You are on page 1of 14

Journal of Wind and Engineering, Vol. 7, No. 1, January 2010, pp.

16-29

16

WIND PRESSURES ON GABLE ROOF BUILDING WITH OVERHANGS


K. Narayan1 , A. Gairola2
1 2

Assistant Professor, Deptt of Civil Engg. I.E.T. Lucknow, India Associate Professor, Deptt. of Civil Engg. I.I.T. Roorkee, India

ABSTRACT
This paper presents the study on a single story gable (pitched) roof building with overhangs on either sides and without any openings. Roof slopes selected for this study are 10, 20 and 25. The natural wind was simulated in the wind tunnel on the basis of full/model scale comparison. The prototype is considered to be situated in Terrain Category 2 as per IS: 875 (Part-3) 1987. Wind pressure coefficients for different zones are presented in this paper. Peak pressure coefficients for overhang zones are significantly higher for roof slope 10, whereas the corner zones have significantly lower values compared with the Code values for roof slopes 20 and 25. Other zones have significantly higher values compared with the Code values. The results of this study indicate that there is need for standardization in the experimental as well as the codification process for assigning the design pressure coefficients.

INTRODUCTION
In the assessment of wind loads on roofs, the effect of eaves (otherwise called canopies or overhangs) is sometimes overlooked. Windward eaves, however, may be loaded severely due to wind, since the deflected flow from the windward wall gives rise to a pressure on the lower eave surface, which reinforces the high suction on the upper eave surface immediately after separation. Unfortunately, only a limited amount of information is available on the subject. This is probably the reason why most Standards and Codes of Practice have poor and often conflicting documentation of provisions related to wind loads on eaves. Leutheusser (1965) has published a comprehensive study on mean wind loads on eaves using models of various geometries, but he performed all his tests under uniform wind conditions (no velocity variation with height). Thus, his test conditions did not simulate natural wind effect, and therefore his results may not be representative. Dreher and Cermak (1973) have examined local pressures on upper and lower eave surfaces in a boundary layer wind tunnel simulating natural wind conditions. Their measurements have been limited to one geometrical configuration only, but they have reported that pressure underneath the overhangs can account for 50% of the total wind acting on the overhangs of a house. Best and Holmes(1978) have examined local pressures on eaves of a low-rise building model and have recognized that combining the negative peak values on the top with the positive peak values under the eave is somewhat conservative, since peak values do not occur together. The lack of pertinent information regarding wind loads on eave sections in combination with reports of severe wind damages on these areas, have led Codes and Standard Committees to significantly increase the wind design specifications for eaves. Stathopoulos (1981) carried out experiments in the Boundary Layer Wind Tunnel Laboratory of the University of Western Ontario. The basic model used for measurement on eaves was of 1:250 scale with roof slope of 1:12 (5).

Journal of Wind and Engineering, Vol. 7, No. 1, January 2010, pp. 16-29

17

Savory et al (1991) have examined the effect of eave geometry- sharp versus curved edge - on the wind pressure distribution. Stathopoulos and Luchian (1994) carried out experimental study on the effects of wind on eaves (perimetric extension of the roof surface) of gable-roof buildings with roof slopes 4:12 and 12:12. The study has found that the 4:12 sloped overhang is the most severely loaded, particularly where suction on the upper surface near the gable is concerned. Positive pressures on the upper surface are higher for the 12:12 slope. However, lower eave surfaces are subjected to higher pressures and suctions for the 4:12 slope. Authors compared the experimental values with the current wind Standards (NBCC-90 & ASCE 788) and found that values are significantly overestimated. National Building Code of Canada in 1995, revised the values based on the above investigations. Tore Wiik and Ernst W.M. Hansen (1997) carried out numerical simulations and compared the results with experimental mean values of the characteristic pressure (Cp) values on the gable wall and on the roof overhang of a low-rise building. Wind pressures, both mean and fluctuating components, on low-rise buildings are not only strongly influenced by building geometry, roof shape and angle of wind incidence but also depend on the surroundings and the wind flow characteristics. Wind tunnel tests on rigid models of low-rise buildings in a simulated flow condition provide detailed information on the effect of various parameters that influence wind pressures on these buildings. Wind pressures on the building roofs are highly fluctuating and random in nature. Pressure coefficients for the mean and fluctuating components can be derived from the pressure history records. Design pressure coefficients can then be deduced from the mean and fluctuating pressure coefficients. The experimental results of wind tunnel testing of various building models in isolated (stand-alone) condition are presented in this paper.

TEST PROGRAMME The building selected for the study is a single story gable (pitched) roof building with overhangs on either side with the openings considered to be closed at the time of a windstorm. The prototype is considered to be situated in a sub-urban terrain with well scattered objects having height between 1.5m and 10m, defined as Terrain Category 2 in IS: 875 (Part-3) 1987. For this terrain type, the variation of hourly mean wind speed with height is assumed to follow a power law with coefficient a = 0.176 (Fig.1). The turbulence intensity at the eave height is 20%, which is close to the mean full-scale value of TTU Mode M04 data. Integral length scale of turbulence L , is ux obtained as 0.45m. Small scale turbulence parameter,S, is necessary to match the peaks not the integral length scale. The work reported by Tieleman et al (1999), gave the best duplication of mean and peak pressure coefficients comparing model/full-scale results, establishing adequacy of the Small scale value achieved. S is defined below: S = (n Su (n)/ Su2 ) (Su / U)2 x106 where, n Su U L Su (n) Frequency of velocity fluctuations Standard deviation of the stream-wise velocity fluctuations Mean velocity evaluated at building eave height, H Characteristic model dimension, which is eaves height in the present study Spectral density function at frequency n

18

Wind Pressure on Cable Roof Building with Overhangs

S is evaluated at frequency n (=10U/L) at model eave height, and its average value has been obtained as 101.This is considered and found to be appropriate for the model study. Wind tunnel studies so far have been made with smaller values of S excepting the one reported by Tieleman. The flow characteristics obtained with the present set of roughness devices closely agreed with the simulation # 8 of Tieleman et al (1999), mentioned above.

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.0 0.0 0.2 0.4 0.6 0.8 1.0

0.0 0.0 5.0 10.0 15.0 20.0 25.0

Fig 1. Mean Velocity and Turbulence Intensity Profiles for the Simulated Flow in Wind Tunnel

A 1:25 scale Perspex model of the TTU Building and three models of a gable roof building with overhangs and having 10, 20 and 25 roof pitches were fabricated. Overhangs have been made as extension of the roof surface. Steel tubes of 1.0 mm inner diameter were inserted into the 2.0 mm gap between two Perspex sheets making up upper and lower surfaces of the overhangs - a technology which was developed for this study. In all 66 pressure points were created on the upper surface of overhangs and 40 pressure points were created on the lower surface of the overhangs. In addition, 88 pressure points were created on the remaining surface of the roof. Pressure taps have been placed at a distance of 0.6 mm from the edge of the roof projection (eave) and 0.4 mm from the ridge as well as the walls (Fig. 2). Building models have been tested in the wind tunnel under the simulated flow conditions for angle of wind incidence of 0 to 90 degrees, with increment of 15 degrees. Pressure measurements are carried out on the roof surface and upper and lower surface of eaves/overhangs. Pressure time histories for all pressure taps have been recorded and mean, minimum, maximum and rms pressure coefficients are deduced form the time history records. Area averaged pressure coefficients for all zones of roof and overhang portion of the building are computed using corresponding pressure taps of the concerned zone. Different roughness devices have been used to meet the wind tunnel simulation requirements and for the development of highly turbulent flow for generating the Atmospheric Surface Layer in a 2.1m x 2.0m Boundary Layer Wind Tunnel (BLWT). For flow measurements with the hot-wire system, a sampling frequency beyond 1 kHz and a 4second length of record was adopted as minimum requirement (Gupta1996). In the present study, instantaneous velocity fluctuations have been recorded using hot-wire probe at a sampling frequency of 4 kHz for a duration of approximately 4 seconds, viz, a total of 16384 samples are recorded at each point for flow characteristic measurement. The mean velocity and longitudinal turbulence intensity variations obtained in the wind tunnel are presented.

Journal of Wind and Engineering, Vol. 7, No. 1, January 2010, pp. 16-29

19

B1 1 B2

B1

Location of Different Zones on Building Roof

Overhang

Main Roof

Main Roof

Overhang

Pressure taps on upper roof surface Pressure taps on lower roof surface

Fig 2. Layout of Locations of Pressure Tapping on Roof of the Building Model

20

Wind Pressure on Cable Roof Building with Overhangs

Modelling of Tubing system


Holmes and Lewis (1986, 1987, and 1989), and Holmes (1995) performed extensive experimental work on the fluctuating pressure measurements using small diameter connecting tubes to transmit the pressure from the tap to the pressure transducer. Their work on tubing systems provided guidelines to develop a range of near optimum systems for the measurement of fluctuating pressures on models of buildings in a wind tunnel. In the present study, surface pressures on the roof of the building models have been measured by connecting steel taps of 1.0 mm internal diameter, flushed to the model surface, to vinyl tubing of 1.2 mm internal diameter. A three stage tubing system is used for the pressure measurements. Total length of tubing system used was 250 mm. A restrictor of 60 mm in length and 0.4 mm in diameter was connected at 150 mm from the pressure tap. Restrictor was connected to a Scanivalve Pressure Scanner with 40 mm long 1.2 mm internal diameter vinyl tube. Amplitude frequency response of the tubing system was approximately flat and phase characteristic was linear up to frequency 200 Hz (Fig. 3).

3.00

Amplitude response function

2.00

1.00

0.00

-1.00 0.00 0.00 100.00 200.00 300.00 400.00 500.00

Frequency Hz
-100.00

Phase Lag (deg)

-200.00

-300.00

-400.00

Fig 3. Amplitude and Phase Response of Three-Stage Tubing System

Journal of Wind and Engineering, Vol. 7, No. 1, January 2010, pp. 16-29

21

Surface Pressure Measurements A 32 channel ZOC12 pressure scanner from Scanivalve Corporation Ltd has been used to measure surface pressures. The output signal in the form of voltage from the pressure scanner has been recorded using PCL206 ADC Card. A computer programme is developed to acquire the voltage signal from Scanivalve through the PCL206 ADC Card. Data has been recorded at a sampling frequency of 400 samples/sec/channel. 12000 samples of pressure data from each channel have been recorded, thus giving a record of approximately 30 seconds. Wind pressures measured on the roof of building models are expressed in the form of a non-dimensional pressure coefficient. RESULTS AND DISCUSSION Variation of Pressure Coefficients On Roof Surface
Spatial distribution of Cpmean, Cpmin and Cprms over the entire roof surface and lower eave surface has been studied for wind directions 0 and 90. Variation of the pressure coefficients is presented in the form of contours in Figs 4.1 to 4.14.

Cpmean, Lower eave surface

Cpmax, Lower eave surface

Cpmean, Upper roof surface

Cpmin, Upper roof surface

Cpmean, Lower eave surface


0

Cpmax, Lower eave surface

Fig. 4.1 Variation of Cpmean for 10 Roof Slope at 0 Angle of Wind Attack

Fig. 4.2 Variation of Cpmin for 10 Roof Slope at 0 Angle of Wind Attack

22
Cprms Lower eave surface

Wind Pressure on Cable Roof Building with Overhangs


Cpmean Lower eave surface

90

Cprms, Upper roof surface

Cpmean, Upper roof surface

Cprms Lower eave surface

Cpmean Lower eave surface

Fig. 4.3 Variation of Cprms for 10 Roof Slope at 0 Angle of Wind Attack
Cpmax lower eave surface

Fig. 4.4 Variation of Cpmean for 10 Roof Slope at 90 Angle of Wind Attack
Cprms, Lower eave surface

90

90

Cpmin, Upper roof surface including eaves


Cprms, Upper roof surface

Cpmax lower eave surface

Cprms, Lower eave surface

Fig. 4.5 Variation of Cpmin for 10 Roof Slope at 90 Angle of Wind Attack

Fig. 4.6 Variation of Cprms for 10 Roof Slope at 90 Angle of Wind Attack

Journal of Wind and Engineering, Vol. 7, No. 1, January 2010, pp. 16-29
Cpmean, Lower eave surface
Cpmax lower eave surface

23

Cpmean, Upper roof surface

Cpmin, Upper roof surface

Cpmean, Lower eave surface

Cpmax, Lower eave surface

Fig. 4.7 Variation of Cpmean for 20 Roof Slope at 0 Angle of Wind Attack
Cprms, Lower eave surface

Fig. 4.8 Variation of Cpmin for 20 Roof Slope at 0 Angle of Wind Attack
Cpmean, Lower eave surface

Cprms, Upper roof surface


Cpmean, Upper roof surface

Cprms, Lower eave surface

Cpmean, Lower eave surface

Fig. 4.9 Variation of Cprms for 20 Roof Slope at 0 Angle of Wind Attack

Fig. 4.10 Variation of Cpmean for 25 Roof Slope at 0 Angle of Wind Attack

24
Cpmax, Lower eave surface

Wind Pressure on Cable Roof Building with Overhangs


Cprms Lower eave surface

Cprms, Upper roof surface


Cpmin, Upper roof surface

Cpmax, Lower eave surface


0

Cprms Lower eave surface

Fig. 4.11 Variation of Cpmin for 25 Roof Slope at 0 Angle of Wind Attack
Cpmean Lower eave surface

Fig. 4.12 Variation of Cprms for 25 Roof Slope at 0 Angle of Wind Attack
Cpmax lower eave surface

90

90

Cpmean, Upper roof surface

Cpmin, Upper roof surface

Cp Lower eave surface

Cpmax lower eave surface

Fig. 4.13 Variation of Cpmean for 25 Roof Slope at 90 Angle of wind Attack

Fig. 4.14 Variation of Cpmin for 25 Roof Slope at 90 Angle of Wind Attack

Journal of Wind and Engineering, Vol. 7, No. 1, January 2010, pp. 16-29
Cprms, Lower eave surface

25

90

Cprms, Upper roof surface

Cprms Lower eave surface

Fig. 4.15 Variation of Cprms for 25 Roof Slope at 90 Angle of Wind Attack

Spatial Variation of Cp on Upper Roof Surface at 0 Angle of Wind Incidence


With increase in roof pitch, the mean suction (Cpmean) significantly decreases on the windward roof slope. For 0 angle of wind incidence, minimum Cpmean is -2.80 and occurs in the roof with 10 slope at the middle of windward eave edge. At the ridge, the value of Cpmean decreases uniformly from -1.60 to -2.20. For 20 roof slope Cpmean decreases from windward eave to windward wall and then increases uniformly up to the ridge, whereas, variation of Cpmean on the leeward roof is not significant. For 25 roof slope, no variation in Cpmean is observed on the roof surface of windward overhang (eave) whereas it decreases from -0.30 at windward wall to -1.70 at the leeward eave. Minimum value of Cpmin of -6.50 is observed at the eave corner of 10 roof slope at the angle of wind incidence 0. Variation of Cpmin for 10 roof slope is significant; it varies from -6.50 at the windward eave corner to -1.50 at the leeward edge of the eave. For 20 roof slope, Cpmin decreases from windward eave to windward wall, then increases upto the ridge. Same value of Cpmin is observed at windward eave edge and at the ridge. Variation of Cpmin on the leeward roof is not significant. For 25 roof slope, same value of Cpmin is observed at windward as well as the leeward eave edge and it decreases towards the ridge. Uniform variation (decrease) is observed for the value of Cprms with roof pitch from 10 to 25.

26

Wind Pressure on Cable Roof Building with Overhangs

Spatial Variation of Cp on Lower Eave Surface at 0 Angle of Wind Incidence


Contours show that the mean pressure coefficients follow the same trend of increasing values from windward edge towards the windward wall for all roof pitches. Maximum value observed is for 20 roof slope. The value of Cpmean also increases from corner to middle of the eave. The minimum value of Cpmean for leeward eave is observed as -1.20 for the 25 roof slope and occurs at the leeward edge. The minimum value of Cpmin is observed as -3.40 for the 10 roof slope and occurs at the windward eave. The Cprms value for the windward eave is higher than for the leeward eave for all roof pitches.

Spatial Variation of Cp on Upper Roof Surface at 90 Angle of Wind Incidence


There is little effect of roof pitch on the mean roof pressure at 90 angle of wind incidence. Minimum value is observed as -3.00 at the windward corner of the building. Minimum value of Cpmin is observed at the eave corner as -8.50 for roof slopes of 10 and 20 whereas it is -8.00 for the 25 roof slope. Variation of Cpmin follows the same pattern for different roof pitches. Maximum value of Cprms is observed as 1.15 at the windward eave corner, with uniform variation towards the leeward edge and has a minimum value of 0.30 at the leeward edge.

Spatial Variation of Cp on Lower Eave Surface at 90 Angle of Wind Incidence


The mean pressure coefficient on the lower eave surface is observed to be negative for all roof pitches. Its value for the 10 roof slope decreases from windward edge towards the leeward edge. Maximum value is observed as -1.35 for the 10 roof slope. The minimum value of Cpmin is observed as -4.00 at the windward eave corner for the 10 roof slope.

Comparison for Extreme Peaks with the Work of Stathopoulos


Stathopoulos and Luchian (1994) carried out experimental study on perimetric extension of roof on 1:3 and 1:1 roof slopes. The 1:3 sloped overhang is the most severely loaded, particularly as far as suction on the upper surface near the gable is concerned. Most critical peak pressure coefficients observed for upper taps near eave edge as -8.5 whereas at wall it varies from -5.0 to -3.0. In present study worst peaks observed for wind direction parallel to ridge for roof slope 10 and 20 as -8.5 whereas for the 25 roof it is as -8.0. In the case of overhang walls peaks varies from 4.4 to 2.5. In case of lower taps on the overhang, Stathopoulos and Luchian observed maximum positive peaks (Cpmax) as 2.7 where in present study it is observed as 3.0. It shows a good comparison with Stathopoulos and Luchian (1994) work on roof extended overhang.

Comparison with Indian Standard Code (IS: 875 Part-3, 1987)


In the Indian Code, the upper roof of the gable building is divided into eight zones. Three edge zones (one gable end and two eave edges) and one parallel to ridge zone are referred to as local pressure zones whereas E, F, G and H are the major pressure zones. For overhangs (eaves) no zoning has been specified. Pressure coefficients for zones E, F, G, and H are given for wind direction 0 and 90 whereas the local zones are independent of wind direction. In the present study, zone A is the gable end zone, B1 and B2 are overhang (eave) zones, C is the corner zone, and D has been referred to as 'parallel to ridge' zone. E, F, G and H are the major zones as in the Indian Code. A comparison has been given in Table 1

Journal of Wind and Engineering, Vol. 7, No. 1, January 2010, pp. 16-29

27

Table 1 Comparison of Design Pressure Coefficients Cpq of Present Study with Indian Standard Code (IS : 875 Part - 3, 1987) 3-sec gust values

Roof Slope ( (a) Upper Roof zone of Study 10o


Angle of Wind Incidence

20o
Angle of Wind Incidence

25o
Angle of Wind Incidence

0
Study s -1.2 -1.6 -1.4 -1.7 -1.7 -1.0 -1.4 -1.0 -0.9

90
IS Code -2.0 -2.0 -1.5 -2.0 -1.2 -1.1 -1.1 -0.6 -0.6 Study -1.6 -0.9 -1.0 -1.9 -1.5 -1.3 -1.0 -1.3 -1.0

0
IS Code -2.0 -2.0 -1.5 -2.0 -1.2 -0.8 -0.6 -0.8 -0.6 Study -0.9 -1.3 -0.9 -1.3 -0.9 -0.5 -0.6 -0.9 -1.0

90
IS Code -1.5 -1.5 -1.5 -1.5 -1.0 -0.7 -0.7 -0.5 -0.5 Study -1.5 -1.0 -1.2 -1.5 -1.4 -1.3 -0.9 -1.3 -0.8

0o
Study -0.7 -1.1 -0.9 -0.8 -0.8 -0.6 -0.6 -0.9 -0.9 IS Code* -1.3 -------1.0 -0.5 -0.5 -0.5 -0.5 -1.5 -1.0 -1.2 -1.3 -1.4 -1.5 -1.0 -1.5 -0.9

90o
Study IS Code* -1.3 -------1.0 -0.8 -0.8 -0.8 -0.8

A B1** B2** C** D E F G H

IS Code -1. 5 -1.5 -1.5 -1.5 -1. 0 -0.8 -0.6 -0.8 -0.6

*The value is interpolated between 20 and 30, since value for 25 roof slope in not given in the IS code. ** Net Design Pressure Coefficients

Comparison with National Building Code of Canada (NBCC 1995)


Zoning in the Canadian Code is different from that in the Indian Code. In the Canadian Code, two sets of zoning are provided. In the first set, three zones c (corner), s ( gable end and gable edge) and r (for others) for roof slope 0<a =10 in gable buildings with same zones for roof with overhang and without overhang are given. While in the second set, one additional zone is defined parallel to the ridge for roof slope 10<a=30. In the present study, zones B1, B2 and C are overhang zones. In the Canadian Code overhang zones include contributions from both upper and lower surfaces of the overhang. Hence, for direct comparison purpose, the area averaged peak pressure coefficients, Cpmin for zones B1, B2 and C factored by 0.8 (i.e., Cpmin80) have been considered for both upper and lower eave (overhang) surfaces. In the present study for the comparison for overhang zones, the net peak pressure coefficients have been derived by adding the pressures on both the surfaces (upper & lower) for any particular wind direction and compared with the Code values in Tables 2.

Comparison for 10 Roof Slope Peak pressure coefficients for Zones C, E, F, G and H are closer to the Code values whereas zone A has higher value than the Code value. Peak pressure coefficients for overhang zones B1 and B2 are significantly higher than the Code values. Peak pressure coefficient for ridge zone D is slightly higher than the Code value.

28

Wind Pressure on Cable Roof Building with Overhangs

Table 2 Comparison of Worst Peak Pressure Coefficients of Present Study (10, 20 & 25 Roof) with National Building Code of Canada (1995)
Comparison Roof Slope Study NBCC Study NBCC
10 0<a=10 20 25 10<a=30 -4.0 -3.9 -4.4 -3.7 -4.1

Different Zones of the present study A


-4.0

B1*
-4.6

B2*
-4.6 -3.2 -4.0 -3.8

C*
-5.9 -5.4 -5.2 -5.4 -7.0

D
-3.8

E
-3.2

F
-3.3

G
-3.1 -3.2

H
-2.8

-2.6 -2.6

-2.4 -2.8

-2.0 -2.0 -1.6

-2.6 -3.0

-2.9 -2.5

*Net peak pressure


Comparison for 20 and 25 Roof Slopes In the case of 10 roof slope in NBCC for eave edge and gable end the same value for both zones is considered, on the same pattern in present case peak pressure coefficients for zones A, B1 and B2 of roof slopes 20 and 25 are considered for comparison. Peak pressure coefficients for zones A, B1 and B2 are closer to the Code values. Peak pressure coefficients for the corner zones have significantly lower values than the Code values for both roof slopes. Other zones have significantly higher values than the Code values.

CONCLUSION
Studies made for comparing the results with different codes and standards have indicated that pressure coefficients obtained from the experimental studies for different zones of the roof are of the same order as given in different codes. However, some difference exists between the Code values and the experimental values, as also amongst the Code values themselves [Krishna (1995)]. These variations can be mainly due to two reasons, as follows: 1. The values given in different standards are usually based on experimental values from different studies, which yield some difference amongst themselves. 2. Either deterministic or different statistical approaches have been used in different Codes of Practice to deduce the design pressure coefficients from the experimentally obtained pressure fluctuations. 3. The results of this study indicate that there is need for standardization in the experimental as well as the codification process for assigning the design pressure coefficients.

ACKNOWLEDGEMENT
The work presented in this paper is part of the Ph. D. Thesis of the first author awarded in the Department of Civil Engineering I. I. T. Roorkee, Roorkee, India.

Journal of Wind and Engineering, Vol. 7, No. 1, January 2010, pp. 16-29

29

REFERENCES
1. Best, R.J. and Holmes, J.D. (1978), "Model study of wind pressures on an isolated single story house", Wind Engineering Report 3178, Department of Civil and Systems Engineering, James Cook University of North Queensland, Australia. 2. Dreher, K.J. and Cermak, J.E. (1973), "Wind loads on a house roof", Report CER 72-73, KJD- JEC22, Colorado State University, College of Engineering, Ft. Collins, Colorado. 3. Gupta, A. (1996), "Wind tunnel studies on aerodynamic interference in tall rectangular buildings", Ph.D. Thesis, University of Roorkee, Roorkee, India. 4. Holmes, J.D., and Lewis, R.E. (1986), "The dynamic response of pressure measurement systems", Proc. of 9th Australian Fluid Mechanics Conference, Aukland, pp 8-12. 5. Holmes, J.D., and Lewis, R.E. (1987), "Optimization of dynamic-pressure-measurement systems - I Single point measurements", Journal of Wind Engineering and Industrial Aero dynamics, Vol. 25, pp 249-273. 6. Holmes, J.D., and Lewis, R.E. (1987), "Optimization of dynamic-pressure-measurement systems - II Parallel tube-manifold system", Journal of Wind Engineering and Industrial Aerodynamics, Vol. 25, pp 275-290. 7. Holmes, J.D., and Lewis, R.E. (1989), "A re-examination of the leaked tube dynamic pressure measurement systems", Proc. of 10th Australian Fluid Mechanics Conference, Melbourne University, pp 5.39-5.42. 8. IS: 875 (Part 3) - 1987 ( Reaffirmed 1997 ) Code of Practice for Design loads (Other than Earthquake Loads) for building and structures, Part 3 Wind Loads, Bureau of Indian Standards New Delhi. 9. Krishna, P. (1995), "Wind loads on low-rise buildings -A review", Journal of Wind Engineering and Industrial Aerodynamics, Vol. 54-55, pp 383-396. 10. Leutheusser, H.J. (1965), "The effect of eaves on the roof pressure-coefficients of block type and cylindrical structures", UT Mech. E TP 6503, Department of Mechanical Engineering, University of Toronto, Toronto, Canada. 11. NBCC (1995), National Building Code of Canada. 12. Savory, E., Dalley , S. and Toy, N. (1991), "The effects of eaves geometry, model scale and approach flow conditions on portal frame building wind loads", Proc. of 8th Int. Conf. on Wind Engineering, The University of Western Ontario, London, Canada, pp 8-12. 13. Stathopoulos, T. (1981), "Wind loads on eaves of low buildings", Journal of Structural Div., ASCE 107, ST 10, pp 1921-1934. 14. Stathopoulos, T. and Luchian, H. (1994), "Wind-induced forces on eaves of low buildings", Journal of Wind Engineering and Industrial Aerodynamics, Vol. 52, pp 249-261. 15. Tieleman, HW., Hajj, M.R. and Reinhold, T.A. (1999), "Pressure characteristics for separated flows", Wind Engineering into the 21st Century, Larsen, Larose & Livesey (Eds.), Balkema, Rotterdam, pp 1731-1738. 16. Wiik, T. and Hansen, E.W.M. (1997), "The assessment of wind loads on roof overhang of low-rise buildings", Journal of Wind Engineering and Industrial Aerodynamics, Vol. 67-68, pp 687-696.

You might also like