Вы находитесь на странице: 1из 172

WIND FLOW CHARACTERISTICS AND THEIR EFFECTS

ON LOW-RISE BUILDINGS

by

ZHONGSHAN ZHAO, B.E., M.E.

A DISSERTATION

IN

CIVIL ENGINEERING

Submitted to the Graduate Faculty


of Texas Tech University in
Partial Fulfillment of
the Requirements for
the Degree of

DOCTOR OF PHILOSOPHY

Approved

Chair~ommittee

hool

December, 1997
ACKNOWLEDGEMENTS

First., I need to express my deep thanks to Dr. Partha P. Sarkar, chairman of my


doctoral advisory committeet for his constant guidance and encouragement throughout
this research. Thanks are also extended to several other committee members, Dr. Kishor
Mehta, Dr. James R. McDonald, Dr. J. Walter Oler, and Dr. Zhimin Zhang, for their
guidance and critical comments.
I wish to thank Dr. Douglas A. Smith and Dr. Jianming Yin for their help and useful
comments. My thanks are due to Thomas Gardner for his help in conducting some of the
experiments. Kim K. Aycock was so kind as to proofread my dissertation. I thank all my
friends in the Wind Engineering Program and from the Civil Engineering Department for
their friendship.
Financial support from the US National Science Foundation (CMS-9409869) is
greatly acknowledged.
My deepest thanks are due to my wife, Qinglin. It is her love, patience and
encouragement that make my life happy and joyful.
It is to my father and my late mother that I wish to dedicate this work.

ii
TABLE OF CONTENTS

ACKNOWLEDGMENTS ii
ABSTRACf vii
UST OFTABLES ix
UST OF FIGURES xi
CHAPTER
1. INTRODUCITON 1
1.1 General Remarks l
1.2 Background Information 2
1.2.1 Introduction 2
1.2.2 Wind Damage 2
1.2.3 Wind Load Codification 3
1.3 Objectives and Scope 4
1.3.1 A Brief Background Review 4
1.3.2 The CUSfiTU Cooperative Program 6
1.3.3 Research Objectives and Scope 6
2. LITERATURE REVIEW 9
2.1 Introduction 9
2.2 Wind-tunnel Simulation of Wind and Loads 10
2.2.1 Leading Edge Phenomenon 10
2.2.2 Wind Turbulence 11
2.2.3 Effects of Non-stationarity of Natural Wind 13
2.2.4 Characteristic Length-scale 14
2.2.5 Velocity Protile and Scale Effect 15
2.2.6 Reynolds Number and Blockage Effect 16
2.3 Separation Bubble and Conical Vortices 17
2.3.1 Separation Bubble 18
2.3.2 Conical Vortices 20

iii
2.4 Summary 23
3. WIND LOADS AND BUILDING PERFORMANCE 25
3.1 Introduction 25
3.2 Atmospheric Boundary Layer 25
3.3 Wind and Turbulence 27
3.4 Wind Effects on Buildings 29
3.4.1 Observations of Wind-induced Damage 30
3.4.2 Air-flow and Building Failure 32
3.4.3 Engineering Attention and Building Performance 35
3.4.4 Improvement of Building Practice 38
4. THE WERFL FACILITY 40
4.1 Introduction 40
4.2 Test Building and Meteorological Tower 41
4.3 Anemometers and Pressure Measurement System 44
4.3.1 Anemometers 44
4.3.2 Pressure Measurement System 44
4.4 Data Acquisition System and Data Validation 46
4.4.1 Data Acquisition System 46
4.4.2 Data Analysis and Validation 47
5. DESIGN OF EXPERIMENTS 48
5.1 Introduction 48
5.2 Incident Wind and Pressures 48
5.3 Wind-flow Measurement 49
5.4 Flow Visualization 51
5.5 Pressure Measurement 52
5.6 Comparison of Velocity Measured with Different Anemometers 53

iv
6. SEPARATION BUBBLE AND ITS EEFECTS ON PRESSURE 55
6.1 Introduction 55
6.2 Structure of Separation Bubble 56
6.2.1 Flow Measurement 56
6.2.2 Velocity Vectors 56
6.2.3 Boundary of Separation Bubble 60
6.3 Wind Flow Characteristics 61
6.4 Pressure Distribution 61
6.5 Mechanism of Peak-pressure Generation 65
6.5.1 Non-conventional Pressure Coefficient 66
6.5.21ncident Wind and Pressure Coefficient 66
6.5.2.1 Incident Wind 67
6.5.2.2 Conventional Pressure Coefficient 68
6.5.2.3 Wind Angle and Pressure Coefficient 69
6.5.2.4 Non-conventional Pressure Coefficient 71
6.6 Spectral Analysis of Pressure 74
6. 7 Summary 77
7. CONICAL VORTICES AND THEIR EFFECTS ON PRESSURE 79
7.1 Introduction 79
7.2 Theoretical Background of Vortex and Pressure 79
7.3 Characteristics of Pressure on Leading Roof-comer 81
7.3.1 Pressure Coefficient for Tap 50101 82
7.3.2 Pressure Coefficient for Tap 50501 89
7.3.3 Predicting Pressure Coefficient from Incident Wind 94
7.4 Pressure Zones and Critical Wind Angles 99
7.5 Features of Conical Vortices 101
7.5.1 Position of Vortex Core 101
7.5.2 Forms and Patterns of Vortex Pair 103

v
7.6 Comparison of Wind Velocity Measured with Different Anemometers 111
7.7 Summary 113
8. CONCLUSIONS AND SUGGESTIONS 117
8.1 About Separation Bubble 117
8.2 About Conical Vortices 118
8.3 Suggestions for Further Study 119
REFERENCES 120
APPENDICES
A. TECHNICAL SPECIFICATION
OF THE SONIC ANEMOMETER 126
B. DATA ACQUISffiON LOG 127
C. PRESSURE COEFFICIENT DATA
FOR SELECTED TAPS FROM MODE-15 145
D. DERIVATION OF EQUATION
FOR PRESSURE PREDICTION 158

vi
ABSTRACf

Two fundamental flow phenomena, the separation bubble (SB) and conical vortex~

over the roof (flat and rectangular) of the Texas Tech test building are studied in terms of
flow characteristics and pressure-generating mechanisms. Major fmdings contribute to
understanding the mechanisms of pressure generation and the roles of turbulence and
other properties of the incident wind in loading effects.
Much of the debate on wind-tunnel simulation priority in the wind engineering
community has failed to distinguish the different effects the incident wind has on two
types of quantities: single events, e.g., minimum (peak) pressures, and statistics~ such as
the mean and the rms pressures. The turbulence intensities, reflecting the gust structure
and the directional fluctuations of the free-stream wind, might bear significant influences
collectively on the pressure statistics. It is the intensity of individual longitudinal gusts
(as well as the incident wind angles) that decides the data-run-wise peak-pressure
coefficients on the roof corner~ while the peak-pressure coefficients associated with the
SB are governed by the lateral directional fluctuation in the incident wind. Proper
simulation of the incident wind profile (boundary layer type) is probably the most
important single input in the wind- structure interaction process.
The separation bubble, having a mean reattachment-point approximately 10 ft from
the leading edge, is oblong and elongated in the horizontal direction. Pressure distribution
on the roof surface is intimately related to the structure of the SB. Conditional sampling
technique indicates that the wind flow associated with strong suctions is highly three-
dimensionaL Introduction of the non-conventional pressure coefficient makes it possible
to isolate the effect of wind speed from that of the wind direction on pressure generation.
The mechanism of peak-pressure generation associated with the SB is governed by
directional fluctuations of incident wind. Primary peaks of pressure coefficient
(conventional) are often an outcome of combined wind gust and fast direction fluctuation.
Non-Gaussian-pressure zone along the short roof-axis is found to extend about l5 ft from

vii
the leading edge. This dimension is comparable to the stream-wise overall size of the
vortex circulation and to the height of the test building (13 ft).
Low pressures (high suctions) on a leading roof comer are always related to strong
vortices. The wind angles favoring vortex formation over a roof comer are bounded in a
50 range symmetric about the diagonaL Within this range, the form and dimension of a
single vortex, and the combination (pattern) of the vonex pair vary significantly with the
incident wind angle. Instantaneous pressure coefficients for taps underneath either of the
vortices can be satisfactorily predicted from the mean pressure coefficient and the wind
direction and speed of the incident wind. Peak pressures on the roof corner are produced
by wind gusts approaching at wind angles conducive to strong vortex formation. This
mechanism is different from that of the SB in that the fluctuations of wind direction (not
the direction itselt) are important in the case of the SB.
A practical means of wind load prediction is established. It is a uhybrid" approach
that combines the mean pressure coefficient.. obtainable from either full-scale experiment
or model test in properly simulated boundary layer wind flow, with easily available site-
specific wind data measured at an appropriate point.

viii
LIST OF TABLES

6.1 Statistics of velocity vectors 58


6.2 Intermittence factors 62
6.3 Coordinates and mean Cp of selected taps along the short roof-axis 63
6.4 Typical statistics of natural wind 66
7.1 Coordinates of selected pressure taps 82
7.2 Statistics of roof comer pressures9 averaged values over 220 - 230 83
7.3 Position of vortex core at 225 wind angle 102
8.1 Sonic data acquisition log (2/15/1997) 131
8.2 Sonic data acquisition log (2/16/1997) 131
8.3 Sonic data acquisition log (2/17/1997) 131
8.4 Sonic data acquisition log (2/21/1997) 132
8.5 Sonic data acquisition log (2/23/1997) 132
8.6 Sonic data acquisition log (2/26/1997) 133
8.7 Sonic data acquisition log (2/27/1997) 133
8.8 Sonic data acquisition log (2/28/1997) 133
8.9 Sonic data acquisition log (3/03/1997) 134
8.10 Sonic data acquisition log (3/06/1997) 135
8.11 Sonic data acquisition log (3/09/1997) 135
8.12 Sonic data acquisition log (3/13/1997) 136
8.13 Sonic data acquisition log (3/16/1997) 137
8.14 Sonic data acquisition log (3/18/1997) 137
8.15 Sonic data acquisition log (3/23/1997) 137
8.16 Sonic data acquisition log (3/24/1997) 137
8.17 Sonic data acquisition log (3/24/1997) 138
8.18 Sonic data acquisition log (9/28/1997) 140
8.19 Sonic data acquisition log (10/05/1997} 141
8.20 Sonic data acquisition log (10/08/1997) 142

ix
B.21 Sonic data acquisition log (10/12/1997) 142
B.22 Sonic data acquisition log (10/12/1997) 143
B.23 Sonic data acquisition log (10/12/1997) 144

X
LIST OF FIGURES

2.1 General flow pattern at normal-to-wall wind 14


2.2 Flow over flat roofs 18
2.3 Pressure distributions for flat roofs 19
2.4 Conical vortices 20
3.1 A sketch of the ABL in suburban areas 26
3.2 Horizontal wind-speed spectrum at about lOOm height 28
3.3 Time series of wind speed and direction 29
3.4 Building components affected by wind 30
3.5 Overall failure-mode of a building 32
3.6 Local airflow and loading effects 33
3. 7 Openings and internal pressure 34
3.8 A high-rise stood after the Lubbock Tornado 36
3.9 A fully engineered hospital survived a tornado 36
3.10 Damage to the roof of a marginally engineered building by a tornado 37
4.1 Trees shaped by winds 41
4.2 Test building and meteorological tower 42
4.3 Schematics of test building and defmitions of position and angle 43
4.4 Data acquisition system 46
5.1 Three-component sonic anemometer 49
5.2 Set-up of sonic anemometer for wind flow measurement on roof 50
5.3 Set-up of grid on roof of test building 51
5.4 Set-up of velocity comparison experiment 54
6.1 Mean structure of separation bubble 57
6.2 Boundary of separation bubble 61
6.3 Tap locations and pressure distribution 64
6.4 Wind flow near leading edge of the SB 65
6.5 Fluctuations of incident wind 68
6.6 Simultaneous traces of incident wind and pressures 70

xi
6. 7 Pressure coefficient versus wind angle 71
6.8 Wind-structure interaction and loading effect 72
6.9 Simultaneous traces of incident wind and pressures 73
6.10 Spectra of pressure measured at six taps along short roof-axis 76
6.11 Pressure zones on flat roofs: Gaussian and non-Gaussian 76
6.12 Typical trace of pressure coefficient: Gaussian 78
6.13 Typical trace of pressure coefficient: non-Gaussian 78
7.1 Pressure distribution under a free point-vortex 80
7.2 Instrumented roof comer and taps 81
7.3 Tap 50101 pressure coefficients versus wind angle 84
7.4 Schematic view of vortices as they affect pressure tap 50101 85
7.5 Straight flow and vortices formed at incident wind around 225 85
7.6 Time series of pressure coefficient and incident wind angle (M49nl60) 87
7.7 Time series of pressure coefficient and incident wind angle (M49n163) 88
7.8 Tap 50501 pressure coefficients versus wind angle 90
7.9 Location of tap 5050 1 with respect to vortices 91
7.10 Simultaneous series (M49n 161) 92
7.11 A close-up look of incident wind and pressure (M49nl6l, tap50501) 93
7.12 Mean Cp versus wind angle (tap 50501, Mode-15) 95
7.13 Traces of variables used in pressure prediction 96
7.14 Comparison of pressure coefficients 97
7.15 Schematic view of variability of upstream wind 98
7.16 Division of pressure zones on tlat roof 100
7.17 Schematic view of flow visualization setup 101
7.18 Position of vortex core derived from flow visualization 102
7.19 Schematics of vortices and incident wind angle 103
7.20 Vortex at the left edge (x =l0'-06n) 105
7.21 Vortex at the left edge (x =6ft) 105
7.22 Effect of wind angle on vortex shape and core position 106

xii
7.23 A pair of near-circular vortices 106
7.24 A big oblong vortex forming at the right edge 107
7.25 A big oblong vortex forming at the left edge 108
7.26 Vortex and its relationship to pressure-tap location 108
7.27 Wind direction and pressure coefficient 110
7.28 Comparison of longitudinal velocity spectra 112
7.29 Comparison of lateral velocity spectra 112
7.30 Comparison of horizontal velocity components 114
7.31 Comparison of vertical velocity component and wind direction 115
B.l Coordinate system 128
B.2 Definition of instrument u- v- w-components 129
B.3 Locations of sonic anemometer and grid 136
B.4 Schematics of experimental setup 138
B.5 Sketch of experimental setup for velocity measurement comparison 139
B.6 Definition of instrumental u- v- w-components (M49n252-259) 140
B.7 Defmition of instrumental u- v- w-components (M49n26l-265) 141
B.8 Defmition of instrumental u- v- w-components (M49n267) 142
B.9 Defmition of instrumental u- v- w-components (M49n269-272) 142
B.10 Defmition of instrumental u- v- w-components (M49n274-283) 143
B.ll Definition of instrumental u- v- w-components (M49n284-287) 144
C.l Tap 50901 pressure coefficient data 146
C.2 Tap 50205 pressure coefficient data 147
C.3 Tap 50505 pressure coefficient data 148
C.4 Tap 50905 pressure coefficient data 149
C.5 Tap 50209 pressure coefficient data 150
C.6 Tap 50509 pressure coefficient data 151
C.7 Tap 50909 pressure coefficient data 152
C.8 Tap 50123 pressure coefficient data 153
C.9 Tap 50523 pressure coefficient data 154

xiii
C.lO Tap 50823 pressure coefficient data 155
C.11 Tap 51123 pressure coefficient data 156
C.12 Tap 51423 pressure coefficient data 157
0.1 Definition sketch 158

xiv
CHAPfERl
INTRODUCTION

1.1 General Remarks


The sphere we call the earth is a magnificent creation of the universe. A thin layer of
gaseous mass known as the atmosphere envelops it and presents it as a beautifuL blue
planet to a viewer in outer space. However.. under this veil of peacefulness hides rage and
violence. Thunderstorms9 earthquakes and volcanic eruptions are occurring every minute
on the earth's surface or deep inside.
Wind is a part of our daily lives. While wind can be beneficial to humankind's
enterprises, it can also be extremely destructive to buildings and other structures we build
on the surface of the earth. The average annual dollar losses caused by natural phenomena
hazards from 1991 to 1995 were over 10 billion~ of which damage inflicted by extreme
winds constitute approximately 70%. With hurricanes Andrew, Iniki and Opal, annual
dollar losses totaled 23 billion and 17 billion in 1992 and 1994, respectively. Predictions
indicate that hurricane activities will increase in the next thirty years. Natural disaster
mitigation is becoming more and more important; the responsibility lies with the science
and engineering community as well as with the general public.
Humanity's exploration of wind began with mythology. Early civilizations saw in the
wind the manifestation of a divine force used by benevolent or malevolent gods.
Mankind's passive acceptance of wind~s destructiveness, however. is giving way to a new
era of understanding. Specific contributions to the knowledge of wind in the lower portion
of the atmosphere near the earth's surface have been made by Ekman, Prandtl, Taylor and
Sutton. A new discipline called wind engineering has emerged in the past three decades. A
stage has now been reached such that engineers can design structures to minimize the
harmful consequences of severe winds once thought to be unavoidable. Notwithstanding,
some fundamental questions remain unclarified. From the viewpoint of structural design,
we need to know wind better; we need to improve the precision with which we determine
the loading effects of wind.
2
1.2 Background Information
1.2.1 Introduction
With hurricanes Hugo (1989), Andrew (1992), Opal (1994) and the Northridge
earthquake (1994), the 1989-1995 period has seen an unusual number of natural hazards
and huge losses in property. Technically, improvement of wind resistance capacity of
buildings and other structures, and thus reduction of property losses inflicted by severe
winds, involves two steps: (1) understanding wind effects on buildings, and (2)
transferring and implementing technical advances into engineering practices. The technical
aspect of the issue, critical as it is, presents only part of a rather complex problem
attending the improvement of wind resistance of buildings and other structures.
Construction practices include intricate relationships among designers, contractors, codes,
insurers and the general public (Minor, 1981). Awareness and preparedness of our society
as a whole is the key to effective mitigation and fast recovery from a natural disaster.

1.2.2 Wind Damage


Although there are variations in wind characteristics of hurricanes, tornadoes and
other extreme winds, their general loading effects on structures are basically the same.
While hurricanes can cause extensive damage to wide coastal areas, damage inflicted by
tornadoes is usually limited to a much smaller scale. Hurricane Hugo caused extensive
roof damage in Charleston, South Carolina and the surrounding areas (Smith &
McDonald, 1990). In the Charleston area, probably in excess of 75% of the roofs suffered
at least minimal damage. The extensive roof damage resulted from two things:
1. Roof systems experienced higher loading than any other elements of a building;
2. Roof systems often did not receive adequate engineering (design) attention.
Assessment of wind-induced damage to structures revealed some consistent patterns
(Minor, & Mehta, 1979):
l. Structures failed, principally, because of wind-induced forces acting on critical
building components;
3
2. Non-engineered and marginally engineered structures were susceptible to failures
at relatively low wind speeds;
3. Small increase in the degree of engineering attention~ using new wind engineering
technology, could produce large dividends in increasing the wind resistance of
structures;
4. The geometry of a structure and its orientation were important detenninants of
wind resistance.
Inadequate design due to a deficiency on the part of standards or codes, and a lack of
understanding of wind-induced forces were believed to be the reasons behind the
structural failures under lower-than-design wind speeds (Mehta~ 1984).
On-site surveys of building performances in windstorms indicate that high-rise
buildings, which receive a significant amount of engineering attention in design and
construction, generally perform satisfactorily in maintaining their structural integrity. Their
frames are adequately designed to sustain lateral wind loads. Component damage is the
main concern with high-rise buildings. On the other hand, low-rise industrial, commercial
and residential buildings may incur damage ranging from component and cladding failure
to catastrophic structural destruction. These structures are usually inadequately engineered
for the action of the lateral forces as well as the uplifting loads (Mehta, 1984). The causes
of component and cladding failure are likely to be twofold: poor construction practices
and under-design. The latter calls for a better understanding of localized intensive wind
forces and for more precise wind-loading specifications in our standards and codes.

1.2.3 Wind Load Codification


Normally. standards or codes specify wind loads in terms of pressures or forces. In
general, two approaches are accepted to determine the wind-loading effects on buildings
and other types of structures. One is a relatively simple procedure, appropriate for use
with the majority of wind loading applications, including the structural design of low and
medium rise buildings and the cladding design. These are situations concerning relatively
4

rigid structures, and can be dealt with by equivalent static loads. Standards and codes
often specify non-dimensional pressure coefficients for different structural elements for
limited variations of buildings, and specify force coefficients for other structures. These
coefficients are used in conjunction with site specific wind data and terrain characteristics
to detennine design wind loads. In most cases, the coefficients are established from wind-
tunnel tests conducted in simulated atmospheric boundary layers. The other approach is
applicable whenever the buildings and other structures are likely to be susceptible to wind-
induced vibrations. For these buildings and structures, customized wind-tunnel tests or
other experimental methods~ or a detailed procedure is to be followed to determine wind
loads. Wind-tunnel tests are more appropriate when more exact definition of dynamic
response is needed, and for determining exterior pressure coefficients of buildings with
complicated geometry.
It is thus obvious that adequate and efficient structural design against wind induced
damage relies heavily on precise specification of wind loads. Wind-tunnel testing
techniques have made significant contributions in the load codification. Engineered
structures such as the main-wind-force-resistance frames of many high-rise buildings have
performed well in severe storms. However, there is still a need to improve the
performance of residential buildings and other types of low-rise structures. Peak pressures
of short duration are important in the case of low-rise buildings, where the damage of
components and cladding often leads to structural destruction. Roof edge and roof comers
sustain the largest peak pressures of all the surfaces of a building. It is important to
understand the mechanism of pressure generation for appropriate design and developing
strategies to reduce the intensity of pressures in those regions.

1.3 Objectives and Scope


1.3.1 A Brief Background Review
Three parallel and yet complementary approaches are taken in wind engineering
research of wind effects on structures. They are full-scale investigation~ wind-tunnel
5
simulation and numerical modeling. Data acquired from full-scale tests can be compared
with wind-tunnel predictions. These two dissimilar approaches see a unique opportunity
for convergence through numerical modeling. Numerical models can be corroborated and
validated by actual field data and wind-tunnel tests.
Research on numerical modeling gained considerable momentum in the past decade
and the its development has now reached a point where a sub-discipline called
computational wind engineering (CWE) can stand on its own merit.
The past two decades have seen extensive research in the area of wind-induced loads
on low-rise buildings~ primarily using the wind-tunnel technique, and the results have had a
significant impact on wind load codification. On the full-scale investigation, the Wind
Engineering Research Field Laboratory (WERFL) at Texas Tech University began to
publish high quality pressure-coefficient data in 1989 (Levitan et al., 1989). Comparisons
of results naturally ensued. Wind-tunnel simulations performed by various researchers
could predict the mean pressures on building surfaces. Numerical simulation of wind loads
was successful in predicting the mean wind pressures for a few cases involving simple
building geometry (Selvam and Konduru~ 1993; Mochida et aL, 1993). Discrepancies
between the full-scale observations and the simulation results still exist and need more
research.
In advance of full-scale data from the WERFL at Texas Tech University, a model-
scale experiment was undertaken to define the characteristics of the local pressures. The
designed test was partly to aid in the definition of the full-scale experiment and partly to
provide an unbiased set of pressure data for comparison (Surry, 1989). Comparisons of
the wind-tunnel data with the earliest sets of full-scale measurements revealed significant
differences in peak pressures measured on the leading roof comer at oblique wind angles.
Specifically, the peak pressures in full scale were much larger than the model results.
Better comparisons were achieved for the taps on the short roof-axis at wind normal to
the ridge. Further, the difference in peak pressures at skewed wind angles extended to
both the windward wall and the wake. This discrepancies strongly suggested that the gust
6
structures of natural winds were significantly different from that of the generated wind,
since these regions were expected to exhibit quasi-steady type of response, non-
stationarity would exaggerate the peak-to-mean ratios (Surry, 1989). However, the mean
pressure coefficients from model tests compared quite well with full-scale measurements;
and the agreement in the rms values of pressure coefficients were reasonably good.
Cochran et al (1992}, Jamieson and Carpenter (1993), Lin et aL (1995), Rofail (1995),
and Tieleman (1995), among others., investigated the discrepancies in wind-tunnel tests.
Different simulation techniques emphasizing various aspects of wind flow simulation were
employed. A need arose for full-scale study to examine the issue in a parallel manner.

1.3.2 The CSUfiTU Cooperative Program


The US National Science Foundation sponsored Colorado State University I Texas
Tech University Cooperative Program in Wind Engineering, hereafter referred to as
CSUIITU-CPWE, is currently in its second five-year phase (CPWE-m. This program
was initiated in 1988, and it started operation in June 1989. The objectives of the first five-
year program (CPWE-1) were: (1) to focus an interdisciplinary effort on selected wind
engineering research programs related to the quality and safety of man's environment, and
(2) to disseminate results of this research during the International Decade for Natural
Disaster Reduction. The CPWE-I included eight tasks (TIU/CSU, 1995a) covering a
variety of topics of wind engineering. The current CSU!ITU CPWE-II (1995- 1999) has
been organized to focus primarily on three technical research topics related to the behavior
of low-rise buildings under the influence of severe winds generated by thunderstorm-scale
winds (TTU/CSU, 1995b). One of the focus topics is wind flow around buildings, and the
research described is part of that topic.

1.3.3 Research Objectives and Scot>k


Over the years, the intense pressures observed at upwind roof corners and leading
edges on flat roofs of low-rise buildings have been a subject of intensive research. It is
7
generally accepted that unsteady pressures on a structure's envelope depend primarily on
the geometry of the structure, the relative direction of wind to the structure and the
incident wind conditions. Although extensive research has been conducted during the last
two decades, the exact mechanisms of the extreme pressures in the critical areas of a
building envelope are not yet clear (Tieleman, 1995).
Damage to roofs was often associated with the comers, ridges and eaves. Roof
damage was the heaviest at the ridge and gable ends for sloped roof systems, and at the
roof comers for flat roofs (Vognild, 1993). In Hurricane Iniki, more than 90% of all
single- or multiple-family dwellings lost substantial portions of their roof coverings. This
was true for the less expensive housing as well as for the more expensive properties
around the Island of Kauai (Chiu et al., 1993). Damage to engineered structures was
generally restricted to the roofing and glazing systems (Levitan et al., 1993). The real-
world situation being such, that questions were raised on the ability and enforcement of
current codes to prevent, or mitigate to an acceptable level.. such damage from recurring in
hurricanes (Vo gnild, 1993).
Mechanisms of peak and the root-mean-square (rms) pressure generation for those
roof regions characterized by the presence of flow separation and formation of strong
vortices are not well understood. These flow phenomena are associated with roof comers,
ridges and eaves, the zones most susceptible to wind induced damage.
Factors bearing effects on the wind loading on structures are the structural geometry
and the properties of incident wind. The latter includes characteristics which are not
always independent from each other: the mean wind speed profile, the turbulence
intensities and the turbulence integral scales associated with the three velocity
components, spectral densities, roughness length, atmospheric stability conditions, non-
stationarity of wind, Reynolds effects in wind-tunnel simulations, and so on. However,
divisions exist as to which parameters are essential or most important to the successful
prediction of wind loads by the wind-tunnel simulation (Melbourne, 1980; Atkins, 1992;
Tieleman, 1992; Cochran, 1992; Dalley, 1993~ Tieleman, 1995). Roughly speaking, one
8
school argued that the mean velocity profile, the longitudinal turbulence intensity and the
turbulence integral scale were the essential parameters to simulate. The other claimed that
the turbulence intensities of lateral and vertical components were equally or even more
important; besides, small-scale turbulence contents must be properly simulated, and the
velocity profile (Zo) and the integral length scale (x") were only secondary in priority.
Fortunately, despite all the divisions among researchers, there seemed to exist a
consensus that the problematic issue is not fundamentaL and it is hoped that clarifications
of the wind - structure interaction would eventually make it possible to precisely predict
wind loads for adequate structural design against wind damage. The full-scale study can
uniquely contribute to a better understanding of wind loading effects.
Synchronized measurement of the incident wind just before it hits the building, and
the external pressures along with the aid of flow visualization technique seemed to be a
promising approach. The objectives are to understand the wind-structure interaction
process and the mechanisms of pressure generation for two flow phenomena:
1. The separation bubble at normal-to-wall wind, which creates large-magnitude
suctions in the roof regions close to the leading edge; and
2. The conical vortices at cornering wind, which produce the worst-case suction at
the leading roof corner.
The scope of this study is to:
1. Document the flow characteristics over flat roofs;
2. Clarify the roles of the wind-speed profile, the turbulence and gust structure, and
the directional fluctuations of the incident wind in generating pressures on low-
rise building roofs;
3. Find a practical means to predict wind loads.
CHAPTER2
LITERATURE REVIEW

2.1 Introduction
Historically~ wind engineering has focused on preventing failure of the main-wind-
force-resistance-system (MWFRS) of structures. However, direct losses from the breach
of the building envelope and the consequential water damage to structural elements and
building contents have been tremendous when severe storms struck the increasingly
constructed and densely populated coastal regions. Recent years have seen a shift of focus
on improving the performance of building envelopes for mitigation of property losses.
The regions of worst suction~ where initial damage starts, on the roofs of low-rise
buildings, comprise only a small fraction of the total roof area and its influence on
structural loads is small; it is, however, critically important for roof-cladding or roof-
covering systems (Kind~ 1986). Failure of roof systems tends to begin at the upwind
corner of the rooftop, and an initial failure could precipitate failures over large areas (Kind
and Wardlaw, 1982). Our ability to design against such component failures underlies
considerable economic significance, and this, to a large degree, depends on how well we
know the complicated wind loading effects on structures.
As aforementioned, the technique we have relied on heavily for determining wind
loads has room for improvement. The staggering amount of property-loss our society has
suffered recently testifies to this. The unpleasant reality, from the technical viewpoint, is
that it is precisely with those critical regions in a building envelope that we have been least
successful in load predicting. Significant discrepancies of wind-induced pressures were
found in modeVfull scale comparisons for those regions.
Two flow phenomena associated with the critical loading can be identified for tlat
roofs of low-rise buildings: ( l) flow separation and possible reattachment, at normal-to-
wall wind; and (2) conical vortices formed over roof corners at oblique wind. In the
situation when the wind approaches normal to one of the walls., the flow above the

9
10
stagnation point on the windward wall will go over the roof. The up-wash flow separates
at the roof edge to form a separated shear layer.. which may reattach to the roof at some
distance downstream if the stream-wise dimension of the roof is large enough, to fonn a
separation bubble (SB). When wind hits a building at skewed angles, a conical vortex or a
pair of vortices may form after flow separation at the two oblique leading roof edges.
Both flow phenomena involve separation and vortex formation, which is closely related to
the pressures developed on the roof. The magnitude of the outward-acting pressure is
strongly dependent on the condition of the vortices, which are themselves very sensitive to
the incident wind, specifically, to the fluctuations of the wind speed and direction.
This chapter intends to present a thorough literature review of the two flow
phenomena and their loading effects on flat roofs of low-rise buildings, and on the state-
of-the-art of wind-tunnel simulation technique.

2.2 Wind-tunnel Simulation of Wind and Loads


In spite of all the progress made in the past few years .. physical simulation of the
wind's loading effects still has plenty of room for improvement. Significant
underestimation of the worst-case peak pressures and the root-mean-squared or nns
pressures near the leading roof-comer and edges remained a problematic issue (Tieleman
et al., 1993). Explicit and definitive guidance toward fundamental resolution is lacking.
The state-of-the-art of wind-tunnel simulation of wind loading on low-rise buildings is
such that a debate of simulation priorities has been going on for quite some time. A
thorough discussion on this matter will be attempted in this section.

2.2.1 Leading Edge Phenomenon


In wind engineering, the term leading edge phenomenon implied some relationship
between pressure fields on a stream-wise surface associated with separated flow
downstream of a sharp leading edge. The occurrence of large magnitude pressures on
these surfaces on which cladding was designed, the cross-wind response of towers and
11
buildings, and the vertical response of a bridge deck or a roof canopy., were all examples
of wind-structure interaction which had their origin in the leading edge phenomenon
(Melbourne., 1992).
Although the mechanisms causing very low pressures on stream-wise surfaces were
not fully understood., the boundary conditions necessary for their occurrence were well
known. High free-stream turbulence combined with an edge discontinuity usually sets the
scene for the occurrence of large magnitude suctions under the reattaching shear layer.
While the general role of turbulence was reasonably understood, it was not clear how the
extreme pressures were generated when the vortex was anchored at one end by an edge
discontinuity (Melbourne, 1992).

2.2.2 Wind Turbulence


The first known documented appreciation of the effects of incident wind turbulence
on the pressure on a stream-wise surface under a separated shear-layer had its origin in
wind engineering study. In a very early work by Bailey (1933), full-scale and model-scale
pressure measurements on small buildings were compared. Discrepancies were attributed
to the low-turbulence in the uniform flow used in the model tests. Tests in the boundary-
layer wind yielded much improved comparisons (Bailey and Vincent, 1943). Jensen (1958)
demonstrated that increased turbulence level at eaves height had significant effects on the
pressure distribution over the roof.
Gartshore (1973) provided the first incisive description of the effects of free-stream
turbulence on the flow around bluff bodies. He suggested that the effects of increasing
free-stream turbulence was to increase the turbulence mixing in the shear layer and
promote entrainment of air in the wake, leading to a decrease of the radius of curvature
radius of the shear layer. The key observation he made was the recognition of the role of
the turbulence in affecting a reduced radius of curvature in the shear layer.
Melbourne (1980) furthered Gartshore's conclusions with respect to the significant
role of the small-scale turbulence. and introduced a parameter to quantify the spectral
12
density of the incident free-stream flow. He showed that increasing the small-scale
turbulence, while maintaining the large-scale turbulence, significantly increased the
magnitude of pressures under the reattaching separated shear layer. Later, he claimed that
incident turbulence certainly played a dual role at both the fine-scale and the large-scale
ends (Melbourne, 1992).
Hillier and Cherry (1981) and Cherry et al. (1984) investigated the effects of free-
stream turbulence on the separation bubble generated by the blunt leading edge of a long
flat plate. They showed that the mean and the fluctuating pressures under the bubble were
very dependent on turbulence intensity, and the mean flow structure responded strongly to
turbulence intensity but was not greatly affected by the integral scale; the mean
reattachment length decreased with increasing free-stream turbulence intensity.
Wind-tunnel experiments were conducted by Saathoff and Melbourne ( 1987) to study
the effects of blockage and free-stream turbulence on stream-wise surface pressures in a
separation bubble. The mean and the fluctuating pressures were measured near the leading
edge of short, axisymmetric cylinders in grid-generated turbulence and in smooth flow.
The mean pressures were found to be strongly dependent on the turbulence intensity but
were little affected by the integral scale. Auctuating pressures were dependent on both the
(longitudinal) turbulence intensity and the turbulence integral scale.
Tieleman's (1992) argument about wind simulation was probably the most explicit
and conclusive up to date:
L Duplication of the small-scale turbulence was more important than the velocity
profile scaling (Zo) or the integral scale (L.ru);
2. Simulation of the surface-layer flows over complex terrain required careful
simulation of both the small-scale turbulence and the lateral turbulence intensity
or the directional fluctuation;
3. New wind-tunnel simulation techniques must be developed to take care of the
requirement of wind direction fluctuations;
13
4. High suction pressures at the edges and comers of roofs could not be predicted
from wind-tunnel tests using conventional simulation methods;
5. Inadequate simulation of secondary flow effects associated with vonex generation
was responsible for the deficiencies.

2.2.3 Effects of Non-stationarity of Natural Wind


Natural wind is highly variable in both speed and direction. Wind speed and direction
both fluctuate in a non-stationary, random manner. Observations of natural winds indicate
that wind speed fluctuations in the order of plus and minus 10 mph is common at averaged
winds of 20 mph. Wind directions typically fluctuate 70 in range, 10 in the root-mean-
squares, for 15-minute records documented by the WERFL of Texas Tech University.
These characteristics are extremely difficult to reproduce in conventional wind-tunnel
simulation and have posed problems for comparative studies of wind loading effects on
low-rise buildings. This is best revealed in the comparison of the primary peak pressures~

which typically realize only a couple of times in 15 minutes. Relevant works are scarce in
the literature. Only a few authors have made brief statements about the non-stationarity of
wind possibly being the reason for full-scale- model-scale mismatch.
Surry ( 1989) compared the mean, rms and peak pressure distributions from the full-
scale data of the WERFL and that from the wind-tunnel results. He used the wind-tunnel
data for an oblique wind angle of 60 and the full-scale data with wind angles in the range
of 50 - 70. Reasonable agreement was found in the mean and rms pressure data, but the
peak pressure data in the full-scale were much larger than the model predictions. The
scatter bands, indicating the span of the results from 8 full-scale runs, showed that the
results for this wind angle were highly variable; significant differences extended to both the
front walls and the wake, in the regions where quasi-steady type of response was
expected. He concluded that this observation was strongly suggestive of significant
differences in the gust structures of natural wind and that of the simulated winds. Lin et at
14

( 1995) aJso pointed out that, among other factors, non-stationary effects in the full-scale
might be the cause of discrepancies.

2.2.4 Characteristic Length-scale


For low-rise buildings having hlb<1J.5, e.g., the WERFL building, the wind-flow finds
it easier to go over the roof rather than around the building, except at the two ends of the
front wall (Figure 2.1). On the contrary, in the case of high-rise buildings, most flow finds
its way going around the building. Cook ( 1990) reasoned that the characteristic length-
scale for the wind flow and loading for low-rise buildings was the height of the building.

Figure 2.1 General flow pattern at nonnal-to-wall wind (hlb<1J.5, after Cook~ 1990)

For the conical vortices developed over the upwind roof-comer, Lin et al. (1995)
argued that the length from the stagnation point on the two windward walls, where a
boundary layer began, to the roof edge, where the boundary layer separated, was an
appropriate characteristic length. Since the length over which the wall boundary developed
was a portion of the wall height (h), h could be taken as a convenient scale length. They
went on to suggest that similar wall-boundary-layers would develop for similar geometry.
In the same paper, Lin et al. used pressure models having plan dimensions 3 times the
length of the long side of a 1:50 model of the WERFL building, with varied heights of 1, 2
and 3 times of the exactly scaled model They found that although the mean, rms and peak
15
values of Cp at the same location, on a rays or at (x, y). changed from model to model. the
pressure coefficients seemed to be independent of the plan dimensions, especially for the
mean pressure coefficient. However, the curves of Cp versus s/h or (xlh, ylh) essentially
collapsed for different model heights (in the same flow) and thus tended to be independent
of the building dimensions. They suggested adopting the height rather than one of the plan
dimensions as the characteristic scale, provided the height to plan dimensions ratio fell in a
certain range (Lin et al, 1995).

2.2.5 Velocity Profile and Scale Effect


The Jensen number, Je = hiZtJ, defined as the ratio of the building height (h) to the
roughness length ( ZtJ) or equivalently the mean-velocity profile of the simulated boundary-
layer, represents the scale effect in the wind tunnel simulation. A variation of it is the ratio
of building height (h) to the thickness of the boundary layer (t)} in which the building
model is immersed.
By obtaining a good collapse of data for a wide range of building height and terrain
roughness, Akins et a1 (1977) had demonstrated that the effect of the Jensen number on
over-all forces was small enough to be ignored, provided it was in the correct range.
Stathopoulos et al. (1978) did not find any dramatic scale effect. Bachlin et al. (1983)
found only minor scale effect in their experiments covering a large range of h/8.
Stathopoulos and Surry ( 1983) reviewed scale effect.
Maximum sensitivity to scale effect appeared to occur where turbulence intensity
and/or scale-length could influence a reattachment process (Kind, 1986). Kind observed
only a weak tendency for suctions to decrease with increasing h/8. The definition of the
pressure coefficient based on the roof-top-level mean velocity eliminated much of the
effect of the variation of the approach wind velocity with the height. However, other
factors related to h/8 could be imponant, notably the non-dimensional velocity gradient,
(h/V)(dV/dz) .. at the rooftop level, the turbulence intensity and the turbulence scale. For

large values of h/8, the rooftop was below the logarithmic portion of the velocity profile;
16
that is., it was in a non-universal portion of the flow where conditions depend on the
specific characteristics of the underlying surface roughness. In such cases.. the roughness
parameter Zo or the ratio hiZD would have little direct physical significance. The non-
dimensional velocity gradient at roof level appeared to result in lower pressures. Because
increased velocity gradient could be expected to result in reduced flow up-wash and thus
reduced suctions.. in a similar way as reduced angle-of-attack of a delta wing leading to
smaller lift on the wing.
L'1 an extensive study by Lin et al. (1995) .. it was observed that the largest magnitude
of Cp generally increased with model height and changed with flow characteristics. It was
also found that., near the roof corner, the mean pressures were much more sensitive to the
change of flow characteristics than the peak or rms pressures. Away from the corner, the
peak and rms pressures became more sensitive.
Although the difference of pressures induced on models immersed in the uniform
flows (Je =ex:) and in the atmospheric flows, typically Je =50 - 500, was remarkable, it
was usually sufficient to ensure that the Jensen number was in the correct range (Cook,
1990).

2.2.6 Reynolds Number and Blockage Effect


The Reynolds number (Re) is a non-dimensional quantity describing the relative
importance of fluid inertia and viscosity. Its values in model tests are usually smaller than
their prototype counterparts by an order of two. Generally, the Reynolds number affects
the flow transition and separation. With the structures having sharp edges, as the majority
of structures are, the separation of flow is defined at those sharp edges, as long as the Re
is large enough for inenial forces to dominate. Recently, however, theRe was suspected
as one of the culprits of the model-full scale mismatch. Hoxey et al. (1996) claimed that
the Re was responsible for the discrepancies of the mean pressure coefficients between the
model tests and the full-scale investigations. They tested a slope-roofed building with
sharp eaves, major differences were found with regions on the roof involving flow
17

separation. In their view, almost all the factors other than the Re were Correctly' scaled
4

and simulated. They claimed that the model test results were consistent with the full-scale
data on the basis of the Re, and the differences could be corrected by extrapolating the
full-scale data to the model results in tenns of the Re. In another work., Savory and Toy
(1996) compared the pressure distributions of a 1:43 Silsoe Structures Building model
tested at the Re ranging between 2.35xl04 to 7.14xl04 , based on the ridge height. They
observed significant reduction of suctions at the leading eaves of the windward slope with
decreasing Re; the greatest dependence of pressure coefficients on the Re was with
regions having surface discontinuities, that is at the windward eaves and the ridge.
Unlike the situation of flow-structure interaction in the open atmosphere, the
streamlines around the model structures in the wind-tunnel tests are not completely free to
extend laterally; they are confined to some degree by the presence of the two side walls
and the ceiling of the wind tunneL Flow is accelerated in the constricted segment between
the model and the walls, increasing the loading of the structure. This is called blockage
effect and needs be minimized by choosing a low ratio of the projected area of the model
to that of the wind tunnel cross section. While approximate corrections to some extent are
possible, it is always a good practice to work with low blockage ratio. Studies at the
Bristol University in UK indicated that ground-mounted models were less susceptible to
blockage effect and 10% blockage might be acceptable without correction (Kirrane and
Steward, 1978), although most workers regarded 5% as the 'safe' limit (Cook, 1990).

2.3 Separation Bubble and Conical Vortices


The separation bubble and the conical vortex are two different flow phenomena; the
former is not a special case of the latter at the normal-to-wall wind. They are equally
important from the wind-load point of view. In the full-scale studies in the WERFL at
Texas Tech University, comparable worst suctions were documented for these two flow
situations, at different azimuth angles. At normal-to-wall winds, peak pressure coefficients
of about -10 were measured at a central cross-section tap close to the leading roof edge.
18
The pressure coefficients are expected to decrease in the lateral direction towards the two
ends of the roof. The WERFL at Texas Tech University documented peak pressure
coefficients as low as -13 on the roof comer for oblique wind directions.

2.3. 1 Separation Bubble


Consider the case of wind approaching a flat-roofed building. The flow over the roof
and the pressure distribution on the underlying roof were very different for uniform
incident wind and for boundary-layer type wind (Cook, 1985). With the uniform incident
wind., the wind flow from the full height of the building rose over the roof. The
downward-curving separated flow might not reattach onto the roof, see Figure 2.2(a). In
the boundary-layer flow, wind flow from the upper third of the building rose over the roof,
the displaced flow had higher kinetic energy because of the mean velocity profile, and the
separated streamline was lower. In this situation, the vorticity in the shear layer along the
separation boundary and the Reynolds stress with the boundary-layer type flow drove the
downward motion of the separated flow, leading to a much earlier reattachment (Figure
2.2(b)).

Momep.\um from
V Reynolds stress
~
.,.__-w~--

(a) Uniform wind flow (b) Boundary-layer wind Flow

Figure 2.2 Flow over flat roofs (reproduced after Cook, 1985)
19

~ Cp=-1.2

. Q.65
...
(a) Uniform wind flow (b) Boundary-layer wind flow

Figure 2.3 Pressure distributions for flat roofs (reproduced after Cook, 1985)

As a direct result of the different flow characteristics over the roof, the pressure
distributions on the roof were very dissimilar (see Figure 2.3). The pressure distribution
changed from nearly unifonn in the unifonn incident flow to increasingly negative toward
the upwind roof edge.
The flow inside the separation bubble at the central cross-section was basically a two-
dimensional flow where the reattachment process depends on the mass balance within the
separation bubble. The turbulence structure of the incident wind affected the entrainment
into the reattaching shear layer, and thus had a strong and direct influence on the mass
balance inside the bubble, and hence on the reattachment process and the pressure
fluctuations (Kind, 1986).
The size of the separation bubble was expected to scale to the across-wind dimension
of the front wall or two times the building height, whichever is smaller. The dynamics of
the vortex governing the loading characteristics on the roof depend on the wind angle, the
slenderness ratio, the roof pitch and so on; it was also strongly influenced by the incident
turbulence (Cook, 1990).
High suctions of short duration were found in regions of the roof close to the leading
edge both in full-scale measurement and in model test with properly simulated boundary-
layer flows (Cook, 1985). Melbourne (1975) speculated that the observed high suctions
were associated with instability of the shear layer triggered by incident turbulence.
20
Wagaman (1993) and Letchford (1995) are among the early researchers in full-scale
investigation of the separation bubble (SB). Wagaman (1993) established the length and
height of the SB over the WERFL test building through flow visualization; Letchford
(1995) made some observations about the dynamics of the SB and hypothesized the
mechanism of peak-pressure generation. Their work will be referred to later for
comparisons with this study.

2.3.2 Conical Vortices


In a real situation., the incident wind is more likely to be skewed at angles other than
normal to one of the walls. At skewed wind., flow separating at the upwind edges of the
roof has a velocity component along the line of separation.

(a) Aow structure (b) Pressure distribution

Figure 2.4 Conical vortices (reproduced after Cook, 1985)

As sketched in Figure 2.2(a), the flow separating along the edge entrains into a
circulation which increases as it moves downstream to create a pair of conical vortices.
These two vortices would have different strength unless they are formed at quartering
wind (Cook, 1985).
21

The conical vortices forming over the upwind roof corner of a low-rise building are
quite similar to the vortex over a delta wing at a moderate angle of attack. The vortices
break down at a certain point downstream. The center of each vonex is a region of high
negative pressure. The two vortices in Figure 2.4(a) produce two lobes of negative
pressure on the underlying roof comer~ as shown in Figure 2.4(b). These vortices~ acting
directly or interacting with the incident turbulence, are the principal cause of high uplift on
roofs (Cook, 1985).
Unlike the vortex of the SB, conical vortices are three-dimensional and highly stable
(Kind, 1986). Fluid particles in the vortices follow helical paths. The three-dimensional
reattachment process is governed by the laws of vortex dynamics and is essentially
independent of the viscous effects or the turbulence structures of the shear layer. This
statement is supported by the fact that inviscid analyses have been highly successful in
predicting this type of flow. Consequently, the turbulence intensity and scale is not likely
to have any significant effect on the three-dimensional conical vortex flow by virtue of
their effects on turbulence entrainment processes (Smith., 1984). In fact Stathopoulos et al.
(1978) found the occurrence of similar worst mean-suctions for very different turbulence
intensities of 35% and 15%. Direction fluctuation in the approaching wind due to the
turbulence causes a ''wandering"' of flow pattern and thus a 'smearing' of the mean
suction-peaks (Kind, 1986).
For the conical vortex phenomenon, the important length scale is likely to be the
distance over which the boundary layer on the windward walls of the building grows. That
is the length from the wall stagnation-point, where a new boundary layer embedding
vorticity begins, to the roof comer or any point on the roof edge, where the boundary
layer separates and the wall boundary-layer vorticity rolls into vortices developing over
the roof. This wall boundary layer would be similar for those models that have similar
frontal walls and roof-comer features, regardless of the plan dimensions, if the models are
exposed to similar flows and reattachment takes place on the roof. That is, the wake
pressure cannot have any substantial upstream influence. Hence~ the wall boundary layer
22
development (WBLD) length appears to be the non-dimensionalizing length of choice (Lin
et al.~ 1995). Since, in similar flows~ the WBLD length is likely to be a fraction of the
building height (h), h itself is an acceptable substitute. However, the WBLD length is
expected to be affected by the approaching boundary-layer flow conditions, which would
influence the entrainment rate in the wall boundary layer and the separating shear layers
over the roof, and hence the reattachment distance on the roof and the pressure.
Therefore, h is not expected to be a global collapsing parameter if the approaching flow
condition changes. A fixed power law velocity profile of the fonn, VIVh= {z/h)a., inherently
implies self-similarity, and the source of the most important residual differences is likely to
be the variation of the incident turbulence characteristics eaves height or along the
stagnation streamline (Lin et al., 1995).
The magnitude of high suctions at the upwind roof comer, and equivalently the
strength of the vortex, changes significantly with the wind angle. However, the location
does not seem to change very much. Each high suction lobe shifts only slightly from b=20o
to b=l0 or symmetrically b=80 to /r-70, when wind angle varies from about 15 to 75,
where b is the angle of the lobe axis with the roof edge (Lin et al., 1995). Within the two
lobes near the roof comer, the worst wind angles were around 30 and 60
correspondingly. For other locations, however, the wind direction corresponding to the
worst suctions was location-specific. Peak pressure coefficients beyond -18 for rough flow
and mean pressure coefficients beyond -12 for smoother flow were observed, at x/h=0.026
and ylh=<l0065 (or at a ray of an angle /r-14 at s/h=0.0268). The roof-level turbulence
intensity was 17%. It was unknown whether the suctions would continue to increase
toward the roof corner (Lin et al, 1995). However, Kind ( 1986) pointed out that the true
worst suctions on the upwind roof corner might have been missed in some wind-tunnel
simulations due to a lack of pressure taps sufficiently close to the roof edge. The suctions
seemed to increase monotonically toward the apex of a flat roof, as close to the apex as
installation of pressure taps allows (Kind, 1986; Jamieson and Carpenter, 1993). The true
worst suction remains to be identified in full-scale studies.
23
In an effon to better understand the conical vortex phenomenon and the pressures
developed on the outside surface of the roof~ the works of Kawaii and Nishimura ( 1996),
Marwood (1996} and Banks et al (1997) are representative. Kawaii and Nishimura (1996)
took detailed measurements of the incident wind and the pressure on the square roof of a
model (120 mm x 120 mm x 60 mm). They calculated the correlation between a reference
tap and the rest of the taps, corning up with a theory to separate the pressure fluctuations
into a low-frequency pan and a high-frequency part caused by different mechanisms.
Marwood (1996) measured the detailed flow field inside the conical vonex on a model
using a Laser Doppler Anemometer (LDA). By developing a simultaneous pressure-
velocity measuring technique and a conditional statistical technique, he could relate the
surface pressure to the velocity events within the vortices. Banks et aL (1997) performed
simultaneous flow visualization and pressure measurement on a 450 mm cuboidal modeL
Among their findings were: ( 1) the relationship between the wind angle and the angular
vortex-core location, (2) a linear change of the vonex-core displacement with the distance
from the roof apex, and (3) a good correlation between the pressure profiles and the
vortical circulation.

2.4 Summary
To conclude this chapter, a brief summary might be usefuL The main points discussed
in this chapter were as follows:
1. With its reasonable agreement with the full-scale measurements of building
pressures, wind-tunnel simulation will continue to be a valuable tool for research
and applications;
2. The turbulence intensity might be a more relevant factor than the integral scale for
wind-loading simulation on low-rise structures; the lateral turbulence seemed to be
equally, if not more, important than the longitudinal turbulence. Usually only the
longitudinal turbulence was reported, but the longitudinal turbulence and the
lateral turbulence are likely to be correlated, due to the inherent three-
24
dimensionality of the turbulence. More realistic simulation of wind~directional

fluctuation might prove essential to improved simulation of the surface-layer flow


over complex terrain. Natural wind is after all very non-stationary in the sense of
directional fluctuation;
3. Comparison of Cp values from different models should be made only between data
obtained under similar testing conditions. The appropriate characteristic length for
non-dimensionalizing the coordinates seems to be the building height for most low-
rise buildings;
4. The non-stationarity of the natural wind might pose a real difficulty for the wind-
flow simulation in the laboratory; to circumvent this, some mathematical
approaches might have to be resorted to. The scale effect of the wind-tunnel
simulation seems to be only minor and secondary in priority;
5. The suction on the leading roof-comer seemed to increase monotonically toward
the apex of the roof comer; the true worst suction remains to be identified in the
full-scale study;
6. The turbulence in the incident wind is responsible for high suctions associated with
flow instability in the separation bubble, but the exact mechanism is unknown;
7. The wind-tunnel simulation technique has room for improvement; enhanced
knowledge of the cause-effect relationship between the incident wind and the
pressures and pressure-generation mechanisms would eventually guide improved
simulation of the wind-loading effects on low-rise buildings.
CHAPTER3
WIND LOADS AND BUILDING PERFORMANCE

3.1 Introduction
An overall picture of the wind loading consists of three fundamental elements: the
climate., the near-surface wind., and the structure. Wind is generated at a considerable
distance above the earth's surface, in the first instance., by weather systems, which are
themselves initiated by pressure gradients resulting from differential heating of the
atmosphere. As it is with the flow of any fluid over a surface, a boundary layer, called the
atmospheric boundary layer (ABL), is formed over the earth's surface. All ground-based
structures are immersed in the ABL.
Structures are usually built to stand various environmental influences. The
performance of the existing buildings and other structures in severe winds is the only
realistic indicator of the success or failure of our design and construction practice against
wind loads. Buildings come in all sizes and shapes, and they are complex entities whose
performance would fare anywhere between a complete success and a total failure in the
event of strong winds. In general, the wind-loading effect on a building is governed by a
few factors: the wind characteristics, the terrain or topographical features, the building
geometry and its orientation with respect to the wind direction. Different parts of a
building envelope receive varying amount of wind loading, and the existence of adjacent
buildings might affect a building unfavorably. The wind damage a particular building
might sustain is the outcome of a competition between the loading effect and the
resistance capacity of the building components.

3.2. Atmospheric Boundary Layer


Wind is the motion of air with respect to the surface of the earth. While the sun is the
ultimate source of energy that drives the motion of the air., it is from the surface of the
eanh where the air directly draws the energy. The atmosphere is to a large extent

25
26

transparent to the solar radiation incident on the earth; but the earth is capable of
absorbing the solar energy. The heated surface emits energy in the fonn of terrestrial
radiation to warm up the air close to the ground. Differential wann-up of the air over the
water bodies and the landscapes gives rise to an uneven spatial temperature distribution
and creates pressure systems initiating the large-scale air motions. The earth~ s surface
poses a boundary for the atmosphere to draw momentum from and to move over.
The retarding frictional effect exerted by the surface of the earth on the moving air
slows down the air motion near the surface; this effect is diffused by the shear forces
(Reynolds stresses) and the turbulent mixing throughout the ABL. The depth of the ABL
ranges from a few hundred meters to several kilometers in the case of neutrally stratified
flows~ depending on the wind intensity, the terrain roughness and the angle of latitude.
Within the boundary layer, the wind speed increases with the elevation; it reaches the
gradient speed at the top of the boundary layer. Beyond the boundary layer, in the free
atmosphere where the surface effect has diminished, the wind flows approximately at the
gradient speed along the isobars (Simiu and Scanlan, 1996). The large-scale air motions
are significantly modified by an apparent force termed the Coriolis force caused by the
rotation of the earth. It is within the lowest portion of the ABL that most of human
being'"s engineering activities take place. A sketch of the ABL in the suburban areas is
shown in Figure 3. I.

Figure 3.1 A sketch of the ABL in suburban areas


27

Characteristics of the ABL are closely related to the terrain features over which the
air moves. The underlying terrain roughness governs the extension and profile of the near-
surface portion of the atmosphere. The terminology exposure category is used in ASCE 7-
95 to classify the roughness characteristics. In the classification, account is taken of the
variation in the surface roughness that arises from the natural topography and vegetation
as well as from the constructed features. The exposure in which a specific building or
other structure is sited is assessed as category A, B, CorD, with corresponding gradient
height of 1500, 1200, 900 or 700 ft, respectively, representing the urban terrain, the
suburban terrain, the flat open terrain and the open waters. In addition, an exponent is
assigned to each category to facilitate a convenient power-law expression of the mean
wind-speed profile of the ABL.

3.3 Wind and Turbulence


Wind constantly changes its magnitude and direction as it blows. Meteorologists
divide the motions of the atmosphere into three broad groups: macroscale, mesoscale, and
m.icroscale. The macroscale motions are those large-scale systems seen on weather maps.
The mesoscale motions occur in horizontal dimensions of 10-500 kilometers, including
valley winds and squall lines. The term microscale refers to all those motions that are
smaller still, or the turbulent part of the wind flow. In the m.icroscale motions, the three
velocity components are of the same order of magnitude, whereas in the larger scales the
vertical motions are much smaller than their horizontal counterparts (Panofsky and
Dutton, 1984).
The widely quoted wind-speed spectrum established by Issac Vander Hoven (1957)
is reproduced in Figure 3.2 to facilitate an easy presentation. The graph was so created
that the variance contributed by a frequency range is equal to the area under that portion
of the curve; the ordinate has a dimension of m2/sec 2 and the abscissa is in cycles/hour.
There are two predominant peaks in the spectrum; one occurs at a period of four days,
and the other at one minute. The former peak (synoptic peak) is due to wind-speed
28
fluctuations caused by migratory pressure systems of synoptic weather map; the latter one
(turbulence peak) corresponds to microscale motions, ie., the turbulence of mechanical
and convective origin.

7
6
.~
en Structure-generated j
c 5
-8 turbulence
g
8.
II)
4
3
12 hours 1 hour 1 min t I<E-<___,\.____--310>.

b
~ 2
cf
l
0
0.001 0.01 0.1 l 10 100 1000
Frequency, cycle/hour

Figure 3.2 Horizontal wind-speed spectrum at about 100-m height


(reproduced after Issac Vander Hoven, 1957)

The synoptic peak is produced by instabilities associated with large-scale horizontal


gradients of the wind and temperature (horizontal scales - l()(X) km) whereas the
turbulence peak is due to instabilities arising from vertical gradients of the wind and
temperature (scales of several hundred meters).
Between the peaks in the spectrum, there exists a notable broad spectral gap centered
at a frequency ranging from one to ten cycles per hour. Vander Hoven (1957) reasoned
that there was a lack of physical process that could support the wind fluctuations in this
frequency range. This spectral gap renders it possible to separate the atmospheric motions
of different temporal and spatial scales in analysis. For surface observations, one-hour
averages are recommended to separate the larger-scale from the micro-scale processes;
shorter periods are used for engineering applications (Panofsky and Dutton, 1984). It is
the small-scale fluctuations of the wind speed and direction that are of significance in
building aerodynamics. Records of 15-minute duration seem to contain all the significant
frequency contents in the microscale.
29
Figure 3.3 represents a one-minute duration trace of the wind speed variation~ along
with that of the wind direction fluctuation. documented by the WERFL of Texas Tech
University. The wind speed at each instance can be considered as the superposition of a
fluctuating component (turbulence) on an average value calculated for the 15-minute time
interval This fluctuating part has significant aerodynamic consequences., but no detailed
discussion will be attempted here.

Wind speed series, M49n072


15.0

10.0
.!!
E
5.0
run mean= 7.58 m/s
0.0
0 10 20 30 40 50 Seconds 60

60.0
Wind direction series M49n072
~------~~~~==~====~~~~------------~

40.0
tn
1 20.0
g 0.0 ~....1'-1.....---v-.::.w-+-t-~~~~~~~~'Vl+...___-+-=1-___---'I:..L-~__,.,,......
20.0
-40.0 "------------------------~
0 10 20 30 40 50seconds60

Figure 3.3 Time series of wind speed and direction (WERFL, T1U)

3.4 Wind Effects on Buildings


The loading effects of the natural wind on low-rise buildings are a rather complicated
interactive process between the wind flow and the various components of the building,
Figure 3.4. For a building designed and constructed with engineered details to resist the
wind loading, the local forces received by different parts of the envelopes are transferred
to the main wind forces resistance system (MWFRS) and ultimately to the foundation.
30

In this section.. some damage patterns of buildings are first established, followed by a
description of the wind-building interactive process; finally, the issue of engineering
attention and performance improvement is discussed. The materials covered in this section
closely follow the works of Minor (1979, 1984) and Mehta (1984).

Figure 3.4 Building components affected by wind (after Minor, 1979)

3.4.1 Observations of Wind-induced Damage


Following the 1970 Lubbock Tornado, which devastated the city of Lubbock in the
state of Texas .. extensive post-disaster damage documentation was conducted by the
Institute for Disaster Research at Texas Tech University. The accumulated database
provided insightful information about the performance of buildings in severe winds. Some
patterns were identified after thorough damage assessments (Minor, 1979):
1. Structures failed.. principally, because of the wind-induced forces acting on the
critical building components;
2. Non-engineered or marginally engineered structures were susceptible to failures at
low wind speeds;
31
3. Small increases in the degree of engineering attention could produce large
differences in perfonnance;
4. Building geometry and building orientation were important factors of the wind
resistance; and
5. Certain types of structures, e.g. mobile homes, frame housing, unreinforced
masonry buildings, were weak in their wind resistance.
Mehta (1984) attributed the wind-induced building damage to one or more of the
following reasons:
1. Speeds significantly higher than the design wind speed, causing loads in excess of
some factor-of-safety used in design;
2. Deficiency in design or construction, creating weak links in the structural system;
3. Possible deficiency in standards or codes, leading to inadequate designs; and
4. Lack of understanding of the wind-loading effects, resulting in under-design.
It was further pointed out that building damage in general was due to a combination
of some causes mentioned above, and if similar damage was found in different storms.
there was a good reason to believe that the damage was due to a combination of the last
two causes. Weak links would precipitate progressive failures., and the severity of the
damage was dependent on the structural redundancy; lack of redundancy incurred severe
progressive failure once the weak links were broken.
After careful examination of the wind-induced failure of various building types with
different degrees of engineering attention (Mehta, 1984), it was found that the structural
frames of high-rise buildings, which received significant engineering attention in design
and construction, generally performed adequately in windstorms. Most damage to them
was with the cladding. In most cases, breakage of window glass by debris impact,
combined with the action of internal pressure., caused a breach of building closures and
subsequent damage to the interior. Typical pre-engineered low-rise metal industrial and
commercial buildings were designed with virtually no redundancy, failure of a weak link
could trigger extensive damage to and even collapse of the structural frame. Residential
32
buildings received relatively little engineering attention in their design; their construction
nonnally followed the construction prescription of building codes. Initiation failure of
residential units was quite often pinpointed to those locations sustaining high wind
pressures, such as the eaves.. the comers of the roof and walls. Wind-induced damage to
residential buildings was extensive; failure of a weak link frequently lead to progressive
damage. Appropriate attention to component design and connection details could
significantly improve their performance and reduce the property losses (Mehta, 1984).

3.4.2 Air-flow and Building Failure


When wind approaches a building., its passage is obstructed and the airflow is forced
to go either around or over the building (see Figure 3.5). In doing so., the airflow detours
along a longer path and is accelerated at the leading edges of the building, such as the
eaves., the ridges, the wall comers and the gable tops., forcing flow separation and vortex
formation to produce outward-acting forces on these components.

Figure 3.5 Overall failure-mode of a building (after Minor, 1979)


33
Usually the windward wall is the only part of the building enclosure to receive
inward-acting forces. The overall effect of the wind loading is to apparently explode the
building., as depicted in Figure 3.5.
The local effects of the external wind pressures on the building envelope are closely
related to the geometry and the orientation of the building with respect to the incident
wind. Different components sustain the most critical loading at different wind angles
(Figure 3.6).

WALL CORNER EAVE

RI)GE ROOF C~NER

Figure 3.6 Local airflow and loading effects (after Minor., 1979)
34
With flat or gently-sloped roofs" the normal-to-ridge wind generates high suctions at
the windward eaves and the ridge. Oblique wind direction usually produces worst-case
uplifting forces at the leading roof comer. Parallel-to-ridge wind causes significant loads
on the roof verge and the gable wall. For most wind directions., generally two of the four
wall-comers are under severe loading effects (see Figure 3.6).
Other than the external pressures acting on the outside surfaces of a building
enclosure, the internal pressure is in most situations different from the static pressure, due
to air penneability present by either design or component failure. The internal pressure
may aggravate the loading effects on the building enclosure. Windward wall openings
usually cause an increase in the pressure within the building. This pressure increase
combines with the outward-acting pressures already acting on the roof, the leeward wall..
and the sidewalls to intensify the forces on these components. This is demonstrated in
Figure 3.7.

Figure 3.7 Openings and internal pressure (after Minor, 1979)

Conversely. openings in the sidewalls or the leeward wall cause a decrease in the
internal pressure. Such a pressure drop collaborates with the inward-acting pressures on
the windward wall However.. the decrease in pressure partly counteracts the outward-
35
acting forces on the external surfaces of the roof.. and on the side and leeward walls to
reduce the effects of pulling apart the building envelope (Minor7 1981 ). Local component
failure could lead to a total disaster when the building loses its integrity (see Figure 3. 7).
The presence of adjacent buildings of comparable height might magnify the suctions
on the side walls7 but pressures on the windward wally on the roof and leeward wall are
basically not changed by the proximity of other buildings (Sachs .. 1978). Building damage
by the impact of flying debris in severe storms is also an issue of serious concern.

3.4.3 Engineering Attention and Building Performance


The performance of structures in windstorms is closely related to the amount of
engineering attention they receive in the design and the construction phases. Minor ( 1984)
classified buildings into four categories, according to the engineering attention they were
given: fully engineered., pre-engineered, marginally engineered, and nonengineered. In that
order., the susceptibility to wind damage increases. In the event of the 1970 Lubbock
Tornado, at least two and possibly three tornadoes rampaged across the center of
Lubbock; one-quarter of the city was impacted, and the affected buildings included
commercial, industrial and residential buildings. Damage surveys showed that the principal
patterns of failure were loss of complete roofs or roof cladding, failure of masonry, loss of
cladding and glazing by the direct wind action or by windbome debris, and collapse of
steel-framed industrial buildings.
Fully engineered buildings are individually designed; considerable attention is usually
paid in the construction phase to ensure quality7 and they usually fare very well in high
winds. It is now a standard procedure in many countries to conduct wind-tunnel testing of
these types of buildings to determine the wind loads for design purpose. Although a 20-
story steel-framed building in the extensively damaged downtown area of Lubbock was
taken past yield point., it did not collapse (Figure 3.8). A fully-engineered hospital survived
the direct impact of the Omaha Tornado in Nebraska on May 6, 1975 (see Figure 3.9).
36

Figure 3.8 A high-rise stood after the Lubbock Tornado ( 1970)


(Documented by the IDR, Texas Tech University)

Figure 3.9 A fully-engineered hospital survived a tornado (Minor, 1979)


37
Pre-engineered buildings are planned as a group before construction and marketed as
individual units. These fabricated buildings are often so designed that all components are
'equally' strong, thus achieving optimum economy in construction. However, their
performance in windstorms was not satisfactory. The problem spots were the overhead
doors, the claddings on wall and roof comers, the eaves and ridges. Failure could occur at
relatively low windspeeds, if care was not exercised in design and construction (Minor,
1979).
Post-disaster surveys seem to indicate that damage to marginally engineered buildings
represents the largest single contribution to losses. Commercial buildings, light-industrial
buildings and some apartments built with combinations of masonry, light-steel framing,
wood framing and concrete constitute this group. More often than not, poor performance
of these buildings in windstonns was preconditioned by the way they were designed or
constructed. The same materials could have been used more effectively in engineered
designs. Buildings containing masonry were particularly susceptible to wind damage
(Minor, 1979).

Figure 3.10 Damage to the roof of a marginally-engineered building by a tornado


(Omaha, Nebraska, May 6, 1975, as documented by the IDR, Texas Tech University)
38
Many single- and multiple-family residential buildings and small commercial buildings
hardly receive engineering attention. Largely of wood-frame construction., these buildings
offer little resistance to lateral loading., to uplifting pressures acting on the roof and roof to
wall and wall to foundation connections are usually weak. They would be damaged or lose
overall structural integrity (Figure 3.10} at wind speeds of75- 125 mph (Minor., 1984).

3.4.4 Improvement of Building Practice


From the view-point of technology, improving the wind resistance capacity of
buildings and other structures, and thus mitigating propeny loss by severe windstorms
involves two aspects: (l) understanding the wind effect on buildings., and (2) transferring
and implementing technical advances into engineering practices.
The wind resistance of a building can be improved by proper provision of anchorage
and connections. In addition.. the design of components such as window glass, metal siding
and purlins can be accomplished so as to resist severe local pressures and missiles impact,
thus preventing breaching of the building enclosure (Minor, 1979). Damage to the building
frames is generally due to the laterally-acting wind forces, while failures of components
and cladding are caused by uplifting or outward-acting forces. Components and cladding
fail either because of high wind pressures or because of inadequate fastenings and
connections. Improvement in design practice and construction details for uplifting and
outward-acting wind pressures, especially in those regions of the building envelope
experiencing severe loading't should mitigate these failures. Standards of practice for wind
resistance need to address the possibility of failures of weak links in high winds to develop
procedures for mitigating progressive damage (Mehta, 1984). Airborne debris impact is a
problem of serious concern in the event of severe windstorms., and the design of glass
windows and claddings should take this into consideration. The construction quality of
residential units needs to be improved.
The design and construction practice is guided by standards and codes. Despite all the
advances in our understanding of the wind loading effects, and the progress made in wind-
39
loading codification, the past few years have seen wind damage of staggering amounts in
the coastal regions of the United Sates. The extensive property damage testifies to our
inability as a society to effectively mitigate the economic losses inflicted by severe
windstorms. The problem is likely to be two-fold:
1. Codes were not effectively enforced in the design and construction practice; and
2. There was a lack of clear understanding of the wind induced forces on buildings
and other structures~ leading to underdesign of buildings., except in a few cases
when higher than the design wind speeds were involved.
Finally, construction practice involves intricate relationships among designers,
contractors., codes, insurers and the general public. Awareness and preparedness of our
society as a whole is the key to effective mitigation and fast recovery.
CHAPTER4
THE WERFL FACILITY

4.1 Introduction
It is perhaps no overstatement that wind engineering at Texas Tech University staned
with the 1970 Lubbock Tornado. In the decade after the tornado, about 50 extreme
windstorms in the United States were documented by the researchers from Texas Tech.
During the same period, other investigators representing the civil defense interests, the
architecture and engineering protessions, the nuclear industry, and the academia were also
involved in the conduct of technical post-storm damage surveys (Minor, 1979). The
decade of 1980, prior to the establishment of the Wind Engineering Research Field
Laboratory (WERFL) at Texas Tech University, saw extensive research in the area of
wind-induced loads on low-rise buildings, primarily using the wind-tunnel testing
technique. The resul[S of the research have had significant impact on wind-loading
codification. Nevertheless.. problems remained, and there was a need to fully validate the
model-scale results for low-rise buildings. A few full-scale experiments and some
comparative studies revealed problems and difficulties. Both the wind-tunnel testing
technique and the data quality of the full-scale measurements were matters of concern
(Surry, 1979)
The WERFL located on the campus of Texas Tech University in Lubbock, Texas is a
unique facility for full-scale studies of wind loading effects on low-rise buildings. Since its
initial operation in 1988, the WERFL has represented one of the best-instrumented full-
scale installations to provide data of the highest quality for comparison with and
verification of model-scale experiments. Located in Lubbock, a city on the West Texas
High Plains, the site of the WERFL experiences persistent wind throughout the year. The
prevailing winds are from the south and southwest. Sustained winds of 20-35 mph are
frequent occurrences in the windy seasons of spring and late fall. Figure 4.1 shows trees in
front of the Texas Tech Museum (about one mile east of the field site) shaped mainly by

40
41

unrelenting winds in the springs. The generally flat and open landscape presents an ideal
terrain characteristics for the wind-tunnel simulatio~ by minimizing the number of
uncertainties to make the simulation of wind flow a less daunting task. All these factors
make the site an ideal location for a full-scale wind engineering research facility.

Figure 4.1 Trees shaped by winds

4.2 Test Building and Meteorological Tower


Detailed description of the facility can be found in Levitan and Mehta (1992a,
1992b). Some essential information is provided here for reference. The WERFL facility
consists of a rotatable metal building (30x45x13 ft), a 160 ft high meteorological tower
and a small concrete block room housing a data acquisition system. The surrounding
terrain of the experimental building is flat and open; the circular area of radius 1.2 miles
centering at the building contains land features such as cultivated field, small shallow playa
lakes, residential houses and high buildings; but the immediate surrounding is open grass
42
land. A photograph showing the test building and the meteorological tower is given in
Figure 4.2.

Figure 4.2 Test building and meteorological tower

The test building is a prefabricated rectangular metal building with a nearly flat
duopitch roof. It is constructed on a rigid undercarriage that rides on a circular steel track
embedded in a concrete slab. When jacked up and supported by the four wheels installed
at the comers of the undercarriage, the building can be rotated to the desired orientation
with respect to approaching wind direction. Temporary anchorage of the building is
provided by bolts embedded at 15 intervals in the concrete foundation slab. A schematic
indicating the dimensions of the test building is shown in Figure 4.3.
Also shown in Figure 4.3 are the definitions of the wind azimuth (a), the building
position (~) and the angle of attack (9). The azimuth angle is measured clockwise from the
true north to the direction of the wind. The building position angle is the angle measured
43
clockwise from the true north to the assigned building north., i.e. ., the outward pointing
nonnal of the wall with the door. The angle of attack is a relative wind direction with
respect to the building orientation; it is measured clockwise from the building north to the
wind direction, i.e. e =a - p, if a > p, or a=360 + a - p, if a < p.

(a)

_L 0.25'

Il2.83'
30.25'

(b)

a- wind azimuth
P- building position
9- angle of attack

Figure 4.3 Schematics of test building and definitions of position and angle
(a) Dimensions of test building, (b) Building position and wind angle

The meteorological tower is a guyed latticed metal tower. It is located about 150 ft
(49 m) to the west (280 azimuth) of the test building, so that it is upstream of the test
44
building at the prevailing wind directions from the south and southwest. The anemometers
are installed on the tower in such a way as to minimize the interference of the wind flow
by the tower.. Instrumentation is selected at six levels on the tower: 3, 8, 13, 33, 70 and
160ft (l, 2.5, 4, 10., 21., and 49 m) above the ground.

4.3 Anemometers and Pressure Measurement System


The two most important parameters in the documentation of wind-load effects on a
building are the pressures acting on the building envelope and the wind inducing the
pressures. Two types of anemometers are used at the WERFL for measuring the wind
speeds and the wind directions at different elevations on the meteorological tower.
Transducers are employed to measure pressures at selected locations on the external
surfaces and internal pressure inside the building. Other parameters such as the air
temperature, the barometric pressure and the relative humidity are also monitored on a
regular basis. The temperature sensor, the barometric pressure and the relative humidity
sensor are mounted at 13ft on the tower.

4.3.1 Anemometers
The anemometer placement on the meteorological tower for Mode-49, a data
acquisition mode incorporating a 3-component sonic anemometer created specifically for
this study, is as follows. One 3-cup anemometer is mounted at 3 ft; five UVW
anemometers of propeller type are installed at 8, 13, 33, 70, and 160ft levels. Two 3-cup
anemometers are mounted, one at 13 ft and the other at 19 ft, on two pules located
between the test building and the meteorological tower to provide additional wind speed
measurements.

4.3.2 Pressure Measurement System


The pressure measurement system is the same as that used by Thomas (1996), and
the following description of the system is from his work. Two types of pressure
45
transducers are employed for external and internal pressure measuremenL They are the
Omega and the Validyne transducers. Currently 17 of Validyne Engineering Model
DP103-22N&-S-4-H differential pressure transducers are in use. These transducers have
a full-scale {FS) range of 1.38 kPa (0.20 psi) or 2.21 kPa (0.32 psi)~ with rated accuracy,
including linearity, hysteresis and repeatability, of 0.25% FS. Differential pressure on the
two sides of a diaphragm causes a displacement to initiate a change in the magnetic field,
which is converted to a voltage output linearly proportional to the applied pressure
difference.
Thirty transducers of model PX 163-005 BD 5V (Omega Engineering, Inc.) are also
in use. The Omega transducers have a range of 1.24 kPa (0.18 psi), a rated linearity of
0.50% FS, and hysteresis and repeatability of 0.25% FS. Unlike the Validyne transducers.,
the Omega transducers convert the strain in the diaphragm to an analog voltage output
proportional to the pressure applied on the membrane.
Signals from the transducers are low-pass filtered. Each transducer, along with a
three-way solenoid valve, is mounted on a small wooden board fixed on the inside surface
of the building. Undisturbed ambient atmospheric pressure from an underground box 75 ft
to the west of the building is used as a reference pressure. The reference pressure is
transmitted to the data acquisition room by means of an 8-inch diameter pipe. The surface
pressure from a tap and the reference pressure are introduced to a transducer through
flexible plastic tubing. The frequency response of the pressure transmission and
measurement system was investigated by Sandri (1992) and Letchford et aL (1992). The
system was found satisfactory for the frequency content involved in the full-scale study.
The system conveys fluctuating pressures having frequencies up to 20 Hz with only
minimal distortion and small phase lag. Although the signals from the transducers are low-
pass filtered at 8Hz and 10Hz for the Omega and the Validyne transducers., respectively,
the system effectively preserves the fast-fluctuating signals of instantaneous peak
pressures. For a sampling frequency of 30 to 40 Hz at a mean wind speed of 20 mph, the
46
attenuation of the peak and the rms pressure was reported to be in the range of 0-8% and
0-3%, respectively (Sandri, 1992).

4.4 Data Acguisition System and Data Validation


A data acquisition system converts conditioned electronic signals from different
instruments to digital information which is sorted and saved by the system (see Figure
4.4). The digital data are later processed to give quantities having specific physical
interpretation. The task of data validation is twofold: ensuring proper operation of the
instruments and getting the correct information about the measured physical quantities.

Sensors .
I / Multiplexor boards

AID

Figure 4.4 Data acquisition system

4.4.1 Data Acquisition System


The concrete block room inside the test building houses the data acquisition system.
The system consists of four multiplexor boards., analog-to-digital (AID) conversion boards
and a host computer. Electronic signals from the outputs of all the instruments were
directed to the four multiplexor boards capable of accommodating 4 x 24 =96 channels.
These analog signals are converted to digital information through the software driven AID
boards before they are stored on the hard drive.
The LabTech Notebook software from the Laboratory Technologies Corporation
running in the Windows environment is used to drive the AID board. This software runs
within a custom shell written in Microsoft Quick.BASIC. The data acquisition system
operates continuously. By monitoring the wind speed at 13 ft on the tower, the system can
47
be triggered automatically when a preset one-minute mean speed is exceeded. Once
triggered., the computer first perfonns a short (20 seconds) pre-test calibration run for
transducer zero drift., then turns on the solenoid valves on and off, followed by a typically
15-minute data run. A post-test zero calibration run is performed and the triggering
program is restarted.
Raw digital data are saved in different files storing different types of information such
as the pressures and the wind speeds. Calibration information of the zero readings is saved
in other files. In each file, data are stored in columns., each corresponding to a specific
instrument unit of the same type.

4.4.2 Data Analysis and Validation


Data obtained in the full-scale investigations serve as a database against which results
from simulation technique are verified and calibrated. One of the most important tasks of
full-scale studies like that of the WERFL is data quality assurance. Routine daily checks of
the laboratory., weekly instrument calibration and maintenance, and three-stage validation
of collected data provide a quality and reliable database to the research community.
Collected data are downloaded from the hard drive of the host computer and
transported back to the campus for processing. Timely analysis of the acquired data helps
locate problems with the instrumentation and the data acquisition system. For processing
the raw data, a preprocessor software converts the raw data into equivalent voltages and
then into engineering units. Data from the pressure transducers are automatically corrected
using the calibration data taken before and after each data run. The processor also
computes summary statistics, plots time series for each instrument to yield a 10-page
printout for each data run. Finally, a three-stage validation is carried out by different
individuals., and the data runs that pass all the three validations are archived to become
part of a permanent database.
CHAPTERS
DESIGN OF EXPERIMENTS

5.1 Introduction
The wind-tunnel simulation technique has long been the main means of determining
wind-loading effects on structures. However, researchers were always aware of the
limitations of model test results for prototype applications. After full-scale wind pressure
data became available from field measurements conducted in laboratories like the
WERFL at Texas Tech University, comparative studies ensued, and discrepancies
between the wind-tunnel predictions and the full-scale data were soon noted. There was
quite a debate on the simulation priorities., and this is the setting of the current study.
In the course of time, exploration of flow mechanisms associated with the peak
pressures took place. Trying to explain the causes of high suctions related to the
separation bubble (SB) flow situation, Melbourne (1977) hypothesized that the high
suctions were due to an instability of the bubble triggered by the incident turbulence.
Research by Letchford (1995), Kawaii and Nishimura {1996), Marwood (1996) and
Banks et al. (1997) on the conical vortices is representative of the renewed effort. Their
works went one step further by looking into the details of the incident wind, the near-
surface wind flow and the pressure on scaled models tested in the wind tunneL

5.2 Incident Wind and Pressures


It was strongly perceived by the author that there might exist some clear cause-effect
relationship between the incident wind characteristics and the pressures on the roof, for
both the SB and the conical vortices. The location for measuring the incident wind should
preferably be close enough to the test building so that an anemometer can measure the
wind as it hits the building. Wind characteristics changes in the course of traveling, the
selected location should have minimum undesirable distorting effect form the building.
Synchronized measurements of the incident wind and the pressures could be insightful.

48
49
Wind measured in the far field" on the meteorological tower might be too dissimilar to
the near-field' wind to permit synchronization with the pressures.

5.3 Wind-flow Measurement


For the SB phenomenon., measurement of the flow inside the bubble was designed to
document the mean structure of the SB., to establish a relationship between the flow-field
and the pressures on the underlying roof. It was also designed to understand some cause-
effect relationships between the incident wind and the pressure fluctuations on the roof
near the leading edge., especially at those instants when high suctions are initiated.
Clarifications of this nature could enhance our knowledge of the pressure generation
mechanisms, especially the role of the incident wind., which is important from a practical
viewpoint. The same methodology is feasible for the case of the conical vortices.

Figure 5.1 Three-component sonic anemometer


50
All measurements of the incident wind flow and the measurements of the flow field
inside the SB were accomplished by using a three-component ultrasonic anemometer.
Generally., sonic anemometers have much faster frequency response than propeller-type
anemometers. The sonic anemometer used in this study is an omni-directional (R3)
anemometer (Model 1210-K-063., Gill Instruments); a photo of the anemometer is given
in Figure 5.1. The detailed technical specification is given in Appendix A
A data acquisition mode., Mode-49., was created to incorporate the signals (three
channels) from the sonic anemometer into the existing data acquisition system. The sonic
anemometer was mounted on a portable support., and could be manually moved to the
desired positions (Figure 5.2). The support was sturdy and undesirable vibrations of the
sonic anemometer were practically negligible.

Figure 5.2 Set-up of sonic anemometer for wind flow measurement on roof
51
5.4 Flow Visualization
As an important element in the experimental research, flow visualization has played
a broad role in improving our physical understanding of complicated flow phenomena.
Effort was made in this study to seek feasible means to visualize the wind-flows over the
roof of the WERFL test building. Three techniques were explored: tuft-grid, smoke
injection and airfoil-grid, each with its own advantages and shortcomings. The tuft-grid
technique was used for most of the flow visualization. These methods were inexpensive,
suitable for full-scale applications and yet effective to provide insightful information
about the wind flow of interest.
The tuft-grid method has been used for flow visualization in wind tunnel tests. A 6.4
m (L) x 2.1 m (H) (21 ft x 7 ft) metallic frame with a square-grid system of 15.2 em (6
in.) was built. Bright-orange colored yarn segments were tied to the grids. The yarns
align with the local wind flow to reveal the location, the overall shape and the dimensions
of the SB or the conical vortices. The grid consists of three detachable panels, each has
dimensions of 7 ft x 7 ft. Figure 5.3 shows the set-up of a grid of 14 ft x 7 ft on the roof
of the test building.

Figure 5.3 Set-up of grid on roof of test building


52
The smoke injection technique~ applied to the flow visualization in the full-scale,
needs a high volume discharge of smoke. An F-100 Performance Smoke Generator was
used (Lightwave Research, Austin, TX). The shortcoming with the generator was its
incapability to supply the desired amount of smoke for a sufficiently long duration. The
smoke generator was positioned at about 15 em below the roof level and right ahead of
the building. Smoke bombs were fiXed onto selected locations of the grid to enhance the
visualization. The grid also served as a background with a scale indicating the dimensions
of the vortex circulation.
The airfoil-grid technique used specifically designed lightweight airfoils. To make
an airfoil, a sheet of Balsa wood 80-lb paper was glued to the perimeter of two parallel
airfoil-shaped wooden pieces 10 em (4 in.) apart The airfoils have a chord length of 16.5
em (4 in.), a thickness of 1.9 em (3/4 in.). Wind-tunnel tests indicated that the airfoils
aligned with the airflow at speeds between 4.5 m/s and 13.4 mls. Automotive reflective
tape was pasted on the sides of the airfoils for better nighttime visualization. The setback
with the airfoil technique was the erroneous alignment of the airfoils at low wind speeds,
as was the case for the wind flow near the core of a vortex, where the speeds of air
motion were low.

5.5 Pressure Measurement


The pressure measurement in this study was synchronized with the measurement of
the incident wind measurement and/or with the flow visualization. Pressure taps of
interest in the study of the SB are 50223, 50523, 50923, 51423, 52323 and 52923. These
are taps along the short roof-axis and were active in the Mode-49 data acquisition
configuration. Information about their locations and the mean pressure coefficients at
incident winds of around 270 will be presented in Chapter 6, when the SB and the
pressures are discussed. The synchronized measurement of the pressures along the short-
axis of the building roof was designed to relate the pressures to the wind flow structure of
the SB. It was also designed to link the properties of the incident wind to that of the
pressures. The pressures were sampled at 30Hz, the same sampling frequency as for the
53
wind flow near the roof. This frequency coincides with the speed of the video camera (30
frames in a second).
In the situation of the conical vortices over the leading roof-comer, the interested
pressure taps available for Mode-49 are 50101, 50501, 50901, 50205, 50505, 50905,
50209, 50509 and 50909. Detailed information about these taps will be given in Chapter
7. Tap 50501 is the most studied by the wind engineering community. The full-scale
wind pressure data for this tap are the most frequently compared with by researchers from
all around the world. Tap 50501 experienced the worst-case uplifting pressures of all the
taps installed on the WERFL testing building, and the true worst-case location on the
leading roof-corner should be somewhere closer to the edge and the apex. Again, as is
with the SB, the ultimate objective of the designed experiment for the conical vortices
was to clarify the mechanisms of pressure generation, and with that achieved, to provide
some guidance for the wind-tunnel simulation of wind loads on low-rise buildings.

5.6 Comparison of Velocity Measured with Different Anemometers


As discussed earlier, the consensus for improving wind load prediction was on the
simulation of incident wind. In the study of wind loading effects on structures, the
documentation of incident wind characteristics is as imponant as that of the pressures.
Propeller-type anemometers have been used at the WERFL. These anemometers have the
shortcoming of not being capable to keep up with the high rate of velocity fluctuation of
the natural wind. Conventional wind simulation technique is inadequate to produce large
and fast directional fluctuations and non-stationary gusts present in the natural wind.
Thomas (1996) compared the spectra of the horizontal wind velocity components of
the simulated data (1:100 Rll) from the Meteorological Wind Tunnel at the Colorado
State University, and the full-scale data measured with different types of anemometers by
the WERFL of the Texas Tech University. In the full-scale measurements, propeller-type
UVW anemometers, 3-Cup/Vane anemometers and two-component sonic anemometers
were employed. The full-scale wind velocity data were taken at different time frames.
and he looked at the statistics of wind velocity.
54

Figure 5.4 Set-up of velocity comparison experiment

In the current study, a direct time series comparison of the incident wind velocities is
sought. The velocity measurements were simultaneous, with closely spaced anemometers
of different types. Specifically, the comparisons are between measurements by the three-
component sonic anemometer and the UVW anemometer at 13ft leveL The experimental
set-up is shown in Figure 5.4. The two anemometers are about 6.5 ft apan in the east-
west direction.
CHAPTER6
SEPARATION BUBBLE AND ITS
EFFECfS ON PRESSURE

6.1 Introduction
Field surveys of building damage inflicted by severe winds indicate that roof eaves
are one of the most vulnerable portions of the cladding. When wind approaches a
building normal to one of its walls~ the obstructed flow is forced to either go over the
building, or to descend along the wall and detour around the wall comers. A conceptual
model of the flow around a cubic obstacle was given by Woo et al. (1977). The up-wash
flow cannot negotiate the sharp eaves to remain attached because of the viscosity of the
air; it separates to form a shear-layer, which may reattach to the roof surface to generate a
vortex called the separation bubble (SB). The SB and the pressures induced on the
underlying roof surface are worth investigating, from the viewpoint of both academic and
practical imponance.
The WERFL at Texas Tech University has documented peak-pressure coefficients
up to -10 at a pressure tap (50123) located one foot from the leading roof edge. The
strongest suction ever documented for the building cladding was from tap 50501, located
on one of the four roof-comers. Tap 50501 recorded peak-pressure coefficients up to -13.
However, worst suctions for the eaves and the roof corners do not occur at the same or
even close wind angles. The SB is not a special case of the conical vortex.
In this chapter, a time-averaged structure of the SB is fust established from flow
measurements. Mean pressures developed on the roof underneath the SB are then
examined in relationship to the flow structure of the SB. A concept termed non-
conventional pressure coefficient is introduced to make possible the separation of the
effect of direction fluctuation from that of the speed variation on the pressure-generating
mechanism. Spectral analysis is performed to explore the characteristics of the pressures
associated with the SB.

55
56
6.2 Structure of Separation Bubble
Systematic measurement of wind flow velocity within the SB was conducted using
the 3-component sonic anemometer. Based on the measurement~ the mean structure of the
SB is established in the form of normalized velocity vectors.

6.2.1 Flow Measurement


With frequent flow reversal in all three directions, flow inside the SB is highly
turbulent. Ordinary hot-wire anemometers would not be applicable because of their
direction ambiguity. Sonic anemometers seem to be the most feasible option for full-scale
applications, because they can operate reliably in an adverse outdoor environment.
The WERFL test building has a dimension of 45x30xl3 ft. The flow measurement
was carried out in a vertical plane along the shon roof-axis (30 ft). One 3-component
sonic anemometer was used for the flow measurement. The sonic anemometer (Model
1210R3, Gill Instruments Ltd.) samples an air volume of about 52 in3 (a cylinder with a
diameter of 4 in. and height of 4.13 in.), enclosed by three pairs of probes. The flight
distance between each pair of probes is 5.75 inches. Adopted for full-scale usage, the
sonic anemometer has spatial resolution comparable with that of hot-wire probes used on
geometrically-scaled low-rise building models (typically 1:50- 1:100). The anemometer
measures axial velocities, which are converted into conventional velocity components.
Typical data runs are of 15-minute duration. The sampling frequency for the flow
measurement was 30Hz, which was the same as that for the pressure measurement. The
setup of the sonic anemometer on the roof of the test building was given in Figure 5.1.

6.2.2 Velocity Vectors


Data from more than one hundred runs were analyzed to yield information about the
magnitude and direction of the velocity vectors within the SB. A sketch of the SB created
from the velocity vectors is presented in Figure 6.1. Detailed statistics are listed in Table
6.1, where x is the distance of a measurement station from the leading edge, h is the
height from the roof, s and a are, respectively, the size and the inclination angle of a
vector.
57

-
\C)
58
Table 6.1 Statistics of velocity vectors
h=0'-06'' h=1 '-04''
Run# X ex s Run X ex s
9 0'-05" 139.8 - 145 0' 146.2 1.078
49 1' 166.5 0.087 146 1' 155 1.204
48 2' -5.8 0.146 147 2' 164.5 0.495
47 3' -4.5 0.187 148 3' 167.1 0.258
46 4' -1.4 0.200 150 4' 171.9 0.180
45 5' -0.5 0.210 151 5' 179.4 0.116
44 6' -5.9 0.143 152 6' -172.7 0.091
43 7' -6.6 0.119 153 7' -165 0.138
38 8' -8.4 0.099 154 8' -161.6 0.136
37 9' -63 0.030 155 9" -164.45 0.180
36 10" -89.5 0.018 156 10' -166.05 0.227
35 11' -153.3 0.038 197 11' -168.6 0.275
31 12' -171.6 0.108 199 12' -172.0 0.301
30 13' -174.8 0.148 200 13' -171.6 0.267
29 14' -176.8 0.203 201 14' -173.8 0.325
26 15' -175.9 0.214 202 15' -175.6 0.340
h=1 '-00" h=1 '-08,..
13 0'-05" 142.4 1.182 50 0' 144 1.057
14 1' 155.8 0.787 51 1 150.5 1.192
15 2' 154.7 0.125 54 2 157.6 0.902
16 3' -31.3 0.063 55 3 165.9 0.551
18 4' -11.5 0.052 58 4 169.3 0.330
19 5' -56.5 0.018 59 5 176.3 0.295
20 6' -96.7 0.029 60 6 -179.2 0.262
22 7' -96.6 0.030 63 7 -173.8 0.276
23 8' -118.9 0.045 64 8 -172.7 0.279
24 9' -143.8 0.077 65 9 -172.4 0.282
25 10' -158.2 0.119 66 10 -172.4 0.298
204 11, -162.8 0.110 67 11 -174.1 0.325
205 12' -168.5 0.155 68 12 -175.6 0.366
206 13' -170.0 0.183 69 13 -176.3 0.396
207 14' -175.2 0.277 70 14 -177.6 0.408
208 15' -175.4 0.258 71 15 -178.6 0.436
59
Table 6.1 Continued
h =2\t-03" h = 3'-00"
Run# X. a s Run X a s
72 0 152.4 1.139 105 0 158.6 0.962
73 0'-06" 151.42 1.234 106 0'-06" 160 1.080
74 1' 154.04 1.312 107 1' 160.6 1.080
81 2' 157.94 1.078 108 2' 162.4 1.127
82 3' 164.63 1.113 109 3' 165.9 1.052
84 4' 169.93 0.816 112 4' 171.8 1.204
88 5' 173.66 0.672 113 5' 175.3 1.094
92 6' 177.31 0.571 115 6' 178.6 1.017
94 7' -178.23 0.601 116 7' -178.8 0.916
95 8' -175.7 0.488 117 8' -176.1 0.833
97 9' -174.96 0.508 119 9' -174.7 0.769
98 10' -175.38 0.606 120 10' -172.6 0.741
99 11' -175.25 0.440 121 11' -172.5 0.633
100 12' -175.51 0.546 122 12' -171.9 0.645
101 13' -175.28 0.542 126 13' -173.7 0.703
103 14' -174.77 0.573 127 14' -172.9 0.712
104 15' -177.03 0.486 128 15' -174.1 0.709
h = 4'-02"
213 0' 163.63 1.038
215 1' 164.34 1.116
216 3' 167.87 1.074
217 5' 172.49 1.128
218 7' 176.71 1.159
219 9' 179.56 1.059

The velocity magnitudes (s) in Table 6.1 are normalized by the corresponding mean
wind speeds, measured at roof-height level on the meteorological tower. For plotting,
sizes of the velocity vectors are scaled by the vector with the largest magnitude, which is
assigned a convenient size.
A few observations can be made about the time-averaged SB:
1. The vortex is oblong and elongated in the horizontal dimension;
2. The mean reattachment point (RP), as identified by a nearly downward-pointing
velocity vector, is approximately 10 ft from the leading edge; the mean height of
the SB is about 4.5 ft.
60

3. The region with predominant flow reversal is confmed close to the roof surface. In
this region, time series records indicate frequent flow reversal and low mean
speed;
4. In the immediate proximity of the roof, there exist four distinct zones of wind
flow. They are the leading edge zone (ZONE-0, where the separated shear layer
meets the reverse flow to create a wedged region, the flow reversal zone (ZONE-
m, the reattachment zone (ZONE-Ill) and the forward flow zone (ZONE-IV);
5. Despite the conceivably random and turbulent nature of the flow inside the SB,
the averaged structure of the SB is amazingly organized and regular.

6.2.3 Boundary of Separation Bubble


The same data presented in Table 6.1 can be used to obtain the vertical proftles of
the u-component velocity, the component parallel to the short axis of the roof. Sarkar et
al (1997) used a parameter called intermittence factor (I.F.) to defme the boundary of the
SB. The intermittence factor was defined as the fraction of time (within certain time
period) the local flow goes in the direction opposite to that of the upstream wind. The LF.
value is also an indicator of flow turbulence and degree of flow reversal inside the SB.
The boundary of the SB was defmed by a curve formed by connecting those points
above which there was no flow reversaL The actual situation turned out to be more
involved. The difficulty is partly caused by the highly non-stationary nature of the
atmospheric wind. That is, occasional large wind-direction excursions from the
perpendicular direction might have added to uncertainty to the small LF. values. Beyond
the mean point of reattachment, the boundary of the SB is not well defmed. With the data
obtained, the boundary of no flow reversal extends almost the whole length of the roof.
Relaxation of the I.F. value by a few percentages dramatically transforms the overall
dimension of the separation bubble in the wind direction, see Table 6.2. In this table, x
andy designate the position of a measurement stations, with x measuring the horizontal
distance from the leading roof-edge, y the height from the roof surface. Therefore, instead
of giving a defmitive answer, the accurate boundary after the mean reattachment-point
61
(RP) is left open and represented by the broken lines in Figure 6.2. Wind tunnel tests
could possibly give a more precise defmition. More information about the data runs listed
in Table 6.2 can be found in Appendix B.

Figure 6.2 Boundary of separation bubble

6.3 Wind Flow Characteristics


As pointed out in Section 6.2.3, the turbulent flow inside the SB can be partially
characterized by the LF. values. Intermittence of wind flow represents alternating change
in the local wind direction. Specifically, the local flow may assume instantaneous
forward direction or backward direction. Here the forward flow moves in the direction of
the incident wind.
It can be seen that the flow inside the SB is highly intermittent and turbulent.
Considering the flow near the roof surface, flow close to the leading edge experiences
more frequent flow reversal, which persists beyond the averaged reattachment point.
With increasing height from the surface, the flow reversal becomes less frequent until a
point of no flow reversal is reached. That point was earlier defmed to be the upper
boundary of the SB.

6.4 Pressure Distribution


Pressure coefficient data (Mode-49) from six taps along the short roof-axis are
available for analysis. These taps are 50223, 50523, 50923, 51423, 52323 and 52923; the
data for tap 50123 are from Mode-15. The location and the mean pressure coefficient
(Cp) for these six taps and tap 50123 are provide in Table 6.3 and depicted in Figure 6.3.
62
Table 6.2 Intermittence factors
X=0'-00'' 0'-06''
X= X= 1'-00" X= 2'-00"
y Run# I.F. Run# I.F. Run# I. F. Run# I.F.
0'-06'' - - 009 0.001 049 0.514 048 0.812
1'-00'' - - 013 0.000 014 0.048 015 0.465
1'-04" 145 0.000 - - 146 0.003 147 0.136
1'-08" 050 0.000 - - 051 0.001 054 0.020
2' -03" 072 0.()00 073 0.000 074 0.000 081 0.003
3'-00" 105 0.000 106 0.000 107 0.000 108 0.000
4'-02" 213 0.000 - - 215 0.000 - -
X= 3'-00'' 4'-00"
X= X= 5'-00'' X= 6'-00''
y Run# I. F. Run# I.F. Run# LF. Run# I.F.
0'-06" 047 0.813 046 0.803 045 0.810 044 0.714
1'-00" 016 0.642 018 0.619 019 0.561 020 0.531
1'-04'' 148 0.301 150 0.374 151 0.432 152 0.458
1'-08"' 055 0.122 058 0.239 059 0.289 060 0.290
2' -03" 082 0.009 84 0.045 088 0.084 092 0.121
3' -00'' 109 0.000 112 0.002 113 0.006 115 0.017
4'-02" 216 0.000 - - 217 0.000 - -
X= 7' -00'' 8'-00"
X= X= 9'-00" X= 10'-00"
y Run# LF. Run# I. F. Run# I. F. Run# I.F.
0' -06" 043 0.680 038 0.662 037 0.550 036 0.538
1'-00" 022 0.534 023 0.506 024 0.463 025 0.398
1, -04" 153 0.407 154 0.399 155 0.366 156 0.316
1'-08" 063 0.279 064 0.275 065 0.263 066 0.242
2' -03" 094 0.107 095 0.123 097 0.127 098 0.111
3' -00" 116 0.024 117 0.038 119 0.051 120 0.056
4'-02" 218 0.002 - - 219 0.003 - -
x = 1e-oo" 12'-00"
X= X= 13'-0Q'' X= 14'-00"
y Run# I.F. Run# I. F. Run# I. F. Run# I. F.
0'-06'' 035 0.505 31 0.389 030 0.349 029 0.264
1'-00" 204 0.421 205 0.366 206 0.302 207 0.222
1'-04" 197 0.263 199 0.243 200 0.248 201 0.206
1'-08" 067 0.205 68 0.187 069 0.168 070 0.130
2'-03" 099 0.143 100 0.132 101 0.097 103 0.074
3' -00" 121 0.092 122 0.068 126 0.048 127 0.037
63
Table 6.2 Continued
X = 15'~-00" X= 17' -00'' X= 19'-00"
y Run# I.F. Run# I. F. Run# LF.
0' -06,.. 026 0.272 211 0.189 212 0.084
1'-00'' 208 0.225 209 0.135 210 0.091
1,-04" 202 0.163 - - - -
1'-08" 071 0.110 - - - -
2'-03'' 104 0.101 - - - -
3'-00" 128 0.037 - - - -
X= 11 '-00" X= 13' -00" X = 15' -00'' X= 17'-00''
y Run# I.F. Run# I. F. Run# LF. Run# I.F.
4'-11" 261 0.()04 262 0.003 263 0.018 264 0.007
X= 19'-00" X= 21'-00" X = 23 '-00' ' X= 25'-00''
y Run# LF. Run# I. F. Run# LF. Run# I.F.
0'-06" - - 274 0.068 275 0.057 276 0.012
1'-04" 283 0.058 - - 282 0.027 - -
3' -08" 284 0.011 - - 285 0.003 - -
4'-11" 269 0.013 270 0.006 272 0.007 - -
X= 27'-00'' X= 29'-00''
y Run# I. F. Run# I. F.
0'-06'' 277 0.023 278 0.006
1'-04"' 281 0.013 280 0.002
3'-08" 286 0.001 - -

Table 6.3 Coord"mates and mean C;p of seIected taps along the short roo f-axtS
.
location
tap X (ft) y (ft) Cp (mean) Remarks
50123 1.00 23.17 -1.18 Mode-15
50223 1.67 23.17 -1.05 Mode-49
50523 4.67 23.17 -0.73
50923 8.67 23.17 -0.72 reattachment
51423 14.18 23.17 -0.43 occurs at l 0 ft
52323 22.58 23.17 -0.25 (3m)
52923 29.25 23.17 -0.21
64

Cp= -1.18

~ -1.05
i i""-..~: 0.73 -0.72

.'
II
I
I
I
I
I
I
I
I
.......- - - -
'I
I
l
l
-0.25 -021

I l I
I I I

50123 50523 50923 51423 52323 52923


50223
X

Figure 6.3 Tap locations and pressure distribution

Since the mean reattachment point (RP) is at about 10 ft (3 m) from the leading
edge, taps 50123, 50223., 50523 and 50923 are located upstream of the reattachment
point; taps 51423, 52323 and 52923 are downstream of the reattachment point.
A closer look at the velocity vectors near the leading edge of the roof reveals that
taps 50123 and 50223 are in the wedge-shaped region where the separated shear layer
meets the reverse flow (Figure 6.4). These two taps experience the strongest suction, this
could possibly be explained by a vacuum-like effect: the air in that wedge-shaped parcel
is drafted upward to create a zone of extremely low pressure (high suction).
Taps 50523 and 50923 are located within the zone of flow reversal and their mean
pressures differ only slightly. Taps 52323 and 52923 are in the forward flow zone and the
difference in the mean pressures is small. Tap 51423 assumes a mean pressure coefficient
falling sontewhere in between the values for the two zones before and after.
65

Wind---..

wedge-shaped region

Figure 6.4 Wind flow near leading edge of the SB

6.5 Mechanism of Peak-pressure Generation


Simultaneously recorded time series of pressure~ incident wind direction and speed
for Runs 50, 72, 105, 145 and 213 of the Mode-49 data were subject to detailed analysis
and comparison. Some statistics about these runs are summarized in Table 6.4. The 3-
component sonic anemometer was placed right above the leading edge, at heights of l '-
08", 2' -03"', 3', 1' and 4' -02"' from the roof surface, for those runs, respectively.
The wind direction measured by the sonic anemometer was verified to represent the
upstream wind direction, by comparison with the direction measured by the UVW
anemometer at 13 ft on the meteorological tower. Table 6.4 gives the comparison of the
mean and rms wind angles of attack (AOA). It can be seen that the mean values compare
favorably, and the rms values measured by the sonic anemometer are consistently larger
presumably because of the insufficient frequency response of the uvw-anemometer on the
tower and the spatial variability of wind field. Of course, the magnitudes of wind speed
measured by the sonic anemometer placed above the leading roof edge were affected by
the presence of the test building. A correction to them is achieved by using a run-specific
factor, a value determined as the ratio of a run-averaged wind speed from the sonic
anemometer to that from the UVW anemometer at 13ft on the meteorological tower. The
calculated correction factors are listed in the last column of Table 6.4.
It is noticed that natural wind changes its direction and speed considerably and
constantly. A typical range of wind direction fluctuations is around 80 in a time period
66
of 15 minutes. A large fluctuation could occur in a very short time. Fluctuations in the
order of 60 could happen in one~ tenth of a second. On the other hand, the fluctuations of
wind speed are relatively slow. Both the wind direction fluctuations and the wind speed
variations are important in the pressure-generation process.

Table 6.4 Typical statistics of natural wind


Meteorological tower ( 4 m level) Sonic Correction
- Wind azimuth~ Wind speed., mph AOA(j AOA(~ factor
Run# mean rms range mean rms mean/rms mean/rms -
50 0.6 10.1 61 18.3 3.35 255.6/10.1 258.5/16.7 1.07
72 142.7 8.78 118 14.3 3.28 262.7/8.8 260.8/13.9 1.15
105 230.6 10.8 70 18.3 4.08 275.6/10.9 279.6/16.8 0.97
145 30.8 11.6 87 21.2 2.95 270.8/11.7 271.0/15.5 1.08
213 233.8 10.8 87 30.4 4.92 263.8/10.8 265.4/12.9 1.07

6.5.1 Non-conventional Pressure Coefficient


Before continuing to the discussion of the pressure-generating mechanism~ a concept
of non-conventional pressure coefficient is introduced. To obtain the non-conventional
coefficient of pressure, Cp (t), an instantaneous dynamic pressure is used for
normalization, instead of a time-averaged or run-averaged dynamic pressure. The
instantaneous dynamic pressure is calculated by using the corrected velocity, measured
with the sonic anemometer placed right above the leading edge. The introduction of the
non-conventional coefficient of pressure makes it possible to isolate the effects of the
wind direction fluctuations from that of the wind speed variations on the pressure-
generating process.

6.5.2 Incident Wind and Pressure Coefficient


Careful and extensive examination and comparison of synchronized time series of
the approaching wind direction and speed, the conventional pressure coefficient and the
67
non-conventional pressure coefficient were conducted. Representative segments of time
series are illustrated for observational and explanatory purposes. The pressure coefficient
dealt with here is from tap 50223, the tap located 1.67 ft from the leading edge. The
pressure fluctuations observed at this tap location respond to the fluctuations of the
incident wind measured with the sonic anemometer with negligibly small time lags.

6.5 .2.1 Incident Wind


Sonic anemometers were originally selected for this full-scale application primarily
for their capability to detect wind flow direction, a feature essential to the measurement
of highly turbulent flow with frequent flow reversal. Sonic anemometers are rugged and
can operate reliably in adverse environments~ therefore suitable for outdoor usage. The
used Modell210-K-063 (Gill Instruments) sonic anemometer can measure wind speed at
rates up to 100 s- 1 (full 3-axis measurement). The anemometer's fast frequency-response
made it an indispensable piece of equipment for this application.
Data from the sonic anemometer indicated that direction fluctuation as large as 60
could happen in one tenth of a second, or even faster (Figure 6.5(a)). Conventional
propeller-type anemometers would have completely failed to detect such fluctuations.
The variation of wind speed in a typical 15-minute run is significant but at much slower
rate than that of the wind direction. It seems that fluctuations of wind speed and wind
direction were not entirely independent. A fast-and-large fluctuation in the direction was
often accompanied by a drop in the wind speed (Figure 6.5). Some typical statistics for
natural wind were summarized in Table 6.4.
The fact that pressure coefficients on the roof underneath the SB did not vary much
with the wind angle suggests factors other than the direction are responsible for high
suctions on the roof surface. In other words, a simple direction shift does not induce high
suction. The transient nature of those observed short-duration pressure peaks seems to
rule out the direction itself as an important factor in the pressure-generating mechanism.
68

Wind direction series, M49n072


80 ~------------------------------------------~
60
40
CD 20
~ O~hr~----~~~~~~~~~~~~~~~~--~
~-20
-40
-60 Seconds
-~ ~------------------------------------------~
560 570 580 590 600 610 620 630 640

(a) Wind direction

Wind speed series, M49n072


12.0
10.0
8.0
~ 6.0
4.0
2.0
0.0
560 570 580 590 600 610 620 630 640

(b) Wind speed

Figure 6.5 Fluctuations of incident wind

6.5.2.2 Conventional Pressure Coefficient


The conventional pressure coefficient, Cp, is a non-dimensional parameter
describing loading effect on buildings and other structures. It is generally location-
specific and is a function of building geometry. Characteristics of the incident wind and
the building's orientation with respect to the incident wind direction are also important
factors affecting the pressure coefficient. The conventional pressure coefficient, Cp, is
obtained by normalizing the pressure by a time-averaged free-stream dynamic pressure
measured at a reference height:
69
where p( t) is the pressure, p is the air density and V is an averaged wind speed over a
time period. Obviously, due to the non-stationarity of wind, both the starting instant and
the duration affect the value of V. Therefore, Cp (t) is virtually wind-speed dependent.
This widely used definition is practically convenient, but not always consistent.
Look at Figure 6.6(b), there is a period of 50 seconds (640 s - 690 s) in which the
conventional Cp takes on relatively small values (low suctions) simply because of the
lower-than-the-average incident wind speed during that interval. The ensuing higher Cp is
mostly a result of wind gusts (Figure 6.6(d)). Although information on the actual pressure
p(t) can be recovered from the Cp values, the conventional pressure coefficient is,
however, somehow misleading and does not accurately reflect the mechanism of pressure
generating, which ideally should be wind-speed independent.

6.5.2.3 Wind Angle and Pressure Coefficient


In general, the pressure coefficient is a function of the wind angle with respect to the
building orientation, or equivalently a function of the wind angle-of-attack (AOA). A
common practice in documenting pressure coefficient is using an averaged wind angle.
Little effort has been made to examine the condition of incident wind at instants of
""interesting eventn when high suctions were produced. This practice is legitimate for the
documentation of statistical quantities, such as the mean and rms pressure coefficients. It
is not justified when individual events such as the minimum coefficient are dealt with.
Pressure coefficient (conventional) data indicate that the mean and minimum pressure
coefficients of tap 50223 are insensitive to the AOA around 270, this is shown in Figure
6. 7. Similar data for other taps along the short roof-axis are provided in Appendix C.
70

Wind direction series, M49n072


150

100

ISO
Eo
-50
Seconds
-100
840 650 660 670 680 690 700 710 720

(a) Wind direction

0.0 r---------__,.....,..-----.------------.
-0.5
-t.O 1----------~-:.------.........::;r.-~~~~~~:i-H
a..-1.5
o.2.o run mean= -1.04
2.5
-3.0 run minimum= -4.79
-3.5 ~...- _ __.__ __.__ __,___---'_ ___,__ _ _.___--..~.__ _ ___,
640 650 660 670 680 690 700 710 720

(b) Conventional pressure coefficient

(c) Non-conventional pressure coefficient

2
Seconds !
0 ~--~----~----~--~----~----~-~----~
640 650 660 670 680 690 700 710 720

(d) Wind speed


Figure 6.6 Simultaneous traces of incident wind and pressures (640s - 720s)
71

Mean & minimum pressure, aaparation bubble

IS
..r -2
.............................._........ --
0+----+----+----+----+----+----~---r----+----r--~

.......... -.-.,........... .
c

~ -4

.!~ .......~~~-'-.!.!~--.-----
..6
II
i -8 i

l -10~----------------------------------------------~
J
~ ~ ~ ~ ~ m ~ ~ ~ ~ ~

wind Wigle

Figure 6.7 Pressure coefficient versus wind angle (tap 50223., Mode-49 data)

6.5.2.4 Non-conventional Pressure Coefficient


As defmed in Section 6.5. L the non-conventional pressure coefficient is wind-speed
independent, i.e.,
cp <t> =P<t>t<p ~<t>t2).
This defmition is free from the ~contaminating' effect of the wind speed on the pressure
coefficient, Cp (t); since, instead of the averaged dynamics pressure, an instantaneous

dynamics pressure is used. The non-conventional pressure coefficient, Cp (t), is


theoretically consistent. It reflects the pressure-generating mechanism governed by
parameters other than the wind speed, such as the upstream wind profile, the incident
wind direction fluctuations and the wind-structure interaction.
Of all those factors other than the incident wind speed, characteristics of the incident
wind are the input; the wind-structure interaction acts like a transfer function; the loading
effect on a structure is the output. Conceptually, this is shown in Figure 6.8. The building
geometry and the AOA are variables of the transfer function. While the input and the
output are usually directly measurable, a transfer function represents a complicated
process involving many elements, and is often only identifiable.
72

Loading ( Cp)

Figure 6.8 Wind - structure interaction and loading effect

Some cause-effect relationship between the incident wind and the pressure
coefficients is of interest. It is assumed that the direction fluctuations (not the direction
itself) are the most important single input; other contributing factors such as the wind
profile are put aside to simplify the problem. It is from this basic assumption that the non-
conventional pressure coefficient was introduced.
Figure 6.9 displays four synchronized segments of 80-second duration for the
incident wind direction, the conventional pressure coefficient., the non-conventional
pressure coefficient and the wind speed. A close comparison of the direction fluctuations
(Figure 6.9 (a)) with the trace of the non-conventional pressure coefficient (Figure 6.9
(c)) reveals that the primary peaks of Cp (t) coincide with large-and-fast fluctuations of
the wind direction. It is difficult, though, at this stage to quantify the phenomenon.
Qualitatively, the rate of a direction fluctuation, the magnitude of the fluctuation, the
immediate history of wind direction tluctuation., including the amount of deviation of the
direction from the windward normal, are among those factors affecting the Cp (t).
The SB is usually considered as a two-dimensional phenomenon; fluctuations of the
incident wind direction disturb a certain delicate balance within the SB. It takes some
time for the balance to establish itself. A fast-and-large wind direction fluctuation about
270 preceding a more or less constant wind direction seems to break this balance and
initiate low pressures (high suctions). This is not the case with the conical vortices, where
the vortices are three-dimensional and stable; the peak pressures are results of combined
gusts and incident wind directions (not the fluctuations) conducive to vortex formation.
73
Wind c:iraction series. M49n072
~~----------------------------------------~
60
40
.20
I o~~~--~~~yq~-w~H+~~~~~~~~
l-20
-40
-80
~ ~----------------------------------------~
560 570 580 590 600 610 640

(a) Wind direction

Pressure coefficient, M49n072


05 ~--------------------~------------------~

-3.0
560 570 580 590 600 610 630 640

(b) Conventional pressure coefficient

Norrconventional pressure coefficient, M49n072


1.0
0.0
-1.0
-2.0
8--3.0
u~--~~~~
II' I ~ I' .- 1 ll I , - ,

-4.0 run mean= .0.99


-s.o run minimum= -9.23
-6.0
Seconds!
560 570 580 590 600 610 620 630 640

(c) Non-conventional pressure coefficient

Wind speed series, M49n072


12.0 ....-----------..:::.....:..!.....:..::..---.::.....:_.:...;__ _ _ _ _ _ _ _ _ _ __

10.0
.~
8.0 .' '>~,_ I

'i '1\r ...... --~~!


e s.o lrt ,...
~ ':;
J
1

; ?"e ' ' ; ,
ill[:; -~ '~.., 1 1."tf.,-...
r-r,.., ' ~ l:
,,.... ~.A.
., j
1 ...,..,/r/\i.ftl..~
1..1'-- T ' ;>--'4<4 ,
4.o \ "f '~"'r ..i,'
run mean= 7.58 t ., i
2.0
0.0 ....___ _....._ _"""'-_ _ ....__--~. _ __..__ _....__ _Seconds
_ _-..J l
560 570 580 590 600 610 620 630

(d) Wind speed

Figure 6.9 Simultaneous traces of incident wind and pressures


74

Observations indicate that the primary peaks of the conventional pressure coefficient in
the SB were often an outcome of combined fast-and-large fluctuations of incident wind
direction and wind gust, with the former playing a more important role (Figure 6.9 (b)
and (c)). Slow (but not necessarily small) fluctuations of the incident wind direction only
induces weak suctions ( Cp (t) of small magnitudes), see segments from 564 second - 580
second in Figures 6.9 (a) ;ud 6.~(c).
It is not clear at this point what was happening in terms of the wind flow inside the
SB at instants of high suction. Simultaneous, multiple-point velocity measurements are
desirable for an affumative answer. Letchford (1995) claimed that some shear-layer
instability, which was strongly influenced by the incident turbulence, was an important
parameter in this respect. He postulated that instability caused by incident turbulence
forces the separated shear layer to prematurely reattach onto the roof, creating peak
suctions.
Effort was made in this study to establish certain flow patterns associated with high
suction, using the conditional sampling technique adopted by Marwood ( 1996) for the
conical vortex situation. No simple and clear pattern emerged, except the observation that
flow inside the SB at moments of high suction was quite three-dimensional Thus the
flow defies a straightforward two-dimensional characterization. Correlation analysis
between the incident wind and the pressure at tap 50223 failed to show good correlation.
More work is needed for quantification of the wind-pressure relationship. However, the
current study suggests an answer to the question posed by Letchford (1995) as to the
cause of shear-layer instability, i.e., fast-and-large directional fluctuation in the incident
wind might be the initiator of the shear-layer instability.

6.6 Spectral Analysis of Pressure


Spectral analysis of pressure for taps 50223, 50523., 50923, 51423, 52323, 52923 in
Mode-49 is performed in this section. These taps are along the short roof-axis of the
WERFL test building, and under different flow zones of the SB.
75
Six 20-minute data runs M49n063, M49n067. M49n082~ M49n088, M49n092 and
M49n114 of the Mode-49 data are used for spectral analysis. These runs have mean wind
AOA within 1 from 270. Commercial software DADiSP (Data Analysis and Digital
Signal Processing Software) is used for spectral analysis. The sampling frequency for the
pressures was 22.5 Hz. Each of the six pressure spectra presented in this section is an
ensemble average of 18 spectra computed from 18 (= 6 runs x 3 segments) non-
overlapping segments of time series of364 (=2 13/22.5=364) seconds.
Figure 6.10 shows the pressure spectra for taps 50223, 50523, 50923, 51423, 52323
and 52923. These taps are located at 1.67, 4.67, 8.67, 14.08, 22.58 and 29.25 ft,
respectively~ from the leading roof-edge.
Two features are readily noticed from the six spectra curves: (1) There is a small but
noticeable growth in spectral amplitude at 0.03 Hz (33 seconds) in a general trend of
decay for pressure taps 50223, 50523, 50923 and 51423; (2) The spectral amplitudes of
pressure taps 52323 and 52923 die out consistently with increasing frequency. These two
observations are in agreement with the fmdings of Kumar and Stathopoulos (1997). Their
results showed that although the spectra changed for different roof geometry, terrain
condition, wind angle and tap location.. the pressure spectra could be classified into two
categories, namely, ( 1) Gaussian spectra, which died out at high frequencies, and (2) non-
Gaussian spectra, which died out up to a certain frequency and then grew to have another
hump prior to their dying.
Figure 6.11 indicates the Gaussian and the non-Gaussian zones of pressure on flat
roofs. Using skewness and kurtosis as criteria, the NBCC (1995) specifies Z to be 10% of
least horizontal dimension or 40% of lower eave height, whichever is lesser (Figure
6.11 ). Accordingly, Z == 3 ft in the case of the WERFL test building. However, using the
criteria given in Kumar and Stathopoulos ( 1997), the spectra of pressure obtained in this
study seem to indicate a non-Gaussian pressure zone of Z = 15 ft for the WERFL test
building, which is equivalent to the roof height (13 ft). Yin (1996) gave a value of
skewness coefficient of - 0. 78 for tap 51423, which again indicates a non-Gaussian
pressure zone extending at least 15 ft from the leading roof edge if the criterion of the
76
absolute value of the skewness coefficient being greater than 0.5 is used. In fact, the
pressures on the roof are never ideally Gaussian, and Yin (1996) found that the shifted
lognormal PDF is a suitable model
Concurrence of the peaks at 0.03 Hz (33 seconds) in the pressure spectra of taps
50223, 50523, 50923 and 51423 is probably associated with some flow phenomenon
inside the SB, e.g., the periodicity of dynamic circulation of the SB, which has a
characteristic time of about 30 seconds.

pressure spectra (separation bubble)


0.16

0.12

~0.08

0.04

0.00
0.001 0.01 0.1 f, Hz 1

Figure 6.10 Spectra of pressure measured at six taps along short roof-axis

non-Gaussian

Wind Gaussian

z
Figure 6.11 Pressure zones on flat roofs: Gaussian and non-Gaussian
(after Kumar and Stathopoulos, 1997)
77
Typical time series of Gaussian and non-Gaussian pressure distributions are
displayed in Figures 6.12 and 6.13, respectively. Wind pressure spectra, representing
energy content in the pressure fluctuation, receive contributions not only from the
mechanical turbulence but also from the building generated turbulence., both varying with
the building configuration, the exposure condition and architectural details (Kumar and
Stathopoulos, 1997).
Recalling the structure of the SB discussed in Section 6.2, the non-Gaussian pressure
zone as represented by taps 50223, 50523, 50923 and 51423 could be related to the wind
flow in the SB. These taps are underneath the overall flow circulation of the SB, and the
pressure fluctuation is non-Gaussian. Of course, the pressure becomes progressively less
non-Gaussian as the location moves away from the leading roof-edge.

6.7 Summary
The main points discussed in this chapter can be summarized as follows.
l. The overall SB is oblong and elongated in the upstream wind direction. Four flow
zones near the roof surface can be identified: (a) the leading edge zone, where the
separated shear-layer meets the reverse flow to create a wedge-shaped region; (b)
the reverse tlow zone; {c) the reattachment zone; and {d) the forwarding flow
zone;
2. The mean point of reattachment (PR) is approximately 10 ft from the leading roof
edge; the mean height of the SB is about 4.5 ft;
3. The pressure distribution on the roof surface underneath the SB is directly related
to the structure of the SB. The uplifting pressure on the roof generally decreases
with distance from the leading edge;
4. The introduction of the non-conventional pressure coefficient made it possible to
separate the effect of wind speed from that of the wind direction on pressure
generation;
5. Time series analysis indicates that fast-and-large fluctuation of the incident wind
direction governs the mechanism of peak-pressure generation;
78
6. The primary peaks of conventional pressure coefficient are often an outcome of
combined wind direction fluctuation and wind gust.
7. The non-Gaussian pressure zone is found to extend at least 15ft from the leading
edge~ and thls dimension is comparable to the height of the test building.

time series, tap 52923 (M49n088)


0.2

0.0

-o.2
c..
0
-o.4

-o.6
seconds
-o.a
0.00 60.00 120.00 180.00

Figure 6.12 Typical trace of pressure coefficient: Gaussian

time series, tap 50223 (M49n145)


0.0
-0.5
-1.0

-1.5
c..
0
-2.0
-2.5
-3.0
seconds
-3.5
0.00 60.00 120.00 180.00 240.00 300.00 360.00

Figure 6.13 Typical trace of pressure coefticient: non-Gaussian


CHAPTER 7
CONICAL VORTICES AND THEIR EFFECfS ON PRESSURE

7.1 Introduction
Post-disaster surveys have repeatedly revealed the wlnerability of low-rise building
roofs to severe wind. It is now well-established knowledge that, at oblique wind, the
leading roof comers of rectangular low-rise buildings sustain outward acting pressures
higher than any other part of the building envelope. Yet, it is precisely with the roof-
corner region that our efforts at load prediction were the least successful. Despite the
progress made over recent years, simulation techniques and results by different
researchers were not always consistent; model-scale predictions of wind loading in
simulated wind flow still have room for improvement. Among others, the works of
Tieleman et al (1993), Kawaii and Nishimura (1996), Marwood (1996) and Banks et aL
(1997) are representative of an effort to understand the conical vortex phenomenon over
the roof and the pressure induced on the underlying roof surface.
Full-scale studies of the incident wind and its relationships with the conical vortex
and the pressure could uniquely contribute to a better understanding of the wind
simulation requirements, and eventually provide guidance for improving wind-loading
simulation.
This chapter intends to discuss the fundamentals of the conical vortices above a flat-
roof corner and their relationship with the surface pressures.

7 .2. Theoretical Background of Vortex and Pressure


By mapping the surface pressure field on a roof-comer, Tieleman et al. (1992)
deduced the location of roof-edge vortices, assuming the loci of minimum roof pressure
being directly beneath the axes of the vortices. They found that, at quartering incident
wind, (two) vortices would form above (two) rays taking an angle of approximately 15
with each of the roof edges.

79
80
The assumption of minimum pressure occurring below the axis of a vortex core is
partly supported by the potential flow theory. In a free vortex, the theory states., the
streamlines are in the form of a nest of concentric circles, and the tangential velocity is
inversely proportional to the distance from the vortex center. The circulation along any
circumferential path in the vortex is independent of its radius. The pressure distribution
on an underlying surface can be estimated from the velocity field using

in which pis the surface pressure at a point, p min is the peak suction, y is the distance of

a point on the surface from the vortex axis, parallel to the roof surface, zh = 1.55 Yuz., is
the height of the vortex core above the surface, y 1n. is the half-width of the vortex, ie.
the value of y at p/p min= 0.5 (Figure 7.1). More detailed derivation can be found in
Marwood (1992). The above formula is only approximate, given the fact that the vortices
above a roof corner are usually oblong (elongated in the horizontal direction) and far
from the ideal circular free-point-vortex. The pressure distribution on an underlying
surface is shown in Figure 7.l.

P min

Figure 7.1. Pressure distribution under a free point-vortex (after Marwood, 1996)
81
In their work of simultaneous pressure measurement and flow visualization on
cuboidal models tested in wind tunneL Banks et aL ( 1997) found that the pressure
distribution under roof-comer vortices was asymmetric with respect to the vortex axis.
Pressures for locations having angular position (from the roof edge) greater than the
vortex axis exceeded the symmetric curve-fit values. However, their fmdings indicated
that the angular position of highest (mean and peak) suctions did follow the vortex core
position closely.

7.3 Characteristics of Pressure on Leading Roof-comer


The Wind Engineering Research Field Laboratory of Texas Tech University has a
huge database of wind pressure coefficient for its test building. One of the roof comers is
heavily instrumented with pressure taps. Among the taps provided~ taps 50101, 50501.
50901, 50205, 50505, 50905, 50209, 50509 and 50909 have been active for most of the
data acquisition modes. This corner region receives the most severe suction in the
building envelope. Comparisons of pressure coefficient with model-scale results have
focused on the taps at this roof corner. A portion of the roof corner accommodating these
taps is sketched in Figure 7.2, and the coordinates of the taps are provided in Table 7.1.
The strongest suction ever recorded was with tap 50501.

(y)
---------------~

0 0 0
50209 50509 50909
50205 50505 50907
0 0 0

50101 5<a01 50901


0 0 (x)

Figure 7.2 Instrumented roof comer and taps


82

Table 7.1 Coordinates of selected pressure taps


Tap# X y
50101 1.17 ft 0.36 m 1.17 ft 0.36 m
50501 4.67 ft L42m 1.17 ft 0.36m
50901 8.67 ft 2.64m 1.17 ft 0.36m
50205 1.5 ft 0.46m 5.17 ft 1.58m
50505 4.5 ft 1.37 m 5.17 ft 1.58 m
50905 8.67 ft 2.64m 5.17 ft 1.58m
50209 1.83 ft 0.56m 9.17 ft 2.79 m
50509 4.83 ft L47m 9.17 ft 2.79 m
50909 8.83 ft 2.69m 9.17 ft 2.79 m

Some statistics of the pressure coefficients for these nine taps are listed in Table 7 .2.
A general trend of the pressure distribution can be identified: as the location moves away
from the two roof edges and from the apex, the intensity of suction drops rapidly. ln other
words., only a small portion of the upwind roof comer sustains severe wind loading. It is
of vital importance, however, that this portion be adequately designed and constructed
against intense local uplifting force to prevent initial damage to the building envelope.
Mode-15 data of pressure coefficient versus the wind angle are reproduced here for
taps 5010 I and 50501 to give a general idea about the characteristics of pressure on flat
roof corners of low-rise buildings. Similar graphs for the other seven taps are included in
the Appendices. Synchronized traces of pressure coefficient and incident wind are also
provided to emphasize the points to be made.

7.3.1 Pressure Coefficient for Tap 50101


For tap 50101, Figure 7.3(a) indicates that there are two peaks for the mean pressure
coefficient at about 190 and 260, respectively; considerable pressure recovery (less
negative) occurs at wind angles around 225. This is due to the fact that tap 50101 is
located on the 225 ray starting from the apex.
83
Table 7.2 Statistics of roof corner pressures, averaged values over 220-230 (Mode-15)
Pressure coefficients, Cp
Tap# mean nns minimum maximum
50101 -0.74 0.57 -6.83 0.38
50501 -2.08 1.27 -8.60 0.19
50901 -1.76 0.75 -6.76 0.01
50205 -1.61 1.16 -6.87 0.27
50505 -0.39 0.27 -4.10 0.49
50905 -0.38 0.21 -3.01 0.42
50209 -1.76 0.89 -5.97 0.14
50509 -0.52 0.55 -4.66 0.53
50909 - - - -

Tap 50101 is either beneath the vortex generated at the left roof-edge with incident
wind angles of around 260~ or under the vortex formed at the right roof-edge with wind
angles around 190 (Figure 7.4). At around 225~ two small vortices are generated but
neither of them is big enough to reach tap 50101, and straight flow passes over the tap.
Figure 7.5 gives a schematic view of the phenomenon. Synchronized traces of the
pressure coefficient and the incident wind direction in Figures 7.6 and 7.7 support this
claim.
Pressure coefficient traces of tap 50101 are characterized by small fluctuations
around a nearly constant value of about - 0.5, which is sporadically interrupted by brief
moments of large fluctuation (Figure 7.6(a)). The same figure shows a generally flat
pressure coefficient, except a brief moment of strong suction lasting about 5 seconds
(290s - 295s). This eventful 5 seconds for the pressure corresponds to incident wind
angles close to 190 (Figure 7.6(b)), Le., the angles at which the vortex formed from the
flow separation at the right edge is big enough to encompass tap 50101. A close look at
the fluctuation of the wind speed within these 5 seconds reveals that the pressure
fluctuation closely follows the wind speed variation.
84

AOAva.U.an
o.s
0--~~-+~~~~~---+------+------r------+-----~----~

~r----,~~-r~~~~~j---~~r-t--=~~---t
c -1~-----+------+-----~~

i -15~-----+------+------+-----
2~-----+------+------+------+------+------+--.---+----~

-25~-----+------+------+------+------+------+-r----+----~

~~----~------~----_.------~-------~---------~-------~-------~
0 90 135 180 270 315
AOA

AOA.va. RMS
1.4
12+------+------+------+------~--~~--~~1-----~-------

~ OB+------+------+-----~-----,~~~~----~~~--~------
:1
~ OB+------+------+-----~--~~--~~~~~--~~
~4+-----~------r---~~
02~~--~~~~~~~~

0+---~-+------+-----~------~----~------~----~------~
0 90 135 180 225 270 315
AOA

A.OAva. Min

10+------+------+-----~------~~--~K-~~~~--~~----~
c
2
15~-----+------+-----~--------~----~------~--------~----~

25

A.OAva. Max

i ~~i'l~t.#~~f\4~
0 90 135 180
AOA
225 270 315

Figure 7.3 Tap 50101 pressure coefficients versus wind angle (Mode-15)
85

' Vortex from right


Vortex from left edge separatioo at 190"
edge separation at 2

Figure 7.4 Schematic view of vortices as they affect pressure tap 50101

(Roof corner)

--tap50101

t
225
Figure 7.5 Straight flow and vortices formed at incident wind around 225

A similar argument accompanies Figure 7.7, except here., instead of one, there are
several short moments of high suction, and the pressure tap was under the influence of
the vortex from flow separation at the left roof-edge, when the wind angles were close to
260. However, a pressure-coefficient peak did occur at 27-second, at wind angles not so
much off225, due to higher-than-the-mean wind speed at the moment.
At this point., it is worthwhile to go back and look at the wind-angle/pressure-
coefficient relationship shown in Figure 7.3. The distribution of the minimum pressure
coefficient has significant scatter and no distinct peak exits, unlike the distribution of the
86
mean pressure coefficient. This is presumably caused by what can be called the "direction
smear'' effect't ie., the minimum pressure coefficient is an individual event, it took on its
values at wind angles that were probably much off the run-averaged wind direction, a
convenient wind direction commonly used for documentation. Therefore, if the
instantaneous wind angle at the instant of pressure peak occurring had been used for
plotting, it might have been a very different picture: with much-reduced scatter and
possibly two distinct peaks, like the case for the mean pressure coefficient. The scatter
with the rms distribution was likely caused by the non-stationary nature of the
atmospheric wind.
To end this subsection, a brief summary is helpful:
1. The pressure coefficient traces at tap 50101 were characterized by small
fluctuations around a nearly constant value of about -0.5, which was sporadically
interrupted by brief moments of large fluctuations.
2. The pattern of pressure fluctuation observed at tap 50101 was presumably
associated with the straight flow passing over the tap, and occasional large
vortices formed from flow separation at one of the roof edges.
87

pressure coefficient trace, tap 501 01 (M49n160)


1.0

Cp
4.0 ~----------------------------~~+-----------~

~.0 ~----------------------------~~+-----------~
seconds
4.0 ~--~--~------~--~--~--------~--~--~--~
255 265 275 285 295 305

(a) Pressure coefficient at tap 50101

wind angle trace (m49n160)


250 -~~---------------,..----r------~

240
230
~220
Q)

~10
"'0
200
190
seconds
180
255 265 275 285 295 305

(b) Incident wind angle

Figure 7.6 Time series of pressure coefficient and


incident wind angle (Mode-49,Run M49n160).
88

pressure coefficient trace, tap 50101 {M49n163)

2.0
1.0
0.0
~ -1.0
-2.0
-3.0
-4.0
0 10 20 30 40 50 60

(a) Pressure coefficient at tap 50101

wind angle trace (m49n 163)


270
260
en
Q)250
~
~240
"'0
230
220
seconds ,
210
0 10 20 30 40 50 60

(b) Incident wind angle

Figure 7. 7 Time series of pressure coefficient and


incident wind angle (Mode 49, Run M49nl63)
89
7 .3.2 Pressure Coefficient for Tap 50501
Figure 7.8 presents the relationship between the wind angle and the pressure
coefficients at tap 50501. The mean., nos and minimum pressure coefficients have only
one recognizable peak at the wind angle around 210. This is different from what was
observed for tap 50101. Tap 50501., because of its location, is basically under the
influence of the vortex formed after separation at the right roof-edge; the vortex formed
at the left edge does not reach tap 5050 1. Figure 7.9 depicts the situation.
Earlier, we discussed the vortex - pressure relationship in Section 7 .2. The peak of
the mean pressure coefficient distribution in Figure 7.8(a) seems to indicate the presence
of a strong vortex over tap 50501, at wind angles of 200 through 210. It is noticed that
the behavior of the pressure coefficient at tap 5090 l is quite similar.
Two synchronized time histories (M49n161) of the pressure coefficient (tap 50501)
and the upstream wind speed are produced in Figure 7.10. A comparison of the traces
suggests some quasi-steady wind-speed - pressure relationship. From 0 second to 300
second~ the pressure coefficient takes on low values (high suctions) because of the
higher-than-the-average wind speed. In the ensuing period from 300 second through 900
second, there are noticeably eight pressure (coefficient) peaks.. and the general fluctuation
is characteristic of pressure coefficients on a flat roof corner under the direct impact of
the conical vortices. These high suctions (Figure 7.l0(a)) concur with gusts in the
incident wind (Figure 7.10(b)).
A close-up look at one segment (395 second - 425 second) of Figure 7.10 is
instrumental in revealing the wind - pressure relationship. Referring to Figure 7.11, there
are two peaks in the pressure coefficient trace at 407 second and 410 second
corresponding to high wind speed at wind angle in the range of 200 - 210. As
mentioned earlier, 200- 210 is the range conducive to intense vortex formation over
tap 5050 l. Concurrence of wind gusts and wind angles favoring vortex-formation leads
to strong vortical circulation, and hence high suctions.
90

AOAva.Man
0.5

-0.5
0
-- -.... -~-..-. ..... ~-
~. ....... 'f'

I;~ ,...4...:" .
c 1

:1 1.5
2
~~
-2.5
-3
_.._
"--:
-3.5 ~-

0 45 90 135 180 225 270 315 360


AOA

AOAvs. RMS
1.8
1.6 ...
1.4
4~ I

1.2 .. ...c~ ~ ....


Ia: 1
jf_~ -
0.8
0.6
0.4
0.2 .... .....
.........
..... ~
....
6

~r
....
.... ...
.... ~ ......
....

,..
I
I

0 J -~
0 45 90 135 180 225 270 315 360
A.OA.

AOAvs. Min
o~w.~~~~~----~----~----r-----r-----r---~
-2~----~~~~~~~~~---+----~~--~~-
-4+----+---+---
c -6+------+------+------+~~~

~ -8~----~-----+------~----~~--~~---4------~----~
-10t-------+------t-------+------4~~~~L-~~+------+---------

12+------+------+------+------+---~-+------+------+---------
-14
0 45 90 135 180 225 270 315
AOA

1.5

= 05 lii~Sil~dt!~+~~~d~--~
~ o~~~~---~--.-~~

~+-----~----~------+-----~

-1
0 45 90 100 270 315
NJA

Figure 7.8 Tap 50501 pressure coefficients versus wind angle (Mode-15)
91

In the same figure, the wind gust at 404 second induces only a weak suction due to
wind angles not favoring vortex formation. The wind at 400 second seems to be in the
right angle range but it fails to create a high suction because of much lower wind speed at
that instant.

Vortex from tbe right edge

Vortex from the left edge


tap 50501

t"" Wind

Figure 7.9 Location of tap 50501 with respect to vortices

The above argument can be generalized to taps 50901, 50205 and 50209., except that
the range of wind angle for taps 50205 and 50209 should be approximate 240 - 250 out
of symmetry of pressure tap location. All these taps are close to the apex and also close to
the roof edges.
Similar to the case with tap 50101, the minimum pressure coefficient of tap 50501
has more scatter than that of the mean pressure coefficient. This can again be attributed to
the udirection smear," ie., the erroneous assignment of the angle-of-attack to some of the
data points. The scatter would have been much less if the actual instantaneous angle,
instead of the run-mean angle, had been used for plotting.
The wind-speed-dependent minimum-pressure coefficient (conventional) is not a
fundamental quantity. It takes on values that fall in a wide range, depending on the
incident wind. The conventional minimum pressure coefficient would reduce to the mean
pressure coefficient, if its wind-speed dependence could be removed by normalization
with respect to the instantaneous dynamic pressure. This is verified through a successful
construction of pressure coefficient series.
(a) Pressure Coefficient

(b) Upstream wind speed (from sonic anemometer)

Figure 7.10 Simultaneous series (Data run M49n 161) '-0


N
93

pressure coefficient (tap 50501), M49n161


0
-1
-2
fr-3
4 ~------------------~~+---------------------~
-5
seconds
-6
390 400 410 420 430

(a) Pressure coefficient (tap 50501)

e~ 10 ~------------~~~~-r~------+--------------

~ : t----IIFUU.-.:,.-:::----1-f/'-L--r:

6 ~--------~----------~----~----~----~----~
390 400 410 420 seconds 430

(b) Incident wind speed

incident wind angle. M49n161


~0 ~-----------------------------------------------
210 ..,._ _ _ _---T=---~ilf-Y---------
'"'~~------~MM~~~~~~~~~~~--~~~~
-3
f190
l180~~~~-------------------------------------
'i
170 t---...,._;.---------------------------
100 ~--------~----~----~----~----~-------~
390 400 410 420 seconds 430

(c) Incident wind angle

Figure 7.11 A close-up look of incident wind and pressure (M49n 161, tap 50501)
94
It seems appropriate at this point to establish the flow pattern of the vortices above a
flat roof comer at oblique wind angles:
1. A pair of vortices form over the roof comer at wind angles close to 225, but they
are relatively small;
2. Only one large vortex would form at a time from flow separation on either side,
depending on the incident wind angle, since the two ranges of wind angle
favoring the formation of large vortices at the two roof edges do not overlap;
3. Pressures on the underlying roof-corner reflect the flow field over the roof-corner.
In conclusion, strong vortical circulation over flat roof-corners and the associated
high suctions on the underlying surface are conditional on favoring incident wind angles
and wind gusts. This viewpoint will be strengthened in the following subsection, where
the pressure is successfully predicted from incident wind, and will be revisited later in the
context of flow visualization.

7.3.3 Predicting Pressure Coefficient from Incident Wind


It was observed in the previous subsection that both speed and direction of the
upstream wind are primary factors governing the pressure generation on roof comers at
oblique wind angles. Inspired by this observation, a numerical experiment was designed
to further reinforce the observation by predicting the pressure coefficient from the
information of the incident wind and the mean pressure coefficient. Ideally, instead of the
mean pressure coefficient, a wind-speed-independent pressure coefficient should be used.
As discussed earlier, of all the commonly documented pressure coefficients, only the
mean pressure coefficient is fundamental. Other coefficients are wind-speed-dependent
and stochastic in nature. The mean pressure coefticient serves as a convenient yet close
substitute for the ideal wind-speed-independent pressure coefficient. A time history of the
pressure coefficient for tap 50501 using the incident wind from data run M49nl58 is
created next. The information of the mean pressure coefficient versus the wind angle for
tap 5050 1 is shown in Figure 7 .12.
95
Here a different defmition of wind angle is used for convenience. The wind angle
perpendicular to the long wall (45 ft) of the WERFL building is defmed as 0~ and the
direction along the diagonal of the roof comer is 45. The data points in Figure 7.12 cover
a wind angle range of 3.3- 108.7, and are curve-fitted to a polynomial of the sixth
order,. to find the mean pressure coefficient (Cp(a)) as a function of the mean wind angle
(a). The expression of this function is used with the time series of the measured wind
angle (<Xi, at time t;) to fmd Cp(a;), which is employed to calculate the instantaneous
Cp(ti), using the following formula., Cp(ti) =Cp(a;) x (u(ti)IU) 2 (see Appendix D for its
derivation), where u(ti) is the measured incident wind speed, and U is the run-averaged
wind speed.

v=O.ooooooooo1saess4x-- o.ooooooo&&397112&r +O.oooooa2asoaso&42x 4


0.0004473223375942~ +0.0096039582928446x' 0.0814064164569079x


Cp 0.4544660256819530
0.0 . . . . . . - - - - - - - - - - - - - - - - - - - - - - - - - - - ,
-o.s r--R~i!J;Ji..-------------.:~
1.0 1---z.-~--...~~------------.
-1.5 1-------....z~--=-------=---a........-;~~
-2.0 t----------.... -------..~...:----::
-2.5 1------------...-E::::!!II.........piG~-----------!
... ~L-------1

-3.0 J__ _ _ _ _ ____::._~~~-=------~


-3.5 "--"--"--...__"--..__....___......._..~..-....__..~..-...._...~.-_.__...~.-_,__-L-__._--a...__._--a...;::...._~
0 10 20 30 40 50 60 70 80 90 100 110

Figure 7.12 Mean Cp versus wind angle (tap 50501, Mode-15)

Traces of the variables used in the above computation are demonstrated in Figures
7.13 and 7.14. Figure 7.l3(a) is the incident wind direction series {<Xi), which has
considerable fluctuation with a range of about 87. Figure 7.13(b) shows the wind speed,
u(tJ, with a run-mean U = 10.33 rn/s. Figure 7.13(c) gives the instantaneous pressure
coefficient, Cp(a;), corresponding to a; at instant ti. The factor [(u(ti)/U)2 ] that accounts
for the wind-speed dependence of the instantaneous pressure coefficient (Cp(ti)) is
presented in Figure 7.13(d). Figure 7.14 compares the measured pressure coefficient with
the predicted one (Cp(ti)). The similarity is amazing, given the simplicity of the model
used.
WtM4

(a) Wind direction (b) Wind speed

-o.5
-1.0

-1. ~

-2. 0

-2. ~

(c) Pressure coefficient (d) Gust factor squared

Figure 7.13 Traces of variables used in pressure prediction (M49n 158)

....0
0'\
(a) Measured pressure coefficient (M49n 158)

~ M49N158.\4

0.0

-2.0

--4. 0

-6.0

-8.0

(b) Predicted pressure coefficient

7.14 Comparison of pressure coefficients


\0
-...J
98
A close-up comparison of the two traces in Figure 7.14 indicates that differences do
exist. The measured pressure coefficient contains more high frequency content,
presumably because of the spatial variation and temporal fluctuation of the incident wind.
Since the vortex at one instant is influenced by wind entrained into the circulation from
different locations along the roof edge, and from a segment of airflow time-wise in the
continuous incident stream trapped in the vortical circulation. This compound effect
introduces more high frequency content to the pressure fluctuation, a situation the simple
model used here could not accommodate. This is schematically depicted in Figure 7.15.

Figure 7.15 Schematic view of variability of upstream wind

In spite of the complexity of the actual situation and the simplicity of the model, the
general agreement of the pressure coefficient is satisfactory. Of course, while the wind
speed and the wind direction are the most important parameters, other factors might also
be important. For instance, the venical profile of wind speed is important in determining
how much airflow would go over the roof when the flow approaches the building.
99

Perceivably, the fundamental features of the profile are preserved when the wind flow
actually fluctuates, and it is probably this fact that renders it possible to predict the
pressure from the upstream wind measured at only one point. The fmdings made here are
important from the viewpoint of practical applications, since a point-measurement of the
approaching wind should not pose a problem in many practical situations, and the wind
tunnel technique can predict the mean pressure coefficients fairly accurately.

7.4 Pressure Zones and Critical Wind Angles


Although the focus of this chapter is on the conical vortices and the pressure on flat
roof corners, it is worth the time to take a look at the overall pressure on rectangular flat
roofs. A thorough look into the pressure database of Mode-15 for the taps plotted in
Figure 7.2 reveals some patterns in terms of the relationship between the mean pressure
coefticient and the wind angle. Five zones are identified. and they are labeled I, II, Ill, IV
and V. Each zone extends 13ft (roof height. a characteristic length) from its edge toward
the center of the roof, Figure 7 .16.
It is observed that there are particular ranges of wind angle that induce the most
severe loading for zones I, It III, and IV. These ranges are 351-45, 45-135,
135-225, and 225-315, respectively. Zone V, which experiences the least uplifting
loading (Cp = -0.3) on the roof. sustains near-constant loading, irrespective of the
approaching wind angle. Differences do exist for different locations within a zone. In
each zone, the most critical wind angles are around the normal-to-wall direction for the
central portion; oblique winds seem to initiate the strongest suction near the corners.
From the standpoint of flow field over the roof, normal-to-wall winds create a separation
bubble whose size diminishes, away from the roof-axis, thus creating less suction at the
comers; oblique winds generate conical vortices, developing the strongest suction on roof
comers.
100

45


CD

30

@Y G) 0 90
,.
270
-(

13 ft (roof height)

15

@)

13 ft (roof height)


0

0 10 20 30

I 180

Figure 7.16 Division of pressure zones on flat roof


101
7.5 Features of Conical Vonices
Flow visualization using the tuft-grid method was performed to understand some
fundamental features of the conical vortices developed over a flat roof comer at skewed
incident winds. The grid tied with yellow yarn segments at its nodes was placed either
perpendicular to the roof-comer diagonal or perpendicular to the long wall. By
systematically moving the grid (placed perpendicular to the long wall) to desired
locations, positions of the vortex core at different planes parallel to the short walls can be
determined. The grid oriented perpendicular to the roof-comer diagonal to visualize the
pattern of the vortex pair. A sketch of the experimental setup is presented in Figure 7.17.

Long wall Short wall

Figure 7.17 Schematic view of flow visualization setup

7.5.1 Position of Vortex Core


The core of a vortex is a continuous three-dimensional curve., which can be
approximated by a series of discrete points. Three parameters are chosen to specify the
position of the vortex core. They are the location of the grid (y), the offset from the long
roof edge (x), and the displacement from the roof surface (z). Table 7.3 summarizes these
parameters for the vortex formed after flow separation at the long roof edge, for incident
wind along the roof comer diagonal (225}.
The discrete points and the curve fitting to them are displayed in Figure 7.18 to aid
envisioning. For a convenient planar description of the spatial curve, two curves are
102
given for the offset (x) and the displacement (z). It can be seen that a straight line
approximates quite well the change of the core displacement with distance measured
from the apex. For the offset of the vortex core from the edge, a linear fitting is suitable
up to about 12 ft from the apex, while a curve of the second order might be a better
approximation up to 20 ft, the range covered in this study. It must be pointed out that the
core position readings from visualization are subject to error in the order of 3 inches (half
of the grid division).

Table 7.3 Position of vortex core at 225 wind angle


Run# Grid location (y) Offset (x) Displacement (z) Mean wind
inches inches Speed., mph
M49nl60 6'-05" 12 6 20
M49nl69, 8' -07" 18 9 19
M49nl58 11 '-00"' 21 9 20
M49nl68 12' -05"' 24 12 20
M49nl61 15"-05n 27 15 16
M49nl67 18' -07"' 30 15 20
M49n162 19' -08"' 30 18 17

3.0 , - - - - - - - - - - - - -
x, z (feet)
offset\ ..... _ .........o .... .o

2.0
X =0.159y + 0.0353
.,....c.- ...
_... ....-ot-
"" x = -o.0064yz + o.2n4y - o.4841
0

1.0 _o..----~~
------o- 0

___ 0~
----o-
~ z = 0.0689y + 0.0937
I
a----
displacement
0.0 .___ _ ___.__.....___... _ _,____..._ _._____.__ ___._ _.._____...___...____.

5 10 15 y (feet) 20

Figure 7.18 Position of vortex core derived from flow visualization


103
Figure 7.18 indicates that the vortex-core displacement rises at a slope of 6.89% with
distance for the short roof edge., and has an initial displacement of 1.1 inches at x=O.
Close to the apex, the projection of the vortex core on the roof takes an angle of about 10
with the long edge.
Banks et al. (1997) performed flow visualization on cuboidal models (45 em to 120
em) in wind-tunnel generated non-boundary-layer-type turbulent flows. In their
experiment, glycerin smoke was introduced through holes near the leading comer, and a
laser light sheet was used to illuminate a vertical plane within the vortex core region.
They found, at 225 wind angle, the vortex core rose at a slope of 7%., and had an initial
displacement of 2 mm above the apex; the vortex core assumed an angle of 12 with the
roof edge. The slope and the angular position of the vortex core are quite comparable
with the findings of the current study.

7.5.2 Forms and Patterns of Vortex Pair


Based on the observations through flow visualization, forms of the conical vortex
and various patterns of the vortex pair combination are discussed in this subsection. For
easier understanding, a sketch defming the incident wind angle is shown in Figure 7.19.

/ Rightedge
It!' (short wall)
180

--
~ 210

t"
Left edge (long wall) j )"

225"
Figure 7.19 Schematics of vortices and incident wind angle
104

In general, vortices forming over the roof-corner take various sizes and forms, and
there are a variety of combinations (patterns) of the vortex pair, depending on the
incident wind angles. Here., the wind angles of concern are in the range of 180-270.
First., a single vortex is examined. The grid is set perpendicular to the long wall to
visualize the vortex developing from flow separation at the long roof edge. Presumably,
the vortex at the short roof edge is symmetrically similar. Figure 7.20 shows the vortex at
a wind angle of about 220, and Figure 7.21 at a wind angle of 210. The wind was
blowing from right to left in the figures.
A comparison of the two vortices indicates that at the larger wind angle (220,
Figure 7.20) the vortex is more oblong, and its core has a larger offset from the roof edge
and a smaller displacement from the roof surface. This is conceptually generalized and
depicted in Figure 7.22 based on extensive observations.
At a certain range of wind angle around 225 (roof comer diagonal), a pair of
vortices would form. To visualize the vortex pair, a grid of 14 ft long was orientated
nearly perpendicular to the diagonal of roof comer; the camera was set downstream of the
grid and the wind was blowing toward the viewer. This setup remains true for the pictures
presented later, unless otherwise specified. Figure 7.23 captures one of those instants.
Unfortunately., the wind velocity i:aata from the sonic anemometer were bad for that
data run, due to a wind speed overshoot. Wind gusts exceeding 30 m/s (60 mph) were
recorded on the meteorological tower at 13 ft, while the speed range for the sonic
anemometer was set at 30 rn/s. The wind angle was presumably about 230- 235 at that
particular instant of the picture, since the vortex on the right-hand side of Figure 7.23
(forming at the left roof edge, as defmed in Figure 7 .19) is slightly oblong shaped.
105

Figure 7.20 Vortex at the left edge (x =10'-06", 220, M49nl59)

Figure 7.21 Vortex at the left edge (x =6', 210, M49nl60)


106

Core position

Roof comer

Change of wind direction

2<Xt
Figure 7.22 Effect of wind angle on vortex shape and core position

Figure 7.23 A pair of near-circular vortices (M49n222~ at about 230-235j


107
It was mentioned earlier in this subsection that the interested range of wind angles is
180-270. However, some bounds of wind angles seem to exit for the formation of the
conical vortices. Beyond those two bounds, no recognizable conical vortices would be
present. The range of wind angles conducive to vortex formation is identified as
200-250. Figures 7.24 and 7.25 represent two near-bound situations. At these extreme
cases., only one quite oblong vortex formed at one of the edges, and straight flow passed
over the rest of the roof comer. This is because the wind direction was close to being
parallel to the other edge and there was no flow wash-up over this roof edge. A
conceptual sketch for the flow condition in Figure 7.24 (wind angle 200~ and the form
and dimension of the vortex and its relationship with pressure taps 50101. 50501 are
shown in Figure 7.26.

Figure 7.24 A big oblong vortex forming at the right edge (M49n220, 200~
108

Figure 7.25 A big oblong vortex forming at the left edge (M49n220, 250,

9.5 ft

50101 ~Left edtze (1om! wall)

9ft

1 50501

Right edge
(short wall)

Figure 7.26 Vortex and its relationship to pressure-tap location


109
A more common situation would be the flow condition between the two extreme
cases given in Figures 7.24 and 7.25, when two vortices would coexist. Starting at 200,
with increasing wind angles, a small, near-circular vortex would form along the left roof
edge (Figure 7 .19) and became increasingly oblong, while a relatively large and oblong
vortex formed along the right edge would shrink and became less oblong, until the wind
angle reached 225. At 225, a pair of vortices with comparable size and near-circular
shapes would take form. With still increasing wind angles beyond 225, the vortex along
the left edge enlarged in size and assumed oblong form, while the originally oblong
vortex along the right edge diminished in size and took shapes that were more circular.
This process would continue until the wind angle of about 250 was reached. This is the
other extreme case, symmetrical to what happened with the right edge at 200 wind
angle; only one oblong shaped vortex formed at the left edge.
Although the conical vortices are stable, high wind with small directional fluctuation
was needed for steady and strong vortices to occur. Vortices tend to break down as they
travel downstream; this seems to happen at a distance of about y =15ft from the apex. A
clearly recognizable vortex at x = 20 ft formed only with high wind speed (presumably
greater than 25 mph). The vortex observed at this location was usually less organized,
and its oscillation was at much reduced paces. The pressure coefficient data for tap 50123
indicate that the conical vortex is no longer the flow condition that induces critical
loading at that portion of the roof surface.
In this paragraph, the vortex-pressure correlation is revisited. Formerly, it was
argued that low pressures (high suctions) were a direct result of vortical circulation.
Recall that the fluctuations of pressure at tap 50101 (Figure 7.6) was characterized by
small fluctuations around a nearly constant value of about --0.5, which was sporadically
interrupted by brief moments of large fluctuation. This pattern was presumably associated
with the straight flow passing over the tap~ and occasional large vortices formed from
flow separation at one of the roof edges.. due to fluctuation in the wind direction.
Non-conventional pressure coefficent(M49n220) --.
2
..... --------------- .......... I

0 'II _ J i . J II I ~ . I_ 1. : I a . : J. ,1_ I
.
{5-2 ...__....
fs:i3:ss 1 --------...
-4 ~~~----~------------------------+---~~~~--~--~
I
I
I
6 I I
I
t f t
-----------------
I
I

360 370 380 390 400 410 420

-~j_r}cident_wind direction (M49n220l__ _ _____


90

60 I II w
1\ \iW I "' II .. I ... , - I
~
e:
0>
Q) I - I I
0 30 ''' -- ._flfll '\.E r I II

seconds
0 --------~--------~-----+~--------~------~--------
360 370 380 390 400 410 420

Figure 7.27 Wind direction and pressure coefficient --


0
lll

Flow visualization supports the above presumption. In Figures 7.24 and 7.25, two
big oblong vortices were observed. Figure 7.26 shows that the vortex (at wind angle of
200, was large enough to encompass tap 5010 L Figure 7.27 gives the pressures
coinciding with the vortex. Referring to Figure 7.27, at 8:13:12, the wind angle was
around 30 (equivalent to 240.. an angle favoring vortex formation at the left edge), a
(non-conventional) pressure coefficient peak occurred at tap 50101. Similar phenomenon
happened at 8:13:58 with a wind angle of 70 (equivalent to 200, an angle favoring
vortex formation at the right edge), except it was with the right edge.

7.6 Comparison of Wind Velocity Measured with Different Anemometers


In the study of wind loading effects on structures, the characteristics of incident
wind, especially the turbulence part of wind fluctuations, are very important,. Findings
made in Chapter 6 and this chapter underline the significant effects of both the
longitudinal turbulence and the lateral turbulence on wind loading.
At the WERFL of Texas Tech University, propeller-type UVW anemometers are
being used to measure wind velocity components; the direction is calculated from the two
horizontal component of wind velocity. Propeller-type anemometers usually do not have
adequate frequency response. Thomas (1996) compared the spectra of the longitudinal
and the lateral wind velocities. Figures 7.28 and 7.29 reproduce his results for velocities
measured at WERFL field site and CSU wind tunneL It can be seen in these two Figures
that the spectra from the UVW anemometer are deficient in the high-frequency end in
comparison with that from the sonic anemometer, mainly due to insufficient frequency
response of the UVW anemometers.
This brief comparative investigation intends to stress the point that the velocity
measurement with UVW anemometers is deficient. Instead of presenting analytical
results, direct comparisons of simultaneously measured segments of velocity traces are
made. The experimental setup was discussed in Chapter 5. These straightforward
comparisons highlight the inadequate frequency response of propeller-type anemometers.
112

0.1
.L
N
<:1
0.01
~:1
:1
rn
L
0..001

0..0001
0.00001 0.000] o.cxn 0.01 0.1 10
rzii.J (cycles]
_ Wind Tunnel J :100 Rll _ Field Sonic _rteldUVW
_ Field 3-Cup _ Univcnal
fia frequency [Hz). z [4m] is roof height.
U and Su "'2 arc mcau and variance.. respec::tively. of longitudinal velocity at z.

Figure 7.28 Comparison of longitudinal velocity spectra (after Thomas 1996)

-
..!...
N
(
0.1

~ 0.01
r::;

J
L
o.cxn

0.0001
0.00001 0.0001 0.001 0.01 0.1 1 10
r'z/U (cycles)
_ WindTunnell:lOORII _Field Sonic _FaeldUVW
_ Field ~-Cup - Universal
(is frequency [Hz], z [ 4mJ is roof beighL
U is mean longitudinal velocity at z, and Sv""" 2 is variance of lateral velocity at z.

Figure 7.29 Comparison of lateral velocity spectra (after Thomas, 1996)


113

In the experimental setup., the instrumental u-, v- and w-components of the UVW
anemometer and the sonic anemometer are consistent. Figures 7.30 compares the
instrumental u- v-components of the wind velocity. No attempt was made to fmd a mean
wind direction to decompose the velocity components accordingly to obtain the two
conventional velocity components in the horizontal plane. Comparisons of the w-
component of the wind velocity and the wind direction are shown in Figure 7 .31. The
most outstanding contrast is that high frequency fluctuations in all three velocity-
components and in the wind direction are missing from the UVW anemometer signals.
Peaks of short duration are left undetected by the UVW anemometer. The fact that the
two anemometers are set to sample at different rates, 10Hz for the UVW anemometer
and 30 Hz for the sonic anemometer, are not be responsible for the observed differences.
The fundamental cause is the retarded frequency response of the moving blades of the
UVW anemometer. Because of the inertia., the blades simply could not speed up or slow
down fast enough to keep up with rapid variations of wind speed. Occasionally.
overshoots of velocity peak would occur, due to abrupt changes of the wind direction.
Finally, it needs to be stressed that the differences observed represent the turbulence
part of natural wind fluctuations, and turbulence fluctuations are indeed important from
the viewpoint of wind loading effects on structures.

7.7 Summary
The major points discussed in this chapter can be summarized as follows.
1. In tenns of local flow-pressure relationship, a fundamental concept is that low
pressures (high suctions) on the roof comers are always related to strong
cornering vortices;
2. Wind angles favoring vortex-formation are bounded by 200 and 250;
3. Instantaneous pressure coefficients for tap 50501 could be satisfactorily predicted
from the mean pressure coefficient data and the incident wind properties. Pressure
prediction for tap 50101 was less successful, presumably because there are two
flow phenomena (straight flow and vortical flow) involved, and as a result, the
u-oornponent (sonic anemometer} v-component (sonic anemometer)
20 6
3
j 15
~..
ll:

0
gb 10 l:'
'Q -3
~ ~ -6
seconds
-9 ~~--~------~----~------~--~~
0 20 40 60 80 100 0 20 40 60 80 100

u-component (UVW anemometer) v-component (UVW anemometer)

20 6

8... 15 II) 3

0 ~ 0
- 10 'Q -3 I
b -- - 'f ~.,.. I
. 5
II t '"

, ~onds I
11.)
> 0 seconds l :: I . ' ' ' ',
0 20 40 60 80 100 0 20 40 60 80 100

Figure 7.30 Comparison of horizontal velocity components

-
~
w-component (sonic anemometer) Wind direction (sonic anemometer)

6
30
20.. J Ill
.., I . j
t. lnu

i 3 a
tn
10
.e.' 0
g
!' r
l -3
-6

seconds
l :~~ I .: ' ."
?> o
-30
' ' ' JC<Ondi
I

0 20 40 60 80 100 0 20 40 60 80 100

w-component (UVW anemometer) Wind direction (UVW anemometer)


6 30
20 ~----------------------------------fl

ii~. 03 n IV'~
rI\ ~~ n.r .~l".1 uLr\u:vvu "ldl>-c .. ,. . n"A.
l.J, -6 I
1 ' ' ' , ' , , scoonds I -30
0 20 40 60 80 100 0 20 40 60 80 100

Figure 7.31 Comparison of vertical velocity component and wind direction

-
U\
116

role of the mean pressure coefficient as a fundamental quantity in the pressure-


generation process is compromised.
4. The form and dimension of a single vortex and the patterns of combination of the
vortex pair varies considerably with the incident wind angle.
CHAPTERS
CONCLUSIONS AND SUGGESTIONS

This study set out to clarify the mechanisms of pressure generation for two important
flow phenomena significantly affecting the wind loading on low-rise building roofs: the
separation bubble and the conical vortices. Findings in this study have partially clarified the
mechanisms of pressure generation for these two flow situations.
In a retrospective view, much of the debate on wind-tunnel simulation priority has
failed to distinguish the effects incident wind has on two different types of quantities:
single events, e.g., minimum (peak) pressures, and statistics, such as the mean and the rms
pressure coefficients. The turbulence intensities, reflecting the gust structure and the
directional fluctuations of the free-stream wind, might bear significant influences
collectively on the pressure statistics. It is the intensity of individual longitudinal gusts that
decides the data-run-wise peak-pressure coefficients on the roof corner, while the peak-
pressure coefficients associated with the SB are governed by the lateral directional
fluctuations in the incident wind. Proper simulation of the incident wind profile (boundary-
layer type) is probably the most important single input in the wind- structure interaction
process.
The major contributions from this study to the understanding of the wind flow and
the related loading effects on low-rise buildings are summarized as follows.

8.1 About Separation Bubble


1. The overall SB is oblong and elongated in the horizontal direction. Despite the
conceivably random and turbulent nature of the airflow inside the SB, the averaged
structure of the SB is amazingly organized and regular to small details.
2. The mean point of reattachment is approximately 10 ft from the leading roof edge,
and the mean height of separation bubble is about 4.5 ft.

117
118
3. The pressure distribution on the roof surface underneath the SB is directly related
to the structure of the SB, with the uplifting pressure on the roof generally
decreasing with distance from the leading edge.
4. The introduction of the non-conventional pressure coefficient makes it possible to
separate the effect of wind speed from that of the wind direction on pressure
generation.
5. Time series analysis indicates that fast-and-large fluctuations of the incident wind
direction govern the mechanism of peak-pressure generation.
6. The primary peaks of the conventional pressure coefficient are often outcomes of
combined wind direction fluctuations and wind gusts.
7. The non-Gaussian pressure zone is found to extend approximately 15 ft from the
leading edge, and this dimension is comparable to the overall longitudinal size of
vortical circulation and the height of the test building.
8. It is not very clear as to what was happening with the flow inside the SB at those
moments of pressure peaks, except that the flow was quite three-dimensional

8.2 About Conical Vortices


1. Low pressures (high suctions) on the outside surface of a roof comer are always
related to strong vortices.
2. Wind angles favoring vortex-formation are bounded by 200 and 250.
3. Instantaneous pressure coefficient for taps (such as tap 50501) underneath either of
the two vortices could be satisfactorily predicted from the mean pressure
coefficient data and the incident wind property.
4. Pressure prediction for the symmetrically located tap (50 101) was less successful,
presumably because there were two flow phenomena (straight flow and vortical
flow) involved, and as a result, the role of the mean pressure coefficient as a
fundamental quantity of the wind-structure interaction process was compromised.
119

Predictions underestimate the pressure peaks while over estimating the mean
pressure.
5. The form and dimension of a single vortex and the pattern of combination of the
vortex pair varied considerably with the incident wind angle.
A practical means of wind load prediction is established. It is a "hybrid" approach that
combines wind-tunnel simulated mean pressure coefficients with easily available
information of site-specific wind data.

8.3 Suggestions for Further Study


It was discovered that the mechanism of pressure generation in the case of SB was
more involved than that of the conical vortex situation. Further work is needed to quantify
the incident-wind/pressure correlation associated with the SB. Simultaneous multiple-
point measurement of the wind flow inside the SB along with the pressure could lead to a
better understanding of what is going on flow-wise, when pressure peaks occur.
In the conical vortex case, quantification of the incident-wind I pressure relationship
for tap 50101 demands more research. Pressure distributions on the roof comer and their
relationship with the structure (core position, shape and size, etc.) of the vortices can be
investigated by installing more rays of closely spaced pressure taps at the roof comer.
Wind flow inside the vortical circulation should be examined quantitatively. Similar studies
on other building geometries might be interesting.
In light of the importance of documenting natural wind't especially the turbulence part
of wind fluctuations, anemometers with sufficient frequency-response should be used. The
worst-case suction on the roof-comer under the conical vortices remains to be identified in
the full-scale study.
REFERENCES

ASCE., 1995: Minimum Design Loads for Buildings and Other Structures., ASCE7-95,
American Society of Civil Engineers.

Akins, R.E., Peterka, J.A., and Cermak, I.E., 1977; "Mean Force and Moment
Coefficients for Buildings in Turbulent Boundary Layers/' Journal of Wind
Engineering and Industrial Aerodynamics, Vol 2'1 pp195-209.

Atkins, R.E., 1992: "Effects of Turbulence on Mean Force and Momentum Coefficients,"
Journal ofWind Engineering and Industrial Aerodynamics, Vol 41, pp701-712.

Bachlin, W . ., Plate, E.J.'I and Kamarga., A.., 1983: "Influence of the Ratio of Building
Height to Boundary Layer Thickness and of the Approach Flow Velocity Profile
on the Roof Pressure Distribution of Cubical Buildings," Journal of Wind
Engineering and Industrial Aerodynamics, VoL 11, pp63-74.

Bailey, A., 1933: "Wind Pressures on Buildings.,n Inst. Civil Eng., selected papers, No.
139.

Bailey, A.., and Vincent., N.D.C., 1943: 'Wind Pressure Experiments at the Severn
Bridge," Journal of Institution of Civil Engineers, VoL 11'1 pp363-.

Banks, D., Meroney, R.N., Sarkar, P.P., and Zhao, Z.'l 1997: ''Visualization of Flow and
Separation about a Low Rise Building: Comer Vortex Zones," submitted to the
Journal of Wind Engineering and Industrial Aerodynamics.

Campbell, J.A, 1995: "Wind Engineering Research Field Laboratory Site


Characterization,, Master's thesis, Department of Civil Engineering, Texas Tech
UDiversity.

Cherry, N. J.., Hillier, R. ., and LaTour M.E.M.P, 1984: "Unsteady Measurements in a


Separated and Reattaching Flow.," Journal of Auid Mechanics, VoL 104, pp123-
146.

Chiu, G.L.F., Perry'l D. C., and Chiu, A.N.L., 1993: "Structural Performance in Hurricane
Iniki.,, Proceedings of the 7th US National Conference on Wind Engineering,
Vol!, pp155-l61, University of California., Los Angeles.

Cochran, L. S., and Cermak, J.E .., 1992: uFull- and Model-scale Cladding Pressure on the
Texas Tech University Experimental Building," Journal of Wind Engineering and
Industrial Aerodynamics~ VoL 41-44, pp 1589-1600.
120
121

Coc~ L.S., 1992: uWind-tunnel Modeling of Low-rise Structures.," Ph.D. dissertation,


Department of Civil Engineering., Colorado State University.

Cook., N. J .., 1985: The Designer"s Guide to Wind Loading of Building Structures. Part I:
Background. Damage Survey, Wind Data and Structural Classification,
Butterworths., London, UK.

Cook., N. J., 1990: The Designer's Guide to Wind Loading of Building Structures. Part II:
Static Structures, Butterworths, London., UK.

Dalley, S., 1993: "''Investigation into the Simulation Priorities for Wind Tunnel Tests on
Low-rise Buildings," Proceedings of First International Conference on Wind
Engineering. European and African Regional Conference, pp95-104.

Gartshore, I.S., 1973: "'The Effects of Free Stream Turbulence on the Drag of Rectangular
Two-dimensional Prism," University ofWestern Ontario, Canada BLWT-4-73.

Hiller, R., and Cherry, N. J., 1981: 'The Effects of Stream Turbulence on Separation
Bubbles,, Journal of Wind Engineering and Industrial Aerodynamics, Vol 14,
pp375-386.

Hoxey, R., Robertson, A., Richardson G., and Short, L., 1996: "Correction of Wind-
tunnel Pressure Coefficients for Reynolds Number Effects,u Abstract, the Third
International Colloguium on Bluff Body Aerodynamics and Applications," ppA-
VI-1, Virginia Tech and State University.

Jamieson, N.J., and Carpenter, P., 1993: "Wind Tunnel Pressure Measurements on the
Texas Tech Building at a Scale of 1:25 and Comparison with Full-scale,n
Proceedings of the 7th US National Conference on Wind Engineering, Vol 1,
pp303-312, University of California, Los Angeles.

Jenson, M., 1958: 'The Model Law for Phenomena in Natural Wind," Ingenicpren
(International Edition), Vol 2, ppl21-128.

Kawaii, H., and Nishimura., G., 1996: "Characteristics of Fluctuating Suction and Conical
Vortices on a Flat Roof in Oblique Flow," Journal of Wind Engineering and
Industrial Aerodynamics. Vol60, pp211-225.

Kind, R.J., 1986: 'Worst Suctions Near Edges of Flat Rooftops on Low-rise Buildings,"
Journal of Wind Engineering and Industrial Aerodynamics, VoL 25, pp31-47.
122
Kind7 R.J., and Wardlaw, R. L., 1982: uFailure Mechanism of Loose-Laid Roof-Insulation
Systems," Journal of Wind Engineering and Industrial Aerodynamics, VoL 9,
pp325-34L

Kirrane, P.P., and Steward S.J., 1978: "The Effect of Blockage on Shear Flow in the
Wind Tunnel,"' Department of Aeronautical Engineering7 Thesis 221, University of
Bristol

Kumar, K.S., and Stathopoulos T., 1997: 'tpower Spectra of Wind Pressures on Low
Building Roofs," Second Europe and Africa Conference on Wind Engineering,
ppl069-1076, July 1997, Genoa, Italy.

Letchford, C.W., 1995: "Simultaneous Flow Visualization and Pressure Measurements on


the Texas Tech Building," Proceedings of the 9th International Conference on
Wind Engineering, pp524-535, New Delhi, India.

Levitan.. M.L., Holmes J. D., Mehta K.C., and Vann W. P., 1989: "Field Measured
Pressures on the Texas Tech Building,'' Proceedings of the 8th Colloquium on
Industrial Aerodynamics, ppl3-24, Aachen., West Germany, September 4-7. 1989.

Levitan, M.L., and Mehta, K.C., 1992a: "Texas Tech Experiments for Wind Loads Part 1:
Building and Pressure Measuring System," Journal of Wind Engineering and
Industrial Aerodynamics, Vol 41-44, ppl565-1576.

Levitan, M.L., and Mehta, K.C., 1992b: ~'Texas Tech Experiments for Wind Loads Part
II: Meteorological Instrumentation and Terrain Parameters," Journal of Wind
Engineering and Industrial Aerodynamics, VoL 41-44, pp1577-1588.

Levitan, M.L., Cochran, L., and Norville, H.S., 1993: ~'Damage Assessment of Hurricane
Andrew in Louisiana. " Proceedings of the 7th US National Conference on Wind
Engineering, Vol 1, pp393-402 . University of California, Los Angeles.

Lin, J.X., Surry, D. and Tieleman, H.W., 1995: 'The Distribution of Pressure Near Roof
Corners of Flat Roof Low Buildings," Journal of Wind Engineering and Industrial
Aerodynamics., VoL56, pp235-265.

Marwood., R., 1996: "An Investigation of Conical Roof Edge Vortices," Ph.D.
dissertation, Lincoln College, University of Oxford.

Mehta, K. C., 1984: "Wind Induced Damage Observations and Their Implications for
Design Practice," Journal of Engineering Structures., ASCE, Vol 6, 1984, pp242-
247.
123

Melbourne, H.W.~ 1975: "The Relevance of Codification to Design,'' Proceedings of the


Fourth International Conference on Wind Effects on Buildings and Structures,
Cambridge, Cambridge University Press, pp785-790.

Melbourne, W.H., 1980: 'Turbulence Effects on Maximum Surface Pressures - A


Mechanism and Possibility of Reduction,u Wind Engineering, J.E. Cermak (Ed.),
pp54l-551, Pergamon Press, Oxford.

Melbourne, W.H.~ 1992: 'Turbulence and the Leading Edge Phenomenon," Proceedings
of the 2nd International ColloQuium on Bluff Body Aerodynamics and
Applications, Melbourne, Australia.

Minor, J.E., and Mehta K.C., 1979: "Wind damage Observations and Implications,"
Journal of the Structural Division, ASCE .. November, 1979., pp2279-229l.

Minor, J.E., 1981: "Wind Effects on Buildings.,H The Thunderstorm in Human Affairs, E.
Kessler (Ed.), pp89-109, University of Oklahoma Press, Norman.

Mochida, A, Murakami, S., Shoji, M., and Ishida, Y., 1993: 'Numerical Simulation of
Flow-field around Texas Tech Building by Large-Eddy Simulation/' Journal of
Wind Engineering and Industrial Aerodynamics, VoL 46-47, pp455-460.

NBCC, 1995: "User's Guide - NBC 1995, Structural Commentaries (Part 4). National
Research Council Canada, Ottawa, Canada.

Panofsky, H.A., and Dutton J. A, 1984: Atmospheric Turbulence- Models and Methods
for Engineering Anplications, John Wiley & Sons, New York. NY.

RofaiL A. W., 1995: "Full-scale/Model-scale Comparisons of Wind Pressures on the TIU


Building," Proceedings of the 9th International Conference on Wind Engineering,
ppl055-1065, New Delhi, India.

Saathoff, P.J., Melbourne, W.H .., 1987: "Free-stream Turbulence and Wind Tunnel
Blockage Effects on Stream-wise Surface Pressures~" Journal of Wind Engineering
and Industrial Aerodynamics, VoL 26, pp353-370.

Sachs, P., 1978: Wind Forces in Engineering, 2nd Ed., Pergamon Press, Elmsford, NY.

Sandri, P., 1992: ''Frequency Response Requirements for Measuring Fluctuating Wind
Pressure on Buildings," Master's thesis, Department of Civil Engineering, Texas
Tech University.
124

Sarkar, P.P., and Zhao .., Z., 1997: ~ow Visualization and Measurement on the Roof of
the Texas Tech Building.," to be published in the Journal of Wind Engineering and
Industrial Aerodynamics.

Savory, E., and Toy., N.., 1996: ''Reynolds Number Effects on Wind Loads on Low Rise
Buildings," Abstract.. the Third International Colloguium on Bluff Body
Aerodynamics and Applications.,"' ppA-VI-5, Virginia Tech and State University.

Selvam, P.P .., and Konduru., P.B., 1993: "'Computational and Experimental Roof Comer
Pressures on the Texas Tech Building," Journal of Wind Engineering and
Industrial Aerodynamics, Vol 46-47., pp449-454.

Sheffield, J.W., 1993: ''A Survey of Building Performance in Hurricane Iniki and Typhoon
Omar.," Proceedings of the 7th US National Conference on Wind Engineering,
Vol 2, pp653-662, University of California, Los Angeles.

Simiu, E., and Scanlan, R. H., 1996: Wind Effects on Structures: Fundamentals and
Applications to Design, 3rd Ed., John Wiley & Sons Inc., New York, NY.

Smith, T.L., and McDonald, J.R., 1990: "Roof Wind Damage Mitigation: Lessons from
Hugo.,"' Department of Civil Engineering, Texas Tech University.

Smith, J.H.B., 1984: "Theoretical Modeling of Three-dimensional Vortex Flows in


Aerodynamics," Aeronaut. J., (1984) pplOI-116.

Stathopoulos, T., Davenport, A.G., and Surry, D., 1978: "The Assessment of Effective
Wind Loads Acting on Flat Roofs.," Proceedings of the 3rd Colloguium on
Industrial Aerodynamics, Part 1, pp225-239.

Stathopoulos, T., and Surry, D., 1983: "Scale Effects in Wind Tunnel Testing of Low
Buildings," Journal of Wind Engineering and Industrial Aerodynamics, Vol.l3,
pp313-326.

Surry, D., 1989: "Pressure Measurements on the Texas Tech Building-II: Wind Tunnel
Measurements and Comparison with Full Scale," Proceedings of the 8th
Colloquium on Industrial Aerodynamics, Aachen, West Germany, September 4-7,
1989, pp25-35.

Thomas, G., 1996: '1denti.fication of Transter Functions for Wind-induced Pressures on


Prismatic Buildings," Ph.D. dissertation, department of Civil Engineering, Texas
Tech University, Lubbock, Texas.
125

Tieleman~ H.W., 1992: uProblems Associated with Flow Modeling Procedures tor Low-
rise Structures~" Journal of Wind Engineering and Industrial Aerodynamics, Vol
41-44, pp923-934.
4
Tieleman, H ..W., Surry, D., and Lin, J.X., 1994: 'Characteristics of Mean and Fluctuating
Pressure Coefficients under Comer (Delta Wing) Vortices," Journal of Wind
Engineerin& and Industrial Aerodynamics, Vol 52, pp263-275.

Tieleman, H.W., 1995: "Simulation of Surface Winds for Assessment of Extreme Wind
Loads on Roofs," Proceedin&s of the 9th International Conference On Wind
Engineering, pp 1162-1169, New Delll India.

TrU/CSU, 1995a: "Five Years of Achievement Under the CSU!ITU Cooperative


Program in Wind Engineering: An Executive Summary Report."

TTU/CSU, 1995b: "CSUITTU Cooperative Program in Wind Engineering: Progress


Report for Technology Assessment and Advisory Council," October 13-14, 1995,
Fort Collins, Colorado.

Van der Hoven, 1957: "'Power Spectrum of Horizontal Wind Speed in the Frequency
Range from 0.0007 to 900 Cycles per Hour.," Journal of Meteorology, Vol 14.
ppl60-164.

Vognild., R.A, 1993: "Are Building Codes Adequate for Hurricane Protection?"
Proceedings of the 7th US National Conference on Wind Engineering, Vol. 2,
pp833-838., University of California, Los Angeles.
4
Wagaman, S. A.., 1993: 'Full-scale Flow Visualization over a Low-rise Building,"
Master's thesis, Department of Civil Engineering., Texas Tech University,
Lubbock, TX.

Woo., H.G.G., Peterka, J. A., and Cennak J.E., 1977: "Wind-tunnel Measurements in the
Wakes of Structures,'' NASA Contractor Report NASA CR-2806, National
Aeronautics and Space Administration.

Yin., J.M., 1996: uProbability Study of Wind Pressures on Buildings," Ph.D. dissertation.
Department of Civil Engineering, Texas Tech University. Lubbock, TX.
APPENDIX A

TECHNICAL SPECIFICATION OF THE SONIC ANEMOMETER

A.l Wind Measurement


Measurement rate: 100 s (Full3-axis measurement)
Data output rate: from 0.4 to 100 s
Wind speed range: 0 to 45 mls
Wind speed accuracy: <1 % nns
Wind speed resolution: 0.01 rn/s
Wind speed offset: 0.0 lm/s
Directional Accuracy: <:: l o nns
Directional resolution: 1o

A.2 Environmental
Operating temperature: -40C to +60C
Storage temperature: -50C to +75oC
Relative humidity: 5 % to 100 %
Precipitation: up to 300 mmlhr
Altitude: 0 to 3000 m
Moisture ingress: IP65

More details are available in the user manual.

126
APPENDIXB
DATA ACQUISffiON LOO

B.l General Guide


This appendix contains general information about data Runs M49n001- M49n287 for
the Mode-49 data acquisition configuration. The Mode-49 data acquisition was setup
from some modifications of the Mode-48 configuration to incorporate three channels of
analog signal from the output of the three-component sonic anemometer. These three
channels correspond to the instrumental u- v- and w-components~ and are stored in three
columns of file M49XXXS.PRN in ASCII format. Generally. three types of wind velocity
data were collected using the sonic anemometer. One is for local wind flow measured
within the separation bubble; the other two measured incident wind velocity either close to
the meteorological tower or at positions close to the test building. Specific information
about the locations of the sonic anemometer and the test building orientation is given for
each successful data run. Beside the velocity measured by the sonic anemometer, data
collected by Mode-49 contained infonnation of wind velocity at 6 levels on the
meteorological tower and pressures at those active taps., data routinely collected for a data
acquisition mode.

B.2 Sampling Frequency


The (non-integer) sampling frequencies of 7.5 Hz and 22.5 Hz were not intentional
They resulted from the replacement of the hard drive after a crash. The software was not
modified accordingly to match the hardware configuration.
Sonic Anemometer. The sampling frequency for runs M49n00 1-055 was 7.5 Hz;
there were 9(X)() data points for each 20-minute run duration. The sampling frequency for
data runs M49n056 - 125 was 22.5 Hz; there are 27,000 data points for each 20-minute
run duration. Wind velocities were sampled at 30Hz for runs M49nl26- 287; these runs
were of 15-minute duration and each contains 27,000 data points.

127
128
Anemometers on the Meteorological Tower. The sampling frequencies were 7.5
Hz and 10Hz for runs M49n001- 125 and runs M49n126-287~ respectively.
Pressures. The sampling frequencies for pressure data were 22.5 Hz and 30 Hz for runs
M49n001-l25 and runs M49nl26-287~ respectively.

B.3 Sonic Anemometer Data Acguisition Log


Information about the location and orientation of the sonic anemometer for those
successful runs is provided in this appendix. The missing data run numbers were aborted
runs and can be ignored.
Data Runs M49n00l-l56, M49nl96-219, M49n252-272 were for local flow
measurements within the separation bubble. Some of those runs were also used as incident
wind data because they measured the incident wind after correction.
For data runs M49n158-169, the sonic anemometer was placed about 5 ft above the
leading roof corner to measure the incident wind speed and direction, along with tlow
visualization experiments for a single cornering vortex. Runs M49n220-225 had the same
setup for the sonic anemometer except that the grid was oriented perpendicular to the roof
diagonal to visualize two vortices. M49n226 visualized the SB.
Data runs M49nl70-195 were automatic data runs with a building position of 165.
These runs had wind coming at angles around 225.
Data runs M49n231-250 were designed for comparative measurements of upstream
wind velocities by the sonic anemometer and the uvw-anemometer at 13ft.
The three instrumental velocity components obey the right-hand-rule (Figure B.l).

v
Figure B.l Coordinate system
u
129
In the following tables7 BLK is the number of firings averaged by the anemometer.
The basic sampling rate for wind velocity is 10ms. The desired output (sampling)
frequency is determined by a user-specified integer value between 1 and 250. BLK=5
indicates an output frequency of 1000 ms I (5x 10 ms) =20 Hz. 'Range' represents the
maximum axial wind speed corresponding to 2.5 Volts analog output. Here a linear
relationship applies. 'x' is the horizontal distance of the sonic anemometer from the
leading roof edge; .. h' is the height of the sonic anemometer from the building roof. 'H'
and 'V" in the column 'sonic' denote the anemometer orientation being horizontal (H) or
vertical (V). 'Bldg' specifies the orientation of the test building; 'U' is the approximate
run-averaged upstream wind speed at 13ft on the meteorological tower.
The positive wind velocities of the instrumental u- v- w-components are indicated in
the following graph for runs M49n00l-156, M49nl96-219.
The following definition of u- v- and w-components are good for runs
M49n158-169, M49n220-226 (Figure B.2).

U+ ~V+ Wmd
~

Figure B.2 Definition of instrument u- v- and w-components (M49n220-226)


130

For other data runs, the definitions are given wherever convenient along with other
information. The sonic anemometer was in operation for automatic data runs
M49nl70-195.
131

Tables B.l - B 11 list details of experimental setup for data runs M49n008-156.

Table B. I Sonic data acquisition log (2115/1997)


Run# BLK Range X b sonic Bldg. start Uympb
M49n008 5 30m/s 0'..()5" o -06., H 315 12:27:06 -
M49n009 . .. .. 12:47:34 -
.. ..
n n

.. ..
>Y

M49n011 1'-04,. 13:14:38 -


.. .. .. .. ..
n

M49n012 1'-09" 14:06:34 15


M49n013 .. .. .. 1t .()()" .. .. 14:43:25 19
M49n014 .. 1'..00... 1' .()()'. .. ... 15:08:26 15
.. ..
n

M49n015 .. , 2' ..()0" .. 15:33:05 13


M49n016 .. .. 3'..00.. .. .. 15:59:28 11
.. .. .
tY

M49n017 .. ., 4' -<xr 16:26:22 12


M49n018 .. .. .. . .. .. 16:49:11 11
M49n019 .. .. 5'..00.. .. .. .. 17:14:40 11
M49n020 .. .. 6' ..()0" .. .. .. - 10

T a bleB .2 S oruc
. d ata acquJStt.J.on o~ {2116/1997)
Run# BLK Range X b sonic Bldg. start U,mpb
M49n022 5 30mls 7'..00.. 1' .()()" H 285 16:36:09 13
M49n023 .. .. 8'..00" .. .. .. 16:59:32 12
M49n024
M49n025
..
..
.... 9'-00,.
10'..()*' ..
'tt ..
..
..
..
17:23:00
17:49:08
12
12

Table B.3 Sonic data acquisition loj (2117/1997)


Run# BLK Range X h Sonic Bldg. start U,mph
M49n026 5 30mls 15' 0' ..()6H H 285 10:51:41 16
M49n027 .. .. .. .. .. 11:14:30 18
.. ..
n

M49n029 .. ., 14' .. 11:43:28 16


M49n030 . .. 13' .. .. .. 12:06:26 15
M49n031 . .. 12' .. . .. 12:37:03 -
M49n032 . .. .. .. . .. 12:58:42 -
M49n033 .. .. .... .. .. .. 13:20:22 -
M49n034 .. .. .. t i .. .. - -
M49n035 .. .. 11' .. .. .. 14:06:52 17
M49n036 .. .. 10' .. ... .. 14:29:12 16
M49n037 .. .. 9' .. .. .. 14:52:09 17
M49n038 .. .. 8' .. .. .. 15:18:30 -
M49n039 .. .. .. .. .. .. 15:40:09 -
M49n040 .. .. .. .. .. .. 16:01:49 -
M49n041 .. .. .. .. .. .. 16:23:29 -
M49n043 .. T .. . . 17:41:32 13
.. ..
n

M49n044 .. 6' .. . 18:04:24 13


132

Table B.4 Sonic data acquisition loJ (212111997)


Run# BLK Ran2e X h Sonic Bldg. start U.mph
M49n045 5 30mls 5~ 0"-06 .. H 105 10:23:51 19
M49n046 ... 4. .. .. . 10:47:02 18
.. .. ..
19

M49n047 ... 3~ 11:09:56 17


.. .. ..
n

M49n048 .. 2' .. 11:32:31 18


M49n049
M49n050
..
..
.... 1'
0'
..
1' -08"
,.
..
.
..
11:54:56
12:20:21
17
18
M49n051 .. .. 1' .. .. .. 12:50:31 17
M49n052 .. .. .. .. .. .. 13:13:26 17
M49n053 .. .,. .. .. .. 13:36:18 -
.. ..
n

M49n054 .. 2' .. .. 14:04:39 14


M49n055 .. n 3' .. .. .. 14:23:40 15
Sonic data were sampled at 7.5 Hz. for Runs M49n008- M49n055; T=20 minutes, 9000 data
points.
Pressure data were sampled at 22.5 Hz for Runs M49n008 - M49n055; T=20 minutes, 27000 data
points.

Sonic data were sampled at 22.5 Hz for Runs M49n058 - M49n 125; T=20 minutes, 27000 data
points
Pressure data were sampled at 22.5 Hz for Runs M49n058 - M49n 125; T=20 minutes, 27000 data
points
M49n058 ., 4' .. .. .. 16:37:29 14
..
u
M49n059 .. s .. .. .. 16:59:56 12

T ableB .5 Sorne
. d ata aCQUJStUon 01 (2123/ 1997)
Run# BLK Ran2e X h Sonic Bldg. start u.mph
M49n060 5 30m/s 6' 1'-08" H 240 12:56:29 18
M49n061 .. .. T .. .. 13:21:07 20
M49n062 . .. . .." . 13:43:57 17
M49n063 .. .. . .." .. .. 14:06:49 17
M49n064 .. .. s . .. . 14:29:42 18
M49n065 .. .. 9' . .. .. 14:52:34 19
M49n066 .. .. 10" .. .. . 15:15:27 18
M49n067 ..
..
..
..
11' .. ..
..
.. 15:38:19 19
M49n068 12. .. .. 16:01:12 14
M49n069 .. .. 13' .. .. .. 16:24:29 ..
M49n070 .. .. 14' .. .. . 16:47:21 ..
M49n071 . .. 15' .. .. .. 17:10:13 15
M49n072 .. .. oo 2'-03" .. .. 17:47:32 17
133
Table B.6 Sonic data acquisition lo! (2/26/1997)
Run# BLK Range X h Sonic Bldg. start U~mph
M49n073 5 30m/s o~ -06'' 2"-03 .. H 15 14:20:07 19
M49n074 .. . 1' .,. .. .. 14:42:58 16
M49n077 .. ... 2' ., . .. 16:18:55 15
M49n078 .. .. .. .. .. 16:41:37 17
. .. .
19

M49n079 , ... n 17:04:19 17

Table B.7 Sonic data acquisition lo~ (2/27/1997)


Run# BLK Range X h Sonic Bldg. start U,mpb
M49n081 5 30m/s 2' 2'-03" H 270 11:55:56 13
M49n082 .. .. 3' ..,., .. .. 12:30:37 .,..
M49n083 .. ... . .. .. 12:56:59
M49n084 .. .. 4' .. . 13:19:43 15
M49n085 . . s .. ..
"
.. 13:42:26 16
M49n086 .. .. .. .. . ,. 14:05:17 16
M49n087 .. . .. .. .. .. 14:29:14 15
M49n088 .. .. .. .. . . 14:51:57 15
M49n089 ,., .. .. .. .. .. 15:16:08 16
M49n090 .. .. .. .. .. . 15:38:51 15
M49n091 .. .. 6' .. .. .. 16:01:33 17
M49n092 . .. .. .. 240 16:24:58 17
M49n094 .."' .. 7' .. .. .. 16:54:26 ..
M49n095 .. .. 8' .. .. . 17:17:08 16
M49n097 .. .. 9' .. .. . 17:43:40 15
M49n098 .. .. 10' .. .. .. 18:13:02 ..
T ableB .8 Soruc
. data acquiSition o~ (2/28/1997)
Run# BLK Range X b Sonic Bldg. start U,mpb
M49n099 5 30m/s 11' 2'-03" H 300 15:00:11 19
M49n100 .. . 12' .. .. .. 15:22:53 20
M49n101 .. .. 13' .. .. . 15:45:39 21
M49n103 . . 14' .. . .. 16:13:22 18
M49nl04 .. .. 15' .. .. 315 16:36:04 21
M49nl05 .. . 0' 3'-00" v .. 17:07:31 18
M49nl06 .. . 0'-06' .. .. . 17:30:14 19
M49n107 .. .. 1' .. .. .. 17:52:58 18
M49n108 . 2' .., .. .. 18:15:42 17
M49n109 . .." 3' .. .. 18:38:26 20
134
T ableB .9 Somc
. data acQwsttion 1o~ (3/03/1997)
Run# BLK Ran~e X h Sonic Bldg. start U,mph
M49n112 5 30m/s 4' 3'-00,. v 30 10:31:19 21
M49nll3 5' ... .. 10:54:02 21
.. .. .. .
n n

..
n

M49n114 6' 11:16:45 20


M49n115 . .. .. . . .. 11:39:29 16
M49n116 .. .. 7' ., . .. 12:02:14 17
M49n117 .. .. 8' .. . . 12:24:58 16
M49n118 .. .. . . . . 12:47:42 16
M49nll9 .,. .. 9' .. .. .. 13:10:26 16
M49n120 . .. 10' .. .
..
.. 13:33:10 13
M49n121 .. .. 11' .. 15 13:56:44 12
M49n122 .. .. 12' .. . .. 14:21:14 14
M49nl23 .. .. .. .. .. ..
..
14:44:16 14
M49nl25 .. .. 13' .. .. 15:28:59 13
Sonic data were sampled at 22.5 Hz for Runs M49n058 - M49nl25; T=20 minutes, 27000 data
paints.
Pressure data were sampled at 22.5 Hz for Runs M49n058- M49nl25; T=20 minutes, 27000 data
paints.
Runs M449n058 - M49nl25 were sampled at 22.5 Hz. for both pressure and sonic. With total data points
of 27000, the run duration was 20 minutes.
135

Runs M49n126 and after were sampled at 30 Hz for both the pressure and the sonic
anemometer. With total data points of 27()(X), the run duration was 15 minutes.

T ableB.10 Somc
. data acqwstnon 1og (3/06/1997)
Run# BLK. Range X h Sonic Bldg. start U,mph
M49n126 5 30mls 13. 3"-00'" v 270 15:50:50 15
M49n127 .. 14' ... . .. 16:09:16 16
M49nl28
"
.. 15 .. .. 16:27:45 16
M49n130 .." . o 1'-04"
"
H .. 17:55:31 15

Runs M49n 131 - 143 are for flow visualization of conical vorte~ but the video turned out not clear

T ableB .11 Somc


. data acqwstnon Iog (3/09/1997)
Run# BLK Range X h Sonic Bldg. start U,mph
M49nl44 5 30m/s o 1'..()4" H 120 14:17:27 19
M49n145 .. . ., .. .. .. 14:35:23 21
M49n146 .. .. r .. .. . 14:53:16 19
M49n147 .. .. 2' . .. .. 15:12:38 19
M49nl48 .. .. 3' .. .. .. 15:30:30 19
M49nl50 .. .. 4' .. .. 135 15:58:37 19
M49n151 .. .. s .., .. . 16:16:30 18
M49n152 .. . 6 .. 16:34:22 17
M49nl53 .. .. 7' .. .. .". 16:52:58 16
M49n154 .. . 8' .. .. .. 17:10:52 15
M49n155 .. .. 9' .. .. .. 17:29:02 14
M49n156 .. .. 10' .. .. .. 17:48:16 14
136

3/13/1997
Runs M49nl58 - 169

The 3-D sonic was placed above the leading roof corner to measure the direction I
magnitude of the upstream wind. Conical vortex was visualized at different location by
moving the grid tied with yarns. Details of experimental setup are given in Table B.l2.
Figure B.3 is a schematics view of the locations of the sonic anemometer and the grid.

Table B.12 Sonic data acquisition log (3/13/1997)


Run# Grid location, x Bldg. position start end
M49n158 10'-06"' 195 19:57:23 20:12:36
M49n159 180 20:16:38 20:31:51
"
M49n160 6'-00H . 20:38:10 20:53:21
M49n161 15'-00" .. 20:57:38 21:12:50
M49n162 19'-03.. 150 21:19:10 21:34:22
M49n164 25'-03"' 21:57:27 22:12:39
"
M49n166 22'-03 ... 165 22:28:55 22:44:07
M49n167 18' -02... . 22:47:15 23:02:27
M49n168 12' -00... .. 23:05:43 23:20:55
M49n169 8'-02.. 23:24:44 23:39:56
"

Wind~
Grid location

(PLAN VIEW of the ROOF CORNER)

Figure B.3 Locations of sonic anemometer and grid {M49n158- 169)

Runs M49nl70-195 were automatic data runs with bldg. position of 165.
137

Tables B.l3- 816 list details of experimental setup for data runs M49nl96-219.

Table 8.13 Sonic data acquisition log (3/16/1997)


Run# BLK Range X h Sonic Bldg. start U.m_ph
M49n196 5 30m/s 1r r-04"* H 270 15:00:47 -
M49n197 .. . .. . , ... 16:10:27 14
M49n198 .. .. , .. 16:28:20 14
.. ..
n

..
n

M49n199 .. 12' , 16:46:13 14

Table 8.14 Sonic data acquisition log (3/18/1997)


Run# BLK Range X h Sonic Bldg. start U,mph
M49n200 5 30m/s 13' 1'-04" H 105 09:54:54 22
M49n201 .. .. 14' .. .. .. 10:12:46 25
M49n202 .. . 15' .. .. ... 10:36:32 24

Run#
M49n203

T ableB .16 Somc


. data acqwstuon 1og (03/24/1997)
Run# BLK Range X h Sonic Bldg. start U,mph
M49n204 5 30m/s 11' 1"-00" H 330 13:16:13 20
M49n205 .. .. 12 .. . .. 13:33:09 20
M49n206 .. 13 .. .. 345 13:50:45 23
.. ..
n

M49n207 ., 14 .. 330 14:07:42 18


M49n208 .. . 15 .. . 345 14:25:25 18
M49n209 .. . 17' .. .. .. 14:42:14 22
M49n210 .. .. 19 .. .. .. 14:59:17 20
M49n211 .. .. 17' 0'-06" 330 15:22:07 22
M49n212 .. .. 19' .. ..v" .. 16:59:37 24
M49n213 .. .. 0' 4"-02,. .. 17:22:24 20
M49n215 .. . 1' . .. ... 17:43:43 23
M49n216 .. .. 3' .. .. 18:02:26 22
M49n217 .. .. 5' .. .. . 18:20:18 21
M49n218 .,. .. 1' .. .. 18:38:11 20
M49n219 .. .. 9' ..'' .. .. 18:56:04 20
138

Date: 3/24/1997
Flow Visualization of conical vortices (M49n220-224); SB (M49n226)

Figure B.4 and Table B.17 give the experimental setup and details.

Table B.17 Sonic data acquisition log (3/24/1997)


Run# BLK Ran~e X h Sonic Bld2. start u.mnh
M49n220 5 30m/s - - v 30 20:07:07 21
M49n222 .. ., - - . .. 20:33:56 23
M49n223 .. .. - - . 20:50:56 24
..
n

M49n224 ,, .. - - .. 21:07:49 18

M49n226 .. .. - - v 3400 21:36:15 SB

Sonic

(ROOF)

Figure B.4 Schematics of experimental setup (M49n220-226)

L Sonic anemometer was placed about 2 ft from each edge, 4' -02" above the roof;
2. At the end of Run 223, the camera was knocked over by a gust over 60 mph. The sonic
was blown off its position by 22 to the right, which was found after tests were over, thus
sonic data must be corrected by this amount for runs 224 and 226, and probably for a
small portion of 223 toward the data run end.
3. The sonic data were bad for runs m49n222, 223, 224, with voltages being close to zero.
probable because the 30 m/s range was exceeded and the anemometer started to
malfunction.
139
Freguency-Res.ponse Comparison of Anemometers
Earlier, large-and-fast direction fluctuations were observed using the three-
component sonic anemometer, and it was suspected that the propeller type UVW
anemometers might not respond adequately to swift direction changes and wind gusts of
short duration. For comparison, the three-component sonic anemometer was setup side by
side with the UVW anemometer on the meteorological tower at 13ft. Figure B.5 shows
the experimental setup.

9/27/1997-9/28/1997: Building position for Runs M49n23l- 250 was 90.


M49n231 (9/27/1997), manually triggered run, started at 16:52:54
M49n232 - M49n250 (09/28/1997), automatic runs, triggered at 15 mph of one-minute
mean. Started at 7:06:26- (M49n232), ended at 12:46:03- (M49n250).

6' -04"'
t< >f

Figure B.5 Sketch of experimental setup for velocity measurement comparison


140

09/28/1997
Outstanding measurement of wind flow inside the SB
Tables B.l8- B23 list details of experimental setup for data runs M49n252-287.

Table 8.18 Sonic data acquisition log (09/28/1997)


Run# BLK Range X. b Sonic Bldg. start U,mpb
M49n252 2 30m/s 0 0'-06.., H 13SO 13:56:03 -
M49n253 .. , .. l' -00" . ,. 14:20:35 -
M49n255 . .. .. 1'-04" .. 150 14:48:54 -
M49n256 ., .. 0'-{)6n . . .. 15:13:44 -
M49n257 .. . .. 1' -08" .. .. 15:33:49 -
M49n258 .. .. .. 0'-06'"' .. .. 16:05:22 -
M49n259 .. .. .. 1'-0" . .. 16:24:04 -

Figure B.6 indicates the definition of instrumental u- v- w-components for data runs
M49n252-259.

D
Figure 8.6 Definition of instrumental u- v- w-components (M49n252-259}

By mistake, the sonic anemometer was not oriented as desired. The actual orientation is
shown in the above sketch. Again, the right-hand-rule applies.
141

10/05/1997
Data runs intended to find the upper boundary of the separation bubble.

Table B.19 Sonic data ac_guisition log (10/05/1997)


Run# BLK Range X h Sonic Bldg. start U,mpb
M49n261 2 30m/s 11' 4'-ll" v 240 16:01:29 -
M49n262 .. .. 13' .. ., 16:19:00 -
. ..
H

M49n263 .. 15' .. 16:37:13 -


.. .. ..
H

M49n264 .. 17' 255 16:57:37 -


M49n265 .. .. 19' .. . .. - -
Figure B. 7 indicates the definition of instrumental u- v- w-components for data runs
M49n261-265.

W+

D
Figure B.7 Definition of instrumental u- v- w-components (M49n261-265)
142
10/08/1997

Run#
M49n267

Figure B.8 indicates the definition of instrumental u- v- w-components for data run
M49n267.

W+

U+._l

D
Figure B.8 Definition of instrumental u- v- w-components (M49n267)

10/12/1997
Runs M49n269 - 272

T able B.21 S omc


. data acquJSttlon 1og (10/12/1997)
Run# BLK Rang X b Sonic Bldg. start U.mpb
M49n269 2 30m/s 19' 4'-11" v 15 12:27:26 -
M49n270 .. .. 21' .. .. .. 12:46:34 -
M49n271 .. .. 21' .. .. .. 13:04:28 -
M49n272 . .. 23' .. .. .. 13:22:20 -
Figure B.9 indicates the definition of instrumental u- v- w-components for data runs
M49n269-272.

D
Figure B.9 Definition of instrumental u- v- w-components (M49n269-272)
143

10/12/1997
Runs M49n274- 283

Table B.22 Sonic data acquisition log (10/12/1997)


Run# BLK Ran~ X h Sonic Bid~. start U,mph
M49n274 2 30m/s 21' 0'-06' H oo 13:49:17
.....
~

M49n275 .. .. 23' . .. 14:06:08


....
~

M49n276 .. .. 25' . 14:23:58 ~

M49n277 .. .. 27'" .. ., 14:41:07 -


M49n278 .. .. 29' .. .. .. 14:58:04 -
M49n280 .. .. 29' 1'..()4'' .. 345 15:22:48 -
M49n281 .. .. 27' .. .. .. 15:40:21 -
M49n282 ... .. 23' .. ... oo 15:58:11 -
M49n283 ... .. 19' .. .. .. 16:16:30 -
Figure B.lO indicates the definition of instrumental u- v- w-components for data runs
M49n274-283.

U+ ~ U+ ---.,.
.-v+ V+ t

f ol f ol
(M49n274 - 275) (M49n276 - 283)

Figure B.lO Definition of instrumental u- v- w-components (M49n274-283)


144
Ta bleB .23 Sorne
. data acqwstuon 1og (10/1211997)
Run# BLK Rang X h Sonic Bld_g. Start U.mjlh
M49n284 2 30m/s 19' 3'-08" v oo 16:38:46 -
M49n285 , .. 23' .. .. .. 16:55:57 -
M49n286 ... .. 27' .. ... .. 17:13:37 -
M49n287 , n 12' 5'..()9,.
"'
... - -
Figure B.ll indicates the definition of instrumental u- v- w-components for data runs
M49n284-287.

W+
"\+
+-+U+ V+ .J
ol
(M49n284-286)
DI
(M49n287)

Figure B.ll Definition of instrumental u- v- w-components (M49n284-287)


APPENDIXC
PRESSURE COEFFICIENT DATA
FOR SELECfED TAPS FROM MODE-15

The configuration of Mode-15 is identical with that of Mode-49. In these two modes,
the test building was in its basic form and without any appurtenances attached.
The selected taps for which their pressure coefficients are provided are taps from the
heavily instrumented roof corner, and taps along the short roof-axis. They are taps 50901,
50205, 50505, 50905, 50209, 50509, 50909, 50123, 50523, 50823, 51123 and 51423.
The mean, rms, min. and max. pressure coefficients are given versus the wind angle of
attack. Similar figures for taps 50101 and 50501 were provide in the text.

145
146

AOA vs. Mean

0 45 90 135 180 225 270 315 360


AOA

AOA vs. RMS

Ol 1 +---+-----f--+----+---,..,---r:""~~-~-----+----i

il 0.5 +---+-----f-~....
0~~~~~~-L~~~~
0 45 90 135 180 225 270 315 360
AOA

AOA vs. Min

a -5+-----+---~~
.....
~ -10+---+----~--~--~~~~--~~--+---~

-15~--------~--~----~--~--~---------

0 45 90 135 180 225 270 315 360


AOA

AOA vs. Max

0 45 90 135 180 225 270 315 360


AOA

Rgure C.1 Tap 50901 pressure coefficient data


147

AOAvs. Mean

~~~~--~~-iM..~==~--~~~ 1~1
0 45 90 135 180 225 270 315 360
AOA

AOAvs. RMS

0 45 90 135 180 225 270 315 360


AOA

AOAvs. Min

0 45 90 135 180 225 270 315 360


AOA

AOAvs. Max

::E 0.5
=
0 + - -..... :;._=-.....;_+---Z....t

0 45 90 135 180 225 270 315 360


AOA

Figure C.2 Tap 50205 pressure coefficient data


148

AOAvs. Mean

c: .0.5 +---+---------i---=-::-11~
I
~ -1+----~-~-----

0 45 90 135 180 270 315 360


AOA

AOAvs.RMS

0.8+-----+-----+-----+-----~----r-~~~----r---~

~ 0.6
a: 0.4 -t----+----+-----::":11111
02.,~~~~~~~~

45 90 135 180 225 270 315 360


AOA

AOAvs.. Min
o~~rr~~~,-----~----~----~----~----~----~
-2 -t-=-~~~~~..__,
.5-4+-----+----t---~
~
~+-----+-----~--~~----~----~----~----~----~
~~----~----~----~----~----~----~----~----~
0 45 90 135 180 225 270 315 360
AOA

0 45 90 135 180 225 270 315 360


AOA

Figure C.3 Tap 50505 pressure coeffiCient data


149

AOAvs. Mean

o . .~~~~~~t----r--~~.-~~~~~
c: ...().5 ~;::;..._....::111111. . . .' - =...........~
I -1+-----r-----~--~
:E 1.5 +---+----+------1~~~~-__,._ _-+__,;:::..---+-----t
-2+-----+-----r----1~------~------~-----+~---+------t

-~5._--------------~~--~----~----~----------~
0 90 135 180 225 270 315 360
AOA

AOAvs. RMS
1.2

fiJ 0.8 +---+---+---+---+----:::.~+-----+----+----1


::s 0.6 +---+----+---"!1
a: 0.4 +----+---+--.~
0.2~~~~--~~~~-r~~i---~

45 90 135 180 225 270 315 360


AOA

AOAvs. Min
OT-~r-~----~----~----~----~----~----~----~
-2 ..pr::----'11111~~~~..... -::;:IE"%=::--+----""'".:'::-

.s -4 +---+----+---'~
::s
~+-----+----+----+----4~~~-+-----~~--~---~

-8~----~----~----~----~---~----~----~-----
0 45 90 180 225 270 315 360
AOA

0 45 90 135 180 225 270 315 360


AOA

Figure C.4 Tap 50905 pressure coefficient data


150

AOAvs. Mean
0~~--~--~~----~----~----~----~----~--~
..0.5 +--.-;:::~~IIIFZ~~+--~
i -1
:I -1.5 +---+---+---+---+--.::IIWII!i---r4!JIK-+----+------I
~+-----+----+-----+-----+---.::11~~-+------+------1

-~5L-----L-----L-----~----~----~----~----~--~
0 45 90 135 180 225 270 315 360
AOA

AOAvs.RMS

0
o.s +----+-----4-----....._---+-~
:E 0.6 +----+----4------+----~~--~----'--:-t-----+------t
a: 0.4 +------+------+---
0.2 ~~--=-~::"'"""":1:=----t---::E:~illlll
o~~~~----~~--~---~---~---~---~----~
0 45 90 135 180 225 270 315 360
AOA

AOAvs. Min
0~--~~~--~----~----~----~-----r----~----~
-2~~-w~~~...._~-.

c: -4+----+------+--___,.,.~--r.-;..;:;;

2 ~T-----T-----;----;-----~~~~~~~-----r----~
~+----+-----+------+-----+---~~----~-----+---~

-10~----~----~----~----~----~----~----~----~
0 45 90 135 180 225 270 315 360
AOA

AOAvs. Max

0 45 90 135 180 225 270 315 360


AOA

Figure C.S Tap 50209 pressure coefficient data


151

AOAvs. Mean
0~----~----~----------~--~----~----~----~
..0.2
..0.4.f---~
I -o.s + - - - + - - - + - - -
:.:~ ..0.8 -1----+----1----~

1+-----+-----+-----~~ ~----~~~~
-1.2 +----f----..J-----J.---Z.--ii-----4-._..Za-J-.:.......--J.--~
-1.4 ..L-----'-----'----"----1----___...IL____;;____._ _---~._ _---J
0 45 90 135 180 225 270 315 360
AOA

AOAvs.RMS
1~----~---~----~----~----~----~----~----~
0.8+-----4-----~----~----~----~~-.--+-----+----~

0 ~6+-----4-----~----~----~----~~~~----+-------
:E
a: 0.4 - 1 - - - - - t - - - - - + - - -
0.2 ..f.IIIE-----.,~::--=~-+-~~

45 90 135 180 225 270 315 360


AOA

AOAvs. Min

0 45 90 135 180 225 270 315


AOA

AOAvs.Max

M
0.5
0+--~~~--+-----~

-o.5~-----'-------'-----~----~----"------L-----~----
o 45 90 135 180 .225 270 315 360
AOA

Figure C.6 Tap 50509 pressure coeffiCient data


152

=-o.5
0
.... ~~.
AOAvs. Mean

... ~~ ~

I
:IE -1
IMeanl
~
-1.5
0 45 90 135 180 225 270 315 360
AOA

AOAvs.RMS
3
2.5 .. !
2 I
UJ !

..
:1 1.5

~

0.5
0
1

0 45
.. 90
....
135
....180
~ 225 270 315
j
!
I
i

360
I+RMSI

AOA

.5-4
:1
0
-2

-6
&.. I .~. ~. 41 ~

~ .
~
AOAvs. Min

..
I
I+Minl
-a
~
l
0 45 90 135 180 225 270 315 360
AOA

AOAvs.Max
5
~ ;
4
..
)( 3
.
j

-
!
I+Maxj

.-~
:12
i
1
0
0 45
.....

90
~..-.....
135
Jf!_
180
......

225 270 315 360


AOA

Figure C.7 Tap 50909 pressure coefficient data


153

AOAvs. Mean
o~--~-=~~~--~--~~----~--~----~--~
-o.5
i -1+---~--------+----~----+-~~

:1 -1.5 +---~--------+--~---+-----t---z;:::;:p~..;.._-+-------t

~+---~--------+-------~----------+--------t----+~~-+-------t

_,5----~----~----~----~---------------------
0 45 90 135 180 225 270 315 360
AOA

AOAvs.RMS

0.8 +------l~---+---+--------t--------+--___. ...__~-----i


c:n 0.6 +--~~---+---+---t----t--.....=...tll~ ~-----~-----------~
:1
= 0.4 +---------+-::----+-------~--+-.
0.2

45 90 135 180 225 270 315 360


AOA

AOAvs. Min

~~~~jili~----~:::r===r=::J==~
:ic ~t=====t====j=====t===~=r~~~~~~~~~!===:j
-6
~+----~--~--------+---+----t-~~~~-~----~

-10------~--~----~----~----~----~----~--~
0 45 90 135 180 225 270 315 360
AOA

0 45 90 135 180 225 270 315 360


AOA

Figure c.a Tap 50123 pressure coefficient data


154

AOAvs.Mean
o~~~~~r2~~--~~~~~r-~----~----~--~
-o.5 ~~-~~U~IQ!!!!!!!
I _,
:::E -1.5 +----+----+---+---+-------+---+--~+------!

~+-----+-----+-----+-----+-----+-----~~--+---~

-~5._---~----~----~----~----~----~----~--~
0 45 90 135 180 225 270 315 360
AOA

AOAvs. RMS
1~----~----~----~----~-------~-----r-------~----~
0.8 + - - - - - + - - - - + - - - - + - - - - + - - - - - + - - -............- - + - - - - - i

= 0.6
a: 0.4 +----+-::----+---------+-----+-~:4
-.r.p
,.
02,.~~~. .~~.a~ ....~~----~---1~-~
0+-------+--------+---~-+-~---+-----+-----~----+-------~

0 45 90 135 180 225 270 315 360


AOA

AOAvs. Min

c -~ +--~~~~~~~~~~~~~!~~~~i~aii~l=~~
-4
5 ~+-----~-----+-----+----~----+-~~~.-~----~
~+-----+--------~----~--~-----+--~----~----~
-10~------~-------~-------~----~----~----~----~---~
0 45 90 135 180 225 270 315 360
AOA

AOAvs. Max
1.5 .,.----~----..,---~--...,.---...,.---..,..-----,------,

1~~-+-----~----~-~~~--~----~~--~---~

= 0.50+--...;.
:::E
..0.5+-----+-----~--~----~----~----~~L-~---~

-1------------------------~----~-------~----~---~
0 90 135 180 225 315
AOA

Figure C.9 Tap 50523 pressure coefficient data


155

AOA vs. Mean


0.5 -----r---r-----r--.....-----r---r----r-----,

&::: -0.~ .iii


I
:E -1.5
-1 + - - - - + - - + - - - - + - - + - - - - + -
+----+--+---+--+---+--~:::----+-----!
J Mean J

-2+----+---+----+--+---+---+-~--+------1
-2.5 ...___ __.__ _.....___ __.__ __.___ __.__ _ _ _ ___.__ ___,

0 45 90 135 180 225 270 315 360


AOA

AOAvs. RMS
1 ----~--~~--~--~--~~--~--~--~
0.8 +----+-----it----+----+--t---~~--+-----1
I'll 0.6 +---+---1------+---+-----:~lr--:::-~~......___---1
r! 0.4 +----1-----+--+-----+----z~

0 45 90 135 180 225 270 315 360


AOA

AOA vs. Min


a~--~----~--~~--~--~--~---~--~
d -2
""* 4+----+--+------+----~~~~~

6._--~--------~----~--~--~~~~--~
0 45 90 135 180 225 270 315 360
AOA

AOA vs. Max

45 90 135 180 225 270 315 360


AOA

Figure C.1 o Tap 50823 pressure coefficient data


156

AOAvs. Mean
0

i
~
~.5+---~~~~~~~~----~~--~

-1.5
1+--------+-----+-----+-----+-----+-----+-----+---~
+---------+---+----+----+---+-----+----+----i
~~----~----~----~----~----~----~----._--~
..

.
0 45 90 135 180 270 315 360
AOA

AOAvs.RMS
0.8 ..,...----..,.......--..,.......---,.-----,----..,......----T----T-----,
0.6 +-----+-------+-----t-----t----t----t-..-...:.r---t-----1
tn

~ 0.2
o. t;~:t~L~:~::l:::J~M~II~;t=~J
4
.~.~._""':16::~........~=--=--+--:---:
0+-----t----~----~~--~----~------~----~----~
0 45 90 135 180 225 270 315 360
AOA

AOAvs. Min
o~----~----~----~~~~-----~~-----~----~--~
-2
=~+------r----~--~----~----~~--~~
~
~+------+------t------+----~---~------r-~--+---~

-8~----~----~----~----~----~~----~----~--~
0 45 90 135 180 225 270 315 360
AOA

AOAvs.Max
2~-----------r------r-----~-----r----r----~---~---~

1.5 +-----t-----+----+------+-----+-------+-----+----1

0.5
0+---__...:ll"f-----~--.::;.._,.;;,~

0 45 90 135 180 225 270 315 360


AOA

Rgure C.11 Tap 51123 pressure coefficient data


157

AOAvs. Mean

0 45 90 135 180 225 270 315


~
360
IMeanl

AOA

AOAvs.RMS
0.8

Cl)
0.6
:1 0.4
a:
0.2 IRMsj
0
0 45 90 135 180 225 270 315 360
AOA

AOAvs. Min

0 45 90 135 180 225 270 315


~360
IMinj

AOA

AOAvs. Max
2
1.5
)( 1

:1 0.5 IMaxl
0
-o.5
0 45 90 135 180 225 270 315 360
AOA

Figure C.12 Tap 51423 pressure coefficient data


APPENDIXD
DERIVATION OF EQUATION FOR PRESSURE PREDICI10N

The Cp(a) in the text was based on wind measurement on the meteorological tower.
In the following derivation for the equation used to predict Cp(t) from wind measured at a
point close to the test building~ reference is made to the definition shown in the following
sketch. The parameters with a bar above are averaged values.

Figure D.l Definition sketch


Given AP(t) =P(t)-P.7
Cp(t) = L\P(t)/q OCy and Cp = AP nr CIC

Assume that time dependence of AP is only a function of time dependence of freestream,

LlP(t) = Cp *Qoc(t).
To avoid the need to compensate for time lag between measured qoc:(t) and corresponding
pressure fluctuations, use the approximation,
Qoc{t) = {q Jq)*q(t),

where the quantity inside the parenthesis is an interference factor.


It follows then
AP(t) = Cp *(q Jq)*q(t).
The corresponding pressure coefficient is
Cp(t) =AP(t)/ q ...= Cp *( qJq)*[q(t)/ q.].
That is, Cp(t) = Cp *(q(t)/ q) = Cp *[u(t)/u] 2

158

Вам также может понравиться