Вы находитесь на странице: 1из 60

DYNAMIC COMFORT CRITERIA FOR STRUCTURES

A review of UK codes, standards and advisory documents


The effects of vibration are becoming an increasingly important issue in the
design of buildings and building elements. Modern construction methods
mean that buildings are becoming lighter and have less structural damping.
The response of such buildings to imposed vibration is therefore increased,
DYNAMIC COMFORT CRITERIA
so people using the buildings are more likely to experience vibration. Over a
number of years, different serviceability criteria have been developed to ensure
that buildings and building elements are suitable for their intended activity.
FOR STRUCTURES
This BRE Trust Report draws together the themes and disparate (sometimes
A review of UK codes, standards and advisory documents
conflicting) acceleration comfort criteria associated with the vibration
of buildings in the UK. It reviews Eurocodes (including the UK National
Annexes), British Standards, International (ISO) Standards and other sources
of information. The UK Building Regulations and advice provided by UK trade
Gordon Breeze
associations and other industry bodies are also considered.

This report is technical in nature, and is aimed primarily at building designers,


consultants, architects and structural engineers. It is hoped that the complex
and interlinking nature of this subject can be understood by presenting all the
relevant acceleration criteria in one document in a logical and concise way.

y x
RELATED TITLES FROM IHS BRE PRESS
BUILDING-MOUNTED MICRO-WIND TURBINES ON HIGH-RISE AND COMMERCIAL BUILDINGS
FB 22, 2010
THE RESPONSE OF STRUCTURES TO DYNAMIC CROWD LOADS, 2004 EDITION
DG 426, 2004
WIND LOADS ON UNCLAD STRUCTURES
SD 5, 2004

IHS BRE Press, Willoughby Road


Bracknell, Berkshire RG12 8FB
www.brebookshop.com
FB 33
DYNAMIC COMFORT CRITERIA
FOR STRUCTURES
A review of UK codes, standards and advisory documents

Gordon Breeze
ii

This work has been funded by BRE Trust. Any views The information and data presented in this report are up
expressed are not necessarily those of BRE Trust. While to date at the time of publication. Over time, values given
every effort is made to ensure the accuracy and quality of in standards and codes of practice change. Nevertheless,
information and guidance when it is first published, BRE the themes underpinning the methodologies described
Trust can take no responsibility for the subsequent use of are likely to remain the same for the foreseeable future.
this information, nor for any errors or omissions it may
contain. For the purpose of drawing comparisons and presenting
different approaches, information from many referenced
The mission of BRE Trust is ‘Through education and sources is quoted in this document. BRE does not
research to promote and support excellence and accept liability for any information presented for this
innovation in the built environment for the benefit of purpose, nor does BRE endorse such information. In all
all’. Through its research programmes the Trust aims to circumstances, it is recommended that the latest, up-to-
achieve: date, primary sources of information are consulted to
• a higher quality built environment ensure compliance with required standards or codes of
• built facilities that offer improved functionality and practice.
value for money
• a more efficient and sustainable construction sector,
with
• a higher level of innovative practice.

A further aim of BRE Trust is to stimulate debate on


challenges and opportunities in the built environment.

BRE Trust is a company limited by guarantee, registered


in England and Wales (no. 3282856) and registered
as a charity in England (no. 1092193) and in Scotland
(no. SC039320).

Registered Office: Bucknalls Lane, Garston, Watford,


Herts WD25 9XX

BRE Trust
Garston, Watford WD25 9XX
Tel: 01923 664743
Email: secretary@bretrust.co.uk
www.bretrust.org.uk

BRE Trust and BRE publications are available from


www.brebookshop.com
or
IHS BRE Press
Willoughby Road
Bracknell RG12 8FB
Tel: 01344 328038
Fax: 01344 328005
Email: brepress@ihs.com

Requests to copy any part of this publication should be


made to the publisher:
IHS BRE Press
Garston, Watford WD25 9XX
Tel: 01923 664761
Email: brepress@ihs.com

Printed on paper sourced from responsibly managed Cover images:


forests Main: BRE laser system used to measure pinnacle
vibration
FB 33 Top right: Grandstands at Everton Football Club (courtesy
© Copyright BRE 2011 of Everton Football Club)
First published 2011 Middle right: Geocentric coordinate axis system
ISBN 978-1-84806-173-6 Bottom right: A ‘grandstand’ rotary shaker used by BRE
CONTENTS iii

CONTENTS

Executive summary � iv

1 INTRODUCTION � 1

2 BACKGROUND INFORMATION � 3

2.1 Preamble 3

2.2 Units of acceleration 3

2.3 Coordinate axis systems 3

2.4 Acceleration vectors 4

2.5 Weighting factors 4

2.6 Measures of acceleration 5

3 GENERAL ACCELERATION CRITERIA � 7


4 HORIZONTAL ACCELERATION CRITERIA � 9



4.1 Preamble 9

4.2 Wind criteria 9

4.3 Criteria not related directly to wind 10

5 VERTICAL ACCELERATION CRITERIA � 18



5.1 Preamble 18

5.2 Eurocodes 18

5.3 International (ISO) Standards 21

5.4 British Standards 22

6 SPECIALISED BUILDINGS � 24

6.1 Grandstands and sports stadia 24

6.2 Hospitals 25

6.3 Car parks 27

7 � GUIDANCE GIVEN BY UK BUILDING REGULATIONS, TRADE ASSOCIATIONS 28



AND OTHER INDUSTRY BODIES

7.1 UK Building Regulations 28

7.2 Trade associations and other industry bodies 28

8 � DYNAMIC TESTING OF BUILDINGS AND BUILDING ELEMENTS 32



8.1 Preamble 32

8.2 Introduction 32

8.3 Commonly used test methods 33

8.4 Practical use of dynamic testing 38

9 � CONCLUSIONS 40

10 � REFERENCES 41

Appendix A Vectors and vector addition 44



Appendix B Buildings with complex vibration motions 46

Appendix C Acceleration weighting factors 49

iv EXECUTIVE SUMMARY

EXECUTIVE SUMMARY

The effects of vibration are becoming an increasingly The present status of dynamic testing (both laboratory
important issue in the design of buildings and building and offsite) is also discussed, as well as the underlying
elements. Modern construction methods mean that principles of commonly used test methods. Information is
buildings are becoming lighter and have less structural provided about what can be measured and what can be
damping. The response of such buildings to imposed inferred from those measurements. This part of the report
vibration is therefore increased, so people using the gives practical advice, and full-scale measurements are
buildings are more likely to experience vibration. Over a presented that illustrate the issues raised.
number of years, different serviceability criteria have been This review is technical in nature, and is aimed
developed to ensure that buildings and building elements primarily at building designers, consultants, architects
are suitable for their intended activity. and structural engineers. It is hoped that the complex
The purpose of this BRE Trust Report is to distil and and interlinking nature of this subject can be understood
draw together the themes and disparate (sometimes by presenting all the relevant acceleration criteria in one
conflicting) acceleration comfort criteria associated with document in a logical and concise way.
the vibration of buildings in the UK. This report considers
the Eurocodes (including the UK National Annexes),
British Standards, International (ISO) Standards and other
sources of information. The UK Building Regulations
and advice provided by UK trade associations and other
industry bodies are also considered.
page header right – page1header
iNtrOdUC
subti
tiON
tle 1

1 iNtrOdUCtiON

The effects of vibration are becoming an increasingly important issue in the
design of buildings and building elements. Modern construction methods
mean that buildings are becoming lighter and have less structural damping.
The response of such buildings to imposed vibration is therefore increased,
so people using the buildings are more likely to experience vibration. If
the vibration is large enough it can cause annoyance, motion sickness and
ultimately panic. Over a number of years, different serviceability criteria have
been developed to ensure that buildings and building elements are suitable for
their intended activity.
Issues concerned with structural integrity (and hence safety) are not
considered in this report. When a properly designed structure is subjected
to vibration, people tend to become uncomfortable well before the limiting
state of the design is reached. Hence, the serviceability requirements tend to
be the critical factor in structures with human occupants. Although there are
serviceability criteria associated with the appearance of building finish and
public health, this report is limited to a consideration of the intrusive effect of
vibration upon people.
The aim of serviceability criteria is to ensure that a structure is suitable for
its intended use. Within the population there is a wide range of sensitivity
to vibration. Generally, children are the most sensitive to vibration and adult
males the least sensitive. Both the activity and posture of a person affect their
perception of vibration. A person’s expectation and exterior cues (such as
sound) are also important factors. This wide number of complicating factors
means that serviceability criteria can never be precise: what one person
judges to be a tolerable level of vibration may be perceived by another
to be annoying. Therefore, judgements often need to be made about the
percentage of the population that would perceive the vibration conditions to
be unacceptable; these judgements can have significant financial implications.
For this reason, many codes of practice recommend that agreement is
reached between the designer and the client beforehand about limiting
vibration threshold levels.
As noted above, vibration serviceability criteria have been developed
over many years. Some of these criteria have been developed for specific
applications, whereas others have more general application. Although there
are many similarities, there are also differences in the approaches used. Some
of the methods relate to specified return time periods (eg one-year, five-year
or 10-year periods), other methods relate to specific periods of vibration (eg
a 16-hour day or eight-hour night). There is uncertainty about how criteria
developed on the basis of simple vibration modes can be extended or
interpreted for cases where a structure vibrates in a more complex way. Some
methods give ranges of limiting accelerations, whereas other methods give a
specific value. Also, some approaches conflict with other methods, and no
two methods give consistent agreement with each other. There is not even
consistent agreement about the units of acceleration, although the units are
easily converted.
Many countries (eg the UK, the USA, Canada, Australia, New Zealand, The
Netherlands, Japan, France, Germany and Denmark) have their own national
2 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

codes of practice and standards, within which are contained building-related


serviceability criteria. It is not the purpose of this report to compare and
contrast the criteria adopted by each country. Instead, this study focuses on
identifying and reviewing information that is relevant to buildings in the UK.
Nevertheless, where appropriate, attention is drawn and references made to
sources of information from other countries.
The UK falls under the scope of the following codes of practice and
standards: the Building Regulations 2000 (England and Wales)[1], Eurocodes
(including the UK National Annexes), International (ISO) Standards and British
Standards. There are also documents giving advice about specific building
applications; these specialised approaches tend not to be consistent with other
sources. The issue as to which standards and documents have precedence
over others is touched upon in this review. However, a resolution of this matter
is outside the scope of this report.
The aims of this review are:
1. To bring together state-of-the-art knowledge about vibration comfort
criteria that are appropriate for three building-related areas in the UK
where vibration has proved to be an important recurring issue, namely
buildings, floors and grandstand risers.
2. To discuss dynamic test methods presently used.
2 BaCKgrOUNd
page header right iNfOrMat
– page header subtiiON
tle 3

2 BaCKgrOUNd iNfOrMatiON

2.1 preaMBle
In the literature, it is generally accepted that people’s perception of vibration
is related to the rate of change of acceleration (or ‘jerk’). However, with one
exception all of the vibration comfort criteria found in the literature are related
to acceleration. The continuance of this approach results from the fact that
acceleration levels can be measured relatively easily. This means that both the
measurement and interpretation of accelerations have been undertaken for
many years. Consequently, practising engineers have become familiar with this
method over time, and are therefore reluctant to change.

2.2 UNitS Of aCCeleratiON


As noted earlier, the units of acceleration differ in many of the assessment
methods used; the units include m/s2, cm/s2, mm/s2, %g and milli-g (note that
g is acceleration due to gravity, usually taken as 9.81 m/s2). The International
System of Units (SI) unit of acceleration is m/s2, and this unit will be used
throughout this review. Where appropriate, units presented in other formats
will be converted into m/s2.

2.3 COOrdiNate axiS SYSteMS


Before proceeding, orthogonal axes need to be defined that can be used
to reference specific directions in three-dimensional space. This is because
acceleration is a vector quantity, and has both magnitude and direction.
Considering first the commonly used basicentric coordinate axis system, the
z-axis is the foot-to-head direction, the x-axis is the front-to-back direction
and the y-axis is the side-to-side direction. This coordinate system is shown in
Figure 1 for a person standing, sitting and lying down.

z
z y z

y
Lying down

Figure 1: Basicentric coordinate axis system


x
y

Sitting

Standing
4 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

Although the basicentric axis system relates to a human body whatever the
activity, the latest trend[2] is a movement towards the use of the geocentric
coordinate axis system. This system relates to the building geometry instead
of the body. In this situation, the z-axis is vertical and the x- and y-axes are in
horizontal (or lateral) directions. Figure 2 illustrates the geocentric coordinate
axis system.

y x

Figure 2: Geocentric coordinate axis


system
(Note that the x- and y-axes are
abritrary, but are usually chosen to align
with the major building axes)

2.4 aCCeleratiON veCtOrS


As noted in the previous section, accelerations are vectors. When two or more
accelerations occur at the same time in different directions, their simultaneous
effect consists of the vector sum of all of the accelerations. The method by
which a vector sum can be calculated is described in Appendix A.
Whichever orthogonal axis system is used, the magnitude (or amplitude) of
the time-varying acceleration, a(t), is defined by Equation 1:

(Equation 1)

where ax(t), ay(t) and az(t) are the time-varying accelerations in the x-, y- and
z-axis directions, respectively.
In many applications (such as the vertical vibration of floors), the principal
direction of the vibration is obvious. However, when considering the
dynamic motion of a building that twists and bends, the combined effect of
the accelerations in the two orthogonal lateral x- and y- directions (which
incorporates the effect of torsion) needs to be considered. Unless it is explicitly
stated otherwise, it will be assumed that the magnitude of acceleration, a(t), is
used as the basis for the vibration comfort criteria.

2.5 WeightiNg faCtOrS


To take into account human perception of vibration, many (but not all)
methods apply weighting factors to measured acceleration signals. The
weighting factors are different depending upon the orientation of the body,
and this reflects the fact that people are generally most sensitive to vibration
when standing (as opposed to sitting or lying down). When required, the
appropriate weighting factors are specified in the methodology used.
Weighting can be applied in either the time or frequency domains, and
weighted accelerations will henceforth be denoted by the subscript w. Hence,
the time-varying weighted magnitude of acceleration is denoted by aw(t).
Section 4 presents base-line curves that show the threshold of acceleration
perception, ie the vibration level at which motion is felt. These curves are
2 BaCKgrOUNd
page header right iNfOrMat
– page header subtiiON
tle 5

a function of the vibration frequency. At high and low frequencies, the


acceleration levels shown increase, and this means that greater levels of
acceleration are required before they are noticed. An alternative way of
looking at this is that people are more sensitive to vibration in a middle range
of frequencies. This sensitivity can be expressed mathematically using the
weighting factor approach described above. The point is that the base-line
curves and the weighting factors both perform the same function; that is, they
both correct the actual acceleration values to give the acceleration levels that
people perceive.
A technical point worth making here is that the weighting factors on the
frequency domain need to be applied to a specific type of acceleration
spectrum, which is sometimes referred to as an ‘amplitude spectrum’. In this
context, such a spectrum has a positive frequency range (ie between zero and
infinity Hz), and this feature is called a ‘one-sided’ spectrum. By definition, the
area under a one-sided amplitude spectra is the root mean square (rms) of the
signal; the rms is a fundamental measure of statistical variation, and is defined
in the following section.

2.6 MeaSUreS Of aCCeleratiON


The following two measures of acceleration are the most common parameters
used to assess vibration levels. These measures are sometimes weighted,
and sometimes not, according to the methodology used. In general the
latest British Standards, International (ISO) Standards and Eurocodes tend
to use appropriate weightings applied to measured acceleration signals; US,
Canadian and Australian standards, older standards and criteria derived from
wind tunnel testing of buildings use un-weighted acceleration values.
• Root mean square (rms) acceleration, arms – For a stationary signal (eg
acceleration), the rms of that signal is independent of the record length
(or time interval). The (un-weighted) rms acceleration is defined by
Equation 2[3]:

(Equation 2)

where T is the time duration of the measurement(s).


• Peak acceleration, â – For a given record length (or time interval), the
peak acceleration level is the maximum value of acceleration measured
or predicted over that length. Note that the time-domain response of
instrumentation used must be sufficiently rapid that it can accurately
determine the peak value.

The Gaussian (otherwise known as ‘Normal’ or ‘Bell-shaped’) distribution is


a very common statistical distribution that is found throughout the natural
world. In the context of this review, a Gaussian distribution would result from
combining a small number of randomly time-varying forces or accelerations
(which themselves may or may not be Gaussian). Two general characteristics of
Gaussian distribution records are that they are continuous (ie not intermittent)
and they tend not have individual large peaks (or ‘spikes’). If an acceleration
signal has a Gaussian statistical distribution, the expected peak acceleration is
related to the rms acceleration by Equation 3:

(Equation 3)

where:
• ln is the natural logarithm, or loge
• g is a factor, often called the ‘peak factor’; this factor must not be confused
with the same symbol that is commonly used to denote the acceleration
due to gravity (ie 9.81 m/s2)
6 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

• n is the natural frequency of structural vibration (Hz)


• T is the time period within which the peak acceleration occurs (s)

In British Standards, the weighted rms acceleration (awrms) is the usual method
used to assess vibration comfort. However, this method tends to significantly
underestimate the perception of vibration if the acceleration contains
impulsive accelerations (or spikes). A measure of the acceleration ‘spikiness’ is
the crest factor, which is defined as âw/awrms. If the crest factor is greater than 6
then the Vibration Dose Value (VDV) method described below should be used
instead.
For a given exposure time period, the VDV is defined by Equation 4[2, 4]:

(Equation 4)

Note that the weighted time-varying acceleration is used in this expression,


and that VDVs have units of m/s1.75. Methods are given in BS 6472-1:2008[2]
that allow measured VDV values to be corrected for different exposure times.
For continuous vibration that is statistically stationary (ie the statistical
characteristics of the acceleration do not vary with time) and has a crest
factor between 3 and 6, the estimated VDV (eVDV) can be determined using
Equation 5[4]:

(Equation 5)

where N is the number of occurrences of vibration lasting t* seconds; each of


these occurrences has a weighted rms acceleration of awrms. Note that the eVDV
method is not recommended for non-stationary accelerations or for shocks.
Low-frequency vibrations (less than about 0.5 Hz) can induce motion
sickness. As well as using the weighted acceleration rms criteria, the Motion
Sickness Dose Value (MSDV) can be determined using Equation 6[4]:

(Equation 6)

Note that the units of MSDV are m/s1.5 and T is the total period(s) during which
motion occurs. An alternative simple method for estimating MSDV is given in
BS 6841:1987[4]. However, compared with Equation 6, which can be used for
any time-varying acceleration signal, the simplified method has a more narrow
application.
3 geNeral
page header right –aCCele
pag eratiON
header Criteria
subtitle 7

3 geNeral aCCeleratiON Criteria



It is important to distinguish between the level of acceleration that can be
perceived and the threshold level of acceleration that is deemed to be suitable
for a given activity. There is typically about an order of magnitude difference
between these levels. The suitability acceleration level is less stringent than the
perception level since a person may experience certain levels of vibration that
do not compromise that person’s activity.
People show a wide range of sensitivity to acceleration. Therefore, the
present approach is to set threshold acceleration levels (or ranges) that a
given percentage of the population find unpleasant. Most of the latest codes
of practice and standards either do not state a percentage, or use threshold
values that 2% of the population are likely to complain about. A notable
exception is the CIRIA approach[5], which is based on work undertaken in the
USA. The CIRIA report states that 12% adverse comment in a five-year return
period* is an acceptable serviceability criterion. The different approaches
mean that it is important that clients recognise and understand the impact
of their choice. For this reason, EN 1990:2002[6] states that ‘usually the
serviceability requirements are agreed for each individual project’. Caution
against the use of making precise judgements on the level of adverse comment
is advised in BS 6472:1992[7].
Most building acceleration criteria can be (or are) expressed in terms
of either an rms acceleration threshold level (or range of values), a peak
acceleration level or a VDV. Equation 3 (Section 2) provides a relationship
between the rms and peak accelerations. However, this relationship is only
appropriate if the statistical distribution is Gaussian (or nearly Gaussian). This
peak factor relationship means that peak and rms acceleration criteria given
in different standards are in fact sometimes identical. For example, it might be
agreed that the rms acceleration in a building must not exceed 0.1 m/s2, and
it might also be agreed that the peak acceleration level of that building must
not exceed 0.4 m/s2. If the peak factor g is 4, then Equation 3 shows that these
two agreed acceleration requirements are the same. In situations where the
acceleration signal is not Gaussian (as characterised by ‘spiky’ or intermittent
signals), rms-based methods should not be considered and the VDV approach
(as defined in Equation 4) should be used instead.
Human response to vibration is related to jerk (ie the rate of change
of acceleration). The levels of both the rms and peak accelerations are
related to the amount of jerk. Since the variation with time is not taken into
account, these measures can only be approximations of the human response.
Nevertheless, compared with the peak value, the rms of a time-varying signal
is a more robust measure of the acceleration signal behaviour. Indeed, the
variation of rms over time can be used to demonstrate whether or not a signal
is stationary†. By contrast, peak accelerations rely on relatively few data points.
When measuring a time-varying acceleration signal, instantaneous sampled
data are taken at uniformly spaced intervals. Unless the peak acceleration
value recorded occurs at exactly the time that the actual peak acceleration
occurs, the peak sampled acceleration will be less than the actual peak level.

* A five-year return period means that an event has a one-in-five chance of occurring in any given
year. It does not mean that an event occurs once every five years.
† The statistical characteristics of a stationary signal do not vary with time.
8 page
dYNaMiC
header
COMfOrt
lef t –Criteria
page header
fOr StrUC
subtitUreS
tle

Hence when using measured data, the peak method always has a tendency to
underestimate slightly the actual peak accelerations that occur.
The threshold levels presented in the literature depend primarily upon
the frequency of the vibration and the human activity. The effect of building
rotation is often mentioned as causing a person’s response to be more adverse.
However, this effect is not quantified. For continuous low levels of vibration,
the period of time to which a person is exposed to that vibration seems to
be a second order effect. However, for intermittent and shock accelerations,
the number of doses of vibration (which increases with time) does become
an important factor; this is consistent with the likelihood (or return period) of
large-magnitude acceleration events occurring. The effects of human activity
and body orientation are taken into account either by appropriate weighting
of the measured acceleration records (this seems to be the preferred, latest
approach) or by means of sensitivity factors applied to the results.
In the literature, even within a given set of codes of practice, there can be
overlapping acceleration criteria. For example, there are (i) general criteria that
apply to people, (ii) criteria that apply to buildings, (iii) criteria for buildings
constructed using specified materials, and (iv) criteria relating to building
elements.
The complexity of the assessment methodology, coupled with different
evaluation methods and threshold levels, means that it is often not clear
whether any of the above criteria are more or less onerous. In this situation,
the acceleration vibration levels need to be assessed for all of the above
criteria to ensure that the requirements of a given code of practice (or
standard) are met fully.
To ensure that a building structure has satisfactory serviceability
performance, designers should be aware that guidance is available from
the following sources: Approved Document A[8] of the Building Regulations
2000 (England and Wales)[1], Eurocodes, British Standards, International
(ISO) Standards, the National House-Building Council (NHBC), the Steel
Construction Institute (SCI), the Concrete Centre and the Timber Research and
Development Association (TRADA). The latest guidance given in these sources
is described in this report. Many of these sources state that the intended usage
and acceptable limits of vibration need to be agreed by the client and the
designer at the start of the project. The general consensus is that the design
should achieve a ‘low probability of adverse comment’.
In several of the specialist publications, expressions are given that enable
a designer to undertake a design process that, in principle, is consistent with
the serviceability requirements given therein. There is no requirement that
such a design is tested once built to ensure that it meets those requirements.
However, in practice what is actually built is not the same as the design. It
is only by dynamic testing that one can be certain that a design performs
according to its intended specification once built.
Finally, there is an unresolved issue about which standard or code of
practice is the most appropriate to use in a given situation. An example is the
situation in the UK where a lightweight steel floor is installed in a commercial
or residential building. Following the phased withdrawal of British Standards
pertaining to the design and construction of civil engineering works, all
structures in the UK should now be designed according to Eurocodes.
However, the SCI and the NHBC both provide different guidance. The
serviceability criteria of these three approaches are different, so which
criteria should a designer or client choose? A pragmatic approach would be
to consider all of the methods and ensure that the design meets every set of
requirements. However, this approach may have financial implications, as well
as necessitating the designer to undertake each set of required calculations.
This is an important issue, and further work is required to consider the
different criteria given, with the ultimate aim of harmonising the different
approaches.
page 4header
hOriZONtal
right –aCCele
pag eratiON
header Criteria
subtitle 9

4 hOriZONtal aCCeleratiON Criteria



4.1 preaMBle
Horizontal (or lateral) acceleration of a building is caused by that building
swaying due to imposed loading acting on the building. Horizontal
acceleration is most commonly caused by external vibration being transmitted
into and through the building, by internal unbalanced rotating machinery
within the building or by external time-varying forces acting on the building
(eg wind or earthquakes).
The wind-induced horizontal acceleration of tall buildings has been
an important subject of investigation for many years. This has led to the
development of wind-induced acceleration comfort criteria that are presented
in a format that is appropriate to this particular field of activity. Humans react
in similar ways to different sources of vibration, and therefore in principle the
comfort levels derived from the basis of these wind studies should be similar
to, and compatible with, more general horizontal acceleration comfort criteria.
However, the statistical nature of wind (with the occurrence of strong wind
‘events’ having a specified return period ) does not appear to be reconcilable
with methods associated with randomly occurring vibrations. This division
is shown clearly in ISO 10137:2007[9], in which there is a separate Annex
relating specifically to wind-induced motion in buildings.
The division between criteria relating to wind and to other sources of
vibration is an important one, and this approach has been adopted below.
However, not all codes of practice and standards adopt the ISO 10137
approach. In particular, it should be noted that neither the Eurocodes nor
the British Standards make this distinction. It could be argued that having
a separate wind approach is not necessary. Nevertheless, the Melbourne
Criteria (see below) use variables that relate to the specific problem being
considered, and are the most commonly used current criteria for the design
and assessment of wind-induced horizontal motion of UK buildings.

4.2 WiNd Criteria


4.2.1 Melbourne Criteria
The swaying motion of tall buildings in strong wind conditions has stimulated
many studies of the effects of building-induced motion on people. These
studies have in turn led to the development of horizontal acceleration
criteria, which are discussed below. Although this report does not present in
detail the historical development of these criteria, it is worth noting that the
original acceleration comfort criteria developed by Irwin[5] were based on the
results of surveys of people in two buildings in the USA, and the 2% and 12%
objection level criteria were established from these surveys. As noted earlier,
people’s response to vibration has a large statistical variation. Therefore, using
results taken from only two studies means that there must be considerable
uncertainty in generalising the findings into reliable and robust estimators of
wind-induced acceleration response.
Irwin’s original comfort criteria were developed and extended into criteria
that have general application by Melbourne[10], and in the UK the Melbourne
10 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

Criteria are now the most commonly used criteria in the field of wind
engineering. Equation 7 relates the (un-weighted) peak horizontal acceleration
threshold level â, the natural frequency of the building n, the return period R
(in years) and the time duration T (in seconds):

(Equation 7)

The peak threshold level of acceleration given by Equation 7 relates to an


objection level of 2% of the people present in the building. The horizontal
peak acceleration criteria curves shown in Figure 3 were obtained using
Equation 7.
The time duration T takes into account the nature of the event. In areas
dominated by thunderstorm activity (such as in the USA), the time period
is usually taken to be 10 minutes (600 seconds). However, in the UK the
cyclonic storm response tends to occur through a longer time period, which is
usually taken as one hour (3600 seconds).
At first sight, Melbourne’s Criteria appear to be simple to use. However,
in practice there are complicating features of dynamic building motion. For
example, in Equation 7 it is not clear how the criteria relate to (i) effects
caused by a building swaying in different orthogonal directions, (ii) effects
caused by torsion, or (iii) effects of higher modes of building vibration. These
three issues are discussed in Appendix B.

4.2.2 Other wind-related criteria


There are many wind-related criteria given in different national standards and
codes of practice. However, as noted in Section 1 this study focuses on those
criteria related directly to the UK, or to UK practice.
In Annex D of ISO 10137:2007[9], guidance is given for the peak lateral
(un-weighted) acceleration levels of offices and residences. ISO 10137 states
that the peak accelerations within a one-year return period should not exceed
the levels shown in Figure D.1 of ISO 10137 (this figure is reproduced in this
report as Figure 4). These accelerations include contributions from both x- and
y- horizontal directions as well as the torsional contribution. ISO 10137 states
that ‘in more complex situations, a combination of modal responses may be
required’. However, no further guidance is given.
As noted above, the guidance given relates only to a one-year return
period. ISO 10137 states that ‘multiplying factors other than those used here
would apply’ for longer return periods. However, the standard does not state
what these factors are, nor does it give equations by which the factors can be
calculated. For frequencies of vibration between 0.063 Hz and 1 Hz, further
information on this matter is given in Annex A of ISO 6897:1984[11]. The
appropriate information presented as Figure 1 in that reference is reproduced
here as Figure 5. This curve represents the 2% complaint level about the worst
10-minute motion within a storm having a return period of five years. It is
stated that lateral vibration in different directions should be added vectorally
(vector addition is described in Appendix A of this report). It is also noted
that the visual effects of rotation (eg movement of far field objects from side
to side) exaggerate the sensation of motion, and that satisfactory levels of
acceleration would be less than those shown in Figure 5.

4.3 Criteria NOt related direCtlY tO WiNd


This section considers horizontal acceleration criteria from Eurocodes,
International (ISO) Standards and British Standards. Both the British Standards
and the International (ISO) Standards have criteria that are not material
dependent. However, the approaches contained within the Eurocodes differ
depending upon the method of building construction.
page 4header
hOriZONtal
right –aCCele
pag eratiON
header Criteria
subtitle 11

1.0 1
10-year return period (T = 3600 s)
5-year return period (T = 3600 s)
10 year return period (T = 3600s)
2-year return5period
year return(T = (T3600
period s)
= 3600s)
2 year return period (T = 3600s)
1-year return 1period
year return(T = 3600
period s)
(T = 3600s)
Peak acceleration (m/s2)
Peak Horizontal Acceleration (m/s )
2

0.10.1

0.01
0.01
Figure 3: Melbourne’s horizontal
0.01 0.1 1 10
0.01 0.1 1.0 10.0
Building Natural Frequency, n (Hz) criteria for occupancy comfort in
Natural frequency of building (Hz) buildings

0.50
0.5

Offices
Residences

0.20
0.2
Peak acceleration (m/s2
)
Peak acceleration (m/s )
2

0.10
0.1

0.05
0.05

Figure 4: Acceptable horizontal motion


0.02
0.02
criteria (one-year return period)
0.06 0.1 0.2 0.5 1 2 5
0.06 0.1 0.2 0.5 1.0 2.0 5.0
First natural frequency of building in bending and torsion (Hz) (Source: Figure D.1 of
First natural frequency of building in bending and torsion (Hz) ISO 10137:2007[9])

0.10
0.10

0.08
0.08
)

0.06
0.06
2
rms acceleration (m/s
rms Acceleration (m/s )
2

0.04
0.04

Figure 5: Suggested satisfactory


0.02
0.02
magnitudes of horizontal motion of
0.1 0.2 0.5 1
0.1 0.2 0.5 1.0
Frequency (Hz) buildings used for general purposes
Frequency (Hz) (Source: Figure 1 of ISO 6897:1984[11])
12 page header
dYNaMiC COMfOrt
left –Criteria
page header
fOr StrUCtUreS
subtitle

4.3.1 eurocodes
Eurocodes consist of a linked set of documents that enable UK buildings
and building elements to be designed. Eurocodes started to replace British
Standards pertaining to the design and construction of civil engineering works
in 2010. An overview of the Eurocodes is presented in Figure 6, which shows
how parts of the Eurocodes relate to each other. Figure 6 is shown here so that
reference to specific parts of the Eurocodes can be understood in the context
of the overall structural design process.

Structural safety,
EN 1990 serviceability
and durability

Actions on
EN 1991 structures

EN 1992 EN 1993 EN 1994


Design and
detailing

EN 1995 EN 1996 EN 1999

Geotechnical
EN 1997 EN 1998 and seismic
design
Figure 6: Overview of Eurocode
framework for structural design

Concrete structures are considered in Section 7.1 of EN 1992-1-1:2004[12]


(ie Eurocode 2). It is stated that ‘Other limit states (such as vibration) may be of
importance in particular structures but are not covered in this standard.’ Note
that this Eurocode covers plain, reinforced and pre-stressed concrete structures.
Timber structures are considered in EN 1995-1-1:2004[13] (ie Eurocode 5).
However, this document gives no guidance relating to the horizontal vibration
of structures.
Steel structures are considered in EN 1993-1-1:2005[14] (ie Eurocode 3).
Paragraph (1)B of Section 7.2.3 of the Eurocode states that ‘With reference
to EN 1990 – Annex A1.4.4, the vibrations of structures on which the public
can walk should be limited to avoid significant discomfort to users, and limits
should be specified for each project and agreed with the client.’ The following
text (Note B) states that ‘The National Annex may specify limits for vibration of
floors.’ The vertical vibration of floors is considered in Section 5 of this report.
In addition to this information, and noting that it is steel structures that are
being considered in this paragraph, the text below is quoted verbatim from
Annex A1.4.4 of EN 1990:2002[6]. It can be seen that no specific criteria are
given, but reference is made to EN 1991-1-1:2002[15], EN 1991-1-4:2005[16]
and ISO 10137:2007[9]. The guidance and criteria given in ISO 10137 are
described in Section 4.3.2 of this report.

Annex A1.4.4 Vibrations


1. To achieve satisfactory vibration behaviour of buildings and their structural members
under serviceability conditions, the following aspects, amongst others, should be
considered:
a. the comfort of the user;
b. the functioning of the structure or its structural members (eg cracks in partitions,
damage to cladding, sensitivity of building contents to vibrations).
Other aspects should be considered for each project and agreed with the client.

2. For the serviceability limit state of a structure or a structural member not to be


exceeded when subjected to vibrations, the natural frequency of vibrations of the
page 4header
hOriZONtal
right –aCCele
pag eratiON
header Criteria
subtitle 13

structure or structural member should be kept above appropriate values which

depend upon the function of the building and the source of the vibration, and

agreed with the client and/or the relevant authority.


3. If the natural frequency of vibrations of the structure is lower than the appropriate
value, a more refined analysis of the dynamic response of the structure, including the
consideration of damping, should be performed.
Note: For further guidance, see EN 1991-1-1, EN 1991-1-4 and ISO 10137.

4. Possible sources of vibration that should be considered include walking, synchronised


movements of people, machinery, ground borne vibrations from traffic, and wind
actions. These, and other sources, should be specified for each project and agreed
with the client.

Actions on structures are considered in EN 1991-1:2002[15]. This document


states that actions causing significant accelerations should be considered
using a dynamic analysis. Clause 3.3.2(4) states that the serviceability limit
should be specified in accordance with the service and conditions, and the
structural performance requirements. EN 1991-1:2002 also states that special
dynamic analysis should be considered for vertical loads, and the associated
National Annex considers only vertical accelerations. Vertical accelerations are
discussed in Section 5 of this report.
The general design procedure is given in Clause 5.1.3 of EN 1990:2002[6]
(ie Eurocode 0). For specific structural materials the design procedures are
given in EN 1992 to EN 1999 (ie Eurocode 2 to Eurocode 9). The information
presented in EN 1992 to EN 1999 is considered later in this report. Guidance
for serviceability limit state criteria is given in Annex A of EN 1990:2002, the
text of which was quoted earlier as coming from Annex A1.4.4 Vibrations.
Wind actions are considered in EN 1991-1-4:2005[16]. With regard to along-
wind buffeting wind loads, methods are given in Annexes B and C that enable
rms and characteristic peak accelerations to be determined. It is stated that
these values can be used for serviceability assessments of a vertical structure.
However, no serviceability criteria are given. Serviceability criteria associated
with rms and peak accelerations are given in International (ISO) Standards, as
discussed in the following section.

4.3.2 international (iSO) Standards


The ISO Standards have horizontal acceleration criteria that are based upon
frequency-weighted accelerations. Details of the appropriate frequency-
weighting functions referred to in these standards are given in Appendix C
of this report, and Figure 7 shows the general behaviour of the frequency-
weighting characteristics.

10.010

1.01
Frequency weightings
Frequency Weightings

0.1
0.1

0.01
0.01 0.1 1 10 100
0.1 1.0 Frequency (Hz) 10.0 100.0 Figure 7: Frequency-weighting curve for
Frequency (Hz) horizontal acceleration
14 page header
dYNaMiC COMfOrt
lef t –Criteria
page header
fOr StrUC
subtitUreS
tle

The appropriate ISO Standard relating to vibration in buildings is


ISO 10137:2007[9]. Section 7.1 of this standard states that ‘The designer shall
decide on the serviceability criterion and its variability.’ In Section 7.2.1 it is
stated that ‘In general, the criteria for the restriction of vibration magnitudes
for ordinary buildings are based on the minimum adverse comments of the
population involved.’ ISO 10137 also states that guidance on the assessment
of probable human response to vibration in buildings in the frequency range
1 Hz to 80 Hz is given in Annex C of ISO 10137, and in the range 0.063 Hz to
1 Hz in ISO 6897:1984[11].

frequency range 1 hz to 80 hz
In Annex C of ISO 10137:2007[9] it is stated that, if the ratio of the peak value
of the (weighted) acceleration to the rms value is less than or equal to 6, then
the rms lateral acceleration criteria are given by the base-line curve shown in
Figure C2 (reproduced in this report as Figure 8) factored by the multiplying
factors given in Table C.1 (presented below as Table 1). These multiplying
factors take into account the usage, time of day and location, and give
‘magnitudes of vibration below which the probability of adverse comments is
low’.
If the ratio of the peak value of the (weighted) acceleration to the rms value
is greater than 6 (ie for spiky signals), then it is stated that the rms acceptance
criteria described above may not be appropriate, and that VDV values can be
used. The derived VDV values can be compared with the acceptance criteria
given in Table C.2 (presented below as Table 2). Note that these criteria relate
only to residential buildings, and no recommendations are given relating to
other usages shown in Table 1 (ie offices, workshops and critical working
areas). For these applications advice is given in the British Standards, as
described in the following section.

1.0 1

0.144
0.144
rms acceleration (m/s2)

0.10.1
rms aceeleration (m/s )
2

0.01
0.01

0.0036
0.0036

Figure 8: Baseline rms horizontal


acceleration comfort criteria curve 0.001
0.001
1 10 100
1 10 100
(Source: Figure C2 of Frequency (Hz)
ISO 10137:2007[9]) Frequency (Hz)
page 4header
hOriZONtal
right –aCCele
pag eratiON
header Criteria
subtitle 15

Table 1: Multiplying factors


(Source: Table C.1 of ISO 10137:2007[9])

Multiplying factors to base curve


Place Time Continuous vibration and Impulsive vibration excitation with
intermittent vibrationa several occurences per day

Critical working areas (eg some Day 1 1


hospital operating theatres,
some precision laboratories) Night 1 1

Residential (eg flats, homes, Day 2 to 4b 30 to 90b, c


hospitals) Night 1.4 1.4 to 20

Day 2 60 to 128d
Quiet office, open plan
Night 2 60 to 128

General office (eg schools, Day 4 60 to 128d


offices) Night 4 60 to 128

Day 8 90 to 128d
Workshopse
Night 8 90 to 128

a Doubling the suggested vibration magnitudes can result in adverse comments and this can increase significantly if the
magnitudes are quadrupled (where available, dose/response curves may be consulted). ‘Continuous vibrations’ are
those with a duration of more than 30 min per 24h; ‘intermittent vibrations’ are those of more than 10 events per
24h.
b Within residential areas, people exhibit wide variations of vibration tolerance. Specific values are dependent upon
social and cultural factors, psychological attitudes and expected degree of intrusion.
c Further advice concern [sic] the effects caused by blasting effect are given in ISO 10137.
d The magnitudes of for [sic] impulsive shock excitation in offices and workshop areas should not be increased without
considering the possibility of significant disruption of working activity.
e Vibration acting on operators of certain processes such as drop forgers [sic] or crushers, which vibrate working places,
may be in a separate category from the workshop areas considered in this table. The vibration magnitudes specified
in ISO 2631-1 would then apply to the operators of the exciting processes.

Table 2: Vibration Dose Values above which various degrees of adverse comments may be expected in residential
buildings
(Sources: BS 6472:1992[7] and ISO 10137:2007[9])

Low probability of adverse Adverse comment Adverse comment


Place
comment (m/s1.75) possible (m/s1.75) probable (m/s1.75)

Residential buildings: 16-hour day 0.2 to 0.4 0.4 to 0.8 0.8 to 1.6

Residential buildings: 8-hour night 0.13 0.26 0.51


16 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

frequency range 0.063 hz to 1 hz


For buildings used for general purposes with horizontal vibration events having
duration in excess of 10 minutes, satisfactory magnitudes of (weighted) rms
acceleration are the same as for buildings having natural frequencies between
1 Hz and 80 Hz. As noted earlier, these magnitudes of rms acceleration are
shown in Figure 5, which is taken from ISO 6897:1984[11]. ISO 6897 states that
horizontal vibration in different directions should be added vectorally (vector
addition is considered in Appendix A of this report). It is noted that visual
effects of rotation exaggerate the sensation of motion, and the satisfactory
magnitudes of acceleration would be less than those shown in Figure 5.
For buildings used for special purposes with horizontal vibration
events having duration in excess of 10 minutes, satisfactory magnitudes of
(weighted) rms acceleration are shown in Figure 9, which is also taken from
ISO 6897:1984[11]. ISO 6897 states that these values ‘are appropriate for
special buildings where routine precision work is carried out’.
For buildings with events with duration less than 10 minutes it is suggested
that analysis of acceleration records begins when the (weighted) rms level
exceeds the Figure 9 levels, and stops when the rms acceleration falls below
this magnitude. However, ISO 6897 does not state how those analysed data
should then be assessed.

0.1

0.05
rms acceleration (m/s2)

0.02

Figure 9: Acceleration values


appropriate for special buildings 0.01
(Source: Curve 1 of Figure 1 of 0.1 0.2 0.5 1.0
ISO 6897:1984[11]) Frequency of building (Hz)

4.3.3 British Standards


The British Standards have horizontal acceleration criteria that are based
upon frequency-weighted accelerations. Details of the appropriate frequency-
weighting functions referred to in these standards are given in Appendix C
of this report, and Figure 7 shows the general behaviour of the frequency-
weighting characteristics.
The appropriate standard relating to vibration in buildings is BS 6472-1:2008[2].
The method of assessment used is the VDV method, and the assessment of
the human response in residential buildings is given in Table 1 of that standard
(reproduced below as Table 3). It can be seen that the values shown in Table 3
are very similar to the residential usage ISO VDV values and ranges presented in
Table 2 of this report. As noted earlier, the ISO VDV levels apply only to residential
buildings. The footnote in Table 3 gives multiplying factors that enable this
approach to cover building usages including offices and workshops.
page 4header
hOriZONtal
right –aCCele
pag eratiON
header Criteria
subtitle 17

Table 3: Vibration Dose Values that might result in various probabilities of adverse comment within residential buildings
(Source: Table 1 of BS 6472-1:2008[2])

Low probability of adverse Adverse comment Adverse comment


Place
comment (m/s1.75) possible (m/s1.75) probable (m/s1.75)

Residential buildings: 16-hour day 0.2 to 0.4 0.4 to 0.8 0.8 to 1.6

Residential buildings: 8-hour night 0.1 to 0.2 0.2 to 0.4 0.4 to 0.8

Note: For offices and workshops, multiplying factors of 2 and 4, respectively, should be applied to the above Vibration Dose
Values for a 16-hour day.

BS 6472-1:2008[2] draws attention to the fact that VDV ranges are


presented, instead of discrete values. These ranges mainly reflect the widely
differing susceptibility to vibration evident among members of the population.
BS 6472-1 states that because there is a range of values for each category, ‘it is
clear that the judgement can never be precise’.
Annex A of BS 6399-1:1996[17] contains examples of structures that are
susceptible to dynamic imposed loads. No acceleration criteria are given,
but reference is made to other specialist guidance documents. Buildings with
areas subjected to dancing and jumping (vertical vibrations) are considered in
Section 5 of this report. For lightweight and long-span structures, BS 6399-1
states that ‘Structural design should be undertaken, with the help of specialist
advice and specialist guidance documents, as required by the appropriate
certifying authority.’ For buildings containing machinery it is stated that
‘designers should seek specialist guidance’.
18 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

5 vertiCal aCCeleratiON Criteria



5.1 preaMBle
For the purposes of this section, it will be assumed that vertical accelerations
are synonymous with floor accelerations. This section considers floors located
in commercial and residential buildings. Floors in specialised types of building
are considered in Section 6.

5.2 eUrOCOdeS
To repeat a point made earlier, the Eurocodes consist of a linked set of
documents that enable UK buildings and building elements to be designed. An
overview of the Eurocodes is presented in Figure 6, and this figure shows how
parts of the Eurocodes relate to each other.

5.2.1 Concrete structures (eurocode 2)


Section 7.1 of EN 1992-1-1:2004[12] states that ‘Other limit states (such as
vibration) may be of importance in particular structures but are not covered
in this standard.’. Note that this Eurocode covers plain, reinforced and pre-
stressed concrete structures.

5.2.2 Steel structures (eurocode 3)


Steel structures are considered in EN 1993-1-1:2005[14]. Paragraph (1)B
of Section 7.2.3 of the Eurocode states that ‘With reference to EN 1990
– Annex A1.4.4, the vibrations of structures on which the public can walk
should be limited to avoid significant discomfort to users, and limits should
be specified for each project and agreed with the client.’ The following text
(Note B) states that ‘The National Annex may specify limits for vibration
of floors.’ Clause NA.2.25 of the National Annex to EN 1993-1-1:2005[18]
states that reference should be made to specialist literature as appropriate.
It also states that floor vibrations are covered in Clause NA.4, which relates
to Non-Contradictory Complementary Information (NCCI). This NCCI was
found at www.steel-ncci.co.uk, and will be discussed below. However, it
should be noted that Clause NA.4 states that ‘Whilst this material is likely to
be technically authoritative, not all of it has been reviewed by the UK national
committee, and users should satisfy themselves of its fitness for their particular
purpose. In particular they should be aware that material indicated as not
having been endorsed by the committee might contain elements that are in
conflict with the Eurocode.’
The NCCI consists of three documents, SN034a[19], SN035a[20] and
SN036a[21]. There is no relevant information in SN035a. SN034a states that the
‘comfort of users may be a serious concern for long-span or shallow floors’,
then states that SN036a ‘provides guidance for floor vibrations’. Acceptable
acceleration limits are given in SN036a, which are described below. The
approach uses weighted acceleration records and, since the direction of the
acceleration is known (vertical), reference to ISO 2631-1:1997[3] is made. The
appropriate vertical direction weighting factors are described in Appendix C
of this report. Provided the ratio of the peak value to the rms value is less than
page header
5 vertiCal
right –aCCele
pag eratiON
header Criteria
subtitle 19

6, the rms values of the (weighted) accelerations can then be compared with
the criteria given below. If the ratio is greater than 6 then the VDV approach
can be used. It is noted that it is not considered appropriate to use the VDV
approach on sensitive floors, such as in an operating theatre. A designer
can also take into account the fact that floor vibrations induced by walking
activities are intermittent; this situation is considered in the rms method
described below.

rms method
The recommended acceptable (weighted) accelerations in buildings are
determined from a multiple of the base rms vertical acceleration levels, which
are given in SN036a[21]. For z-axis (foot-to-head direction) accelerations,
the base rms acceleration is 5 x 10-3 m/s2. Multiplying factors are given in
Table 6.2 of SN036a. With the minor clarifications of usage shown in Table 4,
Table 6.2 is identical to Table 1 presented in this review (which is taken from
ISO 10137:2007[9]). Note that the multiplying factors for all of the usage
categories shown in Table 4 are the same.

Table 4: Minor differences between usage stated in ISO 10137:2007[9] and SN036a[21]

Place (ISO 10137) Place (SN036a)

Critical working areas (eg some hospital operating theatres, Critical working areas (eg some hospital operating theatres,
some precision laboratories some precision laboratories)

Residential (eg flats, homes, hospitals) Residential (eg flats, homes, hospitals)

Quiet office, open plan Quiet office, open plan

General office (eg schools, offices) General office (eg schools, offices, laboratories)

Workshops Busy offices and workshops

If vertical vibrations are caused by walking activities, multiplying factors


are given in Figure 6.6a of SN036a[21], which is reproduced below as
Figure 10. These factors are appropriate for office, residential and general
laboratory environments during a 16-hour day. These multiplying factors
provide equivalent VDVs that have a low probability of adverse comment
(VDV = 0.4 m/s1.75).

88

77

66
Multiplying factor
Multiplying factor

55 Corridor length 5 m
Corridor length 10 m
44 Corridor length 20 m
Corridor length 40 m
33

22

11 Figure 10: Multiplying factors for


10 100 1000 10000
10 100 1000 10,000 walking crossings
Crossings per hour
Crossings per hour (Source: Figure 6.6a of SN036a[21])
20 page header
dYNaMiC COMfOrt
lef t –Criteria
page header
fOr StrUC
subtitUreS
tle

vdv method
VDVs can be determined from a (weighted) vertical acceleration record using
Equation 4 given in SN036a[21]. It is stated in SN036a that, for steel-framed
floors, VDVs can be estimated using Equation 8:

(Equation 8)

where t is the total duration of the vibration exposure time (in seconds). As
shown later, in Section 7.2, the total duration time t shown in Equation 8
can be expressed as the product of the number of times that a given level of
vibration occurs and the time of the walking activity.
The calculated dose values should be less than or equal to the z-axis values
presented in Table 6.3 of SN036a[21], which are reproduced in this report as Table 5.
The VDV values shown in this table correspond to a low probability of adverse
comment.

Table 5: Vibration Dose Values below which there is a low probability of adverse comment
(Source: Table 6.3 of SN036a[21])

Place Time Vibration Dose Value (m/s1.75)

Day 0.2 to 0.4


Residential
Night 0.13

Quiet office, open plan Day 0.2

General offices Day 0.4

Busy offices and workshops Day 0.8

5.2.3 timber structures (eurocode 5)


Section 7.3 of EN 1995-1-1:2004[13] considers the effects of vibration.
Section 7.3.2 states that acceptable levels for continuous vibration from
machinery should be taken from Figure 5a in Appendix A of ISO 2631-2:1989[22]
with a multiplying factor of 1.0. However, this ISO Standard has been superseded
by a more recent edition, ISO 2631-2:2003[23], which does not include this
figure. Nevertheless, it is believed that the relevant information is shown in
Figure C.1 of ISO 10137:2007[9], which is reproduced below as Figure 11.

0.1
0.1
rms acceleration (m/s2
)
rms acceleration (m/s )
2

0.01
0.01

Figure 11: Building vibration z-axis base


curve for acceleration (foot-to-head
vibration direction) 0.001
0.001
1 10 100
(Source: Figure C.1 of 1 10 100
Frequency, f
ISO 10137:2007[9]) Frequency (Hz)
page header
5 vertiCal
right –aCCele
pag eratiON
header Criteria
subtitle 21

Residential floors are considered in Section 7.3.3 of EN 1995-1-1:2004[13].


If the fundamental natural frequency of the floor is less than 8 Hz, then
‘special investigation should be made’. Otherwise two parameters, a and b,
need to be determined. a is the deflection of the floor under a 1 kN point
load and b is the floor width (m). Limits for the values of a and b are given
in Table NA.6 of the National Annex to EN 1995-1-1:2004[24], which are
reproduced in Table 6 of this report.
Note that the National House-Building Council (NHBC) also advocates a
deflection limitation to limit the vibration of residential wooden floors; this
approach is described in more detail in Section 7.2.

Table 6: Limits for a and b given in the National Annex to EN 1995-1-1:2004[24]


(Source: Table NA.6 of the National Annex to EN 1995-1-1:2004)

Parameter Limit

a, deflection of floor under a 1 kN point load 1.8 mm for l ≤ 4000 mm


16,500/l1.2 for l > 4000 mm
where l = joist span in mm

b, constant for the control of unit velocity impulse response for a ≤ 1 mm b = 180 – 60a
for a > 1 mm b = 160 – 40a
Note: The formulae for b correspond to Figure 7.2 of EN 1995-1-1:2004+A1:2008[13]. With a value of 0.02 for the modal
damping ratio, ς, the unit velocity impulse will not normally govern the size of floor joists in residential timber floors.

5.3 iNterNatiONal (iSO) StaNdardS


ISO Standards have vertical acceleration criteria that are based upon frequency-
weighted accelerations. This approach is also adopted in BS 6472:1992[7],
although the multiplying factors associated with a given intended usage are
different. Details of the appropriate frequency-weighting functions referred to
in these standards are given in Appendix C of this report, and Figure 12 shows
the general behaviour of the frequency-weighting characteristics.

10

1.0
Frequency weightings

0.1

0.01
0.1 1.0 10 100 Figure 12: Frequency-weighting curve
Frequency (Hz) for vertical accelerations
22 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

The appropriate ISO Standard relating to vibration in buildings is


ISO 10137:2007[9]. Section 7.1 of this standard states that ‘the designer
shall decide on the serviceability criterion and its variability’. Section 7.2.1
states ‘In general, the criteria for the restriction of vibration magnitudes for
ordinary buildings are based on the minimum adverse comments of the
population involved.’ ISO 10137 also states that guidance on the assessment
of probable human response to vibration in buildings in the frequency range
1 Hz to 80 Hz is given in Annex C, and in the range 0.063 Hz to 1 Hz in
ISO 6897:1984[11]. However ISO 6897 relates only to horizontal accelerations,
which are considered in Section 4 of this report.
The following criteria relate to the frequency range 1 Hz to 80 Hz. In
Annex C of ISO 10137:2007[9], it is stated that if the ratio of the peak value
of the (weighted) acceleration to the rms value is less than or equal to 6,
then the rms vertical acceleration criteria are given by the base-line curve
shown in Figure C.1 (presented earlier in this report as Figure 11) factored by
the multiplying factors given in Table C.1 (presented earlier in this report as
Table 1). These multiplying factors take into account the usage, time of day
and location, and give ‘magnitudes of vibration below which the probability of
adverse comments is low’. Stadia and floors of assembly halls are considered
in Section C.2.3 of ISO 10137; to ensure the comfort of the passive part of the
audience, the multiplying factor is 200. Note that an averaging time of 10 s is
recommended when calculating the (weighted) rms. Although it is related to
the safety of the audience (which has a different set of serviceability criteria),
ISO 10137 states that guidance for design of grandstands is given by the
Institution of Structural Engineers (IStructE)[25].
If the ratio of the peak value of the (weighted) acceleration to the rms value
is greater than 6 (ie for spiky signals), then it is stated that the rms acceptance
criteria described above may not be appropriate, and that VDV values can be
used. The derived VDV values can be compared with the acceptance criteria
given in Table C.2 (presented in Section 4.3.2 of this report as Table 2). Note
that these criteria relate only to residential buildings, and no recommendations
are given relating to the other usages shown in Table C.1 (ie offices, workshops
and critical working areas). This is consistent with the ISO approach to
horizontal accelerations, which are considered in Section 4.3.2. For these
applications, advice about vertical acceleration levels is given in the British
Standards, as described in the following section.

5.4 BritiSh StaNdardS


The British Standards have vertical acceleration criteria that are based upon
frequency-weighted accelerations. Details of the appropriate frequency-
weighting functions referred to in these standards (the Wb weights) are given
in Appendix C, and Figure 12 shows the general behaviour of the frequency-
weighting characteristics.
To repeat here several relevant statements already made in Section 4.3.3, the
primary British Standard relating to vibration in buildings is BS 6472-1:2008[2].
The method of assessment used is the VDV method, and the assessment of
the human response in residential buildings is given in Table 1 of that standard
(reproduced as Table 3 in this report). It can be seen that the values shown
in Table 3 are very similar to the residential usage ISO VDV values and ranges
presented in Table 2. As noted earlier, the ISO VDV levels apply only to
residential buildings. The footnote in Table 3 gives multiplying factors that enable
this approach to cover building usages including offices and workshops. BS 6472
states that, because there is a range of values for each category, ‘it is clear that
the judgement can never be precise’.
Appendix D of BS 6841:1987[4] presents information concerning the
effects of low-frequency vertical vibration (0.1 Hz to 0.5 Hz). For this scenario
the accelerations are frequency weighted using Wg factors described in
Appendix C. The Motion Sickness Dose Value (MSDV) can be calculated using
page header
5 vertiCal
right –aCCele
pag eratiON
header Criteria
subtitle 23

Equation 6. Most people become de-sensitised to motion sickness after initial


exposure. Nevertheless, the percentage of un-adapted people who may vomit,
PV, is given by the following expression:

(Equation 9)

where Km varies according to the exposed population. For a mixed population


of unadapted male and female adults, Km = 1/3. This relationship is based
on exposures to motion lasting from about 20 minutes to six hours with the
prevalence of vomiting varying up to about 70%.
Annex A of BS 6399-1:1996[17] contains examples of structures that are
susceptible to dynamic imposed loads. No acceleration criteria are given, but
reference is made to other specialist guidance documents. For buildings with
areas subjected to dancing and jumping, it is stated that ‘Detailed design should
be undertaken to account for the dynamic response of the structure ... with the
help of specialist advice and specialist guidance documents.’
For lightweight and long-span structures, BS 6399-1:1996[17] states that
‘Structural design should be undertaken, with the help of specialist advice
and specialist guidance documents, as required by the appropriate certifying
authority.’ For buildings containing machinery, it is stated that ‘designers should
seek specialist guidance’.
24 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

6 SpeCialiSed BUildiNgS

6.1 graNdStaNdS aNd SpOrtS Stadia
Recent guidance on this specialist subject is given in a document published
by the Institute of Structural Engineers (IStructE) in 2008[25]. With regard to the
effects of crowd-induced motion of grandstands and stadia, this document
brings together a large amount of research undertaken over many years.
Acceleration comfort criteria are given in Table 1 of the publication[25], and
these criteria are summarised below. The rms acceleration is used to assess
the structure, and attention is drawn to the fact that this publication uses un-
weighted rms values.
The criteria depend upon the four operational scenarios described
below. However, in general if the fundamental natural frequency of the bare
grandstand (f0) is greater than 6 Hz, the grandstand is suitable for all scenarios.
f0 must be greater than 3.5 Hz for all new constructions suitable for sport and
other events; for seating areas with f0 less than 3 Hz, resonance could occur
with consequent large structural movements.
• Scenario 1 – Stand used for viewing of sporting and similar events with less
than maximum attendance. f0 ≥ 3.5 Hz, or accepted at the discretion of
a Listed Engineer. Existing stands with 3 < f0 < 3.5 Hz may be deemed
satisfactory on the basis of past experience and used for less lively sections
of the crowd.
• Scenario 2 – Stand used for classical concerts and typical well-attended
sporting events. Either f0 ≥ 3.5 Hz or rms (un-weighted) acceleration must
be less than 3%g, and rms of maximum dynamic displacement due to
crowd loading should not exceed 7 mm.
• Scenario 3 – Stand used for commonly occurring events including high-
profile sporting events, concerts with medium tempo or cross-generational
appeal. Either f0 ≥ 6 Hz or rms (un-weighted) acceleration must be less
than 7.5%g, and rms of maximum dynamic displacement due to crowd
loading should not exceed 7 mm.
• Scenario 4 – Stand used for more extreme events including high-energy
concerts with periods of high-intensity music. Either f0 ≥ 6 Hz or rms
(un-weighted) acceleration must be less than 20%g, and rms of maximum
dynamic displacement due to crowd loading should not exceed 7 mm.

For grandstands, Table 3 of BRE Digest DG 426[26] provides recommendations


of peak (frequency-weighted) acceleration levels. The values shown in that
table are reproduced below as Table 7.
page header right6–SppeCialiSed
age headerBUildiNgS
subtitle 25

Table 7: Reaction to peak acceleration levels on grandstands for a frequency range less than 10 Hz
(Source: Table 3 of BRE Digest DG 426[26])

Peak acceleration Reaction

< 5%g (< 0.49 m/s2) Reasonable limit for passive persons

< 18%g (< 1.77 m/s )


2
Disturbing

< 35%g (< 3.43 m/s )


2
Unacceptable

> 35%g (> 3.43 m/s2) Probably causing panic

Note: The values in parentheses are calculated from the values presented in BRE Digest DG 426.

The Steel Construction Institute (SCI) recommends acceptance thresholds


for vibrations of steel grandstand floors, which are given in publication P354[27].
The weighted rms acceleration approach given in BS 6472:1992[7] is advised in
P354, and this approach is described in Section 5.3. However, P354 states that
Wg and Wd weighting factors are used, whereas these weights are not given
in BS 6472:1992[7]. It is believed that the Wb and Wd weighting factors given
in BS 6472-1:2008[2] (which supersedes BS 6472:1992[7]) should be used.
P354 recommends that a multiplying factor of 24 is used for stadia. The public
reaction to imposed vertical vibration is reproduced in this report as Table 8.

Table 8: SCI acceptability criteria for grandstands


(Source: Table 5.5 of P354[27])

rms acceleration Reaction

< 3.5%g (< 0.34 m/s2) Reasonable limit for passive persons

< 12.7%g (< 1.25 m/s2) Disturbing

< 24.7%g (< 2.42 m/s )


2
Unacceptable

> 24.7%g (> 2.42 m/s2) Probably causing panic

Note: The values in parentheses are calculated from the values presented in P354.

Comparing the information given in Tables 7 and 8, it can be seen that


the only difference is that BRE Digest DG 426 uses peak acceleration levels,
whereas P354 uses rms values. Using Equation 3, the peak factor calculated
is 1.42, or √2. Sinusoidal signals (having a single frequency) have the property
that the peak amplitude is equal to the √2 times the rms. Hence it would
appear that the assumption that the grandstand vibration is sinusoidal has
been made in P354. Whereas this assumption produces the acceptability
criteria shown in Table 8, it is interesting to note that grandstand acceleration
measurements taken by Ellis and Littler[28] show that the response of
grandstands to crowd loads is not sinusoidal.

6.2 hOSpitalS
6.2.1 information given by the Steel Construction institute
(SCi)
Advice published by the SCI relating to the acceleration serviceability limits
of hospital floors is given in P354[27]. These guidelines are consistent with
achieving the requirements specified in Health Technical Memorandum
(HTM) 08-01[29]. It is noted in HTM 08-01 that, as well as taking into account
requirements associated with normal building floors, hospital floors also need
to take into account the reduced walking speeds that are likely to occur when
a walker is near to or accompanying a patient.
For continuous vibrations, the Wg weighting factor is used (defined in
Appendix C of this report), which is for standing or seated people when hand
control is important. The base (human perception level) acceleration value
26
26 page header
dYNaMiC COMfOrt
lef t –Criteria
page header
fOr StrUC
subtitUreS
tle

is 5 mm/s2 (0.005 m/s2), and the multiplying factors shown in Table 9 are
reproduced from Table 8.1 of P354. P354 states that these criteria are given in
Section 2.132 of HTM 08-01[29]. It is also stated that ‘these multiplying factors
will result in a low probability of adverse comment for continuous activities in
hospitals’.

Table 9: Multiplying factors for continuous vibration of hospital floors


(Source: Table 8.1 of P354[27])

Room type Multiplying factor

Operating theatres, precision laboratories, audiometric testing booths 1

Wards 2

General laboratories, treatment areas 4

Offices, consulting rooms 8

For intermittent vibrations (eg walking activities), the VDV approach


described in Section 2 of this report can be used. For hospitals, the maximum
allowable values for VDVs are given in Section 2.133 of HTM 08-01[29], and
are reproduced in this report as Table 10. Note that the accelerations are
again weighted using the Wg weighting factors described in Appendix C of
this report. For operating theatres and critical working environments (such
as precision laboratories), the VDV assessment method is not appropriate.
This is because a single event in excess of perceptibility could cause a critical
consequence to occur.

Table 10: Limiting VDV values for hospital floors


(Source: Section 2.133 of HTM 08-01[29])

Room type Maximum VDV value (m/s1.75)

Ward, residential – day 0.2

General laboratories, offices 0.4

Workshops 0.8

6.2.2 information given by the Concrete Centre


The Concrete Centre publishes a design guide[30] that gives advice relating to
hospital vibration criteria. This document relates to the NHS requirements
given in HTM 2045[31]. The approach is based on the methodology described
in BS 6472:1992[7] (note that this British Standard has been superseded by
BS 6472-1:2008[2]).
For continuous vibrations, BS 6472:1992[7] gives the baseline (threshold
of perception) rms criteria curve shown in Figure 11, and multiplying factors
that relate to the intended usage. The multiplying factor for residential usage is
1.4, and this value is recommended in the Concrete Centre’s design guide[30]
for ward areas. For intermittent vibration, the VDV approach is used, which
is described in Section 2 of this report. According to BS 6472:1992[7], the
limiting VDV values for residential buildings are shown in Table 3 of this report,
and these values are recommended for ward areas.
The Concrete Centre’s design guide[30] states that the effect of full-height
partitions of masonry, studwork or glass can significantly alter the dynamic
properties of floors, leading to lower floor vibration levels. In hospitals, the
presence of such partitions is sometimes integral to the function of the floor
space. In this circumstance, it is appropriate to consider the effect of the partitions
when assessing the dynamic performance of a floor. Note that part-height and
moveable partitions cannot be relied upon to reduce vibration response.
HTM 2045[31] states that vibration levels in operating theatres must be less
than the baseline rms curve, and this criterion applies to both continuous and
intermittent vibration. Any relaxation of this requirement must be agreed with
the client.
page header right6–SppeCialiSed
age headerBUildiNgS
subtitle 27

6.3 Car parKS


Advice published by the SCI relating to the acceleration serviceability limits
of car parks is given in P354[27]. The vertical response factor is defined as the
ratio of the weighted acceleration rms divided by 0.005 m/s2. The horizontal
response factor is defined as the ratio of the weighted acceleration rms
divided by 0.00357 m/s2. P354 recommends that the response factor of the
empty car park should be less than 65 with a damping ratio of 1.1%. No
recommendation is given for full car parks.
28
28 page header
dYNaMiC COMfOrt
lef t –Criteria
page header
fOr StrUC
subtitUreS
tle

7 � gUidaNCe giveN BY UK BUildiNg


regUlatiONS, trade aSSOCiatiONS
aNd Other iNdUStrY BOdieS
7.1 UK BUildiNg regUlatiONS
Apart from the situation described below, Approved Document A[8] of the
Building Regulations 2000 (England and Wales)[1] does not consider the effects
of dynamic loads or vibrations acting on structures or buildings.
In Section 0.3 of Approved Document A, explicit reference is made to
grandstands and structures erected in places of public assembly. It states
that interim guidance may be found in IStructE’s 2001 publication[32], and
that supplementary advice on dynamic testing of grandstands and seating
decking is published in an advisory note, also published by IStructE[33].
According to IStructE, both of these sources have now been superseded by
another document published in 2008[25]. However, this latest document is
not referenced directly by the UK Building Regulations. The treatment of
grandstands is a specialist area that is considered in detail in Section 6.1 of this
report.

7.2 trade aSSOCiatiONS aNd Other iNdUStrY BOdieS


The Timber Research and Development Association (TRADA) and the Centre
for Timber Engineering (CTE) recommend the usage of Eurocodes. The
Concrete Centre has no stated position, but has produced a design guide[30]
that considers the dynamic behaviour of floors; the vibration criteria given in
this document are described below. The National House-Building Council
(NHBC) and the Steel Construction Institute (SCI) have their own approaches
for floors, which are also described below.

7.2.1 the Concrete Centre


Advice given in the Concrete Centre’s design guide[30] is based on the
methodology described in BS 6472:1992[7]. (Note that this British Standard
has been superseded by BS 6472-1:2008[2].) For continuous vibrations,
BS 6472 gives the baseline (threshold of perception) rms criteria curve
shown in Figure 11, and multiplying factors relating to the intended usage.
The multiplying factors given in BS 6472 for different intended usages are
reproduced in Table 11 below; these factors are associated with vibration
levels where the probability of adverse comment is low.

Table 11: Multiplying factors associated with low probability of adverse comment
(Source: BS 6472:1992[7])

Room type Multiplying factor

Critical working areas (eg hospital operating theatres, precision laboratories) 1

Residential – day 2 to 4

Residential – night 1.4

Offices – day or night 4

Workshops – day or night 8


7 gUidaNCe giveN BY UK BUildiNg regUlatiONS, trade aSSOCiatiONS
page header right
aNd Other
– pageiNdUSt
headerY
r subtitle
BOdieS 29

For commercial buildings (eg offices, retail, restaurants, airports), the


Concrete Centre’s design guide[30] states that for single-person footfall
vibration, a multiplication factor of 8 is ‘almost always satisfactory’. Note that
the Concrete Centre’s recommended footfall level is twice the level proposed
by BS 6472:1992[7] for continuous vibration.
The Concrete Centre’s design guide notes that there have been occasions
where complaints have been recorded at the above footfall vibration levels,
particularly when the vibration occurs regularly. Therefore, it is recommended
that a multiplication factor of 4 is used in the following circumstances:
1. In premium-quality open-plan offices when precision tasks are undertaken.
2. In open-plan offices with busy corridor zones near mid-span.
3. In heavily trafficked public areas with seating.

To take into account the degree to which full-height partitions reduce the
vibration levels of floors, it is suggested that a multiplication factor of 6 is
appropriate for floors with many full-height partitions.
For residential buildings, the Concrete Centre’s design guide[30] states that
the BS 6472:1992[7] criteria for residential buildings are based on experience
with external sources of vibration such as railways and roads. It is argued that
the stringent levels of vibration (especially at night) are ‘not really necessary for
vibration caused by occupants of the residence themselves walking’. Therefore
the design guide recommends that twice the multiplication values (reproduced
in Table 11) are adopted for footfall-induced vibration. Note that permanent
full-height partitions of masonry, studwork or glass can be used to reduce the
floor vibration levels.

7.2.2 the National house-Building Council (NhBC)


The NHBC has two approaches depending upon the floor construction. According
to Page and Sangarapillai[34], the NHBC will accept timber floors that are designed
in accordance with either the National Annex to EN 1995-1-1:2004[24], or where
the maximum instantaneous deflection of unstrutted floors is limited to 12 mm
under dead-plus-imposed loads.
For steel floors, guidance is given in the NHBC Standards[35], and the
relevant criteria given therein are presented verbatim below.

iii. the natural frequency of the floor should be limited to 8 Hz for dead load plus
0.2 x imposed load. This can be achieved by limiting the deflection of a single joist to

5 mm for the given loading.



iv. the deflection of the floor (ie a series of joists plus the floor decking) when subject to
a 1 kN point load should be limited to the following values:

Span (m) Maximum deflection (mm)

3.5 1.7
3.8 1.6
4.2 1.5
4.6 1.4
5.3 1.3
6.2 1.2

The deflection of a single joist is dependent on the overall floor construction and the
number of effective joists that are deemed to share the applied 1 kN point load. The
following table gives typical values:

Number of effective joists

Floor configuration Joist centres

400 mm 600 mm

Chipboard, plywood or oriented strand board 2.5 2.35


Built-up acoustic floor 4 3.5
30 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

7.2.3 the Steel Construction institute (SCi)


The SCI’s recommended acceptance thresholds for vibrations of steel floors
are given in P354[27]. For continuous vibrations the weighted rms acceleration
approach adopted in BS 6472:1992[7] is given in P354, and this approach is
described in Section 5.3. However, P354 states that Wg and Wd weighting
factors are used, whereas these weights are not given in BS 6472:1992[7]. It
is believed that the Wb and Wd weighting factors given in BS 6472-1:2008[2]
(which supersedes BS 6472:1992) should be used.
Instead of using the multiplying factors given in BS 6472, P354[27]
recommends the use of factors that are reproduced in Table 12 of this report.
Note that these factors do not apply to hospital environments, which is a
specialist topic that has already been considered in Section 6. These factors
also do not apply to ‘sensitive processes’. These processes are considered in
Section 5.4 of P354, but are not considered further in this review since they
do not relate to the comfort of people. P354 states that for dance floors in
night clubs (having loud music and low lighting), a multiplying factor of 120 is
acceptable.
For intermittent vertical (z-axis) accelerations, the VDV approach is
recommended in P354[27]. The acceptability criteria are those specified in
BS 6472:1992[7], which are identical to the values shown in Table 2 of this
report (taken from ISO 10137:2007[9]). It is stated in P354 that a purely
sinusoidal response was used to estimate the VDVs in that table and this
approach is unsuitable for walking activities. Nevertheless, the VDV limits
shown in Table 2 still apply.
In P354[27], the VDV of a walking activity of duration Ta that occurs na times
in an exposure period is:

(Equation 10)

Note that the accelerations used in this assessment are frequency-weighted


according to the British Standard criteria given in Appendix C. The VDV
values calculated using Equation 10 can be compared with the acceptability
limits shown in Table 5.4 of P354 (which refer to BS 6472:1992[7] and are
reproduced in Table 2 of this report).
Alternatively, Equation 10 can be re-arranged with na as the subject. For
a given VDV and time period (according to BS 6472:1992[7], the exposure
periods that should be considered are a 16-hour day and an eight-hour night),
a VDV analysis can be considered satisfactory if the floor will be traversed
fewer than na times in the exposure period.
P354[27] also provides advice about the design of lightweight steel floors. It
states that the frequency of floors within buildings should be higher than 8 Hz,
and in linking corridors higher than 10 Hz. The accelerations are weighted
using the Wb curve given in BS 6841:1987[4]; these weights are described
in Appendix C. The vertical response factor is defined as the ratio of the
weighted acceleration rms divided by 0.005 m/s2. If the response factor is less
than 16 there is no need to consider intermittent vibrations. If the design brief
permits the use of VDVs, VDV values that correspond to a ‘low probability
of adverse comment’ are given in Table 8.5, which is reproduced here as
Table 13 (it is assumed that the values presented have units of m/s1.75).
7 gUidaNCe giveN BY UK BUildiNg regUlatiONS, trade aSSOCiatiONS
page header right
aNd Other
– pageiNdUSt
headerY
r subtitle
BOdieS 31

Table 12: Recommended base curve multiplying factors given in P354 based on single-person excitation
(Source: Table 5.3 of P354[27])

Place Multiplying factor for exposure to continuous vibration

Office 8

Shopping mall 4

Dealing floor 4

Stairs – light use (eg offices) 32

Stairs – heavy use (eg public buildings, stadia) 24

Table 13: VDV values for lightweight steel floors


(Source: Table 8.5 of P354[27])

Place Low probability of adverse comment

Buildings: 16-hour day 1.6 m/s1.75

Buildings: 8-hour night 0.51 m/s1.75


32
32 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

8 � dYNaMiC teStiNg Of BUildiNgS


aNd BUildiNg eleMeNtS
8.1 preaMBle
The previous sections of this report have been concerned with establishing
acceleration comfort criteria that are relevant to UK structural design. This
section discusses issues associated with the dynamic testing of structures. The
purpose of this section is to provide information about the different types of
dynamic testing that are available and how dynamic testing can be used in a
practical context. Additional useful and practical advice on this subject is given
by Moore[36], who identifies historic pitfalls that need to be avoided when
undertaking dynamic testing.

8.2 iNtrOdUCtiON
Dynamic testing involves applying time-varying forces to a building (or
structure), measuring the structural response and then drawing inferences from
that response. Hence, dynamic testing is different in nature to static testing
in which time-invariant loads are applied and measured. There is surprisingly
little information in the literature about the test methodologies used, what
can be measured, how the results can be used practically and what is the best
method (or methods) to use in a given situation.
Dynamic testing of buildings ranges from investigating the vibration of the
whole building to testing large or small building elements, or testing vibration
produced by (large or small) machinery within a building. Hence, it can be
seen that the scale of this testing can vary dramatically. The scope of the testing
varies depending on the problem being considered and the budget available
for testing. Specialist skills are necessary to plan and undertake large-scale
testing, and then to analyse and interpret the results. This means that expert
advice is usually sought at an early stage of a project where dynamic testing is
required.
Most dynamic testing of building elements is undertaken in a laboratory.
This is because it is possible to control the experimental parameters, and
this increases the accuracy of the experimental techniques used. Although
laboratory testing of a test specimen (or specimens) is often extremely useful,
such testing does not simulate every aspect of the structural behaviour of
elements when installed in a building. This is because it is not possible to
reproduce exactly the full-scale edge conditions in a laboratory. Therefore,
on-site testing is the only way to demonstrate whether or not the performance
of a building element (eg a building floor) meets its specified dynamic
performance. Nevertheless, laboratory testing can be used to certify that
products do meet specified structural requirements (eg the natural frequency
of a floor is greater than a given frequency).
This review is not a treatise on structural dynamics, but at this point it is
useful to have a brief overview of the relevant parameters that need to be
considered. The motion of a dynamic structure is determined completely
if, for every mode of vibration, the natural frequency, mass distribution (or
structural stiffness), mode shape and damping are known. In theory, any
continuous structure (eg beams or columns) has infinite modes of vibration.
8 dYNaMiC teStiNg
page Of
header
BUildiNg
ri gS
htaNd
– page
BUildiNg
headereleMeNt
subti tS
le 33

Nevertheless, for practical purposes it is only the modes that have the lowest
natural frequencies that need to be considered. The principal aim of dynamic
testing is, for each mode of vibration of practical interest, to measure the
aforementioned parameters, thus enabling the dynamic motion of the system
to be predicted.

8.3 COMMONlY USed teSt MethOdS


8.3.1 impact methods
Impact methods consist of suddenly applying (or releasing) force, and
measuring the resulting decaying motion. It should be noted that the energy
supplied by this approach excites all of the modes of vibration, including
transients. Exciting high-frequency modes and transients is not necessarily
desirable, since it is only the vibrations of the lowest modes that are of
practical importance. It is essential that sufficient energy is imparted into the
structure to ensure that the response of the whole structure is measured.
Different methods can be used to create impact loads. The most common
method used for testing horizontal surfaces (floors) is the so-called ‘heel-
drop’ method. Dropping a known weight onto a floor is also sometimes used.
Walking across a floor generates impact loads, and explosive charges have
been used in the past (particularly in geotechnical studies). Impact hammers
(which may be calibrated) can be used to impart impact loads onto horizontal,
vertical or sloping surfaces; BS 6897-5:1995[37] describes such testing in
detail. Swinging weights, rockets, explosive charges and releasing tightened
cables have all been used to impart impact loads to the walls of buildings and
masts. A final method that can be used to generate impact loads is to excite
the structure continuously (eg by means of mechanical shakers, or using bells
within bell towers) and then to stop that excitation suddenly.
The instrumentation used to measure the response must be sensitive
enough to detect small amplitude motion, yet not so sensitive that it is
damaged when a sudden impact occurs. Hence, these instruments must have
a large ‘dynamic range’, which tends to makes them expensive.
A typical time history of a heel-drop test undertaken on a wooden floor is
shown in Figure 13. It can be seen that after the initial impact the vibration of
the floor quickly returns to the ambient vibration level. The speed at which
the vibration decays is related to a parameter known as the structural damping
ratio, ς.

0.2

0.1
Vertical acceleration (m/s2)

-0.1

-0.2
0 0.5 1.0 1.5 2.0 2.5 3.0 Figure 13: Vertical acceleration time-
Time (s) trace of heel-drop test
34 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

A single degree of freedom (1-DOF) system consists of a vibrating mass, a


damper and a spring. The extension of the spring creates a force that acts on
the mass to oppose the displacement; this force is proportional to the spring
extension. The damping also induces a force that opposes the motion of the
mass, but this force is proportional to the velocity of the mass.
For small levels of system damping (ie ς much less than 1), consider now
a mass being pulled away from its equilibrium position by a distance (or
amplitude) A1. Releasing this mass will cause it to pass through the equilibrium
position, and then come to rest at a distance (or amplitude) A2 away. Note that
A2 will be less than A1 as a result of the energy dissipated through the effect of
damping.
It can be shown that for a lightly damped system, the damping ratio is
proportional to the amplitude ratio of successive peaks, A1/A2. Thus:

(Equation 11)

This approach is commonly termed the ‘logarithmic-decrement method’ (or


‘log-dec method’ for short). The time period between successive peaks, T, can
also be determined from the decaying signal. This enables the fundamental
natural frequency of the system, f0 (= 1/T), to be estimated.
Figure 13 was chosen deliberately to illustrate some of the practical
difficulties associated with this approach. For example, it is not clear where
the impact occurs, and at what exact time the response starts to decay. The
time interval over which the range of amplitude ratios should be taken is not
obvious, and therefore certain arbitrary decisions have to be made. Note that
some structures have damping ratios that are functions of the amplitude of
the decay. In this situation, a single damping ratio value is not an appropriate
descriptor of the vibration behaviour.
The method described above also assumes implicitly that the structure
tested vibrates as a 1-DOF system. If the structure has two close natural
frequencies there is an interaction between the energy contained in the two
vibrating modes. This interaction causes a ‘beating’ of the decaying amplitude,
and prevents an accurate determination of the damping ratio.
Having stated the problems associated with this method, it is only fair to
state that the method is popular and in widespread use. It is relatively easy
to undertake the test and to collect the data. Using phase-locked averaging
techniques, it is possible to reduce the inherent variation associated with
impacts (which by their nature are extreme events). Furthermore, transforming
impact time-varying data into the frequency domain enables all of the natural
frequencies of a structure to be identified quickly. Indeed, with regard to
structural investigation, the ability to identify natural frequencies is believed by
the author to be the most appropriate use of impact test methods.

8.3.2 Continuous vibration methods


Structures can be set into forced vibration either by means of a mechanical
shaker (or shakers) or by ambient conditions (eg wind or traffic). It is worth
noting that, in full-scale building testing, vibrations induced by ambient
conditions can obscure the response signals induced by the forced vibration. It
is therefore important that forced vibration investigations are undertaken in a
test environment that is as windless and vibration-free as possible.
The present ‘method of choice’ for dynamic testing is using mechanical
shakers to induce forced structural vibration. The reason is that this test
methodology provides control over the test environment. Moreover, if the
magnitude of the applied force is known (and the response to that force
measured), this can provide additional information about the dynamic
behaviour of the structure.
There are two types of mechanical shaker that impart force to a structure:
rotary shakers and loudspeaker-type (or moving coil) shakers, which are
described briefly below.
8 dYNaMiC teStiNg
page Of
header
BUildiNg
ri gS
htaNd
– page
BUildiNg
headereleMeNt
subti tS
le 35

Figure 14: A ‘grandstand’ rotary shaker


used by BRE

Figure 15: Grandstands at Everton


Football Club
(Courtesy of Everton Football Club)

Rotary shakers consist of counter-rotating masses mounted within a frame.


The rotation of these weights creates a sinusoidal force, the direction of which
can be changed by mounting the frame in different orientations with respect to
the structure. The rotation of the weights is controlled by a stepper motor. This
stepper motor must be powerful to prevent interaction between the weight
rotating speed (or frequency) and the response from the oscillating structure.
Since the masses and their offset distance from the axis of rotation are known,
this enables the amplitude of the sinusoidal force acting on the structure to be
determined by means of calculation. Rotary shakers come in different sizes.
The shaker that BRE usually uses when testing buildings is the ‘grandstand’
shaker, which is shown in Figure 14.
BRE has been involved in the dynamic testing of grandstands, sports
stadia and rock concert venues for many years. Football stadia tested
include Liverpool, Everton, Nottingham Forest, Boston United, Reading (the
Madejski Stadium), Lincoln City and Luton Town. The Millennium Stadium
in Cardiff was tested to assess its suitability for use as a rock concert venue.
Northamptonshire County Cricket Ground, Earls Court and the Ricoh Arena
(Coventry) have also been tested. A photograph of the grandstands at Everton
is shown in Figure 15.
Loudspeaker-type (or moving coil) shakers convert electrical input signals
into translational motions of an armature, with the armature then being
attached to the structure. The force transmitted to the structure is then related
directly to the magnitude and behaviour of the electrical signal. Hence, if the
electrical signal is sinusoidal, the force imparted to the structure is sinusoidal.
Note that the magnitude of the force needs to be a calibrated function of the
electrical signal gain. This approach means that any time-varying force signal Figure 16: A medium moving coil
shaker used by BRE
can be generated; this has certain advantages, which are described below. A (Note that the size of the shaker can be
photograph of a medium-sized moving coil shaker used by BRE is shown in gauged by the 15 cm steel ruler shown
Figure 16. in the foreground)
36 page header
dYNaMiC COMfOrt
left –Criteria
page header
fOr StrUCtUreS
subtitle

It is generally agreed that the most accurate measurements of dynamic


structural properties are made by means of a ‘frequency sweep’. This involves
vibrating the structure at a fixed frequency by the shaker. Once steady-state
conditions occur, the response of the structure is measured. The frequency
of vibration is adjusted and the process repeated across the whole frequency
range of interest. A frequency sweep measured for a large steel floor mounted
in the BRE structural test laboratory is shown in Figure 17. The curve fitted to
the measured results is a 1-DOF theoretical relationship that relates the applied
sinusoidal force to the measured acceleration response. From this curve, the
natural frequency and damping ratio of the structure can be obtained. Since
the applied force is known, this method also enables the mode-generalised
mass and mode-generalised stiffness to be calculated. The frequency sweep
method allows a rigorous analysis of the statistical uncertainty of this process
to be carried out, which is not possible with any other dynamic test methods.
However, undertaking frequency sweeps is a time-consuming process, and
therefore this approach tends to be expensive. The mode shapes of a structure
can be determined using accelerometers placed at different structural locations.
If the structure is vibrated at or near to a given natural frequency, the relative
measured response enables the mode shape to be established.
acceleration
Non-dimensional Acceleration
Non-dimensional

Theoretical curve
Theoretical curve
Measured
Measureddata
Data

8.0 8.5 9.0 9.5


8.0 8.5 9.0 9.5
Figure 17: Typical frequency sweep of a Frequency (Hz)

large steel floor Frequency (Hz)

A relatively new approach to dynamic testing[27] is to use loudspeaker-type


shakers to apply random white noise forcing functions to a structure, while
measuring the simultaneous frequency response at test points located across
that structure. This method not only identifies all of the natural frequencies,
it also predicts damping ratios, mode shapes of vibration, mode-generalised
mass and mode-generalised stiffness. The advantage of this approach is that
it is relatively quick. However, it requires very sophisticated equipment to
measure and analyse the huge volumes of data needed to obtain statistically
reliable results.

8.3.3 ambient vibration methods


Although the most accurate method of dynamic testing is to use forced
vibration techniques, measured ambient vibrations can also provide useful
information. This method assumes that the ambient conditions have
broadband energy, which is distributed evenly across the frequency range of
interest. The resulting structural vibration is similar to that excited by random
white noise vibration, which is a situation considered in the previous section.
However, when using ambient methods the applied force is not known. This
means that such methods cannot be used to determine the mode-generalised
mass or stiffness. However, they can be used to determine natural frequencies
and mode shapes, and to estimate the damping ratio.
8 dYNaMiC teStiNg
page Of
header
BUildiNg
ri gS
htaNd
– page
BUildiNg
headereleMeNt
subti tS
le 37

The ambient measurement technique was used to measure the behaviour


of a wooden ship’s mast, with the wind providing the excitation energy. The
results of this testing are shown in Figure 18. Since wind varies randomly in
time, it is important that such results are repeated to establish the statistical
variability of the measured results, and two repeated sets of data are shown. It
can be seen that the mode shapes of each repeated run agree very well, but
the damping ratio values show considerable scatter. The large degree of scatter
of the damping values is expected because it is notoriously difficult to measure
structural damping accurately in the absence of controlled experimental
conditions. As an aside, it is believed that the increase in damping ratio seen
towards the deck is caused by frictional losses associated with the vibration
and rubbing together of an increasing amount of the ship’s rigging.

1.0 1.0 Mode 1: Natural frequency f1 = 0.64 Hz


Non-dimensional height (z/H)

0.8 Non-dimensional height (z/H) 0.8

0.6
0.6

0.4
0.4

0.2
0.2
0
-1.0 -0.5 0 0.5 1.0 0
Mode shape 0 10 20 30

Damping ratio (%)

1.0 1.0 Mode 2: Natural frequency f2 = 1.75 Hz


Non-dimensional height (z/H)

Non-dimensional height (z/H)

0.8 0.8

0.6
0.6

0.4
0.4

0.2
0.2
0
-1.0 -0.5 0 0.5 1.0 0
Mode shape 0 5 10 15 20 25 30

Damping ratio (%)

1.0 1.0 Mode 3: Natural frequency f3 = 4.90 Hz


Non-dimensional height (z/H)

Non-dimensional height (z/H)

0.8 0.8

0.6
0.6

0.4
0.4

0.2
0.2
0
-1.0 -0.5 0 0.5 1.0 0
Mode shape 0 5 10 15 20 25 30
Figure 18: Mode shapes and damping
Damping ratio (%) ratios along ship’s mast
38 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

Ambient vibration testing can also be used to identify outlying behaviour.


BRE is engaged in an ongoing programme of work with the Houses of
Parliament to measure the vibration of the pinnacles. Figure 19 shows the
BRE laser system that measures the pinnacle vibration. By measuring this
vibration in light wind conditions, the natural frequencies of each pinnacle
can be identified. Any pinnacles with natural frequencies that are significantly
different from those of the other pinnacles are brought to the attention of
the relevant authorities. This enables a physical inspection of any identified
pinnacles to be carried out, and remediation work undertaken if necessary.

Figure 19: BRE laser system used to


measure pinnacle vibration

8.4 praCtiCal USe Of dYNaMiC teStiNg


The use of dynamic testing to identify outlying structural behaviour is
described in Section 8.3.3. Vibration comfort criteria associated with floors
and grandstands are described in Sections 5 and 6.1, respectively. These
criteria are associated with establishing the natural frequency of the structure,
and this frequency can be established by any of the techniques described
in Section 8.2. For structures with natural frequencies that are lower than
a specified range of frequencies, some codes of practice recommend that
specialised investigations are required. In practice this means that a dynamic
investigation is needed, which is likely to involve dynamic testing. Dynamic
testing can also be used to verify structural performance and thus be part
8 dYNaMiC teStiNg
page Of
header
BUildiNg
ri gS
htaNd
– page
BUildiNg
headereleMeNt
subti tS
le 39

of a structure’s certification. For example, to allow a floor to be certified by


BRE Global, both the fundamental frequency and the damping ratio must be
measured experimentally.
On its own, mode shape behaviour has limited practical usage. Mode
shapes do provide information about the best locations to install dampers.
However, their principal importance is that they allow the mode-generalised
mass and mode-generalised stiffness of a structure (which can be found by
dynamic testing) to be converted into the actual mass and stiffness of the
dynamic system. Since:

(Equation 12)

it can be shown that:

(Equation 13)

For certain dynamic tests, the applied mode-generalised force and mode-
generalised stiffness are measured. Hence, the modal displacement can
be determined using Equation 13. Summing the modal displacements for
all modes of vibration gives the total displacement, so the overall structural
stiffness can be determined by dividing the applied force by the total
displacement. This overall structural stiffness is the same as the static stiffness
of the structure. As shown in the example below, stiffness information can be
useful to compare with specified design deflections.

Example: A common beam deflection criterion used in the construction


industry is that a beam must not deflect more than 1 in 200. If a beam
is 5 m long and the combined live and dead design load at the end
of the beam is 20 kN, using Equation 12 the stiffness must be at least
20 x 200/5 = 800 kN/m.

As described above, dynamic testing can be used to measure all of the


parameters necessary to create dynamic models of a structure. These models
can be used to predict the structural response under any fluctuating load
condition. Hence, such models can determine whether or not specified
deflection or acceleration criteria are exceeded.
40 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

9 CONClUSiONS

Apart from specialised buildings, few actual requirements in the UK codes
of practice relate to acceleration criteria, and instead recommendations are
given. It is emphasised that the designer and client need to agree beforehand
the intended building use and its associated serviceability requirements.
There is no single codified method that can be used to assess acceleration
serviceability. This is because the nature of vibration can be continuous or
intermittent, and both of these situations need to be treated in different ways.
The VDV method is an approach that tries to reconcile this situation, and
seems to be gaining increased acceptance. However, this approach is not
universally adopted. It is not clear whether VDV methods are consistent with
acceleration effects caused by wind.
It is not clear which acceleration criteria given in codes of practice,
by trade associations or by industry bodies have precedence over others.
Acceleration serviceability requirements can be different depending on
whether the building or an element within that building is being considered.
Furthermore, the criteria given in the Eurocodes differ depending on the
construction material used. (The issue of complex building motions is not
treated systematically in the codes of practice; this matter is considered in
Appendix B.)
Dynamic testing can provide useful and practical results in the investigation
of building serviceability. However, this testing is specialised in nature, and
expert help is needed (i) to ensure that the testing is planned and carried
out properly, and (ii) to ensure that the findings of the testing are interpreted
correctly.
page header right – page head
10 reefereNCeS
r subtitle 41

10 refereNCeS

[1] Department for Communities and Local Government (DCLG). The Building
Regulations 2000 (England and Wales) as amended by Statutory Instrument 2006
No. 3318. London, The Stationery Office, 2000.

[2] BSI. Guide to evaluation of human exposure to vibration in buildings – Vibration


sources other than blasting. BS 6472-1:2008. London, BSI, 2008.

[3] International Organization for Standardization (ISO). Mechanical vibration and


shock – Evaluation of human exposure to whole-body vibration – General requirements.
ISO 2631-1:1997. Geneva, ISO, 1997.

[4] BSI. Guide to measurement and evaluation of human exposure to whole-body


mechanical vibration and repeated shock. BS 6841:1987. London, BSI, 1987.

[5] Irwin A W. Design of shear wall buildings. CIRIA Report R 102. London,
Construction Industry Research and Information Association, 1984.

[6] European Committee for Standardization (CEN). Eurocode 0: Basis of structural


design. EN 1990:2002+AC:2008. London, BSI, 2009.

[7] BSI. Guide to evaluation of human exposure to vibration in buildings (1 Hz to


80 Hz). BS 6472:1992 (now superseded by BS 6472-1:2008). London, BSI, 1992.

[8] Office of the Deputy Prime Minister. The Building Regulations 2000, Approved
Document A: Structure. London, NBS, 2004.

[9] International Organization for Standardization (ISO). Basis for design of structures
– Serviceability of buildings and walkways against vibrations. ISO 10137:2007. Geneva,
ISO, 2007.

[10] Melbourne W H and Palmer T R. Accelerations and comfort criteria for buildings
undergoing complex motions. Journal of Wind Engineering and Industrial Aerodynamics,
1992, 41 (1–3) 105–116.

[11] International Organization for Standardization (ISO). Guidelines for the evaluation
of the response of occupants of fixed structures, especially buildings and off-shore
structures, to low-frequency horizontal motion (0.063 Hz to 1 Hz). ISO 6897:1984.
Geneva, ISO, 1984.

[12] European Committee for Standardization (CEN). Eurocode 2: Design of concrete


structures – General rules and rules for buildings. EN 1992-1-1:2004+AC:2008. London,
BSI, 2009.

[13] European Committee for Standardization (CEN). Eurocode 5: Design


of timber structures – General – Common rules and rules for buildings.
EN 1995-1-1:2004+A1:2008+AC:2006. London, BSI, 2009.

[14] European Committee for Standardization (CEN). Eurocode 3: Design of steel


structures – General rules and rules for buildings. EN 1993-1-1:2005+AC:2006+AC:2009.
London, BSI, 2010.

[15] European Committee for Standardization (CEN). Eurocode 1: Actions on


structures – General actions – Densities, self-weight, imposed loads for buildings.
EN 1991-1-1:2002+AC:2004+AC:2009. London, BSI, 2010.
42 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

[16] European Committee for Standardization (CEN). Eurocode 1: Actions on


structures – General actions – Wind actions. EN 1991-1-4:2005. London, BSI, 2005.

[17] BSI. Loading for buildings – Code of practice for dead and imposed loads.
BS 6399-1:1996 (now superseded by BS EN 1991-1-7:2006). London, BSI, 1996.

[18] European Committee for Standardization (CEN). UK National Annex to


Eurocode 3: Design of steel structures – General rules and rules for buildings. NA to
EN 1993-1-1:2005. London, BSI, 2008.

[19] NCCI. Vertical and horizontal deflection limits for multi-storey buildings.
SN034a-EN-EU. Available at: www.steel-ncci.co.uk/Clauses/List_NCCIs.htm (accessed
18 March 2011). Ascot, NCCI, 2006.

[20] NCCI. Practical deflection limits for single-storey buildings. SN035a-EN-EU.


Available at: www.steel-ncci.co.uk/Clauses/List_NCCIs.htm (accessed 18 March 2011).
Ascot, NCCI, 2006.

[21] NCCI. NCCI vibrations. SN036a-EN-EU. Available at: www.steelbiz.org (type


‘Vibrations’ into the search field; accessed 18 March 2011).

[22] International Organization for Standardization (ISO). Mechanical vibration and


shock – Evaluation of human exposure to whole-body vibration –Vibration in buildings
(1 Hz to 80 Hz). ISO 2631-2:1989 (now superseded by ISO 2631-2:2003). Geneva, ISO,
1989.

[23] International Organization for Standardization (ISO).Mechanical vibration and


shock – Evaluation of human exposure to whole-body vibration – Vibration in buildings
(1 Hz to 80 Hz). ISO 2631-2:2003. Geneva, ISO, 2003.

[24] European Committee for Standardization (CEN). UK National Annex to


Eurocode 5: Design of timber structures – General rules and rules for buildings. NA to
EN 1995-1-1:2004+A1:2008. London, BSI, 2006.

[25] Institution of Structural Engineers (IStructE), Department for Communities and


Local Government and Department for Culture, Media and Sport. Dynamic performance
requirements for permanent grandstands subject to crowd action: recommendations for
management, design and assessment. London, IStructE, 2008.

[26] Ellis B R and Ji T. The response of structures to dynamic crowd loads. BRE Digest
DG 426. Bracknell, IHS BRE Press, 2004.

[27] Smith A L, Hicks S J and Devine P J. Design of floors for vibration: a new approach
(revised edition). Publication P354. London, Steel Construction Institute, 2009.

[28] Ellis B R and Littler J D. Response of cantilever grandstands to crowd loads.


Part 2: Load estimation. Proceedings of the Institution of Civil Engineers – Structures and
Buildings, 2004, 157 (SB5) 297–307.

[29] Department of Health. Health Technical Memorandum 08-01: Acoustics. London,


The Stationery Office, 2008.

[30] Willford M R and Young P. A design guide for footfall-induced vibration of


structures. Camberley, The Concrete Centre, 2006.

[31] NHS Estates. Health Technical Memorandum 2045: Acoustics – design


considerations. London, HMSO, 1996.

[32] Institution of Structural Engineers (IStructE), Department for Transport, Local


Government and the Regions and Department for Culture, Media and Sport. Dynamic
performance requirements for permanent grandstands subject to crowd action: interim
guidance on assessment and design. London, IStructE, 2001.

[33] Institution of Structural Engineers (IStructE), Department for Transport, Local


Government and the Regions and Department for Culture, Media and Sport. Advisory
note: dynamic testing of grandstands and seating decks. London, IStructE, 2002.

[34] Page A V and Sangarapillai V G. Design methods for domestic floors constructed
using engineering joists. The Structural Engineer, 2008, 86 (17) 36–41.
page header right – page head
10 reefereNCeS
r subtitle 43

[35] NHBC Standards 2010. Part 6 – Superstructure (excluding roofs) (6 of 9). Milton
Keynes, NHBC, 2010.

[36] Moore J F A. Monitoring building structures. Glasgow, Blackie and Son Ltd, 1992.

[37] BSI. Experimental determination of mechanical mobility – Measurement using


impact excitation with an exciter which is not attached to the structure. BS 6897-5:1995.
London, BSI, 1995.
44 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

appeNdix a veCtOrS aNd veCtOr additiON



A vector has magnitude and direction. On a sheet of paper, as shown in
Figure A1, a vector can be considered as an arrow of length a. A Cartesian (x-,
y-) axis coordinate system is also shown on this figure. Length a represents the
amplitude of the vector, and the direction of the arrow relative to the axes
shown represents the direction of the vector.

y-axis

a
ay
a

ax

Figure A1: Two-dimensional pictorial


representation of a vector x-axis

The vector shown in Figure A1 can be considered as having components in


the x- and y-axis directions. If ax and ay are the magnitudes in the x- and y-axis
directions, then by Pythogoras’ theorem:

a2 = ax2 + ay2 (Equation A1)

The third orthogonal direction comes out vertically from the page (or x-, y-
plane), and this direction will be referred to as the z-axis direction. If az is the
magnitude of the vector in the z-axis direction, then by extension of the above
method it is easy to show that in three dimensions:

a2 = ax2 + ay2 + az2 (Equation A2)

Two vectors in the x, y plane (the green arrows with lengths or magnitudes
a1 and a2) are shown in Figure A2. The sum of these vectors (known as the
resultant) is shown as the yellow arrow, the length of which is aR.

a2x

a2y a2

a1x
aR

a1y a1

Figure A2: Addition of two vectors


page header right – page header
appeNdix
subtitle
a 45

Each of the vectors a1 and a2 have components in the x- and y-axis


directions. Again using Pythagoras’ theorem:

aR2 = (a1x + a2X)2 + (a1y + a2y)2 (Equation A3)

By extension into three dimensions, the resultant magnitude of the vector


addition is shown by Equation A4:

aR2 = (a1x + a2X)2 + (a1y + a2y)2 + (a1z + a2z)2 (Equation A4)

The method of adding two vectors together, as described above, is known


as ‘vector addition’.
46 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

appeNdix B BUildiNgS With COMplex


viBratiON MOtiONS
B1 effeCtS CaUSed BY a BUildiNg SWaYiNg iN differeNt
OrthOgONal direCtiONS
The effects caused by buildings vibrating in different horizontal x- and y-
orthogonal directions are discussed by Melbourne and Palmer[B1]. In that paper,
it is shown that the overall lateral response is modelled best by combining the
vector of the uncorrelated x- and y-direction rms accelerations. That is:

(Equation B1)

where ax and ay are the (un-weighted) rms accelerations measured in the x-


and y-axis directions.
The expected peak acceleration, â, can then be determined using arms and
the gust factor g given in Equation 3 of the main text. Since n is different in the
x- and y-axis directions and g is a function of n, then if gx and gy are the gust
factors associated with vibration in the x- and y-axis directions:

(Equation B2)

B2 effeCtS CaUSed BY tOrSiON


In Annex D of ISO 10137:2007[B2], it is stated that the equivalent translational
acceleration is defined as r × Aθ(t), where r is the distance from the centre
of torsion to the objective point and Aθ(t) is the angular acceleration of the
torsional vibration.
In this Appendix, the symbol A is used to denote amplitude. If θ (in radians)
is the angle of twist and the building rotates with a sinusoidal motion at a
natural frequency of n, then the time-varying amplitude of motion is given by
Equation B3:

(Equation B3)

Differentiating this expression with respect to time gives:

(Equation B4)

The quantity is the angular velocity, and in mechanics is often given the
symbol ω.
Differentiating Equation B4 with respect to time gives:

(Equation B5)

The quantity is the angular acceleration of the torsional vibration, Aθ(t).


Comparing Equations B3 and B5, it can be seen that the amplitude of the
angular acceleration is the amplitude of the angular rotation θ multiplied by
(2πn)2.
page header right – page header
appeNdix
subtitleB 47

Since the equivalent translational acceleration is defined as r x Aθ(t), from


Equation B5 the amplitude of this acceleration, aθ is:

(Equation B6)

Since the vibration is sinusoidal, the rms of equivalent translational


acceleration ar,θ is given by Equation B7:

ar,θ = (2πn)2 x A x (1/√2) (Equation B7)

The rms torsional response determined using Equation B7 can be added


vectorally to the x- and y-axis rms accelerations. Thus:

(Equation B8)

B3 effeCtS Of higher MOdeS Of BUildiNg viBratiON


The underpinning work that Melbourne based his acceleration criteria upon
is described by Daryl Boggs in an unpublished paper (‘Acceleration indices for
human comfort in tall buildings – peak or RMS?’). That study considered only
the fundamental (or first) natural frequency of vibration of the two buildings
studied. In ISO 10137:2007[B2], it is stated clearly that the criteria given
therein (which are not the same as Melbourne’s Criteria) are based upon the
fundamental (or first) natural frequency of vibration. There is no advice given
either by Melbourne or ISO 10137 about how to take into account the effects
of higher modes of building vibration.
At this time, it is worth noting that the first three modes of building
vibration usually comprise (i) swaying about the weaker structural building
axis, (ii) swaying about the stronger structural building axis, and (iii) a
predominantly torsional mode of vibration. For buildings whose shear centre
is a large distance away from the centre of twist, combined sway and torsion
modes occur. The natural frequency of these first three modes of vibration
increases with increasing vibration mode. The point is that using the vector
addition method described above (Equation B6), Melbourne’s Criteria can
(and the author believes should) be used to take into account the effects
of these first three modes of vibration. The author also believes that the
fundamental natural frequency of vibration should be used to determine the
vibration threshold. This approach is consistent with the parametric usage of
the building fundamental frequency in the studies that underpin Melbourne’s
criteria.
It is then possible to extend Melbourne’s approach to higher order vibration
modes than those already considered. The accelerations induced by these
higher order modes can be determined in the same way as the contributions
from the lower modes. Usually these contributions reduce with increasing
mode number, and in practice only a limited number of modes are necessary
to obtain values that are sufficiently accurate for engineering purposes. The
rms of each of the modal accelerations can be added vectorally to the other
contributions from the other modes. That is, if the rms contributions of the
first four modes are a1, a2, a3 and a4, the overall (un-weighted) rms level, arms, is
given by Equation B9:

(Equation B9)

Again, the fundamental natural frequency of the building should be used to


determine the acceptable acceleration threshold level. Despite the logic of this
approach, to the author’s knowledge no statements are given in codes of practice
or standards that allow such a method (or any other higher order method) to be
used. However, without including the contributions from higher order modes,
the predicted rms acceleration will be lower than if all of the modes are included.
This is a non-conservative situation, and therefore not desirable.
48 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

As described earlier, the expected peak acceleration, â, can then be


determined using arms and the gust factor g given in Equation 3 of the main
text. Hence:

(Equation B10)

The approach above can be extended to include the significant


contributions that arise from all building vibration modes.

B4 refereNCeS
[B1] Melbourne W H and Palmer T R. Accelerations and comfort criteria for buildings
undergoing complex motions. Journal of Wind Engineering and Industrial Aerodynamics,
1992, 41 (1–3) 105–116.

[B2] International Organization for Standardization (ISO). Basis for design of structures
– Serviceability of buildings and walkways against vibrations. ISO 10137:2007. Geneva,
ISO, 2007.
page header right – page header subtitleC
appeNdix 49

appeNdix C � aCCeleratiON WeightiNg


faCtOrS
C1 preaMBle
To take into account the frequency dependency of the human perception of
lateral acceleration, weighting factors are applied to the accelerations. This
Appendix presents weighting factors given in the International (ISO) Standards
and British Standards.

C2 hOriZONtal WeightiNg faCtOrS


ISO 10137:2007[C1] states that the weighting parameters depend upon
whether or not the critical vibration direction is specified. The weights for
the horizontal (x- or y-axis) direction accelerations considered here are
given in ISO 2631-1:2009[C2]. Table 1 of ISO 2631-1 states that the Wd
frequency-weighting parameters are used for lateral standing and horizontal
recumbent activities. Note that although different frequency-weighting factors
are specified for sitting, these are less onerous and will therefore not be
considered further. Numerical values of the Wd weights (in one-third octaves)
are given in Table 3 of ISO 2631-1; these values are shown as the light-blue
curve in Figure C1.
The frequency-weighting approach used in the British Standards is described
in BS 6472-1:2008[C3]. BS 6472-1 states that the mathematical definitions of
the frequency weightings are given in BS 6841:1987[C4]. For horizontal (lateral)
accelerations, the Wd weighting function is appropriate, and this curve is shown
as the black curve in Figure C1. Although the mathematical expressions used
to obtain each Wd curve are different, comparing the two Wd functions shown
in Figure C1 it can be seen that the curves are very similar. For all practical
purposes, the International (ISO) Standard and British Standard horizontal
acceleration frequency-weighting curves can be considered the same.

C3 vertiCal WeightiNg faCtOrS


In the International (ISO) Standards, acceleration weighting factors are given
in ISO 2631-1:1997[C2]. In Table 1 of that reference it is stated that for vertical
(z-axis) direction accelerations, the Wk frequency-weighting parameters are
used for standing and sitting. Numerical values of the Wk weights (in one-third
octaves) are given in Table 3 of ISO 2631-1, and these values are plotted as
the red curve in Figure C1.
In the British Standards, the vertical Wb and Wg weighting factors are given in
BS 6841:1987[C4]. The Wb weighting factors relate to people standing, and the
Wg weights relate to people sitting. In BS 6841 these weights are tabulated, and
expressions are given that enable the weights to be determined mathematically.
The shapes of the different frequency-weighting curves are shown in Figure C1;
the green curve is the Wb weights, and the purple curve the Wg weights.
The International (ISO) Standard and British Standard frequency-weighting
factors for standing are shown respectively as the red (Wk) and green (Wb)
curves in Figure C1. It can be seen that although the general trends of the plots
50 dYNaMiC COMfOrt
page header lef t –Criteria fOr StrUC
page header subtitUreS
tle

are similar, there are significant differences in the weighting curves. At lower
frequencies the International (ISO) Standard weights are lower, and at higher
frequencies the British Standard weights are lower. This means that in general
it is not possible to state whether or not either set of weights gives more, or
less, conservative results. An important point to note here is that codes of
practice and standards are designed to be self-consistent. Therefore ‘mixing
and matching’ frequency weights from one set of codes with multiplying
factors or VDV values from another set of codes is not recommended.
According to BS 6841:1987[C4], the appropriate frequency weightings to
apply when considering motion sickness (low-frequency vibration) are the Wf
weighting factors. These are shown as the dark blue curve in Figure C1.

10.0

1.0
Frequency weightings

0.1

Figure C1: International (ISO) Standard


and British Standard frequency-
weighting curves 0.01
0.1 1.0 10.0 100.0
Frequency (Hz)

Vertical accelerations Horizontal accelerations


(Wk curve, ISO 2631-1) (Wd curve, BS 6841)
Vertical accelerations Accelerations causing motion
(Wb curve, BS 6841) sickness (Wf curve, BS 6841)
Horizontal accelerations Vertical sitting accelerations
(Wd curve, ISO 2631-1) (Wg curve, BS 6841)

C4 hOSpital flOOr WeightiNg faCtOrS


According to P354[C5], the weighting factors used in assessing hospital floors
are the Wg weights given in BS 6841:1987[C4]. In BS 6841, these weights are
tabulated and expressions are given that enable the weights to be determined
mathematically. The shape of the Wg weighting curve is shown as the purple
curve in Figure C1.

C5 refereNCeS
[C1] International Organization for Standardization (ISO). Basis for design of structures
– Serviceability of buildings and walkways against vibrations. ISO 10137:2007. Geneva,
ISO, 2007.

[C2] International Organization for Standardization (ISO). Mechanical vibration and


shock – Evaluation of human exposure to whole-body vibration – General requirements.
ISO 2631-1:1997. Geneva, ISO, 1997.

[C3] BSI. Guide to evaluation of human exposure to vibration in buildings – Vibration


sources other than blasting. BS 6472-1:2008. London, BSI, 2008.

[C4] BSI. Guide to measurement and evaluation of human exposure to whole-body


mechanical vibration and repeated shock. BS 6841:1987. London, BSI, 1987.

[C5] Smith A L, Hicks S J and Devine P J. Design of floors for vibration: a new approach
(revised edition). Publication P354. London, Steel Construction Institute, 2009.
page header right – page header subtitle 51

Other reports from BRE Trust


Subsidence damage to domestic buildings: lessons
Micro-wind turbines in urban environments: an

learned and questions remaining. September 2000. FB 1


� assessment. December 2007. FB 17

Potential implications of climate change in the built
Siting micro-wind turbines on house roofs. May 2008.

environment. December 2000. FB 2


� FB 18

Behaviour of concrete repair patches under propped


Automatic fire sprinkler systems: a guide to good

and unpropped conditions: critical review of current


practice. June 2009. FB 19

knowledge and practices. March 2000. FB 3

Complying with the Code for Sustainable Homes:
Construction site security and safety: the forgotten
lessons learnt on the BRE Innovation Park.
costs! December 2002. FB 4
� November 2009. FB 20

New fire design method for steel frames with composite


The move to low-carbon design: are designers taking
floor slabs. January 2003. FB 5
� the needs of building users into account?
December 2009. FB 21

Lessons from UK PFI and real estate partnerships:

drivers, barriers and critical success factors. November


Building-mounted micro-wind turbines on high-rise and

2003. FB 6
� commercial buildings. March 2010. FB 22

An audit of UK social housing innovation. February 2004.
The real cost of poor housing. February 2010. FB 23

FB 7

A guide to the Simplified Building Energy Model (SBEM):

Effective use of fibre reinforced polymer materials in


what it does and how it works. April 2010. FB 24

construction. March 2004. FB 8

Vacant dwellings in England: the challenges and costs of

Summertime solar performance of windows with


bringing them back into use. April 2010. FB 25

shading devices. February 2005. FB 9

Energy efficiency in new and existing buildings:

Putting a price on sustainability. BRE Centre for Sustainable


comparative costs and CO2 savings. September 2010.

Construction and Cyril Sweett. May 2005 FB 10


� FB 26

Modern methods of house construction: a surveyor’s


Health and productivity benefits of sustainable schools: a

guide. June 2005. FB 11


� review. September 2010. FB 27

Crime opportunity profiling of streets (COPS): a quick Integrating BREEAM throughout the design process:

crime analysis – rapid implementation approach. a guide to achieving higher BREEAM and Code for

November 2005, FB 12
� Sustainable Homes ratings through incorporation with

the RIBA Outline Plan of Work and other procurement

Subsidence damage to domestic buildings: a guide to


routes. November 2010. FB 28

good technical practice. June 2007. FB 13

Design fires for use in fire safety engineering.
Sustainable refurbishment of Victorian housing: December 2010. FB 29

guidance, assessment method and case studies.
September 2006. FB 14
� Ventilation for healthy buildings: reducing the impact of

urban pollution. January 2011. FB 30



Putting a price on sustainable schools.
May 2008. FB 15
� Financing UK carbon reduction projects. February 2011.

FB 31

Knock it down or do it up? June 2008. FB 16

The cost of poor housing in Wales. April 2011. FB 32



BRE Connect Online
Build on your foundation of knowledge and expertise

WHAT IS BRE CONNECT ONLINE?


BRE Connect Online gives you access to the unrivalled expertise and insight
of BRE – the UK’s leading centre of excellence on the built environment.
BRE Connect Online is an annual subscription service from IHS BRE Press
giving online access to over 1600 BRE titles

WHAT DO I GET? WHAT’S NEW IN 2011?


More than 50 new titles, including:
ALL NEW AND PUBLISHED BRE TITLES
• Airtightness in commercial and public
650 books, reports and guides – research, buildings
innovation, best practice and case studies,
including: • BREEAM In-Use
• The Green Guide to Specification • Design of durable concrete structures
• Designing quality buildings • Environmental impact of floor finishes
• Complying with the Code for Sustainable
• Low-water-use fittings
Homes
� • Sustainable shopfitting equipment
• Roofs and roofing • Ventilation for healthy buildings
• Site layout planning for daylight and sunlight
250 Digests – authoritative state-of-the-art reviews
550 Information Papers – BRE research and how
to apply it in practice
150 Good Building and Repair Guides – illustrated
practical guides to good building and repair work
and much more ...

All this for an annual subscription of only £349 + VAT


Call now on +44 (0) 1344 328038 to find out more
DYNAMIC COMFORT CRITERIA FOR STRUCTURES
A review of UK codes, standards and advisory documents
The effects of vibration are becoming an increasingly important issue in the
design of buildings and building elements. Modern construction methods
mean that buildings are becoming lighter and have less structural damping.
The response of such buildings to imposed vibration is therefore increased,
DYNAMIC COMFORT CRITERIA
so people using the buildings are more likely to experience vibration. Over a
number of years, different serviceability criteria have been developed to ensure
that buildings and building elements are suitable for their intended activity.
FOR STRUCTURES
This BRE Trust Report draws together the themes and disparate (sometimes
A review of UK codes, standards and advisory documents
conflicting) acceleration comfort criteria associated with the vibration
of buildings in the UK. It reviews Eurocodes (including the UK National
Annexes), British Standards, International (ISO) Standards and other sources
of information. The UK Building Regulations and advice provided by UK trade
Gordon Breeze
associations and other industry bodies are also considered.

This report is technical in nature, and is aimed primarily at building designers,


consultants, architects and structural engineers. It is hoped that the complex
and interlinking nature of this subject can be understood by presenting all the
relevant acceleration criteria in one document in a logical and concise way.

y x
RELATED TITLES FROM IHS BRE PRESS
BUILDING-MOUNTED MICRO-WIND TURBINES ON HIGH-RISE AND COMMERCIAL BUILDINGS
FB 22, 2010
THE RESPONSE OF STRUCTURES TO DYNAMIC CROWD LOADS, 2004 EDITION
DG 426, 2004
WIND LOADS ON UNCLAD STRUCTURES
SD 5, 2004

IHS BRE Press, Willoughby Road


Bracknell, Berkshire RG12 8FB
www.brebookshop.com
FB 33

Вам также может понравиться