mean loads on vaulted canopy roofs: tests in … · mean loads on vaulted canopy roofs: tests in...

13
Mean loads on vaulted canopy roofs: tests in boundary layer wind tunnel M.B. Natalini 1 , C. Morel 2 , B. Natalini 3 , O. Canavesio 4 1 Prof. of Civil Eng., Univ. Nac. del Nordeste, Resistencia, Argentina, [email protected] 2 Assistant Prof. of Civil Eng., Univ. Nac. del Nordeste, Rcia., Argentina, [email protected] 3 Prof. of Mechanical Eng., CONICET-UNNE, Resistencia, Argentina, [email protected] 4 Assistant Prof. of Civil Eng., Univ. Nac. del Nordeste, Rcia., Argentina, [email protected] ABSTRACT The vaulted canopy roof (VCR) is a widespread structure in vast areas of South-America, especially in parts of Argentina, Brazil and Paraguay. Information about the aerodynamics of VCRs however, is scarce in the open literature. This paper presents the results of mean wind loads coefficients on VCRs obtained in boundary layer tunnel tests, which constitute a firm base to assess wind loads on the structure. INTRODUCTION The vaulted canopy roof (VCR) is a widespread structure in vast areas of South-America, especially in parts of Argentina, Brazil and Paraguay where the climate is warm. They can be seen across both urban and rural areas, in a wide variety of uses. For instance, a common landmark in the north-east of Argentina, is a carpenter’s workshop sheltered by a VCR. Information available in the specialized literature about the aerodynamic behavior of VCRs is scarce. Cook [1] summarized this situation when he wrote: “There are only a small amounts of data for curved canopies and these are all early data obtained in smooth uniform flow. For example Irminger and Nokkentved included a barrel vault canopy with a rise of r = W/4 in their studies published in 1936, and Blessmann included domed canopies with rises W/4 y W/8 in his 1971 studies. The validity of the loading coefficient is open to question in the light of current techniques, however the general loading characteristics may be still be useful.”. This situation is also reflected in the codes of practice. The authors of this work found only two codes that provide loads coefficient values for VCRs: the French code NV 65 [2] and the Argentinean code CIRSOC 102 [3]; the former having been superseded by the Eurocode and the later being superseded by a new version that will be shortly come into effect [4]. Both codes suggest using the same mean pressure coefficients of planar canopy roofs while keeping the same relation rise/span. Marighetti et al. [5] though, showed that this approximation is incorrect in their presentation of wind tunnel results of mean pressure distributions on VCRs which they compared with the those corresponding to a planar canopy roof with similar dimensions. Fig. 1 reproduces two figures from their work, corresponding to incident wind angles of 45º and 90º relative to the ridge. It can be seen that the contour plots have different patterns as well as different values, which is not surprising taking into account that it is hardly possible that the flow over a surface with sharp edges would be similar to the flow over a relatively smooth curved surface. Reports on boundary layer wind tunnel tests related to these structures have also not been found in the literature, apart from those of Natalini et al. [6, 7, 8]. Regarding computational modelling the situation is similar, though there have been recently some promising results reported by Balbastro et al. [9, 10, 11].

Upload: tranthuy

Post on 02-Aug-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

Mean loads on vaulted canopy roofs: tests in boundary layer wind tunnel

M.B. Natalini1, C. Morel

2, B. Natalini

3, O. Canavesio

4

1Prof. of Civil Eng., Univ. Nac. del Nordeste, Resistencia, Argentina, [email protected]

2Assistant Prof. of Civil Eng., Univ. Nac. del Nordeste, Rcia., Argentina, [email protected]

3Prof. of Mechanical Eng., CONICET-UNNE, Resistencia, Argentina, [email protected]

4Assistant Prof. of Civil Eng., Univ. Nac. del Nordeste, Rcia., Argentina, [email protected]

ABSTRACT

The vaulted canopy roof (VCR) is a widespread structure in vast areas of South-America,

especially in parts of Argentina, Brazil and Paraguay. Information about the aerodynamics of

VCRs however, is scarce in the open literature. This paper presents the results of mean wind

loads coefficients on VCRs obtained in boundary layer tunnel tests, which constitute a firm base

to assess wind loads on the structure.

INTRODUCTION

The vaulted canopy roof (VCR) is a widespread structure in vast areas of South-America,

especially in parts of Argentina, Brazil and Paraguay where the climate is warm. They can be

seen across both urban and rural areas, in a wide variety of uses. For instance, a common

landmark in the north-east of Argentina, is a carpenter’s workshop sheltered by a VCR.

Information available in the specialized literature about the aerodynamic behavior of VCRs is

scarce. Cook [1] summarized this situation when he wrote: “There are only a small amounts of

data for curved canopies and these are all early data obtained in smooth uniform flow. For

example Irminger and Nokkentved included a barrel vault canopy with a rise of r = W/4 in their

studies published in 1936, and Blessmann included domed canopies with rises W/4 y W/8 in his

1971 studies. The validity of the loading coefficient is open to question in the light of current

techniques, however the general loading characteristics may be still be useful.”. This situation is

also reflected in the codes of practice. The authors of this work found only two codes that

provide loads coefficient values for VCRs: the French code NV 65 [2] and the Argentinean code

CIRSOC 102 [3]; the former having been superseded by the Eurocode and the later being

superseded by a new version that will be shortly come into effect [4]. Both codes suggest using

the same mean pressure coefficients of planar canopy roofs while keeping the same relation

rise/span. Marighetti et al. [5] though, showed that this approximation is incorrect in their

presentation of wind tunnel results of mean pressure distributions on VCRs which they compared

with the those corresponding to a planar canopy roof with similar dimensions. Fig. 1 reproduces

two figures from their work, corresponding to incident wind angles of 45º and 90º relative to the

ridge. It can be seen that the contour plots have different patterns as well as different values,

which is not surprising taking into account that it is hardly possible that the flow over a surface

with sharp edges would be similar to the flow over a relatively smooth curved surface.

Reports on boundary layer wind tunnel tests related to these structures have also not been found

in the literature, apart from those of Natalini et al. [6, 7, 8]. Regarding computational modelling

the situation is similar, though there have been recently some promising results reported by

Balbastro et al. [9, 10, 11].

Curved Planar Curved Planar

Fig. 1: Contour plots of mean pressure coefficients over curved and planar canopy roofs (after Marighetti et

al. [5]).

Considering the lack of reliable data sources for state-of-art wind engineering and that the

structural designer needs design aerodynamic coefficients, the subject was the object of a wind-

tunnel-based studio in the Laboratorio de Aerodinámica de la Facultad de Ingeniería de la

Universidad Nacional del Nordeste (UNNE), Argentina.

This paper presents results of mean wind loads coefficients on VCRs obtained in boundary layer

tunnel tests. The first problem approached was the selection of the most suitable technique for

modeling pressures on curved surfaces. Once the modeling conditions were established, local

and global wind load coefficients were determined on models of six VCRs of different geometry

using wind simulation.

WIND TUNNEL MODELING TECHNIQUES

The study of these structures through wind tunnel tests presents difficulties inherent to both free-

standing canopies and bodies with curved surfaces, for which it was necessary to develop

specific modeling techniques. The first problem approached was the selection of the most

suitable technique for modeling pressures on curved surfaces. The issue of modelling wind loads

on VCRs was approached by Natalini et al. [6], who based in previous work of Ribeiro [12],

Cheung and Melbourne [13], Batham [14], Blackbourn and Melbourne [15], Blessmann and

Loredo Souza [16] and Blessmann [17], tested three 1:75 scale models of similar geometry but

different roughness over the roof. As a result, sand of appropriate size was added to the upper

side of the roof of the models.

On the other hand, Natalini et al. [18] reported a study on the modelling of mean pressures on

planar canopy roofs. The paper presented wind tunnel results of tests on scale models of the full-

scale Dutch barn tested by Robertson et al. [19] at Silsoe, UK. The comparison between both full

and reduced scale barns showed that severe distortions can appear which are linked to scale

effects if the size of the model is below a certain level. In the case of the UNNE tunnel such

distortion did not manifest when using a 1:75 scale model [20]. Based in these experiences the

scale 1:75 was adopted for all the models used in the present study.

EXPERIMENTAL ARRANGEMENT

MODELS

The geometry and size of the six models that were tested were adopted on account of the range

of dimensions of the VCRs more frequently found in the north-east of Argentina, and the size of

the models of enclosed buildings appearing in the literature. Fig. 2 and Table 1 summarize the

geometry of the models.

Fig. 2: Geometry of the models.

Table 1: Dimensions of the models.

The rise, f, and the span, b, was kept similar in all models. Basically, two kinds of models were

built, the deep models (A, B and C) and the short models (D, E and F), which allowed the

assessment of the influence of the depth. By varying the height of the eves, h, (2, 4 and 6 cm),

the six models were produced. It is clear that models C and F, with a h dimension that

corresponds to 1.5 m at full-scale, do not represent any real situation. They have been included in

the tests in order to observe the trend that pressures follow when varying the height of the eves.

Wind load coefficients were measured under wind blowing at angles of 60º, 75º and 90º relative

to the ridge line since, as demonstrated by Natalini et al. [8], these directions produce the most

severe loads.

The roof of the models was made with a 2 mm thick aluminium plate, and the columns with a 2.5

mm diameter steel rod (except one column, the farthest one from the area where taps were

placed, which had a square cross section of 10 × 10 mm). As the models have two axes of

Absolute dimensions Relative dimensions

Mo

del

a

[cm]

b

[cm] h [cm]

f

[cm] b/a h/b f/b f/h

r

[cm] α Re × 10

5

A 60 15 6 3 0.25 0.40 0.20 0.5 10.88 43º58’ 2.09

B 60 15 4 3 0.25 0.27 0.20 0.75 10.88 43º58’ 1.96

C 60 15 2 3 0.25 0.133 0.20 1.50 10.88 43º58’ 1.68

D 30 15 6 3 0.50 0.40 0.20 0.50 10.88 43º58’ 2.09

E 30 15 4 3 0.50 0.27 0.20 0.75 10.88 43º58’ 1.96

F 30 15 2 3 0.50 0.133 0.20 1.50 10.88 43º58’ 1.68

b

h

f

r α

a

symmetry, only a quarter of the models' roofs had pressure taps in place, in this way reducing the

number of tubes needed. In addition, all the tubes were led towards the farthest column, through

which they reached the floor. In this way, the scale distortion in both columns and roof thickness

and the possible interference of the tubes upon the measurements were minimized.

In order to obtain a flow in transcritical conditions, sand was added to the upper side of the roof

of the models, being the relative roughness, dk / , equal to 3.30 × 103.

WIND SIMULATION

The wind tunnel tests were undertaken in the “Jacek P. Gorecki” wind tunnel at the Universidad

Nacional del Nordeste. This is an open return wind tunnel with a working section of 22.4 m in

length × 2.4 m in width × 1.8 m in height and a maximum flow velocity of 25 m/s when the

working section is empty. Further details are given by Wittwer and Möller [21].

All the models were tested under a wind simulation corresponding to a suburban area. The

simulation hardware consisted of two modified Irwin’s spires [22] and 17.1 m of surface

roughness fetch downstream from the spires. In this way, a part-depth boundary layer simulation

of neutrally stable atmosphere was obtained. Mean velocities, when fitted to a potential law, give

an exponent of 0.24. The length scale factor determined according to Cook’s procedure [23] was

about 150. The local turbulence intensity at the level of the roof is 0.25. De Bortoli et al. [24]

gave further details of this simulation, including size, geometry and arrangement of the hardware

and design criteria.

PRESSURE MEASUREMENT SYSTEM

Pressures were measured using a differential pressure electronic transducer Micro Switch

Honeywell 163 PC. A sequential switch Scanivalve 48 D9-1/2, which was driven by means of a

CTLR2 / S2-S6 solenoid controller connected the roof pressure taps to the transducer through

PVC tubes of 1.5 mm in internal diameter and 400 mm in length. No resonance problems were

detected for tubes of that length (the gain factor being around one) therefore restrictors of section

were not used for filtering. The DC transducer output was read with a Keithley 2000 digital

multimeter. The integration time operation rate of the A/D converter was set to produce mean

values over 55 seconds.

Simultaneously to the pressure measurements being taken on the roof, the reference dynamic

pressure, qref, was measured at the eaves height with a Pitot-Prandtl tube connected to a Van

Essen 2500 Betz differential micromanometer of 1 Pa resolution. The probe stayed beside the

model at a distance of about 70 cm to avoid mutual interference. The reference static pressure

was obtained from the static pressure tap of the same Pitot-Prandlt tube.

RESULTS

PRESSURE COEFFICIENTS

The net pressure coefficient, pc , was obtained by subtracting the internal pressure coefficient,

pic , from the external pressure coefficient, pec . The two last coefficients are the rate between

the pressure on the tap, p, which can be either an external or internal pressure, and qref , which is

the reference dynamic pressure measured at the reference height. Here, the tap pressures are

relatives to the static reference pressure, refp , which is obtained from the static tap of the same

Pitot-Prandtl tube used to measure qref. Following the usual convention, negative values of both

pec and pic indicate actions directed out from the surface (suctions). Accordingly, positive

values of pc indicate actions directed into the building. In this work, the most positive and

negative values of a set of pressure coefficient will be referred as the maximum and minimum

values of that set.

The whole set of contour plots were given by Natalini [25], including figures of internal, external

and net pressure coefficients. When comparing pressures on deep and short models, at first

glance they show more similarities than disparities. However, two differences can be noted,

though they are restricted to the central area of the roofs. On the one hand, external pressures are

slightly higher near the upwind eave of the deep models and the zero pressure coefficient line is

shifted towards leeward; on the other hand, internal negative pressures are lower (that is bigger

in absolute value) near the ridge of the deep models.

The second difference can be better appreciated in figures 6 to 8, which show the profiles of the

pressure coefficient distribution on a cross section in the middle of the roof (section II). In

contrast, figures 3 to 5, which show the pressures on a cross section located 5 mm inwards the

upwind edge (Section I), present virtually the same figures for both models.

60º Section I. Model D

1

1.5

-2

-0.5

-1

-1.5

0.5

0

1.5

60º Section I. Model A

1

0.5

0

-0.5

-1

-1.5

-2

External pressure coefficient

Internal pressure coefficient

Net pressure coefficient

Fig. 3: External, internal and net pressure coefficients profiles. θ=60º. Upwind edge.

75º Section I. Model A

-1.5

-1

-0.5

0

0.5

1.5

1

75º Section I. Model D

1

1.5

0.5

0

-0.5

-1.5

-1

Fig. 4: External, internal and net pressure coefficients profiles. θ θ θ θ = 75º. Upwind edge.

90º Section I. Model A

-1

-0.5

0

0.5

1

90º Section I. Model D

1

1.5

0.5

0

-0.5

-1

Fig. 5: External, internal and net pressure coefficients profiles. θθθθ = 90º. Upwind edge.

60º Section II. Model A

1

0.5

0

-0.5

-1

1.5

60º Section II. Model D

1

1.5

0.5

0

-0.5

-1

Fig. 6: External, internal and net pressure coefficients profiles. θ=60º. Central section.

75º Section II. Model A

1.5

1

0.5

0

-0.5

-1

75º Section II. Model D

1

1.5

0.5

0

-0.5

-1

Fig. 7: External, internal and net pressure coefficients profiles. θ = 75º. Central section.

90º Section II. Model A

1

0.5

0

-0.5

-1

90º Section II. Model D

1

1.5

0.5

0

-0.5

-1

Fig. 8: External, internal and net pressure coefficients profiles. θ θ θ θ = 90º. Central section.

Theses differences occur because in the deep models the flow tends to be bidimensional in the

central area. Let us do the following thought experiment: divide the deep model plots in four

parts by means of normal-to-the-ridge vertical planes, remove the two central pieces and join the

two remaining parts. The plots obtained in this way are almost the same than the corresponding

to short models. It shall be seen in the following section how these differences propagate into the

global coefficients.

Note the role that internal pressures play in the resulting net pressures. The contribution of the

internal pressures ranges from 29% to 69% of the net pressures. Internal pressures are highly

dependant on blockage conditions, and as the results presented here correspond to models with

no blockage, loads can be expected to be significantly different under any degree of blockage.

Figures and plots show that high pressures occur near the edges and the ridge. Table 2 presents

the highest and lowest pressures near the edges of every model, from the records of pressure taps

while Table 3 presents the highest and lowest pressures near the ridge.

Table 2: Maxima and minima pressure coefficient near edges.

Model

A B C D E F

Wind

direction Max Min Max Min Max Min Max Min Max Min Max Min

60º 1.8 -1.8 1.7 -2.0 1.6 -2.0 1.8 -1.8 1.9 -1.9 1.7 -2.0

75º 1.5 -1.3 1.4 -1.5 1.2 -1.4 1.5 -1.5 1.5 -1.5 1.3 -1.4

90º 1.1 -0.8 1.1 -0.9 1.2 -0.8 1.2 -0.9 1.2 -0.9 1.2 -1.0

Table 3: Maxima and minima pressure coefficient near the ridge.

Model

A B C D E F

Wind

direction Max Min Max Min Max Min Max Min Max Min Max Min

60º 0 -1.2 -0.3 -1.3 -0.2 -1.2 -0.1 -1.1 -0.2 -1.1 -0.2 -1.3

75º 0.1 -1.2 -0.3 -1.2 -0.2 -1.2 -0.2 -1.1 -0.3 -1.2 -0.3 -1.2

90º -0.1 -0.8 -0.2 -0.9 -0.3 -0.9 -0.2 -0.8 -0.2 -0.8 -0.2 -0.9

FORCE COEFFICIENTS

The roof of the models were divided into four zones as shown in figure 9. Force coefficients

were determined by integrating the pressures on the four areas and on the whole roof.

43

yx

x

21

z

y

Reference area for global

horizontal actionson sections 1 and 3

Reference area

for global vertical

actions on sector 3

A = A

A =Y

Z Z

3

1 3

Fig. 9: Reference areas.

As force coefficients are associated with directions, a convention on the reference axes must be

formulated. The conventional axes most commonly used in aerodynamics are wind axes and

body axes. In this work, body axes are used, as displayed in figure 9. The force coefficients are

then defined as

i

ii

yref

yy

Aq

FC = i = 1,2,3,4 (1)

i

ii

zref

zz

Aq

FC = i = 1,2,3,4 (2)

yref

yy

Aq

FC = (3)

zref

zz

Aq

FC = (4)

Where iyC is the partial force coefficient on zone “i” in y-direction (horizontal), izC is the

partial force coefficient on zone “i” in z-direction (vertical), iyF is the force on zone “i” in y-

direction, izF is the force on zone “i” in z-direction, iyA = f2

a is the vertical reference area for

zone “i”, izA = ba

2 is the horizontal reference area for zone “i”, yC is the total force coefficient

in y-direction, zC is the total force coefficient in z-direction, yF is the force on the whole roof in

y-direction, zF is the force on the whole roof in z-direction, yA = a.f is the total vertical

reference area and zA = a.b is the total horizontal reference area.

The reference dynamic pressure, qref, was measured at the eave eight, as described before.

Table 4 shows the force coefficients acting on every zone (partial coefficients) and the overall

roof (total coefficients). The force coefficients are positive when the action is directed in the

positive direction of the corresponding reference axis (fig. 9).

Table 4: Force coefficients.

h/b = 0.4

Parcial Total Parcial Total θ º

1yC 2yC

3yC 4yC yC

1zC

2zC 3zC

4zC zC

60 0.31 0.43 0.65 0.33 0.86 -0.03 -0.02 0.73 0.36 0.26

75 0.43 0.50 0.54 0.35 0.91 -0.08 -0.10 0.59 0.39 020

A

90 0.35 0.35 0.45 0.45 0.80 -0.01 -0.01 0.50 0.50 0.26

h/b = 0.27

60 0. 22 0.32 0.71 0.37 0.81 0.09 0.08 0.81 0.44 0.36

75 0.32 0.39 0.64 0.47 0.91 0.05 -0.02 0.73 0.55 0.34 B

90 0.40 0.40 0.52 0.52 0.92 0.00 0.00 0.60 0.60 0.30

h/b = 0.133

60 0.34 0.36 0.56 0.36 0.81 0.00 0.02 0.66 0.45 0.28

75 0.36 0.38 0.58 0.44 0.88 0.02 0.03 0.68 0.54 0.32 C

90 0.35 0.35 0.50 0.50 0.85 0.06 0.06 0.62 0.62 0.34

h/b = 0.4

60 0.36 0.51 0.78 0.31 0.98 -0.08 -0.04 0.89 0.35 0.28

75 0.40 0.53 0.71 0.32 0.98 -0.07 -0.09 0.79 0.37 0.25 D

90 0.51 0.51 0.47 0.47 0.98 -0.12 -0.12 0.52 0.52 0.20

h/b = 0.27

60 0.34 0.62 0.78 0.32 1.03 -0.07 -0.13 0.89 0.36 0.26

75 0.39 0.58 0.74 0.35 1.03 -0.04 -0.13 0.80 0.40 0.26 E

90 0.49 0.49 0.50 0.50 0.99 -0.10 -0.10 0.55 0.55 0.23

h/b = 0.133

60 0.31 0.55 0.85 0.35 1.03 0.01 -0.08 0.96 0.43 0.33

75 0.38 0.48 0.75 0.43 1.02 0.01 0.02 0.85 0.52 0.34 F

90 0.49 0.49 0.55 0.55 1.04 -0.08 -0.08 0.64 0.64 0.29

If the yC coefficients of short and deep models are compared, it can be observed that in general

terms the ones corresponding to short models are about 10% higher for any h/b relationship. This

difference becomes even more evident between models B and E, where the disparity is 21%.

It is interesting to view figures 10 and 11, which have been built with values extracted from

Table 4. Figure 10 summarizes the variation of the total force coefficient in the y-direction for all

models in regard to θ and h/b. The coefficients of the A, B and C models are at the left of the

figure, clearly separated from those of the D, E and F models. The yC on the short models are

higher than the corresponding ones of the deep models for every value of θ and h/b, and in this

case they are not sensitive to θ and h/b, the values range between 0.98 and 1.04. Deep models

proved to be more sensitive, being the most severe case when the wind direction is 75º.

In figure 11 the variation of the total force coefficient in the z-direction is shown. In this

example, even though the curves follow the same trend but are inverted, they are not as clearly

split up as the yC values in fig. 10. For h/b = 0.27 the values of zC are higher on deep models,

but in the other two cases they are merged with the ones of the short models. Note that for θ =

90º the variation is practically linear with h/b.

In the previous section, it was pointed out that in the central area of the deep models roofs,

external pressures near the upwind eave are smaller than those of short models. This causes the

net pressure on that area to be smaller on deep roofs, which contributes to both decreased yC

and increased zC in regard to the short models. On the other hand, as internal negative pressures

are lower (that is bigger in absolute value) near the ridge of the deep models, they contribute to

decrease both yC and zC . As both effects add up to influence on yC , they explain the clear

separation in figure 10, and as they have an opposite influence on zC , they also explain the fact

that zC figures are a bit entangled in figure 11.

0,10

0,15

0,20

0,25

0,30

0,35

0,40

0,45

0,78 0,83 0,88 0,93 0,98 1,03Cy

h/b

60 deep 75 deep 90 deep 60 short 75 short 90 short

Fig. 10: Variation of the total force coefficient in the y-direction.

0,10

0,15

0,20

0,25

0,30

0,35

0,40

0,45

0,19 0,21 0,23 0,25 0,27 0,29 0,31 0,33 0,35 0,37Cz

h/b

60 short 75 short 90 short 60 deep 75 deep 90 deep

Fig. 11: Variation of the total force coefficient in the z-direction.

CONCLUSIONS

Marighetti et al. [5] demonstrated that pressures on VCR are different from pressures on planar

canopy roofs, for which reason it is a mistake to estimate the loads on VCRs by approximating

them with those on planar canopy roofs, as advised in some codes of practice. In this paper,

results of mean wind loads coefficients on vaulted canopy roofs obtained in boundary layer

tunnel tests have been presented. These results constitute a firm base to assess wind loads on

VCRs, though it must be noted that tests have been limited to models with a relation rise/span f/b

= 0.20, therefore estimating loads for others relation f/b is not straightforward.

Positive and negative pressures reach important values near the edges and the ridge. Tables 2 and

3 show the maxima and minima pressure coefficients in those zones of all models and for all the

wind directions tested; they can be used to verify cladding and structural parts. Global

coefficients given in Table 4 establish the total force acting on the structure, which is necessary

information for the design of the immediately supporting structure.

Mean aerodynamic coefficients on VCRs are influenced by size relations. They vary with the

column height and the relation span/depth. The maxima mean net pressure coefficients are higher

on the short models, while the minima mean net pressure coefficients occur on deep models. The

absolute value of pressure coefficients show a general trend to increase as the column height

decreases. As for the global coefficients, yC values are higher on short roofs and zC values are,

mostly but not always, higher on deep roofs.

ACKNOWLEDGEMENTS

The authors wish to acknowledge the help of Jose A. Iturri in making the models and preparing

the experimental set-up. The present work is part of an area of research supported by the

Facultad de Ingeniería de la Universidad Nacional del Nordeste, Argentina.

REFERENCES

[1] N.J. Cook, The designer's guide to wind loading of building structures, part 2: static structures,

Building Research Establishment Report, London, 1990.

[2] NV 65 Règles définisant les effets de la neige et du vent sur les constructions et annexes. Societé de

Diffusion des Techniques du Bâtiment et des Travaux Publics, France, 1970.

[3] CIRSOC 102: Acción del viento sobre las construcciones. Instituto Nacional de Tecnología

Industrial, Buenos Aires, 1983.

[4] CIRSOC 102: Reglamento Argentino de acción del viento sobre las construccione. Instituto Nacional

de Tecnología Industrial, Buenos Aires, 2005.

[5] J.O. Marighetti, O. Canavesio, B. Natalini, M.B. Natalini, Comparación entre coeficientes de presión

media en cubiertas aisladas planas y curvas, Memorias de las XVII Jornadas Argentina de Ingeniería

Estructural, Rosario, Argentina, 2002.

[6] M.B. Natalini, O. Canavesio, B. Natalini,M.J. Paluch, Wind tunnel modelling of mean pressures on

curved canopy roofs, Proceeding of the Americas Conference on Wind Engineering, Clemson, USA,

2001.

[7] M.B. Natalini,O. Canavesio,B. Natalini,M.J. Paluch, Pressure distribution on curved canopy roof,

Proceedings of The Second International Symposium on Advances in Wind and Structures, Pusan, Korea

2002.

[8] M.B. Natalini,C, Morel, B., Natalin,. Mean wind loads on vaulted canopy roofs under different wind

directions, Revista Sul-Americana de Engenharia Estrutural. 2 (2005) 29-40.

[9] G. Balbastro, V. Sonzogni, G. Franck, Simulación numérica del viento sobre una cubierta abovedada,

Mecánica Computacional. 24 (2005) 1261-1278.

[10] G. Balbastro,V. Sonzogni, Coeficientes de presión en cubiertas abovedadas aisladas, Anales de XIX

Jornadas Argentinas de Ingeniería Estructural, Mar del Plata, Argentina, 2006.

[11] G. Balbastro,V. Sonzogni, Simulación de un ensayo en túnel de viento aplicando CFD, Mecánica

Computacional, 26 (2007) 3779-3787.

[12] J.L.D. Ribeiro, Effects of surface roughness on the two-dimensional flow past circular cylinders,

Journal of Wind Engineering and Industrial Aerodynamics. 37( 1991) 299-309.

[13] J.C.K. Cheung, W.H. Melbourne, Turbulence effects on some aerodynamics parameters of a circular

cylinder at supercritical numbers, Journal of Wind Engineering and Industrial Aerodynamics. 14 (1983)

399-410.

[14] J.P. Batham, Wind tunnel test on scale models of a large power station chimneys, Journal of Wind

Engineering and Industrial Aerodynamics. 18 (1985) 87-91.

[15] H.M. Blackburn,W.H. Melbourne, The Effect of free-stream turbulence on sectional lift forces on a

circular cylinder, Journal of Fluid Mechanics. 306 (1996) 267-292.

[16] J. Blessmann, J., A.M. Loredo Souza, Açao do vento em coberturas curvas, Caderno Técnico CT-94,

Universidade Federal do Rio Grande do Sul, Brasil, 1998.

[17] J. Blessmann, Wind load on isolated and adjacent industrial pavilion curved roof. In: Wind Effects

on Buildings and Structures (Gramado, Brasil, 1998), Balkema, Rotterdam,1998. pp. 137-171.

[18] B. Natalini, O. Marighetti, M.B. Natalini, Wind tunnel modelling of mean pressures on planar

canopy roof, Journal of Wind Engineering and Industrial Aerodynamics. 90 (2002) 427-439.

[19] A.P Robertson, R.P. Hoxey, P. Moran, A full scale study of wind loads on agricultural ridged

canopy roof structures and proposal for design, Journal of Wind Engineering and Industrial

Aerodynamics. 21 (1985) 167-205.

[20] B. Natalini, O, Marighetti, M.B. Natalini,Wind tunnel test of planar canopy roof model. Memorias

de las XXIX Jornadas Sudamericanas de Ingeniería Estructural, Punta del Este, Uruguay, 2000.

[21] A.R. Wittwer, S.V. Möller, Characteristics of the low-speed wind tunnel of the UNNE, Journal of

Wind Engineering and Industrial Aerodynamics. 84 (2000) 307-320.

[22] H.P.A.H. Irwin, The design of spires for wind simulation, Journal of Wind Engineering and

Industrial Aerodynamics. 7 (1981) 361-366.

[23] N.J Cook,. Determination of the model scale factor in wind-tunnel simulations of the adiabatic

atmospheric boundary layer, Journal of Wind Engineering and Industrial Aerodynamics. 2 (1977/1978)

311-321.

[24] M.E. De Bortoli, B. Natalini, B., M.J. Paluch, M.B. Natalini, Part-depth wind tunnel simulations]of

the atmospheric boundary layer, Journal of Wind Engineering and Industrial Aerodynamics. 90 (2002)

281-291.

[25] M.B. Natalini, Acción del viento sobre cubiertas curvas aisladas, Tesis de Doctorado, Universidad

Nacional del Nordeste, Argentina (in Spanish), 2005.