You are on page 1of 50

JOURNAL OF WIND AND ENGINEERING

Vol. 10 No. 2 July 2013

CONTENTS

1. Wind Induced Mean Interference Effects on Tall Rectangular Buildings 1-17


Achal Kr. Mittal, Nikhil Agrawal, V. K. Gupta

2. A Dual Aerodynamic Data Analysis Framework for tall Buildings 18-40


Workamaw P. Warsido, Girma T. Bitsuamlak

3. Statistical Modelling of Key Cyclone Parameters for the East 41-49


Coast of India
Pradeep K. Goyal , T.K. Datta
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17

WIND INDUCED MEAN INTERFERENCE EFFECTS ON TALL


RECTANGULAR BUILDINGS

1 2 3
Achal Kr. Mittal , Nikhil Agrawal , V. K. Gupta
1
Principal Scientist, Structural Engineering Group, CSIR-CBRI Roorkee (achal_cbri@rediffmail.com)
2
Research Scholar, Department of Civil Engineering, IIT Roorkee (nikky1827@gmail.com)
3
Professor, Department of Civil Engineering, IIT Roorkee (vkgsufce@iitr.ernet.in)

ABSTRACT

Interference effect for two tall rectangular buildings has been studied in a boundary layer wind
tunnel using an aeroelastic model of the principal building and a rigid model of the interfering
building, placed at different upstream and downstream positions with respect to the principal
building. Testing has been carried out in a simulated flow expected over a suburban terrain and the
results of mean along-wind interference effects between the buildings have been presented in this
paper and dynamic response results are presented in a companion paper. Discussion on the effect
of height ratio, reduced velocity, and orientation of principal as well as interfering building model on
the interference factor has been included. Both the shielding and amplification due to interference
are found to be predominant, when the heights of the interfering building range from 67% to 150% of
the height of the principal building. An increase of 100% in mean moment is observed at side-by-
side arrangement in case of the interfering building. Interference influence zones considering the
variations of height ratio, reduced velocity and orientations of principal/interfering building are also
include in this paper.
Keywords: Interference effects, Mean moments, Tall buildings, Wind tunnel test, Height ratio,
Orientation, Velocity ratio

NOTATION

H Height of building (m) Power law index


Ur Reduced velocity HR Height ratio
Mean wind velocity at Natural frequency of principal building
UH nx
top of building (m/s) (Hz)
Iu Turbulent intensity (%) IF Interference factor
SAB Short afterbody LAB Long afterbody
Usual meaning of notation unless/otherwise mentioned in the text

INTRODUCTION
Preliminary assesment of wind loads on buildings is carried out mainly by using wind codes and
standards, whose specifications are generally based on wind tunnel tests, performed on isolated
structures in an open terrain. However, it has been brought out by several researchers that wind
loads on buildings in realistic environments may be considerably different from those measured on
isolated buildings due to the presence of adjacent building(s). The existing adjacent building(s) may
Wind Induced Mean Interference Effects on Tall Rectangular Buildings 2

either decrease or increase the wind loads on a building, depending on various geometrical,
structural and wind parameters, including size, cross-section shape, relative position of the
buildings, reduced wind velocity, number of the adjacent building(s), upstream terrain condition,
orientation of existing and proposed building and so on. The phenomenon, commonly known as the
interference effect, is very complicated and must be evaluated properly (Khanduri et. al. 1998).
For a pair of buildings of equal size in tandem arrangement, Sakamoto and Haniu (1988) found that
the drag force of the downstream building reduced to zero when the upstream building was three
times the building breadth away (centre to centre spacing) from the downstream building, and the
mean drag force could be negative when the spacing was less than the critical distance. The
shielding effect decreases with the increase of the spacing between the two buildings.
However, Taniike (1992) found that the shielding effects could be still noticeable when the upstream
building was located at a place 16 times of the building breadth away from the downstream building.
The results show a mean interference factor of 0.8, or, a shielding of 20% of mean wind loads on the
principal building at above location.
Some studies aimed at providing the general recommendations on the wind-induced interference
effects. English (1993) suggested a third order regression polynomial to predict the shielding
factors for mean interference effects in case of a pair of rectangular prism in tandem arrangement.
Khanduri et. al. (1998) summarized research developments in the area of wind-induced
interference effects and English and Frick (1999) as well as Khanduri et. al. (2000) applied a well-
trained artificial neural network (ANN) to predicate the interference effects on the design loads for
buildings located in a variety of geometrical configurations and boundary layer flows.
Xie and Gu (2004) made detailed discussions on the mean interference factor (MIF) between two
and among three tall buildings by a series of wind tunnel tests. They considered the effects of breath
as well as height ratio (ratio of height of the interfering building to those of principal building, HR) of
building and terrain category. The authors mainly focused on the behaviour of the MIF among three
tall buildings. Larger upstream building(s) could produce more shielding effects on the principal
building and meanwhile, side-by-side interfering buildings with larger size can produce more
serious channelling effect on the principal building. The results show that the interfering building
with height ratio equal to 0.5, produced insignificant shielding effects; on the other hand, HR >1.25
produce almost same value of interference factor. So the mean interference effects may only be
sensitive to the height ratio in the range 0.5HR1.25.

Amin and Ahuja (2009) studied the mean interference effects between a pair of buildings located in
a geometrical configuration of 'L' and 'T' plan shapes. The study detailed the arrangement of
buildings, wind incident angles and level of interference. Pressure measurements were restricted to
open country flow type.
While the interference phenomenon has been investigated, there is still a shortage of information
on the interference effects between two tall rectangular buildings arranged in different orientations,
such that it is not good enough for generalisation. Present paper focuses on the mean interference
effects between two tall rectangular buildings for various parameters i.e. Ur, HR, relative position of
interfering building, and orientation of principal as well as interfering building.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17 3

Furthermore, only the mean moment in the along-wind direction is considered in this study as it
affects safety of building due to the direct impact of the wind, and form the basis for other responses
i.e. rms and peak. The study of dynamic response between two tall buildings will be discussed in
another paper.

This paper summarizes the significant findings of the mean interference effects based on wind
tunnel tests conducted on an aero-elastic model, under suburban terrain conditions.

DESCRIPTION OF EXPERIMENTAL PROGRAMME

Wind Tunnel and Wind Model


The wind tunnel tests were conducted in the Boundary Layer Wind Tunnel of Civil Engineering
Department at Indian Institute of Technology Roorkee (IITR). This is an open circuit, continuous
flow, suction type tunnel using single blower fan (125 HP) having a test section of 2.1 m 2.0 m size.
Wind characteristics corresponding to suburban terrain have been simulated based on Indian
Standard (IS 875 (Part 3)-1987). The power law exponent , of the mean wind velocity profile is
0.19. The measured turbulent intensity Iu, at the top of the building height (H=240m) is about 10%.
Measured longitudinal mean wind velocity profile, turbulent intensity profile, and turbulence
spectrum of the wind simulation are shown in Figure 1 (a-c).

Figure 1 Wind model properties: (a) Mean wind velocity profile (=0.19); (b) Turbulent
intensity profile; (c) Longitudinal turbulence spectrum, (Nikhil et al. 2012)

Building Models

In the present study, a single rigid lumped mass model, pivoted at the base, having two degrees of
freedom, described by Isyumov (1982) is adopted. This is a rectangular building model with
dimensions of 0.05x0.15x0.60m. Tests are conducted for two different orientations, namely, long
afterbody (LAB) and short afterbody (SAB). Orientation of building in which shorter dimension of
building faces the wind, with longer dimension along the wind, corresponds to LAB orientation and
when longer dimension of building faces the wind, with shorter dimension along the wind,
corresponds to SAB orientation. The length scale, velocity scale and frequency scale are also set
as 1/400, 1/4 and 100 respectively. Data are recorded for a sampling duration of 20 seconds
Wind Induced Mean Interference Effects on Tall Rectangular Buildings 4

(corresponding to 33.34 min. at full scale) at a sampling rate of 250 Hz per channel, i.e. 5000 data
points are taken for each channel. Base bending moments are measured in along and across wind
direction with the help of strain gauge transducers (cantilever type). However, only mean moment
results are presented in this paper.
The rigid models, referred to as the interfering building models, are used to produce interference.
Photograph 1 (a) and (b), show aeroelastic and rigid models used in this study, respectively.

(a) (c)

(b) (d)

Photo 1 (a) Aeroelastic model in the wind tunnel (b) Aeroclastic model bottom view
(c) Rigid model in SAB orientation (d) Rigid model in LAB orientation

Experimental Arrangements

Rigid model of the interfering building (I) is located at various locations upstream and downstream
from the aeroelastic model to provide different interference excitation or shielding effects on the
wind induced response of the principal building model (P). Experimental arrangements for all
configurations/orientations of principal as well as interfering building models are shown in Table 1.
Each orientation of principal building and interfering building model consists of different height
ratios of interfering building models, which varies from 0.6 to 1.5. The plan aspect ratio is fixed at
3:1.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17 5

Table 1.Experimental arrangement for the building models


Principal Interfering No. of Positions of the
Terrain Building Building interfering building
Case Height Ratio (H R )
Category Model (P) Model (I) model
Orientation Orientation
0.6 65
0.75 65
I SAB SAB 1 70
1.25 70
1.5 70
0.6 65
0.75 65
II SAB LAB 1 70
1.25 70
Suburban 1.5 70
(=0.19) 0.6 65
0.75 65
III LAB SAB 1 65
1.25 65
1.5 65
0.6 70
0.75 70
IV LAB LAB 1 70
1.25 70
1.5 70

The centre-to-centre spacing between the two models which, varies longitudinally, X, and, laterally,
Y, on the Cartesian coordinates are shown in Figure 2 (a-d). The orientation of principal as well as
interfering building is also displayed in the same figure.

(a). Case I: P=SAB, I=SAB


Wind Induced Mean Interference Effects on Tall Rectangular Buildings 6

(b). Case II: P=SAB, I=LAB

(c). Case III: P=LAB, I=SAB

(d). Case IV: P=LAB, I=LAB

Figure 2 X-Y Coordinate system for the interfering building locations as well as their
orientations for (a) Case I: P=SAB, I=SAB, (b) Case II: P=SAB, I=LAB, (c) Case III:
P=LAB, I=SAB, (d) Case IV: P=LAB I=LAB
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17 7

Incident Wind Velocity


Wind-induced responses of principal building are measured at a range of reduced velocity,
Ur=UH/nx*B, varies between 3 to 5 in SAB orientation and 9 to 12 in LAB orientation of principal
building. where, UH is the wind velocity at the top of the principal building which varies from 9.1m/s
to11.1m/s at model scale and 36.4m/s to 44.4m/s corresponding to full scale. B is the width of
principal building facing the wind. For LAB orientation of principal building B=b, where b is the
normalized width (0.05m).

Interference Factor (IF)


Most of the interference studies quantified interference effects in terms of interference factor (IF),
which was initially introduced by Saunders and Melbourne (1979). For mean responses, IF can be
defined as follows:

Mean base momentof a building with interfering buildingpresent


IF
Mean base momentof an isolated building

Hence, increase in the mean response due to interference effect is proportional to the value of IF
(given in above expression). Measured IFs are usually illustrated in the form of IF contours. The IF
contour is a two dimensional plan which reflects the variation of interference factor with the position
of interfering building. The critical locations for the interfering building and the extent of mean
interference effect can be easily found using these contours.

EXPERIMENTAL RESULTS AND DISCUSSIONS


The most striking characteristics regarding the effects of interference on mean along-wind loads is
the high degree of shielding (reduction in mean along moments) and significant amplification
(increase in mean moments) offered to the principal building by the interfering building.

Comparison of Results for Twin Buildings with Previous Studies:


English (1993) synthesized several existing wind tunnel results that were obtained for different
simulated terrains which concluded into a regression equation to predict the mean along-wind
interference factor of the downstream building for twin buildings in tandem arrangement. The
formula is given in the polynomial form as:

IF 0. 05 0 .65 Z 0 . 29 Z 2 0 . 24 Z 3 (1)

where Z = log[S*(H + B)/H*B], S is the clear spacing between the two buildings, B is the breadth of
the building, and H is the height of the principal building.
To check the reliability of the results of the present experiments, the IF's of two tandem-arranged
buildings under the suburban terrain (=0.19), are compared with equation (1) and presented in
Figure 3. The comparison shows a good agreement between the results obtained from present
Wind Induced Mean Interference Effects on Tall Rectangular Buildings 8

study and calculated from a regression equation (1), for locations (>5.5B). However, larger
differences are observed for close locations (5.5B). It can be seen that the differences of the IF's,
calculated using equation 1 and present experimental results, decrease with the increase of
building spacing.

Figure 3 Interference factors for two buildings in tandem-arrangement

Table 1 show the comparison of IF's (mean to mean ratio) of two identical buildings (HR=1.0),
located in SAB orientation. The results show a good consistency, although some differences
appear due to the use of different building configurations, flow conditions and testing techniques.

Table 1 Comparison of IFs for twin building configuration


Power Interfering Building at (X,Y)
Building
law
Authors Dimensio
exponent (0b, 2.5b) (1.5b, 2.5b) (1.5b, 3.5b) (1b, 4.5b) (3.5b, 4.5b) (6b, 4.5b) (-1b, 3b) (-2b, 4b)
ns Ratio
()
Yahyai et.
1:1.2:6 0.30 0.20 0.50 0.90 1.01 0.98 1.03 0.83 0.83
al. (1992)
Gupta
1:2:12 0.18 0.40 0.85 0.95 1.0 1.0 1.01 1.0 1.02
(1996)
Sakamoto
and Haniu 1:1:3 - 0.15 0.96 1.0 1.03 1.01 1.01 - -
(1988)
Taniike
and
1:1:4.5 0.14 - 0.95 1.0 - 1.02 1.02 - -
Inaoka
(1988)
Present
1:3:12 0.19 0.17 0.60 0.90 1.0 1.02 1.03 0.84 0.90
Study
Here b = 20m (full scale)

Results of Different Orientation of Principal and Interfering Building

Case I: Principal and Interfering Buildings in SAB Orientation:

The interference contours for the mean moments of the principal building under the interference
excitation from adjacent building of height H, in suburban terrain is shown in Figure 4 for reduced
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17 9

velocity (Ur= 3.8). It is evident that mean along-wind moment of the principal building is decreased
by the presence of an interfering building. Substantial reduction in mean moment is recorded when
the two rectangular buildings are in tandem arrangements (locations of interfering building along
the line passing through the centre of principal building in the direction of wind).
As the upstream interfering building moves in lateral direction, away from the longitudinal axis,
shielding effect from the interfering building is decreased, steadily. A downstream interfering
building (X=2.5b to 4.5b, Y=0 to 4.5b) has the insignificant effect on the response of the principal
building at upstream.
For shortest height interfering building (HR=0.60), 30% increase in mean base moments has been
recorded at X=8b. Maximum shielding factor of 90% has been observed when the interfering
building (HR=1.5) is located in front of principal building at a small separation distance (X=2.5b,
Y=0). However, in side-by-side arrangement, the mean moment is increased by 20%. A shielding of
50% at X=20b (400m), and 20% at X=32b (640m) has been noted for interfering building (HR=1.5).

Observations are also recorded when the interfering building is placed at downstream of the
principal building. Deviation in mean response is generally found within 20% for all downstream
interfering building locations (HR=0.6 to 1.5).

Figure 4 IF contours for mean bending moments at Ur= 3.8

Figure 5 Variation of IF with HR at various locations on (X/b, Y/b) coordinates

Variation of IF with respect to the HR of the interfering building, located at various positions is
presented in Figure 5. From Figure 5, it is noted that, shielding increases with the increase in HR for
Wind Induced Mean Interference Effects on Tall Rectangular Buildings 10

various upstream locations of interfering building, but for side-by-side and downstream locations,
there is an amplification with the increase in HR. As the interfering building travel laterally, shielding
decreases with the increase in HR.

Case 2: Principal Building in SAB and Interfering Building in LAB Orientation:


For the above orientation of principal as well as interfering building only shielding effect has been
observed for range of HR from 0.6 to 1.5, throughout the grid region. Maximum reduction of 55% in
mean moment is detected for identical interfering building (HR=1.0). No significant effect of variation
in Ur is noted.

Figure 6 IF contours for mean bending moments for HR=1.25

Shielding effect due to interfering building (HR=1.25) is presented by IF contours in Figure 6. At side-
by-side arrangement of interfering building a 10% reduction in moment is detected. Similarly, for
downstream locations neither shielding nor amplification (almost same value as in case of isolated
building) has been observed.

Figure 7 Variation of IF with HR at various locations on (X/b, Y/b) coordinates

Generally, shielding increases with the increase in HR upto 1.25, after that there is a decrease in the
shielding effect for HR upto 1.5 (Figure 7). Maximum variation of 30% in IF value is obtained due to
the presence of 1.5H interfering building, as it moves in lateral direction from longitudinal axis.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17 11

Case 3: Principal Building in LAB and Interfering Building in SAB Orientation:

For this orientation of principal and interfering building, both the effects i.e. shielding as well as
amplification in mean moment are observed. Maximum reduction of 90% and amplification of 83%
in mean moment is recorded for tallest interfering building (HR= 1.5).

Figure 8 indicate the shielding as well as the amplification due to the presence of interfering building
(HR=1.5) at Ur=11.7.

Figure 8 IF contours for mean bending moments at Ur=11.7

From the contours shown in Figure 8, amplifications have been observed in the dotted zone.

For tandem arrangement of interfering building maximum reduction of 80% in mean moment at
(5.5b, 0) is seen. It is interesting to note that even up-to a distance of 390m (full scale) 30% shielding
is present. Dotted region in Figure 8, show the amplification in mean moments due to the presence
of interfering building. In this region maximum amplification of the order of 80% at side-by-side
location (X=0.5b, Y=3.5b) and 30% at downstream location (X=-2.5b, Y=3.5b) of interfering building
has been noted. In general, as the interfering building moves away from the principal building either
longitudinally or laterally relative increase in amplification can be observed.

Figure 9 Variation of IF with HR of various locations on (X/b, Y/b) coordinates

Representative variation of IF with HR at various upstream and downstream locations is presented


in Figure 9. Shielding as well as amplification in mean moments with the increase in HR at various
Wind Induced Mean Interference Effects on Tall Rectangular Buildings 12

locations in upstream and side-by-side/downstream of principal building respectively is generally


observed from the experimental results.

Case 4: Principal and Interfering Building are in LAB Orientation:


When both the buildings are in the LAB orientation shielding as well as amplification in mean
moment are noted, but not as dominant as in Case 1 and Case 3. Interference contours for HR=1.0
at different Ur's are presented in Figure 10 (a-b) and interference contours for HR=1.25 at Ur=9.6 are
present in Figure 10 (c). From Figure 10 (a-b), marginal amplification of 10-20% and shielding of the
order of 20-30% with the increase in velocity, is noted. Similar trend has been observed for other HR
values at different Ur's. A change (around 75%) from shielding to amplification has been noted at
X=1.5b, Y=4.5b location, when HR is changed from 1 to 1.25. On the other hand, at X=9b, a reverse
change (around 25%) from amplification to shielding has also been observed.

Figure 10 IF contours for mean bending moments at (a) Ur=9.6 and


HR=1; (b) Ur=11.7 and HR=1; (c) Ur=9.6 and HR=1.25
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17 13

INTERFERENCE INFLUENCE ZONES (IIZs):

A concept of interference influence zones has been devised to present an overall simplified yet
comprehensive view of wind-induced interference effects. Structural designers are more
concerned about the amplification in the response due to interference. Therefore, maximum value
of amplification has been picked up for all the variations of HR, Ur and all the locations. They are
presented in Figure 11 (a-d). For Case II i.e. P=SAB, I=LAB, mostly there is shielding. Substantial
amplification for all other three cases is evident.
Maximum increase in mean moment has been observed at side-by-side locations of interfering
building. Significant increase upto 60% in mean moment at downstream of principal building has
been noted for Case III as well as Case IV. Similarly, at X=11.5b and Y=0, there is an increase of
40% in Case I as well as Case IV and 100% in Case III.
It is worthy to note (Figure 11-c) that there is an increase of 40-80% in IF values when X=19.5b
(390m) and Y varies from 2.5b (50m) to 3.5b (70m).
In most of the Cases, maximum response is recorded in side-by-side arrangement of interfering
building. Similar trend of mean response (i.e. strongest mean moments are developed when the
two square prisms are at lateral offset from the longitudinal velocity direction) had also been noted
by Reinhold et. al. (1977).
Wind Induced Mean Interference Effects on Tall Rectangular Buildings 14

Figure 11 Interference influence zones for mean bending moments for (a) Case I: P=SAB,
I=SAB, (b) Case II: P=SAB, I=LAB, (c) Case III: P=LAB, I=SAB, (d) Case IV: P=LAB I=LAB
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17 15

CONCLUSIONS:

The mean interference effects between two buildings with different height ratio, orientation,
reduced velocity and relative position have been studied by a series of wind tunnel tests in suburban
terrain. Comparison of present study and the existing results have been made to validate the
present study results. The main inferences from the study are summarized as follows:

1) For SAB orientation of principal as well as interfering building, an increase in shielding with the
increase in HR for upstream locations is observed except for buildings with HR=0.6 and 0.75 at
higher reduced velocity (4.6). However, for side-by-side and downstream locations
amplification in mean moments is noted. Generally, 5-10% amplification due to higher reduced
velocity is noted, however, the amplification upto 20% is also recorded at a few locations. It is
worthy to note that, for this particular orientation of buildings, 20% shielding even at 640m (full
scale) distance due to presence of 1.5H interfering building is also observed.

2) When principal and interfering buildings are set at SAB and LAB orientations, respectively,
mostly shielding is observed. Reduction in the mean moments with the increase in HR is noted.
Maximum 50% reduction in moment is found due to the presence of tallest interfering building
(1.5H), at X=2.5b. Increase of 5 to 10% in amplification due to increase in reduced velocity is
also observed.

3) For LAB orientation of principal and SAB orientation of interfering building significant amount of
shielding and amplification is noted. Maximum 90% shielding and 83% amplification are
observed for interfering buildings having HR=1.5 and at higher reduced velocity. An increase in
the amplification of 30% due to change in Ur by 20% is noted for HR=1.5.

4) For LAB orientation of principal and interfering buildings, maximum 50% shielding and also
50% amplification is noted at different locations, due to variations in HR. Further, an increase in
mean moment of the order of 20% is observed due to increase in Ur by 13%.

5) In the present study, a concept of interference influence zones (IIZs) has been devised to
present an overall simplified yet comprehensive view of wind-induced interference effects. In
all orientations (Case I, Case II, Case III and Case IV) considered in the study, Case III, is the
worst case giving the maximum interference factor value i.e. 2.0 in IIZ. Case IV and Case I
come next regarding the adverse effect of Interference.

ACKNOWLEDGEMENT

The authors are thankful to the Professor Prem Krishna for his time and support during the study.
Wind Induced Mean Interference Effects on Tall Rectangular Buildings 16

REFERENCES:

1. Ahuja, A. K., Dalui, S. K., Ahuja, R. and Gupta, V. K. (2005), Effect of interference on wind
environment around high-rise buildings, Journal of Wind and Engineering, Vol. 2, No. 1, pp.
1-8.
2. Amin, J. A. and Ahuja, A. K. (2011), Mean interference effects between two buildings: effects
of close proximity, Journal of The Structural Design of Tall and Special Buildings, Vol. 20, No.
7, pp 832-852.
3. Arunachalam, S., Lakshmanan N., Babu Ramesh, G., Rajan Selvi, S. and Selvam Panneer
(2006), Wind-induced interference studies on square building model, Journal of Wind and
Engineering, Vol. 3, No. 1, pp. 25-31.
4. Bailey, P. A. (1984), Interference excitation of tall buildings, M. Engg. Sc. thesis, school of
civil and mining engineering, University of Sydney, Nov. 1984.
5. English, E. C. (1993), Shielding factors for paired rectangular prisms: An analysis of along-
wind mean response data from several sources, International Proceedings of the 7th U.S.
National Conference on Wind Engineering, University of California, Los Angeles, CA, 193-
201.
6. English, E. C. and Frick, F. R. (1999), Interference index and its prediction using a neural
network analysis of wind-tunnel data, Journal of Wind Engineering and Industrial
Aerodynamics, Vol. 83, pp 567-575.
7. Gupta, Abhay (1996), Wind tunnel studies on aerodynamic interference in tall rectangular
buildings, Ph.D. thesis, University of Roorkee, Roorkee, India.
8. Indian Standards IS: 875 -Part 3 (1987), Code of Practice for Design Wind Loads (other than
Earthquake) for Buildings and Structures, Bureau of Indian Standards, New Delhi.
9. Isyumov, N. (1982), The Aero-elastic Modelling of Tall Buildings, Proceeding of
International Wind Symposium on Wind Tunnel Modelling, USA, pp 373-407.
10. Khanduri, A. C., Stathopoulos, T. and Bedard, C. (1998), Wind induced interference effects
on buildings A review of the state-of-the-art, Engineering Structures, Vol. 20, No. (7), pp
617-630.
11. Khanduri, A. C., Stathopoulos, T. and Bedard, C. (2000), Generalization of wind-induced
interference effects for two buildings, Wind and Structures, An International Journal, 3(4), pp
255-266.
12. Lee, B. E. and Fowler, G. R. (1975), The mean wind forces acting on a pair of square prisms,
Building Science, Vol. 10, pp 107-110.
13. Xie, Z. N. and Gu, M. (2004), Mean interference effects among tall buildings, Engineering
Structures, Vol. 26, pp 1173-1183.
14. Reinhold, T. A., Tieleman, H. W. and Maher, F. J. (1977). "Interaction of square prisms in two
flow fields", Journal of Wind Engineering and Industrial Aerodynamics, Vol. 2, pp 223-241.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 1-17 17

15. Sakamoto, H. and Haniu, H. (1988), Aerodynamic forces acting on two square prisms placed
vertically in a turbulent boundary layer, Journal of Wind Engineering and Industrial
Aerodynamics, Vol. 31, pp 41-66.
16. Saunders, J. W. and Melbourne, W. H. (1979). Buffeting effects of upstream buildings,
Proceedings of 5th International Conference on Wind Engineering, Fort Collins, Colorado,
USA, Vol. 1, pp 593-606.
17. Taniike, Yoshihito (1992), Interference mechanism for enhanced wind forces on
neighbouring tall buildings, Journal of Wind Engineering and Industrial Aerodynamics, Vol.
41-44, pp 1073-1083.
18. Taniike, Yoshihito and Inaoka, Hideki (1988), Aeroelastic behaviour of tall buildings in wake,
Journal of Wind Engineering and Industrial Aerodynamics, Vol. 28, pp. 317-327.
19. Yahyai M, Krishen, K., Krishna, P. and Pande, P. K. (1992), Aerodynamic interference in tall
rectangular buildings, Journal of Wind Engineering and Industrial Aerodynamics, Vol. 41-44,
pp 859-866.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40

A DUAL AERODYNAMIC DATA ANALYSIS FRAMEWORK


FOR TALL BUILDINGS
1 2
Workamaw P. Warsido , Girma T. Bitsuamlak
1
Department of Civil and Environmental Engineering, Florida International University, 10555 W Flagler
St, Miami, Florida 33174, USA (wwars001@fiu.edu)
2
Department of Civil and Environmental Engineering, WindEEE Research Institute, University of
Western Ontario in London, ON, Canada N6A 5B9 (gbitsuam@uwo.ca)

ABSTRACT

Aerodynamic wind tunnel testing has been the most commonly used method for estimating wind-
induced responses of tall buildings. Analysis of aerodynamic data for estimating wind-induced
responses of tall buildings can be carried out in the frequency or time domain. Frequency domain
analysis has been the most commonly used approach as it requires relatively less computational
time than carrying out the analysis in the time domain. The present study discusses the frequency
and time domain approaches for aerodynamic data analysis and compares the two analysis
approaches from accuracy and computational time perspectives. Moreover, the significance of
some traditionally used simplifying approximations on the accuracy of estimated wind-induced
responses is also assessed. The study showed that accuracy of results obtained from frequency
domain analysis could be significantly affected by the traditionally used simplifying approximations.
It was also learned that the enhanced computing capabilities of today's computers allow time
domain analysis to be carried out within a reasonable computing time.

Keywords: Wind-induced responses; Tall building; Frequency domain; Time domain.

INTRODUCTION
Tall buildings are among wind sensitive structures which could experience high wind-induced
effects. The degree of sensitivity is continuously increasing with the current trend of architectural
and engineering designs heading towards the construction of buildings with complex shapes using
lightweight materials. Hence, the methodologies used for estimating wind-induced responses of
such wind sensitive structures should meet high levels of accuracy. Although building codes and
standards have been widely used for estimating wind-induced responses of low-rise buildings,
limited wind load provisions are given for tall buildings. Wind tunnel studies are often recommended
to be conducted for most tall buildings and have been quite successful for estimating wind-induced
loads and responses (Cermak 2003; Isyumov 2004). Aerodynamic wind tunnel tests are the most
commonly conducted tests for tall buildings (Boggs 1992; Boggs and Peterka 1989). The two
aerodynamic tests conducted in most wind tunnel laboratories are the High Frequency Force
Balance (HFFB) and the High Frequency Pressure Integration (HFPI). In the HFFB test, base
bending and torsional moments are measured by mounting a building model on a high frequency
and sensitive balance (Tschanz and Davenport 1983; Xie and Irwin 1998). In the HFPI test,
pressure readings are collected by installing a large number of pressure taps on the building
A Dual Aerodynamic Data Analysis Framework for tall Buildings 19

envelope (Steckley et al. 1992; Ueda et al. 1994). The overall aerodynamic loads on the building
can be obtained by integrating the distributed forces obtained by multiplying the recorded pressure
time history with the tributary area assigned to each pressure tap. The data collected from
aerodynamic wind tunnel tests contains the dynamic effects resulting from the turbulence in the
wind. However, design wind effects on tall buildings should also consider dynamic wind effects
resulting from the building inertia. Hence, the aerodynamic data obtained from wind tunnel tests
should be analyzed together with the dynamic properties of the building to get wind effects which
contain dynamic effect of the building inertia. The numerical techniques used to carry out this
analysis can be classified into two as the frequency and time domain analysis approaches.
Theoretically, both the frequency and time domain analysis approaches should give the same
results for linear analysis provided additional approximations are not used in the solution process of
each approach. However, simplifying approximations are traditionally used, mostly in the frequency
domain approach, which could potentially affect accuracy of the analysis results. In this paper, a
dual aerodynamic data analysis framework which works in the time and frequency domains is
presented. The framework allows estimation of modal, resultant and peak values of various wind-
induced responses of a given tall building at each story level. The uncertainties caused by different
traditionally used simplifying approximations are also investigated with a case study on a 50 story
building. Moreover, the feasibility of carrying out time domain analysis is also assessed from the
computational time perspective.

AERODYNAMIC DATA ANALYSIS


Similar to other vibration problems the response of tall buildings to dynamic loads can be expressed
by the three motion parameters, namely, acceleration, velocity and displacement. The summation
of forces associated to each parameter is equated to the excitation force to get the following
governing equation of motion.

mu(t) cu (t) ku(t) p(t) (1)

where m, c and k represent mass, damping and stiffness matrices of the building while (t), (t),
u(t) and p(t) denote acceleration, velocity, displacement and wind load vectors respectively. The
sizes of these matrices and vectors depend on the total number of degrees of freedom defined for
dynamic analysis. Even though infinite degrees of freedom exist for tall buildings, the degrees of
freedom are usually minimized practicality to three, two translations (denoted here by subscripts x
and y) and one rotation (denoted here by subscript ), at each floor level using rigid floor idealization
(Clough et al. 1964). Hence, a total of 3n degrees of freedom are defined where n denotes number
of floors of the building. Eq. (1) can be written in expanded form as follows:

(2)

where m and I are n x n diagonal sub-matrices with lumped mass m and mass moment of inertia I of
each floor as diagonal elements respectively; cxx, cyy, cx, cy, c, kxx, kyy, kx, ky and k are n x n sub-
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 20

matrices; u x (t) , u y (t) , u (t) , u x (t) , u


y (t) , u (t) , u x (t) , uy (t) , u (t) , p x (t) , py (t) and p (t) are nx1
sub-vectors in the corresponding damping and stiffness matrices; ,,,,,,,,,,and are nx1 sub-vectors in
the corresponding acceleration, velocity, displacement and wind load vectors. Using classical
modal analysis Eq. (2) can be converted into a set of uncoupled linear second order differential
equations as

(t) Cq (t) Kq(t) P(t)


Mq (3)

Where M T m , C 2M and K 2M denote the generalized mass, damping and stiffness


matrices respectively; , and 2 are diagonal matrices having, j, j and j2 as diagonal elements
respectively; j, and j denote damping ratio and natural frequency of the jth vibration mode
respectively; P(t) Tp(t) denotes the generalized wind load vector; 1 , 2,...,k
denotes the modal matrix comprised of the mode shape vectors and q q1 , q2 ,..., qk denotes the
T

generalized displacement vector; q T q1 , q2 ,..., qk denotes the generalized velocity vector and
q T q1 , q2 ,..., qk Denotes the generalized acceleration vector. The subscript k denotes the total
number of vibration modes considered for analysis. The generalized displacement vector can be
related to the different responses of interest through the modal participation coefficients (Chen and
Kareem 2005). For instance, the modal participation coefficients of displacement, acceleration,
shear, bending and torsional moment responses can be obtained using Eqs. (4-8).

dlj ljs (s x, y, ) (4)

(5)
alj 2jljs (s x, y, )

n
vlj 2j mi ijs ( s x, y) (6)
i l

n
zi if l 1 (7)
mlj ( z ) 2j m iijs z ( s x, y )
il zi - z l 1 otherwise

n
tlj 2j I i ij
i l (8)

Where is modal participation coefficient; the subscripts d, a, v, m and t represent floor


displacement, floor acceleration, story shear, story bending moment and story torsion responses
th
respectively; z is bending moment arm; , ljx , ljy and lj are values of the j mode shape vector at
the lth floor in the x, y and degrees of freedom respectively. The generalized displacement vector
can be evaluated in the frequency or time domain. In the frequency domain analysis, the
generalized wind load time history is transformed into spectral data from which standard deviations
of the modal generalized displacements can be computed without solving the dynamic equation of
motion. In the time domain analysis, the dynamic equation of motion is solved using the generalized
wind load time history to get the time history of the modal generalized displacements. Computation
A Dual Aerodynamic Data Analysis Framework for tall Buildings 21

of the dynamic wind-induced responses in the frequency and time domain approaches is briefly
explained in the following subsections.

Frequency domain approach

Application of frequency domain analysis for wind-induced vibration problems could be dated back
to the 1950's. Liepmann (1952) used a frequency domain approach for computing wind-induced
buffeting loads on aircrafts. The application was later extended for civil engineering structures in the
1960's (Davenport 1961; 1963; 1964; Harris 1963; Vickery 1965; 1966). The first step in the
frequency domain analysis is to transform the wind load time history data into frequency domain.
Transformation of a given data from time domain to frequency domain can be carried out with a
Fourier transform. When the time domain data is converted into frequency domain, variance and
covariance of the statistical parameters cannot be computed in the conventional way instead they
are computed from the spectral density functions. Power spectral density of the generalized wind
load in Eq. (3) can be obtained from product of the frequency domain data with its conjugate as

S Pjj f Pj f Pj* f (9)

th
where Pj ( f ) and Spji are the j mode generalized wind load data in frequency domain and its power
spectrum respectively. The superscript * denotes the complex-conjugate operator. The cross
spectral density of the ith and jth mode generalized wind loads can be obtained by multiplying the
frequency domain data of one vibration mode with the complex-conjugate of another vibration
mode as follows:

S Pij f Pi f Pj f
*
(10)

th th
Where Pi ( f ) and Pj ( f ) are the i and j mode generalized wind loads in the frequency domain
respectively and Spji ( f ) is their cross spectrum. Spectral density functions of the generalized wind
loads can be related to the corresponding spectral density functions of the generalized
displacements through the mechanical admittance function H ( f ) as follows:

Sq f H j f S P f H *j f
jj jj
(11)

S qij f Hi f SPij f H *j f (12)

where the mechanical admittance function in the jth vibration mode is given by
1
H j f 2
f f
K j 1 2 i j (13)
f j f j


Where Kj and fj = j / 2 denote the jth mode generalized stiffness and frequency; i 1 indicates
imaginary part of the complex number. The power and cross spectral density functions can also be
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 22

obtained from the Fourier transforms of autocorrelation and cross correlation functions
respectively. Variance of the jth mode generalized displacement can be obtained from the area
under the power spectral density curve as

2
q jj S q jj f df H j f S Pjj f H *j f df (14)
0 0

Covariance between the ith and jth mode generalized displacements can also be computed as


q2ij Re Sqij f df Re H i f SPij f H *j f df (15)
0 0

where 'Re' denotes the real operator. Traditionally, the integrals in Eqs.(14) and (15) are evaluated
as the sum of two regions, which correspond to background and resonant response components.
The background component shows the dynamic response resulting from the turbulent nature of
wind alone while the resonant component shows the contribution from dynamic properties of the
building. The background component can be computed by removing the effect of dynamic
properties of the building from Eq. (14) as follows:

1 (16)
q2jjb SP f df
K 2j 0 jj

By subtracting Eq. (16) from Eq. (14) the resonant component can be computed as

q2jjr 2q jj q2jjb (17)

Sometimes, the resonant variance component is computed approximately by simplifying Eq. (14)
with the assumption of White Gaussian process and j in the ranges of frequencies which are
close to the natural frequencies of the building as follows:
1
q2jjr f S (f ) (18)
K 2j 4 j j Pjj j
The approximation used in Eq. (18) has been the basis for simplified equations recommended by
some building codes and standards for estimating resonant components of wind-induced response
of tall buildings. The investigation on the accuracy of White noise approximation on wind-induced
responses is presented in the latter section.
th
Standard deviation of any response of interest (R) in the j mode of vibration can then be obtained
from standard deviation of the generalized displacement as follows:

R R q (19)
jj j jj

Standard deviations of modal responses should be combined to get standard deviation of the
resultant response. The most simplified modal combination rule is the square root of sum of the
squares (SRSS) rule which is suggested by Rosenblueth (1951).
A Dual Aerodynamic Data Analysis Framework for tall Buildings 23

k k


2 (20)
R 2Rjj
j 1 j 1
Rj q jj

As it can be seen from Eq. (20) the cross correlation of modal responses is ignored in this approach.
This approach works well when the modal responses are uncoupled for which the natural
frequencies are well-separated or the ratio of natural frequencies of two consecutive vibration
modes is at least 1.2 (Tse et al. 2008). The most commonly used modal combination rule, which
considers the cross correlation of modal responses, is the complete quadratic combination (CQC)
rule given by
k k k k
R Rii R jj rij
i 1 j 1

i 1 j 1
qii q jj rij
Ri Rj
(21)

where rij is the correlation coefficient which is defined as follows:



Re H i f S Pij f H *j f df
q2ij 0
rij (22)
qii q jj

H f S f H f d f H f S f H f d f
* *
i pii i j p jj j
0 0

Different formulations were suggested for computing the correlation coefficient without directly
solving Eq. (22). The two most commonly used formulations were suggested by Chen and Kareem
(2005) and Huang et al. (2009).
Peak(minimum and maximum) value of any response of interest can be computed from the mean
and standard deviation components with a peak factor approach as follows (Davenport 1967):

R R g R R (23)

where R denotes the peak response; R denotes the mean response which can be computed as
follows (Chen and Kareem 2005):

k
Pj (24)
R Rj q j , qj
j 1 Kj

gR is the peak factor which can be computed with the following closed form equation in Eq. (25)
assuming the response time history to be a normally distributed stationary Gaussian process
(Davenport 1961).
0.5772
gR 2lnT (25)
2lnT
where T is the time interval in seconds over which the peak value is required and is the mean zero
upcrossing rate. The mean upcrossing rate can be computed as follows (Rice 1945):
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 24


(26)
f 2 SR f df S f df
R
0 0

Since spectra of the response R and the generalized displacement show the same variation with
frequency, Eq. (26) can be written as


(27)
f
2
Sq f df S f df
q
0 0

where Sq is a spectrum obtained by combining spectra of the generalized displacements in the


different vibration modes using the CQC modal combination rule as follows:

k k

S ( f )df
q
Sqii ( f )df

Sq jj ( f )df rij

(28)
0 i1 j1 0 0

Similarly, the numerator term in Eq. (27) can be computed as

k k
2
f 2 S q ( f )df f 2 S qii ( f )df f S q jj ( f )df rij (29)

0 i1 j1 0 0

where rij is given by



Re f 2 Sqij f df
rij 0 (30)


f 2 S qii f df f 2 S q jj f df

0 0

Traditionally, the peak response is computed approximately after rearranging Eq. (23) as follows:

(31)
R R g g
b Rb
2
rj R rj
2

where gb and grj are the peak factors corresponding to the background and jth mode resonant
response components respectively. The back ground peak factor is commonly given a value
between 3 and 4(Chen and Kareem 2005). The resonant peak factor for each vibration mode is
computed from Eq. (25) by approximating the mean upcrossing rate parameter with the building
natural frequency (j fj) where fj is frequency of the jth vibration mode in Hz (Chena and Huang
2009).
A Dual Aerodynamic Data Analysis Framework for tall Buildings 25

Time domain approach

In the 1970's and 80's, when the computer processing-speed was limited, analysis of aerodynamic
data had been mostly carried out in the frequency domain owing to its smaller computational
demand compared to the time domain approach. However, with the fast growth of computing
speed, memory and storage capacity of computers, carrying out time domain analysis is becoming
increasingly favorable as it has been underscored by some recent research works (e.g. Flay and Li
2007; Simiu et al. 2008; Spence et al. 2008; Tse et al. 2008). In the time domain analysis, time
history of wind-induced responses are obtained by using the time history of wind load data,
measured in wind tunnel tests, as input in the dynamic equation of motion. Time history of the
generalized displacement vector can be obtained by directly solving Eq. (3) with numerical time
stepping methods. Several numerical time-stepping methods have been developed over the years
which can be generally categorized into two schemes, namely explicit and implicit. In explicit
schemes unknown value of a function can be explicitly defined in terms of known values of a
function, while in implicit schemes coupled system of equations, which contain both unknown and
known values of a function, need to be solved. Explicit schemes are easy to implement compared to
implicit schemes but require relatively smaller time steps to ensure convergence and stability of the
numerical solutions. Implicit schemes involve solving of a system of equations at each time step,
hence they require extra computational effort compared to explicit schemes. In the present study,
time domain analysis is carried out using the Newmark's linear acceleration method which is an
implicit scheme. Since the mathematical background of this numerical scheme is available in
published literatures (e.g. Chapra and Canale 2009; Chopra 2007) it is not presented here. Once
time history of the generalized displacement is obtained, time history of the required modal
response of interest can be obtained using the modal participation coefficient as follows:

R j R j q j (32)

Modal responses can then be directly superimposed to get the resultant time history of response.
Moreover, since time history of a wind-induced response is obtained, the statistical parameters
such as mean and standard deviation can be computed in the conventional way. The resultant time
history also contains peak responses. However taking instantaneous peak values directly from the
time history may not be a good peak estimation approach since such measured peak values may
not be statistically stable. A more stable peak estimation approach would be one that makes use of
information from whole of the time history data. Sadek and Simiu (2002) suggested a peak
estimation approach for low-rise buildings for which the time histories of wind-induced responses
are generally non-Gaussian. The approach involves fitting a combination of gamma and normal
distributions to the time history data together with a standard translation processes approach
(Grigoriu 1995). In the present work, normal probability distribution is fitted to the time history of
wind-induced responses following a similar approach. Goodness of fit is checked by the probability
plot correlation coefficient (PPCC) method (Filliben 1975). Moreover, acceptability of the normal
distribution fit to the time history of wind-induced responses is also checked from the probability and
cumulative distribution function plots. The mathematical formulation of this approach is discussed
as follows.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 26

For a standard Gaussian process X(t) the cumulative distribution function of the largest peak X in a
certain time interval T is given as (Rice 1954):

exp T exp X /2
F X X X
2 (33)

Rearranging Eq. (33) the peak value can be expressed in terms the cumulative distribution function
as:

X T (34)
X 2 ln
ln F X
X


Hence, for a given value of the cumulative distribution function FXk X the maximum and minimum
peak values can be computed as:

X T X T (35)
X max
k
2 ln and X min
k
2 ln
X X
ln F k

ln F k X
X


where X is the mean zero upcrossing rate for process X, which can be computed from Eq. (27).
Peak values of the process X corresponding to a set of predetermined values of the cumulative
distribution function can then be mapped to the corresponding peak values of the process R (time
history of the response of interest) as follows:

R max
k
X max
k
and Rmin
k
X min
k
(36)

Where and are slope and intercept parameters of the least squares regression line obtained by
fitting inverse of the standard normal cumulative distribution to the sorted time history of response.
The final estimated peak values of the response are then computed as the average of peak values
obtained in Eq. (36).


R max f Xk X R max
k

and R min f Xk X R min
k
(37)
X X


where f Xk X is the probability density function of the peak X in a certain time interval T, which can
be computed as:

X T k

f Xk X 2 ln
F X ln F k X (38)
ln F k X
X X

X

CASE STUDY

Experimental setup

HFPI wind tunnel test was carried out on a 50 story steel building which is 182.88m (600ft) high with
a rectangular plan dimensions of 30.48m (100ft) x 45.72m (150ft) (Fig. 1(a)), similar to the
A Dual Aerodynamic Data Analysis Framework for tall Buildings 27

Commonwealth Advisory Aeronautical Research Council (CAARC) building (Melbourne 1980).


The building model was constructed to a 1:400 scale and tested at RWDI's boundary layer wind
tunnel facility (Fig. 1(b)). The tunnel has a cross-section of 2.13m x 2.44m (7ft x 8ft) at 13.3m (43.5ft)
distance down-stream of the tunnel entrance, where the building model was installed on a turn
table. The test was carried out in an open upwind exposure which was generated by the combined
effect of 15.24cm x 33.02cm (6in x 13in) trapezoidal spires and 2.54cm (1.0in) high triangular floor
roughness elements. The mean wind speed and turbulence intensity profiles as well as longitudinal
turbulence spectrum of the simulated open exposure are shown in Fig. 2. Pressure readings were
taken for several wind directions using 280 pressure taps installed on the building surface at 10
levels with 28 taps at each level. The data was collected for a duration of 36sec at a sampling
frequency of 512Hz which is equivalent to 1hr in the full scale. Since the pressure taps were
connected to the pressure scanner with 1.34mm (0.053in) PVC tubes the data collected was low
pass filtered to reduce resonance effects from the tubes (Irwin et al. 1979). Using the pressure data
and an assumed design wind speed of 56.7m/sec at the building height, the full-scale forces and
torsional moment are computed at each floor level.

Fig. 1 (a) Dimensions of the study building in m and pressure tap locations;
(b) HFPI test setup at RWDIs boundary layer wind tunnel laboratory
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 28

Fig. 2 (a) Normalized mean wind speed (U/Uref) and turbulence intensity (TI) profiles and
(b) longitudinal turbulence spectrum at the building height

Classical modal analysis

Classical modal analysis was carried out for the building model using finite element analysis
software (see Fig. 3). The building was designed in such a way that it will have coupled vibration
modes by deliberately orienting the elevator core asymmetrically on the building plan as shown in
Fig. 3(b). Natural frequencies for the first three vibration modes were found to be 0.280 Hz, 0.291 Hz
and 0.562 Hz respectively and the corresponding mode shapes are shown in Fig. 4. It has to be
noted that the torsional mode shape was plotted together with the sway modes after being
multiplied by the radius of gyration (ra) at each floor level. As it can be observed from Fig. 4 as well
as from the natural frequency values, the first two vibration modes are coupled. A 1% damping ratio
is assumed in all the three vibration modes. The aerodynamic data coupled with dynamic properties
of the building was analyzed in the frequency and time domains to obtain wind-induced responses.

Fig. 3 FEM model of the study building: (a) 3D view and (b) typical plan view
A Dual Aerodynamic Data Analysis Framework for tall Buildings 29

Fig. 4 Normalized (a) first, (b) second and (c) third vibration modes

RESULTS AND DISCUSSION


Modal responses
Standard deviations of the generalized displacements corresponding to the three vibration modes
are computed in the frequency and time domain analysis approaches. In the frequency domain
analysis the modal responses are computed with two approaches. In the first approach, resonant
components are computed with the White noise approximation using Eq. (18) and combined with
the background components obtained from Eq. (16). In the second approach, the modal
generalized displacements are computed without approximation by integration of the full spectrum
given in Eq. (14). Fig. 5 shows comparison of the modal generalized displacements obtained from
the frequency domain approaches, with the time domain results, for several wind directions. As it
can be observed, the frequency domain results obtained by integration of the full spectrum closely
match the time domain results in all the three vibration modes. The variations observed between the
full spectrum frequency domain results and the time domain results are less than 5% for all wind
directions. The slight percentage deviations observed between the two approaches can be
attributed to the approximation in the process of numerical integration. The frequency domain
results obtained with the White noise approximation show relatively larger deviations from the time
domain results. As it can be observed, as high as 30% deviation is observed in some wind
directions.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 30

Fig. 5 Standard deviation of generalized displacements in the three vibration modes obtained from
frequency domain (FD) and time domain (TD) approaches (a, c and e); Percentage deviations of
frequency domain results from time domain results(b, d and f)
A Dual Aerodynamic Data Analysis Framework for tall Buildings 31

Modal correlation coefficients


Fig. 6 shows comparison of modal correlation coefficients obtained from the CQC formulations of
Chen and Kareem (2005) and Huang et al. (2009) with the exact correlation coefficients obtained by
directly solving Eq. (22). The correlation coefficients obtained from the formulation of Huang et al.
(2009) show perfect match with the exact correlation coefficients. However, the coefficients
obtained from Chen and Kareem (2005) show some deviations particularly between the highly
coupled first and second modes. The correlation coefficients between the remaining uncoupled
pairs show good agreement. If standard deviations of the generalized displacements are computed
by integration of the full spectrum given in Eq. (14), it will be more convenient to compute the exact
correlation coefficients directly from Eq. (22). This is because the standard deviation values are
already available and the only extra parameter that needs to be computed is the covariance
parameter.

Fig. 6 Modal correlation coefficients (a) r12 (b) r13 and (c) r23
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 32

Resultant responses

Comparison of the standard deviation of resultant base moments obtained from frequency and time
domain analyses are shown in Fig. 7. In the frequency domain analysis, standard deviations of the
modal base moments are computed from integration of the full spectrum and they are combined
with the SRSS and CQC combination rules. In the time domain analysis, standard deviations of the
resultant base moments are computed in the conventional way from the time history of base
moments. As it can be observed from the figures, frequency domain results obtained with the SRSS
rule show relatively larger deviations from the time domain results. For most wind directions the
SRSS resulted in 10% deviations and up to 20% deviations are observed for few wind directions.
Results obtained from the different CQC formulations show close agreement with time domain
results. For most wind directions, the percentage deviations for the different CQC formulations fall
below 5%, while for few wind directions, up to 10% deviations are observed. Larger deviations in the
SRSS rule are observed particularly in the bending moment responses (Mx and My) for which the
major contributions come from the coupled first and second vibration modes. Hence, the variations
observed in the SRSS rule result from ignoring the cross-correlation effects between the first and
second mode responses. The major contribution for the torsional moment response (M) comes
from the third vibration mode which is not coupled with the other two modes. Hence, results
obtained from the SRSS and the CQC rules are relatively close since the cross-correlation effects
are minimal.
A Dual Aerodynamic Data Analysis Framework for tall Buildings 33

Fig. 7 Standard deviation of resultant base moments (a, c and e); Deviations of base moments
obtained with different modal combination
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 34

Peak responses

In the frequency domain analysis, peak values of the base moment responses are computed with
two approaches. In the first approach, the peak factor is computed by approximating the mean
upcrossing rate by the fundamental natural frequency of the building. In the second approach, the
peak factor is computed using the mean cycling rate computed using Eq. (27). For the sake of
convenience, the first approach will be referred as Traditional Peak Factor (TPF) approach and the
second approach will be referred as Accurate Peak Factor (APF) approach. In the time domain
analysis also two sets of peak values are obtained. The first set of peak values referred as
estimated peak values are computed using the statistical approach discussed before, while the
second set of peak values referred as measured peak values are directly taken from the time history
of responses. For the estimated peak responses, the probability plots and cumulative distribution
function plots of the normal distribution fitted to the time history of base moment responses are
shown in Fig. 8. As it can be observed from the plots as well as from the PPCC values, which are
very close to log normal distribution fit to the data is acceptable. The peak bending and torsional
base moments obtained from frequency and time domain analysis are compared in Fig. 9. It can be
observed from the figure that the peak values obtained from the TPF frequency domain approach
are relatively larger than the peak values obtained from the APF frequency domain approach. It can
also be observed that, the peak values obtained with the APF frequency domain approach and the
estimated peak values from time domain analysis show close agreement. The peak responses
obtained from the TPF frequency domain approach are larger than the estimated peak values by
about 15% for most wind directions but reach as high as 25% for few wind directions. Peak values
from the APF frequency domain approach show less than 5% deviations from the estimated peak
values for most wind directions. The measured peak responses in the time domain show random
percentage deviations in different wind directions with magnitudes falling in a wide range of 0-30.
This indicates the relatively unstable nature of measured peak values compared to the estimated
peak values.
A Dual Aerodynamic Data Analysis Framework for tall Buildings 35

Fig. 8 Probability plots (a, c and e) and Cumulative distribution function plots (b, d and f) for base
moment responses at 180o wind direction
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 36

Fig. 9 Absolute peak base moments (a, c and e); Deviations of moments obtained with different
approaches from statistically estimated peak moments in the time domain (b, d and f)
A Dual Aerodynamic Data Analysis Framework for tall Buildings 37

CPU time

The computational time required to carry out time domain analysis is generally expected to be
longer than the time required for frequency domain analysis. In this section the CPU time elapsed
for carrying out the analysis in the time and frequency domains are compared. The case study is
carried out on a laptop computer having 4.0GB RAM and dual core processors running at 2.2 GHz.
It is to be noted that the codes for both analysis techniques are developed in a similar programming
language. Moreover, common programming tasks are handled with similar algorithms in both
analysis techniques. In addition, to assure repeatability of the results obtained, the analysis was
carried out for multiple wind directions. Fig.10 shows a comparison of the CPU times elapsed in
time and frequency domain approaches to compute mean, standard deviation and peak values of
displacement, acceleration, shear, bending moment and torsion responses at each floor level. It
has to be noted that the extra CPU time required for exporting analysis results in the required output
formats are not included in the time values. It can be observed from the figure that to analyze
aerodynamic data from one wind direction it took about 15sec in the frequency domain while it took
about 57 sec in the time domain. Hence, although the CPU time taken for time domain analysis
seems relatively large compared to the CPU time taken for frequency domain analysis it is still
efficient to carryout time domain analysis.

Fig. 10 CPU time elapsed for analyzing aerodynamic data in time and frequency domains

CONCLUSION

An aerodynamic wind tunnel test was carried out on a 50 story building which has the same
dimensions as the CAARC building and the data collected was analyzed in time and frequency
domains. The results obtained were compared at different stages of the solution process starting
from the computation of modal responses up to the estimation of peak responses. Comparison of
standard deviations of modal responses revealed that frequency domain results obtained without
simplifying approximations match time domain results more closely than the results obtained with
simplifying approximations. In the frequency domain analysis, standard deviations of modal
responses were combined with the SRSS and CQC modal combination rules and the resultant
responses were compared with the corresponding time domain results. Results obtained from the
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 38

SRSS rule showed relatively larger deviations from time domain results compared to results
obtained from the CQC rule. A statistical peak response estimation approach was implemented in
time domain analysis which was found to be more stable than the measured peak values.
Moreover, a more accurate procedure was implemented to compute peak factors used in the
frequency domain analysis. Peak responses obtained with accurate peak factors in the frequency
domain showed close agreement with statistically estimated peak values in the time domain.
Hence, accuracy of results obtained from frequency domain analysis can be compromised by the
traditionally used simplifying approximations. The present work also attempted to assess feasibility
of time domain analysis from the computational perspective by comparing the CPU time elapsed for
time and frequency domain analysis. Although, the CPU time taken for time domain analysis may
seem relatively large compared to the CPU time taken for frequency domain analysis it is still
computationally efficient to carryout time domain analysis.

ACKNOWLEDGMENT

This research is based upon work supported by the National Science Foundation under Grant No.
0846811, any opinions, findings, and conclusions or recommendations expressed in this material
are those of the authors and do not necessarily reflect the views of the granting agencies.

REFERENCES

1. Boggs, D. W. (1992). "Validation of the aerodynamic model method." J. Wind Eng. Ind.
Aerodyn., 41-44, 1011 1022.
2. Boggs, D. W., and Peterka, J. A. (1989). "Aerodynamic model tests of tall buildings." J. Eng.
Mech., 15(3), 618-635.
3. Cermak, J. E. (2003). "Wind-tunnel development and trends in applications to civil
engineering." J. Wind Eng. Ind. Aerodyn., 91, 355370.
4. Chapra, S., and Canale, R. (2009). Numerical methods for engineers, 6th Ed., McGraw-Hill
5. Chen, X., and Kareem, A. (2005). "Dynamic wind effects on buildings with 3D coupled modes:
application of high frequency force balance measurements." J. Struct. Eng., 131(11), 1115-
1125.
6. Chena, X., and Huang, G. (2009). "Evaluation of peak resultant response for wind-excited tall
buildings." Eng. Struct., 31, 858-868.
7. Chopra, A. K. (2007). Dynamics of structures- theory and applications to earthquake
engineering, Third Ed., Pearson Prentice Hall, Upper Saddle River, New Jersey.
8. Clough, R. W., King, I. P., and Wilson, E. L. (1964). "Structural analysis of multistory
buildings." ASCE, J. Struct. Div., 90, 19-34.
9. Davenport, A. G. (Year). "The application of statistical concepts to the wind loading of
structures." Proceedings of the Institution of Civil Engineers, 449-472.
A Dual Aerodynamic Data Analysis Framework for tall Buildings 39

10. Davenport, A. G. (Year). "The buffeting of structures by gusts." Proceedings, International


Conference on Wind Effects on Buildings and Structures, Teddington, UK, 358-391.
11. Davenport, A. G. (Year). "Note on the distribution of the largest value of a random function with
application to gust loading." Proceedings of the Institution of Civil Engineers, 187-196.
12. Davenport, A. G. (1967). "Gust loading factors." J. Struct. Eng., 93, 11-34.
13. Filliben, J. J. (1975). "The probability plot correlation coefficient test for normality."
Technometrics, 17(1), 111-117.
14. Flay, R. G. J., and Li, Y. F. (Year). "On predicting accelerations in tall buildings from wind tunnel
model tests using spectral and time history methods." Proc. 12th Int. Conf. on Wind
Engineering, Cairns, Australia, 1215-1222.
15. Grigoriu, M. (1995). Applied non-Gaussian processes, Prentice Hall, Englewood Cliffs, N.J.
16. Harris, R. I. (Year). "The response of structures to gusts." International conference on wind
effects on buildings and structures, Teddington, UK, 394421.
17. Huang, M. F., Chan, C. M., Kwok, K. C. S., and Hitchcock, P. A. (2009). "Cross correlations of
modal responses of tall buildings in wind-induced lateral-torsional motion." J. Eng. Mech.,
135(8), 802-812.
18. Irwin, H. P. A. H., Cooper, K. R., and Girard, R. (1979). "Correction of distortion effects caused
by tubing systems in measurements of fluctuating pressures." Journal of Wind Engineering
and Industrial Aerodynamics, 5(12), 93-107.
19. Isyumov, N. (Year). "Improved performance of tall buildings through wind tunnel testing."
Structures Congress 2000: Advanced Technology in Structural Engineering Philadelphia,
PA, United states.

20. Liepmann, H. W. (1952). "On the application of statistical concepts to the buffeting problem."
J. Aeronaut. Sci., 19(12), 793-800.
21. Melbourne, W. H. (1980). "Comparison of measurements on the CAARC standard tall
building model in simulated model wind flows." J. Wind Eng. Ind. Aerodyn., 6, 73-88.
22. Rice, S. O. (1945). "Mathematical analysis of noise, Pt III." Bell syst. tech., 24, 46-156.
23. Rice, S. O. (1954). Mathematical analysis of random noise. Selected papers on noise and
stochastic processes, Dover, New York.
24. Rosenblueth, E. (1951). "A basis for aseismic design," PhD Thesis University of Illinois,
Urbana, Illinois.
25. Sadek, F., and Simiu, E. (2002). "Peak non-Gaussian wind effects for database-assisted low-
rise building design." J. Eng. Mech., 128(5), 530539.
26. Simiu, E., Gabbai, R. D., and Fritz, W. P. (2008). "Wind-induced tall building response: a time-
domain approach." Wind Struct., 11(6), 427-440.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 18-40 40

27. Spence, S. M. J., Gabbai, R. D., and Simiu, E. (Year). "Time-domain wind-tunnel based
methodology for tall building analysis and optimal design." The 4th International Conference
on Advances in Wind and Structures, Jeju, Korea.
28. Steckley, A., Accardo, M., Gamble, S. L., and Irwin, P. A. (1992). "The use of integrated
pressures to determine overall wind-induced response." J. Wind Eng. Ind. Aerodyn., 41-44,
1023-1034.
29. Tschanz, T., and Davenport, A. G. (1983). "The base balance technique for the determination
of dynamic wind loads." J. Wind Eng. Ind. Aerodyn., 13, 429-439.
30. Tse, K. T., Hitchcock, P. A., and Kwok, K. C. S. (2008). "A time domain technique for
aerodynamic wind tunnel model studies." J. Wind Eng., 5(2), 1-16.
31. Ueda, H., Hibi, K., Tamura, Y., and Fujii, K. (1994). "Multi-channel simultaneous fluctuating
pressure measurement system and its applications." J. Wind Eng. Ind. Aerodyn., 51, 93-104.
32. Vickery, B. J. (1965). "On the flow behind a coarse grid and its use as a model of atmospheric
turbulence in studies related to wind loads on buildings." National Physical Laboratory (UK).
33. Vickery, B. J. (1966). "On the assessment of wind effects on elastic structures." Australian
Civil Engineering Transactions, CE8, 183-192.
34. Xie, J., and Irwin, P. A. (1998). "Application of the force balance technique to a building
complex." J. Wind Eng. Ind. Aerodyn., 77&78, 579-590.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 41-49

STATISTICAL MODELLING OF KEY CYCLONE PARAMETERS


FOR THE EAST COAST OF INDIA
1 2
Pradeep K. Goyal , T.K. Datta
1
Dept. of Civil Engg. Govt. Engineering College, Ajmer, Rajasthan, India, 305001
Email:pk_goyal2003@yahoo.co.in
2
Department. of Civil Engineering, Indian Institute of Technology Delhi, India, 110016
Email: tushar_k_datta@yahoo.com

ABSTRACT

In this study, statistical distributions of key cyclone parameters are obtained by analyzing the
cyclone track records obtained from India meteorological department for east coast region of India.
The dataset of historical landfalling cyclone tracks is put into GIS with latitude /longitude and landfall
locations. The Statistical tests were performed to find a best fit distribution to the track data for each
key cyclone parameter for the states of Andhra Pradesh and Orissa.

Keywords: Tropical cyclones; Cyclone tracks; GIS; Simulation

INTRODUCTION

Cyclone hazard analysis in the coastal region is a crucial issue in pre cyclone disaster mitigation of
any country prone to cyclones. Cyclone wind field simulation techniques have emerged as the most
reliable method for predicting design wind speeds and cyclone hazard risk analysis for the area of
interest in cyclone-prone regions. Therefore, it is important to obtain the probability distribution of
key cyclone parameters for cyclone hazard risk analysis and vulnerability assessment for the
region of interest. Key Cyclone parameters are defined as the parameters which are used in a
cyclonic wind field model to estimate wind speeds.

These key parameters are (i) central pressure difference p, (ii) the radius of maximum wind speed,
Rmax, (iii) the translation velocity of cyclone track, VT, and (iv) heading of cyclone track (). Using
these probability distributions, distribution of the cyclonic wind speed for a particular site in the
coastal region can be obtained.
Mathematical simulation methods have been used to estimate cyclonic wind speeds at the site of
interest (Russell, 1971; Russell and Schueller, 1974; Georgiou, et al., 1983; Vickery and Twisdale,
1995a, 1995b).
In this paper, the statistical distributions of key cyclone parameters are obtained for the states of
Andhra Pradesh and Orissa from the landfalling track records of Indian coastal region.

METHODOLOGY

The hourly latitude/longitude information of cyclone tracks is collected from the Journals/reports
published by India Meteorological Department (IMD) and database of the Joint Typhoon Warning
Statistical Modelling of Key Cyclone Parameters for the East Coast of India 42

Centre (JTWC). The dataset of historical landfalling storm tracks in India from 1975 to 2007 with
latitude /longitude is put into GIS and landfall locations are assigned in a region of study.
The statistical distribution of the key cyclone parameters are derived from actual data of central
pressure difference, translation velocity, radius of maximum wind, heading direction of all
landfalling tropical cyclones on Indian coastal region formed in Bay of Bengal. The
latitude/longitude information for each storm track is mapped into GIS and landfall locations are
assigned as shown in Fig. 1. Standard statistical techniques are employed to determine the best fit
distributions.

Fig. 1: Historical storm tracks in Bay of Bengal

2.1. Procedure for evaluation of velocity using the distribution of the cyclone key
parameters

The most intensive wind generally occurs at the eye wall of cyclones. The wind speed decreases as
the location moves away from cyclone centre. The gradient wind speed at a distance r from the
cyclone centre is given by
B B
rf r 22f p R max R (1)
Vg B exp max
2 4 r r

where f is the Coriolis parameter, r is the radial distance from the storm centre to the point of
interest, is the density of air, is the pressure difference, B is Holland (1980) parameter.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 41-49 43

Empirical relation for B is given by (Hubbert et al.; 1991)

(980) po
B 1.5 (2)
120
where po is the central pressure of a tropical cyclone.
After the translation of cyclone is taken into account, the 1-minute average wind speed at 10-meter
above the sea surface (Chang and Lindell, 2005) is given by

V10 0.8Vg 1.75 VT0.63 cos (3)

where V10 denotes the 1-minute average wind speed at 10-meter above the sea surface, Vg denotes
the gradient wind speed given by Eq. (1), VT denotes translation velocity of the cyclone centre, and
denotes the angle between the hurricane track and surface wind direction in the region of interest
based on wind field model as shown in Fig. 1. Note that no wind field model is available for
Indian coastal region. The wind field model based on Hubbert et al., (1991) and by Chang and
Lindll (2005) has been widely used. Because this model can represent a wide variety of
storms in both size and shape it is not necessary to choose a random value for B
The distribution of wind velocity at the site of interest is simulated using the distributions of the key
cyclone parameters. Simulation procedure consists of the following steps:

(i) Using the distribution of the key cyclone parameters, random values of the key
parameters are generated using MATLAB.
(ii) For each cyclone key parameter, a set of 1000 random values (length of record) are
generated. The sampling interval is taken as 6-h, same as that of the cyclone track
records as obtained from the available data.
(iii) For generating the values of key cyclone parameters p , Rmax, VT , , different trial sets
of random values are obtained using MATLAB. Note that simulated values of the
cyclone key parameters are site specific, since mean and standard deviation values of
the parameters are determined from the actual track records.
(iv) Thus, a set of cyclonic wind speeds at the site of interest is obtained using Eq. (3) and
the CDF of the data set is calculated.

NUMERICAL STUDY

The states of Andhra Pradesh and Orissa are the most disaster prone states in India. Therefore,
these states are taken as example problem for the numerical study. The distributions of the key
cyclone parameters are obtained from the track records using the method described before. The
probability density and cumulative distributions of key cyclonic parameters for Andhra Pradesh and
Orissa are shown in Figs. 3 to 9.
The cumulative distributions and probability density function obtained for the central pressure
difference for the states of Andhra Pradesh and Orissa are shown in Figs. 3 and 4 respectively. Two
Statistical Modelling of Key Cyclone Parameters for the East Coast of India 44

standard distributions, namely Weibull and lognormal, are fitted to the histograms. From the
Chi-square and K-S tests, it was observed that lognormal distribution fits best to the histograms. It
is seen that one of the reported distributions for central pressure difference in the literature
is the lognormal distribution (Batts et al., 1980). The characteristics of the distribution of the
central pressure difference data according to the best fit for east coast of India are given in Table 1.
The percentages indicate that the probability of occurance ofthe central pressure will be less than
the given value.
Since each site individually has too few data for the radius of maximum wind, it is difficult to
get any good distribution from the data for each state. Therefore, data of the sites located in the
eastern side, i.e., Andhra Pradesh, Orissa, Tamil Nadu, West Bengal are grouped together to get
the distribution of key cyclone parameters and the same distributions are assumed to hold good for
all sites in the eastern region. Fig. 5 shows cumulative and probability density function of radius of
maximum wind (RMW). Three distributions, namely lognormal, normal, and logistic are fitted to the
histograms. From the Chi-square and K-S test, lognormal distribution is found to be the best fit for
the radius of maximum wind. It is seen that the distribution as obtained from Indian data is the
same as that observed elsewhere (Georgiou et al., 1983; Vickery and Twisdale, 1995a;
Huang et al., 2001).
The characteristics of the distribution of the data for the radius of maximum wind are given in Table
2. The percentages indicate that the probability of occurance of RMW will be less than the given
value.
CDF and PDF of storm heading direction for the states of Andhra Pradesh and Orissa are
shown in Figs. 6 to 7. Normal distribution is found to be the best fit from Chi-square and K-S tests for
both the states. The distribution as obtained from Indian data is the same as that observed
elsewhere (Huang et al., 2001).
The characteristics of the distribution of the storm heading direction data for east coast of India
according to the best fit are given in Table 3.
CDF and PDF of translation velocity for the states of Andhra Pradesh and Orissa are shown in Fig. 8
to 9 respectively. Three distributions, namely Weibull, lognormal and normal are fitted to the
histograms of translation wind velocity. A Weibull distribution ranks first as the best fit for the
translation speed.
It is seen that there is some variation in the distribution of the key parameter of translation
velocity (for Indian condition) compared to those in the literature. This kind of deviation is
expected because characteristics of the cyclonic wind speed are likely to vary from region
to region and from country to country.
The characteristics of the distribution of translation (forward) speed data for east coast of India are
shown in Table 4. The percentages indicate that the probability of occurance the translation speed
will be less than the given value.
Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 41-49 45

Table 1: Central pressure difference distributions (Units are in millibars)

Percentile Value
Min 0.4941
5% 1.5444
10% 2.1067
25% (Q1) 3
50% (Median) 4.6058
75% (Q3) 5.7506
90% 8.0338
95% 9.6067
Max 20.972

Table 2: RMW data distributions (Units are in km)

Percentile Value
Min 26

5% 26
10% 26
25% (Q1) 29
50% (Median) 32
75% (Q3) 38
90% 44
95% 44
Max 44

Table 3: Storm heading distributions (Units are in degrees)

Percentile Value
Min - 169.07
5% - 101.4
10% - 89.964
25% (Q1) - 89.837
50% (Median) - 57.983
75% (Q3) - 22.445
90% 0
Statistical Modelling of Key Cyclone Parameters for the East Coast of India 46

Table 4 : Forward speed distributions (Units are in m/s)

Percentile Value
Min 0.4941
5% 1.5444
10% 2.1067
25% (Q1) 3
50% (Median) 4.6058
75% (Q3) 5.7506
90% 8.0338
95% 9.6067
Max 20.972

Fig. 2: Wind field model

Fig. 3: Probability distribution function of central pressure for Andhra Pradesh


Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 41-49 47

Fig. 4: Probability distribution function of central pressure for Orissa

Fig. 5: Probability distribution function of radius of maximum wind for Andhra Pradesh and Orissa

Fig. 6 : Probability distribution function of storm heading data for Andhra Pradesh
Statistical Modelling of Key Cyclone Parameters for the East Coast of India 48

(a) PDF (b) CDF

Fig. 7 : Probability distribution function of storm heading data for Orissa

Fig. 8: Probability
Fig. distribution
9: Probability function
distribution (b)
(a)CDF
(a)
PDF
of forward
function
CDF motion
of forward for Andhra
motion Pradesh
for Orissa

Fig. 8 : Probability distribution function of forward motion for Andhra Pradesh

Fig. 9 : Probability distribution function of forward motion for Orissa


Journal of Wind and Engineering, Vol. 10, No. 2, July 2013, pp. 41-49 49

4. Conclusions
The statistical distributions of key cyclone parameters for the states of Andhra Pradesh and
Orissa are obtained from the actual track records. Best fit distributions of key cyclone parameters
are obtained by an elaborate statistical analysis using MATLAB. The lognormal distribution is found
to be the best fit for central pressure difference and radius of maximum wind. The Normal
distribution and Weibull distributions are found to fit best for storm heading direction and translation
velocity respectively.

References

1. Batts, M.E., Russell, L.R. and Simiu, E. (1980). Hurricane Wind Speeds in the United States.
Journal of the Structural Division, ASCE, 106(ST10), pp. 2001-2016.
2. Chang, L. and Lindell, M.K (2005), Hurricane Wind Risk Assessment for Miami-Dade
Country, Florida: A Consequence Based Engineering (CBE) Methodology. CBE Report for
the Mid-America Earthquake (MAE) Center.
3. Georgiou, P.N., Davenport, A.G. and Vickery, B.J. (1983). Design Wind Speeds in Regions
Dominated by Tropical Cyclones. Journal of Wind Engineering and Industrial Aerodynamics,
13(1-3), pp. 139-152.
4. Goyal P.K. and Datta T.K. (2011). Probability Distributions for Cyclone key Parameters and
cyclonic wind speed for the east coast of Indian Region, The International Journal of Ocean
and Climate Systems, Vol.2 (3), Multi science, UK
5. Huang, Z., Rosowsky, D.V. and Sparks, P.R. (2001). Long-term Hurricane Risk Assessment
and Expected Damage to Residential Structures. Reliability Engineering and System Safety,
74(3),pp. 239249.
6. Holland, G. J. (1980), An analytic model of the wind and pressure profiles in hurricance,
Monthly Weather Review, Vol. 108, P 1212 - 128.
7. Holmes, J.D. (2001). Wind Loading of Structures, Spon Press, New York.
8. Hubbert, G.D., Holland, G.J., Leslie, L.M., and Manton, M.J. (1991). A Real Time System for
Forecasting Tropical Cyclone Storm Surges. Weather and. Forecasting, American
Meteorological Society, 6(1), pp. 86-97.
9. Russell, L.R. (1971). Probability Distributions for Hurricane Effects. Journal of Waterways,
Harbors, and Coastal Engineering Division, ASCE, 97(1), pp. 139-154.
10. Russell, L.R. and Schueller, G.F. (1974). Probabilistic Models for Texas Gulf Coast Hurricane
Occurrences. Journal Petroleum Technology, pp. 279-288.
11. Vickery, P.J. and Twisdale, L.A. (1995a). Prediction of Hurricane Wind Speeds in the United
States. Journal of Structural Engineering, ASCE, 121(11), pp. 1691-1699.
12. Vickery, P.J. and Twisdale, L.A. (1995b). Wind-field and Filling Models for Hurricane Wind
Speed Prediction. Journal of Structural Engineering, ASCE, 121(11), pp. 17001709.

You might also like