in
some 60% increase compared to the quasi-static tensile strength. Above ca. E = 0.3 S-l the
increase with strain rate is even more rapid.

Compressive strength of concrete


The compressive strength increases with higher strain rate at a smaller percentage than
the tensile strength. The increase at E= 0.1 S-l amounts to about 10%.
162 Fundamentals of Vibration Theory

Modulus of elasticity of concrete


The effect on Young's modulus is the same as on the compressive strength with about
10% at t 0.1 S-l.

Strain capacity of concrete


The strain capacity is affected by the strain rate to various degree. While it increases with
the strain rate in tension, conclusions for the compression domain are uncertain due to
inconsistent test results.

Strength and modulus of elasticity of reinforcing steel


In the relevant range of strain rates up to about t = 0.1 S-l no strength increase should be
expected. The same applies to the Young's modulus of elasticity.

Strain capacity of reinforcing steel


Especially cold-worked steel proves sensitive to the strain rate. The change in uniform
strain or the elongation at fracture [1.2], however, is not relevant in the present context,
since only elastic deformation is desired.

Fatigue resistance of reinforcing steel


As for concrete, the reader is referred to the national standards and the literature.

".... 'r>HrdJ'n Members


.t1~ej]nt()rced-.fi.n{'>r,p1i,p

The dynamic analysis of reinforced concrete structures, in particular when calculating


the eigenfrequencies and the expected displacements, relies heavily on the flexural rigidity
of the structural members. In general, the state of the structure, whether cracked or
uncracked, is not known, neither whether a transition will occur under the loading in
question. To be conservative, a sensitivity study with bounds on the cracked and the
uncracked zones should be carried out.
The bounds on the flexural rigidity E I of reinforced concrete sections - uncracked and
cracked state can be determined as follows [B8.1]:

uncracked state (<<state I):

(B8.1)

in which:
Dynamic Material Properties 163

== 1.10Ee,stat
dynamic modulus of elasticity of concrete (see Section B 8.1)
If moment of inertia of the uncracked section (possibly including
reinforcing bars)

- completely cracked state (<<state II without tension stiffening):

(B8.2)

in which:
If! moment of inertia of the cracked section (including the rein-
forcing bars in the tension zone)

- effective flexural rigidity of the cracked section with tension stiffening by the concrete in
between cracks:

BII --
BII (B8.3)
eff - 'X

in which the so-called bond coefficient 'X is computed as

MerJ2.
[M
stat
[1 - B~IJ 1
B
:::::; (B8.4)

where:
Mer cracking moment of the cross-section (uncracked stress
distribution according to the tensile strength of concrete)
M stat bending moment due to static loading (at rest).

'~-~ Mdyn,mox
'----.f--\---I--'>- M stot (at rest)
~1""'--H'-7-----'-'----"-L--Mdyn , min
Fig. B8.1
Moment-curvature relation
and secant stiffnesses of
curvature a cracked reinforced-concrete
arctg BI
beam element under dynamic
loading (starting from static load)
164 Fundamentals of Vibration Theory

The flexural rigidity will hardly fall below B~h (towards B II) unless the cracks are spaced
very closely (e. g. due to very small spacing of stirrups) or the bond is destroyed over long
distances due to large bar diameters and a high number of load cycles. Regarding the
moment-curvature diagram in Fig. B8.1, the secant values of the flexural rigidity can be
quite different for the upper and lower load maximum in a cycle. As a reasonable
approximation, one may take the average rigidity corresponding to the static load level
when the structure is at rest.
In a refined analysis one may take into account the existence of uncracked zones where
the stressing is low, e. g. toward the supports, although these zones do not contribute much
to the total deformation.

Inherent structural damping is very important for the dynamic behaviour. Several
mechanisms can be distinguished (e. g. [B8.2]):
- Material damping denotes the energy dissipation inside a monolithic structure.
- System damping results from energy dissipation in appurtenances, for instance friction in
sliding bearings and joints, energy dissipation in nonstructural elements.
- Radiation damping describes the effect that energy is transmitted from the structure
through the foundation and subsequently radiated into the halfspace of the ground in the
form of wave motion.

Each of these mechanisms contribute to the damping value (logarithmic decrement 0)


measured in a vibration decay test on structures in situ. Cracking in reinforced concrete
structures may also contribute a large share [B8.1], [B8.2], [B8.3]. The damping is further
raised by the presence of many people on the vibrating structural member (see Section
2.2).

In the present context the main distinction is to be drawn between the damping in
- pedestrian structures (overpasses, walkways, stairs, etc.)
- buildings (gymnasia and sports halls, dancing and concert halls, industrial buildings,
etc. ).
Pedestrian structures are often lightly damped because of the few nonstructural elements
(surfacing, railing) which could contribute to system damping. Figure B8.2 gives a distribu-
tion of the damping values (logarithmic decrement) found in 44 investigated pedestrian
bridges [2.5]. The damping values for the basic material, for beams with frictionless
bearings and for bridges, respectively, are given in Tab. B8.1. Another investigation gave
similar results [2.6]. Despite a considerable scatter the difference in construction produces
Dynamic Material Properties 165

0.2 all bridges 0.2 composite

0.1 0.1

0.04 0.08 0.12 0.04 0.08 0.12

0.2 concrete 0.2 steel


Fig. B8.2
0.1 0.1
Statistical distribution of
the degree of damping found
0
in 44 different footbridges
0 0.04 0.08 0.12 0 0.04 0.08 0.12
(after [2.5]):
damping (logarithmic decrement) total, and groups according
to construction material

a systematic trend from non-prestressed concrete structures with the highest damping to
welded steel structures with the lowest damping.

basic material beams bridges


steel 0.002 -=- 0.008 0.004 -=- 0.03 0.02 -=- 0.06
concrete 0.01 -=- 0.06 0.02 -=- 0.06 0.02 -=- 0.2
Tab. B8.1 Damping in various construction materials and in structures
due to contributions of system damping (from [2.5]);
values are logarithmic decrements 6 (Eq. (B1.13))

In the absence of better information, the following damping ratios may be assumed for
calculation purposes:
reinforced concrete structures ; 0.7%
- prestressed and composite (concrete slab on steel girder) ; = 0.5%
- steel structures ; = 0.3%
- timber structures ; = 1.5%

In the vast majority of cases these values should give conservative lower bounds, the
average values being possibly twice as high.

The larger damping found in buildings is due to energy dissipating floor cover, cladding,
nonstructural partitions, installations, stored goods, etc. However, the high damping
values quoted in handbooks are stated for earthquake excitation and should never be used
166 Fundamentals of Vibration Theory

for calculating the vibration response to human or mechanical excitation. Without any
better knowledge, the damping in buildings may be assumed to be double the values for
pedestrian structures recommended above. Again, this should result in conservative lower
bounds, while the average values in buildings could be three times the values for pedestrian
structures recommended above.
Provided that advantageous influences such as vibrating frame legs embedded in the
ground, board flooring with motion relative to the structure, numerous partitions and
equipment, etc. are obviously present to a large extent, then the values recommended
could be raised individually.
167

structural dvnalmllCS

[1.1] Ammann W., Muhlematter M., Bachmann H.: Tensile tests on reinforcing and
prestressing steel at elevated strain rates (in German), Institute of Structural
Engineering, Swiss Fed. Inst. of Techn. Zurich, report 7709-1, Birk-
hauser, Basle/Boston/Stuttgart, June 1982.
[1.2] Ammann W.: Reinforced and prestressed concrete structures under impact
loading (in German), doctoral diss., Institute of Structural Engineering, Swiss
Fed. Inst. of Techn. (ETH) Zurich, report 142, Birkhauser, Basle/Boston/Stutt-
gart, June 1983.

Man-induced -.:,;;1........,11-;;"'..... '"

[2.1] Bachmann H. : Vibrations of building structures caused by human activities,


case study of a gymnasium, Nat. Res. Council of Canada, Techn. Transl. 2077,
1984 - Orig. in German, J. SIA (Schweizer Ingenieur und Architekt), vol. 101,
no. 6, pp. 104-110, 1983.
[2.2] Matsumoto Y., Nishioka T., Shiojiri H., Matsuzaki K.: Dynamic design of
footbridges, IABSE Proceedings P-17178, Aug. 1978.
[2.3] Schulze H. : Dynamic effects of the live load on footbridges (in German), Signal
und Schiene, vol. 24, no. 2, pp. 91-93, and no. 3, pp. 143-147, 1980.
[2.4] Kramer H., Kebe H. W.: Man-induced structural vibrations (in German), Der
Bauingenieur, vol. 54, no. 5, pp. 195-199, 1979.
[2.5] Tilly G. P., Cullington D. W., Eyre R.: Dynamic behaviour of footbridges,
IABSE Surveys S-26/84, May 1984.
[2.6] Wheeler J. E.: Prediction and control of pedestrian induced vibration in foot-
bridges, J. Struct. Div. ASCE, vol. 108, ST 9, pp. 2045-2065, 1982.
[2.7] Galbraith F. W., Barton M. V.: Ground loading from footsteps, J. Acoust.
Soc. of America, vol. 48, no. 5 (part 2), pp. 1288-1292,1970.
[2.8] Harper F. C.: The mechanics of walking, Res. Appl. in Industry, vol. 15, no. 1,
pp. 23-28, 1962.
[2.9] Blanchard J., Davies B.L., Smith J.W.: Design criteria and analysis for
dynamic loading of footbridges, Symposium on Dynamic Behaviour of Bridges,
suppl. report 275, TRRL, Berkshire (D. K.), May 1977.
68 References

2.10] Baumann K., Bachmann H.: Dynamic loading induced by persons and its effect
on beam structures (in German), Institute of Structural Engineering, Swiss Fed.
Inst. of Techn. (ETH) Zurich, report 7501-3, Birkhauser, Basle, 1987.
2.11] Jones R. T., Pretlove A. J.: Vibration absorbers and bridges, Highway
Engineer, vol. 26, no. 1, pp. 2-9, 1979.
2.12] Petersen C.: Theory of random vibrations and application (in German), Work
Reports on the Safety Theory of Structures, no. 2/72, Struct. Eng. Laboratory,
Techn. Vniv. Munich (FRG), Nov. 1972.
2.13] GDR State Construction Superv. Board, Regulation SBA 123/82: Traffic and
pedestrian bridges, vibration tests (in German), GDR Ministry of Transporta-
tion, East Berlin, July 1982.
~.14] Rainer J. H.: Dynamic response of a gymnasium floor, Build. Res. Note 213,
Div. of Build. Res., Nat. Res. Council of Canada, Ottawa, May 1984.
~.15] Allen D. E., Rainer J. H., Pernica G.: Vibration criteria for assembly occupan-
cies, Can. J. Civil Eng., vol. 12, no. 3, pp. 617-623,1985.
~.16] National Building Code of Canada, NBC 1985, Commentary A: Serviceability
criteria for deflections and vibrations, 1985.
~.17] Pernica G.: Dynamic live loads at a rock concert, Can. J. Civil Eng., vol. 10,
no. 2, pp. 185-191, 1983.
~.18] Nilsson L.: Impact loads produced by human bodies, Swedish Council for
Build. Res., docum. D20, 1980.
~.19] VEAtc: Directives VEAtc on shock loading on vertical opaque structures (in
French), VEAtc, vol. 228, no. 1768, June 1981.
~.20] Mann W.: Impact forces due to falling objects and persons, and related collapse
of a falsework structure (in German), Die Bautechnik, vol. 56, no. 5, pp. 169-
171, 1979.
~.21] Struck W., Limberger E.: Loading of horizontal structural members by the foot
impact of a person (in German), Die Bautechnik, vol. 58, no. 10, pp. 347-351,
1981.
:.22] Hauck G.F.W.: Hyatt-Regency walkway collapse: Design alternates,
J. Struct. Div. ASCE, vol. 109, ST 5, pp. 1226-1234,1983.
,.23] Gerasch W.J., Natke H. G.: Vibration reduction of two structures, Int. Symp.
on Vibration Protection in Construction, sc. report, vol. 1, pp. 132-142, Lenin-
grad (VSSR) , 1984 .
.24] Vilnay 0.: Active control of machinery foundation, J. Eng. Mech. Div. ASCE,
vol. 110, EM 2, pp. 273-281, 1984.
.25] British Standards Institution, BS 5400, Part 2, Appendix C: Vibration servicea-
bility requirements for foot and cycle track bridges, 1978.
.26] Cantieni R.: Investigations on floor vibrations in an office building (in Ger-
man), report 116/2, Swiss Fed. Lab. for Material Testing (EMPA) Dubendorf,
Sept. 1985.
References 169

[2.27] TGB 1972, NEN 3850: Regulations for the calculation of building structures -
General considerations and loading (in Dutch), Nederlands Normalisatie-
Instituut, Rijswijk (NL), 1972.
[2.28] National Building Code of Canada, NBC 1980, Commentary A: Serviceability
criteria for deflections and vibrations, Suppl. Chapter 4, 1980.
[2.29] Mayer H.: Vibrational behaviour of high-diving platforms in outdoor and
indoor swimming pools (in German), Report on Acad. Year 1970171, Swiss
State Techn. School (HTL) Brugg-Windisch, 1971.
[2.30] Rainer J. H., Pernica G.: Vertical dynamic forces from footsteps, Canadian
Acoustics, vol. 14, no. 2, pp. 12-21, 1986.

l:hapter 3:

[3.1] Institution of German Engs. (VDI), Directive 2060: Criteria for evaluating the
balancing of rotating stiff machinery parts (in German), VDI Verlag, Dussel-
dorf, Oct. 1966.
[3.2] Korenev B. G., Rabinovic 1. M.: Structural dynamics handbook (in German),
VEB Verlag furs Bauwesen, East Berlin (GDR), 1980.
[3.3] Newcomb W. K.: Principles of foundation design for engines and compressors
(incl. discussion), ASME Trans., vol. 73, no. 3, pp. 307-318,1951.
[3.4] Natke H. G., Thiede R., Elmer K.: Investigations of vibrations emitted from
weaving mills and recommendations for planning weaving mill floors (in Ger-
man), Assoc. of the North-West German Textile Industry, publ. no. 66, Munster
(FRG), 1985.
[3.5] Muller F. P.: Calculation and construction of bell towers (in German),
W. Ernst & Sohn, Berlin/Munich, 1968.
[3.6] Muller F. P.: Structural dynamics (in German), Betonkalender 1978, part H.E,
pp. 745-962, W. Ernst & Sohn, Berlin/Munich/Dusseldorf, 1978.
[3.7] Muller F. P., Keintzel E., Charlier H.: Dynamic problems in reinforced con-
crete construction - Part 1: RIC material behaviour under dynamic loading (in
German), German Committee on Reinforced Concrete (DAfStb), vol. 342,
1983.
[3.8] Harris C. M., Crede C. E.: Shock and vibration handbook, 2nd ed., McGraw
Hill Int., New York, 1976.
[3.9] Lipinski J.: Foundations and support structures for machines (in German),
Bauverlag, Wiesbaden (FRG), 1972.
[3.10] German Inst. for Standards, DIN 4178: Bell towers Calculation and construc-
tion (in German), ed. Aug. 1978.
[3.11] Cremer L., Heckl M.: Structure-borne sound (in German), Springer, Berlin/
Heidelberg/New York, 1982.
170 References

[3.12] Studer J.A., Ziegler A.: Soil dynamics Fundamentals, properties, problems
(in German), College Textbook Series, Springer, Heidelberg/New York/Tokyo,
1986.
[3.13] Clough R. W., Penzien J.: Dynamics of structures, McGraw Hill Kogakusha,
Tokyo, 1975.
[3.14] Wolf J. P.: Dynamic soil-structure interaction, Prentice Hall, Englewood Cliffs
/ N.J., 1985.
[3.15] Gazetas G.: Analysis of machine foundations: State of the art, Soil Dyn.
Earthqu. Eng., vol. 2, no. 1, pp. 2-42, 1983.

[4.1] German Inst. for Standards, DIN 4150: Vibrations in civil engineering - Part 2:
Effects on people in buildings (in German), Sept. 1975.
[4.2] Richart F. E., Woods R. D., Hall J. R.: Vibrations of soils and foundations,
Prentice Hall, Englewood Cliffs / N.J., 1970.
[4.3] Raab A., Widmer R.: Effects of vibrations on buildings Reference values in
various countries (in German), J. SIA (Schweizer Ingenieur und Architekt),
vol. 98, no. 4, pp. 45-48, 1980.
[4.4] German Inst. for Standards, DIN 4150: Vibrations in civil engineering Part 3:
Effects on structures (in German), draft, March 1983.
[4.5] Arnold K. H.: The consequences arising from the yellow-print draft of DIN 4150
'Vibrations in civil engineering, part 3: Effects on structures' for carrying out
blasting operations (in German), Bull. Nobel Comp., pp. 58-62, July/Dec. 1983.
:4.6] Institution of Swiss Highway Engs. (VSS), Swiss Standard SN 640312: Vibration
effects on structures (in German), VSS secretariat Zurich, Nov. 1978.
-4.7] GDR Chamber of Technology, Directive KDT 046/72: Effects of blasting oper-
ations on buildings (in German), East Berlin (GDR), 1972.
4.8] Wustenhagen K.: The skeleton of vibration protection in the GDR as rep-
resented in the current KDT Directive 'Effects of blasting operations on build-
ings' (in German), Bull. Nobel Comp., pp. 29-34, Jan./March 1978.
4.9] Int. Standard Organization, ISO/DIS 4866: Mechanical vibrations and shock-
measurement and evaluation of vibration effects on buildings - Guidelines for the
use of basic standard methods, draft, July 1984.
4.10] Wiss F., Parmelee R. A.: Human perception of transient vibrations, J. Struct.
Div. ASCE, vol. 100, ST 4, pp. 773-787, 1974.
4.11] Gierke H. E. von, Goldman D. E.: Effects of shock and vibration on man, in:
Shock and vibration handbook (Eds. Harris C. M., Crede C. E.), 2nd ed.,
McGraw Hill Int., New York, 1976.
4.12] Ministry of Metallurgy and Chern. Eng., Instruction I 200-54: Planning and
calculation of building structures for dynamic loading from machinery (in Rus-
sian), Moscow (USSR), 1955.
References 171

[4.13] Int. Standard Organization, ISO 2631: Guide for the evaluation of human
exposure to whole-body vibration, ISO Standards Handbook 4: Acoustics,
vibration and shock, 1st ed., ISO secretariat, Geneva (Switz.), pp. 493-507,
1980.
[4.14] Institution of German Engs. (VDI), Directive 2057: Evaluation of mechanical
vibration effects on people, Folio 1: Fundamentals, classification, terminol-
ogy, draft, Dec. 1983; Folio 2: Vibration effects on the human body, May
1981; Folio 3: Vibrational strain on man, draft, Feb. 1979; all in German, VDI
Verlag, Dusseldorf (FRG), Feb. 1979.
[4.15] Building Research Establishment, BRE Digest 278: Vibrations: Building and
human response, Building Research Station, Garston, Watford (U. K.), Oct.
1983.
[4.16] Int. Standard Organization, ISO 2372: Mechanical vibration of machines with
operating speeds from 10 to 200 rev/s - Basis for specifying evaluation standards,
ISO Standard Handbook 4: Acoustics, vibration and shock, 1st ed., ISO
secretariat, Geneva (Switz.), pp. 423-431, 1974.
[4.17] Int. Standard Organization, ISO 2373: Mechanical vibration of certain rotating
electrical machinery with shaft heights between 80 and 400 mm. Measurements
and evaluation of the vibration severity, ISO Standard Handbook 4: Acoustics,
vibration and shock, 1st ed., ISO secretariat, Geneva (Switz.), pp. 433-437,
1974.
[4.18] Institution of German Engs. (VDI), Directive 2056: Guideline for evaluating
mechanical vibrations from machinery (in German), VDI Verlag, Dusseldorf
(FRG), Oct. 1964.
[4.19] Institution of German Engs. (VDI) , Directive 2063: Measurement and evalua-
tion of mechanical vibrations from vertical-piston machines (in German), draft,
VDI Verlag, Dusseldorf (FRG), Aug. 1982.
[4.20] Isaacs A. D., Klock F. G., Koop C. D., Mitchell J. D., Snyder W. F., Winchell
J. A.: Recommended environments for standards laboratories, ISA Trans.,
vol. 3, no. 4, pp. 366-377, 1964.
[4.21] Eshleman, R. L. : Vibration standards, in: Shock and vibration handbook
(Eds. Harris C.M., Crede C.E.), 2nd ed., McGraw Hill Int., New York, 1976.
[4.22] Int. Standard Organization: Acoustic and vibration: National and international
standards, ISO secretariat, Geneva (Switz.), 1978.
[4.23] Reiher H., Meister F. J.: The human sensitivity to vibrations (in German),
Forschung im Ingenieurwesen, vol. 2, no. 11, pp. 381-386, 1931.
[4.24] Major A.: Dynamics in civil engineering - Analysis and design, Akademial
Klado, Budapest (Hung.), 1980.
[4.25] British Standards Institution, BS 6472: British Standard guide to evaluation of
human exposure to vibration in buildings (l Hz to 80 Hz)>>, 1984.
172 References

[ALl] Communication from Dr. H. W. Kebe, structural dynamics consultant, Hamburg


(FRG).
[A2.1] Communication from Mr. H. Frey, Frey & Associes Eng. Office, Lausanne
(Switz. ).
[A9.1] Communication from Dr. D. E. Allen, Nat. Res. Council of Canada, Ottawa /
Ontario (Canada).
[AI0.l] Communication from Dr. R. Suter, Institute of Reinforced and Prestressed
Concrete Structures (IBAP), Swiss Fed. Inst. of Techn. (EPF) Lausanne.
[A15.1] Company bulletin no. 25, Stahlton AG, Zurich.
[A20.1] Communication from Mr. A. Iseli, Gartenmann Eng. Office, Berne (Switz.).

lDI)enalX B:

[B1.1] Bathe K. J.: Finite element procedures in engineering analysis, Prentice Hall,
Englewood Cliffs / N. J., 1982.
[B3.1] Pitloun R.: Vibrating beams Computation tables (in German), VEB Verlag
furs Bauwesen, East Berlin (GDR), 1970.
[B3.2] Pitloun R.: Vibrating frames and towers - Computation tables (in German),
VEB Verlag furs Bauwesen, East Berlin (GDR) , 1975.
[B3.3] Unger H.: The air raid shelter handbook (in German), B. G. Teubner, Leipzig
(GDR), 1961.
[B8.1] Bachmann H.: Reinforced concrete, part 1 (in German), lecture notes, Insti-
tute of Structural Engineering, Swiss Fed. Inst. of Techn. (ETH) Zurich, 1985.
[B8.2] Dieterle R.: Models for the damping behaviour of vibrating reinforced concrete
beams in the uncracked and cracked conditions (in German), doctoral diss.,
Institute of Structural Engineering, Swiss Fed. Inst. of Techn. (ETH) Zurich,
report 111, Birkhauser, Basle/Stuttgart, April 1981.
[B8.3] Dieterle R., Bachmann H.: Experiments and models for the damping behaviour
of vibrating reinforced concrete beams in the uncracked and cracked conditions,
IABSE ColI. Delft on Advanced Mechanics of Reinforced Concrete, final report
AK 34, Oct. 1981.
173

acceleration amplitude; slab dimension


real discrete Fourier coefficients
damping constant
ccrit critical damping for aperiodic vibration (= 2-m-w)
d displacement amplitude, deflection
e eccentricity of mass
fb operating frequency of machine
f1,g+p fundamental structural frequency incl. life load
f/f 1 tuning ratio
fm frequency of music beat
fn natural frequency of structure in mode n
fs pacing rate, skipping frequency
g acceleration of gravity (= 9.80665 m/s2)
imaginary number (= v=1)
spring stiffness
dynamic impact factor (= FJG)
length
stride length in walking or running
mass; factor for number of pedestrians
unbalanced rotating mass
speed of revolution of machine parts
order of highest relevant harmonic of the loading function
load per unit length
radius
complex eigenfrequency
safety factor in frequency tuning
time variable; thickness of slab
rise time of load
duration of contact or impulse
time since end of impulse
u complex displacement variable
v velocity amplitude
forward speed of pedestrian
deflection along beam axis
real displacement variable; coordinate along beam axis
y modal vibration coordinate
174 Notation

A cross-sectional area
complex Fourier amplitude
flexural rigidity (=
damping matrix
Young's modulus of elasticity
j-th harmonic force component
amplitude of a forcing function (frequency independent)
amplitude of quadratic excitation (frequency dependent)
dynamic forcing function
G weight force; shear modulus
amplitude of the n-th harmonic load component
momentum of impact load; moment of inertia
torsional moment of inertia
stiffness matrix; intensity of human perception (= KB)
mass matrix; bending moment in statics
cracking moment
moment of resistance in plastic hinge or yield line
( - s over the support, - i l l in midspan)
P concentrated load
R reaction force of structure or support
S dynamic stiffness; tension force in string
T period of oscillation, vibration 1/f)
To time necessary to cross a bridge
Tp return period of impulse
VA,... complex displacement amplitude constants
Vj complex displacement amplitude at frequency Wj
V dynamic reaction force
Vk dynamic magnification for constant-load excitation
Vq dynamic magnification for quadric excitation
Vs dynamic magnification for impact
VR transmissibility, i. e. ratio of reaction force to load amplitude
1-VR isolation effectiveness
WA ,.. deflection shape constants
X A ,.. real displacement amplitude constants

a sector angle
coefficients of Rayleigh damping
vibration shape parameter of beams
aspect ratio of plates (= alb)
logarithmic decrement 2-n-s)
E strain
X bond coefficient in tension stiffening
Notation 175

A flow rate of pedestrians; frequency parameter for beams


[! mass per unit length or unit area
v Poisson's ratio 0.1570.20 for uncracked concrete)
Jt ratio of circle perimeter to diameter 3.14159)
~ damping ratio with respect to critical damping (= c/(2mw))
Q mass per unit volume
CPL,.. factors for substitute SnOF system
CPn frequency parameter for n-th mode of vibration;
phase angle of n-th harmonic (with respect to first)
CPo phase angle of excitation frequency
CPl phase angle of snop system
'X rotational inertia per unit length
't/Jk k-th eigenmode shape vector
'If matrix of eigenmode vectors (modal matrix)
WI circular eigenfrequency of undamped snop system (= Vk/m)
WD circular eigenfrequency of damped snop system
Wj circular eigenfrequency of j-th harmonic load component
Q matrix of eigenfrequencies

adm admissible
c concrete
dyn dynamic
eff effective
order of natural vibration of structure
j order of harmonic component (complex)
k number of mode shape
1 single-degree-of-freedom (SnOP) parameter
max maximum, peak value
min minimum
n order of harmonic component (real)
o excitation parameter
opt optimum
p impulse, impact parameter
s walking, skipping parameter
stat static, at-rest position
X,y,2 spatial components
M machine
S structural system
T tuned vibration absorber

You might also like