Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Wind Loading on Attached Canopies: Codification Study

José D. Candelario, S.M.ASCE 1; Ted Stathopoulos, F.ASCE 2; and Ioannis Zisis, M.ASCE 3

Abstract: A wind tunnel study was performed to examine wind loads on canopies attached to the walls of low-rise buildings. A model of a
building with an attached canopy of geometric scale of 1∶100 was constructed and tested in a simulated open terrain exposure. The attached
canopy model was equipped with pressure taps at both upper and lower surfaces to allow for simultaneous monitoring of wind pressures and
evaluation of the overall load. A total of 63 different building/attached canopy configurations were tested for 28 wind directions. Pressure and
correlation coefficients were generated to provide a better understanding of how the wind-loading patterns at upper and lower surfaces of the
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

attached canopy contribute to the net loading effect. Current design guidelines and building code and standard provisions are assessed and
compared with the experimental results of the present study. The influence of the geometry of each configuration on the experimental net
pressure coefficients was assessed and recommendations for design wind load standards and codes of practice are made. DOI: 10.1061/
(ASCE)ST.1943-541X.0001007. © 2014 American Society of Civil Engineers.
Author keywords: Wind tunnel tests; Wind loads; Low-rise buildings; Load distribution; Standards and codes; Net pressure coefficients;
Canopy; Patio covers; Wind effects.

Introduction tunnel measurements on a building model with an instrumented


attached canopy. A commentary on comparisons with available
An attached canopy is a roof adjacent to the wall of a building. This international studies and design guidelines is provided, along
kind of addition is commonly used in suburban low-rise housing with recommendations for design guidelines for possible imple-
because it provides a space adequate for many social and leisure mentation of future standards and codes of practice.
activities with partial protection from the elements. The wall to This study extends significantly the results of a previous pilot
which the canopy is attached is referred to as the parent wall and study (Zisis and Stathopoulos 2010) also undertaken in the
represents the only enclosed side of the canopy. Building Aerodynamics Laboratory at Concordia University but
Currently, only limited knowledge exists in terms of design that addressed only a very limited number of cases.
loads for these types of structures. Neither the National Building
Code of Canada 2010 (NBCC 2010) nor the ASCE 7-10 standard
(ASCE 2010) contain any wind load provisions for the proper de- Previous Studies and Design Provisions
sign of attached canopies. In North America, only the International
Building Code (IBC 2006a) and the International Residential Code Previous studies of wind loads on attached canopies provided a
(IRC 2006b) contain some provisions to address this issue. How- foundation for the development of the present study. Parameters,
ever, these provisions do not seem to result from a detailed study. procedures, and findings, and the inconsistencies presented among
Given this lack of design provisions in the North American one another, were carefully investigated before defining the test
codes and standards, designers and practitioners have found ways program of the present study.
to obtain approximate design loads using the aforementioned One of the first studies to address wind loads on attached
codes. One such example is to treat the attached canopy as an canopies is that of Jancauskas and Holmes (1985). Experiments
extension of the roof overhang and design it according to the were performed in a boundary layer wind tunnel for a roughness
provisions available in ASCE 7-10 (ASCE 2010) or NBCC that simulated an urban terrain exposure. This study tested a total
(2010). A detailed study dealing with wind force coefficients on of 14 configurations by varying the height of the building, the
canopy roofs was carried out by Uematsu et al. (2007, 2008a, b). height at which the canopy is placed, the length of the canopy,
The purpose of this paper is to present the methodology and its width. This study forms the basis of the provisions available
used and experimental findings obtained from a series of wind from Australian/New Zealand (AS/NZS) 1170.2:2011 [Australia/
New Zealand (AS/NZS) 2011].
1
Graduate Student, Dept. of Building, Civil and Environmental Engi- Paluch et al. (2003) carried out extensive work regarding
neering, Centre for Building Studies, Concordia Univ., Montreal, QC, wind loads on canopies attached to buildings with an arched roof.
Canada H3G 1M8 (corresponding author). E-mail: jose.candelario@ However, pressure coefficients presented in this paper consider s
upr.edu eparately the loads applied on the upper and lower surfaces of the
2
Professor, Dept. of Building, Civil and Environmental Engineering, canopy and not their net effects. Greater focus is noted as being
Centre for Building Studies, Concordia Univ., Montreal, QC, Canada given to the loads acting on the roof as a result of the presence of
H3G 1M8. a canopy than to the loads on the canopy itself.
3
Assistant Professor, Dept. of Civil and Environmental Engineering,
Hölscher et al. (2007) performed an extensive study on wind-
Florida International Univ., Miami, FL 33174.
Note. This manuscript was submitted on April 15, 2013; approved on induced pressures on canopies attached to the walls of flat roof
November 14, 2013; published online on February 12, 2014. Discussion buildings. Experiments were carried out in a boundary layer wind
period open until July 12, 2014; separate discussions must be submitted tunnel both with and without surroundings. For each configuration,
for individual papers. This paper is part of the Journal of Structural En- 24 wind directions were tested. Both local and area-averaged
gineering, © ASCE, ISSN 0733-9445/04014007(14)/$25.00. peak loads were analyzed. The results form the basis of the design

© ASCE 04014007-1 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


guidelines provided in Eurocode 1 (European Committee for Building Aerodynamics Laboratory. An acrylic glass model with
Standardization 2004) for Wind Actions [DIN EN 73 1991-1- a scale of 1∶100 serves as the parent building to which a metallic can-
4:2010-12, Deutsches Institut für Normung (DIN) 2010]. opy model of the same geometric scale is to be attached. The building
Zisis and Stathopoulos (2010) conducted a detailed study on model has external dimensions of 15 by 10 cm (length and width)
wind-induced pressures on attached canopies. The experiments and a ridge height of 12.30 cm, as shown in Fig. 2. The gabled
were performed in a boundary layer wind tunnel for a roughness roof has a slope of 4∶12. The parent wall of the building model con-
simulating an open terrain exposure. The loads were determined sists of five wall segments of unique widths that may be arranged in
using a system of pressure sensitive scanners. For each configura- different ways to provide a slot at different positions to which the
tion, 28 wind directions were tested. This study presents both local canopy model is attached (Fig. 3). In addition, the height of the build-
and area-averaged peak loads, and pressure and correlation coeffi- ing model may be lowered using an adjustable base.
cient contour plots, to provide a better understanding of the loading Two separate canopy models were made out of thin sandwiched
patterns on attached canopies. Three different configurations were metallic plates. One of the models stretches the entire length of the
tested for an attached canopy extending over the entire length of the parent wall (15 cm) and the other stretches over half of the length
parent wall. (7.5 cm). Both models have the same width of 3.65 cm, which may
In addition to the Australian/New Zealand (AS/NZS) be further reduced by pushing it deeper into the slot of the building
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

1170.2:2011 [Australia/New Zealand (AS/NZS) 2011] DIN EN model. Pressure taps were placed at both upper and lower surfaces
1991-1-87 4/NA:2010-12 (DIN 2010), provisions for the wind load of the canopy at almost the same locations to form pressure tap
design of attached canopies are also available in the Austrian Stan- pairs that enable the determination of net pressure coefficients.
dard (ÖNORM B 4014-1), the Swiss Standard (Schweizerische Pressure tap locations for both models are shown in Fig. 4. The
Norm, SIA 261), and the Indian Standard (IS: 875 Part-3-1987) full-length canopy model has a total of 30 pressure taps (15 pres-
code of practice for design loads. Two main geometric parameters sure tap pairs), and the half-length canopy model has a total of 18
are present in most of these design provisions and are commonly pressure taps (nine pressure tap pairs).
considered to have the greatest effect on wind loads, the hc/h ratio To refer to the relevant parameters, this study followed the no-
(height of canopy to eave height), and the hc/wc ratio (height of menclature illustrated in Fig. 5. The eave height (h) and the wall
canopy to width of canopy). Additional parameters were also con- length (l) refer to the dimensions of the parent building, whereas
sidered, but their effect was deemed less significant. the canopy height (hc ), canopy width (wc ), and canopy length (lc )
To provide a quick comparison, three different canopy and parent describe the geometry and location of the attached canopies. The
building geometries were selected and designed in accordance with canopy edge to building edge distance (ed ) accounts for the hori-
the five available design guidelines. Fig. 1 presents the maximum zontal location of the attached canopy. The adjustable model results
(uplifting; −) and minimum (downwards; +) net pressure coefficients in the possibility of testing 63 configurations, that is, each of the
obtained from each geometry. These coefficients were normalized by three cases (Case I, II, and III) shown in Fig. 5 tested for the 21
the dynamic velocity pressure on the basis of the 3-s gust. The incon- configurations described in Table 1.
sistencies that arise depending on design guidelines followed are The dimensionless ratios in Table 1 were selected to cover
clearly noted. Design uplifting pressure coefficients vary by up to geometries commonly found in practice, in combination with other
a 70% difference for canopies located near the eave height. Further- less common configurations. As a result, more comprehensive de-
more, a zero value was obtained in one of the cases for a canopy sign guidelines are obtained and the trends and effects of the rel-
attached at a lower relative height. These discrepancies served as ative geometries are better assessed. The following four ratios were
motivation and demonstrated the pertinence of the present study. tested as principal variables in this study with their corresponding
ranges in brackets:
• hc=h ¼ f0.20 to 0.94g;
Wind Tunnel Study • hc=wc ¼ f0.58 to 8.54g;
• lc=l ¼ f0.5 to 1g; and
Experimentation for the present study was performed in the
Boundary Layer Wind Tunnel (BLWT) at Concordia University’s • ed=lc ¼ f0 to 1g.
Please note that Table 1 also indicates with a star the three con-
figurations tested previously for Case I (Zisis and Stathopoulos
-2
2010), the results of which were used for repeatability checks.
-1.5 The experiments were performed for a simulated open terrain
-1
exposure. The average and root mean square longitudinal wind
Net Pressure Coefficient

SIA
velocities (V̄z and Vz; rms) were measured at different heights
-0.5 DIN
ÖNORM at the center of the wind tunnel test section without the model
AS/NZS in place. The corresponding average velocity (V̄z) and longitudinal
0 IS

0.5

1.5

2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
hc/h

Fig. 1. Local net pressure coefficients for the three buildings illustrated
as obtained from the labeled design code provisions for a velocity
pressure on the basis of 3-s gust wind speed data Fig. 2. Building configuration tested (cm)

© ASCE 04014007-2 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


Fig. 3. Building model with different canopies (images by José D. Candelario)
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Top view of building and canopy model showing pressure tap locations (cm)

Fig. 5. Nomenclature of dimensions and description of three principal cases tested in terms of canopy length (l) and edge-to-edge distance (ed )

© ASCE 04014007-3 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


Table 1. Configurations Tested for Each of the Three Cases for 63 Unique
Configurations
Test h (cm) hc (cm) wc (cm) hc=h hc=wc
a
1 3.5 2.85 3.65 0.81 0.78
2 3.5 2.85 2.38 0.81 1.2
3 3.5 2.85 1.13 0.81 2.52 Fig. 7. Illustration of the generation of net pressure coefficients, Cp;net
4a 7 6.35 3.65 0.91 1.74
5 7 6.35 2.38 0.91 2.67
6 7 6.35 1.13 0.91 5.62
7a 7 3.5 3.65 0.5 0.96 The equipment used for the flow visualization consisted of a
8 7 3.5 2.38 0.5 1.47
Dantec smoke generator and high-speed video equipment. The
9 7 3.5 1.13 0.5 3.1
10 10.3 9.65 3.65 0.94 2.64
instrumentation used to measure the wind flow characteristics
11 10.3 9.65 2.38 0.94 4.05 was a TFI Cobra multihole probe with a frequency response
12 10.3 9.65 1.13 0.94 8.54 exceeding 2,000 Hz, suitable for both turbulent and mean flow
measurements. Pressure taps placed on the model are connected
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

13 10.3 7 3.65 0.68 1.92


14 10.3 7 2.38 0.68 2.94 by urethane tubing to a pressure-sensitive scanner (ZOC33/64
15 10.3 7 1.13 0.68 6.19 Px-2003, Scanivalve Corp.), which is in turn connected to the data
16 10.3 3.5 3.65 0.34 0.96 acquisition system (DSM 3000, Scanivalve Corp.). A Pitot tube
17 10.3 3.5 2.38 0.34 1.47 was placed at the free flow above the boundary layer and was
18 10.3 3.5 1.13 0.34 3.1 connected to the scanning system to measure the dynamic and
19 10.3 2.1 3.65 0.2 0.58
static pressure. The system was set to operate at a scanning fre-
20 10.3 2.1 2.38 0.2 0.88
21 10.3 2.1 1.13 0.2 1.86 quency of 250 Hz generating 8,200 pressure readings in approx-
a
imately 32 s. These values simulate the wind flow characteristics
Configurations previously tested for Case I (Zisis and Stathopoulos 2010). of a full-scale storm event one-hour long. As a result, the data
presented in this study correspond to hourly wind speed data un-
less stated otherwise. A total of 28 wind directions were examined
Vz,rms / Vz for each of the 63 configurations tested. The step size of wind
1
directions ranges from 5° in critical ranges to 25° in ranges in
Experimental Turbulence Intensity
Experimental Mean Wind Speed
which critical pressures are less likely to occur.
0.8 I(z)==c1/Ln(Z/Zo);
I(z) (z/10)^-d Zo = 0.01cm All data acquired by the pressure scanner are presented in
Vz/Vg = (Z/Zg)^α, for α = 0.14 dimensionless pressure coefficient form in accordance with the
following equation:
0.6
Cp ¼ ΔP=qmrh ¼ ðPt − Pa Þ=qmrh ð2Þ
Z/Zg

0.4
where Pt = surface pressure at any tap; Pa = atmospheric pressure;
qmrh = dynamic velocity pressure at mean roof height converted
from qpitot using the power law velocity profile as follows:
0.2
qmrh ¼ qpitot ½ðzmrh =zg Þα 2 ð3Þ

0 Because the canopy is generally a thin element exposed to wind


0 0.2 0.4 0.6 0.8 1 pressures on both upper and lower surfaces, considering the pres-
Vz / Vg sures acting simultaneously on each plane is essential. This analysis
is done using the net pressure coefficients as defined in the follow-
Fig. 6. Velocity and turbulence intensity profiles ing equation:

Cp;net ¼ ΔP;net =qmrh ¼ ðΔP;upper − ΔP;lower Þ=qmrh


turbulence intensity (Vz; rms=V̄z) profiles are shown in Fig. 6.
Plotting the power law with an exponent of α ¼ 0.14 results in ¼ Cp;upper − Cp;lower ð4Þ
the best agreement with the measured values. The experimental
turbulence intensity profile also portrays consistency with the in which ΔP;upper and ΔP;lower are measured at the top and bottom
analytical curve derived from the logarithmic law as components, respectively, of a pressure tap pair, as illustrated in
Fig. 7. Importantly, note that the negative sign represents a pressure
I ðzÞ ¼ ð2.5v Þ=½ðv =0.4Þ lnðz=z0 Þ ¼ 1=½lnðz=z0 Þ ð1Þ directed away from the surface (suction) and a positive sign rep-
resents a pressure directed toward a surface. If this convention is
where z is the height above the surface, zo is defined as the rough- maintained when computing net loads in accordance with Eq. (3),
ness length found to be 0.01 cm for an open terrain exposure, and a negative value for a Cp;net results in a net uplifting load, whereas a
v is the shear velocity. positive value results in a net downward loading.
The turbulence intensities at the mean roof height (mrh) for Throughout this paper, peak pressure coefficients may be iden-
the three different parent building heights tested were found to tified as either local or area-averaged. A local peak Cp refers to the
be 17.9% (mrh ¼ 4.5 cm), 15.5% (mrh ¼ 8 cm), and 14.4% critical value experienced at a single pressure tap (or pressure tap
(mrh ¼ 11.3 cm), thus complying with the importance of proper pair in the case of local Cp;net ). An area-averaged Cp refers to the
simulation of the turbulence intensity when dealing with low-rise peak value that the entire area experiences as determined by the
buildings (Tieleman et al. 1998). average of pressure measured simultaneously at every pressure

© ASCE 04014007-4 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


tap (or pressure tap pair). Both local and area-averaged pressure leeward wall (180°). Clearly, local peak Cp;net values display a
coefficients are referred to as minimum, maximum, or mean, de- great sensitivity to wind direction. In the configuration examined,
fined as follows: the local peak minimum Cp;net obtained at 330° degrees is approx-
⌣ imately five times that encountered at 180°.
Cp;minimum ¼ ðPt − Pa Þ=qmrh ð5Þ In the case presented in Fig. 9 for a canopy extending half the
length of the parent wall and placed eccentrically along the length
of the parent wall, a very different behavior is observed. The high-
Cp;maximum ðPbt − Pa Þ=qmrh ð6Þ
est magnitude local peak minimum Cp;net values are found to occur
for wind directions running parallel to the building ridge 90–270°
Cp;mean ¼ ðP̄t − Pa Þ=qmrh ð7Þ and the highest local peak maximum Cp;net values occur at wind
⌣ directions nearly perpendicular to the ridge at 15° or 345°. As in the
where Pt , Pbt , and Pt are the extreme (peak) minimum, extreme previous case, the lowest magnitudes for both local maximum and
(peak) maximum, and mean pressure recorded during the simulated minimum Cp;net values were found to occur when the canopy is
storm, respectively. Peak pressure coefficients were evaluated as placed at the leeward wall.
the average of the maximum ten measured values picked up from Clearly, wind direction has a significant effect on the local peak
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

a 1-h full-scale equivalent time history of the respective signal. Cp;net . The response of peak Cp;net values to wind direction was
This approach is considered a good approximation to the mode noted to be different for every configuration. For this reason, every
value of a detailed extreme value distribution and was used in
peak Cp and Cp;net value presented subsequently in this paper
previous wind tunnel studies.
refers to its corresponding critical wind direction unless otherwise
For individual surface pressures (at one pressure tap), the mini-
stated.
mum peak is associated with the highest suction and the maximum
peak is associated with the highest positive pressure. For net pres-
sures, the minimum peak is associated with the highest net uplifting Visualization of Flow Around the Canopy
load and the maximum peak is associated with the highest down-
ward loading. Flow visualization experiments were performed for a canopy at-
tached to mid-height and near the eave height of the parent wall.
Figs. 10 and 11 provide still shots for zero-degree wind direction
Results and Discussion for these configurations. For a canopy attached at mid-height, stag-
nation of flow occurs at the parent wall both above and below the
canopy. Stagnation of flow above the canopy results in a downward
Effect of Wind Direction on Local Net Pressure flow running along the surface of the parent wall, which ultimately
Coefficients inflicts a downward force on the upper surface of the canopy.
Pressure coefficients on attached canopies may vary significantly In turn, stagnation of flow in the parent wall underneath the canopy
with wind direction. Figs. 8 and 9 display the local maximum, min- generates flow directed toward the lower surface of the canopy, re-
imum, and mean Cp;net values as a function of wind direction for sulting in an upward force. These two counteracting forces at upper
two different configurations. Given that the highest local Cp;net and lower surfaces are expected to result in a decreased net uplifting
value may occur at a different location of the canopy for different pressure coefficient (min Cp;net ).
wind directions, the local net pressure coefficients presented in In the case of a canopy attached closer to the eave height, flow
Figs. 8 and 9 correspond to the highest value found at any pressure stagnates significantly at the parent wall underneath the canopy
tap pair throughout the entire canopy for each particular wind di- (Fig. 11), resulting in a dominating upward flow that generates
rection. Given symmetry, only azimuths from 180° to 360° are pre- a force directed toward the lower surface of the canopy. In contrast,
sented in Fig. 8. high levels of flow separation occur at the upper surface of the
Fig 8, which shows a canopy extending over the entire length of canopy, resulting in dominating suctions. The combination of high
the parent wall, the highest magnitudes for local peak minimum suctions acting on the upper surface and high pressure on the lower
Cp;net values is seen to occur at the 330° wind direction. In addi- surface is expected to result in an increased uplifting force, i.e., a
tion, the lowest magnitudes for both local maximum and minimum higher net uplifting pressure coefficient.
Cp;net values were found to occur when the canopy is placed at the

-4 -4

-3 Peak Min -3

-2 -2
Local Cp, net
Local Cp, net

-1 -1 Peak Min
Mean
0 0
Mean
1 1
Peak Max Peak Max
2 2
hc/h = 0.94 hc/h = 0.5
3 hc/wc = 2.64 3 hc/wc = 0.96
ed/lc = 0.00 ed/lc = 1
4 4
180 210 240 270 300 330 360 0 30 60 90 120 150 180 210 240 270 300 330 360
Wind Direction ( º ) Wind Direction ( º )

Fig. 8. Effect of wind direction on local peak minimum, peak Fig. 9. Effect of wind direction on local peak minimum, peak
maximum, and mean Cp;net for a canopy extending over the full length maximum, and mean Cp;net for an eccentrically placed canopy
of parent wall extending over half the length of parent wall

© ASCE 04014007-5 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


hc/h = 0.5 wc = 3.65m lc = 15m
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10. Flow visualization for a canopy attached at the midheight of parent wall and zero degree azimuth (image by José D. Candelario)

hc/h = 0.94 wc = 3.65m lc = 15m

Fig. 11. Flow visualization for a canopy attached close to the eave height of parent wall and zero degree azimuth (image by José D. Candelario)

Thus, the expectation is that the higher the location of the pressures acting on upper and lower surfaces of canopies placed
canopy along the parent wall, that is, the higher the hc=h ratio, far from the eave (lower hc=h ratios) is seen to result in significant
the higher the net uplifting pressure will be. Conversely, that higher reductions to the loads experienced on either side, thus resulting in
net downward forces occur for lower hc=h ratios may also be a reduced min Cp;net .
expected. In contrast, Fig. 13 shows that for a canopy placed closer to the
eave height, dominant suctions occur on the upper surface in com-
bination with dominant pressures on the lower surface. The high
Net Pressures and Correlation Coefficients suctions on the upper surface, particularly along the corners and
Instrumentation of the canopy model with pressure taps at both leading edges, are attributed to the phenomenon of flow separation
upper and lower surfaces allows for separate monitoring of pressure (refer to Fig. 11 for flow visualization photograph). The dominant
coefficients at either surface (Cp;upper ; Cp;lower ) or their simultane- high pressures on the lower surface of the canopy result from
ous effect at the same position of the canopy (Cp;net ). This effect the upward flow generated by the stagnation of flow underneath
provides the possibility of a clearer understanding of the wind the surface of the canopy. In this case, a peak min Cp;upper
loading patterns to which the attached canopy is exposed. Figs. 12 of −2.30, in combination with a peak max Cp;lower of 0.60 (not
and 13 show pressure and correlation coefficient contour plots for at the same location), results in a peak min Cp;net of −2.70.
a canopy attached to both the mid-height and eave height of the The correlation coefficients for this configuration range from weak
parent wall, respectively, again for a zero-degree wind direction. to zero, implying that the high suctions occurring at the upper
Fig. 12 shows that for a wind direction perpendicular to the pa- surface of the canopy do not occur in a well-defined pattern with
rent wall, the upper surface of the canopy experiences downward the positive pressures on the lower surface of the canopy. This
loading of higher magnitudes than the suctions. In the lower sur- phenomenon is reflected by the fact that the min Cp;net value is
face, pressures are also significantly higher than suctions. However, increased but only by a small amount more than the individual
note that despite having a peak max Cp;lower of more than 1.20, the contribution of the min Cp;upper and max Cp;lower to the net
peak min Cp;net was found to be −0.20. This phenomenon of coun- uplifting forces on the canopy. Despite their weak correlation,
teracting forces is further observed by the correlation coefficient the combined effect of the high suction and high pressure contri-
contour plot (Fig. 12) exhibiting high positive values between butions at upper and lower surfaces respectively, result in an
upper and lower surfaces of the canopy. The combined effect of increase min Cp;net for higher hc=h ratios.

© ASCE 04014007-6 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 12. Pressure and correlation coefficient contour plots for a canopy attached at the midheight of parent wall and zero degree azimuth

However, peak Cp;net values were found to not necessarily oc- each case: the local minimum and maximum Cp;net (Loc. Min.;
cur for wind directions perpendicular to the length of the canopy. Loc. Max.), which refer to the critical values experienced at a single
Fig. 14 provides pressure and correlation coefficient contour plots pressure tap pair, and the area-averaged minimum and maximum
for a canopy attached at the mid-height of the parent wall for the Cp;net (Area-avg. Min.; Area-avg. Max.), which refer to the peak
critical wind direction. When compared with the contours for a value that the entire area experiences determined by the critical
zero-degree wind direction for the same configuration (refer to simultaneous averages at every pressure tap pair.
Fig. 12) the peak min Cp;upper is seen to significantly increased
at the leftmost corner, and the peak max Cp;lower is decreased.
Furthermore, the correlation coefficient at the same corner was de- Effect of hc=h
creased. As a result, the peak min Cp;net was increased significantly Local and area-averaged Cp;net values as a function of the hc=h
compared with the zero-degree wind direction. ratio are presented in Fig. 16. Clearly, a higher location of the can-
Fig. 15 provides pressure and correlation coefficient contour opy along the parent wall generally results in higher net uplifting
plots for a canopy attached near the eave height of the parent wall pressures (min Cp;net ) on the canopy. Also note that for hc=h ratios
for the critical wind direction. When compared with the contours smaller than 0.5, peak min Cp;net values display little sensitivity to
for a zero-degree wind direction for the same configuration a slight increase in the canopy height. For intermediate heights
(Fig. 13), the min Cp;upper and max Cp;lower are seen to signifi- (0.5 < hc=h < 0.9), a higher sensitivity is observed as the slope
cantly increase at the rightmost corner. Moreover, the correlation
for the local peak min Cp;net is increased. Finally, for hc=h values
coefficient may have slightly increased. As a result, the min Cp;net
greater than or equal to 0.9, a more notable increase occurs. The
is increased by approximately 1.5 times the magnitude of that
change from a low to a high hc=h ratio results in an increase of
found for the zero-degree wind direction. Note that the peak min
more than three times the magnitude of the peak local min Cp;net .
Cp;net ¼ −3.89 recorded was the largest observed for any configu-
Thus, the conclusion is that net uplifting pressures are highly sen-
ration examined.
sitive to the hc=h ratio, which is consistent with the previous
observation from the pressure coefficients contours. In contrast, the
Effect of Geometric Parameters on Critical Net net downward pressure coefficients (max Cp;net ) portray a smaller
Pressure Coefficients sensitivity to the hc=h ratio. Furthermore, Fig. 16 shows that the
This section examines the effect of each parameter on the loads area-averaged max Cp;net shows similar, although less pronounced,
exerted on the canopy by varying a single parameter and keeping characteristics with the local values. For instance, the highest maxi-
the others constant. Four dependent variables are investigated for mum local Cp;net encountered is only 1.3 times larger than the

© ASCE 04014007-7 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. Pressure and correlation coefficient contour plots for a canopy attached close to the eave height of parent wall and zero degree azimuth

smallest. In general, the net downward pressures are less sensitive Fig. 18 shows the local and area-averaged Cp;net values as a
to the hc=h ratio than the net uplifting pressures. function of hc=wc for an hc=h ratio of 0.34. Clearly, at this lower
hc=h ratio, the effect of hc=wc on the net uplifting forces appears
Effect of hc=w c negligible for both local and area-averaged peak Cp;net values. For
the net downward pressures, increasing hc=wc results in a marked
Fig. 17 presents local and area-averaged Cp;net values as a function increase in the max Cp;net values. This trend is contrary to that
of the hc=wc ratio for an hc=h ratio of 0.94 because such a pre- found for the case of high hc=h ratios; however, for lower hc=h
sentation showed as more critical in Fig. 16. When the width of the ratios, the downward flow along the parent wall dominates the up-
canopy (wc) is decreased and the hc=h ratio is maintained as con- ward flow. As a result, the portion of the total canopy area exposed
stant, the net uplifting forces initially increase, which is explained to the strongest downward flow becomes more significant for
by the reduction in the surface area for the reattachment of flow to decreases in wc.
occur. However, further reducing the width causes the vortices gen-
erated in the gap between the canopy and the roof overhang to pos-
Effect of Canopy Length (lc) and Location (ed )
sibly have a greater influence on the net uplift forces, resulting in a
marked reduction in both local and area-averaged min Cp;net val- The effects of the length of the canopy (lc) and the largest edge-to-
ues. However, the highest local min Cp;net encountered is only 1.3 edge distance (ed) on the net pressure coefficients were examined.
times larger than the lowest. Fig. 19 presents local and area-averaged Cp;net values as a function
In contrast, net downward pressure coefficients portray an of the ed=lc ratio for hc=h ¼ 0.94 and wc ¼ 3.65 m. An increase
inversely proportional relationship to the hc=wc ratio, probably in the ed=lc ratio generally corresponds to a decrease in the
attributable to the effect of the stagnation occurring at the portion magnitude of the local min Cp;net , whereas the area-averaged
of the parent wall located underneath the canopy. The portion of the min Cp;net portrays little sensitivity. Also note that the case of
total canopy area exposed to the strongest upward flow becomes the full-length canopy (lc=l ¼ 1) has a significantly higher local
more significant when wc is decreased. As a result, the net down- min Cp;net compared with the other two half-length canopy
ward loading is decreased. The highest local max Cp;net encoun- (lc=l ¼ 0.5) cases. Thus, the lc=l ratio is significant for the net
tered is 2.5 times larger than the lowest. uplifting pressures, whereas the edge distance seems of secondary
Given that the proximity of the canopy to the roof overhang and importance.
the dominance of upward flow are considered critical on the effect For net downward pressure coefficients, no significant differ-
of the hc=wc ratio on a configuration with a high hc=h ratio, in- ence appears to exist between the full-length canopy model and
terest in examining the effect of hc=wc on a lower hc=h ratio arises. the half-length models. The same is true for the distance from

© ASCE 04014007-8 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. Pressure and correlation coefficient contour plots for a canopy attached at the midheight of parent wall and critical wind direction (75°)

the edge, which shows only a slight increase when going from a significant inconsistencies among themselves. Publications on
canopy placed at the center (ed=lc ¼ 0.5) to a canopy placed at wind tunnel experimentation and analyses that led to the develop-
the corner (ed=lc ¼ 1) of the parent wall. ment of the AS/NZS and the DIN design provisions (Jancauskas
Therefore, local uplift forces are sensitive to changes in the lc=l and Holmes 1985; Hölscher et al. 2007) facilitate comparisons
ratio but less sensitive to the largest edge-to-edge distance (ed). of the experimental results of the present study and those from these
In contrast, local downwards wind-induced forces are more sensi- studies. Fig. 20 compares peak negative net pressure coefficients
tive to the ed=lc ratio and display little sensitivity to the lc=l ratio. measured by Jancauskas and Holmes (1985) and those from the
present study. These coefficients are on the basis of dynamic veloc-
Comparisons with Experimental Results of Previous ity pressure evaluated from the mean hourly wind speed. For all
Studies cases considered, the results of the present study generally portray
a good agreement with the previous results.
As previously stated, several wind standards and codes of practice The experimental results and analyses that led to the design
containing provisions for the design of attached canopies show guidelines provided in the DIN (Hölscher et al. 2007) were

© ASCE 04014007-9 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 15. Pressure and correlation coefficient contour plots for canopy attached close to eave height of parent wall and critical wind direction (330°)

compared with those of the present study in Fig. 21. Peak net Toward Codification
pressure coefficients are also on the basis of dynamic velocity
pressure evaluated from the mean hourly wind speed. The com- In addition to the local pressure coefficients presented in this paper,
parison shows that the results of the present study are higher, area-averaged values are necessary for codification purposes. The
area-averaging effect on min and max Cp;net peaks was determined
particularly for higher hc=h ratios, which may be attributed to
by considering single or multiple sets of pressure taps and assigning
the increased number of wind directions examined in the present
them to their corresponding effective surface areas. The three plots
study. For instance, within azimuths ranging from 30° to 90°, the shown in Fig. 22 (for cases I, II, and III; see Fig. 5) contain all
present study considers six different wind directions as opposed to of the experimental results in terms of peak net pressure coefficients
three in Hölscher et al. (2007), which may have a significant effect as a function of the effective area. Each plot contains 21 curves
on the critical peak coefficients. Additional notable differences in- corresponding to all respective runs as defined in Table 1. The
clude the different flow conditions (open country versus suburban) expected monotonically decreasing relationship between the effec-
and the different roof types (gable versus flat) of the low-rise tive area and the magnitudes of the peak pressure coefficients is
building. clearly shown.

© ASCE 04014007-10 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


On the basis of the previous results regarding the effect of the Local Min Area-avg.
Global MinMin. Local Max Area-avg.
Global MaxMax.
-4.0
hc=h ratio on peak Cp;net , all curves shown in Fig. 22 were grouped
-3.0
into one of the following three categories:
• 0.9 ≤ hc=h ≤ 1 -2.0

• 0.5 ≤ hc=h < 0.9

Peak Cp, net


-1.0

• hc=h < 0.5 0.0


Consequently, minimum envelope lines were superimposed 1.0
onto each plot. Please note that the envelope with the highest 2.0
net uplifting peak is found for Case I (ed=lc ¼ 0, whereas the high-
lc/l = 1 lc/l = 0.5 lc/l = 0.5
ed/lc = 0 ed/lc = 0.5 ed/lc = 1
3.0
est net downward loading peak occurs for Case III (ed=lc ¼ 1).
4.0
The case of net downward loading shows that the experimental 0.0 0.2 0.4 0.6 0.8 1.0
curves for different hc=h ranges appear to be intertwined, which hc/h = 0.94 ; wc = 3.65 m ed/lc

Fig. 19. Effect of length and edge distances on local and area-averaged
Cp;net for hc=h ¼ 0.94 and wc ¼ 3.65 m
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Local Min Area-avg.


Global MinMin Local Max Global MaxMax
Area-avg.
-4.0
wc = 3.65m
-3.0

-2.0
Peak Cp, net

-1.0
-4.0
0.0
-3.0
1.0
-2.0
2.0 -1.0
3.0 0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0
4.0
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
hc/h
-4.0
Fig. 16. Effect of hc=h on local and area-averaged Cp;net for
Peak Cp, net

wc ¼ 3.65 m -3.0

-2.0

-1.0

Area-avg. Min. 0.0


Local Min MinGlobal Local Max Area-avg Max
MaxGlobal 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0
-4.0

-3.0

-2.0 -4.0
Peak Cp, net

-1.0 -3.0

0.0 -2.0

1.0 -1.0

2.0 0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
3.0 hc/wc
hc/h = 0.94 ; lc = 7.5m
4.0
2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 Fig. 20. Comparison of results of the present study and results pre-
hc/wc sented in Fig. 5 of Australian study (Jancauskas and Holmes 1985)
for velocity pressure on the basis of 1-h wind speed data
Fig. 17. Effect of hc=wc on local and area-averaged Cp;net for
hc=h ¼ 0.94

Hölscher et al. (2007) Present Study


-4.0
Net Pressure Coefficient Cp, net

Local Min MinGlobal


Area-avg. Min Local Max Area-avg. Max
-3.0
-4.0
-2.0
-3.0

-2.0 -1.0
Peak Cp, net

-1.0 0.0

0.0 1.0

1.0 2.0

2.0 3.0
3.0 4.0
hc/h = 0.34 ; lc = 7.5m 0.0 0.2 0.4 0.6 0.8 1.0
4.0
0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 hc/h
hc/wc

Fig. 18. Effect of hc=wc ratio on local and area-averaged Cp;net for Fig. 21. Comparison of results of the present study and those of
hc=h ¼ 0.34 Hölscher et al. (2007) for similar configurations

© ASCE 04014007-11 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


-5.0
-4.0
-3.0
Case I
-2.0
-1.0

Cp, net
0.0
1.0
2.0
3.0
4.0
-5.0
-4.0
-3.0
Case II -2.0
-1.0
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Cp, net
0.0
1.0
2.0
3.0
4.0
-5.0
-4.0

Case III -3.0


-2.0
Cp, net

-1.0
0.0
1.0
2.0
3.0
4.0
1 10 100 1000
(0.1) (0.9) (9.3) (92.9)
Effective Wind Area, ft ² (m²)

Fig. 22. Area averaging effect for all runs and envelopes for corresponding ranges

reaffirms the observation that net downward loads are less sensitive -2.0
to the hc=h ratio. For this reason, a single maximum envelope is
provided in Fig. 22 for all hc=h ratios. lc/l = 1
Furthermore, the envelopes for Case I (lc=l ¼ 1) are signifi- -1.0

cantly higher than those for Cases II and III (lc=l ¼ 0.5). Addition-
GCp, net

ally, the envelopes of Cases II and III present small differences with
0.0
one another. This result is consistent with those shown in Fig. 19, in
which lc=l has a significant effect on min local Cp;net in contrast
with (ed=lc), which has very little effect.
1.0

-2.0
Recommendations

Fig. 23 provides recommended provisions for the design of lc/l 0.5 -1.0

attached canopies. Note that all pressure coefficients presented


from this point forward were converted to conform to a 3-s
GCp, net

0.0
gust averaging period for codification purposes. The conver-
sions were approximated using the Durst curve for gust duration
(Durst 1960). 1.0
These proposed design provisions were generated from the 1 10 100 1000
envelopes of all experimental data obtained from this parametric (0.1) (0.9) (9.3) (92.9)

study (Fig. 22). The approach followed for the derivation of Effective Wind Area, ft² (m²)
codified provisions was similar to previous codification work
Fig. 23. Proposed guidelines for GCp;net on attached canopies
(Davenport et al. 1985).

© ASCE 04014007-12 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


-2.5
ratios, linear interpolation between the GCp;net values of the two
Net Pressure Coefficient, GCp -2.0 graphs may be used assuming that the lower diagram curves
-1.5
AS/NZS for hc/h = 1
Present:
correspond to lc=l ¼ 0.5.
AS/NZS for hc/h = 0.75
-1.0

-0.5
Comparisons with Current Code Provisions
AS/NZS for hc/h = 0.5

0.0 In this section, comparisons are made between the recommended


0.5
design provisions of the present study and those of the AS/NZS
and DIN.
1.0
1 10 100 1000
(0.1) (0.9) (9.3) (92.9)
Effective Wind Area, ft² (m²)
Australian/New Zealand 1170.2:2011
Fig. 24 presents comparisons between the values proposed by the
Fig. 24. Comparisons of recommended envelopes of the present study AS/NZS and the recommended envelopes of the present study for
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

and AS/NZS provisions the Cp;net as a function of the effective area. In general, good agree-
ment was observed. Note that for hc=h ¼ 1, the AS/NZS recom-
mends considerably higher values than those recommended in
-2.5
the present study for the range 0.9 ≤ hc=h ≤ 1, particularly for
local suctions. This marked increase in the AS/NZS provisions
-2 occurs when multiplying the recommended GCp;net by a local
Net Pressure Coefficient, GCp

Present min net pressure factor of 1.5 in accordance with Table D1 of the
-1.5 AS/NZS. For hc=h ¼ 0.75 in the AS/NZS, good agreement is seen
-1
with the range 0.5 < hc=h < 0.9 from the present recommenda-
DIN min tions. For hc=h ¼ 0.5, the min GCp;net recommended by the AS/
-0.5 NZS is considerably lower than the present recommendations. In
the case of max GCp;net , the three different values corresponding
0 to the AS/NZS recommendations for hc=h ratios of 1, 0.75, and
DIN max
0.5 respectively, were shown to be within the values of the present
0.5 Present max recommendations.
1
0.4 0.5 0.6 0.7 0.8 0.9 1 1.1
hc/h DIN EN 1991-1-4/NA:2010-12
Fig. 25. Comparison of recommended local net pressure coefficients of Fig. 25 presents the comparisons between the design provisions
the present study and DIN provisions for three different hc=h ratios on the DIN and those of the present study as a function of hc=h
for three configurations. The recommended values for local min
GCp;net are seen to generally display good agreement in which
Net pressure coefficient design values as a function of the hc=h the recommendations of the present study are slightly higher than
ratios and the effective area considered are presented in the two the recommendations of the DIN. For the local max GCp;net , the
graphs shown. The first graph corresponds to a canopy extending DIN and the present study recommend the same values regardless
over the entire length of the parent wall (lc=l ¼ 1), and the second of the hc=h ratio for this range. Also note that the recommendations
graph corresponds to a canopy extending over half the length of the of the present study for the highest downward loading are consid-
parent wall or less (lc=l ≤ 0.5). For the case of intermediate lc=l erably higher than that recommended by the DIN.

-3
ASCE 7-10
Present
Net Pressure Coefficient, GCp

-2.5

-2

-1.5
0.9 hc/h 1
-1
hc/h<0.9

-0.5 hc/h < 0.5

0
1 10 100 1000
(0.1) (0.9) (9.3) (92.9)

Effective Wind Area, ft² (m²)

Fig. 26. Comparisons of recommended design values for full-length canopy and ASCE 7-10 provisions for overhangs on gable roof of slope smaller
than 7° (Fig. 30.4-2A)

© ASCE 04014007-13 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007


Comparisons with ASCE 7-10 Provisions for Australia/New Zealand (AS/NZS). (2011). “Australian/New Zealand stan-
Overhangs dards, structural design actions—Part 2, wind actions.” AS/NZS
1170.2:2011, Sydney, Australia.
Given a lack of provisions in the major North American codes and Davenport, A. G., Stathopoulos, T., and Surry, D. (1985). “Reliability of
standards, practitioners often consider the attached canopies as wind loading specifications for low buildings.” Proc., Int. Conf. on
extensions of the roof overhangs and design them in accordance Structural Safety and Reliability, ICOSSAR 85, International Associa-
with the corresponding design guidelines. However, as expected, tion for Structural Safety and Reliability.
such designs may result in significant differences with the exper- Durst, C. S. (1960). “Wind speeds over short periods of time.” Meteorol.
imental values obtained for attached canopies. Fig. 26 presents Mag., 89, 181–186.
the comparisons between the min GCp;net values in the ASCE 7 European Committee for Standardization (CEN). (2004). “Eurocode 1:
provisions for roof overhangs and the present recommendations Actions on structures—General actions—Part 1.4: Wind actions.”
for a canopy extending over the entire length of the parent wall. Pr EN 1991-1-4, Brussels, Belgium.
German Institute for Standardization (DIN). (2010). “Einwirkungen auf
These marked differences denote the significant economic benefit
Tragwerke—teil 4: Windlasten.” DIN1055-4:2005-03, Berlin.
to the design of canopies if the recommended provisions are
Hölscher, N., Hubert, W., and Niemann, H. J. (2007). “Normgemäße
adopted. However, the effect of the roof slope was also seen to
Downloaded from ascelibrary.org by Tenira Forman on 06/12/17. Copyright ASCE. For personal use only; all rights reserved.

Erfassung der Windwirkungen an Vordächern Vordächern.” Fachteil


have an effect on the pressures around the heavily loaded corner WtG, Band 82, April, S.S2–S5.
regions. International Code Council. (2006a). International building code (IBC),
Falls Church, VA.
International Code Council. (2006b). International residential code (IRC),
Summary and Conclusions
Falls Church, VA.
Wind tunnel experimentation was performed on 63 different con- IS (Schweizerischer Ingenieur und Architektenverein). (1987). “Indian
figurations to serve as the basis for a parametric study of wind Standard code of practice for design loads-(Part 3-wind loads) for build-
ings and structures” IS: 875 (Part-3)-1987, New Delhi, India.
loading on attached canopies. The effect of wind direction on
Jancauskas, L., and Holmes, J. (1985). “Wind loads on attached canopies.”
net pressure coefficients was examined and showed that critical
Proc., National Conf. on Wind Engineering, Texas Tech Univ.,
peaks may occur for different wind directions depending on
Lubbock, TX.
the configuration. Local and area-averaged net pressure coeffi- National Research Council of Canada. (2010). National building code of
cients were analyzed as a function of four basic geometric ratios Canada (NBCC), Ottawa.
to examine the trends and relationships that they exhibit with one ÖNORM (Österreichisches Normungsinstitut ). (1993). “Belastungsannah-
another. Pressure and correlation coefficient contour plots were men im Bauwesen—Statische Windeinwirkungen.” ÖNORM B 4014-
also presented to provide a better understanding of the flow 1:1993 05 01, Vienna, Austria.
patterns occurring around the canopy. Among the four geometric Paluch, M. J., Loredo-Souza, A. M., and Blessmann, J. (2003). “Wind loads
ratios considered, hc=h was concluded to have the greatest influ- on attached canopies and their effects on the pressure distribution over arch
ence on the local uplift GCp;net . roof industrial building.” J. Wind. Eng. Ind. Aerodyn., 91(8), 975–994.
Recommended design provisions appropriate for ASCE 7 are SIA (Schweizerischer Ingenieur und Architektenverein). (2003). “Einwir-
presented from the results of the present study. kungen auf Tragwerke, Schweizerischer ingenieur- und Architektenver-
Comparisons were made between the recommended design ein.” SIA 261:2003, Zurich.
provisions and those established in the AZ/NZS 1170.2:2011 Tieleman, H. W., Hajj, M. R., and Reinhold, T. A. (1998). “Wind tunnel
[Australia/New Zealand (AS/NZS) 2011] and DIN EN 1991-1- simulation requirements to asses wind loads on low-rise buildings.”
4/NA:2010-12. These provisions were shown to be similar—higher J. Wind. Eng. Ind. Aerodyn., 74–76, 675–685.
in some cases and lower in others—when compared with the Uematsu, Y., Iizumi, E., and Stathopoulos, T. (2007). “Wind force coeffi-
present recommendations. cients for designing free-standing canopy roofs.” J. Wind. Eng. Ind.
Finally, a comparison between the recommended design provi- Aerodyn., 95(9–11), 1486–1510.
sions of the present study and those of the ASCE 7 for the design of Uematsu, Y., Stathopoulos, T., and Iizumi, E. (2008a). “Wind loads on free-
roof overhangs shows a significant economic benefit if the recom- standing canopy roofs: Part 1. Local wind pressures.” J. Wind. Eng. Ind.
mended provisions are adopted. Aerodyn., 96(6–7), 1015–1028.
Uematsu, Y., Stathopoulos, T., and Iizumi, E. (2008b). “Wind loads on free-
standing canopy roofs: Part 2. Overall wind forces.” J. Wind. Eng. Ind.
References Aerodyn., 96(6–7), 1029–1042.
Zisis, I., and Stathopoulos, T. (2010). “Wind-induced pressures on patio
ASCE. (2010). “Minimum design loads for building and other structures.” covers.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X.0000210,
ASCE/SEI 7-10, Reston, VA. 1172–1181.

© ASCE 04014007-14 J. Struct. Eng.

J. Struct. Eng., 2014, 140(5): 04014007

You might also like