You are on page 1of 14

ARTICLE IN PRESS

Journal of Wind Engineering


and Industrial Aerodynamics 96 (2008) 10291042
www.elsevier.com/locate/jweia

Wind loads on free-standing canopy roofs: Part 2


overall wind forces
Yasushi Uematsua,, Theodore Stathopoulosb, Eri Iizumic
a
New Industry Creation Hatchery Center, Tohoku University, Aoba-ku, Sendai 980-8579, Japan
Department of Building, Civil and Environmental Engineering, Concordia University, Montreal, Canada H3G 1M8
c
Department of Architecture and Building Science, Tohoku University, Aoba-ku, Sendai 980-8579, Japan

Available online 26 July 2007

Abstract
Wind loads on free-standing canopy roofs have been studied in a wind tunnel. Three types of roof
geometries, i.e. gable, troughed and mono-sloped roofs, with roof pitches ranging from 01 to 151, were
tested. Wind pressures were measured simultaneously at many points both on the top and bottom
surfaces of the roof model for various wind directions. The paper describes the characteristics of overall
wind forces and moments acting on the roof with special attention to load combination. Correlation
between wind force and moment coefcients is investigated and wind force coefcients for the design of
main wind force resisting systems are proposed. The roof is assumed rigid and simply supported by four
corner columns, whose axial forces appear as the most important load effect. Two loading patterns that
cause the maximum tension and compression of columns are considered. The proposed values are also
compared with the specications of the Australian/New Zealand Standard [Standards Australia, 2002.
Australian/New Zealand Standard, AS/NZS 1170.2].
r 2007 Elsevier Ltd. All rights reserved.
Keywords: Free-standing canopy roof; Wind tunnel experiment; Overall wind force; Main wind force resisting
system; Design wind load; Codication

1. Introduction
The characteristics of local pressures on free-standing canopy roofs are described in a
companion paper, Part 1 (Uematsu et al., 2007). This paper presents the experimental
results on overall wind forces and moments acting on these roofs.
Corresponding author. Tel./fax: +81 22 795 7875.

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


0167-6105/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jweia.2007.06.026

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1030

Gumley (1984), Letchford and Ginger (1992) and Ginger and Letchford (1994)
measured area-averaged pressures for various areas using a pneumatic averaging
technique. Furthermore, Ginger and Letchford (1994) calculated the overall forces using
the covariance integration technique. Holmes (2001) discusses the application of their
results to load estimation. Recently, Letchford et al. (2000) measured the mean wind forces
on solid and porous canopy roof models using a force balance technique. This is perhaps
the rst attempt to measure the overall forces on canopy roofs directly. More recently,
Altman (2001) has made extensive measurements of the overall forces and moments using
a high-frequency force balance developed at Clemson University, USA.
In the present study, the wind forces and moments are computed from the time history
of pressures measured simultaneously at many points both on the top and bottom surfaces.
Such simultaneous pressure measurements leading to a reliable denition of realistic windinduced forces on canopy roofs have not been made so far, to the authors best knowledge.
Correlation between the wind force and moment coefcients is also investigated. Design
wind force coefcients are based on suitable load combinations for development of the
unbalanced load distribution. The analysis assumes that the roof is rigid and supported by
the four corner columns, and the axial forces induced in the columns are taken as the most
important load effect for determining the wind force coefcients.
2. Denition of wind force and moment coefcients
The notation and sign of the wind forces and moments are schematically illustrated in
Fig. 1, together with the coordinate system. The wind force and moment coefcients are
dened as follows:
CNW

NW
,
qH b l=2

(1a)

CNL

NL
,
qH b l=2

(1b)

+ NW

+ NL
b
+L

+ My

= 0
(Duo-pitched roof)
+N

+L + My
= 0
(Mono-sloped roof)

+Mz

= 0

+ My

+Mx

= 90

Fig. 1. Denition of wind force and moment coefcients and coordinate system. (a) Cross section; (b) Plan view.

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1031

CL

L
,
qH bl

(1c)

CN

N
,
qH b l

(1d)

CMx

Mx
,
qH bl 2

(1e)

CMy

My
,
qH b2 l

(1f)

CMz

Mz
,
qH bl 2

(1g)

where Mx, My and Mz are moments about the x-, y- and z-axis, respectively; qH is the
velocity pressure at the mean roof height H; and b* the actual length of the roof ( b/
cos b, with b being the roof pitch).

3. Wind tunnel experiments


The experimental arrangement and procedures are described in Part 1. In the
measurements, models with the regular tap arrangement (tap arrangement 1) are used.
Using the pressure difference coefcients at 48 points on the roof model, the wind force
and moment coefcients are computed. The statistics (mean, standard deviation,
maximum and minimum) of the wind force and moment coefcients are computed by
applying ensemble average to the nine consecutive runs.

4. Results and discussion


4.1. Mean wind force coefficients
Fig. 2 shows the variation of the mean values of C N W and C N L (C N W and C N L ) with roof
pitch b for gable roofs when y 01, in which the results of previous studies are also plotted
for comparative purposes. The present results generally agree with those of Altman (2001).
Gumleys results (1984) for b 01 show a normal force coefcient approximately equal to
0.2. This is probably due to the effect of blockage induced by the windward four columns
at eaves edge reducing the roof uplift force. It should be noted that the ratio of the column
width to the distance between adjacent columns was as large as 0.18 in the model used in
Gumley (1984).
A similar comparison for the mean value of CN (C N ) on mono-sloped roofs is shown in
Fig. 3. Again, the present results agree well with those of Altman (2001). In fact, the results
of most previous experiments agree well when y 1801, for which the effect of the roof
supporting system on the wind forces seems to be less signicant.

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1032

1.0

0.5
Present
Gumley (1984)
Letchford et al.(1992)
Letchford et al.(2000)
Altman (2001)

0.0

CNL (mean)

CNW (mean)

1.0

-0.5

Present
Gumley (1984,)
Letchford et al. (1992)
Letchford et al. (2000)
Altman (2001)

0.5
0.0
-0.5
-1.0

10

20
30
Roof pitch (deg)

40

10

20
30
Roof pitch (deg)

40

Fig. 2. Mean normal force coefcients on the windward and leeward halves of gable roofs (y 01). (a) Windward
half; (b) Leeward half.

0.0

2.0

CN (mean)

CN (mean)

1.5
-0.5

-1.0

1.0
0.5

Present

0.0

Gumley (1984)
Letchford et al. (2000)
Altman (2001)

-1.5

-0.5
0

10

20
30
Roof pitch (deg)

40

10

20
30
Roof pitch (deg)

40

Fig. 3. Mean normal force coefcients on mono-sloped roofs for y 01 and 1801. (a) Wind direction: 01;
(b) Wind direction: 1801.

4.2. Peak wind force coefficients


Fig. 4 depicts the maximum and minimum peak values of C N W and C N L on gable
roofs for a wind direction range from y 01 to 451. The estimated values from the results
of Ginger and Letchford (1994), who measured maximum and minimum peak values of
the area-averaged pressure over six panel areas, are also plotted. Using these results, the
present study estimated the maximum and minimum peak normal force coefcients on the
windward and leeward halves of the roof by calculating the ensemble average of the peak
pressure coefcients on three panel areas. The estimation is based on the assumption that
the peak area-averaged pressures on these three panel areas occur simultaneously; this of
course may result in an overestimate of the actual peak values. Indeed, the estimated
normal force coefcients are generally larger in magnitude than the present results. This is
also due to differences in ow conditions (open country exposure in the present study and
suburban exposure in the study of Ginger and Letchford, 1994).
4.3. Correlation between wind force and moment coefficients
Fig. 5 shows a phase-plane representation of the C N W  C N L relation for a gable roof.
The dashed lines represent the mean values and the circle represents the condition where

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1
CNL (max & min)

3
CNW (max & min)

1033

2
1
0

0
-1
Present (max)
Present (min)

-2

Ginger & Letchford (1994) (max)


Ginger & Letchford (1994) (min)

-3

-1
0

10

20
30
Roof pitch (deg)

40

10

20
30
Roof pitch (deg)

40

Fig. 4. Maximum and minimum peak wind force coefcients on the windward and leeward halves of gable roof
(y 0451). (a) Windward half; (b) Leeward half.

0.1

CNL

0.0
-0.1
-0.2
-0.3
-0.4
-0.8

-0.4

0.0
CNW

0.4

0.8

Fig. 5. Phase-plane representation of the C N W  C N L relation: gable roof (b 51, y 01).

the C N W value becomes an extreme in each run. The result indicates a poor correlation
between C N W and C N L . Similar features were observed for all cases tested. Therefore, a
combination of the peak values of C N W and C N L , which is often used in code provisions,
may overestimate the design wind loads, because they are not induced simultaneously.
The variation of CL, C M x and C M y (maximum, mean and minimum values) with wind
direction y is plotted in Fig. 6 for a gable roof and in Fig. 7 for a mono-sloped roof, both
with b 101. When yE01, the peak values of CL and C M y are generally large in
magnitude, while the value of jC M x j is relatively small. With an increase in y, the
magnitude of the peak values of CL and C M y decreases, while that of jC M x j increases. In
the mono-sloped roof case, the value of |CL(min)| and C M y max are large when yE1801.
Fig. 8 shows the coefcients of correlation (R) between the wind force and moment
coefcients, plotted as a function of y. When the wind direction is nearly normal to the
eaves (yE01 or 1801), the correlation between CL and C M y is relatively high and that
between CL and C M x is rather low. The coefcient of correlation between CL and C M y is
low for oblique winds.
Fig. 9 shows the variation of the maximum and minimum axial forces induced in the
columns with wind direction y for three roof geometries, where the axial force is reduced

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1034

0.05

1.0

0.00

mean

max

0.0

mean

CMy

CL

0.5

CMx

max

0.1

max

-0.05
min

0.0

mean

-0.1

-0.10

min

min

-0.5

-0.2

-0.15
0

30
60
90
Wind direction (deg)

30
60
90
Wind direction (deg)

30
60
90
Wind direction (deg)

1.5
1.0
0.5
0.0
-0.5
-1.0
-1.5

0.2

Maximum
Mean
Minimum

0.4

Maximum
Mean
Minimum

0.0
-0.1

60
120
180
Wind direction (deg)

0.2
0.1
0.0

-0.2
0

Maximum
Mean
Minimum

0.3
CMy

0.1
CMx

CL

Fig. 6. Variation of the statistics of CL, C M x and C M y with wind direction y: gable roof (b 101). (a) CL;
(b) CMx; (c) CMy.

-0.1
0

60
120
180
Wind direction (deg)

60
120
180
Wind direction (deg)

Fig. 7. Variation of the statistics of CL, C M x and C M y with wind direction y: mono-sloped roof (b 101). (a) CL;
(b) CMx; (c) CMy.

1.0

CL-CMy

0.5
R

0.5
R

1.0
CL-CMy

0.0

0.0

CL-CMx

CL-CMx

30
60
90
Wind direction (deg)

CL-CMx

-1.0

-1.0

-1.0

0.0
-0.5

-0.5

-0.5

CL-CMy

0.5
R

1.0

90
30
60
Wind direction (deg)

60
120 180
Wind direction (deg)

Fig. 8. Coefcient of correlation between the wind force and moment coefcients. (a) Gable roof (b 51);
(b) Gable roof (b 101); (c) Mono-sloped roof (b 51).

by qH(bl/4). Note that the axial forces induced in the columns are easily computed from the
time history of pressures and their inuence coefcients for the axial forces, because the
roof is assumed rigid and supported by four corner columns, as already mentioned. In the
case of gable and troughed roofs, the maximum axial force (maximum tension) occurs at a
wind direction ranging from 01 to approximately 451 and the minimum axial force
(maximum compression) occurs at yE01. In the mono-sloped roof case, the behavior in the
wind direction range from 01 to 901 is similar to that for the gable and troughed roofs. The

ARTICLE IN PRESS

max

0.5
0
min

-0.5
-1
0

Reduced axial force

Reduced axial force

Reduced axial force

Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

max

0.5
0
min

-0.5

30
60
90
Wind direction (deg)

-1
0

2.0

1035

max

1.0
min

0.0
-1.0
-2.0

30
60
90
Wind direction (deg)

60
120
180
Wind direction (deg)

Fig. 9. Variation of the maximum and minimum axial forces with wind direction. (a) Gable roof (b 101);
(b) Troughed roof (b 101); (c) Mono-sloped roof (b 101).

WIND

WIND

WIND

Wind force distribution


due to lift L

Wind force distribution


due to moment My

Wind force distribution


due to L and My

Fig. 10. Equivalent pressure distribution on the windward and leeward halves caused by CL and C M y .

maximum compression is induced at yE1801. It is interesting to note that the minimum


axial force at y 01 and the maximum axial force at y 1801 are both small in magnitude,
nearly equal to zero.
4.4. Design wind force coefficients
Based on the above ndings, the design wind force coefcients can be evaluated
according to the following procedure:
Step 1: The basic values of the wind force coefcients C N W and C N L , denoted as C N W 0
and C N L 0 , are determined from a combination of CL and C M y that produces the maximum
load effect for winds normal to the eaves (y 01 or 1801). The values of C N W 0 and C N L 0
are calculated as follows (see Fig. 10):
C N W 0 C L  4C M y ,

(2a)

C N L 0 C L 4C M y .

(2b)

Step 2: In order to consider the effects of C M x and wind direction on the axial forces
induced in the columns, a correction factor g, which is dened as the ratio of the actual
peak force for y 01451 to that computed from C N W 0 and C N L 0 , is introduced.

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1036

Step 3: The design wind force coefcients C N W and C N L , which give equivalent static
wind loads, are provided as follows:
C N W

gC N W 0
,
Gf

(3a)

C N L

gC N L 0
,
Gf

(3b)

CMy
0.0

0.0

-0.1

-0.1

0.25

CMy

0.1

0.1

0.20

CMy

where Gf represents a gust effect factor, which should be determined based on the load
effect.
Fig. 11 shows a phase-plane representation of the C L  C M y relation for low roof
pitches and y 01. The circles represent the maximum and minimum peak values of CL
during each of the nine runs. The correlation between CL and C M y is generally high for
mono-sloped roofs. The value of C M y at the instant when the maximum CL (CL max) occurs
is nearly equal to the maximum C M y (C M y max ). This feature implies that the maximum
load effect is given by a combination of these peak values, i.e. the maximum tension is
given by C L max C M y max . Similarly, the maximum compression is given by
C L min C M y min ; the sufx min represents the minimum value of the coefcient under
consideration. When the roof pitch is relatively high, such as b 151, for example,
compression is no longer induced in any column. In the case of y 1801, the value of C M y
at the instant when CL min occurs is nearly equal to C M y max . In the gable and troughed roof
cases, on the other hand, the C L  C M y correlation is relatively low and becomes lower as
the roof pitch increases. The peak+peak combination does not always give the maximum
load effect. The envelope of the C L  C M y trajectory is approximated by a hexagon shown
in Fig. 12. The critical condition producing the maximum load effect may be given by one
of the apexes of the hexagon. In order to investigate the load combination effects, the axial
forces induced in the columns are computed for the six combinations of CL and C M y
(Points 16) in Fig. 12. Table 1 summarizes the conditions that give the maximum load
effect for each case. Load cases A and B represent the conditions producing the
maximum tension and compression in the columns, respectively. Note that the column
subjected to the maximum tension or the maximum compression depends on the roof
geometry. Substituting the CL and C M y values (for the cases provided in Table 1) into
Eq. (2), the values of C N W 0 and C N L 0 for each case are obtained.

0.15
0.10

CL
-0.2
-0.2
-0.4 -0.2 0.0 0.2 0.4 0.6
-0.2

0.05
CL
0.0

0.2

0.4

0.00

-0.05
-0.5
0.6

CL
0.0

0.5

1.0

Fig. 11. Phase-plane representation of the CLC M y relation (y 01). (a) Gable roof (b 51); (b) Gable roof
(b 101); (c) Mono-sloped roof (b 51).

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1037

Fig. 12. Model of the envelope of the trajectory.

Table 1
Combination of CL and C M y values producing the maximum load effect for gable and troughed roofs
Roof

Pitch (deg)

Load case
A

C L max C M y max

C L min C M y min

Gable

5
10
15

C L max C M y max
C L max C M y mean
C L max C M y mean

C L min C M y min
C L min C M y min
C L min C M y min

Troughed

5
10
15

C L max C M y max
C L max C M y max
C L max C M y max

C L min C M y mean
C L min C M y mean
C L min C M y mean

Flat

The correction factor g in Eq. (3) is obtained by calculating the ratio of the actual
maximum or the minimum axial force to the predicted value from C N W 0 and C N L 0 .
Fig. 13 shows the gust effect factor Gf, dened as the ratio of the maximum or the
minimum axial force to the mean value induced in the columns, plotted against the
mean reduced axial force N mean for y 01451 (W.D.1) and y 13511801 (W.D.2).
When the value of jN mean j is small, Gf exhibits a large value. However, as jN mean j increases,
the values of Gf collapse into a narrow range around Gf 2.0 (dashed line), which
corresponds to a peak factor of gv 3.0, based on the quasi-steady assumption, i.e.
GfE(1+gvIuH)2 (1+3.0  0.14)2 2.0. Therefore, Gf 2.0 is used for evaluating the
wind force coefcients hereafter.
Plotted on Figs. 1416 are the estimated values of C N W and C N L for gable and monosloped roofs (triangles). For comparative purposes, the maximum and minimum peak
values of C N W and C N L are plotted by circles; these values are divided by Gf to make a
direct comparison with C N W and C N L . Note that they are not induced simultaneously.
Furthermore, the specication of the Australian/New Zealand (AS/NZ) Standard
(Standards Australia, 2002) is also shown by the dashed lines. The wind force coefcients
on the windward half for the two load cases are consistent with the maximum and
minimum peak values. When yE01 (W.D.1), the values of C N W for load cases A and B are
nearly equal to those of C N W min and C N W max and close to the AS/NZ specication.

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1038

10
Flat (W.D. 1)

Gable (W.D. 1)
Troughed (W.D. 1)

6
Gf

Mono-sloped (W.D. 1)
Mono-sloped (W.D. 2)

Gf = 2.0

2
0
0

0.1

0.2

0.3

|N*mean|
Fig. 13. Gust effect factor based on the load effect.

0.5

1.0
AS/NZ (positive)

0.0

0.0
AS/NZ (positive)

CNL*

CNW*

0.5

AS/NZ (negative)

AS/NZ (negative)

-0.5
-0.5

Load case A

Load case B

Load case A

Load case B

Maximum

Minimum

Maximum

Minimum

-1.0

-1.0
0

10
15
Roof pitch (deg)

20

10
15
Roof pitch (deg)

20

Fig. 14. Equivalent static wind force coefcients C N W and C N L (gable roof, W.D.1). (a) Windward half; (b)
Leeward half.

0.5

0.5
AS/NZ (positive)

AS/NZ (positive)

0.0

-0.5

CNL*

CNW*

0.0
AS/NZ (negative)

-1.0

AS/NZ(negative)

-0.5

-1.5

Load case A

Load case B

Load case A

Load case B

Maximum

Minimum

Maximum

Minimum

-1.0

-2.0
0

10
15
Roof pitch (deg)

20

10
15
Roof pitch (deg)

20

Fig. 15. Equivalent static wind force coefcients C N W and C N L (mono-sloped roof, W.D.1). (a) Windward half;
(b) Leeward half.

Similarly, when yE1801 (W.D.2), the values of C N L of mono-sloped roofs for load cases A
and B are nearly equal to those of C N L min and C N L max . Note that C N L , not C N W ,
represents the wind force coefcient on the windward half when yE1801. Regarding the

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

0.5

1.5

1039

AS/NZ (positive)

0.0

0.5

CNW*

CNL*

1.0
AS/NZ (positive)

0.0

-0.5

AS/NZ (negative)

-0.5

AS/NZ (negative)

Load case A

Load case B

Load case A

Load case B

Maximum

Minimum

Maximum

Minimum

-1.0

-1.0
0

10
15
Roof pitch (deg)

20

10
15
Roof pitch (deg)

20

Fig. 16. Equivalent static wind force coefcients C N W and C N L (mono-sloped roof, W.D.2). (a) Windward half;
(b) Leeward half.

0.5

CNW*(Load case A)
CNW*(Load case B)
CNW_Nmax/Gf
CNW_Nmin/Gf

0.5

CNL*, CNL_Npeak/Gf

CNW*, CNW_Npeak/Gf

1.0

0.0

-0.5

CNL*(Load case A)
CNL*(Load case B)
CNL_Nmax/Gf
CNL_Nmin/Gf

0.0

-0.5
0

10
15
Roof pitch (deg)

20

10
15
Roof pitch (deg)

20

Fig. 17. Comparison of C N W and C N L with the actual wind force coefcients producing the maximum and
minimum axial forces (gable roof, W.D.1). (a) Windward half; (b) Leeward half.

leeward half, on the other hand, the wind force coefcients for the two load cases are
similar to each other and close to either of the maximum or the minimum peak value. The
values are signicantly different from the AS/NZ specication. The results for troughed
roofs were found quite similar to those of the gable roofs.
eN and C
eN producing the maximum tension and
The actual wind force coefcients C
W
L
compression in the columns are obtained from the time history analysis. The results are
compared with the C N W and C N L values for the two load cases in Fig. 17; the values of
eN and C
eN are obtained by applying ensemble average to the results of nine runs and
C
W
L
then dividing by Gf. Results are consistent with each other. Therefore, the proposed values
of C N W and C N L correspond approximately to the pressure distribution that causes the
maximum load effect.
The maximum shear force induced in the columns is obtained from the time history
analysis and compared with the predicted values from C N W and C N L together with
Gf 2.0. The calculation of the shear force considers the torsional moment Mz as well as
the drag. The results for gable and troughed roofs are shown in Fig. 18, which indicates
that the predicted values from C N W and C N L capture the actual maximum shear force

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1040

0.4
Reduced shear force

Reduced shear force

0.4
Load case A

0.3

Load case B

0.2

Time history

0.1
0
-0.1

Load case A

0.3

Load case B

0.2

Time history

0.1
0
-0.1

10
15
Roof pitch (deg)

20

10
15
Roof pitch (deg)

20

Fig. 18. Maximum reduced shear forces induced in the columns (W.D.1). (a) Gable roof; (b) Troughed roof.

1.5

3.0
Present (Max)

Present (Max)

1.0

2.0

0.5

N*/Gf

N*/Gf

Present (Min)

AS/NZ (max)

0.0
AS/NZ (min)

AS/NZ (max)

Present (Min)

1.0
0.0

-0.5

AS/NZ (min)

-1.0

-1.0
0

10
15
20
Roof pitch (deg)

25

30

10
15
20
Roof pitch (deg)

25

30

Fig. 19. Comparison for axial forces between the predicted values from C N W and C N L and from the AS/NZ
specications (W.D.1). (a) Troughed roof; (b) Mono-sloped roof.

reasonably well. In fact, the maximum shear force obtained from the time history analysis
lies between the two predicted values corresponding to the load cases A and B.
Finally, the axial forces induced in the columns are computed by using C N W and C N L for
the two load cases and compared with those predicted from the AS/NZ Standard. Sample
results are shown in Fig. 19. The AS/NZ Standard generally provides two values of the
wind force coefcients for each of the windward and leeward halves and this results in four
combinations of C N W and C N L . The maximum and minimum axial forces among the four
values from the AS/NZ Standard are generally consistent with the results for the load cases
A and B, respectively, in spite of the difference in the wind force coefcients (see Figs.
1416). Consequently, the proposed values of C N W and C N L can be used for design
purposes. Table 2 summarizes the proposed wind force coefcients (tentative), which are
obtained from a simplied model of the variation of the wind force coefcients with roof
pitch b.
In the present study, it is assumed that the roof is supported by four corner columns, and
the axial forces induced in the columns are taken as the load effect for discussing the wind
force coefcients. These coefcients can also be used for the cases where the roof is
supported by more columns, say six, because the critical condition may be given when

ARTICLE IN PRESS
Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

1041

Table 2
Wind force coefcients C N W and C N L
Roof

Pitch (deg)

Wind direction

Load case A
C N W

Flat

Load case B
C N L

C N W

C N L

W.D.1

0.40

0.10

0.20

0.10

Gable

5
10
15

W.D.1
W.D.1
W.D.1

0.40
0.20
0

0.12
0.33
0.55

0.35
0.50
0.65

0.08
0.27
0.45

Troughed

5
10
15

W.D.1
W.D.1
W.D.1

0.49
0.58
0.68

0.21
0.33
0.44

0.13
0.07
0

0.27
0.43
0.60

Mono-sloped

5
10
15

W.D.1
W.D.1
W.D.1

0.65
0.90
1.15

0.07
0.03
0

0.13
0.07
0

0.07
0.03
0

Mono-sloped

5
10
15

W.D.2
W.D.2
W.D.2

0.07
0.03
0

0.27
0.13
0

0.13
0.17
0.20

0.47
0.73
1.00

yE01 or 1801 irrespective of the number of columns, and the values of C N W 0 and C N L 0 are
determined based on a combination of CL and C M y when y 01 or 1801. Furthermore, the
effect of C M x on the axial forces, which is considered by the correction factor g, becomes
largest when the roof is supported by four columns. Therefore, the wind force coefcients
proposed in the present study will be conservative in so far as they may somewhat
overestimate the wind-induced loads in the columns when the number of columns is more
than four.
5. Concluding remarks
The overall wind forces and moments acting on free-standing canopy roofs have been
investigated based on a series of wind tunnel experiments. The study provides basic data
for the establishment of design wind loads for this type of roofs. Two load cases yielding
maximum tension and compression in the columns are considered. Based on the
combination of CL and C M y , design wind force coefcients C N W and C N L , which give
equivalent static wind loads, are proposed as a function of roof pitch from 01 to 151. Axial
forces produced by using these coefcients are consistent with those obtained from the
AS/NZ Standard.
Further research is necessary to establish design wind force coefcients for higher roof
pitches.
Acknowledgments
This study was made during the rst authors appointment at the Department of
Building, Civil and Environmental Engineering, Concordia University, Montreal, Quebec,
Canada, as a visiting professor, supported by the Ministry of Education, Culture, Sports,

ARTICLE IN PRESS
1042

Y. Uematsu et al. / J. Wind Eng. Ind. Aerodyn. 96 (2008) 10291042

Science and Technology, Japan, for the period May 2002February 2003. The authors are
much indebted to Mr. Kai Wang, graduate student of Concordia University, for his
assistance with the experiments.
References
Altman, D.R., 2001. Wind uplift forces on roof canopies. M.Sc. Thesis, Clemson University, NC, USA.
Ginger, J.D., Letchford, C.W., 1994. Wind loads on planar canopy roofs, Part 2: uctuating pressure distributions
and correlations. J. Wind Eng. Ind. Aerodyn. 51, 353370.
Gumley, S.J., 1984. A parametric study of extreme pressures for the static design of canopy structures. J. Wind
Eng. Ind. Aerodyn. 16, 4356.
Holmes, J.D., 2001. Wind Loading of Structures. Spon Press.
Letchford, C.W., Ginger, J.D., 1992. Wind loads on planar canopy roofs, Part 1: mean pressure distributions.
J. Wind Eng. Ind. Aerodyn. 45, 2545.
Letchford, C.W., Row, A., Vitale, A., Wolbers, J., 2000. Mean wind loads on porous canopy roofs. J. Wind Eng.
Ind. Aerodyn. 84, 197213.
Standards Australia, 2002. Australian/New Zealand Standard, AS/NZS 1170.2.
Uematsu, Y., Stathopoulos, T., Iizumi, E., 2007. Wind loads on free-standing canopy roofs, Part 1: local
pressures, J. Wind Eng. Ind. Aerodyn., to appear.

You might also like