Você está na página 1de 46

A

PROJECT REPORT ON

STUDY OF STRUCTURAL SHAPES DEPENDING


ON WIND LOADING

SUBMITTED BY:

AKHILESH MADHAV KAJAREKAR

(200711009)

UNDER THE GUIDANCE OF:

DR.PRADEEPKUMAR RAMANCHARLA

INTERNATIONAL INSTITUTE OF INFORMATION TECHNOLOGY


HYDERABAD -500032

MAY 2009
CERTIFICATE

It is certified that work contained in the project titled ―STYDY OF STRUCTURAL SHAPES
DEPEND ON WIND LOADING‖ by Akhilesh Madhav Kajarekar has been carried out under
my/our supervision and it is not published elsewhere for a degree.

Dr. Pradeep Kumar Ramancharla


Coordinator,
Computer Aided Structural Engineering

1
ACKNOELEDGMENTS

I own a great deal of gratitude towards my advisor Associate Professor Dr. Pradeep
Kumar Ramancharla for his generous support through my project at IIIT. He guided
me through this interdisciplinary area of research with his professional insight and
diligence, without which this project would have not come to fruit.

I also thank Professor Dr. Venkateshwarlu and all my teachers at IIIT, who enhanced
my technical wisdom. I also express my thanks to my friends, EERC and every one
who are directly or indirectly have been the source of encouragement, without which it
would not have been possible for me to accomplish this task.

AKHILESH MADHAV KAJAREKAR

(200711009)

2
ABSTRACT

Today, Tall structures are important and dominant landmark of any city or town,
which are required to fulfill the increasing demand of space for various purposes. For
the safety of Life and Property, it is very necessary to consider lateral forces which are
acting on structures. Wind force is one of the important lateral forces to be considered
while designing high rise structures because as height of structure increases, wind
force becomes predominant force over seismic force. In this project, different aspects
of wind are studied. Also effect of wind on various shapes and sizes of structures are
considered. Study of wind forces on Y-Shaped building is taken as Case Study of the
Project.

3
TABLE OF CONTENTS

PAGE

Certificate………………………………………………………………………………....1
Acknowledgements …………............................................................................................2
Abstract ……………………………………….................................................................3

Table of contents……………………………………….....................................................4

CHAPTER 1. Introduction .................................................................................................5


CHAPTER 2. Wind Study...............................................................................................................................6

CHAPTER 3. Study of Wind under Different Environment of Building………………..12


CHAPTER 4. Design Criteria............................................................................................25
CHAPTER 5. Study of Wind on Y-Shaped Structure…………………………………..…26
CHAPTER 6. Case Study...................................................................................................32
CHAPTER 7. Interpretations of Results.............................................................................42
CHAPTER 8. Conclusions.................................................................................................44

REFERENCES.................................................................................................................45

4
1. INTRODUCTION

From the beginning in the middle of the last century and right up to the present
day, high-rise buildings have always been a dominant landmark in the townscape,
visible from far and wide, like the towers of Antiquity and the middle Ages. At the
same time, this sky-scraping construction method has always been an ideal means of
displaying power and influence in the community. Another reason of construction of
Tall buildings in large numbers is, due to scarcity of land and to meet the increasing
demand for space for residential and commercial purposes. The safety of life and
property is main aim of proper design of structure against lateral and gravity loads
which are acting on the structure.

These tall structures being slender light weight and with low structural damping
undergo oscillations due to earthquake and wind loads. These oscillations are in
direction of ground motion in case of Earthquake loading and along wind, across wind
and torsional directions under the influence of strong winds. Safe with good human
comfort and economic design of tall buildings are achieved, if reliable methods to
account for complicated wind or seismic effects are practiced while their design. As
height of structure increases, wind intensity also increases. At a particular height,
wind force is the governing factor of design of structure against lateral loading. Many
of such high rise structures are needed in region where wind intensity is higher. Also
in coastal regions where Tornados, Hurricanes are active, there is a need of wind
design structures. So in this project, wind study is done on various shapes and sizes of
structures to understand the effect of wind forces on the structures.

5
2.1 WIND STUDY

Wind is one of the significant forces of nature that must be considered in the design
of buildings. Structural loads applied by high winds are readily appreciated, even if
the method of determining them is not so easily understood. Other effects that can be
caused even by moderate breezes are commonly overlooked, however, because very
often there is no obvious link between wind and the behavior of a building. Wind
loads have become particularly significant because of the increasing number of high
rise buildings. Some tall buildings that extend into regions of high wind velocity have
swayed excessively in strong winds. Improperly anchored light-weight roofs have
been sucked off bodily by wind forces, and roofing materials have been lifted by high
local suctions and eventually peeled from large areas of roofs. These and many other
problems have emphasized the importance of a clearer understanding of wind and its
effects. With the old simplified approach, the total effect of wind was often
represented merely by a uniform lateral pressure on the windward side of a building
and suction on the leeward wall.
Wind is not constant either with height or with time, is not uniform over the side of
a building, and does not always cause positive pressure. In fact, wind is a very
complicated phenomenon; it is air in turbulent flow, which means that the motion of
individual air particles is so erratic that in studying wind one ought to be concerned
with statistical distributions of speeds and directions rather than with simple averages
or fixed physical quantities. Architects and engineers are concerned with and
responsible for not only structural design, but also the choice of exterior cladding
materials and components, the operation of mechanical services such as heating and
ventilating equipment, and with details of openings to limit infiltration. Wind has
important effects on each of these aspects of design; one might even conclude that of
the manifestations of nature with which the architect has to contend, apart from
gravity, the effects of wind are the most ubiquitous.

The shape of the building is another factor influencing the wind forces actually at
work. When wind meets an obstacle, it normally generates compressive forces on the
windward side of the building and suction forces on the leeward side. In addition, air
6
streaming around the building produces suction forces on the sides parallel to the wind
direction. The shape of the corners and edges of the building is particularly important.
Separation effects can cause suction and compressive forces several times greater than
the original dynamic pressure. The magnitude of these edge and corner forces depends
primarily on the geometry of the building round which the air flows. Basically, it may
be said that the more sharp-edged and irregular the building is, the more irregularly
the wind forces will be distributed. Suction forces cause major problems around the
roof in particular. If the roof structure has not been adequately anchored, parts of the
roof may be lifted off and catapulted away unhindered. In addition to the roof, such
elements as light-metal facades, antennas, promotional signs and water tanks are some
of the parts most seriously threatened by wind on high-rise buildings. The risk of parts
being blown away and flying around is greatest during the construction phase. Such
parts can cause considerable property damage to their surroundings, and harbor
potential for bodily injury which cannot be neglected. Hence Along and Cross loading
of wind concepts are required to understand while designing.

2.2 ALONG AND CROSS-WIND LOADING

Not only is the wind approaching a building a complex phenomenon, but the flow
pattern generated around a building is equally complicated by the distortion of the
mean flow, flow separation, the formation of vortices, and development of the wake.
Large wind pressure fluctuations due to these effects can occur on the surface of a
building. As a result, large aerodynamic loads are imposed on the structural system
and intense localized fluctuating forces act on the facade of such structures. Under the
collective influence of these fluctuating forces, a building tends to vibrate in
rectilinear and torsional modes, as illustrated in Fig.1. The amplitude of such
oscillations is dependant on the nature of the aerodynamic forces and the dynamic
characteristics of the building.

7
Fig.1

A) Along-Wind Loading
The along-wind loading or response of a building due to buffeting by wind can be
assumed to consist of a mean component due to the action of the mean wind speed
(eg, the mean-hourly wind speed) and a fluctuating component due to wind speed
variations from the mean. The fluctuating wind is a random mixture of gusts or eddies
of various sizes with the larger eddies occurring less often (i.e. with a lower average
frequency) than for the smaller eddies. The natural frequency of vibration of most
structures is sufficiently higher than the component of the fluctuating load effect
imposed by the larger eddies i.e., the average frequency with which large gusts occur
is usually much less than any of the structure's natural frequencies of vibration and so
they do not force the structure to respond dynamically. The loading due to those larger
gusts (which are sometimes referred to as "background turbulence") can therefore be
treated in a similar way as that due to the mean wind. The smaller eddies, however,
because they occur more often, may induce the structure to vibrate at or near one (or
more) of the structure's natural frequencies of vibration. This in turn induces a
magnified dynamic load effect in the structure which can be significant. The
separation of wind loading into mean and fluctuating components is the basis of the
so-called "gust-factor" approach, which is treated in many design codes. The mean
load component is evaluated from the mean wind speed using pressure and load

8
coefficients. The fluctuating loads are determined separately by a method which
makes an allowance for the intensity of turbulence at the site, size reduction effects,
and dynamic amplification (Davenport, 1967). The dynamic response of buildings in
the along wind direction can be predicted with reasonable accuracy by the gust factor
approach, provided the wind flow is not significantly affected by the presence of
neighboring tall buildings or surrounding terrain.

B) Cross-Wind Loading
There are many examples of slender structures that are susceptible to dynamic motion
perpendicular to the direction of the wind. Tall chimneys, street lighting standards,
towers and cables frequently exhibit this form of oscillation which can be very
significant especially if the structural damping is small. Crosswind excitation of
modern tall buildings and structures can be divided into three mechanisms and their
higher time derivatives, which are described as follows:

(a) Vortex Shedding:-


Wind is composed of a multitude of eddies of varying sizes and rotational
characteristics carried along in a general stream of air moving relative to the earth‘s
surface. These eddies give wind it‘s gusty or turbulent character. The gustiness of
strong winds in the lower levels of the atmosphere largely arises from interaction with
surface features. The average wind speed over a time period of the order of ten
minutes or more, tends to increase with height, while the gustiness tends to decrease
with height
The most common source of crosswind excitation is that associated with ‗vortex
shedding‘. Tall buildings are bluff (as opposed to streamlined) bodies that cause the
flow to separate from the surface of the structure, rather than follow the body contour
(Fig. 2). For a particular structure, the shed vortices have a dominant periodicity that
is defined by the Strouhal number. Hence, the structure is subjected to a periodic cross
pressure loading, which results in an alternating crosswind force. If the natural
frequency of the structure coincides with the shedding frequency of the vortices, large
amplitude displacement response may occur and this is often referred to as the critical

9
velocity effect. The asymmetric pressure distribution, created by the vortices around
the cross section, results in an alternating transverse force as these vortices are shed. If
the structure is flexible, oscillation will occur, transverse to the wind and the
conditions for resonance would exist if the vortex shedding frequency coincides with
the natural frequency of the structure. This situation can give rise to very large
oscillations and possibly failure.

Fig.2
(b) The incident turbulence mechanism:-
The ‗incident turbulence‘ mechanism refers to the situation where the turbulence
properties of the natural wind give rise to changing wind speeds and directions that
directly induce varying lift and drag forces and pitching moments on a structure over a
wide band of frequencies. The ability of incident turbulence to produce significant
contributions to crosswind response depends very much on the ability to generate a
crosswind (lift) force on the structure as a function of longitudinal wind speed and
angle of attack. In general, this means sections with a high lift curve slope or pitching

10
moment curve slope, such as a streamline bridge deck section or flat deck roof, are
possible candidates for this effect.

(c) Higher derivatives of crosswind displacement:-


There are three commonly recognized displacement dependent excitations, i.e.,
‗galloping‘, ‗flutter‘ and ‗lock-in‘, all of which are also dependent on the effects of
turbulence in as much as turbulence affects the wake development and, hence, the
aerodynamic derivatives. Many formulae are available to calculate these effects
(Holmes, 2001). Recently computational fluid dynamics techniques (Tamura, 1999)
have also been used to evaluate these effects.

11
3. STUDY OF WIND UNDER DIFFERENT ENVIRONMENT OF BUILDING

Many factors will have an impact on the wind conditions around a building. Some of
these parameters include: the ambient wind statistics, local topography, building
massing, nearby foliage and the proximity of similarly tall structures. These all may
influence the resulting winds around the base of a new building and at elevated levels
on balconies and terraces. It is for this reason that many new building designers
evaluate their project in a boundary-layer wind tunnel with the subject building both
installed and removed from the turntable. In this way, the project‘s impact on the local
environment may also be assessed.

3.1BUILDING MASSING AND ORIENTATION

It is well known that the design of a building will influence the quality of the ambient
wind environment at its base. A shear curtain wall to ground level with a rectilinear
floor plan (circular shapes typically do not cause flows of this type) is often a design
which may aggravate street-level winds by allowing the high-elevation, faster winds
to flow down the face of the structure. The mechanism is called downwash (see
Figures 3).

Fig.3

Once the wind reaches the ground it is then accelerated around the ground-level
corners (see Figure 4). A large canopy may interrupt the flow as it moves down the
windward face of the building. This will protect the entrances and sidewalk area by
12
deflecting the downwash at the second storey level (Figure 5). However, this approach
may have the effect of transferring the breezy conditions to the other side of the street.
Large canopies are a common feature near the main entrances of major office
buildings. The architect may elect to use an extensive podium for the same purpose if
there is sufficient land and it complies with the design mandate (Figure 6). This is a
common architectural feature for many major projects in recent years, but it may be
counterproductive if the architect wishes to use the podium roof for long-term
pedestrian activities, such as a pool or tennis court.

Fig.4 Fig.5

Another massing issue, which may be a cause of strong ground-level winds, is an


arcade or thoroughfare opening from one side of the building to the other. This
effectively connects a positive pressure region on the windward side with a negative-
pressure region on the lee side. A strong flow through the opening often results as
illustrated in Figure 7.

13
Fig.6 Fig 7

A similar phenomenon occurs with a high-rise building raised up on columns, a


design popular in the 1960s (Pen warden and Wise, 1975). The uninvitingly windy
nature of these open areas is a contributing reason behind the rarity of this type of
architectural form in modern high-rise buildings. One exception is in calm, tropical
climates where the extra breeze creates a desirable feature. For example, the Hitachi
Building in Singapore has used this approach to provide shaded, cooler areas at the
ground-level entrances with great success. However, the same design in a windy city
with cold winters, like Chicago, would be an unfortunate design choice. An entrance
alcove behind the building line will generally produce a calmer entrance area (Figure
8) at a mid-building location. In some cases a canopy may not be necessary with this
scenario, depending on the local geometry and directional wind characteristics. The
same undercut design at a building corner is usually quite unsuccessful (Figure 9).
This is due to the accelerated flow mechanism described in Figure 4 and the ambient
directional wind statistics - often described graphically using a wind rose. If there is a
strong directional wind preference at the city in question, and the corner door is
shielded from those common stronger winds, then the corner entrance may work.

14
Fig.8 Fig.9

However, it is more common for a corner entrance to be adversely impacted by


this local building geometry and the strong winds that more commonly occur in that
city, both influencing the exposed corner entrance. The result can range from simply
unpleasant conditions to a frequent inability to open the doors (Figure 9). The way in
which a building‘s vertical line is broken up may also have an impact. For example, if
the floor plans have a decreasing area with height the flow down the ‗stepped‘
windward face may be greatly diminished. To a lesser extent the presence of many
balconies can have a similar impact on ground level winds, although this is far less
certain and more geometry dependent. Condominium designs with many elevated
balconies and terrace areas near building ends or corners often attract a windy
environment to those locations. Mid building balconies, on the broad face, are usually
a lot calmer. Corner balconies are generally a lot windier and so the owner is likely to
be selective about when the balcony is used or endeavors to find a protected portion of
the balcony that allows more frequent use, even when the wind is blowing.

3.2THE INFLUENCE OF TERRAIN

The presence of topography around a site will impact the ambient pedestrian
conditions in much the same way as major neighboring buildings do in the vicinity of
proposed new development. Being in the shadow of substantial topography such as
this is renowned for impacting pedestrian-wind conditions and must be physically
15
modeled in the wind tunnel to make a reliable assessment of the higher speed winds
(pedestrian comfort) and lower speed winds (removal of pollution or viruses such as
SARS) around the anthropogenic environment being assessed. In some circumstances
the new building is an integral part of the complex terrain and so flow over the terrain
passes over the balconies of the condominium in much the same manner as the
escarpment itself. Thus, the terrain at a new building site may reduce the wind speeds
by shielding the new development, or it may speed up the local winds due to the
presence of a hill or escarpment. In fact, both phenomena may occur on the same
project, depending on the wind direction being studied.

3.3 USE OF CANOPIES, TRELLISES AND HIGH CANOPY FOLIAGE

Once the flow mechanism at a problem location has been established for the critical
wind directions the remedial solutions may be explored for effectiveness in a
boundary layer wind tunnel. At model scales ranging from about 1:200 to 1:500,
locations of interest may be investigated using a variety of techniques. One of the
most common is the hot-film or hot-wire anemometer shown in Figure 10.

Fig.10

This fine wire is uniquely suited to measuring both the mean and peak gust wind
speeds at the pedestrian locations of interest due to its low thermal inertia. The hot
film (more robust than a hot wire) may be fixed at the desired location for data
collection, or be installed on a computer controlled traverse for movement to
sequential sites on the model (Law Flay and Barthel, 2003). As noted earlier,
downwash off a tower may be deflected away from ground-level pedestrian areas by
large canopies or podium blocks. The downwash then effectively impacts the canopy
or podium roof rather than the public areas at the base of the tower (see Figures 5 and
16
6). Provided that the podium roof area is not intended for long-term recreational use
(e.g. swimming pool, tennis court or putting green), this massing method is typically
quite successful. However, some large recreational areas may need the wind to be
deflected away without blocking the sun, and so a large canopy is not an option.
Downwash deflected over expansive decks like these may often be improved by
installing elevated trellis structures or a dense network of trees to create a high, bushy
canopy over the long-term recreational areas. Various architecturally acceptable ideas
may be explored in the wind tunnel prior to any major financial commitment on the
project site.

3.4 USE OF LANDSCAPING FOLIAGE AND POROUS OR SOLID

SCREENS

Horizontally accelerated flows between two tall towers (Figure 11) may cause an
unpleasant, windy, ground-level pedestrian environment, which could also be locally
aggravated by ground topography. By inspection of the available wind data, the
designer may find a dominant wind direction that can be used to align the buildings on
the site so as to minimize these accelerated flows in highly trafficked pedestrian areas.
In major entrance areas it is rarely an option to use extensive planting or porous
screens to reduce the speed of the flow through the gap. However, these landscaping
techniques are the preferred methodology to ameliorate winds that are principally
horizontal (i.e. not the vertical downwash flow discussed previously). Extensive
planting was used to improve conditions on the pool decks in Figures 12 and 14. The
accelerated flow around the corner of the tower in Figure 10 was retarded using
foliage and tall porous parapet screens on the podium perimeter. In Figure 14 massive
planting and the use of porous screens calmed this hotel recreational area on the beach
in Aruba from the very dominant easterly breezes. Horizontally accelerated flows that
create a windy environment are best dealt with by using porous screens or substantial
landscaping. Large hedges, bushes or other porous media serve to retard the flow and
absorb the energy produced by the wind. A solidity ratio (i.e. proportion of solid area

17
to total area) of about 60-70% has been shown to be most effective in reducing the
flow‘s momentum (Rouse, 1950). These physical changes to the pedestrian areas are
most easily evaluated by a model study in a boundary-layer wind tunnel. Figure 13
shows a tennis court that was part of a Taiwanese condominium development. By
using trees to protect from key wind directions and a 3-m high porous perimeter fence,
the conditions on the court were substantially improved.
A comparative study showing the impact on the site with and without the
ameliorative additions is a useful method of defining their effectiveness, and it also
allows the architect to define the extent and usefulness of the pedestrian space. When
these studies are done as part of the design process, the architect can be more assured
of the success of open space, pool areas, balconies, terraces and entrances around a
new project.

Fig.11

18
Fig.12 Fig.13

Fig.14

19
3.5TORSIONAL WIND LOADING

Torsional wind loading on buildings is not as well understood as lateral or


overturning loads, and is not as amenable to analytical treatment. Because of this,
many designers—and indeed most codes of practice—ignore this aspect of the load,
and simply apply the lateral load at the geometric or elastic center of the structure.
Wind tunnel tests on model buildings have revealed that torsional loads usually exist,
and span a great range of significance. Several causes of the torsion can be identified,
and this categorization aids the designer in providing methods to either reduce or
manage the loading.
Shear and torsional loads on several building models are shown in following Figures.
All are real-life cases but represent somewhat elementary examples in terms of shape
all are essentially prismatic with the cross-section indicated; and exposure all are
essentially isolated buildings except where indicated. All were tested in a wind tunnel
in turbulent shear flow. The loads are in coefficient form, with the coefficient of total
shear defined as

And the torsional moment Mz expressed as a normalized eccentricity,

Where D = maximum building width, H = building height, U = a reference wind


speed, and Q = a reference dynamic pressure. This normalized eccentricity gives a
common and intuitive indication of the additive effect of torsion on the total building
shear, at a given wind direction. Often the maximum eccentricity does not occur at the
same wind direction as the maximum shear, as demonstrated by the graphs. In general
the governing design case is not obvious and will depend on the torsion-resistance
properties of the frame. For example, if torsion is resisted by shear walls placed near
the ends of the building then a given torque may increase the shear stresses only
slightly, and the wind direction of maximum overall building shear may well represent
the design case. However, if the shear walls are concentrated near the core then the
20
same torque will produce a much greater shear stress, and the design condition is more
likely to occur at the direction of maximum eccentricity.

(Angle of contact vs. Normalized base shear for different shapes of buildings)

Fig.15

Fig.16

21
Fig.17

Fig.18

Fig.19

22
3.6 THE EFFECT OF BUILDING SHAPE ON TORSION

Aerodynamic torsion occurs even on cylinders of elementary shape (other than round)
whenever the angle of wind incidence is skewed to an axis of symmetry. This can be
seen with the aid of Figure 2, which shows schematically the streamlines, separated
areas, and pressure distribution on a square cylinder in steady uniform flow. The
positive pressure distribution on the windward wall is slightly unsymmetrical,
resulting in a small anticlockwise torque. The leeward wall is in a separated zone, and
the negative pressure acting on it is more nearly uniform. Perhaps less obvious and
more important, however, is the pressure on the side walls. The flow is entirely
separated from the upper wall, which therefore experiences moderately negative
pressure, nearly uniformly distributed. Flow on the lower wall separates from the
leading corner but eventually reattaches to the surface. Pressure within the separation
bubble is more negative than in the reattached zone, and this non-uniform pressure
results in an anticlockwise torque which reinforces the torque on the windward wall.
Thus a pattern of four cycles of alternating torque occurs within a period of 360
degrees of wind rotation. This pattern can clearly be seen in the eccentricity curves of
Figure 15, buildings A and B. Building A is a square cylinder, and the maximum
eccentricity is only about 6 to 8 percent. Building B is rectangular, and the maximum
eccentricity approaches 10 percent. Also, note that the eccentricity (and also the
torque, since the shear is nearly constant) is larger when the wind is nearly parallel to
the long axis, than when it is nearly parallel to the short axis. This is reasonable in
light of the above discussion of Figure 16, because the torque is affected more by the
separated region on the bottom side wall than by the non-uniform pressure on the
windward wall. In the former case this separated zone occurs at a greater distance
from the building center. The same effect occurs in Building C, but it is further
accentuated by the skewed end walls. The maximum eccentricity is now 25 to 30
percent. In fact this parallelogram is a popular building shape, and it has become
somewhat of a classic case for experiencing unusually high torsion. An even more
extreme case of shape-induced torque occurs in Building D, a rectangle with opposite
corners rounded. For this case, when wind is nearly parallel to the long axis—say 80–
90 degrees—separation occurs on one side but is essentially suppressed at the rounded
23
opposite side. Thus, the negative pressure within the separation bubble has nearly
nothing to counteract it on the opposite side, resulting in a very large torque, and
eccentricity in the neighborhood of 20 percent. Even more interesting, however, is
wind skirting the curved faces (i.e., roughly parallel to the long diagonal), say 30 or
210 degrees. A significant negative pressure develops on the curved walls, while the
flat walls experience positive pressure on the windward side and slight negative
pressure on the leeward side. All of these pressures contribute to a clockwise torque,
and result in eccentricity of 40 to nearly 50 percent. Although building shapes such as
C and D are somewhat special in producing large eccentricities, it must be noted that
the largest eccentricities coincide with a reduced shear of perhaps half the maximum
shear. Conversely, the maximum shear tends to coincide with a relatively minor
eccentricity. Critical design cases for such shapes will be highly dependent on the
nature of the torsional resistance system: those with perimeter stiffness will tend
towards governing by the maximum shear, and those with core stiffness will be more
sensitive to the large eccentricities.

3.7 INTERFERENCE EFFECTS OF OTHER BUILDINGS

It is easily imagined that the torque on a high-aspect-ratio building would increase if


wind on the broad face is partially shielded by an upwind building. Less obvious is the
effect that occurs with an in-line upwind building, as illustrated on Building E in
Figure 3a. This building was tested with and without the upwind interfering building
at azimuths 220–290 degrees. Without this building, with winds from the
neighborhood of 270 degrees, the maximum eccentricity is a modest 5 to 6 percent.
When the building is added, the eccentricity increases to 8–10 percent. This increase
occurs despite a slight reduction in the total shear force. Therefore, the effect of the
interfering building is to provide some (slight) shielding regarding the total force on
the subject building, but the torque is increased. It is conjectured that the interfering
building, when slightly displaced from directly upwind, creates a sheared flow within
its wake which engulfs the subject building. Thus the subject building is impacted by
a mean flow which varies over its width, creating an unbalance.

24
4. DESIGN CRITERIA

In terms of designing a structure for lateral wind loads the following basic design
criteria need to be satisfied.
1) Stability against overturning, uplift and/or sliding of the structure as a whole.
2) Strength of the structural components of the building is required to be sufficient to
withstand imposed loading without failure during the life of the structure.
3) Serviceability for example for buildings, where inter storey and overall deflections
are expected to remain within acceptable limits.

4.2 CONVERSION FROM WIND SPEEDS TO WIND PRESSURES

Wind pressures exerted on a structure depend on the speed of the wind as well as
the interaction between the air flow and the structure. Since wind is air in motion the
pressures it can exert are related to its kinetic energy. If the full kinetic energy is
transformed into pressure then the resulting increase is given by the expression

q = ½ ρV²

Where, ρ is the mass density and V the velocity of the air. This is called the
"stagnation pressure" and is the maximum positive increase over ambient pressure that
can be exerted on a building surface by wind of any given speed. It is the basic
pressure to which all other pressures over the structure are referred. The wind speed to
be used in computing the design pressure depends on the particular component of the
building being designed. For structural purposes the maximum value is required and
will vary with the geographical location. Meteorological records of wind speed are
analysed to yield the most probable maximum that will be equalled or exceeded, on
the average, once during a given period of time comparable to the life of a structure.
In the National Building Code of Canada 1960 the "return period," as it is called, has
been set at 30 years.

25
5. STUDY OF WIND ON Y-SHAPED STRUCTURE

Cross section of buildings or structures often appears in a form of rectangle with


different aspect ratios of breadth to width. They may be aerodynamically unstable
under certain conditions, and the phenomenon of galloping may occur. As flow passes
through a rectangular cylinder, the boundary layer becomes detached or separated
from the rectangular cylinder at some point on its surface. Vortex will form in the
wake because of the instability or interaction between the shear layers and it may lead
to vortex-induced vibrations. While numerous investigations were made of the flow
past single obstacles with various shapes, few studies were concerned with the wake
interference and vortex shedding associated with complex configurations consisting of
multiple obstacles.
In engineering practices, there are many examples of buildings or structures in Y-
shape, which may be regarded as three identical rectangular cylinders arranged in l20
degrees apart either in a connected or a separated way. Therefore, it is both an
academic and practical issue to gain an understanding of the flow around such special
cross section of structures, as well as their wind loads. Vortex shedding from an H-
shape cylinder was studied by Nakagawa, Hayashida and Iwasa studied vortex-
induced vibration of eight different kinds of cross section of high-rise building by
wind tunnel tests, including Y-shapes. The results show that no peak value can be
found in the spectrum of cross wind force and no influence on wind direction. Gu-et-
al. studied the wind loads of three rectangular cylinders arranged in separated Y-shape
in a uniform smooth flow; different patterns of pressure distributions were obtained
for different aspect ratios of component rectangular cylinders under various angles of
incident flow. Galloping phenomenon was observed in a certain case. Recently, in
engineering practice, a strong vibration was reported on a TV tower with its cross
section in a separated Y-shape.
The experiments were conducted in an open-circuit wind tunnel with a working
section of 6m long, 0.6m wide and 0.6m high. The maximum velocity of free stream
is about 36 m/s with a background turbulence intensity of 0.2%. For the high-turbulent
flow test, a bi-planar turbulence grid made of rectangular cylinders (30mm × 30 mm)
with a vertical and horizontal mesh width of 0.114m is installed 1.82m upstream of
26
the model position to generate uniform high turbulent flow. The longitudinal
turbulence intensity, Iu at the model position is then 10% and the longitudinal integral
scale of the U-component, Lx, is 75mm. Two aspect ratios of breadth l to width b, i.e.
l/b = 3.0 and 1.4, for the rectangular cylinders, were chosen to construct two
connected and separated Y-shape models, respectively. As a frequent reference will be
made to the individual model, the connected models of l/b = 3.0 and 1.4 are labeled as
LC and SC, and the separated models of l/b = 3.0 and 1.4 are labeled as LS and SS,
respectively. For the same reason, the individual limb of the connected model or the
cylinder of the separated model is denoted as A, B and C, respectively, as shown in
Fig. 1 together with the definition of length D/2 and β.
The models are made of Plexiglas and placed horizontally in the middle of the test-
section between end plates. The natural frequencies of LC, LS and SC, SS are 43 Hz
and 54 Hz, respectively, with a damping ratio of 0.022, which is close to that in an
actual structure. The loads on the models are measured with strain gauges in one
particular direction enabling a separate measurement of drag and lift forces. Two
identical strain gauges are placed on each surface of a flat profile with rectangular
cross section, which is mounted on one side in a stiff support. On the other side of this
profile, the model is clamped, leading to dominant deformations and, consequently,
strains in the strain gauges, if forces perpendicular to the profile are exerted on the
model. The strain gauges are interconnected in half-Wheatstone Bridge in order to
guarantee a linear bridge output. The signals from the strain gauges through a low
pass filter are amplified by a DC amplifier (6M72). The drag and lift coefficients,
denoted by CD and CL, are defined conventionally and made dimensionless by the
characteristic length D. In order to obtain the specific property of wake fluctuating
velocity, a constant temperature hot-wire anemometer (TSI model 1050) is used. Two
hot-wire probes are placed at the two positions downstream of the model, located on
one side (position S) and on the centerline (position C) in the wake, respectively (see
Fig. 20). The flow visualization experiment was performed in the same wind tunnel.
The smoke-wire technique was employed. One of the four wires was placed in the
front of the model and the remaining three were located behind the model. The

27
pictures were taken with a time-delay, between the generation of smoke and the flash
of camera, of 80 milliseconds.
The Reynolds numbers (Re), which is based on the length of D in the flow
visualization, are l.3×104 and 7.8×103 for models l/b = 3.0 and 1.4, respectively. In
the force measurement experiments, the Reynolds numbers are in the range of
4.8×104 to 1.4×105. It is believed that the aerodynamic characteristics of a
rectangular-cylinder type object, such as vortex shedding frequency, separation points,
shear layers and wake structures, are similar in the region of subcritical Reynolds
numbers.

Fig.20

◦ ◦ ◦
The pictures of flow around the connected models at β = 0 , 30 and 60 are shown in

Fig. 2. For LC model at β = 0 , no evident vortex structure can be observed. The
separated shear layers from the leading edges of the windward surface of limb A
reattach on the limbs B and C, and then re-separate. The angles of re-separated shear
layers are big, as a result, a wide wake, somewhat like a ―dead water region‖, forms
behind the model. Since the two shear-layers deviate from each other, also because of
their self-stability in nature, no regular vortex forms in the near wake flow region.
This phenomenon is verified by the results of the fluctuating force measurement on
the model and the velocity measurement in the near wake, and will be discussed later
on. However, a large size vortex could form in a far wake behind the model because

of the instability of wake structure. At β = 30 , the separated shear layer generated
from the downside of limb A passes over limb C, which is entirely submerged in the
28
wake. A big separating area forms and creates a significant suction on the downside of
the model. It is confirmed by the results of force measurement. On the other hand, a
shear layer forms on limb B. Vortex forms due to the interaction between the two
shear layers. Peak-values can be found in the power spectra of both the lift force on
the model and the velocity in the wake. It suggests that a vortex-induced vibration
occurs, though it is rather weak, which will be discussed in a later section.

At β = 60 , the flow separates from both sharp edges of limbs A and B. The wake flow
region formed is rather broad. Symmetrical reversed flows are established in the wake
regions between limbs B & C and limbs A & C, respectively. The power spectra of
both the force and the velocity show that a strong peak-value appears at the same
frequency.
It means that two individual symmetrical vortices are formed behind the model and
the strong vortex-induced vibration is found. However, it is interesting that no peak
value can be detected in the power spectrum of velocity at the centerline of the wake
(position C, now located behind limb C). It suggests that there is no significant
interference between the two vortices because of the existence of limb C, which

functions as a split plate in the wake. The flow around SC model at β = 0 seems
somewhat unstable compared with that of LC model. The two shear layers separated
from limbs B and C are relatively closer; therefore, the interference between the two

shear layers happens not far away in the wake. At β = 30 , the interaction of the shear
layers is stronger than that of LC model and vortex forms close to the model. The

separating angles of shear layers from both limbs A and B are small at β = 60 , the
reversed flows are found close to the back of limbs A and
B. Limb C now is no longer acting as a split plate because of its shortness in breadth,
and a full interaction between the two shear layers occurs just behind the cylinder.
Figure 21 shows the flow visualization of the separated Y-shape models. For the LS
model at
β = 0◦, the flow is separated from the leading edges of cylinder A and then passes
through the gap between cylinders B and C, the wake structure is disturbed and the

29
―dead water region‖ no longer exists. On the other hand, the oncoming flow can reach
cylinders
B and C directly, and forms two separate shear layers. The wake region is relatively
narrower than that in the LC model. The wake is unstable because of the gap-flow,
which is, in turn, unstable and sensitive to the angular change of incident flow and

usually switches to one side. At β = 30 , the separated shear layer from the downside
of cylinder A is curved and rolled up, the oncoming flow can reach the tip of cylinder
C and then separates. An upward gap flow forms and strongly interferes with the

separated flow from the upside of the model. At β = 60 , compared with the case of
LC, the separated shear layers from cylinders A and B are concave and the wake
structure with the two individual symmetrical vortices are destroyed because of the
gap flow.

For the SS model at β = 0 the strong gap flow plays an important role in the wake

structure. At β = 30 cylinder C is immersed in the wake and the reversed flow can be

found close to the back of cylinder B. At β = 60 the gap flows interact strongly with
the shear layers separated from cylinders A and C. The power spectra of both
fluctuating force on the model and velocity in the wake have significant peak values at
the same frequency. It suggests that rather strong vortex shedding occurs in the wake.
Compared with that of the connected model, the flow pattern of the separated Y-shape
model is more complicated because of the interference of gap flow and the shear
layers in the wake.

30
Fig.21

31
6. CASE STUDY: - STUDY OF Y- SHAPED STRUCTURE FOR WIND
LOADS.

A. PLAN OF BUILDING

Fig.22

B. ANALYSIS TOOL: STAAD PRO 2006


C. AIM OF STUDY
a) To find out at what height wind force is predominant over seismic force.
b) To find orientation of building in which wind force is minimum.
c) Study the effect by changing storey height on different wings.

32
6.1 DETAILS OF STRUCTURE:-

A) Dimensions (In Plan) of structure.

Fig.23

B) Location of structure: Hyderabad


C) Soil Type: Medium Hard strata.
D) Type of structure: Steel Structure.

33
6.2 DETAILS OF ANALYSIS:-

A) Indian Standard methods for Lateral force analysis.


B) Response spectrum method for Seismic Analysis.(IS 1893)
C) Standard wind Analysis. (IS 875 part III).
D) Analysis Tool: - STAAD-PRO 2006.
E) Design Of structure :- By IS-800(1984) (Steel Design Code)
F) Wind speed at Hyderabad (as per IS 875):- 44 m/s.

34
PART I
TO FIND OUT AT WHAT HEIGHT WIND FORCE IS PREDOMINANT OVER
SEISMIC FORCE.

Analysis Steps:-
1) First 10 storey building is designed for gravity loading (Optimal design).
2) Seismic forces are applied on designed building and failed members are
observed.
3) Then Wind forces are applied on designed building and failed members are
observed.
4) By comparing results from steps 2 and 3, predominant force is found out which
is acting on structure.
Note: - For this 10 storey structure it is observed that seismic forces are
predominant over Wind force.
5) For every increasing no. of storey, structure is designed for wind force
(optimum design) and then it is checked for seismic forces. Again, that
structure is designed for seismic forces (optimum design) and then it is checked
for wind force.
6) Above procedure is repeated until it is observed that wind forces are
predominant.

RESULT:-
For above considered structure, it is observed that for 18 storey building, wind
forces are predominant.

35
Fig.24 SEISMIC ANALYSIS IN STAAD PRO

Fig.25 WIND ANALYSIS IN STAAD PRO

36
PART II
TO FIND ORIENTATION OF BUILDING IN WHICH THE WIND FORCES ARE
MINIMUM.

Analysis Steps:-
For finding out orientation here 18 strayed building is considered, which is
designed previously.

1) Wind forces are applied on this structure and base shear and base moments
are found out.

2) By keeping wind direction constant, structure is rotated by 10 and again
base shear and base moments are found out.

3) This procedure is repeated for up to 360 that is 36 iterations are done.
4) Results are plotted (fig.26) and minimum base shear and moments are
found out.

RESULTS:-

Degrees B.S (kN) B.M(kNm) Degrees B.S (kN) B.M(kNm) Degrees B.S (kN) B.M(kNm)
0 -4788.518 42029.22 120 4788.52 42029.22 240 4788.52 42029.22
10 -5205 46063.05 130 -5205 46063.05 250 -5205 46063.05
20 -5326.982 47351.85 140 5326.98 47351.85 260 5326.98 47351.85
30 -5355.294 48656.37 150 5355.29 48656.37 270 5355.29 48656.37
40 -5367.359 50189.49 160 5367.36 50189.49 280 5367.36 50189.49
50 -5224.28 50137.68 170 5224.28 50137.68 290 5224.28 50137.68
60 -4783.602 47278.38 180 -4783.6 47278.38 300 -4783.6 47278.38
70 -5206.251 53065.45 190 5206.25 53065.45 310 5206.25 53065.45
80 -5348.721 56103.94 200 5348.72 56103.94 320 5348.72 56103.94
90 -5404.496 58112.15 210 -5404.5 58112.15 330 -5404.5 58112.15
100 -5323.603 57717.27 220 -5323.6 57717.27 340 -5323.6 57717.27
110 -5265.961 56903.39 230 5265.96 56903.39 350 5265.96 56903.39

37
1.14

1.12
NORMALISED BASE SHEAR

1.1

1.08

1.06

1.04

1.02

0.98
0 50 100 150 200 250 300 350 400
DEGREES

1.6

1.4
NORMALISED BASE MOMENT

1.2

0.8

0.6

0.4

0.2

0
0 50 100 150 200 250 300 350 400
DEGREES

Fig.26

38
PART III
STUDY THE EFFECT BY CHANGING STOREY HEIGHT ON DIFFERENT
WINGS.

Fig.27
Analysis:-
To study this type of effect changes are made in above considered structure. The
storey no. of one wing kept as it they are i.e., 18. The storey no. of another wing
reduced to 16 and storey no. of remaining wing is increased to 20.
1) Wind forces are applied on this structure and base shear and base moments
are found out.

2) By keeping wind direction constant, structure is rotated by 10 and again
base shear and base moments are found out.

3) This procedure is repeated for up to 360 that is 36 iterations are done.
4) Results are plotted (fig.28) and minimum base shear and moments are
found out.

39
RESULTS:-
Degrees B.S (kN) B.M(kNm) Degrees B.S (kN) B.M(kNm) Degrees B.S (kN) B.M(kNm)
0 4899.57 44691.11 120 -4947 54485.25 240 5173.97 56872.34
10 5346.62 49073.53 130 5281.84 58270.89 250 -5592.5 61045.8
20 5414.24 50753.03 140 5234.11 57097.47 260 5735.51 62014
30 5525.02 51688.03 150 5207.82 56131.69 270 5357.09 57924.72
40 -5583.4 53424.27 160 5182.54 55025.73 280 5242.57 55793.31
50 5490.27 53861.71 170 5173.55 54659.35 290 5547.89 58848.96
60 5007.56 50195.9 180 4873.42 51805.09 300 5139.98 54196.21
70 5523.92 57181.2 190 5213.95 55387.05 310 5476.14 56685.66
80 5575.97 59084.54 200 5435.12 57949.29 320 5719.34 57755.01
90 4695.32 44265.87 210 5780.73 62197.74 330 4940.96 47948.88
100 5697.68 62242.99 220 5501.28 59755.86 340 5652.44 54077.38
110 5411.75 59844.83 230 5508.77 60222.33 350 5294.21 49241.16

40
1.4

NORMALISED BASE SHEAR 1.2

0.8

0.6

0.4

0.2

0
0 50 100 150 200 250 300 350 400

DEGREES

1.6
1.4
NORMALISED BASE MOMENT

1.2
1
0.8
0.6
0.4
0.2
0
0 50 100 150 200 250 300 350 400

DEGREES

Fig.28

41
7. INTERPRETATIONS OF RESULTS

Part I
a) Wind intensity increases with the height of a building, at a particular height it
results in predominance of wind forces over seismic forces.
b) Seismic forces are mostly depending on weight of structure. Here steel
structure is considered, so weight of structure is lesser as compare to concrete
structure. In case of R.C.C., the height will be more where predominance
occurs.

Part II
a) In this case, all wings are having 18 floors. So similar pattern is obtained after
◦ ◦
120 and 240 in base shear and base moment graphs (fig. 26).
b) From the fig 26, it is observed that base shear is minimum when structure is at
◦ ◦
60 as in fig 29a and base shear is maximum when structure is at 90 as in fig
29b.

Fig 29a Fig 29b

Part III
a) In this case, all wings are having different floors. It is observed that base forces

pattern is not similar after 120 as in Part II. This is because area exposed to
wind is kept on changing for every orientation.

42
b) From the fig 28, it is observed that base shear is minimum when structure is at
◦ ◦
90 as in fig 30a and base shear is maximum when structure is at 210 as in fig
30b.

Fig 30a Fig 30b

43
8. CONCLUSION
In this Project study of wind forces is done from which it is concluded that Wind
forces acing on a structure are depend on shape of building, exposed area, surrounding
environment, interference effects. The general design requirements for structural
strength and serviceability assume particular importance in the case of tall
building design as significant dynamic response can result from both buffeting and
cross-wind wind loading excitation mechanisms. The aim of case study was to study
orientation of building in which wind force is minimum and studies the effect by
changing storey height on different wings. This study concludes that if lower height is
kept exposed first to wind, then wind forces acting on the structures will be minimum.

44
REFERENCES

1) Sources of Torsional Wind Loading on Tall Buildings: Lessons From the Wind
Tunnel By Daryl W. Boggs, Noriaki Hosoya and Leighton Cochran

2) Design Features to Change and/or Ameliorate Pedestrian Wind Conditions by


Leighton Cochran1 (PhD. CP Eng.)

3) Wind Loading on Tall Buildings by P. Mendis, T. Ngo, N. Haritos, A. Hira

4) Flow around three rectangular cylinders arranged in connected and separated


Y-Shape by Zhifu GU, Yan Li

5) Flow-induced Vibration, New York by Van No strand Reinhold

45

Você também pode gostar