Sie sind auf Seite 1von 28

ARTICLE IN PRESS

Journal of Wind Engineering


and Industrial Aerodynamics 95 (2007) 843–870
www.elsevier.com/locate/jweia

Wind engineering—Past, present and future


C.J. Baker
School of Engineering, University of Birmingham, Edgbaston, Birmingham B15 2TT, UK
Available online 23 March 2007

Abstract

This paper firstly considers the history of wind engineering in five rather arbitrary time periods—
the ‘‘traditional’’ period (up to 1750), the ‘‘empirical’’ period (1750–1900), the ‘‘establishment’’
period (1900–1960), the period of growth (1960–1980), and the modern period (1980 onwards). In
particular it considers the development of the discipline in terms of the socio-economic and
intellectual contexts of the time. This leads to a description of the current state of the discipline and a
forward look at possible developments, again taking into consideration the likely socio-economic
and intellectual changes in the next few decades.
r 2007 Elsevier Ltd. All rights reserved.

Keywords: Wind engineering; History; Wind loads; Wind tunnels; CFD; Future developments

1. Introduction

This paper, written at the request of the organisers of the 4th European and African
Conference on Wind Engineering, considers first of all the history of wind engineering and
attempts to summarise the current state of the discipline, before then attempting to
consider the prospects for the subject over the next few decades. Now it is relatively easy to
write a history of the subject that simply lists the ‘‘facts’’ and applications of the subject to
various situations, and this is the approach that has been taken by several authors
recently—see for example Aynsley et al. (1977). Cook (1985), Davenport (1999), Surrey
(1999), Meroney (1999), Murakami and Mochida (1999). Indeed parts of this paper draw
heavily on the work of these authors. However, all the developments within wind
engineering have taken place within specific socio-political and intellectual contexts and

Tel.: +44 121 414 5067; fax: +44 121 414 36785.
E-mail address: c.j.baker@bham.ac.uk.

0167-6105/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jweia.2007.01.011
ARTICLE IN PRESS
844 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

cannot properly be considered without a consideration of such contexts. In particular


future developments will inevitably be driven by the prevailing social and intellectual
conditions of specific times and places, and thus if the future prospects of wind engineering
are to be properly considered, some attempt must be made to predict the contexts in which
the developments will take place.
Sections 2–6 of this paper thus considers the history of wind engineering, with due
regard to the prevailing social conditions and intellectual environment. To structure this
discussion we consider five periods—the ‘‘traditional’’ period (up to 1750), the ‘‘empirical’’
period (1750–1900), the ‘‘establishment’’ period (1900–1960), the period of growth
(1960–1980), and the modern period (1980 onwards). These periods are to an extent
arbitrary, and should not be regarded as fundamental in any way—but simply serve as a
useful framework for discussion.
But first some justification of what follows is required. This paper approaches the
consideration of wind engineering history from a fundamentally European point of view—
and indeed many readers may well judge that the viewpoint is even more restrictive and
that the approach is excessively anglo-centric. If an apology is needed for such a
perspective, then this emphasis arises of course because of the author’s experience and
background. Further by way of justification, at least in the European context, it may be
argued that the UK experiences the force of the North Atlantic weather systems and
generally higher wind speeds than most other European countries. This has driven many of
the developments in wind engineering that have found application elsewhere. This point
being made however, developments in wind engineering techniques have of course taken
place throughout Europe and around the world, even if all of these are not fully
represented in this paper.

2. The traditional period (up to 1750)

The historic period up to 1750 of course covers a vast range of different social and
intellectual contexts. In wind engineering terms, in most parts of the world, the style of
structures to withstand prevailing wind conditions evolved by experience and the
development of tradition (Aynsley et al., 1977). These styles were thus inevitably localised
and varied significantly from place to place. I would suggest however that these structures
were also heavily influenced by considerations that are foreign to the modern secular
mind—those of religion and ritual. As an early example of this we can consider the Iron
Age of the Atlantic seaboard of Europe, where a particular style of structure developed
known as the round house, with circular plan and a conical roof (Pryor, 2003). Now the
people of this time and region had a highly developed ‘‘ritual’’ system that made extensive
use of solar and lunar observations, and based much of their daily and seasonal routine
around such systems. In this, circular geometries were of major importance—for example
in burial mounds and the earlier Neolithic and bronze age henge monuments, such as
Stonehenge or Avebury. The round houses themselves (Fig. 1), including their internal
arrangement, were part of such ritual systems, and are very often oriented with major solar
alignments. Nonetheless it is possible to conjecture that this style arose at least partly
because in this area, the windiest part of Europe, the conical roof form experienced
significantly lower loadings than the rectangular roof forms, that are more common in
Central and Eastern Europe.
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 845

Fig. 1. A reconstructed Round House at Flag Fen in Cambridgeshire, UK (Flag Fen, 2004).

The ritual/religious context to structural form and to the effects of natural phenomena,
although alien to modern minds, nonetheless persisted well into historic time. As an
example consider the following two entries from the ‘‘Anglo-Saxon Chronicle’’ that
records events in England from the dawn of the common era to around 1150 (Whitlock,
1961).

793 In this year fierce, foreboding omens came over the land of Northumbria and
wretchedly terrified the people. There were excessive whirlwinds, lightning storms
and fiery dragons were seen flying in the sky. The signs were followed by great
famine and shortly afterwards in the same year the ravaging of heathen men
destroyed God’s church at Lindisfarne through brutal robbery and slaughter.
1122 Thereafter, the Tuesday after Palm Sunday, was a very great wind on March 22nd.
Thereafter came many signs far and wide in England, and many illusions were
seen and heard. The night of July 25th there was a very great earthquake over all
Somerset and Gloucester. Later, on September 8th, St Mary’s day, there was a
very great wind from morning to black night. That same year passed away Ralph,
Archbishop of Canterbury.

The use of religious language for specifying dates is obvious, but note that wind
storms are seen here as divine omens—in the first case of the coming of the Vikings to the
north east of England, and in the second of the death of the Archbishop of Canterbury
amid the tumultuous events in the reign of Henry I of Normandy and England that
led to strained relationships with the Pope. In a rather more secular vein consider the entry
for 1114.
ARTICLE IN PRESS
846 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

1114 This year were very great winds in the month of October, but it was immeasurably
great on the night of the Octave of St Martin, November 18th, and that was seen
everywhere in the woods and villages afterwards.

The observation that trees and homes suffer most from windstorms is of course as valid
now as it was then.
Even within this traditional and religious framework of society however developments
took place. In the 12th and 13th centuries across Europe, the desire to glorify God in
architecture (mixed with not a little rivalry between religious communities) led to ever
larger and grander churches being built, which would inevitably experience greater wind
loads than on earlier, lower structures. Roofing methods evolved to ensure that roofs were
not lost in wind storms—and in particular the use of lead on roofs became widespread.
However perhaps the major evolution can be seen in spires (Encyclopedia Britannica,
2004). These were originally simply tall pyramids with four faces. However, over the
centuries they evolved into polygonal and eventually conical structures (such as in Fig. 2).
Such structural form would experience significantly lower wind loads that the sharp edged
pyramid structures (because of the narrower wakes at supercritical Reynolds numbers) and
it must have been observed fairly quickly by those responsible for the building of such

Fig. 2. The spires at Lichfield Cathedral, Staffordshire, UK.


ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 847

structures that polygonal spires blew down less frequently than pyramid shaped spires.
Over the following centuries forms of spire also developed with significant gaps in the
structure—which again would have reduced drag (such as the central spire on Fig. 2).
Another example of the development of traditional forms can be found in the
development of windmills. These seem to have been first introduced in Persia around 950,
and were very much based on the design of water mills, although these were vertical axis
machines with horizontal blades (Hill, 1984) (Fig. 3). The windmill seems to have been
introduced into Europe around 1180, and records of mills exist in Normandy and in
England at around that date. The number of mills increased rapidly around this time, and
in the 13th century there were 120 mills in the vicinity of Ypres alone. Mills are also
recorded in Italy at around this time, and the first illustrations begin to appear. These
illustrations however show a considerable evolution from the original design, with all the
mills being horizontal axis post mills of the type that was to become very familiar across
Europe for the next few centuries (Fig. 3).
Towards the end of this period, the intellectual atmosphere across Europe changed
significantly with the onset of the Renaissance and the period known as the Enlightenment.
In the 17th and 18th centuries we see the birth of modern science, with contributions from
Newton, Euler and Bernoulli that were of very wide significance and were eventually to
play a major role in the discipline of wind engineering. The influence of organised religion
declined significantly and societies across Europe became increasingly secular. Building
techniques were evolving into more modern forms, as the religious drivers of structural
form were becoming of much less significance. As an example of how such forms coped
with strong winds it is instructive to consider the effects of the ‘‘great storm’’ that affected
Western Europe in 1703. RMS (2003) quotes the work of Defoe (1704), the author of
‘‘Robinson Crusoe’’, who wrote a comprehensive account of the storm’s effects in the UK.
In the rest of Western Europe the War of Spanish Succession was underway, and
newspaper reports there were politically coloured, with claims that the opponent’s ports

Fig. 3. Vertical (early Persian) and horizontal (Medieval Dutch) axis windmills (Hill, 1984).
ARTICLE IN PRESS
848 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

were destroyed. The storm consisted of a severe depression of around 95 mbar that
travelled across the English Midlands on December 7 and 8, 1703 and caused major
damage over a wide area. It was at its peak in London at 6:00 a.m. and in Copenhagen at
11:00 a.m. Estimates of maximum wind speeds are in the range of 45–50 m/s. There is some
evidence of ‘‘streaky’’ damage (which will be discussed further below). The return period
can be estimated to be around 200 years. Defoe writes that in London:
‘‘the streets lay so covered with Tiles and Slates, from the Tops of the Houses,
especially in the Out-parts, that the quantity is incredible, and the Houses were so
universally stript, that all the Tiles in Fifty Miles round would be able to repair but a
small Part of it’’
London at the time was a relatively new city, having been rebuilt after being destroyed
by the Great Fire of 1666, and thus much of the construction was ‘‘modern’’. Nonetheless
around 2000 chimney stacks were broken down in the London area and there was major
gable end damage. Around 16–20 roofs of large buildings were lost. In other words there
was very serious structural damage. In London there were 21 fatalities and around 200
serious injuries, with around 120 fatalities nationally. In the UK outside London the
damage was equally as severe with major roof damage, over 1000 outhouses or barns
destroyed and around 400 windmills toppled over. The total loss in London was estimated
to be around £2 million pounds. Given that the estimate for the total value of the buildings
in all of England and Wales was around £100 million at the time, this represents a very
major impact indeed. In the rest of Western Europe there was similar damage, particularly
in Belgium and Holland, with rather lower levels of damage in Denmark. Thus at the close
of the period, it is evident that the vulnerability to wind storm damage was increasing
significantly.

3. The empirical period 1750–1900

Around the start of this period the social phenomenon that became known as the
Industrial revolution began in earnest. Some writers would see this as finding its first
manifestation in Coalbrookdale in Shropshire in the UK, where the ironmasters of the
area became ever more daring in their art, culminating in the building of the first iron
bridge across the River Severn. This structure was constructed in a traditional style that
would have been used for timber bridges. As such it was in modern terms heavily over
engineered (see Fig. 4) and would most certainly not have experienced aeroelastic
oscillations! Close by, in Birmingham James Watt began with Matthew Boulton to
construct steam engines of ever increasing power. (It is of interest to note that the
‘‘heirloom’’ of office of the Head of Engineering at the University of Birmingham is a chair
once owned by James Watt himself). These and industrial developments elsewhere led to
major economic development across Europe, both at a local and regional scale, through
the provision of much improved road systems and bridges, the development of the railway,
etc. On a wider scale this was the period of the growth of European empires, which was
again driven by the new possibilities that were arising from the industrial revolution, all
within a context of largely unrestrained capitalism.
Intellectually this period saw the development of classical hydrodynamics building on
the work of Euler, Newton and Bernoulli and later through Navier’s formulation of the
fundamental equations of fluid flow in 1845. The use of the techniques of potential flow
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 849

Fig. 4. The Ironbridge in Coalbrookdale (Shropshire Tourism, 2004).

were extensively studied mathematically, and very significant work carried out on the flow
around a variety of body shapes—although with the inconvenience that the solutions that
were obtained did not always seem to match reality. Scientific experimentation was also
gaining in respectability and Hadley and Smeaton carried out the first fluid mechanics
model experiments in 1759. Towards the end of this period, work of immense significance
was carried out by Osborne Reynolds in Manchester which revealed the differences
between laminar and turbulent flow in pipes, and, in retrospect, began the undermining of
the edifice of classical hydrodynamics.
It was in this context that the effects of wind loads on transportation systems came to be
of significance. In Naval terms Admiral Beaufort derived the familiar Beaufort Scale in
1808, to enable reasonably accurate weather information to be given to the military and
merchant fleets. The first long span bridges that were built to improve communication
links inevitably suffered from adverse effects of the wind leading to some spectacular
downfalls—such as the 1836 collapse of the Brighton Chain pier due to aeroelastic
oscillations, and, most famously, the collapse of the Tay bridge in 1879 (Fig. 5). This
resulted in perhaps the most appalling poetry in the English language from William
McGonnegal. (The full text of which is given in Appendix as it seems to the author that it is
essential reading for all wind engineers.)

So the train mov’d slowly along the Bridge of Tay,


Until it was about midway,
Then the central girders with a crash gave way,
And down went the train and passengers into the Tay!
The Storm Fiend did loudly bray,
Because ninety lives had been taken away,
On the last Sabbath day of 1879,
Which will be remember’d for a very long time.

More positively the Tay Bridge collapse led to the experiments of Benjamin Baker in
1884 (Baker, 1884), carried out as part of the building of the Forth Bridge, who measured
the wind forces on rectangular plates, and found that smaller plates experience
ARTICLE IN PRESS
850 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

Fig. 5. The Tay Bridge collapse.

proportionally more load than larger plates, and conjectured, quite correctly, that this was
due to the greater coherence of small rather than large gusts.

4. The establishment period 1900–1960

At the beginning of the 20th century, the industrial revolution was coming of age and
beginning to influence every aspect of society. Nowhere was this more true than in the
military sphere and it is an unfortunate fact that many of the technological advances that
were made in the 20th century were largely driven by military considerations, as the
European empires vied for supremacy both within Europe and around the world. For
example in the study of atmospheric dispersion early experimental and theoretical
developments at Porton Down in the UK were driven by the demands to understand
aspects of chemical warfare (Meroney, 1999). Large-scale technological warfare demands a
collective effort and this period saw the development of the large government laboratories
that were ultimately to have such an influence on wind engineering. By contrast, the
increase in personal mobility led to a breakdown of established social patterns, one side
effect of which was an increasing demand for high-level education by all classes of society.
This led throughout this period to the establishment of more and more Universities, many
with a technological bias, that were again to become active in the wind engineering field.
Technological and material advances also led to attempts to construct more and more
challenging structures—such as the first high-rise structures, and ever longer suspension
bridges. In this context the work of Baker was taken forward at the National Physical
Laboratory in Teddington under the direction of Thomas Stanton, who carried out full-
scale measurements with tower arrays to attempt to find the ‘‘size’’ of wind gusts
(Davenport, 1999). This work also included wind measurements on Tower Bridge in
London.
In terms of intellectual concepts, this period was dominated by the rise of boundary
layer aerodynamics as pioneered by Prandtl and von Karman, which clearly demonstrated
the deficiencies of classical hydrodynamics, together with the birth and development of the
statistical theory of turbulence, first set out by Taylor in 1935. In addition statistical
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 851

techniques were also developing rapidly and the extreme value theory that was to become a
dominating paradigm in wind engineering was first elucidated in 1928 by Fisher and
Tippett (1928).
This era also saw the birth of what the author regards as three of the main wind
engineering tools. Firstly there was the development of the wind tunnel—driven by the
nascent aeronautical industry, although the first wind tunnels predated the advent of
aeroplanes, with the pioneering work of Wenham in 1871 (Surrey, 1999). In 1893 Irminger
measured pressure distributions on a variety of shapes using the flow through a chimney.
Irminger (1895). Eiffel made his first wind tunnel measurements in 1909. In the 1930s
Irmiger made measurements on building models in low turbulence wind tunnels. (Irminger
and Nokkentved, 1936).
Secondly there was the development of codes of practice with the realisation of the need
to provide engineers with practical guidance on design to enable environmental loads such
as wind to be properly defined. The first UK code of practice was published in 1944
(British Standards Institution, 1944).
Thirdly this period saw the beginnings of full-scale measurements of wind loads on
structures. It is in the interaction between wind tunnel and full-scale tests that most
progress was made in the field of wind engineering during this period. The experiments of
Irminger mentioned above were carried out in low turbulence wind tunnels, and many
other similar measurements were made in the years that followed and produced loading
data that was used in codes of practice for many decades after this. A similar set of low
turbulence measurements were those made by Dryden and Hill on the Empire State
Building in 1933. In 1940 Rathbun made equivalent full-scale measurements of the same
building, using banks of anemometers that were photographed to give instantaneous
pressures (Rathbun, 1940). Not surprisingly these full-scale pressures were subject to rapid
fluctuations due to turbulence. Nonetheless it became obvious that these results did not
agree at all well with the wind tunnel results. This was resolved in 1944 with wind tunnel
measurements with an attempt to simulate the atmospheric boundary layer within the wind
tunnel, and indicated the absolute need to simulate the wind characteristics in the wind
tunnel if proper modelling was to be achieved and reliable results obtained. Similar
conclusions were drawn from the full-scale measurements on the pressures on a railway
shed made by Bailey (1933) which were compared with wind tunnel model results
immersed in a deep boundary layer by Bailey and Vincent (1943).
A further wind engineering ‘‘milestone’’ of this era was the collapse of the Tacoma
Narrows bridge, about which a very great deal has been written. The cause is still not fully
understood (Wyatt and Walshe, 1992), but the collapse illustrated the importance of a
proper consideration of the aerodynamic stability of long span bridges, and investigations
after the failure led to the derivation of appropriate wind tunnel scaling laws and wind
tunnel methodology by Farquharson in 1949. Work on long span bridges also took place
at NPL in the UK, with a full bridge study of the Severn Bridge (Frazer and Scruton,
1952), and the development of section model testing, in parallel with similar work in the
USA and Japan.
Towards the end of this ‘‘establishment’’ period the final foundations were being laid for
the rapid development of wind engineering over the coming decades. Van der Hoven
carried out the full-scale measurements that enabled him to derive the van der Hoven
spectrum which, because of the existence of the spectral gap, allowed the concepts of
independent small and large scale wind fluctuations to be formulated, which is of
ARTICLE IN PRESS
852 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

Fig. 6. Scruton, Jensen (from Davenport (1999)) and Cermak.

fundamental importance to the developments that followed over the coming decades (van
der Hoven, 1957). At the National Physical Laboratory in London, Scruton further
developed techniques for the wind tunnel modelling of long span bridges, together with
scaling laws for the wind tunnel modelling of dynamic structures—which ultimately
resulted in the widespread acceptance of a dimensionless number that bears his name
(Scruton, 1963). In Denmark Jensen was formulating the scaling laws for proper boundary
layer testing of low rise structures, and again deriving the scaling factor that was to bear his
name (Jensen, 1958). In the USA Cermak set up the Fluid Dynamics laboratory at
Colorado State University, where much fundamental work was carried out on the
simulation of atmospheric flows with and without thermal effects (Fig. 6). By 1960 the
foundations for the discipline had been securely laid on the much older foundations of the
disciplines from which it has grown—meteorology, fluid mechanics and structural
mechanics.

5. The period of growth 1960–1980

During this short 20-year period there were major changes in the nature of society in
Europe and across the world, and these were paralleled with significant advances within
the discipline of wind engineering. In western society at large, the after effects of the
1939–1945 war were beginning to fade, and this was a period of ever increasing prosperity
and economic development. There was a widespread ‘‘social democratic’’ consensus in
western countries, one side effect of which was the further development and growth of
central government research facilities. One British Prime Minister of the 1960s spoke of
‘‘the white heat of technology’’ and this was a period of optimism concerning the potential
benefits of technology to society (by contrast, and rather chillingly this was also the period
when the fear of nuclear annihilation was at its height). High rise structures and other large
infrastructure projects were the order of the day across the developed world. During this
period computer technology began to develop at an accelerating pace, and there were
massive developments in the design of scientific instruments and in data acquisition
technology.
It was within this context that the discipline of wind engineering developed, with ever-
increasing speed, and achieved some level of self-definition. The papers presented at the
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 853

international conferences (see below) indicate that the primary concern has been with
strong wind climatology and wind structure, wind effects on low rise and high rise
buildings, bridges, masts and towers, wind damage, codification and insurance. The
discipline has concerned itself to a much lesser extent with the study of wind effects on
people (pedestrian comfort, air quality, ventilation, etc.), wind effects on the natural
environment (trees and crops, sand and snow drifting) and to studies of wind energy, these
tending to become the preserve of other, related discipline.
From now on in this paper, I will not mention any individuals, as in historical terms I
believe we are too close to fully appreciate any individual contribution to the subject. That
is, I will not mention any individuals other than one—Prof. Alan Davenport (Fig. 7). The
importance of the contribution made by Davenport to the subject cannot be over
emphasised. In 1961 he elucidated the concept of the wind loading chain, which gave a
conceptual framework to the study of wind effects on structures (Davenport, 1961). He
was heavily involved in the construction of the World Trade Centre. The laboratory he
developed at the University of Western Ontario became the dominant force in
international wind engineering terms. Davenport’s major role having thus been noted,
there were of course many others involved in the development of the discipline during this
period.

Fig. 7. Alan Davenport.


ARTICLE IN PRESS
854 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

Davenport’s concept of the wind loading chain, which he applied in the frequency
domain, led to a range of spectral methods for calculating the loads and displace-
ments of high rise buildings, bridges, etc. These have become very widely used
throughout the discipline, and indeed have become the dominant class of
analytical method, although by their nature they implicitly assume linear structural
behaviour.
These decades saw the development of the boundary layer wind tunnel from an
essentially research tool, into a reliable and robust tool for commercial design purposes,
with the increasing realisation of the need to model the turbulence spectrum as
accurately as possible and with the routine use of small pressure transducers with
scanivalves, and the introduction of the base balance techniques. Techniques
for the measurement and prediction of atmospheric pollutants also advanced rapidly
(Meroney, 1999), and in 1961 Pasquil developed his classification of atmospheric stability
that was to remain in use for many decades. Around the world a number of ground
breaking full-scale experiments took place—the Aylesbury house experiment (Eaton
and Mayne, 1975) and the mobile home measurements in the USA (Marshall, 1975).
Perhaps most significantly for the long-term future of the discipline, the series of
international conference began (Table 1). Davenport (1999) noted that the first of these
was somewhat different in character from the rest, with many fewer papers, and much
greater industrial involvement than the conferences that followed. Perhaps something has
been lost in this change.
During this period, the process of codification of wind effects began in earnest,
and a significant number of codes were developed by National Standards Organisations—
for example the updated UK code (British Standards Institution, 1972) and the
Australian Code (Standards Association of Australia, 1973). One particularly
useful development that has been of long-term importance is the production of the
ESDU Wind Engineering series, which has become an invaluable data source for many
designers over recent decades (ESDU, 2004). Finally in 1975 the first edition of the
Journal of Industrial Aerodynamics was published (later to become the Journal of
Wind Engineering and Industrial Aerodynamics) and quickly established itself as the
main journal of record in the field.

Table 1
The development of the International Wind Engineering Conferences (updated from Davenport, 1999)

Year Location Papers Attendees Countries

1963 Teddington 24 300 20


1967 Ottawa 42 85 20
1971 Tokyo 113 215 26
1975 Heathrow 70 200 25
1979 Fort Collins 250 350 25
1983 Gold Coast 120 — —
1987 Aachen 90 — —
1991 London, Ont. 264 385 35
1995 Delhi 200 — 25
1999 Copenhagen 290 — —
2003 Lubbock 238 (+126 posters) 380 32
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 855

6. The modern period 1980–

Whilst all divisions of historical periods in this paper are arbitrary, this is particularly
the case with any division that marks the start of a ‘‘modern’’ period. Developments within
wind engineering were continuous across the boundary. Perhaps the real reason that the
author has chosen this date is that 1980 is the time when he first became involved in the
discipline! However, the date does seem to mark, even if in an arbitrary fashion, the
maturity of the discipline, with the adoption of the Journal of Wind Engineering and
Industrial Aerodynamics as the official journal of the International Association of Wind
Engineering.
More seriously at around that time, major societal changes were beginning. The
prevailing social democratic consensus was beginning to break down, and individualism
and free choice became the dominant themes of political discourse. This was the era of
privatisation in a number of western countries, with the closure or commercialisation of
national laboratories and industries. There was a general lessening of public confidence in
science and technology and development slowed down, although there was growth in the
service industries—particularly in financial services. The discipline of wind engineering was
seriously affected by such changes with the closure/commercialisation of many national
building institutes for example (often with a deplorable loss of corporate memory of past
work). There was a reduction in the number of large developments within the developed
world to be tested at model scale, although this lack of work was often compensated for by
an increased workload on large high-rise structures in the developing world.
These points being made there have been major developments within the subject over
this period. There have been significant advances in wind tunnel testing techniques,
particularly in terms of instrumentation, with the use of large number of simultaneously
monitored pressure transducers and the increasingly frequent use of LDA and PIV
techniques for velocity measurements. Similarly the development of the three component
sonic anemometer has revolutionised full-scale measurements. Further major full-scale
experiments were carried out in South Africa (Milford et al., 1991a, b), Silsoe in the UK
(Richardson et al., 1995; Hoxey et al., 1995) and at Texas Tech (Levitan and Mehta,
1991a, b). At the time of writing an extensive project is underway in the USA to measure
wind conditions and full-scale structural loading during hurricanes, which should yield a
very considerable quantity of information that will be of significant use in design (Gurmley
et al., 2005). All these developments have of course been underpinned by the rapid growth
in IT techniques and computer power which makes high-speed data acquisition and the
analysis of large amounts of experimental data possible. It has also led to the increasing
use of what is now the fourth fundamental tool of wind engineering—CFD techniques.
Computational fluid dynamics has progressed immensely over the past two decades–
through the use of inviscid panel methods; then simple k–e techniques, which were
afterwards refined in various ways to make them more suitable for wind engineering
application; and now increasingly through unsteady flow methods such as LES, DES and
discrete vortex modelling. For a full review of CFD developments in Wind Engineering,
see Murakami and Mochida (1999). The last 20 years have also been extremely busy in
terms of code development and revision across the world. These are well summarised in the
recent series of papers produced by the IAWE Codification Initiative (Wind and
Structures, 2005). In conceptual terms the period has seen an increasing application of
modern analytical methods to wind engineering—particularly advanced probabilistic
ARTICLE IN PRESS
856 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

techniques, wavelet analysis, orthogonal decomposition, etc. (see Kareem, 2003). Of


particular significance has been the gradual trend towards using time domain methods in
the design process. This will be discussed further below.
As the discipline has developed, the regular international conferences have become focal
points for the academic and research communities within Wind Engineering. At the 1983
conference the International Association of Wind Engineering was formed, whose primary
role for the first 15 or 20 years of its existence was the organisation of the International
Conferences (Solari, 2005). These have come to be supplemented by a regular series of
regional conferences for the three IAWE regions—Americas (from 1970), Asia/Pacific
(from 1985) and European/African (from 1993 onwards). Over this period a number of
national or regional wind engineering associations have been formed, and this process has
been greatly stimulated by the more formal constitution recently adopted for the IAWE in
2003. Other conference series have been initiated that are regularly attended by the Wind
Engineering community—the Bluff Body Aerodynamics and its Applications series that
began in Kyoto in 1989, and the Computational Wind Engineering Conferences that began
in 1992.
This period has also seen a seeming increase in the frequency of major wind disasters
both in temperate and in tropical regions. There are a number of possible reasons for this
including the increase in population densities in vulnerable areas, the growth of informal
megacities and, almost certainly, the first effects of anthropogenically induced climate
change. Whatever the reason there have been a number of storms that have had major
impacts on the regions they have affected. It is appropriate to consider the effects of such
storms at this point, since they illustrate something of the current understanding and
preparedness for large wind storms. In particular we consider four ‘‘events’’, two near the
start of the period—the 1987 storm in the UK and Hurricane Gilbert in the Caribbean in
1988—and two near the end—the late 1999 storms Lothar and Martin in Western Europe
and Hurricanes Charley, Francis, Ivan and Jean in the Carribean in 2004.
The October 1987 storm was a major cyclonic depression that passed across the south
east of England, hitting London in the middle of the night at around 3:00 a.m. (Buller,
1988). The full extent of this storm was not well forecasted, largely due to a lack of
offshore meteorological stations, and came as a complete surprise to the public. The
maximum wind speeds are shown in Fig. 8. The ‘‘streaky’’ nature of the damage suggested
the existence of large-scale structures within the main cyclonic depression. These have been
described by Browning (2003) and Browning and Field (2004) and given the name ‘‘Sting
Jets’’ because of the characteristic ‘‘scorpion tale’’ pattern of clouds that are associated
with them. Gust speeds of around 40–45 m/s were recorded in the south of England, with
gusts of up to 50 m/s near the coast and over open sea. The return period was around
150–200 years. In the whole of the UK there were around 18 casualties and very
considerable damage and disruption was caused across the south of England. However,
most of this damage was due to tree fall and power supply failure. Whilst there was
significant cladding damage to buildings, particularly to older buildings that were built
before the development of codes of practice, the lack of major structural damage was very
noticeable (Fig. 9). In other words all the modern structures, which had been designed to
be wind resistant, survived, and the majority of structures which had been designed to
various generations of wind loading code of practice, performed more than adequately.
In late December 1999 the storms Lothar and Martin moved across Northern France
and into Germany in quick succession on December 26 and December 27. (Munich Re,
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 857

Fig. 8. Gust speeds in the 1987 storm (from Meteorological Office, 2005).

2002). The storm tracks are shown in Fig. 10. Maximum wind speeds of 40–50 m/s were
recorded, which one presumes would be near or above the design wind speeds for
engineered structures. The insured losses in the two storms were 11.5 and 4 billion euros
respectively, with 110 fatalities in Lothar and 10 in Martin. In both storms there was very
significant fac- ade and roof damage to buildings (including chimney fall), with cranes and
overhead power lines being blown down. Blackmore and Delpeche (2002) showed that
much of the roof damage was to relatively old tiled roofs and chimneys, whilst a major part
of the fac- ade damage was caused by flying debris. There was very significant disruption to
electricity supplies with millions of consumers being without power in France for several
days, and major disruption to transport services of all types. Much of this disruption was
due to fallen trees and there was significant forest damage. All in all the effects were similar
to the 1987 storm in the UK—much cladding and roof damage, and major disruption to
urban life. However, the warnings for these storms were significantly better than for the
1987 storm, because of the enhanced meteorological networks.
Hurricane Gilbert first made landfall on September 12, 1988, when it struck the island of
Jamaica, with torrential rains and winds in excess of 110 mph (Fig. 11). The hurricane
caused buildings to loose their roofs and destroyed power supply lines, cutting off the
island’s 2.3 million residents from the outside world. The centre of the storm hit the capital
of Kingston, where wind gusts up to 140 mph were recorded, seriously damaging the
Kingston airport and tossing airplanes across the tarmac. In total 26 people were killed
and 500,000 were left homeless. Gilbert then moved on to the Cayman Islands, and
140 mph winds were recorded on September 13. By the time it had reached the Yucatan
peninsula wind speeds of 175 mph were recorded, and a historic low-pressure reading of
ARTICLE IN PRESS
858 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

Fig. 9. Damage in the 1987 storm (Buller, 1988).

888 mbars was measured. On September 14, the eye of the storm passed over the Mexican
resort areas of Cozumel and Cancun with 160 mph winds. As Gilbert passed over the
Yucatan Peninsula, 23-ft waves and continuing high wind speeds were recorded, which
devastated the beachfront resorts and forcing tens of thousands to evacuate the area. In
total 29 people were killed and the hurricane caused more than $880 million in damage and
left nearly 200,000 homeless. The storm then weakened as it moved north over the Gulf of
Mexico and the USA. Whilst the meteorological forecasts were on the whole accurate, in
many instances these were not communicated to those likely to be affects—and there is
colloquial evidence that tour companies were still moving tourists into the path of the
hurricane up to a few hours before the hurricane hit.
The 2004 Caribbean Hurricanes were the most severe series of storms in the region for a
number of decades. They tracked from East to West in succession over a period of a few
weeks—Charley (12–14 August), Francis (1–4 September), Ivan (14–16 September) and
Jean (15–17 September). Storm tracks are shown in Fig. 12. The effect these storms had on
different countries very much depended upon the state of development of these countries,
their preparedness for such events, and to a significant extent the political systems in these
countries. In almost all countries there was some attempt to warn the population before
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 859

Fig. 10. Storm Tracks of Lothar (top) and Martin (bottom) (from Munich Re, 2002).
ARTICLE IN PRESS
860 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

Fig. 11. Hurricane Gilbert storm track.

the storms, and to evacuate vulnerable areas. The efficiency of this process however
varied significantly across the region—with large-scale (repeated) evacuations in Florida
and in Cuba (where the process was more or less enforced on the population). In
other countries there was simply nowhere to be evacuated to, and the population managed
as it could in local shelters. In general however warnings were given far enough in advance
for preparations to be made, and communications were restored fairly quickly after the
events This would not have happened a decade or so ago, and there has clearly been
significant technical progress in the field of forecasts and warnings that has been of major
significance. In all areas however there was major damage to structures, disruption of
power and transport systems for many days, loss of crops, contamination of water
supplies, etc.
In the light of such major effects it has been heartening to see the involvement of wind
engineers in the IDNDR. In writing of this activity Davenport (1995) began to see wind
engineering in its wider context and advocated building design for hurricane resistance,
hazard identification, education, legislation, insurance. This is surely a trend to be
welcomed by the wind engineering community.

7. The future

It has been shown in the preceding sections that to a large extent, developments in the
discipline that has come to be known as Wind Engineering have been driven by various
social and economic circumstances, and by the prevailing intellectual environment. This is
likely to continue to be the case over the next few decades, and thus to consider future
prospects for wind engineering, the likely contexts need to be considered. These contexts
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 861

Fig. 12. Storm tracks of 2004 storms.

are likely to be different in different parts of the world. In parts of the less developed
world, there will be a need for fundamental infrastructure development, in terms of
transport links, energy infrastructure, etc. In other parts of the world (for example China
and various countries in South East Asia) urban development will continue at a pace
dependent upon the growth of local economies. In Europe and North America the socio-
economic drivers for the development of wind engineering are likely to change
significantly, as the infrastructure is well developed—for example there are few long span
bridges left to be built. Whilst there will continue to be a need for wind engineering
involvement in the design of prestige large structures, the focus is likely to turn towards the
need to sustain and maintain the existing infrastructure, to develop energy sources that are
sustainable in the long term. The desire to improve the quality of life will also become an
increasingly important driver. For all parts of the world the effects of climate change on
the frequency and intensity of wind storm are likely to be important.
In intellectual terms, it is likely that the continued and seemingly limitless development
in computing power will become of increasing importance—and thus computational
techniques will become of increasing practical use. Similarly the increase in computing
power will mean that time series analysis of wind loading becomes increasingly attractive,
and may well displace some of the well established statistical and frequency domain
methods that have been used in the past, and will thus be able to deal more satisfactorily
with non-linear structural behaviour.
ARTICLE IN PRESS
862 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

Thus on the basis of the above assumptions for the likely socio-economic and
intellectual contexts, I would see the following trends emerging in the discipline of wind
engineering.

(a) In the less developed world, there will be an increasing need for wind engineering
expertise in the development of appropriate, sustainable, wind energy systems. This
will involve both the development of novel turbines using locally based material, and
also a significant amount of fundamental work to fully specify local wind climates. In
such areas, wind engineering expertise could also be useful in the alleviation of a
number of other problems—such as wind damage to agriculture, shelter belt design,
etc.
(b) In the developed world there will be a continued need for wind engineering input into
the design of new infrastructure and buildings. In Europe and similar areas, the
emphasis of wind engineering will change from involvement in the design of new
structures, to involvement in the maintenance and improving the sustainability of
existing infrastructure. Indeed the major impact made by recent storms in the UK (of
one to two years return period) has been to cause major disruption to transport and
power supply links through tree fall, flying debris, etc. The incidence of structural
damage has been small (Baker, 2004). Where it has occurred, it is usually limited to
cladding damage that (as noted by ABI, 2003) could usually be alleviated by better
maintenance of the building stock.
(c) The requirement for sustainable energy sources, and the need to reduce greenhouse gas
emissions will also lead to an increasing need for the development of wind energy
systems within developed countries. In addition the improvement of air quality will
become of increasing importance. If wind engineers are to become more involved in
both these developments, then more consideration needs to be given to the structure of
low speed winds than has been the case in the past—particularly the effects of
atmospheric stability.
(d) A full understanding of the effects of climate change on wind storm frequency and
intensity is required in order to establish appropriate design risks. Current predictions
from the use of large-scale Global Climate Models are ambivalent (Munich Re, 2002;
ABI, 2003) but the most recent suggest that the major effect of climate change in
European terms is to weaken the high pressure region that lies over the centre of
Europe in the winter period, which will allow a southward drift of storm tracks. This
will result in higher extreme wind speeds over the south of England, France, Germany,
etc (McDonald, 2004). In the author’s opinion one of the pre-requisites for such an
understanding is to determine the reasons for the differences in the design wind speeds
for different countries in Europe (as given in the National Annexes for the
Eurocodem—see Fig. 13 from Miller, 2003) to establish a reliable base for
understanding the likely effects of climate change. Methods based on gradient wind
condition and surface pressure measurement such as that developed by Miller (2003)
would seem to be appropriate here in harmonising design wind speeds across Europe
(Fig. 14).
(e) There is a need to revisit some of the basic assumptions of wind engineering in the light
of recent advances in measurement and analysis techniques. For example there are
recent indications (Jensen, 1999; Courtney and Troon, 1990; Richards et al., 1999) that
the existence of a ‘‘spectral gap’’ in the van den Hoven spectrum (which is the basis of
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 863

Fig. 13. Once in 50 year wind speeds in Ireland, UK, and France from code data (Miller, 2003).

Fig. 14. Once in 50 year wind speeds in Ireland, UK, and France from surface pressure data (Miller, 2003).
ARTICLE IN PRESS
864 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

much wind engineering analysis) is not as clear as might be supposed (Fig. 15), and
there may well be significant interactions between micro- and macro-meteorological
wind conditions, which would have implications for the specification of design wind
speeds. In a similar way recent work by colleagues of the author based on the
conditional sampling of extreme gust velocities from full scale high response sonic
anemometer measurements have indicated that the structure of atmospheric turbulence
at full scale and that simulated in the wind tunnel are somewhat different (Fig. 16).
Also it would be wise to give further consideration to the commonplace assumption
that Reynolds number effects can be neglected provided a minimum value is achieved
in wind tunnel tests—the large Reynolds number full-scale measurements of Hoxey et
al. (1997) for low rise buildings, the measurements of Larose and D’Auteuil (2004) for
large cylinders, and measurements made by the current author for wind loads on
trains, and (Fig. 17) give pause for thought in this regard.

0.3
Normalised spectral density

0.25
0.2
0.15
0.1
0.05
0
1.00E-08 1.00E-06 1.00E-04 1.00E-02 1.00E-00
Frequency (Hz)

Denmark UK

Fig. 15. The ‘‘van der Hoven’’ spectrum as measured by Courtney and Troon (1990) in Denmark and Richards et
al. (1999) in the UK.

Conditional sampling around peak values


(Velocity - mean) / standard deviation
3

2.5

1.5

0.5

0
-10 -8 -6 -4 -2 0 2 4 6 8 10
Time from peak (sec)

Full scale Wind tunnel geometric scale

Fig. 16. Comparison of turbulence structure at full scale and in wind tunnels.
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 865

1 1
Side force coefficient

Lift force coefficient


0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 5 10 15 0 5 10 15
10m wind speed m/s 10m wind speed m/s

30 60 90
120 150

Fig. 17. ‘‘Reynolds number effects’’ on a full-scale train coach (figure shows variation of side and lift force
coefficients for a variety of wind angles, with 901 normal to the vehicle). At 10 m/s the Reynolds number based on
train height is approximately 2  107.

(f) There can be expected to be an ever-increasing use of CFD techniques, to predict both
wind environment and, in the medium to long term, loads, and a corresponding fall in
the use of boundary layer wind tunnels. The CFD techniques that will prove to be of
most use will be those that will faithfully model the turbulence structure within the
atmospheric boundary layer, e.g. LES or DES techniques. The use of RANS based
techniques will decrease over time, although their relative simplicity and economy will
ensure their continued use for many applications. Applications will become widespread
in areas where wind velocities rather than surface pressures are required, such as the
assessment of pedestrian comfort. These trends may well lead to the concentration of
boundary layer wind tunnel testing for complex structures into a smaller number of
institutions over the next few decades.
(g) There will be an increasing realisation of the importance of non-stationary wind
systems, such as thunderstorms, frontal systems, etc., particularly where wind loading
is important for serviceability considerations. New types of physical and numerical
modelling techniques will be developed to simulate a variety of such unsteady, non-
stationary wind effects—for example thunderstorm and tornado generators.
(h) The use of time series analysis will replace the use of statistical techniques to some
extent—for example in the input of fluctuating wind loads directly into unsteady finite
element codes to predict specific load effects. An example of these developments was
presented at a UK Wind Engineering Society meeting in February 2005 (Cunningham,
2005), which described the calculation of frame force and moment time histories using
pressure time histories from a large number of pressure tappings on a wind tunnel
model, and the time domain application of a finite element model. Such techniques will
allow a full consideration of the loads due to the non-stationary wind systems
described above, as well as allowing the effects of structural non-linearity to be
considered. A consequence of this will be a reduced need to carry out loading
calculations using frequency domain methods.
ARTICLE IN PRESS
866 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

(i) Davenport (1999) wrote

As Martin Jensen has reminded us the lack of full scale verification that has been
tolerated is ‘‘embarrassing’’. It is not characteristic of other technologies such as
shipping, transportation or aeronautics.
The quote remains relevant and there is a continuing need for full-scale
experimentation to validate loading data in codes, the use of wind tunnels and
CFD.
(j) It is likely that the above developments will lead to another round of code revision
in two or three decade’s time, and radically different types of code can be
expected that will allow time domain calculations of load using standard wind and
loading datasets.
(k) To enable the wind engineering community to contribute to improving the
maintenance and sustainability of existing infrastructure, more involvement with
other ‘‘stakeholders’’ is required—for example with emergency planners (to develop
appropriate methods for storm prediction, and for reducing building vulnerability),
city planners and engineers (to develop rational methods for maintenance of urban
infrastructure to withstand wind storm), environmental scientists (to study the effect of
the dispersion of pollutants in urban areas, which has a major effect upon quality of
life), and transportation professionals (to reduce the vulnerability of road and rail links
to wind storms), etc.

If the above developments are to come about, the author is of the opinion that the
education sector needs to play a major role in a number of areas.

(a) In the author’s experience, which is of course very largely UK based, there is often a
major vocabulary gap between wind engineers and others with whom they need to
work, and very often much research is not applied as fully as it could be for that
reason. There is a need for wind engineers to be able to communicate the fundamentals
of their subject in a wide variety of ways—through the encouragement of
undergraduate and postgraduate modules in this area, suitably constructed for, say,
architectural students, urban planners, maintenance engineers, etc., through the
provision of short continuing educational courses for a wide range of professional and
managerial groups.
(b) In addition there is probably a place for a small number of specialist Wind Engineering
programmes at Masters cycle level in Europe to train specialists in the field—perhaps
through the ERASMUS MUNDUS scheme.
(c) Associated with such initiatives there is a need that is growing ever more urgent for the
thorough archiving and cataloguing of much of the report literature of the last few
decades, so that the information contained in such reports is available to all involved in
the discipline.
(d) Research funds, at both national and international level, are becoming more and more
directed to multi-disciplinary rather than discipline specific projects. If wind
engineering is to survive within universities, the development of multi-disciplinary
research teams is becoming more and more vital in the effort to obtain research
funding.
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 867

As can be seen from the above, the discipline of wind engineering remains active and
vibrant, with the need for much further research and development. But it is perhaps
appropriate at this point to recall the fundamental rationale of the discipline–to help
societies to cope with one of the most destructive forces of nature. This was perhaps
summed up by McGonnogal over 100 years ago when he concluded his offering on the Tay
Bridge disaster, with the last two lines summing up the whole discipline of wind
engineering.

Oh! ill-fated Bridge of the Silv’ry Tay,


I must now conclude my lay
By telling the world fearlessly without the least dismay,
That your central girders would not have given way,
At least many sensible men do say,
Had they been supported on each side with buttresses,
At least many sensible men confesses,
For the stronger we our houses do build,
The less chance we have of being killed

Appendix. The Tay Bridge Disaster by William McGonnegal

Beautiful Railway Bridge of the Silv’ry Tay!


Alas! I am very sorry to say
That ninety lives have been taken away
On the last Sabbath day of 1879,
Which will be remember’d for a very long time.

‘Twas about seven o’clock at night,


And the wind it blew with all its might,
And the rain came pouring down,
And the dark clouds seem’d to frown,
And the Demon of the air seem’d to say-
‘‘I’ll blow down the Bridge of Tay.’’

When the train left Edinburgh


The passengers’ hearts were light and felt no sorrow,
But Boreas blew a terrific gale,
Which made their hearts for to quail,
And many of the passengers with fear did say-
‘‘I hope God will send us safe across the Bridge of Tay.’’

But when the train came near to Wormit Bay,


Boreas he did loud and angry bray,
And shook the central girders of the Bridge of Tay
On the last Sabbath day of 1879,
Which will be remember’d for a very long time.
ARTICLE IN PRESS
868 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

So the train sped on with all its might,


And Bonnie Dundee soon hove in sight,
And the passengers’ hearts felt light,
Thinking they would enjoy themselves on the New Year,
With their friends at home they lov’d most dear,
And wish them all a happy New Year.

So the train mov’d slowly along the Bridge of Tay,


Until it was about midway,
Then the central girders with a crash gave way,
And down went the train and passengers into the Tay!
The Storm Fiend did loudly bray,
Because ninety lives had been taken away,
On the last Sabbath day of 1879,
Which will be remember’d for a very long time.

As soon as the catastrophe came to be known


The alarm from mouth to mouth was blown,
And the cry rang out all o’er the town,
Good Heavens! the Tay Bridge is blown down,
And a passenger train from Edinburgh,
Which fill’d all the peoples hearts with sorrow,
And made them for to turn pale,
Because none of the passengers were sav’d to tell the tale
How the disaster happen’d on the last Sabbath day of 1879,
Which will be remember’d for a very long time.

It must have been an awful sight,


To witness in the dusky moonlight,
While the Storm Fiend did laugh, and angry did bray,
Along the Railway Bridge of the Silv’ry Tay,
Oh! ill-fated Bridge of the Silv’ry Tay,
I must now conclude my lay
By telling the world fearlessly without the least dismay,
That your central girders would not have given way,
At least many sensible men do say,
Had they been supported on each side with buttresses,
At least many sensible men confesses,
For the stronger we our houses do build,
The less chance we have of being killed.

References

ABI, 2003. The vulnerability of UK property to windstorm damage, Association of British Insurers Report, 47pp.
Aynsley, R.M., Melbourne, W., Vickery, B.J., 1977. Architectural Aerodynamics. Applied Science Publishers
Ltd., London.
ARTICLE IN PRESS
C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870 869

Bailey, A., 1933. Wind pressures on buildings, ICE Selected Papers, 139, London.
Bailey, A., Vincent, N.D.G., 1943. Wind pressures on buildings, including effects of adjacent buildings. J. Inst.
Civil Eng. 20, 243–275.
Baker, B., 1884. The forth bridge. Engineering 38.
Baker, C.J., 2004. Wind effects in the urban environment—considerations for human health, comfort and safety,
ISWE Tokyo.
Blackmore, P., Delpeche, P., 2002. A comparison of wind damage to structures in France and the UK. COST C14
Conference, Nantes.
British Standards Institution, 1944. Code of functional requirements of buildings. Loading Code of Practice CP4
(Chapter V).
British Standards Institution, 1972. Code of basic data for the design of buildings. Chapter V loading. Part 2
Wind loads. Code of Practice CP3.
Browning, K.A., 2003. The sting at the end of the tail: damaging winds, Forecasting Research Technical Report
388, The Met. Office, UK.
Browning, K.A., Field, M., 2004. Evidence from Meteosat imagery of the interaction of sting jets with the
boundary layer. Meteorol. Appl. 11 (4), 277–290.
Buller, P.S.J., 1988. The October gale of 1987: damage to buildings and structures in the south east of England.
Building Research Establishment Report 138.
Cook, N.J., 1985. The Designers Guide to Wind Loading of Building Structures. BRE, Butterworths.
Courtney, M.S., Troon, 1990. Wind speed spectrum for one year of continuous 8 Hz measurements. In:
Proceedings of the 9th AMS Symposium on Turbulence and Diffusion, Boston, pp. 301–304.
Cunningham, A., 2005. The direct use of wind tunnel time history pressure data in structural design. Presentation
made at UK Wind Engineering Society, London 1/2/5.
Davenport, A.G., 1961. The application of statistical concepts to the wind loading of structures. Proc. ICE 19,
449–471.
Davenport, A.G., 1995. The role of wind engineering in the reduction of natural disasters. In: Proceedings of the
9th International Conference on Wind Engineering, vol. 5, Delhi, India, pp. xi–xviii.
Davenport, A.G., 1999. The missing links, In: Proceedings of the 10th International Conference on Wind
Engineering, Copenhagen, pp. 3–15.
Defoe, D., 1704. The Storm.
Eaton, K., Mayne, J.R., 1975. The measurement of wind pressures on two storey houses at Ayelsbury. J. Ind.
Aerodyn. 1 (1), 67–109.
Encyclopedia Britannica, 2004. /http://www.britannica.comS.
ESDU, 2004. Engineering Sciences Data Unit Wind Engineering Series, vols. 1–4 (9 parts).
Fisher, R.A., Tippett, L.H.C., 1928. Limiting forms of the frequency distribution of the largest or smallest
member of a sample. Proc. Cambridge Philos. Soc. 24, 180–190.
Flag Fen, 2004. Flag Fen—Britain’s Bronze Age Centre http://www.flagfen.com/www.flagfen.comS.
Frazer, R.A., Scruton, C., 1952. A summarized account of the Severn Bridge aerodynamic investigation.
Aerodynamics Division National Physical Laboratory NPL/Aero/222.
Gurmley, K., Masters, F., Prevatt, D., Reinhold, T., 2005. Hurricane data collection: FCMP deployments during
the 2004 Atlantic Hurricane Season. In: 10th Americas Conference on Wind Engineering, Baton Rouge,
Louisiana.
Hill, D., 1984. A History of Engineering in Classical and Medieval Times. Routledge, London.
Hoxey, R.P., Richards, P., Richardson, G.M., Robertson, A.P., Short, J.L., 1995. The Silsoe structures building;
the completed experiment, Part 2. In: Proceedings of the 9th International Conference on Wind Engineering,
Delhi, pp. 1115–1126.
Hoxey, R.P., Robertson, A.P., Richardson, G.M., Short, J.L., 1997. Correction of wind tunnel pressure
coefficients for Reynolds number effect. J. Wind Eng. Ind. Aerodyn. 69–71, 547–555.
Irminger, J.O.V., 1895. Nogle Forsog over Trykforholdene paa Planer og Legemer paavirkede of
Luftstromninger. Engineering News and Engineers.
Irminger, J.O.V., Nokkentved, C., 1936. Wind Pressures on Buildings; Experimental Researches. Danmarks
Naturvidenskabelige, Samfund.
Jensen, M., 1958. The model law for phenomenon in the natural wind. Ingenioren Int. Ed. 2.
Jensen, N.O., 1999. Atmospheric boundary layers and turbulence. In: Proceedings of the 10th International
Conference on Wind Engineering, Copenhagen, pp. 29–42.
ARTICLE IN PRESS
870 C.J. Baker / J. Wind Eng. Ind. Aerodyn. 95 (2007) 843–870

Kareem, A., 2003. A tribute to Jack E Cermak—wind effects on structures: a reflection on the past and
outlook for the future. In: Proceedings of the 11th International Conference on Wind Engineering, Lubbock,
pp. 1–28.
Larose, G.L., D’Auteuil, A., 2004. Experiments on 2-D rectangular prisms at high Reynolds numbers in a
pressurized wind tunnel. In: Proceedings of the 4th Conference on Bluff Body Aerodynamics and its
Applications, Ottawa, pp. 177–180.
Levitan, M.L., Mehta, J., 1991a. Texas Tech field experiments for wind loads. Part 1. Building and pressure
measuring system. In: Proceedings of the 8th International Conference on Wind Engineering, London,
Ontario, pp. 1565–1576.
Levitan, M.L., Mehta, J., 1991b. Texas Tech field experiments for wind loads. Part 1. Meteorological
Instrumentation and terrain parameters. In: Proceedings of the 8th International Conference on Wind
Engineering, London, Ontario, pp. 1577–1588.
Marshall, R.D., 1975. A study of wind pressures on a single family dwelling in model and full scale. J. Ind.
Aerodyn. 1 (2), 177–199.
McDonald, R., 2004. Modelling response of mid-latitude storms to climate change. In: Proceedings of the UK
Wind Engineering Society Conference, Cranfield.
Meroney, R.N., 1999. Perspectives on air pollution aerodynamics. In: Proceedings of the 8th International
Conference on Wind Engineering, Copenhagen, pp. 79–90.
Meteorological Office, 2005. http://www.met-office.gov.uk/education/historic/1987.html/www.met-office.
gov.uk/education/historic/1987.htmlS.
Milford, R.V., Waldeck, J.L., Goliger, A.M., 1991a. Jan Smuts experiment—details of the full scale experiment.
In: Proceedings of the 8th International Conference on Wind Engineering, London, Ontario, pp. 1693–1704.
Milford, R.V., Goliger, A.M., Waldeck, J.L., 1991b. Jan Smuts experiment—comparison of full scale and wind
tunnel test results. In: Proceedings of the 8th International Conference on Wind Engineering, London,
Ontario, pp. 1705–1716.
Miller, C., 2003. A once in 50 year wind speed map for Europe derived from mean sea level pressure
measurements. J. Wind Eng. Ind. Aerodyn. 91, 1813–1826.
Munich Re, 2002. Winter Storms in Europe, Munich Re Publication.
Murakami, S., Mochida, A., 1999. Past, present and future of CWE. The view from 1999. In: Proceedings of the
10th International Conference on Wind Engineering, Copenhagen, pp. 91–104.
Pryor, F., 2003. Britain BC. Harper Collins, London.
Rathbun, J.C., 1940. Wind forces on a tall building. ASCE Trans. 105, 2056.
Richards, P.J., Hoxey, R.P., Short, J.L., 1999. Spectral models of the atmospheric surface layer. In: Proceedings
of the 10th International Conference on Wind Engineering, Copenhagen, pp. 315–321.
Richardson, G.M., Hoxey, R.P., Robertson, A.P., Short, J.L., 1995. The Silsoe structures building; the completed
experiment. Part 1. In: Proceedings of the 9th International Conference on Wind Engineering, Delhi,
pp. 1103–1114.
RMS, 2003. 1703 windstorm—a 300 year retrospective. Risk Management Solutions Report, 6pp.
Scruton, C., 1963. On the wind excited oscillation of stacks, towers and masts. In: International Conference on
Wind Effects on Buildings and Structures, National Physical Laboratory, London.
Shropshire Tourism, 2004. http://www.shropshiretourism.info/ironbridge/www.shropshiretourism.info/
ironbridgeS.
Solari, G., 2005. The International Association for Wind Engineering (IAWE): birth, development and
perspectives. Keynote paper. In: Proceedings of the 4th European and African Regional Conference on Wind
Engineering, Prague.
Standards Association of Australia, 1973. SAA loading code Part 2—Wind forces As 1170 Part 2, Sydney.
Surrey, D., 1999. Wind loads on low rise buildings; past, present and future. In: Proceedings of the 10th
International Conference on Wind Engineering, Copenhagen, pp. 105–116.
van der Hoven, I., 1957. Power spectrum of horizontal wind speed in the frequency range from 0.0007 to 900
cycles/hour. J. Meteorol. 14, 160–164.
Whitlock, D., 1961. The Anglo-Saxon Chronicle—a Revised Translation, Eyre and Spottiswoode.
Wind and Structures, 2005. In: Holmes, J. (Ed.), Special Edition on the IAWE Codification Initiative.
Wyatt, T.A., Walshe, D.E., 1992. Bridge aerodynamics 50 years after Tacoma Narrows—1. The Tacoma failure
and after. J. Wind Eng. Ind. Aerodyn. 40, 317–326.

Das könnte Ihnen auch gefallen