You are on page 1of 17

Journal of Wind Engineering

and Industrial Aerodynamics 88 (2000) 101–117

Mean and fluctuating wind loads on rough and


smooth parabolic domes
C.W. Letchford*, P.P. Sarkar
Wind Science and Engineering Research Center, Department of Civil Engineering, Texas Tech University,
P.O. Box 4089-1023, Lubbock, TX 79409, USA
Accepted 5 June 2000

Abstract

Simultaneous pressure measurements have been obtained on rough and smooth parabolic
domes in simulated atmospheric boundary layer flow. Mean and fluctuating pressure
distributions compare favorably with earlier studies for similar shape and Reynolds number.
The effect of surface roughness is to reduce suctions over the apex of the dome and increase
suctions in the wake region on the leeward face. The consequence for mean and fluctuating
overall loads is reduced uplift but increased drag for rougher surfaces. Correlation analysis of
the fluctuating pressures reveals that the first two eigenvectors account for approximately 60%
of the fluctuating pressure energy and follow the mean pressure coefficient and its gradient
with respect to horizontal wind direction as predicted by the quasi-steady theory. Overall base
shear and uplift forces on the domes can be well approximated by this theory. # 2000
Elsevier Science Ltd. All rights reserved.

1. Introduction

Many wind tunnel studies have been undertaken to determine wind loads on
domes and hemispheres in boundary layer flow [1–8]. Maher’s classical study was
conducted on large models (600 mm diameter) but in more or less uniform flow with
little turbulence and no attempt to simulate the earth’s atmospheric boundary layer.
Only the studies by Ogawa and Taylor present measurements of fluctuating pressures
and only Taylor presents contour maps of maximum and minimum point pressures
for hemispheres and truncated spheres. However, the correlation of fluctuating

*Corresponding author.
E-mail address: chris.letchford@coe.ttu.edu (C.W. Letchford).

0167-6105/00/$ - see front matter # 2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 0 0 ) 0 0 0 3 0 - 1
102 C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117

pressures was not obtained and thus effective overall loading distributions producing
the worst eccentric or unsymmetrical loading are not available.
An additional complication with wind tunnel model studies of these types of
structures is due to the curved surface, which leads to Reynolds number effects.
Here, Reynolds number is defined as r DU/ m , where r and m are fluid properties,
D is the base diameter and U is the mean velocity at top of model height. Maher
examined these effects in the range 6  105–18  105, while Taylor dealt with this
issue more systematically by evaluating loading under different turbulence intensities
and Reynolds numbers ranging from 1  105 to 3  105. Ogawa [7,9] also investigated
a range of turbulence intensities with Reynolds numbers ranging from 1.2  105 to
2.1  105. Taylor suggests that as long as the Reynolds numbers exceeds 2  105 and
turbulence intensity exceeds 4% the surface pressure distributions are largely
independent of Reynolds number. Maher suggests a critical Reynolds number of
1.4  106 but his tests were conducted in very low turbulence. Maher did study the
influence of surface roughness, but in the absence of turbulence.
The present study sought to re-examine wind loading of domes and specifically to
determine the effect of Reynolds number, surface roughness, and overall fluctuating
load distributions to aid designers of these structures.

2. Experimental procedure

The wind tunnel tests were undertaken in the Colorado State University’s
Industrial Wind Tunnel. The wind tunnel is a closed circuit, 1.8 m wide with a ceiling
adjustable to  1.8 m. There is an upstream fetch of approximately 15 m for
developing appropriate simulations of the earth’s atmospheric boundary layers. A
simulation of Exposure C at 1 : 300 was achieved by employing 4 spires, a 240 mm
fence, and 18 mm chain at 200 mm spacing right up to the wind tunnel turntable. The
mean velocity and turbulence intensity profiles are shown in Figs. 1 and 2,
respectively, and compared with ASCE 7-98 [10] target values. The comparison with
longitudinal velocity component spectrum at top of dome height is shown in Fig. 3,
where the model-scale integral length scale of turbulence is  700 mm. In all cases
the simulation achieved compares very well with the Exposure C target. For the
pressure measurements, the mean wind speed and turbulence intensity at the top of
the dome was 18 m/s and 15%, respectively.
The nominal design gust wind velocity was 59 m/s at 10 m in flat open terrain.
Using the approach of ASCE 7-98 [10], the design gust wind speed at the top of the
simulated dome (45 m) is expected to be Kz1=2 59 ¼ 1:17  59 ¼ 69 m/s. The mean
velocity at this height would be 0:82  59 ¼ 48 m/s where the ratio of the 3 s gust to
the 1 h mean velocity, G ¼ 1:17=0:82 ¼ 1:43. The model scale gust factor is expected
to be very similar given the good agreement for mean and turbulence intensity
profiles (Figs. 1 and 2). The maximum mean wind speed in the wind tunnel at
150 mm was measured to be  18 m/s and thus the velocity scale was 18/48  1/3.
Combining the length and velocity scales a time scale of ð1=3Þ=ð1=300Þ ¼ 1=100 was
obtained. Thus 36 s in the wind tunnel is equivalent to 1 h at full scale.
C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117 103

Fig. 1. Comparison of wind tunnel and ASCE 7-98 mean velocity profiles.

Fig. 2. Comparison of wind tunnel and ASCE 7-98 [10] turbulence intensity profiles.

A model dome was constructed with a base diameter (D) of 480 mm and height (h)
of 150 mm. The nominal model scale was 1 : 300. The model was constructed from
fiberglass to a parabolic profile that well approximates a sphere of radius 280 mm
(  1%). The exterior surface was covered with ‘gelcoat’ to obtain a very smooth
finish. To obtain a rougher surface the model was covered with a tailored fly screen
mesh of 0.3 mm diameter fiber at 1.5 mm spacing and secured by tape around the
base of the dome. This gave a nominal roughness coefficient of e=D ¼ ð0:321:5Þ=
480  0:001. Although wind tunnel blockage was not great (  1.5%) the wind
104 C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117

Fig. 3. Comparison of wind tunnel and ASCE 7-98 [10] turbulence spectra at top of dome height.

tunnel roof was positioned to minimize the static pressure gradient in the wind
tunnel induced by the simulation and the model.
The dome was instrumented with 85 pressure taps, with one tap at the apex and
seven taps non-uniformly distributed along 12 meridians at 30o increments, to
achieve equal tributary area. Fig. 4 shows a general arrangement of the dome
showing tapping grid and wind direction definition.
The taps were connected by tubing (240 mm long, 0.51 mm diameter), tubing to a
multi-port PSI pressure transducer and sampled simultaneously at 480 Hz for a total
of 96 s. The instrumentation frequency response was linear to 160 Hz, equivalent to
1.6 Hz at full scale. Mean and standard deviation (rms) pressures were obtained from
the full 96-s record. To estimate mean extreme or peak pressures, each record was
divided into 16 segments, equivalent to 10 min full scale, and a Fischer-Tippett type 1
extreme value distribution fitted. From this the 10-min mode and dispersion were
obtained and the mean hourly extreme or peak pressure coefficient was estimated
according to Eq. (1), after Cook [11]:
ðD^p or DpÞ ¼ mode þ ð0:577 þ lnð6ÞÞ  dispersion10 min : ð1Þ
All pressures were then non-dimensionalized by the mean dynamic pressure
ð1=2rU 2 Þ at the top of the dome. The definition of the coefficients is shown below
with positive external pressure acting towards the surface and suction away. Dp is the
instantaneous pressure difference between the surface pressure and a reference
pressure in the wind tunnel.
Dp
Cp ¼ 1 mean pressure coefficient; ð2Þ
2
2 rU
C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117 105

Fig. 4. Tapping arrangement and wind direction definition for single dome tests.

Dprms
Cprms ¼ 1 standard deviation pressure coefficient; ð3Þ
2
2 rU

^ ¼ D^p
Cp mean peak maximum pressure coefficient; ð4Þ
1 2
2 rU

 ¼ Dp
Cp mean peak minimum pressure coefficient: ð5Þ
1 2
2 rU

Using the gust velocity for design purposes, it is often convenient to define pseudo-
steady pressure coefficients which are the peak pressure coefficients defined above
(Eqs. (4) and (5)) divided by the gust velocity/mean velocity ratio squared. In this
way, it is possible to compare directly the mean pressure coefficient and the pseudo-
steady coefficient with the latter being more realistic as it takes into account the
106 C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117

fluctuating pressures over the structure:

^3
U
G¼ gust velocity ratio: ð6Þ
U1 h

The corresponding coefficients are


^
Cp
C~p^ ¼ 2 mean maximum pseudo-steady pressure coefficient; ð7Þ
G

Cp
C~p ¼ 2 mean minimum pseudo-steady pressure coefficient: ð8Þ
G
If the quasi-steady theory of wind loading [12] which attributes the source of all
pressure fluctuations to upwind velocity fluctuations, holds true, then the mean and
pseudo-steady mean maxima (or mean minima depending on the sign of the mean)
should be equal.
Overall fluctuating forces (or other load effects) on the dome are obtained by
appropriate weighting of the pressures and integrating over the entire surface. This
effect is shown in Eq. (9).
X
85
CFf ¼ Cpj fj ðdA=AÞ; ð9Þ
j¼1

where fj is the weighting coefficient or influence line for the force or load effect (CFf )
1
required, dA is the tributary area of the tapping (here 85 of the surface area of the
dome), and A has been set as the projected plan area of the dome.
Thus given the influence lines for any particular load effect (e.g., overall drag,
overall uplift, top of dome deflection, or bending moment), the fluctuating time
history of that effect can be calculated and the statistics, mean, rms, mean maxima
and mean minima estimated.
Here the overall force coefficients in the X (parallel to wind direction a =0o), Y
(perpendicular to X in the horizontal plane), and Z (vertical or uplift), were
calculated because the influence coefficients can be determined without recourse to
structural analysis of the dome. Here fX ¼ cos b cos y, fY ¼ cos b sin y, and
fZ ¼ sin b. b is defined as the angle between horizontal and a line perpendicular
to the surface and y is the tapping angle; both are defined in Fig. 4.

3. Experimental results

A series of tests were conducted varying wind speeds, surface finishes, and wind
direction. They are summarized in Table 1. The five wind directions were used to
obtain a detailed picture of the mean pressure field and to provide a check on the
results.
C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117 107

Table 1
Summary of testing regimes

Wind speed (m/s) Surface finish Wind directions, a (deg)

9 and 18 Smooth 0, 7.5, 15, 30, 180


18 Rough 0, 7.5, 15, 30, 180

Fig. 5. Comparison of mean pressure coefficient along centerline of a smooth dome.

3.1. Comparison with other studies

The earlier studies of Taylor [8] represent the best work for comparison since
similar geometries, in similar boundary layers were tested and mean and fluctuating
pressures were reported. Ogawa [9] also provided data for comparison. Fig. 5
compares the mean pressure coefficient along the centerline of a smooth dome
plotted against angle g, defined in Fig. 4. The results for the domes tested by Maher
[1], that bracket the current h/D ratio, are included. Generally, there is reasonable
agreement for mean pressures even with interpolating (within h/D) Maher’s tests,
where no attempt was made to scale the atmospheric boundary layer. This is
encouraging because Maher’s tests were conducted at four times the Reynolds
number of the present study and lend further support to Taylor’s criteria for valid
pressure measurements on domes.
Figs. 6 and 7 compare the fluctuating pressures measured along the centerline of
the smooth dome with those of Taylor [8]. It is observed that there is excellent
agreement for peak maximum and minimum pressure coefficient, with only a small
discrepancy in distribution of rms pressure coefficient, which is in part due to the
108 C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117

Fig. 6. Comparison of rms pressure coefficient along centerline of a smooth dome.

Fig. 7. Comparison of peak pressure coefficients along centerline of a smooth dome.

different dome shape. Ogawa et al. [7] described but did not present results for the
truncated dome but has kindly provided them for this paper Ogawa [8]. His results
are for a lower turbulence intensity at top of dome height (10%), which is reflected in
the lower rms pressure fluctuations.

3.2. Effects of surface finish on cladding pressures

Maher [1] investigated the effect of surface finish on pressures over a hemisphere.
As described above, his flow was basically uniform with only a shallow boundary
C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117 109

layer on the wind tunnel floor. There was little turbulence in the flow. The flow
Reynolds number was 1.8  106. Mean pressures were obtained on a smooth
hemisphere and on one roughened by sand grains such that e=D ¼ 0:001, where e is a
measure of the roughness height and D the diameter of the hemisphere. His results
indicate a reduction in maximum mean pressure coefficient near the top of the dome
from 1.15 to 0.8,  30%, due to the addition of this roughness. This reduction is to
be expected since surface roughness promotes a turbulent boundary layer over the
dome surface, which will tend to separate earlier than that on the smooth surface.
The consequences of earlier separation are reduced suctions at the point of
separation but higher suction overall in the wake, leading to reduced uplift and
increased drag. This result is born out in Maher’s experiments.
In the present study, similar effects were observed. Figs. 8, 9 and 10 compare the
mean, rms and peak pressure coefficients along the centerline of the smooth and
rough single domes, respectively. Reduced suctions, both mean and fluctuating, over
the apex of the dome are present but to a lesser extent than in Maher’s study due to
smaller h/D, here 0.31, cf. Maher’s hemisphere with h=D ¼ 0:5. In addition, the
greater suctions in the wake lead to increased drag on the rougher dome. Overall
forces on the dome are discussed in Section 3.5.

3.3. Effect of Reynolds number

The pressure distribution was measured on the smooth dome at a wind speed half
that of the previous tests. This was the lowest wind speed that could be run and
maintain reasonable sensitivity from the pressure measurement instrumentation. The
Reynolds numbers of these tests was 2.3  105. The mean, rms and peak pressure
coefficients along the centerline are compared with the earlier results at the higher
Reynolds numbers in Figs. 11, 12 and 13, respectively. It is evident that these results

Fig. 8. Comparison of mean centerline pressure coefficients for rough and smooth domes.
110 C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117

Fig. 9. Comparison of rms centerline pressure coefficients for rough and smooth domes.

Fig. 10. Comparison of peak centerline pressure coefficients for rough and smooth domes.

indicate little sensitivity to Reynolds number over this small range for both mean
and fluctuating pressures. On the graph of mean pressure coefficients, the result of
Taylor [8] and Ogawa [9] are reproduced and the large difference between the present
results and theirs may have been caused by static pressure gradients in their wind
tunnels that remained uncorrected or the slightly different geometry of the dome.
C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117 111

Fig. 11. Comparison of mean pressure coefficients along the centerline of a smooth dome for different
Reynolds number.

Fig. 12. Comparison of rms pressure coefficients along the centerline of a smooth dome for different
Reynolds number.
112 C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117

Fig. 13. Comparison of peak pressure coefficients along the centerline of a smooth dome for different
Reynolds number.

Table 2
First 10 eigenvalues for the smooth and rough domes

Eigenvector number Eigenvalue (%) smooth Eigenvalue (%) rough

1 44 36
2 14 12
3 6.1 7.3
4 5.1 7.1
5 3.5 4.1
6 3.3 3.5
7 2.9 2.5
8 2.2 2.3
9 1.6 1.7
10 1.4 1.5

Sum of first 10 eigenvalues 83 77

3.4. Correlation of fluctuating surface pressures

A covariance matrix was formed from the 85 simultaneous time histories using
MATLAB for both the smooth and rough domes for the 0o wind direction. An
eigenvector/eigenvalue analysis was then undertaken to further interpret the
fluctuating pressure field. The first 10 eigenvalues contribute approximately 80% of
the total fluctuating pressure energy with the first eigenvalue accounting for some 40%
in each case. Table 2 compares the first 10 eigenvalues for the rough and smooth
C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117 113

domes. The first eigenvector is compared with the mean pressure distribution for the
smooth dome in Fig. 14 and for the rough dome in Fig. 15. In each case, there is
excellent correspondence as might be anticipated from the quasi-steady theory. Fig. 16
shows a contour plot, with full-scale dimensions, of the second eigenvector for the

Fig. 14. Comparison of first eigenvector and variation of mean Cp with pressure tapping for the smooth dome.

Fig. 15. Comparison of first eigenvector and variation of mean Cp with pressure tapping for the rough dome.
114 C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117

Fig. 16. Contour plot of second eigenvector for smooth and rough domes.

smooth and rough domes, respectively. The dominant feature is the asymmetry on the
frontal portion, which closely resembles the gradient of mean pressure coefficient
dCp=da. Once again, this result would be anticipated from the quasi-steady theory.

3.5. Overall forces on a dome

As indicated in Section 2, by weighting the individual pressure taps accordingly,


any particular load effect and its statistics can be calculated; all that is required are
C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117 115

Table 3
Comparison of overall force coefficients on smooth and rough domes

Dome Force component mean rms mean maxima mean minina

Smooth X 0.023 0.008 0.057 ÿ 0.008


Y 0.001 0.016 0.056 ÿ 0.061
Z ÿ 0.236 0.053 ÿ 0.079 ÿ 0.430
Rough X 0.058 0.014 0.115 0.019
Y 0.002 0.013 0.056 ÿ 0.053
Z ÿ 0.144 0.041 ÿ 0.017 ÿ 0.319

Table 4
Compares the pseudo-steady maxima and minima force coefficients with the mean force coefficients

Dome Force component mean Pseudo-steady maxima Pseudo-steady minima

Smooth X 0.023 0.028 ÿ 0.004


Y 0.001 0.027 ÿ 0.030
Z ÿ 0.236 ÿ 0.039 ÿ 0.211
Rough X 0.058 0.057 0.010
Y 0.002 0.027 ÿ 0.026
Z ÿ 0.144 ÿ 0.008 ÿ 0.157

the influence coefficients for the load effect in question. For instance, the peak
deflection at the top of the dome or the peak axial force in a chord member could be
calculated.
As an example, the overall forces on a dome in the X, Y and Z directions (defined
in Fig. 4) have been calculated according to Eq. (9). The results are in force
coefficient form using the mean dynamic pressure at the top of the dome and the
projected plan area of the dome base for non-dimensionalizing. Table 3 compares
the X, Y and Z force coefficients for the smooth and rough domes. Both mean and
fluctuating coefficients are presented.
It is evident from Table 3 that the smooth dome has nearly twice the uplift (Z) and
less than half the drag (X) of the rough dome, and these differences are more
pronounced in the mean force coefficients than the peak (maxima and minima) force
coefficients.
Table 4 compares the mean force coefficients with the pseudo-steady force
coefficients as defined by Eqs. (7) and (8). It is interesting to note that the quasi-
steady theory predicts quite well the maximum drag on each dome, but would
slightly overpredict the maximum uplift on the smooth dome while underpredicting
the uplift on the rough dome.

4. Conclusions

Domes are commonly used to enclose large spaces because of their structural
efficiency and consequent economic benefits. Such structures are used to cover
116 C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117

sporting grounds, shopping centers and raw material stockpiles. These structures are
sensitive to load distributions, and whereas they are excellent at resisting symmetric
loading, asymmetric loading may cause structural distress [13,14]. In this study wind
loading of domes has been revisited, and a comprehensive wind tunnel study
produced, for the first time, simultaneous external fluctuating surface pressures over
an entire dome. In a correctly scaled atmospheric boundary layer, this study has
examined the influence of Reynolds number, and surface roughness as well as
presented overall base shear loads on circular domes with height/base diameter
aspect ratio of 1/3.
The specific conclusions of this study are:

* Mean and fluctuating pressure distributions compare favorably with earlier


studies for similar shape, flow characteristics, and Reynolds number. The
parabolic dome is well approximated by a spherical dome of the same height to
diameter ratio.
* Pressure distributions were independent of Reynolds number in the range
2.3  105–4.6  105 defined by velocity at top of dome and base diameter.
* The effect of surface roughness is to reduce suctions over the apex of the dome
and increase suctions in the wake region on the leeward face. The consequence for
mean and fluctuating overall loads is reduced uplift but increased drag for
rougher surfaces.
* Correlation analysis of the fluctuating pressures reveals that the first two
eigenvectors account for approximately 60% of the fluctuating pressure
energy and follow the mean pressure coefficient and its gradient with
respect to horizontal wind direction as predicted by the quasi-steady theory.
Overall base shear and uplift forces on the domes can be well approximated by
this theory.

Acknowledgements

The authors wish to thank Dr. Bob Meroney and Dr. David Neff of Colorado
State University for their assistance during the wind tunnel work. We also thank
Dr. Toshiyuki Ogawa for kindly providing data.

References

[1] F.J. Maher, Wind loads on basic dome shapes, J. Struct. Div. ASCE ST3 (1965) 219–228.
[2] J. Blessmann, Pressures on domes with several wind profiles, Proceedings of the 3rd International
Conference on Wind Effects and Building Structures, Tokyo, 1971, pp. 317–326.
[3] S. Taniguchi, H. Sakamoto, Time-averaged aerodynamic forces acting on a hemisphere immersed in a
turbulent boundary, J. Wind Eng. Ind. Aerodyn. 9 (1981) 257–273.
[4] N. Toy, W.D. Moss, E. Savory, Wind tunnel studies on a dome in turbulent boundary layers, J. Wind
Eng. Ind. Aerodyn. 1 (1983) 201–212.
C.W. Letchford, P.P. Sarkar / J. Wind Eng. Ind. Aerodyn. 88 (2000) 101–117 117

[5] B.G. Newman, U. Ganguli, S.C. Shrivastava, Flow over spherical inflated buildings, J. Wind Eng.
Ind. Aerodyn. 17 (1984) 305–327.
[6] E. Savory, N. Toy, Hemispheres and hemisphere-cylinders in turbulent boundary layers, J. Wind
Eng. Ind. Aerodyn. 23 (1986) 345–364.
[7] T. Ogawa, M. Nakayama, S. Murayama, Y. Sasaki, Characteristics of wind pressures on basic
structures with curved surfaces and their response in turbulent flow, J. Wind Eng. Ind. Aerodyn. 38
(1991) 427–438.
[8] T.J. Taylor, Wind pressures on a hemispherical dome, J. Wind Eng. Ind. Aerodyn. 40 (1991) 199–213.
[9] T. Ogawa, private communication, 1999.
[10] ASCE 7-98, ASCE Minimum Design Loads for Buildings and Other Structures, SEI, Reston, VA,
1999.
[11] N.J. Cook, The Designer’s Guide to Wind Loading of Building Structures: Part 1, BRE/
Butterworths, London.
[12] C.W. Letchford, R.E. Iverson, J.R. McDonald, The application of the Quasi-steady theory to full
scale measurements on the Texas tech building, J. Wind Eng. Ind. Aerodyn. 48 (1993) 111–132.
[13] J.D. Holmes, Determination of wind loads for an arch roof, Civil Eng. Trans. IEAust. CE26 (1984)
247–253.
[14] J.D. Holmes, Optimised peak load distributions, J. Wind Eng. Ind. Aerodyn. 41–44 (1992) 267–276.

You might also like