You are on page 1of 137

WIND EFFECT ON SUPER-TALL BUILDINGS

USING COMPUTATIONAL FLUID DYNAMICS AND STRUCTURAL DYNAMICS

by

Bilal Assaad

A Thesis Submitted to the Faculty of

The College of Engineering and Computer Science

In Partial Fulfillment of the Requirements for the Degree of

Master of Science

Florida Atlantic University

Boca Raton, Florida

May 2015
Copyright 2015 by Bilal Assaad

ii
ACKNOWLEDGEMENTS

The author wishes to express sincere gratitude to his committee members for all of their guidance

and support in the development of this manuscript. Special thanks to my thesis advisor, Dr. M.

Arockiasamy, Professor and Director, Center for Infrastructure and Constructed Facilities, Department of

Civil, Environmental and Geomatics Engineering. Dr. M. Arockiasamy provided the author with incessant

support, persistence, patience and encouragement during the evolution of this thesis. The author is grateful

to Dr. Yan Yong, and Dr. Panagiotis Scarlatos for serving as fair evaluators of the work presented herein.

Thankfulness must also be extended to Dr. Chaouki Ghenai for the enlightenment provided to the author

with his precise and comprehensive expertise in the subject matter of Computational Fluid Dynamics. Last

but not least, Mr. Sijal Ahmed must be acknowledged as a brilliant tutor in the manipulation of the

advanced software package utilized for the analysis of the work presented in this publication.

iv
ABSTRACT

Author: Bilal Assaad

Title: Wind Effect on Super-Tall Buildings Using Computational Fluid Dynamics and
Structural Dynamics

Institution: Florida Atlantic University

Thesis Advisor: Dr. Madasamy Arockiasamy

Degree: Master of Science

Year: 2015

Super-tall buildings located in high velocity wind regions are highly vulnerable to large lateral

loads. Designing for these structures must be done with great engineering judgment by structural

professionals. Present methods of evaluating these loads are typically by the use of American Society of

Civil Engineers 7-10 standard, field measurements or scaled wind tunnel models. With the rise of high

performance computing nodes, an emerging method based on the numerical approach of Computational

Fluid Dynamics has created an additional layer of analysis and loading prediction alternative to

conventional methods. The present document uses turbulence modeling and numerical algorithms by means

of Reynolds-averaged Navier-Stokes and Large Eddy Simulation equations applied to a square prismatic

prototype structure in which its dynamic properties have also been investigated. With proper modeling of

the atmospheric boundary layer flow, these numerical techniques reveal important aerodynamic properties

and enhance flow visualization to structural engineers in a virtual environment.

v
DEDICATION

This manuscript is in part and as whole dedicated to my humble and unconditionally loving

family. Their relentless support during this journey has provided me with the courage to develop,

overcome, and realize this level of intellectuality. I would like to particularly dedicate this thesis to my

Father and Mother for nurturing a family with genuine love, regardless of the circumstances and hardships

they have encountered and surpassed in their destined lives. I will forever be indebted to their kindness,

exemplary courage, and love. This document is also dedicated to the people that have embraced me as a

friend, brother and/or son and provided me with the opportunity to reach this accomplishment. Lastly, the

work presented herein is dedicated to my loving partner that kindly brightens my heart, mind, and soul.

vi
WIND EFFECT ON SUPER-TALL BUILDINGS

USING COMPUTATIONAL FLUID DYNAMICS AND STRUCTURAL DYNAMICS

LIST OF TABLES .......................................................................................................................................... x


LIST OF FIGURES ........................................................................................................................................xi

CHAPTER 1: INTRODUCTION.................................................................................................................... 1
1.1. Wind Engineering Tools ................................................................................................................... 1
1.2. Study Objectives ............................................................................................................................... 4

CHAPTER 2: LITERATURE REVIEW......................................................................................................... 5


2.1. Overview ........................................................................................................................................... 5
2.2. Wind Effect on Structure .................................................................................................................. 5
2.3. Wind Damaged Structures ................................................................................................................ 6
2.4. Computational Wind Engineering..................................................................................................... 9
2.4.1. Microscale and CFD ................................................................................................................ 11
2.4.2. Reduced-scale Wind Tunnel Testing and CWE ...................................................................... 12
2.5. Wind Tunnel Measurements ........................................................................................................... 12
2.6. Overview of Tall Buildings............................................................................................................. 14
2.6.1. Factors Affecting Growth, Height, and Structural Form ......................................................... 14
2.6.2. Criteria for the Definition of Tall Buildings ............................................................................ 16
2.7. Wind Loading on Tall Buildings..................................................................................................... 18

CHAPTER 3: ELEMENTARY PRINCIPLES OF WIND ENGINEERING................................................ 19


3.1. Forces Acting in the Free Atmosphere............................................................................................ 19
3.1.1. Pressure Gradient and the Coriolis Force ................................................................................ 19
3.2. Geostrophic Wind, Gradient Wind, and Frictional Effects ............................................................. 20
3.3. Atmospheric Boundary Layer ......................................................................................................... 21
3.3.1. Mean Wind Profiles................................................................................................................. 21
3.3.1.1. The Logarithmic Law ..................................................................................................... 22
3.3.1.2. The Power Law ............................................................................................................... 23
3.3.2. Turbulence ............................................................................................................................... 24
3.3.2.1. Turbulence Intensities ..................................................................................................... 25
3.3.2.2. Integral Turbulent Length Scale ..................................................................................... 27

vii
CHAPTER 4: THEORY OF COMPUTATIONAL FLUID DYNAMICS ................................................... 29
4.1. Overview ......................................................................................................................................... 29
4.2. Computational Fluid Dynamics ...................................................................................................... 30
4.3. Governing Equations of Fluid Dynamics ........................................................................................ 31
4.3.1. Infinitesimal Fluid Element ..................................................................................................... 31
4.3.2. The Continuity Equation ......................................................................................................... 31
4.3.3. The Momentum Equation ........................................................................................................ 34
4.3.4. The Energy Equation ............................................................................................................... 35
4.4. Turbulence and Modeling for CFD ................................................................................................. 36
4.4.1. Turbulent Scales ...................................................................................................................... 38
4.4.2. Energy Spectrum ..................................................................................................................... 38
4.4.3. Transition from Laminar to Turbulent Flow............................................................................ 40
4.4.4. Turbulent Flow Calculations ................................................................................................... 40
4.4.5. RANS Equations and Classical Models .................................................................................. 41
4.4.5.1. κ-ε Model ........................................................................................................................ 43
4.4.5.2. κ-ω Model ....................................................................................................................... 43
4.4.5.3. SST κ-ω Model ............................................................................................................... 44
4.4.6. Large Eddy Simulation ............................................................................................................ 45
4.4.6.1. Smagorinsky-Lilly Model ............................................................................................... 47
4.4.6.2. Remarks on LES ............................................................................................................. 47
4.4.7. Typical Meshing Information .................................................................................................. 48
4.4.8. Near Wall Treatment ............................................................................................................... 50
4.4.9. Discretization Techniques and Algorithms.............................................................................. 51

CHAPTER 5: THEORY OF STRUCTURAL DYNAMICS ........................................................................ 54


5.1. Basic Concepts of Vibration ........................................................................................................... 54
5.2. Degrees of Freedom and Classification of Vibration ...................................................................... 55
5.3. Equation of Motion and Natural Frequency .................................................................................... 56
5.4. Property of Matrices for Vibrating Systems ................................................................................... 58
5.4.1. Flexibility and Stiffness Matrix ............................................................................................... 58
5.4.2. Mass Matrix ............................................................................................................................. 59
5.5. Eigenproblem Formulation, Natural Vibration Frequencies and Modes......................................... 60

CHAPTER 6: CFD ANALYSIS OF WIND FLOW USING RANS & LES FORMULATIONS ................ 62
6.1. 2D & 3D Setup of Model in CFD ................................................................................................... 62
6.2. Grid Independent Study .................................................................................................................. 64
6.3. Model Validation ............................................................................................................................ 68
6.4. Discussion of Results Based on 2D Plan Model ............................................................................. 71

viii
6.5. 3D LES Model ................................................................................................................................ 73
6.6. Pressures Acting on Slender Structure ............................................................................................ 75

CHAPTER 7: BASIC STRUCTURAL AND DYNAMIC ANALYSIS OF TALL BUILDING ................. 77


7.1. Overview of Idealized Example Structure ...................................................................................... 77
7.2. Assigned Loading and Section Properties to Structural Members .................................................. 78
7.3. Idealized Structure Discretization ................................................................................................... 81
7.4. Vibration Modes of Modeled Structure .......................................................................................... 82
7.5. Deflections of Super-Tall Structure ................................................................................................ 85

CHAPTER 8: CONCLUSIONS & FUTURE RESEARCH.......................................................................... 88


8.1. Final Remarks ................................................................................................................................. 88
8.2. Recommendations for Future Research .......................................................................................... 89

APPENDICES............................................................................................................................................... 91
Appendix A................................................................................................................................................ 92
Appendix B .............................................................................................................................................. 102
Appendix C .............................................................................................................................................. 109
Appendix D.............................................................................................................................................. 111

REFERENCES ............................................................................................................................................ 120

ix
LIST OF TABLES

Table 1: Roughness lengths and Surface drag coefficients per ASCE 7-10 .................................................. 23
Table 2: Power law exponents and gradient height per ASCE 7-10 Standard .............................................. 23
Table 3: Experimental data and derived quantities for various cross sections .............................................. 69
Table 4: Model input criteria for 2D plan CFD ............................................................................................. 72
Table 5: Assumed loads acting on structure .................................................................................................. 79
Table 6: Cross-section and properties of members along the height of the structure .................................... 80

x
LIST OF FIGURES

Figure 1: Typical wind tunnel setup of a model high rise building ................................................................. 3
Figure 2: (a) Moment of collapse of Cooling Tower 2A, U.K. (b) Collapse of midsection of Tacoma
Narrows Bridge, WA ........................................................................................................................... 7
Figure 3: (a) Shiten'noji Pagoda collapse in 1937 and (b) Present day rebuilt structure ................................. 8
Figure 4: (a) One Indiana Square tornado induced damage (b) Remodeled façade of building ...................... 9
Figure 5: (a) Woolworth Building, NY (b) Empire State Building, NY ....................................................... 15
Figure 6: Comparison of tall, supertall and megatall building height criteria ............................................... 17
Figure 7: World's ten tallest buildings according to 'height to architectural top' as of November 2014 ....... 18
Figure 8: Coriolis force results in wind being deflected owing to the rotation of the Earth .......................... 20
Figure 9: Wind profile in different boundary layers ...................................................................................... 22
Figure 10: Comparison of the logarithmic and power law for mean velocity profile for z0 = 0.02 m
and α = 0.128 ..................................................................................................................................... 24
Figure 11: Typical point velocity measurements in turbulent flow ............................................................... 25
Figure 12: Longitudinal turbulence intensity for rural terrain (z0 = 0.04 m) and suburban terrain (z0 =
0.15 m) ............................................................................................................................................... 26
Figure 13: Variation of turbulent length scale as height increases per AIJ code ........................................... 28
Figure 14: Three dimensions of fluid dynamics ............................................................................................ 29
Figure 15: Infinitesimal fluid element model of flow ................................................................................... 31
Figure 16: Model of infinitesimally small element fixed in space including mass flux diagram .................. 32
Figure 17: Model used for the derivation of the x-component of momentum equation ................................ 34
Figure 18: von Kármán vortices representing turbulent flow characteristics forming in clouds
flowing past a volcano ....................................................................................................................... 37
Figure 19: Representation of cascade process with a spectrum eddies ......................................................... 38
Figure 20: Spectrum for turbulent kinetic energy, κ...................................................................................... 39
Figure 21: Transition of flow in a pipe from laminar to turbulent................................................................. 40
Figure 22: Increase of computational cost per type of turbulence modeling ................................................. 41
Figure 23: Spectrum of velocity .................................................................................................................... 46
Figure 24: Typical 2D and 3D Mesh Shapes................................................................................................. 49
Figure 25: Typical geometry meshing process .............................................................................................. 49
Figure 26: Log-Law of the Wall .................................................................................................................... 50
Figure 27: Discretization of grid points......................................................................................................... 51
Figure 28: Idealized multi-story building with five degrees of freedom ....................................................... 56

xi
Figure 29: Model of a simple SDOF and free-body diagram of system ........................................................ 57
Figure 30: Flexibility coefficients for a beam ............................................................................................... 59
Figure 31: Lumped mass coefficients for structural elements with distributed mass .................................... 59
Figure 32: 3D Boundary conditions .............................................................................................................. 63
Figure 33: 2D Plan boundary conditions ....................................................................................................... 63
Figure 34: 2D Elevation boundary conditions ............................................................................................... 64
Figure 35: 157 K Nodes quad-mesh for 2D plan view of structure ............................................................... 65
Figure 36: Quad-mesh for 2D elevation of structure ..................................................................................... 66
Figure 37: Boundary layer development along wind tunnel section ............................................................. 67
Figure 38: Variation of mean pressure coefficient along the windward face for grid analysis ..................... 68
Figure 39: Drag coefficient curve for 44K SST-κω transient model ............................................................. 69
Figure 40: Lift coefficient curve for 44K SST-κω transient model ............................................................... 70
Figure 41: Mean velocity magnitude along centerline .................................................................................. 71
Figure 42: Variation of mean pressure coefficient along faces ..................................................................... 72
Figure 43: 3D mesh with 7.7 M nodes .......................................................................................................... 73
Figure 44: Wind profile generated in computer domain................................................................................ 74
Figure 45: Present LES windward contours compared to experimental data ................................................ 75
Figure 46: Windward estimated pressure variation ....................................................................................... 76
Figure 47: Framing model of 67-story tower used in present study .............................................................. 78
Figure 48: First three modes of vibration of tall structure ............................................................................. 83
Figure 49: Fourth to sixth modes of vibration of tall structure...................................................................... 84
Figure 50: Deflections experienced by the loading acting on super-tall structure......................................... 86

xii
NOMENCLATURE

Acceleration x Standard Deviation 


Atmospheric pressure p Stiffness matrix k 
Boundary layer value of first grid pt. y Surface shear stress Csd
Coef. dependent on terrain roughness  Temperature T
Constant Velocity U Terrain exposure constant 
Coriolis parameter f Time t
Density  Total internal energy e
Density of air a Turbulence frequency 
x Turbulence Intensity in i -th TIi
Displacement
direction
Distance d Turbulent kinetic energy 
Dynamic viscosity  Velocity x
Elevation or height z Viscous damping c
Equivalent height z von Kármán’s constant k
Flexibility matrix a  Natural circular frequency 0
Force F Radius r
Frequency f Rate-of-strain tensor S ij
Friction velocity u* Roughness coefficient of terrain z0
Integral length scale in i -th direction Li Shear stress 
Kinematic velocity v Spring constant or stiffness k
Kinematic viscosity  Standard Deviation 
Kinetic energy dissipation  Stiffness matrix k 
Latitude  Surface shear stress Csd
Local grid scale  Temperature T
Mass matrix m  Terrain exposure constant 
Mean wind speed V Time t
Mixing length for subgrid scales Ls Total internal energy e
Modal matrix   Turbulence frequency 
Natural circular frequency 0 Turbulence Intensity in i -th TIi
direction
Radius r Turbulent kinetic energy 
Rate-of-strain tensor S ij Velocity x
Roughness coefficient of terrain z0 Viscous damping c
Shear stress  von Kármán’s constant k
Spring constant or stiffness k

xiii
CHAPTER 1

INTRODUCTION

The development of new construction techniques in the 20th century has created structures that are

flexible, low in damping, and relatively light in weight which therefore exposes the structure to the effect of

wind acting upon it. Wind engineering has been the field with the aim of primarily developing tools to

better understand the action of the fluid on the structure with origins that could be traced back to the 1960s.

The development of knowledge found in the present literature regarding this subject has lead structural

engineers to design and ensure the performance of the structure subjected to the action of wind to be within

adequate limits during the lifetime of the structure in structural safety and serviceability criteria (Simiu and

Scanlan 1978).

1.1. Wind Engineering Tools

Computational Wind Engineering (CWE) is primarily defined as the use of Computational Fluid

Dynamics (CFD) for wind engineering applications, although it also includes other approaches of computer

modelling and in the broadest sense also field and wind tunnel measurements supporting CWE model and

development. Wind engineering itself is best described as “the rational treatment of interactions between

wind in the atmospheric boundary layer and man and his works on the surface of Earth.” (Cermak 1975).

In spite of these difficulties, in the past decades and driven by the pioneering studies, CWE has

undergone a successful transition from an emerging field into an increasingly established field in wind

engineering research, practice, and education. This transition and the success of CWE are illustrated by: (1)

the establishment of CWE as an individual research and application area in wind engineering with its own

successful conference series, (2) the increasingly wide range of topics covered in CWE, ranging from

pedestrian-level wind conditions over natural ventilation of buildings and wind loads on buildings and

bridges to sports aerodynamics, and (3) the history of review and overview papers in CWE.

CWE/CFD has some particular advantages over experimental (full scale or reduced scale) testing.

It can provide detailed information on the relevant flow variables in the whole calculation domain, under

1
well-controlled conditions and without similarity constraints. The accuracy and reliability of CFD

simulations are of concern and solution verification and validation studies are essential. Therefore,

experiments remain indispensable for CWE.

As already previously mentioned, CWE is complementary to other, more traditional areas of wind

engineering, such as full scale on-site experimentation and reduced scale wind tunnel testing. Each

approach has its specific advantages and disadvantages. CFD has some particular advantages over

experimental testing (full scale or reduced scale), especially the fact that it provides detailed information on

the relevant flow variables in the whole calculation domain (whole flow field data) under well controlled

conditions and without similarity constraints.

The advancement of high-speed digital computer combined with the accurate numerical

algorithms for solving physical problems on these computers revolutionized the way we study and practice

fluid dynamics today. It has introduced a fundamentally important new third approach in fluid dynamics—

the approach of computational fluid dynamics.

Computational fluid dynamics results are analogous to wind tunnel results obtained in a laboratory

in which they represent sets of data for given flow configurations at different Reynolds numbers. In other

words, a computer program with CFD code imbedded into it acts like a “virtual wind tunnel.” Having

access to the program allows the user to carry out some interesting experiments with it. Experiments you

perform with the computer program are numerical experiments instead of physical ones.

Tallness is a relative matter, and tall buildings cannot be defined in specific terms related just to

height or to the number of floors. The tallness of a building is a matter of a person’s or community’s

circumstance and their consequent perception; therefore a measurable definition of a tall building cannot be

universally applied (Stafford Smith and Coull 1991). From the structural engineer’s point of view, a tall

building may be defined as one that because of its height, it is affected by lateral forces due to wind and

earthquake actions to an extent that they play a critical role in the structural design (Stafford Smith and

Coull 1991).

Tall buildings are subject to resonant wind loads. The primary source of the along-wind motion is

the pressure fluctuations in the windward and leeward faces, which are affected by the nature of the

turbulent wind and the interaction with the structure itself. The across-wind motion is caused by the

2
fluctuating separation in the shear layers of the fluid. Torsional motion is caused by the imbalance of

pressure distribution on the faces of the building which could be due to the varying angles of the wind

direction, interference effect of neighboring buildings, or due to varying building geometry and eccentricity

of the center of mass. Kareem (1985) has performed an experimental study in which it was concluded that

tall buildings are subject to greater magnitudes of the across-wind and torsional response when compared to

the along-wind response in terms of serviceability and limit state requirements.

The present engineer uses design codes and procedures to properly account for wind loading on

structures. The American Society of Civil Engineers Standard ASCE 7-10 gives provisions for the design

of Main Wind Force Resisting Systems (MWFRS) for building with common geometric shapes in different

types of exposures. Many present structures with complex geometries do not fall in the categories specified

by the code, and therefore it refers the engineer to implement the use of physical model and testing of the

project in boundary layer wind tunnel facilities. Model scale testing is the most common and widely used

engineering tool. Wind tunnel testing can predict wind-induced effects on structures that address some of

the difficulties in specifying the effect in codes.

Figure 1: Typical wind tunnel setup of a model high rise building


(Credit: RWDI Inc., Canada)

The concerns pertaining the limited provisions by standards and codes, the cost of physical testing

and scarce wind tunnel facilities, have encouraged researchers to adopt hardware and software technology

to investigate the potential of numerical modelling using the theory of computational fluid dynamics. This

provides and additional tool to the engineer and is also an alternative to common practice. The use of these

numerical methods is easily accessible and is expected to gain popularity in the practice of structural design

3
for wind loading on structures. This will result in more resilient and more sustainable systems by allowing

the engineer to adapt aerodynamic and a smart structure design.

The most significant parameter concerning the extent of the dynamic effect a load causes on a

structure is the natural period of vibration (Tedesco et al. 1999). The natural period of vibration simply is

the time required for the structure to go through one complete cycle of vibration (i.e. units of seconds). The

natural frequency is defined as the number of cycles the structure undergoes per second (i.e. units of cycles

per second or Hertz). If the load acting on a structure acts on a sufficiently large time period, the load can

be considered to be static. However, if the load acting on a structure is close to the natural period of the

structure, it will induce a dynamic response. It must be noted that the stresses, strains, and deflections are

generally more severe when loads of given amplitude acts dynamically. Wind loads are dynamic in nature

and they must be accounted for as essential loading criteria to ensure proper design of any given structure.

In short, wind loads are a function of velocity, height of structure, and shape and stiffness characteristics of

the structure.

1.2. Study Objectives

The main objective of this research was to investigate the strategies that can be adapted in the

computational evaluation and assessment of wind loads on tall buildings under turbulent wind flows. The

present study explored the numeric of the aerodynamics of a square slender building. The investigation was

carried using two-dimensional and three-dimensional study of the structure using CFD. Moreover, a basic

structural dynamic analysis was carried on a prototype building designed by the author. The building

analysis was deemed satisfactory based on industry deflection recommendations.

4
CHAPTER 2

LITERATURE REVIEW

2.1. Overview

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 its gusty or

turbulent character. Wind interaction with surface features gives rise to the gustiness of strong winds in the

lower levels of the atmosphere. 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. There are several

different phenomena giving rise to dynamic response of structures in wind. These include buffeting, vortex

shedding, galloping and flutter. Slender structures can be sensitive to dynamic response in line with the

wind direction as a consequence of turbulence buffeting. Transverse or cross-wind response is more likely

to arise from vortex shedding or galloping but may also result from excitation by turbulence buffeting.

Flutter is a coupled motion, often being a combination of bending and torsion, and can result in instability.

For building structures flutter and galloping are generally not an issue. An important problem associated

with wind induced motion of buildings is concerned with human response to vibration and perception of

motion. Humans are very sensitive to vibration to the extent that motions may feel uncomfortable even if

they correspond to relatively low levels of stress and strain. Therefore, for most tall buildings serviceability

considerations govern the design and not strength issues (Mendis et al. 2007).

2.2. Wind Effect on Structure

The greatest probability of damage to structures has been presented by Davenport (1963) to be the

case of strong winds with neutral atmospheric conditions. Davenport suggests that structural response to

repeated loads of successive gusts is an important factor in the design of tall buildings. Repeated loading

may lead to fatigue, failure, foundation settling, excessive deflections causing cracking to building

elements, or induced motion that may affect the comfort of the occupants of the structure. A building can

be considered to have failed if it becomes unserviceable due to the action of repeated loads or the action of

5
a single large load of great magnitude. It is very important that the fluctuating loads caused by wind on a

structure play an important role in the design and analysis of tall buildings, especially structures with large

aspect ratios (Reinhold 1977).

The primary concern for a structural engineer when studying wind phenomena, around a building,

is the mean velocity profile of the wind. Moreover, two aspects of turbulent flows are of interest to the

engineer: (a) the state of turbulence of the natural wind approaching a structure, and (b) the local turbulence

provoked in the wind by the structure itself. Most structures in civil engineering present bluff forms, in

wind engineering studies we focus on the bluff-body aerodynamics aspects of the wind and structure

interaction. This has led the industry to further research on the details of flow effects around bluff bodies

such as tall buildings. This finally leads to the interest of the engineer in the study of the development of

body pressures by the flow acting around a structure (Simiu and Scanlan 1978).

2.3. Wind Damaged Structures

Damage to buildings and other structures by windstorms has been a fact of life for human beings

from the time they moved out of cave dwellings to the present day. Trial and error has played an important

part in the development of construction techniques and roof shapes for small residential buildings, which

have usually suffered the most damage during severe winds. In past centuries, heavy masonry construction,

as used for important community buildings such as churches and temples, was seen, by intuition, as the

solution to resist wind forces. For other types of construction, windstorm damage was generally seen as an

‘act of god’, as it is still viewed today by many insurance companies.

The nineteenth century was important as it saw the introduction of steel and reinforced concrete as

construction materials, and the beginnings of stress analysis methods for the design of structures. The latter

was developed further in the twentieth century, especially in the second half, with the development of

computer methods. During the last two centuries, major structural failures due to wind action have occurred

periodically, and provoked much interest in wind forces by engineers. Long-span bridges often produced

the most spectacular of the failures, with the Brighton Chain Pier, England in 1836, the Tay Bridge,

Scotland in 1879, and Tacoma Narrows Bridge, Washington State, U.S.A. in 1940 being the most notable,

with the dynamic action of wind playing a major role (Holmes 2007).

6
Other large structures have experienced failures as well – for example, the collapse of the

Ferrybridge cooling towers in the United Kingdom in 1965, and the permanent deformation of the columns

of the Great Plains Life Building in Lubbock, Texas during a tornado in 1970. These events were notable,

not only as events in themselves, but also for the part they played as a stimulus to the development of

research into wind loading in their respective countries.

(a) (b)
Figure 2: (a) Moment of collapse of Cooling Tower 2A, U.K. (b) Collapse of midsection of Tacoma Narrows
Bridge, WA
(Credit: Shellard 1965, and Bashford and Thompsons 1940)

Some major windstorms, which have caused large scale damage to residential buildings, as well as

some engineered structures, are also important for the part they played in promoting research and

understanding of wind loads on structures. The effects of Hurricane Andrew in Florida proved to be the

costliest natural disaster in the state’s history. Andrew made landfall near Homestead, Florida on August

24, 1992 as a Category 5 hurricane. Strong winds from the hurricane affected four southeastern counties of

the state in which it damaged or destroyed 730,000 houses and buildings. The hurricane caused about $25

billion in damage and 44 deaths.

The first ‘tall buildings’ to appear in Japan might be the traditional wooden pagodas which are

seen in historic Japanese cities such as Nara and Kyoto. Strong typhoons could cause damage to pagodas.

The 5-story, 47.8 m (157.8 ft.) high Shiten’noji Pagoda collapsed due to typhoon Muroto on September 21,

1934. The maximum peak gust speed was estimated to be more than 60 m/s (134.2 mph) and was

accompanied by a high tidal wave of more than 4 meters (13.1 ft.). Thus, the history of the development of

design and construction methods for tall buildings was a record of fights with strong winds. There are many

7
wind-related problems in construction of tall buildings, but the main problem for engineers is their

capability of resistance to wind forces, because higher altitudes mean higher wind speeds, and consequently

higher wind forces (Tamura 2009).

(a) (b)
Figure 3: (a) Shiten'noji Pagoda collapse in 1937 and (b) Present day rebuilt structure

As well as damage to buildings produced by direct wind forces – either overloads caused by

overstressing under peak loads, or fatigue damage under fluctuating loads of a lower level, a major cause in

severe wind storms is flying debris. Penetration of the building envelope by flying ‘missiles’ has a number

of undesirable results: high internal pressures threatening the building structure, wind and rain penetration

of the inside of the building, the generation of additional flying debris, and the possibility of flying missiles

inside the building endangering the occupants. The area of a building most vulnerable to impact by missiles

is the windward wall region, although impacts can also occur on the roof and side walls. As the air

approaches the windward wall, its horizontal velocity reduces rapidly. Heavier objects in the flow with

higher inertia will probably continue with their velocity little changed until they impact on the wall. Lighter

and smaller objects may lose velocity in this region or even be swept around the building with the flow if

they are not directed at the stagnation point (Holmes 2007).

One Indiana Square is a 36-story (504 ft) tall building located in downtown Indianapolis, Indiana.

The building went exterior remodeling after damage by tornado-strength winds reaching speeds exceeding

130 km/h (80.7 mph) that occurred on April 2, 2006. This particular storm brought winds sufficient to

cause severe damage to the façade and structural elements of 16 out of 36 stories of the tower causing

millions of dollars in monetary loss and the closing of streets and businesses for several days. The nature of

8
the damage prompted debate about whether the damage was caused by tornado, downburst, or extreme

straight wind conditions. The recorded wind speed was very close to typical design wind speed for

buildings as recommended per ASCE 7-10 national standard. The recommended speeds are dependent

upon geographic locations in which the region of southeast Florida has the highest wind speed values. After

the 2006 damage, design of a new façade with curtain wall to be installed over the existing façade was

released in 2007 by the integrated design firm Gensler. The new façade after the re-cladding process

essentially put another layer of skin around the building’s exterior face by expanding it by 18 in. around its

perimeter (Yilmaz and Duffin 2014).

(a) (b)
Figure 4: (a) One Indiana Square tornado induced damage (b) Remodeled façade of building
(Credit: Chriss Barrett via CTBUH)

2.4. Computational Wind Engineering

The historical starting point of CWE could be situated around 1963 when Smagorinsky developed

one of the first successful approaches to Large Eddy Simulation (LES), the Smagorinsky-Lilly model,

which is still intensively used in many areas of fluid mechanics today. The main research area of

Smagorinsky was Numerical Weather Prediction applied at the meteorological macroscale. Of particular

importance for CWE were the pioneering studies by Meroney and his co-workers in which a hybrid

approach was pursued for the systematic comparison of numerical simulations with dedicated wind tunnel

measurements in atmospheric boundary layer wind tunnel (Meroney and Yamada 1971, Yamada and

Meroney 1972, Derickson and Meroney 1977). In Aerospace Engineering, the T3 group at the Los Alamos

9
National Laboratories in 1963 first used computers to model the 2D swirling flow around an object using

the vorticity stream function method, followed by the first 3D application by Hess and Smith (1967) using

the so called panel method. Driven by these early achievements, early efforts in CWE focused on the

determination and analysis of wind velocity and pressure field around buildings (Blocken 2014).

The difference in time between the earliest CFD developments in the 1950s and the later

application of CFD in CWE for wind velocity and pressure fields around buildings is attributed to the

specific difficulties associated with the flow around bluff bodies with sharp edges. Murakami (1998)

diligently outlined some of the difficulties encountered in CWE: (1) high Reynolds numbers in wind

engineering applications, necessitating high grid resolutions, especially near wall regions as well as

accurate wall functions, (2) the complex nature of the 3D flow field with impingement, separation and

vortex shedding, (3) the numerical difficulties associated with flow at sharp corners and consequences of

discretization schemes, and (4) the inflow and outflow boundary conditions which are particularly

challenging for LES. These difficulties were directly linked to limitations in physical modelling and in

computational requirements at those times, but many of those limitations are still to some extent present

today.

CWE is complementary to other, more traditional areas of wind engineering, such as full scale on-

site experimentation and reduced scale wind tunnel testing. Each approach has its specific advantages and

disadvantages. The main advantage of on-site measurements is that they are able to capture the real

complexity of the problem under study. To name a few, important disadvantage are that they are not fully

controllable due to the inherently variable meteorological conditions, that they are not possible state in the

design stages of the building, and that usually only point measurements are performed. The latter

disadvantage also hold true for wind tunnel measurements. Techniques such as Particle Image Velocimetry

(PIV) and Laser-Induced Fluorescence (LIF) in principle allow planar or even full 3D data to be obtained in

wind tunnel tests, but the cost is considerably higher and application for complicated geometries can be

hampered by laser-light shielding by the obstructions constituting the model. Another disadvantage is the

required adherence to similarity criteria in reduced scale testing, which can limit the extent and the range of

problems that can be studied in wind tunnels.

10
In addition, it is widely recognized that the results of CFD simulations can be very sensitive to the

wide range of computational parameters that have to be set by the modeler. For typical simulations, the user

has to select target variables, the approximate form of the governing equations, the turbulence model, the

computational domain, the computational grid, the boundary conditions, the discretization schemes, the

convergence criteria, etc. Therefore this expresses the need for best practice guidelines for CWE. CWE has

grown to a strongly established field in wind engineering research, practice and education. It is employed

daily by probably thousands of researchers, practitioners and teachers all over the world.

2.4.1. Microscale and CFD

At the microscale, the flow around surface mounted obstacles such as buildings is explicitly

resolved, i.e. these obstacles are represented with their actual shape. Yamada and Meroney (1972) studied

2D airflow over a square surface mounted obstacle in a stratified atmosphere, both in the wind tunnel and

with CFD. Hirt and Cook (1972) calculated 3D flow around structures and over rough terrain. CFD

simulations around 3D buildings started with fundamental studies of isolated buildings, often with a cubical

shape, to analyze the velocity pressure fields (Murakami and Mochida 1988, 1989; Baskaran and

Stathopoulos 1989, 1992). Together with later studies they laid the foundations for the current best practice

guidelines by focusing on the importance of grid resolution, the influence of boundary conditions on the

numerical results, and by comparing the performance of various types of turbulence models in steady

RANS simulations. Also steady RANS versus LES studies were performed (Blocken 2014).

In the past, especially the deficiencies of the steady RANS approach with the standard κ-ε model

(Jones and Launder 1972) for wind flow around buildings were addressed. These include the stagnation

point anomaly overestimation of turbulent kinetic energy near the frontal corner and the resulting

underestimation of the size and separation and recirculation regions on the roof and the side faces, and the

underestimation of turbulent kinetic energy in the wake resulting in an overestimation of the size of the

cavity zone and wake. Various revised linear and non-linear κ-ε models and also second-moment closure

models were developed and tested and showed improved performance for several parts of the flow flied.

However the main limitation of steady RANS modelling remained: its incapability to model the inherently

transient features of the flow field such as the separation and recirculation downstream of windward edges

and vortex shedding in the wake. These large-scale features can be explicitly resolved by LES. The studies

11
by Murakami et al. (1987), later by Murakami et al. (1990, 1992) illustrated the intrinsically superior

performance of LES compared to RANS. Nevertheless, LES entails specific disadvantages that are not easy

to overcome, including the strongly increased computational requirements and the difficulty in specifying

appropriate time-dependent inlet and wall boundary conditions (Blocken 2014).

2.4.2. Reduced-scale Wind Tunnel Testing and CWE

In the past decades, often statements have been made that CFD would replace reduced scale wind

tunnel testing and that it would be the numerical wind tunnel. Many scholars such as Castro and Graham

(1999) and Stathopoulos (2002) convincingly denounced the label without recognizing the important

complementary value and potential of CWE.

The complementary aspects of wind tunnel testing and CWE are multifold. Wind tunnel testing

can provide the indispensable high-quality validation data needed for CWE, and CWE can supplement

wind tunnel testing by providing whole flow field data on all relevant parameters. Leitl and Meroney

(1997) indicated the value of CFD to design wind tunnel experiments by using numerical codes that can

help design and setup wind tunnel experiments which can reduce the time required to optimize a physical

model and expensive pre-runs in a wind tunnel. Moonen et al. (2007) developed a series of new indicators

for wind tunnel test section flow quality and applied CFD to illustrate the effectiveness of these indicators.

This approach was adopted by Calautit et al. (2014) for further development of design methodologies of

closed-loop subsonic wind tunnels (Blocken 2014).

2.5. Wind Tunnel Measurements

Wind tunnel tests are powerful tools that give engineers the ability to estimate the nature and

intensity of wind forces acting on complex structures such as tall buildings. Wind tunnel testing is

especially useful when the surrounding terrain and the shape of the structure causes complex wind flows

that are not fully addressed by simplified codes (Samali et al. 2004). Many studies have been performed in

the measurements of wind loads on structures by either using full-scale measurements or by wind tunnel

model studies. As technology has advanced, the estimation of these forces have increased in reliability.

Wind loads are particularly important for flexible structures such as tall-buildings with low damping.

Typically, wind tunnel measurements are performed in boundary-layer wind tunnels that are

capable of developing flow conditions that meet these conditions (Taranath 2005):

12
1) The natural atmospheric boundary layer is modeled as such to account for the variation of

wind speed with height

2) The length scale of atmospheric turbulence is approximately the same scale as of that of the

building

3) The model building and surrounding topography are geometrically similar to the full-scale

4) The pressure gradient in the longitudinal direction is accounted for

5) Reynolds number effects on pressures and forces are kept to a minimum

6) Response characteristics of the instrumentation are consistent with the measurements to be

taken

The wind tunnels have generally these test-section dimensions: width of 9 to 12 feet, height of 8 to

10 feet, and length of 75 to 100 feet. Wind speeds than can be generated in these tunnels can range from 25

to 100 miles per hour (Taranath 2005). Typically there are two types of test models being used to conduct

studies: the first one is the rigid High Frequency Base Balance Model (HFBBM), and the second being the

Aeroelastic Model (AM). The models can be used independently or combined to obtain design loads for a

structure. The HFBBM measures overall fluctuating loads for the determination of dynamic responses. The

aeroelastic model is employed for direct measurements of loads, deflections and accelerations when the

lateral motions of a building are considered to have a large influence on the loading produced by the wind.

Numerous techniques are used in these wind tunnels to generate the turbulence and atmospheric boundary

layer by using tools such as spires and grids. In long wind tunnel sections, turbulent boundary layer is

generated by providing roughness elements in the approaching flow. Although these techniques are

considered to be appropriate, there are concerns in whether the wind turbulence is appropriately modeled.

Typically the scaling used to account for all these variables varies in the order of 1:400 to 1:600 for urban

environments (Taranath 2005).

Reinhold (1977) investigated several problems associated with the measurements of fluctuating

wind loads on tall structures using a number of building orientation and configurations. The author

generated atmospheric winds over urban areas in a short-test section tunnel and presented the results in the

dissertation document. In the study, a simple square prism was used because of its simplicity. Reinhold’s

study extended measurements of these random loads at multiple levels that improved with the respect of the

13
placement of pressure transducers along the model structure. It must be noted that most complete wind

tunnel tests and reports which have been conducted in the past that are of aid to design engineers are often

considered proprietary and are almost never published (Reinhold 1977).

2.6. Overview of Tall Buildings

Tall towers and building have fascinated mankind from the beginning of civilization, their

construction being initially for defense and subsequently for ecclesiastical purposes. The growth in modern

tall building construction which began in the late part of the 19th century has been largely for commercial

and residential purposes. Tall commercial buildings are primarily a response to the demand by business

activities to be as close to each other, and to the city center, as possible, thereby putting intense pressure on

the available land space. Tall commercial buildings are frequently developed in city centers as prestige

symbols for corporate organizations. Furthermore, the business and tourist community has fuelled a need

for more frequently city center hotel accommodations such as high rises.

The rapid growth of the urban population and the consequent pressure of limited space have

influenced city residential developments. The high cost of land , the desire to avoid continuous urban

sprawl, and the need to preserve important agricultural production have all contributed to drive residential

buildings vertically. Also, some topographical conditions make tall buildings the only feasible solution for

housing needs such as the ones encountered in Hong Kong and Rio de Janeiro.

2.6.1. Factors Affecting Growth, Height, and Structural Form

The feasibility and desirability of high-rise structures have always depended on the available

materials, the level of construction technology, and the state of development of the services necessary for

the use of buildings. Significant advances have occurred with the advent of a new material, construction

facility, or form of service.

The socioeconomic problems that followed industrialization in the nineteenth century coupled

with an increasing demand for space in U.S. cities created a strong stimulus to tall building construction.

The growth could not have been sustained without two major technical innovations that occurred in that

century:

1) The development of higher strength and structurally more efficient materials, wrought iron

and thereafter steel.

14
2) The introduction of the elevator

For the first time, this made upper stories as attractive to rent as the lower ones and made the taller

buildings financially viable. The new materials allowed the development of lightweight skeletal structures

permitting buildings of greater height and with larger interior open spaces and windows. Improved design

methods and construction techniques allowed the maximum height of steel frame structures to reach a

height of 60 stories with the construction of the Woolworth Building in 1913. This golden age of

skyscraper construction culminated in 1931 with its crowning glory, the Empire State Building, whose 102-

story brace steel frame reached a height of 1250 ft. (381 m).

(a) (b)
Figure 5: (a) Woolworth Building, NY (b) Empire State Building, NY
(Credit: The Pictorial News Co. and Virginia University)

Reinforced concrete construction began around the turn of the 20th Century but it has only been

used for the construction of multistory buildings approximately after the end of World War I. The inherent

advantages of the composite material which could be readily formed to simultaneously satisfy both

aesthetic and load-carrying requirements were not fully appreciated by then due to limited design

knowledge of the material. The economic depression of the 1930s put a hold to the great skyscraper era and

it was only after some years passed after World War II that the construction of high-rise buildings

recommenced with new structural and architectural solutions.

Different structural systems have gradually evolved for residential and office buildings, reflecting

their differing functional requirements. In modern office buildings, the need to satisfy the differing

15
requirements of individual clients for floor space arrangements led to the provision of large column-free

open areas to accommodate flexibility in planning. Other architectural features of commercial buildings

that have influenced structural form are the large entrances and open lobby areas at ground level, the

multistory atriums, and the high-level restaurants and viewing galleries that may require more extensive

elevator systems and associated sky lobbies. A residential building’s basic functional requirement is the

provision of self-contained individual dwelling units, separated by substantial partitions that provide

adequate acoustic and fire insulation. Because partitions are repeated from floor to floor, modern designs

have utilized them in a structural capacity leading to the shear wall, cross wall, or infilled-frame forms of

construction.

The trends to exposed structure and architectural cutouts, and the provision of setbacks at upper

levels to meet daylight requirements have also been features of modern architecture. The requirement to

provide adequately stiff and strong structures led to the development of a new generation of structural

framing such as braced frames, framed-tube and hull-core structures, wall-frame systems, and outrigger-

braced structures (Stafford Smith and Coull 1991). The latest generation of buildings with their more varied

and irregular external architectural treatment has led to a hybrid double and sometimes triple combinations

of the structural forms for modern buildings.

2.6.2. Criteria for the Definition of Tall Buildings

The Council on Tall Buildings and Urban Habitat (CTBUH) has developed a guideline to define

what constitutes a “tall building” that exhibits some element of height in one of these three categories:

1) Height relative to context: it is not just about height but about the context in which it exists. A

20-story building may not be considered a tall building in a high-rise city such as New York

or Hong Kong, but in a provincial city or suburb this may be distinctly taller than the urban

norm.

2) Proportion: a tall building is not just about height but also proportion. There are a number of

buildings which are slender enough to give appearance of a tall building against the

background of a low urban environment. On the other hand, there are numerous large

footprint which are quite tall but their floor area rules them out as being classified as a tall

building.

16
3) Tall building technologies: If a building contains technologies which may be attributed as

being a product of tallness such as high speed elevators and wind bracing, then this building

can be classified as a tall building.

The number of floors if a poor indicator of defining a tall building due to the changing nature of

floor to floor height between different buildings uses. A building of perhaps 15 or more stories, or over 50

m (165 ft.) in height could be used as a threshold for considering it a “tall building.” However, the CTBUH

defines a “supertall” building over 300 meters (984 ft.) in height, and a “megatall” as a building over 600

meters (1,968 ft.) in which it recognizes building height in three categories. As of August 2014 there exists

82 supertall and 2 megatall buildings that have been completed and are presently occupied (CTBUH 2014).

Figure 6: Comparison of tall, supertall and megatall building height criteria


(Credit: CTBUH)

17
Figure 7: World's ten tallest buildings according to 'height to architectural top' as of November 2014
(Credit: CTBUH)

2.7. Wind Loading on Tall Buildings

Wind is a phenomenon of great complexity arising from the interaction of wind with structures.

Simple quasi-static treatment of wind loading, which is universally applied to design of typical low to

medium-rise structures, can be very conservative for design of very tall buildings. Important factors in wind

design of tall buildings are dynamic response (effects of resonance, acceleration, damping, structural

stiffness), interference from other structures, wind directionality, and cross wind response. Mendis et al.

(2007) considered a number of key factors associated with the design of tall buildings to the effects of wind

loading. The general design requirements for structural strength and serviceability assume particular

importance in the case of tall building design. Significant dynamic response can result from both buffeting

and cross-wind loading excitation mechanisms. Serviceability with respect to occupier perception of lateral

vibration response can govern the design. The authors have suggested a specific purpose-designed damping

system in order to reduce these vibrations to acceptable levels. Dynamic response levels also play an

important role in the detailed design of façade systems. State of the art boundary layer wind tunnel testing,

for determining global and local force coefficients and the effects of wind directionality, topographical

features and nearby structures on structural response are identified to be quite useful to tall building design.

The emerging use of CFD codes, particularly at the concept design stage, is also noted as assuming

increasing importance in the design of tall buildings. The authors have suggested that the design criteria for

lateral wind loads shall consider stability against overturning, uplift and or sliding of the structure as a

whole, strength of the structural components of the building, and serviceability so as to restrict the

interstorey and overall deflections within acceptable limits.

18
CHAPTER 3

ELEMENTARY PRINCIPLES OF WIND ENGINEERING

3.1. Forces Acting in the Free Atmosphere

Wind is air movement relative to the earth, driven by several different forces, especially pressure

differences in the atmosphere, which themselves are produced by differential solar heating of different parts

of the earth’s surface, and forces generated by the rotation of the earth. The differences in solar radiation

between the poles and the equator produce temperature and pressure differences. These, together with the

effects of the earth’s rotation, set up large-scale circulation systems in the atmosphere, with both horizontal

and vertical orientations (Holmes 2007).

Severe tropical cyclones such as hurricanes and typhoons generate extremely strong winds over

some parts of the tropical oceans and coastal regions both north and south of the equator. For these types of

severe storms, the wind is highly turbulent or gusty. The turbulence is produced by eddies or vortices

within the air flow, which are generated by frictional interaction at ground level or shearing action between

air moving in opposite directions with respect to altitude (Holmes 2007).

The two most important forces acting in the free atmosphere, i.e. above the frictional effects of the earth’s

boundary layer, are the pressure gradient and the Coriolis force.

3.1.1. Pressure Gradient and the Coriolis Force

Based on the principles of fluid mechanics, at a point in a fluid in which there is a pressure

gradient, ∂p/∂x, in a given direction, x, in a Cartesian coordinate system, there is a resulting force per unit

mass given by:

 1  p
Pressure gradient per unit mass =    (3.1)
 a  x

where ρa is the density of air and p is the atmospheric pressure. This force acts form a high-pressure region

to a low-pressure region.

19
The Coriolis force is an apparent force due to the rotation of the earth. It acts to the right of the

direction of motion in the northern hemisphere and to the left of the velocity vector in the case of the

southern hemisphere; at the equator, the Coriolis force is zero. Within about 5° of the equatorial region, the

Coriolis is negligible in magnitude thus explaining why tropical cyclones do not form within this region

(Holmes 2007).

Figure 8: Coriolis force results in wind being deflected owing to the rotation of the Earth
Credit: The Atmosphere, 8th Edition, Lutgens and Tarbuck (2001)

3.2. Geostrophic Wind, Gradient Wind, and Frictional Effects

Steady flow under equal and opposite values of the pressure gradient and the Coriolis force is

known as balanced geostrophic flow. Equating the pressure gradient force per unit mass and the Coriolis

force per unit mass given by f∙U, we obtain:

 1  p
U    (3.2)
  a f  x

This is the equation for the geostrophic wind speed, which is directly proportional to the

magnitude of the pressure gradient. In the northern hemisphere the high pressure is to the right of an

observer facing the flow direction; in the southern hemisphere, the high pressure is on the left. This results

in a counter-clockwise rotation of winds around a low pressure in the northern hemisphere, and a clockwise

rotation in the southern half. Rotation about a low-pressure center is known as a cyclone to meteorologists,

which usually produces strong winds (Holmes 2007).

Near the center of tropical cyclones, the centrifugal force acting on the air particles cannot be

neglected due to the significant curvature of the isobars. For flows around a low-pressure center, i.e.

cyclone, the centrifugal force acts in the same direction as the Coriolis force and opposite to the pressure

20
gradient force. The equation of motion for a unit mass of air moving at a constant velocity, U, for a cyclone

is:

U2 1 p
 f U 0 (3.3)
r a r

The quadratic equation represents the gradient wind speed formulation. This equation has two

theoretical solutions, but if the pressure gradient is zero then U must also be zero so that the solution, for a

cyclone, becomes:

f r f 2r 2 r p
U    (3.4)
2 4  a r

where f is the Coriolis parameter ( = 2Ω sin λ), λ is the latitude, and r the radius from the storm center. The

term under the square root is always positive, therefore the wind speed in a cyclone is only limited by the

pressure gradient, i.e. cyclones are associated with strong winds.

As we approach the earth’s ground surface, frictional forces gradually play a larger role through the shear

between layers of air in the atmospheric boundary layer. The frictional force acts in opposite direction to

that of the flow (Holmes 2007).

3.3. Atmospheric Boundary Layer

As the earth’s surface is approached, the frictional forces play an important role in the balance of

forces on the moving air. For larger storms such as extra-tropical depressions, this zone extends up to 500

to 1,000 m height. The region of frictional influence is called the atmospheric boundary layer and it is

similar in many respects to the turbulent boundary layer on a flat plate at high wind speeds.

3.3.1. Mean Wind Profiles

The ABL characteristics vary with the conditions of the terrain considered. Design standards such

as the ASCE 7-10 take into consideration and factor the effect of surrounding structures and terrain in the

load conditions of the structure of interest. There are four different types of terrain that will affect shape

and thickness of the boundary layer as seen in Figure 9.

21
Marine Flat/Open Suburban Urban

Figure 9: Wind profile in different boundary layers

3.3.1.1. The Logarithmic Law

In strong wind conditions, the most accurate mathematical expression is the logarithmic law for

wind profiles. The logarithmic law was originally derived for the turbulent boundary layer on a flat plate by

Prandtl; however it has been found to be valid in an unmodified form in strong wind conditions in the

atmospheric boundary layer near the surface (Holmes 2002). The logarithmic law describes the variation

with height and surface roughness of strong mean speeds with averaging times of 10-min to 1-hr in straight

winds. After mathematical derivation, its expression may be written as:

z
ln
z0
V ( z )  Vzref  (3.5)
zref
ln
z0

where V ( z ) and Vzref is the mean wind speeds at elevation z and zref respectively, and zref is a reference

elevation, and z0 is an empirical measure of the surface roughness called roughness length. Another

measure of terrain roughness is the surface drag coefficient, Csd, which is a non-dimensional surface shear

stress defined as:

2
 
 k 
Csd    (3.6)
 ln  33  
  z0  

where k is known as von Karman’s constant, and has been found experimentally to have a value of ≈ 0.4.

This value is typically based by the mean wind speed measured at a height zref of 10 m (32.8 ft.). Table 1

(Simiu 2011) gives the appropriate value of roughness length and surface drag coefficient, for various types

of terrain types adapted from ASCE 7-10.

22
Table 1: Roughness lengths and Surface drag coefficients per ASCE 7-10
(Credit: Simiu 2011)

On a final note, the logarithmic law has some mathematical constraints which may cause

problems: first, the logarithms of negative numbers to not exist, and secondly, it is less easy to integrate. To

avoid some of these problems, wind engineers have often preferred the use of the power law.

3.3.1.2. The Power Law

The power law has no theoretical basis but is easily integrated over height – a convenient property

when wishing to determine bending moments at the base of a tall structure. To relate the mean wind speed

at any height z with that at 10 m (32.8 ft.) the power law can be represented as:

1/ 
 z 
V ( z )  Vzref   (3.7)
z
 ref 

The exponent α will change with terrain roughness, with height range, and upon averaging time.

The power law applied to 3 sec. gust wind profiles has the same form as Equation 3.7. Table 2 shows

power law exponents and gradient heights specified by the ASCE 7-93 Standard for sustained wind speeds

(including fastest-mile wind speeds) and for ASCE 7-10 Standard for 3-sec gusts (Simiu 2011).

Table 2: Power law exponents and gradient height per ASCE 7-10 Standard
(Credit: Simiu 2011)

23
Figure 10 shows a matching of the two laws for a height range of 100 m using the previous

equations where the average height in the range over which matching is required (i.e. 50 m). It is clear that

the two are relatively close, and the power law can be adequately used for engineering purposes (Holmes

2007).

100

80
Height, z (m)

60

40

20

0
0.6 0.8 1.0 1.2 1.4
V/Vz ref

Logarithmic law Power Law

Figure 10: Comparison of the logarithmic and power law for mean velocity profile for z0 = 0.02 m and α = 0.128

3.3.2. Turbulence

The general level of turbulence or ‘gustiness’ in the wind speed, such as that it can be measured by

its standard deviation, or root-mean-square. First we subtract out the steady or mean component, then

quantify the resulting deviations. Since both positive and negative deviations can occur, we first square the

deviations before averaging them, and finally take the square root to give a quantity with the units of wind

speed. Mathematically, the formula for standard deviation can be expressed as:
1

 1 
T 2 2

 u    U (t )  U  dt  (3.8)
 T 0 

where U(t) is the total velocity component in the direction of the mean wind equal to U  u(t ) , where u(t)

is the ‘longitudinal turbulence component (i.e. in the mean wind direction). Other components of

turbulence in the lateral horizontal direction is denoted by v(t) and the vertical direction by w(t) are

quantified by their standard deviations  v and  w respectively.

24
3.3.2.1. Turbulence Intensities

The ratio of the standard deviation of each fluctuating component to the mean value is known as

the turbulence intensity (TI) of that component. Figure 11 illustrates a typical time dependent measurement

of wind velocities measured in the atmospheric boundary layer. The velocity in the figure is decomposed

into a steady mean value with a fluctuating component (Versteeg and Malalasekera 2007).

u(t)

Figure 11: Typical point velocity measurements in turbulent flow

Therefore the equation for turbulence intensity can be represented as:

u v w
TIu  (longitudinal ) ; TI v  (lateral ) ; TI w  (vertical ) (3.9)
U U U

It has been measured that near the ground by winds produced by large depression systems the

standard deviation of the longitudinal wind is approximately equal to 2.5 u* , where u* is the friction

velocity. Alternatively, the turbulence intensity TIu can be expressed as the following equation:

2u* 1
TI u  
 0.4  ln  z0   z 
u* z
(3.10)
ln  
 z0 

For a rural terrain with a roughness length ( z0 ) of 0.04 m the various longitudinal turbulence

intensities for increasing height above ground is demonstrated in Figure 12, thus it can be concluded that

the turbulence intensity above ground decreases as the height increases.

25
300

250

200

Height, z (m)
150

100

50

0
10% 15% 20% 25% 30% 35% 40%
Turbulence Intensity (Longitudonal)

Rural Terrain Suburban Terrain

Figure 12: Longitudinal turbulence intensity for rural terrain (z0 = 0.04 m) and suburban terrain (z0 = 0.15 m)

The lateral and vertical turbulence components are generally lower than the corresponding

longitudinal value. For well-developed boundary layer winds, simple relationships between standard

deviation and the friction velocity u* have been developed. For well-developed boundary layer winds,

simple relationships between standard deviation and the friction velocity u* have been developed.

Therefore, the standard deviation for the lateral velocity, v, is approximately equal to 2.20 u* and the

vertical, w, component is given by approximately 1.35 u* . The equivalent turbulent intensity equations for

TIv and TIw can be shown to be:

0.88 0.55
TI v  ; TI w 
 z   z  (3.11)
ln   ln  
z
 0  z0 

The ASCE 7-10 code uses a single formulation for the calculation of turbulence intensity based on

the variable present in gust-effect factor equation. The intensity of turbulence at height z is defined as:

1/6 1/6
 33   10 
Iz  c   or I z  c   (3.12)
 z   z

where z is the equivalent height of the structure at 0.6h and not less than zmin. The coefficient c is given by

the Table 26.9-1: Terrain Exposure Constants found in the ASCE 7-10 publication. The equation on the

left represents imperial units, and on the right the international system.

26
3.3.2.2. Integral Turbulent Length Scale

The velocity fluctuations in a flow passing a point may be considered to be caused by an overall

eddy consisting of a superposition of component eddies transported by the mean wind. Each component

eddy is viewed as causing, at that point, a periodic fluctuation with frequency f. Integral turbulence length

scales are measures of the spatial extents of the overall turbulent eddy.

The integral turbulence scale Lu is an indicator of the extent to which an overall eddy is associated

with the longitudinal wind speed fluctuation u will engulf a structure in the along-wind direction, and will

thus affect at the same time both its windward and leeward sides. If Lu is large in relation to the along-wind

dimension of the structure, the gust will engulf both sides. The scales Lv and Lw are measures of the lateral

and vertical spatial extent of the fluctuating longitudinal component u of the wind speed. Mathematically

the integral turbulent length Lu is defined as follows:


1
L  R
x
u u1u2 ( x)dx (3.13)
u2 0

in which u1=u(x1,y1,z1,t), u2=u(x1+x,y1,z1,t), and the denominator is the variance of the longitudinal velocity

fluctuations, a statistic that for a given elevation z is the same throughout the flow. The integrand is the

cross covariance of the signals u1 and u2. The integral length is a measure of the average eddy size (Simiu

2011). Measurements show that Lu increases with height above ground and as the terrain roughness

decreases.

The ASCE 7-10 code uses a single formulation for the calculation of the integral length scale of

turbulence at the equivalent height. The length scale at height z is defined as:

 
 33   10 
Lz     or Lz     (3.14)
 z   z 

The coefficients of the above equations are given by the Table 26.9-1: Terrain Exposure Constants found

in the ASCE 7-10 publication. The equation on the left represents U.S. customary units, and on the right the

SI system.

The Architectural Institute of Japan defines the turbulence length scale to be height dependent but

defined independently of the terrain categories (AIJ 2006).

27
  z 0.5
100 30m  z  z g
Lz    30  (3.15)
 100 z  30m

where z is the height above ground measured in meters, and zg determines the exposure factor as defined in

AIJ’s code Table A6.3. A graphical representation of this formulation can be seen in Figure 13. Based on

AIJ, the turbulent length

300

250
Height, z (m)

200

150

100

50

0
0 50 100 150 200 250 300 350
Turbulent Length Scale, Lu (m)

Figure 13: Variation of turbulent length scale as height increases per AIJ code

28
CHAPTER 4

THEORY OF COMPUTATIONAL FLUID DYNAMICS

4.1. Overview

In the seventeenth century, the foundations of experimental fluid dynamics were laid in France

and England. The eighteenth and nineteenth centuries saw the gradual development of theoretical fluid

dynamics primarily in Europe. As a result, throughout most of the twentieth century the study and practice

of fluid dynamics involved the use of pure theory on the one hand and pure experiment in the other hand. If

a person were learning fluid dynamics in the 1960s, the individual would have been operating in the “two-

world approach” of theory and experiment. As seen in Figure 14, computational fluid dynamics (CFD) is

today an equal partner with pure theory and pure experiment in the analysis and solution of fluid dynamics

problems. CFD provides this new third approach which nicely and synergistically complements the other

two approaches of pure theory and pure experiment; however, it will never replace either of these

approaches (Anderson 1995). There will always be need for theory and experiment. The advancement of

CFD rests upon a proper balance of all three approaches, with computational fluid dynamics helping

understand and interpret the results of pure theory and experiment, and vice versa.

Comp. Fluid
Pure Exp.
Dynamics

Pure Theory

Figure 14: Three dimensions of fluid dynamics

The development of CFD started in the 1970s in which the type of computers and algorithms that

existed at that time limited most practical solutions to basically two-dimensional flows. However the real

world in which fluids exists is three-dimensional. It was only until the 1990s that computer machines

29
increased in storage and speed capacity which has allowed CFD to operate in a three-dimensional

environment. However a great deal of human and computer resources are still frequently needed to

successfully carry out numerically intensive solutions for applications like flow over a complete speed

racing car. These solutions have become more and more prevalent within industry and government

facilities to an extent that some three-dimensional flow solutions have attained standard guidelines as a tool

during the design process of machinery (Anderson 1995). CFD is playing a strong role as a design tool in

which it has become a powerful influence in the way fluid dynamicists and aerodynamicists are engineering

products. There are various types of applications that CFD can be employed as a design tool in several

industrial applications such as:

1. Automobile and Engine

2. Manufacturing

3. Civil Engineering

4. Environmental Engineering

5. Naval Architecture

4.2. Computational Fluid Dynamics

The physical aspects of any fluid flow are based upon three fundamental principles in which they

can be expressed in terms of mathematical equations with a general form of either integral or partial

differential equations.

1. Mass is conserved (Continuity Equation)

2. Newton’s second law (Momentum Equation, F = ma)

3. Energy is conserved (First law of thermodynamics)

Anderson (1995) defines CFD as “the art of replacing the integrals or the partial derivatives in the

above equations with discretized algebraic forms, which in turn are solved to obtain numbers for the flow

field values at discrete points in space and/or time.” The end product of CFD is a collection of numbers.

The tool that has allowed the practical growth of this topic is the high-speed digital computer because it

processes the manipulation of thousands and millions of numbers.

30
4.3. Governing Equations of Fluid Dynamics

All of CFD, in one form or another is based on the fundamental governing equations of fluid

dynamics—the continuity, momentum, and energy equations. These equations speak physics (Anderson

1995). They are mathematical statements of three fundamental physical principles upon which all of fluid

dynamics is based. It is important for the reader to feel comfortable with these equations before continuing

further with applications of CFD to a specific problem. The governing equations can be obtained in various

different forms such as conservation and nonconservation form.

4.3.1. Infinitesimal Fluid Element

Consider the general flow field as represented in the streamlines in Figure 15. Imagine an

infinitesimally small fluid element in the flow with a differential volume dV. The fluid element is

infinitesimal in the same sense of differential calculus however, it is large enough to contain a huge number

of molecules so that it can be viewed as a continuous medium. The element may be moving along a

streamline with a velocity vector V equal to the flow velocity at each point. The fundamental physical

principles are applied to just the infinitesimally small fluid element itself. This application leads directly to

the fundamental equations in partial differential equation (PDE) form. The particular partial differential

equations obtained directly from the moving fluid element are called nonconservation forms of the

equations (Anderson 1995).

(a) Infinitesimal fluid element fixed in space with the (b) Infinitesimal fluid element moving along a
fluid moving through it streamline with the velocity V equal to the local flow at
each point

Figure 15: Infinitesimal fluid element model of flow

4.3.2. The Continuity Equation

In this section we will treat the following case for the physical principle: mass is conserved.

Consider the flow model shown in Figure 15 (a). This fluid element is fixed in space and has the fluid

moving through it. We will adopt a Cartesian coordinate system where the velocity and density are

31
functions of space (x, y, z) and time (t). There is mass flow passing through this fixed element. Consider the

left and right faces of the element which are perpendicular to the x axis as seen in Figure 16.

Figure 16: Model of infinitesimally small element fixed in space including mass flux diagram

The mass flow through the left face with area dy dz is, (ρu) dy dz. Since the velocity and density

are functions of spatial location the values of the mass flux across the right face will be different than that

of the left; therefore, the mass flow can be expressed as  u  ( u) / x x dy dz across the right face. In

similar fashion the expression can be represented for the faces perpendicular to the y and z axes. Note that

u, v, and w are positive by convention in the positive x, y, and z directions, respectively. Hence the net mass

flow out of the element is given by

  (  u )  (  v)  (  w) 
Net mass flow     dx dy dz (4.1)
 x y z 

The total mass of fluid in the infinitesimally small element is ρ (dx dy dz); hence the time rate of

increase of mass inside the element is given by


Time rate of mass increase  (dx dy dz ) (4.2)
t

32
The physical principle that mass is conserved can be expressed in words as follows: the net mass

flow out of the element must equal the time rate of decrease of mass inside the element. In equation terms

this means

   (  u )  (  v)  (  w) 
   0 or
t  x y z 
(4.3)

   ( V)  0
t

where in Cartesian coordinates, the vector operator nabla, , is defined as

  
 i j k (4.4)
x y z

Equation 4.3 is a partial differential equation form of the continuity condition. This represents an

unsteady, three-dimensional mass conservation at a point in a compressible fluid. The infinitesimally small

aspect of the element is why the equation is directly obtained in this form. The fact that the element was

fixed in space leads to the specific form given by the equation 4.3 which is called the conservation form.

Consider the flow model shown in Figure 15 (b); an infinitesimally small fluid element moving

with the flow. This fluid element has fixed mass, but in general its shape and volume will change as it

moves downstream. Denote the fixed mass and variable volume of this moving fluid element by m and δV,

respectively. Since the mass in conserved we can state that the time rate of change of the mass of the fluid

element is zero as the element moves along with the flow. Therefore the equation for this condition

becomes

D
  V  0 (4.5)
Dt

where D/Dt is the substantial derivative, which is physically the time rate of change following a moving

fluid element and it can be generally written as

D 
  ( V) (4.6)
Dt t

 / t is called the local derivative, which is physically the time rate of change at a fixed point;   V is

called the convective derivative, which is physically the time rate of change due to the movement of the

fluid element from one location to another in the flow field where the flow properties are spatially different.

33
Equation 4.6 states physically that the density of the fluid element is changing as the element

sweeps past a point in the flow because at that point the flow-field density itself may be fluctuating with

time, e.g. local derivative, and because the fluid element is simply on its way to another point in the flow-

field where density may be different, e.g. convective derivative.

Equation 4.5 is a partial differential equation form of the continuity equation which was derived

on the basis of an infinitesimally small fluid element moving with the flow. The fact that the element is

moving with the flow leads to the specific differential equation form given by the aforementioned equation

which is called the nonconservation form.

4.3.3. The Momentum Equation

The momentum equation is based on Newton’s second law of motion pertaining to the behavior of

objects for which all the existing forces acting on the body are unbalanced. The physical principle of the

law states that the net force acting on a body is dependent upon two variables—the net force acting on the

body and the mass of the object. Newton’s second law can also be expressed as the rate of change of

momentum of a fluid particle equals the sum of the forces acting on the particle (Anderson 1995).

Figure 17: Model used for the derivation of the x-component of momentum equation
(Credit: Anderson 1995)

The element experiences two types of forces based on Newton’s principle. The first being body

forces, which act directly on the volumetric mass of the fluid element. Typically these forces act at distance

and examples of these are: electric, magnetic, and gravitational forces. The second type is surface forces,

which act directly on the surface of the fluid element. They are due to two sources: (a) the pressure

34
distribution acting on the surface which is imposed by the outside fluid surrounding the fluid element, and

(b) the shear and normal stress distributions acting on the surface which is imposed by the outside fluid

pushing on the surface by means of friction (Anderson 1995). The shear and normal stresses in a fluid are

related to the time rate of change of the deformation of the fluid element.

The scalar equations presented below are called the Navier-Stokes equations which were derived

by Frenchman M. Navier and Englishman G. Stokes who independently obtained the equations in the first

half of the nineteenth century. These equations are partial differential equations obtained directly from an

application of the physical principle of the infinitesimal fluid element that is moving with the flow which

take the nonconservation form.

Du   xx  yx  zx
       fx (4.7a)
Dt x x y z

Dv   xy  yy  zy
       fy (4.7b)
Dt y x y z

Dw   xz  yz  zz
       fz (4.7c)
Dt z x y z

4.3.4. The Energy Equation

This is the third physical principle as outlined before which is based on the first law of

thermodynamics. We present the flow model of an infinitesimally small fluid element moving with the

flow. When applied to the flow model of a fluid element moving with the flow, the first law states

Rate of change of energy = Net flux of heat + Rate of work done on element
(4.8)
inside fluid element into element due to body and surface forces
(A) = (B) + (C)

By evaluating term C, it can be shown that the rate of doing work by a force exerted on a moving

body is equal to the product of the force and the component of velocity in the direction of the force. In total,

the net rate of work done on the moving fluid is the sum of the surface force contribution in the x, y, and z

directions, as well as body force contribution. By evaluating term B, the net flux of heat into the element is

due to (1) volumetric heating such as absorption or emission of radiation and (2) heat transfer across the

surface due to temperature gradients. Term A denotes the time rate of change of energy in the fluid

element. The total energy of a given molecule is the sum of its electronic, vibrational, rotational, and

translational energies; the total energy of the each atom is the sum of its translational and electronic energy
35
(Anderson 1995). The internal energy of the gas systems simply is the energy of each molecule or atom

summed over all the molecules or atoms in the system which is the physical significance of the internal

energy that appears in the first law of thermodynamics. Term A concerns the energy of a moving fluid

element which has two contributions to its energy:

1. The internal energy due to random molecular motion, e (per unit mass).

2. The kinetic energy due to translational motion of the fluid element which simply put is V2/2.

Hence, the moving fluid has both internal and kinetic energy and the sum of these two is the total

energy. The final form of the energy equation can be presented in nonconservation terms which is based on

the total energy e + V2/2.

D  V2    T    T    T   (up )  (vp )  ( wp )
 e     q   k  k  k   
Dt  2  x  x  y  y  z  z  x y z
 (u xx )  (u yx )  (u zx )  (v xy )  (v yy )  (v zy )  ( w xz )  ( w yz )
        (4.9)
x y z x y z x y
 ( w zz )
  f  V
z

All of theoretical and computational fluid dynamics is based upon the equations presented in the

aforementioned section. It is essential that the reader is familiar with them and that their physical

significance is understood. Moreover, there are numerous references in which the reader can rely on for

further clarification of the different equations in which computational fluid dynamics is based upon.

4.4. Turbulence and Modeling for CFD

Almost all fluid flow which we encounter in daily life is turbulent. Typical examples are flow

around cars, airplanes, and buildings. The boundary layers and the wakes around and after these bluff

bodies are turbulent in nature. In turbulent flows we divide the velocities in one time-averaged part, v ,

which is independent of time (i.e. when the mean flow is steady), and one fluctuating part v' so that

v  v v' .

36
Figure 18: von Kármán vortices representing turbulent flow characteristics forming in clouds flowing past a
volcano
(Credit: Gary Davies)

Turbulent flows have no specific definition, however it has a number of characteristics that we use

to describe its properties, such as:

1. Irregularity: turbulent flow is irregular and chaotic; they may seem random, but they are

indeed governed by the Navier-Stokes equations. The flow consists of a spectrum of different

scales, also referred to as eddy sizes. Turbulent eddies exists in a certain region of space for a

certain time and that it undergoes into dissipation. These eddies have velocity and length

scales associated with it.

2. Diffusivity: in these types of flows, diffusivity increases. The turbulence increases the

exchange of momentum in boundary layers, and reduces separation at bluff bodies.

3. Large Reynolds Numbers: turbulent flows occur at high Reynolds number. For example, the

transition to turbulent flow in pipes occurs at Re ≈ 2,300 and in boundary layers at Re ≈

500,000.

4. Three-dimensional: turbulent flow is always unsteady and exists in three dimensional space.

However, we can treat the flow as two-dimensional when the equations are time averaged.

5. Dissipation: turbulent flow is dissipative, which means that its kinetic energy in the small

eddies are transformed into thermal energy. The largest eddies extract their energy from the

mean flow which transfers the energy to smaller eddies in a cascade process.

6. Continuum: even though we have small turbulent scales in the flow they are much larger

than the molecular scale, therefore we can treat the flow as a continuum (Davidson 2015).

37
Figure 19: Representation of cascade process with a spectrum eddies
(Credit: Davidson 2015)

4.4.1. Turbulent Scales

The largest scales of the order of the flow geometry with length scale 0 and velocity scale v0 .

These scales extract kinetic energy from the mean flow which has a time scale comparable to the large

scales. Part of the kinetic energy of the large scales is lost to slightly smaller scales with which the large

scales interact. The kinetic energy is in this way transferred from the largest scale to the smallest scale

through a method referred to as cascade process. The kinetic energy dissipation is denoted by ε which is

energy per unit time and unit mass (ε = m2/s3). The dissipation is proportional to the kinematic velocity, ν,

times the fluctuating velocity gradient to the power of two. The friction forces exist at all scales but they

are largest at the small eddies. The smallest scales where dissipation occurs are called the Kolmogorov

scales whose velocity scale is denoted by v , length scale by  and time scale  . It is assumed that these

scales are determined by viscosity, ν, and dissipation, ε (Davidson 2015).

4.4.2. Energy Spectrum

As mentioned previously, the turbulence fluctuations are composed of a wide range of eddie

scales. The turbulent scales are distributed over a range of scales which extends from the largest scales

which interact with the mean flow to the smallest scales where dissipation occurs. In wavenumber space the

energy of eddies can be expressed as

E ( ) d  (4.10)

38
where the above equation expresses the contribution from the scales with wavenumber between κ + dκ to

the turbulent kinetic energy κ. The dimension of wavenumber is one divided by length (e.g. m-1); thus we

can think of the wavenumber as inversely proportional to the eddy’s diameter. The kinetic energy is

obtained by integrating over the whole wavenumber space. In other words, we compute the kinetic energy

by first sorting all eddies by size, then computing the energy of each eddy size, and finally sum the kinetic

energy for all eddy sizes. The kinetic energy is the sum of the kinetic energy of the three fluctuating

components of velocity

k
2
1 '2 2


2 1
u1  u 2'  u3'  ui' ui'
2
(4.11)

The spectrum of E is shown in Figure 20. Region I, II and III correspond to the following:

I. Large eddies which carry most of the energy in which the eddies’ velocity and length scale

are v0 and  0 , respectively.

II. Dissipation range where eddies are small and isotropic. It is in this region is where true

dissipation occurs.

III. Inertial subrange requires that the Reynolds number is high and the flow is fully turbulent.

The turbulence in this region is also isotropic.

As a final note, the ratio of velocity, length, and time scales of the energy-containing eddies to the

Kolmogorov eddies increases with increasing Reynolds number (Davidson 2015)


Energy of eddies

Turbulent kinetic energy

Figure 20: Spectrum for turbulent kinetic energy, κ


(Credit: Davidson 2015)

39
4.4.3. Transition from Laminar to Turbulent Flow

Osborne Reynolds, a British scientist, was the first to distinguish the difference between the

classification of laminar and turbulent flow in a pipe. Reynolds observed that for small flow rates a drop of

dye remained a well-defined line as it flowed along the fluid (i.e. laminar flow). For intermediate flow rates

the drop fluctuated in time and space and bursts of irregular behavior appear along the fluid flow (i.e.

transition flow). For large flow rates, the drop of dye almost immediately became blurred and spread across

with randomness (i.e. turbulent flow) (Munson et al. 2006).

In general terms, flows are classified as laminar or turbulent. The important parameters that help

us distinguish these two types of flows is largely based on the Reynolds number and their critical values

depend on the specific flow situation involved. For flow in a pipe the Reynolds number must be less than

2,300 for laminar flow, and greater than 4,000 for turbulent flow. The region in between is known as

transition flow. For flow along a flat plate, the transition between laminar to turbulent occurs at a Reynolds

number approximately to 500,000.

Figure 21: Transition of flow in a pipe from laminar to turbulent


(Credit: Munson et al. 2006)

4.4.4. Turbulent Flow Calculations

Turbulence causes the appearance in the flow of eddies with a wide range of length and time

scales that interact dynamically complex way. A substantial amount of research effort has been dedicated to

the development of numerical methods to capture the important effects due to turbulence. These methods

can be grouped in three main categories (Versteeg and Malalasekera 2007).

1. Reynolds-averaged Navier-Stokes (RANS) equations: the numerical attention is focused on

the “mean flow” and the effects of turbulence on the mean flow properties. The Navier-Stokes

40
equations are time averaged prior to the application of numerical techniques. Extra terms

appear in the time-averaged flow equations to represent the interaction between turbulent

fluctuations. These terms are modelled with two “classical” models – κ-ε and κ-ω

formulation. The computer requirements required to arrive at reasonable accurate flow

computations are modest.

2. Large Eddy Simulation (LES): this is considered to be an intermediate form of turbulence

calculations which tracks the behavior of the larger eddies. This method involves space

filtering of the unsteady Navier-Stokes equations prior to the start of computations, which

allows larger eddies to be accounted for and excludes smaller eddies. The effects on the

resolved flow due to the smallest unresolved eddies are included by means of sub-grid scale

modelling. The computer demands are greatly increased due to the solution of unsteady flow

equations.

3. Direct Numerical Simulation (DNS): this type of simulation compute the mean flow and

“all” turbulent velocity fluctuations. The unsteady Navier-Stokes equations are solved on

spatial grids that are very fine to resolve the Kolmogorov length scale at which energy

dissipation takes place and with very small time steps to capture the period of the fastest

fluctuations. These calculations are extremely demanding on computer resources.

DNS
LES

RANS

Figure 22: Increase of computational cost per type of turbulence modeling


(Credit: psc.edu)

4.4.5. RANS Equations and Classical Models

For most engineering studies, it is unnecessary to resolve the details of turbulent fluctuations. CFD

engineers are almost always satisfied with the information about the time-averaged (mean) properties of the

flow of a fluid, such as pressures, velocities, stresses, etc. The majority of turbulent computations are

carried out based on the procedure presented on the Reynolds-averaged Navier-Stokes equations. In order
41
to be able to compute turbulent flows with RANS equations, the general purpose CFD code must have wide

applicability, be simple, and economical to simulate. The most common RANS turbulence models are

classified on the basis of the number of additional transport equations that need to be solved along with the

RANS flow equations. Listed below, there are four types of extra transport equations models (Versteeg and

Malalasekera 2007):

1. Zero equations (e.g. Mixing length model)

2. One equation (e.g. Spalart-Allmaras model)

3. Two equations (e.g. κ-ε model, κ-ω model)

4. Seven equations (e.g. Reynolds stress model)

In Reynolds averaging, the solution variables in the instantaneous Navier-stokes equations are decomposed

into the mean (i.e. ensemble-averaged or time-averaged) and fluctuating components:

    ' (4.12)

where  denotes the mean and ϕ’ the fluctuating velocity components in the x, y, and z directions (i = 1, 2,

3, respectively). Likewise the same formulation as Equation 4.12 applies for scalar quantities such as

pressure, energy, or species concentration (ANSYS 2014).

Substituting expressions of this form for the flow variables into the instantaneous continuity and

momentum equations and taking a time average yields the ensemble-averaged momentum equations which

can be written in Cartesian tensor form as

 
   ui   0 (4.13)
t xi

  ui u j 2 ul   

t

x j
p
  ui     ui u j    

xi x j
     ij  
  x j xi 3 xl   x j

  ui' u 'j  (4.14)

Equations 4.13 and 4.14 are called the Reynolds-averaged Navier-Stokes (RANS) equations. They

have the same general form as the instantaneous Navier-stokes equations, with the velocities and other

variables now representing the ensemble or time-averaged values. Additional terms now appear that

represent the effects of turbulence. These Reynolds stresses,   u i' u 'j , must be modeled in order to close

Equation 4.14.

42
4.4.5.1. κ-ε Model

The κ-ε Model focuses on the mechanisms that affect the turbulent kinetic energy. It must be noted

that the symbol κ represents the turbulent kinetic energy, and ε represent the turbulent dissipation rate. The

two-equation turbulence model allows the determination of both, a turbulent length and time scale by

solving two separate transport equations. The κ-ε model falls within a class of models that has become the

workhorse for practical engineering problems since it was proposed by Launder and Spalding (1974).

(Versteeg and Malalasekera 2007).

The standard κ-ε model is based on the transport equations for the turbulence kinetic energy (κ)

and its dissipation rate (ε). The equation for κ was derived from exactness, however the model transport

equation for ε was obtained by physical reasoning and is of little resemblance to its exact counterpart. The

transport equations for the turbulence kinetic energy, κ, and its rate of dissipation, ε, are represented as

follows:

       
      ui      t    G  Gb    YM  S (4.14)
t xi x j    x j 

         2
      ui      t    C1  G  C G   C   S (4.15)
t xi x j    x j    3 b 2

In these equations, Gκ, represents the generation of turbulence kinetic energy due to the mean

velocity gradients calculated by referring to the model of turbulent production in the κ-ε models. Gb is the

generation of turbulence kinetic energy due to buoyancy. YM represents the contribution of fluctuating

dilatation in compressible turbulence to the overall dissipation rate. C1ε, C2ε, Cμ, are constants equal to 1.44,

1.92, 0.09, respectively; σκ, σε are the Prandtl numbers for κ and ε with values of 1.0 and 1.3, respectively.

These values have been determined from fundamental experiments for turbulent flow which were

frequently encountered in shear flows like boundary layers. They have been found to be work fairly well

for a wide range of wall-bounded and free shear flows (ANSYS 2014).

4.4.5.2. κ-ω Model

The most prominent alternative to the previous model is the κ-ω model which was proposed by

Wilcox (1988). This model uses the turbulence frequency ω = ε/κ as the second variable; ω can also be

referred to as the specific dissipation rate.

43
     
      ui      G  Y  S (4.16)
t xi x j  x j 

     
     ui      G  Y  S (4.17)
t xi x j  x j 

In the above equations, Gκ, represents the generation of turbulence kinetic energy due to mean

velocity gradients. Gω represents the generation of ω. Γκ and Γω represent the effective diffusivity of κ and

ω, respectively. Yk represents the dissipation of κ and Yω the dissipation of ω due to turbulence. These

terms have specific models associated with them that are left to the reader to reference to. Lastly, Sκ and Sω

are user defined source terms.

The κ-ω attracted attention because integration to the wall does not require wall damping

functions in low Reynolds number applications. The value of κ at the wall is set to zero and the frequency

ω tends to infinity at the wall, however we normally specify a very large value. Practical experience with

this model has proved that the results do not depend entirely on the precise details of this condition

(Versteeg and Malalasekera 2007).

4.4.5.3. SST κ-ω Model

The shear-stress transport (SST) κ-ω model was developed by Menter (1994) to effectively

combine the robust and accurate formulation in the near-wall region with the freestream independent of the

κ-ε model in the far field. The SST κ-ω model is equivalent to the standard model, but it includes the

following refinements:

1. The standard κ-ω and the transformed κ-ε model are multiplied by a blending function and

then they are summed together. The function is designed to have a value of one in the near-

wall region, and zero away from the surface.

2. The SST model includes a damped cross-diffusion derivative term in the ω equation.

3. The definition of the turbulent viscosity is modified to account for the transport of shear

stress.

4. The constants in these models have different values.

The major way in which the shear-stress transport (SST) model differs from the standard model

are as follows: (1) gradual change from the standard κ-ω model in the inner region of the boundary layer to

44
a high Reynolds number of the κ-ε model in the outer part of the boundary layer, and (2) modified turbulent

viscosity formulation to account for the transport effects of the principal turbulent shear stress. These

features make the SST κ-ω model more accurate and reliable for a wider class of flows.

     
      ui      G  Y  S (4.18)
t xi x j  x j 

     
     u j      G  Y  D  S (4.19)
t x j x j  x j 

In the above equations, Gκ, represents the generation of turbulence kinetic energy due to mean

velocity gradients. Gω represents the generation of ω. Γκ and Γω represent the effective diffusivity of κ and

ω, respectively. Yk represents the dissipation of κ and Yω the dissipation of ω due to turbulence. Dω

represents the cross-diffusion term. These terms have specific models associated with them that are left to

the reader to reference to. Lastly, Sκ and Sω are user defined source terms (ANSYS 2014).

The field of turbulence modelling provides abundant research activities for the CFD and

engineering communities. The RANS formulations presented in this section are widely available in

commercially available computer codes. Industry experts have widely used these concepts and have

produced useful results in spite of observations relating to its limited capabilities (Versteeg and

Malalasekera 2007). In the present study the SST κ-ω is utilized for the modeling of the wind flow around

the bluff body.

4.4.6. Large Eddy Simulation

A different approach to the computation of turbulent flows accepts that the larger eddies need to

be computed for each problem with a time-dependent simulation. The universal behavior of the smaller

eddies should be easier to capture with a compact model. Instead of time-averaging, LES uses spatial

filtering to separate the larger from smaller eddies. This method uses the selection of a filtering function

and a certain cutoff length scale with the aim of resolving in an unsteady flow computation all those eddies

that have a larger length scale than the cutoff dimension. During the spatial filtering, information related to

the smaller eddies below the cutoff length is destroyed. This interaction effects between larger and smaller

eddies give rise to sub-grid-scale (SGS) stresses. This is the key concept of the large eddy simulation (LES)

approach to the numerical treatment and solution of turbulence in fluids (Versteeg and Malalasekera 2007).

45
Energy of eddies Turbulent kinetic energy

Figure 23: Spectrum of velocity


(Credit: Davidson 2015)

The rationale behind LES can be interpreted as: (1) momentum, mass, energy and other scalars are

transported by larger eddies, (2) larger eddies are problem-dependent; they are governed by the geometry

and boundary conditions of the flow, (3) smaller eddies do not depend on geometry, they are isotropic. By

resolving the large eddies the must use finer meshes than those used in RANS models. LES has to be

typically run for long duration of time to obtain stable statistics for the flow being modeled (ANSYS 2014).

Therefore the computational time is longer and high-performance computing (HPC) is mandatory for LES

simulations. The disadvantage of LES lies in the high resolution requirement for wall boundary layers since

near the wall even larger eddies become small and require a Reynolds number dependent resolution. LES is

typically limited to Reynolds numbers in the range of 104 – 105 and smaller computational domains.

A substantial portion of  ij is attributed to convective momentum transport due to interactions

between the unresolved eddies which are referred to as subgrid-scale stresses. The subgrid-scale turbulence

models usually employ the Boussinesq hypothesis (Hinze 1975). We compute the subgrid-scale turbulent

stresses from

1
 ij   kk  ij  2  t S ij (4.20)
3

where  ij is the subgrid-scale turbulent viscosity. S ij is the rate-of-strain tensor for the resolved scale define

by

1  u i u j 
S ij     (4.21)
2  x j xi 

46
For compressible flows, we must introduce Favre (density-weighted) filtering operator. The compressible

form of the subgrid stress tensor is defined by splitting it into its isotropic and deviatoric parts

1 1
 ij   ij   kk  ij   kk  ij (4.22)
3 3

where the deviatoric part of the subgrid-scale stress tensor is modeled using the compressible form of the

Smagorinsky model (ANSYS 2014)

1  1 
 ij   kk  ij  2t  Sij  Skk  ij  (4.23)
3  3 

4.4.6.1. Smagorinsky-Lilly Model

This model was first developed by Smagorinsky (1963). In this formulation, the eddy-viscosity is

modeled as

t   L2s S (4.24)

where Ls is the mixing length for subgrid scales and S  2 S ij S ij and Ls can be computed using

Ls  min( d , Cs ) (4.25)

where κ is the von Kármán constant ≈ 0.41, d is the distance to the closest wall, Cs is the Smagorinsky

constant, and Δ is the local grid scale based on the element volume. The researcher Lilly derived a value of

Cs = 0.23 for homogenous isotropic turbulence in the inertial subrange.

It must be noted that Cs is not a “universal constant.” The fact that CS has different values is due to

the mean flow strain or shear that gave an indication that the behavior of small eddies is not as universal as

first hypothesized. A different Cs value of around 0.1 has been found to yield very good results for a wide

variety of flows and this value is used in present computer codes. Further formulation of this theory can be

found in computational fluid dynamic textbooks.

4.4.6.2. Remarks on LES

LES has been established in the 1960s, however the computational power to process this model

has not been made available until recently for industrially relevant problems. The inherent unsteady nature

of LES leads to the much higher computational demand than those needed by traditional RANS techniques.

LES is good at resolving certain time dependent features of turbulence with no additional equation since it

is inherent in its own formulation. LES gives a deeper insight into the mean flow and statistics of the
47
resolved fluctuations. The incorporation of LES codes in commercially available softwares has only been

accessible recently to the engineering community. The pace of development and research based on this

model will increase as computing resources become more powerful. Engineers will gain awareness of the

advantages of the LES method to turbulence modelling as more meaningful data is published (Versteeg and

Malalasekera 2007).

4.4.7. Typical Meshing Information

The partial differential equations that govern fluid flow are not usually responsive to analytical

solutions, except for very simple cases. Therefore, in order to analyze fluid flows, flow domains are split

into smaller subdomains. The governing equations are then discretized and solved inside each of these

subdomains. Typically, one of three methods is used to solve the approximate version of the system of

equations: finite volumes, finite elements, or finite differences. Care must be taken to ensure proper

continuity of solution across the common interfaces between two subdomains, so that the approximate

solutions inside various portions can be put together to give a complete picture of fluid flow in the entire

domain. The subdomains are often called nodes, elements, or cells. The collection of all elements or cells is

called a mesh or grid. There mainly are three classifications of meshes that depend on the connectivity of

the nodes:

1. Structured Meshes: these are characterized by regular connectivity that can be expressed as

arrays. These elements are usually quadrilaterals in 2D and hexahedra in 3D.

2. Unstructured Meshes: there are characterized by irregular connectivity and cannot be

expressed as an array in a computer. This allows for the use of any variation of geometric

mesh shapes that can fit the model geometry.

3. Hybrid meshes: there are meshes that contain partially structured and unstructured meshes

around the geometry being investigated.

Meshes can be classified as either 2D or 3D. Common elements for 2-dimensional meshes are

rectangles and triangles. In 3-dimensional geometry, common elements are hexahedral, tetrahedral, square

prisms, and triangular prisms as seen Figure 24.

48
Figure 24: Typical 2D and 3D Mesh Shapes
(Credit: Bakker 2006)

Figure 25 illustrates a typical flow chart in the geometry and meshing generation process that must

be followed for CFD modeling. The user must first create the geometry using a computer aided design

package and the must transfer it to a meshing software. The user then must create the appropriate mesh

based upon the geometry requirements. Once all items are satisfied, the user must export the mesh to a

CFD solver. Explain mesh flow chart.

Figure 25: Typical geometry meshing process


(Credit: ANSYS 2014)

49
4.4.8. Near Wall Treatment

It is critical to capture boundary layer near wall properly. In order to do that, the mesh should be

generated in such a manner that it captures the boundary layer properly. For turbulent flows, calculation of

the y+ value of the first interior grid point helps achieve the capture of the boundary layer. This

dimensionless distance is defined as

u* y
y  (4.26)

where u* is the friction velocity given by


u*  (4.27)

The wall shear stress is usually determined after the simulation has been completed and usually

the engineers must assume a value and then check it with the simulation results. The role of the wall

function becomes unrealistic when the flow velocity increases and a more refined method must be used

such as the three-layer form of the log-law of the wall

 y for  5

u    3.05  5ln y  for 5 < y +  30 (4.28)
5.5  1/   lny  for y  30
+

Figure 26: Log-Law of the Wall


(Credit: McDonough 2007)

50
4.4.9. Discretization Techniques and Algorithms

In essence, discretization is the process by which a closed-form mathematical expression, which

are viewed as having an infinite continuum of values through a domain, can be approximated by analogous

expressions which prescribe values at only a finite number of discrete points or volumes in computational

domain. Numerical solutions can give answers at only discrete points in the domain also referred to as grid

points or nodes as see. Figure 27 represents a typical discretization that may be encountered in a fluid

domain. Uniform spacing such as Δx = Δy is typically used in CFD to allow for the simplification of

numerical computations, saves computer storage, and usually gives results with greater accuracy. This type

of grid system is also referred to as structure grid which reflects a consistent geometrical regularity in the

computational domain.

Figure 27: Discretization of grid points


(Credit: Anderson 1995)

The finite volume method (FVM) is a discretization method which is well suited for the numerical

simulation of various types of conservation laws. It is widely used in engineering fields such as fluid

mechanics. The finite volume method may be used in arbitrary geometries that can lead to robust schemes.

An additional feature of the FVM is that the numerical flux is conserved between neighboring cells which

in fact is a very important feature when modeling fluid mechanics problems (Eymard et al. 2003).

It must be noted that in most CFD applications, a first order accuracy is not sufficient to represent

the numerical techniques. The first order accuracy consists of a finite representation of a partial derivative

and the remaining terms in that equation give the truncation error. The truncation error tells the formulation

what is being neglected in the approximation. A more common technique is the use of the second-order

51
accuracy equation that enhances the numerical approximation as shown in Equation 4.29. The information

used comes from both sides of the grid located at a specific coordinate falling between two adjacent points.

The truncation error in the formulation involves second-order accuracy (Anderson 1995).

 u  ui 1, j  ui 1, j
    O ( x ) 2 (4.29)
 x  i , j 2 x

There are two common types of errors in all CFD techniques that are difficult to be avoided. These

are: (a) discretization error, (b) round-off error. As mentioned in the previous paragraph, discretization

errors are caused by truncation error for the difference equation plus any additional error introduced by the

numerical treatment of boundary conditions. Round-off error is introduced due to the repetitive number of

calculations which the computer is constantly rounding the numbers to some particular significant figure

such as the limits of double precision (Anderson 1995).

There are four different numerical schemes that can be adopted as algorithms for the solution of

CFD problems. These are:

1) Semi-Implicit Method for Pressure-Linked Equations (SIMPLE)

2) SIMPLE-Revised (Patankar 1980) or SIMPLE-Consistent (Van Doormal and Raithby 1984)

3) Pressure Implicit with Splitting of Operators (PISO) by Issa (1986)

4) Coupled

The SIMPLE algorithm gives a method of calculating pressure and velocities for a given flow.

This method is iterative, and when other scalars are coupled to the momentum equations the calculation

need to be performed sequentially. The SIMPLER algorithm developed by Patankar (1980) is an improved

version of the previous. In this algorithm the discretized continuity equation is used to derive a discretized

equation for pressure instead of the pressure correction equations as in SIMPLE. Therefore the intermediate

pressure field is obtained directly without the use of correction factors. However, velocities are still

obtained using corrections. The PISO algorithm is a pressure-velocity computational procedure developed

for non-iterative analysis of unsteady compressible flows. However, it has also been successfully adapted

for the iterative solution of steady state problems. PISO involves one predictor step and two corrector steps

and may be seen as an extension of SIMPLE (Versteeg and Malalasekera 2007). The corrector steps

involve neighbor and skewness correction.

52
Coupled algorithms offer advantages over the segregated approach. The coupled method obtains a

robust and efficient single phase implementation for steady-state flows, with superior performance when

compared to the previous schemes. The pressure-based algorithm offers an alternative over the other

systems. For transient flows, using the coupled algorithm is a proper choice when the quality of the mesh is

poor, or if large time steps are used (ANSYS 2014). It is left to the reader to further explore the formulation

for each of these numerical algorithms as they are readily available in many literature references.

53
CHAPTER 5

THEORY OF STRUCTURAL DYNAMICS

5.1. Basic Concepts of Vibration

Humans became interested in vibration when the first musical instruments were discovered. Since

then, we have applied a critical investigation in the study of the “vibration phenomena.” Galileo discovered

the relationship between the length of a pendulum and its frequency. Galileo also found the resonance of

two bodies which were connected by an energy transfer medium and tuned to the same natural frequency.

Many mathematicians such as Bernoulli, D’Alembert, etc. contributed to the development of vibration

theory. Sauveur coined the term fundamental for the lowest frequency and harmonics for the others. (Rao

1990). Whenever the natural frequency of vibration of a structure coincides with the frequency of an

external excitation load, it produces the phenomenon known as resonance. This phenomenon leads to

excessive deflections and can cause catastrophic failures (Rao 1990). In its broadest term, vibrations or

oscillations can be defined as any motion that repeats itself after an interval of time.

The theory of vibration focuses on the study of the motions of bodies and the forces associated

with them. A vibrating system includes means of having three types of energy. The first being potential

energy, the second being kinetic energy, and third by means of gradual loss of energy. In general, the

vibration of a system involves the transfer of its potential to kinetic energy, and vice-versa. The damping of

the system causes energy dissipation in each cycle of vibration. In order for a system to vibrate infinitely,

the system would have no means of damping or energy dissipation due to friction.

A structural dynamics problem differs greatly from its static equivalent. One difference between

dynamic and static analysis is that dynamic loading is time-varying in nature and the structural response to

that excitement also is time-dependent. In addition, the occurrence of inertia forces is another important

factor that distinguishes dynamic from static analysis. Therefore, there is no single solution for this

problem. The engineer must investigate a solution over a given period of time to fully evaluate the

54
structural response and to design accordingly. Dynamic analysis is more computationally intensive than

static analysis (Tedesco et al. 1999)

5.2. Degrees of Freedom and Classification of Vibration

The minimum number of independent coordinates required to determine completely the position

of all parts of a system an any given instant defines the degree of freedom (DOF) of the system. For

example, a simple pendulum swinging in a plane represents a single degree of freedom system. Two

pendulums connected to each other is an example of two degrees of freedom; three pendulums connected to

each other is an example of three degrees of freedom and so on. A large number of practical systems can be

described with a specific and deterministic number of degrees of freedom. However some systems such as

a continuous elastic member can be said to have an infinite number of degrees of freedom. Systems with a

finite number of degrees of freedom are called lumped parameter/mass systems, and those with infinite

number of degrees of freedom are termed as continuous/distributed systems (Rao 1990). A continuous

model will exhibit the mathematical formulation of a system of partial differential equations. For lumped

systems, the mathematical formulation gives a set of ordinary differential equations (one for each DOF).

Most structures exhibit at least several degrees of freedom. The number of DOF that a structure

contains is equivalent to the number of independent spatial coordinates that can describe its geometric

configuration and motion (Tedesco et al. 1999). For each degree of freedom, there will be an equivalent

number of equations of motion and natural frequencies associated with it. A single node of a structural

element represented in a three dimensional space has a up to six degrees of freedom; three in translation

and three in rotation. The number of degrees of freedom can be reduced based on the boundary condition of

the node, such as fixed, pinned, or roller support.

Buildings can be idealized as lumped parameter systems with multiples degrees of freedom (i.e.

based on the number of stories it contains) as seen in Figure 28. For each DOF exhibited by a structure,

there corresponds a natural frequency of vibration. For each natural frequency the structure will vibrate to a

particular mode of vibration.

55
Figure 28: Idealized multi-story building with five degrees of freedom
(Credit: Rao 1990)

The mass or inertia element is assumed to be a rigid body that can gain or lose kinetic energy

whenever the velocity of the body fluctuates. Based on Newton’s second law of motion, multiplying the

mass by the acceleration of the body gives an equivalent force acting on the body. Work is equal to the

force multiplied by the displacement in the direction of the force acting on the system. The work done is

stored by the system in the form of kinetic energy of the mass. By assuming that the mass of the frame is

negligible compared to the mass of the floor system, the building can be treated as multi-degree of freedom

system as previously stated. The masses at the various floors represent the mass elements, and the vertical

columns are comparable to spring elements which contribute stiffness.

MDOF systems require matrix formulations to clarify the problem and to create a systematic

manner to conduct the response calculations. Matrices provide a convenient format for organizing the

calculations required to analyze MDOF systems. Matrix notations create the option for the engineer to use a

procedure to be able to program digital computers to arrive at solutions. There are three main physical

properties important to every MDOF systems: the mass, stiffness/flexibility, and damping matrices. If the

mode of superposition method is employed for the dynamic analysis, the system must be solved as an

eigenproblem (Tedesco et al. 1999).

5.3. Equation of Motion and Natural Frequency

The mathematical expression that defines the dynamic equilibrium of a system is referred to as the

equation of motion of a structure. An important conclusion that can be derived from the solution of this

56
equation is the displacement-time history of a structure that is subject to a specific time-varying dynamic

load. The basic components of a vibrating system include: mass, stiffness/flexibility, damping, and forcing.

Damping is the energy loss of the system and forcing is the source of loading.

Figure 29: Model of a simple SDOF and free-body diagram of system

The mechanical model for a simple single degree of freedom (SDOF) systems and its

corresponding free-body diagram is depicted in Figure 29. It consists of a rigid body of of mass m, which is

limited to translate in only one direction, in this case the x-axis direction. A spring of stiffness k fixed at the

west end and attached to the body provides an elastic resistance to displacement. Energy dissipating system

is represented by a damper having its own particular coefficient c. The time-varying load being applied to

the systems is represented by F(t).

The motion of the mass is resisted by the force FS that develops in the spring and it is defined by

FS  kx (5.1)

A conservative system (i.e. no dissipation of energy) will continue to oscillate indefinitely even

after the applied load is removed. However, structures in the real world experience energy dissipation due

to the effects of friction or damping that prevents the structure from behaving as a conservative system.

Damping is a very complex phenomenon for which a number of analytical models have been developed to

describe its behavior. The most common model is the linear viscous dashpot model (Craig 1981). The

damping force is proportional to the velocity of the mass and is given by

FD  cx (5.2)

where c is the viscous damping coefficient having units of pounds-seconds per inch (lbf∙sec/in) or newton-

seconds per meter (N∙sec/m).

The inertia force is equivalent to the product of the mass and the acceleration. The negative sign in

the below equation represents the inertial force acting against the acceleration of the mass.

57
FI  mx (5.3)

By using D’Alembert’s principle of dynamic equilibrium, the equations of motion of a SDOF and MDOF

systems can be established. For a given body with a constant mass, the rate of change of momentum is

equal to the product of the mass and its acceleration. By referring to the free-body diagram given in Figure

29, the expression for dynamic equilibrium can be expressed as

mx  cx  kx  F (t ) (5.4)

If we divide equation by m we can obtain the natural circular frequency, ω0, of a system with the

units of radians per second. To obtain the natural frequency, f, of the system we further divide ω0 by 2π as

seen in the below equation. The units for natural frequency is Hertz (Hz) or cycles per second. The natural

frequency of a structure plays an important role in vibration analysis (Tedesco et al. 1999).

0 k/m
f   (5.5)
2 2

5.4. Property of Matrices for Vibrating Systems

In this section the author briefly discusses the flexibility/stiffness and mass matrices since these

matrices are of relevant significance in structural dynamic theory. The document also briefly discusses the

eigenproblem in vibration analysis.

5.4.1. Flexibility and Stiffness Matrix

The displacement and forces that act upon a structure can be related to one another by using

flexibility or stiffness properties. First, the discussion focuses on the concept of flexibility but then move on

the stiffness since the unique relationship between these methods will be evident.

Consider the structural beam illustrated in Figure 30. If we apply a load F2 at the second point of

the beam, deflections at point 1, 2 and 3 will be x1, x2, x3 respectively. The same concept can be done for

points 1 and 3 and the following expression can be arrived at in matrix form

 x1   a11 a12 a13   F1 


    
 x2    a21 a22 a23   F2  (5.6)
 x  a a32 a33   F3 
 3   31

The matrix (a) is known as the flexibility matrix that contains the influence coefficients that relate

the displacements, x, relative to the applied loads, P. For elastic systems, the flexibility matrix is

symmetric.
58
Figure 30: Flexibility coefficients for a beam

It can be easily demonstrated that the inverse of the flexibility matrix results in the stiffness matrix

[k] since the stiffness matrix relate the forces {F} to the deformations {x}. The force vector {F} contains

forces and moments, and the deformation vector {x} contains translations and rotations.

k   a
1

(5.7)
F    k  x
The stiffness can be expressed in verbal terms as the force or moment required to produce a unit

displacement or rotation at a given point in a structure if displacement or rotation is stopped from occurring

at all other nodes. The nodes of interest correspond to the defined DOF of the system.

5.4.2. Mass Matrix

The mass matrix will generally be diagonal in nature and it includes the inertia effect associated

with the translational DOF. The inertial influence for rational DOF may also be included. The lumped mass

matrix for an element of a three-dimensional frame possess coefficients which are equal to one half of the

total inertia of the beam segment. A mass matrix for a uniform beam can be obtained in greater complexity

by including the matrix of axial effects, the matrix for torsional effects, and the matrix for flexural effects

(Paz 2004).

 F1   m1 0 0   x1 
     
 F    0  0   x 
 (5.8)
 F   0 0 m    
 n  n   xn 

mL mL
m1  and m2 
2 2

Figure 31: Lumped mass coefficients for structural elements with distributed mass

59
5.5. Eigenproblem Formulation, Natural Vibration Frequencies and Modes

Analysis of a vibrating system with MDOF is required to go the solution of the standard

eigenproblem which can be expressed as

 DA    A (5.9)

where [D] is a square symmetric matrix, {A} is the solution vector, and λ a scalar quantity equivalent to ω2.

{A} must not be a null vector to allow for Equation 5.9 to be satisfied. The solution is given by the

eigenvalue λr and corresponding eigenvector {A}r where r represents the rth solution. So each solution

consists of an eigenpair with solutions given arrived to be ((λ1,{A}1), …, ((λn,{A}n)) where λ is arranged

from smallest to largest numbers.

In the analysis of structures for vibration, [D] is knowns as to the dynamic matrix and is usually

formulated by the stiffness matrix. The n eigenvalues of the dynamic matrix represent the natural

frequencies of the structure, and the corresponding eigenvectors represent the modes of vibration as normal

or principal. The eigenvector can be multiplied by any chosen constant cr. Equation 5.8 is satisfied if we

multiply both sides and a more general form can be arrived at by introducing the modal vector {Φ}r

corresponding to λr equivalent to cr{A}r. The relative values of the elements in the matrix remain

unmodified and the eigenvector {A}r is normalized by the constant. Therefore by considering all λn’s, the n

number of modal vectors {Φ}r from the modal matrix that defines all the normal modes of vibration of the

system. (Tedesco et al. 1999). The modal matrix [Φ] can be written in terms of the vectors as

11 22  1n 


   2 n 
 nn  1 , 2 ,  , n    21 22
   
(5.10)
 
n1 n 2  nn 

For a system with a large quantity of degrees of freedom, obtaining the solution for the

eigenproblem must be accomplished via computerized numerical eigensolvers to extract the mode shapes

and natural frequencies of the structure. We must also note that a MDOF vibrating systems possess an

important property referred to as orthogonality. The orthogonality of the eigenvectors system is with

respect to the stiffness and mass matrix. For modes with different frequencies, where ωs ≠ ωr, it follows

60
s  m r  0
T

(5.11)
s  k r  0
T

These orthogonality conditions of normal modes are the basic theory that leads to the modal

analysis of a vibrating system; the formulation of the aforementioned system is left for the reader to

explore.

61
CHAPTER 6

CFD ANALYSIS OF WIND FLOW USING RANS & LES FORMULATIONS

6.1. 2D & 3D Setup of Model in CFD

The computational domain defines the region where the flow field is computed numerically used

the assigned models. It should be large enough to accommodate all relevant flow features that will have

potential effects in altering the characteristics of the flow field (Franke et al. 2007). The building geometry

for the computational domain was based on wind tunnel scaling for the study of prototype structures. In the

case of our model, the prototype was a square structure dimensioned as 40 meters (B) by 40 meters (W)

(130 feet × 130 feet) in plan form, and had height of 300 meters (H) (≈ 1,000 feet) in order for it to qualify

as a super-tall building. The structure had a height to base ratio of 7.5 which made it fall in the category of

a slender structure. Typically, slender structures are greatly affected by lateral loading and extra diligence

must be taken by the engineer when designing such structures that are known to be sensitive to lateral

loading.

The wind tunnel scale was chosen to be 1:400 as that is typical in industry practice to use this

factor between prototype and model studies. In the case of the chosen building, this reduced its dimensions

to 0.1 × 0.1 × 0.75 meters. The constant ‘B’ throughout the study was computed to be equivalent to 0.1

meters based on the scaling ratio. The aforementioned constant was used as a common variable to describe

all geometric properties of the model.

CFD problems must be defined by initial and boundary conditions for meaningful numerical

output. It is important for the engineer to understand their vital role in the numerical algorithm. The most

common boundary conditions used are: inlet, outlet, wall, and symmetry. The distribution of the flow

variables must be specified at the inlet along the upstream direction. Outlet boundary conditions are

selected at the downstream direction and distant enough from geometrical disturbances so that the flow can

reach fully developed state. The wall boundary conditions is the most commonly used in confined fluid

flow problems in which the no-slip condition is associated with it. Symmetry is regarded as a condition

62
where normal velocities are set to zero, and the values of all other properties just outside the domain are

equated to their values at the nearest node inside the domain. Figure 32 represents a sketch of the proposed

computational domain dimensions and boundary conditions used for the CFD analysis performed in this

study. The three dimensional boundary conditions were also used as a basis for the configurations used for

the two dimensional studies.

Figure 32: 3D Boundary conditions

The plan form of the geometric properties and boundary conditions can be seen in Figure 33. The

overall computational domain was 21B × by 47B where the structure was centered at 10.5B from the inlet

and 10.5B from the south wall condition. A distance of 36B in the downstream direction was chosen to

allow for the reduction of the disturbances experienced by the flow. The north and south edge were

assigned the wall boundary conditions to simulate a wind tunnel cross section.

Figure 33: 2D Plan boundary conditions

63
The elevation form of the geometric properties and boundary conditions can be seen in Figure 34.

The overall computational domain was 18.75B × by 47B where the structure was centered at 10.5B from

the inlet. Again, a distance of 36B in the downstream direction was chosen to allow for the reduction of the

disturbances experienced by the flow. A symmetry condition was applied to the top edge of the

computational domain.

Figure 34: 2D Elevation boundary conditions

6.2. Grid Independent Study

A grid sensitivity analysis for the model was performed using different number of nodes in

increasing multiples of two. This type of analysis must be performed to reduce the influence of the number

of nodes on the computational results since the solution must be independent of the mesh resolution in the

computational domain. It is good practice to run this study before a more global analysis of the system is

completed. The building geometry for the computational domain was based on wind tunnel scaling for the

study of prototype structures. In the case of our model, the prototype was a square structure dimensioned as

40 meters (B) by 40 meters (W) (130 feet × 130 feet) in plan form, and had roof height of 300 meters (H)

(≈ 1,000 feet) in order for it to qualify as a super-tall building. The structure had a height to base ratio of

7.5 which falls in the category of a slender type structure. Typically, slender structures are greatly affected

by lateral loading and extra diligence must be taken by the engineer when designing such structures that are

known to be sensitive to lateral loading.

The wind tunnel scale was chosen to be 1:400 as that is typical in industry practice to use this

factor between prototype and model. In the case of the chosen building, this reduced its dimensions to 0.1 ×

0.1 × 0.75 meters. The constant ‘B’ throughout the chapter was assigned to be equivalent to 0.1 meters

which was used as a common variable for all geometric properties of the model. Figure 35 displays typical

64
meshing arrangement used for the simulation. As evident, the mesh was concentrated in the immediate

surroundings of the structure to capture the flow separation and consequent fluid behavior. The mesh was

created using the software ICEM CFD in which it allows for the user to have greater control of the meshing

criteria to be utilized for any type of geometries, whether they are regular or irregular shape. ICEM CFD

allows the user to have control of the mesh by utilizing the concept of blocking.

For the simulation involving 157,000 nodes, the model used a very fine grid cell in the near-wall

region and ensured a non-dimensional wall distance y+ to be less than 5 which was well within the inner

sublayer. A bias ratio 1.05 was applied to the mesh between successive grid points to ensure solution

convergence for the numerical methods applied. The blocking of the geometric model in ICEM CFD

allowed the author to place 100 grid points in the upstream direction, and 360 grid points in the

downstream directions. The transverse directions were also assigned 100 grid points. The bias were applied

to the nodes approaching the walls of the square structure. They were also applied in the direction of the

north and south wall of the wind tunnel. The building structure was assigned 100 grid points along the four

faces of the building where were also biased to have a greater concentration along the sharp edges of the

building. The combination of the aforementioned arrangements generated a structure quadrilateral mesh

with near perfect orthogonality. This allowed the CFD model to perform with one less possible source of

error in the computation of results.

Figure 35: 157 K Nodes quad-mesh for 2D plan view of structure

For the grid analysis of 78,000 and 44,000 the same concept as outlined above was applied. The

only difference was in the reduction of the grid points along the faces of the building, the upstream and

65
downstream flow, and transverse directions. This introduced an increase in aspect ratios between the

rectangular mesh elements.

A similar principle was applied for the meshing of the two dimensional elevation study of the

wind flow around the structure. The upstream was divided into 100 nodes, and the downstream was divided

into 360 nodes. Along the height of the building, 150 nodes were specified to better capture the wind

pressures acting in that direction. Directly above the building, the number of nodes was reduced to 41

points. The final meshing resulted in 90,000 nodes with rectangular shapes and it is displayed in Figure 36.

ICEM CFD evaluation of the orthogonality of the meshes proved to be near perfect (i.e. close to a value of

unity).

Figure 36: Quad-mesh for 2D elevation of structure

The grid independent study was validated using two different data outputs from the model. The

first was based on the concept of the development of fluid flow along a “pipe.” A wind tunnel can

essentially be seen as a pipe when a cross section of the tunnel is taken along the longitudinal direction.

The region where the fluid enters the pipe is known as the entrance region. The fluid in this experiment

entered the tunnel with a uniform velocity. As the fluid moved through the pipe, viscous effects caused the

air to stick to the wind tunnel wall therefore a boundary layer was formed along this edge. Figure 37

represents this behavior that occurred and was captured in the CFD model. At the inlet, the normalized

velocity magnitude was constant. However at a distance of 45B from the inlet, the normalized velocity

magnitude taken along a height of 0.2 meters from the face of the south tunnel wall, displayed the

characteristical behavior of “pipe” flows. The boundary layer formed at 45B was captured for the fully

developed flow. That was made possible due to the fine mesh resolution near the wall that was specifically

66
designed by the author. Based on the figure, the boundary layer for three different meshing followed the

same profile. Moreover, the 78 K nodes model was almost identical to the 157 K nodes model thus

indicating that the solution has reached an acceptable grid independence and the engineer can opt to go

with the lowest amount of nodes (i.e. 44 K) to continue the study.

1.2
1.0
V/Vstream (unitless)
0.8
0.6
0.4
0.2
0.0
0 0.05 0.1 0.15 0.2
Position perpendicular to South wall (m)

Inlet 44 K Nodes 78 K Nodes 157 K Nodes

Figure 37: Boundary layer development along wind tunnel section

The second data set chosen to ensure the grid convergence of the study was chosen to be the mean

pressure coefficient (Cp) acting along the windward face of the structure as presented in Figure 38.

p  p
Cp  (6.1)
0.5V2

The pressure coefficient describes the relative pressure throughout a flow field in fluid dynamics

and it is a dimensionless number. In engineering modeling of fluid dynamics problems, this number is of

paramount importance since a model can be tested at a wind tunnel scale where the Cp values can be

determined at critical locations, and these same coefficients can be used with confidence to predict the fluid

pressure at these same critical locations of the full-scale structure.

When comparing the mean Cp values at y/B = 0.5 for the cases with lesser nodes, the percent error

between the values of mean Cp decreases as the number of nodes increases. In other words, the error

between mean Cp at 157 K and 78 K is approximately 2%, and the error between mean Cp 157 K and 44 K

is almost 7%. Thus, giving a clear indication of grid convergence as the number of nodes increased. For the

purpose of this study, a 44 K node analysis was adopted as an adequate meshing option.

67
1.5
1

Coefficient (Cp)
Mean Pressure
0.5
0
-0.5
-1
-1.5
0 0.2 0.4 0.6 0.8 1
y/B

44 K Nodes 78 K Nodes 157 K Nodes

Figure 38: Variation of mean pressure coefficient along the windward face for grid analysis

6.3. Model Validation

Validation of a model being investigated by a CFD user is imperative to perform. Validation is

seen as the process of determining the degree to which a model is an accurate representation of the real

world from the perspective of the intended model uses. The best way to achieve comfort with a CFD model

is to make a direct comparison to published or experimental data. In the case of this study, numerous

physical studies have been performed for the fluid flow around square geometries.

Any object where fluid is moving around it will experience drag, which can be simply defined as a

net force in the direction of flow due to the pressure and shear forces acting on the surface of the obstacle.

Typically an object with any given shape can be defined by its drag coefficient, CD, where mathematically

it can be written as

1
FD  CD U 2 A (6.2)
2

The drag coefficient is also a function of dimensional parameters such as Reynolds number, Mach number,

Froude number, and relative roughness. A square cylinder has been found out to be a CD value of

approximately 2.10 derived from experimental studies as reported in Table 3.

68
Table 3: Experimental data and derived quantities for various cross sections
(Credit: Ahlborn et al. 2002)

The Reynolds number (Re) is considered to be the most famous non-dimensional number in fluid

mechanics. This number demonstrated the relationship of the combined variables to determine whether the

flow was in laminar or turbulent state. The ratio between the inertial forces and the friction force acting on

the volume of a fluid is expressed mathematically as:

VL VL
Re   (6.3)
 

The validation of the model was performed with the fluid behaving in a turbulent manner since its

Reynolds number was in the order of 2.0 × 105. The details of the model set up can be found in Appendix D

as it contains all the means and methods used as input parameters for the model in ANSYS Fluent. Figure

39 presents a time-dependent plot of the coefficient of drag that the square structure possesses. The

comparison of the coefficient of drag computed by the numerical CFD model was in very close agreement

to experimental data.

2.20
2.10
Coefficient of Drag

2.00
1.90
1.80
1.70
1.60
1.50
1.40
0.00 0.20 0.40 0.60 0.80 1.00
Flow time (sec)

Figure 39: Drag coefficient curve for 44K SST-κω transient model

69
The CFD model was also programmed to measure the coefficient of lift, which by definition acts

in perpendicular direction to the drag forces.

2.50
2.00
1.50

Coefficient of Lift
1.00
0.50
0.00
-0.50
-1.00
-1.50
-2.00
-2.50
0.00 0.20 0.40 0.60 0.80 1.00
Flow time (sec)

Figure 40: Lift coefficient curve for 44K SST-κω transient model

An additional validation analysis was performed based on the numerical output of the CFD

analysis based upon the stream line velocity of the fluid along the centerline of the structure as plotted in

Figure 41. It can be seen that the κ-ε predictions by Franke and Rodi (1991) show an unrealistically large

recirculation zone and therefore the discrepancy is noted. The present study utilized the SST-κω model

showed a large improvement in the performance of the formulation. The upstream velocity is predicted

accurately up to the stagnation point that the square structure experiences near its midpoint in the windward

face. The model also shows significant performance along the downstream flow. The SST-κω formulation

captures the immediate wake of the flow velocity occurring within a distance of 1B from the leeward face.

The normalized velocity follows relatively closely the slope as presented by the experimental study

performed by Durão et al. (1998) and Lyn et al. (1995). The fine resolution of meshes in the two

dimensional plane offered adequacy for the CFD model.

70
1.2
1.0
0.8
0.6

U/Uo
0.4
0.2
0.0
-0.2
-0.4
-2 -1 0 1 2 3 4 5 6 7 8
x/B

Durão et al. (1988) Lyn et al. (1995)


Franke and Rodi - Stnd. κ-ε (1991) Present - SST κ-ω - Transient 3 sec

Figure 41: Mean velocity magnitude along centerline

It can be stated with confidence that the use of the SST κ-ω was of adequate choice for the

numerical computation of the wind flow patterns acting near the structure.

6.4. Discussion of Results Based on 2D Plan Model

A multiple level analysis was performed for sections located along the height of the structure in

the wind tunnel CFD model of the slender building. The parameters used for the analysis along the height

are presented in Table 4. The chosen section corresponded to points of interest along the height of the

building. The Reynolds number was based by the characteristic length, L, equivalent to 0.1 meters (B) and

the corresponding wind speeds at the respective section cut heights levels. The wind profile was computed

using the power law with the exponent 1/α = 1/7. The velocity scale was proportioned to the following 1:4

ratio from the prototype wind profile for urban/suburban terrain located in Miami, Florida. The data

presented in Table 4 was deemed acceptable inputs based on computations and scaling assumptions for the

present study.

71
Table 4: Model input criteria for 2D plan CFD

SECTION CFD INPUT FOR ANALYSIS


Turbulence Reynolds
No. %H Height (m) Wind Speed (m/s) Turb. Length Scale (m)
Intensity (%) No.
Cut 1 3.33% 0.0250 19.6 23.9 0.250 1.31E+05
Cut 2 10.0% 0.0750 22.9 18.9 0.250 1.53E+05
Cut 3 30.0% 0.2250 26.8 15.6 0.433 1.79E+05
Cut 4 60.0% 0.4500 29.6 14.1 0.613 1.98E+05
Cut 5 91.7% 0.6875 31.4 13.2 0.758 2.10E+05

A second property was compared to previous experimental studies and CFD analysis performed

by Kawamoto (1997) in which the coefficient of pressures were measured along the four faces of a square

prism. In order for the results in this study to be comparable, the present model was matched to the

Reynolds number used by the work performed by Kawamoto (1997).

1.5
1 2
1
0.5
0/4 3
Coefficient (Cp)
Mean Pressure

0
-0.5
-1
-1.5
-2
-2.5
0 1 2 3 4
Coordinate of Square Prism (d/B)

Kawamoto - Stnd. κ-ε (1997) Kawamoto - Experimental (1997)


Present - SST κ-ω - Transient 3 sec

Figure 42: Variation of mean pressure coefficient along faces

Figure 42 is a chart that compared the present SST κ-ω model mean pressure coefficients along the

windward, leeward and suctions faces of the pyramid. The windward face was represented by the 0 to 1

normalized coordinate system, the leeward face was identified by 2 to 3. Comparing the results to the

experimental and numerical model, the present study provided good agreement with the published

literature. The negative coefficient of pressures followed the same profile although their magnitude was

relatively larger.

72
6.5. 3D LES Model

LES is a multi-scale computational approach that offers a comprehensive way of capturing

unsteady flows. The use of LES as a wind load evaluation tool has been improved significantly in recent

years through the use of numerical techniques. LES holds promise to becoming the future of computational

wind engineering for which turbulent flow is of fundamental importance. In this study the Smagorinsky-

Lilly subgrid-scale was employed.

The discretization of convective terms the central difference scheme gives adequate accuracy

when compared to upwind schemes. The bounded central difference was adapted to minimize the effect of

oscillations arising from high Reynolds number flows in the wake regions. For temporal discretization,

second order schemes are recommended by the CWE community. The PISO algorithm was used for

pressure-velocity coupling and is typically used in LES simulations. This algorithm is based on the higher

degree of the relationship approximations between corrections for velocity and pressure. Figure 43 was a

representation on the meshing criteria used for this model. The same concept as previously presented was

utilized in the creating of this computational domain. The mesh was concentrated around the slender

building structure.

Figure 43: 3D mesh with 7.7 M nodes

The simulation has been carried out by utilizing a newly released High Performance Computer

(HPC) based on Linux Operating System available at Florida Atlantic University. The computations have

73
been performed using 24 physical central processing units (CPUs). The time step chosen for the model was

2 × 10-4 seconds with 25 sub-iterations per time step. A strict convergence criteria was applied to the model

equivalent to 10-5 to ensure convergence of the solution.

In addition to the above specifications, a user defined function (UDF) based on the computer C

language was written by the author to simulate the desired wind profile. The profile was in accordance to

the assumptions mentioned previously. The code was written to accommodate the three dimensional model

and is made available in Appendix C.

Figure 44: Wind profile generated in computer domain

Figure 44 is the computer output of the velocity profile as written by the code developed by the

author. It was based on the power law and carried an exponent of 1/7 as per the prototype model conditions

and ASCE 7-10 recommendations.

The LES model was ran with the HPC facility for 17 days and at a very fine time step and sub-

iterations. No grid independent study was performed since the set-up of the analysis was based to similar

work performed by Dagnew (2012) based on a rectangular building with a ratio of width to length of 2:1.

74
Figure 45: Present LES windward contours compared to experimental data
(Credit: Simiu and Scanlan 1996)

The windward contours have been graphically post-processed from the LES model of this present

study. In Figure 45 the left image represents the contours of the study which was compared to previously

published wind tunnel measurements and the respective pressure coefficient contours acting on a slender

square slender body (Simiu and Scanlan 1996). The results have good agreement as it can be seen from the

approximate 0.9 coefficient of pressure predicted by CFD to the experimental study. The higher

magnitudes are highlighted with red colors, and the lower ones with blue colors. The pressure variation

along the face of the CFD model truly represents the wind behavior acting on the slender body.

6.6. Pressures Acting on Slender Structure

Similarity requires that the reduced frequencies and the Reynolds numbers be the same in the

laboratory and in the prototype. This should hold true regardless of the nature of the frequencies involved,

or of the densities being considered. There are three fundamental requirements concerning mass, length,

and time in which three fixed choices of scale can be made (e.g. length scale, velocity scale, and density

scale).

In principle, for similarity between prototype and wind tunnel flows to be achieved, the respective

Reynolds numbers must be the same, also known as the Reynolds number similarity. In wind tunnels, the

fluid used is air at atmospheric pressure, and Reynolds number similarity is violated. Bodies with sharp

75
edges are assumed to be “independent” of Reynolds number. It is assumed that flows around these bodies

are similar at full scale and in the model scale.

1,000
900
800
700
Height (ft)

600
500
400
300
200
100
0
0 50 100 150 200 250
Pressure (psf)

3D LES ASCE 7-10 SAP2000 2D SST KW

Figure 46: Windward estimated pressure variation

Hand calculations were carried based on ASCE 7-10 Main Wind Force Resisting System

(MWFRS) and presented on Appendix A. The profile of the pressure generated on the windward face can

be seen in Figure 46. The structural analysis software SAP2000 contained a built in wind loading

evaluation tool and the criteria outlined in Appendix A were given as an input. A parabolic pressure profile

was generated by the structural software as is typical in for these types of tall structures. Two different CFD

techniques were utilized to derive the pressures acting along the windward face of the building, the first

was using the two dimensional SST κ-ω formulation, and the second was based on the three dimensional

LES. The maximum pressure acting on the face of the building using ASCE 7-10 hand and software

method were determined to be approximately 150 lb/ft2. However, the two dimensional SST κ-ω CFD

model estimated the pressure to be approximately 220 lb/ft2, which is approximately 46% larger than the

ASCE 7-10 provisions. The three dimensional LES model presented a result of approximately 160 lb/ft2 or

7% greater than the code. The LES profile seen in the above figure represents a reasonable representation

of the true flow and pressure variation acting along the building face. The two-dimensional CFD model is

an over simplified analysis and generates results that overestimate the wind loads on the tall building

model.

76
CHAPTER 7

BASIC STRUCTURAL AND DYNAMIC ANALYSIS OF TALL BUILDING

7.1. Overview of Idealized Example Structure

The structure being considered in this study was a reinforced concrete structure consisting of 67-

stories. The overall plan dimensions of the building were 129.5 by 129.5 feet (40 by 40 meters) containing

7 bays in the horizontal x and y directions. The tower was 1,010 feet (300+ meters) high above the street

level. The façade was assumed to be a glass and aluminum curtain wall system. The tower floors consist of

reinforced concrete beams spanning from the cores to the façade with a concrete two-way slab deck. The

column spacing was at 18.5 feet center-to-center. The typical floor-to-floor height assigned was 15 feet (4.6

meters). However the first story had a floor height of 20 feet (6.1 meters) to produce a typical architectural

features of a tall building. Figure 47 illustrates the typical framing plan of the structure in different views.

The wind resistance structure consisted of the core walls cantilevering from the base of the building to its

full height. In these systems typically 85-95% of the lateral load is carried by the shear wall.

(a) Discretized FEM model 67-Story


(b) Floor plan framing
tower

Figure 47: Framing model of 67-story tower used in present study

77
(c) East elevation framing (typ.) (d) North elevation framing (typ.)
Figure 47 (cont’d): Framing model of 67-story tower used in present study

The frame was designed as moment resisting which consists of horizontal and vertical components

that are rigidly connected together in a planar grid form. This system primarily resists through the flexural

stiffness of the members. A point of contra flexure is usually located near the midpoint of the beams and

columns. The deformation of the frame is typically induced by shear-sway and partially by column

shortening. Some advantages of using this system is the flexibility it provides to architectural planning.

The size of the members used in a moment resisting frame is often controlled by stiffness

characteristics rather than strength. The stiffness of elements greatly control acceptable lateral drifts for a

tall structure. The structures response is mainly a function of column and beam stiffness. Moreover, in

moment resisting frames the size of the column members tend to decrease in size as higher levels are

reached in the building which is in proportion to the lateral shear. Advances in computer technology have

allowed for better analysis of indeterminate moment frames. Optimization techniques when used also

determine the most efficient distribution of material for a given deformation limit analysis.

Reinforced concrete frames have positive aspect in which the connections are casted

monolithically which are well suited for moment resisting systems. Advances in testing has led to

conclusions in improved concrete characteristics, additional reinforcement requirements for the

improvement of ductility, and better frame forming techniques to be used in the field.

7.2. Assigned Loading and Section Properties to Structural Members

The loading applied on the analysis of this building was based on ASCE 7-10 Minimum Design

Loads on Buildings. The building was assumed to exist in Miami, FL in an urban/suburban region. The

function of the building was assumed to be an office high-rise located near the downtown area of this

78
hurricane prone city. Further details for the assumptions used for the design of the building can be found in

Appendix A.

Table 5: Assumed loads acting on structure

Description Type Magnitude (psf)


Dead Load DL Member Size x Unit Weight
Superimposed Dead SDL 10
Live Load LL 50 (Floor) ; 100 (Lobbies) ; 80 (Corridors)
Roof Load RL 20
Wind Loads WL Generated per ASCE 7

Based on the ASCE7/ACI318 there are typically seven load combinations to be used for the

strength design method. Structures, components, and foundations are required to be designed so that their

strength is greater or equal than the effects of the factored load in the following combinations:

1. 1.4D

2. 1.2D + 1.6L + 0.5(Lr or S or R)

3. 1.2D + 1.6(Lr or S or R) + (L or 0.5W)

4. 1.2D + 1.0W + L + 0.5(Lr or S or R)

5. 1.2D + 1.0E + L + 0.2S

6. 0.9D + 1.0W

7. 0.9D + 1.0E

There are a few exceptions and conditions that must be met for each case. The information can be

readily found in published literature and codes. For the case of this structure and the application of the

loading presented in the previous table, there were 74 problem specific combinations that were analyzed for

the structure. The large quantity of combinations is generated due to the study of the wind effect in multiple

directions acting along the building. This provides the analysis a comprehensive overview of the loads

acting on the whole structure. Based on this information, the members can be designed to account for the

worst case bending, shearing, and torsional conditions of the loading being experienced by the structure.

Based on the discussion of the previous section, the concrete building was designed taking the

state of the art guidelines into consideration. The building was subdivided into five categories where

different section properties to the members of the structure where assigned. The first category was

concentrated along the base region of the building where most of the bending and shearing loading is

79
concentrated. The fifth category was located on the last seven floors of the structure where the members

tend to be smaller in dimension.

Table 6: Cross-section and properties of members along the height of the structure

Column Design Shear Wall Beam Design Slab Design Header Beam
Cat. Floor No. Elevation B x W f'c Thick. f'c B xH f'c Thick. f' c B xW f' c
(ft) (in) (psi) (in) (psi) (in) (psi) (in) (psi) (in) (psi)
1 20 50 x 50 10,000 24 10,000 24 x 34 5,000 8 5,000 24 x 48 10,000
1
15 230 50 x 50 10,000 24 10,000 24 x 34 5,000 8 5,000 24 x 48 10,000
16 245 48 x 48 10,000 24 10,000 24 x 34 5,000 8 5,000 24 x 48 10,000
2
30 455 48 x 48 10,000 24 10,000 24 x 34 5,000 8 5,000 24 x 48 10,000
31 470 44 x 44 10,000 24 10,000 24 x 34 5,000 8 5,000 24 x 48 10,000
3
45 680 44 x 44 10,000 24 10,000 24 x 34 5,000 8 5,000 24 x 48 10,000
46 695 40 x 40 8,000 24 10,000 20 x 34 5,000 8 5,000 24 x 48 10,000
4
60 905 40 x 40 8,000 24 10,000 20 x 34 5,000 8 5,000 24 x 48 10,000
61 920 36 x 36 6,000 24 10,000 20 x 34 5,000 8 5,000 24 x 48 10,000
5
67 1010 36 x 36 6,000 24 10,000 20 x 34 5,000 8 5,000 24 x 48 10,000

Table 6 illustrates that columns had the greatest cross section along the height located within the

section named Category 1. The final column design was arrived using an iterative procedure and verified

against the ACI318-08 code requirements using SAP2000 Version 16 as the structural analysis software.

Moreover, the columns had to be design using concrete with relatively high compressive strength (e.g. f’c =

10,000 psi). The cross section of the columns along the Category 5 levels was significantly reduced by

approximately 48% when comparing it to the base floor levels. The compressive strength of the concrete

was also reduced by 40%. The interior and exterior columns were assumed to be equivalent in dimensions

throughout all category levels to simplify computational design time. The reinforcement specified for this

model was ASTM A615 Grade 60 and Grade 75 steel. Grade 60 steel was used in the design of beams and

slabs. Grade 75 steel was used for the design of columns and shear walls.

Beam design along the height of the building was concluded to be almost uniform since the span

of the bays were equal in both longitudinal and transverse direction. The only difference being that

throughout Category 1 and 3, the width of the beam had to be 24 inches (610 mm) wide to accommodate

the larger shearing forces acting at those elevations imposed by the lateral loading acting on the structure.

The slab design remained constant along all levels of the building and was assumed to provide rigid

diaphragm action on the structure. The shear wall was designed to be 24 inches thick and have a

80
compressive strength of 10,000 psi along the full height of the structure. A final element, header beams,

was used to ensure proper framing, stiffness, and interaction among the shear walls in the building. They

were placed in such a manner that they framed physically to all corresponding core walls.

7.3. Idealized Structure Discretization

Beams are the most common type of structural components used in civil engineering. A beam can

be thought of as a bar-like structural element whose function is to support transverse loadings and

effectively transfer the forces to the supports. A beam is significantly larger in magnitude in the

longitudinal direction, than the two cross-section dimensions. A longitudinal plane passes through the beam

axis. A beam resists loads mainly through bending action. Moments acting on the beam produce

compressive stresses on one side, and tensile stresses on the opposite. These two surfaces are separated by a

neutral axis/surface of zero stress. The interaction of these stresses causes an internal bending moment

which is the mechanism that transports the loads to the support. Beams must also support shearing forces

acting upon them to be properly used as a structural component in a system. Columns are vertical members

that carry significant importance as component of a structure. Stacked one on top of the other, they receive

loads that must be finally transmitted to the foundation system. Columns can be considered to be a type of

beam with a different function. Columns must resist axial and flexural loads to be effective.

Beams and columns are analyzed as frame elements. A frame element is modeled as a straight bar

with an assigned cross-section which can deform in the axial and perpendicular directions. The bar is

capable of carrying axial and transverse loads, and moments. A frame element possesses the properties of

truss and bema elements which is encountered in most of real world structural elements. The formulation of

frame elements is based on the equations for beam elements developed by the finite element method. Many

commercial softwares, such as SAP2000, use the theory of frame elements for actual engineering

applications such as the analysis of tall buildings. Frame elements are applicable for the analysis of skeletal

space frames in three dimensions and were therefore used in this study. The formulation of such theory is

left to the reader to explore, and they can be found in finite element textbooks such as Logan (2012).

A plate is considered to be a two-dimensional extension of a beam in simple bending. Beams and

plates support transverse loads to their plan and through bending action. A plate is considered to be flat,

and a shell is considered to be curved, therefore a plate is a shell with an infinite radius of curvature.

81
Kirchhoff plate theory is considered to be the classical derivation of equations in this analytical concept.

For the structural model in this study, slabs were assigned the property of shell-type behavior which

accounts for in-plane membrane stiffness and out-of-plane plate bending stiffness that are provided in the

section. This was deemed adequate since it can receive horizontal loading due to wind action, and can

account for bending caused by gravity loading. The slab was modeled as a thin shell element. The shear

walls were also considered to fall within this acceptable theory, however a thick shell element was assigned

to this structural section.

In structural engineering, a diaphragm is a structural system used to transfer lateral lads to shear

walls primarily through in-plane shear stresses. The diaphragm of a structure usually has an additional

function in where it can also support gravity loading. Diaphragms can also provide lateral support to walls

and parapets. Typically, roofs and floors participate in the distribution of lateral forces to vertical elements

up to the limit of its strength. For a slab to function as a diaphragm, horizontal elements must be

interconnected to transfer shear with relatively stiff connections. Rigid diaphragm distributes the horizontal

forces to the vertical resisting elements in direct proportion the relative rigidities. This element is assumed

to not deform and therefore will cause the vertical elements deflect the same amount. A rigid diaphragm

assumes infinite in-plane stiffness of floors, and therefore reduces the stiffness matrix. The super-tall

building was analyzed using the rigid diaphragm approach for the lateral loads induced by the wind acting

along the faces of the structure.

7.4. Vibration Modes of Modeled Structure

A normal mode of a system is a pattern of motion in which all parts of the system moves in

sinusoidal pattern with the same frequency and a fixed phase relation. The frequencies of the normal modes

of a system are referred to as its natural frequencies. A building has a set of normal modes that mainly

depend on its structure stiffness, material’s mass, and boundary conditions. Every building has a number of

ways, or modes, in which it can vibrate naturally. For each mode, the structure can oscillate naturally to and

fro with a particular distorted shape known as its mode shape. The building can be made to sway at other

frequencies of vibration with different mode shapes than its fundamental (lowest) frequency when the

loading occurs at different levels. One of the main functions that affect the stiffness of a building is its

height. Therefore, taller buildings tend to be more flexible and susceptible to lower natural frequencies. The

82
results presented on Figure 48 and Figure 49 were arrived using the eigenvector modal analysis code

imbedded in SAP2000 which follows the theory presented in Chapter 5. The eigenvector analysis

determines the undamped free-vibration mode shapes and frequencies of the system. The analysis was

limited by the user to six modes of vibration.

(a) Mode 1 – f = 0.1563 Hz (b) Mode 2 – f = 0.1595 Hz (c) Mode 3 – f = 0.2949 Hz

Figure 48: First three modes of vibration of tall structure

Generally speaking, the modes of vibration can be understood in the following simplified terms.

The first mode is a translational in direction. The second mode is also translational. The third mode is a

torsional mode. The rest of the higher modes are typically a combination of the first three. For the study of

83
a building, the first five to six modes give plenty of information to the structural engineer in terms of design

criteria.

(a) Mode 4 – f = 0.5573 Hz (b) Mode 5 – f = 0.5938 Hz (c) Mode 6 – f = 0.8391 Hz

Figure 49: Fourth to sixth modes of vibration of tall structure

In order to mitigate vibrations, the engineer can make several modifications in the design of the

building to avoid possible resonance between the structure and the lateral loading. If the structure vibrates

at frequency which humans are sensitive to, the engineer can alter the structures period of vibration by

altering its mass or stiffness. This can be achieved by: (a) stiffening structural members, (b) increasing the

number of columns, and (c) reduce the decking material to a lighter weight component. Moreover, all

84
structures have some inherent damping. Further method can be employed in which the damping can be

increased thus dissipating energy when the structure moves. However, the employment of mechanical

damping devices comes at a great financial cost.

7.5. Deflections of Super-Tall Structure

There are mainly four types of analysis than can be performed in a framed systems such as of a tall

structure:

1. First Order Elastic Analysis

2. Second Order Elastic Analysis

3. First Order Plastic Analysis

4. Second Order Plastic Analysis (Advanced analysis)

The work presented in this thesis was performed using the first analysis method listed above

which is considered the most basic and fundamental analysis. In a first order analysis the internal forces,

and the displacements of the structure are evaluated in relation to the initial, undeformed shape of the

building. This method is mostly used in the analysis of current structures. The stresses, strains, and

displacements may be obtained using established methods such as force method or displacement method. In

the case of first order analysis the deformations are proportional to the applied loads.

The loading used for the analysis was described previously. The wind load generated by the

structural analysis software was based upon the ASCE 7-10 code and design criteria presented in Appendix

A. The wind loads were applied on the structure as exposure form extents of rigid diaphragms. The loads

varied in a parabolic shape along the height of the structure on the windward face. The wind loading

generated by the software implemented the four different cases for the Main Wind Force Resisting System

presented in the ASCE 7-10 code. Case 1 considered the full design wind pressure acting on the projected

area perpendicular to each principal axis of the structure. Case 2 considered 75% of the design wind

pressure acting on the structure in conjunction of a torsional moment. Case 3 considers a similar criteria to

Case 1 but the loading acts simultaneously. Case 4 considers similar criteria to Case 2 but the loading acts

simultaneously at 75% of the specified value.

85
(c) Deflection in x-direction

(a) Deflection in x-direction (b) Deflection in y-direction (d) Deflection in y-direction

Figure 50: Deflections experienced by the loading acting on super-tall structure

The analysis of the structure was determined to have a total of 5,236 number of joints in which 76

of them were properly restrained with the fixed condition. The model contained 9,913 frame elements

representing the columns and beams. The quantity of shell elements was counted to be 5,165. The software

assembled all the matrices and determined that there was 15,681 number of equilibrium equations with

386,244 number of non-zero stiffness terms. Based on this large number of members, the only feasible

method at arriving at a solution to the matrix analysis of the structure was by using computer hardware

coupled with appropriate software to perform these demanding computations.

Wind drift limits for building design are on typically in the order of H/400. These limits generally

are sufficient to minimize damage to the structure (Taranath 2005). In the case of the present study, it was

verified that the magnitudes of the displacement were within the prescribed limit as shown in Figure 50.

86
The maximum limit imposed was computed to be 30 inches. After analyzing the software outputs, it was

determined that the maximum deflection occurred in the lateral y-direction having a magnitude of

approximately 24 inches, which is within the allowable limit. As a note, there was a number of different

deflection modes not presented in this thesis due to the large number of output generated by the software

analysis. Some results included torsional deflections of the floor systems; they were separately verified to

still abide by the H/400 criteria to ensure that the model had acceptable deflections throughout all the 16

load cases and 74 combinations that were analyzed during the evaluation of the tall building structure.

87
CHAPTER 8

CONCLUSIONS & FUTURE RESEARCH

8.1. Final Remarks

The advancements of computational wind engineering is making the evaluation of wind loads

acting on structures more attractive to researchers due to the numerical evaluation that can be performed

using modern day computational technologies. In general, the use of computational fluid dynamics allows

the engineer to utilize the techniques developed by subject pioneers to study the accuracy of RANS and

LES when combined to building aerodynamics, or bluff bodies, which is distinguished by flow separation

at building corners and high Reynolds numbers. CFD provides continuous visualization of aerodynamic

data on the entire surface areas of the building instead of the discrete point measurements extracted from

experimental wind tunnel analysis. In addition, CFD can provide detailed flow visualizations for the wind

field around the building which information on flow separation and reattachment, recirculation zone, and

vortex shedding can be extracted.

It is imperative to ensure that the accuracy of the simulation is validated rigorously to identify its

positive and negative implications. It is expected that the numerical approach with further developments in

the practical use and analysis of CFD will allow engineers to overcome the existing obstacles. Therefore

this tool might be incorporated in the design process of building engineering whether at the predesign

phases or final phases. This will provide the industry with a new set of tools to use for the aerodynamic

considerations in the building design stages. Presently, most studies are still conducted using wind tunnel

experimental techniques due to the acceptable results that industry engineers have become comfortable

with for the use of loading design purposes of tall buildings and other structures.

The current state of the art of computational wind engineering has been reviewed by several

researchers. Key findings indicate that turbulence modeling, boundary conditions, high Reynolds numbers

turbulent flow, and computational cost are essential parameters when arriving at appropriate computational

predictions. The advance in computational machines and development of parallel computing has enabled

88
the use of CFD for industrial applications such as in tall building design. Generating the atmospheric

boundary layer in the computational domain is a critical step for the wind load evaluation process. Making

data of model and full scale available are very important for the validation of the CFD results. Creating a

time-dependent analysis model can create a time-history of pressure fluctuations similar to what wind

tunnel data provide engineers with. Generally speaking, mean static pressure values showed comparative

agreement with published data and industry accepted design codes.

It must be noted that CFD results are greatly dependent on the input of the mean wind parameters

such as the velocity profile and turbulence intensities in the relevant three dimensional coordinate system.

The engineer must be attentive in defining these properties accurately. This has great implications in the

convergence of the simulation and the time required for the computer machine to generate qualitative and

quantitative aerodynamic data which can be used by the design engineer.

Generally speaking, it can be fairly said that CFD simulations can be used as an alternative tool for

wind pressure loading for at least pre-design stages of a tall structure project. The results of employing this

technique should be taken with caution to ensure that the designer is still abiding by design codes. With

proper boundary conditions applied to the model, and the advancement of computational resources could

become useful for wind loading studies in the future. The present study would like to point out that the

main limitation of using LES is that the machines required to numerically compute this model are

expensive, and the time it takes to arrive at a final computational iteration is more time consuming than

present boundary wind tunnel techniques. Performing RANS simulations can significantly decrease the

computational cost and time to arrive at reasonable results.

8.2. Recommendations for Future Research

Based on the work presented in this thesis, the author would like to suggest future venues for

research to increase the current state-of-the-art knowledge in this subject:

 By performing CFD parametric studies and assimilating the data with experimental wind

tunnel models, the present theory could be enhanced to lead to new mathematical formulation

to better predict CFD data output to real-world values.

89
 A complete fluid-structure interaction between the structural body (i.e. tall building) could be

investigated by assigning the structural properties to the object and utilizing dynamic meshes

to evaluate the response of the fluid and the structure through a desirable time domain.

 Further investigation can be carried by utilizing different numerical techniques in a simplified

model to allow researchers to identify and couple the best present methods that could provide

improved accuracy in results.

 As the computational power provided by new technologies and hardware become available in

the foreseeable market, full-scale modeling at high Reynolds numbers could be researched

with more computationally expensive techniques such as DNS.

90
APPENDICES

91
Appendix A

Wind Loading Computations of Prototype Building per ASCE 7-10

92
93
94
95
96
97
98
99
100
101
Appendix B

ICEM CFD Meshing Script for 2D Plan Geometry

102
103
104
105
106
107
108
Appendix C

User Defined Function for 3D Wind Profile in C-language

109
110
Appendix D

ANSYS Fluent 2D Simulation Criteria of CFD Model

111
112
113
114
115
116
117
118
119
REFERENCES

Ahlborn, B., Seto, M. L., Noack, B. R. (2002). “On drag, Strouhal number and vortex-street structure,”
Japan Society of Fluid Mech., Vol. 30, 379-399.

Anderson, J.D. (1995). Computational Fluid Dynamics: The Basics with Applications. McGraw-Hill Inc.,
New York, N.Y.

ANSYS, Inc. (2014). Fluent – Theory Guide. Canonsburg, PA.

Bailey, A., Vincent, N.D.G. (1943). “Wind pressure on buildings, including effects of adjacent buildings.”
J. Inst. Civil Eng., 20(8), 243-275.

Bashford and Thompson (1940). Tacoma Narrows Bridge midsection collapsing into the waters of the
Tacoma Narrows, November 7, 1940. University of Washington.
http://digitalcollections.lib.washington.edu/cdm/singleitem/collection/farquharson/id/19/rec/2
(Accessed on November 2, 2014)

Baskaran, A. and Stathopoulos, T. (1989). Computational evaluation of wind effects on buildings.


Build.Environ. 24(4), 325–333.

Baskaran, A. and Stathopoulos, T. (1992). Influence of computational parameters on the evaluation of


wind effects on the building envelope. Build. Environ. 27(1), 39–49.

Blocken, B. (2014). “50 years of Computational Wind Engineering: Past, present, and future.” J. Wind Eng.
Ind. Aerodyn., 129, 69-102.

Calautit, J.K., Chaudry, H.N., Hughes, B.R., Sim, L.F. (2014 0. A validated design methodology for a
closed-loop subsonic wind tunnel. J. Wind Eng. Ind. Aerodyn. 125, 180–194.

Castro, I.P. and Graham, J.M.R. (1999). Numerical wind engineering: the way ahead? Proc. Inst. Civil Eng.
– Struct. Build. 134(3), 275–277.

Cermak, J.E. (1975). “Applications of fluid mechanics to wind engineering,” A Freeman Scholar Lecture.
J. Fluids Engng., ASME, 9-38.

Craig, R. R. (1981). Structural Dynamics, John Wiley & Sons, New York, N.Y.
CTBUH (2014). “Criteria for the Defining and Measuring of Tall Buildings.” Council on Tall Buildings
and Urban Habitat, Chicago, IL.
http://www.ctbuh.org/TallBuildings/HeightStatistics/Criteria/tabid/446/language/en-
US/Default.aspx (Accessed on January 19, 2015)

120
Dagnew, A. K. (2012). “Computational Evaluation of Wind Loads on Low- and High-Rise Buildings,”
Dissertation at Florida International University, Miami, FL.

Davenport, A.G. (1963). “The relationship of wind structure to wind loading,” Proceedings of the
Conference on Wind Effects on Buildings and Structures, National Physical Laboratory, England,
53-102.

Davenport, A.G. (2002). “Past, present and future of wind engineering.” J. Wind Eng. Ind. Aerodyn., 90,
1371-1380.

Davidson, L. (2015). Fluid Mechanics, Turbulent Flow, and Turbulence Modeling, Chalmers University of
Technology, Goteborg, Sweden.

Derickson, R.G. and Meroney, R.N. (1977). A simplified physics air flow model for evaluating wind power
sites in complex terrain. In Proceedings of the Summer Computer Simulation Conference, July18–
20, 1977. Hyatt Regency, Chicago, Illinois.

Durão, D., Heitor, M., Pereira, J. (1988). Measurements of turbulent and periodic flows around a square
cross-section cylinder, Exp. Fluids. (6) 298-304.

Eymard, R., Gallouet, T., and Herbin, R. (2003). “Finite Volume Methods,” Handbook of Numerical
Analysis, Marseille, France. Vol. 7, 713-1020.

Flaschbart, O. (1930). Winddruck auf bauwerke, Naturwissenschaften 18.

Franke, J., Hellsten, A., Schlunzen, H., Carissimo, B., (2007). Best Practice Guideline for the CFD
Simulation of Flows in the Urban Environment, COST Action 732.

Franke, R. and Rodi, W. (1991). Calculation of vortex shedding past a square cylinder with various
turbulence models, in: Proc. 8th Symp. Turbulent Shear Flows, 11 September 1991, Tech. Univ.
Munich, Springer Berlin, pp. 189-204.

Hess, J, L. and Smith, A.M.O. (1967). Calculation of potential flow about arbitrary bodies. Prog. Aeronaut.
Sci. 8, 1-138.

Hinze, J. O. (1975). Turbulence. McGraw-Hill Publishing Co., New York, N.Y.

Hirt, C.W. and Cook, J. L. (1972). Calculating three-dimensional flows around structures and over rough
terrain. Journal of Computational Physics. (10), 324–340.

Holmes, J. D. (2007). Wind Loading of Structures, 2nd Ed. Taylor & Francis, New York, N.Y.

Irminger, J.O.V., Nokkentved, C. (1936). Wind Pressure on Buildings, Experimental Researches, 2nd
Series, No. 42, Translated by A.J. Jarvis.

Jensen, M. (1958). “The model law for phenomena in natural wind.” Reprint from Ingenioren (International
Edition), 2(4), 121-128.

121
Jones, W. and Launder, B., (1972). The prediction of laminarization with a two-equation model of
turbulence. Int. J. Heat Mass Transf. (15), 301–314.

Kareem, A. (1985). “Lateral-torsion motion of tall buildings to wind loads,” J. Struct. Eng., ASCE 111
(11), 2479-2496.

Launder, B. E. and Spalding, D. B. (1974). The Numerical Computation of Turbulent Flows, Comput.
Methods Appl. Mech. Eng., Vol. 68(3), 537-566.

Leitl, B.M., Meroney, R.N. (1997). Car exhaust dispersion in a street canyon. Numerical critique of a wind
tunnel experiment. J. Wind Eng. Ind. Aerodyn. 67 & 68, 293–304.

Logan, D. L. (2012). A First Course in the Finite Element Method, 5th Edition, Cengage Learning,
Stamford, CT.

Lyn D., Einav S., Rodi W., Park J. (1995). A laser doppler velocimetry study of ensemble-averaged
characteristics of the turbulent near wake of a square cylinder, J. Fluid Mech. (304) 285-319.

McDonough, J.M. (2007). “Introductory Lectures on Turbulence.” University of Kentucky.

Mendis, P., Ngo, T., Haritos, N., Hira, A., Samali, B., Cheung, J. (2007). “Wind Loading on Tall
Buildings.” Electronic J. of Structural Eng., Special Issue: Loading on Structures, 42-54.

Menter, F. (1994). Two-equation Eddy-Viscosity Turbulence Model for Engineering Applications, AIAA
J., Vol. 32(8), 1598-1605.

Meroney ,R.N. and Yamada,T. (1971). “Wind tunnel and numerical experiments of two-dimensional
stratified airflow over a heated island,” Winter Annual Meeting of ASME, Nov. 28-Dec. 2,
Washington, DC.

Moonen, P., Blocken, B., Carmeliet, J. (2007). Indicators for the evaluation of wind tunnel test section flow
quality and application to a numerical closed-circuit wind tunnel. J. Wind Eng. Ind. Aerodyn.
95(9–11), 1289–1314.

Munson, B. R., Young, F. D., and Okiishi, T.H. (2006). Fundamentals of Fluid Mechanics, 5th Edition,
John Wiley & Sons, Hoboken, N.J.

Murakami, S. (1998). Overview of turbulence models applied in CWE-1997. J.Wind Eng. Ind.Aerodyn.74-
76, 1–24.

Murakami, S. and Mochida, A. (1989). Three-dimensional numerical simulation of turbulent flow around
buildings using the k-ε turbulence model. Build. Environ. 24(1), 51–64.

Murakami, S. and Mochida, A., (1988). 3-D numerical simulation of airflow around a cubic model by
means of the k–ε model. J. Wind Eng. Ind. Aerodyn.31(2–3), 283–303.

Murakami, S., Mochida, A., Hayashi, Y. (1990). Examining the k-ε model by means of a wind tunnel test
and large-eddy simulation of the turbulence structure around a cube. J. Wind Eng. Ind. Aerodyn.
(35), 87–100.
122
Murakami, S., Mochida, A., Hayashi, Y., Sakamoto, S., (1992). Numerical study on velocity-pressure field
and wind forces for bluff bodies by k–ε, ASM and LES. J. Wind Eng. Ind. Aerodyn. 44(1–3),
2841–2852.

Murakami, S., Mochida, A., Hibi, K. (1987). Three-dimensional numerical simulation of airflow around a
cubic model by means of large eddy simulation. J. Wind Eng. Ind.Aerodyn. (25), 291–305.

National Archives of Australia (2015). Cyclone Tracy, Darwin – Fact Sheet 176.
http://www.naa.gov.au/collection/fact-sheets/fs176.aspx (Accessed on January 11, 2015)

Pielke, R. A. and colleagues (2008). Natural Hazard Review.

Rao, S.S. (1990). Mechanical Vibrations, 2nd Edition, Addison-Wesley Publishing Company, Inc., Reading,
MA.

Reinhold, T. A. (1977). “Measurements of Simultaneous Fluctuating Loads at Multiple Levels on a Model


of a Tall Building in a Simulated Boundary Layer.” Doctoral Dissertation, Virginia Polytechnic
Institute and State University, Blacksburg, VA.

Samali, B., Kwok, K. C. S., Wood, G.S., Yang, J.N. (2004). “Wind Tunnel Tests for Wind-Excited
Benchmark Building.” J. Eng. Mech., Vol. 130, 447-450.

Schott, T., Landsea, C., Hafele, G., Lorens, J., Taylor, A., Thurm, H., Ward, W., Willis, M., Zaleski, W.
(2012). The Saffir-Simpson Hurricane Wind Scale. (http://www.nhc.noaa.gov/pdf/sshws.pdf)

Scruton, C., Frazer, R.A. (1952). A Summarized Account of the Severn Bridge Aerodynamic Investigation,
HMSO, London, U.K.

Simiu, E. (2011). Design of Buildings for Wind – A Guide for ASCE 7-10 Standard Users and Designers of
Special Structures, 2nd Editon, John Wiley & Sons, Hoboken, N.J.

Simiu, E., Scanlan, R.H. (1978). Wind Effects on Structures, John Wiley & Sons, New York, N.Y.

Smagorinsky, J. (1963). "General Circulation Experiments with the Primitive Equations. I. The Basic
Experiment," Month. Wea. Rev. 91. 99–164.

Spurr, H.V. (1930). Wind Bracing the Importance of Rigidity in High Towers, 1st Edition, McGraw-Hill,
New York, N.Y.

Stafford Smith, B., Coull, A. (1991). Tall Building Structures: Analysis and Design, John Wiley & Sons,
New York, N.Y.

Stathopoulos,T. (2002). The numerical wind tunnel for industrial aerodynamics: real or virtual in the new
millennium? Wind Struct. 5(2–4), 193–208.

Sutton, O.G. (1949). Atmospheric Turbulence, Mathuen, London.

123
Tamura, Y. (2009). Wind and Tall Buildings. European and African Conferences on Wind Engineering 5th
Convention, Florence, Italy. http://www.iawe.org/Proceedings/5EACWE/K01.pdf (Accessed on
January 11, 2015)

Taranath, B. S. (2005). Wind and Earthquake Resistant Buildings – Structural Analysis and Design. CRC
Press, Taylor & Francis Group, Boca Raton, FL.

Taylor, G.I. (1914). “Eddy motion in the atmospheric.” Philos. Trans. Roy. Soc.

Tedesco, J. W., McDougal, W. G., Ross, C. A. (1999). Structural Dynamics: Theory and Applications.
Addison-Wesley Longman, Inc., Menlo Park, CA.

Versteeg, H. K. and Malalasekera, W. (2007). An Introduction to Computational Fluid Dynamics: The


Finite Volume Method, 2nd Edition, Pearson Education Limited, Essex, England.

Von Karman, T. (1967). The Wind and Beyond, Little, Brown & Company, Boston, M.A.

Wilcox, D.C. (1998). Reassessment of the Scale-determining Equation for Advanced Turbulence Models,
AIAA J., Vol. 26(11), 1299-1310.

Yamada, T. and Meroney, R.N. (1972). Numerical and wind tunnel simulation of airflow over an obstacle.
National Conference on Atmospheric Waves, American Meteorological Society, Salt Lake City,
October 12-15, 1971.

Yilmaz, E. and Duffin, B. (2014). “Computational fluid dynamic assessment of damaging wind loads on
the One Indiana Square tower.” Environ. Fluid Mech., 14, 795-819.

124

You might also like