Anda di halaman 1dari 28

11th Americas Conference on Wind Engineering, San Juan, PR, USA

June 22-26, 2009

Wind Loading Codification in the Americas: Fundamentals for a Renewal


Emil Simiu1
1

NIST Fellow, Building and Fire Research Laboratory, National Institute of Standards and
Technology, Gaithersburg, Maryland, USA,
Distinguished Research Professor, International Hurricane Research Center and Department of
Civil and Environmental Engineering, Florida International University, Miami, Florida, USA
emil.simiu@nist.gov

ABSTRACT
A survey of current wind engineering practices and their reflection in codes and standards
reveals the existence of areas in which substantial progress needs to be made and is now
possible. These areas belong to micrometerology, wind climatology, aerodynamics laboratory
testing, structural dynamics, aeroelasticity, statistics, structural reliability, multi-hazard design,
wind effects estimation practice, and the standards development process from the points of view
of both form and substance. In particular, low-rise building aerodynamics need serious
attention, since both design and loss estimation are likely being seriously affected by the
incorporation in the ASCE 7 Standard of test results that can differ significantly in certain
instances by factors larger than two -- from actual values. It is pointed out that wind effects on
tall buildings must be estimated by correctly accounting for wind directionality, one of the main
factors that caused large differences between recent wind tunnel test results obtained by
independent wind engineering laboratories. Progress in terms of transparency and public
availability of test and wind effects estimation records is also necessary to allow clear
understanding by structural engineers, independent scrutiny, public accountability, and a
defensible basis for wind engineering laboratory certification.

1. INTRODUCTION
This talk is motivated by the belief that it is useful to take a fresh look at our field and its
evolution in the last half-century, and re-examine some of its assumptions, results, and
techniques. This is in our opinion justified by recent findings according to which the state of the
art in wind engineering, as reflected in codes, standards, and professional practice, has
shortcomings that need careful consideration and constructive debate. While not seeking to be
exhaustive, we will attempt in this talk to provide the motivation for research, development, and
actions that we believe are needed to markedly improve the basis for and the process of wind
loading codification. A vast and concerted research effort has achieved remarkable results in the
field of earthquake engineering. It is in our opinion high time for an effort of similar scope in our
field, with a view to designing or retrofitting structures that better conform to the physics of
wind-structure interaction and satisfy more rational structural performance demands in a more
economical manner than is now the case, -- an effort that will help to achieve a more resilient
built environment under wind loads.
We start with a brief historical survey. Denmarks Irminger is probably among the first
scientists or engineers to have produced experimental estimates of wind pressures on buildings.
In praise of his achievements an American structural engineer wrote: In 1894 J. O.V. Irminger,
manager of the Copenhagen Gas Works, made a number of experiments on wind pressure... The
conclusion he drew was that if his experiments on models represent the facts with regard to

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

buildings, the methods with which roofs are commonly calculated for wind pressure need
revision. An enthusiastic admirer of Irminger wrote: It will be due to him that we surely in the
future shall save tons of materials in our roofs [1].
However, as has been known for some time, Irmingers was in fact not the last word. A
large compendium of wind pressures on a wide variety of structures, produced decades later in
Switzerland on the basis of tests performed in uniform, smooth flow and reproduced in a 1961
American Society of Civil Engineers (ASCE) report [2], has long been viewed as obsolete. In the
early 1930s the great Goettingen aerodynamicist Ludwig Prandtl tasked Otto Flachsbart to
perform research on building aerodynamics. Following a series of papers on trusses and other
types of structure, in 1932 Flachsbart produced an internal report on the significant differences
between pressures on low-rise buildings in smooth flow on the one hand and shear flow on the
other [3]. The report was not followed up by a journal paper because Flachsbart, having refused
to divorce his Jewish wife -- a punishable offense under Nazi rules, -- was no longer allowed to
publish. So the sea change that Flachsbarts finding could have been expected to produce in wind
engineering did not come to pass. Excerpts from his report, re-discovered in the library of the
National Bureau of Standards, became widely known only after their 1986 publication in [4].
In the 1950s Jensen again in Denmark performed tests on a small model of a house
and reached, independently, conclusions similar to Flachsbarts. At about the same time Alan G.
Davenport, a Bristol University graduate student working under Professor Sir Alfred Pugsley,
applied the Liepmann [5] spectral approach to the calculation of turbulence-induced structural
response to the dynamic loading of line-like structures immersed in an atmospheric boundary
layer [6]. Davenport also used probabilistic and statistical procedures for estimating extreme
wind speeds developed by Gumbel. The wind flow was described micrometeorologically by
improving upon the Pagon [7] model of mean wind speed profiles, used in conjunction with
models of the atmospheric boundary layer depth and models proposed by Davenport for the
spectral and cross-spectral density of the longitudinal turbulent velocity fluctuations. This
beautiful synthesis started a new era in wind engineering, in which aerodynamic testing of
structures was typically no longer conducted in aeronautical wind tunnels, but rather in shear,
turbulent flows produced in so-called boundary layer wind tunnels. Most wind-tunnel testing in
those facilities was concerned with tall buildings and long-span bridges. However, from the late
1970s on boundary-layer wind tunnels were also used for the aerodynamic testing of low-rise
buildings, for which new wind loading provisions were introduced in the ASCE 7 Standard [8]
(hereinafter ASCE 7).
This very brief historical survey provides a framework for considering the question asked
at the beginning of this talk. Is it appropriate, half a century after Jensens and Davenports
seminal work, to proceed to a fundamental re-examination of current assumptions and
techniques? We will consider this question within the context of codification, with specific
reference to ASCE 7. We will stress the need to document transparently wind engineering
procedures and results, thus enabling effective scrutiny and accountability.
In the following sections we assess selected elements of the wind effects estimation
process that draw on the following inter-related areas: micrometeorology, wind climatology,
aerodynamics, structural dynamics, interfacing of directional wind climate and directional
aerodynamics/structural dynamics, aeroelasticity, statistics, structural reliability, multi-hazard
design, wind effects estimation, and standards development. We will do so in a manner that
emphasizes the relevance of these areas to specific provisions in the ASCE 7 Standard. We
believe our assessment and critique can stimulate fruitful discussions and help to put to good use
the experience and knowledge accumulated to date, while leading to a renewal of our field, that
is, to the development and/or adoption of improved technical approaches, and to fundamental

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

advances in codification in terms of both form and substance. Our concluding section will
summarize topics that, in our opinion, require consideration in the years to come, with a view to
providing the wind and structural engineering communities with standards and tools superior to
their existing counterparts.

2. MICROMETEOROLOGY
2.1 WIND SPEED PROFILES: LOGARITHMIC LAW VERSUS THE POWER LAW. VEERING ANGLE
Standards and codes, and wind tunnel practice, incorporate explicitly or implicitly models of the
variation of wind speed with height. Following Pagons work [6] such models were already used
in U.S. codes in the 1950s.
Unlike the Eurocode [9] and the Australian/New Zealand Standard [10], which use the
logarithmic law description of the wind profile, ASCE 7 uses the power law. One justification
for this choice was the belief that the logarithmic law, now universally used in
micrometeorology, is valid only up to a 50 m-100 m height above the surface. Fluid mechanics
theory showed that for strong winds the logarithmic law is in fact valid up to hundreds of meters
above the surface [11, 12]. Experimental evidence confirming this finding was reported in [13].
Nevertheless, that belief proved to have a long life, and a statement reflecting it was incorporated
into a draft U.S. standard as late as 2006. A second justification, which was claimed to be
relevant in the slide-rule era, was that power law calculations were easier to perform than
calculations based on the logarithmic. Finally, owing to the use of flow management devices
such as, e.g., spires, which are invariably present even in the most modern wind tunnels for civil
engineering applications, wind profiles achieved in the laboratory are not logarithmic. They can
be described, however, by the power law. If it is assumed that atmospheric flows are similar to
wind tunnel flows it follows that the power law holds not only in wind tunnels but in the
atmosphere as well. (It should be remembered, however, that it is not the atmosphere that
attempts to mimic the wind tunnel, but the other way around.)
While it can legitimately be argued that the power law is reasonably adequate for
practical design purposes, it is rarely employed by professional micrometeorologists. Advances
in micrometeorology that wind engineers commonly draw on (e.g., advances on the structure of
wind in regions close to surface roughness changes, or on the description of turbulence intensity
and spectra) are reported in terms of logarithmic law parameters, and their translation into power
law terms introduces unnecessary errors and inconsistencies. It is for this reason that the
logarithmic law has crept its way into parts of ASCE 7, probably the only standard that describes
wind profiles in terms of the power law for some applications and of the logarithmic law for
others. Eliminating this inconsistency from future, modernized versions of ASCE 7 would help
interactions between wind engineering and micrometerology, and render the standard cleaner and
more internally consistent. For a document that has been widely criticized for not being
sufficiently clear and user-friendly this may be a worthwhile goal.
Data on the variation of the veering angle of the mean speed with height above the
surface are scarce and preliminary, see [14, p. 15].
2.2 SUSTAINED WIND SPEEDS VERSUS 3-S PEAK GUSTS
ASCE 7 provisions are based on the use of 3-s peak wind gusts. This choice replaced fastest-mile
speeds, an awkward measure of wind speeds adopted in the first half of the twentieth century on
account of the widespread use in the U.S. of a type of measurement device that is now obsolete.
Movement toward international standards appears to justify the use of 10-min mean speeds, thus

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

conforming to practice sanctioned by the World Meteorological Organization, or of other


internationally acceptable measures of sustained wind speeds,
Perhaps equally importantly, the uncertainty in the estimation of the response to wind
increases unnecessarily and fairly significantly if calculations are based on peak gust speeds
instead of on sustained wind speeds [15]. This is due to the considerable variability of the peak
gust factor and is an additional reason for considering the use in future standards of sustained
wind speeds.
2.3 ATMOSPHERIC TURBULENCE: HURRICANE VS. LARGE-SCALE EXTRATROPICAL STORMS
Progress in the description of atmospheric turbulence was achieved in recent years mainly for
tropical storm winds. Analyses of recent measurements yielded the result that the frequency
content of fluctuating wind spectra, and the ratio between peak 3-s wind speeds and hourly wind
speeds, are greater in tropical storms than in large-scale extratropical storms [16, 17]. This result
is relevant for design and should be incorporated into wind and structural engineering practice
through appropriate standard provisions.
2.4 THUNDERSTORM WINDS
New standard provisions should accommodate the fact that the micrometerological structure of
thunderstorm winds, which dominate the wind climate in various U.S. regions, and the duration
of strong winds in thunderstorms, differ from their counterparts in hurricanes and synoptic
winds.

3. WIND CLIMATOLOGY
ASCE7-05 contains maps of basic wind speeds v. The basic wind speeds are defined by the
requirement that their product by the wind load factor 1.6 yields wind speeds with a nominal
mean recurrence interval (MRI) of about 720 years. For non-hurricane regions estimates of the
basic wind speeds are based on the Extreme Value Type I probabilistic model and correspond to
a nominal 50-yr MRI. For hurricane-prone regions the estimates of the basic wind speeds are
based on a methodology whose basic concept was first developed in 1971 [18]; their nominal
MRIs exceed 50 years and vary as a function of geographical position. The wind maps represent
wind speeds regardless of their direction.
In this section we briefly discuss issues related to the estimation of (a) ASCE 7 wind map
speeds for non-hurricane regions, and (b) hurricane wind speeds; propose the use of recently
developed tools for the improved estimation of wind speeds by using mixed-probability
modeling and newly available data; explain the rationale of and review a recently developed
procedure for augmenting wind speed data samples; and note the importance of the wind
directionality issue. The choice of the MRIs of the design wind speeds is dictated by implicit
structural reliability considerations and is discussed in a subsequent section.
3.1 NON-HURRICANE REGIONS: SUPERSTATION CONSTRUCTION AND MISESTIMATION OF
BASIC WIND SPEEDS
The ASCE 7 basic wind speed map was developed in 1998 [19] (for details see
www.nist.gov/wind, Section I, item 4), and is discussed in detail in by Simiu et al. [20, 21]. To
illustrate the approach used by these authors we show three maps that contain 23 stations from 5

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

states grouped into a large superstation (Fig. 1). The rationale for the grouping of individual
stations into superstations is that the sample size for each station is relatively small. By
creating a superstation the sample size increases, and the sampling errors correspondingly
decrease. For this rationale to be valid it is necessary that the individual stations be homogeneous
micrometeorologically and meteorologically, and should not be too close to each other so that
the respective data be reasonably independent of each other. The estimated 50-yr speeds for the
23 individual stations varied between 28 m/s (Devens, Mass. and Central Park, New York City),
and 55 m/s (Milton, Mass.). Nevertheless all stations were included in a 40 m/s 50-yr wind speed
zone. For numerous examples of similar discrepancies, see [20, 21]. This approach artificially
homogenizes the wind climate, that is, it hides differences between local wind climates, and led
to a division of all non-hurricane regions in the U.S. into two 50-yr wind speed zones: 40 m/s for
all states except California, Oregon, and Washington, and 35 m/s for those three states. The
eastern political boundaries of those states coincide with the boundary between the 40 m/s and
35 m/s zones in the ASCE 7 map.
The superstation approach has been applied in other disciplines, e.g., hydrology, and is
anticipated to have a useful place in the development of a U.S. future wind map. Solid scientific
methodologies have been developed for its use, but were not employed in the work leading to the
current ASCE 7 wind map.
3.2 WIND SPEEDS IN HURRICANE-PRONE REGIONS
For hurricane-prone regions the wind speed data sets were obtained from Monte Carlo
simulations applied to physical and probabilistic models of tropical cyclones and hurricanes
(e.g., [22] and www.nist.gov/wind, for the only public domain directional hurricane wind speed
data sets for the Gulf and Atlantic coasts); [23, 24, 25]. Except for [23], in which reverse Weibull
distributions were used, the estimation of wind speeds with various MRIs was performed by
these authors non-parametrically, using order statistics and estimated storm arrival rates (for
details see, e.g., [14, pp. 33-35].
It is of interest to compare estimates of hurricane wind speeds. For four stations chosen at
random the following speeds as estimated by various authors are listed in Table 1. Entries for
[25] were obtained through division of reported 3-s peak gusts by the 1.64 ratio of 3-s peak gust
to mean hourly speed approximately applicable to hurricane winds speeds [15].
Validation of 2000-yr estimates is not possible in practice, and for such estimates
modeling errors of about 15-20 % would not be surprising given the elusive nature of some
parameters (including, e.g., the Holland beta parameter, see [25]). Such errors are comparable to
the sampling errors inherent in the sample sizes available to date [26]. This suggests that seeking
additional precision in modeling of hurricane wind speeds may at some point yield diminishing
returns, since it may not lead to the significant reduction of the overall response uncertainties due
to the effect of all component uncertainties. A reliability-based approach to the development of
probabilistically-based design criteria would make possible rational decisions on where to stop a
quest for precision that, in the context of interest to structural engineers, may at some point
become irrelevant. Note that the comparisons of Table 1 are not complete, and that they do not
detract from the quality of the dedicated work performed to date in the hurricane wind speed
estimation field, especially in [24, 25].

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

Figure 1: Locations of New Jersey, New York, Connecticut, Massachusetts, and


Rhode Island stations (marked as yellow lozenges) belonging to superstation
99927 [20] (for continuation see next page).

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

Figure 2: Locations of New Jersey, New York, Connecticut, Massachusetts, and Rhode Island stations
(marked as yellow lozenges) belonging to superstation 99927 [20] (continued).
Table 1: Hurricane wind speeds at coastline, hourly mean speeds at 10 m above open terrain, m/s1
Milestone2

200

700

(S. of Brownsv., TX) (E. of N. Orl.)

1400

2300

(Miami)

(N. of C. Hatteras)

Batts et al. (1979)3

34; 37; 47

32; 35; 44

36; 39; 47

30; 33; 44

Whalen et al. (1996)3

35; 38; 46

33; 35; 41

36; 39; 46

31; 34; 43

Vickery, Twisdale (1995)3

30; 34; 45

34; 38; 50

35; 39; 51

29; 32; 45

Vickery et al. (2009)

30; 35; 49

33; 35; 49

35; 38; 52

25; 27; 38

Sets of three numbers separated by semi-colons are 50-yr; 100-yr; 2,000-yr speed, respectively.
For milestone location see locator map [34, p. 110] or www.nist.gov/wind, Sect. I.
3
Hourly mean speeds [34, p. 117].
2

3.3 MIXED WIND SPEED DISTRIBUTIONS FOR SERVICEABILITY DESIGN CRITERIA AND FOR
EXTREME WIND ESTIMATES
For hurricane-prone regions mixed distributions are needed for the estimation of wind speeds
with the relatively small MRIs relevant to serviceability design. For non-hurricane regions with

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

both thunderstorm and large-scale extratropical storm winds, mixed distributions are needed for
the improved estimation of extreme wind speeds [27-32].
In a climate with two types of wind, the probability that during the life of the structure
wind speeds are smaller than a specified value v is equal to the probability that wind speeds of
the first type, denoted by V1, are smaller than v, and winds speeds of the second type, denoted by
V2, are smaller than v, that is,
Prob(V1<v, V2<v) = Prob(V1<v) Prob(V2<v).

(1)

Mixed distributions of thunderstorm and large-scale extratropical storm winds were studied in
[33] for future application to the development of the ASCE 7-15 Standard wind speed map.
Software for extracting separate sets of data for the two types of storm from the National
Weather Service Automated Surface Observing System (ASOS) is available on
www.nist.gov/wind
3.4 AUGMENTATION OF WIND SPEEDS DATA SETS
For rigid structures building codes use the assumption that wind effects are proportional to the
square of the extreme wind speeds, regardless of their direction. This assumption allows the
estimation of wind effects with large MRIs from probabilistic models of the extreme wind
speeds. However, as will be shown later, for structures sensitive to wind directionality effects,
including structures with significant dynamic response, the estimation of wind effects with large
MRIs requires the use of time series of directional wind speeds with size exceeding the length of
the MRI of interest. For this reason it is necessary to construct synthetic directional wind speed
data sets that (1) have any specified sample size, and (2) are statistically consistent with the
available records. The construction of synthetic wind speed data sets involves two steps. First,
probabilistic models are developed for directional wind speeds, and these models are calibrated
to the data. Second, the calibrated models are used to generate synthetic directional wind speed
records. For details see [35].

4. AERODYNAMICS
Aerodynamics in civil engineering is concerned with bluff bodies, typically under high Reynolds
number conditions. It is based on the use of several tools: wind tunnel testing; full-scale testing
in natural winds; full- or large-scale testing in high winds induced by powerful fans in facilities
generically called Walls of Wind; and Computational Fluid Dynamics (CFD). We limit our
comments on CFD to noting that it is beginning to make useful inroads in structural engineering
practice, but cannot as yet be used as a substitute for design methods based on physical testing.
It is fitting at this point to pay tribute to University of Western Ontario (UWO)
researchers who, in papers by Surry et al. [36], Ho et al. [37], and St.Pierre et al. [38], have
addressed issues of aerodynamics testing and codification with uncompromising professionalism
and integrity. They have shown, as have Fritz et al. [39] and Coffman et al. [40] and, in their
pioneering efforts as well as on professional committees, T. Stathopoulos and G. Harris, that at
least for low-rise buildings a massive effort is still needed to achieve the goal of providing
structural engineers with realistic provisions on wind-induced aerodynamic effects. To develop
such provisions two conditions must be satisfied. First, the data underlying the provisions must
be adequate. Second, the provisions must not distort those data to an unacceptable extent.

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

4.1 AERODYNAMIC DATA UNDERLYING ASCE 7 PROVISIONS


Aerodynamic data on which ASCE 7 is based have significant shortcomings. They include data
obtained about three decades ago from tests in which the numbers of pressure taps were small -one to two orders of magnitude smaller than in recent UWO tests [37] in which, in addition,
records were taken for many more mean flow directions using far more advanced pressure
measurement technology. Although some calibration against full-scale measurements appears to
have been performed, no pertinent documentation is available for convenient public scrutiny.
ASCE 7 also contains provisions based on aerodynamic data of unknown source measured under
unknown testing conditions.
4.2 SPECIFICATION OF AERODYNAMIC DATA IN ASCE 7: PSEUDO-PRESSURES
The aerodynamic pressures specified by ASCE 7 (referred to therein as pseudo-pressures)
typically represent envelopes of actual pressures. This means that if the pseudo-pressures are
smaller than pressures measured in recent UWO tests, actual pressures underlying the ASCE
provisions may be unrealistically small. Investigations by [37, 38, 40] for a number of cases
concluded that the ASCE 7 pseudo-pressures can indeed be too small, in some cases by factors
of approximately two. It may be also be the case that for some cases ASCE 7 pressures are too
large. Furthermore, many types of low-rise buildings of interest to designers, builders, and
insurers are not covered by the ASCE 7 aerodynamic provisions.
4.3 DEVELOPMENT OF AERODYNAMIC DATABASES. DATA COMPRESSION
The development of pseudo-pressures from actual pressures was intended to ensure that
critical wind effects (e.g., bending moments at portal frame knees) are estimated conservatively.
However, as the aforementioned investigations have shown, this goal was not achieved, or at
least it was not achieved consistently. Inadequate standard specifications on wind pressures lead
not only to inadequate designs, and therefore to unnecessary losses in high winds and/or
unnecessary consumption of materials, but also to inadequate loss estimates, and therefore to
insurance rates that may not be commensurate with actual losses. A concerted effort may be
warranted aimed at replacing pressure coefficients specified in ASCE 7 with values based on
comprehensive tests yielding full records of aerodynamic data obtained at large numbers of
pressure taps on the exterior surface of the building models for as many as 36 wind directions, so
that effects of winds blowing from aerodynamically unfavorable directions are not missed. The
direct use of such data in design, within a framework referred to as database-assisted design [41],
is permitted by ASCE 7, subject to safeguards ensuring that the quality of the data is satisfactory
Hoerner [42] developed a large and widely used compendium of aerodynamic effects
which, for civil engineering applications, is now obsolete. A 21st century electronic successor to
Hoerners compendium is in our opinion necessary. In view of the important consequences such
a compendium could have on design, retrofitting, codification, and insurance rates, it may well
be in the interest of a broadly based group of stakeholders to support its development, much as
steel industry stakeholders support the development of the AISC manual. Absent such an effort
engineers will continue to base structural design of low-rise buildings on data accumulated by
successive accretions, of uncertain quality, and inadequately documented if documented at all.
Aerodynamic databases are large, and can be substantially compressed with no
significant lossiness (the term used by data compression specialists, see, e.g., [43]) by using a
variety of procedures. The use of compressed databases would be helpful in expanding the use of

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

aerodynamic databases. Standard procedures and provisions need to be developed for the
convenience of developers and users of compression techniques.
4.4 CALIBRATION AND CORRECTION OF WIND TUNNEL TEST RESULTS
As is suggested by Fig. 2, wind tunnel test results need to be calibrated and/or corrected against
results of credible full-scale tests, e.g., [44, 45, 46]. For this to be possible compendiums of fullscale data in user-friendly formats and correction methodologies should be developed. A
consensus on how this should be done has yet to be reached.

Figure 2: Pressures measured at building corner, eave level, Texas Tech University experimental building
[46].

4.5 COMPARISONS OF PEAK WIND EFFECTS MEASURED IN DIFFERENT TESTS. THE NEED FOR
PUBLIC RECORDS.
In the past comparisons between data measured in different tests were hampered by the fact that
peak pressures or other random wind effects are random variables, meaning that a peak
corresponding to, say, a 90 percentage point of its distribution would be compared to its
counterpart corresponding perhaps to a 5 percentage point. For comparisons to be effective
statistics of peak effects (e.g., means, 90 % quantiles) are needed. Such statistics can now be
estimated from single records using upcrossing theory. A user-friendly procedure for doing so
was developed in [47]. The requisite software, with improvements by J.A.Main, is available on
www.nist.gov/wind.
The ASCE 7 uses a peak factor incorporated in the specified aerodynamic pressure or
force coefficients, or used to multiply the root mean square value of the random process of
interest. Except for pressures at some hot spots, the peak factor specified or implicit in ASCE 7
typically has approximately the same value as the gust response factor specified in ASCE 7 for

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

the quasi-static along-wind response induced on a building by wind normal to a building face.
This approach is computationally convenient but when applied by ASCE 7 to, e.g., effects on
roofs, side walls, or leeward walls, it bears little or no relation to the respective actual peak wind
effects.
According to Griffis [48] there have been a large number of projects tested by more than
one wind tunnel laboratory where results were very close, typically within about 10 %. To our
knowledge the comparisons to which this statement refers are not publicly available. The public
availability of records of pressure, base moment, and base shear time histories obtained in
various wind tunnel laboratories for selected types of structure would allow independent public
scrutiny and comparisons, as well as serving as benchmarks for the oversight and certification of
wind tunnel facilities.
4.6 ATMOSPHERIC FLOW SIMULATIONS
The quality of the measured aerodynamic data depends to a large extent upon the laboratory
simulation of the atmospheric boundary layer flow. As was mentioned earlier, the simulation of
the wind speed profile is not a trivial matter, especially for low-rise buildings, because, even for
flow over homogeneous terrain, achieving a logarithmic profile over a sufficient height above the
floor is in practice not possible as long as spires are used as a flow management device, which is
the case for virtually all commercial wind tunnels. Even achieving specified power law profiles
adequate for low-rise building tests should not be taken for granted, as was indeed concluded in
[49], following careful analyses of various laboratory flows.
A recent draft standard on wind tunnels includes in its list of similarity criteria the socalled Jensen law. Jensen similarity states that replication in the wind tunnel of aerodynamic
effects on a full-scale structure requires that the ratio between the surface roughness length z0
and the structures height H be the same in the model and the prototype. Because z0 is typically
not meaningful for defining civil engineering wind tunnel flows, the Jensen similarity criterion is
in fact unusable. While the criterion is of interest from a historical point of view, it is in our
opinion not useful in a standard of practice for wind tunnels.
Modeling of the aerodynamic pressures also requires that the atmospheric turbulence be
reproduced in the laboratory. Fig. 3 shows that this is not uniformly achieved by various wind
tunnels. According to [49], differences between laboratory simulations of mean wind
profiles and of turbulence are responsible in large part for the discrepancies, by factors that can
exceed two, between wind-tunnel based estimates of wind effects on low-rise structures reported
for six wind tunnels in [39], five of which are denoted by block letters in Fig. 3.
4.7 HIGH-FREQUENCY AND LOW-FREQUENCY TURBULENT FLUCTUATIONS
High-frequency fluctuations tend to be weaker in the wind tunnel than in the atmosphere owing
to internal friction effects that are more pronounced at smaller scales. Their main effect is to
transport across separation layers fluid particles with high momentum from outside the layer into
the separation bubbles, thereby helping to overcome the effects of negative pressure gradients
and promote reattachment. It is likely that this small-scale turbulence deficit can in practice be
compensated to a degree sufficient for an adequate reproduction of prototype reattachment

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

Figure 3: Turbulence intensities in simulations by wind tunnels participating in the Fritz et al. (2008)
comparison [49].

patterns by artificially increasing the flows higher-frequency turbulent content. However, this
issue requires careful research.
Low-frequency fluctuations correspond in the turbulence spectrum to frequencies lower
than those approximately associated with Kolmogorovs inertial subrange. To the extent that the
larger-scale eddies fully envelop the structure, their aerodynamic effect tends to be
indistinguishable from the effect of the mean speed. That is, except for its smaller duration, the
aerodynamic effect of the peak gust speed in a flow with both high-and low-frequency
turbulence components may be assumed to a first approximation to be the same as the effect of a
flow with (a) mean speed equal to that peak gust, (b) comparable high-frequency turbulent
fluctuations, and (c) no or weak low-frequency fluctuations. For this to be acceptable, the ratios
between the integral turbulence scales and the corresponding characteristic dimensions of the
structure must be sufficiently large for the turbulent fluctuations to be essentially coherent over
spatial separation distances comparable to those dimensions.
4.8 RESEARCH INTO THE DEVELOPMENT OF A NEW LABORATORY TECHNIQUE FOR MEASURING
PRESSURES ON LOW-RISE BUILDINGS
The preceding observations suggest the possibility of developing flow management devices
resulting in flows with specified mean wind profiles, high-frequency turbulent components
mimicking approximately their prototype counterparts, and no or weak low-frequency content.
The peak gust speed achieved in the laboratory would be equal to the sum of the mean wind
speed and the relatively small peak fluctuating component associated with the weak turbulence
in the laboratory flow. The mean wind speed used in the tests would be such that this sum is
equal to the target peak gust speed in the atmospheric flow. This approach is likely to be fruitful
only for low-rise buildings, or portions thereof, that are relatively small in relation to integral

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

turbulence scales in atmospheric flows. Laboratory flows of the type suggested here would have
both low turbulence intensity and small integral turbulence scales. This would help to reduce
errors in aerodynamic output that are now common owing to the difficulty of faithfully
simulating atmospheric boundary layer flows close to the ground. Facilities generically called
Wall of Wind (Figure 4) are currently capable of producing flows with large specified frequency
content induced by controlled, quasiperiodically fluctuating rotational speeds of the fans [50].
This capability was used by G. Bitsuamlak, A. Gan Chowdhury, R. Li, and the author to test
successfully the hypothesis that an increase in the mean speed is an aerodynamically effective
substitute for the low-frequency fluctuations effect on a small building. A larger and more
modern Wall of Wind facility is currently under construction at Florida International University.

Figure 4. Wall of Wind, International Hurricane Research Center, Florida International University [50].

5. STRUCTURAL DYNAMICS: TIME-DOMAIN METHODS


The frequency domain approach to the calculation of dynamic structural response in turbulent
flow was first proposed in 1952 [5], when computers were too weak to solve routinely in the
time domain problems involving ordinary differential equations. The Fourier transformation of
Newtons equations of motion had the advantage of converting their solution into one involving
algebraic equations instead. Those computational constraints no longer exist, however, and timedomain calculations for wind engineering purposes are now easy to perform. They have over
their frequency domain counterparts the significant advantage of preserving phase information
and rendering unnecessary the awkward estimation of cross-spectral effects. Consider for
example a problem in which the wind effect of interest, denoted by M, is a sum of wind effects,
Mx and My, due to pressures acting in two orthogonal directions. What is the peak value of M? If

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

the phase information is lost the answer must be guessed, and combinations of the wind effects
must be specified accordingly. This would also be necessary when adding individual components
of demand-to-capacity ratios in structural interaction equations accounting for effects of axial
forces and bending moments. In contrast, in the time domain all the phase information is
preserved, so time histories of all the relevant effects can simply be added algebraically, and
standard methods for the estimation of peaks are applied to the resulting sums. A single
combination, which yields accurate results, is required, rather than large numbers of
combinations that yield results of questionable accuracy and unnecessarily complicate the
designers task.
A second advantage of making full use of current computational capabilities is the ability
to calculate internal forces by summing up records of pressure time histories measured
synchronously at large numbers of taps, weighted by appropriate influence coefficients. This
capability was not available until the 1990s. A third advantage is the ability to use time histories
of synchronously recorded pressures to calculate internal forces at any elevation of the structure
by using pressures measured in the laboratory. In contrast, if the High-Frequency Force Balance
(HFFB) is used for buildings (a) affected by aerodynamic interference effects and (b) whose
fundamental modes are not straight lines, the pressures must be guessed by the analyst. A fourth
advantage is the possibility of taking higher modes of vibration into account. Finally, wind
velocity and response directionality can be accounted for jointly in a transparent and physically
unimpeachable manner, as is briefly shown in the following section. For details and software
pertaining to time-domain calculations, see [51, 52].

6. DIRECTIONAL WIND CLIMATOLOGICAL AND AERODYNAMICS/STRUCTURAL


DYNAMICS EFFECTS
6.1 ESTIMATION OF MEAN RECURRENCE INTERVALS OF WIND EFFECTS
To clarify the role of wind directionality we consider directional wind speeds in three successive
wind storms for which we assume that winds blow from just two directions. Let the wind speed
time series vij (in m/s) consist of three storm events (i=1, 2, 3) with two wind directions (j=1, 2):
Event 1: 54 (dir. 1), 47 (dir. 2),
Event 2: 41 (dir. 1), 46 (dir. 2),
Event 3: 47 (dir. 1), 39 (dir. 2).
This set of data describes the wind speeds themselves as well as their dependence on direction.
The description of the wind speeds in the three storm events that considers only the maximum
wind speed in each event is the following:
Event 1: 54
Event 2: 46
Event 3: 47.
Clearly, this description is one that contains less information. This loss of information comes at a
price. To see why, consider the case where the aerodynamic coefficients Cp,j for the two
directions are
0.8 (dir. 1), 1.0 (dir. 2)
The corresponding nominal wind effects are assumed to be, to within a constant dimensional
factor, equal to the quantities Cp,jvij2. For the three events, the quantities [Cp,jvij2]1/2 are
equal to the squares of the quantities [Cp,jvij2]1/2. For the three events, these quantities are
Event 1: 48 (dir. 1), 47 (dir. 2),

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

Event 2: 37 (dir. 1), 46 (dir. 2),


Event 3: 42 (dir. 1), 39 (dir. 2),
(e.g., for event 1, dir. 1, [0.8 x 542]1/2=48).
Since for each storm event it is the largest nominal wind effect that matters, for design purposes
we extract from those quantities the following time series (in m/s):
Event 1: 48
Event 2: 46
Event 3: 42
Assuming further that the rate of occurrence of the storm events is 1/yr, it follows from the above
calculations that the largest and second largest of these quantities are, respectively, 48 and 46
m/s. The wind effects are proportional to the squares of these quantities, that is, to
2304 and 2116 (m/s)2
respectively. If the calculations were performed without accounting for the directionality of the
wind speeds, and the largest pressure coefficient (i.e., Cp=1) were used for all directions, then the
largest and the second largest of the quantities [maxj(Cp,j)maxj(vij)2] would be 542and 472, that is,
2916 and 2209 (m/s)2.
A 0.85 directionality reduction factor has been adopted in ASCE 7 to account summarily for
directionality effects on rigid buildings. However, estimates of extreme wind effects based on
that blanket factor, rather than accounting explicitly for wind directionality, can be in error either
on the conservative or unconservative side. Multiplication of the 2916 (m/s)2 and 2209 (m/s)2
wind effects by 0.85 would yield largest and second largest wind effects of
2480 and 1878 (m/s)2,
rather than 2304 and 2116 (m/s)2, as obtained by using a physically realistic model. Note that the
largest of the wind effects estimated by disregarding wind directionality and using the blanket
directionality reduction factor (i.e., 2480 (m/s)2) is conservative with respect to the largest
physics-based estimate of the wind effect (i.e., 2304 (m/s)2). However, if the second largest wind
effect were of interest, the difference between the result based on the use of the directionality
factor (1878 (m/s)2) and the physics-based result (2116 (m/s)2) would be -11 %, that is, in this
instance the directionality factor approach used in ASCE 7 yields an unconservative estimate).
It is commonly assumed that the MRI of the wind effect associated in ASCE 7 with nondirectional wind speeds is the same as the nominal MRI of those speeds. This assumption is in
general not correct. For example, in our illustration it would be assumed that the ranking of the
wind effect induced by the 47 m/s non-directional speed associated with storm event 3 is two.) In
fact, if directional effects are taken into account, the ranking of the wind effect induced by storm
3 is three.
The approach just described is applicable to any wind effects, including pressures,
dynamic response, and sums of demand-to-capacity ratios used in axial force-bending moments
interaction equations. The approach amounts to converting a multi-dimensional time series
consisting of the effects of m storms in p directions into a one-dimensional time series consisting
of the largest effects induced by the m storms, regardless of their direction. To use terminology
introduced in structural reliability, we do not operate in the space of wind speed variables, but
rather in the space of wind effect variables.
Figure 5 represents a wind speed and direction dependent response surface for the
demand-to-capacity index of a member of tall building analyzed in [52]. Once such a response
surface is developed, a one-dimensional time series of that index can be constructed from the
matrix of directional wind speeds simulated for a large number of storm events. The value of that
index for any MRI can be obtained immediately from that time series and the mean rate of
occurrence of the storm events.

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

Figure 5. Response surface representing a members demand-to-capacity index bij as a function of wind speed
and direction [52].

For critiques of alternative approaches to wind directionality, see [14, 53]. The critiques
addresses the questions of whether the probabilistic basis of such approaches is correct, whether
inferences on rare events can be made from records of non-extreme wind speeds, and whether
the approaches are sufficiently transparent for scrutiny by structural engineers. A critique by the
National Institute of Standards and Technology (NIST) of the method used by one laboratory
was judged in a Skidmore Owings and Merrill (SOM) report [54] to be fairly compelling; on
the upcrossing method used by another laboratory it was commented by SOM that the method
is computationally complex and verification is not possible because sufficient details of the
method used to estimate the return period of extreme events are not provided. This comment
suggests that more transparency in the presentation of wind tunnel laboratory reports is
necessary. That method that it tends to yield unconservative results; also, the low wind speeds it
uses can be meteorologically different from high wind speeds, and may not be suitable for
extrapolation to extreme wind speeds [14].

7. AEROELASTICITY
Aeroelastic wind tunnel modeling techniques for buildings are well documented, see e.g., [55].
However, dependable predictive tools usable for design or at least preliminary design purposes
have yet to be developed. Progress in this direction has been made, e.g. in [56], which can be
helpful for the qualitative estimation of vortex-induced effects on flexible buildings.
For structural reliability purposes documented estimates are needed of speeds at which
galloping or, in some special cases, flutter oscillations could occur. A structure may well
experience no significant aeroelastic effects under wind speeds with a, say, 720-yr MRI. Will

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

that structure be safe will the aerodynamic effects affecting it be acceptable -- under, say, a
2000-yr wind speed? This question is similar to the safety question that should have been asked
by the designers of the Katrina levee. Basic knowledge needed for developing the requisite
estimation tools is available, although the effect of modal shapes on the aeroelastic forces needs
additional study. For galloping oscillations the development of such tools would require static
aerodynamic testing. In special cases entailing the possibility of flutter, tests under oscillatory
conditions similar to those developed for bridges would be required.
Aerodynamic damping is due to the relative speed of the oncoming wind velocity
fluctuations with respect to the windward face of a building as the latter oscillates. This effect
has long been recognized, and as early as the 1960s Davenport proposed a simple test in an
attempt to develop aerodynamic damping estimates. However, estimates based on Davenports
attempt have not found acceptance in practice. The ingredients needed for a reliable estimation
of the aerodynamic damping are: (1) a relation between fluctuating wind speeds impinging on
the buildings windward face and the pressures they induce on that face, (2) experimental
information on the time series of the pressures on the buildings windward face, now readily
available thanks to the development of transducers for measuring pressures simultaneously at
multiple taps; and (3) a time domain model of the building motion due to the action of the
fluctuating speeds, also available (see Section Structural Dynamics ...). Approximate time
histories of the fluctuating speeds associated with the measured pressures can be obtained from
(1) and (2). The requisite estimates are obtained by a simple iterative numerical procedure in
which relative velocities of the wind speed with respect to the oscillating structure are used. This
procedure largely eliminates the need for aerodynamic damping estimates based on wind-tunnel
aeroelastic tests [57].

8. STATISTICS
Statistics plays an important part in the specification of wind effects, and is used for: (1) the
estimation of extreme wind speeds with specified MRIs (discussed in sections Wind
Climatology), (2) the estimation of extreme wind effects with specified MRIs (discussed in
section Directional Wind Climatological and Aerodynamics/Structural Dynamics Effects), (3)
the estimation of peaks of time series of wind effects (discussed in section Comparing Peak
Wind Effects...) (4) the assessment of uncertainties in the estimation of wind speeds and wind
effects.
8.1 ESTIMATION OF UNCERTAINTIES IN THE WIND EFFECTS
Uncertainties in the estimation of wind effects are calculated by accounting for uncertainties in
the atmospheric boundary layer flow characterization, the wind tunnel flow simulation,
aerodynamic pressure coefficient or force coefficient estimates, extreme wind speed estimates,
and peak wind effects estimates; for flexible buildings natural frequencies and modal shapes are
associated with additional uncertainties. ASCE 7 includes no explicit estimates of uncertainties
in the estimation of wind effects. However, these uncertainties are implicitly accounted for via
wind load factors specified in ASCE 7, Chapter 2. A discussion of wind load factors is presented
in the following section.

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

9. STRUCTURAL RELIABILITY
Structures are designed for wind effects that are sufficiently large to ensure that, under
reasonable assumptions regarding structural capacity, their probability of performing
inadequately under wind and other loads is acceptably small. Design criteria aimed at assuring
adequate structural reliability are specified in Chapter 2 of ASCE 7 in terms of wind loads that
induce effects associated with either allowable stress design (ASD) or Strength Design (SD). For
flexible buildings additional criteria pertain to inter-story drift and top floor accelerations (see,
e.g., [14, p. 169].
9.1 CURRENT ASCE 7 DESIGN CRITERIA
The purpose of the design criteria specified in ASCE 7, Chapter 2 is to assure that demand-tocapacity indexes (i.e., sums of demand-to-capacity ratios used in interaction equations) are
adequate, see [14, pp. 160-164]. However, especially for new types of structural system,
meeting the requirements implicit in those design criteria is a mere indication, not a guarantee,
that a structures behavior under very strong winds will be adequate, and that the material
consumption required to ensure adequate performance is not unnecessarily high. In some
engineers opinion the experience of the last decades, during which no spectacular wind-induced
structural collapses of engineering structures were recorded, and engineering judgment based on
that experience, support the belief that current design criteria are safe. However, it should be kept
in mind that before the devastation wrought by Hurricane Katrina it was assumed by some
engineers that the design criteria used for the New Orleans levees were also supported by
experience.
9.2 EXPLORATORY RESEARCH ON NONLINEAR BEHAVIOR AND INCIPIENT COLLAPSE
Given the approximations, simplifications, and uncertainties inherent in current design criteria
for wind, the most careful modeling possible of the relevant phenomena is warranted. Nonlinear
analyses accounting for post-elastic strength reserves under wind loads would be useful in this
regard, but in the present state of the art they can seldom be performed in practice, especially for
structures with significant dynamic and/or potential aeroelastic effects corresponding to longer
MRIs than those that typically induce linear effects. From this point of view structural design is
far less advanced for wind than for earthquakes. Nevertheless, exploratory work was performed
in recent years based on finite elements estimates of nonlinear structural behavior of rigid
structures under wind loads estimated from data measured simultaneously at large numbers of
taps [58, 59]. That work is a first step toward a structural reliability approach that accounts for
the behavior of the entire structure under wind loads up to incipient collapse, rather than just for
the behavior of individual members in the elastic range, as in ASCE 7.
9.3 IS THE ASCE 7 WIND LOAD FACTOR UNIVERSALLY VALID?
ASCE 7 is based on the assumption that a wind load factor of 1.6 applied to the effects
associated with the basic wind effect assures an adequate nominal reliability with respect to wind
loads for all structures. The ASCE 7 Commentary defines the wind load factor as the square of
the ratio between the wind speed, regardless of direction, with an approximately 720-yr MRI and
the basic wind speed specified in the ASCE wind map. This definition is consistent with the
specification of a wind load factor equal to 1.6.
In reality the wind load factor was originally defined as a function of uncertainties in the
estimation of wind effects. In accordance with this definition, the 1.6 value is typically

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

appropriate for rigid buildings in non-hurricane regions [60]. It is recognized in ASCE 7 that an
effective wind load factor larger than 1.6 is warranted for hurricane-prone regions. For those
regions the basic wind speed specified in ASCE 7 has a longer but unspecified MRI, and is
approximately equal to the ratio between the estimated wind speed with an approximately 720-yr
MRI and the square root of 1.6. The adoption of a basic speed with an MRI larger than 50 years
allows the use of the same numerical value of the load factor for both non-hurricane and
hurricane regions, even though the probability distributions of extreme wind speeds have longer
tails for hurricane-prone than for non-hurricane regions and are therefore characterized by larger
ratios of 720-yr to 50-yr wind speeds. While this is a step in the right direction, it does not
account for the fact that it is not just the aleatory variability of the wind speeds that is greater in
hurricanes than in typical non-hurricane winds. In addition, hurricane wind speed estimates are
affected by modeling errors that are typically greater for hurricane than for non-hurricane winds.
Uncertainties inherent in wind effects on structures with significant dynamic effects can
be larger than for rigid buildings. This is due to (a) the presence of uncertainties in the damping
and natural frequencies, which are not relevant for rigid structures, and (b) the stronger effect on
the response of uncertainties in the wind speeds, since wind effects on flexible buildings are
proportional to wind speeds raised to powers larger than two, rather than just two, as is the case
for rigid structures. For this reason wind load factors applied to flexible buildings should
typically be larger than 1.6 (i.e., larger than the square of the typical ratio between the nominal
720-yr winds and the basic wind speeds) even for non-hurricane regions. The use of the 1.6 wind
load factor can therefore lead to situations where the nominal safety level inherent in the
Strength Design approach would be lower for, say, a 500 m tall building than for an engineered
low-rise building.
It has been suggested that using a 720-yr MRI of the wind effect for Strength Design will
assure the same safety level for dynamically active structures and rigid structures. This
suggestion disregards the fact that the nominal 720-yr MRI used for the Strength Design of rigid
structures in non-hurricane regions was calculated by assuming that the wind load factor is 1.6.
Had a wind load factor commensurate with the uncertainties in the dynamic parameters been
estimated instead, the corresponding MRI of the wind effects for Strength Design would
typically have been larger than 720 years. Therefore, rather than relying on magic numbers
(1.6, or its 720 years counterpart), it would be useful for structural designers of tall buildings to
use a procedure aimed at assuring risk-consistency between members of tall structures on the one
hand and members of rigid structures on the other. The procedure should account for all the
uncertainties in the wind effects. Conceptually such a procedure is simple, based as it is on
Monte Carlo simulations that cover all the significant uncertainties in the relevant wind and
response parameters, and was in fact outlined in its simplest form in [60]. However, at the time
computational capabilities were insufficient for its implementation. This has changed, and an
adaptation of the procedure to newly developed tools for estimating dynamic response to wind
was developed in [51, 61]. For the procedure to be broadly applicable a consensus needs to be
reached on the appropriate definition of those uncertainties and on a method for ensuring that the
MRI of the wind effect considered in design, and the percentage point of the probability
distribution of the uncertainty in the wind effect, are consistent with their counterparts for rigid
structures in non-hurricane regions. In other words, Strength Design would be based on a wind
effect with an MRI N and a percentage point p of the uncertainty in the response, where, for the
sake of risk-consistency, N and p are the same as for those counterparts. It would be appropriate,
in our opinion, to assume, e.g., N=50 years, estimate on this basis the percentage point p inherent
in the wind load factor 1.6 used for rigid structures in non-hurricane regions, and use the same
values N and p for flexible structures and for structures in hurricane-prone regions or in other

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

regions where the ASCE 7 probabilistic assumptions concerning extreme wind speeds in nonhurricane regions do not apply.

10. MULTI-HAZARD DESIGN


10.1 MRIS OF STRUCTURES

SUBJECTED TO BOTH WIND AND SEISMIC LOADS

In accordance with the ASCE Standard 7-05, in regions subjected to wind and earthquakes,
structures are designed for loads induced by wind and, separately, by earthquakes, and the final
design is based on the more demanding of these two loading conditions. Implicit in this approach
is the belief that the Standard assures risks of exceedance of the specified limit states that are
essentially identical to the risks inherent in the provisions for regions where only wind or
earthquakes occur. We draw the attention of designers, code writers, and insurers to the fact that
this belief is, in general, unwarranted, and that ASCE 7 provisions are not risk-consistent, i.e., in
regions with significant wind and seismic hazards, risks of exceedance of limit states can be up
to twice as high as those for regions where one hazard dominates. This conclusion is valid even
if the limit states due to wind and earthquake are defined differently, as is the case in ASCE 7.
We propose an approach to modifying ASCE 7 provisions which guarantees that risks implicit in
minimum ASCE 7 requirements for regions where one hazard dominates are not exceeded for
structures in regions with strong wind and seismic hazards.
We now show that implicit in ASCE 7 provisions are risks of exceedance of limit states
due to two distinct hazards that can be greater by a factor of up to two than risks for structures
exposed to only one hazard. An intuitive illustration of this statement follows. Assume that a
motorcycle racer applies for insurance against personal injuries. The insurance company will
calculate an insurance premium commensurate with the risk that the racer will be hurt in a
motorcycle accident. Assume now that the motorcycle racer is also a high-wire artist. In this case
the insurance rate would increase as the risk of injury, within a specified period of time, in either
a motorcycle or a high-wire accident will be larger than the risk due to only one of those two
types of accident. This is true even though the nature of the injuries sustained in a motorcycle
accident and in a high-wire accident may differ. The argument is expressed formally as:

P( s1 s 2 ) P( s1 ) P( s 2 )

(2)

where P(s1) is the annual probability of event s1 (injury in motorcycle accident), P(s2) the annual
probability of event s2 (injury in high-wire accident), and P( s1 s 2 ) is the annual probability of
injury due to a motorcycle or a high-wire accident. (Note that s1 and s2 are, clearly, mutually
exclusive events.)
Equation 2 similarly holds for a structure for which P(s1) is the probability of the event s1
that the wind loads are larger than those required to attain a limit state associated with design for
wind, and P(s2) is the probability of the event s2 that the earthquake loads are larger than those
required to attain the same limit state associated with design for earthquakes. (Note that, as in the
earlier example, it is assumed that s1 and s2 cannot occur at the same time.) P( s1 s 2 ) is the
probability of the event that, in any one year, s1 or s2 occurs. It follows from Equation 2 that
P( s1 s 2 ) > P(s1), and P( s1 s 2 ) > P(s2), i.e., the risk that a limit state will be exceeded is
increased in a multi-hazard situation with respect to the case of only one significant hazard. If

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

P(s1) = P(s2) the increase is twofold [62]. Note that s1 and s2 can differ, as they typically do
under ASCE 7 design provisions. In spite of such differences, it is the case that, both for
earthquakes and wind, inelastic behavior is allowed to occur during the structures lifetime. For
seismic loading, only the MRI of the maximum considered earthquake is specified; the MRI of
the onset of post-elastic behavior is unknown. For wind loading, the MRI of the onset of
nonlinear behavior is specified; however, nonlinear behavior is also possible and allowed to
occur during the structures life. A useful way to make apples and apples comparisons is to
estimate the MRI of the event W that incipient collapse will occur under wind loads and the MRI
of the event E that it will occur under seismic loads. For example, if those MRIs are comparable,
the MRI of the event that either W or E will occur will be approximately half the MRI of the
event W, and approximately half the MRI of event E. Therefore, risks of failure inherent in the
load combinations specified in ASCE 7 can be up to twice as for some structures in regions
exposed to both strong winds and strong earthquakes than for their counterparts where only one
of these hazards is strong. That is, for the former, the minimum requirements with respect to the
requisite safety level are violated by current ASCE 7 provisions. One example of a structure that,
depending on geographical location, may belong to this category is the supporting structure of a
water tower depicted in Figure 6. Tall buildings supported by columns at the lower floors are
another example.

Figure 6: Schematic of water tower.

10.2 OPTIMIZATION AS A MULTI-HAZARD DESIGN TOOL


A multi-disciplinary foundation of multi-hazard design theory naturally includes appropriate
optimization approaches. A useful formulation of the multi-hazard design problem rests on the
fact that optimization under N hazards (N > 1) imposes mi (i=1, 2,.., N) sets of constraints, all of
which are applied simultaneously to the nonlinear programming problem (NLP) associated with
the design. For example, consider in a region exposed to both earthquakes and strong winds a set
of cantilevered columns fixed at the base and supporting at their top a heavy pipe filled with
water. Such a structural assembly is used in some solar energy installations. The wind load
acting on the pipe is largest when the wind direction is normal to the pipe. The seismic load is
independent of direction. What is the optimal column shape? If earthquake was the only hazard a
circular column would be optimal. If wind was the only hazard an elongated cross-sectional

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

shape with the long axis normal to the pipe would be appropriate. Given the two-hazard
constraint, a shape that would be intermediate between the two would be optimal. That shape can
be found effectively and elegantly through NLP. Note that in seeking the shape of the cross
section the stresses induced by loads parallel to the direction of the pipe a single-hazard approach
is appropriate, since wind effects in that direction are small. However, in the direction normal to
the pipe the joint effect of wind and seismic loads can control the design. While this is a toy
problem, a similar optimization approach, albeit more laborious, can be used for a wide variety
of multi-hazard structural problems, including the design of tall building for wind and
earthquakes [63].

11. WIND EFFECTS ESTIMATION


Wind effects are estimated in ASCE 7 through the integration of pertinent micrometerological,
wind climatological, aerodynamics, wind directional, structural reliability, and statistical
elements discussed earlier. The output of the estimation process consists: for ASD, of wind
effects corresponding nominally (though not physically) to the MRIs of the basic wind speeds
used in the calculations; for SD, of the wind effects corresponding nominally to a 720-yr MRI.
The aerodynamic data in ASCE 7 are adversely affected by the specification of drastically
reductive aerodynamic data sets, as well as by the quality of atmospheric boundary layer
simulations achieved in insufficiently documented or undocumented wind tunnel tests.
Directionality effects are accounted for, nominally rather than on physical grounds, by using a
blanket wind directionality reduction factor. The ASCE 7 analytical procedure contains a method
for estimating the along-wind dynamic response of tall buildings, which depends on the wind
loads and on the mechanical properties of the structure (fundamental modal shape and frequency,
and damping ratio in the fundamental mode). The ASCE 7 simplified and analytical procedures
specifically exclude aeroelastically active structures or wind effects due to vortex shedding.
The ASCE 7 provisions for the Wind Tunnel Procedure are not sufficient for
estimating wind effects. For this reason they are complemented by in-house methodologies
which do not always yield mutually consistent results. The best known instance is the
independent estimates by two laboratories of the response to wind of the World Trade Centers
twin towers, which differed by over 40 % [53]. Part of that difference was due to the respective
methods for dealing with the effects of wind directionality. We suggested at the beginning of this
paper that seismic R&D can provide a useful model for wind engineers. One feature of design
for seismic loads is that there is a clear division of work between seismologists and structural
designers. In our opinion a similar division of work is in order for wind engineers and structural
designers of tall buildings. For structures not susceptible to significant aeroelastic effects the
wind engineer needs to provide information on: (1) the wind climate (i.e., matrices of extreme
directional wind speed data for a large number of events, and the mean arrival rate of the events);
(2) the ratio between the wind speeds at the reference height of the appropriate open terrain
exposure meteorological site and the wind speeds at the location of the future building and at the
elevation of its top, (3) the time-histories of the simultaneously measured pressures at large
numbers of taps on the building exterior surface. For cladding design, the wind engineer also
needs to provide (4) recommended internal pressures. With this information, the structural
engineer can, with the help of software already available (see, e.g., www.nist.gov/wind) or
requiring minor adjustments, design members for wind action in accordance with specified
performance criteria on strength and serviceability. Research is also ongoing on optimization of
structural members and connections subjected to wind effects estimated by the realistic models
incorporated in newly developed software.

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

12. STANDARDS DEVELOPMENT


In the 1960s the Uniform Building Code devoted just a few pages to wind loading provisions.
The volume of wind loading provisions has since increased with regularity. ASCE 7-93
contained 15 pages of standard provisions, and a 12-page Appendix; ASCE 7-95, 22 pages, and a
20-page Commentary; ASCE 7-98, 48 pages and a 35-page Commentary; ASCE 7-05, 59 pages
and a 41-page Commentary, an almost fourfold increase in 12 years. ASCE 7-10 will again
increase. To help practitioners struggling with the increasingly complicated ASCE 7 provisions,
ASCE has published several guides to their use, including a 125-page guide to ASCE 7-02.
Nevertheless, standard provisions allowing the design of structures for wind loads by the wind
tunnel method currently do not exist, and in our opinion current efforts to produce a standard on
wind tunnel testing are insufficient in this respect.
Some engineers have asked whether it is desirable to have voluminous standards, or
rather to reduce their volume and provide references to approved research findings and
procedures. Some of the bulk of the standard is due to the decision to include numerous pages
that save the designer the trouble of performing a multiplication (i.e., where pressures are equal
to a constant times the square of the wind speed, large tables are included where the pressures are
listed for several speeds, instead of for just one speed), as well as numerous repetitions of exactly
the same material (e.g., footnotes to tables). In a future standard a more economical and elegant
formatting style, fully consistent with the users needs, may have to be sought.
The computer revolution can have a powerful impact on codification. Efficient design
methods based on large aerodynamic databases can serve the profession well. A useful step in
this direction has been made by introducing provisions in ASCE 7 that allow the use of such
databases, subject to prudent safeguards regarding the quality of the data.
An issue that may have to be considered by standards writers and users is whether
standards should be revised every few years, as has been the practice in recent decades. Because
the standards writing process is slow, periods of a few years are usually not sufficient for the
elaboration and careful scrutiny of new materials being submitted for incorporation in revised
versions of the standards. One example is the ASCE 7 wind map discussed earlier in this paper.
Another example are aerodynamics provisions whose questionable validity is disregarded. A
third example is the wind load factor, which according to the ASCE 7 Commentary should be
about 1.5, but somehow got transformed in the body of the Standard into 1.6. The problem is
equally acute for changes of format. Time is needed to ensure that they result in functional, userfriendly provisions.
It has been argued that the current system for standard development is good enough and
does not need improvement. This view may have to be reconsidered. The unhappiness of many
practitioners over what they perceive as the difficulty of using ASCE 7 provisions for wind loads
is due in part to the insufficiency of the current length of the revision cycle periods, and to
mechanisms and resources that, in practice, do not allow a thorough and effective participation
by the broad users community in the standards critique and development. In our opinion more
effective mechanisms and more resources for ensuring such participation may be warranted. In
considering the possibility of allocating more resources to enable a broadly based scrutiny and
critique of a draft standard it should be kept in mind that saving on such resources can usually
does -- yield a less than satisfactory product. This can result in social costs, both in terms of
users time and less than satisfactory designs, which can far outweigh the costs of testing draft
standards by designated, independent practitioners.

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

13. SUMMARY AND CONCLUSIONS


We have attempted to highlight a number of fundamental issues that we believe need to be
addressed to allow the development of a new, significantly improved new generation of
standards for wind loads. Topics discussed in the paper include: the representation of mean wind
speed profiles, the use in standard provisions of sustained wind speeds in lieu of 3-s peak gusts,
accounting for larger peak gust factors in tropical storm than in large-scale synoptic storm winds,
and the micrometerological description of thunderstorm winds; the use and misuse of
superstations in extreme wind climatology, the estimation of wind speeds in hurricane-prone
regions, the development of mixed-climate models in hurricane-prone regions and in regions
with both synoptic and thunderstorm winds, and the augmentation through modeling and
simulation of wind speed data sets; the quality of the aerodynamic data underlying the ASCE 7
Standard, the development and use of aerodynamic databases, data compression, the calibration
and correction of wind tunnel test results, the need for developing and maintaining transparent
and comprehensive public records of laboratory measurements, wind speed data, and wind
effects estimates, atmospheric flow simulation, and the possibility of developing a new paradigm
for testing of a class of low-rise buildings; replacing where appropriate frequency domain by
time domain methods, accounting correctly for directional wind climatology and
aerodynamics/structural dynamics effects, and estimating mean recurrence intervals (MRIs) of
wind effects as opposed to MRIs of wind speeds; the estimation for structural reliability purposes
of aeroelastic effects for winds with longer MRIs than those inducing linear structural response,
and the development of possible design aids for estimating aeroelastic effects, including
galloping, on typical tall building shapes; uncertainties in the estimation of wind effects; the need
to update current wind speed design criteria to assure risk-consistent performance-based design
and safety margins no less realistic than those employed for seismic design, and studies on
uncertainty-dependent wind load factors and nonlinear behavior to support such updating; multihazard design and its influence on design criteria for regions with wind and earthquake hazards;
and proposed improvements in the standards development process.
We begin our conclusions by quoting from the Skidmore Owings and Merril report
[53]: wind engineering is an emerging technology and there is no consensus on certain
aspects of current practice. Such aspects include the correlation of wind tunnel tests to full-scale
(building) behavior; methods and computational details of treating local (historical) wind data in
overall predictions of structural response; and types of suitable aeroelastic models for extremely
tall and slender structures. In our opinion the topics listed in this talk, among others, deserve to
be carefully debated and researched, and progress in the seismic design field, which has far
outpaced its wind design counterpart, should be vigorously emulated in the future.
Based on the material presented in the talk, it is our opinion necessary and possible to
produce standards that will be clearer, and more accurate, comprehensive, compact, and
responsive to advances in wind engineering than the current ASCE 7 Standard. It also necessary
to achieve significant improvements in the wind tunnel method that render it more transparent,
accurate, and attuned to current computational and measurement capabilities so that time-domain
approaches, multi-modal analyses, and large sets of aerodynamic data can be routinely used, as
well as more attentive to the structural engineers needs and role in estimating structural
response. Finally, it is essential to develop a culture wherein results being used for
standardization and design purposes are documented in a thorough, easily accessible, and userfriendly way. Wind engineering has a great future!

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

ACKNOWLEDGEMENTS
The author gratefully acknowledges valuable contributions by and/or fruitful interactions with G.
Bitsuamlak, C. Crosti, S. M. C. Diniz, D. Duthinh, J.J. Filliben, W.P. Fritz, M. Gioffr, A.
Grazini, R.D. Gabbai, A. Gan Chowdhury, M. D. Grigoriu, G. Harris, N.A. Heckert, E. Ho, M.
Kasperski, F. T. Lombardo, J. A. Main, F. Minciarelli, J.-P. Pinelli, A. Possolo, F.A. Potra, F.
Sadek, C. Schuyler, S. J. M. Spence, R.E. Vega, I. Venanzi, and T.A. Whalen.

REFERENCES
[1] R. Fleming, Wind Stresses, Engineering News, New York, 1915.
[2]Wind Forces on Structures, Transactions, American Society of Civil Engineers, 126, Part 2 (1961),
1124-1196.
[3] O. Flachsbart, Winddruck auf geschlossene und offene Gebaeude, Ergebnisse der Aerodynamischen
Versuchanstalt zu Goettingen, IV Lieferung, L. Prandtl and A. Betz, eds., Verlag von Oldenburg, Munich
and Berlin, 1932.
[4] E. Simiu, R. H. Scanlan, Wind Effects on Structures, 2nd Edition, Wiley, New York, 1986.
[5] H.W. Liepmann, On the Application of Statistical Concepts to the Buffeting Problem, J. Aeronaut.
Sciences 19 (1952) 793-800, 822.
[6] A.G. Davenport, The Application of Statistical Concepts to the Wind Loading of Structures, Proc.
Institution of Civil Engrs. 19 (1961) 449-472.
[7] W.W. Pagon, Wind Velocity in Relation to Height Above Ground, Eng. News Rec. 114 (19350 742745.
[8] ASCE Standard ASCE/SEI 7-05, Minimum Design Loads for Buildings and Other Structures,
American Society of Civil Engineers, 2006.
[9] Eurocode 1: Actions on structure - General Actions Part 1-4: Wind actions, European Committee
for Standardization, Brussels, 2004.
[10] Australian/New Zealand Standard, Structural design actions, Part 2: Wind Actions, Standards
Australia/Standards New Zealand, AS/NZS 1170.2:2002.
[11] G.T. Csanady, On the Resistance Law of a Turbulent Ekman Layer, J. Atmosph. Sciences 24 (1967)
467-471.
[12] E. Simiu, Logarithmic Profiles and Design Wind Speeds, J. Eng. Mech. Div., ASCE 99 (1973) 10731083.
[13] M.D. Powell, P.J. Vickery, T.A. Reinhold, Reduced drag coefficient for high wind speeds in tropical
cyclones, Nature 422 (2003) 279-283.
[14] E. Simiu and T. Miyata, Design of Buildings and Bridges for Wind, Wiley, New York, 2006.
[15] J.J. Filliben, E. Simiu, Tall Building Response Parameters: Sensitivity Study Based on Orthogonal
Factorial Experiment Design Technique, J. Struct. Eng. (submitted).

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

[16] Paulsen, B. M., and J. L. Schroeder, An Examination of Tropical and Extratropical Gust Factors
and the Associated Wind Speed Histograms, J. Appl. Met. 44 (2005) 270-280.
[17] B. Yu, A. Gan Chowdhury, Gust Factors and Turbulence Intensities for the Tropical Cyclone
Environment, J. Appl. Met. And Climatol. 48 (2009) 534-552.
[18] L. Russell, Probability Distributions for Hurricane Effects, J. Waterways, Harbors, and Coastal
Eng. Div, ASCE, 97 (1971) 139-154.
[19] J.A. Peterka, S. Shahid, S., Design Gust Wind Speeds in the United States, J. Struct. Eng., 124
(1998) 207-214.
[20] E. Simiu, R. Wilcox, R., F. Sadek, J.J. Filliben, Wind speeds in ASCE 7 Standard Peak-Gust Map:
Assessment, NIST Building Science Series 178, National Institute of Standards and Technology, 2001.
[21] E. Simiu, R. Wilcox, F. Sadek, J.J. Filliben, Wind speeds in ASCE 7 Standard Peak-Gust Map:
Assessment, J. Struct. Eng. 129 (2003) 427-439.
[22] Batts, M. E., Russell, L.R., and Simiu, E., Hurricane Wind Speeds in the United States, J. Struct.
Div., ASCE 100 (1980) 2001-2015.
[23] E. Simiu, N. Heckert, T. Whalen, Estimates of Hurricane Wind Speeds by Peaks over Threshold
Method, J. Struct. Eng. 124 (1998) 445-449.
[24] P.J. Vickery, L.A. Twisdale, L. A., Prediction of Hurricane Wind Speeds in the United States, J.
Struct. Eng. 121 (1995) 1691-1699.
[25] P.J. Vickery, P.J., D. Wadhera, D., L.A. Twisdale, I. Lavelle, I., U.S. Hurricane Wind Risk and
Uncertainty, J. Struct. Eng., 135 (2009) 310-330.
[26] S. Coles, E. Simiu, Estimating Uncertainty in the Extreme Wind Analysis of Data Generated by a
Hurricane Simulation, J. Eng. Mech. 129 (2003) 1288-1294.
[27] L. Gomes, B.J. Vickery, Extreme wind speeds in mixed climate, J. Ind. Aerodyn. 2 (1978) 331344.
[28] L.A. Twisdale, P.J. Vickery, Research on thunderstorm wind design parameters, J. Wind Eng. Ind.
Aerodyn. 4144 (1992) 545556.
[29] E.C.C. Choi, A. Tanurdjaja, Extreme wind studies in Singapore. An area with mixed weather system,
J. Wind Eng. Ind. Aerodyn. 90 (2000) 16111630.
[30] Cook, N.J., Harris, R.I., Whiting, R,,. Extreme wind speeds in mixed climates revisited. J. Wind Eng.
Ind. Aerodyn. 91 (2003), 403422.
[31] C. Letchford, M ,Ghosalka, Extreme wind speed climatology in the United States Mid-West. The
Sixth UK Conference on Wind Engineering, Cranfield University,UK, 2004.
[32] J.D. Holmes, Wind Loading of Structure, 2nd ed., Taylor and Francis, London, 2007.
[33] F.T. Lombardo, J.A. Main, E. Simiu, Automated extraction and classification of thunderstorm and
non-thunderstorm wind data for extreme-value analysis, J. Wind Eng. Ind. Aerodyn. (in press)
http://dx.doi.org/10.1016/j.jweia.2009.03.001
.

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

[34] E. Simiu, R. H. Scanlan, Wind Effects on Structures, 3rd Edition, Wiley, New York, 1996.
[35] M. Grigoriu, Algorithms for Generating Large Sets of Synthetic Directional Wind Speed Data for
Hurricane, Thunderstorm, and Synoptic Winds, NIST Technical Note 1636, National Institute of
Standards and Technology, Gaithersburg, MD (2009).
[36] D. Surry, T.C.E. Ho, G.A. Kopp, Measuring Pressures is Easy, Isnt It? Proceedings, International
Conf. on Wind Engineering, Texas Tech University, Lubbock, TX, 2 (2003) 2618-2623.
[37] T.C.E. Ho, D. Surry, D. Morrish, G.A. Kopp, The UWO contribution to the NIST aerodynamic
database for wind loads on low buildings: Part I. Archiving format and basic aerodynamic data, J. Wind
Eng. Ind. Aerodyn. 93 (2005) 1-30.
[38] L.M. St. Pierre, G.A. Kopp, D. Surry, T.C.E. Ho, The UWO contribution to the NIST aerodynamic
database for wind loads on low buildings: Part II. Comparison of data with wind load provisions, J.
Wind Eng. Ind. Aerodyn. 93 (2005) 31-59.
[39] W.P. Fritz, B. Bienkiewicz, B. Cui, O. Flamand, T.C.E. Ho, H. Kikitsu, C.W. Letchford, E.Simiu,
International Comparison of Wind Tunnel Estimates of Wind Effects on Low-Rise Buildings: Test-Related
Uncertainties, J. Struct. Eng., 134 (2008)1887-1890.
[40] B.F. Coffman, J.A. Main, D. Duthinh, E. Simiu, Wind effects on low-rise buildings: Databaseassisted design vs. ASCE 7-05 Standard estimates.J. Struct. Eng. (in press).
[41] J.A. Main, W.P. Fritz, Database-Assisted Design for Wind: Concepts, Software, and Examples for
Rigid and Flexible Buildings, Building Science Series 180, Chapter 3, Building Science Series 180, 2006.
[42] S.F. Hoerner, Fluid Dynamic Drag, published by the author, 1965.
[43] A. Possolo, M. Kasperski, E. Simiu, Tunalble Compression of Wind Tunnel Data, Wind and
Structures (submitted).
[44] P.J. Richards, R.P. Hoxey, J.L. Short, Wind Pressures on a 6 m Cube, J. Wind Eng. Ind. Aerodyn. 92
(2004) 1173-2001.

[45] D. A. Smith et al., http://www.wind.ttu.edu/Research/FullScale/ WERFL_4th. php).


[46] F. Long, D.A. Smith, H. Zhu, K. Gilliam, Uncertainties Associated with the Full-scale to Wind
Tunnel Pressure Coefficient Extrapolation, Report submitted to NIST under Department of Commerce
NIST/TTU Cooperative Agreement Award 70NANB3H5003, Texas Tech University, Lubbock, TX, Feb. 9,
2006.
[47] F. Sadek and E. Simiu, Peak Non-Gaussian Wind Effects for Database-Assisted Low-Rise Building
Design, J. Eng. Mech. 128 (2002) 530-539.
[48] L. Griffis, Wind Tunnel Testing Moving Forward, Structure Magazine, 7 (March 2006).
[49] B. Bienkiewicz, M. Endo, J.A. Main, Comparative Inter-laboratory Study of Wind Loading on Lowrise Industrial Buildings, ASCE/SEI Structural Congress, April 30-May 2, 2009, American Society of
Civil Engineers, Austin, Texas.
[50] P. Huang, A. Gan Chowdhury, G. Bitsuamlak, R. Liu, Development of devices and methods for
simulation of hurricane winds in a full-scale testing facility, Wind and Structures 12 (in press).

11th Americas Conference on Wind Engineering, San Juan, PR, USA


June 22-26, 2009

[51] E. Simiu, R.D. Gabbai, W.P. Fritz, Wind-induced tall building response: a time-domain approach,
Wind and Structures 11 (2008) 427-440.
[52] Spence, S.M.J., High-Rise Database-Assisted Design 1.1 (HR_DAD_1.1): Concepts, Software, and
Examples. NIST Building Science Series (in press), 2009.
[53] R.E. Vega [2008], Wind directionality: A reliability-based approach. Doctoral thesis, Texas Tech
University, Lubbock, TX.
[54] Federal Building and Fire Safety Investigation of the World Trade Center Disaster: Final Report of
the National Construction Safety Team on the Collapses of the World Trade Center Tower, NIST,
NCSTAR 1-2: Baseline Structural Performance and Aircraft Impact Damage Analysis of the World Trade
Center Towers, Appendix D, wtc.nist.gov, 2005.
[55] Y. Zhou, A. Kareem, Aeroelastic Balance, J. Eng. Mech. 129 (2003) 283-292.

[56] A. Kareem et al., http://aerodata.ce.nd.edu/pubs.html


[57] R. D. Gabbai and E. Simiu, Aerodynamic Damping in the Along-Wind Response of a Tall Building,
J. Struct. Eng. (submitted).
[58] S. Jang, L.W. Lu, F. Sadek, E. Simiu, Database-Assisted Design Wind Load Capacity

Estimates for Low-rise Steel Frames, J. Struct. Eng, 128 (2002)1594-1603.


[59] D. Duthinh, J. A. Main, A. P. Wright, and E. Simiu, Low-Rise Steel Structures under
Directional Winds: Mean Recurrence Interval of Failure, J. Struct. Engrg. 134 (2008) 1383-1387.
[60] B. Ellingwood, T.V. Galambos, J.G. MacGregor, A.C. Cornell, Development of a Probability Based
Load Criterion for American National Standard A 58, National Bureau of Standards, Washington, D.C.,
1980.
[61] Gabbai, R.D., Fritz, W.P., Wright, A. P., and Simiu, E., Assessment of ASCE 7 Standard Wind
Load Factors for Tall Building Response Estimates, J. Struct. Eng. 134 (2008) 843-845..
[62] D. Duthinh, E. Simiu, A Multi-hazard Approach to the Structural Safety of Buildings Exposed to
Strong Winds and Earthquakes, Proceedings, U.S.-Japan Cooperative Program in Natural Resources,
Panel on Wind and Earthquake Effects, 2006, N. Raufaste, ed., National Institute of Standards and
Technology, Gaithersburg, Maryland, 2008.
[63] F. Potra, E. Simiu, Optimization and Multi-Hazard Structural Design, J. Eng. Mech. (in press).

Anda mungkin juga menyukai