You are on page 1of 25

Journal of Wind Engineering

and Industrial Aerodynamics 82 (1999) 1}25

Review
Wind pressures acting on
low-rise buildings
Yasushi Uematsu *, Nicholas Isyumov
Department of Architecture and Building Science, Tohoku University, Sendai 980-8579, Japan
Boundary Layer Wind Tunnel Laboratory, The University of Western Ontario, London, Ont., Canada N6A 5B9
Received 22 January 1999; accepted 26 February 1999

Abstract

Data gathered from a number of "eld and laboratory experiments concerned with wind
pressures acting on low-rise buildings are reviewed, and selected experimental results are
presented in this paper. Particular attention is paid to works related to cladding design. Only
either full-scale studies or those done under conditions simulating the atmospheric boundary
layer have been considered. Comparisons of the data from various sources are made for the
characteristics of the mean and #uctuating wind pressures. The results indicate that the
statistical properties of #uctuating pressures on the roof edges and corners can be predicted by
a quasi-steady approach. Furthermore, the peak-factor approach is found to perform adequate-
ly in evaluating the design wind loads. The relation between the spatial and time averages is also
discussed.  1999 Elsevier Science Ltd. All rights reserved.

Keywords: Low-rise building; Local wind pressure; Cladding design; Review

1. Introduction

The majority of structures built all over the world can be categorized as low-rise
buildings used for residential, commercial and other purposes. These buildings are
generally susceptible to wind damage caused by typhoons, hurricanes, etc. For
example, Typhoon No. 19 in 1991 (Typhoon 9119) caused severe damage to low-rise
buildings over almost all of Japan. The insurance money that was paid for the loss
caused by this typhoon reached 567.5 billion yen, which was at that time the largest
amount in the world for loss caused by a natural disaster. According to a damage

* Corresponding author. Fax: #81-22-217-7875.


E-mail address: yu@venus.str.archi.tohoku.ac.jp (Y. Uematsu)

0167-6105/99/$ - see front matter  1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 9 9 ) 0 0 0 3 6 - 7
2 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

investigation [1], most wind damage to houses was restricted to the envelope of
buildings, in particular to the roof sheathing. The feature of the wind damage caused
by Hurricanes Hugo (1989) and Andrew (1992) in the United States was quite similar
to that of Typhoon 9119 (see Ref. [2]). These evidences indicate that an improvement
in wind resistance of the building's envelope will result in a signi"cant reduction in
overall economic losses. For this purpose, it may be quite important to understand
su$ciently the wind e!ects on low-rise buildings, and in particular, on roof
sheathing.
Following Jensen's wind tunnel experiment [3], a number of full-scale and wind
tunnel measurements have been made of the wind loads on low-rise buildings.
Stathopoulos [4], Holmes [5] and Krishna [6] presented reviews of the state of the
art, and Kasperski [7] published a review of wind load speci"cations for
low-rise industrial buildings in various wind load codes. Low-rise buildings are
constructed in di!erent types of terrain and topography with various plan forms.
Therefore, despite a number of studies made in the past, there are many problems
remaining unsolved. Most of the previous experiments were made under such
conditions that an isolated building with a gable or #at roof was placed in a uniform
terrain.
Recently, signi"cant improvements in experimental techniques have been made.
These new techniques enable us to understand the structure of the pressure "eld
on buildings in more detail. For example, the authors investigated the temporal
and spatial characteristics of the pressures near the roof and wall edges of a
low-rise building, based on simultaneous pressure measurements in a wind tunnel
(see Ref. [8]).
We carried out a literature survey on the data concerning the wind loads on
low-rise buildings; focus is on the #uctuating and peak pressures acting on the leading
edge and corner regions, from a viewpoint of cladding design. This paper presents
a review of the existing state-of-the-information on the subject, as born out by the
research papers appearing in the journals and the conferences on Wind Engineering
for the last three decades. Comparisons of the data from various sources, including the
code speci"cations, are made for the characteristics of the mean and #uctuating wind
pressures. Application of the quasi-steady and the peak-factor approach to the
evaluation of the design wind loads is discussed. It should be mentioned that this
paper is a revised version of our previous paper published in the Journal of Wind
Engineering, Japan Association for Wind Engineering (JAWE) [9].

2. Survey of the previous papers

In this paper, &low-rise buildings' refer to buildings with B'2H and H(30 m,
with B and H being the width and the reference height (i.e. the mean roof height),
respectively. A number of measurements have been made of the wind loads on
low-rise buildings for the last three decades. The authors collected and arranged more
than 200 research papers on this subject and constructed a database. The main
sources are (i) Journal of Wind Engineering and Industrial Aerodynamics, (ii)
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 3

Proceedings of the International Conference on Wind Engineering, and (iii) Journal


of Wind Engineering, JAWE. The papers are divided into the following four
categories:
(i) Full-scale measurements of wind pressures and/or their load e!ects in natural
winds;
(ii) Laboratory measurements of wind pressures under conditions simulating the
atmospheric boundary layers;
(iii) Discussion of the wind tunnel methodologies suitable for the experiments on
low-rise buildings; and
(iv) Codi"cation of wind pressure coe$cients.
Compared with high-rise buildings, many full-scale measurements have been car-
ried out on low-rise buildings. This may be due to the fact that full-scale measure-
ments can be made more easily for low-rise buildings than for high-rise buildings. In
other words, the determination of the wind loads from wind tunnel experiments is
more di$cult for low-rise buildings. One of the problems encountered in wind
tunnel testing of low-rise buildings is the determination of geometric scale. The length
scale of wind tunnel #ows generated by using naturally grown boundary layers ranges
from 1/200 to 1/500 in most cases. If building models were made with this length scale,
they would be as small as a matchbox. This results in instrumentation problems and
makes it impossible to model the architectural details such as eaves and parapets,
which may play an important role in the wind loading function. Low Reynolds
numbers, resulting from small model size, may cause a distortion of the #ow and the
resulting variations in pressure distributions. Several methodologies have been ap-
plied to the solution of the question of the scale [3,4]. Another problem encountered
in the wind tunnel testing is the appropriate modeling of the atmospheric surface
layer, in particular in its lower reaches wherein low-rise buildings are placed. The
characteristics of the atmospheric surface layer are rather complicated.
Extensive research has been done using the following full-scale experimental build-
ings, which revealed the characteristics of the atmospheric surface layer as well as of
the wind pressures on low-rise buildings in natural winds. The results were also used
for discussing the appropriate wind tunnel methodologies.

2.1. Aylesbury experimental building [10,11]

This experimental building with a gable roof was constructed in Aylesbury, Eng-
land, early in the 1970s. The plan form, represented by the width (B);length (¸), is
7 m;13.3 m, and the eaves height h is 5 m. A unique feature of this building is that the
roof pitch b can be quickly varied to any angle ranging from 53 to 453. Comparative
wind tunnel experiments (IAWE Aylesbury comparative experiment) were conducted
at 17 laboratories worldwide using the identical 1 : 100 model of this building [12].
A comparison between full-scale and wind tunnel measurements indicates that the
traditionally used similarity parameter h/z (building height/roughness length, known

as the &Jensen number') is not su$cient to ensure similarity when signi"cant isolated
local roughness, such as trees and hedges, are present [13}16]. Furthermore, it was
found that the lab-to-lab variation in pressure coe$cients was attributed to
4 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

di!erences in the method of data acquisition and in the measuring point of the
reference static and dynamic pressures [12,16}19].

2.2. Silsoe structures building [20]

This experimental building, with B;¸"12.93 m;24.13 m, h"4.14 m and


b"103 (gable roof), was constructed in Silsoe, England, in the late 1980s. The
structure consists of seven cold-formed steel portal frames at 4 m centers. Two
di!erent eaves cladding details, i.e. a curved eaves detail and a conventional sharp
eaves detail, were tested. Extensive measurements have been made of the wind
pressures at many locations on the envelope as well as of the strains in some structural
members. The full-scale measurements provided a set of benchmark data for verifying
wind tunnel experiments and numerical simulations [20}30]. Application of a quasi-
steady approach to the evaluation of the design wind loads has been investigated
[31}37]; a discussion of this subject will be given in Section 4.1.

2.3. TTU Building [38}40]

A unique experimental building with B;¸;h"9.1 m;13.7 m;4.0 m was con-


structed at Texas Tech University in the late 1980s; the roof of this building is almost
#at. The building can be rotated, permitting positive control over wind angle of
attack. Meteorological conditions are monitored by the instruments mounted on
a 46 m tower located at the site. A signi"cant amount of data on the winds and
building surface pressures has been collected in various acquisition modes. The
recorded data, which are open to all interested researchers, have been used for various
analyses [41}47]. Wind tunnel simulations have been carried out at many laboratories
all over the world [48}55]; comparisons of the data with the full-scale measurements
contributed to a signi"cant improvement in the wind tunnel simulation technique.

Most of the previous studies focus on buildings with gable or #at roofs; approxim-
ately "fty and forty percent of the previous papers we collected are related to gable
and #at roofs, respectively. This is due to the fact that there are a number of studies
regarding the above-mentioned experimental buildings. Therefore, in this paper the
main focus is on gable roofs; #at roof is regarded as a special con"guration of gable
roofs. Furthermore, most experiments were carried out in idealized situations within
homogeneous upstream terrain simulations with the aim toward a better understand-
ing of the fundamental aerodynamic behavior of low-rise buildings. The wind load
speci"cations in codes and standards are generally based on the results of such
experiments. Only a few papers discuss the e!ect of nearby surrounding buildings on
the wind loads. For example, Ho et al. [56}58] investigated the wind loads on
#at-roofed low buildings in realistic environments, using an &experimental Monte
Carlo' simulation. The results indicate that the wind loads on the low-rise buildings in
such environments may be considerably di!erent from those predicted from isolated
building tests and that the coe$cient of variation of the larger loads turns out to be
approximately 0.6}0.7. Based on these results, they claim that the wind load
speci"cations should be determined based on a reliability approach considering such
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 5

a variation in wind loads. At present, however, the data for a variety of surrounding
environments have not been compiled, so we cannot discuss the e!ects quantitatively.
Therefore, we deal with only the isolated building case in this paper.

3. Mean pressures
The codes of practice are generally based on a quasi-steady approach, in which the
wind load speci"cation is determined from the distribution of the mean pressure
coe$cient. The minimum pressure, or the worst suction, occurs in an area near eaves,
verge or ridge, depending on the roof pitch b and the wind direction h. Fig. 1 shows the
notation and the de"nition of the roof regions used in this paper. In the "gure, l'
represents the smaller value of 2H and B; the de"nition is the same as that used in the
AIJ Recommendations for Loads on Buildings [59]. Considering the symmetry of the
buildings, we pay attention to the wind direction h ranging from 03 to 903.
The #ow pattern around a building depends strongly on b and h. Accordingly, the
mean pressure distribution changes signi"cantly with these parameters. Fig. 2 shows
the roof region in which the worst suction occurs for given values of b and h; the data
plotted in the "gure are obtained from the previous publications available [60}68].
When h+03 and b+03, the #ow separates at the leading edge, which causes high
suctions near the leading edge (region &A'). With an increase in b, the magnitude of the
suction gradually decreases. Beyond a critical value of b(+183), the #ow separation
point shifts from the leading edge to the ridge, which causes high suctions near the
ridge (region &E'). In oblique winds (h+203!703), conical vortices are generated
along the leading edges when the value of b is small, such that b(183, for example. In
this case, rather high suctions are induced by these vortices near the windward corner
(region &B'). With an increase in b, the strength of the vortices decreases. When b'22.53,
the worst suction occurs in the region &E' for h(453, while in the region &C' for h'453. In
such a case that b+203 and h+153}453, the region &D' is subjected to the worst suction.
For larger values of h, such as h'703, for example, the worst suction occurs
near the windward verge (region &C') regardless of the roof pitch. This is due to the fact
that high suctions are caused by the #ow separation at the windward verge.
The data on the mean pressure coe$cients for various b and h values have
been obtained for the regions &A', &B' and &E' in the previous studies

Fig. 1. Building dimensioning notation and de"nition of regions.


6 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

Fig. 2. Roof region of the maximum time-averaged suction, as a function of roof pitch b and wind direction h.

[20,21,26,35,41,42,48,50,63,67}71]. On the other hand, the amount of data for the


other regions is quite limited. Fig. 3 shows the minimum mean pressure coe$cients,
irrespective of wind direction, observed in the regions &A', &B' and &E' as a function of b.
The solid lines in the "gure represent the speci"ed values in the AIJ Recommendations
for Loads on Buildings. In the region &E', the speci"ed value for b(103 changes from
!0.8 to !1.0, according to the B/H ratio. Hence, the two extremes are shown in Fig.
3(c). The experimental data are generally very scattered, in particular, in the regions
&A' and &B' for lower roof pitches. In these regions there are many cases where the
experimental values signi"cantly exceed the AIJ speci"cations. When b'203, on the
other hand, the experimental data collapse into a narrow range and agree relatively
well with the AIJ speci"cations.

4. Fluctuating pressures

4.1. Application of quasi-steady approach

4.1.1. Review
The majority of low-rise buildings can be categorized as small-scale structures, such
as residential houses and warehouses. For such buildings, the rigidity of the structure
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 7

Fig. 3. Variation of the minimum pressure coe$cients irrespective of wind direction in the regions A, B and
E with roof pitch b. (a) Region A; (b) region B; (c) region E.

is generally so high that the resonant movement of the building can be neglected.
Furthermore, the characteristic dimension of the building is fairly small compared
with the turbulence integral scale of natural winds. These facts imply that the
maximum load e!ects on the structure and its components can be evaluated from
a quasi-steady approach. In practice, the design wind loads on low-rise buildings are
determined using a modi"ed quasi-steady approach in many codes of practice. The
true quasi-steady approach attributes all #uctuations in pressure to #uctuations in
velocity. In such circumstances, the cross-correlation between velocity and pressure is
unity and the peak pressure or load occurs when the velocity is maximum. Therefore,
the design wind load can be given by the mean pressure coe$cient multiplied by the
peak dynamic pressure q( of the #ow.
This approach fails where there is a signi"cant interference to the #ow by the
building, such as in separated #ow regions where pressure #uctuations are in#uenced
8 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

by building-generated or wake turbulence. The length scale of the building-generated


vortices is much smaller than that of the approach #ow and, consequently, high-
frequency pressure #uctuations are induced in the separated #ow regions. The high-
frequency components may have signi"cant in#uence on small building elements,
such as roof tiles. However, for members with large tributary areas the high-frequency
components will be "ltered out due to a spatial averaging process. This implies that
the quasi-steady approach can be applied to the evaluation of the design wind loads
for such members.
Kawai [72], Letchford et al. [43] and Hoxey et al. [33}37] investigated the
applicability of the quasi-steady approach. The following is the outline of this
approach, based on their results.
In the quasi-steady approach, the pressures on a building respond to atmospheric
turbulence directly, as if they were changes in mean wind speed and direction. Thus,
pressure variations correspond exactly to wind #uctuations. The instantaneous pres-
sure, p(t), may be given by the following equation:
p(t)"q(t) ) C (h(t))"o<(t) ) C (h(t)), (1)
N  N
where o"air density; <(t) and q(t) stand for the instantaneous magnitudes of wind
speed and velocity pressure, respectively; and C (h(t)) is the mean pressure coe$cient
N
for the instantaneous wind direction h(t). Assuming that the peak pressure p( occurs in
accordance with the peak wind speed <K and that the change in wind direction is small,
the peak pressure coe$cient CK , de"ned in terms of the mean velocity pressure q , may
N
be given by
CK +GC (hM ), (2)
N T N
where G represents a gust factor; and hM the mean wind direction. This is the simplest
T
expression for the peak pressure coe$cient derived from the quasi-steady approach.
Since the vertical component w of wind speed #uctuation does not have signi"cant
in#uence on the pressure #uctuations on low-rise buildings, it is generally neglected
for the purpose of simplicity. The wind speed <(t) may be given by the following
equation:
<(t)"(;#u)#v, (3)
where ;"mean wind speed; and u and v represent the longitudinal and lateral
components of wind speed #uctuation, respectively. The lateral component v of wind
speed #uctuation may be replaced by the change h in the wind direction in the
horizontal plane as follows:

 
v
h"tan\ . (4)
;#u
In the "rst-order or linearized quasi-steady vector model, the pressure coe$cient is
approximated by a Taylor polynomial of the "rst order as


dC
C (h(t))+C (hM )#(h(t)!hM ) N . (5)
N N dh
FFM
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 9

Assuming that the #uctuating wind direction is linear for small v and substituting
Eq. (4) into Eq. (5), we obtain


v dC
C (h(t))+C (hM )# N . (6)
N N ; dh
FFM
Substituting Eqs. (3) and (6) into Eq. (1), we obtain the following equation for the
instantaneous pressure coe$cient C (t) after some manipulation:
N

  
p(t) 2u v dC
C (t)" + 1# C (hM )# N . (7)
N 1/2o; ; N ; dh
FFM
One further simpli"cation made is that u and v are assumed independent of each
other and, consequently, the RMS pressure coe$cient C is given by the following
N
equation:

   
p  p dC  
C + 2 S C (hM ) # T N
N ; N ; dh
FFM

  
dC  
" (2I C (hM ))# I N , (8)
S N T dh
FF M

where p and p are the RMS values of u and v, respectively; and I and I represent the
S T S T
longitudinal and lateral turbulence intensities, respectively. When the magnitude of
v is much smaller than that of u, or when the variation of C with h is small, the RMS
N
pressure coe$cient reduces to
C +2I C (hM ). (9)
N S N
The instantaneous pressure coe$cient given by Eq. (7) is a linear combination of the
u and v components. If u and v are normally distributed and independent of each
other, the probability density of C (t) would also be normally distributed. Further-
N
more, the spectrum of C (t) can be given by
N
C(hM ) (dC /dh" M )
S N( f )"4 N S ( f )# N FF S ( f ), (10)
! ; S ; T
where S ( f ) and S ( f ) are the power spectra of u and v, respectively.
S T
4.1.2. Discussion of the application
Stathopoulos et al. [73] investigated the applicability of Eq. (9) to the prediction of
RMS pressure coe$cients, based on a wind tunnel experiment using a 1 : 12 roof-
sloped model (B;¸"24.4;30.5 m, h"4.9 m, in full scale; referred to as &UWO
1 : 12 Model', hereafter). The predicted results were compared with the experimental
ones for the roof pressures. A good agreement was found for the windward edge
regions, immediately adjacent to the #ow separation, when the wind direction was
perpendicular or parallel to the ridge. On the other hand, the agreement was poor for
points in the wake and for oblique winds. Okada and Ha [50], who tested a TTU
Building model, obtained a similar conclusion. Kind [62,74] investigated the
10 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

applicability of Eq. (2) for predicting the peak pressure coe$cients, based on a wind
tunnel experiment using several models of #at-roofed buildings. Their results indicate
that the worst peak suctions, which occur near the windward corner for quartering
wind directions (h+453), can be estimated by Eq. (2) with an assumption that
G "1#3I , where I represents the longitudinal turbulence intensity of the
T S& S&
approach #ow at the roof height. On the other hand, Milford and Waldeck [75]
showed that the gust e!ect factor for the pressures on the roof corner in oblique winds
was much larger than the value of G, based on a full-scale measurement on a large
T
hangar (100 m;150 m;30 m) at Jan Smuts Airport, South Africa. Letchford et al.
[43] investigated the application of the quasi-steady approach to full-scale measure-
ments on the TTU Building. Sample results are shown in Fig. 4. In the case where
conical vortices are generated along the leading edges in oblique winds, the quasi-
steady approach (Eq. (8)) underpredicts the RMS pressures beneath the vortices
despite the high cross-correlation of the pressures.
It has widely been accepted that the discrepancy between the quasi-steady
approach and experiments for C and CK is attributed to the existence of building-
N N
generated turbulence. Recently, Tieleman [76] has shown that the small-scale turbu-
lence of the incident #ow signi"cantly controls the behavior of the separated shear
layer producing the extreme suctions. He suggests that the quasi-steady approach, in
which the e!ect of the small-scale turbulence is not considered, does not apply in the
regions of separated #ow and reattachment. On the other hand, Hoxey et al. [33}37]
claim that the approximation made in the linear quasi-steady model is the main cause
of the discrepancy between the quasi-steady approach and experiments, based on
a detailed full-scale measurement on the Silsoe structures building. They have made
an improvement on the quasi-steady model, in which a higher order of the Taylor
series representation of C (h) is taken into account. According to their results, the
N
RMS pressure coe$cient is given by the following equations:

 
(p!p ) p p
(C )" "[C (hM )] O #pN 1# O , (11)
N q  N q  ! q 

pN"[C (h)!C (hM )], (12)


! N N
where q and p represent the mean and RMS values of the velocity pressure,
O
respectively; and the wind direction h is assumed to be normally distributed. They
compared the results obtained from their model with those of the full-scale measure-
ments on the Silsoe structures building and showed a good agreement between the
two results even for the edge and corner regions. However, the agreement was still
poor for pressures at tapping points on the downstream side where turbulence was
generated in the boundary layer shear #ows on the building surface. At these points,
the magnitudes of the RMS and peak pressure coe$cients are generally small. Hence,
the disagreement is not a serious problem, from the viewpoint of practical design.
Based on these results, they concluded that the quasi-steady pressure coe$cient,
derived from the mean pressure coe$cient and modi"ed for turbulence averaging
e!ects, is the most appropriate for use in design.
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 11

Fig. 4. Comparison between "eld measurement and quasi-steady approach for the TTU Building (quoted
from Ref. [43]). (a) Tapping distribution and angle of attack de"nition; (b) measured and predicted RMS
pressure coe$cients (h"2283); (c) measured and predicted PDF for tapping 50101 (h"2283).

Several researchers have investigated the probability density function of pressure


#uctuations. Some measurements [77,78] have shown that the probability density
functions of wind pressures acting on points or small areas depart signi"cantly from
Gaussian, particularly at their tails in the windward wall region as well as in the
12 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

separated-#ow regions on the side walls and roof. Stathopoulos [77] reported the
non-Gaussian characteristics in velocity #uctuations at heights very close to ground
level; this may be one possible reason for the non-Gaussian characteristics in pressure
#uctuations acting on the windward walls. On the other hand, Kawai [72] and
Holmes [78] showed that non-linear relationships between pressure and velocity
result in probability distributions which depart signi"cantly from Gaussian, even
when the velocity #uctuations have a Gaussian distribution. They obtained a better
agreement between experiment and prediction by taking account of the non-linear
relationships. Letchford et al. [43] compared the results from the modi"ed probability
density function model with those of a full-scale measurement on the TTU Building.
The predicted results were generally in good agreement with the experimental ones.
However, the agreement was still poor for roof corners, where the pressures were quite
sensitive to the variation in wind direction; a sample result is shown in Fig. 4(c). Based
on the wind tunnel data for the &UWO 1 : 12 Model', Stathopoulos [77] showed that
the probability distributions for large positive or negative pressures were adequately
represented by a Weibull model as follows:

  
g I
F(C )"exp ! , (13)
N C

where g"normalized pressure, expressed in the peak factor form. The values of the
Weibull parameters, C and k, were determined from the best "tting to the experi-
mental data for "g"'2.5. In the positive pressure regions, the values were C"0.85
and k"1.10 irrespective of the exposure and location. In the negative pressure
regions, on the other hand, the values of C and k ranged from 0.50 to 0.65 and from
0.95 to 1.20, respectively; they depended on the exposure and location.

4.2. Peak pressures

4.2.1. Evaluation of peak pressures


Commonly used approaches for the assessment of local peak pressures are the
quasi-steady approach, the peak-factor approach and the simpli"ed Cook}Mayne
approach. The peak pressure coe$cient CK is given by Eq. (2) in the quasi-steady
N
approach, while by the following equation in the peak-factor approach:

CK "C $gC , (14)


N N N
where g represents a peak factor. The choice of adding or subtracting the gust
component g ) C is made according to the sign of the mean pressure coe$cient C . In
N N
contrast to these two approaches, the method proposed by Cook and Mayne [79,80]
combines the statistics of the dynamic pressure of extreme wind speeds with those of
peak pressure coe$cients. In the simpli"ed Cook}Mayne approach, the design
pressure coe$cient CH is given by
N
CH"; N#1.4/a N, (15)
N ! !
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 13

Table 1
Calibration of the quasi-steady and peak factor approaches to the assessment of wind loads against the
method of Cook and Mayne for the &UWO 1 : 12 Model' (made by the authors, based on the results of
Ref. [81])

Method Class A Class B Class C


(1 s averaging) (4 s averaging) (16 s averaging)

Quasi-steady approach CH/C 2.56 2.16 1.75


N N
G 2.88 2.38 1.94
T
R 0.951 0.966 0.981

e 0.34 0.24 0.14

Peak factor approach g 5.42 4.74 4.31
p 1.13 0.83 0.70
E
a 0.981 0.981 0.957
R 0.989 0.993 0.996

e 0.16 0.10 0.06


where ; N and 1/a N are the mode and dispersion of the Fisher}Tippet type I extreme
! !
value distribution for the largest peak pressure coe$cients; these values are usually
determined from a wind tunnel experiment.
Cook [81] made the calibration of the quasi-steady and peak-factor approaches to
the assessment of wind loads for equivalent static design against the simpli"ed
Cook}Mayne approach, using model-scale data for a range of typical building shapes.
One of their models was the &UWO 1 : 12 Model'. Table 1 summarizes the results for
this model. Design values derived from the 1-s, 4-s and 16-s averaged extremes are
classi"ed as Class A, B and C design loads, respectively. First, the values of CH at
N
a number of locations on this model were plotted against C for all wind directions at
N
153 increments, and a regression analysis was made for the plotted data. R and e in
 
the table represent the regression correlation coe$cient and the standard deviation of
the error (i.e. the di!erence between the regression line and the experimental value),
respectively. If the quasi-steady assumption held exactly true, the slope of the regres-
sion lines CH/C would be equal to the square of the gust factor, i.e. G. According to
N N T
the results shown in the table, the values of CH/C are generally smaller than those of
N N
G. Although the regression correlation coe$cient remains very high, 0.95}0.98, there
T
is a large scatter about the regression line as expressed by the standard deviations, in
particular, for shorter averaging periods. Then, the values of the peak factor g were
computed by equating Eqs. (14) and (15). In Table 1, g represents the equivalent peak
factor averaged for all locations and wind directions tested, and p is the standard
E
deviation of the variation about the average g . A regression analysis of CH against the
N
peak pressure coe$cients CK derived from the peak-factor approach was made; a,
N
R and e in Table 1 represent the slope of the regression line, the regression
 
correlation coe$cient and the standard deviation of the error, respectively. The value
of R is almost unity, which indicates that the peak pressure coe$cient CK is equal to
 N
the design value from the simpli"ed Cook}Mayne approach on average. Further-
more, the scatter of the data about the regression line is much less than that of the
corresponding quasi-steady calibrations, and the standard deviation of the estimates
14 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

of CH from CK are more than halved. These evidences indicate that the peak-
N N
factor approach performs adequately once the peak factor has been determined
appropriately.

4.2.2. Area-averaged pressures


The results for point pressures are of limited use for estimating the response of
cladding elements or structural components such as rafters. The peak pressure
coe$cient, which should be averaged over the area of the element for the design
purpose, may depend on the element's size. Usually, for cladding design, this e!ect is
considered by low-pass "ltering of the local pressure #uctuations. Since the area-
averaged pressure is regarded as an accumulation of point pressures, it would be more
normally distributed than the individual pressures, as a consequence of the central
limit theorem. Many experimental evidences support this inference. For example,
based on a full-scale measurement on a hangar at the Jan Smuts Airport, Milford and
Waldeck [75] reported that the spatially averaged pressure over an area of approxim-
ately 5 m;5 m at the roof corner was almost Gaussian and the ratio of the observed
peak to the mean pressure coe$cient was approximately equal to the square of the
gust factor, G; the averaging period for wind speed measurements is estimated to be
T
2}3 s, judging from the response of the anemometer they used. Similar results were
obtained by Letchford et al. [43] from a full-scale pressure measurement on the TTU
Building. They showed good agreement between the measured and the predicted
values by the quasi-steady approach, with respect to the other statistical properties,
such as the probability density function.
The e!ect of spatial average on the peak pressure may be related to that of time
average, for example, by using the &TVL formula' proposed by Lawson [82]; a dis-
cussion of this subject will be given later. In Table 1, it is seen that the values of e and

e become smaller, or in other words, the agreement between the results from three

di!erent approaches becomes better, as the averaging period is increased. This feature
indicates that the peak values of the spatially averaged pressure or the resultant
member stresses can be evaluated from the quasi-steady approach. Robertson [32]
investigated the gust loading on the Silsoe structures building, based on a full-scale
measurement of the free-stream dynamic pressure q and the strain responses e of
several structural members. They computed the ratios of peak-to-mean values for
both q and e for a range of digital "ltering (0.5 to 16 s) and then obtained the
integration period which needed to be applied to q in order to predict a given
time-averaged peak strain. The integration period, or e!ective gust duration for q was
found to consistently exceed the strain duration period, because the strain records
were consistently less peaky than the corresponding q record due to the "ltering e!ect.
They obtained some interesting results as follows:
(1) Half-second peak strains for the primary members were virtually identical to the
1 s values, indicating that the gust factors for 1 s peak strains e!ectively embrace all
signi"cant peak loadings generated in the structural frameworks.
(2) The gust duration for 1 s peak strains (or stresses) is 5 to 6 s for the primary
members and 2 to 3 s for the secondary members; the most appropriate value depends
on the location of the member.
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 15

4.2.3. Peak factor


In our previous study [8], we investigated the peak factors g for pressures acting on
low-rise buildings with a 4 : 12 gable roof (b"183). The values of g at many locations
were plotted against the minimum pressure coe$cient C (the negative peak
N 
pressure coe$cient). The results indicated that the scatter of the data was rather large
for small "C " values, while the data collapsed into a relatively narrow range with
N 
an increase in "C "; the representative values of g for the high suction regions on
N 
roofs and walls were 6 and 7, respectively. In the present study, we have made a similar
examination, using the experimental data presented in Refs. [10,75]. Fig. 5 shows the
result for the roof and wall pressures, in which the values of g at many locations are
plotted against the mean pressure coe$cient C ; the data cover a range of b from 53 to
N
22.53. The behavior of g with increasing "C " is quite similar to that observed in our
N
previous study [8]. Furthermore, full-scale and wind tunnel measurements on the
TTU Building for h"03 [41,50] showed that the peak factors for negative pressures
on the roof ranged from 4 to 7. This result is also consistent with the above mentioned
"ndings.
Fig. 6 shows the e!ect of moving average on the peak factor. In the "gure, the ratio
of the peak factor g(¹ ) for some averaging period ¹ to that for no moving average
(¹ "0 s) is plotted against ¹ , based on our previous wind tunnel experiments [8]
and a full-scale measurement on the TTU Building by Levitan et al. [41]. Since the
value of g for ¹ "0 s is not given in Ref. [41], we estimated it from the value for
¹ "0.1 s, using the empirical formula we provided in Ref. [8]. In both cases, the
g(¹ )/g(0) ratio decreases with an increase in ¹ . The ratio of decrease in g(¹ )/g(0) is
somewhat greater for the TTU Building than for our model.

Fig. 5. Variation of peak factor with mean pressure coe$cient. (a) Roof; (b) wall.
16 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

Fig. 6. E!ect of averaging period on the peak factor.

4.2.4. Ewects of spatial and time averages on the peak pressures


Wind pressure distribution on building surfaces changes signi"cantly with time.
According to our previous wind tunnel experiment [8], the pressure distribution,
which causes very severe suctions in the leading edge and corner regions, lasts only 0.2
to 0.3 s and the area of severe suctions is as small as 1 m;2 m. Therefore, it is
expected that the time and/or spatial averages of pressure #uctuations would lead to
a signi"cant reduction in the e!ective peak suction. We investigated the relationship
between the e!ects of spatial and time averages on the peak pressures in the leading
edge and corner regions. The results revealed that the averaging period ¹ for time
average and the area A for spatial average are related to each other through the
following equation:

¹;
(A" & ("lH ), (16)
k 2
N
where ; is the mean wind velocity at the mean roof height H; k the decay constant
& N
for the root coherence of pressure #uctuations acting on two di!erent points in the
area under consideration; and lH ("¹ ; /k ) represents an equivalent length for ¹ .
2 & N
Fig. 7 summarizes the results for the e!ects of time and spatial averages on the peak
pressures in the leading edge and corner regions. The open circles show the ratio of the
minimum pressure coe$cient C (¹ ) for some averaging period ¹ to that for no
N 
moving average as a function of lH (see Ref. [8]). The scatter bars represent the ratio of
2
the minimum area-averaged pressure coe$cient C (0) to the minimum pressure
NJ 
coe$cient C (0) in the area, as a function of (A [8]. The results for the time and
N 
spatial averages correspond well to each other, although the range of (A is limited.
Also shown in the "gure are the experimental results published by Maruta et al. [83]
and the speci"cation in the National Building Code of Canada [84] regarding the
e!ect of the spatial average for the roof corner regions. Both the results of Maruta et
al. and the NBCC speci"cations are consistent with our experimental results.
Various cladding elements of di!erent size and shape are used for low-rise build-
ings. The design wind loads for cladding should be determined by taking account of
the tributary area of the element. The gust duration for evaluating the peak pressures
in the quasi-steady approach should be determined in the same manner. As mentioned
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 17

Fig. 7. E!ects of time and spatial averages on the minimum pressure coe$cient.

above, the information on the root coherence of pressure #uctuations is necessary for
determining the appropriate gust duration for the cladding element. The decay
constant k seems to be dependent on the building shape as well as on the location of
N
the element. Unfortunately, few studies have been made on this subject to date.
Therefore, it is hoped that many experiments will be conducted in order to compile the
data on k for various building shapes.
N

5. E4ects of building geometry on local pressures

Many post-disaster studies have suggested that hip roofs survive better than gable
roofs in severe windstorms. Considering the di!erences in geometric form between hip
and gable roofs, it is not surprising that the aerodynamics di!er signi"cantly. The
better wind resistance of hip roofs is often attributed to lower wind loads on hip roofs
than on gable roofs. However, as pointed out by Meecham et al. [65], the wind
resistance of roofs should be discussed based on the relationship between the pressure
distribution and the underlying structural framing. Fig. 8 shows typical structural
framing of hip and gable roofs used in North America (see Ref. [65]). A gable roof
generally consists of common trusses of triangular form, yielding a constant cross
section throughout the building length. Prefabricated timber gable trusses are usually
located at nominal centers ranging from 0.6 to 1.2 m. The hip roof is framed
di!erently. The central part has the same cross section as the gable roof. However, to
achieve the hip ends, the structural framing usually consists of diagonal hip trusses
spanning from the building corner to a girder truss located at or near the hip/ridge
intersection. The hip trusses support shorter jack trusses. Shown in Fig. 9 are the
mean pressure coe$cient distributions on the 4 : 12 gable and hip roofs of a rectangu-
lar building (B;¸"10 m;20 m, h"3 m) for a wind direction of h"903 [65]. The
solid lines in the "gure represent a typical structural framing pattern. There exists
18 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

Fig. 8. Structural framing of hip, gable roofs generally used in North America (quoted from Ref. [65]).
(a) Gable roof; (b) hip roof.

a marked di!erence in the pressure distribution between these two roofs. However,
according to Meecham et al. [65], the overall lift and overturning loads were almost
the same. This "nding implies that roof shape alone is not responsible for the disparity
in roof performance under extreme winds. The spatial distribution of the pressures
relative to the structural framing should be taken into account. Meecham et al. [65]
found that the spatial organization of the high negative pressures on the full span
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 19

Fig. 9. Mean pressure coe$cient contours superimposed on structural framing (quoted from Ref. [65]).
(a) Gable roof; (b) hip roof.

gable roof trusses was pronounced, whereas the "rst full span hip roof common truss
experiences disorganized pressures of smaller magnitude. As a result, the e!ective
pressures on the full span trusses of the gable roof can be roughly twice those of the
full span hip roof trusses at the same wind speed. These di!erences appear to explain
the improved survival of hip roofs compared with gable roofs.
Several researchers have investigated the wind pressures on roofs of various
con"gurations. For example, Blessmann [85] measured the mean wind pressures on
roofs with a negative pitch (trough roofs) in a smooth uniform and a turbulent shear
#ow. The in#uences of parapet height and wind turbulence on the pressure distribu-
tion were investigated. Stathopoulos and Mohammadian [86] measured the local and
area-averaged pressures on buildings with mono-sloped roofs in a simulated atmo-
spheric boundary layer over an open country terrain, and discussed the e!ects of
height, width and roof pitch of the building on the magnitude of pressures. They found
that both mean and peak pressures were higher than those for buildings with gable
roofs, in particular, on the roof corners. Furthermore, Stathopoulos and Saatho!
[87}89] measured the local and area-averaged pressures on buildings with multi-span
gabled roofs and sawtooth roofs, and investigated the in#uences of roof pitch and
number of spans on the local pressure coe$cients. Comparing the results with the
speci"ed values in the Australian Standard for h"03 [90], they found that peak
suctions and positive pressures were generally larger than those of the speci"cations.
Several architectural features, such as eaves, parapets and edge pro"les, may a!ect
the wind #ow in the building proximity and the resultant wind pressures on the
building. Blackmore [91] experimentally investigated the e!ects of chamfered roof
edges on the local and area-averaged pressures on #at roofs. Their results indicate that
a 303-chamfer can reduce the area-averaged loads on a corner panel by up to 70% and
20 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

Fig. 10. Distributions of the mean pressure coe$cients for the square and curved eaves (quoted from
Ref. [31]).

gives a reduction in overall load of over 30%. The large load reduction on the corner
panel is due to a narrower separation bubble and consequently a much narrower
highly loaded edge region, resulting from the suppression of vortices generated at the
windward corner. It is expected that a curved eaves detail results in a similar e!ect.
Robertson [22] and Hoxey [31] made comparisons of the e!ects of a curved eaves
detail and a conventional sharp line eaves detail on wind pressures over the building,
based on full-scale measurements on the Silsoe structures building. Fig. 10 shows the
distributions of the mean pressure coe$cients over the mid-length line of the building
for the two eaves arrangements when h"03 [31]. The curved eaves reduce sub-
stantially the suctions over the lower third of the windward roof slope; this is due
to the fact that the #ow does not separate at the windward edge of the roof in the
case of the curved eaves detail [22]. However, the suction near the ridge is higher than
that of the sharp eaves case. Thus, the overall lift force acting on the roof is not
reduced so much.
In the case of #at roofs, parapets are also very e!ective in reducing the magnitude of
local suctions near the leading edges and corners. This feature is related to the fact
that parapets lift up the separated shear layer in normal winds and the vortex cones in
oblique winds. The e!ects of parapets on the pressure distribution depend on various
parameters, such as the building geometry and the relative parapet height. According
to the wind tunnel experiments conducted by Kind [74], the worst suction observed
near the windward corner in an oblique wind (h+453) decreases monotonically with
increasing parapet height. However, since parapets also cause a broadening of the
suction peaks, they can increase area-averaged loads on certain areas. Baskaran and
Stathopoulos [92] investigated the e!ects of parapet con"gurations on the suctions
on #at roof corners, based on an idea that the generation of conical vortices would be
suppressed by modifying the parapet con"gurations. Their results indicate that
parapet con"gurations with cuts or slots near the roof corners may signi"cantly
alleviate the high local suctions on roof corners and edges. Several investigators have
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 21

reported similar reductions in the local peak suctions using a variety of aerodynamic
devices. For example, Cochran et al. [93] explored the use of porous screens across
the roof corners on the TTU Building. The impact of various rigid gauge screens on
the formation of conical vortices was investigated by using ten screen shapes. Their
results indicate that each shape reduces the high uplift loads to varying degrees. Surry
and Lin [94] tested the use of rounded edges, rectangular parapets, sawtooth parapets
and roof-top cylinders on a 1 : 50 model of the TTU Building. All roof geometric
modi"cations signi"cantly reduced the high roof suctions near the windward corner
and edges, compared with the normal con"guration results. Among them, porous
parapets with a porosity of 50% provided the most signi"cant e!ects, leading to
a reduction of the high suctions of up to 70% near the corner and a very #at pressure
distribution pattern over the roof.

6. Concluding remarks

This paper reviews the data for the wind pressures acting on low-rise buildings,
derived from various sources, including our previous papers and code speci"cations.
Particular attention is paid to works related to cladding design. Many experimental
evidences indicate that the peak-factor approach performs adequately in evaluating
the design wind loads for cladding. Furthermore, the RMS pressure coe$cient, which
is used in the peak-factor approach, can be evaluated by the quasi-steady approach.
As mentioned above, a number of research e!orts have been conducted for the
determination and codi"cation of wind loads on low-rise buildings. Several elaborate
full-scale measurements of wind pressures etc. in natural winds have been conducted.
These measurements have revealed the characteristics of the wind loads on low-rise
buildings as well as of the atmospheric surface layers. Furthermore, they provided the
bench mark results for discussing the appropriate wind tunnel experiments and
numerical simulations. Despite such e!orts over three decades, the information
available is not comprehensive enough to cover the variables involved, and the code
provisions still need updating. It is hoped that much more data with a high accuracy
will be compiled under various conditions. Fortunately, the quasi-steady approach
can be used for evaluating the design wind loads in many cases. Hence, it would be
e!ective to accumulate the information on mean pressure coe$cients, considering
that the experiments are relatively easy compared with the #uctuating and peak
pressures.

Acknowledgements

A part of this study was made when the "rst author stayed at BLWTL, The
University of Western Ontario, as a research investigator supported by the Ministry
of Education, Science, Sports and Culture, Japan, during the period of March to June
1996. Thanks are due to Mr. T. Miyoshi, a graduate student of Tohoku University, for
collecting and arranging the data from various sources.
22 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

References
[1] Y. Uematsu, T. Miyoshi, K. Sasaki, M. Yamada, Wind damage to cladding of residential houses due
to typhoon in the Tohoku District, Japan, Tohoku J. Natural Disaster Sci. 34 (1998) 155}164 (in
Japanese).
[2] P.R. Sparks, S.D. Schi!, T.A. Reinhold, Wind damage to envelopes of houses and consequent
insurance losses, J. Wind Eng. Ind. Aerodyn. 53 (1994) 145}155.
[3] M. Jensen, Model-law for phenomena in natural wind, Ingenioren } Int. Ed. 2 (1958) 121}123.
[4] T. Stathopoulos, Wind loads on low-rise buildings: a review of the state of the art, Eng. Struct. 6 (1984)
119}135.
[5] J.D. Holmes, Wind Loads on Low-Rise Building } A Review, CSIRO, Division of Building Research,
Hignett, Victoria, Australia, 1993.
[6] P. Krishna, Wind loads on low rise buildings } a review, J. Wind Eng. Ind. Aerodyn. 54/55 (1995)
383}396.
[7] M. Kasperski, Design wind loads for low-rise buildings: a critical review of wind load speci"cations
for industrial buildings, J. Wind Eng. Ind. Aerodyn. 61 (1996) 169}179.
[8] Y. Uematsu, N. Isyumov, Peak gust pressures acting on the roof and wall edges of a low-rise building,
J. Wind Eng. Ind. Aerodyn. 77&78 (1998) 217}231.
[9] Y. Uematsu, N. Isyumov, Wind pressures acting on the roof and wall edges of a low-rise building: Part
4. Review of previous papers, J. Wind Eng. JAWE 78 (1999) 13}27 (in Japanese).
[10] K.J. Eaton, J.R. Mayne, The measurement of wind pressures on two-story houses at Aylesbury, J. Ind.
Aerodyn. 1 (1975) 67}109.
[11] K.J. Eaton, J.R. Mayne, N.J. Cook, Wind loads on low-rise buildings } e!ects of roof geometry, in:
Proceedings of the fourth International Conference on Wind E!ects on Buildings and Structures,
1975, Heathrow, Cambridge University Press, Cambridge, 1977, pp. 95}110.
[12] B.L. Sill, N.J. Cook, C. Fang, The Aylesbury comparative experiment: a "nal report, J. Wind Eng. Ind.
Aerodyn. 41}44 (1992) 1553}1564.
[13] L. Apperley, D. Surry, T. Stathopoulos, A.G. Davenport, Comparative measurements of wind
pressure on a model of the full-scale experimental house at Aylesbury, England, J. Ind. Aerodyn.
4 (1979) 207}228.
[14] P.J. Vickery, D. Surry, The Aylesbury experiments revisited } further wind tunnel tests and compari-
sons, J. Wind Eng. Ind. Aerodyn. 11 (1983) 39}62.
[15] A.E. Holdo, E.L. Houghton, F.S. Bhinder, Some e!ects due to variations in turbulence integral length
scales on the pressure distribution on wind-tunnel models of low-rise buildings, J. Wind Eng. Ind.
Aerodyn. 10 (1982) 103}115.
[16] B.L. Sill, N.J. Cook, P.A. Blackmore, IAWE Aylesbury comparative experiment } preliminary results
of wind tunnel comparisons, J. Wind Eng. Ind. Aerodyn. 32 (1989) 285}302.
[17] S.O. Hansen, E.G. Sorensen, The Aylesbury experiment. Comparison of model and full-scale tests,
J. Wind Eng. Ind. Aerodyn. 22 (1986) 1}22.
[18] P.J. Vickery, D. Surry, A.G. Davenport, Aylesbury and ACE: some interesting "ndings, J. Wind Eng.
Ind. Aerodyn. 23 (1986) 1}17.
[19] S. Mousset, The international Aylesbury collaborative experiment in C.S.T.B., J. Wind Eng. Ind.
Aerodyn. 23 (1986) 19}36.
[20] G.M. Richardson, A.P. Robertson, R.P. Hoxey, D. Surry, Full-scale and model investigations
of pressures on an industrial/agricultural building, J. Wind Eng. Ind. Aerodyn. 36 (1990)
1053}1062.
[21] G.M. Richardson, D. Surry, Comparisons of wind-tunnel and full-scale surface pressure measure-
ments on low-rise pitched-roof buildings, J. Wind Eng. Ind. Aerodyn. 38 (1991) 249}256.
[22] A.P. Robertson, E!ect of eaves detail on wind pressures over an industrial building, J. Wind Eng. Ind.
Aerodyn. 38 (1991) 325}333.
[23] E. Savory, N. Toy, S. Dalley, J. Trussler, Wind loading on a portal frame agricultural building,
J. Wind Eng. Ind. Aerodyn. 38 (1991) 335}345.
[24] S. Dalley, G. Richardson, Reference static pressure measurements in wind tunnels, J. Wind Eng. Ind.
Aerodyn. 41}44 (1992) 909}920.
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 23

[25] G.M. Richardson, D. Surry, The Silsoe building: a comparison of pressure coe$cients and spectra at
model and full-scale, J. Wind Eng. Ind. Aerodyn. 41}44 (1992) 1653}1664.
[26] E. Savory, S. Dalley, N. Toy, The e!ects of eaves geometry, model scale and approach #ow
conditions on portal frame building wind loads, J. Wind Eng. Ind. Aerodyn. 41}44 (1992)
1665}1676.
[27] R.P. Hoxey, P.J. Richards, Flow patterns and pressure "eld around a full-scale building, J. Wind Eng.
Ind. Aerodyn. 50 (1993) 203}212.
[28] G.M. Richardson, D. Surry, The Silsoe structures buildings: comparison between full-scale and
wind-tunnel data, J. Wind Eng. Ind. Aerodyn. 51 (1994) 157}176.
[29] G.M. Richardson, P.A. Blackmore, The Silsoe structures building: comparison of 1 : 100 model-scale
data with full-scale data, J. Wind Eng. Ind. Aerodyn. 57 (1995) 191}201.
[30] S. Dalley, Surface pressure spectra on a model of the Silsoe structures building and comparison with
full-scale, J. Wind Eng. Ind. Aerodyn. 60 (1996) 177}187.
[31] R.P. Hoxey, Structural response of a portal framed building under wind load, J. Wind Eng. Ind.
Aerodyn. 38 (1991) 347}356.
[32] A.P. Robertson, The wind-induced response of a full-scale portal framed building, J. Wind Eng. Ind.
Aerodyn. 41}44 (1992) 1677}1688.
[33] R.P. Hoxey, A.P. Robertson, Pressure coe$cients for low-rise building envelops derived from
full-scale experiments, J. Wind Eng. Ind. Aerodyn. 53 (1994) 283}297.
[34] G.M. Richardson, R.P. Hoxey, A.P. Robertson, J.L. Short, The Silsoe structures buildings: the
completed experiment: Part 1, in: Proceedings of the ninth International Conference on Wind
Engineering, vol. 3, New Delhi, Wiley Eastern Ltd., 1995, pp. 1103}1114.
[35] R.P. Hoxey, P.J. Richards, G.M. Richardson, A.P. Robertson, J.L. Short, The Silsoe structures
buildings: the completed experiment: Part 2, in: Proceedings ninth International Conference on Wind
Engineering, vol. 3, New Delhi, Wiley Eastern Ltd., 1995, pp. 1115}1126.
[36] P.J. Richards, R.P. Hoxey, B.S. Wanigaratne, The e!ect of directional variations on the observed
mean and rms pressure coe$cients, J. Wind Eng. Ind. Aerodyn. 54/55 (1995) 359}367.
[37] R.P. Hoxey, P.J. Richards, Full-scale wind load measurements point the way forward, J. Wind Eng.
Ind. Aerodyn. 57 (1995) 215}224.
[38] M.L. Levitan, K.C. Mehta, Texas Tech "eld experiments for wind loads: Part 1. Building and pressure
measuring system, J. Wind Eng. Ind. Aerodyn. 41}44 (1992) 1565}1576.
[39] M.L. Levitan, K.C. Mehta, Texas Tech "eld experiments for wind loads: Part II. Meteorological
instrumentation and terrain parameters, J. Wind Eng. Ind. Aerodyn. 41}44 (1992) 1577}1588.
[40] B.B. Yeatts, K.C. Mehta, D.A. Smith, Field experiments for wind e!ects on low buildings, in:
Proceedings of the ninth International Conference on Wind Engineering, vol. 1, New Delhi, Wiley
Eastern Ltd., 1995, pp. 500}511.
[41] M.L. Levitan, K.C. Mehta, W.P. Vann, J.D. Holmes, Field measurements of pressures on the Texas
Tech building, J. Wind Eng. Ind. Aerodyn. 38 (1991) 227}234.
[42] K.C. Mehta, M.L. Levitan, R.E. Iverson, J.R. McDonald, Roof corner pressures measured in the "eld
on a low building, J. Wind Eng. Ind. Aerodyn. 41}44 (1992) 181}192.
[43] C.W. Letchford, R.E. Iverson, J.R. McDonald, The application of the quasi-steady theory
to full scale measurements on the Texas Tech building, J. Wind Eng. Ind. Aerodyn. 48 (1993)
111}132.
[44] C.W. Letchford, K.C. Mehta, The distribution and correlation of #uctuating pressures on the Texas
Tech building, J. Wind Eng. Ind. Aerodyn. 50 (1993) 225}234.
[45] D.A. Smith, K.C. Mehta, B.B. Yeatts, S.V. Bhavaraju, Area-averaged and internal pressure coe$cients
measured in the "eld, J. Wind Eng. Ind. Aerodyn. 53 (1994) 89}103.
[46] J.R. McDonald, P.P. Sarkar, H. Gupta, Wind-induced loads on metal edge #ashings, in: Proceedings
of the ninth International Conference on Wind Engineering, vol. 3, New Delhi, Wiley Eastern Ltd.,
1995, pp. 1139}1150.
[47] Y.L. Xu, Model- and full-scale comparison of fatigue-related characteristics of wind pressures on the
Texas Tech building, J. Wind Eng. Ind. Aerodyn. 58 (1995) 147}173.
[48] D. Surry, Pressure measurements on the Texas Tech building: wind tunnel measurements and
comparisons with full scale, J. Wind Eng. Ind. Aerodyn. 38 (1991) 235}247.
24 Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25

[49] L.S. Cochran, J.E. Cermak, Full- and model-scale cladding pressures on the Texas Tech University
experimental building, J. Wind Eng. Ind. Aerodyn. 41}44 (1992) 1589}1600.
[50] H. Okada, Y.C. Ha, Comparison of wind tunnel and full-scale pressure measurement tests on the
Texas tech building, J. Wind Eng. Ind. Aerodyn. 41}44 (1992) 1601}1612.
[51] A.W. Rofail, Full-scale/model-scale comparisons of wind pressures on the TTU Building, in: Proceed-
ings of the ninth International Conference on Wind Engineering, vol. 3, New Delhi, Wiley Eastern
Ltd., 1995, pp. 1055}1066.
[52] H.W. Tieleman, D. Surry, K.C. Mehta, Full/model-scale comparison of surface pressures on the Texas
Tech experimental building, J. Wind Eng. Ind. Aerodyn. 61 (1996) 1}23.
[53] H.W. Tieleman, Model/full scale comparison of pressures on the roof of the TTU experimental
building, J. Wind Eng. Ind. Aerodyn. 65 (1996) 133}142.
[54] H.W. Tieleman, T.A. Reinhold, M.R. Hajj, Importance of turbulence for the prediction of surface
pressures on low-rise structures, J. Wind Eng. Ind. Aerodyn. 69}71 (1997) 519}528.
[55] J.C.K. Cheung, J.D. Holmes, W.H. Melbourne, N. Lakshmanan, P. Bowditch, Pressures
on a 1/10 scale model of the Texas Tech building, J. Wind Eng. Ind. Aerodyn. 69}71 (1997) 529}538.
[56] T.C.E. Ho, D. Surry, A.G. Davenport, The variability of low building wind loads due to surrounding
obstructions, J. Wind Eng. Ind. Aerodyn. 36 (1990) 161}170.
[57] T.C.E. Ho, D. Surry, A.G. Davenport, Variability of low building wind loads due to surroundings,
J. Wind Eng. Ind. Aerodyn. 38 (1991) 297}310.
[58] T.C.E. Ho, D. Surry, A.G. Davenport, Low building wind load variability with application to codes,
J. Wind Eng. Ind. Aerodyn. 41}44 (1992) 1787}1798.
[59] Architectural Institute of Japan, AIJ Recommendations for Loads on Buildings (English version),
1996.
[60] T. Stathopoulos, D. Surry, A.G. Davenport, E!ective wind loads on #at roofs, J. Struct. Div., ASCE
107 (1981) 281}298.
[61] A.G. Davenport, On the assessment of the reliability of wind loading on low buildings, J. Wind Eng.
Ind. Aerodyn. 11 (1983) 21}37.
[62] R.J. Kind, Worst suctions near edges of #at rooftops on low-rise buildings, J. Wind Eng. Ind. Aerodyn.
25 (1986) 31}47.
[63] A.P. Robertson, Codi"cation of local pressure coe$cients for low-rise structures, J. Wind Eng. Ind.
Aerodyn. 30 (1988) 143}152.
[64] D. Surry, Recent and current research into wind loading of low rise buildings at The University of
Western Ontario, J. Wind Eng. Ind. Aerodyn. 36 (1990) 1319}1329.
[65] D. Meecham, D. Surry, A.G. Davenport, The magnitude and distribution of wind-induced pressures
on hip and gable roofs, J. Wind Eng. Ind. Aerodyn. 38 (1991) 257}272.
[66] B. Bienkiewicz, Y. Sun, Local wind loading on the roof of a low-rise building, J. Wind Eng. Ind.
Aerodyn. 45 (1992) 11}24.
[67] H. Kawai, G. Nishimura, Characteristics of #uctuating suction and conical vortices on a #at roof in
oblique #ow, J. Wind Eng. Ind. Aerodyn. 60 (1996) 211}225.
[68] P.C. Case, N. Isyumov, The performance of wood buildings with 4 : 12 pitched roofs under extreme
wind loads, BLWT-SS28-1995, The Univ. of Western Ontario, 1995.
[69] H.J. Gerhardt, C. Kramer, Wind induced loading cycle and fatigue testing of lightweight roo"ng
"xations, J. Wind Eng. Ind. Aerodyn. 23 (1986) 237}247.
[70] H.J. Gerhardt, C. Kramer, E!ect of building geometry on roof windloading, J. Wind Eng. Ind.
Aerodyn. 41}44 (1992) 1765}1773.
[71] B. Bienkiewicz, H.J. Ham, Y. Sun, Proper orthogonal decomposition of roof pressure, J. Wind Eng.
Ind. Aerodyn. 50 (1993) 193}202.
[72] H. Kawai, Pressure #uctuations on square prisms } applicability of strip and quasi-steady theories,
J. Wind Eng. Ind. Aerodyn. 13 (1983) 197}208.
[73] T. Stathopoulos, D. Surry, A.G. Davenport, Some general characteristics of turbulent wind e!ects on
low-rise structures, in: Proceedings of the third Colloq. on Industrial Aerodynamics, Aachen, 1978,
Buildings Aerodynamics Part 1, 1978, pp. 211}224.
[74] R.J. Kind, Worst suctions near edges of #at rooftops with parapets, J. Wind Eng. Ind. Aerodyn. 31
(1988) 251}264.
Y. Uematsu, N. Isyumov / J. Wind Eng. Ind. Aerodyn. 82 (1999) 1}25 25

[75] R.V. Milford, J.L. Waldeck, Statistics of full-scale surface pressures, J. Wind Eng. Ind. Aerodyn. 30
(1988) 35}44.
[76] H.W. Tieleman, Simulation of surface winds for assessment of extreme wind loads on roofs, in:
Proceedings of the ninth International Conference on Wind Engineering, vol. 3, New Delhi, Wiley
Eastern Ltd., 1995, pp. 1162}1169.
[77] T. Stathopoulos, PDF of wind pressures on low-rise buildings, J. Struct. Div. ASCE 106 (1980)
973}990.
[78] J.D. Holmes, Non-Gaussian characteristics of wind pressure #uctuations, J. Wind Eng. Ind. Aerodyn.
7 (1981) 103}108.
[79] N.J. Cook, J.R. Mayne, A novel working approach to the assessment of wind loads for equivalent
static design, J. Ind. Aerodyn. 4 (1979) 149}164.
[80] N.J. Cook, J.R. Mayne, A re"ned working approach to the assessment of wind loads for equivalent
static design, J. Wind Eng. Ind. Aerodyn. 6 (1980) 125}137.
[81] N.J. Cook, Calibration of the quasi-static and peak-factor approaches to the assessment of wind loads
against the method of Cook and Mayne, J. Wind Eng. Ind. Aerodyn. 10 (1982) 315}341.
[82] T.V. Lawson, Wind E!ects on Buildings, vol. 1, Design Applications, Applied Science Publishers,
Basking, 1980.
[83] E. Maruta, H. Ueda, M. Kanda, The local pressure acting on gable roof, The 24th Annual Meeting of
College of Industrial Technology, Nihon University, 1991, pp. 33}36.
[84] The National Building Code of Canada, National Research Council of Canada, 1990.
[85] J. Blessmann, Wind pressures on roofs with negative pitch, J. Wind Eng. Ind. Aerodyn. 10 (1982)
213}230.
[86] T. Stathopoulos, A.R. Mohammadian, Wind loads on low buildings with mono-sloped roofs, J. Wind
Eng. Ind. Aerodyn. 23 (1986) 81}97.
[87] T. Stathopoulos, P. Saatho!, Wind pressure on roofs of various geometries, J. Wind Eng. Ind.
Aerodyn. 38 (1991) 273}284.
[88] T. Stathopoulos, P. Saatho!, Codi"cation of wind pressure coe$cients for sawtooth roofs, J. Wind
Eng. Ind. Aerodyn. 41}44 (1992) 1727}1738.
[89] P.J. Saatho!, T. Stathopoulos, Wind loads on buildings with sawtooth roofs, J. Struct. Eng. ASCE 118
(1992) 429}446.
[90] Australian Standard, Minimum Design Loads on Structures. Part 2: Wind Loads, AS1170.2, 1989.
[91] P.A. Blackmore, Load reduction on #at roofs } the e!ect of edge pro"le, J. Wind Eng. Ind. Aerodyn.
29 (1988) 89}98.
[92] A. Baskaran, T. Stathopoulos, Roof corner wind loads and parapet con"gurations, J. Wind Eng. Ind.
Aerodyn. 29 (1988) 79}88.
[93] L.S. Cochran, J.E. Cermak, E.C. English, Load reduction by modifying the roof corner vortex,
Proceedings of the ninth International Conference on Wind Engineering, vol. 3, New Delhi, Wiley
Eastern Ltd., 1995, pp. 1091}1102.
[94] D. Surry, J.X. Lin, The e!ect of surroundings and roof corner geometric modi"cations on roof
pressures on low-rise buildings, J. Wind Eng. Ind. Aerodyn. 58 (1995) 113}138.

You might also like