Sie sind auf Seite 1von 10

Wind Design of Tall Buildings, Problems, Mistakes and Solutions

Priyan Mendis1, a, Damith Mohotti 2, b & Tuan Ngo 3, c


1
The University of Melbourne
a
pamendis@unimelb.edu.au, b pushpajm@unimelb.edu.au, dtngo@unimelb.edu.au

ABSTRACT

Wind design normally governs the design of tall buildings around the world. The authors will draw on
their more than 25 years of experience in academic research and practical design of tall buildings, to
present the state-of-the-art and problems in designing tall buildings for wind. The recent work
conducted with computational fluid dynamic analysis techniques will be presented. Perceptions of
motion by the occupants and the pedestrian comfort have become very important issues in planning
and designing tall buildings in urban environment. Latest developments in this area will also be
presented in the paper.

KEYWORDS: Supper tall-buildings, Wind loadings, Pedestrian comfort, Computational Fluid


Dynamics (CFD)

1.0 INTRODUCTION
In recent years, there has been a burgeoning growth in tall buildings which is evidently shown in the
recent statistics published by the Council on Tall Buildings and Urban Habitat. As illustrated in
Figure 1, years 2011, 2012 and 2013 recorded as many as 81, 69, and 73 high rise buildings taller than
200m built around the world. The projected number is expected to escalate in 2014 and 2015 (Safarik
and Wood, 2014). This increasing trend can be attributed to population growth, shortage of land, and
the consequent increase in land prices especially in metropolitan areas. For example, due to limited
land space, land plot values in Singapore have multiplied by 30% per year since 2011 based on the
governments auction data (Thakur, 2014). Besides that, many developing countries boost their
countries reputations as financial powerhouses and tourist destinations by building high rise
structures such as the Burj Khalifa in Dubai, Shanghai World Financial Center in China, Taipei 101 in
Taiwan, and the Petronas Twin Towers in Malaysia (Council on Tall Buildings and Urban Habitat,
2014). Furthermore, the development of more advanced construction methods and materials such as
high strength concrete with compressive strength exceeding 100MPa has made construction of tall
buildings more feasible.

120
Number of Buildings 200m+ Completed

105
100
90
81
80 73
70 69

60
48 50

40 31 32 32
26
23 23
18 16 16 17
20 15 13 13
10
7 6 8 5 5
10
7 5 9 11
4 3 5
1
0
1980
1981
1982
1983
1984
1985
1986
1987
1988
1989
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
2010
2011
2012
2013
2014
2015

Year Projected Number

Figure 1: Trend of number of tall buildings constructed per year (Safarik and Wood, 2014)

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


With the increasing trend of tall buildings, it is important to continually develop more accurate
methods for structural design of tall buildings as the collapse of such buildings can lead to
catastrophic consequences with a loss of many human lives. The strength and serviceability of tall
buildings are governed by lateral loads either due to either wind or earthquake loads. The focus of this
research will be on dynamic wind action as it is commonly the governing load case for tall and
slender buildings.

However, due to the lack of understanding of design standards and their limitations designers tend to
over-predict or under-predict the wind induced load on tall buildings. Also recent developments in
stern environmental and safety regulations, one needs to asses all the aspects of the building and its
surrounding in such designs. A large number of design standards have been developed by different
countries to aid the wind design of tall buildings including Australian Standard
(AS/NZS1170.2:2011), Chinese Code (GB50009:2012), American Code (ASCE7-10:2010) EURO
Code (EN1991-1-4:2005), and Japanese Code (AIJ 2004). However, their applicability in tall building
design is limited due to the reasons discussed in Section 2 of this paper.

2.0 LIMITATIONS IN WIND DESIGN CODES AND WIND TUNNEL TESTING

Many developing countries have developed their own design codes for wind actions. However, there
are limitations and possibly over-estimations associated with the wind design codes. Hence, the wind
design codes should be used with considerations regarding the constraints. To understand the key
limitations behind the wind standards, five major international codes were analysed: Australian
Standard (AS/NZS1170.2:2011), Chinese Code (GB50009:2012), American Code (ASCE7-10:2010)
EURO Code (EN1991-1-4:2005), and Japanese Code (AIJ 2004). Six important aspects, where the
codes have limitations were identified and summarised in Table 1.

Table 1: Summary of limitations in wind design codes


Criteria AS/NZS GB50009:2012 EN1991-1-4:2005 AIJ2004 ASCE 7-10
1170.2:2011 (Ministry of (European (Architectu :2010
(Standards Construction of Committee for ral Institute (American
Australia the People's Standardization, of Japan, Society of
Limited/Standards Republic of 2010) 2004) Civil
New Zealand, China, 2012) Engineers,
2011)
2010)
Height No provisions No provisions No provisions for No explicit limit specified
for h>200m for h>550m h>200m

Geometry No provisions for buildings with complicated shapes

Along Wind Geometrical and height restrictions in along wind load calculations
Analysis
Cross Wind Geometrical and height restrictions in across wind load calculations No
Analysis provision
for cross
wind
calculations
Shielding Shielding factor is only based on approximation of terrain and surrounding
Effects environment
Acceleration Only applicable Only applicable No provision for Only No
(comfort) to rectangular to buildings cross wind applicable provision
buildings with uniform acceleration to for cross
shape and mass rectangular wind
along height buildings acceleration

Excluding the Chinese code, it can be observed that most codes only provide analysis of buildings up
to 200m in height. Besides that, all codes have limitations in analysing high-rise structures with

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


complicated shapes. Nguyen (2009) suggested that most major wind codes can only analyse wind
loads and accelerations of tall buildings with square or rectangular cross section and a maximum
aspect ratio of six. Although all examined codes except for the American code provide calculation for
across wind load, there are geometrical restrictions associated. Furthermore, numerous parameters
given by the codes such as the shielding factor are approximated and may result in over conservative
designs.
For the design of super tall buildings or structures with irregular geometry, most major standards
recommend the use of wind tunnel testing (Kwon and Kareem, 2013). While wind tunnel testing
enables more flexibility in mimicking the surroundings of buildings to reality as compared to the
design standards, measurements are only recorded at limited locations on the model and it may suffer
from incompatible similarity requirements due to reduced scale setup (Montazeri and Blocken,
2012). Besides, it is time and cost inefficient to repeat the wind tunnel experiments if there are any
modifications to the structural designs.
In summary, all major wind codes reviewed have limitations in analysing buildings with extremely
tall height, complicated geometry, and with complex terrain and shielding structures surrounding
them. Besides having the ability of modelling and analysing super tall buildings and complicated
shaped buildings, CFD can be adopted in wind design as it is able to model the actual surroundings in
full scale as compared to reduced scale when it is done in wind tunnel experiments. Besides, while
parameters are only measured at selected points on the wind tunnel model, CFD provides detailed
information about wind velocities, wind pressures, and wind concentration at any grid point within the
flow domain (Westbury and Miles, 2002). Additionally, in contrast with the conventional wind tunnel
methods which require material, labour and time resources, CFD offers more flexibility in conducting
parametric studies for different flow conditions, geometries and complex surroundings at a cheaper
cost (Vafaeihosseini et al., 2013). Therefore, it can be deduced that CFD has immense potential in
wind design as it is a practical, flexible and relatively inexpensive tool. The practical example of
using CFD technique in designing of tall buildings is explained in Section 4 of this paper.

In addition to the limitation given in exciting standards and codes, perception of motion and
pedestrian comfort are among two key concerns that Engineers face during the design stage of tall
buildings. The next section of this paper briefly discusses those two issues and alternative ways of
estimating them instead of using costly wind tunnel approach.

3.0 PERCEPTIONS OF MOTION BY THE OCCUPANTS AND THE PEDESTRIAN COMFORT

3.1 Perception of Motion


Perception of motion is a key problem that is encountered by the occupants of tall buildings. A
surevey was conducted recently using 1014 central business district workers in Wellington, New
Zealand, about their experiences of wind-induced building motion, susceptibility to motion sickness,
reported compensatory behaviours, and complaints about building motion. Overall, 41.7% of the
respondents reported that they had felt wind-induced building motion, and 41.6% of those respondents
reported perceptible motion at least once a month. Difficulty in concentrating was the most frequently
reported effect of building motion, reported by 41.9% of the respondents who had felt building motion
(Lamb et al., 2013).
One of the earliest experiments and a survey conducted by Goto (1983) led to following conclusions
on the tolerance of the people. It has shown that not only the acceleration causes the motion sickness
among people, but also the duration of motion can caused the same dilemma. As given in Table 2,
perception threshold value for human has been identified as 0.05 m/s2.
1. People have motion thresholds which can be measured.
2. Human responsiveness, including perception and tolerance, appear to be related to the acceleration
of the head.
3. People can generally perceive motion before furniture and fixtures begin to move or make sounds.
4. Human task performance and movement of objects are affected by the acceleration of the floor.
5. Even when acceleration is not great, the duration of motion can cause people to feel sick.

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


Table 2: The threshold acceleration defined by Goto et al.(1983)

Activity Threshold value


The perception threshold <0.05 m/s2
The limit of psychological and task performance 0.4 m/s2
The limit of walking 0.5-0.7 m/s2
Maximum limit of building motion should not be allowed to exceed 0.80 m/s2
There are many studies performed over the years to define relevant standards for human perception of
tall buildings. However, no internationally agreed guidelines have been developed to quantify an
acceptable level of building motion (Kwok et al., 2009). In addition, non existence of relevant
standards, the new trend towards super-tall buildings which built with the aid of high strength
materials, advanced construction techniques and sophisticated computer analysis make the problem
significantly worst. Further, medium-rise slender structures built with low damping (Table 3) can also
cause similar concerns for the occupants.
Table 3:Average threshold of perception for three towers in Australia (Denoon et al., 2000)

Location Natural frequency Average threshold of perception,


(Hz) peak acceleration (mg)

Brisbane Airport Control Tower 0.54 2.5

Sydney Airport Control Tower 0.94 2.4

Port Operations Airport Control Tower 0.39 2.8

3.2 Pedestrian Comfort


Wind comfort and wind safety for pedestrians are important requirements in the development of urban
areas. Many city authorities request studies of pedestrian wind comfort and wind safety for new
buildings. Bottema (2000) has performed a comprehensive study of the different safety and comfort
criteria used through the recent decades.
Table 4: Observed wind effects on people as a function of gust speed Ug and estimated gust duration tg
(Bottema, 2000)
Ug (m/s) Tg (s) Wind effect
4.0 5 Clothing flaps
5.0-6.0 5 Hair is disturbed/ Hair disarranged
10.0 3 Irregular footsteps, walking difficult to control
3 Difficult to hold umbrella (wind tunnel)
12.0 5 Violent flapping of clothes
10 Appreciably slowed into the wind
15.0 3 Walking difficult; dangerous for elderly person
3 Impossible to hold umbrella (wind tunnel)
16.0 2 Blown sideways
10 Appreciably slowed into the wind
18.0 10 Almost halted in the wind
10 Uncontrolled tottering walking downwind
20.0 3 Great difficulty with balance in gusts
23.0 3 People blown over by gusts
24.0 2 Unbalanced, grabbing at supports

In Table 4, 15 different categories of pedestrian comfort have been identified based on the gust wind
speed and its duration. Therefore, one needs to consider both gust speed and its occurrence time
period in terms of analysing the pedestrian comfort during severe wind action.

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


Also the design criteria have been derived through experimental observation and data collection as
given in Eq.(1) and Eq.(2).

Comfort criteria,
+ > 6/ = 15% (1)

Safety criteria
+ 3 > 20/ = 0.18% (2)

Where U is the mean velocity and is the standard deviation of instantaneous velocity and P max is
the maximum exceedence probability. This yields that the comfort level can be exceeded 1314
hours/year, while the safety level only can be exceeded 16 hours/year.
These parameters are key in designing and planning stages of buildings in order to comply with
environmental and safety guidelines. Therefore, the development of proper understanding about the
wind behavior around the building is a significantly important aspect in property development
process. As an example, the wake regions that can be created in the vicinity of tall buildings can cause
significant discomfort to the pedestrians walking in nearby streets. The shielding effect from tall
structures can create low pressure regions in the vicinity of those buildings. Also, the tunneling effect
created by the presence of two or more nearby tall buildings, can ultimately result in creating high
velocity regions. Therefore, better understanding of wind flow and influence of the proposed structure
on the existing wind pattern are important at the design stage of any building.

As highlighted in sections above, the development of proper methodology to assess wind induce
pressure on buildings, dynamic wind behavior around the building, alterations caused by the new
building to the existing wind pattern have become a necessity in todays building design and town
planning sectors. Due to the high cost and time consuming nature of commonly used wind tunnel
methods, and many changes to the shape and form of the building during the concept stage, it has
become a priority in wind and aero dynamic engineers to search an alternative ways of estimating the
wind induced forces and changes made to the surrounding environment due to new structures.
Computational Fluid Dynamics (CFD) has become a key alternative for the wind tunnel tests where
the virtual wind tunnel can be simulated in a computational environment.

Following section of this paper gives a brief introduction to the history of CFD technique, its
application in many different fields and the work that have been carried out by the authors to simulate
the wind flow around a 500 m tall rectangular building.

4.0 CFD IN WIND ENGINEERING

4.1 Development of CFD technique and its current applications


The basic equations describing fluid dynamics have been formulated as early as the 18th century,
most notably by Daniel Bernoulli and Leonhard Euler who realised mathematical models for flow of
inviscid fluids. In the nineteenth century, Claude-Louis Navier and Sir George Stokes incorporated a
viscosity term in the Eulers equation to obtain the Navier-Stokes equation, an accurate model for
simulating viscous fluid flow, which forms the basis of modern day CFD. With the development of
the Navier-Stokes equation, the possibility of using numerical methods to accurately simulate fluid
flow was conceptually conceived. Nonetheless, no applications of the technique were possible at the
time due to the absence of automated computers. By the turn of the twentieth century, an increase in
demand of the industry for fluid simulation, especially in the areas of aeronautical and naval
engineering, gave rise to numerous approximate and semi-analytical methods. These methods
required significantly shorter computation times, but were extremely limited in their application. It
was not until late 1950s that the development of CFD truly took off. The advancement was pioneered
by the T-3 Group at Los Alamos National Laboratory, led by Francis H. Harlow. In the years 1958 to
1968, the team at Los Alamos developed many of the currently used numerical methods in CFD,
including an early version of the k- turbulence model (Johnson, 1996). The methods developed by
Harlow in the 1960s were capable of solving truly unsolvable problems of the time (Johnson,
1996), opening up many research possibilities. One major advancement that followed was in 1974,

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


when a research group led by Professor Brian Spalding at Imperial College London developed the
Standard k- turbulence model that is used today (Khalil, 2012). The advancements in CFD were
more application driven in 70s and 80s (Johnson, 1996), with the growth being stimulated by ever
increasing computing power. During this time, CFD moved from the academic into the commercial
field with the development of generic commercial tools such as CHAM in 1974, FLUENT in 1983,
and Star-CD in 1987. With the increasing computational power nowadays, the applications of CFD in
the fields of engineering design and product development are limitless.

The cost effectiveness and potential accuracy of CFD has seen it expand into many other fields
including biomedical, chemical, electronic, environmental, and civil applications. For example, a
recent medical study by (Morales et al., 2013) used CFD techniques to simulate postsurgical blood
flow in patients undergoing endovascular coiling treatment in the brain. The research aimed to
determine the effect that blood viscosity has on the blood flow in those patients. Another research
application that extends beyond fluid flow simulation is a study by (Mattarelli et al., 2014) who used a
3 dimensional CFD model to simulate the combustion process in a diesel engine in an effort to reduce
carbon dioxide emissions.

4.2 Case study of rectangular 600 m super tall building

The intention of the following sections is to give the reader a basic overview of the approach that can
be used with CFD. A full scale geometry was modeled using commercial software ANSYS (ANSYS
Inc, 2012). The building has a rectangular prismatic shape with dimensions 100 m (x) by 150 m (y) by
600 m (z) height representing a true scale building in an open terrain. The flow is described in a
Cartesian coordinate system (x, y, z), in which the Z-axis is aligned with the stream flow direction, the
X-axis is in the perpendicular direction and the Y-axis is in the vertical direction. The computational
domain dimensions, boundary conditions and the wind tunnel configurations are given in Figure 2 . A
half of the model has been used in the analysis to save computation cost using the symmetricity of the
building and the fluid domain. As shown in Figure 2 the inlet and outlet boundaries have been
extended to 8 times and 25 times the width of the building in order to obtain the undisturbed flow near
the fluid domain boundaries.

Figure 2: (a)-(c) Geometry of the numerical model (d) Wind velocity profile used in this study,
Mendis et al. (2007)

An engineering wind model for Australia has been developed in Melbourne from the Deaves and
Harris model (D&H model,1978) (Mendis et al., 2007). This model has been developed based on full
scale data and on the classic logarithmic law from which a mean velocity profile in strong winds
applicable in non-cyclonic regions is derived, as given in Eq.(1).

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


2 3 4
u z z z z z
z
V [log e ( ) + 5.75 ( ) 1.88 ( ) 1.33 ( ) + 0.25 ( ) ] (1)
0.4 z0 zg zg zg zg
u z (2)
z
V log e ( )
0.4 z0
The numerical values are based on a mean gradient wind speed of 50 m/s. For values of z<30.0 m, the
z/zg values become insignificant and the Eq.(1) can be simplified to Eq.(2).

Where, Vz is the design hourly mean wind speed at height z, u is the friction velocity as described in
Mendis et al.(2007). This wind profile has been integrated as a user defined function (UDF) in
addition to a constant velocity profile in simulation of incoming wind. A constant tturbulent intensity
of 1% was adopted in the study.
Table 5 Configurations used in the study
Case Configuration Velocity profile
Case 1 Isolated Continuous flow
Case 2 Isolated Use D&H model
Case 3 half height adj. Bldg upwind of PB Use D&H model

In the present study, three different configurations with two different flow fields have been
investigated (Table 5). Case I simulates the wind effects on an isolated building with a constant
velocity profile. Case 2 uses same building model with the D &H velocity profile described above.
Case 3 investigates the effect of wind flow due to the presence of a building with half in height to the
principal building considered. Although not proven in a systematic study, it is believed that, in
general, a half height building could provide maximum interference effects.
4.3 Velocity Development
Flow behaviour around the building under a constant velocity profile is shown in Figure 3. The
preliminary simulations clearly show the vortex formation in the wake of the bluff body. The
simulations were able to capture flow separations and vortex formation quiet well. Even though the
incident velocity is around 40 m/s, the sharp edges near the building corners have caused an increment
in the velocity of the stream lines and jumped approximately closer to 50 m/s in the vicinity of the
front edges. The above results agree with the results published by Baetke et al. (1990).

Figure 3 Velocity stream lines of the flow (Case 1) (a) 3D flow domain (b) At z=150 m (c) At the
symmetrical plane

The pressure coefficients of the building surface along the symmetrical plane of the building were
obtained and presented as shown in Figure 4. The pressure induced on the front surface (A- B) does
not show considerable variation as expected in a real building under a constant velocity profile. This
simulation was only used as a comparison the difference in pressure on the building with the
application of atmospheric boundary layer wind profile (Figure 5 and Figure 6). According to the

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


results obtained from the analysis shows that the building front face has a maximum pressure
coefficient of 1.26 while downwind face has a pressure coefficient of -0.95. These results are in
agreement with the experimental results published by Dagnew et al. (2009).

Obtaining the correct pressure


development on the building surface is
very important in the designing of tall
structures to predict the behaviour of the
structures correctly. This will help the
designers to accurately predict the
acceleration of the building which is one of
the key elements as discussed in previous
sections. The user defined velocity (UDF)
profile was used in the analysis (case 2-3)
to represent the atmospheric boundary
layer wind profile.

Figure 4 Pressure coefficients along the symmetrical


plane (A-B)

Figure 5 Velocity distribution incorporating the D&H model(a) velocity vector formation (b) velocity
contours (Case 2) (c) Velocity stream lines at two different levels of the buildings (a) z=150 m (b)
z=480m

Figure 6 Velocity vectors development around the buildings (case 3)

The boundary layer wind profile generation, relevant pressure distribution for configuration 2 is given
in Figure 5. It also clearly shows the flow separation near the building edges and rejoining them in the
vicinity of the downwind direction of the building. Figure 6 captured from the case 3 where the effect

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


of nearby building were examined. The turbulence generation between two buildings was clearly
visible through the simulations.

CONCLUSIONS
Wind design of tall buildings and trend towards building supper tall building were discussed.
Limitations of current design codes in terms of applying them in analysing wind induced motion of
tall buildings were discussed. Six international wind design codes were investigated and found that
only the American and Japanese codes are not given any restriction over the building height. All the
other major codes are not applicable to buildings over 200m height. In addition to the limitations
given in the standards, two other key concern of the tall building designing were identified, namely;
perception of motion and pedestrian comfort in the vicinity of the buildings. Therefore comprehensive
wind analysis during the planning and design stages of a building showed a significant importance. As
an alternative approach to the time consuming and costly wind tunnel tests, CFD based approach to
predict wind behaviour of tall building were briefly discussed and presented through a case study.
Results show considerable accuracy and potential of using CFD in wind analysis of tall building in the
future.

REFERENCES

American Society of Civil Engineers (2010). ASCE/SEI 7-10 Minimum design loads for buildings and other
structures, American Society of Civil Engineers.
ANSYS Inc (2012). ANSYS CA,USA, Century Dynamics Inc.
Architectural Institute of Japan (2004). RLB recommendations for loads on buildings. Tokyo, Structural
Standards Committee.
Baetke, F., H. Werner, et al. (1990). "Numerical-Simulation of Turbulent-Flow over Surface-Mounted Obstacles
with Sharp Edges and Corners." Journal of Wind Engineering and Industrial Aerodynamics 35(1-3):
129-147.
Bottema, M. (2000). "A method for optimisation of wind discomfort criteria." Building and Environment 35(1):
1-18.
Council on Tall Buildings and Urban Habitat (2014). 100 tallest completed buildings in the world, The
Skyscraper Center.
Denoon, R. O., R. D. Roberts, et al. (2000). Field experiments to investigate occupant perception and tolerance
of wind-induced building motion, Department of Civil Engineering, University of Sydney, Australia.
European Committee for Standardization (2010). EN 1991-1-4:2005 Eurocode 1: Actions on structures - Part 1-
4: General actions - Wind actions.
Goto, T. (1983). "Studies on wind-induced motion of tall buildings based on occupants' reactions." Journal of
Wind Engineering and Industrial Aerodynamics 13(13): 241-252.
Johnson, N. L. (1996). The Legacy and Future of CFD at Los Alamos.
Khalil, E. (2012). "CFD history and applications." CFD Letters 4(2): 43-46.
Kwok, K. C. S., P. A. Hitchcock, et al. (2009). "Perception of vibration and occupant comfort in wind-excited
tall buildings." Journal of Wind Engineering and Industrial Aerodynamics 97(78): 368-380.
Kwon, D. K. and A. Kareem (2013). "Comparitive study of major international wind codes and standards for
wind effects on tall buildings." Engineering Structures 51(23-35).
Lamb, S., K. C. S. Kwok, et al. (2013). "Occupant comfort in wind-excited tall buildings: Motion sickness,
compensatory behaviours and complaint." Journal of Wind Engineering and Industrial Aerodynamics
119(0): 1-12.
Mattarelli, E., C. A. Rinaldini, et al. (2014). "CFD-3D analysis of a light duty dual fuel (diesel/natural Gas)
combustion engine." Energy Procedia 45(0): 929-937.
Mendis, P., T. Ngo, et al. (2007). "Wind Loading on Tall Buildings." Electronic Journal of Structural
Engineering: 41-54.
Ministry of Construction of the People's Republic of China (2012). GB50009-2012 Load code for the design of
building structures. Beijing, China Architecture and Building Press.
Montazeri, H. and B. Blocken (2012). "CFD simulation of wind-induced pressure coefficients on buildings with
and without balconies: Validation and sensitivity analysis." Building and Environment 60: 137-139.
Morales, H. G., I. Larrabide, et al. (2013). "Newtonian and non-Newtonian blood flow in coiled cerebral
aneurysms." Journal of Biomechanics 46(13): 2158-2164.

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)


Nguyen, K. C. (2009). A study of aerodynamic wind loads on tall buildings using wind tunnel tests and
numerical simulations. Department of Civil and Environmental Engineering, The University of
Melbourne: 260.
Safarik, D. and A. Wood (2014). CTBUH year in review: Tall trends of 2013, The Council on Tall Buildings
and Urban Habitat.
Standards Australia Limited/Standards New Zealand (2011). AS/NZS 1170.2:2011 Structural design actions -
Part 2: Wind actions. Sydney, Australia, SAI Global Limited.
Thakur, P. (2014). "Singapore's soaring land prices 'suicidal' for developers." Retrieved 22 May, 2014.
Vafaeihosseini, E., A. Sagheb, et al. (2013). Computational fluid dynamics approach for wind analysis of
highrise buildings. A3C12: Awards Convention and Consultants Colloquium. India.
Westbury, P. and S. Miles (2002). Practical tools to assess wind effects on buildings for environmental and
HVAC design. CIBSE Conference 2002. BRE, Watford.

1st International Conference on Infrastructure Failures and Consequences (ICIFC2014)

Das könnte Ihnen auch gefallen